Hostname: page-component-76fb5796d-5g6vh Total loading time: 0 Render date: 2024-04-26T15:29:16.896Z Has data issue: false hasContentIssue false

Reciprocity sheaves and logarithmic motives

Published online by Cambridge University Press:  13 February 2023

Shuji Saito*
Affiliation:
Graduate School of Mathematical Sciences, University of Tokyo, 3-8-1 Komaba, Tokyo 153-8914, Japan sshuji@msb.biglobe.ne.jp
Rights & Permissions [Opens in a new window]

Abstract

We connect two developments that aim to extend Voevodsky's theory of motives over a field in such a way as to encompass non-$\mathbf {A}^1$-invariant phenomena. One is theory of reciprocity sheaves introduced by Kahn, Saito and Yamazaki. The other is theory of the triangulated category $\operatorname {\mathbf {logDM}}^{{\operatorname {eff}}}$ of logarithmic motives launched by Binda, Park and Østvær. We prove that the Nisnevich cohomology of reciprocity sheaves is representable in $\operatorname {\mathbf {logDM}}^{{\operatorname {eff}}}$.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited. Compositio Mathematica is © Foundation Compositio Mathematica.
Copyright
© 2023 The Author(s)

Introduction

We fix once and for all a perfect base field $k$. The main purpose of this paper is to connect two developments that aim to extend Voevodsky's theory of motives over $k$ in such a way as to encompass non-$\mathbf {A}^1$-invariant phenomena. One is the theory of reciprocity sheaves introduced by Kahn, Saito and Yamazaki [Reference Kahn, Saito and YamazakiKSY16, Reference Kahn, Saito and YamazakiKSY22] and developed in [Reference SaitoSai20, Reference Binda, Rülling and SaitoBRS22]. Voevodsky's theory is based on the category ${\operatorname {\mathbf {PST}}}$ of presheaves with transfers, defined as the category of additive presheaves of abelian groups on the category $\operatorname {\mathbf {Cor}}$ of finite correspondences: $\operatorname {\mathbf {Cor}}$ has the same objects as the category $\operatorname {\mathbf {Sm}}$ of separated smooth schemes of finite type over $k$, and morphisms in $\operatorname {\mathbf {Cor}}$ are finite correspondences. Let $\operatorname {\mathbf {NST}}\subset {\operatorname {\mathbf {PST}}}$ be the full subcategory of Nisnevich sheaves, that is, those objects $F\in {\operatorname {\mathbf {PST}}}$ whose restrictions $F_X$ to the small étale site $X_{{\operatorname {\acute {e}t}}}$ over $X$ are Nisnevich sheaves for all $X\in \operatorname {\mathbf {Sm}}$. Voevodsky proved that $\operatorname {\mathbf {NST}}$ is a Grothendieck abelian category and defined the triangulated category $\operatorname {\mathbf {DM}}^{{\operatorname {eff}}}$ of effective motives as the localization of the derived category $D(\operatorname {\mathbf {NST}})$ of complexes in $\operatorname {\mathbf {NST}}$ with respect to an $\mathbf {A}^1$-weak equivalence (see [Reference Mazza, Voevodsky and WeibelMVW06, Definition 14.1]). It is equipped with a functor $M: \operatorname {\mathbf {Sm}} \to \operatorname {\mathbf {DM}}^{{\operatorname {eff}}}$ associating the motive $M(X)$ of $X\in \operatorname {\mathbf {Sm}}$.

Let $\operatorname {\mathbf {HI}}_{\operatorname {Nis}}\subset \operatorname {\mathbf {NST}}$ be the full subcategory consisting of $\mathbf {A}^1$-invariant objects, namely such $F\in \operatorname {\mathbf {NST}}$ that the projection $\pi _X: X\times \mathbf {A}^1\to X$ induces an isomorphism $\pi _X^* :F(X)\simeq F(X\times \mathbf {A}^1)$ for any $X\in \operatorname {\mathbf {Sm}}$. We say that $F\in \operatorname {\mathbf {HI}}_{\operatorname {Nis}}$ is strictly $\mathbf {A}^1$-invariant if $\pi _X$ induces isomorphisms

\[ \pi_X^* : H^i_{\operatorname{Nis}}(X,F_X) \simeq H^i_{\operatorname{Nis}}(X\times\mathbf{A}^1,F_{X\times\mathbf{A}^1}) \quad\text{for all } i\geq 0. \]

The following theorem plays a fundamental role in Voevodsky's theory.

Theorem 0.1 (Voevodsky [Reference Voevodsky, Friedlander, Suslin and VoevodskyVoe00])

Any $F\in \operatorname {\mathbf {HI}}_{\operatorname {Nis}}$ is strictly $\mathbf {A}^1$-invariant and we have a natural isomorphism

(0.1.1)\begin{equation} H^i_{\operatorname{Nis}}(X,F_X)\simeq \operatorname{Hom}_{\operatorname{\mathbf{DM}}^{{\operatorname{eff}}}}(M(X),L^{\mathbf{A}^1}F[i])\quad \text{for } X\in \operatorname{\mathbf{Sm}}, \end{equation}

where $L^{\mathbf {A}^1} : D(\operatorname {\mathbf {NST}})\to \operatorname {\mathbf {DM}}^{{\operatorname {eff}}}$ is the localization functor.

Notice that there are interesting and important objects of $\operatorname {\mathbf {NST}}$ which do not belong to $\operatorname {\mathbf {HI}}_{\operatorname {Nis}}$. Such examples are given by the sheaves $\Omega ^i$ of (absolute or relative) differential forms; the $p$-typical de Rham–Witt sheaves $W_m\Omega ^i$ of Bloch, Deligne and Illusie; smooth commutative $k$-group schemes with a unipotent part (seen as objects of $\operatorname {\mathbf {NST}}$); and the complexes $R\varepsilon _* \mathbb {Z}/p^r(n)$ with ${\operatorname {ch}}(k)=p>0$, where $\mathbb {Z}/p^r(n)$ is the étale motivic complex of weight $n$ with $\mathbb {Z}/p^r$ coefficients and $\varepsilon$ is the change of site functor from the étale to the Nisnevich topology. For such examples, (0.1.1) fails to hold since $\pi _X:X\times \mathbf {A}^1\to X$ induces an isomorphism $M(X\times \mathbf {A}^1)\simeq M(X)$ in $\operatorname {\mathbf {DM}}^{{\operatorname {eff}}}$ but the maps induced on the cohomology of those sheaves are not isomorphisms.

The category ${\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$ of reciprocity sheaves is a full abelian subcategory of $\operatorname {\mathbf {NST}}$ that contains $\operatorname {\mathbf {HI}}_{{\operatorname {Nis}}}$ as well as the non-$\mathbf {A}^1$-invariant objects mentioned above. Heuristically, its objects satisfy the property that for any $X\in \operatorname {\mathbf {Sm}}$, each section $a\in F(X)$ ‘has bounded ramification at infinity’ and the objects of $\operatorname {\mathbf {HI}}_{\operatorname {Nis}}$ are special reciprocity sheaves with the property that every section $a\in F(X)$ has ‘tame’ ramification at infinity.Footnote 1 Slightly more exotic examples of reciprocity sheaves are given by the sheaves ${\rm Conn}^1$ (for ${\operatorname {ch}}(k)=0$), whose sections over $X$ are rank $1$-connections, or ${\rm Lisse}^1_\ell$ (in case ${\operatorname {ch}}(k)=p>0$), whose sections on $X$ are the lisse $\overline {\mathbb {Q}}_\ell$-sheaves of rank $1$. Since ${\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$ is an abelian category equipped with a lax symmetric monoidal structure by [Reference Rülling, Sugiyama and YamazakiRSY22], many more interesting examples can be manufactured by taking kernels, quotients and tensor products (see [Reference Binda, Rülling and SaitoBRS22, § 11.1] for more examples).

The main purpose of this paper is to establish formula (0.1.1) for all $F\in {\operatorname {\mathbf {RSC}}}_{\operatorname {Nis}}$ in a new category which enlarges $\operatorname {\mathbf {DM}}^{{\operatorname {eff}}}$ (see (0.2)). It is the triangulated category $\operatorname {\mathbf {logDM}}^{{\operatorname {eff}}}$ of logarithmic motives introduced by Binda, Park and Østvær in [Reference Binda, Park and ØstværBPØ22]. Let ${\operatorname {l\mathbf {Sm}}}$ be the category of log smooth and separated $\mathrm {fs}$ log schemes of finite type over $k$, and ${\operatorname {l\mathbf {Cor}}}$ be the category with the same objects as ${\operatorname {l\mathbf {Sm}}}$ and whose morphisms are log finite correspondences (see [Reference Binda, Park and ØstværBPØ22, Definition 2.1.1]). Let ${\operatorname {\mathbf {PSh}}^{\rm ltr}}$ be the category of additive presheaves of abelian groups on ${\operatorname {l\mathbf {Cor}}}$ and ${\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}\subset {\operatorname {\mathbf {PSh}}^{\rm ltr}}$ be the full subcategory consisting of those $\mathcal {F}$ whose restrictions to ${\operatorname {l\mathbf {Sm}}}$ are dividing Nisnevich sheaves (see [Reference Binda, Park and ØstværBPØ22, Definition 3.1.4]). It is shown in [Reference Binda, Park and ØstværBPØ22, Theorem 1.2.1] that ${\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}$ is a Grothendieck abelian category, and $\operatorname {\mathbf {logDM}}^{{\operatorname {eff}}}$ is defined as the localization of the derived category $D({\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}})$ of complexes in ${\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}$ with respect to a ${\overline {\square }}$-weak equivalence, where ${\overline {\square }}$ is $\mathbf {P}^1$ with the log structure associated to the effective divisor $\infty \hookrightarrow \mathbf {P}^1$ (see [Reference Binda, Park and ØstværBPØ22, Definition 5.2.1]).Footnote 2 It is equipped with a functor $M: {\operatorname {l\mathbf {Sm}}} \to \operatorname {\mathbf {logDM}}^{{\operatorname {eff}}}$ associating the logarithmic motive $M(\mathfrak {X})$ of $\mathfrak {X}\in {\operatorname {l\mathbf {Sm}}}$. Thanks to [Reference Binda and MericiBM12, Theorem 1,1], the standard $t$-structure on $D({\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}})$ induces a $t$-structure on $\operatorname {\mathbf {logDM}}^{{\operatorname {eff}}}$ called the homotopy $t$-structure, and its heart is identified with the abelian full subcategory $\operatorname {\mathbf {CI}}^{\rm ltr}_{\rm dNis}\subset {\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}$ consisting of strictly ${\overline {\square }}$-invariant objects in the sense of [Reference Binda, Park and ØstværBPØ22, Definition 5.2.2].Footnote 3 We can now state the main result of this paper.

Theorem 0.2 (Theorems 6.1 and 6.3)

There exists an exact and fully faithful functor

(0.2.1)\begin{equation} {\mathcal{L}og : {\operatorname{\mathbf{RSC}}}_{\operatorname{Nis}} \to \operatorname{\mathbf{CI}}^{\rm ltr}_{\rm dNis}\;:\; F\to F^{\log}=\mathcal{L}og(F) .} \end{equation}

For $X\in \operatorname {\mathbf {Sm}}$ we have a natural isomorphism

(0.2.2)\begin{equation} {H^i_{\operatorname{Nis}}(X,F_X)\simeq \operatorname{Hom}_{\operatorname{\mathbf{logDM}}^{{\operatorname{eff}}}}(M(X,\mathrm{triv}),L^{{\overline{\square}}}F^{\log}[i]),} \end{equation}

where $L^{{\overline {\square }}} : D({\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}})\to \operatorname {\mathbf {logDM}}^{{\operatorname {eff}}}$ is the localization functor and $(X, \mathrm {triv})$ is the log scheme with the trivial log structure.

We remark (see Remark 5.6) that, for $F=\Omega ^i$, $F^{\log }(\mathfrak {X})$ for $\mathfrak {X}\in {\operatorname {l\mathbf {Sm}}}$ whose underlying scheme is smooth agrees with the sheaf of logarithmic differential forms of $\mathfrak {X}$ at least assuming ${\operatorname {ch}}(k)=0$.Footnote 4

We now explain the organization of the paper.

In § 1 we discuss some preliminaries and fix notation. We recall the definitions and basic properties of modulus (pre)sheaves with transfers from [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, Reference Kahn, Miyazaki, Saito and YamazakiKMSY21b, Reference Kahn, Saito and YamazakiKSY22, Reference SaitoSai20]. These are a generalization of Voevodsky's (pre)sheaves with transfers to a version with modulus. The category $\operatorname {\mathbf {\underline {M}Cor}}$ of modulus correspondences is introduced. Its objects are pairs $\mathcal {X} = (\overline {X}, D)$, where $\overline {X}$ is a separated scheme of finite type over $k$ equipped with an effective Cartier divisor $D$ such that the interior $\overline {X}-D = X$ is smooth. The morphisms are finite correspondences on the interiors satisfying admissibility and a properness condition. Let $\operatorname {\mathbf {\underline {M}PST}}$ be the category of additive presheaves of abelian groups on $\operatorname {\mathbf {\underline {M}Cor}}$. A full subcategory $\operatorname {\mathbf {\underline {M}NST}}\subset \operatorname {\mathbf {\underline {M}PST}}$ of Nisnevich sheaves is defined and there is a functor (see § 1(20))

\[ \underline{\omega}^{\operatorname{\mathbf{CI}}} : {\operatorname{\mathbf{RSC}}}_{\operatorname{Nis}} \to \operatorname{\mathbf{\underline{M}NST}}. \]

For every $F\in {\operatorname {\mathbf {RSC}}}_{\operatorname {Nis}}$ and $X\in \operatorname {\mathbf {Sm}}$, it provides an exhaustive filtration on the group $F(X)$ of sections over $X$ which measures the depth of ramification along a boundary of a partial compactification of $X$: for $({\overline {X}},D)\in \operatorname {\mathbf {\underline {M}Cor}}$ with ${\overline {X}}-D=X$, we get the subgroups $\tilde {F}({\overline {X}},D)\subset F(X)$ with $\tilde {F}=\underline {\omega }^{\operatorname {\mathbf {CI}}} F$ such that $\tilde {F}({\overline {X}},D_1)\subset \tilde {F}({\overline {X}},D_2)$ if $D_1\leq D_2$.

In § 2 we prove as a key technical input an analogue of the Zariski–Nagata purity theorem [Reference GrothendieckSGA2, X 3.4] for $\tilde {F}({\overline {X}},D)$ as above. This asserts the exactness of the sequence

\[ 0\to \tilde{F}({\overline{X}},D) \to F(X) \to \underset{\xi\in {{D}^{(0)}}}\bigoplus \frac{F({{{\overline{X}}}^{h}_{|\xi}}-\xi)}{\tilde{F}({{{\overline{X}}}^{h}_{|\xi}},\xi)}, \]

where ${\overline {X}}\in \operatorname {\mathbf {Sm}}$ and $D$ is a reduced simple normal crossing divisor, and where ${{D}^{(0)}}$ is the set of the irreducible components of $D$ and ${{{\overline{X}}}^{h}_{|\xi}}$ is the henselization of ${\overline {X}}$ at $\xi$. In [Reference Rülling and SaitoRS21b] this result is generalized to the case where $D$ may not be reduced under the assumption that ${\overline {X}}$ admits a smooth compactification.

In § 3 we review higher local symbols for reciprocity sheaves constructed in [Reference Rülling and SaitoRS21c]. These are an effective tool with which one can decide when a given element of $F(X)$ with $F\in {\operatorname {\mathbf {RSC}}}_{\operatorname {Nis}}$ and $X\in \operatorname {\mathbf {Sm}}$ belongs to $\tilde {F}({\overline {X}},D)$ as above. The construction of the pairing depends on pushforward maps for the cohomology of reciprocity sheaves constructed in [Reference Binda, Rülling and SaitoBRS22] (which means that Theorem 0.2 depends on the result of [Reference Binda, Rülling and SaitoBRS22]).

In § 4 we prove the following result. Let $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}_{\rm ls}$ be the subcategory of $\operatorname {\mathbf {\underline {M}Cor}}$ whose objects are pairs $(X,D)$ such that $X\in \operatorname {\mathbf {Sm}}$ and the reduced divisor $D_{{\operatorname {red}}}$ underlying $D$ is a simple normal crossing divisor on $X$ and whose morphisms are modulus correspondences satisfying a finiteness conditions instead of the properness condition (see § 1(5)). Then, for $F\in {\operatorname {\mathbf {RSC}}}_{\operatorname {Nis}}$, the association

\[ \tilde{F}^{\log}: (X,D) \to \underline{\omega}^{\operatorname{\mathbf{CI}}} F (X,D_{{\operatorname{red}}}) \]

gives a presheaf on $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}_{\rm ls}$, which gives rise to a cohomology theory $H^i_{\log }(-,\tilde {F}^{\log })$ on $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}_{\rm ls}$, called the $i$th logarithmic cohomology with coefficient $F$ (see Definition 4.4). The higher local symbols for $F$ play a fundamental role in the proof of the result.

In § 5 we prove the invariance of logarithmic cohomology under blowups. Let $\Lambda ^{{\operatorname {fin}}}_{\rm ls}$ be the subcategory of $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}_{\rm ls}$ whose objects are the same as $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}_{\rm ls}$ and whose morphisms are those $\rho : (Y,E) \to (X,D)$ where $E=\rho ^* D$ and $\rho$ are induced by blowups of $X$ in smooth centers $Z\subset D$ which are normal crossing to $D$ (see the beginning of the section). Then, for $F\in {\operatorname {\mathbf {RSC}}}_{\operatorname {Nis}}$ and $\rho :\mathcal {Y}\to \mathcal {X}$ in $\Lambda ^{{\operatorname {fin}}}_{\rm ls}$, we have

\[ \rho^*: {H_{\log}^{i}}(\mathcal{X},F) \cong {H_{\log}^{i}}(\mathcal{Y},F)\quad \forall i\geq 0. \]

In § 6 we prove Theorem 0.2, which is a formal consequence of the theorems in §§ 4 and 5.

1. Preliminaries

We fix once and for all a perfect base field $k$. In this section we recall the definitions and basic properties of modulus sheaves with transfers from [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, Reference SaitoSai20].

  1. (1) Denote by $\operatorname {\mathbf {Sch}}$ the category of separated schemes of finite type over $k$ and by $\operatorname {\mathbf {Sm}}$ the full subcategory of smooth schemes. For $X, Y \in \operatorname {\mathbf {Sm}}$, an integral closed subscheme of $X\times Y$ that is finite and surjective over a connected component of $X$ is called a prime correspondence from $X$ to $Y$. The category $\operatorname {\mathbf {Cor}}$ of finite correspondences has the same objects as $\operatorname {\mathbf {Sm}}$, and for $X, Y \in \operatorname {\mathbf {Sm}}$, $\operatorname {\mathbf {Cor}}(X,Y)$ is the free abelian group on the set of all prime correspondences from $X$ to $Y$ (see [Reference Voevodsky, Friedlander, Suslin and VoevodskyVoe00]). We consider $\operatorname {\mathbf {Sm}}$ as a subcategory of $\operatorname {\mathbf {Cor}}$ by regarding a morphism in $\operatorname {\mathbf {Sm}}$ as its graph in $\operatorname {\mathbf {Cor}}$.

    Let ${\operatorname {\mathbf {PST}}}$ be the category of additive presheaves of abelian groups on $\operatorname {\mathbf {Cor}}$ whose objects are called presheaves with transfers. Let $\operatorname {\mathbf {NST}}\subseteq {\operatorname {\mathbf {PST}}}$ be the category of Nisnevich sheaves with transfers and let

    \[ a^V_{\operatorname{Nis}} :{\operatorname{\mathbf{PST}}}\to \operatorname{\mathbf{NST}} \]
    be Voevodsky's Nisnevich sheafification functor, which is an exact left adjoint to the inclusion $\operatorname {\mathbf {NST}}\to {\operatorname {\mathbf {PST}}}$. Let $\operatorname {\mathbf {HI}}\subseteq {\operatorname {\mathbf {PST}}}$ be the category of $\mathbf {A}^1$-invariant presheaves and put $\operatorname {\mathbf {HI}}_{\operatorname {Nis}}=\operatorname {\mathbf {HI}}\cap \operatorname {\mathbf {NST}}\subseteq \operatorname {\mathbf {NST}}$.
  2. (2) Let $\operatorname {\mathbf {Sm}}^{\rm pro}$ be the category of $k$-schemes $X$ which are essentially smooth over $k$, that is, $X$ is a limit $\mathop {\varprojlim }_{i \in I} X_i$ over a filtered set $I$, where $X_i$ is smooth over $k$ and all transition maps are étale. Note that $\operatorname {Spec} K\in \operatorname {\mathbf {Sm}}^{\rm pro}$ for a function field $K$ over $k$ thanks to the assumption that $k$ is perfect. We define $\operatorname {\mathbf {Cor}}^{\rm pro}$ whose objects are the same as $\operatorname {\mathbf {Sm}}^{\rm pro}$ and whose morphisms are defined as [Reference Rülling and SaitoRS21a, Definition 2,2]. We extend $F\in {\operatorname {\mathbf {PST}}}$ to a presheaf on $\operatorname {\mathbf {Cor}}^{\rm pro}$ by $F(X) := \mathop {\varinjlim }_{i \in I} F(X_i)$ for $X$ as above.

  3. (3) We recall the definition of the category $\operatorname {\mathbf {\underline {M}Cor}}$ from [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, Definition 1.3.1]. A pair $\mathcal {X} = (X,D)$ consisting of $X \in \operatorname {\mathbf {Sch}}$ and an effective Cartier divisor $D$ on $X$ is called a modulus pair if $X - D \in \operatorname {\mathbf {Sm}}$. Let $\mathcal {X}=(X,D_X)$, $\mathcal {Y}=(Y,D_Y)$ be modulus pairs and $\Gamma \in \operatorname {\mathbf {Cor}}(X-D_X,Y-D_Y)$ be a prime correspondence. Let $\overline {\Gamma } \subseteq X \times Y$ be the closure of $\Gamma$, and let $\overline {\Gamma }^N\to X\times Y$ be the normalization. We say that $\Gamma$ is admissible (respectively, left proper) if $(D_X)_{\overline {\Gamma }^N}\geq (D_Y)_{\overline {\Gamma }^N}$ (respectively, if $\overline {\Gamma }$ is proper over $X$). Let $\operatorname {\mathbf {\underline {M}Cor}}(\mathcal {X}, \mathcal {Y})$ be the subgroup of $\operatorname {\mathbf {Cor}}(X-D_X,Y-D_Y)$ generated by all admissible left proper prime correspondences. The category $\operatorname {\mathbf {\underline {M}Cor}}$ has modulus pairs as objects and $\operatorname {\mathbf {\underline {M}Cor}}(\mathcal {X}, \mathcal {Y})$ as the group of morphisms from $\mathcal {X}$ to $\mathcal {Y}$.

  4. (4) Let $\operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}\subset \operatorname {\mathbf {\underline {M}Cor}}$ be the full subcategory of $(X,D)\in \operatorname {\mathbf {\underline {M}Cor}}$ with $X\in \operatorname {\mathbf {Sm}}$ and $|D|$ a normal crossing divisor on $X$.

  5. (5) Let $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}\subset \operatorname {\mathbf {\underline {M}Cor}}$ be the full subcategory of the same objects such that $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}(\mathcal {X}, \mathcal {Y})$ are generated by all admissible finite prime correspondences, where finite prime correspondences are defined by replacing the left properness in (3) by finiteness. We also define $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}_{\rm ls} = \operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}\cap \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$.

  6. (6) There is a canonical pair of adjoint functors $\lambda \dashv \underline {\omega }$:

    \begin{gather*} \lambda: \operatorname{\mathbf{Cor}}\to \operatorname{\mathbf{\underline{M}Cor}}\quad X\mapsto (X,\emptyset),\end{gather*}
    \begin{gather*} \underline{\omega}: \operatorname{\mathbf{\underline{M}Cor}}\to\operatorname{\mathbf{Cor}}\quad (X,D)\mapsto X-D.\end{gather*}
  7. (7) There is a full subcategory $\mathbf {MCor}\subset \operatorname {\mathbf {\underline {M}Cor}}$ consisting of proper modulus pairs, where a modulus pair $(X,D)$ is proper if $X$ is proper. Let $\tau :\operatorname {\mathbf {MCor}}\hookrightarrow \operatorname {\mathbf {\underline {M}Cor}}$ be the inclusion functor and $\omega =\underline {\omega }\tau$.

  8. (8) Let $\operatorname {\mathbf {MPST}}$ (respectively, $\operatorname {\mathbf {\underline {M}PST}}$) be the category of additive presheaves of abelian groups on $\operatorname {\mathbf {MCor}}$ (respectively, $\operatorname {\mathbf {\underline {M}Cor}}$) whose objects are called modulus presheaves with transfers. For $\mathcal {X} \in \operatorname {\mathbf {MCor}}$, let ${\operatorname {\mathbb {Z}_{{\operatorname {tr}}}}}(\mathcal {X})=\operatorname {\mathbf {\underline {M}Cor}}(-,\mathcal {X})$ be the representable object of $\operatorname {\mathbf {\underline {M}PST}}$. We sometimes write $\mathcal {X}$ for ${\operatorname {\mathbb {Z}_{{\operatorname {tr}}}}}(\mathcal {X})$ for simplicity.

  9. (9) In the same manner as (2), the category $\operatorname {\mathbf {\underline {M}Cor}}^{\rm pro}$ is defined and $F\in \operatorname {\mathbf {\underline {M}PST}}$ is extended to a presheaf on $\operatorname {\mathbf {\underline {M}Cor}}^{\rm pro}$ (see [Reference Rülling and SaitoRS21a, § 3.7]).

  10. (10) The adjunction $\lambda \dashv \underline {\omega }$ induces a string of four adjoint functors $(\lambda _!=\underline {\omega }^!,\lambda ^*=\underline {\omega }_!,\lambda _*=\underline {\omega }^*,\underline {\omega }_*)$ (see [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, Proposition 2.3.1]):

    \[ \operatorname{\mathbf{\underline{M}PST}}\begin{smallmatrix}\underline{\omega}^!\\\longleftarrow\\\underline{\omega}_!\\\longrightarrow\\\underline{\omega}^*\\\longleftarrow\\\underline{\omega}_*\\ \longrightarrow\end{smallmatrix}{\operatorname{\mathbf{PST}}} \]
    where $\underline {\omega }_!,\underline {\omega }_*$ are localizations and $\underline {\omega }^!$ and $\underline {\omega }^*$ are fully faithful.
  11. (11) The functor $\tau$ yields a string of three adjoint functors $(\tau _!,\tau ^*,\tau _*)$:

    \[ \operatorname{\mathbf{MPST}}\begin{smallmatrix}\tau_!\\\longrightarrow\\\tau^*\\\longleftarrow\\\tau_*\\ \longrightarrow\end{smallmatrix}\operatorname{\mathbf{\underline{M}PST}} \]
    where $\tau _!,\tau _*$ are fully faithful and $\tau ^*$ is a localization; $\tau _!$ has a pro-left adjoint $\tau ^!$, hence is exact (see [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, Proposition 2.4.1]). We will denote by $\operatorname {\mathbf {\underline {M}PST}}^{\tau }$ the essential image of $\tau _!$ in $\operatorname {\mathbf {\underline {M}PST}}$.
  12. (12) The modulus pair ${\overline {\square }}:=(\mathbf {P}^1,\infty )$ has an interval structure induced by that of $\mathbf {A}^1$ (see [Reference Kahn, Saito and YamazakiKSY22, Lemma 2.1.3]). We say that $F \in \operatorname {\mathbf {MPST}}$ is ${\overline {\square }}$-invariant if $p^* : F(\mathcal {X}) \to F(\mathcal {X} \otimes {\overline {\square }})$ is an isomorphism for any $\mathcal {X} \in \operatorname {\mathbf {MCor}}$, where $p : \mathcal {X} \otimes {\overline {\square }} \to \mathcal {X}$ is the projection. Let $\operatorname {\mathbf {CI}}$ be the full subcategory of $\operatorname {\mathbf {MPST}}$ consisting of all ${\overline {\square }}$-invariant objects and $\operatorname {\mathbf {CI}}^{\tau }\subset \operatorname {\mathbf {\underline {M}PST}}$ be the essential image of $\operatorname {\mathbf {CI}}$ under $\tau _!$.

  13. (13) Recall from [Reference Kahn, Saito and YamazakiKSY22, Theorem 2.1.8] that $\operatorname {\mathbf {CI}}$ is a Serre subcategory of $\operatorname {\mathbf {MPST}}$, and that the inclusion functor $i^{\overline {\square }} : \operatorname {\mathbf {CI}} \to \operatorname {\mathbf {MPST}}$ has a left adjoint $h_0^{\overline {\square }}$ and a right adjoint $h^0_{\overline {\square }}$ given for $F \in \operatorname {\mathbf {MPST}}$ and $\mathcal {X} \in \operatorname {\mathbf {MCor}}$ by

    \begin{align*} &h_0^{\overline{\square}}(F)(\mathcal{X})=\operatorname{Coker}(i_0^* - i_1^* : F(\mathcal{X} \otimes {\overline{\square}}) \to F(\mathcal{X})),\\ &h^0_{\overline{\square}}(F)(\mathcal{X})=\operatorname{Hom}(h_0^{\overline{\square}}(\mathcal{X}), F). \end{align*}
    For $\mathcal {X}\in \operatorname {\mathbf {MCor}}$, we write $h_0^{\overline {\square }}(\mathcal {X})=h_0^{\overline {\square }}({\operatorname {\mathbb {Z}_{{\operatorname {tr}}}}}(\mathcal {X}))\in \operatorname {\mathbf {CI}}$, and by abuse of notation we also write $h_0^{\overline {\square }}(\mathcal {X})$ for $\tau _!h_0^{\overline {\square }}(\mathcal {X})\in \operatorname {\mathbf {CI}}^{\tau }$.
  14. (14) For $F\in \operatorname {\mathbf {\underline {M}PST}}$ and $\mathcal {X}=(X,D)\in \operatorname {\mathbf {\underline {M}Cor}}$, write $F_{\mathcal {X}}$ for the presheaf on the small étale site $X_{{\operatorname {\acute {e}t}}}$ over $X$ given by $U\to F(\mathcal {X}_U)$ for $U\to X$ étale, where $\mathcal {X}_U=(U,D_{|U})\in \operatorname {\mathbf {\underline {M}Cor}}$. We say that $F$ is a Nisnevich sheaf if $F_{\mathcal {X}}$ is also one for all $\mathcal {X}\in \operatorname {\mathbf {\underline {M}Cor}}$ (see [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, § 3]). We write $\operatorname {\mathbf {\underline {M}NST}}\subset \operatorname {\mathbf {\underline {M}PST}}$ for the full subcategory of Nisnevich sheaves and put

    \[ \operatorname{\mathbf{MNST}}^\tau=\operatorname{\mathbf{\underline{M}NST}}\cap \operatorname{\mathbf{MPST}}^\tau,\quad \operatorname{\mathbf{CI}}^{\tau}_{\operatorname{Nis}} =\operatorname{\mathbf{CI}}^{\tau}\cap \operatorname{\mathbf{MNST}}^\tau. \]
    By [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, Proposition 3.5.3] and [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21b, Theorem 2], the inclusion functor $i_{\operatorname {Nis}}: \operatorname {\mathbf {\underline {M}NST}}\to \operatorname {\mathbf {\underline {M}PST}}$ has an exact left adjoint ${\underline {a}}_{\operatorname {Nis}}$ such that ${\underline {a}}_{\operatorname {Nis}}(\operatorname {\mathbf {MPST}}^\tau )\subset \operatorname {\mathbf {MNST}}^\tau$. The functor ${\underline {a}}_{\operatorname {Nis}}$ has the following description. For $F\in \operatorname {\mathbf {\underline {M}PST}}$ and $\mathcal {Y}\in \operatorname {\mathbf {\underline {M}Cor}}$, let $F_{\mathcal {Y},{\operatorname {Nis}}}$ be the usual Nisnevich sheafification of $F_{\mathcal {Y}}$. Then, for $(X,D)\in \operatorname {\mathbf {\underline {M}Cor}}$, we have
    \[ {\underline{a}}_{\operatorname{Nis}} F (X,D) = \mathop{\varinjlim}_{f:Y\to X} F_{(Y,f^*D),{\operatorname{Nis}}}(Y) \]
    where the colimit is taken over all proper maps $f:Y\to X$ that induce isomorphisms $Y-|f^*D|\xrightarrow {\sim }X-|D|$.
  15. (15) By [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21b, Proposition 6.2.1], $\underline {\omega }^*$ and $\underline {\omega }_!$ from (10) respect $\operatorname {\mathbf {\underline {M}NST}}$ and $\operatorname {\mathbf {NST}}$ and induce a pair of adjoint functors (which for simplicity we write $\underline {\omega }_!$ and $\underline {\omega }^*$). Moreover, we have

    \[ \underline{\omega}_!{\underline{a}}_{\operatorname{Nis}}=a^V_{\operatorname{Nis}}\underline{\omega}_!. \]
    By [Reference Kahn, Saito and YamazakiKSY22, Lemma 2.3.1] and [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21b, Proposition 6.2.1a)], for $F\in {\operatorname {\mathbf {PST}}}$, we have $F\in \operatorname {\mathbf {HI}}$ (respectively, $F\in \operatorname {\mathbf {HI}}_{\operatorname {Nis}}$) if and only if $\underline {\omega }^*F\in \operatorname {\mathbf {CI}}^{\tau }$ (respectively, $\underline {\omega }^*F\in \operatorname {\mathbf {CI}}^{\tau }_{\operatorname {Nis}}$).
  16. (16) We say that $F\in \operatorname {\mathbf {\underline {M}PST}}$ is semipure if the unit map

    \[ u: F\to \underline{\omega}^*\underline{\omega}_!F \]
    is injective. For $F\in \operatorname {\mathbf {\underline {M}PST}}$ (respectively, $F\in \operatorname {\mathbf {\underline {M}NST}}$), let $F^{{\rm sp}}\in \operatorname {\mathbf {\underline {M}PST}}$ (respectively, $F^{{\rm sp}}\in \operatorname {\mathbf {\underline {M}NST}}$) be the image of $F\to \underline {\omega }^*\underline {\omega }_! F$ (called the semipurification of $F$. See [Reference SaitoSai20, Lemma 1.30]). For $F\in \operatorname {\mathbf {\underline {M}PST}}$ we have
    \[ {\underline{a}}_{\operatorname{Nis}}(F^{{\rm sp}}) \simeq ({\underline{a}}_{\operatorname{Nis}} F)^{{\rm sp}}. \]
    This follows from the fact that ${\underline {a}}_{\operatorname {Nis}}$ is exact and commutes with $\underline {\omega }^*\underline {\omega }_!$. For $F\in \operatorname {\mathbf {MPST}}^{\tau }$ we have $F^{{\rm sp}}\in \operatorname {\mathbf {MPST}}^{\tau }$ since $\tau$ is exact and $\underline {\omega }^*\underline {\omega }_!\tau _!=\tau _! \omega ^*\omega _!$.
  17. (17) Let $\operatorname {\mathbf {CI}}^{\tau,{\rm sp}}\subset \operatorname {\mathbf {CI}}^{\tau }$ be the full subcategory of semipure objects and consider the full subcategory

    \[ \operatorname{\mathbf{CI}}^{\tau,{\rm sp}}_{\operatorname{Nis}} =\operatorname{\mathbf{CI}}^{\tau,{\rm sp}}\cap \operatorname{\mathbf{MNST}}^\tau \subset \operatorname{\mathbf{CI}}^{\tau}_{\operatorname{Nis}}. \]
    By [Reference SaitoSai20, Theorems 0.1 and 0.4], we have ${\underline {a}}_{\operatorname {Nis}}(\operatorname {\mathbf {CI}}^{\tau,{\rm sp}})\subset \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{\operatorname {Nis}}$.
  18. (18) We write ${\operatorname {\mathbf {RSC}}}\subseteq {\operatorname {\mathbf {PST}}}$ for the essential image of $\operatorname {\mathbf {CI}}$ under $\omega _!$ (which is the same as the essential image of $\operatorname {\mathbf {CI}}^{\tau,{\rm sp}}$ under $\underline {\omega }_!$ since $\omega _!=\underline {\omega }_!\tau _!$ and $\underline {\omega }_!F=\underline {\omega }_!F^{{\rm sp}}$). Put ${\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}={\operatorname {\mathbf {RSC}}}\cap \operatorname {\mathbf {NST}}$. The objects of ${\operatorname {\mathbf {RSC}}}$ (respectively, ${\operatorname {\mathbf {RSC}}}_{\operatorname {Nis}}$) are called reciprocity presheaves (respectively, sheaves). By [Reference SaitoSai20, Theorem 0.1], we have

    (1.0.1)\begin{equation} a^V_{\operatorname{Nis}}({\operatorname{\mathbf{RSC}}})\subset {\operatorname{\mathbf{RSC}}}_{\operatorname{Nis}}. \end{equation}
    We have $\operatorname {\mathbf {HI}}\subseteq {\operatorname {\mathbf {RSC}}}$ which also contains smooth commutative group schemes (which may have non-trivial unipotent part), the sheaf $\Omega ^i$ of Kähler differentials, and the de Rham–Witt sheaves $W_n\Omega ^i$ (see [Reference Kahn, Saito and YamazakiKSY16, Reference Kahn, Saito and YamazakiKSY22]).
  19. (19) $\operatorname {\mathbf {NST}}$ is a Grothendieck abelian category by [Reference Voevodsky, Friedlander, Suslin and VoevodskyVoe00, Lemma 3.1.6] and we can make ${\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$ its full subabelian category as follows. We define the kernel (respectively, cokernel) of a map $\phi : F\to G$ in ${\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$ to be that of $\phi$ as a map in $\operatorname {\mathbf {NST}}$. Here we need (1.0.1) to ensure that the cokernel of $\phi$ in $\operatorname {\mathbf {NST}}$ stays in ${\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$. By definition, a sequence $0\to F\to G\to H\to 0$ is exact in ${\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$ if and only if it is exact in $\operatorname {\mathbf {NST}}$.

  20. (20) By [Reference Kahn, Saito and YamazakiKSY22, Proposition 2.3.7] we have a pair of adjoint functors

    (1.0.2)\begin{equation} \operatorname{\mathbf{CI}}\begin{smallmatrix}\omega^{\operatorname{\mathbf{CI}}}\\\longleftarrow\\\omega_!\\\longrightarrow \end{smallmatrix}{\operatorname{\mathbf{RSC}}}, \end{equation}
    where $\omega ^{\operatorname {\mathbf {CI}}}=h^0_{\overline {\square }}\omega ^*$ and is fully faithful. It induces a pair of adjoint functors
    (1.0.3)\begin{equation} \operatorname{\mathbf{CI}}^{\tau}\begin{smallmatrix}\underline{\omega}^{\operatorname{\mathbf{CI}}}\\\longleftarrow\\\underline{\omega}_!\\\longrightarrow \end{smallmatrix}{\operatorname{\mathbf{RSC}}}, \end{equation}
    where $\underline {\omega }^{\operatorname {\mathbf {CI}}}=\tau _!h^0_{\overline {\square }}\omega ^*$ and is fully faithful. Indeed, let $F=\tau _!\hat {F}$ for $\hat {F}\in \operatorname {\mathbf {CI}}$ and $G \in {\operatorname {\mathbf {RSC}}}$. In view of (13) and the exactness and full faithfulness of $\tau _!$, we have
    \begin{align*} \operatorname{Hom}_{\operatorname{\mathbf{CI}}^{\tau}}(F,\tau_!h^0_{\overline{\square}}\omega^* G) &\simeq\operatorname{Hom}_{\operatorname{\mathbf{CI}}}(\hat{F},h^0_{\overline{\square}}\omega^*G) \\ \simeq \operatorname{Hom}_{\operatorname{\mathbf{MPST}}}(\hat{F},\omega^*G) &\simeq\operatorname{Hom}_{\operatorname{\mathbf{\underline{M}PST}}}(\tau_! \hat{F},\underline{\omega}^*G) \simeq \operatorname{Hom}_{{\operatorname{\mathbf{RSC}}}}(\underline{\omega}_! F,G). \end{align*}
    In view of (15), (1.0.3) induces a pair of adjoint functors
    (1.0.4)\begin{equation} \operatorname{\mathbf{CI}}^{\tau,{\rm sp}}_{\operatorname{Nis}}\begin{smallmatrix}\underline{\omega}^{\operatorname{\mathbf{CI}}}\\\longleftarrow\\\underline{\omega}_!\\\longrightarrow \end{smallmatrix}{\operatorname{\mathbf{RSC}}}_{\operatorname{Nis}}. \end{equation}

2. Purity with reduced modulus

For $F\in \operatorname {\mathbf {\underline {M}PST}}$, we put

\begin{align*} F_{-1}&=\operatorname{Ker}\big(\operatorname{\underline{Hom}}_{\operatorname{\mathbf{\underline{M}PST}}}((\mathbf{P}^1-0,\infty),F)\overset{i_1^*}{\longrightarrow} F\big),\end{align*}
\begin{eqnarray*} \\ F_{-1}^{(1)}&=\operatorname{Ker}\big(\operatorname{\underline{Hom}}_{\operatorname{\mathbf{\underline{M}PST}}}((\mathbf{P}^1,0+\infty),F)\overset{i_1^*}{\longrightarrow} F\big)\end{eqnarray*}

Note that if $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$, then $F_{-1},F_{-1}^{(1)}\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ and

(2.0.1)\begin{equation} \begin{aligned}F_{-1}^{(1)}(\mathcal{X}) & = \operatorname{Hom}_{\operatorname{\mathbf{\underline{M}PST}}}(h_{0,{\operatorname{Nis}}}^{{\overline{\square}},{\rm sp}}(\mathbf{P}^1,0+\infty)^0,\operatorname{\underline{Hom}}_{\operatorname{\mathbf{\underline{M}PST}}}({\operatorname{\mathbb{Z}_{{\operatorname{tr}}}}}(\mathcal{X}),F)),\\ F_{-1}(\mathcal{X}) & =\mathop{\varinjlim}_n \operatorname{Hom}_{\operatorname{\mathbf{\underline{M}PST}}}(h_{0,{\operatorname{Nis}}}^{{\overline{\square}},{\rm sp}}(\mathbf{P}^1,n\cdot 0+\infty)^0,\operatorname{\underline{Hom}}_{\operatorname{\mathbf{\underline{M}PST}}}({\operatorname{\mathbb{Z}_{{\operatorname{tr}}}}}(\mathcal{X}),F)) \end{aligned} \end{equation}

for $\mathcal {X}\in \operatorname {\mathbf {\underline {M}Cor}}$, where

\[ h_{0,{\operatorname{Nis}}}^{{\overline{\square}},{\rm sp}}(\mathbf{P}^1,n\cdot 0+\infty)^0=\operatorname{Coker}\big(\mathbb{Z}={\operatorname{\mathbb{Z}_{{\operatorname{tr}}}}}(\operatorname{Spec} k,\emptyset) \overset{i_1}{\longrightarrow} h_{0,{\operatorname{Nis}}}^{{\overline{\square}},{\rm sp}}(\mathbf{P}^1,n\cdot 0+\infty)\big).\]

Definition 2.1 For $e_1,\dots,e_r\in \{0,1\}$, put

\[ {\tau^{(e_1,\dots,e_r)}}F ={\tau^{(e_r)}}\cdots {\tau^{(e_1)}}F, \]

where

\[ {\tau^{(0)}} F=F_{-1}\quad\text{and}\quad {\tau^{(1)}} F = F_{-1}/F_{-1}^{(1)}, \]

where the quotient is taken in $\operatorname {\mathbf {\underline {M}PST}}$.

The existence of retractions in the following lemma was suggested by A. Merici. It implies ${\tau^{(e_1,\dots,e_r)}}F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ if $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$.

Lemma 2.2 For $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$, the inclusion $F_{-1}^{(1)} \to F_{-1}$ admits a retraction $s_F : F_{-1}\to F_{-1}^{(1)}$ such that for any map $\phi : F\to G$ in $\operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$, the following diagram is commutative:

In particular, ${\tau^{(1)}} F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ if $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$.

Proof. In view of (2.0.1), this follows from [Reference Binda, Rülling and SaitoBRS22, Lemma 2.4].

Theorem 2.3 Let $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$. Let $K\{t_1,\dots,t_n\}$ be the henselization of $K[t_1,\dots,t_n]$ at $(t_1,\dots,t_n)$ and $\mathcal {X}=\operatorname {Spec} K\{t_1,\dots,t_n\}$ and $D= \{t_1^{e_1}\cdots t_n^{e_n}=0\}\subset \mathcal {X}$ with $e_1,\dots,e_n\in \{0,1\}$. For a subset $I\subset [1,n]$ let $i_\mathcal {H} : \mathcal {H} \hookrightarrow \mathcal {X}$ be the closed immersion defined by $\{t_i=0\}_{i\in I}$ and $D_\mathcal {H}=\bigl \{ \prod _{j\in [1,n]-I}t_j^{e_j}=0\bigr \}\subset \mathcal {H}$. Then

(2.3.1)\begin{equation} R^\nu i_\mathcal{H}^! F_{(\mathcal{X},D)} =0\quad \text{for }\nu\not=q:=|I|, \end{equation}

and there is an isomorphism

(2.3.2)\begin{equation} ({\tau^{(e_I)}} F)_{(\mathcal{H},D_\mathcal{H})} \simeq R^q i_\mathcal{H}^! F_{(\mathcal{X},D)}\quad \text{with } e_I=(e_i)_{i\in I}\in \mathbb{Z}_{\geq 0}^q. \end{equation}

Proof. The proof is divided into two steps.

Step 1: we prove (2.3.1) and (2.3.2) for $q=|I|=1$. For $\nu =0$, (2.3.1) follows from the semipurity of $F$ and [Reference SaitoSai20, Theorem 3.1]. Thus, it suffices to show (2.3.1) only for $\nu >1$. Let $J=\{j\in [1,n]\;|\; e_j\not =0\}$ and $r=|J|$. If $\dim (\mathcal {X})=0$, the assertion is trivial. If $r=0$, the assertion follows from [Reference SaitoSai20, Corollary 8.6(3)]. Assume $r>0$ and $\dim (\mathcal {X})\geq 1$, and proceed by the double induction on $r$ and $\dim (\mathcal {X})$. Without loss of generality, we may assume

  • $(\spadesuit )$ $e_1\not =0$, and $\mathcal {H}=\{t_1=0\}$ if $\mathcal {H}\subset |D|$.

Let $\iota :\mathcal {Z}\hookrightarrow \mathcal {X}$ be the closed immersion defined by $\{t_1=0\}$ and $D_\mathcal {Z}=\{t_2^{e_2}\cdots t_r^{e_r}=0\}\subset \mathcal {Z}$ and $D'=\{t_2^{e_2}\cdots t_r^{e_r}=0\}\subset \mathcal {X}$. By [Reference SaitoSai20, Lemma 7.1], we have an exact sequence sheaves on $\mathcal {X}_{{\operatorname {Nis}}}$,

\[ 0\to F_{(\mathcal{X},D')} \to F_{(\mathcal{X},D)} \to \iota_*({F^{(e)}_{-1}})_{(\mathcal{Z},D_\mathcal{Z})} \to 0, \]

which gives rise to a long exact sequence of sheaves on $\mathcal {H}_{{\operatorname {Nis}}}$,

(2.3.3)\begin{equation} \cdots \to R^\nu i_\mathcal{H}^!F_{(\mathcal{X},D')}\to R^\nu i_\mathcal{H}^! F_{(\mathcal{X},D)}\to R^\nu i_\mathcal{H}^!\iota_*({F^{(e)}_{-1}})_{(\mathcal{Z},D_\mathcal{Z})}\to\cdots. \end{equation}

By the induction hypothesis, $R^\nu i_\mathcal {H}^!F_{(\mathcal {X},D')}=0$ for $\nu >1$. If $\mathcal {H}\not =\mathcal {Z}$, we have a Cartesian diagram of closed immersions

and we have an isomorphism

\[ R^\nu i_\mathcal{H}^!\iota_*({F^{(e)}_{-1}})_{(\mathcal{Z},D_\mathcal{Z})} \simeq \iota'_* R^\nu i_{\mathcal{H}\cap \mathcal{Z}}^!({F^{(e)}_{-1}})_{(\mathcal{Z},D_\mathcal{Z})}. \]

By the induction hypothesis, $R^\nu i_{\mathcal {H}\cap \mathcal {Z}}^!({F^{(e)}_{-1}})_{(\mathcal {Z},D_\mathcal {Z})}=0$ for $\nu >1$, noting that ${{F^{(e)}_{-1}}}\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ by Lemma 2.2. So the desired vanishing follows from (2.3.3). Moreover, the assumptions $(\spadesuit )$ and $\mathcal {H}\not =\mathcal {Z}$ imply that $\mathcal {H}\not \subset |D|$. Then (2.3.2) (with $q=1$) follows from [Reference SaitoSai20, Lemma 7.1(2)].

If $\mathcal {Z}=\mathcal {H}$, we have

\[ R^\nu i_\mathcal{H}^!\iota_*({F^{(e)}_{-1}})_{(\mathcal{Z},D_\mathcal{Z})}= R^\nu \iota^!\iota_*({F^{(e)}_{-1}})_{(\mathcal{Z},D_\mathcal{Z})}, \]

which vanishes for $\nu >0$. Hence, (2.3.3) gives the desired vanishing together with an exact sequence

\[ 0\to ({F^{(e)}_{-1}})_{(\mathcal{H},D_\mathcal{H})} \overset{\delta}{\longrightarrow} R^1 i_\mathcal{H}^!F_{(\mathcal{X},D')}\to R^1 i_\mathcal{H}^! F_{(\mathcal{X},D)}\to 0. \]

By [Reference SaitoSai20, Lemma 7.1(2)] we have an isomorphism

\[ (F_{-1})_{(\mathcal{H},D_\mathcal{H})} \simeq R^1 i_\mathcal{H}^!F_{(\mathcal{X},D')} \]

through which $\delta$ is identified with the map induced by the canonical map ${F^{(e)}_{-1}} \to F_{-1}$. This proves the desired isomorphism (2.3.2) for $\mathcal {Z}=\mathcal {H}$ and completes step 1.

Step 2: we prove the theorem by induction on $q$ assuming $q>0$. Let $I=\{i_1,\dots,i_q\}\subset [1,n]$ and $\mathcal {Y}\subset \mathcal {X}$ be the closed subscheme defined by $\{t_{i_1}=0\}$. Let $i_\mathcal {Y}:\mathcal {Y}\hookrightarrow \mathcal {X}$ and $i_{\mathcal {H},\mathcal {Y}}:\mathcal {H}\to \mathcal {Y}$ be the induced closed immersions. By step 1 we have $R^\nu i_{\mathcal {Y}}^! F_{(\mathcal {X},D)} =0$ for $\nu \not =1$ and we have an isomorphism

\[ ({\tau^{(e_{i_1})}} F)_{(\mathcal{Y},D_\mathcal{Y})} \simeq R^1 i_\mathcal{Y}^! F_{(\mathcal{X},D)}\quad {\rm with}\ D_\mathcal{Y}=\{t_1^{e_1}\cdots \overset{\vee}{t_{i_1}^{e_{i_1}}}\cdots t_n^{e_n}=0\}\subset \mathcal{Y}. \]

Note ${\tau^{(e_{i_1})}} F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ by Lemma 2.2. Thus, by the induction hypothesis, we have $R^\nu i_{\mathcal {H},\mathcal {Y}}^! {\tau^{(e_{i_1})}} F_{(\mathcal {Y},D_\mathcal {Y})} =0$ for $\nu \not =q-1$. By the spectral sequence

\[ E_2^{a,b}= R^{b} i_{\mathcal{H},\mathcal{Y}}^! R^a i_{\mathcal{Y}}^! F_{(\mathcal{X},D)} \Rightarrow R^{a+b} i_{\mathcal{H}}^! F_{(\mathcal{X},D)} , \]

we get the desired vanishing (2.3.1) and an isomorphism

\begin{align*} R^q i_{\mathcal{H}}^! F_{(\mathcal{X},D)} &\simeq R^{q-1} i_{\mathcal{H},\mathcal{Y}}^! R^1 i_{\mathcal{Y}}^! F_{(\mathcal{X},D)} \simeq R^{q-1} i_{\mathcal{H},\mathcal{Y}}^! ({\tau^{(e_{i_1})}} F)_{(\mathcal{Y},D_\mathcal{Y})}\\ &\simeq ({\tau^{(e_{i_2},\dots,e_{i_q})}}({\tau^{(e_{i_1})}} F))_{(\mathcal{H},D_\mathcal{H})}\simeq ({\tau^{(e_{i_1}, e_{i_2},\dots,e_{i_q})}} F)_{(\mathcal{H},D_\mathcal{H})}, \end{align*}

where the third isomorphism holds by the induction hypothesis. This completes the proof of the theorem.

We say that $\mathcal {X}=(X,D) \in \operatorname {\mathbf {\underline {M}Cor}}$ is reduced if so is $D$. The following Corollaries 2.4 and 2.5 are immediate consequences of Theorem 2.3.

Corollary 2.4 Take $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ and $(X,D)\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$ reduced. Let $x\in {{X}^{(n)}}$ with $K=k(x)$ and let $\mathcal {X}=X^{h}_{|x}$ be the henselization of $X$ at $x$. Then

\[ H^i_x(X_{{\operatorname{Nis}}},F_{(X,D)})=0 \quad \text{for } i\not=n. \]

Choosing an isomorphism

\[ \varepsilon :\mathcal{X} \simeq \operatorname{Spec} K\{t_1,\dots,t_n\} \]

such that $D_{|\mathcal {X}} = \{t_1^{e_1}\cdots t_n^{e_n}=0\}\subset \mathcal {X}$ with $e_1,\dots,e_n\in \{0,1\}$, there exists an isomorphism depending on $\varepsilon$:

\[ \theta_\varepsilon: {\tau^ {({e_1,e_2,\dots,e_n})}}F(x) \simeq H^n_x(X_{{\operatorname{Nis}}},F_{(X,D)}). \]

Corollary 2.5 For $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ and $\mathcal {X}=(X,D) \in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$ reduced, the following sequence is exact:

\[ 0\to F(X,D) \to F(X-D,\emptyset) \to \underset{\xi\in {{D}^{(0)}}}\bigoplus \frac{F({{{X}}^{h}_{|\xi}}-\xi,\emptyset)}{F({{{X}}^{h}_{|\xi}},\xi)}. \]

The idea of deducing the following corollary from the above is due to A. Merici.

Corollary 2.6 Let $\mathcal {X}=(X,D)\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$ be reduced.

  • (1) Assume given an exact sequence in $\operatorname {\mathbf {\underline {M}NST}}$,

    (2.6.1)\begin{equation} {0\to H \overset{\phi}{\longrightarrow} G \overset{\psi}{\longrightarrow} F}, \end{equation}
    such that $F,G, H\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ and that $\underline {\omega }_!\psi$ is surjective in $\operatorname {\mathbf {NST}}$. If $X$ is henselian local, then
    \[ 0\to H(\mathcal{X}) \to G(\mathcal{X}) \to F(\mathcal{X}) \to 0 \]
    is exact.
  • (2) Let $\gamma : F\to G$ be a map in $\operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ such that $\underline {\omega }_!\gamma$ is an isomorphism. Then $F(\mathcal {X})\to G(\mathcal {X})$ is an isomorphism.

  • (3) For $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$, the unit map $u: F\to \underline {\omega }^{\operatorname {\mathbf {CI}}}\underline {\omega }_! F$ induces an isomorphism $F(\mathcal {X})\cong \underline {\omega }^{\operatorname {\mathbf {CI}}}\underline {\omega }_! F(\mathcal {X})$.

Proof. To show (1), it suffices to show the surjectivity of $G(\mathcal {X}) \to F(\mathcal {X})$. Let $\eta \in X$ be the generic point and consider the following commutative diagram of the Cousin complexes:

By Corollary 2.4, the horizontal sequences are exact. By the assumption, $\psi (\eta )$ is surjective. By a diagram chase we are reduced to showing the following claim.

Claim 2.6.1

  1. (i) For $x\in {{X}^{(1)}}$, the sequence

    \[ H^1_x(X,H_\mathcal{X}) \to H^1_x(X,G_\mathcal{X}) \to H^1_x(X,F_\mathcal{X}) \]
    is exact.
  2. (ii) For $y\in {{X}^{(2)}}$, $H^2_y(\phi )$ is injective.

To show (i), by Corollary 2.4, it suffices to show the exactness of ${\tau^{(e)}} H \to {\tau^{(e)}} G \to {\tau^{(e)}} F$ for $e\in \{0,1\}$. The case $e=0$ follows from the left exactness of the endofunctor $\operatorname {\underline {Hom}}_{\operatorname {\mathbf {\underline {M}PST}}}(\mathcal {X}, -)$ on $\operatorname {\mathbf {\underline {M}NST}}$ for any $\mathcal {X}\in \operatorname {\mathbf {\underline {M}Cor}}$. We have a commutative diagram

where $p_*$ are the projections and $s_*$ is a right inverse of $p_*$ coming from the retractions from Lemma 2.2. We have

\[ \phi\circ p_H= p_G\circ \phi,\quad \psi\circ p_G= p_F\circ \psi, \quad \phi\circ s_H= s_G\circ \phi,\quad \psi\circ s_G= s_F\circ \psi. \]

By a diagram chase, the case $e=1$ follows from the case $e=0$.

To show (ii), by Corollary 2.4, it suffices to show the injectivity of ${\tau^{({{\underline {e}}})}} H \to {\tau^{({{\underline {e}}})}} G$ for ${\underline {e}}\in \{(0,0),(0,1),(1,0),(1.1)\}$. The case ${\underline {e}}=(0,0)$ follows from the same left exactness as above, and the other cases from this case thanks to Lemma 2.2.

To show (2), we may assume $\mathcal {X}$ is henselian local. Then it follows from (1).

Finally, (3) follows from (2) since $\underline {\omega }_! u$ is an isomorphism. This completes the proof of the corollary.

3. Review of higher local symbols

In this section we recall from [Reference Rülling and SaitoRS21c] the higher local symbols for reciprocity sheaves, which are a fundamental tool to prove Theorem 4.2, one of the main theorems of this paper. First we introduce some basic notation. In this section $X$ is a reduced noetherian separated scheme of dimension $d<\infty$ such that $X_{(0)}=X^{(d)}$.

Let $K$ be a field. For an integer $r\geq 0$, let $K^{\rm M}_r(K)$ be the Milnor $K$-group of $K$. Let $A$ be a local domain with the function field $K$. For an ideal $I\subset A$, let $\overline {K}^{\rm M}_r(A,I)\subset K^{\rm M}_r(K)$ denote the subgroup generated by symbols

\[ \{1+a,b_1,\dots,b_{r-1}\} \quad {\rm with}\ a\in I,\, b_i\in A^\times. \]

Let $A$ be a noetherian excellent one-dimensional local domain with function field $K$ and residue field $F$. Let $\tilde {A}$ be the normalization of $A$ and $S$ be the set of the maximal ideals of $\tilde {A}$. For $\mathfrak {m}\in S$, denote $\kappa (\mathfrak {m})=\tilde {A}/\mathfrak {m}$. Then we define

(3.0.1)\begin{equation} \partial_{A}:= \sum_{\mathfrak{m}\in S} \operatorname{Nm}_{\kappa(\mathfrak{m})/F}\circ \partial_{\mathfrak{m}} : K^{\rm M}_r(K)\to K^{\rm M}_{r-1}(F), \end{equation}

where $\partial _{\mathfrak {m}} :K^{\rm M}_r(K)\to K^{\rm M}_{r-1}(\kappa (\mathfrak {m}))$ denotes the tame symbol for the discrete valuation ring $\tilde {A}_{\mathfrak {m}}$, the localization of $\tilde {A}$ at $\mathfrak {m}$, and $\operatorname {Nm}_{\kappa (\mathfrak {m})/F}$ is the norm map.

For $x, y\in X$, we write

\[ y< x:\Longleftrightarrow \overline{\{y\}}\subsetneq \overline{\{x\}},\quad \text{that is, } y\in \overline{\{x\}} \text{ and } y\neq x. \]

A chain on $X$ is a sequence

(3.0.2)\begin{equation} {\underline{x}=(x_0,\ldots, x_n)\quad \text{with } x_0< x_1<\cdots < x_n.} \end{equation}

The chain $\underline {x}$ is a maximal Paršin chain (or maximal chain) if $n=d$ and $x_i\in X_{(i)}$. Note that the assumptions on $X$ imply $x_{i}\in \overline {\{x_{i+1}\}}^{(1)}$. We denote

\[ \mathrm{mc}(X)=\{\text{maximal chains on } X\}. \]

A maximal chain with break at $r\in \{0,\ldots, d\}$ is a chain (3.0.2) with $n=d-1$ and $x_i\in X_{(i)}$, for $i< r$, and $x_i\in X_{(i+1)}$, for $i\ge r$. We denote

\[ \mathrm{mc}_r(X)=\{\text{maximal chain with break at $r$ on $X$}\}. \]

For $\underline {x}=(x_0,\ldots, x_{d-1})\in \mathrm {mc}_r(X)$, we denote by $b(\underline {x})$ the set of $y\in X_{(r)}$ such that

(3.0.3)\begin{equation} {\underline{x}(y):=(x_0,\ldots, x_{r-1}, y, x_{r},\ldots, x_{d-1})\in \mathrm{mc}(X).} \end{equation}

In the rest of this section we fix $F=\underline {\omega }^{\operatorname {\mathbf {CI}}} G\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{\operatorname {Nis}}$ with $G\in {\operatorname {\mathbf {RSC}}}_{\operatorname {Nis}}$ (cf. (1.0.4)). We also fix a function field $K$ over the base field $k$. Let $X$ be an integral scheme of finite type over $K$ and assume $d=\dim (X)\geq 1$. Recall from [Reference Rülling and SaitoRS21c, § 5] that we have a collection of bilinear pairings (cf. the convention from § 1(9))

(3.0.4)\begin{equation} {\big\{ (-,-)_{X/K,\underline{x}}: F(K(X))\otimes K^{\rm M}_d(K(X))\to F(K)\big\}_{\underline{x}\in \mathrm{mc}(X)}.} \end{equation}

The following properties hold for all $a\in F(K(X))$ (see Remark 3.1 below).

  1. (HS1) Let $X\hookrightarrow X'$ be an open immersion where $X'$ is an integral $K$-scheme of dimension $d$. Then, for all $\beta \in K^{\rm M}_d(K(X))$,

    \[ (a,\beta)_{X/K,\underline{x}}=(a,\beta)_{X'/K,\underline{x}}. \]
  2. (HS2) Let $\underline {x}=(x_0,\ldots,x_{d-1}, x_d)\in \mathrm {mc}(X)$ and $Z\subset X$ be the closure of $z=x_{d-1}$, and set $\underline {x}'=(x_0,\ldots, x_{d-1})\in \mathrm {mc}(Z)$. Assume $a\in F(\mathcal {O}_{X, z})$ and let $a(z)\in F(K(Z))$ be the restriction of $a$. Then

    \[ (a, \beta)_{X/K, \underline{x}} = (a(z), \partial_{z}\beta)_{Z/K, \underline{x}'}\quad \text{for $\beta\in K^{\rm M}_d(K(X))$}, \]
    where $\partial _{z}: K^{\rm M}_d(K(X)) \to K^{\rm M}_{d-1}(K(Z))$ is the map (3.0.1) for $A=\mathcal {O}_{X, z}$.
  3. (HS3) Let $D\subset X$ be an effective Cartier divisor with $I_D\subset \mathcal {O}_X$ its ideal sheaf. Assume that $X{\setminus} D$ is regular so that $(X,D)\in \operatorname {\mathbf {\underline {M}Cor}}^{\rm pro}$ and that $a\in F(X,D)$. For $\underline {x}=(x_0,\ldots, x_{d-1},x_d)\in \mathrm {mc}(X)$, we have

    \[ (a, \beta)_{X/K,\underline{x}}=0\quad \text{for $\beta\in \overline{K}^{\rm M}_d(\mathcal{O}_{X, x_{d-1}},I_D \mathcal{O}_{X, x_{d-1}})$.} \]
  4. (HS4) Let $\underline {x}'\in \mathrm {mc}_r(X)$ with $0\leq r\leq d-1$. For $\beta \in K^{\rm M}_d(K(X))$,

    \[ (a, \beta)_{X/K, \underline{x}'(y)}=0\quad \text{for almost all } y\in _{-}(\underline{x}'). \]
    Assume that either $r\ge 1$ or $r=0$, $X$ is quasi-projective, and the closure of $x_1$ in $X$ is projective over $K$, where $\underline {x}'=(x_1,\ldots, x_d)$. Then
    \[ \sum_{y\in b(\underline{x}')} (a, \beta)_{X/K, \underline{x}'(y)}=0. \]

Remark 3.1 Properties (HS1)–(HS4) are slight variants of the (stronger) properties (HS1)–(HS4) in [Reference Rülling and SaitoRS21c, Proposition 5.3], where the Milnor $K$-group $K^{\rm M}_{d}(K^h_{X,\underline {x}})$ of the iterated henselization $K^h_{X,\underline {x}}$ of $K(X)$ along the chain $\underline {x}$ is used instead of $K^{\rm M}_{d}(K(X))$. The version stated here follows easily using the natural maps $\iota _{\underline {x}}: K(X) \to K^h_{X,\underline {x}}$ and the commutative diagram in the situation of (HS2),

and the commutative diagram in the situation of (HS4),

where $\partial _{\underline {x}}$ (respectively, $\iota _y$) is defined in [Reference Rülling and SaitoRS21c, (4.1.1)] (respectively, [Reference Rülling and SaitoRS21c, (3.2.3)]). We also note that $\overline {K}^{\rm M}_d(\mathcal {O}_{X, x_{d-1}},I_D \mathcal {O}_{X, x_{d-1}})$ in (HS2) coincides with the Zariski stalk at $x_{d-1}$ of the sheaf $\overline {V}_{d,X|D}$ defined in [Reference Rülling and SaitoRS21c, 4.4].

For a scheme $Z$ over $k$, write $Z_K=Z\otimes _k K$. If $Z_K$ is integral, we denote by $K(Z)$ the function field of $Z_K$. We quote the following result from [Reference Rülling and SaitoRS21c, Proposition 7.3]. It is a key tool in the proof of Theorem 4.2.

Proposition 3.2 Let $X\in \operatorname {\mathbf {Sm}}$ and assume $D$ is a reduced simple normal crossing divisor on $X$ with $I_D\subset \mathcal {O}_X$ its ideal sheaf. Let $U\subset X$ be an open subset containing all the generic points of $D$. Let $a\in F(X{\setminus} D)$. Assume that, for all function fields $K/k$ and for all $\underline {x}=(x_0,\ldots,x_{d-1}, x_d)\in \mathrm {mc}(U_K)$ with $x_{d-1}\in D_K^{(0)}$, we have

\[ (a,\beta)_{X_K/K, \underline{x}}=0\quad \text{for all } \beta\in \overline{K}^{\rm M}(\mathcal{O}_{X, x_{d-1}},I_D \mathcal{O}_{X, x_{d-1}}). \]

Then $a\in F(X,D)$.

4. Logarithmic cohomology of reciprocity sheaves

For $\mathcal {X}=(X,D)\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$, we write $\mathcal {X}_{{\operatorname {red}}}=(X,D_{{\operatorname {red}}})\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$. We say that $\mathcal {X}=(X,D)\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$ is reduced if $\mathcal {X}=\mathcal {X}_{{\operatorname {red}}}$.

Definition 4.1 Let $F\in \operatorname {\mathbf {\underline {M}PST}}$.

  1. (1) We say that $F$ is log-semipure if for any $\mathcal {X}\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$, the map $F(\mathcal {X}_{{\operatorname {red}}}) \to F(\mathcal {X})$ is injective. Note that if $F$ is semipure, $F$ is log-semipure (cf. § 1(16)).

  2. (2) We say that $F$ is logarithmic if it is log-semipure and satisfies the condition that for $\mathcal {X},\mathcal {Y}\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$ with $\mathcal {X}$ reduced and $\alpha \in \operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}(\mathcal {Y},\mathcal {X})$, the image of $\alpha ^*: F(\mathcal {X}) \to F(\mathcal {Y})$ is contained in $F(\mathcal {Y}_{{\operatorname {red}}})\subset F(\mathcal {Y})$.

Let $\operatorname {\mathbf {\underline {M}PST}}_{\log }$ be the full subcategory of $\operatorname {\mathbf {\underline {M}PST}}$ consisting of logarithmic objects and put $\operatorname {\mathbf {\underline {M}NST}}_{\log }=\operatorname {\mathbf {\underline {M}NST}}\cap \operatorname {\mathbf {\underline {M}PST}}_{\log }$.

Theorem 4.2 Any $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ is logarithmic, that is, $\operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}\subset \operatorname {\mathbf {\underline {M}NST}}_{\log }$.

We need a preliminary lemma for the proof of the theorem.

Lemma 4.3 Let $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$. Let $\mathbf {A}^n_K=\operatorname {Spec} K[x_1,\dots,x_n]$ be the affine space over a function field $K$ over $k$ and $V=\operatorname {Spec} K\{x_1,\dots,x_n\}$ be the henselization of $\mathbf {A}^n_K$ at the origin and $\mathcal {L}_i=\{x_i=0\}\subset V$ for $i\in [1,n]$. For an integer $0< r\leq n$, the natural map

\[ K\{x_{r+1},\dots,x_{n}\} [x_1,\dots,x_r] \to K\{x_1,\dots,x_n\} \]

induces a map in $\operatorname {\mathbf {\underline {M}Cor}}^{\rm pro}$ (cf. § 1(9)):

\[ \rho_r: (V,\mathcal{L}_1+\cdots +\mathcal{L}_r) \to (\mathbf{A}^r_S,\{x_1\cdots x_r=0\})\simeq (\mathbf{A}^1,0)^{\otimes r}\otimes (S,\emptyset) , \]

where $S=\operatorname {Spec} K\{x_{r+1},\dots,x_{n}\}$. It induces

(4.3.1)\begin{equation} \rho_r^*: F(\mathbf{A}^r_S,\{x_1\cdots x_r=0\} ) \to F(V,\mathcal{L}_1+\cdots + \mathcal{L}_r). \end{equation}

Then $F(V,\mathcal {L}_1+\cdots + \mathcal {L}_r)$ is generated by the image of $\rho _r^*$ and

\[ F(V,\mathcal{L}_1+\cdots+\overset{\vee}{\mathcal{L}_i}+\cdots \mathcal{L}_{r}) \quad\text{for } i=1,\dots, r. \]

Proof. For $\mathcal {Y}\in \operatorname {\mathbf {\underline {M}Cor}}$, let $F^\mathcal {Y}\in \operatorname {\mathbf {\underline {M}PST}}$ be defined by $F^\mathcal {Y}(\mathcal {Z})=F(\mathcal {Y}\otimes \mathcal {Z})$. Clearly, we have $F^\mathcal {Y}\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ for $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$. We prove the lemma by induction on $r$. The case $r=1$ holds since by [Reference SaitoSai20, Lemmas 7.1 and 5.9], $\rho _1$ induces an isomorphism

\[ F^{(\mathbf{A}^1,0)}(S)/ F^{(\mathbf{A}^1,\emptyset)}(S) \overset{\simeq}\longrightarrow F(V,\mathcal{L}_1)/F(V). \]

By definition $\mathcal {L}_1=\operatorname {Spec} K\{x_2,\dots,x_n\}$ and we have a map in $\operatorname {\mathbf {\underline {M}Cor}}^{\rm pro}$,

\[ (V,\mathcal{L}_1+\cdots +\mathcal{L}_r) \to (\mathbf{A}^1,0)\otimes(\mathcal{L}_1,\mathcal{L}_1\cap (\mathcal{L}_2+\cdots +\mathcal{L}_r)), \]

induced by the natural map $K\{x_2,\dots,x_{n}\} [x_1] \to K\{x_1,\dots,x_n\}$. By [Reference SaitoSai20, Lemmas 7.1 and 5.9], it induces an isomorphism

\[ F^{(\mathbf{A}^1,0)}(\mathcal{L}_1,E)/ F^{(\mathbf{A}^1,\emptyset)}(\mathcal{L}_1,E) \overset{\simeq}\longrightarrow F(V,\mathcal{L}_1+\cdots +\mathcal{L}_r)/F(V,\mathcal{L}_2+\cdots +\mathcal{L}_r) \]

with $E=\mathcal {L}_1\cap (\mathcal {L}_2+\cdots +\mathcal {L}_r)$. By the induction hypothesis, $F^{(\mathbf {A}^1,0)}(\mathcal {L}_1,E)$ is generated by $F^{(\mathbf {A}^1,0)}(\mathcal {L}_1,E_j)$ with $E_j=\mathcal {L}_1\cap (\mathcal {L}_2\cdots +\overset {\vee }{\mathcal {L}_j}+\cdots \mathcal {L}_{r})$ for $j=2,\dots, r$ together with the image of the map

\[ (F^{(\mathbf{A}^1,0)})^{(\mathbf{A}^1,0)^{\otimes r-1} }(S) =F^{ (\mathbf{A}^1,0)^{\otimes r} }(S) \to F^{(\mathbf{A}^1,0)}(\mathcal{L}_1,E) \]

induced by

\[ (\mathcal{L}_1,E) \to (\mathbf{A}^{r-1}_{S},\{x_2\cdots x_r=0\})\simeq(\mathbf{A}^1,0)^{\otimes r-1}\otimes (S,\emptyset) \]

coming from the map $K\{x_{r+1},\dots,x_n\}[x_2,\dots,x_r] \to K\{x_2,\dots,x_d\}$. This proves the lemma.

Proof of Theorem 4.2 By Corollary 2.6(3), we may assume $F=\underline {\omega }^{\operatorname {\mathbf {CI}}} G$ for $G\in {\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$. Take $\mathcal {X}=(X,D),\mathcal {Y}=(Y,E)\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$ with $\mathcal {X}$ reduced, and let $\alpha \in \operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}(\mathcal {Y},\mathcal {X})$ be an elementary correspondence. We need to show that $\alpha ^*(F(\mathcal {X}))\subset F(\mathcal {Y}_{{\operatorname {red}}})$. The question is Nisnevich local over $X$ and $Y$. Hence, we may assume $(X,D)=(V,\mathcal {L}_1+\cdots +\mathcal {L}_r)\in \operatorname {\mathbf {\underline {M}Cor}}^{\rm pro}$ in the notation of Lemma 4.3. If $r=0$, we have $\alpha \in \operatorname {\mathbf {\underline {M}Cor}}((Y,\emptyset ),(X,\emptyset ))$ by the assumption $\alpha \in \operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}(\mathcal {Y},\mathcal {X})$ so that

\[ \alpha^*(F(\mathcal{X}))=\alpha^*(F(X,\emptyset)) \subset F(Y,\emptyset)\subset F(\mathcal{Y}_{{\operatorname{red}}}). \]

Assume $r>0$ and proceed by induction on $r$. By Lemma 4.3, we may then assume

\[ (X,D)= \mathcal{M}:=(\mathbf{A}^1,0)^{\otimes r}\otimes(S,\emptyset) \quad {\rm for}\ S\in \operatorname{\mathbf{Sm}}^{\rm pro}. \]

On the other hand, by Corollary 2.5, we have an exact sequence

\[ 0\to F(Y,E_{{\operatorname{red}}}) \to F(Y-E_{{\operatorname{red}}},\emptyset) \to \underset{\xi\in {{E}^{(0)}}}\bigoplus \frac{F({{{Y}}^{h}_{|\xi}}-\xi,\emptyset)}{F({{Y}^{h}_{|\xi}},\xi)}. \]

Hence, we may replace $Y$ with its Nisnevich neighborhood of a generic point $\xi$ of $E$. Using the assumption that $k$ is perfect, we may then assume the following condition $(\spadesuit )$. Recall that $\alpha$ is by definition an integral closed subscheme of $(Y-E)\times (X-D)$ finite surjective over $Y-E$, and its closure $\overline {\alpha }$ in $Y\times X$ is finite surjective over $Y$.

  • $(\spadesuit )$ Let $Y'$ be the normalization of $\overline {\alpha }$ and $E':=E\times _Y Y'$. Then $X$, $Y$, $E$ and $E'$ are irreducible, and $\alpha$, $Y'$, $E_{{\operatorname {red}}}$ and $E'_{{\operatorname {red}}}$ are essentially smooth over $k$.

Let $g: Y'\to Y$ and $f: Y'\to X$ be the induced maps. We have $E'=g^*E \geq f^*D$ as Cartier divisors on $Y'$ by the modulus condition for $\alpha$. Hence, these maps induce

\[ F(X,D)\overset{f^*}{\longrightarrow} F(Y',E') \overset{g_*}{\longrightarrow} F(Y,E). \]

We claim that $\alpha ^*:F(X,D)\to F(Y,E)$ agrees with this map. Indeed, this follows from the equality

\[ \Gamma_f \circ ^t\Gamma_g= \alpha \in \operatorname{\mathbf{Cor}}(Y-E,X-D), \]

where $^t\Gamma _g\in \operatorname {\mathbf {Cor}}(Y-E,Y'-E')$ is the transpose of the graph of $g$ and $\Gamma _f\in \operatorname {\mathbf {Cor}}(Y'-E', X-D)$ is the graph of $f$. By definition this follows from the equality

\[ ^t\Gamma_g\times_{Y'-E'} \Gamma_f =\alpha\subset (Y-E)\times (X-D) \]

which one can check easily, noting that $Y'\to \overline {\alpha }$ is an isomorphism over $\alpha$ since $\alpha$ is regular by $(\spadesuit )$. Then we get a commutative diagram

where the top inclusion comes from the inequality $E_{{\operatorname {red}}} \times _Y Y'\geq E'_{{\operatorname {red}}}$ as Cartier divisors on $Y'$ thanks to the semipurity of $F$ (cf. § 1(16)). Hence, it suffices to show $f^*(F(X,D))\subset F(Y',E'_{{\operatorname {red}}})$. By replacing $(Y,E)$ with $(Y',E')$, we may now assume that $\alpha$ is induced by a morphism $f: Y\to X=\mathbf {A}^r\times S$. Then $\alpha$ factors in $\operatorname {\mathbf {\underline {M}Cor}}$ as

\[ (Y,E) \overset{i}{\longrightarrow} (\mathbf{A}^1,0)^{\otimes r}\otimes (Y,\emptyset) \to (\mathbf{A}^1,0)^{\otimes r}\otimes(S,\emptyset), \]

where the first map is induced by the map

\[ i=(pr_{\mathbf{A}^r}\circ f,id_Y): Y \to \mathbf{A}^r\times Y, \]

and the second is induced by

\[ id_{\mathbf{A}^r}\times (pr_S\circ f): \mathbf{A}^r\times Y \to \mathbf{A}^r\times S. \]

Note that $i$ is a section of the projection $\mathbf {A}^r\times Y \to Y$. Thus, we are reduced to showing $i^*(F((\mathbf {A}^1,0)^{\otimes r}\otimes (Y,\emptyset )))\subset F(Y,E_{{\operatorname {red}}})$. By Proposition 3.2 this follows from the following claim.

Claim 4.3.1 Take $a\in F((\mathbf {A}^1,0)^{\otimes r}\otimes (Y,\emptyset ))$. There exists an open neighborhood $U\subset Y$ of the generic point of $E$ such that for every function field $K$ over $k$ and every $\delta =(\delta _0,\dots,\delta _{e-1},\delta _{e})\in \mathrm {mc}(U_K)$ with $\xi :=\delta _{e-1}\in E_K^{(0)}$ and $e=\dim (Y)$, we have

\[ (i^*(a)_K,\gamma)_{Y_K/K,\delta}=0\quad\forall \gamma\in \overline{K}^{\rm M}_e(\mathcal{O}_{Y_K,\xi},\mathfrak{m}_\xi) \]

for the pairing from (3.0.4):

\[ (-,-)_{Y_K/K,\delta}: F(K(Y))\otimes K^{\rm M}_d(K(Y))\to F(K). \]

Proof. After replacing $Y$ by an open neighborhood of the generic point of $E$, we may assume that $Y=\operatorname {Spec}(A)$ is affine and $E_{{\operatorname {red}}}=\operatorname {Spec}(A/(\pi ))$ for $\pi \in A$ and, moreover, that writing

\[ \mathbf{A}^r\times Y=\operatorname{Spec} A[x_1,\dots,x_r],\quad (\mathbf{A}^1,0)^{\otimes r}\otimes(Y,\emptyset) =(\mathbf{A}^r_Y,\{x_1\cdots x_r=0\}), \]

we have

\[ i(Y) =\underset{1\leq i\leq r}{\bigcap} \{x_i-u_i\pi^{m_i}=0\} \quad {\rm with}\ m_i\in \mathbb{Z}_{\geq 0},\; u_i\in A^\times. \]

Let $\delta =(\delta _0,\dots,\delta _e)$ be as in the claim and put $\delta '=(\delta _0,\dots,\delta _{e-1})\in \mathrm {mc}((E_{{\operatorname {red}}})_K)$. Put $\tilde {X}_K=\mathbf {A}^r\times Y_K$ and $F=\{\pi =0\}\subset \tilde {X}_K$. Note $d:=\dim (\tilde {X}_K)=e+r$. Let $z_j$ for $e\leq j\leq d-1$ be the generic point of

\[ Z_j=\underset{1\leq i\leq d-j}{\bigcap} \{x_i-u_i\pi^{m_i}=0\} \subset \tilde{X}_K \]

which lies over $\delta _e$,Footnote 5 and $w_j$ for $e-1\leq j\leq d-2$ be the generic point of

\[ W_j=F\cap Z_{j+1}=\{\pi=x_1=\cdots=x_{d-j-1}=0\} \]

which is contained in the closure of $z_{j+1}$. Note $\dim (Z_j)=\dim (W_j)=j$ and the section $i$ induces isomorphisms

(4.3.2)\begin{equation} {Y_K\simeq Z_e \quad\text{and}\quad (E_{{\operatorname{red}}})_K\simeq W_{e-1}.} \end{equation}

Let $\sigma =(i(\delta '),w_e,\dots,w_{d-2},\eta _1,\nu )\in \mathrm {mc}(\tilde {X}_K)$, where $\nu$ is the generic point of $\tilde {X}_K$ lying over $\delta _e$, $\eta _1$ is the generic point of $D_1=\{x_1=0\}\subset \tilde {X}_K$ contained in the closure of $\nu$, and $i(\delta ')\in \mathrm {mc}(W_{e-1})$ is the image of $\delta '$ under (4.3.2). Take any $\gamma \in \overline {K}^{\rm M}_e(\mathcal {O}_{Y_K,\xi },\mathfrak {m}_\xi )$ and put

(4.3.3)\begin{equation} \beta=\bigg\{\iota(\gamma),\frac{u_1\pi^{m_1}-x_1}{u_1\pi^{m_1}},\dots,\frac{u_r\pi^{m_r}-x_r}{u_r\pi^{m_r}}\bigg\}\in K^{\rm M}_d(\mathcal{O}_{\tilde{X}_K,\nu}), \end{equation}

where $\iota : K^{\rm M}_e(\mathcal {O}_{Y_K,\delta _e}) \to K^{\rm M}_e(\mathcal {O}_{\tilde {X}_K,\nu })$ is induced by the projection $\tilde {X}_K \to Y_K$. For $a\in F((\mathbf {A}^1,0)^{\otimes r}\otimes (Y,\emptyset ))$ and its restriction $a_K\in F((\mathbf {A}^1,0)^{\otimes r}\otimes (Y_K,\emptyset ))$, we have

\begin{align*} 0=(a_K,\beta)_{\tilde{X}_K/K,\sigma} &=-\underset{\substack{\tau\in \tilde{X}_K^{(1)}-\{\eta_1\}\\\tau>w_{d-2}}}{\sum} (a_K,\beta)_{\tilde{X}_K/K,(i(\delta'),w_e,\dots,w_{d-2},\tau,\nu)} \\ &=-(a_K,\beta)_{\tilde{X}_K/K,(i(\delta'),w_e,\dots,w_{d-2},z_{d-1},\nu)} \\ &=\pm ((a_K)_{|Z_{d-1}},\beta_1)_{Z_{d-1}/K,(i(\delta'),w_e,\dots,w_{d-2},z_{d-1})}, \end{align*}

\[ \beta_1=\bigg\{\iota_1(\gamma),\frac{u_2\pi^{m_2}-x_2}{u_2\pi^{m_2}},\dots,\frac{u_r\pi^{m_r}-x_r}{u_r\pi^{m_r}}\bigg\}\in K^{\rm M}_{d-1}(\mathcal{O}_{Z_{d-1},z_{d-1}}), \]

where $\iota _1: K^{\rm M}_e(\mathcal {O}_{Y_K,\delta _e}) \to K^{\rm M}_e(\mathcal {O}_{Z_{d-1},z_{d-1}})$ is induced by the dominant map $Z_{d-1}\to Y_K$ induced by the projection $\tilde {X}_K \to Y_K$. The first equality follows from § 3 (HS3) applied to $D_1\subset \tilde {X}_K$, noting that $\beta$ lies in $\overline {K}^{\rm M}_d(\mathcal {O}_{\tilde {X}_K, \eta _1},\mathfrak {m}_{\eta _1})$ since $(u_1\pi ^{m_1}-x_1)/u_1\pi ^{m_1} \in 1+ x_1 \mathcal {O}_{\tilde {X}_K,\eta _1}$. The second follows from (HS4). The third equality holds since $z_{d-1}$ is the unique $\tau \in \tilde {X}_K^{(1)}-\{\eta _1\}$ such that $\tau >w_{d-2}$ and $(a_K,\beta )_{\tilde {X}_K/K,(i(\delta '),w_e,\dots,w_{d-2},\tau,\nu )}$ may not vanish, which follows from (HS2), noting that $\iota (\gamma )_{|F}=0$. The final equality follows from (HS2). When $r=1$, the last term in the above formula is equal to $((a_K)_{|Y_K},\gamma )_{Y_K/K,\delta }$ by (4.3.2), so that the proof is complete. When $r>1$, we further get

\begin{align*} 0&=((a_K)_{|Z_{d-1}},\beta_1)_{Z_{d-1}/K,(i(\delta'),w_e,\dots,w_{d-2},z_{d-1})}\\ &=-\underset{\substack{\tau\in Z_{d-1}^{(1)}-\{w_{d-2}\}\\\tau>w_{d-3}}}{\sum} ((a_K)_{|Z_{d-1}},\beta_1)_{Z_{d-1}/K,(i(\delta'),w_e,\dots,w_{d-3},\tau,z_{d-1})} \\ &=-((a_K)_{|Z_{d-1}},\beta_1)_{Z_{d-1}/K,(i(\delta'),w_e,\dots,w_{d-3},z_{d-2},z_{d-1})} \\ &=\pm ((a_K)_{|Z_{d-2}},\beta_2)_{Z_{d-2}/K,(i(\delta'),w_e,\dots,w_{d-3},z_{d-2})} , \end{align*}
\[ \beta_2=\bigg\{\iota_2(\gamma),\frac{u_3\pi^{m_3}-x_3}{u_3\pi^{m_3}},\dots,\frac{u_r\pi^{m_r}-x_r}{u_r\pi^{m_r}}\bigg\}\in K^{\rm M}_{d-1}(\mathcal{O}_{Z_{d-2},z_{d-2}}), \]

where $\iota _2: K^{\rm M}_e(\mathcal {O}_{Y_K,\delta _e}) \to K^{\rm M}_e(\mathcal {O}_{Z_{d-2},z_{d-2}})$ is induced by the dominant map $Z_{d-2}\to Y_K$ induced by the projection $\tilde {X}_K \to Y_K$. The above equalities hold by the same arguments as above, except that for the third equality there are a priori two $\tau \in Z_{d-1}^{(1)}-\{w_{d-2}\}$ with $\tau >w_{d-3}$ for which $((a_K)_{|Z_{d-1}},\beta _1)_{Z_{d-1}/K,(i(\delta '),w_e,\dots,w_{d-3},\tau,z_{d-1})}$ may not vanish. One is $z_{d-2}$ and the other is the generic point $\eta _2$ of $Z_{d-1}\cap D_2$ with $D_2= \{x_2=0\}\subset \tilde {X}_K$ which is contained in the closure of $z_{d-1}$. But $((a_K)_{|Z_{d-1}},\beta _1)_{Z_{d-1}/K,(i(\delta '),w_e,\dots,w_{d-3},\eta _2,z_{d-1})}=0$. Indeed, $(a_K)_{|Z_{d-1}}\in F(\operatorname {Spec}(\mathcal {O}_{Z_{d-1},\eta _2}),\eta _2)$ since $Z_{d-1}$ and $D_2$ intersect transversally in $\tilde {X}_K$. Hence, the vanishing follows from (HS3) applied to $Z_{d-1}\cap D_2\subset Z_{d-1}$, noting that $\big ((u_2\pi ^{m_2}-x_2)/u_2\pi ^{m_2}\big )_{|Z_{d-1}} \in 1+ x_2 \mathcal {O}_{Z_{d-1},\eta _2}$ so that $\beta _1\in K^{\rm M}_d(\mathcal {O}_{Z_{d-1},\eta _2},\mathfrak {m}_{\eta _2})$. Repeating the same arguments, we finally get

\[ 0=((a_K)_{|Z_e},\iota_r(\gamma))_{Z_e/K,(i(\delta'),z_e)} =((a_K)_{|Y_K},\gamma)_{Y_K/K,\delta}, \]

where $\iota _r: K^{\rm M}_e(\mathcal {O}_{Y_K,\delta _e}) \to K^{\rm M}_e(\mathcal {O}_{Z_e,z_e})$ is induced by the isomorphism $Z_e\to Y_K$ induced by the projection $\tilde {X}_K \to Y_K$ and the second equality follows from (4.3.2). This completes the proof of the claim and Theorem 4.2.

Definition 4.4 For $F\in \operatorname {\mathbf {\underline {M}NST}}_{\log }$ and an integer $i\geq 0$, consider the association

\[ {H_{\log}^{i}}(-,F) : \operatorname{\mathbf{\underline{M}Cor}}^{{\operatorname{fin}}}_{\rm ls} \to \operatorname{\mathbf{Ab}}\;;\; (X,D) \to H^i(X_{\operatorname{Nis}}, F_{(X,D_{{\operatorname{red}}})}). \]

By the definition this gives a presheaf on $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}_{\rm ls}$, which we call the $i$th logarithmic cohomology with coefficient $F$.

5. Invariance of logarithmic cohomology under blowups

Retain the notation of § 4.

Definition 5.1 Let $\Lambda ^{{\operatorname {fin}}}_{\rm ls}$ be the class of morphisms $\rho : (Y,E) \to (X,D)$ in $\operatorname {\mathbf {\underline {M}Cor}}^{{\operatorname {fin}}}_{\rm ls}$ satisfying the following conditions.

  1. (a) $\rho$ is induced by a proper morphism $\rho : Y\to X$ inducing an isomorphism $Y{\setminus} E\overset {\simeq }\longrightarrow X{\setminus} D$ and $E=\rho ^* D$.

  2. (b) Zariski locally on $X$, $\rho :Y \to X$ is the blowup of $X$ in a smooth center $Z\subset D$ which is normal crossing to $D$.

Here, a smooth $Z$ contained in $D$ is normal crossing to $D$ if, letting $D_1,\dots,D_n$ be the irreducible components of $D$, there exists a subset $I\subset \{1,\dots,n\}$ such that $Z\subset {\bigcap }_{i\in I} D_i$ and $Z$ is not contained in $D_j$ for any $j\not \in I$ and intersects $\sum _{{j\not \in I}} D_j$ transversally. Note that the condition is equivalent to that called strict normal crossing in [Reference Binda, Park and ØstværBPØ22, Definition 7.2.1].

Theorem 5.2 For $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ and $\rho :\mathcal {Y}\to \mathcal {X}$ in $\Lambda ^{{\operatorname {fin}}}_{\rm ls}$, we have

(5.2.1)\begin{equation} \rho^*: {H_{\log}^{i}}(\mathcal{X},F) \cong {H_{\log}^{i}}(\mathcal{Y},F) \quad\forall i\geq 0. \end{equation}

Proof. Write $\mathcal {Y}=(Y,E)$ and $\mathcal {X}=(X,D)$. First we prove the theorem for $i=0$. We may assume that $D$ is reduced and $E=\rho ^*D$. By [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, Proposition 1.9.2 b)], $\rho$ is invertible in $\operatorname {\mathbf {\underline {M}Cor}}$, so that $\rho ^*:F(\mathcal {X})\cong F(\mathcal {Y})$. Since this factors through $F(Y,E_{{\operatorname {red}}})$ by Theorem 4.2, we get (5.2.1) for $i=0$.

To show (5.2.1) for $i>0$, it suffices to prove $R^i\rho _* F_{(Y,E_{{\operatorname {red}}})} =0$. The problem is Nisnevich local, so we may assume that $\rho$ is induced by a blowup $\rho :Y\to X$ in a smooth center $Z\subset D$ normal crossing to $D$. By [Reference Kelly and SaitoKS21, Corollary 9], Nisnevich locally around a point of $Z$, $(X,D)$ is isomorphic to

\[ (\mathbf{A}^c,L_1+\cdots +L_r)\otimes \mathcal{W} \quad {\rm with}\ \mathcal{W}=(W,W^\infty)\in \operatorname{\mathbf{\underline{M}Cor}}_{\rm ls}, \]

where $\mathbf {A}^c=\operatorname {Spec} k[t_1,\dots,t_c]$ with $c={\operatorname {codim}}_z(Z,X)$ and $L_i=V(t_i)$ for $i=1,\dots,r$ with $1\leq r\leq c$, and $Z$ corresponds to $0\times W$. Hence, the theorem follows from the following proposition.

Proposition 5.3 Let $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$ and $\mathcal {W}=(W,W^\infty )\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$. Let $\mathbf {A}^n=\operatorname {Spec} k[t_1,\dots,t_n]$ and put $L_i=V(t_i)$ for $1\leq i\leq n$. Let $\rho : Y\to \mathbf {A}^n$ be the blowup at the origin $0\in \mathbf {A}^n$ and $\tilde {L}_i\subset Y$ be the strict transforms of $L_i$ for $1\leq i\leq n$ and $E=\rho ^{-1}(0)\subset Y$. For any $1\leq r\leq n$, we have

(5.3.1)\begin{equation} R^i\rho_{W*} F_{(Y, \tilde{L}_1 +\cdots +\tilde{L}_r + E)\otimes \mathcal{W}}=0\quad \text{for } i\ge 1, \end{equation}

where $\rho _W:=\rho \times {\operatorname {id}}_W: Y\times W\to \mathbf {A}^2\times W$.

Lemma 5.4 Proposition 5.3 holds for $n=2$.

Proof. The case $r=1$ is proved in [Reference Binda, Rülling and SaitoBRS22, Lemma 2.13] and we show the case $r=2$.Footnote 6 Put $D= L_1+ L_2$. By the case $i=0$ of Theorem 5.2, we get

(5.4.1)\begin{equation} F_{(\mathbf{A}^2,D)\otimes \mathcal{W}} \cong \rho_{W*} F_{(Y, \tilde{L}_1+\tilde{L}_2 + E)\otimes \mathcal{W}}. \end{equation}

Set

\[ \mathcal{F}:=F_{(Y, \tilde{L}_1+\tilde{L}_2 + E)\otimes \mathcal{W}}, \]

and $\mathbf {A}^2_W=\mathbf {A}^2\times W$ with the projection $p:A^2_W \to W$. Since $R^i\rho _{W*}\mathcal {F}$ for $i\ge 1$ is supported in $0\times W$, we have

\begin{align*} R^i\rho_{W*}\mathcal{F}=0&\Longleftrightarrow p_*R^i\rho_{W*}\mathcal{F}=0\\ &\Longleftrightarrow (p_*R^i\rho_{W*}\mathcal{F})_w=0\quad\forall w\in W\\ &\Longleftrightarrow H^0(\mathbf{A}^2_{W_w}, R^i\rho_{W*}\mathcal{F})=0\quad\forall w\in W, \end{align*}

where $W_w$ is the henselization of $W$ at $w$. Hence, it suffices to show $H^0(\mathbf {A}^2_W, R^i\rho _{W*}\mathcal {F})=0$, assuming $W$ is henselian local. Then we have

\[ H^j(\mathbf{A}^2_W, R^i\rho_{W*}\mathcal{F})=0, \quad\forall i,j\ge 1. \]

By (5.4.1) and [Reference Binda, Rülling and SaitoBRS22, Lemma 2.10],

\[ H^i(\mathbf{A}^2_W, \rho_{W*}\mathcal{F})=H^i(\mathbf{A}^2_W, F_{(\mathbf{A}^2, D)\otimes \mathcal{W}})=0. \]

Thus, the Leray spectral sequence yields

\[ H^0(\mathbf{A}_W^2, R^i\rho_{W*}\mathcal{F})= H^i(Y\times W, \mathcal{F}), \quad i\ge 0, \]

and we have to show that this group vanishes for $i\ge 1$. We can write

\[ \mathbf{A}^2=\operatorname{Spec} k[x,y]\quad\text{and}\quad L_1=V(x),\ L_2=V(y) \subset \mathbf{A}^2. \]

Then we have

\[ Y=\operatorname{Proj} k[x,y][S,T]/(xT-yS)\subset \mathbf{A}^2\times \mathbf{P}^1. \]

Denote by

\[ \pi_0: Y\hookrightarrow \mathbf{A}^2\times \mathbf{P}^1 \to \mathbf{P}^1=\operatorname{Proj} k[S,T] \]

the morphism induced by projection, and let $\pi : Y\times W \to \mathbf {P}^1_W$ be its base change. Then $\pi _0$ induces an isomorphism $E\simeq \mathbf {P}^1$, and we have

(5.4.2)\begin{equation} \tilde{L}_1=\pi_0^{-1}(0),\quad \tilde{L}_2=\pi_0^{-1}(\infty). \end{equation}

Set $s=S/T=x/y$ and write

\[ \mathbf{P}^1{\setminus} \{\infty\}= \mathbf{A}^1_s:= \operatorname{Spec} k[s],\quad \mathbf{P}^1{\setminus}\{0\}= \operatorname{Spec} k[\tfrac{1}{s}]. \]

Set $U:=\mathbf {A}^1_s \times W$, $V:=(\mathbf {P}^1{\setminus} \{0\})\times W$ and

\[ \mathcal{U}:=(\mathbf{A}^1_s, 0)\otimes\mathcal{W}, \quad \mathcal{V}:=(\mathbf{P}^1{\setminus}\{0\},\infty) \otimes \mathcal{W}. \]

We have

\[ \pi^{-1}(U)= \mathbf{A}^1_y\times U, \quad \pi^{-1}(V)=\mathbf{A}^1_x\times V, \]

and the restriction of $\pi$ to these open subsets is given by projection. Furthermore, $E\times W\subset Y$ is defined by $y=0$ on $\pi ^{-1}(U)$ and by $x=0$ on $\pi ^{-1}(V)$. In view of (5.4.2), we have

(5.4.3)\begin{equation} \mathcal{F}_{|\pi^{-1}(U)}=F_{(\mathbf{A}^1_y, 0)\otimes \mathcal{U}},\quad \mathcal{F}_{|\pi^{-1}(V)}= F_{(\mathbf{A}^1_x, 0)\otimes \mathcal{V}}. \end{equation}

Thus, [Reference Binda, Rülling and SaitoBRS22, Lemma 2.10] yields

\[ R^j\pi_*\mathcal{F}=0 \quad {\rm for}\ j\ge 1, \]

and it remains to show

(5.4.4)\begin{equation} {H^i(\mathbf{P}^1_W, \pi_*\mathcal{F})=0 \quad {\rm for}\ i\ge 1,} \end{equation}

where $\mathbf {P}^1_W=\mathbf {P}^1\times W$. For this consider the map

\[ a_0: Y \to \mathbf{A}^1_x \times \mathbf{P}^1 \]

which is the closed immersion $Y\hookrightarrow \mathbf {A}^2\times \mathbf {P}^1$ followed by the projection $\mathbf {A}^2\to \mathbf {A}^1_x$. Let $a: Y\times W\to \mathbf {A}^1_x\times \mathbf {P}^1\times W$ be its base change. In view of (5.4.2), the map $a$ induces a morphism in $\operatorname {\mathbf {\underline {M}Cor}}$,

\[ \alpha: (Y,\tilde{L}_1+\tilde{L}_2+E)\otimes \mathcal{W}\to (\mathbf{A}^1_x, 0)\otimes (\mathbf{P}^1,\infty)\otimes \mathcal{W}, \]

which is an isomorphism over $(\mathbf {A}^1_x, 0)\otimes (\mathbf {P}^1\backslash \{0\},\infty )\otimes \mathcal {W}$. Setting

\[ F_1:=\operatorname{\underline{Hom}}({\operatorname{\mathbb{Z}_{{\operatorname{tr}}}}}(\mathbf{A}^1_x, 0), F )\in \operatorname{\mathbf{CI}}^{\tau,{\rm sp}}_{{\operatorname{Nis}}}, \]

it induces a map of Nisnevich sheaves on $\mathbf {P}^1_W$,

\[ \pi_*(\alpha^*): F_{1, (\mathbf{P}^1,\infty)\otimes \mathcal{W}}\to \pi_*\mathcal{F}, \]

which becomes an isomorphism over $(\mathbf {P}^1-\{0\})\times W$. Hence, (5.4.4) follows from

\[ H^i(\mathbf{P}^1_W, F_{1, (\mathbf{P}^1,\infty)\otimes \mathcal{W}})=0 \quad {\rm for}\ i\ge 1, \]

which follows from [Reference SaitoSai20, Theorem 0.6].

Lemma 5.5 Let $N>2$ be an integer and assume that Proposition 5.3 holds for $n< N$. Let $(X,D)\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$ and $Z\subset X$ be a smooth integral closed subscheme with $2\le {\operatorname {codim}}(Z,X)=:c < N$. Assume

\[ D= D_1+\cdots + D_r + D' \quad {with}\ r\leq c , \]

where $D_1,\dots,D_r$ are distinct and reduced irreducible components of $D$ containing $Z$, and $D'$ is an effective divisor on $X$ such that none of the component of $D'$ contains $Z$ and $Z$ is transversal to $|D'|$. Let $\rho : Y\to X$ be the blowup of $X$ in $Z$, let $\tilde {D}_i$ and $\tilde {D}'\subset Y$ be the strict transforms of $D_i$ and $D'$ respectively, and let $E_Z=\rho ^{-1}(Z)$. Then, for all $\mathcal {W}=(W,W^\infty )\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$,

\[ R^i\rho_{W*} F_{(Y, \tilde{D}_1+\cdots + \tilde{D}_r +E_Z + \tilde{D}' )\otimes \mathcal{W}}=0 \quad\text{for } i\ge 1, \]

where $\rho _W: Y\times W \to X\times W$ denotes the base change of $\rho$.

Proof. This proof is adapted from [Reference Binda, Rülling and SaitoBRS22, Lemma 2.14]. The question is Nisnevich local around the points in $Z\times W$. Let $z\in Z\times W$ be a point and set $A:=\mathcal {O}_{X\times W,z}^h$. For $V\subset Y\times W$, we denote $V_{(z)}:= V\times _{X\times W} \operatorname {Spec} A$. By assumption we find a regular system of local parameters $t_1,\ldots, t_m$ of $A$, such that

\[ (D_i\times W)_{(z)}=V(t_i) \quad {\rm for}\ 1\leq i\leq r,\quad (Z\times W)_{(z)}=V(t_1,\ldots, t_c), \]
\[ (D'\times W)_{(z)}=V(t_{c+1}^{e_{c+1}}\cdots t_{m_0}^{e_{m_0}}) \quad {\rm with}\ c+1\leq m_0\le m, \]
\[ (X\times W^\infty)_{(z)}=V(t_{m_0+1}^{e_{m_0+1}}\cdots t_{m_1}^{e_{m_1}}) \quad {\rm with}\ m_0\le m_1\leq m. \]

Letting $K$ be the residue field of $A$, we can choose a ring homomorphism $K\hookrightarrow A$ which is a section of $A \to K$. Then we obtain an isomorphism

\[ K\{t_1,\ldots, t_m\}\xrightarrow{\simeq} A. \]

Let $\rho _1:\widetilde {\mathbf {A}^c}\to \mathbf {A}^c$ be the blowup in $0$. By the above,

\[ \rho_W: (Y,\tilde{D}_1+\cdots + \tilde{D}_r +E_Z + \tilde{D}' )\otimes \mathcal{W}\to (X, D)\otimes \mathcal{W} \]

is Nisnevich locally around $z$ isomorphic over $k$ to the morphism

\[ (\widetilde{\mathbf{A}^c},\tilde{L}_1+\cdots+\tilde{L}_r + E)\otimes \mathcal{W}' \to (\mathbf{A}^c, L_1+\cdots+ L_r)\otimes \mathcal{W}', \]
\[ \bigg(\mathcal{W}'=\bigg(\mathbf{A}_K^{m-c}, \bigg(\prod_{i=c+1}^{m_1} t_{i}^{e_{i}}\bigg)\bigg)\bigg) \]

induced by a map $(\widetilde {\mathbf {A}^c},\tilde {L}_1+\cdots +\tilde {L}_r + E)\to (\mathbf {A}^c, L_1+\cdots + L_r)$ as in Proposition 5.3. Hence, the statement follows from the proposition for $n=c< N$.

Proof of Proposition 5.3 The proof is by induction on $n\geq 2$. The case $n=2$ follows from Lemma 5.4. Assume that $n>2$ and that the proposition is proven for $\mathbf {A}^m$ with $m< n$. For $r=1$, Proposition 5.3 is proved in [Reference Binda, Rülling and SaitoBRS22, Theorem 2.12]. Assume that $r\geq 2$. Let $Z:=L_1\cap L_2\subset \mathbf {A}^n$ and $\tilde {Z}\subset Y$ be the strict transform of $Z$. Denote by $\rho ': Y'\to Y$ the blowup of $Y$ in $\tilde {Z}$, let $\tilde {L}_i', E'\subset Y'$ be the strict transforms of $\tilde {L}, E$ respectively, and let $E''=(\rho ')^{-1}(\tilde {Z})$. Note that $\tilde {Z}=\tilde {L}_1\cap \tilde {L}_2$ intersecting transversally with $\tilde {L}_3+\cdots +\tilde {L}_r+ E$ and ${\operatorname {codim}}(\tilde {Z}, Y)=2$. Hence, by Lemma 5.5,

\[ R^i\rho'_{W*} F_{(Y', \tilde{L}_1'+\cdots+\tilde{L}_r' + E' + E'')\otimes \mathcal{W}}=0\quad {\rm for}\ i\ge 1. \]

Since Theorem 5.2 has been proved for $i=0$, we have

\[ \rho'_* F_{(Y', \tilde{L}_1'+\cdots+\tilde{L}_r' + E' + E'')\otimes \mathcal{W}} = F_{(Y, \tilde{L}_1+\cdots+\tilde{L}_r + E)\otimes \mathcal{W}}. \]

Hence, we obtain

(5.5.1)\begin{equation} {R^i\rho_{W*} F_{(Y, \tilde{L}_1+\cdots+\tilde{L}_r + E)\otimes \mathcal{W}}= R^i(\rho\rho')_{W*} F_{(Y', \tilde{L}_1'+\cdots+\tilde{L}_r' + E'+ E'')\otimes \mathcal{W}}. } \end{equation}

Denote by $\sigma :\hat {Y} \to \mathbf {A}^n$ the blowup in $Z$, let $\hat {L}_i\subset \hat {Y}$ be the strict transform of $L_i$, and let $\Xi =\sigma ^{-1}(Z)$. By Lemma 5.5 we get

(5.5.2)\begin{equation} {R^i\sigma_{W*} F_{(\hat{Y}, \hat{L}_1+\cdots+\hat{L}_r+ \Xi)\otimes \mathcal{W}}=0 \quad {\rm for}\ i\ge 1.} \end{equation}

Denote by $\sigma ':\hat {Y}'\to \hat {Y}$ the blowup in $\hat {Z}=\sigma ^{-1}(0)\subset \Xi$, let $\hat {L}'_i,\; \Xi '\subset \hat {Y}'$ be the strict transforms of $\hat {L}_i,\; \Xi$ respectively, and let $\Xi ''=\sigma '^{-1}(\hat {Z})$. Note that $\hat {Z}\subset \hat {L}_3\cap \cdots \cap \hat {L}_n\cap \Xi$ and ${\operatorname {codim}}(\hat {Z}, \hat {Y})= n-1$ and $\hat {Z}$ intersects transversally with $\hat {L}_1+ \hat {L}_2$. Thus, by Lemma 5.5 and the case $i=0$ of Theorem 5.2, we obtain

(5.5.3)\begin{equation} {R\sigma'_{W*}F_{(\hat{Y}',\hat{L}_1'+\cdots+\hat{L}_r'+ \Xi'+\Xi'')\otimes \mathcal{W}}= F_{(\hat{Y}, \hat{L}_1+\cdots+\hat{L}_r+ \Xi)\otimes \mathcal{W}}.} \end{equation}

Finally, by [Reference Binda, Rülling and SaitoBRS22, Lemma 2.15], there is an isomorphism of $\mathbf {A}^n\times W$-schemes

(5.5.4)\begin{equation} { (\hat{Y}',\hat{L}_1',\dots,\hat{L}_r,\Xi',\Xi'')\cong (Y', \tilde{L}_1',\dots,\tilde{L}_r',E',E'').} \end{equation}

Altogether we obtain, for $i\ge 1$,

\begin{align*} R^i\rho_{W*} F_{(Y, \tilde{L}_1+\cdots+\tilde{L}_r + E)\otimes \mathcal{W}} &= R^i(\rho\rho')_{W*} F_{(Y', \tilde{L}_1'+\cdots+\tilde{L}_r' + E' + E'')\otimes \mathcal{W}}, & & \text{by (5.5.1)},\\ & = R^i(\sigma\sigma')_{W*} F_{(\hat{Y}',\hat{L}_1'+\cdots+\hat{L}_r'+ \Xi'+\Xi'')\otimes \mathcal{W}}, & & \text{by (5.5.4)},\\ & = R^i\sigma_{W*} F_{(\hat{Y}, \hat{L}_1+\cdots+\hat{L}_r+ \Xi)\otimes \mathcal{W}}, & & \text{by (5.5.3)},\\ & = 0, & & \text{by (5.5.2)}. \end{align*}

This completes the proof of the proposition.

Remark 5.6 For simplicity, we write

\[ {H_{\log}^{i}}(-,F)={H_{\log}^{i}}(-,\underline{\omega}^{\operatorname{\mathbf{CI}}} F)\quad {\rm for}\ F\in {\operatorname{\mathbf{RSC}}}_{{\operatorname{Nis}}}. \]

By [Reference Rülling and SaitoRS21a, Corollary 6.8], if ${\operatorname {ch}}(k)=0$ and $F=\Omega ^i$, we have

\[ {H_{\log}^{i}}(-,\Omega^i)=H^i(X,\Omega^i(\log |D|) \quad {\rm for}\ (X,D)\in \operatorname{\mathbf{\underline{M}Cor}}_{\rm ls}\!. \]

Hence, ${H_{\log}^{i}}(-,F)$ for $F\in {\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$ is a generalization of cohomology of sheaves of logarithmic differentials.

6. Relation with logarithmic sheaves with transfers

In this section we use the same notation as [Reference Binda, Park and ØstværBPØ22].

Let ${\operatorname {l\mathbf {Sm}}}$ be the category of log smooth and separated $\mathrm {fs}$ log schemes of finite type over the base field $k$ and ${\operatorname {\mathbf {Sm}l\mathbf {Sm}}}\subset {\operatorname {l\mathbf {Sm}}}$ be the full subcategory consisting of objects whose underlying schemes are smooth over $k$. Let ${\operatorname {l\mathbf {Cor}}}$ be the category with the same objects as ${\operatorname {l\mathbf {Sm}}}$ and whose morphisms are log correspondences defined in [Reference Binda, Park and ØstværBPØ22, Definition 2.1.1]. Let ${\operatorname {l\mathbf {Cor}}}_{{\operatorname {\mathbf {Sm}l\mathbf {Sm}}}}\subset {\operatorname {l\mathbf {Cor}}}$ be the full subcategory consisting of all objects in ${\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$.

Let ${\operatorname {\mathbf {PSh}}^{\rm ltr}}$ be the category of additive presheaves of abelian groups on ${\operatorname {l\mathbf {Cor}}}$ and ${\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}\subset {\operatorname {\mathbf {PSh}}^{\rm ltr}}$ be the full subcategory consisting of those $\mathcal {F}$ whose restrictions to ${\operatorname {l\mathbf {Sm}}}$ are dividing Nisnevich sheaves (see [Reference Binda, Park and ØstværBPØ22, Definition 3.1.4]). It is shown in [Reference Binda, Park and ØstværBPØ22, Theorem 1.2.1 and Proposition 4.7.5] that ${\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}$ is a Grothendieck abelian category and there is an equivalence of categories

(6.0.1)\begin{equation} {{\operatorname{\mathbf{Shv}}^{\rm ltr}_{\rm dNis}} \simeq {\operatorname{\mathbf{Shv}}^{\rm ltr}_{\rm dNis}}({\operatorname{\mathbf{Sm}l\mathbf{Sm}}}),} \end{equation}

where the right-hand side denotes the full subcategory of the category ${\operatorname {\mathbf {PSh}}^{\rm ltr}}({\operatorname {\mathbf {Sm}l\mathbf {Sm}}})$ of additive presheaves of abelian groups on ${\operatorname {l\mathbf {Cor}}}_{{\operatorname {\mathbf {Sm}l\mathbf {Sm}}}}$ consisting of those $\mathcal {F}$ whose restrictions to ${\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$ are dividing Nisnevich sheaves.

We now construct a functor

(6.0.2)\begin{equation} {\mathcal{L}og : \operatorname{\mathbf{\underline{M}NST}}_{\log} \to {\operatorname{\mathbf{Shv}}^{\rm ltr}_{\rm dNis}}.} \end{equation}

For $\mathfrak {X}=(X,\mathcal {M})\in {\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$, we put $\mathfrak {X}^{\rm MP}=(X, \partial \mathfrak {X})$, where $\partial \mathfrak {X}\subset X$ is the closed subscheme consisting of the points where the log structure $\mathcal {M}$ is not trivial. By [Reference Binda, Park and ØstværBPØ22, Lemma A.5.10], $\partial \mathfrak {X}$ with reduced structure is a normal crossing divisor on $X$, so that we can view $\mathfrak {X}^{\rm MP}$ as an object of $\operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$. For $F\in \operatorname {\mathbf {\underline {M}PST}}_{\log }$ and $\mathfrak {X}\in {\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$, we put

(6.0.3)\begin{equation} {F^{\log}(\mathfrak{X})= F(\mathfrak{X}^{\rm MP}).} \end{equation}

Take $\mathfrak {Y} \in {\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$ and $\alpha \in {\operatorname {l\mathbf {Cor}}}(\mathfrak {Y},\mathfrak {X})$. By [Reference Binda, Park and ØstværBPØ22, Definition 2.1.1 and Remark 2.1.2(iii)], we have

\[ \alpha\in \operatorname{\mathbf{\underline{M}Cor}}^{{\operatorname{fin}}}((Y,n\cdot \partial \mathfrak{Y}),(X,\partial \mathfrak{X}))\quad\text{for some } n>0, \]

where $n\cdot \partial \mathfrak {Y}\hookrightarrow Y$ is the $n$th thickening of $\partial \mathfrak {Y}\hookrightarrow Y$. By the assumption $F\in \operatorname {\mathbf {\underline {M}PST}}_{\log }$, the induced map

\[ F^{\log}(\mathfrak{X})=F(\mathfrak{X}^{\rm MP}) \overset{\alpha^*}{\longrightarrow} F(Y,n\cdot \partial \mathfrak{Y}) \]

factors through $F^{\log }(\mathfrak {Y})=F(Y,\partial \mathfrak {Y})\subset F(Y,n\cdot \partial \mathfrak {Y})$ and we get a map

\[ \alpha^{*\log} : F^{\log}(\mathfrak{X})\to F^{\log}(\mathfrak{Y}). \]

Moreover, for a map $\gamma : F\to G$ in $\operatorname {\mathbf {\underline {M}PST}}_{\log }$, the diagram

is obviously commutative. Hence, the assignment $\mathcal {X}\to F^{\log }(\mathcal {X})$ gives an object $F^{\log }$ of ${\operatorname {\mathbf {PSh}}^{\rm ltr}}({\operatorname {\mathbf {Sm}l\mathbf {Sm}}})$ and we get a functor

(6.0.4)\begin{equation} { \mathcal{L}og : \operatorname{\mathbf{\underline{M}PST}}_{\log} \to {\operatorname{\mathbf{PSh}}^{\rm ltr}}({\operatorname{\mathbf{Sm}l\mathbf{Sm}}}),\quad F \to F^{\log}.} \end{equation}

By the definitions of sheaves ([Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, Definition 1], [Reference Binda, Park and ØstværBPØ22, Definition 3.1.4] and [Reference Kahn, Miyazaki, Saito and YamazakiKMSY21a, Proposition 1.9.2]), this induces a functor

\[ \operatorname{\mathbf{\underline{M}NST}}_{\log} \to {\operatorname{\mathbf{Shv}}^{\rm ltr}_{\rm dNis}}({\operatorname{\mathbf{Sm}l\mathbf{Sm}}}) \]

which induces the desired functor (6.0.2) using (6.0.1). By the construction, for $F\in \operatorname {\mathbf {\underline {M}NST}}_{\log }$ and $\mathfrak {X}\in {\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$ with $\mathcal {X}=\mathfrak {X}^{\rm MP}\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$, we have

(6.0.5)\begin{equation} {H^i_{\operatorname{Nis}}(X,F_{\mathcal{X}}) = H^i_{s\mathrm{Nis}}(\mathfrak{X},F^{\log})\; (F^{\log}=\mathcal{L}og(F)),} \end{equation}

where the right-hand side is the cohomology for the strict Nisnevich topology (see [Reference Binda, Park and ØstværBPØ22, Definition 4.3.1]).

Theorem 6.1 For $F\in \operatorname {\mathbf {CI}}^{\tau,{\rm sp}}_{{\operatorname {Nis}}}$, $F^{\log }=\mathcal {L}og(F)\in {\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}$ is strictly ${\overline {\square }}$-invariant in the sense of [Reference Binda, Park and ØstværBPØ22, Definition 5.2.2]. For $\mathfrak {X}\in {\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$ with $\mathcal {X}=\mathfrak {X}^{\rm MP}\in \operatorname {\mathbf {\underline {M}Cor}}_{\rm ls}$, we have a natural isomorphism

(6.1.1)\begin{equation} { H^i_{\operatorname{Nis}}(X,F_{\mathcal{X}})\simeq \operatorname{Hom}_{\operatorname{\mathbf{logDM}}^{{\operatorname{eff}}}}(M(\mathfrak{X}),F^{\log}[i]),} \end{equation}

where $\operatorname {\mathbf {logDM}}^{{\operatorname {eff}}}$ is the triangulated category of logarithmic motives defined in [Reference Binda, Park and ØstværBPØ22, Definition 5.2.1].

Proof. Let $\mathfrak {X}_{\rm div}^{\rm Sm}$ be the category of log modifications $\mathfrak {Y} \to \mathfrak {X}$ such that $\mathfrak {Y}\in {\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$ (see [Reference Binda, Park and ØstværBPØ22, Definition A.11.12]) and $\mathfrak {X}_{\rm divsc}^{\rm Sm}\subset \mathfrak {X}_{\rm div}^{\rm Sm}$ be the full subcategory given by those maps $\mathfrak {Y}\to \mathfrak {X}$ that are isomorphic to compositions of log modifications along smooth centers (see [Reference Binda, Park and ØstværBPØ22, Definitions 4.4.4 and A.14.10]). We have isomorphisms

\begin{align*} H^i_{\operatorname{Nis}}(X,F_{\mathcal{X}}) &\overset{(6.0.5)}{\simeq} H^i_{s\mathrm{Nis}}(\mathfrak{X},F^{\log}) \overset{(*1)}{\simeq} \mathop {\varinjlim}_{\mathfrak{Y}\in \mathfrak{X}_{\rm divsc}^{\rm Sm}} H^i_{s\mathrm{Nis}}(\mathfrak{Y},F^{\log}) \\ &\overset{(*2)}{\simeq} \mathop {\varinjlim}_{\mathfrak{Y}\in \mathfrak{X}_{\rm div}^{\rm Sm}} H^i_{s\mathrm{Nis}}(\mathfrak{Y},F^{\log}) \overset{(*3)}{\simeq} H^i_{d\mathrm{Nis}}(\mathfrak{X},F^{\log}), \end{align*}

where $(*2)$ follows from [Reference Binda, Park and ØstværBPØ22, Corollary 4.4.5] and $(*3)$ from [Reference Binda, Park and ØstværBPØ22, Theorem 5.1.8], and $(*1)$ is a consequence of Theorem 5.2 in view of (6.0.5) and the fact that a log modification of $\mathfrak {X}=(X,\mathcal {M})\in {\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$ along smooth center is induced Zariski locally by a blowup of $X$ in an intersection of irreducible components of $\partial \mathfrak {X}$ so that it corresponds to a morphism in $\Lambda ^{{\operatorname {fin}}}_{\rm ls}$ from Definition 5.1.

Hence, the strict ${\overline {\square }}$-invariance of $F^{\log }$ follows from [Reference SaitoSai20, Theorem 0.6]. Finally, (6.1.1) follows from [Reference Binda, Park and ØstværBPØ22, Proposition 5.2.3].

We now consider the composite functor

\[ \mathcal{L}og': {\operatorname{\mathbf{RSC}}}_{\operatorname{Nis}} \overset{\underline{\omega}^{\operatorname{\mathbf{CI}}}}{\longrightarrow} \operatorname{\mathbf{CI}}^{\tau,{\rm sp}}_{{\operatorname{Nis}}} \overset{\mathcal{L}og}{\longrightarrow} \operatorname{\mathbf{CI}}^{\rm ltr}_{\rm dNis}, \]

where $\operatorname {\mathbf {CI}}^{\rm ltr}_{\rm dNis}\subset {\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}$ is the full subcategory consisting of strictly ${\overline {\square }}$-invariant objects. By [Reference Binda and MericiBM12, Theorem 5.7], $\operatorname {\mathbf {CI}}^{\rm ltr}_{\rm dNis}$ is a Grothendieck abelian category.

Lemma 6.2 $\mathcal {L}og$ and $\mathcal {L}og'$ have the same essential image.

Proof. This follows directly from the construction and Corollary 2.6(3).

In what follows, we let

(6.2.1)\begin{equation} { \mathcal{L}og : {\operatorname{\mathbf{RSC}}}_{\operatorname{Nis}} \to \operatorname{\mathbf{CI}}^{\rm ltr}_{\rm dNis}\;:\; F\to F^{\log} } \end{equation}

denote $\mathcal {L}og'$ defined as above. By (6.0.3), we have

(6.2.2)\begin{equation} { F^{\log}(X,\mathrm{triv})= F(X)\quad {\rm for}\ F\in {\operatorname{\mathbf{RSC}}}_{{\operatorname{Nis}}},\; X\in \operatorname{\mathbf{Sm}},} \end{equation}

where $(X,\mathrm {triv})$ denotes the log scheme with the trivial log structure.

Theorem 6.3 $\mathcal {L}og$ is exact and fully faithful.

Proof. First we prove the full faithfulness. Faithfulness follows from (6.2.2). Let $F,G\in {\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$ and $\gamma : F^{\log }\to G^{\log }$ be a map in ${\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}$. By (6.2.2) it induces maps $\gamma _X: F(X) \to G(X)$ for all $X\in \operatorname {\mathbf {Sm}}$. They are compatible with the action of $\operatorname {\mathbf {Cor}}$ since by [Reference Binda, Park and ØstværBPØ22, Example 2.1.3(3)],

\[ \operatorname{\mathbf{Cor}}(Y,X)= {\operatorname{l\mathbf{Cor}}}(Y,\mathrm{triv}),(X,\mathrm{triv}))\quad {\rm for}\ X,Y\in \operatorname{\mathbf{Sm}}\!. \]

Thus, $\gamma _X$ for $X\in \operatorname {\mathbf {Sm}}$ give a map $\gamma _{{\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}}:F\to G$ in ${\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$. To see $\mathcal {L}og(\gamma _{{\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}})=\gamma$, it suffices by (6.0.1) to show that $\mathcal {L}og(\gamma _{{\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}})$ and $\gamma$ induce the same map $F^{\log }(\mathfrak {X}) \to G^{\log }(\mathfrak {X})$ for $\mathfrak {X}\in {\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$. If $\mathfrak {X}$ has the trivial log structure, this follows immediately from the construction of $\gamma _{{\operatorname {\mathbf {RSC}}}}$. The general case follows from this in view of the commutative diagram

where $j^*$ are induced by the natural map $(X\backslash \partial \mathfrak {X},\mathrm {triv})\to \mathfrak {X}$ of log schemes and are injective by the construction and the semipurity of $\underline {\omega }^{\operatorname {\mathbf {CI}}} F$. This completes the proof of the full faithfulness.

Next we show the exactness of $\mathcal {L}og$. It suffices to show the following claim.

Claim 6.3.1 Given an exact sequence $0\to F\to G\to H\to 0$ in ${\operatorname {\mathbf {RSC}}}_{{\operatorname {Nis}}}$, the induced sequence

\[ 0\to F^{\log}(\mathfrak{X})\to G^{\log}(\mathfrak{X})\to H^{\log}(\mathfrak{X})\to 0 \]

is exact for every $\mathfrak {X}\in {\operatorname {\mathbf {Sm}l\mathbf {Sm}}}$ with $X$ henselian local.

Indeed, by the definition of $\mathcal {L}og$, this is reduced to the exactness of

\[ 0\to \underline{\omega}^{\operatorname{\mathbf{CI}}} F(\mathfrak{X}^{\rm MP})\to \underline{\omega}^{\operatorname{\mathbf{CI}}} G(\mathfrak{X}^{\rm MP})\to \underline{\omega}^{\operatorname{\mathbf{CI}}} H(\mathfrak{X}^{\rm MP})\to 0, \]

which follows from Corollary 2.6(2). This completes the proof of Theorem 6.3.

Acknowledgements

The author would like to thank Kay Rülling, F. Binda and A. Merici for many valuable discussions and comments. He is also grateful to A. Merici to whom he owes crucial ideas for § 2. The author also thanks the referee for a very careful reading and numerous suggestions improving this paper.

Footnotes

The author is supported by JSPS KAKENHI grant no. 20H01791.

1 This heuristic viewpoint is manifested in [Reference Rülling and SaitoRS21a, Theorem 2].

2 In fact it is defined in [Reference Binda, Park and ØstværBPØ22, Definition 5.2.1] as the localization of the homotopy category of complexes in ${\operatorname {\mathbf {Shv}}^{\rm ltr}_{\rm dNis}}$ with respect to a ${\overline {\square }}$-local descent model structure.

3 It is a logarithmic analogue of Voevodsky's strict $\mathbf {A}^1$-invariance.

4 The assumption is necessary to use [Reference Rülling and SaitoRS21a, Corollary 6.8] proved in the case ${\operatorname {ch}}(k)=0$. We expect that it can be dispensed with by using a forthcoming work of K. Rülling extending [Reference Rülling and SaitoRS21a, Corollary 6.8] to the case ${\operatorname {ch}}(k)>0$.

5 Although $Y$ is assumed to be irreducible, $Y_K$ may not be so and possibly a finite product of schemes essentially smooth over $k$, noting that $k$ is perfect.

6 The following argument is adopted from [Reference Binda, Rülling and SaitoBRS22, Lemma 2.13], but the present case is easier.

References

Binda, F. and Merici, A., Connectivity and purity for logarithmic motives, J. Inst. Math. Jussieu, to appear. Preprint (2012), axXiv:2012.08361.Google Scholar
Binda, F., Park, D. and Østvær, P. A., Triangulated categories of logarithmic motives over a field, Astérisque 433 (2022).Google Scholar
Binda, F., Rülling, K. and Saito, S., On the cohomology of reciprocity sheaves, Forum Math. Sigma 10 (2022), E72.CrossRefGoogle Scholar
Kahn, B., Miyazaki, H., Saito, S. and Yamazaki, T., Motives with modulus, I: modulus sheaves with transfers, Épijournal Géom. Algébrique 5 (2021), article 1.Google Scholar
Kahn, B., Miyazaki, H., Saito, S. and Yamazaki, T., Motives with modulus, II: modulus sheaves with transfers for proper modulus pairs, Épijournal Géom. Algébrique 5 (2021), article 2.Google Scholar
Kahn, B., Saito, S. and Yamazaki, T., Reciprocity sheaves, I, Compos. Math. 152 (2016), 18511898.CrossRefGoogle Scholar
Kahn, B., Saito, S. and Yamazaki, T., Reciprocity sheaves, II, Homology Homotopy Appl. 24 (2022), 7191.CrossRefGoogle Scholar
Kelly, S. and Saito, S., Smooth blowups square for motives with modulus, Bull. Pol. Acad. Sci. Math. 69 (2021), 97106.CrossRefGoogle Scholar
Mazza, C., Voevodsky, V. and Weibel, C., Lecture notes on motivic cohomology, Clay Mathematics Monographs, vol. 2 (American Mathematical Society, Providence, RI, 2006).Google Scholar
Rülling, K. and Saito, S., Reciprocity sheaves and their ramification filtrations, J. Inst. Math. Jussieu (2021), doi:10.1017/S1474748021000074.Google Scholar
Rülling, K. and Saito, S., Ramification theory of reciprocity sheaves, I: Zariski-Nagata purity, J. Reine Angew., to appear. Preprint (2021), arXiv:2111.01459.Google Scholar
Rülling, K. and Saito, S., Ramification theory of reciprocity sheaves, II: Higher local symbols, Preprint (2021), arXiv:2111.13373.Google Scholar
Rülling, K., Sugiyama, R. and Yamazaki, T., Tensor structures in the theory of modulus presheaves with transfers, Math. Z. 300 (2022), 929977.Google Scholar
Saito, S., Purity of reciprocity sheaves, Adv. Math. 365 (2020), 107067.CrossRefGoogle Scholar
Grothendieck, A., Cohomologie local des faisceaux cohérents et théorèmes de Lefschetz locaux et globaux (SGA2), Advanced Studies in Pure Mathematics, vol. 2 (North-Holland Publishing Co., Amsterdam, 1968).Google Scholar
Voevodsky, V., Triangulated categories of motives over a field, in Cycles, transfers and motivic cohomology theories, eds. Friedlander, E., Suslin, A. and Voevodsky, V., Annals of Mathematics Studies, vol. 143 (Princeton University Press, Princeton, NJ, 2000), 188238.Google Scholar