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

A motivic construction of the de Rham-Witt complex

Junnosuke Koizumi Graduate School of Mathematical Sciences, University of Tokyo, 3-8-1 Komaba, Meguro-ku, Tokyo 153-8914, Japan jkoizumi@ms.u-tokyo.ac.jp  and  Hiroyasu Miyazaki NTT Institute for Fundamental Mathematics, NTT Communication Science Laboratories, Nippon Telegraph and Telephone Corporation, 3-9-11 Midori-cho, Musashino-shi, Tokyo 180-8585, Japan hiroyasu.miyazaki@ntt.com
Abstract.

The theory of reciprocity sheaves due to Kahn-Saito-Yamazaki is a powerful framework to study invariants of smooth varieties via invariants of pairs (X,D)(X,D) of a variety XX and a divisor DD. We develop a generalization of this theory where DD can be a \mathbb{Q}-divisor. As an application, we provide a motivic construction of the de Rham-Witt complex, which is analogous to the motivic construction of the Milnor KK-theory due to Suslin-Voevodsky.

Key words and phrases:
reciprocity sheaves, cube invariance, modulus pairs, de Rham-Witt complex
1991 Mathematics Subject Classification:
14F42(primary), 13F35, 14F30, 19E15(secondary).
The first author is supported by JSPS KAKENHI Grant (22J20698). The second author is supported by JSPS KAKENHI Grant (19K23413, 21K13783).

Introduction

In Voevodsky’s theory of mixed motives, the notion of 𝔸1\mathbb{A}^{1}-invariant sheaf played a fundamental role (see [Voe00], [Voe98], [MVW06] etc.). An 𝔸1\mathbb{A}^{1}-invariant sheaf is a sheaf with transfers satisfying 𝔸1\mathbb{A}^{1}-invariance. A sheaf with transfers is a Nisnevich sheaf of abelian groups on the category of finite correspondences, denoted Cork\operatorname{\mathrm{Cor}}_{k}, where kk is the fixed base field. The objects of Cork\operatorname{\mathrm{Cor}}_{k} are smooth schemes over kk, and a morphism XYX\to Y in Cork\operatorname{\mathrm{Cor}}_{k} is given by an algebraic cycle on X×YX\times Y whose components are finite surjective over a connected component of XX. We say that a sheaf with transfers FF is 𝔸1\mathbb{A}^{1}-invariant if pr1:F(X)F(X×𝔸1){\operatorname{pr}}_{1}^{*}\colon F(X)\to F(X\times\mathbb{A}^{1}) is an isomorphism for any XCorkX\in\operatorname{\mathrm{Cor}}_{k}.

On the other hand, in a series of papers [KSY16], [KSY22], Kahn-Saito-Yamazaki developed the theory of reciprocity sheaves. This is a vast generalization of the theory of 𝔸1\mathbb{A}^{1}-invariant sheaves over a field, and it captures ramification-theoretic information of invariants of schemes (see e.g. [RS21]). The class of reciprocity sheaves includes many interesting examples that are not 𝔸1\mathbb{A}^{1}-invariant, such as the sheaf of differential forms, the Hodge-Witt sheaf, and all commutative algebraic groups.

Let us recall what reciprocity sheaves are. The key idea is to replace smooth schemes by proper modulus pairs. A proper modulus pair over kk is a pair (X,DX)(X,D_{X}) of a proper kk-scheme XX and an effective Cartier divisor DXD_{X} on XX such that X|DX|X\setminus|D_{X}| is smooth over kk. For example, the pair ¯:=(1,[]){\overline{\square}}:=(\mathbb{P}^{1},[\infty]) is a proper modulus pair, which we call the cube. We can define a category of proper modulus pairs MCork\operatorname{\mathrm{MCor}}_{k} similar to Cork\operatorname{\mathrm{Cor}}_{k} by taking into account the information of Cartier divisors. An additive presheaf F:(MCork)opAbF\colon(\operatorname{\mathrm{MCor}}_{k})^{\operatorname{op}}\to\operatorname{\mathrm{Ab}} is said to be cube invariant if for any modulus pair 𝒳=(X,DX)\mathcal{X}=(X,D_{X}), the map

pr1:F(𝒳)F(𝒳¯),𝒳¯:=(X×1,pr1DX+pr2[]){\operatorname{pr}}_{1}^{*}\colon F(\mathcal{X})\to F(\mathcal{X}\otimes{\overline{\square}}),\quad\mathcal{X}\otimes{\overline{\square}}:=(X\times\mathbb{P}^{1},{\operatorname{pr}}_{1}^{*}D_{X}+{\operatorname{pr}}_{2}^{*}[\infty])

is an isomorphism. A presheaf with transfers (resp. sheaf with transfers) FF is said to be a reciprocity presheaf (resp. reciprocity sheaf) if it belongs to the essential image of

(cube invariantpresheaves)PSh(MCork)ω!PSh(Cork),\left(\begin{tabular}[]{c}\text{cube invariant}\\ \text{presheaves}\end{tabular}\right)\hookrightarrow\operatorname{PSh}(\operatorname{\mathrm{MCor}}_{k})\xrightarrow{\omega_{!}}\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}),

where ω!\omega_{!} is the left Kan extension of (X,DX)X|DX|(X,D_{X})\mapsto X\setminus|D_{X}| 111 Originally, the notion of reciprocity sheaf given in [KSY16] looks quite different from the above one. However, in [KSY22], it is shown that the above definition coincides with the original one, provided that the base field kk is perfect. An advantage of the above description is that we can think of a reciprocity sheaf as the “shadow” of a cube invariant presheaf. Since the definition of cube invariant presheaves is very similar to that of 𝔸1\mathbb{A}^{1}-invariant presheaves, one can prove many important properties of cube invariant presheaves by following Voevodsky’s classical methods, at least partly. .

In the first half of this paper, we will generalize the theory of modulus pairs and sheaves on them to allow the effective divisor DXD_{X} to have rational coefficients (see applications below for the reason why we need this generalization). We will define the category of proper \mathbb{Q}-modulus pairs MCork\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k} over kk (Definition 2.1). We can define the notion of cube invariant presheaf and the functor ω!:PSh(MCork)PSh(Cork)\omega_{!}\colon\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k})\to\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) as before. In fact, the resulting notion of \mathbb{Q}-reciprocity presheaf coincides with the usual one. This allows us to use \mathbb{Q}-modulus pairs in the theory of reciprocity sheaves.

In the latter half of this paper, we will give some applications of our theory. The main results include a motivic construction of the de Rham-Witt complex, which we will sketch below.

It is well-known that the multiplicative group 𝔾m\mathbb{G}_{m}, which is an important example of an 𝔸1\mathbb{A}^{1}-invariant sheaf, has a motivic presentation. That is, there exists a canonical isomorphism 𝔾mh0𝔸1(tr(𝔸1{0})/)\mathbb{G}_{m}\simeq h_{0}^{\mathbb{A}^{1}}(\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})/\mathbb{Z}) in PSh(Corkaff)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}^{\mathrm{aff}}), where Corkaff\operatorname{\mathrm{Cor}}_{k}^{\mathrm{aff}} is the full subcategory of Cork\operatorname{\mathrm{Cor}}_{k} consisting of affine schemes, tr:CorkPSh(Cork)\mathbb{Z}_{\operatorname{tr}}\colon\operatorname{\mathrm{Cor}}_{k}\to\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) is the Yoneda embedding, and h0𝔸1h_{0}^{\mathbb{A}^{1}} is the 0-th Suslin homology:

h0𝔸1F(X)=Coker(F(X×𝔸1)i0i1F(X)).h_{0}^{\mathbb{A}^{1}}F(X)=\operatorname{Coker}(F(X\times\mathbb{A}^{1})\xrightarrow{i_{0}^{*}-i_{1}^{*}}F(X)).

Moreover, the group structure on 𝔾m\mathbb{G}_{m} is simply induced by the multiplication morphism (𝔸1{0})×(𝔸1{0})𝔸1{0}(\mathbb{A}^{1}\setminus\{0\})\times(\mathbb{A}^{1}\setminus\{0\})\to\mathbb{A}^{1}\setminus\{0\}. Suslin-Voevodsky proved more generally that the sheaf of unramified Milnor KK-theory admits a motivic construction [SV00, Theorem 3.4]. These results are fundamental for various computations in the theory of mixed motives.

It is a natural idea to extend this to other invariants of schemes. We start with a motivic presentation of the ring of big Witt vectors 𝕎n(A)=1+tA[t]/(tn+1)\mathbb{W}_{n}(A)=1+tA[t]/(t^{n+1}) of a ring AA. For n0n\geq 0, we define 𝕎n+PSh(MCork)\mathbb{W}_{n}^{+}\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}) by

𝕎n+=limε>0tr(1,(n+ε)[])/,\mathbb{W}_{n}^{+}=\varinjlim_{\varepsilon>0}\mathbb{Z}_{\operatorname{tr}}(\mathbb{P}^{1},(n+\varepsilon)[\infty])/\mathbb{Z},

where tr:MCorkPSh(MCork)\mathbb{Z}_{\operatorname{tr}}\colon\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}\to\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}) is the Yoneda embedding. Then there are operations Fs,VsF_{s},V_{s} on 𝕎n+\mathbb{W}_{n}^{+} for each s>0s>0 which are induced by the morphism ttst\mapsto t^{s} on 𝔸1\mathbb{A}^{1} and its transpose. These operations are called the Frobenius and the Verschiebung. Moreover, the multiplication on 𝔸1\mathbb{A}^{1} induces a multiplication on 𝕎n+\mathbb{W}_{n}^{+}. The 0-th Suslin homology h0𝔸1h_{0}^{\mathbb{A}^{1}} generalizes to PSh(MCork)\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}):

h0\scaleobj0.8¯F(𝒳)=Coker(F(𝒳¯)i0i1F(𝒳)).h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F(\mathcal{X})=\operatorname{Coker}(F(\mathcal{X}\otimes{\overline{\square}})\xrightarrow{i_{0}^{*}-i_{1}^{*}}F(\mathcal{X})).

We write h0=ω!h0\scaleobj0.8¯h_{0}=\omega_{!}h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}. First we prove the following result.

Theorem 1 (Theorem 5.10, Proposition 5.11).

There is an isomorphism h0𝕎n+𝕎nh_{0}\mathbb{W}_{n}^{+}\xrightarrow{\sim}\mathbb{W}_{n} in PSh(Corkaff)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}^{\mathrm{aff}}) which preserves the multiplication, Frobenius, and the Verschiebung.

In the proof of this theorem, we use the fact that the group h0tr(1,r[])(r>0)h_{0}\mathbb{Z}_{\operatorname{tr}}(\mathbb{P}^{1},r[\infty])\;(r\in\mathbb{Q}_{>0}) depends only on r\lceil r\rceil. This is a consequence of the motivic Hasse-Arf theorem which we prove in Theorem 4.14. It is an analogue of the classical Hasse-Arf theorem which states that the upper ramification group G(r)G^{(r)} of an abelian extension of a local field depends only on r\lceil r\rceil.

Theorem 1 has a non-trivial application to reciprocity presheaves. Imitating the construction in [Miy19], we define NF(X)=Ker(i0:F(X×𝔸1)F(X))NF(X)=\operatorname{Ker}(i_{0}^{*}\colon F(X\times\mathbb{A}^{1})\to F(X)) for a reciprocity presheaf FF. Then we have NF=0NF=0 if and only if FF is 𝔸1\mathbb{A}^{1}-invariant. Using the above theorem, we can prove that 𝕎=limn0𝕎n\mathbb{W}=\varprojlim_{n\geq 0}\mathbb{W}_{n} acts on NFNF. As a consequence, we obtain a short proof of the following result of Binda-Cao-Kai-Sugiyama [Bin+17, Theorem 1.3].

Theorem 2 (Corollary 5.15).

Let FF be a reciprocity presheaf over kk and assume that FF is separated for the Zariski topology.

  1. (1)

    If ch(k)=0\operatorname{ch}(k)=0 and F=0F\otimes\mathbb{Q}=0, then FF is 𝔸1\mathbb{A}^{1}-invariant.

  2. (2)

    If ch(k)=p>0\operatorname{ch}(k)=p>0 and FF is pp-torsion-free, then FF is 𝔸1\mathbb{A}^{1}-invariant.

Next we give a motivic presentation of the ring of pp-typical Witt vectors Wn(A)W_{n}(A) of a (p)\mathbb{Z}_{(p)}-algebra AA. For n0n\geq 0, we define 𝕎^n+PSh(MCork)\widehat{\mathbb{W}}_{n}^{+}\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}) by

𝕎^n+=limε>0tr(1,ε[0]+(n+ε)[]).\widehat{\mathbb{W}}_{n}^{+}=\varinjlim_{\varepsilon>0}\mathbb{Z}_{\operatorname{tr}}(\mathbb{P}^{1},\varepsilon[0]+(n+\varepsilon)[\infty]).

As in the case of 𝕎n+\mathbb{W}_{n}^{+}, we have operations Fs,VsF_{s},V_{s} for s>0s>0 and a multiplication on 𝕎^n+\widehat{\mathbb{W}}_{n}^{+}. We define Wn+W_{n}^{+} to be the quotient of 𝕎^pn1+(p)\widehat{\mathbb{W}}_{p^{n-1}}^{+}\otimes\mathbb{Z}_{(p)} by the images of the idempotents 1VF\ell^{-1}V_{\ell}F_{\ell} for all prime numbers p\ell\neq p. The operations Fp,VpF_{p},V_{p} descend to operations F,VF,V on Wn+W_{n}^{+} called the Frobenius and the Verschiebung, and the multiplication also descends to Wn+W_{n}^{+}. We prove the following

Theorem 3 (Corollary 5.12).

There is an isomorphism h0Wn+Wnh_{0}W_{n}^{+}\xrightarrow{\sim}W_{n} in PSh(Corkaff)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}^{\mathrm{aff}}) which is compatible with the Frobenius, the Verschiebung, and the multiplication.

In order to treat the de Rham-Witt complex, we have to assume that kk is perfect and p3p\geq 3. We define 𝔾m+PSh(MCork)\mathbb{G}_{m}^{+}\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}) by

𝔾m+=limε>0tr(1,ε[0]+ε[])/.\mathbb{G}_{m}^{+}=\varinjlim_{\varepsilon>0}\mathbb{Z}_{\operatorname{tr}}(\mathbb{P}^{1},\varepsilon[0]+\varepsilon[\infty])/\mathbb{Z}.

Using the tensor structure on PSh(MCork)\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}) (see §2), we obtain an object Wn+𝔾m+qW_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q} of PSh(MCork)\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}). Our main result is the following:

Theorem 4 (Theorem 6.21).

Let kk be a perfect field of characteristic p3p\geq 3. Then there is an isomorphism

aNish0(Wn+𝔾m+q)WnΩqa_{{\operatorname{Nis}}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})\xrightarrow{\sim}W_{n}\Omega^{q}

in ShNis(Cork)\operatorname{Sh}_{\operatorname{Nis}}(\operatorname{\mathrm{Cor}}_{k}), where aNisa_{\operatorname{Nis}} denotes the Nisnevich sheafification.

The proof goes as follows: first we prove that the left hand side admits a Witt complex structure (Theorem 6.11). Since the de Rham-Witt complex is initial in the category of Witt complexes, we obtain a unique morphism from the right hand side to the left hand side. To prove that this morphism is an isomorphism, we construct an inverse by using the transfer structure on WnΩqW_{n}\Omega^{q}.

We note that a similar motivic presentation for Ωq\Omega^{q} is obtained by Rülling-Sugiyama-Yamazaki [RSY22] and the first author [Koi]. Also, a presentation of the big de Rham-Witt complex 𝕎nΩq\mathbb{W}_{n}\Omega^{q} using additive higher Chow groups is obtained by Rülling [R0̈7a] for fields, and by Krishna-Park [KP21] for regular semilocal algebras over a field. We expect that these results will be connected by some comparison isomorphism between Suslin homology and additive higher Chow groups.

The structure of the present paper is as follows. In §1, we prepare preliminary results on \mathbb{Q}-Cartier divisors. In §2, we define \mathbb{Q}-modulus pairs and prove some basic properties. In §3, we define and study the notion of cube invariant presheaf and the notion of reciprocity presheaf following [KSY22]. In §4, we formulate and prove the motivic Hasse-Arf theorem, by comparing the 0-th Suslin homology group and the Chow group of relative 0-cycles of a modulus curve. This comparison can be seen as a generalization of the result in [RY16]. In §5, we apply our machinery to give a motivic construction of the ring of Witt vectors and basic operations on it. As an application, we give a short proof of the result of Binda-Cao-Kai-Sugiyama [Bin+17] on torsion and divisibility of a reciprocity sheaf. In §6, we give a motivic construction of the de Rham-Witt complex.

Acknowledgements

We would like to thank Shuji Saito for his interest on this work. We also appreciate his valuable comments on earlier versions of this paper, which encouraged the authors to improve the main results.

Notations and conventions

  • -

    For a scheme XX and xXx\in X, we write k(x)k(x) for the residue field of XX at xx. We write X(d)X^{(d)} for the set of points on XX of codimension dd. An element of X(0)X^{(0)} is called a generic point.

  • -

    We say that a morphism of schemes f:XYf\colon X\to Y is pseudo-dominant if it takes generic points to generic points, i.e., f(X(0))Y(0)f(X^{(0)})\subset Y^{(0)}.

  • -

    For an integral scheme XX, we write k(X)k(X) for its function field. If XX is a noetherian normal integral scheme and fk(X)×f\in k(X)^{\times}, we write div(f)\operatorname{div}(f) for the Weil divisor on XX defined by ff.

  • -

    We write Smk\operatorname{\mathrm{Sm}}_{k} for the category of smooth separated kk-schemes of finite type. We define Smkaff\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}} to be the full subcategory of Smk\operatorname{\mathrm{Sm}}_{k} spanned by affine schemes.

  • -

    For an additive category 𝒞\mathcal{C}, we write PSh(𝒞)\operatorname{PSh}(\mathcal{C}) for the category of additive functors 𝒞opAb\mathcal{C}^{\operatorname{op}}\to\operatorname{\mathrm{Ab}}.

  • -

    For a category 𝒞\mathcal{C}, we write Pro(𝒞)\operatorname{Pro}(\mathcal{C}) for the category of pro-objects in 𝒞\mathcal{C}. A pro-functor F:𝒞𝒟F\colon\mathcal{C}\dashrightarrow\mathcal{D} is a functor F:𝒞Pro(𝒟)F\colon\mathcal{C}\to\operatorname{Pro}(\mathcal{D}). A pro-functor F:𝒞𝒟F\colon\mathcal{C}\dashrightarrow\mathcal{D} is said to be pro-left adjoint to G:𝒟𝒞G\colon\mathcal{D}\to\mathcal{C} if there is a natural isomorphism HomPro(𝒟)(F(c),d)Hom𝒞(c,G(d))\operatorname{Hom}_{\operatorname{Pro}(\mathcal{D})}(F(c),d)\simeq\operatorname{Hom}_{\mathcal{C}}(c,G(d)).

1. Preliminaries

1.1. Finite correspondences

First we recall Suslin-Voevodsky’s category of finite correspondences Cork\operatorname{\mathrm{Cor}}_{k}. It is an additive category having the same objects as Smk\operatorname{\mathrm{Sm}}_{k}, and its group of morphisms Cork(X,Y)\operatorname{\mathrm{Cor}}_{k}(X,Y) is the group of algebraic cycles on X×YX\times Y whose components are finite pseudo-dominant over XX. The fiber product over kk gives a symmetric monoidal structure on Cork\operatorname{\mathrm{Cor}}_{k}. For any morphism f:XYf\colon X\to Y in Smk\operatorname{\mathrm{Sm}}_{k}, the graph of ff gives a morphism f:XYf\colon X\to Y in Cork\operatorname{\mathrm{Cor}}_{k}. If ff is finite pseudo-dominant, then the transpose of the graph of ff gives a morphism ft:YX{}^{t}f\colon Y\to X in Cork\operatorname{\mathrm{Cor}}_{k}. For SSmkS\in\operatorname{\mathrm{Sm}}_{k} and smooth SS-schemes X,YX,Y, we write CorS(X,Y)Cork(X,Y)\operatorname{\mathrm{Cor}}_{S}(X,Y)\subset\operatorname{\mathrm{Cor}}_{k}(X,Y) for the subgroup of cycles supported on X×SYX\times_{S}Y.

A presheaf with transfers is an additive presheaf F:(Cork)opAbF\colon(\operatorname{\mathrm{Cor}}_{k})^{\operatorname{op}}\to\operatorname{\mathrm{Ab}}. If f:XYf\colon X\to Y is a finite pseudo-dominant morphism in Smk\operatorname{\mathrm{Sm}}_{k} and FPSh(Cork)F\in\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}), then we write ff_{*} or TrX/Y\operatorname{Tr}_{X/Y} for the map (ft):F(X)F(Y)({}^{t}f)^{*}\colon F(X)\to F(Y). We say that FF is a Nisnevich sheaf if for any XSmkX\in\operatorname{\mathrm{Sm}}_{k}, the presheaf FX:UF(U)F_{X}\colon U\mapsto F(U) on XNisX_{\operatorname{Nis}} is a Nisnevich sheaf. We write ShNis(Cork)PSh(Cork)\operatorname{Sh}_{\operatorname{Nis}}(\operatorname{\mathrm{Cor}}_{k})\subset\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) for the full subcategory spanned by Nisnevich sheaves. The inclusion functor ShNis(Cork)PSh(Cork)\operatorname{Sh}_{\operatorname{Nis}}(\operatorname{\mathrm{Cor}}_{k})\to\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) admits an exact left adjoint aNis:PSh(Cork)ShNis(Cork)a_{\operatorname{Nis}}\colon\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k})\to\operatorname{Sh}_{\operatorname{Nis}}(\operatorname{\mathrm{Cor}}_{k}), so ShNis(Cork)\operatorname{Sh}_{\operatorname{Nis}}(\operatorname{\mathrm{Cor}}_{k}) is Grothendieck abelian.

1.2. \mathbb{Q}-Cartier divisors

We introduce the notion of \mathbb{Q}-Cartier divisor over a general noetherian scheme, which will play a fundamental role in this paper.

Definition 1.1.

For a noetherian scheme XX, we write CDiv(X)\operatorname{CDiv}(X) for the group of Cartier divisors on XX. We define the group of \mathbb{Q}-Cartier divisors on XX by CDiv(X):=CDiv(X)\operatorname{CDiv}_{\mathbb{Q}}(X):=\operatorname{CDiv}(X)\otimes_{\mathbb{Z}}\mathbb{Q}.

If XX is a noetherian normal scheme, then the group CDiv(X)\operatorname{CDiv}(X) is embedded into the free abelian group of Weil divisors on XX. In particular, CDiv(X)\operatorname{CDiv}(X) is a free abelian group in this case. Let f:YXf\colon Y\to X be a morphism of noetherian schemes and DD be a \mathbb{Q}-Cartier divisor on XX. We say that the pullback fDf^{*}D of DD by ff exists if there is an ordinary Cartier divisor EE on XX and rr\in\mathbb{Q} with D=rED=rE such that the pullback fEf^{*}E exists. In this situation, we define fD:=rfECDiv(Y)f^{*}D:=rf^{*}E\in\operatorname{CDiv}_{\mathbb{Q}}(Y). This does not depend on the choice of EE and rr.

Let XX be a noetherian scheme and DD be a \mathbb{Q}-Cartier divisor on XX. We say that DD is \mathbb{Q}-effective if there is an ordinary effective Cartier divisor EE on XX and r>0r\in\mathbb{Q}_{>0} such that D=rED=rE in CDiv(X)\operatorname{CDiv}_{\mathbb{Q}}(X). For \mathbb{Q}-Cartier divisors D,DD,D^{\prime}, we write DDD\leq D^{\prime} if DDD^{\prime}-D is \mathbb{Q}-effective.

If an ordinary Cartier divisor DD is effective, then it is \mathbb{Q}-effective. The converse is not true in general, but it is true for normal schemes:

Lemma 1.2.

Let XX be a noetherian normal scheme and DD be an ordinary Cartier divisor on XX. Then DD is effective if and only if DD is \mathbb{Q}-effective.

Proof.

We may assume that X=SpecAX=\operatorname{Spec}A with AA a normal domain and D=div(a)D=\operatorname{div}(a) with aFrac(A)×a\in\operatorname{Frac}(A)^{\times}. If DD is \mathbb{Q}-effective, then there is some non-zero-divisor bAb\in A and n1n\geq 1 such that div(a)=(1/n)div(b)\operatorname{div}(a)=(1/n)\operatorname{div}(b) in CDiv(X)\operatorname{CDiv}_{\mathbb{Q}}(X). Since XX is normal, the map CDiv(X)CDiv(X)\operatorname{CDiv}(X)\to\operatorname{CDiv}_{\mathbb{Q}}(X) is injective and hence div(an)=div(b)\operatorname{div}(a^{n})=\operatorname{div}(b) in CDiv(X)\operatorname{CDiv}(X). This implies that anub=0a^{n}-ub=0 for some uA×u\in A^{\times}, so aAa\in A since AA is normal. Therefore DD is effective. ∎

Corollary 1.3.

Let XX be a noetherian scheme and D,DD,D^{\prime} be ordinary Cartier divisors on XX. Then we have D|XND|XND|_{X^{N}}\leq D^{\prime}|_{X^{N}} as \mathbb{Q}-Cartier divisors if and only if D|XND|XND|_{X^{N}}\leq D^{\prime}|_{X^{N}} as ordinary Cartier divisors.

Definition 1.4.

Let XX be a noetherian scheme and DD be a \mathbb{Q}-effective \mathbb{Q}-Cartier divisor on XX. Suppose that D=rED=rE in CDiv(X)\operatorname{CDiv}_{\mathbb{Q}}(X) where EE is an ordinary effective Cartier divisor on XX and r>0r\in\mathbb{Q}_{>0}. Then the support |D||D| of DD is defined to be |E||E|. This is well-defined since |E|=|nE||E|=|nE| for any ordinary effective Cartier divisor EE and n1n\geq 1.

2. \mathbb{Q}-modulus pairs

2.1. \mathbb{Q}-Modulus pairs

Recall from [Kah+21] that a modulus pair (over kk) is a pair 𝒳=(X,DX)\mathcal{X}=(X,D_{X}) where XX is a separated kk-scheme of finite type and DXD_{X} is an effective Cartier divisor on XX such that X:=X|DX|X^{\circ}:=X\setminus|D_{X}| is smooth over kk.

Definition 2.1.

A \mathbb{Q}-modulus pair 𝒳\mathcal{X} is a pair (X,DX)(X,D_{X}) where XX is a separated kk-scheme of finite type and DXD_{X} is a \mathbb{Q}-effective \mathbb{Q}-Cartier divisor on XX, such that X:=X|DX|X^{\circ}:=X\setminus|D_{X}| is smooth over kk.

In order to avoid confusion, we use the word “\mathbb{Z}-modulus pairs” to indicate modulus pairs in the sense of [Kah+21]. In what follows, we fix Λ{,}\Lambda\in\{\mathbb{Z},\mathbb{Q}\} and develop the theory of Λ\Lambda-modulus pairs.

An ambient morphism of Λ\Lambda-modulus pairs f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} is a morphism of kk-schemes f:XYf\colon X\to Y such that f(X)Yf(X^{\circ})\subset Y^{\circ} and DX|XNfDY|XND_{X}|_{X^{N}}\geq f^{*}D_{Y}|_{X^{N}} hold. It is called minimal if DX=fDYD_{X}=f^{*}D_{Y} holds.

Let 𝒳,𝒴\mathcal{X},\mathcal{Y} be Λ\Lambda-modulus pairs over kk and let VX×YV\subset X^{\circ}\times Y^{\circ} be an integral closed subscheme. We say that VV is left proper if the closure V¯\overline{V} of VV in X×YX\times Y is proper over XX. We say that VV is admissible if (pr1DX)|V¯N(pr2DY)|V¯N({\operatorname{pr}}_{1}^{*}D_{X})|_{\overline{V}^{N}}\geq({\operatorname{pr}}_{2}^{*}D_{Y})|_{\overline{V}^{N}} holds. We write M¯CorkΛ(𝒳,𝒴)\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(\mathcal{X},\mathcal{Y}) for the subgroup of Cork(X,Y)\operatorname{\mathrm{Cor}}_{k}(X^{\circ},Y^{\circ}) consisting of cycles whose components are left proper and admissible. For any ambient morphism f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} over kk, the graph of ff^{\circ} gives a morphism f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} in M¯CorkΛ\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}. If ff is a proper minimal ambient morphism such that ff^{\circ} is finite pseudo-dominant, then the transpose of the graph of ff^{\circ} gives a morphism ft:𝒴𝒳{}^{t}f\colon\mathcal{Y}\to\mathcal{X} in M¯CorkΛ\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}. Note that for \mathbb{Z}-modulus pairs 𝒳,𝒴\mathcal{X},\mathcal{Y}, we have M¯Cork(𝒳,𝒴)=M¯Cork(𝒳,𝒴)\operatorname{\mathrm{\underline{M}Cor}}^{\mathbb{Z}}_{k}(\mathcal{X},\mathcal{Y})=\operatorname{\mathrm{\underline{M}Cor}}^{\mathbb{Q}}_{k}(\mathcal{X},\mathcal{Y}) by Corollary 1.3.

Lemma 2.2.

Let 𝒳,𝒴,𝒵\mathcal{X},\mathcal{Y},\mathcal{Z} be Λ\Lambda-modulus pairs over kk and αM¯CorkΛ(𝒳,𝒴)\alpha\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(\mathcal{X},\mathcal{Y}), βM¯CorkΛ(𝒴,𝒵)\beta\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(\mathcal{Y},\mathcal{Z}). Then we have βαM¯CorkΛ(𝒳,𝒵)\beta\circ\alpha\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(\mathcal{X},\mathcal{Z}), where \circ denotes the composition in Cork\operatorname{\mathrm{Cor}}_{k}.

Proof.

For Λ=\Lambda=\mathbb{Z}, this is proved in [Kah+21, Proposition 1.2.4 and Proposition 1.2.7]. For Λ=\Lambda=\mathbb{Q}, we may replace DX,DY,DZD_{X},D_{Y},D_{Z} by nDX,nDY,nDZnD_{X},nD_{Y},nD_{Z} for n>0n\in\mathbb{Z}_{>0}, so we may assume that 𝒳,𝒴,𝒵\mathcal{X},\mathcal{Y},\mathcal{Z} are \mathbb{Z}-modulus pairs. Then we have M¯Cork(𝒳,𝒴)=M¯Cork(𝒳,𝒴)\operatorname{\mathrm{\underline{M}Cor}}^{\mathbb{Z}}_{k}(\mathcal{X},\mathcal{Y})=\operatorname{\mathrm{\underline{M}Cor}}^{\mathbb{Q}}_{k}(\mathcal{X},\mathcal{Y}) etc., so the claim follows from the case Λ=\Lambda=\mathbb{Z}. ∎

We define M¯CorkΛ\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} to be the category of Λ\Lambda-modulus pairs over kk, where the morphisms are given by M¯CorkΛ(𝒳,𝒴)\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(\mathcal{X},\mathcal{Y}). The canonical functor M¯CorkM¯Cork\operatorname{\mathrm{\underline{M}Cor}}^{\mathbb{Z}}_{k}\to\operatorname{\mathrm{\underline{M}Cor}}^{\mathbb{Q}}_{k} is fully faithful by Corollary 1.3. We set

𝒳𝒴=(X×Y,pr1DX+pr2DY).\mathcal{X}\otimes\mathcal{Y}=(X\times Y,{\operatorname{pr}}_{1}^{*}D_{X}+{\operatorname{pr}}_{2}^{*}D_{Y}).

The next lemma shows that this gives a symmetric monoidal structure on M¯CorkΛ\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}:

Lemma 2.3.

Let 𝒳1,𝒴1,𝒳2,𝒴2M¯CorkΛ\mathcal{X}_{1},\mathcal{Y}_{1},\mathcal{X}_{2},\mathcal{Y}_{2}\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} and αM¯CorkΛ(𝒳1,𝒴1)\alpha\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(\mathcal{X}_{1},\mathcal{Y}_{1}), βM¯CorkΛ(𝒳2,𝒴2)\beta\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(\mathcal{X}_{2},\mathcal{Y}_{2}). Then we have α×βM¯CorkΛ(𝒳1𝒳2,𝒴1𝒴2)\alpha\times\beta\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(\mathcal{X}_{1}\otimes\mathcal{X}_{2},\mathcal{Y}_{1}\otimes\mathcal{Y}_{2}).

Proof.

For Λ=\Lambda=\mathbb{Z}, this is proved in [Kah+22, Lemma 2.1.3]. For Λ=\Lambda=\mathbb{Q}, we may reduce to the case Λ=\Lambda=\mathbb{Z} as in the proof of Lemma 2.2. ∎

We define MCorkΛM¯CorkΛ\operatorname{\mathrm{MCor}}^{\Lambda}_{k}\subset\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} to be the full subcategory consisting of proper Λ\Lambda-modulus pairs, i.e., 𝒳=(X,DX)\mathcal{X}=(X,D_{X}) with XX proper over kk. There are natural functors

τ:MCorkΛM¯CorkΛ;𝒳𝒳,\displaystyle\tau\colon\operatorname{\mathrm{MCor}}^{\Lambda}_{k}\to\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k};\quad\mathcal{X}\mapsto\mathcal{X},
ω¯:M¯CorkΛCork;𝒳X.\displaystyle\underline{\omega}\colon\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}\to\operatorname{\mathrm{Cor}}_{k};\quad\mathcal{X}\to X^{\circ}.

A Λ\Lambda-modulus presheaf is an additive presheaf F:(M¯CorkΛ)opAbF\colon(\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k})^{\operatorname{op}}\to\operatorname{\mathrm{Ab}}. The category PSh(M¯CorkΛ)\operatorname{PSh}(\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}) of Λ\Lambda-modulus presheaves admits a symmetric monoidal structure \otimes which extends \otimes on M¯CorkΛ\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} by colimits. The functor τ:MCorkΛM¯CorkΛ\tau\colon\operatorname{\mathrm{MCor}}^{\Lambda}_{k}\to\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} induces an adjunction

τ!:PSh(MCorkΛ)PSh(M¯CorkΛ):τ\tau_{!}\colon\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k})\rightleftarrows\operatorname{PSh}(\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k})\colon\tau^{*}

where τ\tau^{*} is the restriction functor and τ!\tau_{!} is the left Kan extension of τ\tau. Similarly, the functor ω:M¯CorkΛCork\omega\colon\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}\to\operatorname{\mathrm{Cor}}_{k} induces an adjunction

ω¯!:PSh(M¯CorkΛ)PSh(Cork):ω¯\underline{\omega}_{!}\colon\operatorname{PSh}(\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k})\rightleftarrows\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k})\colon\underline{\omega}^{*}

where ω¯F(𝒳)=F(X)\underline{\omega}^{*}F(\mathcal{X})=F(X^{\circ}) and ω¯!F(X)=F(X,)\underline{\omega}_{!}F(X)=F(X,\emptyset). We write ω!:=ω¯!τ!\omega_{!}:=\underline{\omega}_{!}\tau_{!} and ω=τω¯\omega^{*}=\tau^{*}\underline{\omega}^{*}:

ω!:PSh(MCorkΛ)PSh(Cork):ω.\omega_{!}\colon\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k})\rightleftarrows\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k})\colon\omega^{*}.

In order to write down τ!\tau_{!} and ω!\omega_{!} explicitly, we use the notion of compactification. A compactification of 𝒳M¯CorkΛ\mathcal{X}\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} is a triple (X¯,D¯,Σ)(\overline{X},\overline{D},\Sigma) where X¯\overline{X} is a proper kk-scheme and D¯,Σ\overline{D},\Sigma are Λ\Lambda-effective Λ\Lambda-Cartier divisors on X¯\overline{X}, equipped with an isomorphism XX¯|Σ|X\simeq\overline{X}\setminus|\Sigma| over kk such that DX=D¯|XD_{X}=\overline{D}|_{X}. We say that a compactification (X¯1,D¯1,Σ1)(\overline{X}_{1},\overline{D}_{1},\Sigma_{1}) dominates (X¯2,D¯2,Σ2)(\overline{X}_{2},\overline{D}_{2},\Sigma_{2}) if the diagonal ΔXX×X\Delta_{X^{\circ}}\subset X^{\circ}\times X^{\circ} gives a morphism (X¯1,D¯1+Σ1)(X¯2,D¯2+Σ2)(\overline{X}_{1},\overline{D}_{1}+\Sigma_{1})\to(\overline{X}_{2},\overline{D}_{2}+\Sigma_{2}) in MCorkΛ\operatorname{\mathrm{MCor}}^{\Lambda}_{k}. This defines a poset CompΛ(𝒳)\operatorname{\mathrm{Comp}}^{\Lambda}(\mathcal{X}) of compactifications of 𝒳\mathcal{X}.

Lemma 2.4.

Let 𝒳,𝒴M¯CorkΛ\mathcal{X},\mathcal{Y}\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}.

  1. (1)

    The poset CompΛ(𝒳)\operatorname{\mathrm{Comp}}^{\Lambda}(\mathcal{X}) is cofiltered.

  2. (2)

    If (X¯,D¯,Σ)CompΛ(𝒳)(\overline{X},\overline{D},\Sigma)\in\operatorname{\mathrm{Comp}}^{\Lambda}(\mathcal{X}), then {(X¯,D¯,nΣ)}n>0\{(\overline{X},\overline{D},n\Sigma)\}_{n>0} is cofinal in CompΛ(𝒳)\operatorname{\mathrm{Comp}}^{\Lambda}(\mathcal{X}).

  3. (3)

    For any αM¯CorkΛ(𝒳,𝒴)\alpha\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(\mathcal{X},\mathcal{Y}) and (Y¯,E¯,Ξ)CompΛ(𝒴)(\overline{Y},\overline{E},\Xi)\in\operatorname{\mathrm{Comp}}^{\Lambda}(\mathcal{Y}), there exists (X¯,D¯,Σ)CompΛ(𝒳)(\overline{X},\overline{D},\Sigma)\in\operatorname{\mathrm{Comp}}^{\Lambda}(\mathcal{X}) such that

    αMCorkΛ((X¯,D¯+Σ),(Y¯,E¯+Ξ)).\alpha\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}((\overline{X},\overline{D}+\Sigma),(\overline{Y},\overline{E}+\Xi)).
Proof.

For Λ=\Lambda=\mathbb{Z}, this is proved in [Kah+21, Section 1.8]. For Λ=\Lambda=\mathbb{Q}, we may reduce to the case Λ=\Lambda=\mathbb{Z} as in the proof of Lemma 2.2. ∎

By Lemma 2.4, we get a fully faithful symmetric monoidal functor

M¯CorkΛPro(MCorkΛ);𝒳``\varlim@"(X¯,D¯,Σ)CompΛ(𝒳)(X¯,D¯+Σ)\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}\to\operatorname{Pro}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k});\quad\mathcal{X}\mapsto\mathop{``\mathchoice{\varlim@\displaystyle{\leftarrowfill@\textstyle}}{\varlim@\textstyle{\leftarrowfill@\textstyle}}{\varlim@\scriptstyle{\leftarrowfill@\textstyle}}{\varlim@\scriptscriptstyle{\leftarrowfill@\textstyle}}"}\displaylimits_{(\overline{X},\overline{D},\Sigma)\in\operatorname{\mathrm{Comp}}^{\Lambda}(\mathcal{X})}(\overline{X},\overline{D}+\Sigma)

which is pro-left adjoint to the inclusion τ:MCorkΛM¯CorkΛ\tau\colon\operatorname{\mathrm{MCor}}^{\Lambda}_{k}\to\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}. Therefore the functor τ!\tau_{!} can be written explicitly as

τ!F(𝒳)=lim(X¯,D¯,Σ)CompΛ(𝒳)F(X¯,D¯+Σ).\tau_{!}F(\mathcal{X})=\varinjlim_{(\overline{X},\overline{D},\Sigma)\in\operatorname{\mathrm{Comp}}^{\Lambda}(\mathcal{X})}F(\overline{X},\overline{D}+\Sigma).

For XSmkX\in\operatorname{\mathrm{Sm}}_{k}, we can choose 𝒴MCorkΛ\mathcal{Y}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k} with YXY^{\circ}\simeq X. Then {(Y,,nDY)}n>0\{(Y,\emptyset,nD_{Y})\}_{n>0} is cofinal in CompΛ(X,)\operatorname{\mathrm{Comp}}^{\Lambda}(X,\emptyset), so we get

ω!F(X)=limn>0F(Y,nDY).\omega_{!}F(X)=\varinjlim_{n>0}F(Y,nD_{Y}).
Proposition 2.5.

The following assertions hold:

  1. (1)

    The functor ω\omega^{*} is exact and fully faithful.

  2. (2)

    The functor ω!\omega_{!} is exact and symmetric monoidal.

Proof.

Both functors are clearly exact. Moreover, the above formula for ω!\omega_{!} implies that ω!ωid\omega_{!}\omega^{*}\simeq{\operatorname{id}} and hence ω\omega^{*} is fully faithful. Since ω\omega is symmetric monoidal and ω!\omega_{!} is its extension by colimits, ω!\omega_{!} is also symmetric monoidal. ∎

3. Cube invariance and reciprocity

Fix Λ{,}\Lambda\in\{\mathbb{Z},\mathbb{Q}\}. Following [KSY22], we introduce a class of Λ\Lambda-modulus presheaves called cube invariant presheaves which is an analogue of the class of 𝔸1\mathbb{A}^{1}-invariant presheaves used in the classical theory of motives. This leads to the notion of Λ\Lambda-reciprocity presheaf, which is a presheaf with transfers that can be “lifted” to a cube invariant presheaf. We show that the notion of \mathbb{Q}-reciprocity presheaf is actually the same as the notion of (\mathbb{Z}-)reciprocity presheaf.

3.1. Cube invariance

The object ¯:=(1,[])MCorkΛ{\overline{\square}}:=(\mathbb{P}^{1},[\infty])\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k} is called the cube over kk. We write π:¯(Speck,)\pi\colon{\overline{\square}}\to(\operatorname{Spec}k,\emptyset) for the ambient morphism given by the canonical projection 1Speck\mathbb{P}^{1}\to\operatorname{Spec}k, and iε:(Speck,)¯(ε=0,1)i_{\varepsilon}\colon(\operatorname{Spec}k,\emptyset)\to{\overline{\square}}\;(\varepsilon=0,1) for the ambient morphism given by ε:Speck1\varepsilon\colon\operatorname{Spec}k\to\mathbb{P}^{1}.

Let 𝒳,𝒴MCorkΛ\mathcal{X},\mathcal{Y}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}. Two morphisms α0,α1MCorkΛ(𝒳,𝒴)\alpha_{0},\alpha_{1}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}(\mathcal{X},\mathcal{Y}) are called cube homotopic if there is some γMCorkΛ(𝒳¯,𝒴)\gamma\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}(\mathcal{X}\otimes{\overline{\square}},\mathcal{Y}) such that γ(idiε)=αε(ε=0,1)\gamma\circ({\operatorname{id}}\otimes i_{\varepsilon})=\alpha_{\varepsilon}\;(\varepsilon=0,1). In this case we write α0α1\alpha_{0}\sim\alpha_{1} and call γ\gamma a cube homotopy between α0\alpha_{0} and α1\alpha_{1}. We say that αMCorkΛ(𝒳,𝒴)\alpha\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}(\mathcal{X},\mathcal{Y}) is a cube homotopy equivalence if there is some βMCorkΛ(𝒴,𝒳)\beta\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}(\mathcal{Y},\mathcal{X}) such that βαid𝒳\beta\circ\alpha\sim{\operatorname{id}}_{\mathcal{X}} and αβid𝒴\alpha\circ\beta\sim{\operatorname{id}}_{\mathcal{Y}} hold. We call such β\beta a cube homotopy inverse of α\alpha.

Lemma 3.1.

The relation \sim is an equivalence relation.

Proof.

Let 𝒳,𝒴MCorkΛ\mathcal{X},\mathcal{Y}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k} and αiMCorkΛ(𝒳,𝒴)(i=0,1,2)\alpha_{i}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}(\mathcal{X},\mathcal{Y})\;(i=0,1,2). Let γ01MCorkΛ(𝒳¯,𝒴)\gamma_{01}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}(\mathcal{X}\otimes{\overline{\square}},\mathcal{Y}) (resp. γ12MCorkΛ(𝒳¯,𝒴)\gamma_{12}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}(\mathcal{X}\otimes{\overline{\square}},\mathcal{Y})) be a cube homotopy between α0\alpha_{0} and α1\alpha_{1} (resp. α1\alpha_{1} and α2\alpha_{2}). Then γ01+γ12α1(idπ):𝒳¯𝒴\gamma_{01}+\gamma_{12}-\alpha_{1}\circ({\operatorname{id}}\otimes\pi)\colon\mathcal{X}\otimes{\overline{\square}}\to\mathcal{Y} gives a cube homotopy between α0\alpha_{0} and α2\alpha_{2}. ∎

Lemma 3.2.

Consider the multiplication map μ:𝔸1×𝔸1𝔸1;(x,y)xy\mu\colon\mathbb{A}^{1}\times\mathbb{A}^{1}\to\mathbb{A}^{1};\;(x,y)\mapsto xy. Then the graph of μ\mu gives an element of MCork(¯¯,¯)=MCork(¯¯,¯)\operatorname{\mathrm{MCor}}^{\mathbb{Z}}_{k}({\overline{\square}}\otimes{\overline{\square}},{\overline{\square}})=\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}({\overline{\square}}\otimes{\overline{\square}},{\overline{\square}}).

Proof.

See [Kah+22, Lemma 5.1.1]. ∎

Lemma 3.3.

For any 𝒳MCorkΛ\mathcal{X}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}, the modulus correspondence idπMCorkΛ(𝒳¯,𝒳){\operatorname{id}}\otimes\pi\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}(\mathcal{X}\otimes{\overline{\square}},\mathcal{X}) is a cube homotopy equivalence.

Proof.

Let us prove that idi0:𝒳𝒳¯{\operatorname{id}}\otimes i_{0}\colon\mathcal{X}\to\mathcal{X}\otimes{\overline{\square}} gives a cube homotopy inverse. The composition (idπ)(idi0)({\operatorname{id}}\otimes\pi)\circ({\operatorname{id}}\otimes i_{0}) is the identity. Set γ:=idμ:𝒳¯¯𝒳¯\gamma:={\operatorname{id}}\otimes\mu\colon\mathcal{X}\otimes{\overline{\square}}\otimes{\overline{\square}}\to\mathcal{X}\otimes{\overline{\square}}. Then γ(idi0)=(idi0)(idπ)\gamma\circ({\operatorname{id}}\otimes i_{0})=({\operatorname{id}}\otimes i_{0})\circ({\operatorname{id}}\otimes\pi) and γ(idi1)=id\gamma\circ({\operatorname{id}}\otimes i_{1})={\operatorname{id}}, so we have (idi0)(idπ)id({\operatorname{id}}\otimes i_{0})\circ({\operatorname{id}}\otimes\pi)\sim{\operatorname{id}}. ∎

Definition 3.4.

We say that FPSh(MCorkΛ)F\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}) is cube invariant if the map (idπ):F(𝒳)F(𝒳¯)({\operatorname{id}}\otimes\pi)^{*}\colon F(\mathcal{X})\to F(\mathcal{X}\otimes{\overline{\square}}) is an isomorphism for all 𝒳MCorkΛ\mathcal{X}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}.

Lemma 3.5.

For FPSh(MCorkΛ)F\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}), the following conditions are equivalent:

  1. (1)

    FF is cube invariant.

  2. (2)

    For any 𝒳MCorkΛ\mathcal{X}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}, the map (idi0):F(𝒳¯)F(𝒳)({\operatorname{id}}\otimes i_{0})^{*}\colon F(\mathcal{X}\otimes{\overline{\square}})\to F(\mathcal{X}) is injective.

  3. (3)

    For any 𝒳MCorkΛ\mathcal{X}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}, the map (idi0)(idi1):F(𝒳¯)F(𝒳)({\operatorname{id}}\otimes i_{0})^{*}-({\operatorname{id}}\otimes i_{1})^{*}\colon F(\mathcal{X}\otimes{\overline{\square}})\to F(\mathcal{X}) is 0.

  4. (4)

    For any two cube homotopic morphisms α0,α1\alpha_{0},\alpha_{1} in MCorkΛ\operatorname{\mathrm{MCor}}^{\Lambda}_{k}, we have α0=α1\alpha_{0}^{*}=\alpha_{1}^{*} on FF.

Proof.

(1) \iff (2) follows from πi0=id\pi\circ i_{0}={\operatorname{id}}. (1)\implies(3) follows from πi0=πi1=id\pi\circ i_{0}=\pi\circ i_{1}={\operatorname{id}}. Let us prove (3)\implies(4). Let γ:𝒳¯𝒴\gamma\colon\mathcal{X}\otimes{\overline{\square}}\to\mathcal{Y} be a cube homotopy between α0\alpha_{0} and α1\alpha_{1}. Then we have α0=(idi0)γ=(idi1)γ=α1\alpha_{0}^{*}=({\operatorname{id}}\otimes i_{0})^{*}\gamma^{*}=({\operatorname{id}}\otimes i_{1})^{*}\gamma^{*}=\alpha_{1}^{*} on FF. This shows that (3)\implies(4). If (4) holds, then for any cube homotopy equivalence α:𝒳𝒴\alpha\colon\mathcal{X}\to\mathcal{Y} in MCorkΛ\operatorname{\mathrm{MCor}}^{\Lambda}_{k}, the map α:F(𝒴)F(𝒳)\alpha^{*}\colon F(\mathcal{Y})\to F(\mathcal{X}) is an isomorphism. Therefore (1) follows from Lemma 3.3. ∎

Lemma 3.6.

The class of cube invariant objects in PSh(MCorkΛ)\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}) is closed under taking subobjects, quotients and extensions.

Proof.

The claim for subobjects follows from the equivalence of (1) and (2) in Lemma 3.5. The remaining assertions then follow by the five lemma. ∎

There are two canonical ways to make a presheaf on MCorkΛ\operatorname{\mathrm{MCor}}^{\Lambda}_{k} cube invariant: one is to take the maximal cube invariant quotient, and the other is to take the maximal cube invariant subobject. The former is called the cube-localization and the latter is called the cube invariant part.

Definition 3.7.

The cube-localization of FPSh(MCorkΛ)F\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}) is defined by

h0\scaleobj0.8¯F(𝒳):=Coker(F(𝒳¯)i0i1F(𝒳)).h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F(\mathcal{X}):=\operatorname{Coker}(F(\mathcal{X}\otimes{\overline{\square}})\xrightarrow{i_{0}^{*}-i_{1}^{*}}F(\mathcal{X})).

There is a canonical epimorphism Fh0\scaleobj0.8¯FF\twoheadrightarrow h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F. We write h0\scaleobj0.8¯(𝒳)h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(\mathcal{X}) for h0\scaleobj0.8¯tr(𝒳)h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}\mathbb{Z}_{\operatorname{tr}}(\mathcal{X}).

Lemma 3.8.

The following assertions hold.

  1. (1)

    For any FPSh(MCorkΛ)F\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}), the quotient h0\scaleobj0.8¯Fh_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F of FF is cube invariant.

  2. (2)

    Let F,GPSh(MCorkΛ)F,G\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}). If GG is cube invariant, then the canonical homomorphism Hom(h0\scaleobj0.8¯F,G)Hom(F,G)\operatorname{Hom}(h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F,G)\to\operatorname{Hom}(F,G) is an isomorphism. In other words, h0\scaleobj0.8¯Fh_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F is the maximal cube invariant quotient of FF.

  3. (3)

    For any 𝒳MCorkΛ\mathcal{X}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}, the morphism h0\scaleobj0.8¯(𝒳¯)h0\scaleobj0.8¯(𝒳)h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(\mathcal{X}\otimes{\overline{\square}})\to h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(\mathcal{X}) induced by idπ{\operatorname{id}}\otimes\pi is an isomorphism.

Proof.

(1) follows from the equivalence of (1) and (3) in Lemma 3.5. (2) The canonical morphism Gh0\scaleobj0.8¯GG\to h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}G is an isomorphism by Lemma 3.5. The claim follows from this and (1). (3) By the Yoneda lemma, it suffices to prove that if FPSh(MCorkΛ)F\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}) is cube invariant then

Hom(h0\scaleobj0.8¯(𝒳),F)Hom(h0\scaleobj0.8¯(𝒳¯),F)\operatorname{Hom}(h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(\mathcal{X}),F)\to\operatorname{Hom}(h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(\mathcal{X}\otimes{\overline{\square}}),F)

is an isomorphism. By (2), this map can be identified with (idπ):F(𝒳)F(𝒳¯)({\operatorname{id}}\otimes\pi)^{*}\colon F(\mathcal{X})\to F(\mathcal{X}\otimes{\overline{\square}}), which is an isomorphism since FF is cube invariant. ∎

Definition 3.9.

The cube invariant part of FPSh(MCorkΛ)F\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}) is defined by

h0,\scaleobj0.8¯F(𝒳):=Hom(h0\scaleobj0.8¯(𝒳),F).h^{0,{\scaleobj{0.8}{\overline{\square}}}}F(\mathcal{X}):=\operatorname{Hom}(h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(\mathcal{X}),F).
Lemma 3.10.

The following assertions hold.

  1. (1)

    For any FPSh(MCorkΛ)F\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}), the subobject h0,\scaleobj0.8¯Fh^{0,{\scaleobj{0.8}{\overline{\square}}}}F of FF is cube invariant.

  2. (2)

    Let F,GPSh(MCorkΛ)F,G\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}). If FF is cube invariant, then the canonical homomorphism Hom(F,h0,\scaleobj0.8¯G)Hom(F,G)\operatorname{Hom}(F,h^{0,{\scaleobj{0.8}{\overline{\square}}}}G)\to\operatorname{Hom}(F,G) is an isomorphism. In other words, h0,\scaleobj0.8¯Fh^{0,{\scaleobj{0.8}{\overline{\square}}}}F is the maximal cube invariant subobject of FF.

  3. (3)

    The functor h0,\scaleobj0.8¯:PSh(MCorkΛ)PSh(MCorkΛ)h^{0,{\scaleobj{0.8}{\overline{\square}}}}\colon\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k})\to\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}) is right adjoint to h0\scaleobj0.8¯h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}.

Proof.

(1) follows from Lemma 3.8 (3). (2) If FPSh(MCorkΛ)F\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}) is cube invariant, then the canonical morphism h0,\scaleobj0.8¯FFh^{0,{\scaleobj{0.8}{\overline{\square}}}}F\to F is an isomorphism. The claim follows from this and (1). (3) follows from (2) and Lemma 3.8 (2). ∎

Remark 3.11.

In [KSY22], h0\scaleobj0.8¯h_{0}^{{\scaleobj{0.8}{\overline{\square}}}} and h0,\scaleobj0.8¯h^{0,{\scaleobj{0.8}{\overline{\square}}}} are defined to be functors taking values in the category of cube invariant presheaves. On the other hand, we define h0\scaleobj0.8¯h_{0}^{{\scaleobj{0.8}{\overline{\square}}}} and h0,\scaleobj0.8¯h^{0,{\scaleobj{0.8}{\overline{\square}}}} as endofunctors of PSh(MCorkΛ)\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}). This has the advantage that h0,\scaleobj0.8¯h^{0,{\scaleobj{0.8}{\overline{\square}}}} becomes right adjoint to h0\scaleobj0.8¯h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}.

Lemma 3.12.

Let F,GPSh(MCorkΛ)F,G\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}).

  1. (1)

    If GG is cube invariant, then om(F,G)\mathcal{H}\mathrm{om}(F,G) is also cube invariant.

  2. (2)

    The canonical epimorphism h0\scaleobj0.8¯(FG)h0\scaleobj0.8¯((h0\scaleobj0.8¯F)G)h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(F\otimes G)\twoheadrightarrow h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}((h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F)\otimes G) is an isomorphism.

Proof.

(1) For any 𝒳MCorkΛ\mathcal{X}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}, we have om(F,G)(𝒳)=Hom(tr(𝒳)F,G)\mathcal{H}\mathrm{om}(F,G)(\mathcal{X})=\operatorname{Hom}(\mathbb{Z}_{\operatorname{tr}}(\mathcal{X})\otimes F,G). Since GG is cube invariant, om(F,G)\mathcal{H}\mathrm{om}(F,G) is cube invariant when FF is representable. Since om(,G)\mathcal{H}\mathrm{om}({-},G) turns colimits into limits, the same is true for a general FF.

(2) For any HPSh(MCorkΛ)H\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}), we have

Hom(h0\scaleobj0.8¯(FG),H)\displaystyle\operatorname{Hom}(h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(F\otimes G),H)\simeq{} Hom(F,om(G,h0,\scaleobj0.8¯H))\displaystyle\operatorname{Hom}(F,\mathcal{H}\mathrm{om}(G,h^{0,{\scaleobj{0.8}{\overline{\square}}}}H)) (h0\scaleobj0.8¯h0,\scaleobj0.8¯h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}\dashv h^{0,{\scaleobj{0.8}{\overline{\square}}}})
\displaystyle\simeq{} Hom(h0\scaleobj0.8¯F,om(G,h0,\scaleobj0.8¯H))\displaystyle\operatorname{Hom}(h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F,\mathcal{H}\mathrm{om}(G,h^{0,{\scaleobj{0.8}{\overline{\square}}}}H)) (by (1))
\displaystyle\simeq{} Hom(h0\scaleobj0.8¯((h0\scaleobj0.8¯F)G),H).\displaystyle\operatorname{Hom}(h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}((h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F)\otimes G),H). (h0\scaleobj0.8¯h0,\scaleobj0.8¯h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}\dashv h^{0,{\scaleobj{0.8}{\overline{\square}}}})

Therefore we get the desired result by the Yoneda lemma.∎

Definition 3.13.

We define two functors h0h_{0} and h0h^{0} as follows.

  1. (1)

    We define h0:=ω!h0\scaleobj0.8¯:PSh(MCorkΛ)PSh(Cork)h_{0}:=\omega_{!}h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}\colon\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k})\to\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}). Since h0\scaleobj0.8¯h_{0}^{{\scaleobj{0.8}{\overline{\square}}}} is lax symmetric monoidal and ω!\omega_{!} is symmetric monoidal, it follows that h0h_{0} is lax symmetric monoidal. We write h0(𝒳)h_{0}(\mathcal{X}) for h0tr(𝒳)h_{0}\mathbb{Z}_{\operatorname{tr}}(\mathcal{X}).

  2. (2)

    We define h0:=h0,\scaleobj0.8¯ω:PSh(Cork)PSh(MCorkΛ)h^{0}:=h^{0,{\scaleobj{0.8}{\overline{\square}}}}\omega^{*}\colon\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k})\to\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}). It follows from the adjunctions h0\scaleobj0.8¯h0,\scaleobj0.8¯h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}\dashv h^{0,{\scaleobj{0.8}{\overline{\square}}}} and ω!ω\omega_{!}\dashv\omega^{*} that h0h^{0} is right adjoint to h0h_{0}.

Note that, by applying ω!\omega_{!} to the canonical epimorphism tr(𝒳)h0\scaleobj0.8¯(𝒳)\mathbb{Z}_{\operatorname{tr}}(\mathcal{X})\twoheadrightarrow h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(\mathcal{X}), we obtain a canonical epimorphism tr(X)h0(𝒳)\mathbb{Z}_{\operatorname{tr}}(X^{\circ})\twoheadrightarrow h_{0}(\mathcal{X}).

Remark 3.14.

By Lemma 3.12, we have h0\scaleobj0.8¯(FG)h0\scaleobj0.8¯((h0\scaleobj0.8¯F)(h0\scaleobj0.8¯G))h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(F\otimes G)\simeq h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}((h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F)\otimes(h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}G)) for any F,GPSh(MCorkΛ)F,G\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}). The latter is a quotient of h0\scaleobj0.8¯Fh0\scaleobj0.8¯Gh_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F\otimes h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}G, so we get an epimorphism h0\scaleobj0.8¯Fh0\scaleobj0.8¯Gh0\scaleobj0.8¯(FG)h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}F\otimes h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}G\twoheadrightarrow h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}(F\otimes G). In particular, h0\scaleobj0.8¯h_{0}^{{\scaleobj{0.8}{\overline{\square}}}} admits a canonical lax symmetric monoidal structure. Applying the exact symmetric monoidal functor ω!\omega_{!}, we also see that there is a canonical epimorphism h0Fh0Gh0(FG)h_{0}F\otimes h_{0}G\twoheadrightarrow h_{0}(F\otimes G).

Remark 3.15.

In [KSY22], a functor named ωCI\omega^{\operatorname{CI}} is defined by the same formula as h0h^{0}, but it is regarded as a functor taking values in the category of cube invariant presheaves rather than PSh(MCorkΛ)\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}).

Lemma 3.16.

Let 𝒳MCorkΛ\mathcal{X}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k} and YSmkY\in\operatorname{\mathrm{Sm}}_{k}. The canonical epimorphism tr(X)h0(𝒳)\mathbb{Z}_{\operatorname{tr}}(X^{\circ})\twoheadrightarrow h_{0}(\mathcal{X}) induces an isomorphism h0(𝒳)(Y)Coker(M¯CorkΛ(Y¯,𝒳)i0i1Cork(Y,X))h_{0}(\mathcal{X})(Y)\simeq\operatorname{Coker}(\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}(Y\otimes{\overline{\square}},\mathcal{X})\xrightarrow{i_{0}^{*}-i_{1}^{*}}\operatorname{\mathrm{Cor}}_{k}(Y,X^{\circ})).

Proof.

This follows from the definition of h0h_{0} and the isomorphism tr(𝒳)(Y)tr(X)(Y)\mathbb{Z}_{\operatorname{tr}}(\mathcal{X})(Y)\simeq\mathbb{Z}_{\operatorname{tr}}(X^{\circ})(Y). ∎

The functors we have defined so far can be summarized as follows.

PSh(MCorkΛ)\textstyle{\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h0\scaleobj0.8¯\scriptstyle{h_{0}^{{\scaleobj{0.8}{\overline{\square}}}}}h0\scriptstyle{h_{0}}PSh(MCorkΛ)\textstyle{\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ω!\scriptstyle{\omega_{!}}h0,\scaleobj0.8¯\scriptstyle{h^{0,{\scaleobj{0.8}{\overline{\square}}}}} \dashv PSh(Cork)\textstyle{\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ω\scriptstyle{\omega^{*}} \dashv h0\scriptstyle{h^{0}}

3.2. Reciprocity sheaves

We define the notion of Λ\Lambda-reciprocity sheaf following [KSY22].

Definition 3.17.

Let XSmkX\in\operatorname{\mathrm{Sm}}_{k}, FPSh(Cork)F\in\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) and aF(X)a\in F(X). Take 𝒴MCorkΛ\mathcal{Y}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k} with YXY^{\circ}\simeq X. We say that 𝒴\mathcal{Y} is a Λ\Lambda-modulus for aa if a:tr(X)Fa\colon\mathbb{Z}_{\operatorname{tr}}(X)\to F factors through the canonical epimorphism tr(X)h0(𝒴)\mathbb{Z}_{\operatorname{tr}}(X)\twoheadrightarrow h_{0}(\mathcal{Y}). This is equivalent to saying that a:tr(𝒴)ωFa\colon\mathbb{Z}_{\operatorname{tr}}(\mathcal{Y})\to\omega^{*}F factors through h0FωFh^{0}F\subset\omega^{*}F.

Lemma 3.18.

Let FPSh(Cork)F\in\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}). The following conditions are equivalent:

  1. (1)

    The counit morphism h0h0FFh_{0}h^{0}F\to F is an isomorphism.

  2. (2)

    For any XSmkX\in\operatorname{\mathrm{Sm}}_{k} and any aF(X)a\in F(X), there exists 𝒴MCorkΛ\mathcal{Y}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k} with YXY^{\circ}\simeq X such that ah0F(𝒴)a\in h^{0}F(\mathcal{Y}).

  3. (3)

    Every section of FF admits a Λ\Lambda-modulus.

Proof.

The equivalence of (1) and (2) follows from the formula

h0h0F(X)=ω!h0F(X)=lim𝒴MCorkΛ,YXh0F(𝒴)F(X).h_{0}h^{0}F(X)=\omega_{!}h^{0}F(X)=\varinjlim_{\mathcal{Y}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k},Y^{\circ}\simeq X}h^{0}F(\mathcal{Y})\subset F(X).

The equivalence of (2) and (3) is clear from the definition. ∎

Definition 3.19.

We say that FF has Λ\Lambda-reciprocity or FF is a Λ\Lambda-reciprocity presheaf if it satisfies the equivalent conditions in Lemma 3.18. If FF is moreover a Nisnevich sheaf, then we say that FF is a reciprocity sheaf. We define RSCkΛ\operatorname{\mathrm{RSC}}^{\Lambda}_{k} (resp. RSCk,NisΛ\operatorname{\mathrm{RSC}}^{\Lambda}_{k,{\operatorname{Nis}}}) to be the full subcategory of PSh(Cork)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) spanned by Λ\Lambda-reciprocity presheaves (resp. Λ\Lambda-reciprocity sheaves).

The category RSCkΛ\operatorname{\mathrm{RSC}}^{\Lambda}_{k} is closed under taking subobjects and quotients in PSh(Cork)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}). In particular, RSCkΛ\operatorname{\mathrm{RSC}}^{\Lambda}_{k} is an abelian category and the inclusion functor RSCkΛPSh(Cork)\operatorname{\mathrm{RSC}}^{\Lambda}_{k}\to\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) is exact.

Lemma 3.20.

We have RSCk=RSCk\operatorname{\mathrm{RSC}}^{\mathbb{Z}}_{k}=\operatorname{\mathrm{RSC}}^{\mathbb{Q}}_{k}.

Proof.

Let XSmkX\in\operatorname{\mathrm{Sm}}_{k}, FPSh(Cork)F\in\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) and aF(X)a\in F(X). Then it is clear that every \mathbb{Z}-modulus 𝒴\mathcal{Y} for aa is also a \mathbb{Q}-modulus for aa, since the description of h0(𝒴)h_{0}(\mathcal{Y}) (Lemma 3.16) is the same for Λ=,\Lambda=\mathbb{Z},\mathbb{Q}. Conversely, if 𝒴\mathcal{Y} is a \mathbb{Q}-modulus for aa, then so is (Y,nDY)(Y,nD_{Y}) for n>0n\in\mathbb{Z}_{>0}, so there exists a \mathbb{Z}-modulus for aa. ∎

We thus write RSCk\operatorname{\mathrm{RSC}}_{k} for RSCk=RSCk\operatorname{\mathrm{RSC}}^{\mathbb{Z}}_{k}=\operatorname{\mathrm{RSC}}^{\mathbb{Q}}_{k} and call its objects reciprocity presheaves over kk.

Lemma 3.21.

If FPSh(MCorkΛ)F\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\Lambda}_{k}) is cube invariant, then ω!FRSCk\omega_{!}F\in\operatorname{\mathrm{RSC}}_{k}.

Proof.

The composition h0Fh0(unit)h0h0h0Fcounith0Fh_{0}F\xrightarrow{h_{0}(\text{unit})}h_{0}h^{0}h_{0}F\xrightarrow{\text{counit}}h_{0}F is the identity. Since the counit morphism h0h0GGh_{0}h^{0}G\to G is a monomorphism for any GPSh(CorS)G\in\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{S}), it follows that the counit morphism h0h0h0Fh0Fh_{0}h^{0}h_{0}F\to h_{0}F is an isomorphism. Therefore ω!F=h0F\omega_{!}F=h_{0}F has reciprocity. ∎

Remark 3.22.

Let FRSCkF\in\operatorname{\mathrm{RSC}}_{k} and 𝒳MCorkΛ\mathcal{X}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k}. The group h0F(𝒳)h^{0}F(\mathcal{X}) can be thought of as the subgroup of ωF(𝒳)=F(X)\omega^{*}F(\mathcal{X})=F(X^{\circ}) consisting of elements whose “ramification” is bounded by DXD_{X}. This actually recovers several classical notions in ramification theory such as the Artin conductor or the irregularity [RS21]. For further developments of the ramification theory of reciprocity sheaves, see [RS21], [RS23], [RS23a], and [RS22].

4. Modulus curves and motivic Hasse-Arf theorem

Fix Λ{,}\Lambda\in\{\mathbb{Z},\mathbb{Q}\}. In this section, we describe h0(𝒳)h_{0}(\mathcal{X}) for a Λ\Lambda-modulus curve (Definition 4.6) 𝒳\mathcal{X} over kk, using the Chow group of relative 0-cycles (Definition 4.9). As a corollary, we prove a motivic analogue of the Hasse-Arf theorem.

4.1. Admissible rational functions

Definition 4.1.

A Λ\Lambda-modulus pair 𝒳M¯CorkΛ\mathcal{X}\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} is called normal integral if XX is a normal integral scheme.

Definition 4.2.

Let 𝒳M¯CorkΛ\mathcal{X}\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} be a normal integral Λ\Lambda-modulus pair. We say that fk(X)×f\in k(X)^{\times} is admissible with respect to DXD_{X} if there is some open neighborhood UU of |DX||D_{X}| such that

  1. (1)

    ff is regular and invertible on UU, and

  2. (2)

    div(f1)DX\operatorname{div}(f-1)\geq D_{X} holds on UU (including the case f=1f=1).

We define Adm(𝒳)k(X)×\operatorname{Adm}(\mathcal{X})\subset k(X)^{\times} to be the subset consisting of rational functions which are admissible with respect to DXD_{X}. It follows from the identity fg1=(f1)g+(g1)fg-1=(f-1)g+(g-1) that Adm(𝒳)\operatorname{Adm}(\mathcal{X}) is a subgroup of k(X)×k(X)^{\times}.

Lemma 4.3.

Let 𝒳M¯CorkΛ\mathcal{X}\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} be a normal integral Λ\Lambda-modulus pair. Then fk(X)×f\in k(X)^{\times} is admissible with respect to DXD_{X} if and only if for any discrete valuation ring RR with fraction field k(X)k(X) and any morphism ρ:SpecRX\rho\colon\operatorname{Spec}R\to X extending Speck(X)X\operatorname{Spec}k(X)\to X such that ρDX\rho^{*}D_{X}\neq\emptyset, the inequality vR(f1)vR(ρDX)v_{R}(f-1)\geq v_{R}(\rho^{*}D_{X}) holds in Λ\Lambda.

Proof.

The “only if” part is easy. Let us prove the “if” part. Let x|DX|x\in|D_{X}|. Our assumption implies that for any discrete valuation ring RR with fraction field k(X)k(X) dominating 𝒪X,x\mathcal{O}_{X,x}, we have fR×f\in R^{\times}. Since 𝒪X,x\mathcal{O}_{X,x} is a noetherian normal domain, this shows that f𝒪X,x×f\in\mathcal{O}_{X,x}^{\times}. Therefore ff is invertible on some open neighborhood UU of |DX||D_{X}|. The condition (1) in Definition 4.2 is satisfied for this UU. Let us verify the condition (2). Let RR be a valuation ring with fraction field k(X)k(X) and ρ:SpecRU\rho\colon\operatorname{Spec}R\to U be a morphism extending Speck(X)U\operatorname{Spec}k(X)\to U. If ρDX\rho^{*}D_{X} is trivial, then we have ρdiv(f1)0=ρDX\rho^{*}\operatorname{div}(f-1)\geq 0=\rho^{*}D_{X}, since f1f-1 is regular on UU. Otherwise, our assumption implies that ρdiv(f1)ρDX\rho^{*}\operatorname{div}(f-1)\geq\rho^{*}D_{X}. Since UU is normal, we get div(f1)DX\operatorname{div}(f-1)\geq D_{X} on UU. ∎

Lemma 4.4.

Let (K,v)(K,v) be a discrete valuation field and L/KL/K be a finite extension. Let γ>0\gamma\in\mathbb{Q}_{>0}, fL×f\in L^{\times} and suppose that for every discrete valuation w:L×w\colon L^{\times}\to\mathbb{Q} extending vv, we have w(f1)γw(f-1)\geq\gamma. Then we have v(NmL/K(f)1)γv(\operatorname{Nm}_{L/K}(f)-1)\geq\gamma.

Proof.

Let ii denote the inseparable degree of L/KL/K. Fix an algebraic closure K¯\overline{K} of KK. Then vv can be extended to a valuation on K¯\overline{K}. Let us fix such an extension v~\widetilde{v}. We have

NmL/K(f)=σσ(f)i=σ(1+σ(f1))i\operatorname{Nm}_{L/K}(f)=\textstyle\prod_{\sigma}\sigma(f)^{i}=\prod_{\sigma}(1+\sigma(f-1))^{i}

where σ\sigma runs over the set of distinct KK-embeddings of LL into K¯\overline{K}. Now v~σ\widetilde{v}\circ\sigma gives a valuation on LL extending vv for each σ\sigma, so we have v~(σ(f1))γ\widetilde{v}(\sigma(f-1))\geq\gamma by our assumption. By the above expression for NmL/K(f)\operatorname{Nm}_{L/K}(f), we get the desired inequality. ∎

Lemma 4.5.

Let 𝒳,𝒞MCorkΛ\mathcal{X},\mathcal{C}\in\operatorname{\mathrm{MCor}}^{\Lambda}_{k} be normal integral Λ\Lambda-modulus pairs and q:𝒳𝒞q\colon\mathcal{X}\to\mathcal{C} be an ambient morphism such that XCX\to C is proper and generically finite. Then for any fAdm(𝒳)f\in\operatorname{Adm}(\mathcal{X}), we have Nmk(X)/k(C)(f)Adm(𝒞)\operatorname{Nm}_{k(X)/k(C)}(f)\in\operatorname{Adm}(\mathcal{C}).

Proof.

By Lemma 4.3, it suffices to show that for any discrete valuation ring RR with fraction field k(C)k(C) and any morphism ρ:SpecRC\rho\colon\operatorname{Spec}R\to C extending Speck(C)C\operatorname{Spec}k(C)\to C such that ρDC\rho^{*}D_{C}\neq\emptyset, we have vR(Nmk(X)/k(C)(f)1)vR(ρDC)v_{R}(\operatorname{Nm}_{k(X)/k(C)}(f)-1)\geq v_{R}(\rho^{*}D_{C}). Let w:k(X)×w\colon k(X)^{\times}\to\mathbb{Q} be an arbitrary extension of vRv_{R} and RR^{\prime} be its valuation ring. Since q:XCq\colon X\to C is proper, there is a unique dashed arrow φ\varphi in the following diagram which makes it commute:

Speck(X)\textstyle{\operatorname{Spec}k(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SpecR\textstyle{\operatorname{Spec}R^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q}Speck(C)\textstyle{\operatorname{Spec}k(C)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SpecR\textstyle{\operatorname{Spec}R\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ\scriptstyle{\rho}C.\textstyle{C.}

By Lemma 4.3, we have w(f1)w(φDX)w(φqDC)=vR(ρDC)w(f-1)\geq w(\varphi^{*}D_{X})\geq w(\varphi^{*}q^{*}D_{C})=v_{R}(\rho^{*}D_{C}). By Lemma 4.4, we get vR(Nmk(X)/k(C)(f)1)vR(ρDC)v_{R}(\operatorname{Nm}_{k(X)/k(C)}(f)-1)\geq v_{R}(\rho^{*}D_{C}), as was to be shown. ∎

4.2. Modulus curves

In this subsection, we fix SSmkS\in\operatorname{\mathrm{Sm}}_{k} which is connected.

Definition 4.6.

A Λ\Lambda-modulus curve over SS is a modulus pair 𝒞M¯CorkΛ\mathcal{C}\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k} equipped with a proper smooth morphism CSC\to S of relative dimension 11 such that CSC^{\circ}\to S is quasi-affine.

Lemma 4.7.

Let 𝒞\mathcal{C} be a Λ\Lambda-modulus curve over SS. If VV is a closed subscheme of CC contained in CC^{\circ}, then VV is finite over SS.

Proof.

Since CSC\to S is proper, VSV\to S is also proper. For any xSx\in S, the fiber VxV_{x} of VV is a closed subscheme of CxC^{\circ}_{x} which is proper over Speck(x)\operatorname{Spec}k(x). By our assumption that CxC^{\circ}_{x} is quasi-affine, there is an open immersion CxBC^{\circ}_{x}\hookrightarrow B into an affine scheme BB. The properness of VxV_{x} implies that the image of VxV_{x} in BB is closed. It follows that VxV_{x} is finite over k(x)k(x) and hence VSV\to S is quasi-finite. By Zariski’s main theorem, we conclude that VSV\to S is finite. ∎

Lemma 4.8.

Let 𝒞\mathcal{C} be a Λ\Lambda-modulus curve over SS. For an integral closed subscheme VCV\subset C^{\circ}, the following conditions are equivalent:

  1. (1)

    VSV\to S is finite surjective.

  2. (2)

    VV has codimension 11 and is closed in CC.

Proof.

(1)\implies(2): Suppose that VSV\to S is finite surjective. Then VV is finite over CC and hence is closed in CC. Let ξ\xi be the generic point of VV and η\eta its image in SS. Then the fiber dimension theorem for flat morphisms [GW10, Corollary 14.95] implies codimC(ξ)=codimS(η)+codimCη(ξ){\operatorname{codim}}_{C^{\circ}}(\xi)={\operatorname{codim}}_{S}(\eta)+{\operatorname{codim}}_{C^{\circ}_{\eta}}(\xi). Since VSV\to S is finite surjective, we have codimS(η)=0{\operatorname{codim}}_{S}(\eta)=0 and codimCη(ξ)=1{\operatorname{codim}}_{C^{\circ}_{\eta}}(\xi)=1. This implies codimC(ξ)=1{\operatorname{codim}}_{C^{\circ}}(\xi)=1.

(2)\implies(1): Suppose that VV has codimension 11 and is closed in CC. Then VV is finite over SS by Lemma 4.7. We have codimC(ξ)=1{\operatorname{codim}}_{C^{\circ}}(\xi)=1 and codimCη(ξ)=1{\operatorname{codim}}_{C^{\circ}_{\eta}}(\xi)=1, so the fiber dimension theorem implies codimS(η)=0{\operatorname{codim}}_{S}(\eta)=0. Therefore VSV\to S is surjective. ∎

Definition 4.9.

Let 𝒞\mathcal{C} be a Λ\Lambda-modulus curve over SS with CC connected. Then for any fAdm(𝒞)f\in\operatorname{Adm}(\mathcal{C}), the components of the Weil divisor div(f)\operatorname{div}(f) on CC^{\circ} are closed in CC and hence div(f)CorS(S,C)\operatorname{div}(f)\in\operatorname{\mathrm{Cor}}_{S}(S,C^{\circ}) by Lemma 4.8. We define the Chow group of relative 0-cycles of 𝒞\mathcal{C} by

CH0(𝒞/S):=Coker(Adm(𝒞)divCorS(S,C)).{\operatorname{CH}}_{0}(\mathcal{C}/S):=\operatorname{Coker}(\operatorname{Adm}(\mathcal{C})\xrightarrow{\operatorname{div}}\operatorname{\mathrm{Cor}}_{S}(S,C^{\circ})).

If 𝒞\mathcal{C} is a Λ\Lambda-modulus curve over SS with CC non-connected, then we define CH0(𝒞/S){\operatorname{CH}}_{0}(\mathcal{C}/S) by taking a direct sum over connected components.

Remark 4.10.

When S=SpeckS=\operatorname{Spec}k and Λ=\Lambda=\mathbb{Z}, the above definition of CH0(𝒞/S){\operatorname{CH}}_{0}(\mathcal{C}/S) coincides with the definition of the relative Chow group of 0-cycles C(C,DC)\mathrm{C}(C,D_{C}) from [KS16, Definition 1.6]. The latter group is also defined for higher dimensional varieties, and it was used to establish ramified higher dimensional class field theory.

When Λ=\Lambda=\mathbb{Z}, we have the following comparison result between the Chow group of relative 0-cycles and the relative Picard group.

Lemma 4.11.

Let 𝒞\mathcal{C} be a \mathbb{Z}-modulus curve over SS such that DCD_{C} is contained in some affine open subset of CC. Then we have CH0(𝒞/S)Pic(C,DC){\operatorname{CH}}_{0}(\mathcal{C}/S)\simeq\operatorname{Pic}(C,D_{C}).

Proof.

We may assume that CC is connected. Then Adm(𝒞)\operatorname{Adm}(\mathcal{C}) consists of rational functions ff which is regular and invertible on some neighborhood of DCD_{C} and f|DC=1f|_{D_{C}}=1. On the other hand, Pic(C,DC)\operatorname{Pic}(C,D_{C}) is by definition the group of isomorphism classes of pairs (L,α)(L,\alpha) where LL is a line bundle on CC and α\alpha is a nowhere-vanishing section of L|DCL|_{D_{C}}. By our assumption that DCD_{C} has an affine open neighborhood, any such α\alpha can be extended to a rational section of LL which is regular and invertible on some neighborhood of |DC||D_{C}|. For two such extensions α~1,α~2\widetilde{\alpha}_{1},\widetilde{\alpha}_{2}, the quotient α~1/α~2\widetilde{\alpha}_{1}/\widetilde{\alpha}_{2} is admissible with respect to DCD_{C}. Therefore Pic(C,DC)\operatorname{Pic}(C,D_{C}) can be identified with the cokernel of div:Adm(𝒞)Q\operatorname{div}\colon\operatorname{Adm}(\mathcal{C})\to Q where QQ is the group of Cartier divisors on CC whose support is disjoint from |DC||D_{C}|. We have Q=CorS(S,C)Q=\operatorname{\mathrm{Cor}}_{S}(S,C^{\circ}) by Lemma 4.8, so the claim follows. ∎

The following result generalizes [RY16, Theorem 1.1] to Λ=\Lambda=\mathbb{Q}.

Lemma 4.12.

Let 𝒞\mathcal{C} be a Λ\Lambda-modulus curve over SS. Let M¯CorSΛ(¯×S,𝒞)\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{S}({\overline{\square}}\times S,\mathcal{C}) be the intersection of M¯CorkΛ(¯×S,𝒞)\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}({\overline{\square}}\times S,\mathcal{C}) and CorS(𝔸1×S,C)\operatorname{\mathrm{Cor}}_{S}(\mathbb{A}^{1}\times S,C^{\circ}). Then we have

CH0(𝒞/S)Coker(M¯CorSΛ(¯×S,𝒞)i0i1CorS(S,C)).{\operatorname{CH}}_{0}(\mathcal{C}/S)\simeq\operatorname{Coker}(\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{S}({\overline{\square}}\times S,\mathcal{C})\xrightarrow{i_{0}^{*}-i_{1}^{*}}\operatorname{\mathrm{Cor}}_{S}(S,C^{\circ})).
Proof.

We may assume that CC is connected. Throughout this proof we identify ¯{\overline{\square}} with (1,[1])(\mathbb{P}^{1},[1]) via the isomorphism ttt1t\mapsto\frac{t}{t-1}, and set :=1{1}\square:=\mathbb{P}^{1}\setminus\{1\}. It suffices to prove that

Im(Adm(𝒞)divCorS(S,C))=Im(M¯CorSΛ(¯×S,𝒞)i0iCorS(S,C)).\operatorname{Im}(\operatorname{Adm}(\mathcal{C})\xrightarrow{\operatorname{div}}\operatorname{\mathrm{Cor}}_{S}(S,C^{\circ}))=\operatorname{Im}(\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{S}({\overline{\square}}\times S,\mathcal{C})\xrightarrow{i_{0}^{*}-i_{\infty}^{*}}\operatorname{\mathrm{Cor}}_{S}(S,C^{\circ})).

We will prove a stronger statement that there is a surjective homomorphism Φ:M¯CorSΛ(¯×S,𝒞)Adm(𝒞)\Phi\colon\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{S}({\overline{\square}}\times S,\mathcal{C})\twoheadrightarrow\operatorname{Adm}(\mathcal{C}) which makes the following diagram commutative:

(4.5)

Let [V]M¯CorSΛ(¯×S,𝒞)[V]\in\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{S}({\overline{\square}}\times S,\mathcal{C}). Let V¯\overline{V} denote the scheme-theoretic closure of VV in 1×S×C\mathbb{P}^{1}\times S\times C and p:V¯1p\colon\overline{V}\to\mathbb{P}^{1}, q:V¯Cq\colon\overline{V}\to C be the canonical projections. Then we have p[1]|V¯NqDC|V¯Np^{*}[1]|_{\overline{V}^{N}}\geq q^{*}D_{C}|_{\overline{V}^{N}} and hence pAdm(V¯N,qDC|V¯N)p\in\operatorname{Adm}(\overline{V}^{N},q^{*}D_{C}|_{\overline{V}^{N}}). We define

Φ([V])={Nmk(V)/k(C)(p)(VC is generically finite),1(otherwise).\Phi([V])=\begin{cases}\operatorname{Nm}_{k(V)/k(C)}(p)&(V\to C\text{ is generically finite}),\\ 1&(\text{otherwise}).\end{cases}

Lemma 4.5 shows that this gives a homomorphism Φ:M¯CorSΛ(¯×S,𝒞)Adm(𝒞)\Phi\colon\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{S}({\overline{\square}}\times S,\mathcal{C})\to\operatorname{Adm}(\mathcal{C}).

Let us prove that (4.5) is commutative. Let KK be the function field of SS. Since the operations appearing in (4.5) are compatible with base change to SpecK\operatorname{Spec}K and the map CorS(S,C)CorK(SpecK,CK)\operatorname{\mathrm{Cor}}_{S}(S,C^{\circ})\to\operatorname{\mathrm{Cor}}_{K}(\operatorname{Spec}K,C^{\circ}_{K}) is injective, we may assume that S=SpecKS=\operatorname{Spec}K. In this case the claim follows from a standard computation of cycles.

It remains to show that Φ\Phi is surjective. Let fAdm(𝒞)f\in\operatorname{Adm}(\mathcal{C}). If f=1f=1, then Φ(0)=f\Phi(0)=f. Otherwise, we take an open neighborhood UU of |DC||D_{C}| satisfying the conditions (1) and (2) in Definition 4.2. Then ff is regular and invertible on UU. Define V¯1×S×C\overline{V}\subset\mathbb{P}^{1}\times S\times C to be the closure of the graph of f|U:U𝔸1×Sf|_{U}\colon U\to\mathbb{A}^{1}\times S. Let p:V¯1p\colon\overline{V}\to\mathbb{P}^{1} and q:V¯Cq\colon\overline{V}\to C be the canonical projections. Let VV¯V\subset\overline{V} denote the inverse image of =1{1}\square=\mathbb{P}^{1}\setminus\{1\} under pp. Since f=1f=1 holds on |DC||D_{C}|, VV is contained in ×C\square\times C^{\circ}. We will prove that [V][V] gives an element of M¯CorSΛ(¯×S,𝒞)\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{S}({\overline{\square}}\times S,\mathcal{C}) such that Φ([V])=f\Phi([V])=f.

Since f=1f=1 holds on |DC||D_{C}| and f1f\neq 1, VV is dominant over ×S\square\times S. Applying Lemma 4.7 to the modulus curve ×𝒞\square\times\mathcal{C} over ×S\square\times S, we see that VV is finite over ×S\square\times S. The admissibility of VV follows from the assumption that div(f1)πDC\operatorname{div}(f-1)\geq\pi^{*}D_{C} holds on UU. Finally, we have Φ([V])=Nmk(V)/k(C)(p)=f\Phi([V])=\operatorname{Nm}_{k(V)/k(C)}(p)=f by construction. ∎

Theorem 4.13.

Let 𝒞\mathcal{C} be a Λ\Lambda-modulus curve over kk. Then for any connected SSmkS\in\operatorname{\mathrm{Sm}}_{k}, the canonical surjection Cork(S,C)=CorS(S,S×C)CH0(S×𝒞/S)\operatorname{\mathrm{Cor}}_{k}(S,C^{\circ})=\operatorname{\mathrm{Cor}}_{S}(S,S\times C^{\circ})\twoheadrightarrow{\operatorname{CH}}_{0}(S\times\mathcal{C}/S) induces an isomorphism

h0(𝒞)(S)CH0(S×𝒞/S).h_{0}(\mathcal{C})(S)\xrightarrow{\sim}{\operatorname{CH}}_{0}(S\times\mathcal{C}/S).

Moreover, if Λ=\Lambda=\mathbb{Z} and SS is affine, then we also have CH0(S×𝒞/S)Pic(S×C,S×DC){\operatorname{CH}}_{0}(S\times\mathcal{C}/S)\simeq\operatorname{Pic}(S\times C,S\times D_{C}).

Proof.

Applying Lemma 4.12 to the Λ\Lambda-modulus curve S×𝒞S\times\mathcal{C} over SS, we get

CH0(𝒞/S)\displaystyle{\operatorname{CH}}_{0}(\mathcal{C}/S)\simeq{} Coker(M¯CorSΛ(¯S,S×𝒞)i0i1CorS(S,S×C))\displaystyle\operatorname{Coker}(\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{S}({\overline{\square}}_{S},S\times\mathcal{C})\xrightarrow{i_{0}^{*}-i_{1}^{*}}\operatorname{\mathrm{Cor}}_{S}(S,S\times C^{\circ}))
\displaystyle\simeq{} Coker(M¯CorkΛ(¯S,𝒞)i0i1Cork(S,C)).\displaystyle\operatorname{Coker}(\operatorname{\mathrm{\underline{M}Cor}}^{\Lambda}_{k}({\overline{\square}}_{S},\mathcal{C})\xrightarrow{i_{0}^{*}-i_{1}^{*}}\operatorname{\mathrm{Cor}}_{k}(S,C^{\circ})).

The last term is isomorphic to h0(𝒞)(S)h_{0}(\mathcal{C})(S) by Lemma 3.16. The second assertion follows from Lemma 4.11. ∎

4.3. Motivic Hasse-Arf theorem

Theorem 4.14.

Let 𝒞\mathcal{C} be a \mathbb{Q}-modulus curve over kk. Then we have

h0(C,DC)h0(C,DC).h_{0}(C,\lceil D_{C}\rceil)\xrightarrow{\sim}h_{0}(C,D_{C}).
Proof.

We want to prove that for any connected SSmkS\in\operatorname{\mathrm{Sm}}_{k}, the canonical map h0(C,DC)(S)h0(C,DC)(S)h_{0}(C,\lceil D_{C}\rceil)(S)\to h_{0}(C,D_{C})(S) is an isomorphism. By Theorem 4.13, it suffices to show that

CH0((S×C,S×DC)/S)CH0((S×C,S×DC)/S){\operatorname{CH}}_{0}((S\times C,S\times\lceil D_{C}\rceil)/S)\to{\operatorname{CH}}_{0}((S\times C,S\times D_{C})/S)

is an isomorphism. We have S×DC=S×DCS\times\lceil D_{C}\rceil=\lceil S\times D_{C}\rceil since SS is smooth over kk. Now the claim follows from the fact that div(f1)\operatorname{div}(f-1) has integral coefficients for any rational function ff on S×CS\times C. ∎

The following corollary can be seen as a motivic analogue of the Hasse-Arf theorem.

Corollary 4.15.

Let 𝒞\mathcal{C} be a \mathbb{Q}-modulus curve over kk. Then for any FRSCkF\in\operatorname{\mathrm{RSC}}_{k} we have

h0F(C,DC)h0F(C,DC).h^{0}F(C,D_{C})\xrightarrow{\sim}h^{0}F(C,\lceil D_{C}\rceil).
Proof.

This follows from Theorem 4.14. ∎

Remark 4.16.

We have tr(X,DX)tr(X,DX)\mathbb{Z}_{\operatorname{tr}}(X,D_{X})\neq\mathbb{Z}_{\operatorname{tr}}(X,\lceil D_{X}\rceil) in general, so Corollary 4.15 is not obvious from the definition. Actually, Corollary 4.15 is false for modulus pairs of higher dimensions.

5. Motivic construction of the ring of Witt vectors

In this section, we present a construction of the ring of Witt vectors using \mathbb{Q}-modulus pairs.

5.1. Usual construction

First we recall the usual definition of the ring of Witt vectors; see [Hes15] for details. Let AA be a ring and n0n\geq 0. The group of big Witt vectors of length nn of AA is defined by

𝕎n(A):=1+tA[t]/(tn+1)(A[t]/(tn+1))×.\mathbb{W}_{n}(A):=1+tA[t]/(t^{n+1})\subset(A[t]/(t^{n+1}))^{\times}.

For aAa\in A, we write [a][a] for the element 1at𝕎n(A)1-at\in\mathbb{W}_{n}(A). The presheaf X𝕎n(𝒪(X))X\mapsto\mathbb{W}_{n}(\mathcal{O}(X)) on Smk\operatorname{\mathrm{Sm}}_{k} is represented by an algebraic group whose underlying scheme is isomorphic to 𝔸n\mathbb{A}^{n}, so it can be regarded as an object of PSh(Cork)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) (see [BVK16, Lemma 1.4.4]). There are several important morphisms in PSh(Cork)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}):

U\displaystyle U :𝕎n,\displaystyle\colon\mathbb{Z}\to\mathbb{W}_{n}, \displaystyle\star :𝕎n𝕎n𝕎n,\displaystyle\colon\mathbb{W}_{n}\otimes\mathbb{W}_{n}\to\mathbb{W}_{n},
Fs\displaystyle F_{s} :𝕎sn𝕎n(s1),\displaystyle\colon\mathbb{W}_{sn}\to\mathbb{W}_{n}\quad(s\geq 1), Vs\displaystyle V_{s} :𝕎n𝕎sn(s1).\displaystyle\colon\mathbb{W}_{n}\to\mathbb{W}_{sn}\quad(s\geq 1).

They are called the unit, the multiplication, the Frobenius, and the Verschiebung. The unit UU is given by 1[1]1\mapsto[1], and the Verschiebung VsV_{s} is induced by ttst\mapsto t^{s}. The multiplication \star is characterized by [a][b]=[ab][a]\star[b]=[ab], and the Frobenius FsF_{s} is characterized by Fs([a])=[as]F_{s}([a])=[a^{s}]. We define

𝕎^n:=𝕎n𝔾mPSh(Cork).\widehat{\mathbb{W}}_{n}:=\mathbb{W}_{n}\oplus\mathbb{G}_{m}\oplus\mathbb{Z}\in\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}).

For a ring AA and aA×a\in A^{\times}, we write [a][a] for the element ([a],a,1)𝕎^n(A)([a],a,1)\in\widehat{\mathbb{W}}_{n}(A). We extend the natural transformations UU, \star, FsF_{s}, and VsV_{s} to 𝕎^n\widehat{\mathbb{W}}_{n} by setting

U(1)=[1],(α,a,m)(α,a,m)=(αα,amam,mm),\displaystyle U(1)=[1],\quad(\alpha,a,m)\star(\alpha^{\prime},a^{\prime},m^{\prime})=(\alpha\star\alpha^{\prime},a^{m^{\prime}}{a^{\prime}}^{m},mm^{\prime}),
Fs(α,a,m)=(Fs(α),as,m),Vs(α,a,m)=(Vs(α),a,sm).\displaystyle F_{s}(\alpha,a,m)=(F_{s}(\alpha),a^{s},m),\quad V_{s}(\alpha,a,m)=(V_{s}(\alpha),a,sm).
Lemma 5.1.

The following assertions hold for both 𝕎n\mathbb{W}_{n} and 𝕎^n\widehat{\mathbb{W}}_{n}:

  1. (1)

    \star is commutative, associative, and unital with unit UU.

  2. (2)

    F1=V1=idF_{1}=V_{1}={\operatorname{id}}, FsFr=FsrF_{s}F_{r}=F_{sr}, VsVr=VsrV_{s}V_{r}=V_{sr}.

  3. (3)

    FsVs=sidF_{s}V_{s}=s\cdot{\operatorname{id}}. If (s,r)=1(s,r)=1, then FsVr=VrFsF_{s}V_{r}=V_{r}F_{s}.

  4. (4)

    Vs(idFs)=(Vsid)V_{s}\circ\star\circ({\operatorname{id}}\otimes F_{s})=\star\circ(V_{s}\otimes{\operatorname{id}}).

  5. (5)

    (FsFs)=Fs\star\circ(F_{s}\otimes F_{s})=F_{s}\circ\star.

Proof.

The statement for 𝕎n\mathbb{W}_{n} is well-known (see e.g. [Hes15]), and the statement for 𝕎^n\widehat{\mathbb{W}}_{n} follows easily from this. ∎

Lemma 5.2.

For any n0n\geq 0, s1s\geq 1, and 0msn0\leq m\leq sn, the morphism VsFs:𝕎sn𝕎snV_{s}F_{s}\colon\mathbb{W}_{sn}\to\mathbb{W}_{sn} descends to 𝕎m𝕎m\mathbb{W}_{m}\to\mathbb{W}_{m}. The same also holds for 𝕎^n\widehat{\mathbb{W}}_{n}.

Proof.

Since s(m/s+1)m+1s(\lfloor m/s\rfloor+1)\geq m+1, the morphism Vs:𝕎m/s𝕎sm/sV_{s}\colon\mathbb{W}_{\lfloor m/s\rfloor}\to\mathbb{W}_{s\lfloor m/s\rfloor} lifts to 𝕎m/s𝕎m\mathbb{W}_{\lfloor m/s\rfloor}\to\mathbb{W}_{m}. Composing with 𝕎m𝕎sm/sFs𝕎m/s\mathbb{W}_{m}\twoheadrightarrow\mathbb{W}_{s\lfloor m/s\rfloor}\xrightarrow{F_{s}}\mathbb{W}_{\lfloor m/s\rfloor}, we get the desired morphism. ∎

Suppose that ch(k)=p>0\operatorname{ch}(k)=p>0. For a prime number \ell different from pp, we have an endomorphism 1VF\ell^{-1}V_{\ell}F_{\ell} of (𝕎^n)(p)=𝕎^n(p)(\widehat{\mathbb{W}}_{n})_{(p)}=\widehat{\mathbb{W}}_{n}\otimes\mathbb{Z}_{(p)}. By Lemma 5.1, 1VF\ell^{-1}V_{\ell}F_{\ell} is idempotent and hence defines a direct summand Im(1VF)\operatorname{Im}(\ell^{-1}V_{\ell}F_{\ell}) of (𝕎^n)(p)(\widehat{\mathbb{W}}_{n})_{(p)}. For n1n\geq 1, the presheaf of pp-typical Witt vectors of length nn is defined by

Wn:=(𝕎^pn1)(p)/pIm(1VF)PSh(Cork)W_{n}:=(\widehat{\mathbb{W}}_{p^{n-1}})_{(p)}/\textstyle\sum_{\ell\neq p}\operatorname{Im}(\ell^{-1}V_{\ell}F_{\ell})\in\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k})

where the sum is taken over all prine numbers different from pp.

Lemma 5.3.

The above definition of WnW_{n} coincides with the usual one (e.g. [Hes15]).

Proof.

Since 1VF(0,a,m)=(0,a,m)\ell^{-1}V_{\ell}F_{\ell}(0,a,m)=(0,a,m), the subpresheaf (0𝔾m)(p)(0\oplus\mathbb{G}_{m}\oplus\mathbb{Z})_{(p)} of (𝕎^pn1)(p)(\widehat{\mathbb{W}}_{p^{n-1}})_{(p)} is contained in Im(1VF)\operatorname{Im}(\ell^{-1}V_{\ell}F_{\ell}). Moreover, 𝕎pn1\mathbb{W}_{p^{n-1}} is a presheaf of (p)\mathbb{Z}_{(p)}-modules since ch(k)=p\operatorname{ch}(k)=p. Therefore we have

Wn=𝕎pn1/pIm(1VF).W_{n}=\mathbb{W}_{p^{n-1}}/\textstyle\sum_{\ell\neq p}\operatorname{Im}(\ell^{-1}V_{\ell}F_{\ell}).

Now the claim follows from the pp-typical decomposition of 𝕎pn1\mathbb{W}_{p^{n-1}} (see [Hes15, Proposition 1.10]). ∎

By Lemma 5.1, the homomorphisms UU, \star, FpF_{p}, and VpV_{p} descend to

U\displaystyle U :(p)Wn,\displaystyle\colon\mathbb{Z}_{(p)}\to W_{n}, \displaystyle\star :WnWnWn,\displaystyle\colon W_{n}\otimes W_{n}\to W_{n},
F\displaystyle F :Wn+1Wn,\displaystyle\colon W_{n+1}\to W_{n}, V\displaystyle V :WnWn+1\displaystyle\colon W_{n}\to W_{n+1}

and we have the following

Lemma 5.4.

The following assertions hold for WnW_{n}:

  1. (1)

    \star is commutative, associative, and unital with unit UU.

  2. (2)

    FV=pidFV=p\cdot{\operatorname{id}}.

  3. (3)

    V(idF)=(Vid)V\circ\star\circ({\operatorname{id}}\otimes F)=\star\circ(V\otimes{\operatorname{id}}).

  4. (4)

    (FF)=F\star\circ(F\otimes F)=F\circ\star.

5.2. Motivic construction

Definition 5.5.

For n0n\geq 0, we define 𝕎n+,𝕎^n+PSh(MCork)\mathbb{W}^{+}_{n},\widehat{\mathbb{W}}^{+}_{n}\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}) by

𝕎n+:=limε>0tr(1,(n+ε)[])/,𝕎^n+:=limε>0tr(1,ε[0]+(n+ε)[])\mathbb{W}^{+}_{n}:=\varinjlim_{\varepsilon>0}\mathbb{Z}_{\operatorname{tr}}(\mathbb{P}^{1},(n+\varepsilon)[\infty])/\mathbb{Z},\quad\widehat{\mathbb{W}}^{+}_{n}:=\varinjlim_{\varepsilon>0}\mathbb{Z}_{\operatorname{tr}}(\mathbb{P}^{1},\varepsilon[0]+(n+\varepsilon)[\infty])

where \mathbb{Z} is viewed as a direct summand of tr(1,(n+ε)[])\mathbb{Z}_{\operatorname{tr}}(\mathbb{P}^{1},(n+\varepsilon)[\infty]) via i0:Speck𝔸1i_{0}\colon\operatorname{Spec}k\to\mathbb{A}^{1}.

Definition 5.6.

Let n0n\geq 0. We define the unit, the multiplication, the Frobenius, and the Verschiebung on 𝕎n+\mathbb{W}^{+}_{n} as follows.

  1. (1)

    We define U:𝕎n+U\colon\mathbb{Z}\to\mathbb{W}^{+}_{n} to be the morphism induced by i1:Speck𝔸1i_{1}\colon\operatorname{Spec}k\to\mathbb{A}^{1}.

  2. (2)

    We define :𝕎n+𝕎n+𝕎n+\star\colon\mathbb{W}^{+}_{n}\otimes\mathbb{W}^{+}_{n}\to\mathbb{W}^{+}_{n} to be the morphism induced by the multiplication map μ:𝔸1×𝔸1𝔸1;(x,y)xy\mu\colon\mathbb{A}^{1}\times\mathbb{A}^{1}\to\mathbb{A}^{1};\;(x,y)\mapsto xy (see Lemma 3.2).

  3. (3)

    For s1s\geq 1, we define Fs:𝕎sn+𝕎n+F_{s}\colon\mathbb{W}^{+}_{sn}\to\mathbb{W}^{+}_{n} to be the morphism induced by ρs:𝔸1𝔸1;xxs\rho_{s}\colon\mathbb{A}^{1}\to\mathbb{A}^{1};\;x\mapsto x^{s}.

  4. (4)

    For s1s\geq 1, we define Vs:𝕎n+𝕎sn+V_{s}\colon\mathbb{W}^{+}_{n}\to\mathbb{W}^{+}_{sn} to be the morphism induced by ρst{}^{t}\rho_{s}.

Similarly, we define the unit, the multiplication, the Frobenius, and the Verschiebung on 𝕎^n+\widehat{\mathbb{W}}^{+}_{n}. For X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}} and aAa\in A (resp. aA×a\in A^{\times}), we write [a][a] for the element of h0𝕎n+(X)h_{0}\mathbb{W}^{+}_{n}(X) (resp. h0𝕎^n+(X)h_{0}\widehat{\mathbb{W}}^{+}_{n}(X)) represented by the morphism a:X𝔸1a\colon X\to\mathbb{A}^{1} (resp. a:X𝔸1{0}a\colon X\to\mathbb{A}^{1}\setminus\{0\}).

Lemma 5.7.

The following assertions hold for both 𝕎n+\mathbb{W}^{+}_{n} and 𝕎^n+\widehat{\mathbb{W}}^{+}_{n}:

  1. (1)

    \star is commutative, associative, and unital with unit UU.

  2. (2)

    F1=V1=idF_{1}=V_{1}={\operatorname{id}}, FsFr=FsrF_{s}F_{r}=F_{sr}, VsVr=VsrV_{s}V_{r}=V_{sr}.

  3. (3)

    FsVs=sidF_{s}V_{s}=s\cdot{\operatorname{id}}. If (s,r)=1(s,r)=1, then FsVr=VrFsF_{s}V_{r}=V_{r}F_{s}.

  4. (4)

    Vs(idFs)=(Vsid)V_{s}\circ\star\circ({\operatorname{id}}\otimes F_{s})=\star\circ(V_{s}\otimes{\operatorname{id}}).

  5. (5)

    (FsFs)=Fs\star\circ(F_{s}\otimes F_{s})=F_{s}\circ\star.

Proof.

(1) follows from the corresponding properties of μ:𝔸1×𝔸1𝔸1\mu\colon\mathbb{A}^{1}\times\mathbb{A}^{1}\to\mathbb{A}^{1}. (2) follows from ρ1=id\rho_{1}={\operatorname{id}} and ρrρs=ρrs\rho_{r}\circ\rho_{s}=\rho_{rs}. The first assertion in (3) follows from ρstρs=sid\rho_{s}{}^{t}\circ\rho_{s}=s\cdot{\operatorname{id}}. The second assertion follows from the fact that ρstρr\rho_{s}{}^{t}\circ\rho_{r} and ρrtρs{}^{t}\rho_{r}\circ\rho_{s} are both represented by the cycle on 𝔸2\mathbb{A}^{2} defined by xs=yrx^{s}=y^{r}. To prove (4), it suffices to show that the following diagram in Cork\operatorname{\mathrm{Cor}}_{k} is commutative:

𝔸1×𝔸1\textstyle{\mathbb{A}^{1}\times\mathbb{A}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρst×id\scriptstyle{{}^{t}\rho_{s}\times{\operatorname{id}}}id×ρs\scriptstyle{{\operatorname{id}}\times\rho_{s}}𝔸1×𝔸1\textstyle{\mathbb{A}^{1}\times\mathbb{A}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu}𝔸1×𝔸1\textstyle{\mathbb{A}^{1}\times\mathbb{A}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu}𝔸1\textstyle{\mathbb{A}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρst\scriptstyle{{}^{t}\rho_{s}}𝔸1.\textstyle{\mathbb{A}^{1}.}

One can easily check that both compositions are represented by the cycle on 𝔸3\mathbb{A}^{3} defined by xys=zsxy^{s}=z^{s}. Finally, (5) follows from ρsμ=μ(ρs×ρs)\rho_{s}\circ\mu=\mu\circ(\rho_{s}\times\rho_{s}). ∎

Lemma 5.8.

For any n0n\geq 0, s1s\geq 1 and 0msn0\leq m\leq sn, the morphism VsFs:𝕎sn+𝕎sn+V_{s}F_{s}\colon\mathbb{W}^{+}_{sn}\to\mathbb{W}^{+}_{sn} descends to 𝕎m+𝕎m+\mathbb{W}^{+}_{m}\to\mathbb{W}^{+}_{m}. The same also holds for 𝕎^n\widehat{\mathbb{W}}_{n}.

Proof.

To prove the first statement, it suffices to show that the finite correspondence ρstρsCork(𝔸1,𝔸1){}^{t}\rho_{s}\circ\rho_{s}\in\operatorname{\mathrm{Cor}}_{k}(\mathbb{A}^{1},\mathbb{A}^{1}) is contained in MCork((1,r[]),(1,r[]))\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}((\mathbb{P}^{1},r[\infty]),(\mathbb{P}^{1},r[\infty])) for every r>0r\in\mathbb{Q}_{>0}. This follows from the fact that ρstρs{}^{t}\rho_{s}\circ\rho_{s} is represented by the cycle on 𝔸2\mathbb{A}^{2} defined by xs=ysx^{s}=y^{s}, which is symmetric in xx and yy. The second statement can be proved similarly. ∎

Suppose that ch(k)=p>0\operatorname{ch}(k)=p>0. For a prime number \ell different from pp, we have an endomorphism 1VF\ell^{-1}V_{\ell}F_{\ell} of (𝕎^n+)(p)(\widehat{\mathbb{W}}^{+}_{n})_{(p)}. By Lemma 5.7, 1VF\ell^{-1}V_{\ell}F_{\ell} is an idempotent and hence defines a direct summand Im(1VF)\operatorname{Im}(\ell^{-1}V_{\ell}F_{\ell}) of (𝕎^n+)(p)(\widehat{\mathbb{W}}^{+}_{n})_{(p)}. For n1n\geq 1, we define

(5.1) Wn+:=(𝕎^pn1+)(p)/pIm(1VF)PSh(MCork).W^{+}_{n}:=(\widehat{\mathbb{W}}^{+}_{p^{n-1}})_{(p)}/\textstyle\sum_{\ell\neq p}\operatorname{Im}(\ell^{-1}V_{\ell}F_{\ell})\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}).

For X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}} and aA×a\in A^{\times}, we write [a][a] for the image of [a]h0𝕎^pn1+(X)[a]\in h_{0}\widehat{\mathbb{W}}^{+}_{p^{n-1}}(X) in h0Wn+(X)h_{0}W^{+}_{n}(X). By Lemma 5.7, the morphisms UU, \star, FpF_{p}, and VpV_{p} descend to

U\displaystyle U :(p)Wn+,\displaystyle\colon\mathbb{Z}_{(p)}\to W^{+}_{n}, \displaystyle\star :Wn+Wn+Wn+,\displaystyle\colon W^{+}_{n}\otimes W^{+}_{n}\to W^{+}_{n},
F\displaystyle F :Wn+1+Wn+,\displaystyle\colon W^{+}_{n+1}\to W^{+}_{n}, V\displaystyle V :Wn+Wn+1+\displaystyle\colon W^{+}_{n}\to W^{+}_{n+1}

and we have the following

Lemma 5.9.

The following assertions hold for Wn+W^{+}_{n}:

  1. (1)

    \star is commutative, associative, and unital with unit UU.

  2. (2)

    FV=pidFV=p\cdot{\operatorname{id}}.

  3. (3)

    V(idF)=(Vid)V\circ({\operatorname{id}}\otimes F)=\star\circ(V\otimes{\operatorname{id}}).

  4. (4)

    (FF)=F\star\circ(F\otimes F)=F\circ\star.

5.3. Comparison

Theorem 5.10.

Let n0n\geq 0.

  1. (1)

    There is an isomorphism φ:h0𝕎n+𝕎n\varphi\colon h_{0}\mathbb{W}^{+}_{n}\xrightarrow{\sim}\mathbb{W}_{n} in PSh(Corkaff)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}^{\mathrm{aff}}). For X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}, the image of div(xm+a1xm1++am)h0𝕎n+(X)\operatorname{div}(x^{m}+a_{1}x^{m-1}+\dots+a_{m})\in h_{0}\mathbb{W}^{+}_{n}(X) is 1+a1t++amtm𝕎n(X)1+a_{1}t+\dots+a_{m}t^{m}\in\mathbb{W}_{n}(X). In particular, we have φX([a])=[a]\varphi_{X}([a])=[a] for aAa\in A.

  2. (2)

    There is an isomorphism φ^:h0𝕎^n+𝕎^n\widehat{\varphi}\colon h_{0}\widehat{\mathbb{W}}^{+}_{n}\xrightarrow{\sim}\widehat{\mathbb{W}}_{n} in PSh(Corkaff)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}^{\mathrm{aff}}). For X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}, the image of div(xm+a1xm1++am)h0𝕎^n+(X)\operatorname{div}(x^{m}+a_{1}x^{m-1}+\dots+a_{m})\in h_{0}\widehat{\mathbb{W}}^{+}_{n}(X) is (1+a1t++amtm,(1)mam,m)𝕎^n(X)(1+a_{1}t+\dots+a_{m}t^{m},(-1)^{m}a_{m},m)\in\widehat{\mathbb{W}}_{n}(X). In particular, we have φ^X([a])=[a]\widehat{\varphi}_{X}([a])=[a] for aA×a\in A^{\times}.

Proof.

We prove only (1); the proof for (2) is similar. First we construct an isomorphism

φX:h0𝕎n+(X)𝕎n(X)\varphi_{X}\colon h_{0}\mathbb{W}^{+}_{n}(X)\xrightarrow{\sim}\mathbb{W}_{n}(X)

for X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}. We may assume that XX is connected. For ε(0,1)\varepsilon\in(0,1)\cap\mathbb{Q} we have

h0(1,(n+ε)[])(X)\displaystyle h_{0}(\mathbb{P}^{1},(n+\varepsilon)[\infty])(X) h0(1,(n+1)[])(X)\displaystyle\simeq h_{0}(\mathbb{P}^{1},(n+1)[\infty])(X) (Theorem 4.14)
Pic(X1,(n+1)[]).\displaystyle\simeq\operatorname{Pic}(\mathbb{P}^{1}_{X},(n+1)[\infty]). (Theorem 4.13)

The last term can be easily identified with 𝕎n(X)\mathbb{W}_{n}(X)\oplus\mathbb{Z} (see e.g. [Koi, Proposition 1.1]). Therefore we have h0𝕎n+(X)(𝕎n(X))/𝕎n(X)h_{0}\mathbb{W}^{+}_{n}(X)\simeq(\mathbb{W}_{n}(X)\oplus\mathbb{Z})/\mathbb{Z}\simeq\mathbb{W}_{n}(X).

It remains to show that this isomorphism is compatible with transfers. Let X,YSmkaffX,Y\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}} and αCork(X,Y)\alpha\in\operatorname{\mathrm{Cor}}_{k}(X,Y). It suffices to prove that the following diagram is commutative:

(5.6)

We may assume that X,YX,Y are connected. Let KK be an algebraic closure of k(X)k(X), and consider the same problem for XK,YKSmkaffX_{K},Y_{K}\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}} and αKCorK(XK,YK)\alpha_{K}\in\operatorname{\mathrm{Cor}}_{K}(X_{K},Y_{K}):

(5.11)

Then there is a natural morphism of diagrams from (5.6) to (5.11). Since the composite

𝕎n(X)𝕎(XK)ΔK𝕎n(K)\mathbb{W}_{n}(X)\to\mathbb{W}(X_{K})\xrightarrow{\Delta_{K}^{*}}\mathbb{W}_{n}(K)

is injective, we may assume that kk is algebraically closed and X=SpeckX=\operatorname{Spec}k. In this case, α\alpha can be written as a \mathbb{Z}-linear combination of morphisms SpeckY\operatorname{Spec}k\to Y, so the assertion is obvious. ∎

Proposition 5.11.

The isomorphisms φ:h0𝕎n+𝕎n\varphi\colon h_{0}\mathbb{W}^{+}_{n}\xrightarrow{\sim}\mathbb{W}_{n} and φ^:h0𝕎^n+𝕎^n\widehat{\varphi}\colon h_{0}\widehat{\mathbb{W}}^{+}_{n}\xrightarrow{\sim}\widehat{\mathbb{W}}_{n} in PSh(Corkaff)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}^{\mathrm{aff}}) from Theorem 5.10 are compatible with the unit, the multiplication, the Frobenius, and the Verschiebung.

Proof.

We prove only the statement for φ\varphi; the proof for φ^\widehat{\varphi} is similar. It suffices to prove that for each XSmkaffX\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}, the isomorphism φX:h0𝕎n+(X)𝕎n(X)\varphi_{X}\colon h_{0}\mathbb{W}^{+}_{n}(X)\xrightarrow{\sim}\mathbb{W}_{n}(X) is compatible with these operations. As in the proof of Theorem 5.10, we may assume that kk is algebraically closed and X=SpeckX=\operatorname{Spec}k. In this case, 𝕎n(k)\mathbb{W}_{n}(k) is generated by elements of the form [a][a] with ak×a\in k^{\times}. For a,bk×a,b\in k^{\times}, we have

U(1)=[1],[a][b]=[ab],Fs([a])=[as],Vs([a])=1atsU(1)=[1],\quad[a]\star[b]=[ab],\quad F_{s}([a])=[a^{s}],\quad V_{s}([a])=1-at^{s}

in 𝕎n(k)\mathbb{W}_{n}(k) and

U(1)=[1],[a][b]=[ab],Fs([a])=[as],Vs([a])=div(xsa)U(1)=[1],\quad[a]\star[b]=[ab],\quad F_{s}([a])=[a^{s}],\quad V_{s}([a])=\operatorname{div}(x^{s}-a)

in h0𝕎n+(k)h_{0}\mathbb{W}^{+}_{n}(k). This proves the claim. ∎

Corollary 5.12.

Suppose that ch(k)=p>0\operatorname{ch}(k)=p>0. For n1n\geq 1, there is an isomorphism

φ(p):h0Wn+Wn\varphi^{(p)}\colon h_{0}W^{+}_{n}\xrightarrow{\sim}W_{n}

in PSh(Corkaff)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}^{\mathrm{aff}}) which is compatible with the unit, the multiplication, the Frobenius, and the Verschiebung. If X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}, then we have φX(p)([a])=[a]\varphi^{(p)}_{X}([a])=[a] for aA×a\in A^{\times}.

5.4. Application to torsion and divisibility of reciprocity sheaves

In this subsection, we give an application of the motivic presentation of the ring of big Witt vectors to reciprocity sheaves.

Let FRSCkF\in\operatorname{\mathrm{RSC}}_{k}. Imitating the construction in [Miy19], we define NnF,NFPSh(Cork)N_{n}F,NF\in\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) by

NnF(X)\displaystyle N_{n}F(X) :=limε>0Ker(h0F((X,)(1,(n+ε)[]))i0F(X)),\displaystyle:=\varprojlim_{\varepsilon>0}\operatorname{Ker}(h^{0}F((X,\emptyset)\otimes(\mathbb{P}^{1},(n+\varepsilon)[\infty]))\xrightarrow{i_{0}^{*}}F(X)),
NF(X)\displaystyle NF(X) :=Ker(F(X×𝔸1)i0F(X)).\displaystyle:=\operatorname{Ker}(F(X\times\mathbb{A}^{1})\xrightarrow{i_{0}^{*}}F(X)).

Here, we regard (X,)(X,\emptyset) as a pro-object in MCork\operatorname{\mathrm{MCor}}_{k}^{\mathbb{Q}} via the compactification functor. In other words, we define NnF:=h0om(𝕎n+,h0F)N_{n}F:=h_{0}\mathcal{H}\mathrm{om}(\mathbb{W}^{+}_{n},h^{0}F) and NF:=limn0NnFNF:=\varinjlim_{n\geq 0}N_{n}F. The presheaf NFNF measures how far FF is from being 𝔸1\mathbb{A}^{1}-invariant. The composition

𝕎n+𝕎n+om(𝕎n+,h0F)id𝕎n+om(𝕎n+,h0F)evh0F\mathbb{W}^{+}_{n}\otimes\mathbb{W}^{+}_{n}\otimes\mathcal{H}\mathrm{om}(\mathbb{W}^{+}_{n},h^{0}F)\xrightarrow{\star\otimes{\operatorname{id}}}\mathbb{W}^{+}_{n}\otimes\mathcal{H}\mathrm{om}(\mathbb{W}^{+}_{n},h^{0}F)\xrightarrow{\mathrm{ev}}h^{0}F

induces an action

𝕎n+om(𝕎n+,h0F)om(𝕎n+,h0F)\mathbb{W}^{+}_{n}\otimes\mathcal{H}\mathrm{om}(\mathbb{W}^{+}_{n},h^{0}F)\to\mathcal{H}\mathrm{om}(\mathbb{W}^{+}_{n},h^{0}F)

by adjunction. Taking h0h_{0}, we obtain an action h0𝕎n+NnFNnFh_{0}\mathbb{W}^{+}_{n}\otimes N_{n}F\to N_{n}F. Since we have an isomorphism of rings h0𝕎n+(X)𝕎n(X)h_{0}\mathbb{W}^{+}_{n}(X)\simeq\mathbb{W}_{n}(X) for XSmkaffX\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}} by Theorem 5.10 and Proposition 5.11, we obtain the following

Theorem 5.13.

Let FRSCkF\in\operatorname{\mathrm{RSC}}_{k} and XSmkaffX\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}. Then NnF(X)N_{n}F(X) has a canonical structure of a 𝕎n(X)\mathbb{W}_{n}(X)-module, which is natural in XX and nn. In particular, NF(X)NF(X) has a canonical structure of a 𝕎(X)\mathbb{W}(X)-module, where 𝕎(X)=limn0𝕎n(X)\mathbb{W}(X)=\varprojlim_{n\geq 0}\mathbb{W}_{n}(X).

Corollary 5.14.

Let FRSCkF\in\operatorname{\mathrm{RSC}}_{k}, XSmkaffX\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}, and let pp be a prime number.

  1. (1)

    If pp is invertible in kk, then NnF(X)N_{n}F(X) and NF(X)NF(X) are uniquely pp-divisible.

  2. (2)

    If ch(k)=p>0\operatorname{ch}(k)=p>0, then NnF(X)N_{n}F(X) and NF(X)NF(X) are pp-groups.

Proof.

This follows from the fact that if pp is invertible (resp. nilpotent) in a ring AA, then pp is also invertible (resp. nilpotent) in 𝕎n(A)\mathbb{W}_{n}(A); see [Hes15, Lemma 1.9 and Proposition 1.10]. ∎

Corollary 5.15 ([Bin+17, Theorem 1.3]).

Let FRSCkF\in\operatorname{\mathrm{RSC}}_{k} and assume that FF is separated for the Zariski topology.

  1. (1)

    If ch(k)=0\operatorname{ch}(k)=0 and F=0F\otimes\mathbb{Q}=0, then FF is 𝔸1\mathbb{A}^{1}-invariant.

  2. (2)

    If ch(k)=p>0\operatorname{ch}(k)=p>0 and FF is pp-torsion-free, then FF is 𝔸1\mathbb{A}^{1}-invariant.

6. Motivic construction of the de Rham-Witt complex

Throughout this section, we assume that kk is perfect and ch(k)=p3\operatorname{ch}(k)=p\geq 3 222 Here, the perfectness of kk is assumed in order to use the transfer structure on the de Rham-Witt complex. The assumption on ch(k)\operatorname{ch}(k) will be used in the proofs of Lemma 6.10 and Theorem 6.11. . In this section, we construct the de Rham-Witt complex of smooth kk-schemes using \mathbb{Q}-modulus pairs.

6.1. De Rham-Witt complex

First we recall the definition of the de Rham-Witt complex. Here, we follow the axiomatization due to [HM04].

Definition 6.1.

Let AA be a (p)\mathbb{Z}_{(p)}-algebra. A Witt complex over AA is a tuple (E,F,V,λ)(E_{\bullet}^{*},F,V,\lambda) where

  1. (1)

    En𝑅En1E0\cdots\to E_{n}^{*}\xrightarrow{R}E_{n-1}^{*}\to\cdots\to E_{0}^{*} is a sequence of CDGAs,

  2. (2)

    F:En+1EnF\colon E_{n+1}^{*}\to E_{n}^{*} is a graded ring homomorphism compatible with RR,

  3. (3)

    V:EnEn+1V\colon E_{n}^{*}\to E_{n+1}^{*} is a graded group homomorphism compatible with RR,

  4. (4)

    λ:Wn(A)En0\lambda\colon W_{n}(A)\to E_{n}^{0} is a ring homomorphism compatible with RR, FF, and VV,

such that the following relations hold:

V(xF(y))=V(x)y,FdV=d,FV=pid,F(dλ[a])=λ[ap1]dλ[a](aA).V(x\star F(y))=V(x)\star y,\quad FdV=d,\quad FV=p\cdot{\operatorname{id}},\quad F(d\lambda[a])=\lambda[a^{p-1}]\star d\lambda[a]\;(a\in A).

The category of Witt complexes over AA is known to have an initial object WnΩ(A)W_{n}\Omega^{*}(A) and it is called the de Rham-Witt complex of AA. We have WnΩ0(A)Wn(A)W_{n}\Omega^{0}(A)\simeq W_{n}(A). It follows from the construction that WnΩq(A)W_{n}\Omega^{q}(A) is a quotient of ΩWn(A)q\Omega^{q}_{W_{n}(A)}.

The presheaf SpecAWnΩq(A)\operatorname{Spec}A\mapsto W_{n}\Omega^{q}(A) on Smkaff\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}} extends to an étale sheaf WnΩqW_{n}\Omega^{q} on Smk\operatorname{\mathrm{Sm}}_{k} having global injectivity 333 This means that for any XSmkX\in\operatorname{\mathrm{Sm}}_{k} and a dense open subset UXU\subset X, the map WnΩq(X)WnΩq(U)W_{n}\Omega^{q}(X)\to W_{n}\Omega^{q}(U) is injective. This follows from [Ill79, I, Corollaire 3.9]. . Moreover, since kk is assumed to be perfect, WnΩqW_{n}\Omega^{q} can be regarded as an object of PSh(Cork)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}) [KSY16, Theorem B.2.1]. It is compatible with the trace maps defined in [R0̈7a, Theorem 2.6] (see [RS21, Section 7.9]).

6.2. Motivic construction

Definition 6.2.

We define 𝔾m+PSh(MCork)\mathbb{G}_{m}^{+}\in\operatorname{PSh}(\operatorname{\mathrm{MCor}}^{\mathbb{Q}}_{k}) by

𝔾m+:=limε>0tr(1,ε[0]+ε[])/\mathbb{G}_{m}^{+}:=\varinjlim_{\varepsilon>0}\mathbb{Z}_{\operatorname{tr}}(\mathbb{P}^{1},\varepsilon[0]+\varepsilon[\infty])/\mathbb{Z}

where \mathbb{Z} is viewed as a direct summand of tr(1,ε[0]+ε[])\mathbb{Z}_{\operatorname{tr}}(\mathbb{P}^{1},\varepsilon[0]+\varepsilon[\infty]) via i1:Speck𝔸1{0}i_{1}\colon\operatorname{Spec}k\to\mathbb{A}^{1}\setminus\{0\}.

Lemma 6.3.

The following assertions hold.

  1. (1)

    Let Fs:𝔾m+𝔾m+F_{s}\colon\mathbb{G}_{m}^{+}\to\mathbb{G}_{m}^{+} be the morphism induced by ρs\rho_{s}. Then we have h0(Fs)=sidh_{0}(F_{s})=s\cdot{\operatorname{id}}.

  2. (2)

    Let Vs:𝔾m+𝔾m+V_{s}\colon\mathbb{G}_{m}^{+}\to\mathbb{G}_{m}^{+} be the morphism induced by ρst{}^{t}\rho_{s}. If ss is odd, then h0(Vs)=idh_{0}(V_{s})={\operatorname{id}}.

  3. (3)

    For X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}} and aA×a\in A^{\times}, we write [a][a] for the element of h0(𝔾m+)(X)h_{0}(\mathbb{G}_{m}^{+})(X) represented by the morphism a:X𝔸1{0}a\colon X\to\mathbb{A}^{1}\setminus\{0\}. Then we have [a]+[b]=[ab][a]+[b]=[ab].

  4. (4)

    Let τ:𝔾m+2𝔾m+2\tau\colon\mathbb{G}_{m}^{+\otimes 2}\to\mathbb{G}_{m}^{+\otimes 2} be the morphism induced by (𝔸1{0})2(𝔸1{0})2;(x,y)(y,x).(\mathbb{A}^{1}\setminus\{0\})^{2}\to(\mathbb{A}^{1}\setminus\{0\})^{2};\;(x,y)\mapsto(y,x). Then we have h0(τ)=idh_{0}(\tau)=-{\operatorname{id}}.

  5. (5)

    Let δ:𝔾m+𝔾m+2\delta\colon\mathbb{G}_{m}^{+}\to\mathbb{G}_{m}^{+\otimes 2} be the morphism induced by 𝔸1{0}(𝔸1{0})2;x(x,x).\mathbb{A}^{1}\setminus\{0\}\to(\mathbb{A}^{1}\setminus\{0\})^{2};\;x\mapsto(x,x). Then we have 2h0(δ)=02\cdot h_{0}(\delta)=0.

Proof.

(1) Define Γ(𝔸1{0})×𝔸1×(𝔸1{0})=Speck[x±,t,y±]\Gamma\subset(\mathbb{A}^{1}\setminus\{0\})\times\mathbb{A}^{1}\times(\mathbb{A}^{1}\setminus\{0\})=\operatorname{Spec}k[x^{\pm},t,y^{\pm}] by the equation

(1t)(yxs)(y1)s1+t(yx)s=0.(1-t)(y-x^{s})(y-1)^{s-1}+t(y-x)^{s}=0.

Regarding the left hand side as a polynomial in yy, the leading term and the constant term have invertible coefficients. Therefore Γ\Gamma is finite locally free over (𝔸1{0})×𝔸1(\mathbb{A}^{1}\setminus\{0\})\times\mathbb{A}^{1} and hence defines a finite correspondence [Γ]Cork((𝔸1{0})×𝔸1,𝔸1{0})[\Gamma]\in\operatorname{\mathrm{Cor}}_{k}((\mathbb{A}^{1}\setminus\{0\})\times\mathbb{A}^{1},\mathbb{A}^{1}\setminus\{0\}). The closure Γ¯1×1×1\overline{\Gamma}\subset\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} of Γ\Gamma satisfies the following condition for ε>0\varepsilon\in\mathbb{Q}_{>0}:

pr1(sε[0]+sε[])|Γ¯N+pr2[]|Γ¯Npr3(ε[0]+ε[])|Γ¯N.{\operatorname{pr}}_{1}^{*}(s\varepsilon[0]+s\varepsilon[\infty])|_{\overline{\Gamma}^{N}}+{\operatorname{pr}}_{2}^{*}[\infty]|_{\overline{\Gamma}^{N}}\geq{\operatorname{pr}}_{3}^{*}(\varepsilon[0]+\varepsilon[\infty])|_{\overline{\Gamma}^{N}}.

Therefore Γ\Gamma induces a morphism 𝔾m+tr(¯)𝔾m+\mathbb{G}_{m}^{+}\otimes\mathbb{Z}_{\operatorname{tr}}({\overline{\square}})\to\mathbb{G}_{m}^{+} whose restriction to t=0t=0 (resp. t=1t=1) is FsF_{s} (resp. sids\cdot{\operatorname{id}}). This shows that h0(Fs)=sidh_{0}(F_{s})=s\cdot{\operatorname{id}}.

(2) Define Γ(𝔸1{0})×𝔸1×(𝔸1{0})=Speck[x±,t,y±]\Gamma\subset(\mathbb{A}^{1}\setminus\{0\})\times\mathbb{A}^{1}\times(\mathbb{A}^{1}\setminus\{0\})=\operatorname{Spec}k[x^{\pm},t,y^{\pm}] by the equation

(1t)(ysx)+t(yx)(y1)s1=0.(1-t)(y^{s}-x)+t(y-x)(y-1)^{s-1}=0.

Regarding the left hand side as a polynomial in yy, the leading term and the constant term have invertible coefficients if ss is odd. Therefore Γ\Gamma is finite locally free over (𝔸1{0})×𝔸1(\mathbb{A}^{1}\setminus\{0\})\times\mathbb{A}^{1} and hence defines a finite correspondence [Γ]Cork((𝔸1{0})×𝔸1,𝔸1{0})[\Gamma]\in\operatorname{\mathrm{Cor}}_{k}((\mathbb{A}^{1}\setminus\{0\})\times\mathbb{A}^{1},\mathbb{A}^{1}\setminus\{0\}). As in the proof of (1), induces a morphism 𝔾m+tr(¯)𝔾m+\mathbb{G}_{m}^{+}\otimes\mathbb{Z}_{\operatorname{tr}}({\overline{\square}})\to\mathbb{G}_{m}^{+} whose restriction to t=0t=0 (resp. t=1t=1) is VsV_{s} (resp. id{\operatorname{id}}). This shows that h0(Vs)=idh_{0}(V_{s})={\operatorname{id}}.

(3) Define ΓX×𝔸1×(𝔸1{0})=SpecA[t,x±]\Gamma\subset X\times\mathbb{A}^{1}\times(\mathbb{A}^{1}\setminus\{0\})=\operatorname{Spec}A[t,x^{\pm}] by the equation

(1t)(xa)(xb)+t(xab)(x1)=0.(1-t)(x-a)(x-b)+t(x-ab)(x-1)=0.

Regarding the left hand side as a polynomial in xx, the leading term and the constant term have invertible coefficients. Therefore Γ\Gamma is finite locally free over SpecA[t]\operatorname{Spec}A[t] and hence defines a finite correspondence [Γ]Cork(X×𝔸1,𝔸1{0})[\Gamma]\in\operatorname{\mathrm{Cor}}_{k}(X\times\mathbb{A}^{1},\mathbb{A}^{1}\setminus\{0\}). This induces a morphism tr(X)tr(¯)𝔾m+\mathbb{Z}_{\operatorname{tr}}(X)\otimes\mathbb{Z}_{\operatorname{tr}}({\overline{\square}})\to\mathbb{G}_{m}^{+} whose restriction to t=0t=0 (resp. t=1t=1) is [a]+[b][a]+[b] (resp. [ab][ab]). This shows that [a]+[b]=[ab][a]+[b]=[ab].

(4) Define Γ(𝔸1{0})2×𝔸1×(𝔸1{0})2=Speck[x±,y±,t,z±,w±]\Gamma\subset(\mathbb{A}^{1}\setminus\{0\})^{2}\times\mathbb{A}^{1}\times(\mathbb{A}^{1}\setminus\{0\})^{2}=\operatorname{Spec}k[x^{\pm},y^{\pm},t,z^{\pm},w^{\pm}] by the equations

z+w=(1t)(x+y)+t(xy+1),zw=xy.z+w=(1-t)(x+y)+t(xy+1),\quad zw=xy.

One can check that Γ\Gamma is finite locally free over (𝔸1{0})2×𝔸1(\mathbb{A}^{1}\setminus\{0\})^{2}\times\mathbb{A}^{1} and hence defines a finite correspondence [Γ]Cork((𝔸1{0})2×𝔸1,(𝔸1{0})2)[\Gamma]\in\operatorname{\mathrm{Cor}}_{k}((\mathbb{A}^{1}\setminus\{0\})^{2}\times\mathbb{A}^{1},(\mathbb{A}^{1}\setminus\{0\})^{2}). This induces a morphism 𝔾m+2tr(¯)𝔾m+2\mathbb{G}_{m}^{+\otimes 2}\otimes\mathbb{Z}_{\operatorname{tr}}({\overline{\square}})\to\mathbb{G}_{m}^{+\otimes 2} whose restriction to t=0t=0 (resp. t=1t=1) is id+τ{\operatorname{id}}+\tau (resp. 0). This shows that h0(id+τ)=0h_{0}({\operatorname{id}}+\tau)=0 and hence h0(τ)=idh_{0}(\tau)=-{\operatorname{id}}.

(5) We have τδ=δ\tau\circ\delta=\delta. Since h0(τ)=idh_{0}(\tau)=-{\operatorname{id}} by (4), we get h0(δ)=h0(δ)-h_{0}(\delta)=h_{0}(\delta) and hence 2h0(δ)=02\cdot h_{0}(\delta)=0. ∎

Remark 6.4.

Actually, we can prove that h0(𝔾m+q)KqMh_{0}(\mathbb{G}_{m}^{+\otimes q})\cong K_{q}^{M} holds in PSh(Cork)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}), where KqMK_{q}^{M} is the unramified sheaf of Milnor KK-groups; it is essentially a corollary of [SV00, Theorem 3.4].

Definition 6.5.

Let n1n\geq 1.

  1. (1)

    For q,r0q,r\geq 0, we define the multiplication on Wn+𝔾m+W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes*} by

    :(Wn+𝔾m+q)(Wn+𝔾m+r)\displaystyle\star\colon(W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes q})\otimes(W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes r}) Wn+Wn+𝔾m+q𝔾m+r\displaystyle\xrightarrow{\sim}W^{+}_{n}\otimes W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes q}\otimes\mathbb{G}_{m}^{+\otimes r}
    idWn+𝔾m+(q+r).\displaystyle\xrightarrow{\star\otimes{\operatorname{id}}}W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes(q+r)}.

    For each XSmkaffX\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}, this makes h0(Wn+𝔾m+)(X)h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes*})(X) into a graded ring. By Lemma 6.3 (4), this multiplication is graded commutative. Moreover, it is a (p)\mathbb{Z}_{(p)}-algebra since h0Wn+(X)Wn(X)h_{0}W_{n}^{+}(X)\simeq W_{n}(X) is so.

  2. (2)

    For q0q\geq 0, we define the Frobenius by F:=Fid:Wn+1+𝔾m+Wn+𝔾m+F:=F\otimes{\operatorname{id}}\colon W^{+}_{n+1}\otimes\mathbb{G}_{m}^{+\otimes*}\to W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes*} and the Verschiebung by V:=Vid:Wn+𝔾m+Wn+1+𝔾m+V:=V\otimes{\operatorname{id}}\colon W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes*}\to W^{+}_{n+1}\otimes\mathbb{G}_{m}^{+\otimes*}. By Lemma 5.9, FF is a graded ring homomorphism, and we have V(xF(y))=V(x)yV(x\star F(y))=V(x)\star y, FV=pFV=p.

  3. (3)

    We define λ:Wnh0Wn+\lambda\colon W_{n}\xrightarrow{\sim}h_{0}W^{+}_{n} in PSh(Corkaff)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}^{\mathrm{aff}}) to be the inverse of the isomorphism φ(p)\varphi^{(p)} given in Corollary 5.12. This is a ring homomorphism compatible with FF and VV.

Definition 6.6.

Let Γ\Gamma be the graph of the diagonal morphism Δ:𝔸1{0}(𝔸1{0})2\Delta\colon\mathbb{A}^{1}\setminus\{0\}\to(\mathbb{A}^{1}\setminus\{0\})^{2}. Then the closure Γ¯1×1×1\overline{\Gamma}\subset\mathbb{P}^{1}\times\mathbb{P}^{1}\times\mathbb{P}^{1} of Γ\Gamma satisfies the following condition for ε>0\varepsilon\in\mathbb{Q}_{>0}:

pr1(2ε[0]+(n+2ε)[])|Γ¯Npr2(ε[0]+(n+ε)[])|Γ¯N+pr3(ε[0]+ε[])|Γ¯N.{\operatorname{pr}}_{1}^{*}(2\varepsilon[0]+(n+2\varepsilon)[\infty])|_{\overline{\Gamma}^{N}}\geq{\operatorname{pr}}_{2}^{*}(\varepsilon[0]+(n+\varepsilon)[\infty])|_{\overline{\Gamma}^{N}}+{\operatorname{pr}}_{3}^{*}(\varepsilon[0]+\varepsilon[\infty])|_{\overline{\Gamma}^{N}}.

We define d:𝕎^n+𝕎^n+𝔾m+d\colon\widehat{\mathbb{W}}^{+}_{n}\to\widehat{\mathbb{W}}^{+}_{n}\otimes\mathbb{G}_{m}^{+} to be the morphism induced by Γ\Gamma.

Lemma 6.7.

For any prime number \ell different from pp, the following diagram becomes commutative after applying h0h_{0}:

𝕎^pn1+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}VF\scriptstyle{\ell V_{\ell}F_{\ell}}𝕎^pn1+𝔾m+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}VFid\scriptstyle{\ell V_{\ell}F_{\ell}\otimes{\operatorname{id}}}𝕎^pn1+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}𝕎^pn1+𝔾m+.\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}.}
Proof.

Recall from Remark 3.14 that we have a canonical epimorphism h0(𝕎^pn1+)h0(𝔾m+)h0(𝕎^pn1+𝔾m+)h_{0}(\widehat{\mathbb{W}}^{+}_{p^{n-1}})\otimes h_{0}(\mathbb{G}_{m}^{+})\twoheadrightarrow h_{0}(\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}). By Lemma 6.3 (1), we have h0(idF)=idh_{0}({\operatorname{id}}\otimes F_{\ell})=\ell\cdot{\operatorname{id}} on h0(𝕎^pn1+𝔾m+)h_{0}(\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}). Therefore it suffices to show that the following diagram is commutative:

𝕎^pn1+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}VF\scriptstyle{V_{\ell}F_{\ell}}𝕎^pn1+𝔾m+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idF\scriptstyle{{\operatorname{id}}\otimes F_{\ell}}𝕎^pn1+𝔾m+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}VFid\scriptstyle{V_{\ell}F_{\ell}\otimes{\operatorname{id}}}𝕎^pn1+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}𝕎^pn1+𝔾m+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idF\scriptstyle{{\operatorname{id}}\otimes F_{\ell}}𝕎^pn1+𝔾m+.\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}.}

This is further reduced to the commutativity of the following diagram in Cork\operatorname{\mathrm{Cor}}_{k}:

𝔸1{0}\textstyle{\mathbb{A}^{1}\setminus\{0\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ\scriptstyle{\Delta}ρtρ\scriptstyle{{}^{t}\rho_{\ell}\circ\rho_{\ell}}(𝔸1{0})2\textstyle{(\mathbb{A}^{1}\setminus\{0\})^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id×ρ\scriptstyle{{\operatorname{id}}\times\rho_{\ell}}(𝔸1{0})2\textstyle{(\mathbb{A}^{1}\setminus\{0\})^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(ρtρ)×id\scriptstyle{({}^{t}\rho_{\ell}\circ\rho_{\ell})\times{\operatorname{id}}}𝔸1{0}\textstyle{\mathbb{A}^{1}\setminus\{0\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ\scriptstyle{\Delta}(𝔸1{0})2\textstyle{(\mathbb{A}^{1}\setminus\{0\})^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id×ρ\scriptstyle{{\operatorname{id}}\times\rho_{\ell}}(𝔸1{0})2.\textstyle{(\mathbb{A}^{1}\setminus\{0\})^{2}.}

Both compositions are given by the cycle on (𝔸1{0})3(\mathbb{A}^{1}\setminus\{0\})^{3} defined by x=y=zx^{\ell}=y^{\ell}=z. ∎

Definition 6.8.

Let n1n\geq 1. We define the differential

d:h0(Wn+𝔾m+q)h0(Wn+𝔾m+(q+1))d\colon h_{0}(W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes q})\to h_{0}(W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes(q+1)})

to be the morphism induced by did:𝕎^pn1+𝔾m+q𝕎^pn1+𝔾m+(q+1)d\otimes{\operatorname{id}}\colon\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+\otimes q}\to\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+\otimes(q+1)} via Lemma 6.7.

Lemma 6.9.

The following diagram becomes commutative after applying h0h_{0}:

𝕎^pn1+2\textstyle{\widehat{\mathbb{W}}_{p^{n-1}}^{+\otimes 2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\star}did+idd\scriptstyle{d\otimes{\operatorname{id}}+{\operatorname{id}}\otimes d}𝕎^pn1+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}𝕎^pn1+2𝔾m+\textstyle{\widehat{\mathbb{W}}_{p^{n-1}}^{+\otimes 2}\otimes\mathbb{G}_{m}^{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}id\scriptstyle{\star\otimes{\operatorname{id}}}𝕎^pn1+𝔾m+.\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}.}
Proof.

It suffices to show that the following diagram becomes commutative after composing with the canonical epimorphism tr(𝔸1{0})2h0(𝕎^pn1+𝔾m+)\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 2}\to h_{0}(\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}):

tr(𝔸1{0})2\textstyle{\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μ\scriptstyle{\mu}q1+q2\scriptstyle{q_{1}+q_{2}}tr(𝔸1{0})\textstyle{\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ\scriptstyle{\Delta}tr(𝔸1{0})3\textstyle{\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 3}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μid\scriptstyle{\mu\otimes{\operatorname{id}}}tr(𝔸1{0})2.\textstyle{\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 2}.}

Here, q1q_{1} (resp. q2q_{2}) denotes the morphism (x,y)(x,y,x)(x,y)\mapsto(x,y,x) (resp. (x,y)(x,y,y)(x,y)\mapsto(x,y,y)). The two compositions are identified with

tr(𝔸1{0})2𝛽tr(𝔸1{0})4μμtr(𝔸1{0})2,\displaystyle\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 2}\xrightarrow{\beta}\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 4}\xrightarrow{\mu\otimes\mu}\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 2},
tr(𝔸1{0})2𝛽tr(𝔸1{0})4μ(pr1+pr2)tr(𝔸1{0})2\displaystyle\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 2}\xrightarrow{\beta}\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 4}\xrightarrow{\mu\otimes({\operatorname{pr}}_{1}+{\operatorname{pr}}_{2})}\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 2}

where β\beta is the morphism (x,y)(x,y,x,y)(x,y)\mapsto(x,y,x,y). Therefore it suffices to show that the two morphisms

μ,pr1+pr2:tr(𝔸1{0})2tr(𝔸1{0})\mu,{\operatorname{pr}}_{1}+{\operatorname{pr}}_{2}\colon\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes 2}\to\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})

coincide after composing with the canonical epimorphism tr(𝔸1{0})h0𝔾m+\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})\twoheadrightarrow h_{0}\mathbb{G}_{m}^{+}. This follows from Lemma 6.3. ∎

Lemma 6.10.

The following diagram becomes commutative after applying h0h_{0}:

𝕎^pn1+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Vp\scriptstyle{V_{p}}d\scriptstyle{d}𝕎^pn+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}𝕎^pn1+𝔾m+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}}𝕎^pn+𝔾m+.\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n}}\otimes\mathbb{G}_{m}^{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}Fpid\scriptstyle{F_{p}\otimes{\operatorname{id}}}
Proof.

By Lemma 6.3 (2) and our assumption that pp is odd, we have h0(idVp)=idh_{0}({\operatorname{id}}\otimes V_{p})={\operatorname{id}} on h0(𝕎^pn1+𝔾m+)h_{0}(\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}). Therefore it suffices to show that the following diagram is commutative:

𝕎^pn1+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Vp\scriptstyle{V_{p}}d\scriptstyle{d}𝕎^pn+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}𝕎^pn1+𝔾m+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idVp\scriptstyle{{\operatorname{id}}\otimes V_{p}}𝕎^pn1+𝔾m+\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n-1}}\otimes\mathbb{G}_{m}^{+}}𝕎^pn+𝔾m+.\textstyle{\widehat{\mathbb{W}}^{+}_{p^{n}}\otimes\mathbb{G}_{m}^{+}\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}Fpid\scriptstyle{F_{p}\otimes{\operatorname{id}}}

This is further reduced to the commutativity of the following diagram in Cork\operatorname{\mathrm{Cor}}_{k}:

𝔸1{0}\textstyle{\mathbb{A}^{1}\setminus\{0\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρpt\scriptstyle{{}^{t}\rho_{p}}Δ\scriptstyle{\Delta}𝔸1{0}\textstyle{\mathbb{A}^{1}\setminus\{0\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ\scriptstyle{\Delta}(𝔸1{0})2\textstyle{(\mathbb{A}^{1}\setminus\{0\})^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idρpt\scriptstyle{{\operatorname{id}}\otimes{}^{t}\rho_{p}}(𝔸1{0})2\textstyle{(\mathbb{A}^{1}\setminus\{0\})^{2}}(𝔸1{0})2.\textstyle{(\mathbb{A}^{1}\setminus\{0\})^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}ρpid\scriptstyle{\rho_{p}\otimes{\operatorname{id}}}

Both compositions are given by the cycle on (𝔸1{0})3(\mathbb{A}^{1}\setminus\{0\})^{3} defined by x=y=zpx=y=z^{p}. ∎

Theorem 6.11.

Let X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}. Then the tuple (aNish0(W+𝔾m+)(X),F,V,λ)(a_{\operatorname{Nis}}h_{0}(W^{+}_{\bullet}\otimes\mathbb{G}_{m}^{+\otimes*})(X),F,V,\lambda) is a Witt complex over AA. In particular, we have a unique homomorphism of Witt complexes

θ:WΩ(A)aNish0(W+𝔾m+)(X)\theta\colon W_{\bullet}\Omega^{*}(A)\to a_{\operatorname{Nis}}h_{0}(W^{+}_{\bullet}\otimes\mathbb{G}_{m}^{+\otimes*})(X)

which extends to a morphism θ:WnΩqaNish0(Wn+𝔾m+q)\theta\colon W_{n}\Omega^{q}\to a_{\operatorname{Nis}}h_{0}(W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes q}) in ShNis(Smk)\operatorname{Sh}_{\operatorname{Nis}}(\operatorname{\mathrm{Sm}}_{k}).

Proof.

By Lemma 6.3 (5), we have 2d2=02d^{2}=0 and hence d2=0d^{2}=0. By Lemma 6.9 and Lemma 6.3 (4), we see that dd satisfies the Leibniz rule. The relation FdV=dFdV=d holds by Lemma 6.10. Let us show that the relation

F(dλ[a])=λ[ap1]dλ[a]F(d\lambda[a])=\lambda[a^{p-1}]\star d\lambda[a]

holds for any aAa\in A. We use the injectivity theorem for reciprocity sheaves [KSY16, Theorem 6 & 7]; since h0(Wn+𝔾m+q)h_{0}(W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes q}) is a reciprocity presheaf, its Nisnevich sheafification has global injectivity. This allows us to assume that aa is invertible in AA. In this case, both sides of the claimed formula are represented by the morphism (ap,a):SpecA(𝔸1{0})2(a^{p},a)\colon\operatorname{Spec}A\to(\mathbb{A}^{1}\setminus\{0\})^{2}. ∎

6.3. Compatibility with transfers

In the last subsection, we have constructed a morphism θ:WnΩqaNish0(Wn+𝔾m+q)\theta\colon W_{n}\Omega^{q}\to a_{\operatorname{Nis}}h_{0}(W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes q}) in ShNis(Smk)\operatorname{Sh}_{\operatorname{Nis}}(\operatorname{\mathrm{Sm}}_{k}). In this subsection, we prove that θ\theta is compatible with transfers.

Lemma 6.12 (Projection formula).

Let XSmkX\in\operatorname{\mathrm{Sm}}_{k}, YSmXY\in\operatorname{\mathrm{Sm}}_{X} and let π:YX\pi\colon Y\to X be the structure morphism. Let aWnΩq(X)a\in W_{n}\Omega^{q}(X) and bWnΩq(Y)b\in W_{n}\Omega^{q}(Y). For any αCorX(X,Y)\alpha\in\operatorname{\mathrm{Cor}}_{X}(X,Y), we have

α(πab)=aαb.\alpha^{*}(\pi^{*}a\star b)=a\star\alpha^{*}b.
Proof.

Since the transfer structure of WnΩqW_{n}\Omega^{q} is compatible with the trace maps defined in [R0̈7a, Theorem 2.6], the claim follows from the projection formula for the trace maps. ∎

Lemma 6.13.

For any X=SpecASmkX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k} and αCorX(X,𝔸X1)\alpha\in\operatorname{\mathrm{Cor}}_{X}(X,\mathbb{A}^{1}_{X}), the following diagram is commutative:

WnΩq(A[t])\textstyle{W_{n}\Omega^{q}(A[t])\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}θ\scriptstyle{\theta}α\scriptstyle{\alpha^{*}}aNish0(Wn+𝔾m+q)(A[t])\textstyle{a_{\operatorname{Nis}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})(A[t])\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha^{*}}WnΩq(A)\textstyle{W_{n}\Omega^{q}(A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}θ\scriptstyle{\theta}aNish0(Wn+𝔾m+q)(A).\textstyle{a_{\operatorname{Nis}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})(A).}

The following proof is inspired by the proof of [KP21, Proposition 5.19].

Proof.

We prove by induction on qq. For q=0q=0, this follows from Corollary 5.12. Suppose that q1q\geq 1. We prove by induction on nn. We say that an element ωWnΩA[t]q\omega\in W_{n}\Omega^{q}_{A[t]} is traceable if αθ(ω)=θ(αω)\alpha^{*}\theta(\omega)=\theta(\alpha^{*}\omega) holds. Since FF, VV, and dd are compatible with α\alpha^{*}, if ω\omega is traceable then so is F(ω)F(\omega), V(ω)V(\omega), and dωd\omega. The group WnΩA[t]qW_{n}\Omega^{q}_{A[t]} is a quotient of ΩWn(A[t])q\Omega^{q}_{W_{n}(A[t])} and hence generated by elements of the form

ω=Vj0([a0])dVj1([a1])dVjq([aq])\omega=V^{j_{0}}([a_{0}])\star dV^{j_{1}}([a_{1}])\star\dots\star dV^{j_{q}}([a_{q}])

where a0,,aqA[t]a_{0},\dots,a_{q}\in A[t]. Let us prove that ω\omega is traceable.

(1) If j0==jq=0j_{0}=\dots=j_{q}=0, then we can write ω\omega as a \mathbb{Z}-linear combination of elements of the form

η\displaystyle\eta =[c0tk0]d[c1]d[cq],\displaystyle=[c_{0}t^{k_{0}}]\star d[c_{1}]\star\dots\star d[c_{q}],
ξ\displaystyle\xi =[c0tk0]d[t]d[c1]d[cq1]\displaystyle=[c_{0}t^{k_{0}}]\star d[t]\star d[c_{1}]\star\dots\star d[c_{q-1}]

where c0,,cqAc_{0},\dots,c_{q}\in A. First we show that η\eta is traceable. By the projection formula, it suffices to show that [c0tk0][c_{0}t^{k_{0}}] is traceable. This follows from the case q=0q=0. Next we show that ξ\xi is traceable. We write k0+1=mpek_{0}+1=mp^{e} with pmp\nmid m. Then we have

Fe(d[tm])\displaystyle F^{e}(d[t^{m}]) =Fe1([tm(p1)]d[tm])\displaystyle=F^{e-1}([t^{m(p-1)}]\star d[t^{m}])
=Fe2([tmp(p1)][tm(p1)]d[tm])\displaystyle=F^{e-2}([t^{mp(p-1)}]\star[t^{m(p-1)}]\star d[t^{m}])
==[tm(pe1)]d[tm]\displaystyle=\cdots=[t^{m(p^{e}-1)}]\star d[t^{m}]
=m[tmpe1]d[t]=m[tk0]d[t]\displaystyle=m[t^{mp^{e}-1}]\star d[t]=m[t^{k_{0}}]\star d[t]

and hence

ξ=m1Fe(d[tm])[c0]d[c1]d[cq1].\xi=m^{-1}F^{e}(d[t^{m}])\star[c_{0}]\star d[c_{1}]\star\dots\star d[c_{q-1}].

By the projection formula, it suffices to show that m1Fe(d[tm])m^{-1}F^{e}(d[t^{m}]) is traceable. This follows from the case q=0q=0.

(2) If j01j_{0}\geq 1, then we have ω=V(ω)\omega=V(\omega^{\prime}) for some ωWn1ΩA[t]q\omega^{\prime}\in W_{n-1}\Omega^{q}_{A[t]}, so the induction hypothesis on nn shows that ω\omega is traceable.

(3) If j0=0j_{0}=0 and j11j_{1}\geq 1, then we have

[a0]dVj1([a1])=d([a0]Vj1([a1]))Vj1([a1])d([a0]).[a_{0}]\star dV^{j_{1}}([a_{1}])=d([a_{0}]\star V^{j_{1}}([a_{1}]))-V^{j_{1}}([a_{1}])\star d([a_{0}]).

We set ω0=[a0]Vj1([a1])\omega_{0}=[a_{0}]\star V^{j_{1}}([a_{1}]), ω1=Vj1([a1])d([a0])\omega_{1}=V^{j_{1}}([a_{1}])\star d([a_{0}]), and ω2=dVj2([a2])dVjq([aq])\omega_{2}=dV^{j_{2}}([a_{2}])\star\dots\star dV^{j_{q}}([a_{q}]), so that ω=d(ω0ω2)ω1ω2\omega=d(\omega_{0}\star\omega_{2})-\omega_{1}\star\omega_{2}. Then d(ω0ω2)d(\omega_{0}\star\omega_{2}) is traceable by the induction hypothesis on qq, and ω1ω2\omega_{1}\star\omega_{2} is traceable by (1).

(4) If j0=0j_{0}=0 and ji1j_{i}\geq 1 for some 2iq2\leq i\leq q, then ω\omega is traceable by the same argument as in (3). ∎

Lemma 6.14.

Let X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}} and set AN=A[t1,,tN]A_{N}=A[t_{1},\dots,t_{N}]. For any αCorX(X,𝔸XN)\alpha\in\operatorname{\mathrm{Cor}}_{X}(X,\mathbb{A}^{N}_{X}), the following diagram is commutative:

WnΩq(AN)\textstyle{W_{n}\Omega^{q}(A_{N})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}θ\scriptstyle{\theta}α\scriptstyle{\alpha^{*}}aNish0(Wn+𝔾m+q)(AN)\textstyle{a_{\operatorname{Nis}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})(A_{N})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha^{*}}WnΩq(A)\textstyle{W_{n}\Omega^{q}(A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}θ\scriptstyle{\theta}aNish0(Wn+𝔾m+q)(A).\textstyle{a_{\operatorname{Nis}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})(A).}
Proof.

Since the target of θ\theta has global injectivity, we may assume that A=KA=K where KK is a function field. We proceed by induction on NN, the case N=1N=1 being Lemma 6.13. We may assume that α=[x]\alpha=[x] where xx is a closed point of 𝔸KN\mathbb{A}^{N}_{K}. Let π:𝔸KN𝔸KN1\pi\colon\mathbb{A}^{N}_{K}\to\mathbb{A}^{N-1}_{K} be the canonical projection and set y=π(x)y=\pi(x). Then [y][y] defines an element of CorK(SpecK,𝔸KN1)\operatorname{\mathrm{Cor}}_{K}(\operatorname{Spec}K,\mathbb{A}^{N-1}_{K}). Since xx is a closed point of 𝔸k(y)1\mathbb{A}^{1}_{k(y)}, it is defined by some monic equation

tNm+am1tNm1++a0=0t_{N}^{m}+a_{m-1}t_{N}^{m-1}+\dots+a_{0}=0

where a0,,am1k(y)a_{0},\dots,a_{m-1}\in k(y). Let bib_{i} be a lift of aia_{i} to KN1K_{N-1} for i=0,,m1i=0,\dots,m-1. Define Γ𝔸KN\Gamma\subset\mathbb{A}^{N}_{K} by the equation

tNm+bm1tNm1++b0=0.t_{N}^{m}+b_{m-1}t_{N}^{m-1}+\dots+b_{0}=0.

Then [Γ][\Gamma] defines an element of Cor𝔸KN1(𝔸KN1,𝔸KN)\operatorname{\mathrm{Cor}}_{\mathbb{A}^{N-1}_{K}}(\mathbb{A}^{N-1}_{K},\mathbb{A}^{N}_{K}) with [x]=[Γ][y][x]=[\Gamma]\circ[y]. Therefore the claim follows from the induction hypothesis and Lemma 6.13. ∎

Lemma 6.15.

For any X,YSmkX,Y\in\operatorname{\mathrm{Sm}}_{k} and αCork(X,Y)\alpha\in\operatorname{\mathrm{Cor}}_{k}(X,Y), the following diagram is commutative:

WnΩq(Y)\textstyle{W_{n}\Omega^{q}(Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}θ\scriptstyle{\theta}α\scriptstyle{\alpha^{*}}aNish0(Wn+𝔾m+q)(Y)\textstyle{a_{\operatorname{Nis}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})(Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha^{*}}WnΩq(X)\textstyle{W_{n}\Omega^{q}(X)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}θ\scriptstyle{\theta}aNish0(Wn+𝔾m+q)(X).\textstyle{a_{\operatorname{Nis}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})(X).}

In other words, θ\theta is a morphism in ShNis(Cork)\operatorname{Sh}_{\operatorname{Nis}}(\operatorname{\mathrm{Cor}}_{k}).

Proof.

The finite correspondence α\alpha can be factored as

X(id,α)X×Ypr2Y.X\xrightarrow{({\operatorname{id}},\alpha)}X\times Y\xrightarrow{{\operatorname{pr}}_{2}}Y.

Therefore we may assume that YSmXY\in\operatorname{\mathrm{Sm}}_{X} and αCorX(X,Y)\alpha\in\operatorname{\mathrm{Cor}}_{X}(X,Y). Since θ\theta is a morphism of Nisnevich sheaves, we may assume that X=SpecAX=\operatorname{Spec}A, Y=SpecBY=\operatorname{Spec}B. Take a closed immersion SpecBSpecAN\operatorname{Spec}B\to\operatorname{Spec}A_{N} where AN=A[t1,,tN]A_{N}=A[t_{1},\dots,t_{N}]. Since WnΩ(AN)WnΩ(B)W_{n}\Omega(A_{N})\to W_{n}\Omega(B) is surjective, we may assume that B=ANB=A_{N}, so the claim follows from Lemma 6.14. ∎

6.4. Comparison

Definition 6.16.

Consider the sections

[t]:tr(𝔸1{0})Wn,t:tr(𝔸1{0})𝔾m[t]\colon\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})\to W_{n},\quad t\colon\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})\to\mathbb{G}_{m}

where tt is the coordinate function of 𝔸1{0}\mathbb{A}^{1}\setminus\{0\}. One can easily see that these morphisms descend to

[t]:ω!Wn+Wn,t:ω!𝔾m+𝔾m.[t]\colon\omega_{!}W^{+}_{n}\to W_{n},\quad t\colon\omega_{!}\mathbb{G}^{+}_{m}\to\mathbb{G}_{m}.

Taking a tensor product and composing with

Wn𝔾mqWnΩq;xa1aqxdlog[a1]dlog[aq],W_{n}\otimes\mathbb{G}_{m}^{\otimes q}\to W_{n}\Omega^{q};\quad x\otimes a_{1}\otimes\dots\otimes a_{q}\mapsto x\star\operatorname{dlog}[a_{1}]\star\dots\star\operatorname{dlog}[a_{q}],

we get a morphism

η:ω!(Wn+𝔾m+q)WnΩq\eta\colon\omega_{!}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})\to W_{n}\Omega^{q}

in PSh(Cork)\operatorname{PSh}(\operatorname{\mathrm{Cor}}_{k}). In other words, η\eta is the morphism induced by the section [t0]dlog[t1]dlog[tq][t_{0}]\star\operatorname{dlog}[t_{1}]\star\cdots\star\operatorname{dlog}[t_{q}] of WnΩqW_{n}\Omega^{q} on (𝔸1{0})q+1(\mathbb{A}^{1}\setminus\{0\})^{q+1}, where t0,,tqt_{0},\dots,t_{q} are the coordinate functions of (𝔸1{0})q+1(\mathbb{A}^{1}\setminus\{0\})^{q+1}.

Lemma 6.17.

The morphism η:ω!(Wn+𝔾m+q)WnΩq\eta\colon\omega_{!}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})\to W_{n}\Omega^{q} descends to h0(Wn+𝔾m+q)h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q}).

Note that η\eta further factors through aNish0(Wn+𝔾m+q)a_{\operatorname{Nis}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q}) since WnΩqW_{n}\Omega^{q} is a Nisnevich sheaf.

Proof.

We identify η\eta with the corresponding section of WnΩqW_{n}\Omega^{q} on (𝔸1{0})q+1(\mathbb{A}^{1}\setminus\{0\})^{q+1}. Fix ε>0\varepsilon\in\mathbb{Q}_{>0}. Let 𝒲=(1,ε[0]+(pn1+ε)[])\mathcal{W}=(\mathbb{P}^{1},\varepsilon[0]+(p^{n-1}+\varepsilon)[\infty]) and 𝒢=(1,ε[0]+ε[])\mathcal{G}=(\mathbb{P}^{1},\varepsilon[0]+\varepsilon[\infty]). It suffices to show that for any connected XSmkX\in\operatorname{\mathrm{Sm}}_{k} and

αCork(X(1,{1}),𝒲𝒢q),\alpha\in\operatorname{\mathrm{Cor}}_{k}(X\otimes(\mathbb{P}^{1},\{1\}),\mathcal{W}\otimes\mathcal{G}^{\otimes q}),

we have (i0i)αη=0WnΩq(X)(i_{0}^{*}-i_{\infty}^{*})\alpha^{*}\eta=0\in W_{n}\Omega^{q}(X). Since the homomorphism WnΩq(X)WnΩq(k(X))W_{n}\Omega^{q}(X)\to W_{n}\Omega^{q}(k(X)) is injective, we may assume that X=SpecKX=\operatorname{Spec}K where KK is a function field. Moreover, we may assume that α=[Γ]\alpha=[\Gamma] for some integral closed subscheme ΓK×K(𝔸K1{0})q+1\Gamma\subset\square_{K}\times_{K}(\mathbb{A}^{1}_{K}\setminus\{0\})^{q+1} which is finite surjective over K\square_{K}. Let CC denote the normalization of Γ\Gamma. Let g:CKg\colon C\to\square_{K} and (f0,,fq):C(𝔸K1{0})q+1(f_{0},\dots,f_{q})\colon C\to(\mathbb{A}^{1}_{K}\setminus\{0\})^{q+1} denote the canonical projections. Then α(i0i)\alpha\circ(i_{0}-i_{\infty}) is equal to the composite

SpecKdiv(g)C(f0,,fq)(𝔸K1{0})q+1.\operatorname{Spec}K\xrightarrow{\operatorname{div}(g)}C\xrightarrow{(f_{0},\dots,f_{q})}(\mathbb{A}^{1}_{K}\setminus\{0\})^{q+1}.

Therefore it suffices to show that

xC(1)nxTrK(x)/K([f0]dlog[f1]dlog[fq])=0\textstyle\sum_{x\in C^{(1)}}n_{x}\operatorname{Tr}_{K(x)/K}([f_{0}]\star\operatorname{dlog}[f_{1}]\star\dots\star\operatorname{dlog}[f_{q}])=0

where div(g)=xC(1)nx{x}\operatorname{div}(g)=\sum_{x\in C^{(1)}}n_{x}\{x\}. Using the residue map ([R0̈7a], [R0̈7]), we can rewrite the left hand side as

xC(1)ResC/K,x([f0]dlog[f1]dlog[fq]dlog[g]).\textstyle\sum_{x\in C^{(1)}}\operatorname{Res}_{C/K,x}([f_{0}]\star\operatorname{dlog}[f_{1}]\star\dots\star\operatorname{dlog}[f_{q}]\star\operatorname{dlog}[g]).

By the residue theorem ([R0̈7a, Theorem 2.19], [R0̈7, Theorem 2]), the above sum is zero if we replace CC with the regular projective model C¯\overline{C} of CC. Therefore it suffices to show that [f0]dlog[f1]dlog[fq]dlog[g][f_{0}]\star\operatorname{dlog}[f_{1}]\star\dots\star\operatorname{dlog}[f_{q}]\star\operatorname{dlog}[g] is regular at every xC¯Cx\in\overline{C}\setminus C. Let π\pi be a local parameter at xx and write fi=uiπmif_{i}=u_{i}\pi^{m_{i}} and g=1+uπmg=1+u\pi^{m} where ui,u𝒪C¯,x×u_{i},u\in\mathcal{O}_{\overline{C},x}^{\times}. Note that we always have m>0m>0. The admissibility of Γ\Gamma implies

mεmax{m0,0}+(pn1+ε)max{m0,0}+ε(|m1|++|mq|)>pn1m0.m\geq\varepsilon\max\{m_{0},0\}+(p^{n-1}+\varepsilon)\max\{-m_{0},0\}+\varepsilon(|m_{1}|+\cdots+|m_{q}|)>-p^{n-1}m_{0}.

Hence the result follows from the next lemma. ∎

Lemma 6.18.

Let (K,v)(K,v) be a henselian discrete valuation field of characteristic pp, 𝒪K\mathcal{O}_{K} be its valuation ring and 𝔪K\mathfrak{m}_{K} be the maximal ideal of 𝒪K\mathcal{O}_{K}. Let f0,,fqK×f_{0},\dots,f_{q}\in K^{\times}, a𝔪Ka\in\mathfrak{m}_{K} and assume that v(a)>pn1v(f0)v(a)>-p^{n-1}v(f_{0}). Then we have [f0]dlog[f1]dlog[fq]dlog[1+a]WnΩq(𝒪K)[f_{0}]\star\operatorname{dlog}[f_{1}]\star\cdots\star\operatorname{dlog}[f_{q}]\star\operatorname{dlog}[1+a]\in W_{n}\Omega^{q}(\mathcal{O}_{K}).

Proof.

Let π\pi be a uniformizer of KK. Write fi=uiπmif_{i}=u_{i}\pi^{m_{i}} and a=uπma=u\pi^{m} where ui,u𝒪K×u_{i},u\in\mathcal{O}_{K}^{\times}. Then we have m>pn1m0m>-p^{n-1}m_{0}. By [R0̈7a, Lemma 3.4], we can write

[1+a]\displaystyle[1+a] =[1]+i=0n1Vi([πm]ci)\displaystyle=[1]+\textstyle\sum_{i=0}^{n-1}V^{i}([\pi^{m}]\star c_{i})

where ciWn(𝒪K)c_{i}\in W_{n}(\mathcal{O}_{K}). Therefore it suffices to show that

[f0]dlog[f1]dlog[fq]dVi([πm]ci)[f_{0}]\star\operatorname{dlog}[f_{1}]\star\cdots\star\operatorname{dlog}[f_{q}]\star dV^{i}([\pi^{m}]\star c_{i})

is regular for i=0,1,,n1i=0,1,\dots,n-1. By the Leibniz rule, it suffices to show that

  1. (1)

    [f0]dlog[f1]dlog[fq]Vi([πm]ci)[f_{0}]\star\operatorname{dlog}[f_{1}]\star\cdots\star\operatorname{dlog}[f_{q}]\star V^{i}([\pi^{m}]\star c_{i}),

  2. (2)

    d[f0]dlog[f1]dlog[fq]Vi([πm]ci)d[f_{0}]\star\operatorname{dlog}[f_{1}]\star\cdots\star\operatorname{dlog}[f_{q}]\star V^{i}([\pi^{m}]\star c_{i})

are regular. For (1), we have

[f0]dlog[f1]dlog[fq]Vi([πm]ci)\displaystyle[f_{0}]\star\operatorname{dlog}[f_{1}]\star\cdots\star\operatorname{dlog}[f_{q}]\star V^{i}([\pi^{m}]\star c_{i})
=\displaystyle= Vi(Fi[f0]dlog[f1]dlog[fq][πm]ci)\displaystyle V^{i}(F^{i}[f_{0}]\star\operatorname{dlog}[f_{1}]\star\cdots\star\operatorname{dlog}[f_{q}]\star[\pi^{m}]\star c_{i})
=\displaystyle= Vi([u0piπm+pim0]dlog[f1]dlog[fq]ci)\displaystyle V^{i}([u_{0}^{p^{i}}\pi^{m+p^{i}m_{0}}]\star\operatorname{dlog}[f_{1}]\star\cdots\star\operatorname{dlog}[f_{q}]\star c_{i})

and hence the claim follows from m>pim0m>-p^{i}m_{0}. For (2), we have d[f0]=[f0]dlog[f0]d[f_{0}]=[f_{0}]\star\operatorname{dlog}[f_{0}] and hence the claim follows from the regularity of (1). ∎

Lemma 6.19.

The morphism η:h0(Wn+𝔾m+q)WnΩq\eta\colon h_{0}(W^{+}_{n}\otimes\mathbb{G}_{m}^{+\otimes q})\to W_{n}\Omega^{q} is compatible with \star, dd, FF, VV, and λ\lambda.

Proof.

The claim for ,F,V,λ\star,F,V,\lambda follows from Proposition 5.11. Let us prove the claim for dd. Define δ:(𝔸1{0})q+1(𝔸1{0})q+2\delta\colon(\mathbb{A}^{1}\setminus\{0\})^{q+1}\to(\mathbb{A}^{1}\setminus\{0\})^{q+2} by (t0,,tq)(t0,t0,t1,,tq)(t_{0},\dots,t_{q})\mapsto(t_{0},t_{0},t_{1},\dots,t_{q}). Then we have the following commutative diagram:

tr(𝔸1{0})(q+1)\textstyle{\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes(q+1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δ\scriptstyle{\delta}tr(𝔸1{0})(q+2)\textstyle{\mathbb{Z}_{\operatorname{tr}}(\mathbb{A}^{1}\setminus\{0\})^{\otimes(q+2)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h0(Wn+𝔾m+q)\textstyle{h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}h0(Wn+𝔾m+(q+1)).\textstyle{h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes(q+1)}).}

Therefore it suffices to show that δη=dη\delta^{*}\eta=d\eta holds as a section of WnΩq+1W_{n}\Omega^{q+1} on (𝔸1{0})q+1(\mathbb{A}^{1}\setminus\{0\})^{q+1}. This follows from the equality [t]dlog[t]=d[t][t]\star\operatorname{dlog}[t]=d[t] in WnΩ1(𝔸1{0})W_{n}\Omega^{1}(\mathbb{A}^{1}\setminus\{0\}). ∎

Lemma 6.20.

We have θ(η)=id\theta(\eta)={\operatorname{id}}, where the right hand side denotes the section of aNish0(Wn+𝔾m+q)a_{\operatorname{Nis}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q}) on (𝔸1{0})q+1(\mathbb{A}^{1}\setminus\{0\})^{q+1} represented by the identity morphism.

Proof.

The section η\eta can be written as

η=[t0t11tq1]d[t1]d[tq]\eta=[t_{0}t_{1}^{-1}\dots t_{q}^{-1}]\star d[t_{1}]\star\dots\star d[t_{q}]

where t0,,tqt_{0},\dots,t_{q} are the coordinate functions of (𝔸1{0})q+1(\mathbb{A}^{1}\setminus\{0\})^{q+1}. Since θ\theta is a morphism of Witt complexes, θ(η)\theta(\eta) is given by the same formula in h0(Wn+𝔾m+q)((𝔸1{0})q+1)h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})((\mathbb{A}^{1}\setminus\{0\})^{q+1}). The element [t0t11tq1][t_{0}t_{1}^{-1}\dots t_{q}^{-1}] is represented by the morphism

t0t11tq1:(𝔸1{0})q+1𝔸1{0}t_{0}t_{1}^{-1}\dots t_{q}^{-1}\colon(\mathbb{A}^{1}\setminus\{0\})^{q+1}\to\mathbb{A}^{1}\setminus\{0\}

and the element d[ti]d[t_{i}] is represented by the morphism

(𝔸1{0})q+1ti𝔸1{0}Δ(𝔸1{0})2.(\mathbb{A}^{1}\setminus\{0\})^{q+1}\xrightarrow{t_{i}}\mathbb{A}^{1}\setminus\{0\}\xrightarrow{\Delta}(\mathbb{A}^{1}\setminus\{0\})^{2}.

By definition of the multiplication on h0(Wn+𝔾m+)h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes*}), we see that θ(η)\theta(\eta) is represented by the identity morphism. ∎

Theorem 6.21.

Let kk be a perfect field of characteristic p3p\geq 3. Then there is an isomorphism

η:aNish0(Wn+𝔾m+q)WnΩq\eta\colon a_{{\operatorname{Nis}}}h_{0}(W_{n}^{+}\otimes\mathbb{G}_{m}^{+\otimes q})\xrightarrow{\sim}W_{n}\Omega^{q}

in ShNis(Cork)\operatorname{Sh}_{\operatorname{Nis}}(\operatorname{\mathrm{Cor}}_{k}).

Proof.

Lemma 6.20 shows that θη=id\theta\circ\eta={\operatorname{id}}. Let X=SpecASmkaffX=\operatorname{Spec}A\in\operatorname{\mathrm{Sm}}_{k}^{\mathrm{aff}}. Since WnΩq(A)W_{n}\Omega^{q}(A) is initial in the category of Witt complexes over AA, and ηθ\eta\circ\theta gives an endomorphism of WnΩq(A)W_{n}\Omega^{q}(A) as a Witt complex over AA by Lemma 6.19, we get ηθ=id\eta\circ\theta={\operatorname{id}}. ∎

References

  • [Bin+17] F. Binda, J. Cao, W. Kai and R. Sugiyama “Torsion and divisibility for reciprocity sheaves and 0-cycles with modulus” In Journal of Algebra 469 Academic Press Inc., 2017, pp. 437–463 DOI: 10.1016/j.jalgebra.2016.07.036
  • [BVK16] Luca Barbieri-Viale and Bruno Kahn “On the derived category of 1-motives” 381, Asterisque Societe Mathematique de France, 2016
  • [GW10] Ulrich Görtz and Torsten Wedhorn “Algebraic Geometry I: Schemes With Examples and Exercises” Wiesbaden: Vieweg+Teubner, 2010 DOI: 10.1007/978-3-8348-9722-0˙15
  • [Hes15] Lars Hesselholt “The big de Rham-Witt complex” In Acta Mathematica 214, 2015, pp. 135–207 DOI: 10.1007/s11511-015-0124-y
  • [HM04] Lars Hesselholt and Ib Madsen “On the de Rham-Witt complex in mixed characteristic” In Annales scientifiques de l’École Normale Supérieure Ser. 4, 37.1 Elsevier, 2004, pp. 1–43 DOI: 10.1016/j.ansens.2003.06.001
  • [Ill79] Luc Illusie “Complexe de de Rham-Witt et cohomologie cristalline” In Annales scientifiques de l’École Normale Supérieure 4e série, 12.4 Elsevier, 1979, pp. 501–661 DOI: 10.24033/asens.1374
  • [Kah+21] Bruno Kahn, Hiroyasu Miyazaki, Shuji Saito and Takao Yamazaki “Motives with modulus, I: Modulus sheaves with transfers for non-proper modulus pairs” In Épijournal Géom. Algébrique 5, 2021, pp. Art. 1, 46 DOI: 10.1137/19m1272494
  • [Kah+22] Bruno Kahn, Hiroyasu Miyazaki, Shuji Saito and Takao Yamazaki “Motives with modulus, III: The categories of motives” In Annals of K-Theory 7.1 MSP, 2022, pp. 119 –178 DOI: 10.2140/akt.2022.7.119
  • [Koi] Junnosuke Koizumi “Steinberg symbols and reciprocity sheaves” arXiv:2108.04163, to appear in Ann. K-Theory
  • [KP21] Amalendu Krishna and Jinhyun Park “De Rham–Witt sheaves via algebraic cycles” In Compositio Mathematica 157.10 London Mathematical Society, 2021, pp. 2089–2132 DOI: 10.1112/S0010437X21007478
  • [KS16] Moritz Kerz and Shuji Saito “Chow group of 0-cycles with modulus and higher-dimensional class field theory” In Duke Mathematical Journal 165.15 Duke University Press, 2016, pp. 2811 –2897 DOI: 10.1215/00127094-3644902
  • [KSY16] Bruno Kahn, Shuji Saito and Takao Yamazaki “Reciprocity sheaves (with two appendices by Kay Rülling)” In Compositio Mathematica 152.9 London Mathematical Society, 2016, pp. 1851–1898 DOI: 10.1112/S0010437X16007466
  • [KSY22] Bruno Kahn, Shuji Saito and Takao Yamazaki “Reciprocity sheaves, II” In Homology Homotopy Appl. 24.1, 2022, pp. 71–91 DOI: 10.4310/hha.2022.v24.n1.a4
  • [Miy19] Hiroyasu Miyazaki “Cube invariance of higher Chow groups with modulus” In J. Algebraic Geom. 28, 2019, pp. 339–390
  • [MVW06] Carlo Mazza, Vladimir Voevodsky and Charles Weibel “Lecture Notes on Motivic Cohomology” 2, Clay Mathematics Monographs, 2006
  • [R0̈7] Kay Rülling “Erratum to: “The generalized de Rham-Witt complex over a field is a complex of zero-cycles”” In J. Alg. Geom. 16, 2007, pp. 793–795
  • [R0̈7a] Kay Rülling “The generalized de Rham-Witt complex over a field is a complex of zero-cycles” In J. Alg. Geom. 16, 2007, pp. 109–169
  • [RS21] Kay Rülling and Shuji Saito “Reciprocity sheaves and their ramification filtrations” In Journal of the Institute of Mathematics of Jussieu Cambridge University Press, 2021, pp. 1–74 DOI: 10.1017/S1474748021000074
  • [RS22] Kay Rülling and Shuji Saito “Ramification theory for reciprocity sheaves, III, Abbes-Saito formula” arXiv, 2022 DOI: 10.48550/ARXIV.2204.10637
  • [RS23] Kay Rülling and Shuji Saito “Ramification theory of reciprocity sheaves, I, Zariski-Nagata purity” In Journal für die reine und angewandte Mathematik (Crelles Journal) 2023.797, 2023, pp. 41–78 DOI: doi:10.1515/crelle-2022-0094
  • [RS23a] Kay Rülling and Shuji Saito “Ramification theory of reciprocity sheaves, II, Higher local symbols” In European Journal of Mathematics 9.56, 2023 DOI: https://doi.org/10.1007/s40879-023-00655-8
  • [RSY22] Kay Rülling, Rin Sugiyama and Takao Yamazaki “Tensor structures in the theory of modulus presheaves with transfers” In Mathematische Zeitschrift 300, 2022, pp. 929–977 DOI: 10.1007/s00209-021-02819-2
  • [RY16] Kay Rülling and Takao Yamazaki “Suslin homology of relative curves with modulus” In J. Lond. Math. Soc. 93, 2016, pp. 567–589
  • [SV00] Andrei Suslin and Vladimir Voevodsky “Bloch-Kato Conjecture and Motivic Cohomology with Finite Coefficients” In The Arithmetic and Geometry of Algebraic Cycles Dordrecht: Springer Netherlands, 2000, pp. 117–189 DOI: 10.1007/978-94-011-4098-0˙5
  • [Voe00] Vladimir Voevodsky “Triangulated categories of motives over a field” In Cycles, transfers, and motivic homology theories 143, Ann. of Math. Stud. Princeton Univ. Press, Princeton, NJ, 2000, pp. 188–238
  • [Voe98] Vladimir Voevodsky 𝔸1\mathbb{A}^{1}-homotopy theory” In Proceedings of the International Congress of Mathematicians, Vol. I (Berlin, 1998), 1998, pp. 579–604