0.25cm \cellspacebottomlimit0.25cm \titlecontentssection [1.5em] \contentslabel1.5em \contentspage \titlecontentssubsection [2.3em] \contentslabel2.3em \contentspage
Crystalline condition for –cohomology and ramification bounds
Abstract.
For a prime and a smooth proper –adic formal scheme over where is a –adic field, we study a series of conditions (), that partially control the –action on the image of the associated Breuil–Kisin prismatic cohomology inside the –prismatic cohomology . The condition () is a criterion for a Breuil–Kisin–Fargues –module to induce a crystalline representation used by Gee and Liu in [EmertonGee2, Appendix F], and thus leads to a proof of crystallinity of that avoids the crystalline comparison. The higher conditions () are used to adapt the strategy of Caruso and Liu from [CarusoLiu] to establish ramification bounds for the mod representations for arbitrary and , which extend or improve existing bounds in various situations.
1. Introduction
Let be a perfect field of characteristic and the associated absolutely unramified field. Let be a totally ramified finite extension with ramification index , and denote by its absolute Galois group. The goal of the present paper is to provide new bounds for ramification of the mod representations of that arise as the étale cohomology groups in terms of and , where is a smooth proper –adic formal scheme over (and is the geometric adic generic fiber). Concretely, let us denote by the –th ramification group of in the upper numbering (in the standard convention, e.g. [SerreLocalFields]) and . The main result is as follows.
Theorem 1.1 (Theorem LABEL:thm:FinalRamification).
Set
Then:
-
(1)
The group acts trivially on when
-
(2)
Denote by the field where is the kernel of the –representation given by . Then
where denotes the different of the extension and denotes the additive valuation on normalized so that
In particular, unlike in previous results of this type (discussed below), there are no restrictions on the size of and with respect to .
Remark 1.2.
As the constants appearing in Theorem 1.1 are quite complicated, let us draw some non–optimal, but more tractable consequences. The group acts trivially on when one of the following occurs:
-
(1)
and
- (2)
-
(3)
( are arbitrary) and
Let us briefly summarize the history of related results. Questions of this type originate in Fontaine’s paper [Fontaine], where he proved that for a finite flat group scheme over that is annihilated by , acts trivially on when ; this is a key step in his proof that there are no non–trivial abelian schemes over . In the same paper, Fontaine conjectured that general –torsion cohomology would follow the same pattern: given a proper smooth variety over with good reduction, should act trivially on when .
This conjecture has been subsequently proved by Fontaine himself ([Fontaine2]) in the case when and by Abrashkin ([Abrashkin]; see also [Abrashkin2]) when and is arbitrary. This is achieved by using Fontaine–Laffaille modules (introduced in [FontaineLaffaille]), which parametrize quotients of pairs of –stable lattices in crystalline representations with Hodge–Tate weights in (such as ). The (duals of the) representations are included among these thanks to a comparison theorem of Fontaine–Messing ([FontaineMessing]). Similarly to the orginal application, these ramification bounds lead to a scarcity result for existence of smooth proper –schemes.
Various extensions to the semistable case subsequently followed. Under the asumption (and arbitrary ), Hattori proved in [Hattori] a ramification bound for –torsion quotients of lattices in semistable representations with Hodge–Tate weights in the range using (a variant of) Breuil’s filtered –modules. Thanks to a comparison result between log–crystalline and étale cohomology by Caruso ([CarusoLogCris]), this results in a ramification bound for when is proper with semistable reduction, assuming when and when 222Recently, in [LiLiu] Li and Liu extended Caruso’s result to the range regardless of , for proper and smooth (formal) scheme. In view of this, results of [Hattori] should apply in these situations as well..
These results were further extended by Caruso and Liu in [CarusoLiu] for all –torsion quotients of pairs of semistable lattices with Hodge–Tate weights in , without any restriction on or . The proof uses the theory of –modules, which are objects suitable for description of lattices in semistable representations. Roughly speaking, a –module consists of a Breuil–Kisin module and the datum of an action of on where is a suitable subring of Fontaine’s period ring (and is a fixed choice of a uniformizer). An obstacle to applying the results of [CarusoLiu] to the torsion étale cohomology groups is that it is not quite clear when (duals of) such representations come as a quotient of two semistable lattices with Hodge–Tate weights in This is indeed the case in the situation when , and has good reduction by the aforementioned Fontaine–Messing theorem, and it was also shown in the case (no restriction on ) for with semistable reduction by Emerton and Gee in [EmertonGee1], but in general the question seems open.
Nevertheless, the idea of the proof of Theorem 1.1 is to follow the general strategy of Caruso and Liu. While one does not necessarily have semistable lattices and the associated –modules to work with, a suitable replacement comes from the recently developed cohomology theories of Bhatt–Morrow–Scholze and Bhatt–Scholze ([BMS1, BMS2, BhattScholze]). Concretely, to a smooth –adic formal scheme one can associate the “-torsion prismatic cohomologies”
where are the prismatic avatars of the – and Breuil–Kisin cohomologies from [BMS1] and [BMS2], resp. Taking and Li and Liu showed in [LiLiu] that is a –torsion Breuil–Kisin module, is a –torsion Breuil–Kisin–Fargues –module, and that these modules recover the étale cohomology group essentially due to the étale comparison theorem for prismatic cohomology from [BhattScholze]. The pair then serves as a suitable replacement of a –module in our context.
The most significant deviation from the strategy of [CarusoLiu] then stems from the fact that the pair obtained this way is “inherently –torsion”, that is, it does not come equipped with any apparent lift to analogous objects in characteristic . This is not the case in [CarusoLiu], where all torsion modules ultimately originate from a free –module . A key technical input in loc. cit. is to establish a partial control on the Galois action on inside namely, a condition of the form
(1.1) |
Here are certain ideals (that are shrinking with growing ). This is a “rational” fact, in the sense that this claim is a consequence of the description of the Galois action in terms of the monodromy operator on the associated Breuil module (cf. [Breuil], [LiuLatticesNew, §3.2]), a vector space over the characteristic field .
As a starting point for replacing (1.1) in our context, we turn to a result by Gee and Liu in [EmertonGee2, Appendix F] (see also [Ozeki, Theorem 3.8]). Given a finite free Breuil–Kisin module (of finite height) and a compatible structure of Breuil–Kisin–Fargues –module on such that the image of under the natural map lands in , the étale realization of is crystalline if and only if
() |
Here denotes the Teichmüller lift and are the elements of given by a collection of (compatible) –th roots of unity and a collection of –th roots of the chosen uniformizer , resp. We call condition () the crystalline condition. As the considered formal scheme is assumed to be smooth over , it is reasonable to expect that the same condition applies to the pair and , despite the fact that the Breuil–Kisin and Breuil–Kisin–Fargues modules coming from prismatic cohomology are not necessarily free.
This is indeed the case and, moreover, it can be shown that the crystalline condition even applies to the embedding of the chain complexes : to make sense of this claim, we model the cohomology theories by their associated Čech–Alexander complexes. These were introduced in [BhattScholze] in the case that is affine, but can be extended to (at least) arbitrary separated smooth –adic formal schemes. We are then able to verify the condition termwise for this pair of complexes. More generally, we introduce a decreasing series of ideals , where and then formulate and prove the analogue of () for and the action of As a consequence, we obtain:
Theorem 1.3 (Theorem 4.1, Corollary 4.5, Proposition 4.8).
Let be a smooth separated –adic formal scheme over .
-
(1)
Fix a compatible choice of Čech–Alexander complexes that compute and , resp. Then for all , the pair satisfies (termwise) the condition
() -
(2)
The associated prismatic cohomology groups satisfy the crystalline condition, that is, the condition
-
(3)
For all pairs of integers with , the –torsion prismatic cohomology groups satisfy the condition
Theorem 1.3 (3) specialized to provides the desired analogue of the property (1.1) of –modules and allows us to carry out the proof of Theorem 1.1.
As a consequence of Theorem 1.3 (2), we obtain a proof of crystallinity of the cohomology groups in the proper case partially by means of “formal” –adic Hodge theory (Corollary 4.6). This fact is usually established via a a direct comparison between crystalline and étale cohomology, and in this generality is originally due to Bhatt, Morrow and Scholze ([BMS1]). Of course, since our setup relies on the machinery of prismatic cohomology and especially the étale comparison, the proof can be considered independent of the one from [BMS1] only in that it avoids the crystalline comparison theorem for (prismatic) –cohomology.
The bounds of Theorem 1.1 compare to the already known bounds as follows. Whenever the bounds of “semistable type” are known to apply to the situation of (e.g. [CarusoLiu] when , [Hattori] when and is a scheme), the bounds from Theorem 1.1 agree with those bounds. The bounds tailored to crystalline representations ([Fontaine2, Abrashkin]) are slightly better but their applicability is quite limited ( and ).
The fact that the cohomology groups have an associated Breuil–Kisin module yields one more source of ramification estimates: in [Caruso], Caruso provides a very general bound for –torsion –modules based on their restriction to via Fontaine’s theory of étale –modules. Using the Breuil–Kisin module attached to , this bound becomes explicit (as discussed in more detail in Remark 5.6). Comparing this result to Theorem 1.1 is more ambiguous due to somewhat different shapes of the estimates, but roughly speaking, the estimate of Theorem 1.1 is approximately the same for , becomes worse when is absolutely tamely ramified with large ramification degree, and is expected to outperform Caruso’s bound in case of large wild absolute ramification (rel. to the tame part of the ramification).
In future work, we intend to extend the result of Theorem 1.1 to the case of arbitrary . This seems plausible thanks to the full statement of Theorem 1.3 (3). In a different direction, we plan to extend the results of the present paper to the case of semistable reduction, using the log–prismatic cohomology developed by Koshikawa in [Koshikawa]. An important facts in this regard are that the –log–prismatic cohomology groups are still Breuil–Kisin–Fargues –modules by a result of Česnavičius and Koshikawa ([CesnaviciusKoshikawa]) and that by results of Gao, a variant of the condition () might exist in the semistable case ([GaoBKGK]; see Remark 3.12 (3) below for details).
The outline of the paper is as follows. In §2 we establish some necessary technical results. Namely, we discuss non–zero divisors and regular sequences on derived complete and completely flat modules with respect to the weak topology of , and establish Čech–Alexander complexes in the case of a separated and smooth formal scheme. Next, §3 introduces the conditions (), studies their basic algebraic properties and discusses in particular the crystalline condition () in the case of Breuil–Kisin–Fargues –modules. In §4 we prove the conditions () for the Alexander–Čech complexes of a separated smooth –adic scheme over and , and draw some consequences for the inidividual cohomology groups (especially when is proper), proving Theorem 1.3. Finally, in §5 we establish the ramification bounds for mod étale cohomology, proving Theorem 1.1. Subsequently, we discuss in detail how the bounds from Theorem 1.1 compare to the various known bounds from the literature.
Let us set up some basic notation used throughout the paper. We fix a perfect field of characteristic and a finite totally ramified extension of degree where . We fix a uniformizer , and a compatible system of –th roots of in , the completion of algebraic closure of . Setting , the choice of determines a surjective map via ; the kernel of this map is generated by an Eisenstein polynomial of degree . is endowed with a Frobenius lift (hence a –structure) extending the one on by .
Denote where denotes the Witt vectors construction and is the tilt of , . The choice of the system describes an element , and hence an embedding of into via where denotes the Teichmüller lift. Under this embedding, is sent to a generator of the kernel of the canonical map that lifts the canonical projection Consequently, is a map of prisms. It is known that under such embedding, is faithfully flat over (see e.g. [EmertonGee2, Proposition 2.2.13]).
Similarly, we fix a choice of a compatible system of primitive –th roots of unity . This defines an element of in an analogous manner, and the embedding extends to a map (actually still an embedding by [Caruso, Proposition 1.14]) by additionally setting . Additionally, we denote by the element . It is well–known that this is another generator of , therefore is a unit in .
The choices of and , hence also the maps and , remain fixed throughout. For this reason, we often refer to as , as , as etc.
Throughout the paper, we use freely the language of prisms and –rings from [BhattScholze], and we adopt much of the related notation and conventions. In particular, a formal scheme over a –adically complete ring always means a –adic formal scheme, and it is called smooth if it is locally of the form for a (derived333As we will always consider the base to have bounded –torsion, there is no distinction between derived –completion and –adic completion in this case.) –completely smooth –algebra – that is, a –complete –algebra such that is a smooth –algebra and for all . By the results of Elkik [Elkik] and the discussion in [BhattScholze, §1.2], is equivalently the –adic completion of a smooth -algebra.
Acknowledgements. I would like to express my gratitude to my Ph.D. advisor Tong Liu for suggesting the topic of this paper, his constant encouragement and many comments, suggestions and valuable insights. Many thanks go to Deepam Patel and Shuddhodan Kadattur Vasudevan for organizing the prismatic cohomology learning seminar at Purdue University in Fall 2019, and to Donu Arapura for a useful discussion of Čech theory. I would like to thank Xavier Caruso, Shin Hattori and Shizhang Li for reading an earlier version of this paper, and for providing me with useful comments and questions. The present paper is an adapted version of the author’s Ph.D. thesis at Purdue University. During the preparation of the paper, the author was partially supported by the Ross Fellowship and the Bilsland Fellowship of Purdue University, as well as Graduate School Summer Research Grants of Purdue University during summers 2019–2021.
2. Preparations
2.1. Regularity on –completely flat modules
The goal of this section is to prove that every –complete and –completely flat –module is torsion–free, and that any sequence with is regular on such modules.
Regarding completions and complete flatness, we adopt the terminology of [stacks, 091N], [BhattScholze], but since we apply these notions mostly to modules as opposed to objects of derived categories, our treatment is closer in spirit to [Positselski], [Rezk] and [YekutieliFlatness]. Given a ring and a finitely generated ideal , the derived –completion444That is, this is derived –completion of as a module. This will be sufficient to consider for our purposes. of an –module is
(2.2) |
is said to be derived –complete if the natural map is an isomorphism. This is equivalent to the vanishing of for and all (equivalently, for for all ), and as a consequence, it can be shown that the category of derived –complete modules forms a full abelian subcategory of the category of all –modules with exact inclusion functor (and the derived –completion is its left adjoint; in particular, derived –completion is right exact as a functor on –modules). Another consequence is that derived –completeness is equivalent to derived –completeness when are two finitely generated ideals and . There is always a natural surjection where stands for –adic completion, which will be reffered to as classical –completion for the rest of the paper. Just like for classsicaly –complete modules, if is derived –complete, then implies (this is referred to as derived Nakayama lemma).
A convenient consequence of the completion formula (2.2) is that in the case when is a derived –complete –algebra, the isomorphism picks a preferred representative in for the power series symbol as the preimage of the class represented by . This gives an algebraically well–behaved notion of power series summation despite the fact that is not necessarily –adically separated555This operation further leads to the notion of contramodules, discussed e.g. in [Positselski]..
An –module is said to be –completely (faithfully) flat if for all and is a (faithfully) flat –module. Just like for derived completeness, –complete flatness is equivalent to –complete flatness when is another finitely generated ideal with 666However, note that while (derived) –completeness more generally implies (derived) –completeness when is a finitely generated ideal contained in , the “opposite” works for flatness, i.e. –complete flatness implies –complete flatness when when is a finitely generated ideal with ..
Let us start by a brief discussion of regular sequences on derived complete modules in general. For that purpose, given an –module and , denote by the usual Koszul complex and let denote the -th Koszul homology of with respect to .
The first lemma is a straightforward generalization of standard facts about Koszul homology (e.g. [Matsumura, Theorem 16.5]) and regularity on finitely generated modules.
Lemma 2.1.
Let be a ring, a finitely generated ideal and let be a nonzero derived -complete module. Let . Then
-
(1)
forms a regular sequence on if and only if for all if and only if .
-
(2)
In this situation, any permutation of is also a regular sequence on .
Proof.
As Koszul homology is insensitive to the order of the elements , part (2) follows immediately from (1).
To prove (1), the forward implications are standard and hold in full generality (see e.g. [Matsumura, Theorem 16.5]). It remains to prove that the sequence is regular on if . We proceed by induction on . The case is clear ( by definition, and follows by derived Nakayama). Let , and denote the truncated sequence . Then we have which produces a short exact sequence
of chain complexes. Taking homologies results in a long exact sequence
(as in [Matsumura, Theorem 7.4]). By assumption, and thus, where . Upon observing that is obtained from finite direct sum of copies of by repeatedly taking kernels and cokernels, it is derived –complete. Thus, derived Nakayama implies that as well, and by induction hypothesis, is a regular sequence on . Finally, the above exact sequence also implies that is injective on and is satisfied thanks to derived Nakayama again. This finishes the proof. ∎
Corollary 2.2.
Let be a derived –complete ring for an ideal where is a regular sequence on , and let be a nonzero derived –complete –module that is –completely flat. Then is a regular sequence on and consequently, each is a non–zero divisor on .
Proof.
By Lemma 2.1 (1), for all , hence is a free resolution of . Thus, on one hand, the complex computes , hence is acyclic in positive degrees by –complete flatness; on the other hand, this complex is by definition . We may thus conclude that for all . By Lemma 2.1, is a regular sequence on , and it remains regular on after arbitrary permutation. This proves the claim. ∎
Now we specialize to the case at hand, that is, . Recall that this is a domain and so is (which is a rank valuation ring).
Lemma 2.3.
For every element forms a regular sequence, and for all we have the equality . Furthermore, the ideal is equal to , the unique maximal ideal of . In particular, given two choices as above, we have .
In particular, the equalities “” imply that all the –adic topologies (for as above) are equivalent to each other; this is the so–called weak topology on (usually defined as –adic topology in our notation), and it is standard that is complete with respect to this topology.
Proof.
By assumption, the image of in is non–zero and non–unit in (non–unit since and ). Thus, is a non–zero divisor both on and on , hence the claim that follows for every . The element is itself non–zero divisor on and thus, is a regular sequence.
To obtain for general , one can e.g. use induction on using the fact that is a non-zero divisor on (or simply note that one may replace elements in regular sequences by arbitrary positive powers).
To prove the second assertion, note that since is a rank valuation ring. It follows that is the unique maximal ideal of above , hence the unique maximal ideal since , and that is equal to this ideal. ∎
We are ready to prove the claim mentioned at the beginning of the section.
Corollary 2.4.
Let be a derived –complete and –completely flat –module, and let . Then is a regular sequence on . In particular, for each , we have . Consequently, is a torsion–free –module.
Proof.
By Lemma 2.3, and are derived –complete and is –completely flat over , and is a regular sequence on . Corollary 2.2 then proves the claim about regular sequence. The sequence is then also regular on , and the claim follows. To prove the “consequently” part, let be a non–zero and non–unit element of . Since is classically –complete, we have , and so there exist such that with . If is a unit, then is a non–zero divisor on since so is . Otherwise , so is a regular sequence on , and so is (e.g. by Lemma 2.1). In particular are both non–zero divisors on , and hence so is . ∎
Finally, we record the following consequence on flatness of –completely flat modules modulo powers of that seems interesting on its own.
Corollary 2.5.
Let , and let be a derived –complete and –completely (faithfully) flat –module. Then is derived –complete and –completely (faithfully) flat. In particular, is a flat –module for every .
Proof.
The fact that is derived –complete is clear since it is derived –complete. We need to show that is a flat –module and that for all . The second claim is a consequence of the fact that is a non–zero divisor on both and by Corollary 2.4. For the first claim, note that is a valuation ring and therefore it is enough to show that is a torsion–free –module. This follows again by Corollary 2.4.
For the ‘faithful’ version, note that both the statements that is faithfully flat over and that is faithfully flat over are now equivalent to the statement where is the unique maximal ideal of . ∎
2.2. Čech–Alexander complex
Next, we discuss the construction of Čech–Alexander complexes for computing prismatic cohomology, introduced in [BhattScholze] in the affine case, in a global situation. Throughout this section, let be a fixed bounded base prism, and let be a smooth separated –adic formal scheme over Recall that denotes the site whose underlying category is the opposite of the category of bounded prisms over together with a map of formal schemes over . Covers in are given by the opposites of faithfully flat maps of prisms, meaning that is –completely flat over . The prismatic cohomology is then defined as the sheaf cohomology ( where is the terminal sheaf) for the sheaf on defined by .
Additionally, let us denote by the site of all bounded prisms, i.e the opposite of the category of all bounded prisms and their maps, with topology given by faithfully flat maps of prisms.
In order to discuss the Čech–Alexander complex in a non-affine situation, a slight modification of the topology on is convenient. The following proposition motivates the change.
Proposition 2.6.
Let be a bounded prism.
-
(1)
Given a collection of maps of (bounded) prisms the canonical map is a map of (bounded) prisms.
-
(2)
is flat over if and only if each is flat over . In that situation, is faithfully flat prism over if and only if the family of maps of formal spectra is jointly surjective.
-
(3)
Let be an element. Then , where stands for the derived (equivalently, classical) –completion, is a bounded prism777We do consider the zero ring with its zero ideal a prism, hence allow the possibility of , which occurs e.g. when Whether the zero ring satisfies Definition 3.2 of [BhattScholze] depends on whether the inclusion of the empty scheme to itself is considered an effective Cartier divisor; following the usual definitions pedantically, it indeed seems to be the case. Also some related claims, such as [BhattScholze, Lemma 3.7 (3)] or [BhattNotes, Lecture 5, Corollary 5.2], suggest that the zero ring is allowed as a prism., and the map is a flat map of prisms.
-
(4)
Let be a collection of elements generating the unit ideal. Then the canonical map is a faithfully flat map of (bounded) prisms.
Proof.
The proof of (1) is more or less formal. The ring has a unique ––algebra structure since the forgetful functor from –rings to rings preserves limits, and is as product of –complete rings –complete. Clearly is an invertible ideal since each is. In particular, , hence a prism by [BhattScholze, Lemma 3.5]. Assuming that all are bounded, from we have for big enough so that for all , showing that is bounded.
The (–complete) flatness part of (2) is clear. For the faithful flatness statement, note that , hence is faithfully flat if and only if the map of spectra is surjective.
Let us prove (3). Since has the equality shows that for each is a unit in . Consequently, as in [BhattScholze, Remark 2.16], for , and the latter has a unique –structure extending that of by [BhattScholze, Lemmas 2.15 and 2.17]. In particular, is a –completely flat ––algebra, hence is flat prism over by [BhattScholze, Lemma 3.7 (3)].
Part (4) follows formally from parts (1)–(3). ∎
Construction 2.7.
Denote by the site whose underlying category is . The covers on are given by the opposites of finite families of flat maps of prisms such that the associated maps are jointly surjective. Let us call these “faithfully flat families” for short. The covers of the initial object 888That is, corresponds to the zero ring, which we consider to be a prism as per the previous footnote. are the empty cover and the identity. We similarly extend to , that is, we proclaim the identity cover and the empty cover to be covers of , and generally proclaim (finite) faithfully flat families to be covers.
Clearly isomorphisms as well as composition of covers are covers in both cases. To check that and are sites, it thus remains to check the base change axiom. This is trivial for situations involving so it remains to check that given a faithfully flat family and a map of prisms , the fibre products999Here we mean fibre products in the variance of the site, i.e. “pushouts of prisms”. We use the symbol to denote this operation. in exist and the collection is a faithfully flat family; the existence and –complete flatness follows by the same proof as in [BhattScholze, Corollary 3.12], only with “–completely faithfully flat” replaced by “–completely flat” throughout, and the fact that the family is faithfully flat follows as well, since (and using Remark 2.8 (1) below).
Remark 2.8.
-
(1)
Note that for a finite family of objects in the structure map of the product together with the map of formal spectra (induced from the maps for individual ’s)
makes into an object of that is easily seen to be the coproduct of ’s. In view of Proposition 2.6 (2), one thus arrives at the equivalent formulation
That is, is the (finitely) superextensive site having covers of as singleton covers. (Similar considerations apply to and .)
-
(2)
The two sites are honestly different in that they define different categories of sheaves. Namely, for every finite coproduct , the collection of canonical maps forms a –cover, and the sheaf axiom forces upon the identity which is not automatic101010For example, every constant presheaf is a sheaf for a topology given by singleton covers only, which is not the case for . In fact, can be identified with the full category of consisting of all sheaves compatible with finite disjoint unions in the sense above. In particular, the structure sheaf is a sheaf for the –topology. (Again, the same is true for and , including the fact that is a sheaf.)
Despite the above fine distinction, for the purposes of prismatic cohomology, the two topologies are interchangeable. This is a consequence of the following lemma.
Lemma 2.9.
Given an object one has for .
Proof.
The sheaf on has vanishing positive Čech cohomology essentially by the proof of [BhattScholze, Corollary 3.12]: one needs to show acyclicity of the Čech complex for any –cover but the resulting Čech complex is identical to that for the –cover , for which the acyclicity is proved in [BhattScholze, Corollary 3.12]. By a general result (e.g. [stacks, 03F9]), this implies vanishing of for all bounded prisms and all .
Now we make use of the fact that cohomology of an object can be computed as the cohomology of the corresponding slice site, [stacks, 03F3]. Let After forgetting structure, we may view as an object of as well, and then [stacks, 03F3] implies that for every we have the isomorphisms
(where for a site and denotes the slice site). Upon noting that the slice sites are equivalent sites (in a manner that identifies the two versions of the sheaf ), the claim follows. ∎
Corollary 2.10.
One has
Proof.
The coverings of contain the coverings of so we are in the situation of [stacks, 0EWK], namely, there is a morphism of sites given by the identity functor of the underlying categories, the pushforward functor being the natural inclusion and the (exact) inverse image functor is the sheafification with respect to the “”-topology. One has
(where denotes the inclusion of abelian sheaves in this context), hence
and to conclude it is enough to show that . But is the sheafification of the presheaf given by ([stacks, 072W]), which is by Lemma 2.9. Thus, , which proves the claim. ∎
For an open –adic formal subscheme , denote by the functor sending to the set of factorizations of the implicit map through that is,
Let correspond to a morphism in . If factors through then so does . It follows that forms a presheaf on (with transition maps given by when , and the empty map otherwise). Note that is the terminal sheaf.
Proposition 2.11.
is a sheaf on .
Proof.
Consider a cover in which is given by a faithully flat family . One needs to check that the sequence
is an equalizer sequence. All the terms have at most one element; consequently, there are just two cases to consider, depending on whether the middle term is empty or not. In both cases, the pair of maps on the right necessarily agree, and so one needs to see that the map on the left is an isomorphism. This is clear in the case when the middle term is empty (since the only map into an empty set is an isomorphism). It remains to consider the case when the middle term is nonempty, which means that for all . In this case we need to show that . Since the maps are jointly surjective and each lands in , it follows that so does the map . Thus, , which finishes the proof. ∎
When is affine, one can cover the sheaf by a representable sheaf. Note that the construction of the representing object is essentially equivalent to Construction 4.17 of [BhattScholze].
Construction 2.12 (Čech–Alexander cover of ).
Let us additionally assume that is affine. Choose a surjection where is a –completed free –algebra. Denote by the kernel of the surjection. Then there is a commutative diagram with exact rows
where stands for derived –completion. Here for an –algebra , denotes the “–envelope” of , that is, the –algebra initial among –algebras endowed with an ––algebra structure. Note that , where is the polynomial algebra before completion; in particular, since is a flat –algebra (essentially by [BhattScholze, Lemma 2.11]), it follows that is –completely flat –algebra. Consequently, the completions in the lower row of the diagram can be equivalently taken as classical -completions (cf. [BhattScholze, Lemma 3.7]).
Denote by the image of the map i.e. the –complete ideal of topologically generated by . Then we have a short exact sequence
Let be the prismatic envelope of . It follows from [BhattScholze, Proposition 3.13, Example 3.14] that exists and is given by a flat prism over . The map
of –complete rings corresponds to the map of formal schemes . This defines an object of which we call a Čech–Alexander cover of .
Remarks 2.13.
-
(1)
Note that is equivalently the prismatic envelope of . Moreover, when the ideal is finitely generated, one has the equality
-
(2)
Since the ring in Construction 2.12 is a -completely smooth –algebra, it is in particular a –completion of a finitely presented –algebra. It follows that the map may be chosen so that is the (derived) –completion of a polynomial –algebra of finite type, with the kernel finitely generated. While such a choice may be preferable, we formulate the construction without imposing it, as it may be convenient to allow non–finite–type free algebras in the construction e.g. for the reasons of functoriality (see the remark at the end of [BhattScholze, Construction 4.17]).
Proposition 2.14.
Denote by the sheaf represented by the object . There exists a unique map of sheaves , and it is an epimorphism.
Proof.
If with this means that factors through since it factors through Thus, we also have , and so the (necessarily unique) map is defined. When is empty, the map is still defined and unique, namely given by the empty map. Thus, the claimed morphism of sheaves exists and is unique.
We show that this map is an epimorphism. Let such that , i.e. factors through , and consider the map associated to the map . Since is a –completed free –algebra surjecting onto and is –complete, the map admits a lift . This induces an ––algebra map which gives a morphism of –pairs , and further the map of prisms using the universal properties of objects involved. It is easy to see that this is indeed (the opposite of) a morphism in . This shows that is nonempty whenever is. Thus, the map is an epimorphism. ∎
Let be an affine open cover of . For and a multi–index denote by the intersection . As is assumed to be separated, each is affine and we write .
Remark 2.15 (Binary products in ).
For , let us denote their binary product by . Let us describe it explicitly at least under the additional assumptions that
-
(1)
are flat prisms over
-
(2)
there are affine opens such that .
Set and denote the rings corresponding to the affine open sets and by and resp. Then any object with maps both to and lives over i.e. satisfies . This justifies the following construction. Consider the following commutative diagram, where denotes the pushout of –complete commutative rings, i.e. taking the classically –completed tensor product (and is the derived, but equivalently classical, –completion of ):
Let be the kernel of the map Then the product is given by the prismatic envelope of the –pair .
Proposition 2.16.
The Čech–Alexander covers can be chosen so that for all we have
Proof.
Clearly it is enough to show the statement for binary products. More precisely, given two affine opens and an arbitrary initial choice of and we show that can be chosen so that the resulting Čech–Alexander cover of is equal to . For the purposes of this proof, let us refer to a prismatic envelope of a –pair also as “the prismatic envelope of the arrow ”.
Consider as in Construction 2.12, and set . Then one has the induced surjection , which can be followed by the induced map . This latter map is surjective as well since is separated, and therefore the composition of these two maps is surjective, with the kernel that contains . We may construct a diagram analogous to the one from Remark 2.15, which becomes the diagram
where the expected arrow in the central column is replaced by an isomorphic one, namely the map obtained from the surjection by the procedure as in Construction 2.12. Now is obtained as the prismatic envelope of this composed central arrow, while is obtained the same way, but only after replacing the downward arrows on the left and right by their prismatic envelopes. Comparing universal properties, one easily sees that the resulting central prismatic envelope remains unchanged, proving the claim. ∎
Remark 2.17.
Suppose that for each , the initial choice of the map has been made as in Remark 2.13 (2), that is, is the –completion of a finite type free –algebra and the ideal is finitely generated. If now is the –completed free –algebra for obtained by iterating the procedure in the proof of Proposition 2.16, it is easy to see that in this case, the algebra is still the –completion of a finite type free –algebra, and it can be shown that the corresponding ideal is finitely generated.
In more detail, given a ring and a finitely generated ideal , Let us call a –algebra –completely finitely presented if is derived –complete and there exists a map from the polynomial ring in finitely many variables such that the derived –completed map is surjective and with a finitely generated kernel. Then the algebra corresponding to is –completely finitely presented by Remark 2.13 (2), and since is the –completion of a finite type polynomial –algebra, the following lemma shows that is finitely generated.
Lemma 2.18.
Let be a –completely finitely presented –algebra, and consider a –algebra map from a polynomial algebra in finitely many variables such that is surjective. Then the kernel of is finitely generated.
Proof.
The proof is an adaptation of the proof of [stacks, 00R2], which is a similar assertion about finitely presented algebras. Consider as in Remark 2.17, and additionally let us fix a generating set of .
For let us choose such that . Then one can define a surjective map
and it is easy to see that That is, we have an exact sequence
where the map on the left is a module map determined by the finite set of generators of . After taking the derived –completion, the sequence becomes the exact sequence
That is, we have a surjective map , which is determined on topological generators by and the kernel of is .
Next, we choose elements such that for each . Then we have a surjective map given by and , which has the property that . That is,
and therefore since is surjective. But is finitely generated by the previous, and hence so is .∎
Proposition 2.19.
The map (where denotes the coproduct in ) to the final object is an epimorphism, hence so is the map .
Proof.
It is enough to show that for a given object there is a faithfully flat family in such that for all where denotes the coproduct of presheaves.
With that aim, let us first consider the preimages of each under the map . This is an open cover of that corresponds to an open cover of . One can then choose such that refines this cover, i.e. every corresponds to an open subset of for some index .
The elements generate the unit ideal of since they do so modulo which is contained in Thus, the family
is easily seen to give the desired faithfully flat family, with each nonempty, since each factors through by construction. ∎
Remark 2.20.
The proof of Proposition 2.19 is the one step where we used the relaxation of the topology, namely the fact that the faithfully flat cover can be replaced by the family .
Finally, we obtain the Čech–Alexander complexes in the global case.
Proposition 2.21.
With the notation for as above and the choice of Čech–Alexander covers as in Proposition 2.16, is modelled by the Čech–Alexander complex
() |
(that is, the complex associated to the cosimplicial ring ).
Proof.
By [stacks, 079Z], the epimorphism of sheaves from Proposition 2.19 implies that there is a spectral sequence with -page
converging to where we implicitly used Corollary 2.10 and the fact that as in Proposition 2.16, and similarly for higher multi–indices.
By Lemma 2.9, for every and every multi–index . The first page is therefore concentrated in a single row of the form and thus, the spectral sequence collapses on the second page. This proves that the cohomologies of are computed as cohomologies of , but in fact, this yields a quasi–isomorphism of the complexes themselves. (For example, analyzing the proof of [stacks, 079Z] via [stacks, 03OW], the double complex of the above spectral sequence comes with a natural map and a natural quasi–isomorphism when the spectral sequence collapses as above, is also a quasi–isomorphism). ∎
Remarks 2.22.
-
(1)
Just as in the affine case, the formation of Čech–Alexander complexes is compatible with “termwise flat base–change” on the base prism essentially by [BhattScholze, Proposition 3.13]. That is, if is a Čech–Alexander complex modelling and is a flat map of prisms, then the complex is a Čech–Alexander complex that computes .
-
(2)
Let now be the prism and let be of the form where is a smooth separated formal –scheme. A convenient way to describe the –action on is via base–change: given the action of on gives a map of prisms , and since comes from . Base–change theorem for prismatic cohomology [BhattScholze, Theorem 1.8 (5)] then gives an –linear map untwisting by on the left, this gives an ––semilinear action map . The exact same procedure defines the –action on the Čech–Alexander complexes modelling the cohomology theories since they are base–change compatible in the sense above.
3. The conditions ()
3.1. Definition and basic properties
In order to describe the conditions (), we need to fix more notation. For a natural number , denote by the field (where is the compatible chain of –th roots of chosen before, i.e. so that in ), and set . Further set and for , set . Note that the field is the Galois closure of . Denote by the Galois group and by the group for .
The group is generated by its two subgroups and (by [LiuBreuilConjecture, Lemma 5.1.2]). The subgroup is normal, and its element is uniquely determined by its action on the elements , which takes the form , with the integers unique modulo and compatible with each other as increases. It follows that , with a topological generator given by (where, again, ’s are chosen as before, so that ).
Similarly, the image of in is the subgroup . Clearly contains and the intersection of with is Just as in the case, is generated by these two subgroups, with the subgroup normal and topologically generated by the element .
There is a natural –action on extended functorially from the natural action on . This action makes the map –equivariant, in particular, the kernel is –stable. The –action on the –closure of in factors through . Note that the subgroup of acts trivially on elements of , and the action of the subgroup is determined by the equality .
For an integer and between and , denote by the element
(recall that ), and set
For convenience of notation, we further set and .
We are concerned with the following conditions.
Definition 3.1.
Let be an –module endowed with a ––semilinear action, let be an –module and let be an –linear map. Let be an integer or .
-
(1)
An element is called a ()–element if for every ,
-
(2)
We say that the pair satisfies the condition () if for every element , the image of in is ().
-
(3)
An element is called a ()–element if for every , there is an element such that
-
(4)
We say that the pair satisfies the condition () if for every element , the image of in is ().
-
(5)
Aditionally, we call ()–elements crystalline elements and we call the condition () the crystalline condition.
Remarks 3.2.
-
(1)
Since the crystalline condition equivalently states that for all and all in the image of
The reason for the extra terminology in the case is that the condition is connected with a criterion for certain representations to be crystalline, as discussed in §3.2. The higher conditions () will on the other hand find application in computing bounds on ramification of –torsion étale cohomology. The conditions () serve an auxillary purpose. Clearly () implies (). The conditions (), () are clearly both equivalent to the condition .
-
(2)
Strictly speaking, one should talk about the crystalline condition (or ()) for the map , but we choose to talk about the the crystalline condition (or ()) for the pair instead, leaving the datum of the map implicit. This is because typically we consider the situation that is an –submodule of and via the natural map (or the derived –completed variant, ). Also note that satisfies the condition () if and only if does.
Lemma 3.3.
For any integer , the ideals and are –stable.
Proof.
It is enough to prove that the ideals and are –stable. Note that the –stability of implies –stability of for any since is a –equivariant automorphism of . Once we know this, we know that equals to times a unit for every and , the same is then true of , hence also of all the elements and it follows that is –stable.
Given , for an ineger unique modulo and such that . It follows that for a –adic integer (). (The –exponentiation used here is defined by and the considered limit is with respect to the weak topology.) Thus, is –stable.
Similarly, we have for integers coprime to , unique modulo and compatible with each other as grows. It follows that for , and so The resulting expression is still divisible by . To see that, fix the integers to have all positive representatives. Then the claim follows from the formula
upon noting that the sequence of elements is still –adically (i.e. weakly) convergent thanks to Lemma 2.3. ∎
Let be as in Definition 3.1. Lemma 3.3 shows that the modules and have a well–defined –action. Consequently, we get the following restatement of the conditions (), ().
Lemma 3.4.
Given as in Definition 3.1, the pair satisfies the condition () ((), resp.) if and only if the image of in (, resp.) lands in .
In the case of the above–mentioned condition the –closure of in is contained in the –submodule , and thus, the –action on it factors through . Under mild assumtions on , the –action on the elements of is ultimately determined by , the topological generator of . Consequently, the conditions () are also determined by the action of this single element:
Lemma 3.5.
Let be as in Definition 3.1. Additionally, assume that is classically –complete and –completely flat, and that the pair satisfies (). Then the action of on elements of makes sense, and satisfies () if and only if
Proof.
Clearly the stated condition is necessary. To prove sufficiency, assume the above condition for . By the fixed–point interpretation of the condition () as in Lemma 3.4, it is clear that the analogous condition holds for every element .
Next, assume that is an element of , the –adic closure of . Then can be written as , with the sequence of integers –adically convergent. For by continuity we have , which is equal to with . Since the sequence is still convergent (using the fact that the –adic topology is the –adic topology, and that is a regular sequence on ), we have that where .
To conclude, note that a general element of is of the form where and Then for , by the assumption we have , and so the condition () is proved by the previous part. ∎
Let us now discuss some basic algebraic properties of the conditions () and (). The basic situation when they are satisfied is the inclusion itself.
Lemma 3.6.
The pair satisfies the conditions () (hence also ()) for all .
Proof.
Note that satisfies the assumptions of Lemma 3.5, so it is enough to consider the action of the element . For an element we have
and thus,
Since divides for each , the obtained series has coefficients in , showing that as desired. ∎
The following lemma shows that in various contexts, it is often sufficient to verify the conditions (), () on generators.
Lemma 3.7.
Fix an integer . Let (C) be either the condition () or ().
-
(1)
Let be an –module with a ––semilinear action. The set of all (C)–elements forms an –submodule of .
-
(2)
Let be an –algebra endowed with a –semilinear action. The set of (C)–elements of forms an –subalgebra of .
-
(3)
If the algebra from (2) is additionally ––algebra such that acts by –maps (i.e. for all ) then the set of all (C)–elements forms a ––subalgebra of .
-
(4)
If the algebra as in (2) is additionally derived –complete and is a map of –algebras that satisfies the condition (C), then so does where is the derived –completion of . In particular, the set of all (C)–elements in forms a derived –complete –subalgebra of .
Proof.
Let be the ideal if (C)=() and the ideal if (C)=(). In view of Lemma 3.4, the sets described in (1),(2) are obtained as the preimages of (ring , resp.) under the canonical projection ( resp.). As these –fixed points form an –module (–algebra, resp.) by Lemma 3.6, this proves (1) and (2).
Similarly, to prove (3) we need to prove only that the ideal is a –ideal and therefore the canonical projection is a map of –rings.
Let us argue first in the case (). As , we have
Recall that divides . The numerator of the last fraction is thus divisible by and since by Lemma 2.3, divides the whole fraction in . (We note that this is true for every integer , in particular , as well.)
Let us now prove that the ideal (hence also ) is a –ideal. For any between and , we have
The numerator is divisible by , and so is the whole fraction thanks to Lemma 2.3. Thus, we have that is a multiple of . Finally, when , we have , and is thus a multiple of by the previous. Consequently, is divisible by . This shows that is a –ideal.
Finally, let us prove (4). Note that , hence even as ideals of ; consequently, the derived –completion agrees with the derived –completion both for – and –modules. We may therefore replace –completions with –completions throughout.
Since is derived –complete, any power series of the form
with defines a unique111111Here we are using the preferred representatives of powers series as mentioned at the beginning of §2.1. element , and comes from if and only if the coefficients may be chosen in the image of the map . Assuming this, for we have
where is the –unit such that . Thus, it is clearly enough to show, upon assuming the condition (C) for , that the terms and are in when . (Note that an element with is itself in , since is finitely generated.)
We have by assumption, so it remains to treat the term . Since is divisible by , it is also divisible by by Lemma 3.6. Thus, the terms are divisible by when , and are when ; in either case, they are members of .
To prove the second assertion of (4), let now be the –subalgebra of all crystalline elements. By the previous, the map satisfies (C), and hence the image of this map consists of (C)–elements. Thus, we have and hence, is derived –complete since so is . ∎
Remark 3.8.
One consequence of Lemma 3.7 is that the –subalgebra of formed by all crystalline elements (or even ()–elements) forms a prism, with the distinguished invertible ideal . As Lemma 3.6 works for any choice of Breuil–Kisin prism associated to in , contains all of these (in particular, it contains all –translates of ).
For future use in applications to –torsion modules, we consider the following approximation of the ideals appearing in the conditions ().
Lemma 3.9.
Consider a pair of integers with . Set . Then the image of the ideal in the ring is contained in the ideal . That is, we have
Proof.
When there is nothing to prove, therefore we may assume that . In the definition of , we may replace the elements
by the elements
since the quotients are –units.
It is thus enough to show that for every with the element
taken modulo is divisible by .
This is clear when , and so it remains to discuss the cases when Write (with ). Then it is enough to show that
() |
taken modulo is divisible by
Since we are interested in the product ( ‣ 3.1) only modulo , in expanding the brackets we may ignore the terms that use the expressions of the form at least times. Each of the remaining terms contains the product of at least distinct terms from the following list:
Thus, each of the remaining terms is divisible by (at least)
which is more than needed. This finishes the proof. ∎
3.2. Crystalline condition for Breuil–Kisin–Fargues –modules
The situation of central interest regarding the crystalline condition is the inclusion such that is an isomorphism, where is a Breuil–Kisin module and is a Breuil–Kisin–Fargues –module. The version of these notions used in this paper is tailored to the context of prismatic cohomology. Namely, we have:
Definition 3.10.
-
(1)
A Breuil–Kisin module is a finitely generated –module together with a –linear isomorphism
For a positive integer , the Breuil–Kisin module is said to be of height if is induced (by linearization and localization) by a –semilinear map such that, denoting its linearization, there exists an –linear map such that both the compositions and are multiplication by . A Breuil–Kisin module is of finite height if it is of height for some .
-
(2)
A Breuil–Kisin–Fargues module is a finitely presented –module such that is a free –module, together with an –linear isomorphism
Similarly, the Breuil–Kisin–Fargues module is called of height if comes from a –semilinear map admitting an –linear map such that and are multiplication maps by , where is the inearization of . A Breuil–Kisin–Fargues module is of finite height if it is of height for some .
-
(3)
A Breuil–Kisin–Fargues –module (of height , of finite height, resp.) is a Breuil–Kisin–Fargues module (of height , of finite height, resp.) that is additionally endowed with an –semilinear –action that makes –equivariant (that makes also –equivariant in the finite height cases).
That is, the definition of a Breuil–Kisin module agrees with the one in [BMS1], and is a Breuil–Kisin–Fargues module in the sense of the above definition if and only if is a Breuil–Kisin–Fargues module in the sense of [BMS1]121212This is to account for the fact that while Breuil–Kisin–Fargues modules in the sense of [BMS1] appear as –cohomology groups of smooth proper formal schemes, Breuil–Kisin–Fargues modules in the above sense appear as prismatic –cohomology groups of smooth proper formal schemes.. The notion of Breuil–Kisin module of height agrees with what is called “(generalized) Kisin modules of height ” in [LiLiu]. The above notion of finite height Breuil–Kisin–Fargues modules agrees with the one from [EmertonGee2, Appendix F] except that the modules are not assumed to be free. Also note that under these definitions, for a Breuil–Kisin module (of height resp.), the –module is a Breuil–Kisin–Fargues module (of height resp.), without the need to twist the embedding by .
The connection between Breuil–Kisin–, Breuil–Kisin–Fargues –modules and the crystalline condition (justifying its name) is the following theorem.
Theorem 3.11 ([EmertonGee2, Appendix F], [GaoBKGK]).
Let be a free Breuil–Kisin–Fargues –module which admits as an –submodule a free Breuil–Kisin module of finite height, such that (as Breuil–Kisin–Fargues modules) via the natural map, and such that the pair satisfies the crystalline condition. Then the étale realization of ,
is a crystalline representation.
Remarks 3.12.
-
(1)
Theorem 3.11 is actually an equivalence: If is crystalline, it can be shown that the pair satisfies the crystalline condition. We state the theorem in the one direction since this is the one that we use. However, the converse direction motivates why it is resonable to expect the crystalline condition for prismatic cohomology groups that is discussed in Section 4.
-
(2)
Strictly speaking, in [EmertonGee2, Appendix F] one assumes extra conditions on the pair (“satisfying all descents”); however, these extra assumptions are used only for a semistable version of the statement. Theorem 3.11 in its equivalence form is therefore only implicit in the proof of [EmertonGee2, Theorem F.11]. (See also [Ozeki, Theorem 3.8] for a closely related result.)
-
(3)
On the other hand, Theorem 3.11 in the one–sided form as above is a consequence of [GaoBKGK, Proposition 7.11] that essentially states that is crystalline if and only if the much weaker condition
is satisfied. We note a related result of loc. cit.: is semistable if and only if
This semistable criterion above might be a good starting point in generalizing the results of Sections 4 and 5 of the present paper to the case of semistable reduction, using the log–prismatic cohomology developed in [Koshikawa]. Thus, a natural question to ask is: Similarly to how the crystalline condition is a stronger version of the crystallinity criterion from [GaoBKGK], what is an analogous stronger (while still generally valid) version of the semistability criterion from [GaoBKGK]?
It will be convenient later to have version of Theorem 3.11 that applies to not necessarily free Breuil–Kisin and Breuil–Kisin–Fargues modules. Recall that, by [BMS1, Propostition 4.3], any Breuil–Kisin module is related to a free Breuil–Kisin module by a functorial exact sequence
where is a -torsion module for some and is supported at the maximal ideal . Taking the base–change to one obtains an analogous exact sequence
(also described by [BMS1, Proposition 4.13]) where is a free Breuil–Kisin–Fargues module. Clearly the maps and become isomorphisms after inverting .
Assume that is endowed with a –action that makes it a Breuil–Kisin–Fargues –module. The functoriality of the latter exact sequence implies that the –action on induces a –action on , endowing it with the structure of a free Breuil–Kisin–Faruges –module. In more detail, given , the semilinear action map induces an –linear map where . As is an isomorphism fixing , up to unit and the ideal it is easy to see that is itself a Breuil–Kisin–Fargues module, and the exact sequence from [BMS1, Proposition 4.13] for can be identified with the upper row of the diagram
where the second vertical map is the linearization of and the rest is induced by functoriality of the sequence. Finally, untwisting the third vertical map induces a semilinear map . Note that the module inherits the –action from ; it is easy to see that the -action on agrees with the one on when viewing as its submodule.
Proposition 3.13.
Assume that the pair satisfies the crystalline condition. Then so does the pair
Proof.
Notice that the crystalline condition is satisfied for and by [BMS1, Propositions 4.3, 4.13], this map can be identified with . Thus, the following lemma finishes the proof. ∎
Lemma 3.14.
Let be a free –module endowed with –semilinear –action and let be a free –submodule such that satisfies the crystalline condition. Then the pair satisfies the crystalline condition.
Proof.
Fix an element and . The crystalline condition holds after inverting , and so we have the equality
with . In other words (using that is a non-zero divisor on ), we have
where the last equality follows by Lemma 2.3 since is a free module. In particular, we have
for yet another element . As is a non–zero divisor on , hence on we may cancel out to conclude
as desired. ∎
Theorem 3.15.
The “free” assumption in Theorem 3.11 is superfluous. That is, given a Breuil–Kisin–Fargues –module together with its Breuil–Kisin––submodule of finite height such that and such that the pair satisfies the crystalline condition, the representation
is crystalline.
Proof.
With the notation as above, upon realizing that and agree, the result is a direct consequence of Proposition 3.13. ∎
4. Conditions () for cohomology
4.1. () for Čech–Alexander complexes
Let be a smooth separated –adic formal scheme over . Denote by a Čech–Alexander complex that models and set , computed termwise – by Remark 2.22, this is a Čech–Alexander complex modelling . We aim to prove the following.
Theorem 4.1.
For every and , the pair satisfies the condition ().
Let be an affine open formal subscheme. Then it is enough to prove the content of Theorem 4.1 for where and are the Čech–Alexander covers of and with respect to the base prism and , respectively, since the Čech–Alexander complexes termwise consist of products of such covers. Let .
Let us fix a choice of the free –algebra whose –completion is the algebra as in Construction 2.12, with being the kernel of the surjection . Then the corresponding choices at the –level are and and the associated –completed “–envelopes” are also related by the completed base change; that is, we have a diagram with exact rows
(4.3) |
By Remark 2.17, we may and do assume that the set of variables is finite, and that the ideal is finitely generated. Consequently, after replacing the maps on the left by their respective images (and invoking Remark 2.13 (1)), diagram (4.3) becomes
(4.4) |
where the rows are exact. The prescription determines uniquely a continuous, –semilinear Galois action by –maps on (and, in particular, this action satisfies for all and all ). Similarly, the term is given the (linear) –action prescribed by for every , and coming from the first row. This makes the map –equivariant, and therefore the kernel –stable. As a consequence, the action extends to the prismatic envelope where the action obtained this way agrees with the one indicated in Remark 2.22. Upon taking the prismatic envelope of the –pair , we arrive at the situation for which we wish to verify the conditions ().
With the goal of understanding the –action on even more explicitly, in similar spirit to the proof of [BhattScholze, Proposition 3.13] we employ the following approximation of the prismatic envelope.
Definition 4.2.
Let be a –ring, an ideal with a fixed generating set and let be an element. Denote by be the kernel of the –algebra map
and let be the -ideal in generated by . Then we denote by the –ring , and call it weak –blowup algebra of and .
That is, the above construction adjoins (in –sense) the fractions to together with all relations among them that exist in , making it possible to naturally compute with fractions.
Note that if is a map of ––algebras such that and this ideal is invertible, the fact that the localization map is injective shows that there is a unique map of ––algebras . (In fact, if happens to be a non–zero divisor on , then is initial among all such ––algebras; this justifies the name ’weak –blowup algebra’.)
The purpose of the construction is the following.
Proposition 4.3.
Let be a bounded orientable prism with an orientation . Consider a map of –pairs and assume that is a prismatic envelope for that is classically –complete. Let be a system of generators of . Then there is a surjective map of –rings , where denotes the classical –completion.
Note that the assumptions apply to a Čech–Alexander cover in place of since it is –completely flat over the base prism, hence classically –complete by [BhattScholze, Proposition 3.7].
Proof.
Since and and is a non–zero divisor on , there is an induced map and hence a map of –rings (using [BhattScholze, Lemma 2.17]).
To see that this map is surjective, let denote its image in , and denote by the inclusion of into . Then is (derived, and, consequently, clasically) –complete ––algebra with . It follows that is a prism by [BhattScholze, Lemma 3.5] and thus, by the universal property of , there is a map of ––algebras which is easily seen to be right inverse to . Hence, is surjective, proving the claim. ∎
Finally, we are ready to prove the following proposition which, as noted above, proves Theorem 4.1.
Proposition 4.4.
The pair satisfies the conditions () for every .
Proof.
Fix a generating set of , and set . We obtain a commutative diagram
(4.5) |
where the vertical maps are the surjective maps from Proposition 4.3, and the horizontal maps come from the –completed base change .
The –action on naturally extends to by the rule on generators
where is the –unit such that (note that the fraction on the right–hand side makes sense as ). The action can be again extended continuously to the –adic completion, and this action makes the right vertical map –equivariant.
It is therefore enough to prove the validity of the conditions () for the pair . By Lemma 3.7 (3),(4), it is enough to check the conditions for the topological generators of as an ––algebra, which are and
Fix an integer . Since the elements satisfy for every , by Lemma 3.7 the pair satisfies the stronger condition (). In particular, () holds not only for the variables , but also for since they come from .
Let us now fix an index and an element . We may write
for some (that are equal up to a multiplication by an –unit). Similarly, we have
with (again equal up to a unit).
Thus, regarding the generator we have that
This shows that each of the generators is a ()–element, which finishes the proof. ∎
4.2. Consequences for cohomology groups
Let us now use Theorem 4.1 to draw some conclusions for individual cohomology groups. The first is the crystalline condition for the prismatic cohomology groups and its consequence for –adic étale cohomology. As before, let be a separated smooth –adic formal scheme over . Denote by the base change , and by the geometric adic generic fiber.
Corollary 4.5.
For every the pair satisfies the conditions () and ().
Proof.
By the results of Section 2.2, we may and do model the cohomology theories by the Čech–Alexander complexes
and by Theorem 4.1 the conditions () and () termwise hold for this pair. The condition () for thus follows immediately, and it remains to verify the crystalline condition.
Each of the terms is –completely flat over , which means in particular that the terms are torsion–free by Corollary 2.4. Denote the differentials on by and , resp.
To prove the crystalline condition for cohomology groups, it is clearly enough to verify the condition at the level of cocycles. Given denote by its image in . For we have for some . As we have
and the torsion–freeness of implies that . Thus, as well, showing that as desired. ∎
When is proper over , we use Corollary 4.5 to reprove the result from [BMS1] that the étale cohomology groups are in this case crystalline representations.
Corollary 4.6.
Assume that is additionally proper over Then for any the –adic étale cohomology is a crystalline representation.
Proof.
It follows from [BhattScholze, Theorem 1.8] (and faithful flatness of ) that and are Breuil–Kisin and Breuil–Kisin–Fargues modules, resp., such that . Moreover, has the structure of a Breuil–Kisin–Fargues –module with
as –representations. By Corollary 4.5, the pair satisfies all the assumptions of Theorem 3.15. The claim thus follows. ∎
For the purposes of obtaining a bound on ramification of –torsion étale cohomology in §5, let us recall the notion of torsion prismatic cohomology as defined in [LiLiu], and discuss the consequences of the conditions () in this context.
Definition 4.7.
Given a bounded prism and a smooth –adic formal scheme over , the –torsion prismatic cohomology of is defined as
We denote the cohomology groups of by (and refer to them as –torsion prismatic cohomology groups).
Proposition 4.8.
Let be a pair of integers satisfying . Set Then the torsion prismatic cohomology groups satisfy the following condition:
Proof.
The proof is a slightly refined variant of the proof of Corollary 4.5. Consider again the associated Čech–Alexander complexes over and ,
Both of these complexes are given by torsion–free, hence –flat, modules by Corollary 2.4. Consequently, is modelled by , and similarly for and . That is, the considered maps between cohomology groups are obtained as the maps on cohomologies for the base–change map of chain complexes
and as in the proof of Corolary 4.5, it is enough to establish the desired condition for the respective groups of cocycles.
Set . Note that by Lemma 3.9, the condition () for the pair of complexes implies the condition
(meant termwise as usual), and since the terms of the complex are –complete and –completely flat, is a non–zero divisor on the terms of by Corollary 2.4.
So pick any element . The image of in lies in and for any chosen we have for some . Now lies in so satisfies
Since is a non–zero divisor on , it follows that , that is, lies in We thus infer that as desired. ∎
5. Ramification bounds for mod étale cohomology
5.1. Ramification bounds
We are ready to discuss the implications to the question of ramification bounds for mod étale cohomology groups when is smooth and proper –adic formal scheme over .
We define an additive valuation on by where is the valuation on normalized so that , and is the multiplicative lift of . This way, we have and . For a real number , denote by ( resp.) the ideal of formed by all elements with (, resp.).
Similarly, we fix an additive valuation of normalized by . Then for an algebraic extension and a real number , we denote by the ideal consisting of all elements with (and similarly, for ’’ as well).
For a finite extensions and a real number , let us recall (a version of131313Fontaine’s original condition uses the ideals instead. Up to changing some inequalities from ‘’ to ‘’ and vice versa, the conditions are fairly equivalent.) Fontaine’s property :
We also recall the upper ramification numbering in the convention used in [Fontaine, CarusoLiu]. For and a non–negative real number set
where is again the additive valuation of normalized by .
For set
(which makes sense as for all ). Then is a piecewise–linear increasing continuous concave function. Denote by its inverse, and set
Denote by the infimum of all such that and by the infimum of all such that Clearly one has
Remark 5.1.
Let us compare the indexing conventions with [SerreLocalFields] and [Fontaine], as the results therein are (implicitly or explicitly) used. If are the upper–index ramification groups in [SerreLocalFields] and [Fontaine], resp., and similarly we denote and the lower–index ramification groups, then we have
where is the ramification index of .
In particular, since the enumeration differs from the one in [SerreLocalFields] only by a shift by one, the upper indexing is still compatible with passing to quotients, and it make sense to set
where varies over finite Galois extensions contained in a fixed algebraic closure of (and is the absolute Galois group).
Regarding , the following transitivity formula is useful.
Lemma 5.2 ([CarusoLiu, Lemma 4.3.1]).
Let be a pair of finite extensions with both and Galois. Then we have
The property is connected with the ramification of the field extension as follows.
Proposition 5.3.
Let be finite extensions of fields with Galois and let be a real number. If the property holds, then:
-
(1)
([Yoshida, Proposition 3.3]) In fact, is the infimum of all such that is valid.
-
(2)
([CarusoLiu, Corollary 4.2.2]) where denotes the different of the field extension .
Corollary 5.4.
Both the assumptions and the conclusions of Proposition 5.3 are insensitive to replacing by any unramified extension of contained in .
Proof.
Let be an unramified extension such that . The fact that is equivalent to is proved in [Yoshida, Proposition 2.2]. To show that also the conclusions are the same for and , it is enough to observe that and . The first two equalities are clear since is unramified. The third equality follows from upon noting that is the unit ideal. Finally, by Lemma 5.2, we have As is unramified, we have and for all . The fourth equality thus follows as well. ∎
Let be a proper and smooth –adic formal scheme over . Fix the integer , and denote by the Galois module . Let be the splitting field of , i.e. where is the associated representation. The goal is to provide an upper bound on , and a constant such that acts trivially on for all .
To achieve this, we follow rather closely the strategy of [CarusoLiu]. The main difference is that the input of –modules attached to the discussed –respresentations in [CarusoLiu] is in our situation replaced by a –torsion Breuil–Kisin module and a Breuil–Kisin–Fargues –module that arise as the –torsion prismatic – and –cohomology, resp. Let us therefore lay out the strategy, referring to proofs in [CarusoLiu] whenever possible, and describe the needed modifications where necessary. To facilitate this approach further, the notation used will usually reflect the notation of [CarusoLiu], except for mostly omitting the index throughout (which in our situation is always equal to ).
The relation of the above–mentioned –torsion prismatic cohomologies to the –torsion étale cohomology is as follows.
Proposition 5.5 ([LiLiu, Proposition 7.2, Corollary 7.4, Remark 7.5]).
Let be a smooth and proper –adic formal scheme over . Then
-
(1)
is a –torsion Breuil–Kisin module of height , and we have
as –modules.
-
(2)
is a –torsion Breuil–Kisin–Fargues –module of height , and we have
as –modules.
-
(3)
We have and the natural map has the image contained in .
So let and , so that Observe further that, since is a unit of we have and where and are again a Breuil–Kisin module and a Breuil–Kisin–Fargues –module, resp., of height . Since is faithfully flat, it is easy to see that the isomorphism remains true. Furthermore, the pair satisfies the conditions
(5.6) |
for all , since the pair satisfies the analogous conditions by Proposition 4.8. Finally, the module is finitely generated and –torsion–free –module, hence a finite free –module (and, consequently, is a finite free –module).
Instead of using directly, we work with the dual module
instead; this is equivalent, as the splitting field of is still . Note that
as a –module.
Remark 5.6 (Ramification bounds of [Caruso]).
Similarly to the discussion above we may take, for any and . Then the –module
is the restriction of to . Denoting by the –adic completion of , then becomes an étale –module over in the sense of [Fontaine3, §A], with the natural map injective; thus, in terminology of [Caruso], serves as a –lattice of height dividing . Upon observing that is the –respresentation associated with (see e.g. [Caruso, §2.1.3]), Theorem 2 of [Caruso] implies the ramification bound
Here are constants that depend on the field and that generally grow with increasing . (Their precise meaning is described in § LABEL:subsec:Comparisons.)
We employ the following approximations of the functors and .
Notation 5.7.
For a real number , we define
We further set and . Given with we denote by (