Hostname: page-component-745bb68f8f-hvd4g Total loading time: 0 Render date: 2025-02-04T22:43:24.605Z Has data issue: false hasContentIssue false

On $K(1)$-local $\mathrm {TR}$

Published online by Cambridge University Press:  04 May 2021

Akhil Mathew*
Affiliation:
Department of Mathematics, University of Chicago, 5734 S. University Ave., Chicago, IL60605, USAamathew@math.uchicago.edu
Rights & Permissions [Opens in a new window]

Abstract

We discuss some general properties of $\mathrm {TR}$ and its $K(1)$-localization. We prove that after $K(1)$-localization, $\mathrm {TR}$ of $H\mathbb {Z}$-algebras is a truncating invariant in the Land–Tamme sense, and deduce $h$-descent results. We show that for regular rings in mixed characteristic, $\mathrm {TR}$ is asymptotically $K(1)$-local, extending results of Hesselholt and Madsen. As an application of these methods and recent advances in the theory of cyclotomic spectra, we construct an analog of Thomason's spectral sequence relating $K(1)$-local $K$-theory and étale cohomology for $K(1)$-local $\mathrm {TR}$.

Type
Research Article
Copyright
© The Author(s) 2021

1. Introduction

The topological cyclic homology, $\mathrm {TC}(R)$, of a ring (or ring spectrum) $R$ is a basic invariant introduced by Bökstedt, Hsiang, and Madsen [Reference Bökstedt, Hsiang and MadsenBHM93] (see also Dundas, Goodwillie, and McCarthy [Reference Dundas, Goodwillie and McCarthyDGM13] and Nikolaus and Scholze [Reference Nikolaus and ScholzeNS18]) with many applications in algebraic $K$-theory. Its $p$-adic completion $\mathrm {TC}(R; \mathbb {Z}_p)$ arises as the fixed points of an operator called Frobenius on another invariant $\mathrm {TR}(R; \mathbb {Z}_p)$, which plays a central role in the approach to $\mathrm {TC}$ via equivariant stable homotopy theory. The construction $\mathrm {TR}(-; \mathbb {Z}_p)$ is often of arithmetic significance; for instance, the foundational calculations [Reference Hesselholt and MadsenHM03, Reference Hesselholt and MadsenHM04, Reference Geisser and HesselholtGH06] of the $p$-adic $K$-theory of local fields $F$ are based on the relationship between $\mathrm {TR}(\mathcal {O}_F; \mathbb {Z}_p)$ and the de Rham–Witt complex with log poles of $\mathcal {O}_F$.

In this paper, we prove some structural results about $\mathrm {TR}$ and how it relates to its $K(1)$-localization, throughout at an implicit prime number $p$. We will only consider $p$-typical $\mathrm {TR}$ and after $p$-adic completion. The operation of $K(1)$-localization when applied to algebraic $K$-theory is dramatically simplifying, as shown by Thomason [Reference ThomasonTho85, Reference Thomason and TrobaughTT90]; in particular, $K(1)$-local $K$-theory satisfies étale descent and admits a descent spectral sequence from étale cohomology under mild hypotheses; see [Reference Clausen and MathewCM19] for a modern account. Here we study analogs of some of these properties for $L_{K(1)} \mathrm {TR}(-)$. Our starting point is the following theorem.

Theorem 1.1 As a functor on connective $H\mathbb {Z}$-algebras, $L_{K(1)} \mathrm {TR}(-)$ is a truncating invariant in the Land–Tamme sense [Reference Land and TammeLT19]. In other words, if $R$ is a connective $H\mathbb {Z}$-algebra, then $L_{K(1) } \mathrm {TR}(R) \xrightarrow {\sim } L_{K(1)} \mathrm {TR}(\pi _0 R)$.

Theorem 1.1 refines results of [Reference Land, Mathew, Meier and TammeLMMT20, Reference Bhatt, Clausen and MathewBCM20], where it is shown that $L_{K(1)} K(-)$ and (equivalently) $L_{K(1)} \mathrm {TC}(-)$ are truncating invariants of connective $H\mathbb {Z}$-algebras, which ultimately follows from the claim

(1)\begin{equation} L_{K(1)} K(\mathbb{Z}/p^n) =0, \quad n \geq 1, \end{equation}

and more generally (and consequently) for any $H\mathbb {Z}$-algebra $R$,

(2)\begin{equation} L_{K(1)} K(R) \xrightarrow{\sim} L_{K(1)} K(R[1/p]). \end{equation}

In [Reference Bhatt, Clausen and MathewBCM20], (1) is proved via a calculation in prismatic cohomology; in [Reference Land, Mathew, Meier and TammeLMMT20], (1) is proved using some unstable chromatic homotopy theory. Our proof of Theorem 1.1 (which also gives a new proof of (1)) is based on a direct $\mathrm {TC}$-theoretic argument via estimation of exponents of nilpotence of the Bott element; in fact, it yields a slightly stronger result (Theorem 3.1 below).

The property of being truncating yields many pleasant features of the construction $L_{K(1)} \mathrm {TR}(-)$: by [Reference Land and TammeLT19], one obtains $cdh$-descent and excision. Since we are working $K(1)$-locally, we can combine this with results of [Reference Clausen, Mathew, Naumann and NoelCMNN20] to obtain $h$-descent.

Theorem 1.2 Any $K(1)$-local localizing invariant which is truncating, such as $L_{K(1)} \mathrm {TR}(-)$, satisfies $h$-descent on quasi-compact and quasi-separated (qcqs) schemes.

In particular, $L_{K(1)}\mathrm {TR}(-)$ satisfies étale descent. This is not so surprising, since $\mathrm {TR}(-; \mathbb {Z}_p)$ itself (like all Hochschild-theoretic invariants) actually satisfies flat descent; see [Reference Bhatt, Morrow and ScholzeBMS19, § 3]. However, Theorem 1.2 (together with (1)) leads to étale descent in the generic fiber. Since $\mathrm {TR}(R; \mathbb {Z}_p)$ of a ring $R$ depends only on the (derived) $p$-adic completion of $R$, we can informally view $L_{K(1)} \mathrm {TR}(R)$ as an invariant of the ‘rigid space’ associated to $R[1/p]$.

Example 1.3 (Galois descent in the generic fiber) Let $R \to S$ be a finite, finitely presented map of rings. Suppose we have a finite group $G$ acting on $S$ such that $R[1/p] \to S[1/p]$ is $G$-Galois. Then, for any $K(1)$-local localizing invariant $E$ which is truncating, we have $E(R) \xrightarrow {\sim } E(S)^{hG}$.

Recall that the Lichtenbaum–Quillen conjecture, refined by the Beilinson–Lichtenbaum conjecture proved by Voevodsky and Rost (see [Reference Haesemeyer and WeibelHW19] for an account), predicts that for $\mathbb {Z}[1/p]$-algebras $A$ satisfying mild finiteness conditions, the $p$-adic $K$-theory $K(A; \mathbb {Z}_p)$ is ‘asymptotically $K(1)$-local’; that is, the map $K(A; \mathbb {Z}_p) \to L_{K(1)} K(A; \mathbb {Z}_p)$ is an equivalence in high enough degrees. We next discuss analogs of such statements for $\mathrm {TR}(R; \mathbb {Z}_p)$ for $p$-adic rings $R$. Indeed, in [Reference Hesselholt and MadsenHM03, Reference Hesselholt and MadsenHM04] it is shown that if $R$ is smooth of relative dimension $d$ over a discrete valuation ring $\mathcal {O}_K$ of mixed characteristic with perfect residue field of characteristic $p > 2$, then $\mathrm {TR}(R; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {F}_p)$ is $d$-truncated; more precisely, this is a consequence of the relationship shown in [Reference Hesselholt and MadsenHM03, Reference Hesselholt and MadsenHM04] with the absolute de Rham–Witt complex. We prove this asymptotic $K(1)$-locality more generally for regular rings satisfying $F$-finiteness hypotheses from the Beilinson–Lichtenbaum conjecture applied to the generic fiber as well as the connection between $\mathrm {TR}$ and $p$-typical curves [Reference HesselholtHes96]. We expect that there should be a purely $p$-adic proof of this result (as in [Reference Hesselholt and MadsenHM03, Reference Hesselholt and MadsenHM04] in the smooth case).

Theorem 1.4 Let $R$ be a $p$-torsion-free excellent regular noetherian ring. Suppose that $R/p$ is finitely generated as a module over its subring of $p$th powers. Suppose, furthermore, that for all $\mathfrak {p} \in \mathrm {Spec}(R)$ containing $(p)$, we have $\dim R_{\mathfrak {p}} + \log _p [ \kappa (\mathfrak {p}):\kappa (\mathfrak {p})^p] \leq d$ for some $d \geq 0$. Then the map $\mathrm {TR}(R; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {F}_p)$ is $(d-1)$-truncated.

Using the Antieau–Nikolaus theory of topological Cartier modules [Reference Antieau and NikolausAN21], we relate the property of $\mathrm {TR}(R; \mathbb {F}_p)$ being ‘asymptotically $K(1)$-local’ to the extensively studied Segal conjecture for $\mathrm {THH}(R)$, that is, the condition that the cyclotomic Frobenius $\varphi \colon \mathrm {THH}(R; \mathbb {F}_p) \to \mathrm {THH}(R; \mathbb {F}_p)^{tC_p}$ should be an equivalence in high degrees. In particular, we obtain a version of the Segal conjecture for $\mathrm {THH}(R)$ when $R$ is regular. We expect that there should be a filtered version of this statement, using the motivic filtrations of Bhatt, Morrow, and Scholze [Reference Bhatt, Morrow and ScholzeBMS19].

Corollary 1.5 Let $R$ be as in Theorem 1.4. Then the cyclotomic Frobenius $\varphi \colon \mathrm {THH}(R; \mathbb {F}_p) \to \mathrm {THH}(R; \mathbb {F}_p)^{tC_p}$ is $(d-1)$-truncated.

Finally, we study the analog of Thomason's spectral sequence [Reference ThomasonTho85, Reference Thomason and TrobaughTT90] from étale cohomology to $K(1)$-local algebraic $K$-theory. For a scheme $X$ over $\mathbb {Z}[1/p]$ satisfying mild finiteness conditions (to wit: $X$ should be qcqs of finite Krull dimension, with a uniform bound on the mod $p$ virtual cohomological dimensions of the residue fields [Reference Rosenschon and ØstværRØ06, Reference Clausen and MathewCM19]), one has a convergent spectral sequence

\[ E_2^{s,t} =H^s(X, \mathbb{Z}_p(t)) \implies \pi_{2t-s} L_{K(1)} K(X). \]

We can construct a similar spectral sequence for $L_{K(1)} \mathrm {TR}$ under significantly stronger finiteness and regularity conditions, arising from a natural filtration. To formulate the $E_2$-term (or the graded pieces of this filtration), we use the $\mathrm {arc}_{{p}}$-topology of [Reference Bhatt and MathewBM18].

Definition 1.6 (The $\mathrm {arc}_{{p}}$-topology and $\mathrm {arc}_{{p}}$-cohomology)

We say that a map of derived $p$-complete rings $R \to R'$ is an $\mathrm {arc}_{{p}}$-cover if, for every map $R \to V$ for $V$ a rank $1$ valuation ring which is $p$-complete and such that $p \neq 0$, there exist an extension of rank $1$ valuation rings $V \to W$ and the following commutative diagram.

The $\mathrm {arc}_{{p}}$-topology is the finitary Grothendieck topology on the opposite of the category of derived $p$-complete rings defined such that a family $\{R \to R'_\alpha \}_{\alpha \in A}$ is a covering family if and only if there exists a finite subset $A' \subset A$ such that $R \to \prod _{\alpha \in A'} R'_\alpha$ is an $\mathrm {arc}_{{p}}$-cover.

Given any functor $F$ from derived $p$-complete rings to abelian groups, we let $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), F(-))$ denote the $\mathrm {arc}_{{p}}$-cohomology of $F(-)$ on a derived $p$-complete ring $R$.Footnote 1

Example 1.7 (i) We can consider the $\mathrm {arc}_{{p}}$-cohomology of the structure presheaf $\mathcal {O}$, $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), \mathcal {O})$. This is closely related to the perfectoidizations considered in [Reference Bhatt and ScholzeBS19, §§ 7 and 8], which work with the $p$-complete $\mathrm {arc}$-topology rather than the $\mathrm {arc}_{{p}}$-topology. For $R = \mathbb {Z}_p$, it is not difficult to see that this is the continuous $\mathrm {Gal}(\mathbb {Q}_p)$-homotopy invariants of the derived saturation $(\mathcal {O}_C)_*$ for $C = \widehat {\overline {\mathbb {Q}_p}}$ (in the sense of almost ring theory [Reference Gabber and RameroGR03]).

(ii) We consider the $\mathrm {arc}_{{p}}$-cohomology of the Witt vector presheaf $W(\mathcal {O})$, denoted $R \mapsto R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), W(\mathcal {O}))$, as well as its $p$-adic Tate twists $W(\mathcal {O})(i)$ for $i \in \mathbb {Z}$.

Theorem 1.8 Let $K$ be a complete nonarchimedean field of mixed characteristic $(0, p)$ with ring of integers $\mathcal {O}_K$ and residue field $k$ with $[k: k^p] < \infty$. Suppose either $K$ is discretely valued or $K$ is perfectoid. Let $R$ be a formally smooth $\mathcal {O}_K$-algebra. Then there exists a natural convergent, exhaustive $\mathbb {Z}$-indexed descending filtration $\mathrm {Fil}^{\geq \ast } L_{K(1)} \mathrm {TR}(R)$ on $L_{K(1)} \mathrm {TR}(R)$ such that

(3)\begin{equation} \mathrm{gr}^{i} L_{K(1)} \mathrm{TR}(R) \simeq R \Gamma_{\mathrm{arc}_{{p}}}( \mathrm{Spec}(R), W(\mathcal{O})(i))[2i]. \end{equation}

Theorem 1.8 is effectively an étale hyperdescent (in the generic fiber) result together with the calculation for perfectoids. For illustration, we specialize to the case where $R = \mathcal {O}_K$, for $K$ discretely valued with perfect residue field. Then the filtration (3) arises via a type of pro-Galois descent in the generic fiber. If $L/K$ is $G$-Galois, Example 1.3 and Theorem 1.4 imply that $\mathrm {TR}(\mathcal {O}_K; \mathbb {F}_p) \to \mathrm {TR}(\mathcal {O}_L; \mathbb {F}_p)^{hG}$ is a $0$-truncated map. However, this does not help with passage to $\bar {K}$ since $\mathrm {TR}$ does not commute with filtered colimits; note that $\mathrm {TR}$ has a simple form for $\mathcal {O}_{\bar {K}}$; see [Reference HesselholtHes06]. Using again the theory of topological Cartier modules as a ‘decompletion’ of the theory of cyclotomic spectra (see [Reference Antieau and NikolausAN21]), we prove the following pro-Galois result (see Example 6.9).

Theorem 1.9 Let $K$ be a complete, discretely valued field of characteristic $0$ with perfect residue field $k$ of odd characteristic $p$. Let $\mathrm {TR}(\mathcal {O}_K|K)$ denote the cofiber of the transfer map $\mathrm {TR}(k) \to \mathrm {TR}(\mathcal {O}_K)$. Then the natural map induces an equivalence

\[ \mathrm{TR}(\mathcal{O}_K|K; \mathbb{F}_p) \to \tau_{{\geq} 0}\mathrm{Tot} \Big( \mathrm{TR}( \mathcal{O}_{\bar{K}}; \mathbb{F}_p) \rightrightarrows \mathrm{TR}( \mathcal{O}_{\bar{K} \otimes_K \bar{K}}; \mathbb{F}_p) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big). \]

The idea that $\mathrm {TR}$ should satisfy this type of pro-Galois descent in the generic fiber is expressed in [Reference HesselholtHes02]; in particular, [Reference HesselholtHes02, Conjecture 5.1] predicts a related (but stronger) conclusion at the level of homotopy groups (in particular, the vanishing of higher Galois cohomology groups in the associated descent spectral sequence with $\mathbb {Q}_p/\mathbb {Z}_p$ coefficients).

Conventions

We write ${\rm Sp}$ for the $\infty$-category of spectra and $\mathbb {S}$ for the sphere spectrum. We use the theory of cyclotomic spectra in the form developed in [Reference Nikolaus and ScholzeNS18], as well as the theory of topological Cartier modules developed in [Reference Antieau and NikolausAN21]; we write $\mathrm {CycSp}$ for the $\infty$-category of cyclotomic spectra. We write $\mathrm {TR}$ for $p$-typical $\mathrm {TR}$. Given an $E_\infty$-ring $B$, a homogeneous element $x \in \pi _*(B)$, and a $B$-module $M$, we often write $M/x$ for the cofiber of multiplication by $x$ on $M$. In the case $x =p$, we will often write this as ‘$; \mathbb {F}_p$’; for example, $\mathrm {THH}(B; \mathbb {F}_p)$ refers to the cofiber of $p$ on $\mathrm {THH}(B)$. A $B$-algebra always refers, unless otherwise specified, to an $E_1$-algebra in $B$-modules.

2. Generalities on $K(1)$-local truncating invariants

Let $B$ be a base connective $E_\infty$-ring. In this section we work with localizing invariants on small $B$-linear idempotent-complete stable $\infty$-categories. Unlike in [Reference Blumberg, Gepner and TabuadaBGT13], we do not assume compatibility with filtered colimits, so for us a localizing invariant is simply a functor from (small, idempotent-complete) $B$-linear stable $\infty$-categories to spectra which carries Verdier quotient sequences to cofiber sequences. Following [Reference Land and TammeLT19], we say that such a localizing invariant $E$ is truncating if for every connective $B$-algebra $A$, we have $E(A) \xrightarrow {\sim } E(H \pi _0 A)$. This implies [Reference Land and TammeLT19, Theorem B] that $E$ satisfies excision.

Example 2.1 The constructions $L_{K(1)} K(-), L_{K(1)} \mathrm {TC}(-)$ are truncating on connective $H\mathbb {Z}$-algebras, as verified in [Reference Land, Mathew, Meier and TammeLMMT20]. For commutative $p$-complete rings, the two invariants actually agree (we do not know if this is true for noncommutative $p$-complete rings; see [Reference Bhatt, Clausen and MathewBCM20, Question 2.20]). Below we will show that $L_{K(1)} \mathrm {TR}(-)$ is truncating.

In the rest of the section we will assume for simplicity of notation that $B$ is discrete; by the assumption of truncatedness, this does not affect any of the results.

Proposition 2.2 Let $E$ be a $K(1)$-local localizing invariant of $B$-linear $\infty$-categories which is truncating. Then, on the category of discrete $B$-algebras, we have the following assertions.

  1. (i) $E$ is nilinvariant.

  2. (ii) $E$ annihilates any $B$-algebra $C$ which is annihilated by a power of $p$.

  3. (iii) Let $A \to A'$ be a map of $B$-algebras which is a $p$-isogeny. Then $E(A) \to E(A')$ is an equivalence.

Proof. For (i), the fact that $E$ is nilinvariant follows from [Reference Land and TammeLT19, Theorem B]. For (ii), since $E$ is nilinvariant, we may assume $C$ is an $\mathbb {F}_p$-algebra, so that $E(C)$ is by the theory of noncommutative motives [Reference Blumberg, Gepner and TabuadaBGT13] a $K(\mathbb {F}_p; \mathbb {Z}_p) = H\mathbb {Z}_p$-module (the last identification by [Reference QuillenQui72]); since $E$ is $K(1)$-local we get $E(C) = 0$.

For (iii), the kernel of $A \to A'$ is annihilated by a power of $p$, so by (ii) (and excision) we can assume that $A \subset A'$. Let $n \gg 0$, so $p^n A' \subset A$. Then the diagram

is a Milnor square of rings. Applying the localizing invariant $E$ and using excision and (ii) again, we conclude (iii).

In the next result, we use Voevodsky's $h$-topology for possibly non-noetherian schemes; in other words, the topology generated by finitely presented $v$-covers (also called universally subtrusive morphisms); see [Reference RydhRyd10] or [Reference Bhatt and ScholzeBS17, § 2].

Theorem 2.3 ($h$-descent for truncating $K(1)$-local invariants)

Let $E$ be a $K(1)$-local localizing invariant on $B$-linear $\infty$-categories which is truncating. Then $E$ satisfies $h$-descent on qcqs $B$-schemes.

Proof. By the results of [Reference Land and TammeLT19, Appendix A], $E$ satisfies $cdh$-descent, and in particular satisfies excision for abstract blow-up squares. The results of [Reference Clausen, Mathew, Naumann and NoelCMNN20] imply that $E$ satisfies finite locally free descent. Since $E$ also satisfies Nisnevich descent (as does any localizing invariant [Reference Thomason and TrobaughTT90]), we obtain that $E$ satisfies fppf descent thanks to [Sta20, Tag 05WN]. By [Reference Bhatt and ScholzeBS17, Theorem 2.9], $h$-descent (for any sheaf) is implied by fppf descent and excision for abstract blow-up squares. Combining these facts, we conclude.

Example 2.4 Let $E$ be as above. Let $\pi \colon X' \to X$ be a finitely presented proper morphism (e.g. a finitely presented closed immersion) of qcqs $B$-schemes such that $X'[1/p] \xrightarrow {\sim } X[1/p]$. Then $E(X) \xrightarrow {\sim } E(X')$. In fact, by $cdh$-descent, we have a pullback square

and the terms on the bottom vanish by Proposition 2.2.

In the next result we use the notion of nilpotence of a group action; see [Reference MathewMat18, § 4.1] or [Reference Clausen and MathewCM19, Definition 2.17] for accounts, or [Reference Mathew, Naumann and NoelMNN17] for the general setup in equivariant stable homotopy theory. Let $G$ be a finite group. The collection of nilpotent objects of the $\infty$-category $\mathrm {Fun}(BG, {\rm Sp} )$ is the thick subcategory generated by the objects which are induced from the trivial subgroup. For an algebra object of $\mathrm {Fun}(BG, {\rm Sp} )$, nilpotence holds if and only if the Tate construction vanishes. A module over a nilpotent algebra object in $\mathrm {Fun}(BG, {\rm Sp} )$ is itself nilpotent.

Corollary 2.5 (Galois descent in the generic fiber) Let $E$ be a $K(1)$-local localizing invariant of $B$-linear $\infty$-categories which is truncating. Let $R \to S$ be a finite and finitely presented map of (commutative, discrete) $B$-algebras. Let $G$ be a finite group acting on $S$ via $R$-algebra maps. Suppose that $R[1/p] \to S[1/p]$ is $G$-Galois. Then the natural map induces an equivalence $E(R) \xrightarrow {\sim } E(S)^{hG}$. Moreover, the $G$-action on $E(S)$ is nilpotent.

Proof. Replacing $S$ with $S \times R/p$, we may assume without loss of generality that $R \to S$ is an $h$-cover. Then, by Theorem 2.3, we have

(4)\begin{equation} E(R) \simeq \mathrm{Tot}( E(C(R \to S)^\bullet)) = \mathrm{Tot} \Big(E(S) \rightrightarrows E(S \otimes_R S ) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big). \end{equation}

Since the group $G$ acts on $S$, we have a natural map of cosimplicial rings from the Čech nerve $C(R \to S)^\bullet$ to the standard resolution $S \rightrightarrows \prod _G S \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix}$ for $G$ acting on $S$ (which calculates $S^{hG}$). This map of cosimplicial rings is an isogeny in each degree; for example, in degree $1$, $S \otimes _R S \to \prod _G S$ is an isogeny because it is a map of finitely presented $R$-modules (see [Sta20, Tag 0564]) which induces an isomorphism after inverting $p$ thanks to the Galois hypothesis. Therefore, by Proposition 2.2, we find that the map induces an equivalence after applying $E$, and we find from (4) that

\[ E(R) \xrightarrow{\sim} \mathrm{Tot} \bigg( E(S) \rightrightarrows E \biggl( \prod_G S \biggr) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \bigg) = E(S)^{hG}, \]

which is the desired claim. Finally, to see that the $G$-action on $E(S)$ is nilpotent, we use that $E(S)$ is a module $G$-equivariantly over $L_{K(1)} K(S) \xrightarrow {\sim } L_{K(1)} K(S[1/p])$ (via (2)), and the $G$-action on $L_{K(1)} K(S[1/p])$ is nilpotent by [Reference Clausen, Mathew, Naumann and NoelCMNN20, Theorem 5.6] (see also [Reference Clausen and MathewCM19, Lemma 4.20]).

3. The truncating property of $L_{K(1)} \mathrm {TR}(-)$

In this section we prove the following basic result. Throughout, we fix a connective, $K(1)$-acyclic $E_\infty$-ring $B$ (e.g. $H\mathbb {Z}$).

Theorem 3.1 The construction $L_{K(1)} \mathrm {TR}(-)$ is truncating on connective $E_1$-$B$-algebras. More generally, for any set $Q$,Footnote 2 the construction $L_{K(1)} (\prod _Q \mathrm {TR}(-))$ is truncating on connective $E_1$-$B$-algebras.

The proof of Theorem 3.1 will rely on a $K(1)$-acyclicity criterion for cyclotomic spectra (Proposition 3.6), which will use some elementary estimates for exponents of nilpotence with respect to the Bott element $\beta$.

In the following, we let $ku$ denote the connective topological $K$-theory spectrum, so $\pi _*(ku) = \mathbb {Z}[\beta ]$ with $|\beta | = 2$. Since $B$ is $K(1)$-acyclic, the associative ring spectrum $B \otimes ku/p$ is annihilated by a power of $\beta$. Our strategy is roughly based on bounding the exponents of nilpotence for $\beta$ in the fixed points $\mathrm {THH}(B)^{C_{p^n}} \otimes ku/p$ and, in particular, showing that they are $O(p^n)$.

Recall that $ku$ is complex-oriented, leading to the following result.

Proposition 3.2 Let $M$ be a $ku$-module equipped with an $S^1$-action. Then, for each $n \geq 1$, the natural map induces an equivalence

(5)\begin{equation} M^{hS^1} \otimes_{ku^{BS^1}} ku^{BC_{p^n}} \xrightarrow{\sim} M^{hC_{p^n}}. \end{equation}

Proof. Compare [Reference Mathew, Naumann and NoelMNN17, § 7.4]. In fact, via the projection formula, $M^{hC_{p^n}} \simeq (M \otimes _{ku} ku^{S^1/C_{p^n}})^{hS^1}$. Here $ku^{S^1/C_{p^n}}$ denotes the $ku$-valued function spectrum of $S^1/C_{p^n}$ with the corresponding $S^1$-action, and the tensor product is taken in $\mathrm {Fun}(BS^1, \mod (ku))$. Let $V_n$ denote the one-dimensional complex representation of $S^1$ where $z \in S^1$ acts by multiplication by $z^{p^n}$, and let $S(V_n)$ denote the unit circle in $V_n$ as an $S^1$-space, so $S(V_n) \simeq S^1/C_{p^n}$. The Spanier–Whitehead dual of the Euler sequence in $\mathrm {Fun}(BS^1, {\rm Sp} )$, $S(V_n)_+ \to S^0 \to S^{V_n}$ and the complex orientability of $ku$ together show that $ku^{S^1/C_{p^n}} \in \mathrm {Fun}(BS^1, \mod (ku))$ belongs to the thick subcategory generated by the unit. Therefore, applying the right adjoint $(-)^{hS^1} \colon \mathrm {Fun}(BS^1, \mod (ku)) \to \mod (ku^{BS^1})$, we find that the natural map

\[ M^{hS^1} \otimes_{ku^{BS^1}} (ku^{S^1/C_{p^n}})^{hS^1} \to (M \otimes_{ku} ku^{S^1/C_{p^n}})^{hS^1} = M^{hC_{p^n}} \]

is an equivalence, whence the result.

By complex orientability, we have

\[ \pi_*( ku^{BS^1}) = \mathbb{Z}[\beta][[x]], \quad |\beta| = 2, |x| ={-}2 . \]

Consider the formal group law over $\mathbb {Z}[\beta ]$ given by $f(u,v) = u + v + \beta uv$; this is the formal group law associated to the complex-oriented ring spectrum $ku$, and is homogeneous of degree $-2$ if $|u|, |v| = -2$. We have an equivalence $ku^{BS^1}/x \simeq ku$. More generally, let $[p^n](x) \in \pi _* ( ku^{BS^1})$ denote the $p$-series of the formal group law $f$; modulo $p$, we have $[p^n](x) \equiv \beta ^{p^n-1} x^{p^n}$ since $[p^n] (x)$ is homogeneous of degree $-2$ and recovers the multiplicative formal group law under the specialization $\beta \mapsto 1$. By the Eilenberg–Moore spectral sequence or Gysin sequence (of which (5) is a form), we have

(6)\begin{equation} ku^{BC_{p^n}} = ku^{BS^1}/ ([p^n](x)). \end{equation}

Proposition 3.3 Let $A$ be a connective $E_\infty$-ring spectrum with $S^1$-action. Suppose that $\beta ^r = 0$ in $\pi _*(A \otimes ku/p)$. Then in $\pi _*((A \otimes ku)^{hC_{p^n}}/p)$ we have $\beta ^{p^{n} -1 + rp^n} = 0$.

Proof. Let $R = A \otimes ku/p$, so $R$ is an associative $ku$-algebra spectrum with $S^1$-action. Consider the $S^1$-homotopy fixed points $R^{hS^1}$, which is a $ku^{BS^1}$-algebra. Since $p = 0$ in $\pi _0 R$, we have by Proposition 3.2 and (6),

\[ R^{hC_{p^n}} = R^{hS^1} \otimes_{ku^{BS^1}} ku^{BC_{p^n}} = R^{hS^1}/ [p^n](x) = R^{hS^1}/ \big(\beta^{p^n-1}x^{p^n}\big). \]

Our assumption is that $\beta ^r =0$ in $\pi _*(R)$, which means that we can write $\beta ^r = x v \in \pi _*(R^{hS^1})$ for some $v$ since $R = R^{hS^1}/x$. It follows that in $\pi _*(R^{hC_{p^n}})$, we have $\beta ^{p^n -1 + rp^n} = \beta ^{p^n-1} (xv)^{p^n} = 0$.

Now we apply the above to the $E_\infty$-ring $\mathrm {THH}(B)$ equipped with its $S^1$-action; recall that $B$ is assumed connective and $K(1)$-acyclic.

Proposition 3.4 There exists a constant $\kappa >0$ such that for each $i > 0$ we have the following assertions.

  1. (i) We have that $\beta ^{ \kappa i} =0$ in $(ku/p \otimes \mathrm {THH}(B))^{hC_{i}}$.

  2. (ii) Let $N$ be a $\mathrm {THH}(B) \otimes ku$-module in $\mathrm {Fun}(BS^1, {\rm Sp} )$. Suppose that $N$ (as a $\mathrm {THH}(B) \otimes ku$-module in $\mathrm {Fun}(BS^1, {\rm Sp} )$) is induced from the cyclic group $C_i \subset S^1$. Then, for any $t$, $N_{hC_t}/p$ is annihilated by $\beta ^{\kappa i}$.

Proof. To prove (i), using the transfer and restricting to a $p$-Sylow subgroup, we may reduce to the case where $i$ is a power of $p$, say $i = p^n$. Then the claim follows from Proposition 3.3, since $\beta$ is nilpotent in $ku/p \otimes \mathrm {THH}(B)$ since $B$ is $K(1)$-acyclic. It follows that we can find a $\kappa _1$ such that $\beta ^{\kappa _1 i} = 0$ in $(ku/p \otimes \mathrm {THH}(B))^{hC_i}$ for all $i > 0$.

For (ii), consider a $\mathrm {THH}(B) \otimes ku$-module $N$ in $\mathrm {Fun}(BS^1, {\rm Sp} )$ which is the induction of a $\mathrm {THH}(B) \otimes ku$-module $N'$ in $\mathrm {Fun}(BC_{i}, {\rm Sp} )$. It follows that $N_{hS^1} /p = (N' /p)_{hC_i}$ is a module over $(ku/p \otimes \mathrm {THH}(B))^{hC_i}$, writing homotopy orbits as a module over homotopy fixed points; this is therefore annihilated by $\beta ^{\kappa _1 i}$ by the previous part of the result. For any $t$, let $W_t$ be the one-dimensional complex representation of $S^1$ where $z$ acts by multiplication by $z^t$, so the unit circle $S(W_t)_+ = (S^1/C_t)_+$ as $S^1$-spaces. We have $N_{hC_t} = (N \otimes S(W_t)_+)_{hS^1}$ by the projection formula for induction and restriction along $C_t \subset S^1$. Then the Euler sequence $S(W_t)_+ \to S^0 \to S^{W_t}$ (as in the proof of Proposition 3.2) and the complex orientability of $ku$ yield a fiber sequence $N_{h C_t} \to N_{hS^1} \to \Sigma ^2 N_{hS^1}$. Therefore, $N_{hC_t} /p$ is annihilated by $\beta ^{2 \kappa _1 i}$; taking $\kappa = 2\kappa _1$ we conclude.

For the next result, we use the notion of a (nonnegatively) graded cyclotomic spectrum’ see [Reference Antieau, Mathew, Morrow and NikolausAMMN20, § 3 and Appendix A] or [Reference BrunBru01]. A graded cyclotomic spectrum consists of a graded spectrum $X = \bigoplus _{i \geq 0} X_i$ equipped with an $S^1$-action together with a graded $S^1$-equivariant cyclotomic Frobenius $\varphi _i \colon X_i \to X_{pi}^{tC_p}$ for each $i$. Given a nonnegatively graded $E_1$-ring $R$, the topological Hochschild homology $\mathrm {THH}(R)$ acquires the structure of a graded cyclotomic spectrum. Given a graded cyclotomic spectrum $X$, we can consider a graded cyclotomic spectrum $X_{\geq i}$ where we only consider the graded summands in degrees $i$ or higher; this gives any graded cyclotomic spectrum a natural descending filtration. The filtration quotients $X_{\geq i}/X_{\geq pi}$ have trivialized Frobenius because of the grading, and their $\mathrm {TR}$ can be thus described explicitly.

Construction 3.5 ($\mathrm {TR}$ of cyclotomic spectra with zero Frobenius)

Suppose $X \in \mathrm {CycSp}_{\geq 0}$ is $p$-complete with trivialized Frobenius. Then, as in [Reference Antieau and NikolausAN21, Remark 2.5] and [Reference Nikolaus and ScholzeNS18, Corollary II.4.7], we obtain a natural equivalence

(7)\begin{equation} \mathrm{TR}(X; \mathbb{Z}_p) \simeq \prod_{i \geq 0} X_{hC_{p^i}}. \end{equation}

For this, we also use the identification for $n \geq 1$, $(X^{tC_p})^{hC_{p^{n-1}}} \simeq X^{tC_{p^n}}$, see [Reference Nikolaus and ScholzeNS18, Lemma II.4.1].

Proposition 3.6 ($K(1)$-acyclicity criterion)

Let $X$ be a positively graded cyclotomic spectrum with the structure of a $\mathrm {THH}(B)$-module. Suppose, for each $i > 0$, that the following assertions hold.

  1. (i) $X_i$ is $(2p+ 1) \kappa i$-connected (where $\kappa$ is as in Proposition 3.4).

  2. (ii) As a $\mathrm {THH}(B)$-module in $\mathrm {Fun}(BS^1, {\rm Sp} )$, $X_i$ is induced from the cyclic group $C_i \subset S^1$.

Then, for any set $Q$, $\prod _Q \mathrm {TR}(X)$ is $K(1)$-acyclic.

Proof. For simplicity of notation, we write $\mathrm {TR}_Q(-) = \prod _Q \mathrm {TR}(-)$. The construction $\mathrm {TR}_Q(-)$ is exact and commutes with geometric realizations on $\mathrm {CycSp}_{\geq 0}$; therefore, it commutes with tensoring with $ku$. Without loss of generality, we can therefore assume that $X$ is a $\mathrm {THH}(B) \otimes ku$-module in graded cyclotomic spectra. Here we regard $ku$ as a trivial cyclotomic spectrum, that is, via the image of the unique symmetric monoidal functor ${\rm Sp} \to \mathrm {CycSp}$.

We consider the descending filtration $\{ \mathrm {Fil}^{\geq n } X = X_{\geq p^{n}}\}_{n \geq 0}$ on the cyclotomic spectrum $X$; note that the associated graded terms have trivialized Frobenii for degree reasons. This yields a filtration on the spectrum $\mathrm {TR}_Q(X)$, with $\mathrm {Fil}^{\geq n} \mathrm {TR}_Q(X) = \mathrm {TR}_Q(X_{\geq p^n})$. Since $\mathrm {TR}(-)$ preserves connectivity, our assumptions imply that $\mathrm {Fil}^{\geq n} \mathrm {TR}_Q(X)$ is $(2p + 1) \kappa p^n$-connective.

We have by (7) that $\mathrm {gr}^n \mathrm {TR}_Q(X) = \bigoplus _{p^{n} \leq i < p^{n+1}} \prod _Q\prod _{t \geq 0} (X_i)_{hC_{p^t}}$, since the cyclotomic Frobenius is trivial on $\mathrm {gr}^n X$ for grading reasons. Since $X_i$ is induced from $C_i \subset S^1$, it follows from Proposition 3.4 that the $ku$-module $\mathrm {gr}^n \mathrm {TR}_Q(X)/p$ is annihilated by $\beta ^{\kappa p^{n+1}}$.

Now $|\beta | = 2$, and $2\sum _{i = 0}^{n-1} \kappa p^{i+1} < (2p+1)\kappa p^{n}$ with the difference tending to $\infty$ as $n \to \infty$. Applying Lemma 3.7 below, we conclude that inverting $\beta$ annihilates $\mathrm {TR}_Q(X) /p$, whence $\mathrm {TR}_Q(X)$ is $K(1)$-acyclic as desired.

Lemma 3.7 Let $R$ be a connective $E_1$-ring spectrum, and let $x \in \pi _t(R)$ be an element. Let $\{\mathrm {Fil}^{\geq n} Y\}_{n \geq 0}$ be a filtered $R$-module spectrum. Suppose that there exist functions $f, g\colon \mathbb {N} \to \mathbb {N}$ such that:

  1. (i) $\mathrm {gr}^n (Y)$ is annihilated by $x^{f(n)}$;

  2. (ii) $\mathrm {Fil}^{\geq n}(Y)$ is $g(n)$-connective;

  3. (iii) $g(n) - t\sum _{i = 0}^{n-1} f(i) \to \infty$ for $n \to \infty$.

Then $Y[1/x] = 0$.

Proof. Let $y \in \pi _s(Y)$. For each $n > 0$, the class $x^{f(0) + f(1) + \cdots + f(n-1)} y \in \pi _*(Y)$ naturally lifts to $\pi _{s + t(f(0) + \cdots + f(n-1))}( \mathrm {Fil}^{\geq n} Y)$. But for $n \gg 0$, the connectivity of $\mathrm {Fil}^{\geq n} Y$ forces this last group to vanish. Therefore, the image of $y$ in $\pi _s(Y[1/x])$ vanishes as desired.

We now use the following basic calculation of $\mathrm {THH}$ of a free associative algebra, as a spectrum equipped with $S^1$-action. Versions of this are classical in ordinary Hochschild homology; see [Reference LodayLod98, § 3.1]. In the language of factorization homology, this result is a special case of the calculation of the factorization homology of a free algebra ([Reference Ayala, Francis and TanakaAFT17, Proposition 4.13] and [Reference Ayala and FrancisAF15, Proposition 5.5]).

Theorem 3.8 ($\mathrm {THH}$ of a free associative algebra)

Let $M$ be a spectrum, and let $T(M) = \bigoplus _{n \geq 0}M^{\otimes n}$ be the free $E_1$-algebra spectrum generated by $M$. Then there is a natural equivalence in $\mathrm {Fun}(BS^1, {\rm Sp} )$,

(8)\begin{equation} \mathrm{THH}( T(M)) \simeq \bigoplus_{n \geq 0} \mathrm{Ind}_{C_n}^{S^1} (M^{{\otimes} n}) , \end{equation}

where we use the natural $C_n$-action on $M^{\otimes n}$ by permuting the factors.

Proof. The results of [Reference Ayala, Francis and TanakaAFT17, Reference Ayala and FrancisAF15] (applied to the framed manifold $S^1$) imply that $\mathrm {THH}( T(M)) \simeq \int _{S^1} T(M) = \bigoplus _{n \geq 0} (\mathrm {Conf}_n(S^1)_+ \otimes M^{\otimes n})_{h \Sigma _n}$, for $\mathrm {Conf}_n(S^1)$ the configuration space of $n$ ordered points on the circle. One checks now (see [Reference Crabb and JamesCJ98, Example II.14.4]) that $\mathrm {Conf}_n(S^1)$, as a space with $S^1 \times \Sigma _n$-action, is homotopy equivalent to $(S^1 \times \Sigma _n)/C_n$ (with $C_n$ embedded diagonally), whence the claim.

Proposition 3.9 Let $M, N$ be connective spectra. Then the map of cyclotomic spectra

\[ \mathrm{THH}( T(M \oplus N)) \otimes \mathrm{THH}(B) \to \mathrm{THH}(T(M)) \otimes \mathrm{THH}(B) \]

induces an equivalence on $L_{K(1)} ( \prod _Q \mathrm {TR}(-))$ for any set $Q$ if $N$ is at least $\kappa (2p+1)$-connective.

Proof. We can consider the tensor algebra $T(M \oplus N)$ as a graded $E_1$-ring spectrum where $M$ is placed in degree $0$ and $N$ is placed in degree $1$. In this case, if we collect the terms in (8), we find that the $i$th graded piece of $\mathrm {THH}( T(M \oplus N))$ is the component which is $i$-homogeneous. Explicitly, for any subset $I \subset \langle n\rangle = \{1, 2, \ldots , n\}$, we write $(M, N)^{\otimes (\langle n \rangle \setminus I, I)}$ for the ordered tensor product of $n$ factors, where the $j$th factor is $M$ if $j \notin I$ and $N$ if $j \in I$. Expanding (8) gives

(9)\begin{equation} \mathrm{THH}( T(M \oplus N))_i = \bigoplus_{n \geq 0} \mathrm{Ind}_{C_n}^{S^1} \bigg( \bigoplus_{I \subset \langle n \rangle, |I| = i} (M, N)^{{\otimes} (\langle n \rangle\setminus I, I)} \bigg). \end{equation}

Here the $C_n$-action on the parenthesized term in (9) permutes the various summands. Note in particular that the stabilizer of $I \subset \langle n\rangle$ in the $n$th summand is a subgroup of a cyclic group $C_i \subset C_n$ since $|I| = i$. In particular, it follows that the $i$th graded piece of $\mathrm {THH}( T(M \oplus N))$ is induced from $C_i \subset S^1$. Furthermore, since $N$ is at least $\kappa (2p+1)$-connective, it follows that the $i$th graded piece of $\mathrm {THH}( T( M \oplus N))$ is at least $\kappa (2p+1)$-connective. By Proposition 3.6, it follows that the positively graded part of $\mathrm {THH}(T(M \oplus N)) \otimes \mathrm {THH}(B)$ has vanishing $K(1)$-local $\prod _Q \mathrm {TR}(-)$ for any set $Q$, whence the result.

For the next result, if $R$ is an $E_1$-ring spectrum and $N$ an $(R, R)$-bimodule, we let $T_R(N) = \bigoplus _{n \geq 0} N \otimes _R N \otimes _R \cdots \otimes _R N$ be the free $E_1$-algebra under $R$ generated by $N$.

Proposition 3.10 Let $R$ be a connective $B$-algebra. Let $N$ be an at least $\kappa (2p+1)$-connective $(R, R)$-bimodule. Then the map $\mathrm {THH}( T_R(N)) \to \mathrm {THH}(R)$ induces an equivalence on $L_{K(1)} (\prod _Q \mathrm {TR}(-))$ for any set $Q$.

Proof. Simplicially resolving $N$ by free $(R, R)$-bimodules (since $\prod _Q \mathrm {TR}(-)$ commutes with geometric realizations), we may assume that $N$ is free on generators (possibly infinitely many) in degrees $\kappa (2p+1)$ or higher. Simplicially resolving $R$ by free $B$-algebras, we may assume that $R$ is free as well, on some classes in degrees at least $0$. In this case, the result follows from Proposition 3.9.

Lemma 3.11 Let $E$ be a localizing invariant of $B$-linear $\infty$-categories. Suppose that there exists $k \geq 0$ such that for every connective $B$-algebra $R$, we have $E( R) \xrightarrow {\sim } E( \tau _{\leq k} R)$. Then $E$ is truncating.

Proof. We show by descending inductionFootnote 3 that for any $i \geq 0$ and for any connective $B$-algebra $R$, the map $R \to \tau _{\leq i} R$ induces an equivalence on $E$; taking $i = 0$ gives the theorem. For $i \geq k$, we already know the claim by assumption. Suppose we know the claim for $i+1$; to prove the claim for $i$, we need to see that $\tau _{\leq i+1} R \to \tau _{\leq i} R$ induces an equivalence on $E$. Now $\tau _{\leq i+1} R \to \tau _{\leq i} R$ is a square-zero extension, that is, we have the following pullback diagram of $B$-algebras.

Since $E$ is a localizing invariant, the main result of Land and Tamme [Reference Land and TammeLT19] yields a $B$-algebra $\tilde {R}$ with underlying spectrum $H(\pi _0 R) \otimes _{\tau _{\leq i+1} R } \tau _{\leq i} R$ fitting into a commutative diagram of $B$-algebras

which is carried to a pullback by $E$. But the map $H(\pi _0 R) \to \tilde {R}$ is an equivalence in degrees up to $i+1$ and therefore induces an equivalence on $E$ by the inductive hypothesis. Therefore, $E( \tau _{\leq i+1} R) \to E( \tau _{\leq i} R)$ is an equivalence, whence the inductive step and the result.

Proof of Theorem 3.1 As before, we write $\mathrm {TR}_Q(-) = \prod _Q \mathrm {TR}(-)$ for a set $Q$. For any connective $B$-algebra $R$, we claim that $R \to \tau _{\leq \kappa (2p+1)} R$ induces an equivalence on $L_{K(1)} \mathrm {TR}_Q(-)$. Indeed, this follows from Proposition 3.10 because we can simplicially resolve $\tau _{\leq \kappa (2p+1)} R$ using free $R$-algebras over free $(R, R)$-bimodules on classes in degrees $\kappa (2p+1)$ or higher and since $\mathrm {TR}_Q(-)$ commutes with geometric realizations. The result now follows from Lemma 3.11.

Corollary 3.12 (cf. [Reference Land, Mathew, Meier and TammeLMMT20])

The invariants $L_{K(1)} K(-), L_{K(1)} \mathrm {TC}(-)$ are truncating on connective $B$-algebras.

Proof. The result for $L_{K(1)} \mathrm {TC}(-)$ follows from Theorem 3.1 by taking Frobenius fixed points. The result for $L_{K(1)} K(-)$ is a formal consequence since the fiber of the trace $K(-) \to \mathrm {TC}(-)$ is truncating by the Dundas–Goodwillie–McCarthy theorem [Reference Dundas, Goodwillie and McCarthyDGM13].

Corollary 3.13 [Reference Land, Mathew, Meier and TammeLMMT20, Reference Bhatt, Clausen and MathewBCM20]

For any $n$, we have $L_{K(1)} K(\mathbb {Z}/p^n) = 0$.

Proof. Take $B = H\mathbb {Z}$ in Theorem 3.1, so $L_{K(1)} \mathrm {TC}(-)$ is truncating and therefore nilinvariant on connective $H\mathbb {Z}$-algebras. Then the result follows because the Dundas–Goodwillie–McCarthy theorem and comparison with $\mathbb {F}_p$ yield $L_{K(1)} K(\mathbb {Z}/p^n) = L_{K(1)} \mathrm {TC}(\mathbb {Z}/p^n)$, but the above shows that this equals $L_{K(1)} \mathrm {TC}(\mathbb {F}_p) = 0$.

4. Asymptotic $K(1)$-locality

In this section we show (Theorem 4.8) that $\mathrm {TR}$ is asymptotically $K(1)$-local for a regular ring satisfying mild hypotheses, using the Beilinson–Lichtenbaum conjecture. This result is due to Hesselholt and Madsen in the case of smooth algebras over a DVR with perfect residue field of characteristic greater than $2$, which we begin by reviewing.

Theorem 4.1 (Hesselholt–Madsen [Reference Hesselholt and MadsenHM03, Reference Hesselholt and MadsenHM04])

Let $K$ be a complete, discretely valued field of characteristic $0$ with ring of integers $\mathcal {O}_K \subset K$ and perfect residue field $k$ of characteristic $p > 2$. Let $R$ be a smooth $\mathcal {O}_K$-algebra of relative dimension $d$. Then the map

\[ \mathrm{TR}(R; \mathbb{F}_p) \to L_{K(1)} \mathrm{TR}(R; \mathbb{F}_p) \]

is $d$-truncated.

Proof. Without loss of generality, we can assume that $\mu _p \subset K$, since otherwise $\mathrm {TR}(R; \mathbb {F}_p)$ is a retract of $\mathrm {TR}(R[\zeta _p]; \mathbb {F}_p)$ via the transfer. In this case, the result follows from [Reference Hesselholt and MadsenHM04, Theorem E]. Indeed, [Reference Hesselholt and MadsenHM04, Theorem E] gives a calculation of the cofiber $\mathrm {TR}(R|R_K; \mathbb {F}_p)$ of the transfer map $\mathrm {TR}(R \otimes _{\mathcal {O}_K} k; \mathbb {F}_p) \to \mathrm {TR}(R; \mathbb {F}_p)$. Since $\mathrm {TR}(R \otimes _{\mathcal {O}_K} k; \mathbb {F}_p)$ is $K(1)$-acyclic and $d$-truncated in view of the identification [Reference HesselholtHes96] with the de Rham–Witt complex of $R \otimes _{\mathcal {O}_K} k$, it suffices to verify the (stronger) claim that $\mathrm {TR}(R|R_K; \mathbb {F}_p) \to L_{K(1)}\mathrm {TR}(R|R_K; \mathbb {F}_p)$ is $(d-2)$-truncated. Equivalently (see Lemma 5.8 below), it suffices to show that the cofiber of the Bott element on $\mathrm {TR}(R|R_K; \mathbb {F}_p)$ is $(d+1)$-truncated. In fact, this follows from the calculation in [Reference HesselholtHes96], once we know that the absolute de Rham–Witt complex $W \Omega ^{\ast }_{(R, M_R)}$ is $p$-divisible in degrees $d+2$ or higher. This in turn follows from the case where $R = \mathcal {O}_K$ (see [Reference Hesselholt and MadsenHM03, Corollary 3.2.7]; the case of polynomial rings via the functor $P$ (see [Reference Hesselholt and MadsenHM04, Lemma 7.1.4] and its proof); and finally, étale base-change [Reference Hesselholt and MadsenHM04, Lemma 7.1.1].

Construction 4.2 ($\mathrm {TR}$ as $p$-typical curves)

Let $R$ be an animated ring.Footnote 4 We let

(10)\begin{equation} C(R) = \varprojlim_n \Omega K( R[x]/x^n, (x)) \end{equation}

denote the spectrum of curves on the $K$-theory of $R$, defining a functor from animated rings to spectra. By [Reference HesselholtHes96, Theorem 3.1.9], if $R$ is a discrete $\mathbb {Z}/p^j$-algebra for some $j$, we have a natural expression of $\mathrm {TR}(R; \mathbb {Z}_p)$ as a summand of $C(R; \mathbb {Z}_p)$ (note that we are considering $p$-typical $\mathrm {TR}$, while in [Reference HesselholtHes96, Theorem 3.1.9] global $\mathrm {TR}$ is considered). Left Kan extending both sides to animated rings (since both $\mathrm {TR}(-)$ and $C(-)$ commute with geometric realizations),Footnote 5 we obtain that $\mathrm {TR}(R; \mathbb {Z}_p)$ is naturally a summand of $C(R)$ for $R$ an animated $\mathbb {Z}/p^j$-algebra. Passing to the limit over $j$ and using the $p$-adic continuity of $K$-theory [Reference Clausen, Mathew and MorrowCMM21, Theorem 5.21], we find that for any $p$-henselian animated ring $R$ (i.e. $\pi _0(R)$ is $p$-henselian), $\mathrm {TR}(R; \mathbb {Z}_p)$ is a summand of $C(R; \mathbb {Z}_p)$.

Next, we discuss the comparison between the $p$-adic $K$-theory of $R$ and $R[1/p]$; see [Reference NiziołNiz08, Lemma 3.5] for this argument.

Proposition 4.3 Let $R$ be a regular noetherian ring of finite Krull dimension. Suppose, for every $x \in \mathrm {Spec}(R/p)$, that we have $[\kappa (x): \kappa (x)^p] \leq p^d$. Then the map $K(R; \mathbb {F}_p) \to K(R[1/p]; \mathbb {F}_p)$ is $d$-truncated.

Proof. The homotopy fiber of the map in question is, by Quillen's dévissage theorem, the mod $p$ $K$-theory of the abelian category of finitely generated $R/p$-modules, that is, the mod $p$ $G$-theory of $R/p$, $G(R/p; \mathbb {F}_p)$. So it suffices to show that $G(R/p; \mathbb {F}_p)$ is concentrated in degrees up to $d$. Using the filtration by codimension of support and dévissage (see [Reference QuillenQui73, Theorem 5.4]), we find that $G(R/p; \mathbb {F}_p)$ has a filtration whose associated graded terms are direct sums of the $K( \kappa (x); \mathbb {F}_p)$ for $x \in \mathrm {Spec}(R/p)$; it therefore suffices to show that these terms are $d$-truncated. But now the Geisser–Levine theorem [Reference Geisser and LevineGL00] implies that for any field $E$ of characteristic $p$, there is an embedding $K_*(E; \mathbb {F}_p) \hookrightarrow \Omega ^{\ast }_{E/\mathbb {F}_p}$. This implies that $K(\kappa (x); \mathbb {F}_p)$ is $\dim _{\kappa (x)} \Omega ^1_{\kappa (x)/\mathbb {F}_p}$-truncated. But $\dim _{\kappa (x)} \Omega ^1_{\kappa (x)/\mathbb {F}_p} = \log _p [\kappa (x): \kappa (x)^p] \leq d$ by the theory of $p$-bases [Reference MatsumuraMat86, Theorem 26.5]. Combining these facts, the result follows.

Proposition 4.4 (Rosenschon–Østvær [Reference Rosenschon and ØstværRØ06])

Let $R$ be a regular noetherian $\mathbb {Z}[1/p]$-algebra of finite Krull dimension. Suppose that $\mathrm {vcd}_p(\kappa (x)) \leq d$ for all $x \in \mathrm {Spec}(R[1/p])$. Then the map $K(R; \mathbb {F}_p) \to L_{K(1)} K(R; \mathbb {F}_p)$ has $(d-3)$-truncated homotopy fiber.

Proof sketch Using Nisnevich descent [Reference Thomason and TrobaughTT90], we reduce to the case where $R$ is henselian local, and even a field of characteristic not equal to $p$ by Gabber–Suslin rigidity [Reference GabberGab92]; note that we do not have to worry about the distinction between connective and nonconnective $K$-theory by regularity. Then the result follows from the Beilinson–Lichtenbaum conjecture (proved by Voevodsky and Rost; see [Reference Haesemeyer and WeibelHW19]) describing the associated graded terms of the motivic filtration on $K(R; \mathbb {F}_p)$ (see, for example, [Reference Clausen and MathewCM19, § 6.2] for an account).

Proposition 4.5 Let $R$ be an excellent normal domain which is henselian along an ideal $I \subset R$ containing $(p)$. Suppose that, for all $\mathfrak {p} \in \mathrm {Spec}(R)$ containing $I$, we have $\dim R_{\mathfrak {p}} + \log _p [\kappa (\mathfrak {p}): \kappa (\mathfrak {p})^p] \leq d$. Then, for all $\mathfrak {q} \in \mathrm {Spec}(R[1/p])$, the residue field $\kappa (\mathfrak {q})$ has $p$-cohomological dimension at most $d + 1$.

Proof. By standard continuity arguments, it suffices to show that for any affine open $U = \mathrm {Spec}( R[1/f]) \subset \mathrm {Spec}(R[1/p])$ and any constructible $p$-torsion sheaf $\mathcal {F}$ on $U$, we have $H^n( U, \mathcal {F}) = 0$ for $n > d +1$. Denote by $j\colon U \hookrightarrow \mathrm {Spec}(R)$ the open inclusion, and denote by $i_0\colon \mathrm {Spec}(R/I) \hookrightarrow \mathrm {Spec}(R)$ the closed embedding. Using that $\mathrm {Spec}(R/I)$ has $p$-cohomological dimension at most $1$ and the affine analog of proper base change [Reference GabberGab94, Reference HuberHub93] applied to the henselian ideal $I \subset R$, we see that it suffices to show that $i_0^* R^n j_* \mathcal {F} = 0$ for $n > d$.

Working stalkwise on $R$ and using the compatibility of étale cohomology with filtered colimits, we can now reduce to the case where $R$ is an excellent, strictly henselian normal local domain with residue field $k$ (and $I \subset R$ is contained in the maximal ideal), since excellence and normality are preserved by strict henselization (see [Reference GrecoGre76] for the former). The statement then becomes that $H^n( U, \mathcal {F}) = 0$ for $n > d$, which follows from the Gabber–Orgogozo bound [Reference Gabber and OrgogozoGO08, Theorem 6.1] (noting that the $p$-dimension of the residue field $k$ is $\log _p [k:k^p]$ since $k$ is separably closed).

Remark 4.6 Suppose that $R_0$ is a ring such that $R_0/p$ is finite type over a perfect field $k$ of characteristic $p$. Then, for every prime ideal $\mathfrak {p} \in \mathrm {Spec}(R_0)$ containing $(p)$, we have $\dim (R_0)_{\mathfrak {p}} + \log _p [\kappa (\mathfrak {p}): \kappa (\mathfrak {p})^p] \leq \dim (R_0)$. Indeed, we have $\dim (R_0)_{\mathfrak {p}} + \dim (R_0/\mathfrak {p}) \leq \dim (R_0)$. Therefore, it suffices to prove $\log _p [ \kappa ( \mathfrak {p}): \kappa (\mathfrak {p})^p] = \dim (R_0/\mathfrak {p})$. But in this case, both are the transcendence degree of the field $\kappa ( \mathfrak {p})$ over $k$; see [Reference MatsumuraMat80, Theorem 27.B].

Lemma 4.7 Let $R_0$ be an $\mathbb {F}_p$-algebra. Suppose that for all residue fields $\kappa$ of $R_0$, we have $[\kappa : \kappa ^p] \leq p^d$ for some $d \geq 0$. Then, for all residue fields $\kappa '$ of $R_0[[x]]$, we have $[\kappa ': \kappa '^p] \leq p^{d+1}$.

Proof. We have that $R_0[[x]]$ is generated by $p$ elements over the subring generated by its $p$th powers and by $R_0$. For every residue field $\kappa$ of $R_0$, it follows that $R_0[[x]] \otimes _{R_0} \kappa$ is generated by $p$ elements over the subring generated by $p$th powers together with $\kappa$. In particular, it is generated by $p^{d+1}$ elements as a module over its $p$th powers. It follows that the same holds for any residue field of $R_0[[x]] \otimes _{R_0} \kappa$ and, varying $\kappa$, we obtain the conclusion for every residue field of $R_0$.

Theorem 4.8 Let $R$ be an excellent, $p$-torsion-free regular noetherian ring. Suppose $R/p$ is finitely generated as a module over its subring of $p$th powers. Suppose, furthermore, that for all $\mathfrak {p} \in \mathrm {Spec}(R)$ containing $(p)$, we have $\dim R_{\mathfrak {p}} + \log _p [ \kappa (\mathfrak {p}):\kappa (\mathfrak {p})^p] \leq d$ for some $d \geq 0$. Then the map $\mathrm {TR}(R; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {F}_p)$ is $(d-1)$-truncated.

Proof. Without loss of generality, we can assume that $R$ is $p$-henselian (since henselization preserves excellence; see [Reference GrecoGre76]), so $\dim R \leq d$. By Construction 4.2, it suffices to show that $C(R)/p \to L_{K(1)} C(R)/p$ is $(d-1)$-truncated. Now $C(R)$ is the desuspension of the fiber of the map $\varprojlim _n K( R[x]/x^n) \to K(R)$. With $p$-adic coefficients, the fact that $R/p$ is finitely generated as a module over its $p$th powers allows us to pass to the limit [Reference Clausen, Mathew and MorrowCMM21, Theorem F] and we obtain

\[ C(R; \mathbb{F}_p) = \Omega K(R[[x]], (x); \mathbb{F}_p). \]

Since $K(R)$ is a retract of $K(R[[x]])$, it suffices to show that the fiber of $K(R[[x]]; \mathbb {F}_p) \to L_{K(1)}K(R[[x]]; \mathbb {F}_p)$ is $d$-truncated. We will verify this by comparing both sides with the intermediate term $K( R[[x]][1/p]; \mathbb {F}_p)$.

First, for every characteristic $p$ residue field $\kappa$ of $R$, corresponding to a prime ideal $\mathfrak {p} \subset R$ containing $(p)$, we have $[\kappa : \kappa ^p] \leq p^{d-1}$ by our assumption, since $\dim R_{\mathfrak {p}} \geq 1$. Now $R[[x]]$ is also a $p$-torsion-free regular ring of dimension $\dim (R) + 1$. For every characteristic $p$ residue field $\kappa '$ of $R[[x]]$, we have $[\kappa ': \kappa '^p] \leq p^{d}$ by Lemma 4.7. By Proposition 4.3, it follows that $K(R[[x]]; \mathbb {F}_p) \to K(R[[x]][1/p]; \mathbb {F}_p)$ is $d$-truncated.

Second, we apply Proposition 4.5 to the ring $R[[x]]$ and the ideal $I = (p, x)$. The power series ring $R[[x]]$ remains excellent (and regular) since $R$ is excellent, thanks to [Reference Kurano and ShimomotoKS16]. For any prime ideal $\mathfrak {p} \in \mathrm {Spec}( R[[x]])$ containing $I$, we let $\mathfrak {p}_0 = \mathfrak {p} \cap R \subset R$, so that $R[[x]]_{\mathfrak {p}}/(x) = R_{\mathfrak {p}_0}$. Then $\dim ( R[[x]]_{\mathfrak {p}}) = \dim ( R_{\mathfrak {p}_0})+ 1$ and $\kappa ( \mathfrak {p}) = \kappa (\mathfrak {p}_0)$. Thus, Proposition 4.5 applies to $R[[x]]$ (with $d$ replaced by $d+1$) and we find that the characteristic $0$ residue fields of $R[[x]]$ have $p$-cohomological dimension at most $d + 2$. Therefore, by Proposition 4.4, $K( R[[x]][1/p]; \mathbb {F}_p) \to L_{K(1)} K( R[[x]][1/p])$ is $(d-1)$-truncated.

Combining the above, we find that the composite map $K(R[[x]]; \mathbb {F}_p) \to L_{K(1)} K(R[[x]]; \mathbb {F}_p) \simeq L_{K(1)} K( R[[x]][1/p]; \mathbb {F}_p)$ (the last identification by (2)) is $d$-truncated, whence the result.

This recovers in particular Theorem 4.1. More generally, we have the following example.

Example 4.9 Let $R$ be a $d$-dimensional regular, excellent $p$-torsion-free noetherian ring with $R/p$ finitely generated over its $p$th powers. Suppose that $(R/p)_{\mathrm {red}}$ is finite type over a perfect field (and necessarily of dimension $d-1$). Then Theorem 4.8 (in view of Remark 4.6) applies to show that $\mathrm {TR}(R; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {F}_p)$ is $(d-1)$-truncated.

5. The Segal conjecture

In this section we discuss the relationship between the following two properties of a cyclotomic spectrum $X$:

  1. (i) $\mathrm {TR}(X)$ agrees with its $K(1)$-localization in high enough degrees;

  2. (ii) the cyclotomic Frobenius $\varphi _X \colon X \to X^{tC_p}$ is an equivalence in high enough degrees.

Property (i) is the Lichtenbaum–Quillen style statement discussed in the previous section, and verified for $\mathrm {THH}(R)$ under regularity and finiteness hypotheses. Property (ii) is often referred to as the ‘Segal conjecture’ since for $X = \mathrm {THH}(\mathbb {S})$, the Frobenius $\mathbb {S} \to \mathbb {S}^{tC_p}$ is a $p$-adic equivalence by the Segal conjecture for $C_p$, proved in [Reference LinLin80, Reference GunawardenaGun80]. The Segal conjecture has been studied extensively for $\mathrm {THH}(R)$ for $R$ a ring (or ring spectrum).

We first show the implication $({\rm i}) \implies ({\rm ii})$. We use the Antieau–Nikolaus theory of topological Cartier modules [Reference Antieau and NikolausAN21], which we begin by briefly reviewing.

Definition 5.1 A topological Cartier module $M$ is an object of $\mathrm {Fun}(BS^1, {\rm Sp} )$ together with maps $V \colon M_{hC_p} \to M$ and $F \colon M \to M^{hC_p}$ in $\mathrm {Fun}(BS^1, {\rm Sp} )$ together with a homotopy between the composite and the norm map $M_{hC_p} \to M^{hC_p}$ (considered $S^1 \simeq S^1/C_p$-equivariantly). The collection of topological Cartier modules is naturally organized into a presentable stable $\infty$-category.

Given a bounded-below, $p$-typical cyclotomic spectrum $X$, we can consider $\mathrm {TR}(X)$ as a topological Cartier module, and we have an identification $X \simeq \mathrm {cofib}(V)$. Under these identifications, the cyclotomic Frobenius $X \to X^{tC_p}$ is obtained from $F\colon \mathrm {TR}(X) \to \mathrm {TR}(X)^{hC_p}$ by taking cofibers by $V$ on both domain and codomain and identifying $\mathrm {TR}(X)^{tC_p} \simeq X^{tC_p}$ as $(\mathrm {TR}(X)_{h{C_p}})^{tC_p} = 0$ [Reference Nikolaus and ScholzeNS18, Lemma I.2.1]. On bounded-below objects, this construction establishes a fully faithful embedding from cyclotomic spectra into topological Cartier modules, with image given by the $V$-complete objects [Reference Antieau and NikolausAN21, Theorem 3.21].

Proposition 5.2 Let $X$ be a connective, $p$-complete cyclotomic spectrum whose underlying spectrum is $K(1)$-acyclic. Suppose the map $\mathrm {TR}(X) \to L_{K(1)} \mathrm {TR}(X)$ is $d$-truncated. Then the Frobenius $\varphi \colon X \to X^{tC_p}$ is $d$-truncated.

In the case $L_{K(1)} \mathrm {TR}(X) =0$, the result is [Reference Antieau and NikolausAN21, Proposition 2.25].

Proof. Since $X$ is $K(1)$-acyclic, it follows that $V\colon \mathrm {TR}( X)_{hC_p} \to \mathrm {TR}(X)$ is $K(1)$-locally an equivalence. The $K(1)$-localization $L_{K(1)} \mathrm {TR}(X)$ acquires the structure of a topological Cartier module as well by $K(1)$-localizing $F, V$ and using the comparison map $L_{K(1)} ( \mathrm {TR}(X)^{hC_p}) \to (L_{K(1)} \mathrm {TR}(X))^{hC_p}$. The composite map $(L_{K(1)} \mathrm {TR}(X))_{hC_p} \to (L_{K(1)} \mathrm {TR}(X))^{hC_p}$ is an equivalence after $p$-completion since Tate constructions vanish in the $K(1)$-local category. Since we saw that $V$ is an equivalence on $L_{K(1)} \mathrm {TR}(X)$ after $p$-completion, it follows that the Frobenius on $L_{K(1)} \mathrm {TR}(X)$ induces an equivalence

\[ L_{K(1)} \mathrm{TR}( X) \xrightarrow{\sim} (L_{K(1)} \mathrm{TR}(X))^{hC_p}. \]

For a topological Cartier module $Y$, we consider the fiber of $F = F_Y\colon Y \to Y^{hC_p}$, which we denote $\mathrm {fib}(F)$. As we saw above, $\mathrm {fib}(F)$ is contractible for the topological Cartier module $L_{K(1)} \mathrm {TR}(X)$. Moreover, $\mathrm {fib}(F)$ is $d$-truncated for the topological Cartier module $\mathrm {fib}( \mathrm {TR}(X) \to L_{K(1)} \mathrm {TR}(X))$ because this topological Cartier module is itself $d$-truncated. In particular, we find that $\mathrm {fib}(F\colon \mathrm {TR}(X) \to \mathrm {TR}(X)^{hC_p})$ is $d$-truncated. Taking the cofiber on both the domain and codomain of the Verschiebung, we find that $\mathrm {fib}(F\colon \mathrm {TR}(X) \to \mathrm {TR}(X)^{hC_p}) = \mathrm {fib}(\varphi \colon X \to X^{tC_p})$ which is therefore $d$-truncated as desired.

Combining Proposition 5.2 and Theorem 4.8, we obtain the following result. Versions of the Segal conjecture have been studied by many authors. For instance, it is known that $\mathrm {THH}(\mathbb {Z}_p; \mathbb {F}_p) \to \mathrm {THH}(\mathbb {Z}_p; \mathbb {F}_p)^{tC_p}$ is an equivalence on connective covers [Reference Bökstedt and MadsenBM94, Lemma 6.5]. Compare [Reference Hesselholt and MadsenHM03] for the more general case of a DVR $\mathcal {O}_K$ of mixed characteristic and perfect residue field of characteristic $p>2$. The Segal conjecture for smooth algebras in characteristic $p$ appears as [Reference HesselholtHes18, Proposition 6.6] and (at the filtered level) [Reference Bhatt, Morrow and ScholzeBMS19, Corollary 8.18]. The Segal conjecture has also been verified for certain ring spectra as well; see [Reference Ausoni and RognesAR02, Reference Lunøe-Nielsen and RognesLR11, Reference Angelini-Knoll and QuigleyAQ21].

Corollary 5.3 (The Segal conjecture for regular rings)

Let $R$ be a $p$-torsion-free excellent regular noetherian ring with $R/p$ finitely generated over its $p$th powers. Suppose that for all $\mathfrak {p} \in \mathrm {Spec}(R)$ containing $(p)$, we have $\dim R_{\mathfrak {p}} + \log _p [ \kappa (\mathfrak {p}):\kappa (\mathfrak {p})^p] \leq d$. Then the map $\mathrm {THH}(R; \mathbb {F}_p) \to \mathrm {THH}(R; \mathbb {F}_p)^{tC_p}$ is $(d-1)$-truncated.

Question 5.4 Is it possible to prove a filtered version of this result, with respect to the motivic filtrations [Reference Bhatt, Morrow and ScholzeBMS19] on both sides?

We next discuss the converse direction. Here we only prove the result under a more restrictive hypothesis, namely for cyclotomic spectra which are $ku$-modules.

Example 5.5 Let $C = \widehat {\overline {\mathbb {Q}_p}}$. If $R$ is a $\mathcal {O}_C$-algebra, then the cyclotomic trace makes $\mathrm {THH}(R; \mathbb {Z}_p)$ into a $ku$-module in view of the equivalence $ku_{\hat {p}} \simeq K(\mathcal {O}_C; \mathbb {Z}_p)$; see [Reference NiziołNiz98, Lemma 3.1], [Reference SuslinSus83], and [Reference HesselholtHes06]. One can improve this slightly: $K( \widehat {\mathbb {Z}_p[\zeta _{p^\infty }]}; \mathbb {Z}_p)$ has the structure of an $E_\infty$-algebra under $ku$, so the same applies to any $\widehat {\mathbb {Z}_p[\zeta _{p^\infty }]}$-algebra $R$. This follows from three facts. First, $K( \widehat {\mathbb {Z}_p[\zeta _{p^\infty }]}; \mathbb {Z}_p) \simeq K(\widehat {\mathbb {Q}_p(\zeta _{p^\infty })}; \mathbb {Z}_p)$; see [Reference Hesselholt and NikolausHN19, Lemma 1.3.7] for this argument (which only uses that the field $\widehat {\mathbb {Q}_p(\zeta _{p^\infty })}$ is perfectoid). Second, since $\widehat {\mathbb {Q}_p(\zeta _{p^\infty })}$ is perfectoid, it has $p$-cohomological dimension at most $1$ by the tilting equivalence, whence $K(\widehat {\mathbb {Q}_p(\zeta _{p^\infty })}; \mathbb {Z}_p) \to L_{K(1)}K(\widehat {\mathbb {Q}_p(\zeta _{p^\infty })}; \mathbb {Z}_p)$ is the connective cover map; see Proposition 4.4. Third, $L_{K(1)} K(\mathbb {Z}[\zeta _{p^\infty }])$ has the structure of an $E_\infty$-algebra under $KU$ and hence $ku$; see [Reference Bhatt, Clausen and MathewBCM20, Construction 3.7].

Lemma 5.6 Let $M \in \mathrm {Fun}(BS^1, \mod (H\mathbb {F}_p))$ be an $r$-connected object. Suppose that there exists a map $f\colon M \to M'$ in $\mathrm {Fun}(BS^1, \mod (H\mathbb {F}_p))$ such that $M'$ is induced (as an object with $S^1$-action) and such that $f$ is an equivalence on $\tau _{\geq r}$. Then there exists a map $\tilde {M} \to M$ of $r$-connected objects in $\mathrm {Fun}(BS^1, \mod (H \mathbb {F}_p))$ which induces an equivalence on $\tau _{\geq r+1}$ and such that $\tilde {M}$ is induced.

Proof. The homotopy groups $\pi _*(M), \pi _*(M')$ form graded modules over the ring $\pi _*( \mathbb {F}_p[S^1]) = \mathbb {F}_p[\epsilon ]/\epsilon ^2, \ |\epsilon | = 1$. By assumption, $\pi _*(M')$ is a free graded $\mathbb {F}_p[\epsilon ]/\epsilon ^2$-module and $\pi _*(M)$ is the submodule of those elements in degree $r$ or higher. Choosing an $\mathbb {F}_p$-subspace $V$ of $\pi _r(M) = \pi _r(M')$ which is complementary to $\epsilon \pi _{r-1}(M') \subset \pi _r(M')$, we can modify $M$ to form $\tilde {M}$ with $\pi _r(\tilde {M}) = V$; it is now easy to see that $\pi _*(\tilde {M})$ is a free graded $\mathbb {F}_p[\epsilon ]/\epsilon ^2$-module, so we can conclude.

Proposition 5.7 Let $X$ be a connective, $p$-complete $ku$-module in $\mathrm {CycSp}$. Suppose that $\varphi \colon X \to X^{tC_p}$ is $d$-truncated. Then the fiber of $\mathrm {TR}(X)/p \to L_{K(1)} \mathrm {TR}(X)/p$ is $(d + 3)$-truncated.

Proof. It follows that the cyclotomic spectrum $Y = X/(p, \beta ) = X \otimes _{ku} H \mathbb {F}_p \in \mathrm {CycSp}_{\geq 0}$ has the property that $\varphi \colon Y \to Y^{tC_p}$ is $(d+4)$-truncated. It follows that the comparison map $Y^{C_p} \to Y^{hC_p}$ is $(d+4)$-truncated as well, via the fiber square

Now, for each $n \geq 1$, we have the following pullback diagram [Reference Nikolaus and ScholzeNS18, Lemma II.4.5].

(11)

The bottom horizontal map is $Y^{C_{p^{n-1}}} \to Y^{hC_{p^{n-1}}} \xrightarrow {\varphi ^{hC_{p^{n-1}}}} (Y^{tC_p})^{hC_{p^{n-1}}}$, and the last term is identified with $Y^{tC_{p^n}}$ using the Tate orbit lemma; see [Reference Nikolaus and ScholzeNS18, Lemma II.4.1].

Now $Y^{C_p} \to Y^{hC_p}$ is an equivalence in degrees at least $d+ 6$, hence $Y^{C_{p^{n-1}}} \to Y^{hC_{p^{n-1}}}$ is an equivalence in degrees at least $d+ 6$ by [Reference Nikolaus and ScholzeNS18, Corollary II.4.9] (a generalization of results of Tsalidis [Reference TsalidisTsa98] and Bökstedt, Bruner, Lunø, Nielsen, and Rognes [Reference Bökstedt, Bruner, Lunøe-Nielsen and RognesBBLR14]). Since $\varphi \colon Y \to Y^{tC_p}$ is an equivalence in degrees at least $d+6$, it follows that the bottom horizontal map in (11) is an equivalence in degrees at least $d+ 6$. Note also that $Y^{tC_p}$ is a module over $\mathbb {F}_p^{tC_p}$ in $\mathrm {Fun}(BS^1, \mod (H\mathbb {F}_p))$; since the latter has induced $S^1$-action, it follows that the former does too.

By Lemma 5.6, we can find an at least $(d+6)$-connected $Y'$ with an $S^1$-equivariant map $Y' \to Y$ which induces an equivalence in degrees at least $d + 7$ and such that $Y'$ is induced as an object of $\mathrm {Fun}(BS^1, \mod (H \mathbb {F}_p))$. It follows that the map $Y'^{hC_{p^n}} \to Y^{hC_{p^n}}$ is an equivalence in degrees at least $d + 7$. But the commutative square

now shows that any $\alpha \in \pi _r( Y^{hC_{p^n}})$ for $r \geq d + 7$ has vanishing image in $\pi _r (Y^{tC_{p^n}})$. Using the commutative square (11) again (since the horizontal maps are equivalences in degrees at least $d+6$), it follows that the restriction map $Y^{C_{p^n}} \to Y^{C_{p^{n-1}}}$ induces the zero map in degrees at least $d + 7$. Consequently, taking the inverse limit over $R$, we find that $\mathrm {TR}(Y) \in {\rm Sp} _{\leq d + 6}$, whence the fiber of $\mathrm {TR}(X)/p \to L_{K(1)} \mathrm {TR}(X)/p$ belongs to ${\rm Sp} _{\leq d + 3}$ by Lemma 5.8.

Lemma 5.8 Let $M$ be a $ku$-module spectrum. Then the following assertions are equivalent.

  1. (i) $M/(p, \beta )$ is concentrated in degrees at most $d + 3$.

  2. (ii) The fiber of $M/p \to L_{K(1)} M/p$ is concentrated in degrees at most $d$.

Proof. Given a $ku$-module $N$, it suffices to show that the fiber of $N \to N[1/\beta ]$ (which after $p$-completion is $L_{K(1)} N$) is concentrated in degrees at most $d$ if and only if $N/\beta$ is concentrated in degrees at most $d+3$; then the result will follow by taking $N = M/p$. To prove the claim, first replace $N$ by $\mathrm {fib}( N \to N[1/\beta ])$; thus we may assume that $N$ is actually $\beta$-power torsion.

Then the claim follows from the observation that a $\beta$-power torsion $ku$-module $N$ is concentrated in degrees at most $d$ if and only if $N/\beta$ is concentrated in degrees at most $d+3$. The ‘only if’ direction is evident, so we verify the ‘if’ direction. In fact, if there exists a nonzero $x \in \pi _i(N)$ for $i > d$, then by multiplying by a power of $\beta$ we may assume $\beta x = 0$, whence there exists a nonzero class in $\pi _{i+3}(N/\beta )$ from the long exact sequence.

Remark 5.9 Let $X = \mathrm {THH}(\mathbb {F}_p)$; in this case, we have that $\mathrm {THH}(\mathbb {F}_p) \to \mathrm {THH}(\mathbb {F}_p)^{tC_p}$ is $(-3)$-truncated. This computation is due to Hesselholt and Madsen [Reference Hesselholt and MadsenHM97, § 5], and is refined in [Reference Nikolaus and ScholzeNS18, Appendix IV-4]. Meanwhile $\mathrm {TR}(\mathbb {F}_p) = H\mathbb {Z}_p\to L_{K(1)} \mathrm {TR}(\mathbb {F}_p) = 0$ is $0$-truncated modulo $p$. Thus, the bound of Proposition 5.7 is the best possible.

Proposition 5.10 Let $R_0$ be a $p$-torsion-free perfectoid ring. Let $R$ be a formally smooth $R_0$-algebra (with respect to the $p$-adic topology) of relative dimension $d$. Then the map $\varphi \colon \mathrm {THH}(R; \mathbb {F}_p) \to \mathrm {THH}(R; \mathbb {F}_p)^{tC_p}$ is $(d-3)$-truncated, and the map $\mathrm {TR}(R; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {F}_p)$ is $(d-1)$-truncated.

Proof. Let $X_i, i \in \mathcal {C}$ be a diagram of spectra. Suppose that for each $i \in \mathcal {C}$, the map $X_i \to L_{K(1)} X_i$ is $m$-truncated for some $m$. Then the map $\varprojlim _i X_i \to L_{K(1)} ( \varprojlim _i X_i)$ is $m$-truncated, and $L_{K(1)} ( \varprojlim _i X_i) \to \varprojlim _i L_{K(1)} X_i$ is an equivalence. This follows because $L_{K(1)}(-)$ annihilates bounded-above spectra. Applying the above observation, and using flat descent of $\mathrm {TR}(-; \mathbb {Z}_p)$ (which follows from [Reference Bhatt, Morrow and ScholzeBMS19, § 3]) and $\mathrm {THH}(-; \mathbb {Z}_p), \mathrm {THH}(-; \mathbb {Z}_p)^{tC_p}$, we reduce to the case where $R$ is a $\widehat {\mathbb {Z}_p[\zeta _{p^\infty }]}$-algebra (e.g. using André's lemma; see [Reference Bhatt and ScholzeBS19, Theorem 7.12]), so $\mathrm {THH}(R; \mathbb {Z}_p)$ is a $ku_{\hat {p}}$-module in $\mathrm {CycSp}_{\geq 0}$ (Example 5.5). With this reduction in mind, the first claim implies the second in view of Proposition 5.7 (applied to the cyclotomic spectrum $\mathrm {THH}(R; \mathbb {F}_p)$).

Thus, we prove the first claim (i.e. the Segal conjecture for $\mathrm {THH}(R; \mathbb {F}_p)$); this result appears as [Reference Bhatt, Morrow and ScholzeBMS19, Corollary 9.12] in the case where $R_0$ is the ring of integers in a complete, algebraically closed nonarchimedean field of mixed characteristic, and in [Reference Bhatt, Morrow and ScholzeBMS19, § 6] when $R = R_0$.

Let $(A, I)$ be the perfect prism associated to the perfectoid ring $R_0$, and let $\tilde {\xi } \in I$ be a generator. We use the prismatic cohomology ${{\mathbb {\Delta }}}_{R/A}$ of [Reference Bhatt and ScholzeBS19] and its Nygaard completion $\hat {{{\mathbb {\Delta }}}}_{R/A}$. For a quasisyntomic $R_0$-algebra $S$, there are [Reference Bhatt, Morrow and ScholzeBMS19] complete, exhaustive descending $\mathbb {Z}$-indexed filtrations on $\mathrm {TC}^-(S; \mathbb {Z}_p), \mathrm {TP}(S; \mathbb {Z}_p)$ with

\[ \mathrm{gr}^i \mathrm{TC}^-( S; \mathbb{Z}_p) \simeq \mathcal{N}^{{\geq} i} \hat{{{\mathbb{\Delta}}}}_{S/A}[2i], \quad \mathrm{gr}^i \mathrm{TP}(S; \mathbb{Z}_p) \simeq \hat{{{\mathbb{\Delta}}}}_{S/A}[2i], \]

and the map $\varphi \colon \mathrm {TC}^-(S; \mathbb {Z}_p) \to \mathrm {TP}(S; \mathbb {Z}_p)$ on graded pieces is given by the prismatic divided Frobenius

(12)\begin{equation} \varphi/\tilde{\xi}^i \colon \mathcal{N}^{{\geq} i} \hat{{{\mathbb{\Delta}}}}_{S/A}[2i] \to \hat{{{\mathbb{\Delta}}}}_{S/A}[2i]; \end{equation}

see [Reference Bhatt and ScholzeBS19, § 13] for the comparison between prismatic cohomology and the objects of [Reference Bhatt, Morrow and ScholzeBMS19]. Here we have trivialized the Breuil–Kisin twists involved since we are over the base perfectoid ring $R_0$, and we can compute the Nygaard-completed absolute prismatic cohomology as the Nygaard completed relative prismatic cohomology over $R_0$.

Now the map $\varphi \colon \mathrm {TC}^-(S; \mathbb {Z}_p) \to \mathrm {TP}(S; \mathbb {Z}_p)$ arises by taking $S^1$-invariants on the cyclotomic Frobenius $\varphi \colon \mathrm {THH}(S; \mathbb {Z}_p) \to \mathrm {THH}(S; \mathbb {Z}_p)^{tC_p}$. Both sides of this map are filtered (again, as in [Reference Bhatt, Morrow and ScholzeBMS19]) and the associated graded pieces are

\[ \mathrm{gr}^i \mathrm{THH}(S; \mathbb{Z}_p) \simeq \mathcal{N}^{i} \hat{{{\mathbb{\Delta}}}}_{S/A}[2i], \quad \mathrm{gr}^i \mathrm{THH}(S; \mathbb{Z}_p)^{tC_p} = (\hat{{{\mathbb{\Delta}}}}_{S/A})/\tilde{\xi} [2i], \]

and the cyclotomic Frobenius is induced from (12). But since $R$ is formally smooth over $R_0$, the Nygaard filtration is complete and $\varphi /\tilde {\xi }^i$ induces an equivalence $\mathcal {N}^i {{\mathbb {\Delta }}}_{R/A} \xrightarrow {\sim } \tau ^{\leq i} ( {{\mathbb {\Delta }}}_{R/A}/ \tilde {\xi })$ (see [Reference Bhatt and ScholzeBS19, § 12.4]), where the right-hand side is the $i$th stage of the conjugate filtration on ${{\mathbb {\Delta }}}_{R/A}/\tilde {\xi }$. Using the Hodge–Tate comparison [Reference Bhatt and ScholzeBS19, Theorem 4.10], it follows that the cohomology groups of ${{\mathbb {\Delta }}}_{R/A}/\tilde {\xi }$ are $p$-torsion-free. It follows that

\[ \mathrm{gr}^i \varphi \colon \mathrm{gr}^i \mathrm{THH}(R; \mathbb{F}_p) \to \mathrm{gr}^i \mathrm{THH}(R; \mathbb{F}_p)^{tC_p} \]

exhibits the domain as the $i$-connective cover of the codomain, and therefore has $(i -2)$-truncated homotopy fiber for all $i$. The codomain is $(2i-d)$-connective by the Hodge–Tate comparison, since $R/R_0$ is formally smooth of relative dimension $d$; therefore, $\mathrm {gr}^i \varphi$ is an equivalence for $i \geq d$. Therefore, the fiber of $\varphi \colon \mathrm {THH}(R; \mathbb {F}_p) \to \mathrm {THH}(R; \mathbb {F}_p)^{tC_p}$ has a complete filtration such that $\mathrm {gr}^i$ is $(i-2)$-truncated for all $i$ and contractible for $i \geq d$. This proves that $\varphi \colon \mathrm {THH}(R; \mathbb {F}_p) \to \mathrm {THH}(R; \mathbb {F}_p)^{tC_p}$ is $(d-3)$-truncated, whence the result.

6. Pro-Galois descent

In this section, we prove a type of pro-Galois descent for $\mathrm {TR}$ in the generic fiber, which is related to the conjecture in [Reference HesselholtHes02]. The basic example is when $K$ is a characteristic $0$ local field, and one tries to relate $\mathrm {TR}(\mathcal {O}_K; \mathbb {Z}_p)$ (computed by [Reference Hesselholt and MadsenHM03]) with the ‘continuous’ homotopy fixed points for the Galois group $\mathrm {Gal}(\bar {K}/K)$ on $\mathrm {TR}( \mathcal {O}_{\bar {K}}; \mathbb {Z}_p)$, before or after $K(1)$-localization. The advantage is that the latter is much more tractable; see [Reference HesselholtHes06] for the calculation of $\mathrm {TR}( \mathcal {O}_{\bar {K}}; \mathbb {Z}_p)$. For finite Galois extensions in the generic fiber, these claims follow from § 2. However, there are some additional subtleties to extend to pro-Galois descent because $\mathrm {TR}$ fails to commute with filtered colimits.

6.1 An auxiliary construction

Let $B$ be a base ring. Let $E$ be a $K(1)$-local, localizing invariant on $B$-linear $\infty$-categories which is truncating. Let $R_0$ be a $p$-adically complete $B$-algebra. Given a $K(1)$-local truncating invariant $E$, we now describe a construction of a sheaf on the finite étale site of $R_0[1/p]$. For every finite étale $R_0[1/p]$-algebra $S$, we can choose a ‘ring of integers’ $S_0$ which is finite and finitely presented over $R_0$ with $S_0[1/p] = S$ and consider $E(S_0)$. It is not difficult to see that this only depends on $S$ and that it defines the desired sheaf; to make the functoriality precise, we use left Kan extension.

Construction 6.1 ($E$ as a sheaf on the finite étale site of the generic fiber)

Let $R_0$ be a $p$-adically complete $B$-algebra and let $R = R_0[1/p]$. Using the $K(1)$-local, localizing invariant $E$ which is assumed to be truncating, we define a sheaf $\mathcal {F}_E$ of spectra on the finite étale site of $\mathrm {Spec}(R)$ as follows. Let $\mathcal {C}_0$ denote the category of finite, finitely presented $R_0$-algebras $S_0$ with $S_0[1/p]$ étale over $R$, and let $\mathcal {C}$ denote the category of finite étale $R$-algebras. Consider the functor $F\colon \mathcal {C}_0 \to \mathcal {C}$ given by inverting $p$. Note the following observations.

  1. (i) $F$ is essentially surjective. That is, given any finite étale $R$-algebra $S$, there exists a finite, finitely presented $R_0$-algebra $S_0$ with $S = S_0[1/p]$. In fact, consider any finite $R_0$-subalgebra $S_0' \subset S$ with $S_0'[1/p] = S$. The algebra $S_0'$ is not necessarily finitely presented, but it is a directed colimit of finite, finitely presented $R_0$-algebras $S_0^{(\alpha )}$ under surjective maps; one of them will have $S_0^{(\alpha )}[1/p] = S$, and can be taken for $S_0$.

  2. (ii) If $S \in \mathcal {C}$, then $\mathcal {C}_0 \times _{\mathcal {C}}\mathcal {C}_{/S}$ is a filtered category. In fact, the subcategory of $\mathcal {C}_0 \times _{\mathcal {C}}\mathcal {C}_{/S}$ spanned by those $S_0$ such that the structure map $S_0[1/p] \to S$ is an isomorphism is itself filtered and cofinal. This follows similarly by comparing $S_0$ with its image in $S$.

We consider the functor $S_0 \mapsto E(S_0)$ on the category $\mathcal {C}_0$ of finite, finitely presented $R_0$-algebras $S_0$ with $S_0[1/p]$ finite étale over $R$. Then, to define $\mathcal {F}_E$ on finite étale $R$-algebras, we left Kan extend $E(-)$ along the functor $\mathcal {C}_0 \to \mathcal {C}$. Explicitly, it follows from Example 2.4 that if $S$ is a finite étale $R$-algebra and $S_0$ is a finite, finitely presented $R_0$-algebra with $S_0[1/p] \simeq S$, then we have a canonical equivalence $\mathcal {F}_E(S) \simeq E(S_0)$. It also follows from $h$-descent that $\mathcal {F}_E$ is indeed a sheaf of $K(1)$-local spectra on the finite étale site of $\mathrm {Spec}(R)$. Note that it is a sheaf of modules over the sheaf $S \mapsto L_{K(1)} K(S)$, which is also the sheaf $\mathcal {F}_{L_{K(1)} K}$ because $L_{K(1)} K(-)$ is insensitive to inverting $p$ on $H\mathbb {Z}$-algebras as in (2).

Proposition 6.2 (Hypercompleteness of $\mathcal {F}_E$)

With notation as above, suppose there is a uniform bound on the mod $p$ cohomological dimensions of the residue fields of $R$ and $R$ has finite Krull dimension. Then $\mathcal {F}_E$ defines a hypercomplete sheaf on the finite étale site of $\mathrm {Spec}(R)$.

Proof. Our hypotheses imply that $\mathrm {Spec}(R)$ has finite mod $p$ étale cohomological dimension; see [Reference Clausen and MathewCM19, Corollary 3.29].

Now we use the $K(\pi , 1)$ property, which states that for any $p$-henselian ring $A$ and any $\mathbb {F}_p$-local system $\mathcal {L}$ on $\mathrm {Spec}(A[1/p])$, the cohomology of $\mathcal {L}$ on the étale sites and the finite étale sites of $\mathrm {Spec}(A[1/p])$ coincide. For $A$ noetherian and $p$-complete, this is proved in [Reference ScholzeSch13, Theorem 4.9]. The case of $A$ noetherian and $p$-henselian then follows using the Fujiwara–Gabber theorem (see [Reference Bhatt and MathewBM18, Theorem 6.11] for an account), and then one can pass to the limit to obtain the arbitrary $p$-henselian case using the commutation of cohomology (on either the étale or finite étale sites) and filtered colimits.

From the above two paragraphs, it follows that the étale fundamental group $\pi _1^{\mathrm {et}}( \mathrm {Spec}( R))$ has finite mod $p$ cohomological dimension, which implies that the notion of hypercompleteness for sheaves of $p$-complete spectra on the finite étale site of $\mathrm {Spec}(R)$ can be made explicit in terms of exponents of nilpotence [Reference Clausen and MathewCM19, § 4.1]. Now $\mathcal {F}_E$ is a module over the hypercomplete sheaf $S \mapsto L_{K(1)}K(S)$ (see [Reference Clausen and MathewCM19, Theorem 7.14] for hypercompleteness), hence a hypercomplete sheaf itself, thanks to [Reference Clausen and MathewCM19, Corollary 4.26].

We also need the following variant of the above with respect to a fixed profinite group.

Construction 6.3 ($\mathcal {F}_E$ relative to a profinite group)

Let $S_0$ be an $R_0$-algebra equipped with the action of a profinite group $\Gamma$ which is continuous (with respect to the discrete topology on $S_0$), and such that $R \to S := S_0[1/p]$ is $\Gamma$-Galois. Suppose that there exist a cofinal collection of open normal subgroups $N_i \subset \Gamma$, $i \in I$, for which the fixed points $S_0^{N_i}$ form a finite, finitely presented $R_0$-algebra. It follows from the above that we obtain a sheaf on the category of finite continuous $\Gamma$-sets which carries $\Gamma /N_i \mapsto E( S_0^{N_i})$. This is just the restriction of $\mathcal {F}_E$ to the site of finite continuous $\Gamma$-sets (which maps to the finite étale site of $R$).

When is the sheaf of Construction 6.3 hypercomplete? When $\Gamma$ has finite cohomological dimension, hypercompleteness is smashing [Reference Clausen and MathewCM19, § 4.1], so hypercompleteness holds if the sheaf of spectra which sends $\Gamma /H \mapsto L_{K(1)} K( S^H)$ (for any cofinal collection of open normal subgroups $H$) is hypercomplete.

Example 6.4 Suppose $\Gamma = \mathbb {Z}_p^n$ and the $\Gamma$-extension of $R_0$ is obtained by adding compatible systems of $p$-power roots. Explicitly, suppose $R_0$ is a $\mathbb {Z}[ \zeta _{p^\infty }, t_1^{\pm 1}, \ldots , t_n^{\pm 1}]$-algebra and $S_0 =R_0 \otimes _{\mathbb {Z}[ \zeta _{p^\infty }, t_1^{\pm 1}, \ldots , t_n^{\pm 1}]} \mathbb {Z}[ \zeta _{p^\infty }, t_1^{\pm 1/p^\infty }, \ldots , t_n^{\pm 1/p^\infty }] ,$ with the evident $\Gamma$-action. In this case, the sheaf of Construction 6.3 is hypercomplete. Again since hypercompleteness is smashing, this follows because any $K(1)$-local localizing invariant yields a hypercomplete étale sheaf on the site of finite étale (or even all étale) $\mathbb {Z}[1/p, \zeta _{p^\infty }, t_1^{\pm 1}, \ldots , t_n^{\pm 1}]$-algebras; see [Reference Clausen and MathewCM19, Theorem 7.14]. In particular, we take as the localizing invariant $A \mapsto L_{K(1)} K(A \otimes _{\mathbb {Z}[1/p, \zeta _{p^\infty }, t_1^{\pm 1}, \ldots , t_n^{\pm 1}]} R)$.

6.2 Completion of topological Cartier modules

Here again we use the ‘decompletion’ of the theory of cyclotomic spectra given by the topological Cartier modules of Antieau and Nikolaus [Reference Antieau and NikolausAN21] and an amplitude property of the completion. We recall that $\mathrm {TR}(-)$ gives a fully faithful right adjoint embedding from bounded-below cyclotomic spectra into bounded-below topological Cartier modules, with image the $V$-complete objects; see [Reference Antieau and NikolausAN21, Theorem 3.21].

Proposition 6.5 Let $M$ be a topological Cartier module which is $d$-truncated and bounded below. Then the $V$-completion of $M$ is $(d+3)$-truncated. If $M$ is $p$-complete, then the $V$-completion of $M$ is $(d+2)$-truncated.

Proof. We reduce by dévissage to the case where $M$ is concentrated in degree $0$, and $d = 0$. The completion of $M$ is given by $\mathrm {cofib}( \varprojlim M_{hC_{p^n}} \to M)$ where the maps $M_{hC_{p^n}} \to M_{hC_{p^{n-1}}}$ are given by $V_{hC_{p^{n-1}}}$; see [Reference Antieau and NikolausAN21, Proposition 3.22].

Now since $M$ is discrete, the Verschiebung is simply given by a map $V\colon M \to M$ of abelian groups, and the $S^1$-action is trivial. We can (as in the proof of [Reference Antieau and NikolausAN21, Lemma 3.25]) form a $\mathbb {Z}_{\geq 0} \times \mathbb {Z}_{\geq 0}$-indexed diagram $Y_{i, j} = M \otimes BC_{p^j + }$ such that the transition maps in the $i$-direction are given by $V$ and in the $j$-direction are given by the canonical projections. By the above, the completion of $M$ is given by the cofiber of the map $\varprojlim _{i, j} Y_{i, j} \to M$.

Now a simple computation shows that for any abelian group $A$, $\varprojlim _{j} HA\otimes BC_{p^j + }$ are concentrated in degrees at most $2$, and degrees at most $1$ if $A$ is derived $p$-complete. This claim will imply the result. Indeed, for the first part, it suffices to show that $\varprojlim _{j} HA\otimes BC_{p^j + }$ is concentrated in degrees at most $1$ when $A$ is torsion-free; this follows because the pro abelian group $\{H_*(BC_{p^j}; \mathbb {Z}) \}_{j \geq 0}$ is pro-zero for $\ast \geq 2$. For the second part, we also observe that the pro-abelian group $\{H_1( BC_{p^j}; \mathbb {Z})\}_{j \geq 0}$ is simply the tower $\cdots \to \mathbb {Z}/p^2 \to \mathbb {Z}/p \to 0$, and our assumption that $A$ is derived $p$-complete gives that $A \simeq \varprojlim A \otimes ^{\mathbb {L}}_{\mathbb {Z}} \mathbb {Z}/p^j$ is in particular discrete.

Remark 6.6 We can give another proof of Proposition 6.5 using the results from the previous section in the case where $M$ is annihilated by a power of $p$. By dévissage, we can reduce to the case where $M$ is an $H\mathbb {F}_p$-module (in topological Cartier modules). Let $X = M/M_{hC_p} = \mathrm {cofib}(V)$ be the associated cyclotomic spectrum, so $\mathrm {TR}(X)$ is the derived $V$-completion of $M$ by the correspondence between bounded-below $V$-complete Cartier modules and bounded-below cyclotomic spectra; see [Reference Antieau and NikolausAN21, Theorem 3.21]. Then the proof of Proposition 5.2 shows that the cyclotomic Frobenius $\varphi \colon X \to X^{tC_p}$ is $d$-truncated, since $\mathrm {fib}(\varphi ) = \mathrm {fib}(F\colon M \to M^{hC_p})$. Using Proposition 5.7, we find that $\mathrm {TR}(X)/p$ is $(d+3)$-truncated, whence $\mathrm {TR}(X)$ is $(d+2)$-truncated.

Corollary 6.7 Let $R_i, i \in I$ be a filtered system of rings and let $R = \varinjlim R_i$. Suppose that the map $\mathrm {TR}(R_i; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R_i ; \mathbb {F}_p)$ is $d$-truncated for all $i \in I$. Then $\mathrm {TR}(R; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {F}_p)$ is $(d+2)$-truncated.

Proof. Let $v\colon \Sigma ^u (S^0/p) \to (S^0/p)$ be a $v_1$-self map (so we can take $u = 2p-2$ for $p$ odd and $u = 8$ for $p = 2$). Let $X$ be any spectrum. Then we observe that the following assertions are equivalent.

  1. (i) The map $X/p \to L_{K(1)} X/p$ is $d$-truncated.

  2. (ii) The spectrum $X \otimes S^0/(p, v)$ is $(d+u+1)$-truncated.

The equivalence is proved analogously to Lemma 5.8, noting that $L_{K(1)} (X/p) = X \otimes (S^0/p)[v^{-1}]$ by the telescope conjecture at height $1$ [Reference MahowaldMah81, Reference MillerMil81].

Therefore, it suffices to show that $\mathrm {TR}(R; \mathbb {Z}_p) \otimes S^0/(p, v)$ is $(d + u+3)$-truncated. To this end, let $M = \varinjlim \mathrm {TR}(R_i; \mathbb {Z}_p)$, so $M$ is a connective topological Cartier module which is not necessarily derived $V$-complete; the $V$-completion of its $p$-completion is $\mathrm {TR}(R; \mathbb {Z}_p)$ since $\mathrm {THH}(-)$ commutes with filtered colimits (as a functor from rings to $\mathrm {CycSp}_{\geq 0}$). Now the topological Cartier module $M \otimes S^0/(p, v)$ is $(d + u + 1)$-truncated as a filtered colimit of $(d + u + 1)$-truncated objects. Taking the $V$-completion and using Proposition 6.5, we find that $\mathrm {TR}(R; \mathbb {Z}_p)/(p, v)$ is $(d + u + 3)$-truncated. Therefore, $\mathrm {TR}(R; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {F}_p)$ is $(d+2)$-truncated.

6.3 The main pro-Galois result

In this subsection we prove the following pro-Galois descent result.

Theorem 6.8 (Pro-Galois descent in the generic fiber) Fix $d \geq 0$. Let $R$ be a $p$-complete ring such that $\mathrm {TR}(R; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {F}_p)$ is $d$-truncated. Let $S$ be an $R$-algebra.

Let $G$ be a profinite group of finite $p$-cohomological dimension which acts continuously on the $R$-algebra $S$ (given the discrete topology). Suppose that the following assertions hold.

  1. (i) $R[1/p] \to S[1/p]$ is a $G$-Galois extension.

  2. (ii) There is a cofinal collection of open normal subgroups $N_i \subset G$, $i \in I$, such that $S_i := S^{N_i}$ is a finite, finitely presented $R$-algebra and such that $\mathrm {TR}(S_i; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(S_i; \mathbb {F}_p)$ is $d$-truncated.

  3. (iii) The induced sheaf of spectra on finite continuous $G$-sets given by

    \[T \mapsto L_{K(1)} ( K( \mathrm {Fun}_G(T, S[1/p])))\]
    (for $\mathrm {Fun}_G(T, S[1/p])$ denoting $G$-equivariant functions $T \to S$) is hypercomplete. For example, this holds if $R[1/p]$ has finite Krull dimension and there is a uniform bound on the mod $p$ cohomological dimensions of the residue fields, but also in cases such as Example 6.4.

Then the map

(13)\begin{equation} L_{K(1)}\mathrm{TR}(R) \to (L_{K(1)}\mathrm{TR}(S))^{h G_{\mathrm{cts}}} := \mathrm{Tot} \Big( L_{K(1)} \mathrm{TR}(S) \rightrightarrows L_{K(1)} \mathrm{TR}( \mathrm{Fun}_{\mathrm{cts}}(G, S)) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big)\end{equation}

is an equivalence and the map

\[ \mathrm{TR}(R; \mathbb{F}_p) \to \mathrm{Tot} \Big( \mathrm{TR}(S; \mathbb{F}_p) \rightrightarrows \mathrm{TR}( \mathrm{Fun}_{\mathrm{cts}}(G, S); \mathbb{F}_p) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big) \]

is $(d+2)$-truncated.

Proof. Let $\mathcal {F}$ be any product-preserving presheaf from finite continuous $G$-sets to $p$-torsion spectra. We let $\mathcal {F}^{\mathrm {disc}}$ denote the left Kan extension to profinite $G$-sets, so if $S$ is a profinite $G$-set which can be written $S \simeq \varprojlim S_j$ for some finite continuous $G$-sets $S_j$, then $\mathcal {F}^{\mathrm {disc}}(S) \simeq \varinjlim \mathcal {F}(S_j)$. We let

\[ R \Gamma({\ast}, \mathcal{F}) = \mathrm{Tot} \Big(\mathcal{F}^{\mathrm{disc}}(G) \rightrightarrows \mathcal{F}^{\mathrm{disc}}(G \times G) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big) . \]

As in [Reference Clausen and MathewCM19, § 4.1], $R \Gamma ( \ast , \mathcal {F})$ is the value of the hypersheafification or Postnikov sheafification of $\mathcal {F}$ (with respect to the topology on finite continuous $G$-sets where covering families are jointly surjective ones) at $\ast$. When we work with $p$-torsion spectra, our assumption that $G$ has finite cohomological dimension implies that $R \Gamma (\ast , -)$ commutes with all colimits.

Now we take $\mathcal {F}$ to be the presheaf which sends a finite continuous $G$-set $T$ to $\mathrm {TR}( \mathrm {Fun}_G(T, S); \mathbb {F}_p)$, where $\mathrm {Fun}_G(T, S)$ is (as in the statement) the ring of $G$-equivariant functions $T \to S$. Unwinding the definitions and hypotheses, we find from Corollary 2.5 that if $T = G/N_j$ for one of the distinguished normal subgroups $N_j$, we have that

\[ L_{K(1)}\mathcal{F}({\ast}) \xrightarrow{\sim} \mathrm{Tot} \Big( L_{K(1)}\mathcal{F}(T) \rightrightarrows L_{K(1)}\mathcal{F}(T \times T) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big) \]

is an equivalence. Note $L_{K(1)} \mathcal {F}$ is an example of Construction 6.3 for the profinite group $G$ (though we denote by $R$ the $p$-complete ring). In particular, our assumptions imply hypercompleteness of $L_{K(1)} \mathcal {F}$, so, passing to the limit, we find that

(14)\begin{equation} L_{K(1)} \mathcal{F}({\ast}) \simeq R \Gamma({\ast}, L_{K(1)}\mathcal{F}) . \end{equation}

Note here that $\mathcal {F}^{\mathrm {disc}}$ is not given by $\mathrm {TR}(-; \mathbb {F}_p)$ because $\mathrm {TR}$ does not commute with filtered colimits; instead, for instance, $\mathcal {F}^{\mathrm {disc}}( G) = \varinjlim _{i \in I} \mathrm {TR}(S_i; \mathbb {F}_p)$. By assumption (ii), $\mathcal {F}^{\mathrm {disc}}(G), \mathcal {F}^{\mathrm {disc}}(G \times G) , \ldots$ are spectra with the property that the map to their $K(1)$-localization is $d$-truncated; therefore, $R \Gamma (\ast , \mathcal {F}) \to L_{K(1)}R \Gamma (\ast , \mathcal {F})= R \Gamma (\ast , L_{K(1)} \mathcal {F})$ is $d$-truncated.

Now we apply two-out-of-three to the sequence of maps $\mathcal {F}(\ast ) \to R \Gamma (\ast , \mathcal {F}) \to R \Gamma (\ast , L_{K(1)} \mathcal {F})$. We just showed that the second map is $d$-truncated, while the composite map is by (14) identified with $\mathrm {TR}(R; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {F}_p)$, which is $d$-truncated by assumption. Therefore, by two-out-of-three, we find that $\mathcal {F}(\ast ) \to R \Gamma (\ast , \mathcal {F})$ is $d$-truncated. In other words,

(15)\begin{equation} \mathcal{F}({\ast}) \to \mathrm{Tot} \Big( \mathcal{F}^{\mathrm{disc}}(G) \rightrightarrows \mathcal{F}^{\mathrm{disc}}(G \times G) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big) \end{equation}

is $d$-truncated. Here both sides of (15) have the structure of topological Cartier modules since the forgetful functor from topological Cartier modules to spectra commutes with limits and colimits [Reference Antieau and NikolausAN21, Proposition 3.11].

Now we take the $V$-completion of both sides in (15) (considered as topological Cartier modules). The left-hand side of (15) is already $V$-complete. To analyze the right-hand-side, observe that the totalization in (15) commutes with $(-)_{hC_{p^{n-1}}}$ because $R\Gamma ( \ast , -)$ commutes with colimits. The limit over $n$ in computing the $V$-completion clearly commutes with the totalization. Thus, the $V$-completion of the right-hand side of (15) is given by

\[ \mathrm{Tot} \Big( \mathrm{TR}(S; \mathbb{F}_p) \rightrightarrows \mathrm{TR}( \mathrm{Fun}_{\mathrm{cts}}(G, S); \mathbb{F}_p) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big). \]

Note finally that the right-hand side of (15) is bounded below by the finiteness of the $p$-cohomological dimension. Taking the $V$-completion in (15), we find from Proposition 6.5 that

\[ \mathrm{TR}(R; \mathbb{F}_p) \to \mathrm{Tot} \Big( \mathrm{TR}(S; \mathbb{F}_p) \rightrightarrows \mathrm{TR}( \mathrm{Fun}_{\mathrm{cts}}(G, S); \mathbb{F}_p) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big) \]

is $(d+2)$-truncated, whence the last claim of the theorem. Finally, $\mathrm {TR}(S; \mathbb {F}_p), \mathrm {TR}( \mathrm {Fun}_{\mathrm {cts}}(G, S); \mathbb {F}_p)$ map via $(d+2)$-truncated maps to their $K(1)$-localizations, via Corollary 6.7, so the $K(1)$-local descent statement follows.

Example 6.9 (Discrete valuation rings) Let $K$ be a complete, discretely valued field of characteristic $0$ whose residue field $k$ is of characteristic $p$ with $[k: k^p] \leq p^d$. It follows that if $k'/k$ is any finite extension, then $[k':k'^p] \leq p^d$; see [Reference Gabber and OrgogozoGO08, Lemma 2.1.1].

By the main result of [Reference Gabber and OrgogozoGO08] (which is due to [Reference KatoKat82] in this case), it follows that the Galois group $\mathrm {Gal}(\bar {K}/K)$ has $p$-cohomological dimension at most $d + 2$. Moreover, by Theorem 4.8, it follows that if $L$ is any finite extension of $K$ and $\mathcal {O}_L \subset L$ the ring of integers (which is excellent as a complete local ring), then $\mathrm {TR}( \mathcal {O}_L; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(\mathcal {O}_L; \mathbb {F}_p)$ is $d$-truncated.

We can now apply Theorem 6.8 to conclude that

\[ \mathrm{TR}( \mathcal{O}_K; \mathbb{F}_p) \to \mathrm{TR}(\mathcal{O}_{\bar{K}}; \mathbb{F}_p)^{h \mathrm{Gal}_K^{\mathrm{cts}}} := \mathrm{Tot} \Big( \mathrm{TR}(\mathcal{O}_{\bar{K}}; \mathbb{F}_p) \rightrightarrows \mathrm{TR}(\mathcal{O}_{\bar{K} \otimes_K \bar{K}}; \mathbb{F}_p) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big) \]

is $d$-truncated. Note that the theorem gives a priori that the map is $(d+2)$-truncated, but we can upgrade the conclusion to $d$-truncated as follows: the analogous comparison map with $L_{K(1)} \mathrm {TR}(-)$ everywhere is an equivalence; the maps $\mathrm {TR}( \mathcal {O}_{\bar {K} \otimes _K \cdots \otimes _K\bar {K}}; \mathbb {F}_p) \to \mathrm {TR}( \mathcal {O}_{\bar {K} \otimes _K \cdots \otimes _K\bar {K}}; \mathbb {F}_p)$ are $(-1)$-truncated by Proposition 5.10 since these are $p$-torsion-free rings whose completions are perfectoid; and the map $\mathrm {TR}(\mathcal {O}_K; \mathbb {F}_p) \to L_{K(1)} \mathrm {TR}(\mathcal {O}_K; \mathbb {F}_p)$ is $d$-truncated (Theorem 4.8).

Suppose, in particular, that $k$ is perfect and of characteristic greater than $2$.Footnote 6 In this case, the results of [Reference Hesselholt and MadsenHM03] (recalled in the proof of Theorem 4.1) show that $\mathrm {TR}(\mathcal {O}_K|K; \mathbb {F}_p)$ is the connective cover of its $K(1)$-localization, which is $L_{K(1)} \mathrm {TR}(\mathcal {O}_K; \mathbb {F}_p)$. Again, $\mathcal {O}_{\bar {K}}, \mathcal {O}_{\bar {K} \otimes _K \bar {K}}, \ldots$ have $p$-completions which are perfectoid rings, so for them $\mathrm {TR}(-; \mathbb {F}_p)$ agrees with the connective cover of its $K(1)$-localization (Proposition 5.10). It follows that $\mathrm {TR}(\mathcal {O}_K|K; \mathbb {F}_p) \simeq \tau _{\geq 0} \mathrm {TR}(\mathcal {O}_{\bar {K}} ; \mathbb {F}_p)^{h \mathrm {Gal}_K^{\mathrm {cts}}}$.

Example 6.10 Let $R_0$ be a $p$-torsion-free perfectoid ring containing a system of primitive $p$-power roots of unity. Let $R$ be a formally smooth $R_0$-algebra, which is formally étale over the formal torus (i.e. $p$-completed Laurent polynomial algebra) $R_0 \langle t_1^{\pm 1}, \ldots , t_n^{\pm 1} \rangle$.

Let $G =\mathbb {Z}_p(1)^n$ and let $S = R \otimes _{R_0\langle t_1^{\pm 1}, \ldots , t_n^{\pm 1} \rangle } \varinjlim _r R_0 \langle t_1^{\pm 1/p^r}, \ldots , t_n^{\pm 1/p^r}\rangle$ with the evident $G$-action by roots of unity. Using Proposition 5.10, one sees that the hypothesis of Theorem 6.8(ii) applies. The hypothesis of Theorem 6.8(iii) applies thanks to Example 6.4. It follows that the comparison map (13) is an equivalence. In particular, the $K(1)$-local $\mathrm {TR}$ of $R$ is expressed as the inverse limit of a diagram of the $K(1)$-local $\mathrm {TR}$ of various rings whose $p$-completion is perfectoid.

7. The analog of Thomason's spectral sequence

In this section, we construct in certain cases an analog of Thomason's étale descent spectral sequence for $L_{K(1)} K(-)$ in terms of étale cohomology (see [Reference ThomasonTho85, Reference Thomason and TrobaughTT90] for $L_{K(1)} \mathrm {TR}(-)$). Our construction splits into two parts. First, we give a formula for $\mathrm {TR}$ and its $K(1)$-localization on the category of perfectoid rings. Second, we hypersheafify $L_{K(1)} \mathrm {TR}(-)$ on all rings in the $\mathrm {arc}_{{p}}$-topology and compare $L_{K(1)} \mathrm {TR}(-)$ to this hypersheafification.

7.1 $\mathrm {arc}_{{p}}$-cohomology

In this section we discuss the cohomology with respect to the $\mathrm {arc}_{{p}}$-topology, mentioned briefly in the introduction (Definition 1.6). This is a variant of the following topology (see [Reference Česnavičius and ScholzeČS19, § 2.2.1]).Footnote 7

Definition 7.1 (The $p$-complete $\mathrm {arc}$-topology)

The $p$-complete $\mathrm {arc}$-topology or $\mathrm {arc}_{\hat {p}}$-topology is the finitary Grothendieck topology (see Remark A.1) on the opposite of the category of derived $p$-complete commutative rings such that a map $R \to R'$ is a cover if, for every $p$-complete valuation ring $V$ of rank at most $1$ and map $R \to V$, there is an extension $V \to W$ of $p$-complete valuation rings of rank at most $1$ and a map $R' \to W$ fitting into the following commutative diagram.

We will also consider restrictions of the $\mathrm {arc}_{\hat {p}}$-topology to appropriate subcategories of the category of derived $p$-complete rings, such as the category $\mathrm {Perfd}$ of perfectoid rings [Reference Bhatt, Morrow and ScholzeBMS18, § 3]. The $\mathrm {arc}_{{p}}$-topology is defined similarly, but we only consider rank $1$ $p$-complete valuation rings where $p \neq 0$.

The $\mathrm {arc}_{\hat {p}}$-topology behaves well with respect to perfectoid rings; one knows that the structure presheaf is a sheaf of rings with respect to this topology, and one even has no higher cohomology [Reference Bhatt and ScholzeBS19, § 8]. The $\mathrm {arc}_{{p}}$-topology is a variant where in some sense we also impose derived saturatedness conditions. Note that a functor is an $\mathrm {arc}_{{p}}$-sheaf if and only if it is an $\mathrm {arc}_{\hat {p}}$-sheaf and annihilates any $\mathbb {F}_p$-algebra. Any derived $p$-complete ring can be covered in the $\mathrm {arc}_{\hat {p}}$-topology (respectively, the $\mathrm {arc}_{{p}}$-topology) by a product of $p$-complete absolutely integrally closed rank $1$ valuation rings (respectively, $p$-complete absolutely integrally closed rank $1$ valuation rings where $p \neq 0$).

We now give some examples of $\mathrm {arc}_{{p}}$-cohomology. To begin with, we consider the simplest case of constant sheaves (or $p$-adically constant ones); the result is that one essentially recovers the $p$-adic étale cohomology of the generic fiber.

Construction 7.2 (Constant sheaves in the $\mathrm {arc}_{{p}}$-topology)

Given an abelian group $M$, we can consider the associated constant sheaf of abelian groups in the $\mathrm {arc}_{{p}}$-topology; its value on a derived $p$-complete ring $R$ is given by $H^0_{\mathrm {\acute {e}t}}(\mathrm {Spec}(R[1/p]), M)$. To see this, we observe that the presheaf of abelian groups $R \mapsto H^0_{\mathrm {\acute {e}t}}( \mathrm {Spec}(R[1/p]), M)$ is a sheaf in the $\mathrm {arc}_{{p}}$-topology on derived $p$-complete rings [Reference Bhatt and MathewBM18, Corollary 6.17] (which assumes $M$ torsion; however, this is not necessary since we are only working with $H^0$). It admits a map from the constant presheaf $M$, and the kernel and cokernel vanish locally in the $\mathrm {arc}_{{p}}$-topology.

Construction 7.3 ($p$-adically constant sheaves and Tate twists in the $\mathrm {arc}_{{p}}$-topology)

Consider the sheaf of rings $\mathbb {Z}_p$ in the $\mathrm {arc}_{{p}}$-topology, defined as the $p$-completion of the constant sheaf associated to $\mathbb {Z}_p$, or equivalently as the inverse limit of the constant sheaves $\mathbb {Z}/p^n \mathbb {Z}$; explicitly, $\mathbb {Z}_p(R) = H^0_{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), \mathbb {Z}_p) = H^0_{\mathrm {pro\acute {e}t}}( \mathrm {Spec}(R[1/p]), \mathbb {Z}_p)$. We construct an invertible $\mathbb {Z}_p$-module $\mathbb {Z}_p(1)$ as the $\mathrm {arc}_{{p}}$-sheafification of the presheaf $R \mapsto \mathbb {Z}_p(1)^{\mathrm {pre}}(R) := T_p(R^{\times })$. To check that this is an invertible module, we observe (see [Reference Bhatt and MathewBM18, Proposition 3.30]) that $\mathbb {Z}_p(1)^{\mathrm {pre}}(R)$ is an invertible $\mathbb {Z}_p(R)$-module whenever $R$ is a product of absolutely integrally closed valuation rings of mixed characteristic $(0, p)$. Note that $\mathbb {Z}_p(1)/p$ can equally be described as the $\mathrm {arc}_{{p}}$-sheaf associated to the presheaf $R \mapsto \mu _p(R) = R^{\times }[p]$, since $\mathrm {arc}_{{p}}$-locally all elements have $p$th roots.

Proposition 7.4 ($\mathrm {arc}_{{p}}$-cohomology as $p$-adic vanishing cycles)

Let $R$ be any derived $p$-complete ring. Then there is a natural equivalence

(16)\begin{equation} R \Gamma_{\mathrm{arc}_{{p}}}( \mathrm{Spec}(R), \mathbb{Z}_p(i)) \simeq R \Gamma_{\mathrm{pro\acute{e}t}}( \mathrm{Spec}(R[1/p]), \mathbb{Z}_p(i)), \end{equation}

where $\mathbb {Z}_p(i)$ on the right-hand side refers to the usual Tate twist on the pro-étale site of $\mathrm {Spec}(R[1/p])$.

Proof. We know that the right-hand side is a $D(\mathbb {Z}_p)^{\geq 0}$-valued sheaf for the $\mathrm {arc}_{{p}}$-topology by [Reference Bhatt and MathewBM18, Corollary 6.17], and we obtain a map from the presheaf $R \mapsto T_p( R^{\times })^{\otimes i}$ to the right-hand side. Sheafifying in the $\mathrm {arc}_{{p}}$-topology and $p$-completing, we obtain a map from the left-hand side to the right-hand side as in (16). To see that (16) is an equivalence of $\mathrm {arc}_{{p}}$-sheaves, it suffices to check on products of rank $1$ absolutely integrally closed, $p$-complete valuation rings of mixed characteristic as these form a basis (see [Reference Bhatt and MathewBM18, Proposition 3.29]), which is handled by the next lemma.

Lemma 7.5 Let $R$ be a product $\prod _{t \in T} V_t$ of absolutely integrally closed, $p$-complete valuation rings. Then:

  1. (i) $H^j_{\mathrm {pro\acute {e}t}}( \mathrm {Spec}(R[1/p]), \mathbb {Z}_p(i)) = 0$ for $j > 0$;

  2. (ii) the map $T_p(R^{\times }) \to H^0_{\mathrm {pro\acute {e}t}}( \mathrm {Spec}(R[1/p]), \mathbb {Z}_p(1))$ is an isomorphism.

Proof. It suffices to prove both claims after reducing modulo $p$. Consider the functor $A \mapsto F_j(A) := H^j( \mathrm {Spec}(\hat {A}_p[1/p]; \mathbb {F}_p(i))$. This functor commutes with finite products and filtered colimits by the Gabber–Fujiwara theorem; see [Reference Bhatt and MathewBM18, Theorem 6.11] for an account. To prove that $F_j(R) = 0$ for $j > 0$ which implies (i), it suffices as in [Reference Bhatt and MathewBM18, Corollary 3.17] to show that $F$ vanishes on every ultraproduct of the $\{V_t\}$ for each ultrafilter on $T$. But these ultraproducts are all absolutely integrally closed, $p$-henselian valuation rings, whence (i). Claim (ii) is proved similarly using the map $A^{\times }[p] \to H^0_{\mathrm {pro\acute {e}t}}( \mathrm {Spec}(A[1/p]), \mathbb {F}_p(1))$.

Next we consider the cohomology of the structure presheaf.

Construction 7.6 ($\mathrm {arc}_{{p}}$-cohomology of the structure presheaf)

For a derived $p$-complete commutative ring $R$, we let $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), \mathcal {O})$ denote the $\mathrm {arc}_{{p}}$-cohomology of the structure presheaf.

The $\mathrm {arc}_{{p}}$-topology restricts to a topology on the opposite of $\mathrm {Perfd}$. Our first goal is to identify concretely $\mathrm {arc}_{{p}}$-cohomology on $\mathrm {Perfd}$.

Construction 7.7 (Saturation) Let $R$ be a perfectoid ring. We have the natural quotient $R \to R' := R/\mathrm {rad}((p))$, which is the universal map from $R$ to a perfect $\mathbb {F}_p$-algebra. Note that $R' \otimes ^L_R R'$ is discrete (and equivalent to $R'$) since relative tensor products of perfectoid rings are $p$-completely discrete (and perfectoid).

Let $J = \mathrm {rad}(p)$. We have $J \otimes ^L_R J \xrightarrow {\sim } J$. It follows that one is in the setup of almost mathematics, and we have a derived almostification functor $(-)_* \colon D(R) \to D(R)$, which is also given by $R \text{Hom} _R( J, -)$. We claim that in fact $J$ can be made explicit, and in particular that it has projective dimension at most $1$, so that for any discrete $R$-module $M$, $M_* \in D(R)^{[0, 1]}$. Indeed, there exists an element $\omega \in R$ such that $\omega$ is a unit multiple of $p$ and such that $\omega$ admits a compatible system of $p$-power roots $\{\omega ^{1/p^n}\}_{n \geq 0}$; see [Reference Bhatt, Morrow and ScholzeBMS18, Lemma 3.9]. We claim that as an $R$-module, there is an equivalence

(17)\begin{equation} J \simeq \varinjlim \Big( R \stackrel{\omega^{1-1/p}}{\to} R \stackrel{\omega^{1/p - 1/p^2}}{\to} \ldots \Big). \end{equation}

To see this, we observe first that the filtered colimit on the right-hand side is a submodule of $R$ given by the ideal $J' := {\bigcup} _{n} (\omega ^{1/p^n})$ (i.e. given by multiplication by $\omega ^{1/p^n}$ on the $n$th term); we can see this by $\mathrm {arc}_{\hat {p}}$-descent to reduce to the case where $R$ is a product of valuation rings of rank at most $1$, in which case the claim is clear. Now clearly $J' \subset J$ and $R/ J'$ is a ring of characteristic $p$ on which the Frobenius is surjective. To obtain $J' = J$, we use [Reference Bhatt, Morrow and ScholzeBMS18, Lemma 3.10] to see that the Frobenius induces an isomorphism $R/\omega ^{1/p^n} \xrightarrow {\sim } R/\omega ^{1/p^{n-1}}$ for $n \geq 1$. This implies that $R/J'$ is perfect, whence $J = J'$ as desired.

Definition 7.8 (Spherically complete fields) We recall (see [Reference van RooijvR78, Chapter 4] for an account) that a nonarchimedean field is called spherically complete if every decreasing sequence of closed disks has nonempty intersection; in particular, any such field is complete. Any nonarchimedean field admits an extension which is spherically complete and algebraically closed.

Lemma 7.9 Let $C$ be spherically complete with ring of integers $\mathcal {O}_C$. Given any tower $\{M_i\}_{i \geq 1}$ of cyclic $\mathcal {O}_C$-modules, we have $\varprojlim ^1 M_i = 0$.

Proof. Writing $\{M_i\}$ as the cokernel of an injective map $\{N_i\}_{i \geq 1} \to \{N_i'\}_{i \geq 1}$ of inverse systems with the $N_i'$ levelwise isomorphic to $\mathcal {O}_{C}$ and using the long exact sequence, we reduce to the case that the $M_i$ are individually isomorphic to $\mathcal {O}_C$ (in particular, torsion-free). Moreover, upon passing to a cofinal subfamily, we may assume the $\{M_i\}_{i \geq 1}$ form a descending sequence of cyclic ideals $\{I_i\}_{i \geq 1} \subset \mathcal {O}_C$; using the short exact sequence of inverse systems $0 \to \{I_i\}_{i \geq 1} \to \{\mathcal {O}_C\}_{i \geq 1} \to \{\mathcal {O}_C/I_i\}_{i \geq 1} \to 0$ (with the middle sequence constant), we see that it suffices to show that $\mathcal {O}_C \to \varprojlim _i \mathcal {O}_C/I_i$ is surjective. But this is precisely the definition of spherical completeness.

Proposition 7.10 ($\mathrm {arc}_{{p}}$-cohomology of perfectoids)

The functor $R \mapsto R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), \mathcal {O})$ restricted on $\mathrm {Perfd}$ agrees with $R \mapsto R_*$. In particular, for $R \in \mathrm {Perfd}$, $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), \mathcal {O}) \in D(R)^{[0, 1]}$.

Proof. Let $R$ be a perfectoid ring. First, we claim that $\mathrm {arc}_{{p}}$-cohomology of $R$ with $\mathcal {O}$-coefficients can be calculated on either all derived $p$-complete rings or the subcategory of perfectoid rings (endowed with the $\mathrm {arc}_{{p}}$-cohomology). To see this, we use that perfectoid rings form a basis for the $\mathrm {arc}_{{p}}$-topology. In fact, any ring admits an $\mathrm {arc}_{{p}}$-cover by a semiperfectoid ring $S$, and then one can take the perfectoidization [Reference Bhatt and ScholzeBS19, § 8] of $S$, which is an $\mathrm {arc}_{{p}}$-cover of $S$ (e.g. maps of $S$ or its perfectoidization into $p$-complete absolutely integrally closed valuation rings are the same). The rest of the claim is a general argument following from the fact that perfectoid rings form a basis for the $\mathrm {arc}_{{p}}$-topology; see Proposition A.6 below.

Now it suffices to prove that $R \mapsto R_*$ is a $D(\mathbb {Z})^{\geq 0}$-valued $\mathrm {arc}_{{p}}$-sheaf on $\mathrm {Perfd}$ and that the map $R \to R_*$ is locally an equivalence. The fact that it is an $\mathrm {arc}_{{p}}$-sheaf follows because the structure presheaf is a $D(\mathbb {Z})^{ \geq 0}$-valued sheaf on the $\mathrm {arc}_{\hat {p}}$-topology on $\mathrm {Perfd}$ and $R \mapsto R_*$ annihilates perfect $\mathbb {F}_p$-algebras. The fact that $R \to R_*$ is locally an equivalence follows by taking an $\mathrm {arc}_{{p}}$-cover by a product of rings of integers in various spherically complete, algebraically closed nonarchimedean fields of mixed characteristic; for such rings, $R = R_*$.

Let $R_0$ be a perfectoid base ring, and let $R$ be a derived $p$-complete $R_0$-algebra. One has the construction of the perfectoidization $R_{\mathrm {perfd}}$ of $R$, a coconnective $E_\infty$-algebra under $R$ which has the property that if $R$ is semiperfectoid, then $R_{\mathrm {perfd}}$ is discrete and is the universal perfectoid ring that $R$ maps to. As shown in [Reference Bhatt and ScholzeBS19, § 8], $R \mapsto R_{\mathrm {perfd}}$ is the $\mathrm {arc}_{\hat {p}}$-cohomology of the structure presheaf on the category of derived $p$-complete $R_0$-algebras.

Proposition 7.11 ($\mathrm {arc}_{{p}}$-cohomology as saturated perfectoidization)

For $R$ an $R_0$-algebra which is derived $p$-complete, we have a natural equivalence of $E_\infty$-algebras in $D(R_0)$, $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), \mathcal {O}) \simeq (R_{\mathrm {perfd}})_*$ (where almost mathematics is taken relative to $R_0$ and the ideal $\mathrm {rad}(p)$).

Proof. This is proved similarly to Proposition 7.10. We have a natural (in $R$) map $R \to (R_{\mathrm {perfd}})_*$. To show that it is the $\mathrm {arc}_{{p}}$-sheafification, it suffices to show that the codomain is an $\mathrm {arc}_{{p}}$-sheaf, and that the map $R \to (R_{\mathrm {perfd}})_*$ has cofiber which vanishes locally in the $\mathrm {arc}_{{p}}$-topology; note also that we can compute the $\mathrm {arc}_{{p}}$-cohomology either over all derived $p$-complete rings or over $R_0$-algebras (Example A.7). The codomain is an $\mathrm {arc}_{\hat {p}}$-sheaf because of the identification of perfectoidization with $\mathrm {arc}_{\hat {p}}$-cohomology [Reference Bhatt and ScholzeBS19, § 8], and hence is an $\mathrm {arc}_{{p}}$-sheaf since it annihilates $\mathbb {F}_p$-algebras. The cofiber of $R \to (R_{\mathrm {perfd}})_*$ vanishes locally in the $\mathrm {arc}_{{p}}$-topology, as one sees by working with perfectoid $R$ and using the argument of Proposition 7.10.

Example 7.12 Suppose $R$ is the $p$-completion of a ring which is integral over the perfectoid ring $R_0$. Then the perfectoidization $R_{\mathrm {perfd}}$ is discrete [Reference Bhatt and ScholzeBS19, Theorem 10.11]. It thus follows that $R\Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), \mathcal {O}) \in D(R)^{[0, 1]}$.

Construction 7.13 (Witt vector cohomology in the $\mathrm {arc}_{{p}}$-topology)

We consider the presheaf $W(\mathcal {O})$ given by $R \mapsto W(R)$ and its $\mathrm {arc}_{{p}}$-cohomology (with Tate twists) $R \Gamma _{\mathrm {arc}_{{p}}}(\mathrm {Spec}(R), W(\mathcal {O})(i))$. Since the Witt vector functor is endowed with Frobenius and Verschiebung operators, so is the construction $R \mapsto R \Gamma _{\mathrm {arc}_{{p}}}(\mathrm {Spec}(R), W(\mathcal {O})(i))$.

In our setting, we can think of the Witt vector cohomology considered above as a one-parameter (along $V$) deformation of the structure presheaf cohomology, especially in light of the next result.

Proposition 7.14 For any ring $R$, $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), W(\mathcal {O})(i))$ is $p$-complete and complete with respect to the Verschiebung.

Proof. By base-changing to $\mathbb {Z}_p[\zeta _{p^\infty }]$, we may assume without loss of generality that $i = 0$. Then this follows from Proposition A.10 below, since the presheaf $W(\mathcal {O})$ commutes with finite products and is complete with respect to $(p, V)$.

Corollary 7.15 Let $R$ be finite and finitely presented over a perfectoid ring. Then $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), W(\mathcal {O})(i)) \in D(\mathbb {Z}_p)^{[0, 1]}$ for any $i$.

Proof. By $p$-completeness and $V$-completeness (Proposition 7.14), it suffices to prove the analogous statement for $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), \mathcal {O}(i)/p)$. Replacing $R$ with $R[\zeta _p]$ and using $(\mathbb {Z}/p)^{\times }$-Galois descent along this extension, we may assume that $R$ contains a primitive $p$th root of unity. This lets us reduce to the case $i = 0$, whence the result follows from Example 7.12.

We note, finally, that one can recover the $p$-adic nearby cycles (see Proposition 7.4) as the fixed points of Frobenius on Witt vector $\mathrm {arc}_{{p}}$-cohomology. This will not be used in the sequel.

Proposition 7.16 For any ring $R$, there is a natural equivalence for each $i$,

(18)\begin{equation} R \Gamma_{\mathrm{arc}_{{p}}}( \mathrm{Spec}(R), \mathbb{Z}_p(i)) \simeq R \Gamma_{\mathrm{arc}_{{p}}}( \mathrm{Spec}(R), W(\mathcal{O})(i))^{F = 1}. \end{equation}

Proof. We have a natural map $\mathbb {Z}_p \to W(\mathcal {O})^{F=1}$ of presheaves. Twisting and sheafifying in the $\mathrm {arc}_{{p}}$-topology, we obtain the map from left to right in (18). To see that it is an equivalence, it suffices to check that it is an equivalence on products (over some indexing set $T$) of rings of the form $\mathcal {O}_C$, for $C$ spherically complete and algebraically closed of mixed characteristic $(0, p)$; now we can trivialize the Tate twists, so can assume $i = 0$. Using Lemma 7.5, we find that the left-hand side is discrete and simply given by $\prod _T \mathbb {Z}_p$. The right-hand side is also given by $\prod _T \mathbb {Z}_p$ by Lemma 7.17 below.

Lemma 7.17 Let $C$ be an algebraically closed complete nonarchimedean field. Then the natural map $\mathbb {Z}_p \to W(\mathcal {O}_C)^{F=1}$ is an equivalence.

Proof. It suffices to work modulo $p$, that is, to show that the natural map $\mathbb {F}_p \to (W(\mathcal {O}_C)/p)^{F = 1}$ is an equivalence. But the Witt vector Frobenius reduces modulo $p$ to the ordinary Frobenius on the ring $W(\mathcal {O}_C)/p$. By the Artin–Schreier sequence, $(W(\mathcal {O}_C)/p)^{F = 1}$ is the mod $p$ étale cohomology of $W(\mathcal {O}_C)/p$. This is unchanged by taking the quotient by the ideal $V W(\mathcal {O}_C)/p$, which squares to zero; therefore, it is also the étale cohomology of $\mathcal {O}_C/p$, or equivalently of the residue field $k$; this in turn is $\mathbb {F}_p$ as $C$ is algebraically closed and hence so is $k$.

7.2 $\mathrm {THH}^{C_{p^r}}$ of perfectoid rings

Here we calculate the fixed points of $\mathrm {THH}$ of perfectoid rings, generalizing the main result of [Reference HesselholtHes06]; these results are known to experts. Recall that for any $p$-complete ring $R$, we have $\pi _0 \mathrm {TR}(R; \mathbb {Z}_p) \simeq W(R)$; see [Reference Hesselholt and MadsenHM97, Theorem F].

Our strategy is to treat the case of a $p$-torsion-free perfectoid ring by a direct spectral sequence argument and appeal to known results for a perfect $\mathbb {F}_p$-algebra; we will glue both cases together using an excision argument. To begin with, we consider the characteristic $p$ case. For perfect fields, this result appears in [Reference Hesselholt and MadsenHM97, § 5].

Lemma 7.18 Let $R$ be a perfect $\mathbb {F}_p$-algebra. Then:

  1. (i) $\pi _*(\mathrm {THH}(R; \mathbb {Z}_p)^{C_{p^r}}) \simeq W_{r+1}(R)[\sigma _r]$ for $|\sigma _r| = 2$;

  2. (ii) $\mathrm {TR}(R; \mathbb {Z}_p) \simeq H W(R)$.

Proof. The Segal conjecture holds for $R$: in fact, $\varphi \colon \mathrm {THH}(R; \mathbb {Z}_p) \to \mathrm {THH}(R ; \mathbb {Z}_p)^{tC_p}$ is $(-3)$-truncated; see [Reference Bhatt, Morrow and ScholzeBMS19, § 6]. Therefore, we have that $\mathrm {THH}(R; \mathbb {Z}_p)^{C_{p^r}} \to \mathrm {THH}(R; \mathbb {Z}_p)^{hC_{p^r}} \xrightarrow {\varphi ^{hC_{p^r}}}\mathrm {THH}(R; \mathbb {Z}_p)^{tC_{p^{r+1}}}$ are equivalences on connective covers; see [Reference Nikolaus and ScholzeNS18, Corollary II.4.9]. Using the homotopy fixed point spectral sequence (or [Reference Bhatt, Morrow and ScholzeBMS19, § 6]), we find $\pi _*( \mathrm {THH}(R; \mathbb {Z}_p)^{tS^1}) \simeq W(R)[ \sigma ^{\pm 1}]$. Now $\mathrm {THH}(R; \mathbb {Z}_p)^{tC_{p^{r+1}}} \simeq \mathrm {THH}(R; \mathbb {Z}_p)^{tS^1}/p^{r+1}$ by [Reference Nikolaus and ScholzeNS18, Lemma IV.4.12]. Combining these facts, we see that claim (i) follows as $W_{r+1}(R) = W(R)/p^{r+1}$.

For claim (ii), note that $\mathrm {THH}(R; \mathbb {Z}_p)$ is a module in $\mathrm {CycSp}$ over $K(\mathbb {F}_p; \mathbb {Z}_p) \simeq H\mathbb {Z}_p$. By Proposition 5.7, we have that $\mathrm {TR}(R; \mathbb {Z}_p) \to L_{K(1)} \mathrm {TR}(R; \mathbb {Z}_p)$ is $0$-truncated; since the target vanishes, it follows that $\mathrm {TR}(R; \mathbb {Z}_p)$ is $0$-truncated, whence the result as we know $\pi _0\mathrm {TR}(R; \mathbb {Z}_p)$.

In the following, we will say that an $E_\infty$-ring $A$ is weakly even periodic if the odd homotopy groups of $A$ vanish, and if for all $m, n \in \mathbb {Z}$, the map $\pi _{2m}(A) \otimes _{\pi _0(A)} \pi _{2n}(A) \to \pi _{2(n+m)}(A)$ is an isomorphism; in particular, $\pi _2(A)$ is an invertible $\pi _0(A)$-module, and Zariski locally $\pi _*(A)$ is a Laurent polynomial algebra over $\pi _0(A)$ on a degree $2$ class.

Lemma 7.19 Let $A, B, C$ be weakly even periodic $E_\infty$-rings and fix maps $A\to C, B \to C$ such that $\pi _0(A) \to \pi _0(C)$ is surjective. Then $A \times _C B$ is weakly even periodic.

Proof. It follows that $\pi _*(A) \otimes _{\pi _0(A)} \pi _0(C) \xrightarrow {\sim } \pi _*(C)$ and, similarly, $\pi _*(A) \otimes _{\pi _0(A)} \pi _0(B) \xrightarrow {\sim } \pi _*(B)$. Now since $\pi _0(A \times _C B), \pi _0(A), \pi _0(B), \pi _0(C)$ form a Milnor square, the category of finitely generated projective $\pi _0(A \times _C B)$-modules is the homotopy pullback of the categories of finitely generated projective $\pi _0(A)$-modules, $\pi _0(B)$-modules, and $\pi _0(C)$-modules; see [Reference MilnorMil71, § 2]. The result now follows.

Construction 7.20 ($p$-torsion-free quotients of perfectoid rings)

Let $R$ be a perfectoid ring, and let $I \subset R$ be the ideal of $p$-power torsion. Then $p I = 0$; see [Reference Bhatt, Morrow and ScholzeBMS19, Proposition 4.19].

We have in fact that $R/I$ is perfectoid as well. To see this, we use the theory of perfectoidizations [Reference Bhatt and ScholzeBS19, §§ 7 and 8]. Let $R' = (R/p)_{\mathrm {perf}}$. For every $p$-complete valuation ring $V$ with a map $R \to V$ which does not annihilate $I \subset R$, we find that $V$ has characteristic $p$ and we obtain a unique extension $R' \to V$. Therefore, applying [Reference Bhatt and ScholzeBS19, Corollary 8.11], we obtain a pullback square of perfectoid rings

(19)

Since the vertical arrows are surjective (see [Reference Bhatt and ScholzeBS19, Theorem 7.4]), this is a Milnor square of rings. Since the terms on the right-hand side are both $\mathbb {F}_p$-algebras, it follows that the kernel of $R \to (R/I)_{\mathrm {perfd}}$ is annihilated by $p$; since it contains $I$, it must be equal to $I$ and we have $(R/I)_{\mathrm {perfd}} = R/I$.

In particular, this shows that any perfectoid ring naturally fits into a Milnor square involving a $p$-torsion-free perfectoid ring and a map of perfect $\mathbb {F}_p$-algebras. The diagram (19) is in addition a homotopy pushout square of $E_\infty$-rings, that is, there are no higher $\mathrm {Tor}$ terms, since everything involved is perfectoid. It follows by [Reference Land and TammeLT19] that (19) induces a pullback after applying any localizing invariant.

Proposition 7.21 Let $R$ be a perfectoid ring. Then, for each $r \geq 1$, we have the following assertions.

  1. (i) $\pi _{*} \mathrm {THH}(R; \mathbb {Z}_p)^{C_{p^r}}$ is concentrated in even degrees.

  2. (ii) $\pi _{2} \mathrm {THH}(R; \mathbb {Z}_p)^{C_{p^r}}$ is an invertible module over $\pi _0 \mathrm {THH} (R; \mathbb {Z}_p)^{C_{p^r}} \simeq W_{r+1}(R)$.

  3. (iii) The multiplication map $\mathrm {Sym}^t (\pi _{2} \mathrm {THH}(R; \mathbb {Z}_p)^{C_{p^r}}) \to \pi _{2t}\mathrm {THH}(R; \mathbb {Z}_p)^{C_{p^r}}$ is an isomorphism for all $t \geq 0$.

Proof. Recall [Reference Bhatt, Morrow and ScholzeBMS19, § 6] that we have a noncanonical isomorphism $\pi _* \mathrm {THH}(R; \mathbb {Z}_p)^{tC_p} \simeq R[\sigma ^{\pm 1}]$ and that $\varphi \colon \mathrm {THH}(R; \mathbb {Z}_p) \to \mathrm {THH}(R; \mathbb {Z}_p)^{tC_p}$ exhibits the source as the connective cover of the target. Therefore, we have that $\mathrm {THH}(R; \mathbb {Z}_p)^{C_{p^r}} \to \mathrm {THH}(R; \mathbb {Z}_p)^{hC_{p^r}} \xrightarrow {\varphi ^{hC_{p^r}}}\mathrm {THH}(R; \mathbb {Z}_p)^{tC_{p^{r+1}}}$ are equivalences on connective covers; see [Reference Nikolaus and ScholzeNS18, Corollary II.4.9].

Our next claim is that $\mathrm {THH}(R; \mathbb {Z}_p)^{tC_{p^{r+1}}}$ is a weakly even periodic $E_\infty$-ring for any $r \geq 0$. If $R$ is $p$-torsion-free, this follows from the (degenerate) Tate spectral sequence applied to $\mathrm {THH}(R; \mathbb {Z}_p)$, whose homotopy groups form a polynomial algebra over $R$ on a class in degree $2$. If $R$ is an $\mathbb {F}_p$-algebra, the claim also follows from Lemma 7.18. We now treat the case of a general perfectoid $R$. In view of the Milnor square (19) which is also a homotopy pushout of $E_\infty$-rings and the main result of [Reference Land and TammeLT19], we find that $\mathrm {THH}(-; \mathbb {Z}_p)^{tC_{p^{r+1}}}$ carries (19) to a pullback of $E_\infty$-ring spectra. Moreover, on $\pi _0(-)$ this square yields a Milnor square since $\pi _0 \mathrm {THH}(-; \mathbb {Z}_p)^{tC_{p^{r+1}}} \simeq W_{r+1}(-)$ for perfectoids via the previous paragraph and [Reference Hesselholt and MadsenHM97, Theorem F]. Via Lemma 7.19, the claim of weak even periodicity follows.

Now the above paragraphs show that $\mathrm {THH}(R; \mathbb {Z}_p)^{C_{p^r}}$ is the connective cover of a weakly even periodic $E_\infty$-ring, and we already know its $\pi _0$ is given by $W_{r+1}(R)$, whence the result.

7.3 Description of $\mathrm {TR}$ of perfectoids

Let $C$ be spherically complete and algebraically closed of mixed characteristic $(0, p)$. Then we recall some of the additional rings attached to $C$. We have the tilt ${C}^{\flat }$, the ring of integers $\mathcal {O}_{C^{\flat }} \subset C^{\flat }$, and the Fontaine ring $A_{\inf } = A_{\inf }(\mathcal {O}_C) = W(\mathcal {O}_{C^{\flat }})$. Choosing a compatible system $(1, \zeta _p, \zeta _{p^2}, \ldots )$ of primitive $p^n$th roots in $C$, we obtain an element $\epsilon \in \mathcal {O}_{C^{\flat }}$ and let $[\epsilon ]$ denote the corresponding element of $A_{\inf }$. We have the map $\theta \colon A_{\inf } \to \mathcal {O}_C$, which exhibits the source as the universal $p$-adically pro-nilpotent thickening of the latter; in particular, this leads to a map $A_{\inf } \to W(\mathcal {O}_C)$, which is a surjection in this case [Reference Bhatt, Morrow and ScholzeBMS18, Lemma 3.23].

Lemma 7.22 ($\mathrm {TR}$ in the spherically complete case)

Let $C$ be a spherically complete, algebraically closed nonarchimedean field of mixed characteristic $(0, p)$ and let $\mathcal {O}_C \subset C$ be the ring of integers. Let $\beta \in \pi _2 \mathrm {TR}(\mathcal {O}_C; \mathbb {Z}_p)$ be the Bott element (arising from the image of the cyclotomic trace). Then $\mathrm {TR}_*(\mathcal {O}_C; \mathbb {Z}_p) \simeq W(\mathcal {O}_C)[\beta ]$.

Proof. Since everything is $p$-complete, it suffices to see that $\mathrm {TR}(\mathcal {O}_C; \mathbb {F}_p)/\beta$ is $0$-truncated. Now for each $n$, each homotopy group in $\pi _*(\mathrm {THH}^{C_{p^n}}(\mathcal {O}_C; \mathbb {F}_p))$ is a cyclic module over $W(\mathcal {O}_C)/p$ (Proposition 7.21), which in turn is a quotient of $A_{\inf }(\mathcal {O}_C)/p = \mathcal {O}_{C^{\flat }}$. Moreover, the homotopy groups are concentrated in even degrees.

Now $C^{\flat }$ is spherically complete; see the proof of [Reference Bhatt, Morrow and ScholzeBMS18, Lemma 3.23]. Using the previous paragraph, Lemma 7.9, and the Milnor exact sequence, it follows that $\pi _*( \mathrm {TR}( \mathcal {O}_C; \mathbb {F}_p))$ is concentrated in even degrees. By Lemma 5.8 and Proposition 5.10, we also find that the cofiber of $\beta$ is $2$-truncated, so $\pi _*( \mathrm {TR}( \mathcal {O}_C; \mathbb {F}_p)/\beta )$ is concentrated in degrees $0$ and $2$, and is given by $W(\mathcal {O}_C)/p$ in degree $0$. It thus suffices to show that the zeroth Postnikov section $\mathrm {TR}(\mathcal {O}_C; \mathbb {Z}_p)/\beta \to H W(\mathcal {O}_C)$ induces an isomorphism on $\pi _0 (-)^{tS^1}$. But we have by [Reference Antieau and NikolausAN21, Corollary 10] that $\mathrm {TR}(\mathcal {O}_C; \mathbb {Z}_p)^{tS^1} = \mathrm {THH}(\mathcal {O}_C; \mathbb {Z}_p)^{tS^1}$, and on homotopy groups this is $A_{\inf }[\sigma ^{\pm 1}]$ (see [Reference Bhatt, Morrow and ScholzeBMS19, § 6]). Moreover, with respect to the above identifications, we have that $\beta$ is a $\mathbb {Z}_p^{\times }$-multiple of $( [\epsilon ] -1 )\sigma$; see [Reference Hesselholt and NikolausHN19, Theorem 1.3.6]. Therefore, we have $\pi _0 ( \mathrm {TR}(\mathcal {O}_C; \mathbb {Z}_p)/\beta )^{tS^1} \simeq A_{\inf }/( [\epsilon ] - 1) [\sigma ^{\pm 1}]$, whence the result since $A_{\inf }/( [\epsilon ] -1 ) \xrightarrow {\sim } W(\mathcal {O}_C)$ by [Reference Bhatt, Morrow and ScholzeBMS18, Lemma 3.23].

Proposition 7.23 ($\mathrm {TR}$ of perfectoid rings)

For $R \in \mathrm {Perfd}$ and for $i > 0$, we have natural equivalences

(20)\begin{equation} \tau_{[2i-1, 2i]} \mathrm{TR}(R; \mathbb{Z}_p) \simeq R \Gamma_{\mathrm{arc}_{{p}}}( \mathrm{Spec}(R), W(\mathcal{O})(i))[2i]. \end{equation}

For $R \in \mathrm {Perfd}$ we have, for $i \in \mathbb {Z}$,

(21)\begin{equation} \tau_{[2i-1, 2i]} \big( L_{K(1)}\mathrm{TR}(R; \mathbb{Z}_p) \big) \simeq R \Gamma_{\mathrm{arc}_{{p}}}( \mathrm{Spec}(R), W(\mathcal{O})(i))[2i] . \end{equation}

Proof. The structure presheaf on $\mathrm {Perfd}$ defines an $\mathrm {arc}_{\hat {p}}$-sheaf [Reference Bhatt and ScholzeBS19, Proposition 8.9]. Taking the inverse limit in Proposition 7.21 over the restriction maps (and accounting for possible $\lim ^1$ terms), we find in particular that $R \mapsto \tau _{[2i-1, 2i]} \mathrm {TR}(R; \mathbb {Z}_p)$ defines an $\mathrm {arc}_{\hat {p}}$-sheaf of spectra on $\mathrm {Perfd}$. Since, for $i > 0$, this functor annihilates perfect $\mathbb {F}_p$-algebras, it in fact defines an $\mathrm {arc}_{{p}}$-sheaf on $\mathrm {Perfd}$.

The cyclotomic trace gives a map of $\mathbb {Z}_p$-modules $T_p( R^{\times }) \to \pi _2 \mathrm {TR}(R; \mathbb {Z}_p)$. Since $\pi _0( \mathrm {TR}(R; \mathbb {Z}_p) ) \simeq W(R)$, we obtain natural maps $f_i \colon W(R) (i)\to \pi _{2i}( \mathrm {TR}(R; \mathbb {Z}_p))$ for $i > 0$. To complete the proof, it suffices to show that:

  1. (a) $f_i$ is an isomorphism for a cofinal (in the $\mathrm {arc}_{{p}}$-topology) collection of $R \in \mathrm {Perfd}$;

  2. (b) $\pi _{2i-1}( \mathrm {TR}(R; \mathbb {Z}_p)) =0$ for a cofinal (in the $\mathrm {arc}_{{p}}$-topology) collection of $R \in \mathrm {Perfd}$.

We will take this collection to be the set of products of perfectoid rings of the form $\mathcal {O}_C$, for $C$ a spherically complete and algebraically closed nonarchimedean field of mixed characteristic $(0, p)$. First, Proposition 7.21 shows that the construction $R \mapsto \mathrm {THH}(R; \mathbb {Z}_p)$ commutes with products for $R \in \mathrm {Perfd}$; inductively and taking the limit, we find that $R \mapsto \mathrm {TR}(R; \mathbb {Z}_p)$ commutes with products for $R \in \mathrm {Perfd}$. Thus, it suffices to verify the above claims for $R = \mathcal {O}_C$ itself, and that follows from Lemma 7.22.

For the claim about $L_{K(1)} \mathrm {TR}(R; \mathbb {Z}_p)$, note first that the first part of the proof and Proposition 5.10 (and (19) to reduce to the $p$-torsion-free case) show easily that $R \mapsto L_{K(1)} \mathrm {TR}(R; \mathbb {Z}_p)$ defines an $\mathrm {arc}_{{p}}$-sheaf on $\mathrm {Perfd}$ (see also the remarks at the beginning of the proof of Proposition 5.10, concerning when $L_{K(1)}(-)$ commutes with homotopy limits). It then follows from the above (by inverting the Bott element over perfectoid rings containing $p$-power roots of unity) that the homotopy groups of the $\mathrm {arc}_{{p}}$-sheaf $L_{K(1)} \mathrm {TR}(R; \mathbb {Z}_p)$ on $\mathrm {Perfd}$ are given by $\pi _{2i} \simeq W(\mathcal {O})(i)$, and we obtain a filtration with associated graded terms the right-hand side of (21) via the Postnikov filtration as $\mathrm {arc}_{{p}}$-sheaves. Note that the Postnikov filtration is always exhaustive, and it is complete because it is complete on the $\mathrm {arc}_{{p}}$-sheaf $R \mapsto \tau _{\geq 1} L_{K(1)} \mathrm {TR}(R) \simeq \tau _{\geq 1} \mathrm {TR}(R; \mathbb {Z}_p)$. Since the relevant associated graded pieces are in homological degrees $[2i-1, 2i]$ by Corollary 7.15, the result follows.

7.4 The main results

In this subsection we prove Theorem 1.8 from the introduction. Our strategy is to compare $L_{K(1)} \mathrm {TR}(-)$ with its $\mathrm {arc}_{{p}}$-hypersheafification.

Construction 7.24 (The invariant $\big ( L_{K(1)} \mathrm {TR}(-) \big )^{\sharp }$)

We define the functor $(L_{K(1)} \mathrm {TR}(-))^{\sharp }$ on derived $p$-complete rings as the $\mathrm {arc}_{{p}}$-hypersheafification of the functor $L_{K(1)} \mathrm {TR}(-)$ on derived $p$-complete rings. We have a comparison map

(22)\begin{equation} L_{K(1)} \mathrm{TR}(-) \to \big( L_{K(1)} \mathrm{TR}(-) \big)^{\sharp}. \end{equation}

For a ring which is not necessarily derived $p$-complete, we define $( L_{K(1)} \mathrm {TR}(-) )^{\sharp }$ as that of its derived $p$-completion.

To analyze this construction, we will use the results about $L_{K(1)} \mathrm {TR}$ of perfectoids proved in the previous section, as well as some general tools from the appendix. On $\mathrm {Perfd}$, Proposition 7.23 implies that $L_{K(1)} \mathrm {TR}(-)$ defines an $\mathrm {arc}_{{p}}$-hypersheaf. Using Proposition A.6, it follows that the restriction of $(L_{K(1)} \mathrm {TR}(-))^{\sharp }$ to $\mathrm {Perfd}$ is just $L_{K(1)} \mathrm {TR}(-)$ again, that is, (22) is an equivalence.Footnote 8

Proposition 7.25 For any ring $R$, there is a natural, complete exhaustive $\mathbb {Z}_{}$-indexed filtration on $( L_{K(1)} \mathrm {TR}(R))^{\sharp }$, denoted $\mathrm {Fil}^{\geq \ast }( L_{K(1)} \mathrm {TR}(R))^{\sharp }$, with associated graded terms $\mathrm {gr}^i ( L_{K(1)} \mathrm {TR}(R))^{\sharp } \simeq R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), W(\mathcal {O})(i))[2i]$.

Proof. The filtration in question is the $\mathrm {arc}_{{p}}$-Postnikov filtration (on the derived $p$-completion of $R$). Note that Postnikov towers converge for hypercomplete $\mathrm {arc}_{{p}}$-sheaves (Proposition A.10). For the identification of the graded pieces (or equivalently, the sheafified homotopy groups), it suffices by descent (see Example A.9) to work with the subcategory $\mathrm {Perfd}$, where the result follows from (21).

We do not know in general for which $R$ the comparison map (22) is an equivalence; for such $R$, one obtains a ‘motivic’ filtration on $L_{K(1)} \mathrm {TR}(R)$ with associated graded terms the $\mathrm {arc}_{{p}}$-cohomology complexes $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), W(\mathcal {O})(i))[2i]$. Here we will show that the comparison map is an equivalence in certain formally smooth cases, using the pro-Galois descent result (Theorem 6.8).

Theorem 7.26 Let $R_0$ be a $p$-torsion-free perfectoid ring. Let $R$ be a formally smooth $p$-complete $R_0$-algebra. Then $L_{K(1)} \mathrm {TR}(R) \xrightarrow {\sim } (L_{K(1)} \mathrm {TR}(R))^{\sharp }$.

Proof. First, we reduce to the case where $R_0$ admits a compatible system of $p$-power roots of unity. By André's lemma [Reference Bhatt and ScholzeBS19, Theorem 7.12], we know that there exists a $p$-completely faithfully flat perfectoid $R_0$-algebra $R_0'$ which has this property (e.g. is absolutely integrally closed). Now we have by descent

(23)\begin{equation} \mathrm{TR}(R; \mathbb{Z}_p) \simeq \mathrm{Tot} \Big( \mathrm{TR}(R \otimes_{R_0} R_0' ; \mathbb{Z}_p) \rightrightarrows \mathrm{TR}( R \otimes_{R_0}R_0' \otimes_{R_0} R_0'; \mathbb{Z}_p) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big). \end{equation}

Since in high degrees all of the terms in the above totalization agree with their $K(1)$-localization by Proposition 5.10, it follows that the descent property (23) holds for $L_{K(1) } \mathrm {TR}(-)$ as well (see the beginning of the proof of Proposition 5.10). Moreover, $(L_{K(1)} \mathrm {TR}(-))^{\sharp }$ satisfies descent for the map $R \to R \otimes _{R_0} R_0'$ by construction. Therefore, we reduce to the case where $R_0$ contains $\widehat {\mathbb {Z}_p[\zeta _{p^\infty }]}$.

Working locally on $\mathrm {Spf}(R)$, we can assume that $R$ receives a map from $R_0\langle t_1^{\pm 1}, \ldots , t_n^{\pm 1} \rangle$ which is étale mod $p$. We consider the extension $R_\infty = R \otimes _{\mathbb {Z}[\zeta _{p^\infty }, t_1^{\pm 1}, \ldots , t_n^{\pm 1}]} \mathbb {Z}[\zeta _{p^\infty }, t_1^{\pm 1/p^\infty }, \ldots , t_n^{\pm 1/p^\infty }]$ and the evident $\mathbb {Z}_p(1)^n$-action on $R_\infty$. As in Example 6.10, we find from Theorem 6.8 that the natural map induces an equivalence

\[ L_{K(1)} \mathrm{TR}(R) \xrightarrow{\sim} \mathrm{Tot} \Big( L_{K(1)} \mathrm{TR}(R_\infty) \rightrightarrows L_{K(1)} \mathrm{TR}( \mathrm{Fun}_{\mathrm{cts}}( \mathbb{Z}_p(1)^n, R_\infty)) \begin{smallmatrix} \to \\ \to \\\to \end{smallmatrix} \ldots \Big). \]

Now this is also true for $(L_{K(1)}\mathrm {TR}(-))^{\sharp }$, because the above augmented cosimplicial ring is an $\mathrm {arc}_{{p}}$-hypercover (strictly speaking, for that we replace all rings by their derived $p$-completions). But now this gives $L_{K(1)} \mathrm {TR}(R) \xrightarrow {\sim } (L_{K(1)}\mathrm {TR}(R))^{\sharp }$, since they agree on perfectoids and every term in the above cosimplicial resolution has perfectoid $p$-completion.

Theorem 7.27 Suppose $R$ is a formally smooth $\mathcal {O}_K$-algebra, where $K$ is a complete discretely valued field of mixed characteristic $(0, p)$ whose residue field $k$ satisfies $[k:k^p]< \infty$. Then $L_{K(1)} \mathrm {TR}(R) \xrightarrow {\sim } ( L_{K(1)} \mathrm {TR}(R))^{\sharp }$.

Proof. We let $G = \mathrm {Gal}(\bar {K}/K)$ and consider the $G$-action on $S = R \otimes _{\mathcal {O}_K} \mathcal {O}_{\bar {K}}$. As in Example 6.9, our assumptions imply that $G$ has finite cohomological dimension and, moreover, that the map (13) is an equivalence. Therefore, it suffices to show that $L_{K(1)} \mathrm {TR}( A) \xrightarrow {\sim } (L_{K(1)} \mathrm {TR}( A))^{\sharp }$ when $A$ is one of $S, \mathrm {Fun}_{\mathrm {cts}}(G, S), \ldots$. But these are all (up to $p$-completions) formally smooth over perfectoids, whence the claim by Theorem 7.26 and descent.

Remark 7.28 Suppose $L_{K(1)} \mathrm {TR}(R) \xrightarrow {\sim } (L_{K(1)} \mathrm {TR}(R) )^{\sharp }$. On Frobenius fixed points, one recovers the Thomason [Reference ThomasonTho85, Reference Thomason and TrobaughTT90] filtration (i.e. the pro-étale Postnikov filtration) on $L_{K(1)} \mathrm {TC}(R) \simeq L_{K(1)} K(R[1/p])$ (see [Reference Bhatt, Clausen and MathewBCM20] for this identification), whose associated graded terms are given by $\mathrm {gr}^i \simeq R \Gamma _{\mathrm {pro\acute {e}t}}( \mathrm {Spec}(R[1/p]), \mathbb {Z}_p(i))[2i]$; see Proposition 7.16.

Remark 7.29 The above filtration on $L_{K(1)} \mathrm {TR}(-)$ and the calculations of $\mathrm {TR}$ of smooth algebras over a DVR in [Reference Hesselholt and MadsenHM03, Reference Hesselholt and MadsenHM04, Reference Geisser and HesselholtGH06] suggest that the cohomology $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), W(\mathcal {O})(i))$ should be related to the absolute de Rham–Witt complex. The work [Reference MorrowMor18] also suggests that for smooth algebras over a perfectoid base containing all $p$-power roots of unity, $R \Gamma _{\mathrm {arc}_{{p}}}( \mathrm {Spec}(R), W(\mathcal {O}))$ should be related to the relative de Rham–Witt complex (over the perfectoid base).

Acknowledgements

I would like to thank B. Antieau, L. Hesselholt, M. Morrow, T. Nikolaus, W. Nizioł, and P. Scholze for many helpful conversations related to the subject of this paper. I would especially like to thank B. Bhatt and D. Clausen for their collaboration in [Reference Bhatt, Clausen and MathewBCM20] and many helpful discussions. Finally, I would like to thank B. Antieau, L. Meier, G. Tamme, and the referee for several comments and corrections on an earlier version.

Topological preliminaries

In this appendix, we record some basic topological preliminaries about (hyper)sheaves of spectra.

Remark A.1 (Conventions for sites) We will for simplicity work only with sites of the following nice form (see [Reference LurieLur18, § A.3.2]). Let $\mathcal {C}$ be a category with pullbacks and finite coproducts, such that coproducts distribute over pullbacks and are disjoint. Suppose $\mathcal {C}$ is equipped with a class of morphisms $S = S_{\mathcal {C}}$ which contains all isomorphisms and is stable under composition and pullback. We equip $\mathcal {C}$ with the Grothendieck topology where a collection $\{X_i \to X\}_{i \in I}$ is a covering if there exists a finite subset $I' \subset I$ such that ${\bigsqcup} _{i \in I'} X_i \to X$ can be refined by a map belonging to $S$. In this case, a presheaf on $\mathcal {C}$ with values in an $\infty$-category $\mathcal {D}$ with all small limits is a sheaf if and only if it carries finite coproducts in $\mathcal {C}$ to finite products in $\mathcal {D}$ and if it satisfies Čech descent for maps in $S$; see [Reference LurieLur18, § A.3.3].

Example A.2

  1. (i) The small or big étale site of a qcqs scheme (where we only allow qcqs schemes) is an example, with $S$ the class of étale surjections.

  2. (ii) The $\mathrm {arc}_{{p}}$-topology or $\mathrm {arc}_{\hat {p}}$-topology on the opposite of the category of derived $p$-complete rings is an example, with $S$ the class of $\mathrm{arc}_p$ or $\mathrm{arc}_{\hat{p}}$-covers.

  3. (iii) The $\mathrm {arc}_{{p}}$-topology or $\mathrm {arc}_{\hat {p}}$-topology on the opposite of the category of perfectoid rings is an example, with $S$ the class of $\mathrm{arc}_p$ or $\mathrm{arc}_{\hat{p}}$-covers.

Remark A.3 (Examples of continuous functors) Let $\mathcal {C}, \mathcal {C}'$ be sites as in Remark A.1. Let $u \colon \mathcal {C}' \to \mathcal {C}$ be a functor which preserves finite coproducts and pullbacks, as well as morphisms in the respective classes $S_{\mathcal {C}}, S_{\mathcal {C}'}$. It follows that if $\mathcal {F}$ is a sheaf (with values in any $\infty$-category $\mathcal {D}$ with all small limits) on $\mathcal {C}$, then $\mathcal {F} \circ u$ is a sheaf on $\mathcal {C}'$. These are examples of continuous functors; the notion can be defined for more general sites; see [Reference Artin, Grothendieck and VerdierSGA4, Exp. III.1].

Construction A.4 (Sheaves of spectra) Given a site $\mathcal {C}$ (as in Remark A.1), we let $\mathrm {PSh}(\mathcal {C}, {\rm Sp} )$ denote the $\infty$-category of presheaves of spectra on $\mathcal {C}$ and $\mathrm {Shv}(\mathcal {C}, {\rm Sp} ) \subset \mathrm {PSh}(\mathcal {C}, {\rm Sp} )$ denote the subcategory of sheaves of spectra [Reference LurieLur18, § 1.3]. We equip these both with their canonical $t$-structures; see [Reference LurieLur18, § 1.3.2].

Given $u \colon \mathcal {C}' \to \mathcal {C}$ as in Remark A.3, we obtain a right adjoint and left $t$-exact functor $(-) \circ u \colon \mathrm {Shv}(\mathcal {C}, {\rm Sp} ) \to \mathrm {Shv}(\mathcal {C}', {\rm Sp} )$. It has a left adjoint $u_! \colon \mathrm {Shv}(\mathcal {C}', {\rm Sp} ) \to \mathrm {Shv}(\mathcal {C}, {\rm Sp} )$ given by left Kan extension along $u$ followed by sheafification. By adjunction, necessarily $u_!$ is right $t$-exact.

Construction A.5 (Hypersheaves of spectra) Let $\mathcal {C}$ be a site as in Remark A.1. Any presheaf $\mathcal {F}$ of spectra on $\mathcal {C}$ fits into a unique fiber sequence of spectra

(A.1)\begin{equation} \mathcal{F}_{\mathrm{null}} \to \mathcal{F} \to \mathcal{F}^{\sharp}, \end{equation}

where $\mathcal {F}_{\mathrm {null}}$ has trivial sheafified homotopy groups and $\mathcal {F}^{\sharp }$ is a hypercomplete sheaf of spectra, that is, for every presheaf $\mathcal {G}$ with trivial sheafified homotopy groups, we have $\text{Hom} (\mathcal {G}, \mathcal {F}^{\sharp }) =0$. If $\mathcal {F}$ is a sheaf of spectra, then the cofiber sequence shows that $\mathcal {F}_{\mathrm {null}}$ is also a sheaf of spectra; it is then $\infty$-connective with respect to the $t$-structure on $\mathrm {Shv}(\mathcal {C}, {\rm Sp} )$. We let $\mathrm {Shv}_{\mathrm {hyp}}( \mathcal {C}, {\rm Sp} ) \subset \mathrm {Shv}(\mathcal {C}, {\rm Sp} )$ denote the full subcategory of hypercomplete sheaves. We refer to [Reference Clausen and MathewCM19, § 2] for an exposition of some of these constructions, which go back to [Reference JardineJar87, Reference Dugger, Hollander and IsaksenDHI04].

Proposition A.6 Let $u \colon \mathcal {C}' \to \mathcal {C}$ be a morphism of sites as in Remark A.1 preserving finite coproducts and pullbacks and carrying the class of arrows $S_{\mathcal {C}'}$ into $S_{\mathcal {C}}$. Suppose that for any object $X' \in \mathcal {C}'$ and a finite covering family $\{ Y_i \to u(X')\}_{i \in I}$ in $\mathcal {C}$, there is a refinement which is the image under $u$ of a finite covering family $\{Y_i' \to X'\}_{i \in I}$ in $\mathcal {C}'$.

Then the restriction functor $(-) \circ u \colon \mathrm {PSh}(\mathcal {C}, {\rm Sp} ) \to \mathrm {PSh}(\mathcal {C}', {\rm Sp} )$ commutes with hypersheafification (and, in particular, preserves hypercomplete sheaves).

Proof. Using the cofiber sequence (A.1), we see that it suffices to prove that $(-) \circ u \colon \mathrm {Shv}( \mathcal {C}, {\rm Sp} ) \to \mathrm {Shv}(\mathcal {C}', {\rm Sp} )$ preserves both the subclasses of objects with trivial homotopy groups and hypercomplete objects.

Our hypotheses imply that if a presheaf of abelian groups on $\mathcal {C}$ has trivial sheafification, then its pullback to $\mathcal {C}'$ has trivial sheafification. Therefore, $(-) \circ u$ preserves objects with trivial homotopy groups. It thus suffices to show that if $\mathcal {F} \in \mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C}, {\rm Sp} )$, then the sheaf $\mathcal {F} \circ u \in \mathrm {Shv}(\mathcal {C}', {\rm Sp} )$ is also hypercomplete; this will not use the assumption in the second sentence of the statement. Equivalently, given $\mathcal {G} \in \mathrm {Shv}(\mathcal {C}', {\rm Sp} )$ which is $\infty$-connective, it suffices to show that $\text{Hom} _{\mathrm {Shv}(\mathcal {C}', {\rm Sp} )}(\mathcal {G}, \mathcal {F} \circ u) =0$. By adjointness, this is $\text{Hom} _{\mathrm {Shv}(\mathcal {C}, {\rm Sp} )} ( u_! \mathcal {G}, \mathcal {F})$; since $u_! \colon \mathrm {Shv}(\mathcal {C}', {\rm Sp} ) \to \mathrm {Shv}(\mathcal {C}, {\rm Sp} )$ is right $t$-exact and therefore preserves $\infty$-connective objects and $\mathcal {F}$ is hypercomplete, this is contractible.

Example A.7 (Overcategories) Suppose $a \in \mathcal {C}$ and $\mathcal {C}' = \mathcal {C}_{/a}$ is the overcategory of $a$, with the induced topology. Then the natural forgetful functor $\mathcal {C}' \to \mathcal {C}$ clearly satisfies the conditions of Proposition A.6. In particular, the hypersheafification of a presheaf on $\mathcal {C}$ when restricted to $\mathcal {C}_{/a}$ is the hypersheafification of the restriction to $\mathcal {C}_{/a}$.

A direct consequence is that given an appropriate site with a ‘basis’, hypersheaves of spectra can be entirely recovered from their values on the basis. The result is a direct analog of [Reference Artin, Grothendieck and VerdierSGA4, Theorem 4.1, Exp. III], and appears (for sheaves of spaces) in [Reference AokiAok20, Appendix A].

Proposition A.8 Let $\mathcal {C}$ be a site as in Remark A.1. Let $\mathcal {C}' \subset \mathcal {C}$ be a full subcategory closed under finite coproducts and fiber products, and define $S_{\mathcal {C}'}$ to be the intersection of $S_{\mathcal {C}}$ with $\mathcal {C}'$. Suppose every object $X \in \mathcal {C}$ admits a map $Y \to X$ in $S_{\mathcal {C}}$ with $Y \in \mathcal {C}'$. Then the restriction functor $\mathrm {Shv}(\mathcal {C}, {\rm Sp} ) \to \mathrm {Shv}(\mathcal {C}', {\rm Sp} )$ restricts to an equivalence on hypercomplete objects.

Proof. Let $u \colon \mathcal {C}' \subset \mathcal {C}$ be the inclusion, so we have a restriction functor $(-) \circ u \colon \mathrm {PSh}(\mathcal {C}, {\rm Sp} ) \to \mathrm {PSh}(\mathcal {C}', {\rm Sp} )$. By Proposition A.6, it restricts to a functor on hypercomplete sheaves, so we have $(-) \circ u \colon \mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C}, {\rm Sp} ) \to \mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C}', {\rm Sp} )$. This last functor has a left adjoint $L \colon \mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C}', {\rm Sp} ) \to \mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C}, {\rm Sp} )$, given by $\mathcal {F} \mapsto L\mathcal {F} := (\mathrm {Lan}_u \mathcal {F})^{\sharp }$, that is, $L$ is obtained by applying the left Kan extension $\mathrm {Lan}_u \colon \mathrm {PSh}(\mathcal {C}', {\rm Sp} ) \to \mathrm {PSh}(\mathcal {C}, {\rm Sp} )$ followed by hypersheafification $(-)^{\sharp }$. We now show that $L$ is fully faithful. Indeed, we have, for $\mathcal {F}, \mathcal {G} \in \mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C}', {\rm Sp} )$,

(A.2)\begin{equation} \text{Hom}_{\mathrm{Shv}_{\mathrm{hyp}}(\mathcal{C}, {\rm Sp}) }( L \mathcal{F}, L \mathcal{G}) = \text{Hom}_{\mathrm{Shv}_{\mathrm{hyp}}( \mathcal{C}', {\rm Sp})}\big(\mathcal{F}, (\mathrm{Lan}_u \mathcal{G})^{\sharp} \circ u \big). \end{equation}

Now since $u$ is fully faithful, since hypersheafification commutes with $(-) \circ u$ by Proposition A.6, and since $\mathcal {G}$ is already hypercomplete, we have $(\mathrm {Lan}_u \mathcal {G})^{\sharp } \circ u = \mathcal {G}$. Therefore, the right-hand side of (A.2) simplifies to $\text{Hom} _{\mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C}', {\rm Sp} )}(\mathcal {F}, \mathcal {G})$ as desired.

Our hypotheses imply that the restriction functor $(-) \circ u$ is conservative on hypercomplete sheaves, because it is conservative on sheaves of abelian groups. Since the restriction functor is conservative and has a fully faithful left adjoint, the result follows.

Example A.9 (Unfolding in the $\mathrm {arc}_{{p}}$-topology)

The natural forgetful functor establishes an equivalence of $\infty$-categories between $\mathrm {arc}_{{p}}$-hypersheaves of spectra on all derived $p$-complete rings and $\mathrm {arc}_{{p}}$-hypersheaves of spectra on perfectoid rings.

Finally, we include a basic observation about commuting hypersheafification and certain products when the site $\mathcal {C}$ is sufficiently large (e.g. the $\mathrm {arc}_{{p}}$-site). This fact is closely related to the theory of replete topoi; see [Reference Bhatt and ScholzeBS15, § 3]. Compare [Reference Bhatt and ScholzeBS15, Proposition 3.3.3] for the second part of the next result for the derived category of abelian sheaves, or equivalently hypercomplete $H\mathbb {Z}$-module sheaves of spectra. Note, in particular, that it applies to the $\mathrm {arc}_{{p}}$-site. This follows because $\mathrm {arc}$-covers in the sense of [Reference Bhatt and MathewBM18] are closed under filtered colimits of rings [Reference Bhatt and MathewBM18, Corollary 2.20] and because a map of derived $p$-complete rings $R \to R'$ is an $\mathrm {arc}_{{p}}$-cover if and only if $R \to R' \times R/p \times R[1/p]$ is an $\mathrm {arc}$-cover.

In the following, we write $\mathrm {PSh}_{\sqcup }(\mathcal {C}) \subset \mathrm {PSh}(\mathcal {C})$ for the subcategory of presheaves which carry finite coproducts to finite products.

Proposition A.10 Let $\mathcal {C}$ be a site as in Remark A.1. Suppose $\mathcal {C}$ admits countable filtered limits. Suppose, moreover, that if $\cdots \to X_i \to X_{i-1} \to \cdots \to X_0$ is a sequence of arrows in $S$, then $\varprojlim _i X_i \to X_0$ belongs to $S$. Then the following assertions hold.

  1. (i) The hypersheafification functor, $(-)^{\sharp } \colon \mathrm {PSh}_{\sqcup }(\mathcal {C}) \to \mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C})$ commutes with countable products and limits along $\mathbb {Z}_{\geq 0}$-indexed towers.

  2. (ii) For any $\mathcal {G} \in \mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C}) \subset \mathrm {Shv}(\mathcal {C})$, the Postnikov tower of $\mathcal {G}$ (as a sheaf of spectra) converges.

Proof. For (i), since hypersheaves are always closed (inside presheaves) under arbitrary limits and since $\mathbb {Z}_{\geq 0}$-indexed limits can be built from countable products, it suffices to show (via the unique cofiber sequence (A.1)) that if $\{\mathcal {F}_i\}_{i \in \mathbb {N}}$ is a countable family of presheaves in $\mathrm {PSh}_{\sqcup }(\mathcal {C})$ with trivial hypersheafification (i.e. trivial sheafified homotopy groups), then the product presheaf $\prod _{i \in \mathbb {N}} \mathcal {F}_i$ has trivial hypersheafification. But this is just a claim about the presheaves of abelian groups $\{\pi _j( \mathcal {F}_i)\}_{i \in \mathbb {N}}$ for each $j$. Explicitly, given a class $\alpha = (\alpha _i) \in \prod _{i \in \mathbb {N}} \pi _j( \mathcal {F}_i(X))$ for some $X \in \mathcal {C}$, we find for each $i$ a cover $X_i \to X$ in $S$ which annihilatesFootnote 9 $\alpha _i \in \pi _j( \mathcal {F}_i(X))$ and then form the cover $X_1 \times _X X_2 \times _X \ldots$ of $X$, which annihilates $\alpha$.

Now let $\mathcal {G}$ be a hypercomplete sheaf of spectra on $\mathcal {C}$. The Postnikov tower of $\mathcal {G} \in \mathrm {Shv}(\mathcal {C})$ is obtained by taking the presheaf truncations $\tau ^{\mathrm {pre}}_{\leq n} \mathcal {G}$ and applying the hypersheafification (or sheafification, since these objects are truncated), that is, one forms $\{ (\tau ^{\mathrm {pre}}_{\leq n} \mathcal {G})^{\sharp } \}$, which is a tower in $\mathrm {PSh}_{\sqcup }(\mathcal {C})$. Since Postnikov towers converge for presheaves of spectra, that is, $\mathcal {G} \simeq \varprojlim _n \tau ^{\mathrm {pre}}_{\leq n} \mathcal {G}$, and we have just seen that $(-)^{\sharp } \colon \mathrm {PSh}_{\sqcup }(\mathcal {C}) \to \mathrm {Shv}_{\mathrm {hyp}}(\mathcal {C})$ commutes with limits along $\mathbb {Z}_{\geq 0}$-indexed towers, we find $\mathcal {G} \simeq \varprojlim _n (\tau _{\leq n}^{\mathrm {pre}} \mathcal {G})^{\sharp }$ as desired.

Footnotes

This work was done while the author was a Clay Research Fellow.

1 For set-theoretic reasons, to define $\mathrm {arc}_{{p}}$-cohomology we should fix a cutoff cardinal. We will only consider situations where the choice of cutoff cardinal does not affect the result.

2 This states informally that if $A$ is a connective $B$-algebra, then the fiber of the map $\mathrm {TR}(A; \mathbb {F}_p) \to \mathrm {TR}(H\pi _0(A); \mathbb {F}_p)$ has the property that each degree is annihilated by a power of $v_1$ (depending on the degree; note that this condition is slightly stronger than the fiber simply being $K(1)$-acyclic).

3 This type of argument is also used in [Reference Land, Mathew, Meier and TammeLMMT20, § 3.2].

4 Also called a simplicial commutative ring; see [Reference Česnavičius and ScholzeČS19] for a discussion of this terminology.

5 For $C(-)$, this follows from the expression $K(R[x]/x^n, (x)) = \mathrm {TC}(R[x]/x^n, (x))$, using that $\mathrm {TC}$ of connective ring spectra commutes with geometric realizations, and passing to the limit along $n$.

6 We expect that this works when $k$ has characteristic $2$ as well.

7 Throughout, for set-theoretic reasons one should impose cardinality bounds, that is, assume all the rings one allows into the site to be of cardinality less than some uncountable strong limit cardinal $\kappa$, so that the sites will be small. However, the choice of $\kappa$ will not matter in all the constructions we consider and we will consequently suppress it.

8 Alternatively, one could build $( L_{K(1)} \mathrm {TR}(-))^{\sharp }$ using the unfolding construction of Example A.9, by starting with the $\mathrm {arc}_{{p}}$-hypersheaf $L_{K(1)} \mathrm {TR}(-)$ on $\mathrm {Perfd}$.

9 Here we use that we are working with presheaves which carry finite coproducts to finite products, so that we can reduce to working with covers consisting of one morphism.

References

Angelini-Knoll, G. and Quigley, J. D., The Segal conjecture for topological Hochschild homology of Ravenel spectra, J. Homotopy Relat. Struct. 16 (2021), 4160.CrossRefGoogle Scholar
Antieau, B., Mathew, A., Morrow, M. and Nikolaus, T., On the Beilinson fiber square, Preprint (2020), arXiv:2003.12541.Google Scholar
Antieau, B. and Nikolaus, T., Cartier modules and cyclotomic spectra, J. Amer. Math. Soc. 34 (2021), 178.CrossRefGoogle Scholar
Aoki, K., Tensor triangular geometry of filtered objects and sheaves, Preprint (2020), arXiv:2001.00319.Google Scholar
Ausoni, C. and Rognes, J., Algebraic $K$-theory of topological $K$-theory, Acta Math. 188 (2002), 139.10.1007/BF02392794CrossRefGoogle Scholar
Ayala, D. and Francis, J., Factorization homology of topological manifolds, J. Topol. 8 (2015), 10451084.CrossRefGoogle Scholar
Ayala, D., Francis, J. and Tanaka, H. L., Factorization homology of stratified spaces, Selecta Math. (N.S.) 23 (2017), 293362.10.1007/s00029-016-0242-1CrossRefGoogle Scholar
Bhatt, B., Clausen, D. and Mathew, A., Remarks on $K (1)$-local $K$-theory, Selecta Math. (N.S.) 26 (2020), Paper No. 39.CrossRefGoogle Scholar
Bhatt, B. and Mathew, A., The arc-topology, Duke Math. J., to appear. Preprint (2018), arXiv:1807.04725.Google Scholar
Bhatt, B., Morrow, M. and Scholze, P., Integral $p$-adic Hodge theory, Publ. Math. Inst. Hautes Études Sci. 128 (2018), 219397.CrossRefGoogle Scholar
Bhatt, B., Morrow, M. and Scholze, P., Topological Hochschild homology and integral $p$-adic Hodge theory, Publ. Math. Inst. Hautes Études Sci. 129 (2019), 199310.CrossRefGoogle Scholar
Bhatt, B. and Scholze, P., The pro-étale topology for schemes, Astérisque 369 (2015), 99201.Google Scholar
Bhatt, B. and Scholze, P., Projectivity of the Witt vector affine Grassmannian, Invent. Math. 209 (2017), 329423.CrossRefGoogle Scholar
Bhatt, B. and Scholze, P., Prisms and prismatic cohomology, Preprint (2019), arXiv:1905.08229.Google Scholar
Blumberg, A. J., Gepner, D. and Tabuada, G., A universal characterization of higher algebraic $K$-theory, Geom. Topol. 17 (2013), 733838.CrossRefGoogle Scholar
Bökstedt, M., Bruner, R. R., Lunøe-Nielsen, S. and Rognes, J., On cyclic fixed points of spectra, Math. Z. 276 (2014), 8191.10.1007/s00209-013-1187-0CrossRefGoogle Scholar
Bökstedt, M., Hsiang, W. C. and Madsen, I., The cyclotomic trace and algebraic $K$-theory of spaces, Invent. Math. 111 (1993), 465539.CrossRefGoogle Scholar
Bökstedt, M. and Madsen, I., Topological cyclic homology of the integers, $K$-theory (Strasbourg, 1992), Astérisque 226 (1994), 7–8, 57–143.Google Scholar
Brun, M., Filtered topological cyclic homology and relative $K$-theory of nilpotent ideals, Algebr. Geom. Topol. 1 (2001), 201230.CrossRefGoogle Scholar
Česnavičius, K. and Scholze, P., Purity for flat cohomology, Preprint (2019), arXiv:1912.10932.Google Scholar
Clausen, D. and Mathew, A., Hyperdescent and étale K-theory, Invent. Math., to appear. Preprint (2019), arXiv:1905.06611.Google Scholar
Clausen, D., Mathew, A. and Morrow, M., $K$-theory and topological cyclic homology of henselian pairs, J. Amer. Math. Soc. 34 (2021), 411473.Google Scholar
Clausen, D., Mathew, A., Naumann, N. and Noel, J., Descent in algebraic $K$-theory and a conjecture of Ausoni–Rognes, J. Eur. Math. Soc. (JEMS) 22 (2020), 11491200.CrossRefGoogle Scholar
Crabb, M. and James, I., Fibrewise homotopy theory, Springer Monographs in Mathematics (Springer, London, 1998).CrossRefGoogle Scholar
Dugger, D., Hollander, S. and Isaksen, D. C., Hypercovers and simplicial presheaves, Math. Proc. Cambridge Philos. Soc. 136 (2004), 951.CrossRefGoogle Scholar
Dundas, B. I., Goodwillie, T. G. and McCarthy, R., The local structure of algebraic K-theory, Algebra and Applications, vol. 18 (Springer, London, 2013).Google Scholar
Gabber, O., $K$-theory of Henselian local rings and Henselian pairs, in Algebraic $K$-theory, commutative algebra, and algebraic geometry (Santa Margherita Ligure, 1989), Contemporary Mathematics, vol. 126 (American Mathematical Society, Providence, RI, 1992), 5970.CrossRefGoogle Scholar
Gabber, O., Affine analog of the proper base change theorem, Israel J. Math. 87 (1994), 325335.CrossRefGoogle Scholar
Geisser, T. and Hesselholt, L., The de Rham-Witt complex and $p$-adic vanishing cycles, J. Amer. Math. Soc. 19 (2006), 136.CrossRefGoogle Scholar
Geisser, T. and Levine, M., The $K$-theory of fields in characteristic $p$, Invent. Math. 139 (2000), 459493.10.1007/s002220050014CrossRefGoogle Scholar
Gabber, O. and Orgogozo, F., Sur la $p$-dimension des corps, Invent. Math. 174 (2008), 4780.10.1007/s00222-008-0133-yCrossRefGoogle Scholar
Gabber, O. and Ramero, L., Almost ring theory, Lecture Notes in Mathematics, vol. 1800 (Springer, Berlin, 2003).CrossRefGoogle Scholar
Greco, S., Two theorems on excellent rings, Nagoya Math. J. 60 (1976), 139149.CrossRefGoogle Scholar
Gunawardena, J., Segal's Burnside ring conjecture for cyclic groups of odd prime order (JT Knight prize essay, Cambridge, 1980).Google Scholar
Haesemeyer, C. and Weibel, C. A., The norm residue theorem in motivic cohomology, Annals of Mathematics Studies, vol. 200 (Princeton University Press, Princeton, NJ, 2019).Google Scholar
Hesselholt, L., On the $p$-typical curves in Quillen's $K$-theory, Acta Math. 177 (1996), 153.CrossRefGoogle Scholar
Hesselholt, L., Algebraic $K$-theory and trace invariants, in Proceedings of the International Congress of Mathematicians, Vol. II (Beijing, 2002) (Higher Education Press, Beijing, 2002), 415425.Google Scholar
Hesselholt, L., On the topological cyclic homology of the algebraic closure of a local field, in An alpine anthology of homotopy theory, Contemporary Mathematics, vol. 399 (American Mathematical Society, Providence, RI, 2006), 133162.CrossRefGoogle Scholar
Hesselholt, L., Topological Hochschild homology and the Hasse–Weil zeta function, in An alpine bouquet of algebraic topology, Contemporary Mathematics, vol. 708 (American Mathematical Society, Providence, RI, 2018), 157180.CrossRefGoogle Scholar
Hesselholt, L. and Madsen, I., On the $K$-theory of finite algebras over Witt vectors of perfect fields, Topology 36 (1997), 29101.CrossRefGoogle Scholar
Hesselholt, L. and Madsen, I., On the $K$-theory of local fields, Ann. of Math. (2) 158 (2003), 1113.CrossRefGoogle Scholar
Hesselholt, L. and Madsen, I., On the de Rham-Witt complex in mixed characteristic, Ann. Sci. Éc. Norm. Supér. (4) 37 (2004), 143.CrossRefGoogle Scholar
Hesselholt, L. and Nikolaus, T., Topological cyclic homology, in Handbook of homotopy theory, ed. H. Miller (CRC Press, Boca Raton, FL, 2019).Google Scholar
Huber, R., Étale cohomology of Henselian rings and cohomology of abstract Riemann surfaces of fields, Math. Ann. 295 (1993), 703708.CrossRefGoogle Scholar
Jardine, J. F., Simplicial presheaves, J. Pure Appl. Algebra 47 (1987), 3587.CrossRefGoogle Scholar
Kato, K., Galois cohomology of complete discrete valuation fields, in Algebraic $K$-theory, Part II (Oberwolfach, 1980), Lecture Notes in Mathematics, vol. 967 (Springer, Berlin, 1982), 215238.CrossRefGoogle Scholar
Kurano, K. and Shimomoto, K., Ideal-adic completion of quasi-excellent rings (after Gabber), Preprint (2016), arXiv:1609.09246.Google Scholar
Land, M., Mathew, A., Meier, L. and Tamme, G., Purity in chromatically localized algebraic $K$-theory, Preprint (2020), arXiv:2001.10425.Google Scholar
Land, M. and Tamme, G., On the $K$-theory of pullbacks, Ann. of Math. (2) 190 (2019), 877930.CrossRefGoogle Scholar
Lin, W. H., On conjectures of Mahowald, Segal and Sullivan, Math. Proc. Cambridge Philos. Soc. 87 (1980), 449458.CrossRefGoogle Scholar
Loday, J.-L., Cyclic homology, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 301, second edition (Springer, Berlin, 1998); appendix E by M. O. Ronco, Chapter 13 by the author in collaboration with Teimuraz Pirashvili.CrossRefGoogle Scholar
Lunøe-Nielsen, S. and Rognes, J., The Segal conjecture for topological Hochschild homology of complex cobordism, J. Topol. 4 (2011), 591622.10.1112/jtopol/jtr015CrossRefGoogle Scholar
Lurie, J., Spectral algebraic geometry, Preprint (2018), http://math.ias.edu/~lurie/.Google Scholar
Mahowald, M., $b\textrm {o}$-resolutions, Pacific J. Math. 92 (1981), 365383.CrossRefGoogle Scholar
Mathew, A., Examples of descent up to nilpotence, in Geometric and topological aspects of the representation theory of finite groups, Springer Proceedings in Mathematics & Statistics, vol. 242 (Springer, Cham, 2018), 269311.CrossRefGoogle Scholar
Mathew, A., Naumann, N. and Noel, J., Nilpotence and descent in equivariant stable homotopy theory, Adv. Math. 305 (2017), 9941084.10.1016/j.aim.2016.09.027CrossRefGoogle Scholar
Matsumura, H., Commutative algebra, Mathematics Lecture Note Series, vol. 56, second edition (Benjamin/Cummings, Reading, MA, 1980).Google Scholar
Matsumura, H., Commutative ring theory, Cambridge Studies in Advanced Mathematics, vol. 8 (Cambridge, Cambridge University Press, 1986); translated from the Japanese by M. Reid.Google Scholar
Miller, H. R., On relations between Adams spectral sequences, with an application to the stable homotopy of a Moore space, J. Pure Appl. Algebra 20 (1981), 287312.CrossRefGoogle Scholar
Milnor, J., Introduction to algebraic K-theory, Annals of Mathematics Studies, vol. 72 (Princeton University Press, Princeton, NJ; University of Tokyo Press, Tokyo, 1971).Google Scholar
Morrow, M., $p$-adic vanishing cycles as Frobenius-fixed points, Preprint (2018), arXiv:1802.03317.Google Scholar
Nikolaus, T. and Scholze, P., On topological cyclic homology, Acta Math. 221 (2018), 203409.CrossRefGoogle Scholar
Nizioł, W., Crystalline conjecture via $K$-theory, Ann. Sci. Éc. Norm. Supér. (4) 31 (1998), 659681.CrossRefGoogle Scholar
Nizioł, W., Semistable conjecture via $K$-theory, Duke Math. J. 141 (2008), 151178.CrossRefGoogle Scholar
Quillen, D., On the cohomology and $K$-theory of the general linear groups over a finite field, Ann. of Math. (2) 96 (1972), 552586.CrossRefGoogle Scholar
Quillen, D., Higher algebraic $K$-theory. I, in Algebraic $K$-theory, I: Higher $K$-theories (Proc. Conf., Battelle Memorial Inst., Seattle, Wash., 1972), Lecture Notes in Mathematics, vol. 341 (Springer, Berlin, 1973), 85147.CrossRefGoogle Scholar
Rosenschon, A. and Østvær, P. A., Descent for $K$-theories, J. Pure Appl. Algebra 206 (2006), 141152.CrossRefGoogle Scholar
Rydh, D., Submersions and effective descent of étale morphisms, Bull. Soc. Math. France 138 (2010), 181230.CrossRefGoogle Scholar
Scholze, P., $p$-Adic Hodge theory for rigid-analytic varieties, Forum Math. Pi 1 (2013), e1.10.1017/fmp.2013.1CrossRefGoogle Scholar
Artin, M., Grothendieck, A. and Verdier, J. L. (eds), Séminaire de Géométrie Algébrique du Bois Marie – 1963–64 – Théorie des topos et cohomologie étale des schémas (SGA 4), Vol. 1, Lecture Notes in Mathematics, vol. 269 (Springer, Berlin, 1972).Google Scholar
The Stacks Project Authors, The Stacks Project (2020), https://stacks.math.columbia.edu.Google Scholar
Suslin, A., On the $K$-theory of algebraically closed fields, Invent. Math. 73 (1983), 241245.CrossRefGoogle Scholar
Thomason, R. W., Algebraic $K$-theory and étale cohomology, Ann. Sci. Éc. Norm. Supér. (4) 18 (1985), 437552.CrossRefGoogle Scholar
Thomason, R. W. and Trobaugh, T., Higher algebraic $K$-theory of schemes and of derived categories, in The Grothendieck Festschrift, Vol. III, Progress in Mathematics, vol. 88 (Birkhäuser, Boston, 1990), 247435.10.1007/978-0-8176-4576-2_10CrossRefGoogle Scholar
Tsalidis, S., Topological Hochschild homology and the homotopy descent problem, Topology 37 (1998), 913934.CrossRefGoogle Scholar
van Rooij, A. C. M., Non-Archimedean functional analysis, Monographs and Textbooks in Pure and Applied Mathematics, vol. 51 (Marcel Dekker, New York, 1978).Google Scholar