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

\RS@ifundefined

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

Shachar Carmeli Department of Mathematics, Weizmann Institute of Science.    Tomer M. Schlank Einstein Institute of Mathematics, Hebrew University of Jerusalem.    Lior Yanovski Max Planck Institute for Mathematics.
Abstract

We introduce and study the notion of semiadditive height for higher semiadditive \infty-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 \infty-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 PrL\Pr^{L}, which we call modes. Using this theory, we introduce and study the universal stable \infty-semiadditive \infty-category of semiadditive height nn, and give sufficient conditions for a stable 11-semiadditive \infty-category to be \infty-semiadditive.

Refer to caption
Nurse measures height of child for Tecumseh Health study, 1959, HS15323, Alumni Association (University of Michigan) records, UM News Service, Bentley Historical Library, University of Michigan.

1 Introduction

1.1 Background & Overview

Chromatic homotopy theory springs from the deep and surprising connection between the \infty-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 πMU\pi_{*}\text{\emph{MU}} carries the universal formal group law. Formal group laws admit a notion of (pp-typical) height for every prime pp. This notion can be defined in terms of a certain sequence of classes vnπ2(pn1)MUv_{n}\in\pi_{2(p^{n}-1)}MU as follows: If v0,,vn1v_{0},\dots,v_{n-1} vanish, then the height is n\geq n, and if vnv_{n} is invertible, then the height is n\leq n. This algebraic filtration has a spectrum level manifestation in the form of the Morava KK-theories K(n)K(n), which are certain MU-algebras with the property πK(n)𝔽p[vn±1]\pi_{*}K(n)\simeq\mathbb{F}_{p}[v_{n}^{\pm 1}]. This suggests that the K(n)K(n)-s are concentrated at height exactly nn, and the corresponding Bousfield localizations SpK(n)Sp\operatorname{Sp}_{K(n)}\subseteq\operatorname{Sp} can then be considered as the “monochromatic layers” of the chromatic height filtration. The process of K(n)K(n)-localization can be loosely thought of as completion with respect to v0,,vn1v_{0},\dots,v_{n-1} followed by the inversion of vnv_{n}.

By the work of Hopkins, Devinatz, and Smith (see [HS98, Theorems 9 and 4.12]), the vnv_{n}-operations can be inductively lifted to finite spectra, without MU-module structure. More precisely, a finite pp-local spectrum FF is said to have type nn, if nn is the lowest integer for which the K(n)K(n)-localization of FF does not vanish. Given a type nn finite spectrum F(n)F(n), there exists a self map ΣdF(n)F(n)\Sigma^{d}F(n)\to F(n), which induces a power of vnv_{n} on K(n)K(n)-homology. The cofiber of this self map is then a type (n+1)(n+1) spectrum. This procedure allows us to construct a sequence of F(n)F(n)-s as iterated Moore spectra, which we may suggestively write as follows:

F(n)𝕊/(pr0,v1r1,,vn1rn1).F(n)\coloneqq\mathbb{S}/(p^{r_{0}},v_{1}^{r_{1}},\dots,v_{n-1}^{r_{n-1}}).

Just as localization with respect to 𝕊/pr\mathbb{S}/p^{r} (for any rr) has the effect of pp-completion, one can think of the localization with respect to F(n)F(n), as completion with respect to v0,,vn1v_{0},\dots,v_{n-1}. Furthermore, localization with respect to the spectrum T(n)=F(n)[vn1]T(n)=F(n)[v_{n}^{-1}], can be thought of as completion with respect to v0,,vn1v_{0},\dots,v_{n-1}, followed by the inversion of vnv_{n}. It is known that the K(n)K(n)-localization factors through the T(n)T(n)-localization, and that they coincide for MU-modules (and also in general when n=0,1n=0,1). Furthermore, the localizations SpT(n)\operatorname{Sp}_{T(n)} 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 SpK(n)SpT(n)\operatorname{Sp}_{K(n)}\subseteq\operatorname{Sp}_{T(n)} is strict for n2n\geq 2 is open (known as the Telescope Conjecture), both candidates for the “monochromatic layers” play a pivotal role in homotopy theory.

The localizations SpK(n)\operatorname{Sp}_{K(n)} and SpT(n)\operatorname{Sp}_{T(n)} 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 11-semiadditivity, and vastly generalize it by showing that the \infty-categories SpK(n)\operatorname{Sp}_{K(n)} are \infty-semiadditive. In turn, this is exploited to obtain new structural results for SpK(n)\operatorname{Sp}_{K(n)}. In [CSY18, Theorem B], the authors extended on [HL13] by classifying all the higher semiadditive localizations of Sp\operatorname{Sp} with respect to homotopy rings. First, for all such localizations, 11-semiadditivity was shown to be equivalent to \infty-semiadditivity. Second, the telescopic localizations SpT(n)\operatorname{Sp}_{T(n)}, for various primes pp and heights nn, were shown to be precisely the maximal examples of such localizations (while the SpK(n)\operatorname{Sp}_{K(n)} 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 \infty-categories, which in the examples SpT(n)\operatorname{Sp}_{T(n)} and SpK(n)\operatorname{Sp}_{K(n)} reproduces the usual chromatic height nn, 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 \infty-category, which controls many aspects of its higher semiadditive structure, and the behavior of local systems on π\pi-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 \infty-categories, we show that higher semiadditive \infty-categories decompose completely according to the semiadditive height, which accounts for the monochromatic nature of the higher semiadditive localizations of Sp\operatorname{Sp}. Finally, building on the work of Harpaz [Har17], we introduce and study universal constructions of stable \infty-semiadditive \infty-categories of height nn, 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 SpT(n)\operatorname{Sp}_{T(n)}, 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 AA with respect to an \infty-category 𝒞\mathcal{C}, that allows “integrating” AA-families of morphisms between pairs of objects in 𝒞\mathcal{C} in a canonical way [HL13, Construction 4.0.7]. In particular, integrating the constant AA-family on the identity morphism of each object, produces a natural endomorphism |A||A| of the identity functor of 𝒞\mathcal{C}. We call |A||A| the 𝒞\mathcal{C}-cardinality of AA, and think of it as multiplication by the “size of AA” (the actual meaning of which depends on 𝒞\mathcal{C}).

An \infty-category 𝒞\mathcal{C} is called mm-semiadditive if every mm-finite space is 𝒞\mathcal{C}-ambidextrous. Our notion of semiadditive height is defined in terms of cardinalities of such spaces. For starters, let us begin with a 0-semiadditive (i.e. semiadditive) pp-local \infty-category 𝒞\mathcal{C}. If pp is invertible in 𝒞\mathcal{C}, then 𝒞\mathcal{C} is rational and we consider it to be of “height 0”. In contrast, if all objects of 𝒞\mathcal{C} are pp-complete, we consider it to be of “height >0>0”. To proceed, let us assume that 𝒞\mathcal{C} is mm-semiadditive for some m0m\geq 0. In such a case, we can consider the 𝒞\mathcal{C}-cardinalities of Eilenberg-MacLane spaces:

p=|Cp|,|BCp|,|B2Cp|,,|BmCp|.p=|C_{p}|,|BC_{p}|,|B^{2}C_{p}|,\dots,|B^{m}C_{p}|.

The definition of semiadditive height uses the maps |BnCp||B^{n}C_{p}| in a manner which is analogous to how the vnv_{n}-self maps are used in the definition of the chromatic height:

Definition (Semiadditive Height, 3.1.6, 3.1.11).

For every 0nm0\leq n\leq m, we write

  1. (1)

    Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n, if |BnCp||B^{n}C_{p}| is invertible in 𝒞\mathcal{C}.

  2. (2)

    Ht(𝒞)>n\mathrm{Ht}(\mathcal{C})>n, if 𝒞\mathcal{C} is complete with respect to |Cp|,|BCp|,,|BnCp||C_{p}|,|BC_{p}|,\dots,|B^{n}C_{p}|111by 3.1.9, Ht(𝒞)>n\mathrm{Ht}(\mathcal{C})>n, if and only if 𝒞\mathcal{C} is |BnCp||B^{n}C_{p}|-complete..

  3. (3)

    Ht(𝒞)=n\mathrm{Ht}(\mathcal{C})=n, if 𝒞\mathcal{C} is of height n\leq n and >n1>n-1.

To show that the semiadditive height of SpT(n)\operatorname{Sp}_{T(n)} and SpK(n)\operatorname{Sp}_{K(n)} is indeed nn, we need to get a handle on the T(n)T(n)-local and K(n)K(n)-local cardinalities of the Eilenberg-Maclane spaces BkCpB^{k}C_{p}. In [CSY18, Lemma 5.3.3], we have already shown that

|BkCp|=p(n1k)π0En,|B^{k}C_{p}|=p^{\binom{n-1}{k}}\quad\in\quad\pi_{0}E_{n},

for the \infty-category of K(n)K(n)-local EnE_{n}-modules. Thus, this \infty-category is of height nn. Since tensoring with EnE_{n} is conservative on K(n)K(n)-local spectra, this also readily implies that SpK(n)\operatorname{Sp}_{K(n)} is of height nn. However, to show that SpT(n)\operatorname{Sp}_{T(n)} is of height nn, one has to know that the map π0𝕊T(n)π0𝕊K(n)\pi_{0}\mathbb{S}_{T(n)}\to\pi_{0}\mathbb{S}_{K(n)}, induced by K(n)K(n)-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 SpT(n)\operatorname{Sp}_{T(n)} is of height nn as well.

The notion of semiadditive height allows us to contextualize various aspects of the \infty-categories SpK(n)\operatorname{Sp}_{K(n)} and SpT(n)\operatorname{Sp}_{T(n)} pertaining to the chromatic height. At the bottom of the hierarchy, the \infty-category SpT(0)=SpK(0)=Sp\operatorname{Sp}_{T(0)}=\operatorname{Sp}_{K(0)}=\operatorname{Sp}_{\mathbb{Q}} can be shown to be \infty-semiadditive by elementary arguments. This is strongly related to the fact that all connected π\pi-finite spaces are \mathbb{Q}-acyclic and the cardinality of any (non-empty) π\pi-finite space AA is invertible. Thus, the higher semiadditive structure of Sp\operatorname{Sp}_{\mathbb{Q}} is in a sense “trivial”. The higher semiadditivity of SpK(n)\operatorname{Sp}_{K(n)} and SpT(n)\operatorname{Sp}_{T(n)} for n1n\geq 1 is more subtle precisely because not all connected π\pi-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 𝒞\mathcal{C} be an nn-semiadditive pp-local \infty-category, which admits all π\pi-finite limits and colimits. If 𝒞\mathcal{C} is of height n\leq n, then

  1. (1)

    𝒞\mathcal{C} is \infty-semiadditive.

  2. (2)

    For every (n1)(n-1)-connected nilpotent π\pi-finite space AA, the map |A||A| is invertible.

  3. (3)

    For every nn-connected π\pi-finite space AA and X𝒞X\in\mathcal{C}, the fold map AXXA\otimes X\to X is invertible.

  4. (4)

    For every principal fiber sequence of π\pi-finite spaces

    FAB,F\to A\to B,

    if FF is (n1)(n-1)-connected and nilpotent, then |A|=|F||B||A|=|F|\cdot|B|.

Informally speaking, A states that the invertibility of |BnCp||B^{n}C_{p}| has the effect of “trivializing” the higher semiadditive structure at levels n\geq n. In particular, it shows that it exists, which is point (1). From point (2), we deduce that having height n\leq n implies having height n+1\leq n+1, 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 n\leq n, 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 𝒞=SpT(n)\mathcal{C}=\operatorname{Sp}_{T(n)}. Finally, point (4) can be used to reduce the computation of the 𝒞\mathcal{C}-cardinalities of nilpotent π\pi-finite spaces to those of nn-finite ones, under the assumption that 𝒞\mathcal{C} is of height n\leq n. The case n=0n=0 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 Bn1CpptBnCpB^{n-1}C_{p}\to\operatorname{pt}\to B^{n}C_{p}, is precisely the obstruction for 𝒞\mathcal{C} to have height n1\leq n-1.

In their work on algebraic KK-theory of ring spectra, Ausoni and Rognes have discovered a phenomena which they dubbed “chromatic redshift”. Roughly speaking, it is the tendency of K(R)K(R), which is a spectrum constructed from the \infty-category of perfect RR-modules, to be of chromatic complexity larger by one, than the ring spectrum RR (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 XX in an \infty-semiadditive \infty-category 𝒞\mathcal{C} is of height n\leq n for some nn, if |BnCp||B^{n}C_{p}| acts invertibly on XX and of height >n>n, if it is complete with respect to |Cp|,|BCp|,,|BnCp||C_{p}|,|BC_{p}|,\dots,|B^{n}C_{p}|. Second, we exploit the fact that the \infty-category Cat

-
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
, of \infty-semiadditive \infty-categories and π\pi-finite colimit preserving functors, is itself \infty-semiadditive. Thus, given an \infty-semiadditive \infty-category 𝒞\mathcal{C}, we can consider the height of 𝒞\mathcal{C} being lower equal (resp. greater than) nn, as an object of Cat

-
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
which we shall denote by ht(𝒞)n\mathrm{ht}(\mathcal{C})\leq n (resp. ht(𝒞)>n\mathrm{ht}(\mathcal{C})>n).

Theorem B (Semiadditive Redshift, 3.3.2).

Let 𝒞\mathcal{C} be an \infty-semiadditive \infty-category. We have that Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n (resp. Ht(𝒞)>n\mathrm{Ht}(\mathcal{C})>n), if and only if ht(𝒞)\mathrm{ht}(\mathcal{C}) n+1\leq n+1 (resp. ht(𝒞)>n+1\mathrm{ht}(\mathcal{C})>n+1).

The higher semiadditive structure of Cat

-
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
is essentially given by taking colimits over π\pi-finite spaces. Hence, B is closely related to point (3) of A. As a concrete example, we can consider for a T(n)T(n)-local ring spectrum RR, the \infty-category of T(n)T(n)-local left RR-modules. The space of objects of this \infty-category is a commutative monoid for the direct sum operation. Moreover, the higher semiadditivity of the \infty-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 n+1\leq n+1 in the \infty-category of higher commutative monoids. In a future work, we shall investigate the implications of this to the chromatic redshift in algebraic KK-theory in the sense of Ausani-Rognes.

Our main interest in the notion of higher semiadditivity is in its application to stable \infty-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 𝒞\mathcal{C} be a stable idempotent complete mm-semiadditive \infty-category for some mm\in\mathbb{N}. There is a canonical equivalence

𝒞𝒞0××𝒞m1×𝒞>m1,\mathcal{C}\simeq\mathcal{C}_{0}\times\cdots\times\mathcal{C}_{m-1}\times\mathcal{C}_{>m-1},

were 𝒞0,,𝒞m1\mathcal{C}_{0},\dots,\mathcal{C}_{m-1} and 𝒞>m1\mathcal{C}_{>m-1} are the full subcategories of objects of height 0,,m10,\dots,m-1 and >m1>m-1 respectively.222We also treat the case m=m=\infty, 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 \infty-categories 𝒞\mathcal{C} of height exactly nn. In [HL13, Section 5.4] it is shown that the behavior of local systems of K(n)K(n)-local spectra on a π\pi-finite space AA, strongly depends on the level of connectedness of AA compared with nn. We show that some of these results hold for general stable \infty-categories 𝒞\mathcal{C} of height exactly nn. First of all, from A(3), it can be deduced that for an nn-connected π\pi-finite space AA, the inclusion functor 𝒞𝒞A\mathcal{C}\hookrightarrow\mathcal{C}^{A} of constant local-systems is fully faithful. The right orthogonal complement 𝒞𝒞A\mathcal{C}^{\perp}\subseteq\mathcal{C}^{A} consists of local-systems whose global sections object (i.e. limit over AA) vanishes. We prove the following:

Theorem D (Semisimplicity, 4.3.2).

Let 𝒞\mathcal{C} be a stable \infty-semiadditive \infty-category such that Ht(𝒞)=n\mathrm{Ht}(\mathcal{C})=n, and let AA be an nn-connected π\pi-finite space. There is a canonical equivalence 𝒞A𝒞×𝒞\mathcal{C}^{A}\simeq\mathcal{C}\times\mathcal{C}^{\perp}.

This result can be seen as a generalization of the “semisimplicity” of SpK(n)\operatorname{Sp}_{K(n)}-valued local systems on nn-connected π\pi-finite spaces (compere [Lurb]). We also provide an explicit formula for the composition 𝒞A𝒞𝒞A\mathcal{C}^{A}\twoheadrightarrow\mathcal{C}\hookrightarrow\mathcal{C}^{A}, as a “symmetrization” of the action of the \infty-group G=ΩAG=\Omega A. Intuitively, the “order” of GG, by which one has to divide, is precisely the 𝒞\mathcal{C}-cardinality of GG, which is invertible by the assumption on the height of 𝒞\mathcal{C} and the connectivity of AA (A).

Based on the classification of higher semiadditive localizations of Sp\operatorname{Sp} with respect to homotopy rings in [CSY18, Theorem B], the authors proposed the conjecture that every stable pp-local presentable 11-semiadditive \infty-category is in fact \infty-semiadditive [CSY18, Conjecture 1.1.5]. In this paper, we prove a partial result in the direction of this conjecture. Given a stable pp-local presentable \infty-category 𝒞\mathcal{C}, we say that an object X𝒞X\in\mathcal{C} is of finite stable height if there exists a non-zero finite pp-local spectrum FF, such that FX=0F\otimes X=0. We also denote by 𝒞st𝒞\mathcal{C}_{\infty^{\mathrm{st}}}\subseteq\mathcal{C} the full subcategory of objects Y𝒞Y\in\mathcal{C}, for which Map(X,Y)=pt\operatorname{Map}(X,Y)=\operatorname{pt} for all XX of finite stable height.

Theorem E (Bounded Bootstrap, 5.5.17).

Let 𝒞\mathcal{C} be a stable pp-local presentable \infty-category. If 𝒞\mathcal{C} is 11-semiadditive and 𝒞st=0\mathcal{C}_{\infty^{\mathrm{st}}}=0, then it is \infty-semiadditive. Moreover, in this case 𝒞n𝒞n\mathcal{C}\simeq\prod_{n\in\mathbb{N}}\mathcal{C}_{n}.

The condition 𝒞st=0\mathcal{C}_{\infty^{\mathrm{st}}}=0 is satisfied if for example for every X,Y𝒞X,Y\in\mathcal{C}, the mapping spectrum hom(X,Y)\hom(X,Y) is LnfL_{n}^{f}-local for some integer nn. 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 \infty-category PrL\Pr^{L} of presentable \infty-categories and colimit preserving functors. Moreover, he showed that many familiar properties of presentable \infty-categories can be characterized as having a (necessarily unique) module structure over certain idempotent algebras in PrL\Pr^{L} [Lura, Section 4.8.2]. We call such idempotent presentable \infty-categories modes. This notion was also considered in [GGN16] from the perspective of smashing localizations of PrL\Pr^{L}. Given a mode \mathcal{M}, it is a property of a presentable \infty-category 𝒞\mathcal{C} to have a structure of a module over \mathcal{M}. 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 Sp\operatorname{Sp}. Consequently, every stable presentable \infty-category is canonically enriched in Sp\operatorname{Sp} and colimit preserving functors between stable presentable \infty-categories preserve this enrichment. This structure naturally plays a significant role in the study of stable \infty-categories. In [Har17, Lemma 5.20], Harpaz showed that mm-semiadditivity is similarly characterized by having a module structure over the idempotent algebra CMonm\operatorname{CMon}_{m} of mm-commutative monoids. The case m=0m=0 recovers the usual \infty-category of commutative (i.e. 𝔼\mathbb{E}_{\infty}) monoids in spaces, which classifies ordinary semiadditivity. The mapping spaces of an mm-semiadditive \infty-category obtain a canonical mm-commutative monoid structure, by analogy with the Sp\operatorname{Sp}-enrichment of stable \infty-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 \infty-categories. First, by the general theory of modes, CMonmSp\operatorname{CMon}_{m}\otimes\operatorname{Sp} is also a mode, which classifies the property of being at the same time stable and mm-semiadditive. Furthermore, using C, we show:

Theorem F (5.3.6).

For every n0n\geq 0, 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, pp-local, \infty-semiadditive and of height nn.

It is natural to compare n with SpT(n)\operatorname{Sp}_{T(n)}, which is in a sense the universal pp-local height nn localization of spectra. Since SpT(n)\operatorname{Sp}_{T(n)} is also \infty-semiadditive and of semiadditive height nn, the theory of modes implies the existence of a unique colimit preserving symmetric monoidal functor L:nSpT(n)L\colon{}_{n}\to\operatorname{Sp}_{T(n)}. In the case n=0n=0, the functor L:0SpT(0)L\colon{}_{0}\to\operatorname{Sp}_{T(0)} is an equivalence and hence 0Sp{}_{0}\simeq\operatorname{Sp}_{\mathbb{Q}} (5.3.7). In general, we show that LL exhibits SpT(n)\operatorname{Sp}_{T(n)} as a smashing localization of n in the sense that LL admits a fully faithful right adjoint SpT(n)n\operatorname{Sp}_{T(n)}\hookrightarrow{}_{n} and there is a canonical isomorphism LX𝕊T(n)XLX\simeq\mathbb{S}_{T(n)}\otimes X for all XnX\in{}_{n} (5.5.14). For n1n\geq 1, the \infty-category n also resembles SpT(n)\operatorname{Sp}_{T(n)} in that the unique colimit preserving symmetric monoidal functor Spn\operatorname{Sp}\to{}_{n} vanishes on all bounded above spectra (5.3.9), and that the right adjoint of the unique colimit preserving symmetric monoidal functor 𝒮n\mathcal{S}\to{}_{n} is conservative (5.3.10). We consider n to be a natural extension of SpT(n)\operatorname{Sp}_{T(n)}, which is a universal home for phenomena of height nn.

In a previous draft of this paper, we proposed the conjecture that for every n0n\geq 0, the unique colimit preserving symmetric monoidal functor L:nSpT(n)L\colon{}_{n}\to\operatorname{Sp}_{T(n)} is an equivalence. However, this conjecture was soon disproved by Allen Yuan already in the case n=1n=1. More precisely, using the Segal Conjecture (now a theorem [Car84]), he has constructed a higher commutative monoid structure of height 11 on the pp-complete sphere, as an object of the \infty-category of pp-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 \infty-categories, but also to enforce them in a universal way. For every mode \mathcal{M} and a presentable \infty-category 𝒞,\mathcal{C}, we can view 𝒞\mathcal{M}\otimes\mathcal{C} as the universal approximation of 𝒞\mathcal{C} by a presentable \infty-category which satisfies the property classified by \mathcal{M}. For example, Sp𝒞Sp(𝒞)\operatorname{Sp}\otimes\,\mathcal{C}\simeq\operatorname{Sp}(\mathcal{C}) is the stabilization of 𝒞\mathcal{C} [Lura, Example 4.8.1.23], and similarly, CMonm𝒞\operatorname{CMon}_{m}\otimes\,\mathcal{C} is the “mm-semiadditivization” of 𝒞\mathcal{C} [Har17, Corollary 5.18]. As alluded to above, the non-trivial gluing in the chromatic fracture square, prevents LnfSpL_{n}^{f}\operatorname{Sp} from being higher semiadditive for n1n\geq 1. Employing the additive pp-derivation δ\delta on the rings π0𝕊T(n)\pi_{0}\mathbb{S}_{T(n)} constructed in [CSY18, Section 4], we show that forcing even 11-semiadditivity on LnfSpL_{n}^{f}\operatorname{Sp}, has the effect of “dissolving the glue” in the chromatic fracture squares:

Theorem G (1-Semiadditive Splitting, 5.4.10).

For every n0n\geq 0, there is a unique equivalence of presentably symmetric monoidal \infty-categories

CMon1LnfSpk=0nSpT(k).\operatorname{CMon}_{1}\otimes L_{n}^{f}\operatorname{Sp}\simeq\prod_{k=0}^{n}\operatorname{Sp}_{T(k)}.

In particular, we see that forcing 11-semiadditivity on LnfSpL_{n}^{f}\operatorname{Sp} makes it automatically \infty-semiadditive. Noticing that both sides of G are modes, we can reinterpret it in terms of the properties classified by them. Namely, that every 11-semiadditive stable presentable \infty-category whose mapping spectra are LnfL_{n}^{f}-local, is \infty-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 π\pi-finite space in a higher semiadditive \infty-category and the corresponding notion of amenability. We then give several examples of these notions in various higher semiadditive \infty-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 \infty-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 \infty-categories. After a general discussion on recollement, we show that a stable higher semiadditive \infty-category splits as a product according to height (C). We then study local systems valued in a stable higher semiadditive \infty-category of height nn 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 \infty-categories. We show how algebraic operations on modes, such as tensor product and localization, translate into operations on the properties of presentable \infty-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 \infty-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. (1)

    We slightly diverge from [Lur09] and [Lura] in the following points:

    1. (a)

      We use the term isomorphism for an invertible morphism in an \infty-category (i.e. equivalence).

    2. (b)

      We denote by 𝒞𝒞\mathcal{C}^{\simeq}\subseteq\mathcal{C} the maximal \infty-subgroupoid of an \infty-category 𝒞\mathcal{C}.

    3. (c)

      We write Pr\Pr for the \infty-category of presentable \infty-categories and colimit preserving functors denoted in [Lur09] by PrL\Pr^{L}.

    4. (d)

      We denote by CatstCat\mathrm{Cat}_{\mathrm{st}}\subset\mathrm{Cat} the subcategory spanned by stable \infty-categories and exact functors. Similarly, we denote by PrstPr\Pr_{\mathrm{st}}\subseteq\Pr the full subcategory spanned by stable presentable \infty-categories.

  2. (2)

    We say that a space A𝒮A\in\mathcal{S} is

    1. (a)

      mm-finite for m2m\geq-2, if m=2m=-2 and AA is contractible, or m1m\geq-1, the set π0A\pi_{0}A is finite and all the fibers of the diagonal map Δ:AA×A\Delta\colon A\to A\times A are (m1)(m-1)-finite444For m0m\geq 0, this is equivalent to AA having finitely many components, each of them mm-truncated with finite homotopy groups..

    2. (b)

      π\pi-finite or \infty-finite, if it is mm-finite for some integer m2m\geq-2. For 2m-2\leq m\leq\infty, we denote by 𝒮m-fin𝒮\mathcal{S}_{m\text{-}\mathrm{fin}}\subseteq\mathcal{S} the full subcategory spanned by mm-finite spaces.

    3. (c)

      pp-space, if all the homotopy groups of AA are pp-groups.

  3. (3)

    Given an \infty-category 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty},

    1. (a)

      For every map of spaces A𝑞BA\xrightarrow{q}B, we write q:𝒞B𝒞Aq^{*}\colon\mathcal{C}^{B}\to\mathcal{C}^{A} for the pullback functor and q!q_{!} and qq_{*} for the left and right adjoints of qq^{*} whenever they exist.

    2. (b)

      Whenever convenient we suppress the canonical equivalence of \infty-categories 𝒮/pt𝒮\mathcal{S}_{/\operatorname{pt}}\overset{\sim}{\longrightarrow}\mathcal{S} by identifying a space AA with the terminal map A𝑞ptA\xrightarrow{q}\operatorname{pt}. In particular, for every \infty-category 𝒞\mathcal{C}, we write AA^{*} for qq^{*} and similarly A!A_{!} and AA_{*} for q!q_{!} and qq_{*} whenever they exist.

    3. (c)

      For every X𝒞X\in\mathcal{C} we write X[A]X[A] for A!AXA_{!}A^{*}X and denote the fold (i.e. counit) map by X[A]XX[A]\xrightarrow{\nabla}X. Similarly, we write XAX^{A} for AAXA_{*}A^{*}X and denote the diagonal (i.e. unit) map by XΔXAX\xrightarrow{\Delta}X^{A}.

  4. (4)

    Given a map of spaces q:ABq\colon A\to B, we denote for every bBb\in B, the homotopy fiber of qq over bb by q1(b)q^{-1}(b). We say that

    1. (a)

      an \infty-category 𝒞\mathcal{C} admits all qq-limits (resp. qq-colimits) if it admits all limits (resp. colimits) of shape q1(b)q^{-1}(b) for all bBb\in B.

    2. (b)

      a functor F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D} preserves qq-colimits (resp. qq-limits) if it preserves all colimits (resp. limits) of shape q1(b)q^{-1}(b) for all bBb\in B.

  5. (5)

    For every 2m-2\leq m\leq\infty,

    1. (a)

      by mm-finite (co)limits we mean (co)limits indexed by an mm-finite space.

    2. (b)

      We let Catm-finCat\mathrm{Cat}_{m\text{-fin}}\subset\operatorname{Cat}_{\infty} (resp. Catm-finCat\mathrm{Cat}^{m\text{-fin}}\subset\operatorname{Cat}_{\infty}) be the subcategory spanned by \infty-categories which admit mm-finite colimits (resp. limits) and functors preserving them.

    3. (c)

      For 𝒞,𝒟Catm-fin\mathcal{C},\mathcal{D}\in\mathrm{Cat}_{m\text{-fin}} (resp. Catm-fin\mathrm{Cat}^{m\text{-fin}}) we write Funm-fin(𝒞,𝒟)\operatorname{Fun}_{m\text{-}\mathrm{fin}}(\mathcal{C},\mathcal{D}) (resp. Funm-fin(𝒞,𝒟)\operatorname{Fun}^{m\text{-}\mathrm{fin}}(\mathcal{C},\mathcal{D})) for the full subcategory of Fun(𝒞,𝒟)\operatorname{Fun}(\mathcal{C},\mathcal{D}) spanned by the mm-finite colimit (resp. limit) preserving functors.

    4. (d)

      We let Cat

      -
      m
      Cat
      \mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}m}\subset\operatorname{Cat}_{\infty}
      be the subcategory spanned by the mm-semiadditive \infty-categories and mm-semiadditive (i.e. mm-finite colimit preserving) functors.

    5. (e)

      Given an \infty-operad 𝒪\mathcal{O}, we say that 𝒞Alg𝒪(Cat)\mathcal{C}\in\operatorname{Alg}_{\mathcal{O}}(\operatorname{Cat}_{\infty}) is compatible with 𝒦\mathcal{K}-indexed colimits for some collection of \infty-categories 𝒦\mathcal{K}, if 𝒞\mathcal{C} admits 𝒦\mathcal{K}-indexed colimits and every tensor operation :𝒞n𝒞\otimes\colon\mathcal{C}^{n}\to\mathcal{C} of 𝒪\mathcal{O} preserves 𝒦\mathcal{K}-indexed colimits in each variable.

    6. (f)

      An mm-semiadditively 𝒪\mathcal{O}-monoidal \infty-category is an 𝒪\mathcal{O}-monoidal mm-semiadditive \infty-category which is compatible with mm-finite colimits.

  6. (6)

    If 𝒞\mathcal{C} is a monoidal \infty-category and 𝒟\mathcal{D} is an \infty-category enriched in 𝒞\mathcal{C}, we write hom𝒟𝒞(X,Y)\hom_{\mathcal{D}}^{\mathcal{C}}(X,Y) for the 𝒞\mathcal{C}-mapping object of X,Y𝒟X,Y\in\mathcal{D}. We omit the subscript or superscript when they are understood from the context. In particular, when 𝒞\mathcal{C} is closed, hom𝒞(X,Y)\hom_{\mathcal{C}}(X,Y) means hom𝒞𝒞(X,Y)\hom_{\mathcal{C}}^{\mathcal{C}}(X,Y). For every \infty-category 𝒞\mathcal{C} we have hom𝒞𝒮(X,Y)=Map𝒞(X,Y)\hom_{\mathcal{C}}^{\mathcal{S}}(X,Y)=\operatorname{Map}_{\mathcal{C}}(X,Y).

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 π\pi-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 π\pi-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 \infty-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 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty}. A π\pi-finite map A𝑞BA\xrightarrow{q}B is called:

  1. (1)

    weakly 𝒞\mathcal{C}-ambidextrous if it is an isomorphism, or AΔqA×BAA\xrightarrow{\Delta_{q}}A\times_{B}A is 𝒞\mathcal{C}-ambidextrous.

  2. (2)

    𝒞\mathcal{C}-ambidextrous if it is weakly 𝒞\mathcal{C}-ambidextrous, 𝒞\mathcal{C} admits all qq-limits and qq-colimits and the the norm map q!Nmqqq_{!}\xrightarrow{\operatorname{Nm}_{q}}q_{*} is an isomorphism.

2.1.1 should be understood inductively on the level of truncatedness of qq. A (2)(-2)-finite map, i.e. an isomorphism, is always 𝒞\mathcal{C}-ambidextrous. If qq is mm-finite, then the diagonal map

AΔqA×BAA\xrightarrow{\Delta_{q}}A\times_{B}A

is (m1)(m-1)-finite and the ambidexterity of Δq\Delta_{q} allows in turn the definition of Nmq\operatorname{Nm}_{q} by [HL13, Construction 4.1.8] (see also [CSY18, Definition 3.1.3]).

Remark 2.1.2.

A map A𝑞BA\xrightarrow{q}B is 𝒞\mathcal{C}-ambidextrous if and only if all the fibers of qq are 𝒞\mathcal{C}-ambidextrous spaces [HL13, Corollary 4.2.6(2), Corollary 4.3.6]. Moreover, the fibers of the diagonal AA×AA\to A\times A are the path spaces of AA. In other words, AA is weakly 𝒞\mathcal{C}-ambidextrous if and only if the path spaces of AA are 𝒞\mathcal{C}-ambidextrous. Thus, 𝒞\mathcal{C}-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 𝒞\mathcal{C} be an \infty-category and let A𝑞BA\xrightarrow{q}B be a π\pi-finite map. The map qq is 𝒞\mathcal{C}-ambidextrous if and only if the following hold:

  1. (1)

    qq is weakly 𝒞\mathcal{C}-ambidextrous.

  2. (2)

    𝒞\mathcal{C} admits all qq-limits and qq-colimits.

  3. (3)

    Either qq_{*} preserves all qq-colimits or q!q_{!} preserves all qq-limits.

Proof.

By 2.1.2, we may assume that B=ptB=\operatorname{pt}, 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 𝒞\mathcal{C} admits, and qq_{*} (resp. q!q_{!}) preserves, all small colimits (resp. limits), the proof uses only qq-colimits (resp. qq-limits). ∎

As a consequence, we can easily deduce that ambidexterity enjoys the following closure properties with respect to the \infty-category:

Proposition 2.1.4.

Let 𝒞\mathcal{C} be an \infty-category and let AA be a π\pi-finite 𝒞\mathcal{C}-ambidextrous space. The space AA is also 𝒟\mathcal{D}-ambidextrous for:

  1. (1)

    𝒟=𝒞op\mathcal{D}=\mathcal{C}^{\text{\emph{op}}} the opposite \infty-category of 𝒞\mathcal{C}.

  2. (2)

    𝒟=Fun(,𝒞)\mathcal{D}=\operatorname{Fun}(\mathcal{I},\mathcal{C}) for an \infty-category \mathcal{I}.

  3. (3)

    𝒟𝒞\mathcal{D}\subseteq\mathcal{C} containing the final object and closed under ΩakA\Omega_{a}^{k}A-limits for all aAa\in A and k0k\geq 0.

  4. (4)

    𝒟𝒞\mathcal{D}\subseteq\mathcal{C} containing the initial object and closed under ΩakA\Omega_{a}^{k}A-colimits for all aAa\in A and k0k\geq 0.

Proof.

First, (4) follows from (3) and (1), so it suffices to consider (1)-(3). In all cases, we proceed by induction on mm, so we may assume by induction that AA is weakly 𝒟\mathcal{D}-ambidextrous. By 2.1.3, it suffices to verify that AA-limits and AA-colimit exist in 𝒟\mathcal{D} and that the functor AA_{*} preserves AA-colimits or that A!A_{!} preserves AA-limits. For (1), the claim follows from the fact that limits in 𝒞op\mathcal{C}^{\text{\emph{op}}} are computed as colimits in 𝒞\mathcal{C} and vice versa. For (2), we use the fact that limits and colimits in Fun(,𝒞)\operatorname{Fun}(\mathcal{I},\mathcal{C}) are computed pointwise. For (3), since AA-limits in 𝒞\mathcal{C} coincide with AA-colimits in 𝒞\mathcal{C}, it follows that AA-limits and AA-colimits are computed in 𝒟\mathcal{D} in the same way as in 𝒞\mathcal{C}. ∎

Cardinality

The main feature of ambidexterity is that it allows us to integrate families of morphisms in 𝒞\mathcal{C}. That is, given a 𝒞\mathcal{C}-ambidextrous map A𝑞BA\xrightarrow{q}B and X,Y𝒞BX,Y\in\mathcal{C}^{B} we have a map (see [CSY18, Definition 2.1.11])

q:Map𝒞(qX,qY)Map𝒞(X,Y).\int_{q}\colon\operatorname{Map}_{\mathcal{C}}(q^{*}X,q^{*}Y)\to\operatorname{Map}_{\mathcal{C}}(X,Y).

When B=ptB=\operatorname{pt}, we can think of an element of Map𝒞(qX,qY)\operatorname{Map}_{\mathcal{C}}(q^{*}X,q^{*}Y) as a map A𝑓Map𝒞(X,Y)A\xrightarrow{f}\operatorname{Map}_{\mathcal{C}}(X,Y), and of AfMap𝒞(X,Y)\int_{A}f\in\operatorname{Map}_{\mathcal{C}}(X,Y) as the sum of ff over the points of AA. In particular, we can integrate the identity morphism:

Definition 2.1.5.

Let 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty} and let A𝑞BA\xrightarrow{q}B be a 𝒞\mathcal{C}-ambidextrous map. We have a natural transformation Id𝒞B|q|𝒞Id𝒞B\operatorname{Id}_{\mathcal{C}^{B}}\xrightarrow{|q|_{\mathcal{C}}}\operatorname{Id}_{\mathcal{C}^{B}} given by the composition

Id𝒞BuqqNmq1q!qc!Id𝒞B.\operatorname{Id}_{\mathcal{C}^{B}}\xrightarrow{u_{*}}q_{*}q^{*}\xrightarrow{\operatorname{Nm}_{q}^{-1}}q_{!}q^{*}\xrightarrow{c_{!}}\operatorname{Id}_{\mathcal{C}^{B}}.

For a 𝒞\mathcal{C}-ambidextrous space AA, we write Id𝒞|A|𝒞Id𝒞\operatorname{Id}_{\mathcal{C}}\xrightarrow{|A|_{\mathcal{C}}}\operatorname{Id}_{\mathcal{C}} and call |A|𝒞|A|_{\mathcal{C}} the 𝒞\mathcal{C}-cardinality of AA.

The name “𝒞\mathcal{C}-cardinality” can be explained as follows. For a given object X𝒞X\in\mathcal{C}, the map X|A|XXX\xrightarrow{|A|_{X}}X equals AIdX\int_{A}\operatorname{Id}_{X}. Thus, we think of |A|X|A|_{X} as the sum of the identity of XX with itself “AA times”. Or in other words, as the result of multiplying by the “cardinality of AA” on XX. The basic example which motivates the terminology and notation is the following:

Example 2.1.6.

Let 𝒞\mathcal{C} be a semiadditive \infty-category. For a finite set AA, viewed as a 0-finite space, the operation |A|𝒞|A|_{\mathcal{C}} is simply the multiplication by the natural number, which is the cardinality of AA in the usual sense.

Remark 2.1.7.

For a general 𝒞\mathcal{C}-ambidextrous map A𝑞BA\xrightarrow{q}B, the transformation |q|𝒞|q|_{\mathcal{C}} can be understood as follows. Let B𝑋𝒞B\xrightarrow{X}\mathcal{C} be a BB-family of objects in 𝒞\mathcal{C}. For each bBb\in B, let us denote the fiber of qq at bb by AbA_{b} and the evaluation of XX at bb by XbX_{b}. By [CSY18, Proposition 3.1.13], the map X|q|𝒞XX\xrightarrow{|q|_{\mathcal{C}}}X acts on each XbX_{b} by |Ab|𝒞|A_{b}|_{\mathcal{C}}. In other words, |q|𝒞|q|_{\mathcal{C}} is just the BB-family of 𝒞\mathcal{C}-cardinalities of the BB-family of spaces AbA_{b}.

For a 𝒞\mathcal{C}-ambidextrous space AA, the AA-limits and AA-colimits in 𝒞\mathcal{C} are canonically isomorphic. This can be used to show the following:

Proposition 2.1.8.

Let 𝒞,𝒟Cat\mathcal{C},\mathcal{D}\in\operatorname{Cat}_{\infty} and let AA be a 𝒞\mathcal{C}- and 𝒟\mathcal{D}-ambidextrous space. A functor F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D} preserves all AA-limits if and only if it preserves all AA-colimits. Moreover, if FF preserves all AA-(co)limits, then F(|A|𝒞)=|A|𝒟.F(|A|_{\mathcal{C}})=|A|_{\mathcal{D}}.

Proof.

[CSY18, Corollary 3.2.4] shows that FF preserves mm-finite limits if and only if it preserves mm-finite colimits and [CSY18, Corollary 3.2.7] shows that in this case F(|A|𝒞)=|A|𝒟F(|A|_{\mathcal{C}})=|A|_{\mathcal{D}} for every mm-finite space AA. One easily checks that all the arguments are valid for an individual space AA as well. ∎

The 𝒞\mathcal{C}-cardinality is additive in the following sense:

Proposition 2.1.9.

Let 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty} and A𝑞BA\xrightarrow{q}B be a map of spaces. If BB and qq are 𝒞\mathcal{C}-ambidextrous then AA is 𝒞\mathcal{C}-ambidextrous and for every X𝒞X\in\mathcal{C},

|A|X=B|q|BX.|A|_{X}=\int_{B}|q|_{B^{*}X}.

Informally, 2.1.9 says that the cardinality of the total space AA is the “sum over BB” of the cardinalities of the fibers AbA_{b} of qq.

Proof.

This follows from the Higher Fubini’s Theorem ([CSY18, Propostion 2.1.15]) applied to the identity morphism. ∎

Remark 2.1.10.

By 2.1.9 and 2.1.7, for every pair of 𝒞\mathcal{C}-ambidextrous spaces AA and BB we have

|A×B|𝒞=|A|𝒞|B|𝒞End(Id𝒞).|A\times B|_{\mathcal{C}}=|A|_{\mathcal{C}}|B|_{\mathcal{C}}\quad\in\quad\operatorname{End}(\operatorname{Id}_{\mathcal{C}}).

Similarly, by [HL13, Remark 4.4.11], for every 𝒞\mathcal{C}-ambidextrous space AA we have

|A|𝒞=aπ0A|Aa|𝒞End(Id𝒞).|A|_{\mathcal{C}}=\sum_{a\in\pi_{0}A}|A_{a}|_{\mathcal{C}}\quad\in\quad\operatorname{End}(\operatorname{Id}_{\mathcal{C}}).

The various naturality properties enjoyed by the operations |A|𝒞|A|_{\mathcal{C}} allow for useful abuses of notation:

  1. (1)

    Given an AA-colimit preserving functor F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D}, if AA is 𝒞\mathcal{C}- and 𝒟\mathcal{D}-ambidextrous, we get by 2.1.8 that F(|A|𝒞)=|A|𝒟F(|A|_{\mathcal{C}})=|A|_{\mathcal{D}}. We therefore write |A||A|, dropping the subscript 𝒞\mathcal{C}, whenever convenient. As a special case, let 𝒞𝒞\mathcal{C}_{\circ}\subseteq\mathcal{C} be a full subcategory which is closed under AA-(co)limits. The cardinality |A|𝒞|A|_{\mathcal{C}_{\circ}} coincides with the restriction of |A|𝒞|A|_{\mathcal{C}} to 𝒞\mathcal{C}_{\circ}.

  2. (2)

    When 𝒞\mathcal{C} is monoidal and the tensor product preserves AA-colimits in each variable, the action of |A||A| on any object X𝒞X\in\mathcal{C} can be identified with tensoring 𝟙|A|𝟙\mathds{1}\xrightarrow{|A|}\mathds{1} with XX (see [CSY18, Lemma 3.3.4]). We therefore sometimes identify |A||A| with an element of π0𝟙π0Map(𝟙,𝟙)\pi_{0}\mathds{1}\coloneqq\pi_{0}\operatorname{Map}(\mathds{1},\mathds{1}).

  3. (3)

    Furthermore, for RAlg(𝒞)R\in\operatorname{Alg}(\mathcal{C}), the map R|A|RR\xrightarrow{|A|}R can be also identified with multiplication by the image of |A|π0𝟙|A|\in\pi_{0}\mathds{1} under the map π0𝟙π0R\pi_{0}\mathds{1}\to\pi_{0}R, which we also denote by |A||A|.

All these abuses of notation are compatible with standard conventions when AA is a finite set (see 2.1.6).

Higher commutative monoids

Of particular interest are mm-semiadditive \infty-categories, i.e. those for which all mm-finite spaces are ambidextrous. For m=0m=0, we recover the ordinary notion of a semiadditive \infty-category. The central feature of semiadditive \infty-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 mm-commutative monoids is developed and applied to the study of mm-semiadditivity for all 2m<-2\leq m<\infty. In this section, we recall from [Har17] a part of this theory of higher commutative monoids and extend it to the case m=m=\infty.

Definition 2.1.11.

[Har17, Definition 5.10, Proposition 5.14] Let 2m<-2\leq m<\infty, for 𝒞Catm-fin\mathcal{C}\in\mathrm{Cat}^{m\text{-}\mathrm{fin}}, the \infty-category of mm-commutative monoids in 𝒞\mathcal{C} is given by

CMonm(𝒞)Funm-fin(Span(𝒮m-fin)op,𝒞).\operatorname{CMon}_{m}(\mathcal{C})\coloneqq\operatorname{Fun}^{m\text{-fin}}(\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}})^{\text{\emph{op}}},\mathcal{C}).

In the case 𝒞=𝒮\mathcal{C}=\mathcal{S}, we simply write CMonm\operatorname{CMon}_{m} and refer to its objects as mm-commutative monoids555For m=0m=0, one indeed recovers the usual notion of a commutative (i.e. 𝔼\mathbb{E}_{\infty}) monoids in spaces by comparison with Segal objects [Lura, Section 2.4.2]..

In the case m=2m=-2, evaluating at pt\operatorname{pt}, the unique object of Span(𝒮(2)-fin)\operatorname{Span}(\mathcal{S}_{(-2)\text{-}\mathrm{fin}}), gives an equivalence CMon2(𝒞)𝒞\operatorname{CMon}_{-2}(\mathcal{C})\simeq\mathcal{C}.

Remark 2.1.12.

We can understand the definition of CMonm\operatorname{CMon}_{m} as follows. An object XCMonmX\in\operatorname{CMon}_{m} consists of the “underlying space” X(pt)X(\operatorname{pt}), together with a collection of coherent operations for summation of mm-finite families of points in it. Indeed, for every A𝒮m-finA\in\mathcal{S}_{m\text{-}\mathrm{fin}}, there is a canonical equivalence X(A)X(pt)AX(A)\simeq X(\operatorname{pt})^{A}. Given ABA\to B in 𝒮m-fin\mathcal{S}_{m\text{-}\mathrm{fin}}, the “right way” map XBXAX^{B}\to X^{A} is given simply by restriction, while the “wrong way” map XAXBX^{A}\to X^{B} 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 2m<-2\leq m<\infty and let 𝒞Cat(m+1)-fin\mathcal{C}\in\mathrm{Cat}^{(m+1)\text{-}\mathrm{fin}}. The restriction along the inclusion functor

ιm:Span(𝒮m-fin)Span(𝒮(m+1)-fin)\iota_{m}\colon\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}})\hookrightarrow\operatorname{Span}(\mathcal{S}_{(m+1)\text{-}\mathrm{fin}})

induces a limit preserving functor

ιm:CMonm+1(𝒞)CMonm(𝒞).\iota_{m}^{*}\colon\operatorname{CMon}_{m+1}(\mathcal{C})\to\operatorname{CMon}_{m}(\mathcal{C}).
Proof.

It suffices to show that the ιm\iota_{m} preserves mm-finite colimits. By [Har17, Corollary 2.16], it suffices to show that the composite

F:𝒮m-finSpan(𝒮m-fin)Span(𝒮(m+1)-fin)F\colon\mathcal{S}_{m\text{-}\mathrm{fin}}\to\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}})\hookrightarrow\operatorname{Span}(\mathcal{S}_{(m+1)\text{-}\mathrm{fin}})

preserves mm-finite colimits. Indeed, FF factors also as the composite

𝒮m-fin𝒮(m+1)-finSpan(𝒮(m+1)-fin).\mathcal{S}_{m\text{-}\mathrm{fin}}\hookrightarrow\mathcal{S}_{(m+1)\text{-}\mathrm{fin}}\to\operatorname{Span}(\mathcal{S}_{(m+1)\text{-}\mathrm{fin}}).

The first functor clearly preserves mm-finite colimits while the second one preserves mm-finite colimits by [Har17, Proposition 2.12]. ∎

We now extend the definition of CMonm\operatorname{CMon}_{m} to m=m=\infty.

Definition 2.1.14.

For 𝒞Cat-fin\mathcal{C}\in\mathrm{Cat}^{\infty\text{-}\mathrm{fin}}, we denote

CMon(𝒞)limmCMonm(𝒞).\operatorname{CMon}_{\infty}(\mathcal{C})\coloneqq\underleftarrow{\operatorname{lim}\,}_{m}\operatorname{CMon}_{m}(\mathcal{C}).

As above we write CMon:=CMon(𝒮)\operatorname{CMon}_{\infty}:=\operatorname{CMon}_{\infty}(\mathcal{S}).

When 𝒞\mathcal{C} is presentable, CMonm(𝒞)\operatorname{CMon}_{m}(\mathcal{C}) is presentable for all mm, by [Har17, Lemma 5.17]. Moreover, CMon(𝒞)\operatorname{CMon}_{\infty}(\mathcal{C}) can then be described as the colimit of CMonm(𝒞)\ \operatorname{CMon}_{m}(\mathcal{C}) in Pr\Pr.

Lemma 2.1.15.

For 𝒞Pr\mathcal{C}\in\Pr, the forgetful functors

ιm:CMonm+1(𝒞)CMonm(𝒞),\iota_{m}^{*}\colon\operatorname{CMon}_{m+1}(\mathcal{C})\to\operatorname{CMon}_{m}(\mathcal{C}),

admit left adjoints and the colimit of the sequence

𝒞CMon2CMon1(𝒞)CMonm(𝒞)\mathcal{C}\simeq\operatorname{CMon}_{-2}\xrightarrow{}\operatorname{CMon}_{-1}(\mathcal{C})\xrightarrow{}\cdots\to\operatorname{CMon}_{m}(\mathcal{C})\xrightarrow{}\cdots

in Pr\Pr is CMon(𝒞)\operatorname{CMon}_{\infty}(\mathcal{C}). In particular, CMon(𝒞)\operatorname{CMon}_{\infty}(\mathcal{C}) is presentable.

Proof.

Since CMonm(𝒞)\operatorname{CMon}_{m}(\mathcal{C}) is presentable for all mm, by the adjoint functor theorem [Lur09, Corollary 5.5.2.9], the functor ιm\iota_{m}^{*} admits a left adjoint if and only if it is accessible and limit preserving. The functor ιm\iota_{m}^{*} is κ\kappa-accessible for any κ\kappa large enough such that (m+1)(m+1)-finite limits commute with κ\kappa-filtered colimits in 𝒞\mathcal{C}, and by 2.1.13, is limit preserving. The second claim follows from the description of colimts in Pr\Pr (see [Lur09, Theorem 5.5.3.18, Corollary 5.5.3.4]). ∎

In the theory of mm-semiadditivity, the \infty-category CMonm\operatorname{CMon}_{m} plays an analogous role to that of the \infty-category CMon\operatorname{CMon} of commutative (i.e. 𝔼\mathbb{E}_{\infty}) monoids in spaces, in the the theory of ordinary (i.e. 0) semiadditivity. In particular, the mapping space between every two objects in an mm-semiadditive \infty-category has a canonical mm-commutative monoid structure. To see this, we begin by recalling the fundamental universal property of CMonm\operatorname{CMon}_{m} from [Har17]. For each 𝒞Catm-fin\mathcal{C}\in\mathrm{Cat}^{m\text{-}\mathrm{fin}}, we have a forgetful functor CMonm(𝒞)CMon2(𝒞)=𝒞\operatorname{CMon}_{m}(\mathcal{C})\to\operatorname{CMon}_{-2}(\mathcal{C})=\mathcal{C}, given by evaluation at pt𝒮m-fin\operatorname{pt}\in\mathcal{S}_{m\text{-}\mathrm{fin}}.

Proposition 2.1.16.

Let 2m-2\leq m\leq\infty. For every 𝒞Cat

-
m
\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}m}
and 𝒟Catm-fin\mathcal{D}\in\mathrm{Cat}^{m\text{-}\mathrm{fin}}, post-composition with the forgetful functor induces an equivalence of \infty-categories

Funm-fin(𝒞,CMonm(𝒟))Funm-fin(𝒞,𝒟).\operatorname{Fun}^{m\text{-}\mathrm{fin}}(\mathcal{C},\operatorname{CMon}_{m}(\mathcal{D}))\simeq\operatorname{Fun}^{m\text{-}\mathrm{fin}}(\mathcal{C},\mathcal{D}).
Proof.

The case m<m<\infty is proved in [Har17, Proposition 5.14]. For m=m=\infty we have

Fun-fin(𝒞,CMon(𝒟))limkFunk-fin(𝒞,CMon(𝒟))\operatorname{Fun}^{\infty\text{-}\mathrm{fin}}(\mathcal{C},\operatorname{CMon}_{\infty}(\mathcal{D}))\simeq\underleftarrow{\operatorname{lim}\,}_{k}\operatorname{Fun}^{k\text{-}\mathrm{fin}}(\mathcal{C},\operatorname{CMon}_{\infty}(\mathcal{D}))\simeq
limklimFunk-fin(𝒞,CMon(𝒟))limkFunk-fin(𝒞,CMonk(𝒟))\underleftarrow{\operatorname{lim}\,}_{k}\underleftarrow{\operatorname{lim}\,}_{\ell}\operatorname{Fun}^{k\text{-}\mathrm{fin}}(\mathcal{C},\operatorname{CMon}_{\ell}(\mathcal{D}))\simeq\underleftarrow{\operatorname{lim}\,}_{k}\operatorname{Fun}^{k\text{-}\mathrm{fin}}(\mathcal{C},\operatorname{CMon}_{k}(\mathcal{D}))\simeq
limkFunk-fin(𝒞,𝒟)Fun-fin(𝒞,𝒟).\underleftarrow{\operatorname{lim}\,}_{k}\operatorname{Fun}^{k\text{-}\mathrm{fin}}(\mathcal{C},\mathcal{D})\simeq\operatorname{Fun}^{\infty\text{-}\mathrm{fin}}(\mathcal{C},\mathcal{D}).

As a corollary, for every mm-semiadditive \infty-category we have a unique lift of the Yoneda embedding to a CMonm\operatorname{CMon}_{m}-enriched Yoneda embedding.

Corollary 2.1.17.

Let 2m-2\leq m\leq\infty. For every 𝒞Cat

-
m
\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}m}
, there is a unique fully faithful mm-semiadditive functor

HCMonm:𝒞Fun(𝒞op,CMonm)\text{\char 72\relax}^{\operatorname{CMon}_{m}}\colon\mathcal{C}\hookrightarrow\operatorname{Fun}(\mathcal{C}^{\text{\emph{op}}},\operatorname{CMon}_{m})

whose composition with the forgetful functor CMonm𝒮\operatorname{CMon}_{m}\to\mathcal{S} is the Yoneda embedding.

Proof.

Taking 𝒟=𝒮\mathcal{D}=\mathcal{S} in 2.1.16, shows that the ordinary Yoneda embedding

H:𝒞Funm-fin(𝒞op,𝒮)Fun(𝒞op,𝒮)\text{\char 72\relax}\colon\mathcal{C}\hookrightarrow\operatorname{Fun}^{m\text{-}\mathrm{fin}}(\mathcal{C}^{\text{\emph{op}}},\mathcal{S})\subseteq\operatorname{Fun}(\mathcal{C}^{\text{\emph{op}}},\mathcal{S})

lifts uniquely to a fully faithful mm-finite limit preserving functor

HCMonm:𝒞Funm-fin(𝒞op,CMonm)Fun(𝒞op,CMonm).\text{\char 72\relax}^{\operatorname{CMon}_{m}}\colon\mathcal{C}\hookrightarrow\operatorname{Fun}^{m\text{-}\mathrm{fin}}(\mathcal{C}^{\text{\emph{op}}},\operatorname{CMon}_{m})\subseteq\operatorname{Fun}(\mathcal{C}^{\text{\emph{op}}},\operatorname{CMon}_{m}).

The CMonm\operatorname{CMon}_{m}-enriched Yoneda embedding HCMonm\text{\char 72\relax}^{\operatorname{CMon}_{m}} corresponds to a functor

homCMonm(,):𝒞×𝒞opCMonm,\hom^{\operatorname{CMon}_{m}}(-,-)\colon\mathcal{C}\times\mathcal{C}^{\text{\emph{op}}}\to\operatorname{CMon}_{m},

whose composition with the forgetful functor CMonm𝒮\operatorname{CMon}_{m}\to\mathcal{S} is the functor Map𝒞(,)\operatorname{Map}_{\mathcal{C}}(-,-). Thus, we obtain a canonical structure of an mm-commutative monoid on each mapping space in 𝒞\mathcal{C}. Informally, the “wrong way” maps for A𝑞BA\xrightarrow{q}B, in the higher commutative monoid structure on Map𝒞(X,Y)\operatorname{Map}_{\mathcal{C}}(X,Y), are given by integration

q:Map𝒞(X,Y)AMap𝒞(X,Y)B.\int_{q}\colon\operatorname{Map}_{\mathcal{C}}(X,Y)^{A}\to\operatorname{Map}_{\mathcal{C}}(X,Y)^{B}.
Remark 2.1.18.

It is overwhelmingly likely that an mm-semiadditive \infty-category 𝒞\mathcal{C} can be canonically enriched in CMonm\operatorname{CMon}_{m} (for e.g. in the sense of [GH15] or [Hin16]), such that the CMonm\operatorname{CMon}_{m}-valued mapping objects coincide with our definition above. In case 𝒞\mathcal{C} is further assumed to be presentable, this follows from the fact that 𝒞\mathcal{C} is left tensored over CMonm\operatorname{CMon}_{m} (see [Har17, Lemma 5.20], 5.3.1 and [Lura, Proposition 4.2.1.33]).

2.2 Examples

We now review some examples of mm-semiadditive \infty-categories and the behavior of cardinalities of mm-finite spaces in them.

Universal

It is proved in [Har17], that the following is the universal example of an mm-semiadditive \infty-category. In particular, it shows that in general, the operations |A||A| need not reduce to something “classical”:

Example 2.2.1 (Universal case).

By [Har17, Corollary 5.7], for 2m<-2\leq m<\infty the symmetric monoidal \infty-category of spans 𝒞=Span(𝒮m-fin)\mathcal{C}=\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}}) is the universal mm-semiadditive \infty-category. For every A𝒮m-finA\in\mathcal{S}_{m\text{-}\mathrm{fin}}, we have

|A|pt=(ptApt)π0MapSpan(𝒮m-fin)(pt,pt).|A|_{\operatorname{pt}}=(\operatorname{pt}\leftarrow A\to\operatorname{pt})\quad\in\quad\pi_{0}\operatorname{Map}_{\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}})}(\operatorname{pt},\operatorname{pt}).

Moreover, π0MapSpan(𝒮m-fin)(pt,pt)\pi_{0}\operatorname{Map}_{\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}})}(\operatorname{pt},\operatorname{pt}) is the set of isomorphism classes of mm-finite spaces with the semiring structure given by (see 2.1.10):

|A|+|B|=|AB|,|A||B|=|A×B|.|A|+|B|=|A\sqcup B|,\quad|A|\cdot|B|=|A\times B|.

Informally, the universality of 2.2.1 is reflected in its construction as follows. The collection of spaces 𝒮m-fin𝒮\mathcal{S}_{m\text{-}\mathrm{fin}}\subseteq\mathcal{S} is generated under mm-finite colimits from the point pt𝒮\operatorname{pt}\in\mathcal{S}. The “right way” maps in Span(𝒮m-fin)\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}}) encode the usual covariant functoriality of these colimits. The “wrong way” maps in Span(𝒮m-fin)\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}}) encode the contravariant functoriality arising from “integration along the fibers”.

A closely related example is the \infty-category CMonm\operatorname{CMon}_{m} of mm-commutative monoids, which is shown in [Har17, Corollary 5.19, Corollary 5.21], to be the universal presentable mm-semiadditive \infty-category. The Yoneda embedding induces a fully faithful mm-semiadditive (symmetric monoidal) functor

Span(𝒮m-fin)CMonm,\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}})\hookrightarrow\operatorname{CMon}_{m},

taking each mm-finite space AA to the “free mm-commutative monoid” on AA. From this we get that cardinalities in CMonm\operatorname{CMon}_{m} are computed essentially in the same way as in Span(𝒮m-fin)\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}})777The relation between Span(𝒮m-fin)\operatorname{Span}(\mathcal{S}_{m\text{-}\mathrm{fin}}) and CMonm\operatorname{CMon}_{m} is somewhat analogues to the relation between the \infty-category Spω\operatorname{Sp}^{\omega} of finite spectra and the \infty-category Sp\operatorname{Sp} of all spectra. . In section 5, we shall discuss more systematically the universality of CMonm\operatorname{CMon}_{m} (see 5.3.1).

Rational

There are however some situations in which the operations |A||A| can be expressed in terms of classical invariants.

Example 2.2.2 (Homotopy cardinality).

For a π\pi-finite space AA, Baez and Dolan [Bae] define the homotopy cardinality of AA to be the following non-negative rational number

|A|0aπ0(A)n1|πn(A,a)|(1)n0.|A|_{0}\coloneqq\sum_{a\in\pi_{0}(A)}\prod_{n\geq 1}|\pi_{n}(A,a)|^{(-1)^{n}}\quad\in\quad\mathbb{Q}_{\geq 0}. (1)

This notion can be seen as a special case of the cardinality of a π\pi-finite space in a higher semiadditive \infty-category as follows. We say that an \infty-category 𝒞\mathcal{C} is semirational if it is 0-semiadditive and for each nn\in\mathbb{N}, multiplication by nn is invertible in 𝒞\mathcal{C} (e.g. 𝒞=Sp\mathcal{C}=\operatorname{Sp}_{\mathbb{Q}} or Vec\mathrm{Vec}_{\mathbb{Q}}). We shall see that a semirational \infty-category which admits all 1-finite colimits is automatically \infty-semiadditive and for every π\pi-finite space AA, we have

|A|𝒞=|A|00End(Id𝒞).|A|_{\mathcal{C}}=|A|_{0}\quad\in\quad\mathbb{Q}_{\geq 0}\subseteq\operatorname{End}(\operatorname{Id}_{\mathcal{C}}).

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 π\pi-finite spaces

FAB,F\to A\to B,

such that BB is connected, we have |A|=|F||B||A|=|F||B|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 𝒞\mathcal{C} be a 0-semiadditive \infty-category, which admits π\pi-finite colimits, and let AA be a π\pi-finite space. If AA satisfies the following condition:

  • ()(*)

    The orders of the homotopy groups of AA are invertible on all objects of 𝒞.\mathcal{C}.

Then AA is 𝒞\mathcal{C}-ambidextrous and |A|𝒞=|A|0|A|_{\mathcal{C}}=|A|_{0}.

From the perspective of the theory we are about to develop, semirational \infty-categories are 0-semiadditive \infty-categories of “height 0”. One of our goals is to generalize the above phenomena to “higher heights” (see 3.2.2 and 3.2.5).

Chromatic

Examples of \infty-semiadditive \infty-categories of “higher height” appear naturally in chromatic homotopy theory. For a given prime pp and 0n<0\leq n<\infty, let K(n)K(n) be the Morava KK-theory spectrum of height nn at the prime pp. One of the main results of [HL13] is that the localizations SpK(n)\operatorname{Sp}_{K(n)} are \infty-semiadditive. In particular, we can consider K(n)K(n)-local cardinalities of π\pi-finite spaces. For n=0n=0, we have SpK(n)Sp\operatorname{Sp}_{K(n)}\simeq\operatorname{Sp}_{\mathbb{Q}} and we recover the homotopy cardinality (2.2.2). Similarly, since SpK(n)\operatorname{Sp}_{K(n)} is pp-local for all nn, for every π\pi-finite space AA whose homotopy groups have cardinality prime to pp, the K(n)K(n)-local cardinality of AA coincides with its homotopy cardinality for all nn (see 2.2.3). In particular, it is independent of nn. However, for n1n\geq 1 the prime pp is not invertible in SpK(n)\operatorname{Sp}_{K(n)}. Thus, there are π\pi-finite spaces AA (e.g. π\pi-finite pp-spaces), which are ambidextrous even though they do not satisfy condition ()(*) of 2.2.3. For such spaces AA, the K(n)K(n)-local cardinality does depend on nn and in general does not (and can not) agree with the homotopy cardinality.999Note that the rationalization functor L:SpK(n)SpL_{\mathbb{Q}}\colon\operatorname{Sp}_{K(n)}\to\operatorname{Sp}_{\mathbb{Q}} does not preserve colimits in general and so does not preserve cardinalities. It does however preserve colimits which are indexed on π\pi-finite spaces whose homotopy groups have order prime to pp.

To study the K(n)K(n)-local cardinalities of π\pi-finite spaces, it is useful to consider their image in Morava EE-theory. For every integer n1n\geq 1, we let EnE_{n} be the Morava EE-theory associated with some formal group of height nn over 𝔽¯p\overline{\mathbb{F}}_{p}, viewed as an object of CAlg(SpK(n))\operatorname{CAlg}(\operatorname{Sp}_{K(n)}). In particular, we have a (non-canonical) isomorphism

πEn𝕎(𝔽¯p)[[u1,,un1]][u±1]|ui|=0,|u|=2.\pi_{*}E_{n}\simeq\mathbb{W}(\overline{\mathbb{F}}_{p})[[u_{1},\dots,u_{n-1}]][u^{\pm 1}]\qquad|u_{i}|=0,\quad|u|=2.
Example 2.2.4 (Chromatic cardinality).

The \infty-category Θn=ModEn(SpK(n))\Theta_{n}=\operatorname{Mod}_{E_{n}}(\operatorname{Sp}_{K(n)}) is \infty-semiadditive [CSY18, Theorem 5.3.1] and hence one can consider cardinalities of π\pi-finite spaces in π0En\pi_{0}E_{n}. We define the (pp-typical) height nn cardinality of AA to be

|A|n|A|Θnπ0En.|A|_{n}\coloneqq|A|_{\Theta_{n}}\quad\in\quad\pi_{0}E_{n}.

It makes sense to consider ¯\overline{\mathbb{Q}} as E0E_{0}, in which case we recover the homotopy cardinality (2.2.2). The technology of [Lur19] allows one to derive a rather explicit formula for |A|n|A|_{n}, for heights n>0n>0 as well. Let L^pAMap(Bp,A)\widehat{L}_{p}A\coloneqq\operatorname{Map}({B\mathbb{Z}}_{p},A) be the pp-adic free loop space of AA. One can show that the element |A|nπ0En|A|_{n}\in\pi_{0}E_{n} belongs to the subring (p)π0En\mathbb{Z}_{(p)}\subseteq\pi_{0}E_{n} and satisfies |A|n=|L^pA|n1|A|_{n}=|\widehat{L}_{p}A|_{n-1}. Applying this relation inductively we obtain the formula

|A|n=|Map(Bpn,A)|0(p).|A|_{n}=|\operatorname{Map}({B\mathbb{Z}}_{p}^{n},A)|_{0}\quad\in\quad\mathbb{Z}_{(p)}. (2)

If AA happens to be a pp-space, then L^pA\widehat{L}_{p}A coincides with the ordinary free loop space LAMap(S1,A)LA\coloneqq\operatorname{Map}(S^{1},A). Thus, |A|n|A|_{n} can be computed as the homotopy cardinality of the space of maps from the nn-dimensional torus to AA.

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 k,n0k,n\geq 0 we have |BkCp|n=p(n1k)|B^{k}C_{p}|_{n}=p^{\binom{n-1}{k}}.101010For height n=0n=0 this should be interpreted via the identity (1k)=(1)k.\binom{-1}{k}=(-1)^{k}.

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 π0𝕊K(n)\pi_{0}\mathbb{S}_{K(n)} is not entirely understood in general, it follows from [BG18] and [BGH17] that:

Proposition 2.2.6.

For all pp and nn, the image of the unit map π0𝕊K(n)𝑢π0En\pi_{0}\mathbb{S}_{K(n)}\xrightarrow{u}\pi_{0}E_{n} is pπ0En\mathbb{Z}_{p}\subseteq\pi_{0}E_{n} and the kernel is precisely the nil-radical.

Proof.

Let Γn\Gamma_{n} be the Morava stabilizer group associated with EnE_{n}. We have an action of Γn\Gamma_{n} on EnE_{n} by commutative algebra maps and thus, the map uu factors through the fixed points (π0En)Γnπ0En(\pi_{0}E_{n})^{\Gamma_{n}}\subseteq\pi_{0}E_{n}. By [BG18, Lemma 1.33], we have

(π0En)Γn=Hc0(Γ;π0En)=pπ0En.(\pi_{0}E_{n})^{\Gamma_{n}}=H_{c}^{0}(\Gamma;\pi_{0}E_{n})=\mathbb{Z}_{p}\subseteq\pi_{0}E_{n}.

By [BGH17, Theorem 2.3.5], the EE_{\infty}-page of the descent spectral sequence

Hcs(Γ;(En)t)πts(𝕊K(n))H_{c}^{s}(\Gamma;(E_{n})_{t})\implies\pi_{t-s}(\mathbb{S}_{K(n)})

has a horizontal vanishing line. Since the spectral sequence is multiplicative, this implies that all elements in π0𝕊K(n)\pi_{0}\mathbb{S}_{K(n)} with positive filtration degree are nilpotent. Finally, since 𝕊K(n)\mathbb{S}_{K(n)} admits a ring map from the pp-complete sphere, the map π0𝕊K(n)𝑢p\pi_{0}\mathbb{S}_{K(n)}\xrightarrow{u}\mathbb{Z}_{p} is surjective. ∎

Thus, for every π\pi-finite space AA, the identity

|A|SpK(n)=|A|n|A|_{\operatorname{Sp}_{K(n)}}=|A|_{n}

holds up to nilpotents. We do not know, however, whether it holds in π0𝕊K(n)\pi_{0}\mathbb{S}_{K(n)}.

Categorical

Another family of examples of higher semiadditive \infty-categories arises from category theory itself.

Proposition 2.2.7.

For every 2m-2\leq m\leq\infty the \infty-category Catm-fin\mathrm{Cat}_{m\text{-fin}} is mm-semiadditive.

Proof.

The case m<m<\infty is exactly [Har17, Proposition 5.26]. We now wish to show that Cat-fin\mathrm{Cat}_{\infty\text{-fin}} is kk-semiadditive for every k<k<\infty. By [Lura, Remark 4.8.1.6], both Catk-fin\mathrm{Cat}_{k\text{-fin}} and Cat-fin\mathrm{Cat}_{\infty\text{-fin}} admit closed symmetric monoidal structures, and by [Lura, Proposition 4.8.1.3], there exists a symmetric monoidal functor 𝒫:Catk-finCat-fin\mathcal{P}\colon\mathrm{Cat}_{k\text{-fin}}\to\mathrm{Cat}_{\infty\text{-fin}}. By [Lura, Remark 4.8.1.8] and [Lur09, Proposition 5.3.6.2(2)], 𝒫\mathcal{P} admits a right adjoint and thus preserves colimits. Hence, Cat-fin\mathrm{Cat}_{\infty\text{-fin}} is kk-semiadditive by [CSY18, Corollary 3.3.2(2)].∎

Example 2.2.8 (Categorical cardinality).

Let 2m-2\leq m\leq\infty and let 𝒞Catm-fin\mathcal{C}\in\mathrm{Cat}_{m\text{-fin}}. For every mm-finite space AA the mm-semiadditive structure of Catm-fin\mathrm{Cat}_{m\text{-fin}} gives rise to a functor |A|𝒞:𝒞𝒞.|A|_{\mathcal{C}}\colon\mathcal{C}\to\mathcal{C}. When m<m<\infty it is shown in [Har17, Section 5.2] that |A|𝒞|A|_{\mathcal{C}} is given by taking the constant colimit on AA. That is, it takes an object X𝒞X\in\mathcal{C} to the object X[A]𝒞X[A]\in\mathcal{C}. Since the forgetful functor Cat-finCatm-fin\mathrm{Cat}_{\infty\text{-fin}}\to\mathrm{Cat}_{m\text{-fin}} preserves limits, and hence m-semiadditive, the same holds for m=m=\infty. This is very suggestive of the idea that “multiplication by |A||A| on 𝒞\mathcal{C}” is given by “summing each object X𝒞X\in\mathcal{C} with itself AA times”. A closely related example is discussed in [HL13, Example 4.3.11], where it is shown that Pr\Pr is \infty-semiadditive (in fact, every mm-truncated space, not necessarily π\pi-finite, is Pr\Pr-ambidextrous).

Remark 2.2.9.

There is a different approach to the higher semiadditivity of Catm-fin\mathrm{Cat}_{m\text{-fin}}, based on the notion of ambidextrous adjunctions of (,2)(\infty,2)-categories. We sketch the argument to demonstrate the role of higher categorical structures as a useful perspective on ambidexterity phenomena. Given an mm-finite space AA, the adjunction

A:Catm-finCatm-finA:AA^{*}\colon\mathrm{Cat}_{m\text{-}\mathrm{fin}}\leftrightarrows\mathrm{Cat}_{m\text{-}\mathrm{fin}}^{A}\colon A_{*}

can be naturally promoted to an adjunction of (,2)(\infty,2)-categories. Moreover, the unit uu and counit cc of AAA^{*}\dashv A_{*}, as 11-morphisms in the respective (,2)(\infty,2)-categories of endofunctors, can be shown to have left adjoints uLu^{L} and cLc^{L} 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 uLu^{L} and cLc^{L} exhibit AA_{*} as a left adjoint of AA^{*} (see [HSSS18, Remark 2.2]). Hence, by 2.1.3, the space AA is Catm-fin\mathrm{Cat}_{m\text{-fin}}-ambidextrous and so Catm-fin\mathrm{Cat}_{m\text{-}\mathrm{fin}} is mm-semiadditive.

The functor that takes an \infty-category to its opposite induces an equivalence Catm-finCatm-fin\mathrm{Cat}_{m\text{-}\mathrm{fin}}\simeq\mathrm{Cat}^{m\text{-}\mathrm{fin}}. Hence, the \infty-category Catm-fin\mathrm{Cat}^{m\text{-}\mathrm{fin}} is mm-semiadditive as well and the higher semiadditive structure is given by taking limits. For every \infty-category 𝒞\mathcal{C} with finite (co)products, the (co)product endows the space of objects 𝒞\mathcal{C}^{\simeq} with a (co)Cartesian commutative monoid structure. Using the mm-semiadditivity of Catm-fin\mathrm{Cat}_{m\text{-}\mathrm{fin}} and Catm-fin\mathrm{Cat}^{m\text{-}\mathrm{fin}} together with the CMonm\operatorname{CMon}_{m}-enriched Yoneda embedding provided by 2.1.17, this too can be generalized to all 0m0\leq m\leq\infty. Given 𝒞Catm-fin\mathcal{C}\in\mathrm{Cat}_{m\text{-}\mathrm{fin}}, by the mm-semiadditivity of Catm-fin\mathrm{Cat}_{m\text{-}\mathrm{fin}}, the mapping space Mapm-fin(𝒮m-fin,𝒞)\operatorname{Map}_{m\text{-}\mathrm{fin}}(\mathcal{S}_{m\text{-}\mathrm{fin}},\mathcal{C}) admits a canonical structure of an mm-commutative monoid. On the other hand, since 𝒮m-fin\mathcal{S}_{m\text{-}\mathrm{fin}} is freely generated from a point under mm-finite colimits [Lura, Notation 4.8.5.2], we have

Mapm-fin(𝒮m-fin,𝒞)Map(pt,𝒞)𝒞.\operatorname{Map}_{m\text{-}\mathrm{fin}}(\mathcal{S}_{m\text{-}\mathrm{fin}},\mathcal{C})\simeq\operatorname{Map}(\operatorname{pt},\mathcal{C})\simeq\mathcal{C}^{\simeq}.
Definition 2.2.10.

For 𝒞Catm-fin\mathcal{C}\in\mathrm{Cat}_{m\text{-}\mathrm{fin}}, we refer to the above mm-commutative monoid structure on the space of objects 𝒞\mathcal{C}^{\simeq} as the coCartesian structure. A completely analogous construction endows the space of objects of each 𝒞Catm-fin\mathcal{C}\in\mathrm{Cat}^{m\text{-}\mathrm{fin}} with a Cartesian mm-commutative monoid structure.

As explained in [Har17, Section 5.2], the integration operations for the (co)Cartesian mm-commutative monoid structure on 𝒞\mathcal{C}^{\simeq} are given by taking mm-finite (co)limits. Finally, the \infty-category Cat

-
m
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}m}
of mm-semiadditive \infty-categories and mm-semiadditive functors is a full subcategory of both Catm-fin\mathrm{Cat}_{m\text{-}\mathrm{fin}} and Catm-fin\mathrm{Cat}^{m\text{-}\mathrm{fin}}. Moreover, we have the following:

Proposition 2.2.11.

Let 2m-2\leq m\leq\infty. The full subcategory Cat

-
m
Catm-fin
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}m}\subseteq\mathrm{Cat}_{m\text{-}\mathrm{fin}}
(resp. Catm-fin\mathrm{Cat}^{m\text{-}\mathrm{fin}}) is closed under colimits and in particular is mm-semiadditive.

Proof.

The functor ()op:Catm-finCatm-fin(-)^{\mathrm{op}}\colon\mathrm{Cat}_{m\text{-}\mathrm{fin}}\to\mathrm{Cat}^{m\text{-}\mathrm{fin}} that takes an \infty-category to its opposite is an equivalence. By 2.1.4, ()op(-)^{\mathrm{op}} restricts to an involution of Cat

-
m
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}m}
. It thus suffices to consider only the inclusion ι:Cat

-
m
Catm-fin
\iota\colon\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}m}\hookrightarrow\mathrm{Cat}^{m\text{-}\mathrm{fin}}
. By 2.1.16, ι\iota admits a right adjoint given by CMonm()\operatorname{CMon}_{m}(-) and thus preserve colimits. ∎

2.3 Amenability

Let A𝑞BA\xrightarrow{q}B be a 𝒞\mathcal{C}-ambidextrous map. Recall from [CSY18, Definition 3.1.7], that qq is called 𝒞\mathcal{C}-amenable if |q|𝒞|q|_{\mathcal{C}} is invertible. As with ambidexterity, amenability is a fiber-wise property ([CSY18, Corollary 3.1.16]). Thus, we shall be mainly interested in 𝒞\mathcal{C}-amenable spaces. i.e. those whose 𝒞\mathcal{C}-cardinality is invertible.

Remark 2.3.1.

We warn the reader not to confuse the condition that the natural transformation |q|𝒞|q|_{\mathcal{C}} is invertible with the condition that the natural transformation Nmq\operatorname{Nm}_{q} is invertible, which is equivalent to qq being 𝒞\mathcal{C}-ambidextrous (which is itself a prerequisite for defining |q|𝒞|q|_{\mathcal{C}}).

In this section, we extend some results from [CSY18] regarding amenability. The main point is that while AA-ambidexterity allows us to sum over AA-families of maps, AA-amenability allows us to average over AA-families of maps, which in turn facilitates “transfer arguments” along AA. We shall explore how this condition affects the higher semiadditive structure.

Closure properties

For a map of spaces, the condition of 𝒞\mathcal{C}-amenability, as the condition of 𝒞\mathcal{C}-ambidexterity, is fiber-wise. However, unlike 𝒞\mathcal{C}-ambidextrous maps, 𝒞\mathcal{C}-amenable maps are not closed under composition. To understand the situation, it suffices to consider the case A𝑞BptA\xrightarrow{q}B\to\operatorname{pt}. By the additivity of cardinality (2.1.9), we get |A|=B|q||A|=\int_{B}|q|. Assume for simplicity that BB is connected and that the fiber of qq is FF. The transformation |q||q| equals |F||F| at each point bBb\in B. Thus, it is tempting to presume that |A|=|B||F||A|=|B|\cdot|F| and hence if both |B||B| and |F||F| are invertible, then so is |A||A|. However, for this reasoning to hold we need to know that |q||q| is constant on BB with value |F||F|. Alas, in general |q||q| is not constant, even when |F||F| is invertible, and |A||A| need not equal |B||F||B|\cdot|F| (see 2.3.6). On the positive side, we show that |q||q| must be constant if we require in addition to the invertibility of |F||F| that FA𝑞BF\to A\xrightarrow{q}B is a principal fiber sequence.

Definition 2.3.2.

We call a map A𝑞BA\xrightarrow{q}B of spaces principal if it can be extended to a fiber sequence A𝑞B𝑓EA\xrightarrow{q}B\xrightarrow{f}E.

We note that for a principal map all the fibers are isomorphic even if the target is not connected.

Proposition 2.3.3.

Let 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty} and let A𝑞BA\xrightarrow{q}B be a principal 𝒞\mathcal{C}-ambidextrous map of 𝒞\mathcal{C}-ambidextrous spaces with fiber FF. For X𝒞X\in\mathcal{C}, if |F|X|F|_{X} is invertible, then

|q|BX=|F|BXEnd(BX),|q|_{B^{*}X}=|F|_{B^{*}X}\quad\in\quad\operatorname{End}(B^{*}X),

and

|A|X=|F|X|B|X.|A|_{X}=|F|_{X}|B|_{X}.
Proof.

The base change of qq along itself is a map A×BAq~AA\times_{B}A\xrightarrow{\tilde{q}}A, which is a principal map with a section. Therefore q~\widetilde{q} is isomorphic to the projection F×AπAA.F\times A\xrightarrow{\pi_{A}}A. Hence, q×Bqq×BπBq\times_{B}q\simeq q\times_{B}\pi_{B}, where F×BπBBF\times B\xrightarrow{\pi_{B}}B is the projection. We get from [CSY18, Corollary 3.1.14] that

|q|2=|q×Bq|=|q×BπB|=|q||πB|End(Id𝒞B).|q|^{2}=|q\times_{B}q|=|q\times_{B}\pi_{B}|=|q||\pi_{B}|\quad\in\quad\operatorname{End}(\operatorname{Id}_{\mathcal{C}^{B}}).

If |F|X|F|_{X} is invertible, then |q|BX|q|_{B^{*}X} is invertible, and thus by the above, we get ([CSY18, Proposition 3.1.13])

|q|BX=|πB|BX=B(|F|X)=|F|BXEnd(BX).|q|_{B^{*}X}=|\pi_{B}|_{B^{*}X}=B^{*}(|F|_{X})=|F|_{B^{*}X}\quad\in\quad\operatorname{End}(B^{*}X).

We can now integrate along BB and get

|A|X=B|q|BX=BB(|F|X)=|F|X|B|XEnd(X).|A|_{X}=\int\limits_{B}|q|_{B^{*}X}=\int\limits_{B}B^{*}(|F|_{X})=|F|_{X}|B|_{X}\quad\in\quad\operatorname{End}(X).

As a simple application, we deduce that when the homotopy groups of a π\pi-finite space have invertible cardinality in 𝒞\mathcal{C}, the notion of cardinality degenerates to the homotopy cardinality (2.2.3).

Proposition 2.3.4.

Let 𝒞Cat

-
0
\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}0}
and let AA be a π\pi-finite space. Assume that 𝒞\mathcal{C} admits AA-colimits and the order of each homotopy group of AA is invertible in End(Id𝒞)\operatorname{End}(\operatorname{Id}_{\mathcal{C}}). Then, AA is 𝒞\mathcal{C}-ambidextrous and |A|𝒞=|A|0|A|_{\mathcal{C}}=|A|_{0}. In particular, if AA is connected, then it is 𝒞\mathcal{C}-amenable.

Proof.

By 2.1.10, we may assume that AA is connected. We prove the claim for all connected mm-finite AA by induction on m0m\geq 0, where the case m=0m=0 is trivial. We prove the claim for some m>0m>0, assuming it holds for m1m-1. Choose a base point in AA and consider the connected component (ΩA)ΩA(\Omega A)^{\circ}\subseteq\Omega A of the identity loop. The space (ΩA)(\Omega A)^{\circ} satisfies the assumptions of the inductive hypothesis. Hence, (ΩA)(\Omega A)^{\circ} is amenable and |(ΩA)|𝒞=|(ΩA)|0|(\Omega A)^{\circ}|_{\mathcal{C}}=|(\Omega A)^{\circ}|_{0}. Now, ΩA\Omega A is just the coproduct of |π1A||\pi_{1}A| copies of (ΩA)(\Omega A)^{\circ}. Hence, by 2.1.10 and the inductive hypothesis, we have

|ΩA|𝒞=|π1A||(ΩA)|𝒞=|π1A||(ΩA)|0=|ΩA|0.|\Omega A|_{\mathcal{C}}=|\pi_{1}A|\cdot|(\Omega A)^{\circ}|_{\mathcal{C}}=|\pi_{1}A|\cdot|(\Omega A)^{\circ}|_{0}=|\Omega A|_{0}.

Moreover, |ΩA|𝒞|\Omega A|_{\mathcal{C}} is invertible as |π1A||\pi_{1}A| is invertible by assumption and |(ΩA)|𝒞|(\Omega A)^{\circ}|_{\mathcal{C}} is invertible by the inductive hypothesis. That is, ΩA\Omega A is 𝒞\mathcal{C}-amenable. Finally, consider the principal fiber sequence

ΩAptA.\Omega A\to\operatorname{pt}\to A.

Since ΩA\Omega A is 𝒞\mathcal{C}-amenable, by [CSY18, Proposition 3.1.17], the space AA is 𝒞\mathcal{C}-ambidextrous and by 2.3.3, we have

|A|𝒞=|ΩA|𝒞1=|ΩA|01=|A|0.|A|_{\mathcal{C}}=|\Omega A|_{\mathcal{C}}^{-1}=|\Omega A|_{0}^{-1}=|A|_{0}.

From 2.3.3, we also deduce that 𝒞\mathcal{C}-amenable maps are partially closed under composition:

Corollary 2.3.5.

Let 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty} and let A𝑓B𝑔CA\xrightarrow{f}B\xrightarrow{g}C be a pair of composable maps of spaces. If ff and gg are 𝒞\mathcal{C}-amenable and ff is principal, then gfg\circ f is 𝒞\mathcal{C}-amenable.

Proof.

We can check that gfg\circ f is 𝒞\mathcal{C}-amenable, by pulling back along ptC\operatorname{pt}\to C for every point of CC. In other words, we can assume that C=ptC=\operatorname{pt}. Taking FF to be the fiber of ff, we have a principal fiber sequence FABF\to A\to B where FF and BB are 𝒞\mathcal{C}-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 |F||F| is not invertible, then the identity |A|=|F||B||A|=|F||B| 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 𝒞\mathcal{C}-cardinality need not be multiplicative even when the fiber and base space are 𝒞\mathcal{C}-amenable. Moreover, in such case, the total space need not even be 𝒞\mathcal{C}-amenable and so in particular, 𝒞\mathcal{C}-amenable maps are not closed under composition.

Example 2.3.6.

Let pp be an odd prime and let ΘnModEn(SpK(n))\Theta_{n}\coloneqq\operatorname{Mod}_{E_{n}}(\operatorname{Sp}_{K(n)}). We consider the map B2Cp𝑓B4CpB^{2}C_{p}\xrightarrow{f}B^{4}C_{p} classifying the cup-square operation xxxx\mapsto x\cup x on mod pp cohomology, and the associated fiber sequence

FB2Cp𝑓B4Cp.F\to B^{2}C_{p}\xrightarrow{f}B^{4}C_{p}.

The only non-trivial homotopy groups of FF are π2Fπ3FCp\pi_{2}F\simeq\pi_{3}F\simeq C_{p}, but the Postnikov invariant represented by ff in H4(B2Cp;Cp)H^{4}(B^{2}C_{p};C_{p}) is non-zero. Using 2.2.4, we have |F|n=|LnF|0|F|_{n}=|L^{n}F|_{0} and we can compute it using the fiber sequence

LnFLnB2CpLnB4Cp,L^{n}F\to L^{n}B^{2}C_{p}\to L^{n}B^{4}C_{p},

via the induced long exact sequence on homotopy groups. The only complication arises at the level of π0\pi_{0}, where we need to compute the size of the kernel of the cup-square map

π0LnB2Cp=H2(Tn;Cp)H4(Tn;Cp)=π0LnB4Cp.\pi_{0}L^{n}B^{2}C_{p}=H^{2}(T^{n};C_{p})\to H^{4}(T^{n};C_{p})=\pi_{0}L^{n}B^{4}C_{p}.

Namely, the number of nn-dimensional 2-forms over CpC_{p} 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:

|F|n=p(n13)p3n+pnp1p21.|F|_{n}=p^{\binom{n-1}{3}}\cdot\frac{p^{3-n}+p^{n}-p-1}{p^{2}-1}.

In particular, taking n=4n=4, we get |F|4=p3+p1|F|_{4}=p^{3}+p-1, which is an invertible element in π0E4\pi_{0}E_{4}. It follows that FF is Θ4\Theta_{4}-amenable. Nevertheless,

|F|4|B4Cp|4=p3+p1|F|_{4}|B^{4}C_{p}|_{4}=p^{3}+p-1

which differs from

|B2Cp|4=p(32)=p3.|B^{2}C_{p}|_{4}=p^{\binom{3}{2}}=p^{3}.

Moreover, FF and B4CpB^{4}C_{p} are both Θ4\Theta_{4}-amenable, but B2CpB^{2}C_{p} is not. Thus, the maps B2Cp𝑓B4CpB^{2}C_{p}\xrightarrow{f}B^{4}C_{p} and B4CpptB^{4}C_{p}\to\operatorname{pt} are Θ4\Theta_{4}-amenable, but their composition is not.

The next example shows that 𝒞\mathcal{C}-cardinality need not be multiplicative when the fiber and total space are 𝒞\mathcal{C}-amenable. Moreover, the base space in this case need not be 𝒞\mathcal{C}-amenable and so in particular the class of 𝒞\mathcal{C}-amenable maps does not satisfy “left cancellation” (compare [CSY18, Theorem 2.4.5]).

Example 2.3.7.

Let p=2p=2 and Θ1=ModE1(SpK(1))\Theta_{1}=\operatorname{Mod}_{E_{1}}(\operatorname{Sp}_{K(1)}). Consider the (non-principal) fiber sequence

Σ3/C2BC2BΣ3.\Sigma_{3}/C_{2}\to BC_{2}\to B\Sigma_{3}.

It can be shown using 2.2.4, that for p=2p=2 we have

|BΣ3|1=|L^pBΣ3|0=23.|B\Sigma_{3}|_{1}=|\widehat{L}_{p}B\Sigma_{3}|_{0}=\frac{2}{3}.

In particular,

|Σ3/C2||BΣ3|=21=|BC2|1.|\Sigma_{3}/C_{2}||B\Sigma_{3}|=2\neq 1=|BC_{2}|_{1}.

Moreover, Σ3/C2\Sigma_{3}/C_{2} and BC2BC_{2} are Θ1\Theta_{1}-amenable, but BΣ3B\Sigma_{3} is not. Thus, the map BC2BΣ3BC_{2}\to B\Sigma_{3} and the composition BC2BΣ3ptBC_{2}\to B\Sigma_{3}\to\operatorname{pt} are Θ1\Theta_{1}-amenable, but BΣ3ptB\Sigma_{3}\to\operatorname{pt} 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 A𝑞BA\xrightarrow{q}B be a map of spaces and let 𝒞\mathcal{C} be an \infty-category. We say that qq is 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial) if 𝒞\mathcal{C} admits all qq-limits and qq-colimits and q:𝒞B𝒞Aq^{*}\colon\mathcal{C}^{B}\to\mathcal{C}^{A} is fully faithful (resp. an equivalence).

It is a standard fact about adjoints, that qq^{*} is fully faithful, if and only if the counit q!qc!qId𝒞Bq_{!}q^{*}\xrightarrow{c_{!}^{q}}\operatorname{Id}_{\mathcal{C}^{B}} is an isomorphism and if and only if the unit Id𝒞Buqqq\operatorname{Id}_{\mathcal{C}^{B}}\xrightarrow{u_{*}^{q}}q_{*}q^{*} is an isomorphism. Similarly, qq^{*} is an equivalence, if furthermore qqcqId𝒞Aq^{*}q_{*}\xrightarrow{c_{*}^{q}}\operatorname{Id}_{\mathcal{C}^{A}} is an isomorphism, or equivalently Id𝒞Au!qqq!\operatorname{Id}_{\mathcal{C}^{A}}\xrightarrow{u_{!}^{q}}q^{*}q_{!} is an isomorphism. Like ambidexterity and amenability, acyclicity and triviality are fiber-wise conditions:

Proposition 2.4.2.

Let 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty}. A map of spaces A𝑞BA\xrightarrow{q}B is 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial) if and only if each fiber of qq is 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial).

Proof.

For each map B~𝑓B\widetilde{B}\xrightarrow{f}B we can form the following pullback square of spaces and the induced commutative square of \infty-categories

A~\textstyle{\widetilde{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q~\scriptstyle{\tilde{q}}f~\scriptstyle{\tilde{f}}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}B~\textstyle{\widetilde{B}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}B\textstyle{B},      𝒞B\textstyle{\mathcal{C}^{B}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q^{*}}f\scriptstyle{f^{*}}𝒞B~\textstyle{\mathcal{C}^{\widetilde{B}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q~\scriptstyle{\tilde{q}^{*}}𝒞A\textstyle{\mathcal{C}^{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f~\scriptstyle{\tilde{f}^{*}}𝒞A~.\textstyle{\mathcal{C}^{\widetilde{A}}.}

By [CSY18, Lemma 2.2.3], we have

f(c!q)=c!q~Map(q~!q~,Id𝒞B~).f^{*}(c_{!}^{q})=c_{!}^{\tilde{q}}\quad\in\quad\operatorname{Map}(\tilde{q}_{!}\tilde{q}^{*},\operatorname{Id}_{\mathcal{C}^{\tilde{B}}}).
f~(u!q)=u!q~Map(Id𝒞A~,q~q~!).\tilde{f}^{*}(u_{!}^{q})=u_{!}^{\tilde{q}}\quad\in\quad\operatorname{Map}(\operatorname{Id}_{\mathcal{C}^{\tilde{A}}},\tilde{q}^{*}\tilde{q}_{!}).

Thus, if qq is 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial), then q~\tilde{q} is 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial). If ff (and hence also f~\tilde{f}) is surjective, then by the conservativity of ff^{*} (and f~\tilde{f}^{*}), the converse holds as well. Namely, if q~\tilde{q} is 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial), then qq is 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial). Applying this to any section of Bπ0B=B~B\to\pi_{0}B=\widetilde{B} yields the claim. ∎

The collections of 𝒞\mathcal{C}-acyclic and 𝒞\mathcal{C}-trivial spaces are also closed under extensions:

Corollary 2.4.3.

Let 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty} and let A𝑞BA\xrightarrow{q}B. If BB is 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial) and all the fibers of qq are 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial), then AA is 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-trivial).

Proof.

By 2.4.2, it suffices to show that 𝒞\mathcal{C}-acyclic (resp. 𝒞\mathcal{C}-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 AA be a space and let 𝒞Alg(Cat)\mathcal{C}\in\operatorname{Alg}(\operatorname{Cat}_{\infty}) which is compatible with AA-colimits. The space AA is 𝒞\mathcal{C}-acyclic if and only if the fold map 𝟙[A]𝟙\mathds{1}[A]\xrightarrow{\nabla}\mathds{1} is an isomorphism.

Proof.

Assume 𝟙[A]𝟙\mathds{1}[A]\xrightarrow{\nabla}\mathds{1} is an isomorphism. By assumption, for every X𝒞X\in\mathcal{C}, tensoring the isomorphism 𝟙[A]𝟙\mathds{1}[A]\xrightarrow{\nabla}\mathds{1} with XX gives the fold map X[A]XX[A]\xrightarrow{\nabla}X. Hence, A!AX=X[A]XA_{!}A^{*}X=X[A]\xrightarrow{\nabla}X is an isomorphism for all X𝒞X\in\mathcal{C}. ∎

The following is the prototypical example of acyclicity:

Example 2.4.5 (Bousfield).

For ESpE\in\operatorname{Sp}, we can consider the \infty-category SpE\operatorname{Sp}_{E} of EE-local spectra. By 2.4.4, a space is SpE\operatorname{Sp}_{E}-acyclic if and only if it is EE-acyclic in the sense of Bousfield, i.e. has the EE-homology of a point.

Remark 2.4.6.

For an \infty-category 𝒞\mathcal{C}, a space AA is 𝒞\mathcal{C}-acyclic if the following equivalent conditions hold:

  1. (1)

    The fold map X[A]XX[A]\xrightarrow{\nabla}X is an isomorphism for all X𝒞X\in\mathcal{C}.

  2. (2)

    The diagonal map XΔXAX\xrightarrow{\Delta}X^{A} is an isomorphism for all X𝒞X\in\mathcal{C}.

We warn the reader that for an individual object XX, it can happen that the map XΔXAX\xrightarrow{\Delta}X^{A} is an isomorphism, but X[A]XX[A]\xrightarrow{\nabla}X is not (and vise versa). As a trivial example, consider 𝒞=𝒮\mathcal{C}=\mathcal{S} with X=ptX=\operatorname{pt} and any AptA\neq\operatorname{pt}.

Relation to amenability

Under suitable ambidexterity assumptions, the notion of 𝒞\mathcal{C}-acyclicity turns out to be closely related to that of 𝒞\mathcal{C}-amenability. To begin with, recall that for a 𝒞\mathcal{C}-ambidextrous space AA and an object X𝒞X\in\mathcal{C}, the map |A|X|A|_{X} is given by the composition

XΔXANmA1X[A]X.X\xrightarrow{\Delta}X^{A}\xrightarrow{\operatorname{Nm}_{A}^{-1}}X[A]\xrightarrow{\nabla}X.

Thus, if AA is 𝒞\mathcal{C}-acyclic, then it is in particular 𝒞\mathcal{C}-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 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty} and let AA be a connected 𝒞\mathcal{C}-ambidextrous space. For every X𝒞X\in\mathcal{C}, the following are equivalent:

  1. (1)

    The fold map X[A]XX[A]\xrightarrow{\nabla}X is an isomorphism.

  2. (2)

    The diagonal XΔXAX\xrightarrow{\Delta}X^{A} is an isomorphism.

  3. (3)

    |ΩA|X|\Omega A|_{X} is invertible.

Moreover, |A|X|A|_{X} is then the inverse of |ΩA|X|\Omega A|_{X}.

Proof.

Let pt𝑒A\operatorname{pt}\xrightarrow{e}A be a base point. First, we show that if the diagonal map XΔXAX\xrightarrow{\Delta}X^{A} is an isomorphism, then |A|X|ΩA|X=IdX|A|_{X}|\Omega A|_{X}=\operatorname{Id}_{X}. We begin by reducing the claim to the fact that the map

|e|AX:AXAX|e|_{A^{*}X}\colon A^{*}X\to A^{*}X

equals the constant map on |ΩA|X|\Omega A|_{X}. Indeed, given that, by integrating along AA, we get ([CSY18, Propositions 2.1.15 and 3.1.13])

IdX=AeIdeAX=A(eIdeAX)=A|e|AX=AA(|ΩA|X)=|A|X|ΩA|X.\operatorname{Id}_{X}=\int\limits_{A\circ e}\operatorname{Id}_{e^{*}A^{*}X}=\int\limits_{A}\left(\int\limits_{e}\operatorname{Id}_{e^{*}A^{*}X}\right)=\int\limits_{A}|e|_{A^{*}X}=\int\limits_{A}A^{*}(|\Omega A|_{X})=|A|_{X}|\Omega A|_{X}.

Now, recall that the diagonal XΔXA=AAXX\xrightarrow{\Delta}X^{A}=A_{*}A^{*}X is the unit uAu_{*}^{A} of the adjunction A:𝒞𝒞A:A.A^{*}\colon\mathcal{C}\leftrightarrows\mathcal{C}^{A}\colon A_{*}. Thus, if uAu_{*}^{A} is an isomorphism at XX, then the map

Map(X,X)Map(AX,AX)\operatorname{Map}(X,X)\to\operatorname{Map}(A^{*}X,A^{*}X)

is an isomorphism. Since eA=Ide^{*}A^{*}=\operatorname{Id}, it follows by 2-out-of-3 that the map

Map(AX,AX)Map(eAX,eAX)=Map(X,X)\operatorname{Map}(A^{*}X,A^{*}X)\to\operatorname{Map}(e^{*}A^{*}X,e^{*}A^{*}X)=\operatorname{Map}(X,X)

is an isomorphism as well. Thus, it suffices to check that the maps |e|AX|e|_{A^{*}X} and A(|ΩA|X)A^{*}(|\Omega A|_{X}) coincide after applying ee^{*}. The pullback square of spaces

ΩA\textstyle{\Omega A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pt\textstyle{\operatorname{pt}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e\scriptstyle{e}pt\textstyle{\operatorname{pt}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e\scriptstyle{e}A\textstyle{A}

gives by [CSY18, Proposition 3.1.13]

e(|e|AX)=|ΩA|eAX=|ΩA|X.e^{*}(|e|_{A^{*}X})=|\Omega A|_{e^{*}A^{*}X}=|\Omega A|_{X}.

This concludes the proof of (2)\implies(3) and that |A|X|A|_{X} is the inverse of |ΩA|X|\Omega A|_{X}. A completely symmetric argument using the adjunction A!AA_{!}\dashv A^{*} instead of AAA^{*}\dashv A_{*} shows that (1)\implies(3) and that |A|X|A|_{X} is the inverse of |ΩA|X|\Omega A|_{X}. The implication (3)\implies(1) in the case that |ΩA|X|\Omega A|_{X} is invertible for all X𝒞X\in\mathcal{C} 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 |ΩA||\Omega A| is invertible. In particular, each of the conditions (1), (2), and (3) implies that |A|X|A|_{X} is the inverse of |ΩA|X|\Omega A|_{X} and so, in particular, is invertible. Finally, the composition

XΔXANmAX[A]XX\xrightarrow{\Delta}X^{A}\xrightarrow{\operatorname{Nm}_{A}}X[A]\xrightarrow{\nabla}X

is |A|X|A|_{X}. Thus, by 2-out-of-3, we also have the implication (1)\implies(2). ∎

From this we get:

Corollary 2.4.8.

Let 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty} and let AA be a connected space. The following conditions are equivalent:

  1. (1)

    AA is 𝒞\mathcal{C}-acyclic and 𝒞\mathcal{C}-ambidextrous.

  2. (2)

    ΩA\Omega A is 𝒞\mathcal{C}-amenable and 𝒞\mathcal{C} admits AA-colimits.

In which case |A|=|ΩA|1|A|=|\Omega A|^{-1}.

Proof.

This follows from 2.4.7 and the fact that if ΩA\Omega A is 𝒞\mathcal{C}-amenable, then AA is 𝒞\mathcal{C}-ambidextrous by [CSY18, Proposition 3.1.17]. ∎

We note that the assumption of 𝒞\mathcal{C}-ambidexterity in (1) of 2.4.8 can not be relaxed to weak 𝒞\mathcal{C}-ambidexterity. The following is a simple counter-example:

Example 2.4.9.

Let 𝒞=Vec𝔽p\mathcal{C}=\mathrm{Vec}_{\mathbb{F}_{p}} be the 11-category of 𝔽p\mathbb{F}_{p}-vector spaces and A=BCpA=BC_{p}. It is clear that AA is weakly 𝒞\mathcal{C}-ambidextrous (as Vec𝔽p\mathrm{Vec}_{\mathbb{F}_{p}} is semiadditive) and that BCpBC_{p} is 𝒞\mathcal{C}-acyclic. However, BCpBC_{p} is not 𝒞\mathcal{C}-ambidextrous (and CpC_{p} is not 𝒞\mathcal{C}-amenable).

Remark 2.4.10.

The notion of acyclicity can be considered for a general (not necessarily π\pi-finite) space in any \infty-category, without any assumptions on ambidexterity. However, in the presence of ambidexterity, 2.4.8 allows us to deduce the acyclicity of a given π\pi-finite space from the amenability of its loop space. This strategy was already employed in the proof of [CSY18, Theorem E] (exploiting the \infty-semiadditivity of SpT(n)\operatorname{Sp}_{T(n)}).

We conclude this subsection by showing an analogue of the equivalence of (1) and (3) in 2.4.7, for the notions of 𝒞\mathcal{C}-acyclicity and 𝒞\mathcal{C}-triviality. As before, it is clear that a 𝒞\mathcal{C}-trivial space is, in particular, 𝒞\mathcal{C}-acyclic, but there is a better statement:

Proposition 2.4.11.

Let AA be a connected space and let 𝒞\mathcal{C} be an \infty-category which admits ΩA\Omega A-colimits. Then, AA is 𝒞\mathcal{C}-trivial if and only if ΩA\Omega A is 𝒞\mathcal{C}-acyclic.

Proof.

Let pt𝑒A\operatorname{pt}\xrightarrow{e}A be a base point. The composition

𝒞A𝒞Ae𝒞\mathcal{C}\xrightarrow{A^{*}}\mathcal{C}^{A}\xrightarrow{e^{*}}\mathcal{C}

is the identity. This implies that AA^{*} is an equivalence if and only if ee^{*} is. Moreover, it also implies that ee^{*} is essentially surjective and hence an equivalence if and only if it is fully faithful. Namely, if and only if ee is 𝒞\mathcal{C}-acyclic. Since 𝒞\mathcal{C}-acyclicity is a fiber-wise condition (2.4.2), ee is 𝒞\mathcal{C}-acyclic if and only if ΩA\Omega A is 𝒞\mathcal{C}-acyclic. ∎

3 Height

In this section, we introduce the notion of “semiadditive height” for objects in higher semiadditive \infty-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 \infty-category depends on a choice of a prime pp. In fact, it suffices to have a certain “pp-typical” version of mm-semiadditivity, in which one requires ambidexterity only for mm-finite pp-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 \infty-category 𝒞\mathcal{C} which is 0-semiadditive and pp-local, the “pp-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.

pp-Typical semiadditivity

We begin with a definition of the following “pp-typical” version of higher semiadditivity:

Definition 3.1.1.

Let pp be a prime and 0m0\leq m\leq\infty. We say that

  1. (1)

    An \infty-category 𝒞\mathcal{C} is pp-typically mm-semiadditive if all mm-finite pp-spaces are 𝒞\mathcal{C}-ambidextrous.

  2. (2)

    A functor F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D} between such is pp-typically mm-semiadditive if it preserves all mm-finite pp-space colimits.

  3. (3)

    An 𝒪\mathcal{O}-monoidal \infty-category 𝒞\mathcal{C}, for some \infty-operad 𝒪\mathcal{O}, is pp-typically mm-semiadditively 𝒪\mathcal{O}-monoidal if it is pp-typically mm-semiadditive and is compatible with mm-finite pp-space colimits.

We denote by Catp-mCat\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}m}\subset\operatorname{Cat}_{\infty} the subcategory of pp-typically mm-semiadditive \infty-categories and pp-typically mm-semiadditive functors.

It is clear that an mm-semiadditive \infty-category or functor are also pp-typically mm-semiadditive for every prime pp. It is useful to know that to verify pp-typical mm-semiadditivity, it suffices to consider only the “building blocks”:

Proposition 3.1.2.

Let 0m0\leq m\leq\infty.

  1. (1)

    An \infty-category 𝒞Cat

    -
    0
    \mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}0}
    is pp-typically mm-semiadditive if and only if BkCpB^{k}C_{p} is 𝒞\mathcal{C}-ambidextrous for all k=1,,mk=1,\dots,m.

  2. (2)

    For 𝒞,𝒟Catp-m\mathcal{C},\mathcal{D}\in\operatorname{Cat}_{\infty}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}m}, a 0-semiadditive functor F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D} is pp-typically mm-semiadditive if and only if it preserves BkCpB^{k}C_{p}-(co)limits for all k=1,,mk=1,\dots,m.

Proof.

(1) The “only if” part is clear. Conversely, we need to show that if BkCpB^{k}C_{p} is 𝒞\mathcal{C}-ambidextrous for all k=1,,mk=1,\dots,m, then every mm-finite pp-space AA is 𝒞\mathcal{C}-ambidextrous. Since 𝒞\mathcal{C} is 0-semiadditive, we are reduced to the case that AA is connected. The Postnikov tower of AA can be refined to a tower of principal fibrations

A=ArA1A0=pt,A=A_{r}\to\dots\to A_{1}\to A_{0}=\operatorname{pt},

such that the fiber of each AiAi1A_{i}\to A_{i-1} is of the form BkiCpB^{k_{i}}C_{p} for some 1kim1\leq k_{i}\leq m. To show that AA is 𝒞\mathcal{C}-ambidextrous, it suffices to show that each AiAi1A_{i}\to A_{i-1} is 𝒞\mathcal{C}-ambidextrous (as 𝒞\mathcal{C}-ambidextrous maps are closed under composition). Finally, since 𝒞\mathcal{C}-ambidexterity is a fiber-wise condition, this follows from the fact that BkiCpB^{k_{i}}C_{p} is 𝒞\mathcal{C}-ambidextrous.

(2) Follows by an analogous argument to (1). ∎

In a pp-typically mm-semiadditive \infty-category 𝒞\mathcal{C}, one can discuss cardinalities of mm-finite pp-spaces. As one might expect, the Eilenberg-MacLane spaces BnCpB^{n}C_{p} play a fundamental role, and so deserve a special notation:

Definition 3.1.3.

Let 𝒞Catp-m\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}m}. For every integer 0nm0\leq n\leq m, we define p(n)𝒞|BnCp|p_{(n)}^{\mathcal{C}}\coloneqq|B^{n}C_{p}| as a natural endomorphism of the identity functor of 𝒞\mathcal{C}. We shall omit the superscript “𝒞\mathcal{C}”, whenever 𝒞\mathcal{C} is clear from the context.

A fundamental example to keep in mind is the following:

Example 3.1.4.

For Θn=ModEn(SpK(n))\Theta_{n}=\operatorname{Mod}_{E_{n}}(\operatorname{Sp}_{K(n)}), we have by 2.2.5:

p(k)Θn|BkCp|n=p(n1k)p_{(k)}^{\Theta_{n}}\coloneqq|B^{k}C_{p}|_{n}=p^{\binom{n-1}{k}}

for all n,k0n,k\geq 0.

Semiadditive height

In what follows it will be convenient to use the following terminology:

Definition 3.1.5.

Let 𝒞Cat\mathcal{C}\in\operatorname{Cat}_{\infty} and let α:Id𝒞Id𝒞\alpha\colon\operatorname{Id}_{\mathcal{C}}\to\operatorname{Id}_{\mathcal{C}} be a natural endomorphism. An object X𝒞X\in\mathcal{C} is called

  1. (1)

    α\alpha-divisible if αX\alpha_{X} is invertible.

  2. (2)

    α\alpha-complete if Map(Z,X)pt\operatorname{Map}(Z,X)\simeq\operatorname{pt} for all α\alpha-divisible ZZ.

We denote by 𝒞[α1]\mathcal{C}[\alpha^{-1}] and 𝒞^α\widehat{\mathcal{C}}_{\alpha} the full subcategories of 𝒞\mathcal{C} spanned by the α\alpha-divisible and α\alpha-complete objects respectively.

Using the operations p(n)p_{(n)} we can now define the semiadditive height:

Definition 3.1.6.

Let 𝒞Catp-m\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}m} and let 0nm<0\leq n\leq m<\infty. For every X𝒞X\in\mathcal{C} we define and denote the (semiadditive) height of XX as follows:

  1. (1)

    ht𝒞(X)n\mathrm{ht}_{\mathcal{C}}(X)\leq n, if XX is p(n)𝒞p_{(n)}^{\mathcal{C}}-divisible.

  2. (2)

    ht𝒞(X)>n\mathrm{ht}_{\mathcal{C}}(X)>n, if XX is p(n)𝒞p_{(n)}^{\mathcal{C}}-complete.

  3. (3)

    ht𝒞(X)=n\mathrm{ht}_{\mathcal{C}}(X)=n, if ht𝒞(X)n\mathrm{ht}_{\mathcal{C}}(X)\leq n and ht𝒞(X)>n1\mathrm{ht}_{\mathcal{C}}(X)>n-1.

We also extend the definition to n=m=n=m=\infty as follows. For every X𝒞X\in\mathcal{C}, we write ht𝒞(X)=\mathrm{ht}_{\mathcal{C}}(X)=\infty if and only if ht𝒞(X)>k\mathrm{ht}_{\mathcal{C}}(X)>k for all kk\in\mathbb{N}. Additionally, by convention 1<ht𝒞(X)-1<\mathrm{ht}_{\mathcal{C}}(X)\leq\infty for all XX, and ht𝒞(X)1\mathrm{ht}_{\mathcal{C}}(X)\leq-1 or ht𝒞(X)>\mathrm{ht}_{\mathcal{C}}(X)>\infty if and only if X=0X=0. We shall drop the subscript 𝒞\mathcal{C} in ht𝒞\mathrm{ht}_{\mathcal{C}}, when the \infty-category is clear from the context.

Remark 3.1.7.

We emphasize that the notation ht(X)n\mathrm{ht}(X)\leq n (and similarly ht(X)>n\mathrm{ht}(X)>n etc.) asserts that XX satisfies a certain property, and does not mean that ht(X)\mathrm{ht}(X) is a well-defined number, which can be compared with nn. We note that the only object in 𝒞\mathcal{C}, which can simultaneously have height n\leq n and >n>n is the zero object.

The motivating example for the definition of height is the following:

Example 3.1.8.

Let 𝒞\mathcal{C} be a 0-semiadditive \infty-category. An object X𝒞X\in\mathcal{C} is of height 0 if and only if p=p(0)p=p_{(0)} acts invertibly on XX, and of height >0>0 if it is pp-complete.

The first thing to show is that the notion of height behaves as the terminology suggests:

Proposition 3.1.9.

Let 𝒞Catp-m\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}m} and let 0n0n1m0\leq n_{0}\leq n_{1}\leq m be some integers. For every X𝒞X\in\mathcal{C},

  1. (1)

    If ht(X)n0\mathrm{ht}(X)\leq n_{0}, then ht(X)n1\mathrm{ht}(X)\leq n_{1}.

  2. (2)

    If ht(X)>n1\mathrm{ht}(X)>n_{1}, then ht(X)>n0\mathrm{ht}(X)>n_{0}.

Proof.

To prove (1), it suffices to show that if ht(X)n\mathrm{ht}(X)\leq n for some nm1n\leq m-1, then ht(X)n+1\mathrm{ht}(X)\leq n+1. We consider the principal fiber sequence

BnCpptBn+1Cp.B^{n}C_{p}\to\operatorname{pt}\to B^{n+1}C_{p}.

All maps and spaces in this sequence are 𝒞\mathcal{C}-ambidextrous by assumption. Since ht(X)n\mathrm{ht}(X)\leq n, we have that |BnCp|X|B^{n}C_{p}|_{X} is invertible. By 2.3.3, we get

|Bn+1Cp|X|BnCp|X=|pt|X=IdX.|B^{n+1}C_{p}|_{X}|B^{n}C_{p}|_{X}=|\operatorname{pt}|_{X}=\operatorname{Id}_{X}.

Thus, |Bn+1Cp|X|B^{n+1}C_{p}|_{X} is invertible as well, and hence ht(X)n+1\mathrm{ht}(X)\leq n+1. Claim (2) now follows from (1) by definition. ∎

Remark 3.1.10.

In any stable \infty-category 𝒞\mathcal{C}, a non-zero object can have height >0>0 for at most one prime pp. In particular, if 𝒞\mathcal{C} is pp-local, every object has height 0 for every prime p\ell\neq p (in fact, this is ‘if an only if’). Nevertheless, in this case the notion of \ell-height with respect to different primes \ell allows one to treat “prime to pp phenomena” as “height 0 phenomena” for primes p\ell\neq p.

It is also useful to consider the corresponding subcategories of objects having height in a certain range:

Definition 3.1.11.

Let 𝒞Catp-m\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}m} and let 0nm0\leq n\leq m\leq\infty. We define

𝒞n=𝒞[p(n)1],𝒞>n=𝒞^p(n),\mathcal{C}_{\leq n}=\mathcal{C}[p_{(n)}^{-1}]\quad,\quad\mathcal{C}_{>n}=\widehat{\mathcal{C}}_{p_{(n)}},
𝒞n=𝒞n𝒞>n1=𝒞[p(n)1]^p(n1)=𝒞^p(n1)[p(n)1].\mathcal{C}_{n}=\mathcal{C}_{\leq n}\cap\mathcal{C}_{>n-1}=\widehat{\mathcal{C}[p_{(n)}^{-1}]}_{p_{(n-1)}}=\widehat{\mathcal{C}}_{p_{(n-1)}}[p_{(n)}^{-1}].

the full subcategories of 𝒞\mathcal{C} spanned by objects of height n\leq n, >n>n and nn respectively. We also write Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n, >n>n or nn, if 𝒞=𝒞n\mathcal{C}=\mathcal{C}_{\leq n}, 𝒞>n\mathcal{C}_{>n} or 𝒞n\mathcal{C}_{n} respectively.

The above defined subcategories are themselves pp-typically mm-semiadditive:

Proposition 3.1.12.

Let 𝒞Catp-m\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}m}. For every n=0,,mn=0,\dots,m, the full subcategories 𝒞n,\mathcal{C}_{\leq n}, 𝒞>n\mathcal{C}_{>n} and 𝒞n\mathcal{C}_{n} are closed under limits in 𝒞\mathcal{C}. In particular, they are pp-typically mm-semiadditive, and are furthermore mm-semiadditive if 𝒞\mathcal{C} is.

Proof.

An object X𝒞X\in\mathcal{C} belongs to 𝒞n\mathcal{C}_{\leq n} if and only if XX is p(n)p_{(n)}-divisible. Thus 𝒞n\mathcal{C}_{\leq n} is closed under all limits which exist in 𝒞\mathcal{C}. By definition, 𝒞>n\mathcal{C}_{>n} and 𝒞\mathcal{C}_{\infty} are closed under limits in 𝒞\mathcal{C} as well. Thus, for n<n<\infty we have that 𝒞n=𝒞n𝒞>n1\mathcal{C}_{n}=\mathcal{C}_{\leq n}\cap\mathcal{C}_{>n-1} is also closed under limits in 𝒞\mathcal{C}. Finally, by 2.1.4(3), it follows that all these subcategories are pp-typically mm-semiadditive and are furthermore mm-semiadditive if 𝒞\mathcal{C} 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 F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D} be a map in Catp-m\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}m}. For all X𝒞X\in\mathcal{C} and 0nm0\leq n\leq m, if ht𝒞(X)n\mathrm{ht}_{\mathcal{C}}(X)\leq n then ht𝒟(F(X))n\mathrm{ht}_{\mathcal{D}}(F(X))\leq n. If FF is conservative, then the converse holds as well.

Proof.

This follows immediately from the fact that FF maps p(n)𝒞p_{(n)}^{\mathcal{C}} to p(n)𝒟p_{(n)}^{\mathcal{D}}. ∎

In contrast, the following example shows that a higher semiadditive functor need not preserve lower bounds on height:

Example 3.1.14.

The 0-semiadditive functor L:Sp(p)SpL_{\mathbb{Q}}\colon\operatorname{Sp}_{(p)}\to\operatorname{Sp}_{\mathbb{Q}} maps the pp-complete sphere 𝕊^p\widehat{\mathbb{S}}_{p}, which is of height >0>0, to a non-zero object 𝕊^p\mathbb{Q}\otimes\widehat{\mathbb{S}}_{p} of height 0.

For an inclusion of a full subcategory, we can say a bit more:

Proposition 3.1.15.

Let 𝒞Catp-m\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}m} and let 𝒞𝒞\mathcal{C}^{\prime}\subseteq\mathcal{C} be a full subcategory closed under mm-finite pp-space (co)limits. Given X𝒞X\in\mathcal{C}^{\prime} and 0nm0\leq n\leq m, we have

  1. (1)

    ht𝒞(X)n\mathrm{ht}_{\mathcal{C}^{\prime}}(X)\leq n if and only if ht𝒞(X)n\mathrm{ht}_{\mathcal{C}}(X)\leq n.

  2. (2)

    ht𝒞(X)>n\mathrm{ht}_{\mathcal{C}}(X)>n implies ht𝒞(X)>n\mathrm{ht}_{\mathcal{C}^{\prime}}(X)>n.

Proof.

(1) follows from 3.1.13 applied to the inclusion 𝒞𝒞\mathcal{C}^{\prime}\hookrightarrow\mathcal{C}. For (2), if ht𝒞(X)>n\mathrm{ht}_{\mathcal{C}}(X)>n, then for every Z𝒞nZ\in\mathcal{C}_{\leq n}, we have Map(Z,X)pt\operatorname{Map}(Z,X)\simeq\operatorname{pt}. We now observe that by (1), we have 𝒞n=𝒞𝒞n\mathcal{C}_{\leq n}^{\prime}=\mathcal{C}^{\prime}\cap\mathcal{C}_{\leq n}. Thus, for every Z𝒞nZ^{\prime}\in\mathcal{C}_{\leq n}^{\prime}, we have Map(Z,X)pt\operatorname{Map}(Z^{\prime},X)\simeq\operatorname{pt}, which by definition means ht𝒞(X)>n\mathrm{ht}_{\mathcal{C}^{\prime}}(X)>n. ∎

In presence of a pp-typically mm-semiadditively monoidal structure, an upper bound on the height of the unit implies an upper bound on the height of the \infty-category:

Corollary 3.1.16.

Let 𝒞\mathcal{C} be pp-typically mm-semiadditively monoidal \infty-category. For every 0nm0\leq n\leq m, we have Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n if and only if ht𝒞(𝟙)n\mathrm{ht}_{\mathcal{C}}(\mathds{1})\leq n.

Proof.

Given X𝒞X\in\mathcal{C}, the functor X():𝒞𝒞X\otimes(-)\colon\mathcal{C}\to\mathcal{C} is pp-typically mm-semiadditive. Thus, the claim follows from 3.1.13. ∎

3.2 Bounded Height

In this subsection, we study the implications for a higher semiadditive \infty-category of having bounded height. These results generalize the previously discussed facts regarding the \infty-semiadditive structure of semirational \infty-categories (i.e. of height 0), 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 0-semiadditive pp-local \infty-category, there is no difference between mm-semiadditivity and pp-typical mm-semiadditivity.

Amenability & acyclicity

The results on amenability and acyclicity from 2.4 imply the following equivalent characterizations of height n\leq n:

Proposition 3.2.1.

Let 𝒞Catp-n\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}n}, which admits Bn+1CpB^{n+1}C_{p}-(co)limits. The following properties are equivalent:

  1. (1)

    Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n (i.e. BnCpB^{n}C_{p} is 𝒞\mathcal{C}-amenable).

  2. (2)

    Bn+1CpB^{n+1}C_{p} is 𝒞\mathcal{C}-acyclic.

  3. (3)

    Bn+2CpB^{n+2}C_{p} is 𝒞\mathcal{C}-trivial.

We can therefore characterize the height of an \infty-category in ways that do not make an explicit reference to the higher semiadditive structure.

Proof.

The equivalence of (1) and (2) follows from 2.4.8 and the equivalence of (2) and (3) from 2.4.11. ∎

The following can be seen as a (pp-typical) generalization of the fact that a semirational \infty-category is automatically \infty-semiadditive:

Proposition 3.2.2.

Let 𝒞Catp-n\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}n} such that Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n, and assume 𝒞\mathcal{C} admits Bn+1CpB^{n+1}C_{p}-(co)limits. Then, 𝒞\mathcal{C} is pp-typically \infty-semiadditive.

Proof.

By 3.1.2(1), it suffice to prove that

  • ()(*)

    BkCpB^{k}C_{p} is 𝒞\mathcal{C}-amenable (and ,in particular, 𝒞\mathcal{C}-ambidextrous) and 𝒞\mathcal{C} admits all Bk+1CpB^{k+1}C_{p}-colimits.

holds for all knk\geq n. We shall prove this by induction on kk. The base case k=nk=n is given by assumption. Assume ()(*) holds for some knk\geq n and consider the fiber sequence

BkCpptBk+1Cp.B^{k}C_{p}\to\operatorname{pt}\to B^{k+1}C_{p}.

By the inductive hypothesis, |BkCp||B^{k}C_{p}| is invertible and 𝒞\mathcal{C} admits Bk+1CpB^{k+1}C_{p}-(co)limits. Therefore by [CSY18, Propostion 3.1.17], the space Bk+1CpB^{k+1}C_{p} is 𝒞\mathcal{C}-ambidextrous. Moreover, by 2.3.3, |Bk+1Cp||B^{k+1}C_{p}| is invertible as well. Finally, by 3.2.1(3), the diagonal functor 𝒞𝒞Bk+2Cp\mathcal{C}\to\mathcal{C}^{B^{k+2}C_{p}} is an equivalence and hence in particular 𝒞\mathcal{C} admits Bk+2CpB^{k+2}C_{p}-(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 𝒞Catp-\mathcal{C}\in\operatorname{Cat}_{\infty}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}\infty} such that Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n. For every π\pi-finite pp-space AA, the following hold:

  1. (1)

    If AA is (n1)(n-1)-connected, then AA is 𝒞\mathcal{C}-amenable.

  2. (2)

    If AA is nn-connected, then AA is 𝒞\mathcal{C}-acyclic.

  3. (3)

    If AA is (n+1)(n+1)-connected, then AA is 𝒞\mathcal{C}-trivial.

Proof.

For (1), let AA be an (n1)(n-1)-connected π\pi-finite pp-space. The Postnikov tower of AA can be refined to a tower of principal fibrations

A=ArA1A0=pt,A=A_{r}\to\dots\to A_{1}\to A_{0}=\operatorname{pt},

such that the fiber of each AiAi1A_{i}\to A_{i-1} is of the form BkiCpB^{k_{i}}C_{p} for some kink_{i}\geq n. Since we assumed Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n, we also have Ht(𝒞)ki\mathrm{Ht}(\mathcal{C})\leq k_{i} (3.1.9), and hence all the spaces BkiCpB^{k_{i}}C_{p} are 𝒞\mathcal{C}-amenable. By 2.3.5, the class of 𝒞\mathcal{C}-amenable spaces is closed under principal extensions, and therefore AA is 𝒞\mathcal{C}-amenable. Now, (2) follows from 2.4.8 and (1) applied to ΩA\Omega A. Similarly, (3) follows from 2.4.11 and (2) applied to ΩA\Omega A. ∎

Example 3.2.4.

Let 𝒞\mathcal{C} be semirational, and so in particular of height 0 (such as Vec\mathrm{Vec}_{\mathbb{Q}} or Sp\operatorname{Sp}_{\mathbb{Q}}). By 2.2.2, for every π\pi-finite pp-space AA, the cardinality |A||A| is a sum of positive rational numbers. Thus, |A||A| is invertible if and only if AA is non-empty (i.e. (1)(-1)-connected). The map XXAX\to X^{A} is an equivalence for all XX, if and only if AA is connected, and 𝒞𝒞A\mathcal{C}\to\mathcal{C}^{A} is an equivalence if and only if AA 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 𝒞Catp-\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}_{p}\text{-}\infty} such that Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n. Given a principal fibration of π\pi-finite pp-spaces

FAB,F\to A\to B,

if FF is (n1)(n-1)-connected, then |A|=|F||B||A|=|F|\cdot|B|. In particular,

p(k)=p(n)(1)kn,kn.p_{(k)}=p_{(n)}^{(-1)^{k-n}},\quad\forall k\geq n.
Proof.

By 3.2.3(1), the space FF is 𝒞\mathcal{C}-amenable. Thus, by 2.3.3, we have |A|=|F||B||A|=|F|\cdot|B|. By an inductive application of this to the fiber sequence

BkCpptBk+1Cp,B^{k}C_{p}\to\operatorname{pt}\to B^{k+1}C_{p},

we obtain the formula p(k)=p(n)(1)knp_{(k)}=p_{(n)}^{(-1)^{k-n}} for all knk\geq n. ∎

Furthermore, using 3.2.5 together with a principal refinement of the Postnikov tower (as in the proof of 3.2.2), one can reduce the computation of the 𝒞\mathcal{C}-cardinality of all π\pi-finite pp-spaces to those of connected nn-finite pp-spaces.

The pp-local Case

In many situations of interest, the \infty-category 𝒞\mathcal{C} under consideration is 0-semiadditive, pp-local and admits all 11-finite colimits. In this case, 𝒞\mathcal{C} is automatically \ell-typically \infty-semiadditive for all primes p\ell\neq p (3.2.2). Moreover, in this case there is essentially no difference between higher semiadditivity and pp-typical higher semiadditivity:

Proposition 3.2.6.

Let 𝒞\mathcal{C} be a 0-semiadditive pp-local \infty-category which admits all 11-finite limits and colimits. The \infty-category 𝒞\mathcal{C} is pp-typically mm-semiadditive if and only if it is mm-semiadditive.

Proof.

Assume that 𝒞\mathcal{C} is pp-typically mm-semiadditive. By [HL13, Proposition 4.4.16] and the fact that the space BCpBC_{p} is 𝒞\mathcal{C}-ambidextrous, we get that 𝒞\mathcal{C} is 11-semiadditive. To get higher semiadditivity, we first have by assumption that 𝒞\mathcal{C} is pp-typically mm-semiadditive. Moreover, since 𝒞\mathcal{C} is pp-local, it is \ell-typically mm-semiadditive for all p\ell\neq p (3.2.2). Thus, BkCB^{k}C_{\ell} is 𝒞\mathcal{C}-ambidextrous for all primes \ell and all integers k=2,mk=2,\dots m. By inductive application of [HL13, Proposition 4.4.19], it follows that 𝒞\mathcal{C} is mm-semiadditive. We note that in both [HL13, Proposition 4.4.16] and [HL13, Proposition 4.4.19], one assumes that 𝒞\mathcal{C} admits all small limits and colimits. However, the proofs use only the limits and colimits which we assumed in the statement. ∎

When applied to pp-local \infty-categories, the main results of this subsection can be summarized as follows:

Theorem 3.2.7.

Let 𝒞\mathcal{C} be a 0-semiadditive pp-local \infty-category, which admits all (n+1)(n+1)-finite limits and colimits. If 𝒞\mathcal{C} is pp-typically nn-semiadditive such that Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n, then 𝒞\mathcal{C} is \infty-semiadditive. Moreover, for every π\pi-finite space AA:

  1. (1)

    If AA is (n1)(n-1)-connected and nilpotent, then AA is 𝒞\mathcal{C}-amenable.

  2. (2)

    If AA is nn-connected, then AA is 𝒞\mathcal{C}-acyclic.

  3. (3)

    If AA is (n+1)(n+1)-connected, then AA is 𝒞\mathcal{C}-trivial.

Proof.

Since Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n and 𝒞\mathcal{C} admits Bn+1CpB^{n+1}C_{p}-limits and colimits, it follows that 𝒞\mathcal{C} is pp-typically \infty-semiadditive (3.2.2). Since 𝒞\mathcal{C} is pp-local, it follows that 𝒞\mathcal{C} is in fact \infty-semiadditive (3.2.6). For (1), we observe that if AA is π\pi-finite and nilpotent, then A=A()A=\prod_{\ell}A_{(\ell)} where \ell ranges over primes and A()A_{(\ell)} is a π\pi-finite \ell-space which is contractible for almost all \ell (e.g. [PS17, Theorem 5.7]). Since we have |A|=|A()||A|=\prod_{\ell}|A_{(\ell)}| (2.1.10), the 𝒞\mathcal{C}-amenability of AA follows from the 𝒞\mathcal{C}-amenability of all the A()A_{(\ell)}-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 ΩA\Omega A 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 n=1n=1. However, in this case it can not be dropped. Indeed, by 2.3.7, at the prime p=2p=2 we have |BΣ3|1=23,|B\Sigma_{3}|_{1}=\frac{2}{3}, which is not-invertible. With a little more effort one can show that if Ht(𝒞)1\mathrm{Ht}(\mathcal{C})\leq 1, then |A||A| is invertible for every connected π\pi-finite space AA, such that that the pp-Sylow subgroup of π1A\pi_{1}A is normal. In other words, the pp-primary fusion in the fundamental group is the only obstruction for the invertibility of |A||A|.

3.3 Semiadditive Redshift

By 2.2.8, the \infty-category Catm-fin\mathrm{Cat}_{m\text{-}\mathrm{fin}} is mm-semiadditive and hence given 𝒞Catm-fin\mathcal{C}\in\mathrm{Cat}_{m\text{-}\mathrm{fin}}, we can discuss ht(𝒞)\mathrm{ht}(\mathcal{C}) for various primes pp, which is the height of 𝒞\mathcal{C} as an object of Catm-fin\mathrm{Cat}_{m\text{-}\mathrm{fin}}. However, if 𝒞\mathcal{C} itself is pp-typically higher semiadditive, then we have also defined Ht(𝒞)\mathrm{Ht}(\mathcal{C}), as the height of the objects of 𝒞\mathcal{C}. These two notions of height do not coincide. Rather, we shall now show that ht(𝒞)\mathrm{ht}(\mathcal{C}) exceeds Ht(𝒞)\mathrm{Ht}(\mathcal{C}) 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 F:𝒞𝒟:GF\colon\mathcal{C}\leftrightarrows\mathcal{D}\colon G be an adjunction. If GFGF is an equivalence, then FF is fully faithful.

Proof.

The functor FF is fully faithful if and only if the unit Id𝑢GF\operatorname{Id}\xrightarrow{u}GF is an isomorphism. The unit is part of a monad structure on GFGF making it a monoid in the homotopy category of Fun(𝒞,𝒞)\operatorname{Fun}(\mathcal{C},\mathcal{C}), whose monoidal structure is given by composition. It therefore suffices to show that given a monoid MM in any monoidal (ordinary) category, if MM is invertible with respect to the monoidal structure, then the unit map 𝟙𝑢M\mathds{1}\xrightarrow{u}M is an isomorphism. We observe that the multiplication map MM𝑚MM\otimes M\xrightarrow{m}M is always a left inverse of 1u1\otimes u. Thus, since M()M\otimes(-) is an equivalence, it follows that uu admits a left inverse as well. To conclude the proof, it remains to show that uu admits a right inverse. Since MM is invertible, it is in particular dualizable, and the dual MM^{\vee} of MM is its inverse. More precisely, the duality datum maps 𝟙𝜂MM\mathds{1}\xrightarrow{\eta}M\otimes M^{\vee} and MM𝜀𝟙M^{\vee}\otimes M\xrightarrow{\varepsilon}\mathds{1} are isomorphisms and exhibit MM^{\vee} as the inverse of MM. Consider the following commutative diagram

M𝟙\textstyle{M\otimes\mathds{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1η1\scriptstyle{1\otimes\eta\otimes 1}\scriptstyle{\wr}1u\scriptstyle{1\otimes u}MM\textstyle{M\otimes M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\wr}1η1\scriptstyle{1\otimes\eta\otimes 1}MMM𝟙\textstyle{M\otimes M\otimes M^{\vee}\otimes\mathds{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}m11\scriptstyle{m\otimes 1\otimes 1}111u\scriptstyle{1\otimes 1\otimes 1\otimes u}MMMM\textstyle{M\otimes M\otimes M^{\vee}\otimes M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}m11\scriptstyle{m\otimes 1\otimes 1}11ε\scriptstyle{1\otimes 1\otimes\varepsilon}MM\textstyle{M\otimes M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}m\scriptstyle{m}MM𝟙\textstyle{M\otimes M^{\vee}\otimes\mathds{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}11u\scriptstyle{1\otimes 1\otimes u}MMM\textstyle{M\otimes M^{\vee}\otimes M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}1ε\scriptstyle{1\otimes\varepsilon}M\textstyle{M}

The clockwise composition is the identity and hence so is the counter-clockwise composition. It follows that the map

MM𝟙11uMMMM\otimes M^{\vee}\otimes\mathds{1}\xrightarrow{1\otimes 1\otimes u}M\otimes M^{\vee}\otimes M

admits a right inverse. Since the functor MM()M\otimes M^{\vee}\otimes(-) is an equivalence, it follows that uu admits a right inverse. ∎

We can now state and prove the following “Semiadditive Redshift” result, which can be informally summarized as “ht=Ht+1\mathrm{ht}=\mathrm{Ht}+1”.

Theorem 3.3.2 (Semiadditive Redshift).

Let 0nm0\leq n\leq m\leq\infty be integers and let 𝒞Cat(m+1)-fin\mathcal{C}\in\mathrm{Cat}_{(m+1)\text{-}\mathrm{fin}}. If 𝒞\mathcal{C} is pp-typically mm-semiadditive, then

  1. (1)

    Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n if and only if ht(𝒞)n+1\mathrm{ht}(\mathcal{C})\leq n+1.

  2. (2)

    Ht(𝒞)>n\mathrm{Ht}(\mathcal{C})>n if and only if ht(𝒞)>n+1\mathrm{ht}(\mathcal{C})>n+1.

In particular, an \infty-semiadditive \infty-category is of height nn if and only if, as an object of Cat-fin\mathrm{Cat}_{\infty\text{-}\mathrm{fin}} (and hence also Cat

-
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
), it is of height n+1n+1.

Proof.

Denote B=Bn+1CpB=B^{n+1}C_{p}. For (1), we observe that Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n if and only if BB is 𝒞\mathcal{C}-acyclic (3.2.1), namely that BB^{*} is fully faithful. On the other hand, the natural endomorphism p(n+1)=|B|p_{(n+1)}=|B| of the identity functor of Cat(m+1)-fin\mathrm{Cat}_{(m+1)\text{-}\mathrm{fin}}, acts on 𝒞\mathcal{C} by the functor B!B:𝒞𝒞B_{!}B^{*}\colon\mathcal{C}\to\mathcal{C} (2.2.8). Thus, ht(𝒞)n+1\mathrm{ht}(\mathcal{C})\leq n+1, if and only if B!BB_{!}B^{*} is invertible. Now, if BB^{*} is fully faithful, then B!B=Id𝒞B_{!}B^{*}=\operatorname{Id}_{\mathcal{C}} and is in particular invertible. On the other hand, if B!BB_{!}B^{*} is invertible, then BB^{*} is fully faithful by the dual of 3.3.1. Thus, Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n if and only if ht(𝒞)n+1\mathrm{ht}(\mathcal{C})\leq n+1.

For (2), we first assume that Ht(𝒞)>n\mathrm{Ht}(\mathcal{C})>n, and show that ht(𝒞)>n+1\mathrm{ht}(\mathcal{C})>n+1. Given 𝒟Cat(m+1)-fin\mathcal{D}\in\mathrm{Cat}_{(m+1)\text{-}\mathrm{fin}} such that ht(𝒟)n+1\mathrm{ht}(\mathcal{D})\leq n+1, we need to show that every (m+1)(m+1)-finite colimit preserving functor F:𝒟𝒞F\colon\mathcal{D}\to\mathcal{C} must be zero. For every X𝒟X\in\mathcal{D}, we have X[B]=B!BXXX[B]=B_{!}B^{*}X\overset{\sim}{\longrightarrow}X. Since FF is (m+1)(m+1)-finite colimit preserving, we get also F(X)[B]F(X)F(X)[B]\overset{\sim}{\longrightarrow}F(X). Thus, by 2.4.7, we get ht(F(X))n\mathrm{ht}(F(X))\leq n. Since Ht(𝒞)>n\mathrm{Ht}(\mathcal{C})>n, the only object of height n\leq n in 𝒞\mathcal{C} is zero and hence F(X)=0F(X)=0. Thus, FF is zero, which proves that ht(𝒞)>n+1\mathrm{ht}(\mathcal{C})>n+1. Conversely, assume ht(𝒞)>n+1\mathrm{ht}(\mathcal{C})>n+1, to show that Ht(𝒞)>n\mathrm{Ht}(\mathcal{C})>n, consider the full subcategory 𝒞n𝒞\mathcal{C}_{\leq n}\subseteq\mathcal{C} spanned by objects of height n\leq n in 𝒞\mathcal{C}, which is also pp-typically mm-semiadditive (3.1.12). By definition, Ht(𝒞n)n\mathrm{Ht}(\mathcal{C}_{\leq n})\leq n, and hence by (1), ht(𝒞n)n+1\mathrm{ht}(\mathcal{C}_{\leq n})\leq n+1, so the inclusion functor 𝒞n𝒞\mathcal{C}_{\leq n}\hookrightarrow\mathcal{C} must be zero. It follows that 𝒞n\mathcal{C}_{\leq n} is zero and therefore Ht(𝒞)>n\mathrm{Ht}(\mathcal{C})>n.

It follows from (1) and (2) that if Ht(𝒞)=n\mathrm{Ht}(\mathcal{C})=n, then ht(𝒞)=n+1\mathrm{ht}(\mathcal{C})=n+1 when 𝒞\mathcal{C} is considered as an object of Cat-fin\mathrm{Cat}_{\infty\text{-}\mathrm{fin}}. The parenthetical remark follows from 3.1.15 and 2.2.11. ∎

Recall from 2.2.10, that for every 𝒞Cat

-
m
\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}m}
, the space of objects 𝒞\mathcal{C}^{\simeq} is endowed with an mm-commutative coCartesian monoid structure making it an object of the mm-semiadditive \infty-category CMonm\operatorname{CMon}_{m}. 3.3.2 has the following corollary:

Corollary 3.3.3.

Let 𝒞Cat

-
\mathcal{C}\in\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
, such that Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n. The space of objects 𝒞\mathcal{C}^{\simeq} with the higher coCartesian structure satisfies ht(𝒞)n+1\mathrm{ht}(\mathcal{C}^{\simeq})\leq n+1, as an object of CMon\operatorname{CMon}_{\infty}.

Proof.

We have seen in 3.3.2, that when we consider 𝒞\mathcal{C} as an object of Cat-fin\mathrm{Cat}_{\infty\text{-}\mathrm{fin}}, we have ht(𝒞)n+1\mathrm{ht}(\mathcal{C})\leq n+1. The space of objects 𝒞CMon\mathcal{C}^{\simeq}\in\operatorname{CMon}_{\infty} can be identified with the image of 𝒞\mathcal{C} under the \infty-semiadditive functor

hom(𝒮-fin,):Cat-finCMon.\hom(\mathcal{S}_{\infty\text{-}\mathrm{fin}},-)\colon\mathrm{Cat}_{\infty\text{-}\mathrm{fin}}\to\operatorname{CMon}_{\infty}.

Thus, by 3.1.13, we have ht(𝒞)n+1\mathrm{ht}(\mathcal{C}^{\simeq})\leq n+1. ∎

Example 3.3.4.

Let RR be a T(n)T(n)-local ring spectrum. By 4.4.3 and 4.4.5, the \infty-commutative monoid ModR(SpT(n))\operatorname{Mod}_{R}(\operatorname{Sp}_{T(n)})^{\simeq} is of height n+1\leq n+1 in CMon\operatorname{CMon}_{\infty}. In particular, this applies to Morava EE-theory R=EnR=E_{n}. This suggests a relation between the “semiadditive redshift” of 3.3.2 and the “chromatic redshift” in algebraic KK-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 BnCpB^{n}C_{p} is 𝒞\mathcal{C}-amenable if and only if Bn+1CpB^{n+1}C_{p} is 𝒞\mathcal{C}-acyclic (2.4.8). In 2.4.11 we categorified this fact by showing that Bn+1CpB^{n+1}C_{p} is 𝒞\mathcal{C}-acyclic, if and only if Bn+2CpB^{n+2}C_{p} is 𝒞\mathcal{C}-trivial. Similarly, 3.3.2 can be categorified as follows. Let Catn-htCat-fin\mathrm{Cat}^{n\text{-}\mathrm{ht}}\subseteq\mathrm{Cat}_{\infty\text{-}\mathrm{fin}} be the full subcategory spanned by the \infty-semiadditive \infty-categories 𝒞\mathcal{C} such that Ht(𝒞)=n\mathrm{Ht}(\mathcal{C})=n.

Lemma 3.3.5.

The full subcategory Catn-htCat-fin\mathrm{Cat}^{n\text{-}\mathrm{ht}}\subseteq\mathrm{Cat}_{\infty\text{-}\mathrm{fin}} is closed under π\pi-finite colimits.

Proof.

We have

Catn-htCat

-
Cat-fin
.
\mathrm{Cat}^{n\text{-}\mathrm{ht}}\subseteq\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}\subseteq\mathrm{Cat}_{\infty\text{-}\mathrm{fin}}.

By 3.3.2, we have

Catn-ht=(Cat

-
)
n+1
.
\mathrm{Cat}^{n\text{-}\mathrm{ht}}=(\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}\infty})_{n+1}.

Thus, by 3.1.12, Catn-ht\mathrm{Cat}^{n\text{-}\mathrm{ht}} is closed under limits in Cat

-
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
. Additionally, Cat

-
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
is closed under limits in Cat-fin\mathrm{Cat}_{\infty\text{-}\mathrm{fin}} by 2.2.11. Therefore, Catn-ht\mathrm{Cat}^{n\text{-}\mathrm{ht}} is closed under limits in Cat-fin\mathrm{Cat}_{\infty\text{-}\mathrm{fin}}. Since Cat-fin\mathrm{Cat}_{\infty\text{-}\mathrm{fin}} is \infty-semiadditive, it follows that Catn-ht\mathrm{Cat}^{n\text{-}\mathrm{ht}} is also closed under π\pi-finite colimits in Cat-fin\mathrm{Cat}_{\infty\text{-}\mathrm{fin}}. ∎

Hence, in particular, Catn-ht\mathrm{Cat}^{n\text{-}\mathrm{ht}} admits π\pi-finite colimits and is therefore an object of the \infty-semiadditive \infty-category Cat^-fin\widehat{\mathrm{Cat}}_{\infty\text{-}\mathrm{fin}} of large \infty-categories, which admit π\pi-finite colimits and functors preserving them.

Proposition 3.3.6.

The \infty-category Catn-ht\mathrm{Cat}^{n\text{-}\mathrm{ht}} is an object of height n+2n+2 in Cat^-fin\widehat{\mathrm{Cat}}_{\infty\text{-}\mathrm{fin}}.

Proof.

By 3.3.5, Catn-ht\mathrm{Cat}^{n\text{-}\mathrm{ht}} is closed under π\pi-finite colimits in Cat-fin\mathrm{Cat}_{\infty\text{-}\mathrm{fin}} and hence is \infty-semiadditive by 2.1.4. Thus, by 3.3.2, it suffices to show that for every 𝒞Catn-ht\mathcal{C}\in\mathrm{Cat}^{n\text{-}\mathrm{ht}} we have ht(𝒞)=n+1\mathrm{ht}(\mathcal{C})=n+1 as an object of Cat-fin\mathrm{Cat}_{\infty\text{-}\mathrm{fin}}, and hence also as an object of Catn-ht\mathrm{Cat}^{n\text{-}\mathrm{ht}} (3.1.15). This follows again from 3.3.2 and the fact that Ht(𝒞)=n\mathrm{Ht}(\mathcal{C})=n. ∎

4 Stability

So far, we have been considering general higher semiadditive \infty-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 \infty-category decomposes completely according to height. Second, inspired by [Lurb], we study semisimplicity properties of local systems valued in general stable \infty-categories of semiadditive height nn. Finally, we show that SpK(n)\operatorname{Sp}_{K(n)} and SpT(n)\operatorname{Sp}_{T(n)} are indeed of semiadditive height nn.

4.1 Recollement

A central tool, which will be used several times in the following subsections, is that of a recollement of stable \infty-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 \infty-categories following the exposition in [Lura, Section A.8.1].

Definition 4.1.1.

Let 𝒞\mathcal{C} be a stable \infty-category and 𝒞𝒞\mathcal{C}_{\circ}\subseteq\mathcal{C} a full stable subcategory. We define the right orthogonal complement 𝒞𝒞\mathcal{C}_{\circ}^{\bot}\subseteq\mathcal{C} to be the full subcategory consisting of objects Y𝒞,Y\in\mathcal{C}, such that Map(X,Y)pt\operatorname{Map}(X,Y)\simeq\operatorname{pt} for all X𝒞X\in\mathcal{C}_{\circ}.

Recall from [Lura, Proposition A.8.20], that if the inclusion 𝒞𝒞\mathcal{C}_{\circ}\hookrightarrow\mathcal{C} admits both a left adjoint LL and a right adjoint RR, then 𝒞\mathcal{C} is a recollement of 𝒞\mathcal{C}_{\circ} and 𝒞\mathcal{C}_{\circ}^{\perp} in the sense of [Lura, Section A.8.1]. In particular, the \infty-category 𝒞\mathcal{C} can be identified with the \infty-category of sections of the cartesian fibration over Δ1\Delta^{1}, classified by the functor L|𝒞:𝒞𝒞.L|_{\mathcal{C}_{\circ}^{\perp}}\colon\mathcal{C}_{\circ}^{\perp}\to\mathcal{C}_{\circ}.

Definition 4.1.2.

We shall say that an inclusion of stable \infty-categories 𝒞𝒞\mathcal{C}_{\circ}\hookrightarrow\mathcal{C}, that admits both a left and a right adjoint, exhibits 𝒞\mathcal{C} as a recollement of 𝒞\mathcal{C}_{\circ} and 𝒞\mathcal{C}_{\circ}^{\perp}.

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 𝒞=Sp(p)\mathcal{C}=\operatorname{Sp}_{(p)}, the inclusion of the full subcategory 𝒞=SpSp(p)\mathcal{C}_{\circ}=\operatorname{Sp}_{\mathbb{Q}}\subseteq\operatorname{Sp}_{(p)} admits both a left and a right adjoint, and we have

Sp=Sp^pSp(p),\operatorname{Sp}_{\mathbb{Q}}^{\bot}=\widehat{\operatorname{Sp}}_{p}\subseteq\operatorname{Sp}_{(p)},

is the full subcategory spanned by pp-complete spectra. The recollement statement in this case recovers the classical pp-local arithmetic square for spectra. More generally, the full subcategory

LnfSpSpk=0nT(k)Sp(p)L_{n}^{f}\operatorname{Sp}\coloneqq\operatorname{Sp}_{\oplus_{k=0}^{n}T(k)}\subseteq\operatorname{Sp}_{(p)}

exhibits Sp(p)\operatorname{Sp}_{(p)} as a recollement of LnfSpL_{n}^{f}\operatorname{Sp} and SpF(n+1)\operatorname{Sp}_{F(n+1)}, where F(n+1)F(n+1) is a finite spectrum of type n+1n+1. In particular, the inclusion Ln1fSpLnfSpL_{n-1}^{f}\operatorname{Sp}\subseteq L_{n}^{f}\operatorname{Sp} exhibits LnfSpL_{n}^{f}\operatorname{Sp} as a recollement of Ln1fSpL_{n-1}^{f}\operatorname{Sp} and SpT(n)\operatorname{Sp}_{T(n)}. This recovers the classical telescopic fracture square at height nn. By the Smash Product Theorem,

LnSpSpk=0nK(k)Sp(p)L_{n}\operatorname{Sp}\coloneqq\operatorname{Sp}_{\oplus_{k=0}^{n}K(k)}\subseteq\operatorname{Sp}_{(p)}

is also a smashing localization, and similarly, the inclusion Ln1SpLnSpL_{n-1}\operatorname{Sp}\subseteq L_{n}\operatorname{Sp} exhibits LnSpL_{n}\operatorname{Sp} as a recollement of Ln1SpL_{n-1}\operatorname{Sp} and SpK(n)\operatorname{Sp}_{K(n)}. This recovers the classical chromatic fracture square at height nn.

Given a recollement 𝒞𝒞\mathcal{C}_{\circ}\subseteq\mathcal{C}, the functor L|𝒞:𝒞𝒞L|_{\mathcal{C}_{\circ}^{\bot}}\colon\mathcal{C}_{\circ}^{\bot}\to\mathcal{C}_{\circ} encodes the “gluing data” in the construction of 𝒞\mathcal{C} from the \infty-categories 𝒞\mathcal{C}_{\circ} and 𝒞\mathcal{C}_{\circ}^{\bot}. A particularly simple instance of a recollement is when this gluing data is trivial:

Proposition 4.1.4.

Given a recollement 𝒞𝒞\mathcal{C}_{\circ}\subseteq\mathcal{C}, the following are equivalent:

  1. (1)

    The functor L|𝒞:𝒞𝒞L|_{\mathcal{C}_{\circ}^{\bot}}\colon\mathcal{C}_{\circ}^{\bot}\to\mathcal{C}_{\circ} is zero.

  2. (2)

    The left adjoints of the inclusions 𝒞,𝒞𝒞\mathcal{C}_{\circ},\mathcal{C}_{\circ}^{\perp}\subseteq\mathcal{C} induce an equivalence 𝒞𝒞×𝒞\mathcal{C}\overset{\sim}{\longrightarrow}\mathcal{C}_{\circ}\times\mathcal{C}_{\circ}^{\bot}.

  3. (3)

    The left and right adjoints LL and RR of 𝒞𝒞\mathcal{C}_{\circ}\hookrightarrow\mathcal{C} are isomorphic.

Proof.

We prove (1)\implies(2)\implies(3)\implies(1). The implication (1)\implies(2) follows from the identification of 𝒞\mathcal{C} with the \infty-category of sections of the cartesian fibration classified by L|𝒞L|_{\mathcal{C}_{\circ}^{\bot}}. For the implication (2)\implies(3), observe that the inclusion 𝒞𝒞×𝒞\mathcal{C}_{\circ}\hookrightarrow\mathcal{C}_{\circ}\times\mathcal{C}_{\circ}^{\bot} can be identified with the functor (Id,0)(\operatorname{Id},0), for which the projection 𝒞×𝒞𝒞\mathcal{C}_{\circ}\times\mathcal{C}_{\circ}^{\bot}\to\mathcal{C}_{\circ} is both a left and a right adjoint. Finally, for every X𝒞X\in\mathcal{C}_{\circ}^{\bot}, we have R(X)=0R(X)=0. Thus, assuming (3) we have RLR\simeq L and so L|𝒞=0L|_{\mathcal{C}_{\circ}^{\bot}}=0, which proves (1). ∎

Definition 4.1.5.

We say that a recollement 𝒞𝒞\mathcal{C}_{\circ}\hookrightarrow\mathcal{C} 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 𝒞Catst\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}}, we say that a descending chain of full subcategories,

𝒞(n)𝒞(2)𝒞(1)𝒞(0)𝒞\dots\subseteq\mathcal{C}_{(n)}\subseteq\dots\subseteq{\cal C}_{(2)}\subseteq{\cal C}_{(1)}\subseteq{\cal C}_{(0)}\subseteq{\cal C}

is a recollement chain, if each inclusion 𝒞(n)𝒞\mathcal{C}_{(n)}\subseteq\mathcal{C} exhibits 𝒞\mathcal{C} as a recollement of 𝒞(n)\mathcal{C}_{(n)} and 𝒞(n)\mathcal{C}_{(n)}^{\perp}. We also set

𝒞()n𝒞(n).\mathcal{C}_{(\infty)}\coloneqq\bigcap_{n\in\mathbb{N}}{\cal C}_{(n)}.

It turns out that under mild conditions, 𝒞()𝒞\mathcal{C}_{(\infty)}\subseteq\mathcal{C} is itself a recollement.

Lemma 4.1.7.

Let 𝒞Catst{\cal C}\in\mathrm{Cat}_{\mathrm{st}}, which admits sequential limits and colimits. For a recollement chain 𝒞(2)𝒞(1)𝒞(0)𝒞\dots\subseteq{\cal C}_{(2)}\subseteq{\cal C}_{(1)}\subseteq{\cal C}_{(0)}\subseteq{\cal C}, the inclusion 𝒞()𝒞\mathcal{C}_{(\infty)}\subseteq\mathcal{C} exhibits 𝒞\mathcal{C} as a recollement of 𝒞()\mathcal{C}_{(\infty)} and 𝒞()\mathcal{C}_{(\infty)}^{\perp}.

Proof.

It suffices to show that the inclusion 𝒞()𝒞\mathcal{C}_{(\infty)}\hookrightarrow\mathcal{C} 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 X𝒞X\in\mathcal{C} we need to construct an object LX𝒞()L_{\infty}X\in\mathcal{C}_{(\infty)} and a morphism X𝜂LXX\xrightarrow{\eta}L_{\infty}X, such that for all Y𝒞()Y\in\mathcal{C}_{(\infty)}, the map

Map(LX,Y)()ηMap(X,Y)\operatorname{Map}(L_{\infty}X,Y)\xrightarrow{(-)\circ\eta}\operatorname{Map}(X,Y)

is an isomorphism. Let Ln:𝒞𝒞(n)L_{n}\colon\mathcal{C}\to\mathcal{C}_{(n)} be the left adjoint of the inclusion 𝒞(n)𝒞\mathcal{C}_{(n)}\hookrightarrow\mathcal{C} and denote by IdηnLn\operatorname{Id}\xrightarrow{\eta_{n}}L_{n} the corresponding unit (i.e. localization) map, where we suppress the embedding functor 𝒞(n)𝒞\mathcal{C}_{(n)}\subseteq\mathcal{C}. Since the 𝒞(n)\mathcal{C}_{(n)}-s are nested, we have canonical isomorphisms LnLn1LnL_{n}L_{n-1}\overset{\sim}{\longrightarrow}L_{n}, and we abuse notation by denoting the composition Ln1ηnLnLn1LnL_{n-1}\xrightarrow{\eta_{n}}L_{n}L_{n-1}\overset{\sim}{\longrightarrow}L_{n} also by ηn\eta_{n}. We now define

LXlim(Xη1L1Xη2L2Xη3)L_{\infty}X\coloneqq\underrightarrow{\operatorname{lim}}\,(X\xrightarrow{\eta_{1}}L_{1}X\xrightarrow{\eta_{2}}L_{2}X\xrightarrow{\eta_{3}}\dots)

and take X𝜂LXX\xrightarrow{\eta}L_{\infty}X to be the cone map from the first object to the colimit in the diagram defining LXL_{\infty}X. For every Y𝒞()𝒞(n)Y\in\mathcal{C}_{(\infty)}\subseteq\mathcal{C}_{(n)}, the map

Map(LnX,Y)()ηnMap(X,Y)\operatorname{Map}(L_{n}X,Y)\xrightarrow{(-)\circ\eta_{n}}\operatorname{Map}(X,Y)

is an isomorphism for each nn\in\mathbb{N}. Thus, by taking the limit over nn, we get an isomorphism

limMap(LnX,Y)limMap(X,Y)Map(X,Y).\underleftarrow{\operatorname{lim}\,}_{\mathbb{N}}\operatorname{Map}(L_{n}X,Y)\overset{\sim}{\longrightarrow}\underleftarrow{\operatorname{lim}\,}_{\mathbb{N}}\operatorname{Map}(X,Y)\simeq\operatorname{Map}(X,Y).

Precomposing with the isomorphism

Map(LX,Y)=Map(limLnX,Y)limMap(LnX,Y),\operatorname{Map}(L_{\infty}X,Y)=\operatorname{Map}(\underrightarrow{\operatorname{lim}}\,_{\mathbb{N}}L_{n}X,Y)\overset{\sim}{\longrightarrow}\underleftarrow{\operatorname{lim}\,}_{\mathbb{N}}\operatorname{Map}(L_{n}X,Y),

we get an isomorphism

Map(LX,Y)Map(X,Y).\operatorname{Map}(L_{\infty}X,Y)\overset{\sim}{\longrightarrow}\operatorname{Map}(X,Y).

Unwinding the definitions, this isomorphism is given by precomposition with η\eta. ∎

To identify the right orthogonal complement of 𝒞()\mathcal{C}_{(\infty)} in 𝒞\mathcal{C}, we need the following general categorical fact:

Lemma 4.1.8.

Let F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D} be a functor in Catst\mathrm{Cat}_{\mathrm{st}}, and denote by ker(F)𝒞\ker(F)\subseteq\mathcal{C} the full subcategory spanned by the objects XX, for which F(X)=0F(X)=0. If FF admits a fully faithful right adjoint G:𝒟𝒞G\colon\mathcal{D}\hookrightarrow\mathcal{C}, then Im(G)=ker(F)\mathrm{Im}(G)=\ker(F)^{\perp}.

Proof.

In one direction, for XIm(G)X\in\mathrm{Im}(G), we have X=G(Y)X=G(Y) for some Y𝒟Y\in\mathcal{D}. Hence, for every Zker(F)Z\in\ker(F) we have

Map(Z,X)Map(Z,G(Y))Map(F(Z),Y)Map(0,Y)pt.\operatorname{Map}(Z,X)\simeq\operatorname{Map}(Z,G(Y))\simeq\operatorname{Map}(F(Z),Y)\simeq\operatorname{Map}(0,Y)\simeq\operatorname{pt}.

Thus, Im(G)ker(F)\mathrm{Im}(G)\subseteq\ker(F)^{\perp}. Conversely, let Id𝑢GF\operatorname{Id}\xrightarrow{u}GF and FG𝑐IdFG\xrightarrow{c}\operatorname{Id} be the unit and counit of the adjunction respectively. Since GG is fully faithful, cc is an isomorphism. By the zig-zag identities and 2-out-of-3, the map F(u)F(u) is also an isomorphism. Now, for every X𝒞X\in\mathcal{C} consider the fiber sequence

X0X𝑢GF(X).X_{0}\to X\xrightarrow{u}GF(X).

On the one hand, since F(u)F(u) is an isomorphism, F(X0)=0F(X_{0})=0 and hence X0ker(F)X_{0}\in\ker(F). On the other hand, if Xker(F)X\in\ker(F)^{\perp}, then since GF(X)Im(G)ker(F)GF(X)\in\mathrm{Im}(G)\subseteq\ker(F)^{\perp}, we also have X0ker(F)X_{0}\in\ker(F)^{\perp} and thus X0=0X_{0}=0. This implies that uu is an isomorphism and so XGF(X)Im(G)X\simeq GF(X)\in\mathrm{Im}(G). ∎

Given a recollement chain as in 4.1.6, for each nn\in\mathbb{N}, we have a fully faithful embedding 𝒞(n)𝒞\mathcal{C}_{(n)}^{\perp}\hookrightarrow\mathcal{C} with left adjoint Pn:𝒞𝒞(n)P_{n}\colon\mathcal{C}\to\mathcal{C}_{(n)}^{\perp}. We abuse notation by suppressing the inclusion 𝒞n𝒞\mathcal{C}_{n}^{\perp}\subseteq\mathcal{C} and the canonical isomorphisms PnPn+1PnP_{n}P_{n+1}\simeq P_{n}. We thus obtain a tower

Pn+1𝒞(n)PnP2𝒞(2)P1𝒞(1)P0𝒞(0)\dots\xrightarrow{P_{n+1}}\mathcal{C}_{(n)}^{\perp}\xrightarrow{P_{n}}\dots\xrightarrow{P_{2}}{\cal C}_{(2)}^{\perp}\xrightarrow{P_{1}}{\cal C}_{(1)}^{\perp}\xrightarrow{P_{0}}{\cal C}_{(0)}^{\perp}

of \infty-categories under 𝒞\mathcal{C}, which induces a functor

P:𝒞lim𝒞(n).P_{\infty}\colon\mathcal{C}\to\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\perp}.
Proposition 4.1.9.

Let 𝒞Catst{\cal C}\in\mathrm{Cat}_{\mathrm{st}} which admits sequential limits and colimits. Given a recollement chain 𝒞(2)𝒞(1)𝒞(0)𝒞\dots\subseteq{\cal C}_{(2)}\subseteq{\cal C}_{(1)}\subseteq{\cal C}_{(0)}\subseteq{\cal C}, the functor P:𝒞lim𝒞(n){\displaystyle P_{\infty}\colon{\cal C}\to\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot}} admits a fully faithful right adjoint, whose essential image is 𝒞()\mathcal{C}_{(\infty)}^{\perp}. Thus, 𝒞{\cal C} is a recollement of 𝒞(){\displaystyle\mathcal{C}_{(\infty)}} and lim𝒞(n)\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot}.

Proof.

By 4.1.7, the inclusion 𝒞()𝒞\mathcal{C}_{(\infty)}\hookrightarrow\mathcal{C} exhibits 𝒞\mathcal{C} as a recollement of 𝒞()\mathcal{C}_{(\infty)} and its right orthogonal complement and hence it suffices to identify 𝒞()\mathcal{C}_{(\infty)}^{\perp}. The objects of 𝒞()\mathcal{C}_{(\infty)} are precisely the X𝒞X\in\mathcal{C} for which P(X)=0P_{\infty}(X)=0. Thus, by 4.1.8, it suffices to show that PP_{\infty} admits a fully faithful right adjoint. Since PnP_{n} is a left adjoint for all nn, by [HY17, Theorem B], the functor PP_{\infty} is a left adjoint and we denote its right adjoint by G:lim𝒞(n)𝒞G_{\infty}\colon\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot}\to\mathcal{C}. We show that GG_{\infty} is fully faithful using the explicit description of the adjunction PGP_{\infty}\dashv G_{\infty} given in [HY17]. An object of lim𝒞(n)\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot} consists of a sequence of objects Xn𝒞(n)X_{n}\in\mathcal{C}_{(n)}^{\perp} together with structure isomorphisms PnXn+1XnP_{n}X_{n+1}\overset{\sim}{\longrightarrow}X_{n}. We shall write {Xn}lim𝒞(n)\{X_{n}\}\in\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot} suppressing the structure isomorphisms. Composing the structure isomorphisms of {Xn}\{X_{n}\} with the corresponding unit (i.e. localization) maps IdunPn\operatorname{Id}\xrightarrow{u_{n}}P_{n}, we get maps as follows:

fn:Xn+1unPnXn+1Xn.f_{n}\colon X_{n+1}\xrightarrow{u_{n}}P_{n}X_{n+1}\overset{\sim}{\longrightarrow}X_{n}.

By [HY17, Theorem B], the functor GG_{\infty} can be described explicitly on objects by the following formula:

G({Xn})lim(fnXnfn1f2X2f1X1f0X0).G_{\infty}(\{X_{n}\})\simeq\underleftarrow{\operatorname{lim}\,}(\dots\xrightarrow{f_{n}}X_{n}\xrightarrow{f_{n-1}}\dots\xrightarrow{f_{2}}X_{2}\xrightarrow{f_{1}}X_{1}\xrightarrow{f_{0}}X_{0}).

To prove that GG_{\infty} is fully faithful, it suffices to show that the counit PG𝑐IdP_{\infty}G_{\infty}\xrightarrow{c}\operatorname{Id} is an isomorphism. Since the collection of projection functors πk:lim𝒞(n)𝒞(k)\pi_{k}\colon\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot}\to\mathcal{C}_{(k)}^{\bot} for all kk\in\mathbb{N} is jointly conservative, it suffices to show that

Pk(limXn)πk(c)XkP_{k}(\underleftarrow{\operatorname{lim}\,}X_{n})\xrightarrow{\pi_{k}(c)}X_{k}

is an isomorphism for all kk\in\mathbb{N} and {Xn}lim𝒞(n)\{X_{n}\}\in\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot}. By [HY17, Theorem 5.5], we can describe πk(c)\pi_{k}(c) as the composition

Pk(limXn)Pk(Xk)Xk,P_{k}(\underleftarrow{\operatorname{lim}\,}X_{n})\to P_{k}(X_{k})\overset{\sim}{\longrightarrow}X_{k},

where the first map is induced by the canonical projection limXnXk\underleftarrow{\operatorname{lim}\,}X_{n}\to X_{k}. By cofinality, we can assume that the limit is taken over nkn\geq k. By definition, for each nn\in\mathbb{N}, the fiber of Xn+1fnXnX_{n+1}\xrightarrow{f_{n}}X_{n} lies in 𝒞(n)𝒞(k)\mathcal{C}_{(n)}\subseteq\mathcal{C}_{(k)}. Since 𝒞(k)\mathcal{C}_{(k)} is closed under sequential limits, it follows that the fiber of limXnXk\underleftarrow{\operatorname{lim}\,}X_{n}\to X_{k} lies in 𝒞(k)\mathcal{C}_{(k)} and hence becomes an isomorphism after applying PkP_{k}. This concludes the proof that GG_{\infty} is fully faithful and hence the proof of the claim. ∎

Corollary 4.1.10.

Let 𝒞Catst{\cal C}\in\mathrm{Cat}_{\mathrm{st}} which admits sequential limits and colimits with a recollement chain 𝒞(2)𝒞(1)𝒞(0)𝒞\dots\subseteq{\cal C}_{(2)}\subseteq{\cal C}_{(1)}\subseteq{\cal C}_{(0)}\subseteq{\cal C}. If 𝒞()=0\mathcal{C}_{(\infty)}=0 then 𝒞lim𝒞(n){\cal C}\simeq\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot}.

Proof.

By 4.1.9, 𝒞{\cal C} is a recollement of 𝒞()=0\mathcal{C}_{(\infty)}=0 and of lim𝒞(n)\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot}, so that 𝒞lim𝒞(n){\cal C}\simeq\underleftarrow{\operatorname{lim}\,}{\cal C}_{(n)}^{\bot}. ∎

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 \infty-category 𝒞\mathcal{C} and Id𝒞𝛼Id𝒞\operatorname{Id}_{\mathcal{C}}\xrightarrow{\alpha}\operatorname{Id}_{\mathcal{C}}, we have the full subcategories 𝒞[α1]\mathcal{C}[\alpha^{-1}] and 𝒞^α=𝒞[α1]\widehat{\mathcal{C}}_{\alpha}=\mathcal{C}[\alpha^{-1}]^{\perp} of 𝒞\mathcal{C} (3.1.5). Assuming 𝒞\mathcal{C} admits sequential limits and colimits, the inclusion 𝒞[α1]𝒞\mathcal{C}[\alpha^{-1}]\hookrightarrow\mathcal{C} admits both a left adjoint LL and a right adjoint RR, given respectively by “inverting α\alpha” on XX

LX=lim(X𝛼X𝛼X𝛼)LX=\underrightarrow{\operatorname{lim}}\,(X\xrightarrow{\alpha}X\xrightarrow{\alpha}X\xrightarrow{\alpha}\dots)

and by taking the “α\alpha-divisible part” of XX

RX=lim(𝛼X𝛼X𝛼X).RX=\underleftarrow{\operatorname{lim}\,}(\dots\xrightarrow{\alpha}X\xrightarrow{\alpha}X\xrightarrow{\alpha}X).
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 𝒞\mathcal{C} 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 Y𝒞Y\in\mathcal{C} is α\alpha-complete if and only if RY=0RY=0 if and only if YlimY/αrY\simeq\underleftarrow{\operatorname{lim}\,}Y/\alpha^{r}. In fact, the α\alpha-completion functor

YY^αlimY/αr,Y\mapsto\widehat{Y}_{\alpha}\coloneqq\underleftarrow{\operatorname{lim}\,}Y/\alpha^{r},

is the left adjoint to the inclusion 𝒞^α𝒞\widehat{\mathcal{C}}_{\alpha}\hookrightarrow\mathcal{C}.

Proposition 4.1.12.

Let 𝒞Catst\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}} which admits sequential limits and colimit and let Id𝒞𝛼Id𝒞\operatorname{Id}_{\mathcal{C}}\xrightarrow{\alpha}\operatorname{Id}_{\mathcal{C}}. Then 𝒞\mathcal{C} is a recollement of 𝒞[α1]\mathcal{C}[\alpha^{-1}] and 𝒞^α\widehat{\mathcal{C}}_{\alpha}.

Proof.

It follows from the discussion above that 𝒞[α1]𝒞\mathcal{C}[\alpha^{-1}]\hookrightarrow\mathcal{C} admits both adjoints. ∎

Our next goal is to give a characterization of when the said recollement is split in terms of the natural endomorphism α\alpha.

Definition 4.1.13.

We say that a natural endomorphism β:Id𝒞Id𝒞\beta\colon\operatorname{Id}_{\mathcal{C}}\to\operatorname{Id}_{\mathcal{C}} is a semi-inverse of α\alpha, if for every α\alpha-divisible XX, the map βX\beta_{X} is an inverse of αX\alpha_{X}.

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 𝒞Catst\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}} which admits sequential limits and colimits. For every α,β:Id𝒞Id𝒞\alpha,\beta\colon\operatorname{Id}_{\mathcal{C}}\to\operatorname{Id}_{\mathcal{C}}, if an object Y𝒞Y\in\mathcal{C} is α\alpha-complete, then it is (1αβ)(1-\alpha\beta)-divisible. If β\beta is a semi-inverse of α\alpha, then the converse holds as well.

Proof.

Note that all natural endomorphisms of Id𝒞\operatorname{Id}_{\mathcal{C}} commute by the interchange law, so in particular αβ=βα\alpha\beta=\beta\alpha. For an α\alpha-complete object Y𝒞Y\in\mathcal{C} we have Y=limY/αrY=\underleftarrow{\operatorname{lim}\,}Y/\alpha^{r}. For every rr\in\mathbb{N}, the map α2r\alpha^{2r} is zero on Y/αrY/\alpha^{r} and hence 1αβ1-\alpha\beta is invertible on Y/αrY/\alpha^{r}. By passing to the limit, 1αβ1-\alpha\beta is invertible on YY. Conversely, assume that (1αβ)(1-\alpha\beta) acts invertibly on YY. If β\beta is a semi-inverse of α\alpha, then for every α\alpha-divisible XX, the map (1αβ)(1-\alpha\beta) acts as zero on XX. Thus the pointed space Map(X,Y)\operatorname{Map}(X,Y) must be contractible as (1αβ)(1-\alpha\beta) acts both invertibly and as zero on it. This implies that YY is α\alpha-complete. ∎

The above lemma leads us to the following characterization of split recollement:

Proposition 4.1.15.

Let 𝒞\mathcal{C} be a stable \infty-category which admits sequential limits and colimits. The recollement associated with a natural endomorphism Id𝒞𝛼Id𝒞\operatorname{Id}_{\mathcal{C}}\xrightarrow{\alpha}\operatorname{Id}_{\mathcal{C}} is split if and only if α\alpha admits a semi-inverse. In which case,

𝒞𝒞[α1]×𝒞^α.\mathcal{C}\simeq\mathcal{C}[\alpha^{-1}]\times\widehat{\mathcal{C}}_{\alpha}.
Proof.

Let β\beta be a semi-inverse of α\alpha. To show that LY=0LY=0 for all Y𝒞^αY\in\widehat{\mathcal{C}}_{\alpha} it suffices to show that Map(Y,X)\operatorname{Map}(Y,X) is contractible for all X𝒞[α1]X\in\mathcal{C}[\alpha^{-1}]. By definition, 1αβ1-\alpha\beta is zero on XX, so it suffices to observe that 1αβ1-\alpha\beta is invertible on YY by 4.1.14. Conversely, if 𝒞𝒞[α1]×𝒞^α\mathcal{C}\simeq\mathcal{C}[\alpha^{-1}]\times\widehat{\mathcal{C}}_{\alpha}, we have for every X𝒞X\in\mathcal{C} a natural decomposition XX[α1]X^αX\simeq X[\alpha^{-1}]\oplus\widehat{X}_{\alpha} with X[α1]𝒞[α1]X[\alpha^{-1}]\in\mathcal{C}[\alpha^{-1}] and X^α𝒞^α\widehat{X}_{\alpha}\in\widehat{\mathcal{C}}_{\alpha}. In this case, the map β=(α|X[α1])10X^α\beta=(\alpha|_{X[\alpha^{-1}]})^{-1}\oplus 0_{\widehat{X}_{\alpha}} is a semi-inverse of α\alpha. ∎

Remark 4.1.16.

In 4.1.3, the recollement SpSp(p)\operatorname{Sp}_{\mathbb{Q}}\subseteq\operatorname{Sp}_{(p)} corresponds to the endomorphism IdSp(p)𝑝IdSp(p)\operatorname{Id}_{\operatorname{Sp}_{(p)}}\xrightarrow{p}\operatorname{Id}_{\operatorname{Sp}_{(p)}}. However, not every recollement arises in such a way. For example, for n1n\geq 1 the recollement LnSpSp(p)L_{n}\operatorname{Sp}\subseteq\operatorname{Sp}_{(p)} is not induced by any endomorphism α\alpha of the identity functor. We do note however, that every split recollement 𝒞𝒞\mathcal{C}_{\circ}\subseteq\mathcal{C} must arise from an endomorphism α\alpha of Id𝒞\operatorname{Id}_{\mathcal{C}}, because we can take α\alpha to be the idempotent ε:Id𝒞Id𝒞\varepsilon\colon\operatorname{Id}_{\mathcal{C}}\to\operatorname{Id}_{\mathcal{C}} projecting onto 𝒞\mathcal{C}_{\circ}. In this case, ε\varepsilon itself is a semi-inverse of ε\varepsilon.

4.2 Height Decomposition

Let 𝒞\mathcal{C} be now a stable mm-semiadditive \infty-category. We shall use the general machinery of (split) recollement to show that 𝒞\mathcal{C} splits into a product of \infty-categories according to height. By definition, an object X𝒞X\in\mathcal{C} is of height n\leq n if it is p(n)p_{(n)}-divisible. Similarly, XX is of height >n>n if it is p(n)p_{(n)}-complete, which by 3.1.9, is if and only if it is complete with respect to all of p=p(0),p(1),,p(n)p=p_{(0)},p_{(1)},\dots,p_{(n)}. Accordingly,

Proposition 4.2.1.

Let 𝒞Catst

-
m
\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}m}
and let 0nm0\leq n\leq m. If 𝒞\mathcal{C} admits sequential limits and colimits, then 𝒞\mathcal{C} is a recollement of 𝒞n\mathcal{C}_{\leq n} and 𝒞>n\mathcal{C}_{>n}.

Proof.

The full subcategory 𝒞n𝒞\mathcal{C}_{\leq n}\subseteq\mathcal{C} consists of the p(n)p_{(n)}-divisible objects and 𝒞>n𝒞\mathcal{C}_{>n}\subseteq\mathcal{C} is the full subcategory of p(n)p_{(n)}-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 𝒞Catst

-
m
\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}m}
and assume it admits sequential limits and colimits. For all n=0,,m1n=0,\dots,m-1, the map p(n+1)p_{(n+1)} is a semi-inverse of p(n)p_{(n)}. In particular, for X𝒞X\in\mathcal{C}, we have ht(X)>n\mathrm{ht}(X)>n if and only if XX is (1p(n)p(n+1))(1-p_{(n)}p_{(n+1)})-divisible.

Proof.

If

p(n)=|BnCp|=|ΩBn+1Cp|p_{(n)}=|B^{n}C_{p}|=|\Omega B^{n+1}C_{p}|

is invertible on X𝒞X\in\mathcal{C}, then by 2.4.7, p(n+1)=|Bn+1Cp|p_{(n+1)}=|B^{n+1}C_{p}| is the inverse of p(n)p_{(n)} on XX. Thus, p(n+1)p_{(n+1)} is a semi-inverse of p(n)p_{(n)}. Therefore, the claim follows from 4.1.14. ∎

Using 4.2.2 we can improve on 3.1.13 in the stable case as follows:

Corollary 4.2.3.

Let F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D} be a functor in Cat

-
m
\mathrm{Cat}^{\scalebox{0.6}{$\oplus$}\text{-}m}
and assume 𝒞\mathcal{C} and 𝒟\mathcal{D} are stable. For every X𝒞X\in\mathcal{C} and 0nm0\leq n\leq m, if XX is of height n\leq n or >n1>n-1, then so is F(X)F(X). The converse holds if FF is conservative.

Proof.

The facts about height n\leq n follow from 3.1.13. The facts about height >n1>n-1 follow similarly using 4.2.2. Namely, that XX is of height >n1>n-1, if and only if (1p(n1)p(n))(1-p_{(n-1)}p_{(n)}) acts invertibly on it. ∎

Remark 4.2.4.

4.2.3 does not cover the case ht(X)>m\mathrm{ht}(X)>m. This indeed can not be guaranteed even in the stable case as witnessed by 3.1.14.

Similarly, we can improve on 3.1.16 as follows:

Corollary 4.2.5.

Let 𝒞Catst\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}} be pp-typically mm-semiadditively monoidal \infty-category. For every 0nm0\leq n\leq m, we have Ht(𝒞)n\mathrm{Ht}(\mathcal{C})\leq n or Ht(𝒞)>n1\mathrm{Ht}(\mathcal{C})>n-1 if and only if ht𝒞(𝟙)n\mathrm{ht}_{\mathcal{C}}(\mathds{1})\leq n or ht𝒞(𝟙)>n1\mathrm{ht}_{\mathcal{C}}(\mathds{1})>n-1 respectively.

Proof.

Given X𝒞X\in\mathcal{C}, the functor X():𝒞𝒞X\otimes(-)\colon\mathcal{C}\to\mathcal{C} is pp-typically mm-semiadditive. Thus, the claim follows from 4.2.3. ∎

Remark 4.2.6.

Again, in 4.2.5 it is not true that if ht(𝟙)>m\mathrm{ht}(\mathds{1})>m then Ht(𝒞)>m\mathrm{Ht}(\mathcal{C})>m. For example, Mod𝕊^p(Sp)\operatorname{Mod}_{\widehat{\mathbb{S}}_{p}}(\operatorname{Sp}) is a presentably symmetric monoidal 0-semiadditive \infty-category, whose unit 𝕊^p\widehat{\mathbb{S}}_{p} has height >0>0, although the \infty-category itself does not.

The following is our main structure theorem for stable higher semiadditive \infty-categories:

Theorem 4.2.7 (Height Decomposition).

Let 𝒞Catst

-
m
\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}m}
for some 0m0\leq m\leq\infty.

  1. (1)

    For m<m<\infty, and 𝒞\mathcal{C} which is idempotent complete, the inclusions 𝒞0,,𝒞m1,𝒞>m1𝒞\mathcal{C}_{0},\dots,\mathcal{C}_{m-1},\mathcal{C}_{>m-1}\subseteq\mathcal{C} determine an equivalence of \infty-categories

    𝒞𝒞0××𝒞m1×𝒞>m1.\mathcal{C}\simeq\mathcal{C}_{0}\times\dots\times\mathcal{C}_{m-1}\times\mathcal{C}_{>m-1}.

    Moreover, 𝒞0\mathcal{C}_{0} is pp-typically \infty-semiadditive and 𝒞1,,𝒞m1\mathcal{C}_{1},\dots,\mathcal{C}_{m-1} are \infty-semiadditive.

  2. (2)

    For m=m=\infty, and 𝒞\mathcal{C} which admits sequential limits and colimits, 𝒞\mathcal{C} is a recollement of 𝒞\mathcal{C}_{\infty} and n𝒞n\prod_{n\in\mathbb{N}}\mathcal{C}_{n}. In particular, if 𝒞=0\mathcal{C}_{\infty}=0, then 𝒞n𝒞n.\mathcal{C}\simeq\prod_{n\in\mathbb{N}}\mathcal{C}_{n}.

If in addition 𝒞\mathcal{C} is mm-semiadditively 𝒪\mathcal{O}-monoidal for some \infty-operad 𝒪,\mathcal{O}, then 𝒞>m1\mathcal{C}_{>m-1} and 𝒞n\mathcal{C}_{n} for all n=0,,m1n=0,\dots,m-1, are compatible with the 𝒪\mathcal{O}-monoidal structure and the equivalences in both (1) and (2) promote naturally to an equivalence of 𝒪\mathcal{O}-monoidal \infty-categories.

Proof.

(1) We first prove the claim under the additional assumption that 𝒞\mathcal{C} admits all sequential limits and colimits (and hence in particular idempotent complete). First, by 4.2.2, the map p(m)p_{(m)} is a semi-inverse of p(m1)p_{(m-1)}. Hence, by 4.1.15, we obtain a direct product decomposition 𝒞𝒞m1×𝒞>m1\mathcal{C}\simeq\mathcal{C}_{\leq m-1}\times\mathcal{C}_{>m-1}. The category 𝒞m1\mathcal{C}_{\leq m-1} is itself mm-semiadditive (3.1.12) and admits all sequential limits and colimits. Thus, we can continue decomposing 𝒞m1\mathcal{C}_{\leq m-1} inductively and get 𝒞n𝒞n1×𝒞n\mathcal{C}_{\leq n}\simeq\mathcal{C}_{\leq n-1}\times\mathcal{C}_{n} for all n=0,,m1n=0,\dots,m-1. Finally, by 3.2.2, each 𝒞n\mathcal{C}_{\leq n} is in fact pp-typically \infty-semiadditive. By 3.1.12, 𝒞n\mathcal{C}_{n} is also pp-typically \infty-semiadditive and since it is also pp-complete for all n1n\geq 1, it is in particular pp-local, and hence \infty-semiadditive by 3.2.6.

For a general 𝒞\mathcal{C} 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 mm-semiadditive \infty-category 𝒞^\widehat{\mathcal{C}} and an mm-semiadditive fully faithful embedding 𝒞𝒞^\mathcal{C}\hookrightarrow\widehat{\mathcal{C}}. By 4.2.3, for each n=0,,m1n=0,\dots,m-1, we have fully faithful embedding 𝒞n𝒞^n\mathcal{C}_{n}\hookrightarrow\widehat{\mathcal{C}}_{n} and 𝒞>m1𝒞^>m1\mathcal{C}_{>m-1}\hookrightarrow\widehat{\mathcal{C}}_{>m-1}, hence also

𝒞0××𝒞m1×𝒞>m1𝒞^0××𝒞^m1×𝒞^>m1𝒞^.\mathcal{C}_{0}\times\dots\times\mathcal{C}_{m-1}\times\mathcal{C}_{>m-1}\hookrightarrow\widehat{\mathcal{C}}_{0}\times\dots\times\widehat{\mathcal{C}}_{m-1}\times\widehat{\mathcal{C}}_{>m-1}\simeq\widehat{\mathcal{C}}.

By the left cancellation property of fully faithful embeddings, we get a fully faithful embedding

𝒞0××𝒞m1×𝒞>m1𝒞.\mathcal{C}_{0}\times\dots\times\mathcal{C}_{m-1}\times\mathcal{C}_{>m-1}\hookrightarrow\mathcal{C}.

For each object X𝒞X\in\mathcal{C}, the height n=0,,m1n=0,\dots,m-1 and >m1>m-1 components of XX in 𝒞^\widehat{\mathcal{C}}, are retracts of XX in 𝒞^\widehat{\mathcal{C}}. Thus, if 𝒞\mathcal{C} is idempotent complete, these components belong to 𝒞\mathcal{C}. It follows that the above fully faithful embedding is also essentially surjective.

(2) For every n<n<\infty, we have by (1), that 𝒞𝒞n×𝒞>n{\cal C}\simeq{\cal C}_{\leq n}\times{\cal C}_{>n}. Hence, we can switch the roles of 𝒞n\mathcal{C}_{\leq n} and 𝒞>n\mathcal{C}_{>n}, and consider the embedding 𝒞>n𝒞\mathcal{C}_{>n}\subseteq\mathcal{C} as exhibiting 𝒞\mathcal{C} as a recollement of 𝒞>n\mathcal{C}_{>n} and (𝒞>n)=𝒞n(\mathcal{C}_{>n})^{\bot}=\mathcal{C}_{\leq n}. We thus obtain a recollement chain

𝒞>2𝒞>1𝒞>0𝒞.\dots\subseteq\mathcal{C}_{>2}\subseteq{\cal C}_{>1}\subseteq{\cal C}_{>0}\subseteq{\cal C}.

By definition, 𝒞=n𝒞>n\mathcal{C}_{\infty}=\bigcap\limits_{n\in\mathbb{N}}\mathcal{C}_{>n}, and so by 4.1.9, 𝒞\mathcal{C} is a recollement of 𝒞\mathcal{C}_{\infty} and

lim𝕟(𝒞n)limn(0kn𝒞k)n𝒞n.\underleftarrow{\operatorname{lim}\,}_{\mathbb{n\in N}}(\mathcal{C}_{\leq n})\simeq\underleftarrow{\operatorname{lim}\,}_{n\in\mathbb{N}}\left(\prod_{0\leq k\leq n}\mathcal{C}_{k}\right)\simeq\prod_{n\in\mathbb{N}}\mathcal{C}_{n}.

Finally, assume that 𝒞\mathcal{C} is mm-semiadditively 𝒪\mathcal{O}-monoidal. The full subcategories 𝒞n\mathcal{C}_{\leq n} and 𝒞>n\mathcal{C}_{>n} for all n=0,,m1n=0,\dots,m-1 consist of objects which are p(n)p_{(n)}-divisible and (1p(n)p(n+1))(1-p_{(n)}p_{(n+1)})-divisible respectively. It follows that 𝒞n\mathcal{C}_{\leq n} and 𝒞>n\mathcal{C}_{>n} are compatible with the 𝒪\mathcal{O}-monoidal structure and hence so is 𝒞n\mathcal{C}_{n} for all n=0,,m1n=0,\dots,m-1. Thus, by [Lura, Proposition 2.2.1.9], the 𝒞n\mathcal{C}_{n}-s and 𝒞>m1\mathcal{C}_{>m-1} inherit an 𝒪\mathcal{O}-monoidal structure such that the projections 𝒞𝒞n\mathcal{C}\to\mathcal{C}_{n} and 𝒞𝒞>m1\mathcal{C}\to\mathcal{C}_{>m-1} are 𝒪\mathcal{O}-monoidal. ∎

We conclude with some remarks regarding the sharpness of 4.2.7. In the case m<m<\infty, the fact that 𝒞\mathcal{C} is mm-semiadditive also implies that 𝒞\mathcal{C} is a recollement of 𝒞m\mathcal{C}_{\leq m} and 𝒞>m\mathcal{C}_{>m}, but there is no guaranty that the “gluing data” is trivial. That is, that 𝒞\mathcal{C} decomposes as a direct product of 𝒞m\mathcal{C}_{\leq m} and 𝒞>m\mathcal{C}_{>m}. Indeed, consider the 0-semiadditive \infty-category 𝒞=Sp(p).\mathcal{C}=\operatorname{Sp}_{(p)}. The case m=0m=0 corresponds to the recollement SpSp(p)\operatorname{Sp}_{\mathbb{Q}}\subseteq\operatorname{Sp}_{(p)} of 4.1.3. In this case 𝒞>0=Sp^p\mathcal{C}_{>0}=\widehat{\operatorname{Sp}}_{p} and the gluing data is not trivial, as the rationalization of the pp-completion does not vanish in general. Having said that, for m1m\geq 1 we do not know whether there even exists a stable pp-local presentable \infty-category that is mm-semiadditive, but not (m+1)(m+1)-semiadditive [CSY18, Conjecture 1.1.5].

In the case m=m=\infty, we do not know whether there exists a stable \infty-semiadditive \infty-category 𝒞\mathcal{C} for which 𝒞0\mathcal{C}_{\infty}\neq 0. We hence propose the following:

Conjecture 4.2.8 (Height Finiteness).

For every 𝒞Prst

-
\mathcal{C}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
, the full subcategory

𝒞n0𝒞>n𝒞,\mathcal{C}_{\infty}\coloneqq\bigcap_{n\geq 0}\mathcal{C}_{>n}\subseteq\mathcal{C},

of objects of height \infty, is trivial.

4.3 Semisimplicity

Classical representation theory tells us that in characteristic 0, representations of a finite group GG are semisimple. The \infty-category of 𝒞\mathcal{C}-representations of GG in any \infty-category 𝒞\mathcal{C} is equivalent to the \infty-category 𝒞BG\mathcal{C}^{BG} of 𝒞\mathcal{C}-valued local systems on BGBG. From the point of view of higher semiadditivity, characteristic 0 corresponds to semiadditive height 0, and so it is natural to consider the analogous situation for higher heights. We shall show that given a stable \infty-semiadditive \infty-category 𝒞\mathcal{C} of height nn, certain analogous semisimplicity phenomena hold for 𝒞A\mathcal{C}^{A} for every π\pi-finite nn-connected space AA (e.g. A=Bn+1GA=B^{n+1}G). For the case 𝒞=ModK(n)(Sp)\mathcal{C}=\operatorname{Mod}_{K(n)}(\operatorname{Sp}) 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 𝒞Catst\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}} and let A𝑞BA\xrightarrow{q}B be a 𝒞\mathcal{C}-acyclic and weakly 𝒞\mathcal{C}-ambidextrous map. The functor q:𝒞B𝒞Aq^{*}\colon\mathcal{C}^{B}\to\mathcal{C}^{A} is fully faithful and exhibits 𝒞A\mathcal{C}^{A} as a recollement of 𝒞B\mathcal{C}^{B} and (𝒞B)𝒞A(\mathcal{C}^{B})^{\bot}\subseteq\mathcal{C}^{A}. The recollement is split if and only if qq is 𝒞\mathcal{C}-ambidextrous, in which case there is a canonical equivalence:

𝒞A𝒞B×(𝒞B).\mathcal{C}^{A}\simeq\mathcal{C}^{B}\times(\mathcal{C}^{B})^{\bot}.
Proof.

By definition of acyclicity, qq^{*} is fully faithful. Hence, the recollement is split if and only the left and right adjoints q!q_{!} and qq_{*} respectively of qq^{*} are isomorphic (4.1.4). If qq is 𝒞\mathcal{C}-ambidextrous then q!q_{!} and qq_{*} are isomorphic and the recollement is split. Conversely, if we have q!qq_{!}\simeq q_{*}, then qq_{*} preserves all qq-colimits and hence qq is 𝒞\mathcal{C}-ambidextrous (2.1.3). ∎

As a special case we obtain:

Theorem 4.3.2.

Let 𝒞Catst

-
\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
be pp-local such that Ht(𝒞)=n\mathrm{Ht}(\mathcal{C})=n. For every map of spaces A𝑞BA\xrightarrow{q}B with nn-connected π\pi-finite fibers, we have a canonical equivalence

𝒞A𝒞B×(𝒞B).\mathcal{C}^{A}\simeq\mathcal{C}^{B}\times(\mathcal{C}^{B})^{\bot}.
Proof.

Since 𝒞\mathcal{C} is of height nn, the map qq is 𝒞\mathcal{C}-acyclic by 3.2.7(2) and 2.4.2. Hence, the claim follows from 4.3.1. ∎

In the case B=ptB=\operatorname{pt}, we can interpret 4.3.2 from the perspective of “higher representation theory” (see [Lurb]). For every space AA, we call an object X𝒞AX\in\mathcal{C}^{A} unipotent if it belongs to the full subcategory 𝒞unipA𝒞A\mathcal{C}_{\mathrm{unip}}^{A}\subseteq\mathcal{C}^{A} generated by colimits from the trivial representations (i.e. constant local systems). It can be shown that every object X𝒞AX\in\mathcal{C}^{A} fits into an essentially unique cofiber sequence

XunipXX0()X_{\mathrm{unip}}\to X\to X^{0}\qquad(*)

with Xunip𝒞unipAX_{\mathrm{unip}}\in\mathcal{C}_{\mathrm{unip}}^{A} and X0(𝒞unipA)X^{0}\in(\mathcal{C}_{\mathrm{unip}}^{A})^{\perp}. If AA is nn-connected, then 4.3.2 implies that XunipX_{\mathrm{unip}} is constant and ()(*) canonically splits.

Transfer idempotents

Given a stable \infty-category 𝒞\mathcal{C} and an equivalence 𝒞𝒞×𝒞′′\mathcal{C}\simeq\mathcal{C}^{\prime}\times\mathcal{C}^{\prime\prime}, every object X𝒞X\in\mathcal{C} has an essentially unique decomposition XXX′′X\simeq X^{\prime}\oplus X^{\prime\prime}, such that X𝒞X^{\prime}\in\mathcal{C}^{\prime} and X′′𝒞′′X^{\prime\prime}\in\mathcal{C}^{\prime\prime}. This allows us to define a natural endomorphism ε:Id𝒞Id𝒞\varepsilon\colon\operatorname{Id}_{\mathcal{C}}\to\operatorname{Id}_{\mathcal{C}} by the formula

XX′′IdX0X′′XX′′.X^{\prime}\oplus X^{\prime\prime}\xrightarrow{\operatorname{Id}_{X^{\prime}}\oplus 0_{X^{\prime\prime}}}X^{\prime}\oplus X^{\prime\prime}.

The natural endomorphism ε\varepsilon is idempotent and realizes internally to 𝒞\mathcal{C} the projection onto the essential image of 𝒞\mathcal{C}^{\prime} in 𝒞\mathcal{C}. Our next goal is to provide an explicit description of the idempotent ε\varepsilon for the split recollement q:𝒞B𝒞Aq^{*}\colon\mathcal{C}^{B}\hookrightarrow\mathcal{C}^{A} of 4.3.1. To help guide the intuition, we begin with a closely related elementary example:

Example 4.3.3.

Let 𝒞=Vec\mathcal{C}=\mathrm{Vec}_{\mathbb{Q}} and let GG be a finite group with a normal subgroup NGN\triangleleft G. The map q:BGB(G/N)q\colon BG\to B(G/N) induces a fully faithful embedding

q:VecB(G/N)VecBGq^{*}\colon\mathrm{Vec}_{\mathbb{Q}}^{B(G/N)}\to\mathrm{Vec}_{\mathbb{Q}}^{BG}

which is the “inflation” functor that takes a vector space VV with a G/NG/N-action to VV itself with the GG-action induced via GG/NG\to G/N. The essential image of qq^{*} consists of GG-representations on which the subgroup NGN\triangleleft G acts trivially. The adjoints q!q_{!} and qq_{*} can be identified with the vector spaces of NN-coinvairant and NN-invariants respectively, equipped with the residual G/NG/N-action. Thus, the full subcategory (VecB(G/N))VecBG(\mathrm{Vec}_{\mathbb{Q}}^{B(G/N)})^{\bot}\subseteq\mathrm{Vec}_{\mathbb{Q}}^{BG} is spanned by the GG-representations without non-trivial NN-fixed vectors and we have an equivalence

VecBGVecB(G/N)×(VecB(G/N)).\mathrm{Vec}_{\mathbb{Q}}^{BG}\simeq\mathrm{Vec}_{\mathbb{Q}}^{B(G/N)}\times(\mathrm{Vec}_{\mathbb{Q}}^{B(G/N)})^{\perp}.

This equivalence is realized explicitly as follows. Since VecBG\mathrm{Vec}_{\mathbb{Q}}^{BG} is semi-simple, we can split each GG-representation VV uniquely as VVNVN-freeV\simeq V^{N}\oplus V^{N\text{-}\mathrm{free}} with VNV^{N} the space of NN-fixed vectors and VN-free(VecB(G/N))V^{N\text{-}\mathrm{free}}\in(\mathrm{Vec}_{\mathbb{Q}}^{B(G/N)})^{\bot}. Consider the natural endomorphism:

α:Id()NNmq()NId,\alpha\colon\operatorname{Id}\to(-)_{N}\xrightarrow{\operatorname{Nm}_{q}}(-)^{N}\to\operatorname{Id},

and its normalization ε|N|1α\varepsilon\coloneqq|N|^{-1}\alpha. Unwinding the definitions, we get

ε(x)=1|N|gNgxV,\varepsilon(x)=\frac{1}{|N|}\sum_{g\in N}gx\quad\in V,

which is an explicit formula for the GG-equivariant projection ε:VV\varepsilon\colon V\to V onto the subspace VNVV^{N}\subseteq V.

In a similar fashion, we have:

Proposition 4.3.4.

Let 𝒞Catst\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}} and let A𝑞BA\xrightarrow{q}B be a 𝒞\mathcal{C}-acyclic and 𝒞\mathcal{C}-ambidextrous map. Consider

α:Id𝒞Au!qq!NmqqqcId𝒞A.\alpha\colon\operatorname{Id}_{\mathcal{C}^{A}}\xrightarrow{u_{!}}q^{*}q_{!}\xrightarrow{\operatorname{Nm}_{q}}q^{*}q_{*}\xrightarrow{c_{*}}\operatorname{Id}_{\mathcal{C}^{A}}.

The natural endomorphism εq(|q|)α\varepsilon\coloneqq q^{*}(|q|)\alpha of Id𝒞A\operatorname{Id}_{\mathcal{C}^{A}} is idempotent, and it realizes the projection onto the essential image of q:𝒞B𝒞Aq^{*}\colon\mathcal{C}^{B}\hookrightarrow\mathcal{C}^{A}.

Proof.

First, for every Y(𝒞B),Y\in(\mathcal{C}^{B})^{\bot}, we have qY=0q_{*}Y=0 and hence α|(𝒞B)=0\alpha|_{(\mathcal{C}^{B})^{\bot}}=0. Next, consider the commutative diagram (see 2.1.5):

q\textstyle{q^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u!\scriptstyle{u_{!}}qq!q\textstyle{q^{*}q_{!}q^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\wr}c!\scriptstyle{c_{!}}\scriptstyle{\sim}Nmq\scriptstyle{\operatorname{Nm}_{q}}qqq\textstyle{q^{*}q_{*}q^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c\scriptstyle{c_{*}}q\textstyle{q^{*}}q\textstyle{q^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sim}|q|1\scriptstyle{|q|^{-1}}q\textstyle{q^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u\scriptstyle{u_{*}}\scriptstyle{\wr}

The composition along the top row computes the restriction of α\alpha along qq^{*}, which is therefore invertible and coincides with q(|q|1)q^{*}(|q|^{-1}). It follows that q(|q|)αq^{*}(|q|)\alpha is the identity on the essential image of 𝒞B\mathcal{C}^{B} and is zero on (𝒞B)(\mathcal{C}^{B})^{\bot}. Thus, q(|q|)αq^{*}(|q|)\alpha equals the idempotent ε\varepsilon which projects onto the essential image of 𝒞A\mathcal{C}^{A}. ∎

Remark 4.3.5.

We note that if we furthermore assume that A𝑞BA\xrightarrow{q}B is principal and BB is connected, then by 2.3.3, we have

q|q|=qB|F|=A|F|.q^{*}|q|=q^{*}B^{*}|F|=A^{*}|F|.

Namely, q|q|q^{*}|q| is constant with value |F||F|. Moreover, by 2.4.8 and 2.4.2, the space ΩF\Omega F is 𝒞\mathcal{C}-amenable and thus we get |F|=|ΩF|1|F|=|\Omega F|^{-1}. To conclude, we can also write ε=|ΩF|1α\varepsilon=|\Omega F|^{-1}\alpha.

Our main motivation for 4.3.4, is to generalize 4.3.3 to higher heights:

Example 4.3.6.

Let 𝒞Catst

-
\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
be of height nn. For an abelian pp-group GG with a subgroup NGN\leq G, we consider the fiber sequence

Bn+1NBn+1G𝑞Bn+1(G/N).B^{n+1}N\to B^{n+1}G\xrightarrow{q}B^{n+1}(G/N).

The map qq satisfies the conditions of 4.3.2, and hence 𝒞Bn+1G\mathcal{C}^{B^{n+1}G} splits as a product of 𝒞Bn+1(G/N)\mathcal{C}^{B^{n+1}(G/N)} and its right orthogonal complement. Moreover, by 4.3.4 and 4.3.5, the idempotent ε\varepsilon is given by ε=|BnN|1α\varepsilon=|B^{n}N|^{-1}\alpha. The case n=0n=0 is a “derived version” of 4.3.3.

We note that 4.3.6 is essentially the only interesting case of 4.3.2, as in view of 3.2.3(3), for any nn-connected space AA, we have 𝒞A𝒞Bn+1πn+1A\mathcal{C}^{A}\simeq\mathcal{C}^{B^{n+1}\pi_{n+1}A}.

4.4 Chromatic Examples

The main source of stable higher semiadditive \infty-categories is chromatic homotopy theory. In this subsection, we shall address the semiadditive height of such \infty-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 F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D} in Alg(Prst)\operatorname{Alg}(\Pr_{\mathrm{st}}) nil-conservative, if for every ring RAlg(𝒞)R\in\operatorname{Alg}(\mathcal{C}), if F(R)=0F(R)=0 then R=0R=0111111This 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 F:𝒞𝒟F\colon{\cal C}\to{\cal D} be a map in Alg(Prst)\operatorname{Alg}(\Pr_{\mathrm{st}}). If 𝒞\mathcal{C} is mm-semiadditive, then so is 𝒟\mathcal{D}. Furthermore, for every 0nm0\leq n\leq m, if 𝒞{\cal C} is of height n\leq n or >n1>n-1 then so is 𝒟{\cal D} and the converse holds if FF is nil-conservative.

Proof.

Since 𝒞\mathcal{C} is mm-semiadditive and F:𝒞𝒟F\colon{\cal C}\to{\cal D} is a monoidal mm-semiadditive functor, by [CSY18, Corollary 3.3.2(2)], 𝒟\mathcal{D} is mm-semiadditive as well. By 4.2.5, the height of 𝒞\mathcal{C} (resp. 𝒟\mathcal{D}) is determined by the height of 𝟙𝒞\mathds{1}_{\mathcal{C}} (resp. 𝟙𝒟\mathds{1}_{\mathcal{D}}). Moreover, ht(𝟙)n\mathrm{ht}(\mathds{1})\leq n if and only if p(n)π0𝟙p_{(n)}\in\pi_{0}\mathds{1} is invertible and ht(𝟙)>n1\mathrm{ht}(\mathds{1})>n-1, if and only if (1p(n1)p(n))π0𝟙(1-p_{(n-1)}p_{(n)})\in\pi_{0}\mathds{1} is invertible. Thus, the claim follows from [CSY18, Corollary 4.4.5]. ∎

As a special case, we get:

Corollary 4.4.3.

Let 𝒞CAlg(Prst)\mathcal{C}\in\operatorname{CAlg}(\Pr_{\mathrm{st}}) be mm-semiadditive and let RCAlg(𝒞)R\in\operatorname{CAlg}(\mathcal{C}). For every integer 0nm0\leq n\leq m, if 𝒞{\cal C} is of height n\leq n or >n1>n-1 then so is ModR(𝒞)\operatorname{Mod}_{R}(\mathcal{C}). The converse holds if tensoring with RR is nil-conservative.

Proof.

The claim follows from 4.4.2 for the colimit preserving symmetric monoidal functor R():𝒞ModR(𝒞)R\otimes(-)\colon\mathcal{C}\to\operatorname{Mod}_{R}(\mathcal{C}). ∎

We now apply the above to higher semiadditive \infty-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 \infty-category Θn=ModEn(SpK(n))\Theta_{n}=\operatorname{Mod}_{E_{n}}(\operatorname{Sp}_{K(n)}) satisfies Ht(Θn)=n\mathrm{Ht}(\Theta_{n})=n.

Proof.

For n=0n=0, the claim is clear, so assume n1n\geq 1. This follows from the explicit formula p(k)=p(n1k)p_{(k)}=p^{\binom{n-1}{k}} given in 2.2.5. Indeed, for knk\geq n, we have p(k)=1p_{(k)}=1 and hence invertible. For 0k<n0\leq k<n, the element p(k)p_{(k)} is a non-zero power of pp, and hence every XΘnX\in\Theta_{n} is complete with respect to it as the \infty-category Θn\Theta_{n} is pp-complete. ∎

In [CSY18, Theorem C] we have shown that for a homotopy ring spectrum RR, the \infty-category SpR\operatorname{Sp}_{R} is 11-semiadditive if and only if it is \infty-semiadditive if and only if supp(R)={n}\mathrm{supp}(R)=\{n\} for some integer nn. We now show that this nn is in fact the semiadditive height of SpR\operatorname{Sp}_{R}.

Theorem 4.4.5.

Let RR be a homotopy ring spectrum121212In fact, it suffices to assume that RR is a weak ring in the sense of [CSY18, Definition 5.1.4].. If supp(R)={n}\mathrm{supp}(R)=\{n\} for some integer nn, then Ht(SpR)=n\mathrm{Ht}(\operatorname{Sp}_{R})=n. In particular, we have

Ht(SpK(n))=Ht(SpT(n))=n.\mathrm{Ht}(\operatorname{Sp}_{K(n)})=\mathrm{Ht}(\operatorname{Sp}_{T(n)})=n.
Proof.

We first consider the case R=K(n)R=K(n). This follows by applying 4.4.3 to the faithful commutative algebra EnCAlg(SpK(n))E_{n}\in\operatorname{CAlg}(\operatorname{Sp}_{K(n)}) and the fact that Ht(Θn)=n\mathrm{Ht}(\Theta_{n})=n (4.4.4). For a general homotopy ring RR with supp(R)={n}\mathrm{supp}(R)=\{n\} (such as R=T(n)R=T(n)), consider the functor LK(n):SpRSpK(n)L_{K(n)}\colon\operatorname{Sp}_{R}\to\operatorname{Sp}_{K(n)}. It is nil-conservative by [CSY18, Proposition 5.1.15]. Since SpK(n)\operatorname{Sp}_{K(n)} is of height nn, the claim follows from 4.4.2. ∎

5 Modes

In this section, we use the theory of idempotent algebras in Pr\Pr, 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 \infty-categories, as a means to encode properties, which induce “canonical structure”.

Definitions & characterizations

Following [Lura, Definition 4.8.2.1], given a symmetric monoidal \infty-category 𝒞{\cal C}, we say that a morphism 𝟙𝑢X\mathds{1}\xrightarrow{u}X in 𝒞\mathcal{C} exhibits XX as an idempotent object of 𝒞{\cal C}, if

XX𝟙1uXXX\simeq X\otimes\mathds{1}\xrightarrow{1\otimes u}X\otimes X

is an isomorphism. By [Lura, Proposition 4.8.2.9], an idempotent object 𝟙𝑢X\mathds{1}\xrightarrow{u}X admits a unique commutative algebra structure for which uu is the unit. Conversely, the unit 𝟙𝑢R\mathds{1}\xrightarrow{u}R of a commutative algebra RR, exhibits it as an idempotent object if and only if the multiplication map RR𝑚RR\otimes R\xrightarrow{m}R is an isomorphism. In this case we call RR an idempotent algebra. More precisely, the functor CAlg(𝒞)𝒞𝟙/\operatorname{CAlg}(\mathcal{C})\to\mathcal{C}_{\mathds{1}/} which forgets the algebra structure and remembers only the unit map, induces an equivalence of \infty-categories from the full subcategory of idempotent algebras CAlgidem(𝒞)CAlg(𝒞)\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C})\subseteq\operatorname{CAlg}(\mathcal{C}) to the full subcategory of idempotent objects [Lura, Proposition 4.8.2.9]. The fundamental feature of an idempotent algebra RR, is that the forgetful functor ModR(𝒞)𝒞\operatorname{Mod}_{R}(\mathcal{C})\to\mathcal{C} is fully faithful, and its essential image consists of those objects for which the map Y𝟙1uYRY\otimes\mathds{1}\xrightarrow{1\otimes u}Y\otimes R is an isomorphism [Lura, Proposition 4.8.2.10]. Thus, it is a property of an object in 𝒞\mathcal{C} to have the structure of an RR-module. We shall say that RR classifies the property of being an RR-module.

Example 5.1.1.

For 𝒞=Ab\mathcal{C}=\mathrm{Ab}, the idempotent algebras are classically known as solid rings [BK72, Definition 2.1]. These include for example \mathbb{Q} and 𝔽p\mathbb{F}_{p}. We note that for 𝒞=Sp\mathcal{C}=\operatorname{Sp}, the ring \mathbb{Q} is still idempotent, classifying the property of being rational, but 𝔽p\mathbb{F}_{p} is not idempotent. The idempotent rings in Sp\operatorname{Sp} correspond precisely to the smashing localizations.

Given an idempotent ring RCAlg(𝒞),R\in\operatorname{CAlg}(\mathcal{C}), the forgetful functor ModR(𝒞)𝒞\operatorname{Mod}_{R}(\mathcal{C})\to\mathcal{C} admits a left adjoint

R():𝒞ModR(𝒞).R\otimes(-)\colon\mathcal{C}\to\operatorname{Mod}_{R}(\mathcal{C}).

This is a localization functor, which can be thought of as forcing the property classified by RR in a universal way. In line with the standard terminology for localizations of spectra, we set:

Definition 5.1.2.

Let L:𝒞𝒟L\colon\mathcal{C}\to\mathcal{D} be a map in CAlg(Cat)\operatorname{CAlg}(\operatorname{Cat}_{\infty}). We say that LL is a smashing localization if there exists an idempotent algebra RR in 𝒞\mathcal{C} and an isomorphism ModR(𝒞)𝒟\operatorname{Mod}_{R}(\mathcal{C})\overset{\sim}{\longrightarrow}\mathcal{D} in CAlg(Cat)\operatorname{CAlg}(\operatorname{Cat}_{\infty}), such that LL is the composition

𝒞R()ModR(𝒞)𝒟.\mathcal{C}\xrightarrow{R\otimes(-)}\operatorname{Mod}_{R}(\mathcal{C})\overset{\sim}{\longrightarrow}\mathcal{D}.

We note that for a smashing localization L:𝒞𝒟L\colon\mathcal{C}\to\mathcal{D}, there is always a fully faithful right adjoint F:𝒟𝒞F\colon\mathcal{D}\to\mathcal{C}, which is lax symmetric monoidal [Lura, Corollary 7.3.2.7], and we can identify the idempotent algebra RR with FL(𝟙𝒞)FL(\mathds{1}_{\mathcal{C}}). To characterize smashing localizations, we first introduce some terminology. Let L:𝒞𝒟L\colon\mathcal{C}\to\mathcal{D} in CAlg(Cat)\operatorname{CAlg}(\operatorname{Cat}_{\infty}), which admits a (lax symmetric monoidal) right adjoint F:𝒟𝒞F\colon\mathcal{D}\to\mathcal{C}. For every X𝒞X\in\mathcal{C} and Y𝒟Y\in\mathcal{D}, we have a natural map

XF(Y)𝛼F(L(X)Y),X\otimes F(Y)\xrightarrow{\alpha}F(L(X)\otimes Y),

which is the mate of

L(XF(Y))L(X)LF(Y)L(X)cYL(X)Y,L(X\otimes F(Y))\simeq L(X)\otimes LF(Y)\xrightarrow{L(X)\otimes c_{Y}}L(X)\otimes Y,

where c:LFIdc\colon LF\to\operatorname{Id} is the counit of the adjunction. We say that the adjunction LFL\dashv F satisfies the projection formula if the map α\alpha is an isomorphism for all X,YX,Y. The map α\alpha is compatible with the unit of the adjunction Id𝑢FL\operatorname{Id}\xrightarrow{u}FL in the following sense:

Lemma 5.1.3.

For all X,Z𝒞X,Z\in\mathcal{C}, the following diagram is commutative

XZ\textstyle{X\otimes Z\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}XuZ\scriptstyle{X\otimes u_{Z}}uXZ\scriptstyle{u_{X\otimes Z}}XFL(Z)\textstyle{X\otimes FL(Z)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha\quad}F(L(X)L(Z))\textstyle{F(L(X)\otimes L(Z))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\quad\sim}FL(XZ)\textstyle{FL(X\otimes Z)}
Proof.

Passing to the mates under the adjunction LFL\dashv F 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 L:𝒞𝒟L\colon\mathcal{C}\to\mathcal{D} in CAlg(Cat)\operatorname{CAlg}(\operatorname{Cat}_{\infty}) is a smashing localization if and only if it admits a fully faithful right adjoint F:𝒟𝒞F\colon\mathcal{D}\to\mathcal{C} (i.e. it is a localization) and the adjunction LFL\dashv F satisfies the projection formula.

Proof.

In the “only if” direction, the right adjoint of R():𝒞ModR(𝒞)R\otimes(-)\colon\mathcal{C}\to\operatorname{Mod}_{R}(\mathcal{C}) is the forgetful functor F:ModR(𝒞)𝒞F\colon\operatorname{Mod}_{R}(\mathcal{C})\to\mathcal{C}, which is fully faithful since RR is idempotent. For X𝒞X\in\mathcal{C} and MModR(𝒞)M\in\operatorname{Mod}_{R}(\mathcal{C}), the projection formula transformation is the composition

XF(M)u(XF(M))FL(XF(M))F(L(X)LF(M))L(X)cMF(L(X)M).X\otimes F(M)\xrightarrow{u_{(X\otimes F(M))}}FL(X\otimes F(M))\overset{\sim}{\longrightarrow}F(L(X)\otimes LF(M))\xrightarrow{L(X)\otimes c_{M}}F(L(X)\otimes M).

The map cMc_{M} is an isomorphism since FF is fully faithful. The map u(XF(M))u_{(X\otimes F(M))} is an isomorphism because XF(M)X\otimes F(M) admits a structure of an RR-module and hence in the essential image of LL.

For the “if” direction, consider the commutative ring

R=FL(𝟙𝒞)=F(𝟙𝒟).R=FL(\mathds{1}_{\mathcal{C}})=F(\mathds{1}_{\mathcal{D}}).

By the projection formula, we have a natural isomorphism

XRXF(𝟙𝒟)F(L(X)𝟙𝒟)FL(X).X\otimes R\simeq X\otimes F(\mathds{1}_{\mathcal{D}})\simeq F(L(X)\otimes\mathds{1}_{\mathcal{D}})\simeq FL(X).

Since FF is fully faithful, the unit map Id𝑢FL\operatorname{Id}\xrightarrow{u}FL exhibits FL:𝒞𝒞FL\colon\mathcal{C}\to\mathcal{C} as a localization (as in [Lur09, Proposition 5.2.7.4]). Moreover, by 5.1.3, the unit map Id𝑢FL\operatorname{Id}\xrightarrow{u}FL is induced from tensoring with the unit 𝟙𝒞𝑢R\mathds{1}_{\mathcal{C}}\xrightarrow{u}R of RR. Thus, by [Lura, Proposition 4.8.2.4], RR is an idempotent ring. Moreover, by [Lura, Proposition 4.8.2.10], the forgetful functor ModR(𝒞)𝒞\operatorname{Mod}_{R}(\mathcal{C})\to\mathcal{C} is a symmetric monoidal equivalence onto the essential image of FLFL with the localized symmetric monoidal structure, which is equivalent to 𝒟\mathcal{D}. ∎

Remark 5.1.5.

Let 𝒞\mathcal{C} be presentably symmetric monoidal and let RR be an idempotent algebra in 𝒞\mathcal{C}. The fully faithful forgetful functor F:ModR(𝒞)𝒞F\colon\operatorname{Mod}_{R}(\mathcal{C})\hookrightarrow\mathcal{C} admits also a right adjoint. Hence, if 𝒞\mathcal{C} is moreover stable, then FF exhibits 𝒞\mathcal{C} as a recollement of ModR(𝒞)\operatorname{Mod}_{R}(\mathcal{C}) and its right orthogonal complement.

Poset structure

Another nice characterization of idempotent algebras is as the (1)(-1)-cotruncated objects of CAlg(𝒞)\operatorname{CAlg}(\mathcal{C}):

Proposition 5.1.6.

Let 𝒞CAlg(Cat)\mathcal{C}\in\operatorname{CAlg}(\operatorname{Cat}_{\infty}). A commutative algebra RCAlg(𝒞)R\in\operatorname{CAlg}(\mathcal{C}) is idempotent, if and only if for all SCAlg(𝒞)S\in\operatorname{CAlg}(\mathcal{C}) the space MapCAlg(𝒞)(R,S)\operatorname{Map}_{\operatorname{CAlg}(\mathcal{C})}(R,S) is either empty or contractible. Moreover, it is non-empty if and only if SS is an RR-module.

Proof.

By [Lura, Proposition 3.2.4.7], the tensor product of commutative algebras is the coproduct in CAlg(𝒞)\operatorname{CAlg}(\mathcal{C}). Moreover, the multiplication map

RR=RR𝑚RR\sqcup R=R\otimes R\xrightarrow{m}R

is the categorical fold map. Thus, it follows by the dual of [Lur09, Lemma .5.5.6.15], that RR is idempotent if and only if it is (1)(-1)-cotruncated. We note that [Lur09, Lemma .5.5.6.15] requires the \infty-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 RSR\to S in CAlg(𝒞)\operatorname{CAlg}(\mathcal{C}), then SS is an RR-algebra and in particular an RR-module. Conversely, for every SCAlg(𝒞)S\in\operatorname{CAlg}(\mathcal{C}), we have maps

R=R𝟙1uSRSR=R\otimes\mathds{1}\xrightarrow{1\otimes u_{S}}R\otimes S
S=𝟙SuR1RSS=\mathds{1}\otimes S\xrightarrow{u_{R}\otimes 1}R\otimes S

in CAlg(𝒞)\operatorname{CAlg}(\mathcal{C}). If SS is an RR-module, then the map uR1u_{R}\otimes 1 is an isomorphism, and thus

(uR1)1(1uS):RS(u_{R}\otimes 1)^{-1}\circ(1\otimes u_{S})\colon R\to S

is a map in CAlg(𝒞)\operatorname{CAlg}(\mathcal{C}). ∎

We also have the following non-commutative analogue of 5.1.6:

Proposition 5.1.7.

Let 𝒞CAlg(Cat)\mathcal{C}\in\operatorname{CAlg}(\operatorname{Cat}_{\infty}) and let RCAlgidem(𝒞)R\in\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}). For every algebra SAlg(𝒞)S\in\operatorname{Alg}(\mathcal{C}), the space MapAlg(𝒞)(R,S)\operatorname{Map}_{\operatorname{Alg}(\mathcal{C})}(R,S) is either empty or contractible and it is non-empty if and only if SS is an RR-module.

Proof.

As in the proof of 5.1.6, SS is an RR-module if and only if there exists a map RSR\to S in Alg(𝒞)\operatorname{Alg}(\mathcal{C}). If SS is an RR-module, then SAlg(ModR(𝒞))S\in\operatorname{Alg}(\operatorname{Mod}_{R}(\mathcal{C})) and

MapAlg(𝒞)(R,S)MapAlg(ModR(𝒞))(R,S)pt,\operatorname{Map}_{\operatorname{Alg}(\mathcal{C})}(R,S)\simeq\operatorname{Map}_{\operatorname{Alg}(\operatorname{Mod}_{R}(\mathcal{C}))}(R,S)\simeq\operatorname{pt},

since RR is the initial object of AlgR(𝒞)=Alg(ModR(𝒞))\operatorname{Alg}_{R}(\mathcal{C})=\operatorname{Alg}(\operatorname{Mod}_{R}(\mathcal{C})). ∎

As a consequence of 5.1.6, the \infty-category CAlgidem(𝒞)\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}) of idempotent algebras is a poset. This poset admits binary meets:

Proposition 5.1.8.

Let 𝒞CAlg(Cat)\mathcal{C}\in\operatorname{CAlg}(\operatorname{Cat}_{\infty}) and R,SCAlgidem(𝒞)R,S\in\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}). Then RSR\otimes S is an idempotent algebra which classifies the conjunction of the properties classified by RR and SS.

Proof.

Let 𝟙uRR\mathds{1}\xrightarrow{u_{R}}R be the unit of RR and 𝟙uSS\mathds{1}\xrightarrow{u_{S}}S the unit of SS. We consider the composition

uRS:𝟙𝟙𝟙uR1R𝟙1uSRS.u_{R\otimes S}\colon\mathds{1}\simeq\mathds{1}\otimes\mathds{1}\xrightarrow{u_{R}\otimes 1}R\otimes\mathds{1}\xrightarrow{1\otimes u_{S}}R\otimes S.

After tensoring with RR the first map becomes an isomorphism and after tensoring with SS the second map becomes an isomorphism. Thus, uRSu_{R\otimes S} exhibits RSR\otimes S as an idempotent object. Given an RSR\otimes S-module MM, tensoring 𝟙uRSRS\mathds{1}\xrightarrow{u_{R\otimes S}}R\otimes S with MM is an isomorphism. Thus, MRSMM\simeq R\otimes S\otimes M and hence MM is both an RR-module and an SS-module. Conversely, tensoring 𝟙uRSRS\mathds{1}\xrightarrow{u_{R\otimes S}}R\otimes S with MM is given by the composition:

𝟙𝟙MuR11R𝟙M1uS1RSM\mathds{1}\otimes\mathds{1}\otimes M\xrightarrow{u_{R}\otimes 1\otimes 1}R\otimes\mathds{1}\otimes M\xrightarrow{1\otimes u_{S}\otimes 1}R\otimes S\otimes M

Hence, if MM is both an RR-module and an SS-module, then tensoring 𝟙uRR\mathds{1}\xrightarrow{u_{R}}R and 𝟙uSS\mathds{1}\xrightarrow{u_{S}}S with MM is an isomorphism, and so tensoring 𝟙uR𝒮RS\mathds{1}\xrightarrow{u_{R\otimes\mathcal{S}}}R\otimes S with MM is an isomorphism as well. Thus, MM is an RSR\otimes S-module. ∎

Idempotent objects are also closed under sifted colimits in 𝒞\mathcal{C}.

Proposition 5.1.9.

Let 𝒞CAlg(Cat)\mathcal{C}\in\operatorname{CAlg}(\operatorname{Cat}_{\infty}), and let 𝒥\mathcal{J} be a sifted \infty-category such that 𝒞\mathcal{C} is compatible with 𝒥\mathcal{J}-indexed colimits. Then

  1. (1)

    The \infty-category CAlgidem(𝒞)\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}) admits 𝒥\mathcal{J}-indexed colimits.

  2. (2)

    The forgetful functor CAlgidem(𝒞)𝒞\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C})\to\mathcal{C} preserves 𝒥\mathcal{J}-indexed colimits.

  3. (3)

    Given a functor F:𝒥CAlgidem(𝒞)F\colon\mathcal{J}\to\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}), the idempotent algebra limF\underrightarrow{\operatorname{lim}}\,F classifies the conjunction of the properties classified by F(j)F(j) for all j𝒥j\in\mathcal{J}.

Proof.

By [Lura, Corollary 3.2.3.2] the \infty-category CAlg(𝒞)\operatorname{CAlg}(\mathcal{C}) admits 𝒥\mathcal{J}-indexed colimits and the forgetful functor CAlg(𝒞)𝒞\operatorname{CAlg}(\mathcal{C})\to\mathcal{C} preserves 𝒥\mathcal{J}-indexed colimits. Thus, to prove (1) and (2) it suffices to show that CAlgidem(𝒞)CAlg(𝒞)\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C})\subseteq\operatorname{CAlg}(\mathcal{C}) is closed under 𝒥\mathcal{J}-indexed colimits. Since 𝒥\mathcal{J} is sifted, the diagonal map 𝒥𝒥×𝒥\mathcal{J}\to\mathcal{J}\times\mathcal{J} is cofinal. Therefore, given a functor F:𝒥CAlg(𝒞)F\colon\mathcal{J}\to\operatorname{CAlg}(\mathcal{C}), the multiplication map

limFlimFlimF\underrightarrow{\operatorname{lim}}\,F\otimes\underrightarrow{\operatorname{lim}}\,F\to\underrightarrow{\operatorname{lim}}\,F

can be identified with the colimit of the multiplication maps F(j)F(j)F(j)F(j)\otimes F(j)\to F(j) for j𝒥j\in\mathcal{J}. Hence, if F(j)F(j) is an idempotent algebra for every j𝒥j\in\mathcal{J}, so is limF\underrightarrow{\operatorname{lim}}\,F. We shall now prove (3). First, for every j𝒥j\in\mathcal{J} there is a canonical algebra map F(j)limFF(j)\to\underrightarrow{\operatorname{lim}}\,F. Thus, every (limF\underrightarrow{\operatorname{lim}}\,F)-module admits an F(j)F(j)-module structure for every j𝒥j\in\mathcal{J}. It remains to prove that if M𝒞M\in\mathcal{C} admits an F(j)F(j)-module structure for every j𝒥j\in\mathcal{J}, then

M𝟙1uMlimFM\otimes\mathds{1}\xrightarrow{1\otimes u}M\otimes\underrightarrow{\operatorname{lim}}\,F

is an isomorphism. Since 𝒞\mathcal{C} is compatible with 𝒥\mathcal{J}-indexed colimits, the map 1u1\otimes u above is the colimit of the isomorphisms M𝟙1ujMF(j)M\otimes\mathds{1}\xrightarrow{1\otimes u_{j}}M\otimes F(j). ∎

Consequently, under mild conditions on 𝒞\mathcal{C}, the poset CAlgidem(𝒞)\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}) admits arbitrary meets.

Corollary 5.1.10.

Let 𝒞CAlg(Cat)\mathcal{C}\in\operatorname{CAlg}(\operatorname{Cat}_{\infty}) which is compatible with filtered colimits. Then the poset CAlgidem(𝒞)\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}) is cocomplete. Moreover, given a functor F:𝒥CAlgidem(𝒞)F\colon\mathcal{J}\to\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}), the idempotent algebra

limF=supj𝒥F(j)\underrightarrow{\operatorname{lim}}\,F=\sup\limits_{j\in\mathcal{J}}F(j)

classifies the conjunction of the properties classified by F(j)F(j) for all j𝒥j\in\mathcal{J}.

Proof.

First, CAlgidem(𝒞)\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}) admits an initial object which is 𝟙𝒞\mathds{1}_{\mathcal{C}}. Second, by 5.1.8, CAlgidem\operatorname{CAlg}^{\mathrm{idem}} admits binary coproducts. Since every filtered \infty-category is sifted by [Lur09, Example 5.5.8.3] we get by 5.1.9 that CAlgidem(𝒞)\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}) admits filtered colimits. Since CAlgidem(𝒞)\operatorname{CAlg}^{\mathrm{idem}}(\mathcal{C}) 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 𝒞CAlg(Cat)\mathcal{C}\in\operatorname{CAlg}(\operatorname{Cat}_{\infty}) which is compatible with all small colimits and is 0-semiadditive. Let R,SR,S be idempotent algebras in 𝒞\mathcal{C}. If RS0R\otimes S\simeq 0, then R×SR\times S is an idempotent algebra, which classifies the property of an object X𝒞X\in\mathcal{C} to be of the form XRXSX_{R}\oplus X_{S}, where XRX_{R} is an RR-module and XSX_{S} is an SS-module.

Proof.

By assumption, the tensor product preserves binary coproducts in each variable. Since 𝒞\mathcal{C} is 0-semiadditive, we get that the tensor product also preserves binary products in each variable. Thus,

(RR)×(SS)(R×S)(R×S)mR×SR×S(R\otimes R)\times(S\otimes S)\simeq(R\times S)\otimes(R\times S)\xrightarrow{m_{R\times S}}R\times S

coincides with mR×mSm_{R}\times m_{S}, which is an isomorphism. Thus, mR×Sm_{R\times S} is an isomorphism, and therefore, R×SR\times S is an idempotent algebra. The projection maps R×SRR\times S\to R and R×SSR\times S\to S induce the extension of scalars functors

ModR×S(𝒞)\textstyle{\operatorname{Mod}_{R\times S}(\mathcal{C})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FR\scriptstyle{F_{R}}FS\scriptstyle{F_{S}}ModR(𝒞)\textstyle{\operatorname{Mod}_{R}(\mathcal{C})}ModS(𝒞)\textstyle{\operatorname{Mod}_{S}(\mathcal{C})}

which in turn induce a functor

F:ModR×S(𝒞)ModR(𝒞)×ModS(𝒞).F\colon\operatorname{Mod}_{R\times S}(\mathcal{C})\to\operatorname{Mod}_{R}(\mathcal{C})\times\operatorname{Mod}_{S}(\mathcal{C}).

Since FRF_{R} and FSF_{S} are left adjoints, by [HY17, Theorem B] the functor FF admits a right adjoint GG, which is given object-wise by G(XR,XS)XR×XSG(X_{R},X_{S})\simeq X_{R}\times X_{S}. To complete the proof of the claim, it would suffice to show that GG is an equivalence. We do this by showing that GG is conservative and FF is fully faithful. By 3.3.1, in order to show that FF is fully faithful, it suffices to show that GFGF is an equivalence. For every (R×S)(R\times S)-module XX, we have

GF(X)=(RR×SX)×(SR×SX)(R×S)R×SXX.GF(X)=(R\otimes_{R\times S}X)\times(S\otimes_{R\times S}X)\simeq(R\times S)\otimes_{R\times S}X\simeq X.

To show that GG is conservative, it suffices to observe that the underlying 𝒞\mathcal{C}-object of G(XR,XS)G(X_{R},X_{S}) is the direct sum XRXSX_{R}\oplus X_{S} and both XRX_{R} and XSX_{S} are retracts of XRXSX_{R}\oplus X_{S}. ∎

5.2 Theory of Modes

We now specialize the notion of idempotent algebras to the \infty-category Pr\Pr of presentable \infty-categories and colimit preserving functors.

Tensor of presentable \infty-categories

Recall from [Lura, Proposition 4.8.1.15], that the \infty-category Pr\Pr admits a closed symmetric monoidal structure. The unit is 𝒮Pr\mathcal{S}\in\Pr, and for every 𝒞,𝒟Pr\mathcal{C},\mathcal{D}\in\Pr, the internal hom and tensor product are given respectively by

hom(𝒞,𝒟)=FunL(𝒞,𝒟),𝒞𝒟=FunR(𝒞op,𝒟)FunR(𝒟op,𝒞).\hom(\mathcal{C},\mathcal{D})=\operatorname{Fun}^{L}(\mathcal{C},\mathcal{D}),\qquad\mathcal{C}\otimes\mathcal{D}=\operatorname{Fun}^{R}(\mathcal{C}^{\text{\emph{op}}},\mathcal{D})\simeq\operatorname{Fun}^{R}(\mathcal{D}^{\text{\emph{op}}},\mathcal{C}).

It is worth spelling out in what sense the above formula for the tensor product is functorial. Given a functor 𝒟1𝐹𝒟2\mathcal{D}_{1}\xrightarrow{F}\mathcal{D}_{2} in Pr\Pr with right adjoint GG, the induced functor 𝒞𝒟1Id𝒞F𝒞𝒟2\mathcal{C}\otimes\mathcal{D}_{1}\xrightarrow{\operatorname{Id}_{\mathcal{C}}\otimes F}\mathcal{C}\otimes\mathcal{D}_{2} is the left adjoint of

FunR(𝒞op,𝒟2)G()FunR(𝒞op,𝒟1).\operatorname{Fun}^{R}(\mathcal{C}^{\text{\emph{op}}},\mathcal{D}_{2})\xrightarrow{G\circ(-)}\operatorname{Fun}^{R}(\mathcal{C}^{\text{\emph{op}}},\mathcal{D}_{1}).

From this we get that tensoring with 𝒞\mathcal{C} preserves reflective localizations:

Lemma 5.2.1.

Let 𝒞\mathcal{C} and 𝒟1𝐹𝒟2\mathcal{D}_{1}\xrightarrow{F}\mathcal{D}_{2} in Pr\Pr. If FF admits a fully faithful (resp. conservative) right adjoint, then so does Id𝒞F\operatorname{Id}_{\mathcal{C}}\otimes F .

Proof.

By the above formula for Id𝒞F\operatorname{Id}_{\mathcal{C}}\otimes F, its right adjoint is given by

FunR(𝒞op,𝒟2)G()FunR(𝒞op,𝒟1),\operatorname{Fun}^{R}(\mathcal{C}^{\text{\emph{op}}},\mathcal{D}_{2})\xrightarrow{G\circ(-)}\operatorname{Fun}^{R}(\mathcal{C}^{\text{\emph{op}}},\mathcal{D}_{1}),

where GG is the right adjoint of FF. Thus, if GG is fully faithful (resp. conservative), then post composition with GG is fully faithful (resp. conservative) as well. ∎

Remark 5.2.2.

On the other hand, if 𝒟1𝐹𝒟2\mathcal{D}_{1}\xrightarrow{F}\mathcal{D}_{2} is itself fully faithful or conservative, then Id𝒞F\operatorname{Id}_{\mathcal{C}}\otimes F need not be. For example, let SpcnSp\operatorname{Sp}^{\mathrm{cn}}\subseteq\operatorname{Sp} be the full subcategory of connective spectra. One can show that SpcnSetAb\operatorname{Sp}^{\mathrm{cn}}\otimes\operatorname{Set}\simeq\mathrm{Ab}, while SpSet0\operatorname{Sp}\otimes\operatorname{Set}\simeq 0 (e.g. by 5.2.10). Thus, tensoring the fully faithful inclusion SpcnSp\operatorname{Sp}^{\mathrm{cn}}\hookrightarrow\operatorname{Sp} with the category Set\operatorname{Set}, produces the zero functor Ab0\mathrm{Ab}\to 0.

Another general fact which we shall require is the preservation of recollements under base change. Recall that PrstPr\Pr_{\mathrm{st}}\subseteq\Pr is the full subcategory of stable presentable \infty-categories.

Proposition 5.2.3.

Let 𝒞Prst\mathcal{C}\in\Pr_{\mathrm{st}} and assume it is a recollement of 𝒞𝒞\mathcal{C}_{\circ}\subseteq\mathcal{C} and 𝒞𝒞\mathcal{C}_{\circ}^{\perp}\subseteq\mathcal{C}. For every 𝒟Pr\mathcal{D}\in\Pr, the morphism 𝒞𝒟𝒞𝒟\mathcal{C}_{\circ}\otimes\mathcal{D}\to\mathcal{C}\otimes\mathcal{D} exhibits 𝒞𝒟\mathcal{C}\otimes\mathcal{D} as a recollement of 𝒞𝒟\mathcal{C}_{\circ}\otimes\mathcal{D} and (𝒞𝒟)𝒞𝒟(\mathcal{C}_{\circ}\otimes\mathcal{D})^{\perp}\simeq\mathcal{C}_{\circ}^{\perp}\otimes\mathcal{D}.

Proof.

Let 𝒞𝐹𝒞\mathcal{C}_{\circ}\xrightarrow{F}\mathcal{C} be the inclusion functor. We denote by 𝒞𝐿𝒞\mathcal{C}\xrightarrow{L}\mathcal{C}_{\circ} and 𝒞𝑅𝒞\mathcal{C}\xrightarrow{R}\mathcal{C}_{\circ} the left and right adjoints of FF respectively. We observe that 𝒞\mathcal{C}_{\circ} is presentable as an accessible localization of 𝒞\mathcal{C}, and hence both functors FF and LL are morphisms in Pr\Pr. The adjunction FRF\dashv R induces an adjunction

F():Fun(𝒟op,𝒞)Fun(𝒟op,𝒞):R().F\circ(-)\colon\operatorname{Fun}(\mathcal{D}^{\text{\emph{op}}},\mathcal{C}_{\circ})\leftrightarrows\operatorname{Fun}(\mathcal{D}^{\text{\emph{op}}},\mathcal{C})\colon R\circ(-).

Since FF and RR are both right adjoints, this adjunction restricts to an adjunction on the full subcategories spanned by the right adjoints on both sides,

F():𝒟𝒞FunR(𝒟op,𝒞)FunR(𝒟op,𝒞)𝒟𝒞:R().F\circ(-)\colon\mathcal{D}\otimes\mathcal{C}_{\circ}\simeq\operatorname{Fun}^{R}(\mathcal{D}^{\text{\emph{op}}},\mathcal{C}_{\circ})\leftrightarrows\operatorname{Fun}^{R}(\mathcal{D}^{\text{\emph{op}}},\mathcal{C})\simeq\mathcal{D}\otimes\mathcal{C}\colon R\circ(-).

On the other hand, the left adjoint of R()R\circ(-) is F𝒟F\otimes\mathcal{D} and the left adjoint of F()F\circ(-) is L𝒟L\otimes\mathcal{D}. It follows that L𝒟L\otimes\mathcal{D} is the left adjoint of F𝒟F\otimes\mathcal{D}. To conclude, F𝒟F\otimes\mathcal{D} admits a right adjoint (as a morphism in Pr)\Pr) and also a left adjoint, given by L𝒟L\otimes\mathcal{D}. Furthermore, by 5.2.1, F𝒟F\otimes\mathcal{D} is also fully faithful. The \infty-categories 𝒞𝒟\mathcal{C}_{\circ}\otimes\mathcal{D} and 𝒞𝒟\mathcal{C}\otimes\mathcal{D} are stable by [Lura, Proposition 4.8.2.18] because 𝒞\mathcal{C}_{\circ} and 𝒞\mathcal{C} are. Hence, we deduce that F𝒟F\otimes\mathcal{D} exhibits 𝒞𝒟\mathcal{C}\otimes\mathcal{D} as a recollement of 𝒞𝒟\mathcal{C}_{\circ}\otimes\mathcal{D} and (𝒞𝒟).(\mathcal{C}_{\circ}\otimes\mathcal{D})^{\perp}.

It remains to identify (𝒞𝒟)(\mathcal{C}_{\circ}\otimes\mathcal{D})^{\perp} with 𝒞𝒟\mathcal{C}_{\circ}^{\perp}\otimes\mathcal{D}. Note that (𝒞𝒟)(\mathcal{C}_{\circ}\otimes\mathcal{D})^{\perp} is the full subcategory of 𝒞𝒟\mathcal{C}\otimes\mathcal{D} spanned by objects on which the right adjoint of F𝒟F\otimes\mathcal{D} is zero. This right adjoint is given by

R():FunR(𝒟op,𝒞)FunR(𝒟op,𝒞).R\circ(-)\colon\operatorname{Fun}^{R}(\mathcal{D}^{\text{\emph{op}}},\mathcal{C})\to\operatorname{Fun}^{R}(\mathcal{D}^{\text{\emph{op}}},\mathcal{C}_{\circ}).

On the other hand, the inclusion F:𝒞𝒞F^{\perp}\colon\mathcal{C}_{\circ}^{\perp}\hookrightarrow\mathcal{C} is right adjoint to its left adjoint L:𝒞𝒞L^{\perp}\colon\mathcal{C}\to\mathcal{C}_{\circ}^{\perp}, and hence the right adjoint of L𝒟L^{\perp}\otimes\mathcal{D} is given by

F():FunR(𝒟op,𝒞)FunR(𝒟op,𝒞).F^{\perp}\circ(-)\colon\operatorname{Fun}^{R}(\mathcal{D}^{\text{\emph{op}}},\mathcal{C}_{\circ}^{\perp})\to\operatorname{Fun}^{R}(\mathcal{D}^{\text{\emph{op}}},\mathcal{C}).

This is a fully faithful functor whose essential image consists precisely of objects on which R()R\circ(-) is zero. Thus, we have canonically identified (𝒞𝒟)(\mathcal{C}_{\circ}\otimes\mathcal{D})^{\perp} with 𝒞𝒟\mathcal{C}_{\circ}^{\bot}\otimes\mathcal{D}. ∎

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 Pr\Pr. We denote by

ModeCAlgidem(Pr)CAlg(Pr)\mathrm{Mode}\coloneqq\operatorname{CAlg}^{\mathrm{idem}}(\Pr)\subseteq\operatorname{CAlg}(\Pr)

the full subcategory spanned by modes.

Applying the preceding results on idempotent algebras to Pr\Pr, we get the following:

Proposition 5.2.5.
  1. (1)

    Mode\mathrm{Mode} is a (large) poset.

  2. (2)

    Mode\mathrm{Mode} 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. (3)

    𝒮Mode\mathcal{S}\in\mathrm{Mode} is the initial mode and it classifies the empty property (which is always satisfied).

  4. (4)

    𝒩=𝒩\mathcal{M}\sqcup\mathcal{N}=\mathcal{M}\otimes\mathcal{N} for all ,𝒩Mode\mathcal{M},\mathcal{N}\in\mathrm{Mode}.

  5. (5)

    The forgetful functor ModePr\mathrm{Mode}\to\Pr preserves sifted colimits.

  6. (6)

    0Mode0\in\mathrm{Mode} is the terminal mode and it classifies the property of being equivalent to 0.

  7. (7)

    For ,𝒩Mode\mathcal{M},\mathcal{N}\in\mathrm{Mode}, if 𝒩=0\mathcal{M}\otimes\mathcal{N}=0, then their join is given by ×𝒩\mathcal{M}\times\mathcal{N}, and it classifies the property of being a direct product of an \mathcal{M}-module and an 𝒩\mathcal{N}-module.

Proof.

(1) follows from 5.1.6. (2) follows from 5.1.10 and the fact that Pr\Pr is closed symmetric monoidal. (3) follows from the fact that 𝒮\mathcal{S} is the unit of Pr\Pr. (4) follows from 5.1.8. (5) follows from 5.1.9. (6) follows from the fact that 0 is a zero object of Pr\Pr. (7) follows from 5.1.11, since Pr\Pr is 0-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. (1)

    (01)(0\to 1) is the boolean mode which classifies the property of being equivalent to a poset (i.e. the mode of propositional logic).

  2. (2)

    Set\operatorname{Set} 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. (3)

    Ab\mathrm{Ab} is the discrete additive mode, which classifies the property of being equivalent to an ordinary additive category.

  4. (4)

    𝒮\mathcal{S}_{*} is the pointed mode, which classifies the property of having a zero object.

  5. (5)

    Sp\operatorname{Sp} is the stable mode, which classifies the property of being stable.

Given a mode \mathcal{M}, the fully faithful forgetful functor Mod(Pr)Pr\operatorname{Mod}_{\mathcal{M}}(\Pr)\hookrightarrow\Pr admits a left adjoint (i.e. localization), which is given by

𝒞𝒞=FunR(op,𝒞).\mathcal{C}\mapsto\mathcal{M}\otimes\mathcal{C}=\operatorname{Fun}^{R}(\mathcal{M}^{\text{\emph{op}}},\mathcal{C}).

This procedure should be thought of as forcing 𝒞\mathcal{C} to be in the mode \mathcal{M} in a universal way.

Example 5.2.7.

For the stable mode Sp\operatorname{Sp}, the \infty-category Sp𝒞\operatorname{Sp}\otimes\,\mathcal{C} is the stabilization Sp(𝒞)Prst\operatorname{Sp}(\mathcal{C})\in\Pr_{\mathrm{st}} [Lura, Example 4.8.1.23]. Similarly, for the discrete mode Set\operatorname{Set}, the \infty-category Set𝒞\operatorname{Set}\otimes\,\mathcal{C} is the 0-truncation τ0𝒞\tau_{\leq 0}\mathcal{C} [Lura, Remark 4.8.2.17].

The general results for idempotent algebras have the following implication:

Proposition 5.2.8.

Let \mathcal{M} be a mode and 𝒞Alg(Pr)\mathcal{C}\in\operatorname{Alg}(\Pr) which is an \mathcal{M}-module. The fully faithful embedding Mod(Pr)Pr\operatorname{Mod}_{\mathcal{M}}(\Pr)\hookrightarrow\Pr induces an equivalence of \infty-categories

LMod𝒞(Mod(Pr))LMod𝒞(Pr).\operatorname{LMod}_{\mathcal{C}}(\operatorname{Mod}_{\mathcal{M}}(\Pr))\overset{\sim}{\longrightarrow}\operatorname{LMod}_{\mathcal{C}}(\Pr).

In particular, every 𝒟LMod𝒞(Pr)\mathcal{D}\in\operatorname{LMod}_{\mathcal{C}}(\Pr) is an \mathcal{M}-module.

Proof.

Since 𝒞\mathcal{C} is an \mathcal{M}-module, by 5.1.7, there is a map of algebras 𝒞\mathcal{M}\to\mathcal{C} and the claim follows. ∎

The (,2)(\infty,2)-categorical structure of Pr\Pr allows further constructions of modes beyond those provided by 5.2.5. Primarily, modes can be localized.

Localization of modes

Given a mode \mathcal{M}, every \mathcal{M}-module 𝒞\mathcal{C} is by definition left-tensored over \mathcal{M}, and hence in particular enriched over \mathcal{M} [Lura, Proposition 4.2.1.33]. For every X,Y𝒞,X,Y\in\mathcal{C}, we denote by hom(X,Y)\hom^{\mathcal{M}}(X,Y) the corresponding hom-object in \mathcal{M}. The \mathcal{M}-enrichment of an \mathcal{M}-module 𝒞\mathcal{C} can be explicitly described via the \mathcal{M}-enriched Yoneda embedding:

𝒞𝒞FunR(𝒞op,)Fun(𝒞op,).\mathcal{C}\simeq\mathcal{C}\otimes\mathcal{M}\simeq\operatorname{Fun}^{R}(\mathcal{C}^{\text{\emph{op}}},\mathcal{M})\hookrightarrow\operatorname{Fun}(\mathcal{C}^{\text{\emph{op}}},\mathcal{M}).
Definition 5.2.9.

Let \mathcal{M} be a mode, let \mathcal{M}_{\circ}\subseteq\mathcal{M} be a reflective full subcategory, and let 𝒞\mathcal{C} be an \mathcal{M}-module. We say that an object X𝒞X\in\mathcal{C} is \mathcal{M}_{\circ}-local if for every Z𝒞Z\in\mathcal{C}, the enriched hom-object hom(Z,X)\hom^{\mathcal{M}}(Z,X) lies in .\mathcal{M}_{\circ}. Furthermore, we say that 𝒞\mathcal{C} itself is \mathcal{M}_{\circ}-local, if every object of 𝒞\mathcal{C} is \mathcal{M}_{\circ}-local.

Proposition 5.2.10.

Let \mathcal{M} be a mode and \mathcal{M}_{\circ}\subseteq\mathcal{M} an accessible reflective full subcategory, which is compatible with the symmetric monoidal structure of 𝒞\mathcal{C}. Let LL be the left adjoint of the inclusion \mathcal{M}_{\circ}\hookrightarrow\mathcal{M}. The composition

𝒮𝑢𝐿\mathcal{S}\xrightarrow{u}\mathcal{M}\xrightarrow{L}\mathcal{M}_{\circ}

exhibits \mathcal{M}_{\circ} as a mode. Moreover, for every \mathcal{M}-module 𝒞,\mathcal{C}, the \infty-category 𝒞\mathcal{M}_{\circ}\otimes\mathcal{C} can be canonically identified with the full subcategory of \mathcal{M}_{\circ}-local objects in 𝒞\mathcal{C}. In particular, \mathcal{M}_{\circ} classifies the property of being an \mathcal{M}_{\circ}-local \mathcal{M}-module.

Proof.

By [Lura, Proposition 2.2.1.9], the \infty-category \mathcal{M}_{\circ} admits a canonical symmetric monoidal structure, such that the left adjoint 𝐿\mathcal{M}\xrightarrow{L}\mathcal{M}_{\circ} promotes to a symmetric monoidal functor. In particular, the unit of this symmetric monoidal structure 𝒮u\mathcal{S}\xrightarrow{u_{\circ}}\mathcal{M}_{\circ} is the composition

𝒮𝑢𝐿,\mathcal{S}\xrightarrow{u}\mathcal{M}\xrightarrow{L}\mathcal{M}_{\circ},

where uu is the unit of the symmetric monoidal structure of \mathcal{M}. We need to show that uu_{\circ}\otimes\mathcal{M}_{\circ} is an equivalence, or equivalently, that its right adjoint GG_{\circ} is an equivalence. Let m\mathcal{M}_{\circ}\otimes\mathcal{M}_{\circ}\xrightarrow{m_{\circ}}\mathcal{M}_{\circ} be the tensor product functor. Since the composition

𝒮um\mathcal{S}\otimes\mathcal{M}_{\circ}\xrightarrow{u_{\circ}\otimes\mathcal{M}_{\circ}}\mathcal{M}_{\circ}\otimes\mathcal{M}_{\circ}\xrightarrow{m_{\circ}}\mathcal{M}_{\circ}

is an equivalence, so is the composition of the right adjoints. It follows that GG_{\circ} is essentially surjective. To complete the proof, we shall show that GG_{\circ} is also fully faithful. Write uu_{\circ}\otimes\mathcal{M}_{\circ} as the composition

𝒮uL.\mathcal{S}\otimes\mathcal{M}_{\circ}\xrightarrow{u\otimes\mathcal{M}_{\circ}}\mathcal{M}\otimes\mathcal{M}_{\circ}\xrightarrow{L\otimes\mathcal{M}_{\circ}}\mathcal{M}_{\circ}\otimes\mathcal{M}_{\circ}.

Let GG be the right adjoint of uu\otimes\mathcal{M}_{\circ}. It follows that GG_{\circ} is the composition of the right adjoint of LL\otimes\mathcal{M}_{\circ} and GG. Since LL admits a fully faithful right adjoint, by 5.2.1, the functor LL\otimes\mathcal{M}_{\circ} has a fully faithful right adjoint as well. Thus, it suffices to show that GG is fully faithful. For this, consider the commutative diagram

𝒮\textstyle{\mathcal{S}\otimes\mathcal{M}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u1\scriptstyle{u\otimes 1}\scriptstyle{\wr}1L\scriptstyle{1\otimes L}𝒮\textstyle{\mathcal{S}\otimes\mathcal{M}_{\circ}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u1\scriptstyle{u\otimes 1}\textstyle{\mathcal{M}\otimes\mathcal{M}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1L\scriptstyle{1\otimes L}.\textstyle{\mathcal{M}\otimes\mathcal{M}_{\circ}.}

Taking the right adjoints, we see that the composition of GG with the right adjoint of L\mathcal{M}\otimes L, which is fully faithful by 5.2.1, is fully faithful. It follows that GG must be fully faithful as well. This concludes the proof that uu_{\circ} exhibits \mathcal{M} as a mode.

Now, we want to analyze the property classified by \mathcal{M}_{\circ}. Since LL is symmetric monoidal, \mathcal{M}_{\circ} is a commutative algebra over \mathcal{M}. Thus, every \mathcal{M}_{\circ}-module is, in particular, an \mathcal{M}-module. Given an \mathcal{M}-module 𝒞\mathcal{C}, it is an \mathcal{M}_{\circ}-module if and only if the composition

𝒞𝒮𝒞u𝒞𝒞L𝒞\mathcal{C}\otimes\mathcal{S}\xrightarrow{\mathcal{C}\otimes u}\mathcal{C}\otimes\mathcal{M}\xrightarrow{\mathcal{C}\otimes L}\mathcal{C}\otimes\mathcal{M}_{\circ}

is an equivalence. The first functor is an equivalence since 𝒞\mathcal{C} is an \mathcal{M}-module. Thus, by 2-out-of-3, the composition is an equivalence if and only if 𝒞L\mathcal{C}\otimes L is an equivalence. The functor 𝒞L\mathcal{C}\otimes L admits by 5.2.1 a fully faithful right adjoint. To describe its essential image, we consider the commutative diagram of Yoneda embeddings

𝒞\textstyle{\mathcal{C}\otimes\mathcal{M}_{\circ}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\wr}𝒞\textstyle{\mathcal{C}\otimes\mathcal{M}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\wr}FunR(𝒞op,)\textstyle{\operatorname{Fun}^{R}(\mathcal{C}^{\text{\emph{op}}},\mathcal{M}_{\circ})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}FunR(𝒞op,)\textstyle{\operatorname{Fun}^{R}(\mathcal{C}^{\text{\emph{op}}},\mathcal{M})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fun(𝒞op,)\textstyle{\operatorname{Fun}(\mathcal{C}^{\text{\emph{op}}},\mathcal{M}_{\circ})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fun(𝒞op,).\textstyle{\operatorname{Fun}(\mathcal{C}^{\text{\emph{op}}},\mathcal{M}).}

We see that 𝒞\mathcal{C}\otimes\mathcal{M}_{\circ} is identified with the full subcategory of \mathcal{M}_{\circ}-local objects in 𝒞\mathcal{C}. ∎

Remark 5.2.11.

5.2.10 need not hold for a reflective subcategory \mathcal{M}_{\circ}\subseteq\mathcal{M}, which is not assumed in advance to be compatible with the symmetric monoidal structure. Indeed, the inclusion of co-connective spectra SpcocnSp\operatorname{Sp}^{\mathrm{co-cn}}\subseteq\operatorname{Sp} is reflective with a left adjoint τ0\tau_{\leq 0}. However, we have

SpcocnSpcocn0.\operatorname{Sp}^{\mathrm{co-cn}}\otimes\operatorname{Sp}^{\mathrm{co-cn}}\simeq 0.

Indeed, SpcocnSp\operatorname{Sp}^{\mathrm{co-cn}}\otimes\operatorname{Sp} is the \infty-category of spectrum objects in Spcocn\operatorname{Sp}^{\mathrm{co-cn}}, and thus is zero. But by 5.2.1, the functor

0=SpcocnSpIdτ0SpcocnSpcocn0=\operatorname{Sp}^{\mathrm{co-cn}}\otimes\operatorname{Sp}\xrightarrow{\operatorname{Id}\otimes\tau_{\leq 0}}\operatorname{Sp}^{\mathrm{co-cn}}\otimes\operatorname{Sp}^{\mathrm{co-cn}}

admits a fully faithful right adjoint, so SpcocnSpcocn\operatorname{Sp}^{\mathrm{co-cn}}\otimes\operatorname{Sp}^{\mathrm{co-cn}} must be zero as well.

Many further examples of modes can be constructed using 5.2.10:

Example 5.2.12.

Consider the subcategory 𝒮d𝒮\mathcal{S}_{\leq d}\subseteq\mathcal{S} of dd-truncated spaces. Every 𝒞Pr\mathcal{C}\in\Pr is an 𝒮\mathcal{S}-module and an object X𝒞X\in\mathcal{C} is 𝒮d\mathcal{S}_{\leq d}-local if and only if it is dd-truncated. Thus, 𝒮d\mathcal{S}_{\leq d} is a mode which classifies the property that every object is dd-truncated. Namely, the property of being equivalent to a (d+1)(d+1)-category (compare with 5.2.11). The cases d=2,d=-2, 1-1 and 0, reproduce the terminal mode 0, the boolean mode (01)(0\to 1), and the discrete mode Set\operatorname{Set} respectively.

Another important family of examples is the Bousfield localizations:

Example 5.2.13.

For every ESp,E\in\operatorname{Sp}, the full subcategory SpESp\operatorname{Sp}_{E}\subseteq\operatorname{Sp} of EE-local spectra is a mode and we have that SpE1SpE2\operatorname{Sp}_{E_{1}}\simeq\operatorname{Sp}_{E_{2}} in CAlg(Pr)\operatorname{CAlg}(\Pr), if and only if E1E_{1} and E2E_{2} are Bousfield equivalent. For every 𝒞Prst\mathcal{C}\in\Pr_{\mathrm{st}}, we write 𝒞ESpE𝒞\mathcal{C}_{E}\coloneqq\operatorname{Sp}_{E}\otimes\,\mathcal{C} and 𝒞(p)Sp(p)𝒞\mathcal{C}_{(p)}\coloneqq\operatorname{Sp}_{(p)}\otimes\,\mathcal{C} for a prime pp.

Remark 5.2.14.

Given E1,E2SpE_{1},E_{2}\in\operatorname{Sp}, if SpE1SpE2=0\operatorname{Sp}_{E_{1}}\otimes\operatorname{Sp}_{E_{2}}=0, then SpE1×SpE2\operatorname{Sp}_{E_{1}}\times\operatorname{Sp}_{E_{2}} is a mode (5.2.5(7)). However, it is usually not a localization of Sp\operatorname{Sp}, and in particular, it is not the same as SpE1E2\operatorname{Sp}_{E_{1}\oplus E_{2}}. For example, for every n1n\geq 1 we have

LnfSpSpk=0nT(k)≄k=0nSpT(k).L_{n}^{f}\operatorname{Sp}\simeq\operatorname{Sp}_{\oplus_{k=0}^{n}T(k)}\not\simeq\prod_{k=0}^{n}\operatorname{Sp}_{T(k)}.

Note that the right-hand side is \infty-semiadditive, while the left-hand side is not even 11-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 L:𝒩L\colon\mathcal{M}\to\mathcal{N} is smashing, if and only if the (fully faithful) right adjoint of LL admits a further right adjoint.

Proof.

In one direction, the forgetful functor ModR()\operatorname{Mod}_{R}(\mathcal{M})\to\mathcal{M} admits a right adjoint for every RCAlg()R\in\operatorname{CAlg}(\mathcal{M}) by [Lura, Remark 4.2.3.8]. Conversely, by 5.1.4, it suffices to show that if the right adjoint FF of LL admits a further right adjoint, the adjunction FLF\dashv L satisfies the projection formula. Since FF is then colimit preserving, the natural transformation in the projection formula

XF(Y)𝛼F(L(X)Y)X\otimes F(Y)\xrightarrow{\alpha}F(L(X)\otimes Y)

is a natural transformation between two functors ×𝒩\mathcal{M}\times\mathcal{N}\to\mathcal{M}, which are colimit preserving in each variable. Equivalently, these are colimit preserving functors 𝒩\mathcal{M}\otimes\mathcal{N}\to\mathcal{M} [Lura, Section 4.8.1]. Thus, it suffices to check that α\alpha becomes an isomorphism after whiskering along the equivalence

𝒩𝒮𝒩𝒩.\mathcal{N}\simeq\mathcal{S}\otimes\mathcal{N}\overset{\sim}{\longrightarrow}\mathcal{M}\otimes\mathcal{N}.

This amounts to verifying the case X=𝟙X=\mathds{1}_{\mathcal{M}}, in which, by unwinding the definitions, α\alpha is the identity and so in particular an isomorphism. ∎

Remark 5.2.16.

Every mode \mathcal{M} provides by definition a smashing localization PrMod(Pr)\Pr\to\operatorname{Mod}_{\mathcal{M}}(\Pr). Furthermore, every map of modes 𝒩\mathcal{M}\to\mathcal{N} induces a smashing localization of the \infty-categories of modules Mod(Pr)Mod𝒩(Pr)\operatorname{Mod}_{\mathcal{M}}(\Pr)\to\operatorname{Mod}_{\mathcal{N}}(\Pr). 5.2.15 characterizes those smashing localizations of Mod(Pr)\operatorname{Mod}_{\mathcal{M}}(\Pr), which arise from smashing localizations of .\mathcal{M}.

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 𝒞Pr\mathcal{C}\in\Pr and Id𝒞α,Id𝒞\operatorname{Id}_{\mathcal{C}}\xrightarrow{\alpha,}\operatorname{Id}_{\mathcal{C}}. The \infty-categories 𝒞[α1]\mathcal{C}[\alpha^{-1}] and 𝒞^α\widehat{\mathcal{C}}_{\alpha} are accessible localizations of 𝒞\mathcal{C} and hence in particular presentable. If moreover 𝒞CAlg(Pr)\mathcal{C}\in\operatorname{CAlg}(\Pr), and α\alpha is given by tensoring with 𝟙α𝟙𝟙\mathds{1}\xrightarrow{\alpha_{\mathds{1}}}\mathds{1}, then the full subcategories 𝒞[α1],𝒞^α𝒞\mathcal{C}[\alpha^{-1}],\widehat{\mathcal{C}}_{\alpha}\subseteq\mathcal{C} are compatible with the symmetric monoidal structure of 𝒞\mathcal{C} and are thus symmetric monoidal localizations of 𝒞\mathcal{C}.

Proof.

To show that 𝒞[α1]\mathcal{C}[\alpha^{-1}] and 𝒞^α\widehat{\mathcal{C}}_{\alpha} are accessible localizations of 𝒞\mathcal{C}, we use [Lur09, Propostion 5.5.4.15], by which it suffices to realize them as the full subcategories of SS-local objects with respect to a suitable (small) set of morphisms in 𝒞\mathcal{C}. Since 𝒞\mathcal{C} is presentable, it is κ\kappa-compactly generated for some cardinal κ\kappa. In particular, the subcategory 𝒞κ𝒞\mathcal{C}^{\kappa}\subseteq\mathcal{C} of κ\kappa-compact objects is essentially small and generates 𝒞\mathcal{C} under colimits. For 𝒞[α1]\mathcal{C}[\alpha^{-1}], we take SS to be the collection of maps X𝛼XX\xrightarrow{\alpha}X for XX in (a set of representatives of) 𝒞κ\mathcal{C}^{\kappa}. For 𝒞^α\widehat{\mathcal{C}}_{\alpha}, we can take SS^{\prime} to be the collection of maps 0Z0\to Z for ZZ in (a set of representatives of) τ\tau-compact objects for τ\tau large enough so that 𝒞[α1]\mathcal{C}[\alpha^{-1}] is τ\tau-compactly generated and the inclusion 𝒞[α1]𝒞\mathcal{C}[\alpha^{-1}]\hookrightarrow\mathcal{C} is τ\tau-accessible.

For 𝒞CAlg(Pr)\mathcal{C}\in\operatorname{CAlg}(\Pr), the assumption on α\alpha implies that for all X,Y𝒞X,Y\in\mathcal{C} we have αXY=XαY\alpha_{X\otimes Y}=X\otimes\alpha_{Y} and similarly for the adjoint αhom(X,Y)=hom(X,αY)\alpha_{\hom(X,Y)}=\hom(X,\alpha_{Y}). In particular, the class of α\alpha-divisible objects is closed under tensoring and exponentiation by any object of 𝒞\mathcal{C}. Now, the class of morphisms S¯\overline{S} in 𝒞\mathcal{C}, which are mapped to isomorphisms under the localization 𝒞𝒞[α1]\mathcal{C}\to\mathcal{C}[\alpha^{-1}], is the set of maps X𝑓YX\xrightarrow{f}Y in 𝒞\mathcal{C} such that for every W𝒞[α1]W\in\mathcal{C}[\alpha^{-1}], the map

Map(Y,W)()fMap(X,W)\operatorname{Map}(Y,W)\xrightarrow{(-)\circ f}\operatorname{Map}(X,W)

is an isomorphism. For any Z𝒞Z\in\mathcal{C}, we have that hom(Z,W)𝒞[α1]\hom(Z,W)\in\mathcal{C}[\alpha^{-1}]. Thus, by adjointness, ZfS¯Z\otimes f\in\overline{S}. The argument for 𝒞^α\widehat{\mathcal{C}}_{\alpha} is similar but simpler. It again suffices to show that for Z𝒞Z\in\mathcal{C} and W𝒞^αW\in\widehat{\mathcal{C}}_{\alpha}, the object hom𝒞(W,Z)\hom^{\mathcal{C}}(W,Z) is α\alpha-complete. This follows from the fact that for any α\alpha-divisible XX, the object WXW\otimes X is α\alpha-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, \infty-semiadditive and of semiadditive height nn, and compare it with SpT(n)\operatorname{Sp}_{T(n)}.

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

CMonmCMon2𝒮\operatorname{CMon}_{m}\to\operatorname{CMon}_{-2}\simeq\mathcal{S}

admits a left adjoint 𝟙Pr=𝒮CMonm\mathds{1}_{\Pr}=\mathcal{S}\to\operatorname{CMon}_{m}. We consider CMonm\operatorname{CMon}_{m} as an object of Pr𝟙/\Pr_{\mathds{1}/} via this left adjoint.

Proposition 5.3.1.

Let 2m-2\leq m\leq\infty. CMonm\operatorname{CMon}_{m} is a mode, which classifies mm-semiadditivity. Moreover, for every 𝒞Pr\mathcal{C}\in\Pr, there is a canonical equivalence141414This can be compared with the fact that the stabilization Sp𝒞\operatorname{Sp}\otimes\,\mathcal{C} can be identified with Sp(𝒞)\operatorname{Sp}(\mathcal{C}), the \infty-category of spectrum objects in 𝒞\mathcal{C}.

CMonm𝒞CMonm(𝒞).\operatorname{CMon}_{m}\otimes\,\mathcal{C}\simeq\operatorname{CMon}_{m}(\mathcal{C}).
Proof.

We first treat the case m<m<\infty. The fact that CMonm\operatorname{CMon}_{m} is a mode, which classifies mm-semiadditivity is exactly [Har17, Corollary 5.21]. Consider now the inclusion ι:Pr

-
m
Pr
\iota\colon\Pr^{\scalebox{0.6}{$\oplus$}\text{-}m}\hookrightarrow\Pr
. To prove that CMonm𝒞CMonm(𝒞)\operatorname{CMon}_{m}\otimes\,\mathcal{C}\simeq\operatorname{CMon}_{m}(\mathcal{C}), 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 ι\iota.

We now consider the case m=m=\infty. For verious kk, the left adjoints of the forgetful functors CMonk+1CMonk\operatorname{CMon}_{k+1}\to\operatorname{CMon}_{k} can be considered as maps in CAlgidem(Pr)Pr𝟙/\operatorname{CAlg}^{\mathrm{idem}}(\Pr)\subseteq\Pr_{\mathds{1}/}. Thus, by 2.1.15 and 5.2.5(5), we have that

CMonlimkCMonkCAlgidem(Pr)\operatorname{CMon}_{\infty}\simeq\underrightarrow{\operatorname{lim}}\,_{k}\operatorname{CMon}_{k}\quad\in\quad\operatorname{CAlg}^{\mathrm{idem}}(\Pr)

is a mode, classifying the property of being kk-semiadditive for every kk. In other words, CMon\operatorname{CMon}_{\infty} is a mode classifying \infty-semiadditivity. Finally, since Pr\Pr is closed symmetric monoidal, for every 𝒞Pr\mathcal{C}\in\Pr we have

CMon𝒞(limkCMonk)𝒞limk(CMonk𝒞)limkCMonk(𝒞)CMon(C).\operatorname{CMon}_{\infty}\otimes\,\mathcal{C}\simeq(\underrightarrow{\operatorname{lim}}\,_{k}\operatorname{CMon}_{k})\otimes\mathcal{C}\simeq\underrightarrow{\operatorname{lim}}\,_{k}(\operatorname{CMon}_{k}\otimes\,\mathcal{C})\simeq\underrightarrow{\operatorname{lim}}\,_{k}\operatorname{CMon}_{k}(\mathcal{C})\simeq\operatorname{CMon}_{\infty}(C).

Remark 5.3.2.

The mere fact that mm-semiadditivity is classified by a mode has already non-trivial implications. For example, given 𝒞𝐹𝒟\mathcal{C}\xrightarrow{F}\mathcal{D} in Alg(Pr)\operatorname{Alg}(\Pr), we get by 5.2.8, that if 𝒞\mathcal{C} is mm-semiadditive then so is 𝒟\mathcal{D} (compare [CSY18, Corollary 3.3.2(2)]).

To study the interaction of higher semiadditivity and stability we introduce the stable mm-semiadditive mode:

Definition 5.3.3.

For every 0m0\leq m\leq\infty, we define the \infty-category of mm-commutative monoids in spectra:

[m]CMonm(Sp).{}^{[m]}\coloneqq\operatorname{CMon}_{m}(\operatorname{Sp}).

We denote [∞] simply by (and observe that [0]Sp{}^{[0]}\simeq\operatorname{Sp}).

It is an immediate consequence of 5.3.1 and 5.2.5(4), that for all values 0m0\leq m\leq\infty, the \infty-category [m] is a mode which classifies the property of being mm-semiadditive and stable.

The mode [m] plays an analogous role to CMonm\operatorname{CMon}_{m} for stable mm-semiadditive \infty-categories (even for the non-presentable ones). In particular, the CMonm\operatorname{CMon}_{m}-enriched Yoneda embedding of an mm-semiadditive \infty-category can be further lifted to [m], provided that the \infty-category is stable. To see this, for a stable \infty-category 𝒞\mathcal{C}, we already have a limit preserving spectral Yoneda embedding (e.g. as can be deduced from [Lura, Corollary 1.4.2.23])

HSp:𝒞Fun(𝒞op,Sp).\text{\char 72\relax}^{\operatorname{Sp}}\colon\mathcal{C}\hookrightarrow\operatorname{Fun}(\mathcal{C}^{\text{\emph{op}}},\operatorname{Sp}).

By 2.1.16, if 𝒞\mathcal{C} is in addition mm-semiadditive, we have a canonical equivalence

Funm-fin(𝒞op,CMonm(Sp))Funm-fin(𝒞op,Sp)).\operatorname{Fun}^{m\text{-}\mathrm{fin}}(\mathcal{C}^{\text{\emph{op}}},\operatorname{CMon}_{m}(\operatorname{Sp}))\simeq\operatorname{Fun}^{m\text{-}\mathrm{fin}}(\mathcal{C}^{\text{\emph{op}}},\operatorname{Sp})).

Thus, we get a [m]-enriched Yoneda embedding functor

H[m]:𝒞Funm-fin(𝒞op,)[m]Fun(𝒞op,)[m].\text{\char 72\relax}^{{}^{[m]}}\colon\mathcal{C}\to\operatorname{Fun}^{m\text{-}\mathrm{fin}}(\mathcal{C}^{\text{\emph{op}}},{}^{[m]})\subseteq\operatorname{Fun}(\mathcal{C}^{\text{\emph{op}}},{}^{[m]}).

We note that this functor is fully faithful, exact and mm-semiadditive. Using the [m]-enriched Yoneda embedding, we can characterize the semiadditive height of an object.

Proposition 5.3.4.

Let 𝒞Catst

-
m
{\cal C}\in\mathrm{Cat}_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}m}
and let X𝒞X\in{\cal C}. For every 0nm,0\leq n\leq m, the object XX is of height n\leq n (resp. >n1>n-1) if and only if hom[m](Z,X)[m]\hom^{{}^{[m]}}(Z,X)\in{}^{[m]} is of height n\leq n (resp. >n1>n-1), for every object Z𝒞Z\in\mathcal{C}.

Proof.

Using the [m]-enriched Yoneda embedding for 𝒞op\mathcal{C}^{\text{\emph{op}}}, the functors

hom[m](Z,):𝒞[m],Z𝒞\hom^{{}^{[m]}}(Z,-)\colon\mathcal{C}\to{}^{[m]}\quad,\quad Z\in\mathcal{C}

are mm-semiadditive and jointly conservative. Thus, by 4.2.3, XX is of height n\leq n or >n1>n-1 if and only if hom[m](Z,X)\hom^{{}^{[m]}}(Z,X) is so for all Z𝒞Z\in\mathcal{C}. ∎

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 𝒞Catst

-
m
\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}m}
, there exists 𝒞^Prst

-
m
\widehat{\mathcal{C}}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}m}
and a fully faithful, mm-semiadditive and exact embedding 𝒞𝒞^\mathcal{C}\hookrightarrow\widehat{\mathcal{C}}.

Proof.

We take

H[m]:𝒞Fun(𝒞op,)[m]𝒞^,\text{\char 72\relax}^{{}^{[m]}}\colon\mathcal{C}\hookrightarrow\operatorname{Fun}(\mathcal{C}^{\text{\emph{op}}},{}^{[m]})\eqqcolon\widehat{\mathcal{C}},

which satisfies all the required properties. ∎

Modes of semiadditive height

We can further concentrate on stable higher semiadditive \infty-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 pp and 0n0\leq n\leq\infty, there exists a mode n which classifies the property of being stable pp-local \infty-semiadditive of height nn151515We keep pp implicit in the notation n, by analogy with SpT(n)\operatorname{Sp}_{T(n)} and SpK(n)\operatorname{Sp}_{K(n)}.. Moreover, n can be canonically identified with [m](p),n{}_{(p),n}^{[m]} for every nmn\leq m\leq\infty.

Proof.

We first consider the case n<n<\infty. For every mnm\geq n, we have that

=(p),n[m]()(p)[m]^p(n1)[p(n)1]{}_{(p),n}^{[m]}=\widehat{({}_{(p)}^{[m]})}_{p_{(n-1)}}[p_{(n)}^{-1}]

is a symmetric monoidal localization of [m](p){}_{(p)}^{[m]} (5.2.17). By 5.3.4, for every pp-local 𝒞Prst

-
m
\mathcal{C}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}m}
, an object X𝒞X\in\mathcal{C} is [m](p),n{}_{(p),n}^{[m]}-local, if and only if ht(X)=n\mathrm{ht}(X)=n. Hence, we can apply 5.2.10 to deduce that [m](p),n{}_{(p),n}^{[m]} is itself a mode classifying the property of being stable pp-local mm-semiadditive and of height nn. Finally, by 3.2.7, every such \infty-category is \infty-semiadditive and hence we can take n to be [m](p),n{}_{(p),n}^{[m]}.

For n=n=\infty, we have

=(p),n.(p),>n{}_{(p),\infty}=\bigcap_{n\in\mathbb{N}}{}_{(p),>n}.

Each =(p),>n((p)^)p(n){}_{(p),>n}=(\widehat{{}_{(p)}})_{p_{(n)}} is a symmetric monoidal localization of [m](p){}_{(p)}^{[m]} (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 pp-local \infty-semiadditive and of height >n>n for all nn, which is the same as being of height \infty. ∎

Example 5.3.7.

In the case n=0n=0, we have

0(p),0[0]Sp(p),0Sp.{}_{0}\simeq{}_{(p),0}^{[0]}\simeq\operatorname{Sp}_{(p),0}\simeq\operatorname{Sp}_{\mathbb{Q}}.

As with the SpT(n)\operatorname{Sp}_{T(n)}-s, the modes n are disjoint for different nn-s.

Proposition 5.3.8.

For all nkn\neq k, we have n=k0{}_{n}\otimes{}_{k}=0.

Proof.

The \infty-category nk{}_{n}\otimes{}_{k} is \infty-semiadditive in which every object is of height both nn and kk, and hence must be the zero object. ∎

Another aspect in which n resembles SpT(n)\operatorname{Sp}_{T(n)}, is that it kills all bounded above spectra:

Proposition 5.3.9.

For 1n<1\leq n<\infty, the map of modes Spn\operatorname{Sp}\to{}_{n} vanishes on all bounded above spectra.

Proof.

Denote by F:SpnF\colon\operatorname{Sp}\to{}_{n} the map of modes. The class of spectra on which FF vanishes is closed under colimits and desuspensions in Sp\operatorname{Sp}. Hence, by a standard devissage argument, it suffices to show that FF vanishes on \mathbb{Q} and 𝔽\mathbb{F}_{\ell} for all primes \ell. First, \mathbb{Q} and 𝔽\mathbb{F}_{\ell} for p\ell\neq p are pp-divisible. Since FF is 0-semiadditive, F()F(\mathbb{Q}) and F(𝔽)F(\mathbb{F}_{\ell}) are pp-divisible as well, but all objects of n are pp-complete, and so F()=F(𝔽)=0F(\mathbb{Q})=F(\mathbb{F}_{\ell})=0. Thus, it remains to show that F(𝔽p)=0F(\mathbb{F}_{p})=0. For every kk\in\mathbb{N} we denote by 𝕊¯[BkCp]\overline{\mathbb{S}}[B^{k}C_{p}] the fiber of the fold map 𝕊[BkCp]𝕊.\mathbb{S}[B^{k}C_{p}]\to\mathbb{S}. Applying FF, we get a fiber sequence

F(𝕊¯[BkCp])𝟙n[BkCp]𝟙n.F(\overline{\mathbb{S}}[B^{k}C_{p}])\to\mathds{1}_{{}_{n}}[B^{k}C_{p}]\to\mathds{1}_{{}_{n}}.

Since n is by definition of height nn, it follows from 3.2.3(2), that the second map above is an isomorphism for kn+1k\geq n+1 and hence F(𝕊¯[BkCp])=0F(\overline{\mathbb{S}}[B^{k}C_{p}])=0. We observe that 𝔽p\mathbb{F}_{p} can be written as a filtered colimit 𝔽p=limΣk𝕊¯[BkCp]\mathbb{F}_{p}=\underrightarrow{\operatorname{lim}}\,\Sigma^{-k}\overline{\mathbb{S}}[B^{k}C_{p}]. Thus, we also get

F(𝔽p)=limΣkF(𝕊¯[BkCp])=0.F(\mathbb{F}_{p})=\underrightarrow{\operatorname{lim}}\,\Sigma^{-k}F(\overline{\mathbb{S}}[B^{k}C_{p}])=0.

Corollary 5.3.10.

For 1n<1\leq n<\infty, the right adjoint of the map of modes 𝒮n\mathcal{S}\to{}_{n} is conservative.

Proof.

The map of modes 𝒮n\mathcal{S}\to{}_{n} is given by the composition

𝒮𝕊[]SpF1Sp(p)F2F3(p)[n].n\mathcal{S}\xrightarrow{\mathbb{S}[-]}\operatorname{Sp}\xrightarrow{F_{1}}\operatorname{Sp}_{(p)}\xrightarrow{F_{2}}{}_{(p)}^{[n]}\xrightarrow{F_{3}}{}_{n}.

The right adjoints G1G_{1} and G3G_{3} of F1F_{1} and F3F_{3} respectively, are fully faithful embeddings and hence in particular conservative. The right adjoint G2G_{2} of F2F_{2} can be identified with the functor

=(p)[n]CMonn(Sp(p))Sp(p){}_{(p)}^{[n]}=\operatorname{CMon}_{n}(\operatorname{Sp}_{(p)})\to\operatorname{Sp}_{(p)}

which evaluates at the point. Thus, G2G_{2} is conservative by [Har17, Lemma 5.17]. Combining the above, the right adjoint of the functor

F=F3F2F1:SpnF=F_{3}F_{2}F_{1}\colon\operatorname{Sp}\to{}_{n}

is given by G=G1G2G3G=G_{1}G_{2}G_{3}, and is therefore conservative.

Now, the right adjoint of 𝒮n\mathcal{S}\to{}_{n}, is given by the composition of the right adjoints 𝐺nSpΩ𝒮{}_{n}\xrightarrow{G}\operatorname{Sp}\xrightarrow{\Omega^{\infty}}\mathcal{S}. Let X𝑓YX\xrightarrow{f}Y be a map in n with fiber ZZ, such that ΩG(f)\Omega^{\infty}G(f) is an isomorphism. It follows that ΩG(Z)=0\Omega^{\infty}G(Z)=0 and hence G(Z)G(Z) is co-connective. By 5.3.9, we get FG(Z)=0FG(Z)=0 and hence GFG(Z)=0GFG(Z)=0. By the zig-zag identities, G(Z)G(Z) is a retract of GFG(Z)GFG(Z) and hence we also get G(Z)=0G(Z)=0. Finally, since GG is conservative, we get Z=0Z=0 and hence X𝑓YX\xrightarrow{f}Y is an isomorphism. This concludes the proof that ΩG\Omega^{\infty}G, the right adjoint of 𝒮n\mathcal{S}\to{}_{n}, is conservative. ∎

Remark 5.3.11.

4.2.8 is equivalent to the statement that =0{}_{\infty}=0.

5.4 1-Semiadditive Decomposition

As we recalled in 4.1.3, while for n1n\geq 1 the \infty-category LnfSpL_{n}^{f}\operatorname{Sp} itself is not even 11-semiadditive, it is an iterated recollement of the \infty-categories SpT(k)\operatorname{Sp}_{T(k)} for k=0,,nk=0,\dots,n, which are \infty-semiadditive. The theory of modes allows us to enforce mm-semiadditivity on LnfSpL_{n}^{f}\operatorname{Sp} in a universal way, by tensoring it with the mm-semiadditive mode CMonm\operatorname{CMon}_{m}. We shall show that enforcing even 11-semiadditivity on LnfSpL_{n}^{f}\operatorname{Sp}, “dissolves the glue” which holds the monochromatic layers together, and decomposes it into a product of SpT(k)\operatorname{Sp}_{T(k)} for k=0,,nk=0,\dots,n.

Remark 5.4.1.

The fact that LnfSpL_{n}^{f}\operatorname{Sp} is a recollement of SpT(k)\operatorname{Sp}_{T(k)} for k=0,,nk=0,\dots,n, is closely related to the fact that for k0<k1k_{0}<k_{1}, every exact functor SpT(k0)SpT(k1)\operatorname{Sp}_{T(k_{0})}\to\operatorname{Sp}_{T(k_{1})} must be zero. The fact that the recollement becomes split after imposing 11-semiadditivity follows from the fact that every 11-semiadditive functor SpT(k1)SpT(k0)\operatorname{Sp}_{T(k_{1})}\to\operatorname{Sp}_{T(k_{0})} must be zero (5.4.9). Note that by 4.2.3, a k1k_{1}-semiadditive functor SpT(k1)SpT(k0)\operatorname{Sp}_{T(k_{1})}\to\operatorname{Sp}_{T(k_{0})} must be zero because every object of SpT(k1)\operatorname{Sp}_{T(k_{1})} is of height k1k_{1}, and thus must be sent to the only object of height k1k_{1} in SpT(k0)\operatorname{Sp}_{T(k_{0})}, 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 11-semiadditivity.

Divisible and complete cardinalities

We begin with some general observations and constructions. Every 𝒞Prst

-
1
\mathcal{C}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}1}
is a module over [1] and thus, every απ0(𝟙[1])\alpha\in\pi_{0}(\mathds{1}_{{}^{[1]}}) induces a natural transformation Id𝒞α𝒞Id𝒞\operatorname{Id}_{\mathcal{C}}\xrightarrow{\alpha_{\mathcal{C}}}\operatorname{Id}_{\mathcal{C}}. If in addition 𝒞\mathcal{C} is presentably symmetric monoidal, then there is a unique functor [1]𝒞{}^{[1]}\to\mathcal{C} in CAlg(Pr)\operatorname{CAlg}(\Pr) (see 5.1.6), which induces a map π0(𝟙[1])π0(𝟙𝒞)\pi_{0}(\mathds{1}_{{}^{[1]}})\to\pi_{0}(\mathds{1}_{\mathcal{C}}). The natural transformation α𝒞\alpha_{\mathcal{C}} can be identified with the one induced by the image of α\alpha in π0(𝟙𝒞)\pi_{0}(\mathds{1}_{\mathcal{C}}). We now restrict to a particular subset of π0(𝟙[1])\pi_{0}(\mathds{1}_{{}^{[1]}}) consisting of elements which have a simple description, as they are of an unstable origin.

Definition 5.4.2.

Let 1π0(𝟙[1])\mathcal{R}_{1}\subseteq\pi_{0}(\mathds{1}_{{}^{[1]}}) be the \mathbb{Z}-linear span of elements of the form |BG|π0(𝟙[1])|BG|\in\pi_{0}(\mathds{1}_{{}^{[1]}}), for GG a finite group.

We observe that the action of 1\mathcal{R}_{1} on the objects of any stable 11-semiadditive \infty-category 𝒞\mathcal{C} is natural with respect to 11-semiadditive functors.

Proposition 5.4.3.

Let F:𝒞𝒟F\colon\mathcal{C}\to\mathcal{D} be a 11-semiadditive functor between stable presentable 11-semiadditive \infty-categories. For every α1\alpha\in\mathcal{R}_{1}, we have F(α𝒞)=α𝒟F(\alpha_{\mathcal{C}})=\alpha_{\mathcal{D}}.

Proof.

By 0-semiadditivity, it suffices to consider the case α=|BG|\alpha=|BG| for GG a finite group, which follows from the fact that FF is 11-semiadditive. ∎

Remark 5.4.4.

If FF is further assumed to be colimit preserving, then F(α𝒞)=α𝒟F(\alpha_{\mathcal{C}})=\alpha_{\mathcal{D}} for all elements απ0(𝟙[1])\alpha\in\pi_{0}(\mathds{1}_{{}^{[1]}}). However, we shall be interested in applying 5.4.3 to functors FF, which are not a priori colimit preserving (e.g. right adjoints).

Recall from [CSY18, Theorem 4.3.2], that for every 11-semiadditive stable presentably symmetric monoidal \infty-category 𝒞\mathcal{C}, the ring π0(𝟙𝒞)\pi_{0}(\mathds{1}_{\mathcal{C}}) is equipped with a canonical additive pp-derivation

δ:π0(𝟙𝒞)π0(𝟙𝒞).\delta\colon\pi_{0}(\mathds{1}_{\mathcal{C}})\to\pi_{0}(\mathds{1}_{\mathcal{C}}).
Proposition 5.4.5.

The subset 1π0(𝟙[1])\mathcal{R}_{1}\subseteq\pi_{0}(\mathds{1}_{{}^{[1]}}) is closed under multiplication and the additive pp-derivation δ\delta inside π0(𝟙[1])\pi_{0}(\mathds{1}_{{}^{[1]}}). Consequently, 1\mathcal{R}_{1} is a semi-δ\delta-ring and the inclusion 1π0(𝟙[1])\mathcal{R}_{1}\hookrightarrow\pi_{0}(\mathds{1}_{{}^{[1]}}) is a homomorphism of semi-δ\delta-rings.

Proof.

The closure under multiplication follows from the identity (2.1.10)

|BG||BH|=|B(G×H)||BG||BH|=|B(G\times H)|

and the closure under δ\delta follows from the identity

δ|BG|=|BGCp||B(Cp×G)|,\delta|BG|=|BG\wr C_{p}|-|B(C_{p}\times G)|,

the formula (see [CSY18, Definition 4.1.1(1)])

δ(x+y)=δ(x)+δ(y)+xp+yp(x+y)pp\delta(x+y)=\delta(x)+\delta(y)+\frac{x^{p}+y^{p}-(x+y)^{p}}{p}

and the closure of 1\mathcal{R}_{1} under multiplication. ∎

11-semiadditive decomposition

We would now like to apply the above to the \infty-categories SpT(n)\operatorname{Sp}_{T(n)}. By [CSY18, Proposition 4.3.4], the construction of the additive pp-derivation δ\delta is functorial with respect to colimit preserving symmetric monoidal functors. In particular, assuming that n1n\geq 1, both maps

π0(𝟙[1])π0𝕊T(n)𝑢π0En\pi_{0}(\mathds{1}_{{}^{[1]}})\to\pi_{0}\mathbb{S}_{T(n)}\xrightarrow{u}\pi_{0}E_{n}

commute with δ\delta. By 2.2.6, the image of uu is the canonical p\mathbb{Z}_{p} copy obtained from the map

pπ0𝕊^pπ0En.\mathbb{Z}_{p}\simeq\pi_{0}\widehat{\mathbb{S}}_{p}\hookrightarrow\pi_{0}E_{n}.

Since pπ0En\mathbb{Z}_{p}\subset\pi_{0}E_{n} is the image of a semi-δ\delta-ring map we obtain a surjective map of semi-δ\delta-rings

u:π0𝕊T(n)p.u\colon\pi_{0}\mathbb{S}_{T(n)}\to\mathbb{Z}_{p}.

The semi-δ\delta-ring structure on p\mathbb{Z}_{p} can be explicitly described.

Lemma 5.4.6.

The unique semi-δ\delta-ring structure on p\mathbb{Z}_{p} is given by

δ(a)=aapp,ap.\delta(a)=\frac{a-a^{p}}{p},\qquad\forall a\in\mathbb{Z}_{p}.

In particular, if vp(a)>0v_{p}(a)>0, then vp(δ(a))=vp(a)1v_{p}(\delta(a))=v_{p}(a)-1.

Proof.

For apa\in\mathbb{Z}_{p} we define ϕ(a)=aappδ(a)\phi(a)=\frac{a-a^{p}}{p}-\delta(a). By [CSY18, Definition 4.1.1(1)], the function ϕ:pp\phi\colon\mathbb{Z}_{p}\to\mathbb{Z}_{p} is additive. By [CSY18, Proposition 4.1.11], ϕ\phi factors trough a map p/(p)p{\mathbb{Z}_{p}}/{\mathbb{Z}_{(p)}}\to\mathbb{Z}_{p}. Since p/(p){\mathbb{Z}_{p}}/{\mathbb{Z}_{(p)}} is pp-divisible, all such maps are zero. ∎

For aπ0𝕊T(n)a\in\pi_{0}\mathbb{S}_{T(n)}, we denote by vp(a)v_{p}(a) the pp-adic valuation of u(a)pπ0Enu(a)\in\mathbb{Z}_{p}\subseteq\pi_{0}E_{n}.

Proposition 5.4.7.

For every n1n\geq 1 and aπ0𝕊T(n)a\in\pi_{0}\mathbb{S}_{T(n)} we have:

  1. (1)

    If vp(a)=0v_{p}(a)=0, then SpT(n)\operatorname{Sp}_{T(n)} is aa-divisible.

  2. (2)

    If vp(a)>0v_{p}(a)>0, then SpT(n)\operatorname{Sp}_{T(n)} is aa-complete.

Proof.

For (1), if vp(a)=0v_{p}(a)=0, then u(a)π0Enu(a)\in\pi_{0}E_{n} is invertible. It follows that aπ0𝕊T(n)a\in\pi_{0}\mathbb{S}_{T(n)} is invertible by nil-conservativity [CSY18, Corollary 5.1.17]. For (2), let aπ0𝕊T(n)a\in\pi_{0}\mathbb{S}_{T(n)} such that vp(a)>0v_{p}(a)>0. We need to show that for every 0XSpT(n)0\neq X\in\operatorname{Sp}_{T(n)}, the element aa acts non-invertibly on XX. Since tensoring with T(n)T(n) is conservative, we may replace XX by T(n)XT(n)\otimes X. Without loss of generality, we may choose T(n)T(n) to be a ring spectrum. Thus, T(n)XT(n)\otimes X is a T(n)T(n)-module and the action of aa on it is by its image under the map π0𝕊T(n)π0T(n)\pi_{0}\mathbb{S}_{T(n)}\to\pi_{0}T(n), which we denote by a¯\overline{a}. Thus, it would suffice to show that a¯π0T(n)\overline{a}\in\pi_{0}T(n) is nilpotent. By the Nilpotence Theorem, it suffices to show that the image of a¯\overline{a} under the map

π0T(n)π0(T(n)En)\pi_{0}T(n)\to\pi_{0}(T(n)\otimes E_{n})

is nilpotent, as T(n)K(j)=0T(n)\otimes K(j)=0 for all j=n+1,,j=n+1,\dots,\infty. Consider the commutative diagram

π0𝕊T(n)\textstyle{\pi_{0}\mathbb{S}_{T(n)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u\scriptstyle{u}π0En\textstyle{\pi_{0}E_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π0T(n)\textstyle{\pi_{0}T(n)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π0(T(n)En)\textstyle{\pi_{0}(T(n)\otimes E_{n})}

We see that the image of a¯\overline{a} in π0(T(n)En)\pi_{0}(T(n)\otimes E_{n}) is contained in the image of uu, which lies in pπ0En\mathbb{Z}_{p}\subseteq\pi_{0}E_{n} (2.2.6). Since T(n)T(n) is pp-power torsion, the map

pπ0Enπ0(T(n)En)\mathbb{Z}_{p}\subseteq\pi_{0}E_{n}\to\pi_{0}(T(n)\otimes E_{n})

factors through a finite quotient p/pr\mathbb{Z}_{p}\twoheadrightarrow\mathbb{Z}/p^{r} and hence the image of a¯\overline{a} in π0(T(n)En)\pi_{0}(T(n)\otimes E_{n}) is nilpotent. ∎

Proposition 5.4.8.

For every k0k\geq 0, there exists an element α1\alpha\in\mathcal{R}_{1}, such that SpT(n)\operatorname{Sp}_{T(n)} is α\alpha-complete for nkn\leq k and α\alpha-divisible for n>kn>k.

Proof.

First, we observe that it suffices to construct elements β(k)\beta_{(k)} such that SpT(k)\operatorname{Sp}_{T(k)} is β(k)\beta_{(k)}-complete and SpT(n)\operatorname{Sp}_{T(n)} is β(k)\beta_{(k)}-divisible for n>kn>k. Indeed, we can then define α=β(0)β(k)\alpha=\beta_{(0)}\cdots\beta_{(k)}, which satisfies the required properties. To construct ββ(k)\beta\coloneqq\beta_{(k)} for a specific k0k\geq 0, we proceed as follows. For k=0k=0, we take β=p|BCp|1\beta=p|BC_{p}|-1. We get that for SpT(0)=Sp\operatorname{Sp}_{T(0)}=\operatorname{Sp}_{\mathbb{Q}}, the element β\beta acts as zero, while for n>0n>0, the \infty-category SpT(n)\operatorname{Sp}_{T(n)} is pp-complete, and hence by 4.1.14, it is β\beta-divisible. Now, we treat the case k1k\geq 1. For every a1π0(𝟙[1])a\in\mathcal{R}_{1}\subseteq\pi_{0}(\mathds{1}_{{}^{[1]}}), we shall denote by ana_{n} its image in pπ0En\mathbb{Z}_{p}\subseteq\pi_{0}E_{n} under the map π0(𝟙[1])π0En\pi_{0}(\mathds{1}_{{}^{[1]}})\to\pi_{0}E_{n}. By 5.4.7, it suffices to construct β\beta, such that vp(βk)>0v_{p}(\beta_{k})>0 and vp(βn)=0v_{p}(\beta_{n})=0 for all nk+1n\geq k+1. Using 5.4.5, we define the element

γ=δk1|BCp|1.\gamma=\delta^{k-1}|BC_{p}|\in\mathcal{R}_{1}.

By 2.2.5, we have |BCp|n=pn1|BC_{p}|_{n}=p^{n-1} for all nn\in\mathbb{N}. Thus, by 5.4.6, we have vp(γn)=nkv_{p}(\gamma_{n})=n-k for all nkn\geq k, and in particular vp(γk)=0v_{p}(\gamma_{k})=0. It follows that there exists an integer bb\in\mathbb{Z} with vp(b)=0v_{p}(b)=0, such that vp(γkb)>0v_{p}(\gamma_{k}-b)>0. We define βγb1.\beta\coloneqq\gamma-b\in\mathcal{R}_{1}. On the one hand, we have by construction vp(βk)>0v_{p}(\beta_{k})>0, and on the other, vp(βn)=0v_{p}(\beta_{n})=0 for n>kn>k by the ultrametric property of the pp-adic valuation. ∎

Corollary 5.4.9.

For all n>k0n>k\geq 0, every 11-semiadditive functor F:SpT(n)SpT(k)F\colon\operatorname{Sp}_{T(n)}\to\operatorname{Sp}_{T(k)} is zero.

Proof.

By 5.4.8, there exists an element a1a\in\mathcal{R}_{1}, such that SpT(n)\operatorname{Sp}_{T(n)} is aa-divisible and SpT(k)\operatorname{Sp}_{T(k)} is aa-complete. Since FF is 11-semiadditive, by 5.4.3, it must preserve the action of aa. It follows that for every XSpT(n)X\in\operatorname{Sp}_{T(n)}, the object F(X)F(X) is both aa-divisible and aa-complete and hence must be zero. ∎

We are now ready to prove that 11-semiadditivity forces LnfSpL_{n}^{f}\operatorname{Sp} to decompose into its monochromatic pieces.

Theorem 5.4.10 (1-Semiadditive Decomposition).

For every n0n\geq 0, there is an equivalence of modes

CMon1LnfSpk=0nSpT(k).\operatorname{CMon}_{1}\otimes L_{n}^{f}\operatorname{Sp}\simeq\prod_{k=0}^{n}\operatorname{Sp}_{T(k)}.
Proof.

For n=0n=0, the claim is CMon1SpSp,\operatorname{CMon}_{1}\otimes\operatorname{Sp}_{\mathbb{Q}}\simeq\operatorname{Sp}_{\mathbb{Q}}, which is true because Sp\operatorname{Sp}_{\mathbb{Q}} is 11-semiadditive. It thus suffices to prove by induction on nn\in\mathbb{N}, that the functor

LnfSpLn1fSp×SpT(n),L_{n}^{f}\operatorname{Sp}\to L_{n-1}^{f}\operatorname{Sp}\times\operatorname{Sp}_{T(n)},

given by the product of the respective symmetric monoidal localizations, becomes an equivalence after tensoring with CMon1\operatorname{CMon}_{1}. We know that LnfSpL_{n}^{f}\operatorname{Sp} is a recollement of Ln1fSpLnfSpL_{n-1}^{f}\operatorname{Sp}\subseteq L_{n}^{f}\operatorname{Sp} and SpT(n)\operatorname{Sp}_{T(n)}. It follows that CMon1LnfSp\operatorname{CMon}_{1}\otimes L_{n}^{f}\operatorname{Sp} is a recollement of CMon1Ln1fSp\operatorname{CMon}_{1}\otimes L_{n-1}^{f}\operatorname{Sp} and CMon1SpT(n)=SpT(n)\operatorname{CMon}_{1}\otimes\operatorname{Sp}_{T(n)}=\operatorname{Sp}_{T(n)} (5.2.3). By the inductive hypothesis, we have

CMon1Ln1fSpk=0n1SpT(k).\operatorname{CMon}_{1}\otimes L_{n-1}^{f}\operatorname{Sp}\simeq\prod_{k=0}^{n-1}\operatorname{Sp}_{T(k)}.

Thus, CMon1LnfSp\operatorname{CMon}_{1}\otimes L_{n}^{f}\operatorname{Sp} is a recollement of k=0n1SpT(k)\prod_{k=0}^{n-1}\operatorname{Sp}_{T(k)} and SpT(n)\operatorname{Sp}_{T(n)}. It is therefore suffices to show that the gluing data given by the functor

L:SpT(n)k=0n1SpT(k)L\colon\operatorname{Sp}_{T(n)}\to\prod_{k=0}^{n-1}\operatorname{Sp}_{T(k)}

is zero (4.1.4). We observe that LL is given as a composition of a right and a left adjoint between 11-semiadditive \infty-categories and hence is 11-semiadditive. Thus, by 5.4.9, we must have L=0L=0 and hence the corresponding recollement is split. To conclude, the localization functors LnfSpSpT(k)L_{n}^{f}\operatorname{Sp}\to\operatorname{Sp}_{T(k)} for k=0,,nk=0,\dots,n induce a functor LnfSpk=0nSpT(k)L_{n}^{f}\operatorname{Sp}\to\prod_{k=0}^{n}\operatorname{Sp}_{T(k)}, which becomes an equivalence after tensoring with CMon1\operatorname{CMon}_{1}. In particular, this is a symmetric monoidal equivalence and hence an equivalence of modes. ∎

Remark 5.4.11.

By tensoring 5.4.10 with LnSpL_{n}\operatorname{Sp}, we also get

CMon1LnSpk=0nSpK(k).\operatorname{CMon}_{1}\otimes L_{n}\operatorname{Sp}\simeq\prod_{k=0}^{n}\operatorname{Sp}_{K(k)}.

5.5 Semiadditive vs. Stable Height

As we recalled in the introduction, for every n0n\geq 0, the localization functors of spectra LnfL_{n}^{f}, LF(n)L_{F(n)} and LT(n)L_{T(n)} can be thought of as restriction to heights n\leq n, n\geq n and nn respectively, as measured by the vnv_{n}-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 LnfL_{n}^{f}, LF(n)L_{F(n)} and LT(n)L_{T(n)} (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 11-semiadditivity vs. \infty-semiadditivity for stable presentable \infty-categories.

Stable Height

By 5.2.13, the \infty-categories LnfSpL_{n}^{f}\operatorname{Sp}, SpF(n)\operatorname{Sp}_{F(n)} and SpT(n)\operatorname{Sp}_{T(n)} 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 \infty-category 𝒞\mathcal{C}, for every X𝒞X\in\mathcal{C}, we define and denote the stable height of XX as follows:

  1. (1)

    htst(X)n\mathrm{ht}_{\mathrm{st}}(X)\leq n, if F(n+1)X=0F(n+1)\otimes X=0 for some (hence any) finite spectrum F(n+1)F(n+1) of type n+1n+1.

  2. (2)

    htst(X)>n\mathrm{ht}_{\mathrm{st}}(X)>n, if Map(Z,X)pt\operatorname{Map}(Z,X)\simeq\operatorname{pt} for every ZZ of stable height n\leq n.

  3. (3)

    htst(X)=n\mathrm{ht}_{\mathrm{st}}(X)=n, if htst(X)n\mathrm{ht}_{\mathrm{st}}(X)\leq n and htst(X)>n1\mathrm{ht}_{\mathrm{st}}(X)>n-1.

By convention, htst(X)>1\mathrm{ht}_{\mathrm{st}}(X)>-1 for all XX, and htst(X)1\mathrm{ht}_{\mathrm{st}}(X)\leq-1 if and only if X=0X=0. We also extend the definition to n=n=\infty as follows: For every X𝒞X\in\mathcal{C}, we write htst(X)=\mathrm{ht}_{\mathrm{st}}(X)=\infty if and only if htst(X)>n\mathrm{ht}_{\mathrm{st}}(X)>n for all nn.

Remark 5.5.2.

Since F(n+2)F(n+2) can be constructed as a cofiber of a self map of F(n+1)F(n+1), it is clear that htst(X)n\mathrm{ht}_{\mathrm{st}}(X)\leq n implies htst(X)n+1\mathrm{ht}_{\mathrm{st}}(X)\leq n+1. Consequently, htst(X)>n\mathrm{ht}_{\mathrm{st}}(X)>n also implies htst(X)>n1\mathrm{ht}_{\mathrm{st}}(X)>n-1.

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 𝒞Catst\mathcal{C}\in\mathrm{Cat}_{\mathrm{st}} and let 0n0\leq n\leq\infty. We denote by 𝒞stn\mathcal{C}_{\leq^{\mathrm{st}}n}, 𝒞>stn\mathcal{C}_{>^{\mathrm{st}}n}, and 𝒞nst\mathcal{C}_{n^{\mathrm{st}}} the full subcategories of 𝒞\mathcal{C} spanned by objects of stable height n\leq n, >n>n, and nn. In addition, we write Htst(𝒞)n\mathrm{Ht}_{\mathrm{st}}(\mathcal{C})\leq n, >n>n, or nn, if 𝒞=𝒞stn\mathcal{C}=\mathcal{C}_{\leq^{\mathrm{st}}n}, 𝒞>stn\mathcal{C}_{>^{\mathrm{st}}n}, or 𝒞nst\mathcal{C}_{n^{\mathrm{st}}} respectively.

Example 5.5.4.

In the special case 𝒞=Sp(p),\mathcal{C}=\operatorname{Sp}_{(p)}, we have by definition

Sp(p),stn=LnfSp,Sp(p),>stn1=SpF(n),Sp(p),nst=SpT(n).\operatorname{Sp}_{(p),\leq^{\mathrm{st}}n}=L_{n}^{f}\operatorname{Sp},\qquad\operatorname{Sp}_{(p),>^{\mathrm{st}}n-1}=\operatorname{Sp}_{F(n)},\qquad\operatorname{Sp}_{(p),n^{\mathrm{st}}}=\operatorname{Sp}_{T(n)}.

We note that the subcategory Sp(p),stSp\operatorname{Sp}_{(p),\infty^{\mathrm{st}}}\subseteq\operatorname{Sp} is rather large. First, for every XLnfSpX\in L_{n}^{f}\operatorname{Sp}, we have

LnfX𝔽pLnfXLnf𝔽pLnfX00.L_{n}^{f}X\otimes\mathbb{F}_{p}\simeq L_{n}^{f}X\otimes L_{n}^{f}\mathbb{F}_{p}\simeq L_{n}^{f}X\otimes 0\simeq 0.

Therefore, for every 𝔽p\mathbb{F}_{p}-module spectrum MM, we have

MapSp(LnfX,M)MapMod𝔽p(Sp)(𝔽pLnfX,M)MapMod𝔽p(Sp)(0,M)0,\operatorname{Map}_{\operatorname{Sp}}(L_{n}^{f}X,M)\simeq\operatorname{Map}_{\operatorname{Mod}_{\mathbb{F}_{p}}(\operatorname{Sp})}(\mathbb{F}_{p}\otimes L_{n}^{f}X,M)\simeq\operatorname{Map}_{\operatorname{Mod}_{\mathbb{F}_{p}}(\operatorname{Sp})}(0,M)\simeq 0,

which implies that MSp(p),stM\in\operatorname{Sp}_{(p),\infty^{\mathrm{st}}}. Since the subcategory Sp(p),stSp\operatorname{Sp}_{(p),\infty^{\mathrm{st}}}\subseteq\operatorname{Sp} is closed under limits, it contains all bounded below pp-complete spectra. In contrast, the \infty-category 𝒞=nSpT(n)\mathcal{C}=\prod_{n\in\mathbb{N}}\operatorname{Sp}_{T(n)} satisfies 𝒞st=0\mathcal{C}_{\infty^{\mathrm{st}}}=0, even though it contains many objects of unbounded height.

We now show that the modes LnfSpL_{n}^{f}\operatorname{Sp}, SpF(n)\operatorname{Sp}_{F(n)} and SpT(n)\operatorname{Sp}_{T(n)} classify the properties of having the corresponding stable height.

Proposition 5.5.5.

For every 0n<0\leq n<\infty, the modes LnfSpL_{n}^{f}\operatorname{Sp}, SpF(n)\operatorname{Sp}_{F(n)} and SpT(n)\operatorname{Sp}_{T(n)} classify the properties of being stable pp-local of stable height n\leq n, >n1>n-1 and nn respectively.

Proof.

We begin with LnfSpL_{n}^{f}\operatorname{Sp}. It follows from 5.2.10, that LnfSpL_{n}^{f}\operatorname{Sp} is a mode, which classifies stable LnfSpL_{n}^{f}\operatorname{Sp}-local \infty-categories. Thus, it suffice to show that an object XX in a stable \infty-category 𝒞\mathcal{C} is LnfSpL_{n}^{f}\operatorname{Sp}-local, if and only if htst(X)n\mathrm{ht}_{\mathrm{st}}(X)\leq n. By definition, XX is LnfSpL_{n}^{f}\operatorname{Sp}-local, if for every ZSpZ\in\operatorname{Sp}, the mapping spectrum hom𝒞Sp(Z,X)\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,X) belongs to LnfSpL_{n}^{f}\operatorname{Sp}. Since the corepresentable functor hom𝒞Sp(Z,):𝒞Sp\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,-)\colon\mathcal{C}\to\operatorname{Sp} is exact, we have a canonical isomorphism

hom𝒞Sp(Z,F(n+1)X)F(n+1)hom𝒞Sp(Z,X).\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,F(n+1)\otimes X)\simeq F(n+1)\otimes\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,X).

It follows that if htst(X)n\mathrm{ht}_{\mathrm{st}}(X)\leq n, then XX is LnfSpL_{n}^{f}\operatorname{Sp}-local. Since the collection of functors hom𝒞Sp(Z,)\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,-) for all Z𝒞Z\in\mathcal{C} is also jointly conservative, the converse holds as well.

We now move to SpF(n)\operatorname{Sp}_{F(n)}. We first show that if htst(X)>n1\mathrm{ht}_{\mathrm{st}}(X)>n-1, then hom𝒞Sp(Z,X)\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,X) is F(n)F(n)-local for all Z𝒞Z\in\mathcal{C}. For AA an F(n)F(n)-acyclic spectrum, AZA\otimes Z is F(n)F(n)-acyclic as well and hence htst(AZ)n1\mathrm{ht}_{\mathrm{st}}(A\otimes Z)\leq n-1. Since htst(X)>n1\mathrm{ht}_{\mathrm{st}}(X)>n-1, it follows that

MapSp(A,hom𝒞Sp(Z,X))Map𝒞(AZ,X)pt,\operatorname{Map}_{\operatorname{Sp}}(A,\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,X))\simeq\operatorname{Map}_{\mathcal{C}}(A\otimes Z,X)\simeq\operatorname{pt},

and hence XX is SpF(n)\operatorname{Sp}_{F(n)}-local. Conversely, assume that XX is SpF(n)\operatorname{Sp}_{F(n)}-local, and let Z𝒞Z\in\mathcal{C} such that htst(Z)n1\mathrm{ht}_{\mathrm{st}}(Z)\leq n-1. We have

0hom𝒞Sp(F(n)Z,X)𝔻F(n)hom𝒞Sp(Z,X),0\simeq\hom_{\mathcal{C}}^{\operatorname{Sp}}(F(n)\otimes Z,X)\simeq\mathbb{D}F(n)\otimes\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,X),

where 𝔻F(n)\mathbb{D}F(n) is the Spanier-Whithead dual of F(n)F(n), which is itself of type nn. Since hom𝒞Sp(Z,X)\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,X) is F(n)F(n)-local and hence 𝔻F(n)\mathbb{D}F(n)-local, we get hom𝒞Sp(Z,X)=0\hom_{\mathcal{C}}^{\operatorname{Sp}}(Z,X)=0. This implies that Map𝒞(Z,X)=pt\operatorname{Map}_{\mathcal{C}}(Z,X)=\operatorname{pt} and so htst(X)>n1\mathrm{ht}_{\mathrm{st}}(X)>n-1.

Finally, for SpT(n)\operatorname{Sp}_{T(n)} we observe that a spectrum is T(n)T(n)-local if and only if it is both F(n)F(n)-local and k=0nT(k)\bigoplus_{k=0}^{n}T(k) -local. Thus,

SpT(n)LnfSpSpF(n).\operatorname{Sp}_{T(n)}\simeq L_{n}^{f}\operatorname{Sp}\otimes\operatorname{Sp}_{F(n)}.

Hence, it classifies the property of being stable pp-local of both stable height n\leq n and >n1>n-1, i.e exactly nn. ∎

Similarly, we also treat the case of stable height \infty.

Corollary 5.5.6.

The \infty-category Sp(p),st\operatorname{Sp}_{(p),\infty^{\mathrm{st}}} is a mode, which classifies the property of being stable of stable height \infty.

Proof.

We have

Sp(p),st=nSpF(n)Sp(p),\operatorname{Sp}_{(p),\infty^{\mathrm{st}}}=\bigcap_{n\in\mathbb{N}}\operatorname{Sp}_{F(n)}\subseteq\operatorname{Sp}_{(p)},

and thus is an accessible reflective subcategory of Sp(p)\operatorname{Sp}_{(p)}, which is compatible with the symmetric monoidal structure [Lur09, Proposition .5.4.7.10]. It follows from 5.2.10, that Sp(p),st\operatorname{Sp}_{(p),\infty^{\mathrm{st}}} is a mode. Moreover, Sp(p),st\operatorname{Sp}_{(p),\infty^{\mathrm{st}}} classifies the property that the \infty-category is stable such that every object XX is SpF(n)\operatorname{Sp}_{F(n)}-local for all nn\in\mathbb{N}. By 5.5.5, the said condition on XX is equivalent to htst(X)n\mathrm{ht}_{\mathrm{st}}(X)\geq n for all nn, which is the same as htst(X)=\mathrm{ht}_{\mathrm{st}}(X)=\infty. Thus, Sp(p),st\operatorname{Sp}_{(p),\infty^{\mathrm{st}}} classifies the property of being stable of stable height \infty. ∎

5.5.5 has the following immediate corollaries regarding stable height for stable presentable \infty-categories:

Corollary 5.5.7.

For every pp-local 𝒞Prst\mathcal{C}\in\Pr_{\mathrm{st}}, we have canonical equivalences

𝒞stn𝒞LnfSp,𝒞>stn𝒞SpF(n+1),𝒞nst𝒞SpT(n)\mathcal{C}_{\leq^{\mathrm{st}}n}\simeq\mathcal{C}\otimes L_{n}^{f}\operatorname{Sp},\qquad\mathcal{C}_{>^{\mathrm{st}}n}\simeq\mathcal{C}\otimes\operatorname{Sp}_{F(n+1)},\qquad\mathcal{C}_{n^{\mathrm{st}}}\simeq\mathcal{C}\otimes\operatorname{Sp}_{T(n)}

for every nn\in\mathbb{N}.

Proof.

This is a direct consequence of 5.5.5 and 5.2.10. ∎

Corollary 5.5.8.

Every pp-local 𝒞Prst{\cal C}\in\Pr_{\mathrm{st}} is a recollement of 𝒞stn{\cal C}_{\leq^{\mathrm{st}}n} and 𝒞>stn{\cal C}_{>^{\mathrm{st}}n} for every nn\in\mathbb{N}.

Proof.

Since Sp\operatorname{Sp} is a recollement of LnfSpL_{n}^{f}\operatorname{Sp} and (LnfSp)=SpF(n+1)(L_{n}^{f}\operatorname{Sp})^{\perp}=\operatorname{Sp}_{F(n+1)}, we have by 5.2.3, that 𝒞𝒞Sp\mathcal{C}\simeq{\cal C}\otimes\operatorname{Sp} is a recollement of 𝒞LnfSp{\cal C}\otimes L_{n}^{f}\operatorname{Sp} and 𝒞SpF(n+1)\mathcal{C}\otimes\operatorname{Sp}_{F(n+1)}. By 5.5.7, we get that 𝒞\mathcal{C} is a recollement of 𝒞stn{\cal C}_{\leq^{\mathrm{st}}n} and 𝒞>stn{\cal C}_{>^{\mathrm{st}}n}. ∎

Let 𝒞\mathcal{C} be a stable presentable \infty-category. For each n,n\in\mathbb{N}, let 𝒞Rn𝒞stn\mathcal{C}\xrightarrow{R_{\leq n}}\mathcal{C}_{\leq^{\mathrm{st}}n} be the right adjoint of the inclusion 𝒞stn𝒞\mathcal{C}_{\leq^{\mathrm{st}}n}\hookrightarrow\mathcal{C} and 𝒞stnLn𝒞nst\mathcal{C}_{\leq^{\mathrm{st}}n}\xrightarrow{L_{n}}\mathcal{C}_{n^{\mathrm{st}}} the left adjoint of the inclusion 𝒞nst𝒞stn\mathcal{C}_{n^{\mathrm{st}}}\hookrightarrow\mathcal{C}_{\leq^{\mathrm{st}}n}. We define

Pn:𝒞Rn𝒞stnLn𝒞nst.P_{n}\colon\mathcal{C}\xrightarrow{R_{\leq n}}\mathcal{C}_{\leq^{\mathrm{st}}n}\xrightarrow{L_{n}}\mathcal{C}_{n^{\mathrm{st}}}.
Proposition 5.5.9.

For every pp-local 𝒞Prst\mathcal{C}\in\Pr_{\mathrm{st}} and X𝒞X\in\mathcal{C}. If Pn(X)=0P_{n}(X)=0 for all nn\in\mathbb{N}, then X𝒞stX\in\mathcal{C}_{\infty^{\mathrm{st}}}. In particular, if 𝒞st=0\mathcal{C}_{\infty^{\mathrm{st}}}=0, the collection of functors Pn:𝒞𝒞nstP_{n}\colon\mathcal{C}\to\mathcal{C}_{n^{\mathrm{st}}} is jointly conservative.

Proof.

Let X𝒞X\in\mathcal{C}, such that Pn(X)=0P_{n}(X)=0 for all nn. We shall show by induction that X𝒞>stnX\in\mathcal{C}_{>^{\mathrm{st}}n} for all nn\in\mathbb{N}, and hence X𝒞stX\in\mathcal{C}_{\infty^{\mathrm{st}}}. Assuming by induction that X𝒞>stn1X\in\mathcal{C}_{>^{\mathrm{st}}n-1} we have

Rn1Rn(X)=Rn1(X)=0.R_{\leq n-1}R_{\leq n}(X)=R_{\leq n-1}(X)=0.

Therefore Rn(X)𝒞>stn1R_{\leq n}(X)\in\mathcal{C}_{>^{\mathrm{st}}n-1} and hence

Rn(X)𝒞nst𝒞>stn1=𝒞nst.R_{\leq n}(X)\in\mathcal{C}_{\leq{}^{\mathrm{st}}n}\cap\mathcal{C}_{>^{\mathrm{st}}n-1}=\mathcal{C}_{n^{\mathrm{st}}}.

It follows that

Pn(X)=LnRn(X)=Rn(X).P_{n}(X)=L_{n}R_{\leq n}(X)=R_{\leq n}(X).

Hence, for all Z𝒞stnZ\in\mathcal{C}_{\leq^{\mathrm{st}}n} we get

Map(Z,X)=Map(Z,RnX)=Map(Z,PnX)=Map(Z,0)=pt\operatorname{Map}(Z,X)=\operatorname{Map}(Z,R_{\leq n}X)=\operatorname{Map}(Z,P_{n}X)=\operatorname{Map}(Z,0)=\operatorname{pt}

and so X𝒞>stnX\in\mathcal{C}_{>^{\mathrm{st}}n}. We can take the base of the induction to be n=1n=-1, in which there is nothing to prove. ∎

When further assuming 11-semiadditivity, we get the following:

Proposition 5.5.10.

For every pp-local 𝒞Prst

-
1
\mathcal{C}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}1}
, we have a canonical equivalence:

𝒞stn𝒞0st×𝒞1st××𝒞nst.\mathcal{C}_{\leq^{\mathrm{st}}n}\simeq\mathcal{C}_{0^{\mathrm{st}}}\times\mathcal{C}_{1^{\mathrm{st}}}\times\cdots\times\mathcal{C}_{n^{\mathrm{st}}}.
Proof.

By tensoring the equivalence of 5.4.10 with 𝒞\mathcal{C}, we get an equivalence

𝒞CMon1LnfSp𝒞k=0nSpT(n)k=0n(𝒞SpT(n)).\mathcal{C}\otimes\operatorname{CMon}_{1}\otimes L_{n}^{f}\operatorname{Sp}\simeq\mathcal{C}\otimes\prod_{k=0}^{n}\operatorname{Sp}_{T(n)}\simeq\prod_{k=0}^{n}(\mathcal{C}\otimes\operatorname{Sp}_{T(n)}).

Since 𝒞\mathcal{C} is already 11-semiadditive, 𝒞CMon1𝒞\mathcal{C}\otimes\operatorname{CMon}_{1}\simeq\mathcal{C}. Thus, the claim follows from 5.5.7. ∎

Comparing heights

For a stable higher semiadditive \infty-category, it is natural to compare the stable height with the semiadditive height. First,

Lemma 5.5.11.

Let 𝒞Prst

-
\mathcal{C}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
be pp-local. For all n,k{}n,k\in\mathbb{N}\cup\{\infty\}, we have (𝒞nst)k(𝒞k)nst(\mathcal{C}_{n^{\mathrm{st}}})_{k}\simeq(\mathcal{C}_{k})_{n^{\mathrm{st}}}.

Proof.

Using 5.3.6 and 5.5.5 (or 5.5.6 for n=n=\infty) we get

(𝒞nst)k𝒞Sp(p),nstk(𝒞k)nst.(\mathcal{C}_{n^{\mathrm{st}}})_{k}\simeq\mathcal{C}\otimes\operatorname{Sp}_{(p),n^{\mathrm{st}}}\otimes{}_{k}\simeq(\mathcal{C}_{k})_{n^{\mathrm{st}}}.

We next observe that for nn\in\mathbb{N} the \infty-category Sp(p),nst=SpT(n)\operatorname{Sp}_{(p),n^{\mathrm{st}}}=\operatorname{Sp}_{T(n)}, which is the mode of stable height nn, is \infty-semiadditive of semiadditive height nn (4.4.5). Therefore, there is a map of modes nSpT(n){}_{n}\to\operatorname{Sp}_{T(n)}, making SpT(n)\operatorname{Sp}_{T(n)} an algebra over n.

Proposition 5.5.12.

Let 𝒞Prst

-
\mathcal{C}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
be pp-local. For all nn\in\mathbb{N} and k{}k\in\mathbb{N}\cup\{\infty\}, we have

(𝒞nst)k(𝒞k)nst{𝒞nstk=n0kn(\mathcal{C}_{n^{\mathrm{st}}})_{k}\simeq(\mathcal{C}_{k})_{n^{\mathrm{st}}}\simeq\begin{cases}\mathcal{C}_{n^{\mathrm{st}}}&k=n\\ 0&k\neq n\end{cases}
Proof.

On the one hand, for knk\neq n we have n=k0{}_{n}\otimes{}_{k}=0 (5.3.8) and so SpT(n)=k0\operatorname{Sp}_{T(n)}\otimes{}_{k}=0. On the other, SpT(n)nSpT(n)\operatorname{Sp}_{T(n)}\otimes{}_{n}\simeq\operatorname{Sp}_{T(n)} as SpT(n)\operatorname{Sp}_{T(n)} is a n-module. Tensoring these with 𝒞\mathcal{C} yields the claim. ∎

Given 𝒞Prst

-
,
\mathcal{C}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}\infty},
tensoring the map of modes nSpT(n){}_{n}\to\operatorname{Sp}_{T(n)} with 𝒞\mathcal{C}, yields a map 𝒞n𝒞nst\mathcal{C}_{n}\to\mathcal{C}_{n^{\mathrm{st}}}.

Proposition 5.5.13.

Let 𝒞Prst

-
\mathcal{C}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}\infty}
be pp-local. For every nn\in\mathbb{N}, the map 𝒞n𝒞nst\mathcal{C}_{n}\to\mathcal{C}_{n^{\mathrm{st}}} admits a fully faithful right adjoint 𝒞nst𝒞n\mathcal{C}_{n^{\mathrm{st}}}\hookrightarrow\mathcal{C}_{n}, which exhibits 𝒞n\mathcal{C}_{n} as a recollement of 𝒞nst\mathcal{C}_{n^{\mathrm{st}}} and (𝒞n)st(\mathcal{C}_{n})_{\infty^{\mathrm{st}}} (=(𝒞st)n=(\mathcal{C}_{\infty^{\mathrm{st}}})_{n}). In particular, if 𝒞st=0\mathcal{C}_{\infty^{\mathrm{st}}}=0, then 𝒞n𝒞nst\mathcal{C}_{n}\simeq\mathcal{C}_{n^{\mathrm{st}}}.

Proof.

By 5.5.8, for every NnN\geq n, the \infty-category 𝒞n\mathcal{C}_{n} is a recollement of (𝒞n)stN(\mathcal{C}_{n})_{\leq^{\mathrm{st}}N} and (𝒞n)>stN(\mathcal{C}_{n})_{>^{\mathrm{st}}N}. Applying 5.5.10 and 5.5.12 to 𝒞n\mathcal{C}_{n} we obtain

(𝒞n)stN(𝒞n)0st×(𝒞n)1st××(𝒞n)Nst𝒞nst.(\mathcal{C}_{n})_{\leq^{\mathrm{st}}N}\simeq(\mathcal{C}_{n})_{0^{\mathrm{st}}}\times(\mathcal{C}_{n})_{1^{\mathrm{st}}}\times\cdots\times(\mathcal{C}_{n})_{N^{\mathrm{st}}}\simeq\mathcal{C}_{n^{\mathrm{st}}}.

Consequently,

(𝒞n)>stN=(𝒞n)>stN+1=(𝒞n)>stN+2==(𝒞n)st(\mathcal{C}_{n})_{>^{\mathrm{st}}N}=(\mathcal{C}_{n})_{>^{\mathrm{st}}N+1}=(\mathcal{C}_{n})_{>^{\mathrm{st}}N+2}=\dots=(\mathcal{C}_{n})_{\infty^{\mathrm{st}}}

and hence, 𝒞n\mathcal{C}_{n} as a recollement of 𝒞nst=(𝒞n)stN\mathcal{C}_{n^{\mathrm{st}}}=(\mathcal{C}_{n})_{\leq^{\mathrm{st}}N} and (𝒞n)st=(𝒞n)>stN(\mathcal{C}_{n})_{\infty^{\mathrm{st}}}=(\mathcal{C}_{n})_{>^{\mathrm{st}}N}. ∎

As a consequence, we get a tight connection between SpT(n)\operatorname{Sp}_{T(n)} and n.

Corollary 5.5.14.

The map of modes nSpT(n){}_{n}\to\operatorname{Sp}_{T(n)} is a smashing localization.

Proof.

This follows from 5.2.15 and 5.5.13. ∎

Bootstrap of Semiadditivity

Based on the classification of higher semiadditive localizations of Sp\operatorname{Sp} with respect to homotopy rings, the authors proposed in [CSY18, Conjecture 1.1.5] the “Bootstrap Conjecture”, stating that if a presentable stable pp-local \infty-category is 11-semiadditive, then it is automatically \infty-semiadditive. Using the 11-semiadditive decomposition of 5.4.10 we now provide some partial results in the direction of proving this conjecture. First, given a pp-local 𝒞Prst\mathcal{C}\in\Pr_{\mathrm{st}} and nn\in\mathbb{N}, the \infty-category 𝒞nst\mathcal{C}_{n^{\mathrm{st}}} is a module over SpT(n)\operatorname{Sp}_{T(n)} and hence over n. It follows that 𝒞nst\mathcal{C}_{n^{\mathrm{st}}} is \infty-semiadditive and of height nn. More generally, if 𝒞\mathcal{C} is 11-semiadditive, then by 5.5.10, we have 𝒞stnk=0n𝒞kst\mathcal{C}_{\leq^{\mathrm{st}}n}\simeq\prod_{k=0}^{n}\mathcal{C}_{k^{\mathrm{st}}}. From this one can deduce that, if 𝒞\mathcal{C} is 11-semiadditive and every object of 𝒞\mathcal{C} is of bounded stable height, then 𝒞\mathcal{C} is \infty-semiadditive. However, we shall show that having such a bound on the stable height of the objects of 𝒞\mathcal{C} is an unnecessarily strong restriction, and it in fact suffices to assume merely that 𝒞st=0\mathcal{C}_{\infty^{\mathrm{st}}}=0.

Since every stable presentable \infty-category of stable height exactly nn (for some nn) is \infty-semiadditive, it has an action of π0[1]\pi_{0}{}^{[1]}, and a fortiori of the subring 1π0[1]\mathcal{R}_{1}\subseteq\pi_{0}{}^{[1]} (see 5.4.2). We begin with a generalization of 5.4.8.

Proposition 5.5.15.

For every nn, there exists a1a\in\mathcal{R}_{1}, such that every 𝒞Prst

-
1
\mathcal{C}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}1}
of stable height kk is aa-complete if knk\leq n and aa-divisible if k>nk>n.

Proof.

By 5.4.8, there exists a1a\in\mathcal{R}_{1}, such that SpT(k)\operatorname{Sp}_{T(k)} is aa-complete if knk\leq n and aa-divisible if k>nk>n. Since 𝒞\mathcal{C} is an SpT(k)\operatorname{Sp}_{T(k)}-module, the action of a1a\in\mathcal{R}_{1} on 𝒞\mathcal{C} is via its image in π0𝕊T(k)\pi_{0}\mathbb{S}_{T(k)}. Thus, it suffices to show that for every aπ0𝕊T(k)a\in\pi_{0}\mathbb{S}_{T(k)}, if SpT(k)\operatorname{Sp}_{T(k)} is aa-complete (resp. aa-divisible), then 𝒞\mathcal{C} is also aa-complete (resp. aa-divisible). We observe that SpT(k)\operatorname{Sp}_{T(k)} is aa-divisible if and only if aa is invertible, in which case 𝒞\mathcal{C} is aa-divisible as well. On the other hand, if SpT(k)\operatorname{Sp}_{T(k)} is aa-complete, then tensoring with A𝕊T(k)/aA\coloneqq\mathbb{S}_{T(k)}/a is conservative on SpT(k)\operatorname{Sp}_{T(k)}. The object AA is dualizable with dual AΣ1AA^{\vee}\simeq\Sigma^{-1}A and so tensoring with AA^{\vee} is also conservative. Finally, to show that 𝒞\mathcal{C} is aa-complete, it suffices to show that tensoring with AA is conservative on 𝒞\mathcal{C}, via the left tensoring of 𝒞\mathcal{C} over SpT(k)\operatorname{Sp}_{T(k)}. For every X,Y𝒞X,Y\in\mathcal{C}, we have isomorphisms

hom𝒞SpT(k)(AX,Y)homSpT(k)SpT(k)(A,hom(X,Y)𝒞SpT(k))Ahom𝒞SpT(k)(X,Y).\hom_{\mathcal{C}}^{\operatorname{Sp}_{T(k)}}(A\otimes X,Y)\simeq\hom_{\operatorname{Sp}_{T(k)}}^{\operatorname{Sp}_{T(k)}}(A,\hom{}_{\mathcal{C}}^{\operatorname{Sp}_{T(k)}}(X,Y))\simeq A^{\vee}\otimes\hom_{\mathcal{C}}^{\operatorname{Sp}_{T(k)}}(X,Y).

Hence, if AX0A\otimes X\simeq 0, then by the conservativity of A()A^{\vee}\otimes(-), we get that hom𝒞SpT(k)(X,Y)=0\hom_{\mathcal{C}}^{\operatorname{Sp}_{T(k)}}(X,Y)=0 for all YY and hence X=0X=0. ∎

From this we derive a strengthening of 5.4.9.

Proposition 5.5.16.

Let 𝒞,𝒟Prst

-
1
{\cal C},{\cal D}\in\Pr_{\mathrm{st}}^{\scalebox{0.6}{$\oplus$}\text{-}1}
be pp-local. If Htst(𝒟)n\mathrm{Ht}_{\mathrm{st}}({\cal D})\leq n, Htst(𝒞)>n\mathrm{Ht}_{\mathrm{st}}({\cal C})>n and 𝒞st=0\mathcal{C}_{\infty^{\mathrm{st}}}=0, then every 1-semidditive functor F:𝒞𝒟F\colon{\cal C}\to{\cal D} is zero.

Proof.

We shall construct an element a1a\in\mathcal{R}_{1} such that 𝒞\mathcal{C} is aa-divisible and 𝒟\mathcal{D} is aa-complete. Since FF is 11-semiadditive, we shall get that it takes every X𝒞X\in{\cal C} to an object of 𝒟{\cal D} which is both aa-complete and aa-divisible (5.4.3). This will imply that F(X)=0F(X)=0 for all X𝒞X\in\mathcal{C} and hence that FF is zero. By 5.5.15, there is an a1a\in{\cal R}_{1} such that 𝒞k\mathcal{C}_{k} is aa-divisible for k>nk>n and 𝒟k\mathcal{D}_{k} is aa-complete for knk\leq n. By 5.5.10, we have

𝒟=𝒟stn0kn𝒟kst\mathcal{D}=\mathcal{D}_{\leq^{\mathrm{st}}n}\simeq\prod_{0\leq k\leq n}\mathcal{D}_{k^{\mathrm{st}}}

and hence 𝒟\mathcal{D} itself is aa-complete. As for 𝒞\mathcal{C}, by 5.5.9, we have a jointly conservative collection of functors Pn:𝒞𝒞nstP_{n}\colon\mathcal{C}\to\mathcal{C}_{n^{\mathrm{st}}} for nn\in\mathbb{N}. Moreover, all the PnP_{n}-s are 11-semiadditive, as a composition of a left and a right adjoint, and thus 𝒞\mathcal{C} is aa-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 𝒞Prst{\cal C}\in\Pr_{\mathrm{st}} be pp-local, such that 𝒞st=0\mathcal{C}_{\infty^{\mathrm{st}}}=0. If 𝒞\mathcal{C} is 11-semiadditive, then 𝒞{\cal C} is \infty-semiadditive. Moreover, in this case 𝒞nst=𝒞n\mathcal{C}_{n^{\mathrm{st}}}=\mathcal{C}_{n} for all 0n0\leq n\leq\infty and there is a canonical decomposition

𝒞n𝒞nstn𝒞n.\mathcal{C}\simeq\prod_{n\in\mathbb{N}}\mathcal{C}_{n^{\mathrm{st}}}\simeq\prod_{n\in\mathbb{N}}\mathcal{C}_{n}.
Proof.

By 5.5.8, 𝒞{\cal C} is a recollement of 𝒞stn{\cal C}_{\leq^{\mathrm{st}}n} and 𝒞>stn{\cal C}_{>^{\mathrm{st}}n}. Since 𝒞>stn{\cal C}_{>^{\mathrm{st}}n} is 11-semiadditive and

(𝒞>stn)st𝒞st=0,({\cal C}_{>^{\mathrm{st}}n})_{\infty^{\mathrm{st}}}\simeq{\cal C}_{\infty^{\mathrm{st}}}=0,

every 1-semiadditive functor F:𝒞>stn𝒞stnF\colon{\cal C}_{>^{\mathrm{st}}n}\to{\cal C}_{\leq^{\mathrm{st}}n} is zero by 5.5.16. It follows that 𝒞{\cal C} is a split recollement of 𝒞stn{\cal C}_{\leq^{\mathrm{st}}n} and 𝒞>stn{\cal C}_{>^{\mathrm{st}}n}, and hence a recollement of 𝒞>stn{\cal C}_{>^{\mathrm{st}}n} and 𝒞stn{\cal C}_{\leq^{\mathrm{st}}n} (i.e. we may switch the roles). By 4.1.10, we get

𝒞limn(𝒞stn)limn(kn𝒞kst)n𝒞nst.{\cal C}\simeq\underleftarrow{\operatorname{lim}\,}_{n\in\mathbb{N}}({\cal C}_{\leq^{\mathrm{st}}n})\simeq\underleftarrow{\operatorname{lim}\,}_{n\in\mathbb{N}}(\prod_{k\leq n}{\cal C}_{k^{\mathrm{st}}})\simeq\prod_{n\in\mathbb{N}}{\cal C}_{n^{\mathrm{st}}}.

For every nn\in\mathbb{N}, the \infty-category 𝒞nst{\cal C}_{n^{\mathrm{st}}} is \infty-semiadditive of semiadditive height nn, hence 𝒞{\cal C} itself is \infty-semiadditive and for every 0k0\leq k\leq\infty we have (5.5.12)

𝒞k(n𝒞nst)kn(𝒞nst)k𝒞kst.{\cal C}_{k}\simeq(\prod_{n\in\mathbb{N}}{\cal C}_{n^{\mathrm{st}}})_{k}\simeq\prod_{n\in\mathbb{N}}({\cal C}_{n^{\mathrm{st}}})_{k}\simeq{\cal C}_{k^{\mathrm{st}}}.

References

  • [AKQ19] Gabriel Angelini-Knoll and JD Quigley. Chromatic complexity of the algebraic K-theory of y(n)y(n). 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 K(2)K(2)-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 K(2){K}(2)-local sphere at p=2p=2. 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 \infty-categories via non-symmetric \infty-operads. Advances in mathematics, 279:575–716, 2015.
  • [GS96] John P.C. Greenlees and Hal Sadofsky. The Tate spectrum of vnv_{n}-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 K(n)K(n)-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 \infty-groups. Topology and its Applications, 222:121–138, 2017.