1. Introduction
Let k be a field of characteristic zero. Following the vision of Beilinson, Deligne, Grothendieck and others, there should exist an abelian category of mixed motives
$\mathcal{MM}(k)$
, the target of the universal cohomology theory

on k-varieties in the sense that any other reasonable cohomology theory on k-varieties would factor uniquely through
$M^*$
. This is beyond the reach of the current technology. However, two constructions have almost all the required properties to provide a category of mixed motives. The first one is Voevodsky, Levine, Hanamura and others’ triangulated category of geometric motives
$\mathrm{DM}_\mathrm{gm}(k,\mathbb{Q})$
[Reference Voevodsky, Suslin and FriedlanderVSF00], a candidate for the bounded derived category of mixed motives. The second is the abelian category of Nori motives
$\mathcal{M}_\mathrm{Nori}(k)$
[Reference FakhruddinFak00], a candidate for
$\mathcal{MM}(k)$
. These categories have realisation functors factoring the known cohomology theories of varieties. Indeed, if k is a subfield of
$\mathbb{C}$
, Huber [Reference HuberHub00], Levine [Reference LevineLev05] and Nori have constructed a Hodge realisation functor

to the derived category of polarisable rational mixed Hodge structures that realises the Hodge cohomology of varieties and factors through the derived category of the
$\mathbb{Q}$
-linear abelian category of motives constructed by Nori thanks to the work of Harrer [Reference HarrerHar16] and Choudhury and Gallauer [Reference Choudhury and de SouzaCGAdS17]. When k is arbitrary they also constructed a
$\ell$
-adic realisation functor with values in the derived category of
$\ell$
-adic Galois representations.
There exist relative versions of the categories of Voevodsky and Nori. The triangulated category of constructible rational étale motives
$\mathrm{DM}^{\mathrm{\acute{e}t}}_c(-)$
constructed by Ayoub [Reference AyoubAyo14a] and Cisinski and Déglise [Reference Cisinski and DégliseCD16] gives a triangulated category of étale motivic sheaves generalising Voevodsky’s category, and the abelian category of perverse Nori motives
${\mathcal{M}_{\mathrm{perv}}}(-)$
constructed by Ivorra and S. Morel [Reference Ivorra and MorelIM22] gives a category of motivic sheaves generalising the abelian category of Nori. Both settings have a six-functor formalism, the tensor product on perverse Nori motives being constructed by Terenzi [Reference TerenziTer24]. Note that another approach based on constructible sheaves and using Nori’s ideas has been considered by Arapura in [Reference ArapuraAra13, Reference ArapuraAra23] but his construction does not include all six operations. If X is a quasi-projective smooth k-variety and k is a subfield of
$\mathbb{C}$
, Ivorra [Reference IvorraIvo16] constructed a realisation functor

to Saito’s derived category of mixed Hodge modules [Reference SaitoSai90]. It also factors through the derived category of perverse Nori motives and computes the relative Hodge homology of X-schemes. Unfortunately, Ivorra’s functor has no obvious compatibility with the six operations that exist on
$\mathrm{DM}^\mathrm{\acute{e}t}_c(-)$
and
$\mathrm{D}^b(\mathrm{MHM}(-))$
. The reason for this difficulty is that the functor is built by constructing an explicit complex
$K^\bullet\in\mathrm{Ch}^b(\mathrm{MHM}(X))$
that computes the relative homology sheaf
$f_!f^!\mathbb{Q}_X\in\mathrm{D}^b(\mathrm{MHM}(X))$
of a smooth affine X-scheme
$f:Y\to X$
, and that we do not know how to lift the six operations explicitly on complexes: a priori they are only defined over the derived category.
On the other hand, we know since Robalo’s thesis [Reference RobaloRob15] that the
$\infty$
-category
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
(which is a stable
$\infty$
-category lifting
$\mathrm{DM}^\mathrm{\acute{e}t}_c(X)$
) has an universal property that enables one to construct realisation functors to any stable
$\infty$
-category that has reasonable properties. Indeed, he shows that
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)=\mathrm{Ind}\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
is the target of the universal symmetric monoidal functor from
$\mathrm{Sm}_X$
to a stable
$\mathbb{Q}$
-linear presentably symmetric monoidal
$\infty$
-category, satisfying
$\mathbb{A}^1$
-invariance, étale hyperdescent and
$\mathbb{P}^1$
-stability. The work of Drew and Gallauer [Reference Drew and GallauerDG22] shows that this universal property works very well in families, so that any contravariant functor
$\mathcal{C}$
from finite-type k schemes to symmetric monoidal
$\mathbb{Q}$
-linear presentable
$\infty$
-categories that satisfies non-effective étale descent and
$\mathbb{A}^1$
-invariance, together with
$\mathbb{P}^1$
-stability and smooth base change, receives a natural transformation from
$\mathcal{DM}^\mathrm{\acute{e}t}(-)$
compatible with smooth base change functoriality. Moreover, Cisinski and Déglise (see also [Reference AyoubAyo07, Scholie 1.4.2]) proved in [Reference Cisinski and DégliseCD19, Theorem 4.4.25] that the full subcategory of geometric objects
$\mathcal{C}_{\mathrm{gm}}$
of any such functor
$\mathcal{C}$
has the six operations, and that any natural transformation
$\mathcal{DM}^{\mathrm{\acute{e}t}}\to\mathcal{C}$
between such functors would induce a natural transformation between the categories of geometric objects that commutes with the six operations. In other words, to construct a family of realisation functors
$\mathrm{Hdg}^*:\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to\mathcal{D}^b(\mathrm{MHM}(X))$
compatible with the six operations, it suffices to lift the six operations on mixed Hodge modules to the
$\infty$
-categorical setting. This is what we do.
There is another approach to the Hodge realisation of étale motives considered by Brad Drew in [Reference DrewDre18]. For each finite-type k-scheme X, he constructs a new category
$\mathcal{DH}_c(X)$
of motivic Hodge modules that has a tautological functor
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to \mathcal{DH}_c(X)$
commuting with all the operations. If
$X = \mathrm{Spec} k$
is the spectrum of a field, then
$\mathcal{DH}_c(\mathrm{Spec} k)\hookrightarrow \mathcal{D}^b(\mathrm{MHS}^p_k)$
embeds in the derived category of mixed Hodge structures. Moreover, one can compute Deligne cohomology as a Hom-group in
$\mathcal{DH}_c(X)$
. However, it was unclear how to relate
$\mathcal{DH}_c(X)$
with
$\mathcal{D}^b(\mathrm{MHM}(X))$
as it is hard to construct an adequate t-structure on
$\mathcal{DH}_c(X)$
. Using our result on the existence of the Hodge realisation and a description of geometric mixed Hodge modules as modules in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
(see Proposition 6.16 for the case of Nori motives), it should not be too hard to prove that Drew’s category
$\mathcal{DH}_c(X)$
embeds fully faithfully in the derived category of mixed Hodge modules. We plan to pursue this in a sequel to this paper.
Our main result is the following theorem.
Theorem 1.1 (Theorem 6.1, Theorem 6.5 and Proposition 6.9). The six operations on the categories
$\mathrm{D}^b(\mathrm{MHM}(-))$
and
$\mathrm{D}^b({\mathcal{M}_{\mathrm{perv}}}(-))$
constructed by Saito, Ivorra, Morel and Terenzi admit
$\infty$
-categorical lifts and are defined over any finite-type k-scheme. For every finite-type k-scheme X there exists an
$\infty$
-functor

and if k is a subfield of
$\mathbb{C}$
then there exist
$\infty$
-functors

and

The three functors commute with the six operations and we have
$R_H\circ\mathrm{Nor}^*\simeq \mathrm{Hdg}^*$
. Moreover,
$\mathrm{Nor}^*$
is compatible with the Betti and
$\ell$
-adic realisation functors,
$R_H$
is t-exact and all functors are compatible with weights.
How to give
$\infty$
-categorical lifts of triangulated functors? In general this is a hard question. One general solution would be to use the formalism developed by Liu and Zheng in [Reference Liu and ZhengLZ15]. In our particular case we have a simpler method. Indeed, it is very easy to construct
$\infty$
-categorical lifts of derived functors. Most of the six operations are not t-exact (even on one side) for the perverse t-structure. However, pullback functors are t-exact for the constructible t-structure, that is, the t-structure whose heart behaves like the abelian category of constructible sheaves. In [Reference NoriNor02], Nori proves that the triangulated category of cohomologically constructible sheaves
$\mathrm{D}^b_c(X(\mathbb{C}),\mathbb{Q})$
on the
$\mathbb{C}$
-points of an algebraic variety X over a subfield of
$\mathbb{C}$
is the derived category of its constructible heart. We show that one can adapt his argument to mixed Hodge modules and perverse Nori motives, and obtain the following result.
Theorem 1.2 (Corollary 4.19). Let X be a quasi-projective k-variety. Denote by
$\mathrm{MHM}_c(X)$
(respectively, by
${\mathcal{M}_{\mathrm{ct}}}(X)$
) the heart of the constructible t-structure on
$\mathrm{D}^b(\mathrm{MHM}(X))$
(respectively, on
$\mathrm{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
). The canonical
$\infty$
-functors

if k is a subfield of
$\mathbb{C}$
, and

for an arbitrary field k, are equivalences of
$\infty$
-categories.
This theorem enables us to make the handy change of variables
$\mathcal{D}^b({\mathcal{M}_{\mathrm{ct}}}(X))\simeq\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
. As pullback functors
$f^*$
are easily seen to be
$\infty$
-functors this implies they are derived functors of their restrictions to the constructible hearts, so that the functoriality of the constructible heart, which can be written by hand, immediately gives the functoriality of the derived
$\infty$
-category! This consideration was the starting idea of this paper.
Let
$\mathcal{DN}(X)=\mathrm{Ind}\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
be the indization of the bounded derived
$\infty$
-category of
${\mathcal{M}_{\mathrm{perv}}}(X)$
. This is a compactly generated stable presentably symmetric monoidal
$\infty$
-category whose category of compact objects is
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
. The functor
$\mathrm{Nor}^*:\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
extends to a functor

which preserves colimits, and hence has a right adjoint
$\mathrm{Nor}_*$
. For each finite-type k-scheme, set
$\mathscr{N}_X:=\mathrm{Nor}_*\mathrm{Nor}^*\mathbb{Q}_X\in\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
. As
$\mathrm{Nor}_*$
is automatically lax symmetric monoidal, we have that
$\mathscr{N}_X$
is an
$\mathbb{E}_\infty$
-algebra of
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
. Using the proof of the same result for the Betti realisation in [Reference AyoubAyo22, Theorem 1.93], we prove the following proposition.
Proposition 1.3 (Proposition 6.16). For each scheme X of finite type over some field k of characteristic zero, denote by
$p_X:X\to\mathrm{Spec}\mathbb{Q}$
the unique morphism. Then
$p_X^*\mathscr{N}_\mathbb{Q}\simeq\mathscr{N}_X$
and the natural functor

is an equivalence of categories that commutes with pullback functors.
Although we do not go into the details, the same arguments would prove that the
$\infty$
-category
$\mathcal{DH}^\mathrm{geo}(X)\subset\mathrm{Ind}\mathcal{D}^b(\mathrm{MHM}(X))$
of objects of geometric origin in the indization of the derived category of mixed Hodge modules is also the category of modules over some algebra
$\mathscr{H}_X$
in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
(see [Reference TubachTub24] for more details). This shows that Drew’s constructions in [Reference DrewDre18] indeed provide a module presentation of objects of geometric origin in mixed Hodge modules (and even, by considering motives enriched in mixed Hodge structures, a presentation of a bigger full subcategory of all mixed Hodge modules as modules in a motivic category).
As the functor
$(\mathcal{A},\mathcal{C})\mapsto\mathrm{Mod}_\mathcal{A}(\mathcal{C})$
that sends a symmetric monoidal
$\infty$
-category pointed at an algebra object
$\mathcal{A}\in\mathrm{CAlg}(\mathcal{C})$
to the
$\infty$
-category of modules over
$\mathcal{A}$
preserves colimits, we obtain the following corollary.
Corollary 1.4 (Corollary 6.18). The
$\infty$
-functor
$X\mapsto\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
has a unique extension
$\mathcal{DN}_c$
to quasi-compact and quasi-separated schemes of characteristic zero such that for every projective system
$(X_i)_i$
of such schemes with affine transitions morphisms the natural functor

is an equivalence of
$\infty$
-categories.
Now that a well-behaved comparison functor
$\mathrm{Nor}^*:\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to\mathcal{DN}_c(X)$
is available, one wants to see to what extent we can compare both categories. We have a result in that direction, conditional on the existence of a motivic t-structure on
$\mathrm{DM}_\mathrm{gm}(K)$
for every field K of characteristic zero, a deep conjecture that we know implies all standard conjectures in characteristic zero [Reference BeilinsonBei12]. Moreover, by the work of Bondarko in [Reference BondarkoBon15, Theorem 3.1.4], the existence of motivic t-structure for all fields of characteristic zero implies that for all finite-type k-schemes, we have a perverse and a constructible t-structure on
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
.
Theorem 1.5 (Theorem 6.22). Assume that a motivic t-structure exists for all fields of characteristic zero. Let X be a finite-type k scheme for such a field k. Then the heart of the perverse t-structure of
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
is canonically equivalent to the category
${\mathcal{M}_{\mathrm{perv}}}(X)$
of perverse Nori motives. Moreover, the functor
$\mathrm{Nor}^* : \mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
is an equivalence of stable
$\infty$
-categories. This implies that
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
is the derived category of both its perverse and its constructible heart.
In particular, over a field we would have
$\mathrm{DM}_\mathrm{gm}(k,\mathbb{Q})\simeq\mathrm{D}^b(\mathcal{MM}(k))$
, showing that the ‘even more optimistic’ expectation of [Reference AndréAnd04, 21.1.8] is in fact no more optimistic than expecting the motivic t-structure to exist.
2. Organisation of the paper
After some recollections on motivic constructions in § 3, in § 4 we review Nori’s proof in [Reference NoriNor02] that the category of cohomologically constructible complexes of sheaves on the
$\mathbb{C}$
-points of a quasi-projective variety X is equivalent to the derived category of the abelian category of constructible sheaves and show that it gives a proof that the derived categories of the perverse and constructible hearts on the categories of mixed Hodge modules (or of perverse Nori motives) are equivalent. The only new idea here was to replace local systems with dualisable objects and cohomology of the global sections with
$\mathrm{Hom}_{\mathrm{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))}(\mathbb{Q}_X,-[q])$
. The main point of the proof is to show effaceability of cohomology of the constant object on an affine n-space.
We divide § 5 into two parts. In the first part we prove general results about the functoriality of the functor
$\mathcal{D}^b(-)$
sending an abelian category to its bounded derived
$\infty$
-category. In the second part we prove that the six operations on mixed Hodge modules and on perverse Nori motives can be lifted to
$\infty$
-categorical functors of stable
$\infty$
-categories.
In § 6, we use what has been proven on
$\infty$
-categorical enhancements to construct the realisation functors. We then prove that Nori motives are modules over an algebra in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
. The arguments more or less follow Ayoub’s proof of the similar statement for the Betti realisation in [Reference AyoubAyo22]. Finally, we prove the result related to the t-structure conjecture. The argument is the repeated use of the universal properties of
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
and
${\mathcal{M}_{\mathrm{perv}}}(X)$
that forces the Nori realisation functor to be an equivalence when
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
has a perverse t-structure.
2.1 Forthcoming and future work
In [Reference TubachTub24], using the
$\infty$
-categorical enhancement of mixed Hodge modules, we will extend them to Artin stacks, together with the operations, t-structures and weights. In joint work with Raphaël Ruimy [Reference Ruimy and TubachRT24] we construct an integral version of the derived category of perverse Nori motives. This gives an Abelian category of motivic sheaves with integral coefficients over any finite-dimensional scheme of characteristic zero.
2.2 Notations and conventions
Throughout the paper, k is a field of characteristic zero. When k is a subfield of
$\mathbb{C}$
, if X is a finite-type k-scheme, we denote by
$\mathrm{D}_c^b(X,\mathbb{Q})$
the category of constructible complexes of sheaves on
$X(\mathbb{C})$
. By [Reference BeilinsonBei87], it is also the derived category of the abelian category of perverse sheaves
$\mathrm{Perv}(X,\mathbb{Q})$
defined in [Reference Beilinson, Bernstein and DeligneBBD82]. For a general field k, we denote by
$\mathrm{D}^b_c(X,\mathbb{Q}_\ell)$
the category of constructible bounded complexes of
$\mathbb{Q}_\ell$
-adic étale sheaves on X as in [Reference Hemo, Richarz and ScholbachHRS23].
We use the language of
$\infty$
-categories as developed in [Reference LurieLur18a] by Lurie, after Joyal’s work on quasi-categories. We fix an universe and call an
$\infty$
-category ‘small’ if its mapping spaces and core are in this universe. We denote by
$\Delta$
the simplicial category. We denote by
$\mathrm{Cat}_\infty$
the
$\infty$
-category of small
$\infty$
-categories and by
$\mathrm{Pr}^L$
the
$\infty$
-category
$\infty$
-category of presentable
$\infty$
-categories. If
$\mathcal{C}$
is a symmetric monoidal
$\infty$
-category we denote by
$\mathrm{CAlg}(\mathcal{C})$
the commutative algebras in
$\mathcal{C}$
, and if
$A\in\mathcal{C}$
is a commutative algebra, we denote by
$\mathrm{Mod}_A(\mathcal{C})$
the category of modules over A in
$\mathcal{C}$
. For example, the category of
$H\mathbb{Q}$
-modules in spectra
$\mathrm{Mod}_\mathbb{Q} := \mathrm{Mod}_{H\mathbb{Q}}(\mathrm{Sp})$
is the unbounded derived category of the abelian category of
$\mathbb{Q}$
-vector spaces. We always try to put the index
$\mathcal{C}$
when mentioning Hom, as in
$\mathrm{Hom}_\mathcal{C}(A,B)$
. When
$\mathcal{C}$
is an ordinary category, this is the Hom set, and when
$\mathcal{C}$
is an
$\infty$
-category this is the
$\pi_0$
of the mapping space
$\mathrm{Map}_\mathcal{C}(A,B)$
. In a stable
$\infty$
-category
$\mathcal{C}$
, the mapping spectra will be denoted by
$\mathrm{map}_\mathcal{C}(-,-)$
.
3. Recollections
3.1 Categories of étale motives
Ayoub [Reference AyoubAyo10, Reference AyoubAyo14a], Cisinski and Déglise [Reference Cisinski and DégliseCD19, Reference Cisinski and DégliseCD16] and Voevodsky [Reference VoevodskyVoe02] have constructed triangulated categories of motivic sheaves that are the relative versions of the triangulated category of geometric motives
$\mathrm{DM}_{\mathrm{gm}}(k)$
. In this paper, we are interested in the étale versions. We use the stable
$\infty$
-categorical construction due to Robalo [Reference RobaloRob15, Section 2.4]. Also, we only consider
$\mathbb{Q}$
-linear categories unless otherwise specified.
Let S be a scheme. One starts with the category
$\mathrm{Sm}_S$
of smooth S-schemes, and considers the
$\infty$
-category
$\mathrm{PSh}(\mathrm{Sm}_S,\mathrm{Sp})$
of presheaves of spectra on it. The category of rational effective étale motivic sheaves is the full subcategory of
$\mathrm{PSh}(\mathrm{Sm}_S,\mathrm{Sp})$
whose objects are the presheaves
$\mathcal{F}$
that are
$\mathbb{Q}$
-linear (i.e., whose image lands in
$\mathcal{D}(\mathbb{Q})\simeq\mathrm{Mod}_{H\mathbb{Q}}\subset \mathrm{Sp}$
),
$\mathbb{A}^1$
-invariant (i.e., the natural map
$\mathcal{F}(X)\to\mathcal{F}(\mathbb{A}^1_X)$
is an equivalence for any smooth S-scheme X), and satisfy étale hyper-descent (i.e., for any étale hypercover
$U_\bullet\to X$
of a smooth S-scheme X, the natural map
$\mathcal{F}(X)\to \lim_{\Delta}\mathcal{F}(U_n)$
is an equivalence). It is a symmetric monoidal stable
$\infty$
-category, the tensor product being inherited from the monoidal structure of
$\mathrm{Sm}_S$
given by the fiber product. Inverting the object
$\mathbb{Q}(1):=\mathrm{cofib}(S\xrightarrow{1}\mathrm{Gm}_S)$
for the tensor product gives the category that we will denote by
$\mathcal{DM}^{\mathrm{\acute{e}t}}(S)$
and call the category of Voevodsky (étale) motives.
In symbols, we have

This is a stable presentably symmetric monoidal
$\infty$
-category.
We will denote by
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(S)$
the full subcategory of
$\mathcal{DM}^\mathrm{\acute{e}t}(S)$
whose objects are compact objects (i.e., the M in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(S)$
such that
$\mathrm{Hom}_{\mathcal{DM}^{\mathrm{\acute{e}t}}}(M,-)$
commutes with small filtered colimits). If X is a smooth S-scheme and j is an integer, we denote by M(X)(j) the image of X under the Yoneda functor
$\mathrm{Sm}_S\to\mathcal{DM}^\mathrm{\acute{e}t}(S)$
, tensored by
$\mathbb{Q}(1)^{\otimes j}$
. As we work rationally, one can show [Reference AyoubAyo14a, Proposition 8.3] that the subcategory of compact objects coincides with the idempotent complete stable subcategory generated by motives of the form
$\mathrm{M}(X)(-n)$
for X a smooth S-scheme and
$n\in\mathbb{N}$
. As
$\mathcal{DM}^{\mathrm{\acute{e}t}}(S)$
is compactly generated, we have
$\mathrm{Ind}\mathcal{DM}^{\mathrm{\acute{e}t}}_c(S)=\mathcal{DM}^{\mathrm{\acute{e}t}}(S)$
. Moreover, if
$S =\mathrm{Spec} k$
is the spectrum of a field, then there is a natural equivalence
$\mathrm{ho}(\mathcal{DM}^{\mathrm{\acute{e}t}}_c(\mathrm{Spec} k))\simeq \mathrm{DM}_\mathrm{gm}(k,\mathbb{Q})$
between the homotopy category of étale motives and the classical triangulated category of geometric Voevodsky motives over a field.
Let S be a scheme. The categories
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
for X ranging through S-schemes can be put together to give a functor
$\mathrm{Sch}_S^\mathrm{op}\to\mathrm{CAlg}(\mathrm{Pr}^L)$
from the category of finite-type S-schemes to the
$\infty$
-category of presentable symmetric monoidal
$\infty$
-categories (see [Reference RobaloRob14, Section 9.1]). Moreover, this functor
$\mathcal{DM}^{\mathrm{\acute{e}t}}$
is a coefficient system in the sense of Drew and Gallauer (see Definition 7.1), and the
$\infty$
-category of constructible objects
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(-)$
is part of a six-functor formalism (see [Reference Cisinski and DégliseCD19, Section A.5] for a detailed definition), which includes the fact that all pullbacks
$f^*$
have right adjoints
$f_*$
, that there exists exceptional functoriality consisting of an adjoint pair
$(f_!,f^!)$
, such that
$f^!\simeq f^*$
for f étale and
$p_*\simeq p_!$
for p proper. If one replaces the étale topology by the Nisnevich topology in the definition of
$\mathcal{DM}^\mathrm{\acute{e}t}(S)$
given in (3.1), one obtains
$\mathcal{SH}_\mathbb{Q}(S)$
, the rationalisation of the stable motivic homotopy category
$\mathcal{SH}(S)$
, which can be obtained by further replacing in (3.1) the
$\infty$
-category
$\mathcal{D}(\mathbb{Q})$
with the
$\infty$
-category Sp of spectra and which is the stabilisation of the
$\mathbb{A}^1$
-homotopy category introduced in [Reference Morel and VoevodskyMV99]. Except for the computation at
$S= \mathrm{Spec}(k)$
, the results stated in this paragraph and the previous one also hold for
$\mathcal{SH}(S)$
and
$\mathcal{SH}_\mathbb{Q}(S)$
.
When S is of finite type over a subfield of the complex numbers, Ayoub constructed in [Reference AyoubAyo10] the Betti realisation of
$\mathcal{DM}^{\mathrm{\acute{e}t}}(S)$
with values in the derived category of sheaves on the analytification
$S^{an}$
of S which restricts to
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(S)$
and then lands into the stable category
$\mathcal{D}_c^b(S(\mathbb{C}),\mathbb{Q})$
of bounded complexes of sheaves with constructible cohomology. This category is both the derived category of the abelian category perverse sheaves [Reference BeilinsonBei87] and of the abelian category of constructible sheaves [Reference NoriNor02]. One can find in [Reference AyoubAyo14a] and [Reference Cisinski and DégliseCD16] the construction of the
$\ell$
-adic realisation functors on étale motives with values in the derived category of
$\ell$
-adic sheaves. An
$\infty$
-categorical version of this functor can be found in [Reference Richarz and ScholbachRS20, 2.1.2]. This is a functor
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(S)\to\mathcal{D}_c^b(S,\mathbb{Q}_\ell)$
, the latter category being enhanced using condensed coefficients in [Reference Hemo, Richarz and ScholbachHRS23] where it is denoted by
$\mathcal{D}_{\mathrm{cons}}(S,\mathbb{Q}_\ell)$
. The
$\ell$
-adic realisation functor also has a version on the big category
$\mathcal{DM}^{\mathrm{\acute{e}t}}(S)$
by taking indization. All the realisation functors commute with the six operations.
3.2 Perverse Nori motives
Let X be a quasi-projective k-scheme with k a field of characteristic zero. In [Reference Ivorra and MorelIM22], Ivorra and Morel introduced an abelian category of perverse Nori motives
${\mathcal{M}_{\mathrm{perv}}}(X)$
. The construction of the category
${\mathcal{M}_{\mathrm{perv}}}(X)$
is as follows.
Let X be a quasi-projective k-scheme. Pick your favourite prime number
$\ell$
. We can compose the
$\ell$
-adic realisation functor

with the perverse cohomology functor

to obtain a homological functor
${\ ^{\mathrm{p}}\mathrm{H}}^0_\ell\colon \mathcal{DM}_c(X)\to \mathrm{Perv}(X,\mathbb{Q}_\ell).$
Let
$\mathcal{R}(\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X))$
be the abelian category of coherent modules over
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
: it is the full subcategory of additive presheaves
$\mathcal{F}:\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)^\mathrm{op}\to\mathrm{Ab}$
that fit in exact sequences

where
$M\to N$
is a morphism in
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
.
The functor
$\mathrm{H}^0_\ell$
factors through
$\mathcal{R}(\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X))$
, giving a functor

Denote by
$Z_\ell$
the kernel of
$\widetilde{{\ ^{\mathrm{p}}\mathrm{H}}^0_\ell}$
, that is, the objects mapping to zero. The category of perverse Nori motives
${\mathcal{M}_{\mathrm{perv}}}(X)$
is the quotient
$\mathcal{R}(\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X))/Z_\ell$
. It comes with a universal functor
$\mathrm{H}_\mathrm{univ}\colon \mathcal{DM}^\mathrm{\acute{e}t}_c(X)\to{\mathcal{M}_{\mathrm{perv}}}(X)$
.
By construction,
${\mathcal{M}_{\mathrm{perv}}}(X)$
has the following universal property: any homological functor
$H:\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to\mathcal{A}$
to an abelian category
$\mathcal{A}$
, such that for all
$M\in\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
we have
$H(M)=0$
whenever
${\ ^{\mathrm{p}}\mathrm{H}}^0_\ell(M)=0$
, factors uniquely through the universal functor

A priori this category depends on the chosen
$\ell$
, but one can show, using a continuity argument, that this is not the case. Indeed, for finite-type fields (or more generally, for fields embeddable into
$\mathbb{C}$
), the Betti realisation can also be used to define
${\mathcal{M}_{\mathrm{perv}}}(X)$
and the comparison between the
$\ell$
-adic realisation functor and the Betti realisation functor (see, for example, the proof of [Reference AyoubAyo24, Proposition 6.10.] that works over any finite-type
$\mathbb{C}$
-scheme) gives that the two constructions of perverse Nori motives agree.
In fact, every realisation functor existing on
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
gives us a realisation functor on
${\mathcal{M}_{\mathrm{perv}}}(X)$
: we have the
$\ell$
-adic realisation
$R_\ell$
and the Betti realisation
$R_B$
. Ivorra and Morel have proven that
${\mathcal{M}_{\mathrm{perv}}}(\mathrm{Spec} k)$
coincides with the category of cohomological motives constructed by Nori, and thus also has an Hodge realisation functor.
Ivorra and Morel show in [Reference Ivorra and MorelIM22, Theorem 5.1] that the triangulated derived category of the abelian category of perverse motives
$\mathrm{D}^b({\mathcal{M}_{\mathrm{perv}}}(-))$
is part of a homotopical 2-functor in the sense of Ayoub [Reference AyoubAyo07, Definition 1.4.1]. This means that they constructed four of the six operations, namely for f a morphism of quasi-projective varieties we have
$f^*,f_*,f_!,f^!$
, together with Verdier duality
$\mathbb{D}$
, verifying
$f^!\simeq \mathbb{D}\circ f^*\circ\mathbb{D}$
and
$f_!\simeq\mathbb{D}\circ f_*\circ\mathbb{D}$
. In his PhD thesis Terenzi [Reference TerenziTer24] constructed the remaining two operations: tensor product and internal
$\mathscr{{H om}}\,$
. This proves that
$\mathrm{D}^b({\mathcal{M}_{\mathrm{perv}}}(-))$
is part of a full six-functor formalism. The six operations commute with the realisation functors
$R_\ell$
and
$R_B$
making them morphisms of symmetric monoidal stable homotopical 2-functors. The construction of the operations on perverse Nori motives is very similar to Saito’s construction of mixed Hodge modules.
3.3 Mixed Hodge modules
Let X be a separated finite-type
$\mathbb{C}$
-scheme. In [Reference SaitoSai90], Saito constructed an abelian category
$\mathrm{MHM}(X)$
of mixed Hodge modules. He also constructed the six operations on
$\mathrm{D}^b(\mathrm{MHM}(-))$
. There is a faithful exact functor
$\mathrm{MHM}(X)\to \mathrm{Perv}(X^\mathrm{an},\mathbb{Q})$
to the abelian category of perverse sheaves on the analytification of X. This induces a functor
$\mathrm{D}^b(\mathrm{MHM}(X))\to\mathrm{D}^b(\mathrm{Perv}(X^\mathrm{an},\mathbb{Q}))\to\mathrm{D}^b_c(X^\mathrm{an},\mathbb{Q})$
and gives a natural transformation
$\mathrm{D}^b(\mathrm{MHM}(-))\to\mathrm{D}^b_c(-,\mathbb{Q})$
that commutes with the six operations. If
$X=\mathrm{Spec} \mathbb{C}$
, then there is a natural equivalence
$\mathrm{MHM}(\mathrm{Spec} \mathbb{C})\simeq\mathrm{MHS}^p_\mathbb{Q}$
between polarisable mixed Hodge structures and mixed Hodge modules. One of the most interesting properties of the category of mixed Hodge modules is that its lisse objects, that is, the objects such that the underlying perverse sheaf is a shift of a locally constant sheaf, are exactly the polarisable variations of mixed Hodge structures. We will not go into details about the construction of Saito’s category, and mostly need to know that the functor taking a complex of mixed Hodge modules to its underlying complex of perverse sheaves is conservative and commutes with the operations.
4. Adapting Nori’s argument to perverse Nori motives
4.1 Hypothesis on the coefficient system
Let k be a field of characteristic zero. For this section we will denote by
$\mathrm{Var}_k$
either the category of quasi-projective k-varieties or that of separated reduced k-schemes of finite type, and we will call objects of
$\mathrm{Var}_k$
varieties. Choose your favourite
$\mathbb{Q}$
-linear triangulated category of coefficients
$\mathrm{D}:\mathrm{Var}_k^\mathrm{op}\to \mathrm{Triang}$
with a six-functor formalism (by this we mean a symmetric monoidal stable homotopy 2-functor as in Ayoub [Reference AyoubAyo07]), and a conservative realisation functor
$R:\mathrm{D}\to\mathrm{D}^b_c$
compatible with the operations, where
$\mathrm{D}^b_c$
is either the derived category of constructible sheaves on the analytification (if k is a subfield of
$\mathbb{C}$
), or the derived category of
$\ell$
-adic constructible sheaves. For
$X\in\mathrm{Var}_k$
, we will refer to the elements of
$\mathrm{D}(X)$
as ‘motives’. For example, D could be the derived category of perverse Nori motives ([Reference Ivorra and MorelIM22, Theorem 5.1] and [Reference TerenziTer24, Main theorem]) or of mixed Hodge modules [Reference SaitoSai90, Theorem 0.1]. Note that in the mixed Hodge modules situation the realisation is nothing more that the functor forgetting the structure of a mixed Hodge module, and is called ‘taking the
$\mathbb{Q}$
-structure’ by Saito. We assume that for each variety X the category
$\mathrm{D}(X)$
has a perverse t-structure for which R is t-exact when
$\mathrm{D}^b_c(X)$
is endowed with its perverse t-structure of heart the abelian category of perverse sheaves
$\mathrm{Perv}(X)$
.
Recall the following definition.
Definition 4.1. The constructible t-structure on
$\mathrm{D}(X)$
is the unique t-structure on
$\mathrm{D}(X)$
that makes all pullback functors t-exact. Explicitly, it is the t-structure given by

and

That this indeed gives a t-structure can be proven by gluing, using that over a smooth variety, a lisse object (see Definition 4.2) is positive (respectively, negative) for the constructible t-structure if and only it is positive (respectively, negative) for the perverse t-structure when shifted by the dimension.
Denote by
${\mathcal{M}_{\mathrm{ct}}}(X)$
the heart of the constructible t-structure on
$\mathrm{D}(X)$
. As R is t-exact and conservative, all known t-exactness results about the six functors are true in D. We will call elements of
${\mathcal{M}_{\mathrm{ct}}}(X)$
constructible motives. Denote by
$\mathscr{{H om}}\,(-,-)$
and
$\mathbb{Q}_X$
the internal Hom and unit object of
$\mathrm{D}(X)$
.
We assume that
$\mathrm{D}(X)$
has enough structure so that there exists a natural triangulated functor
$\mathrm{D}^b({\mathcal{M}_{\mathrm{ct}}}(X))\to\mathrm{D}(X)$
extending the inclusion of the constructible heart (e.g., if
$\mathrm{D}(X)$
is a derived category as shown by Beilinson, Bernstein, Deligne and Gabber in [Reference Beilinson, Bernstein and DeligneBBD82, section 3.1] or more generally if
$\mathrm{D}(X)$
is the homotopy category of some stable
$\infty$
-category
$\mathcal{D}(X)$
as proven in [Reference Bunke, Cisinski, Kasprowski and WingesBCKW24, Remark 7.4.13]), and we assume that the natural functor
$\mathrm{D}^b({\mathcal{M}_{\mathrm{ct}}}(\mathrm{Spec}k))\to\mathrm{D}(\mathrm{Spec} k)$
is an equivalence. This is true for perverse Nori motives and mixed Hodge modules.
From now on in this section by the ‘t-structure’ we mean the constructible t-structure, unless specified otherwise.
Definition 4.2. Let X be a k-variety. A motive L is a lisse object if its realisation R(L) is a local system, that is, a locally constant bounded complex of finite-dimensional
$\mathbb{Q}$
-vector spaces. We denote by
$\mathrm{D}^{liss}(X)$
the full subcategory of lisse objects in
$\mathrm{D}(X)$
.
We will denote by
$\mathbb{Q}_X\in{\mathcal{M}_{\mathrm{ct}}}(X)$
the unit object for the tensor structure. Its realisation is the constant sheaf.
Remark 4.3. Note that as
$\mathrm{D}(X)$
is closed as a tensor category, given an object
$M\in\mathrm{D}(X)$
, we have a natural candidate for the strong dual of M: it is
$\mathscr{{Hom}}\,(M,\mathbb{Q}_X)$
. Therefore, an object M is dualisable if and only if the map

obtained by the
$\otimes$
-
$\mathscr{{Hom}}\,$
adjunction from the map
$\mathrm{ev}_M\otimes \mathrm{Id}_M : \mathscr{{Hom}}\,(M,\mathbb{Q}_X)\otimes M\otimes M\to \mathbb{Q}\otimes M\simeq M$
, is an isomorphism. We will denote by
$M^\vee = \mathscr{{Hom}}\,(M,\mathbb{Q}_X)$
the dual of M when M is dualisable.
Lemma 4.4. Let X be a k-variety. An object
$K\in \mathrm{D}(X)$
is lisse if and only if it is dualisable. In particular, for every object
$N\in \mathrm{D}(X)$
, there exists a dense open on which N is dualisable. If an object
$L\in{\mathcal{M}_{\mathrm{ct}}}(X)$
is lisse, then its dual
$L^\vee$
is also an object of
${\mathcal{M}_{\mathrm{ct}}}(X)$
.
Proof. For the first point, dualisable objects in
$\mathrm{D}^b_c(X,\mathbb{Q})$
are exactly local systems by Ayoub [Reference AyoubAyo22, Lemma 1.24] in the analytic case, and by [Reference Hemo, Richarz and ScholbachHRS23, Proposition 7.6] together with [Reference Martini and WolfMW22, Corollary 7.4.12.] in the
$\ell$
-adic case. As the realisation functor is conservative, we can test if the map (4.1) is an isomorphism after applying the realisation functor. If
$L\in{\mathcal{M}_{\mathrm{ct}}}(X)$
is dualisable, the functor
$\mathscr{{Hom}}\,(L,-) = -\otimes L^\vee$
is t-exact because it is left and right adjoint to the exact functor
$-\otimes L$
. Therefore,
$L^\vee=\mathscr{H{ om}}\,(L,\mathbb{Q}_X)\in{\mathcal{M}_{\mathrm{ct}}}(X)$
lies in the heart.
The second point follows from the fact that after realisation any constructible motive is locally constant on a stratification, hence, on some dense open subset.
4.2 Cohomology of motives over an affine variety
Definition 4.5. (i) An additive functor
$F:\mathcal{A}\to \mathcal{B}$
between abelian categories is called effaceable if for any
$M\in \mathcal{A}$
there is an injection
$M\to N$
in
$\mathcal{A}$
such that the induced map
$F(M)\to F(N)$
is the zero map.
(ii) An object
$M\in \mathrm{D}(X)$
is called admissible if the functor
$\mathrm{Hom}_{\mathrm{D}(X)}(M,-[q]) : {\mathcal{M}_{\mathrm{ct}}}(X)\to \mathrm{Vect}_\mathbb{Q}$
is effaceable for every
$q>0$
.
Lemma 4.6. Let X be a variety and M be a constructible motive on
$\mathbb{A}^1_X$
. Assume that there is a smooth open subset
$U\subset \mathbb{A}^1_X$
on which M is lisse and such that the restriction of
$\pi:\mathbb{A}^1_X\to X$
to
$Z = \mathbb{A}^1_X\setminus U$
is finite and surjective. Assume that
${M}_{\mid {Z}} = 0$
. Then there exists
$N\in {\mathcal{M}_{\mathrm{ct}}}(\mathbb{A}^1_X)$
such that
$\pi_*N = 0$
and that M injects into N.
Proof. Assume that k is a subfield of
$\mathbb{C}$
. Then the following assertion holds. Denote by
$p_i:\mathbb{A}^2_X\to \mathbb{A}^1_X$
the projections, and by
$\Delta:\mathbb{A}^1_X\to \mathbb{A}^2_X$
the diagonal. The map
$\alpha : p_1^*M\to \Delta_* M$
in
${\mathcal{M}_{\mathrm{ct}}}(X)$
(both functors are t-exact) obtained by adjunction from
$\mathrm{id}:\Delta^*p_1^*M\simeq M\to M$
is surjective because it is surjective after realisation. We therefore have an exact sequence in
${\mathcal{M}_{\mathrm{ct}}}(\mathbb{A}^1_X)$
,

Let
$N:=\mathrm{H}^1(p_2)_*\ker \alpha$
. Then the long exact sequence of cohomology obtained after applying
$(p_2)_*$
to the above sequence gives a map
$M \simeq (p_2)_*\Delta_* M \simeq \mathrm{H}^0((p_2)_*\Delta_* M) \to N$
, which is injective by [Reference NoriNor02, Proposition 2.2].
Moreover, Nori [Reference NoriNor02] proves that for all
$p\geqslant 0$
we have
$\mathrm{H}^p(\pi_*N)=0$
, which is equivalent to
$\pi_*N=0$
.
If k is too big to be embedded in
$\mathbb{C}$
, one may use [Reference BarrettBar21, Proposition 3.2] and the
$\ell$
-adic realisation, noting that the construction of N is the same.
Nori’s method gives in our setting a slightly less good result than in the case of constructible sheaves. It will thankfully be sufficient for us.
Theorem 4.7 (Cf. [Reference NoriNor02, Theorem 1]). For every
$n\in\mathbb{N}$
, the motive
$\mathbb{Q}_{\mathbb{A}^n_k}\in{\mathcal{M}_{\mathrm{ct}}}(\mathbb{A}^n_k)$
is admissible.
Proof. We argue as in Nori’s paper.
We proceed by induction on n to prove the theorem. First the case of a point. If
$n=0$
and
$M\in {\mathcal{M}_{\mathrm{ct}}}(\mathrm{Spec} k)$
, as
$\mathrm{D}(\mathrm{Spec} k)= \mathrm{D}^b({\mathcal{M}_{\mathrm{ct}}}(\mathrm{Spec} k))$
, the functors
$\mathrm{Hom}_{\mathrm{D}(\mathrm{Spec} k)}(M,-[q])$
are effaceable by [Reference Beilinson, Bernstein and DeligneBBD82, Proposition 3.1.16].
Suppose that
$n\geqslant 1$
and that the result is known for
$n-1$
. Let
$M\in {\mathcal{M}_{\mathrm{ct}}}(\mathbb{A}^n_k)$
and
$q>0$
. Let
$f\in k[T_1,\dots,T_n]$
be a polynomial such that M is lisse when restricted to the open subset
$U=D(f)\subset \mathbb{A}^n_k$
. Up to a change of coordinates, we may assume that the projection
$\pi:\mathbb{A}^n_k\to\mathbb{A}^{n-1}_k$
on the
$n-1$
first coordinates gives a finite map
$\pi_{|V(f)}:V(f)\to \mathbb{A}^{n-1}_k$
when restricted to
$Z=V(f)$
. Denote by j and i the inclusions of U and Z in
$\mathbb{A}^n_k$
.
By Lemma 4.6,
$j_!{M}_{\mid {U}}$
is a submotive of a constructible motive N’ such that
$\pi_*N'=0$
. Taking N to be the pushout of the maps
$j_!{M}_{\mid {U}}\to M$
and
$j_!{M}_{\mid {U}}\to N'$
we get a morphism of exact sequences

with
$\gamma$
also injective.
Also, as
$\pi_*N' = 0$
, we see that
$\pi_*N\to \pi_*i_*{M}_{\mid {Z}}=({\pi}_{\mid {Z}})_*{M}_{\mid {Z}}$
is an isomorphism because of the distinguished triangle
$N'\to N\to i_*M_Z\xrightarrow{+1}$
. As
${\pi}_{\mid {Z}}$
is finite, its pushforward
$(\pi_{\mid Z})_*$
is t-exact, hence,
$\pi_*N\in {\mathcal{M}_{\mathrm{ct}}}(\mathbb{A}^{n-1}_k)$
.
By the induction hypothesis, we can find an injection
$g:\pi_*N\to K$
with K a constructible motive such that
$\mathrm{Hom}_{\mathrm{D}(\mathbb{A}^{n-1}_k)}(\mathbb{Q}_{\mathbb{A}^{n-1}_k},g[q]) = 0$
. Let L be the pushout of the counit
$\pi^*\pi_*N \to N$
and the injection
$\pi^*g:\pi^*\pi_*N\to \pi^*K$
(note that
$\pi^*$
is t-exact). We get a morphism of exact sequences.

The unit map
$\pi_*N \to \pi_*\pi^*\pi_*N$
is an isomorphism by
$\mathbb{A}^1$
-invariance. Therefore, in the diagram above the left and the right vertical maps become isomorphisms after applying
$\pi_*$
, thus this is also the case for the middle map
$\iota:\pi^*K\to L$
.
Using the adjunction
$(\pi^*,\pi_*)$
and the fact that
$\pi_*N \to \pi_*\pi^*\pi_*N$
is an isomorphism, one gets that the map

is an isomorphism. Moreover, the map

induced by
$\pi^*g$
, vanishes as was already the case before applying
$\pi^*$
.
Therefore, when applying
$\mathrm{Hom}_{\mathrm{D}(\mathbb{A}^n_k)}(\mathbb{Q}_{\mathbb{A}^n_k},-[q])$
to the left square of (4.2), we see that the map

induced by h, factors through the zero map and hence is zero.
Hence, we see that the injection
$M\xrightarrow{\gamma} N\xrightarrow{h} L$
induces the zero map after applying the functor
$\mathrm{Hom}_{\mathrm{D}(\mathbb{A}^n_k)}(\mathbb{Q}_{\mathbb{A}^n_k},-[q])$
, finishing the induction.
Corollary 4.8. If X is affine, the object
$\mathbb{Q}_X$
is admissible.
Proof. Let
$q>0$
. Take
$i:X\to \mathbb{A}^n_k$
a closed immersion and
$M\in{\mathcal{M}_{\mathrm{ct}}}(X)$
. By Theorem 4.7, there is an injection
$i_*M\to K$
with K a constructible motive such that
$\mathrm{Hom}_{\mathrm{D}(\mathbb{A}^{n})}({\mathbb{Q}}_{\mathbb{A}^{n}_k},i_*M[q])\to \mathrm{Hom}_{\mathrm{D}(\mathbb{A}^{n})}({\mathbb{Q}}_{\mathbb{A}^{n}_k},K[q])$
is the zero map. We have an inclusion
$f:i^*i_*M\simeq M\to N:=i^*K$
, and a commutative diagram

giving that
$f\circ-$
is zero, thus finishing the proof.
4.3 Admissibility of constant motives on any variety
Notation 4.9. If
$j:U\hookrightarrow X$
is an open subset of a variety X, and
$M\in {\mathcal{M}_{\mathrm{ct}}}(X)$
is a constructible motive, we will use the notation
$M[U]:=j_!{M}_{\mid {U}}$
. Note that by localisation we have
$M/M[U]\simeq i_*{M}_{\mid {Z}}$
with
$i:Z=X\setminus U\to X$
the closed complement.
Lemma 4.10 (Stability of admissibility). Let

be an exact sequence of constructible motives on X.
-
(i) Assume that M” is admissible and at least one of M’, M is admissible. Then all three are admissible.
-
(ii) Assume that M’ and M are admissible. If the functor
$$\mathrm{coker}\ (\mathrm{Hom}_{\mathrm{D}(X)}(M,-)\to\mathrm{Hom}_{\mathrm{D}(X)}(M',-))$$
Proof. These are formal properties of abelian categories, proven in [Reference NoriNor02, Lemma 3.2].
Corollary 4.11 [Reference NoriNor02, Lemma 3.5] Let
$M\in{\mathcal{M}_{\mathrm{ct}}}(X)$
.
-
(i) If
$U\subset X$ is open, and if M and M[U] are admissible, then so is
$M/M[U]$ .
-
(ii) If V,W are open subsets of X, and if
$M[V],M[W],M[V\cap W]$ are admissible, then the same is true for
$M[V\cup W]$
Proof. We use part (ii) of Lemma 4.10. For part (i), it suffices to show that any
$P\in {\mathcal{M}_{\mathrm{ct}}}(X)$
is a subsheaf of a
$Q\in{\mathcal{M}_{\mathrm{ct}}}(X)$
such that the map
$f\colon\mathrm{Hom}_{\mathrm{D}(X)}(M,Q)\to\mathrm{Hom}_{\mathrm{D}(X)}(M[U],Q)$
is surjective. Denote by
$j:U\to X$
the open immersion and by
$i:Z\to X$
its closed complement. Then define
$Q = i_*i^*P\oplus j_*j^*P$
. Localisation ensures that the map
$P\to Q$
is injective,

and

thus is f surjective as it is just the projection on the second factor.
For part (ii), first one has the Mayer–Vietoris exact sequence

We let
$U=V\cap W$
. For a given P we take the same Q as above for U that gives surjectivity of
$\mathrm{Hom}_{\mathrm{D}(X)}(M,Q)\to \mathrm{Hom}_{\mathrm{D}(X)}(M[U],Q)$
. Now the map
$a:M[U]\to M$
factors through M[V], thus the map

is surjective. Finally, the map induced by
$M[U]\to M[V]\oplus M[W]$
is also surjective, as is one of its direct summands.
Proposition 4.12 (Avoiding the use of injective objects, after Nori [Reference NoriNor02, Remark 3.8]). Let
$\mathcal{A},\mathcal{B}$
and
$\mathcal{C}$
be abelian categories. Let
$G:\mathcal{B}\to\mathcal{A}$
be a functor which has an exact left adjoint. Let
$H:\mathcal{A}\to \mathcal{C}$
be an effaceable functor. Then HG is also effaceable.
Proof. Let F be a left adjoint of G. Let B be an object of
$\mathcal{B}$
. By assumption on H, there is an injection
$u:G(B)\to A$
such that the induced map
$H(u):HG(B)\to H(A)$
vanishes. By exactness of F, the morphism
$F(u):FG(B)\to F(A)$
is a monomorphism. We also have the counit of the adjunction
$\varepsilon:FG(B)\to B$
. Let B’ be the pushout of these two maps.

The map
$v:B\to B'$
is a monomorphism and we have a commutative diagram (with
$\eta$
the unit of the adjunction).

The composition
$G(\varepsilon)\circ\eta_{G(B)}$
is the identity, therefore, the map G(v) factors as
$w\circ u$
with
$w = G(t)\circ \eta_A$
. Now,
$HG(v)=H(w) \circ H (u) = 0$
, hence HG is effaceable.
Corollary 4.13. Let X and Y be varieties. Let
$F:\mathrm{D}(X)\to \mathrm{D}(Y)$
be a t-exact functor which is left adjoint to a t-exact functor. If
$M\in{\mathcal{M}_{\mathrm{ct}}}(X)$
is admissible, then so is
$F(M)\in{\mathcal{M}_{\mathrm{ct}}}(Y)$
.
Proof. Let
$q>0$
and let G be a right adjoint of F. The functor
$H = \mathrm{Hom}_{\mathrm{D}(X)}(M,-[q])$
is effaceable by definition. By Proposition 4.12,
$H\circ G$
is therefore effaceable. We have a natural isomorphism
$H\circ G \simeq \mathrm{Hom}_{\mathrm{D}(Y)}(F(M),-[q])$
because G is t-exact, which gives the claim.
Corollary 4.14. Let X be a variety. Let
$M\in{\mathcal{M}_{\mathrm{ct}}}(X)$
be an admissible motive.
-
(i) Let
$j:X\to Y$ be an étale morphism. Then
$j_!M$ is admissible.
-
(ii) Let
$L\in{\mathcal{M}_{\mathrm{ct}}}(X)$ be a lisse motive. Then
$M\otimes L$ is admissible.
Proof. We use Corollary 4.13. For the first point, we have the adjunction
$(j_!,j^*)$
where
$j_!$
and
$j^*$
are t-exact. Let
$L\in{\mathcal{M}_{\mathrm{ct}}}(X)$
be a lisse motive. The functor
$-\otimes L$
is left adjoint to
$\mathscr{{H om}}\,(L,-)$
and right adjoint to
$\mathscr{H{ om}}\,(L^\vee,-)$
, hence is t-exact. The same can be said about
$\mathscr{{H om}}\,(L,-)=-\otimes L^\vee$
.
Proposition 4.15 [Reference NoriNor02, Proposition 3.6]. Let
$U\subset X$
an open of a variety X. Then the constructible motive
$\mathbb{Q}_X[U]$
is admissible.
Proof. Let
$j:U\to X$
be the open immersion. We prove the theorem by induction on the number n of affines needed to cover U. If U is affine, Corollary 4.8 ensures that
$\mathbb{Q}_U$
is admissible, hence by Corollary 4.14,
$j_!\mathbb{Q}_U=\mathbb{Q}_X[U]$
is admissible. For the induction step, one can write
$U = V\cup W$
with V affine and W covered by
$n-1$
affines, and separateness of X gives that
$V\cap W$
is covered by
$n-1$
affines. Therefore, by induction
$\mathbb{Q}_X[V]$
,
$\mathbb{Q}_X[V\cap W]$
and
$\mathbb{Q}_X[W]$
are admissible. Then Corollary 4.11 ensures that
$\mathbb{Q}_X[U]$
is admissible.
Corollary 4.16 [Reference NoriNor02, Theorem 2]. Let X be a variety over k. Then the constant motive
$\mathbb{Q}_X$
is admissible, that is, for every
$q>0$
, every constructible motive
$M\in {\mathcal{M}_{\mathrm{ct}}}(X)$
can be embedded in a constructible motive
$N\in {\mathcal{M}_{\mathrm{ct}}}(X)$
such that the map

vanishes.
Proof. One takes
$U=X$
in the previous proposition.
4.4 Lisse motives enter the stage
Proposition 4.17. Let X be a variety. Any lisse motive in
${\mathcal{M}_{\mathrm{ct}}}(X)$
is admissible.
Proof. By Corollary 4.16 the unit object
$\mathbb{Q}_X$
on X is admissible. By Corollary 4.14,
$L \simeq \mathbb{Q}_X\otimes L$
is admissible.
Theorem 4.18. Let X be a variety. Then any
$M\in {\mathcal{M}_{\mathrm{ct}}}(X)$
is admissible, that is, for every
$q>0$
, the functor
$\mathrm{Hom}_{\mathrm{D}(X)}(M,-[q]):{\mathcal{M}_{\mathrm{ct}}}(X)\to \mathrm{Vect}_\mathbb{Q}$
is effaceable.
Proof. We prove the result by Noetherian induction on the support S of M in X, that is, the complement of the largest open subset V of X such that
${M}_{\mid {V}} = 0$
. We follow the proof of [Reference NoriNor02, Proposition 3.10].
Let
$S^d$
be the union of the irreducible components of maximal dimension of S. We denote by d the dimension of S. Let
$U_\mathrm{liss}$
be the largest open subset of
$S^d$
on which M is lisse. Then there is an affine open subset
$U_1$
of X such that the intersection
$U_1\cap S^d$
is contained in
$U_\mathrm{liss}$
and is non-empty. Let
$U_2'$
be the open subset of X obtained by removing from
$U_1$
the irreducible components of S that are not of maximal dimension. We have that
$U_2'\cap S$
is non-empty and contained in
$U_\mathrm{liss}$
. We choose a non-empty affine open subset
$U_2$
of
$U_2'$
such that
$U_2\cap S$
is non-empty and a Noether normalisation
$g_2\colon U_2\cap S\to \mathbb{A}^d_k$
. It extends to a map
$f_2\colon U_2\to \mathbb{A}^d_k$
. Indeed, because
$U_2\cap S$
is closed in
$U_2$
and the three schemes
$U_2$
,
$U_2\cap S$
and
$\mathbb{A}^d_k$
are affine, we may choose in
$\mathscr{O}_{U_2}(U_2)$
lifts of the images of the coordinates of
$\mathbb{A}^d_k$
in
$\mathscr{O}_{U_2\cap S}(U_2\cap S)$
. Let
$W\subset \mathbb{A}^d_k$
be a non-empty open subset over which
$g_2$
is smooth. As
$g_2$
is of relative dimension 0 in fact
$g_2$
is étale over W. We set
$U := f_2^{-1}(W)$
.
Let
$Z = U\cap S$
and
$g\colon Z\to W$
be the restriction of
$g_2$
to Z. By construction we have
$Z = g_2^{-1}(W)$
, thus the map
$g\colon Z\to W$
is finite and étale. We also let
$f:U\to W$
be the restriction of
$f_2$
to U. The situation may be summarised in a commutative diagram

where we have denoted by i the inclusion of Z inside U.
Note that because the restriction of M to Z is lisse and g is finite étale, the object
$g_*M_{\mid Z}$
is still lisse, so that
$L:=f^*g_*M_{\mid Z}$
is lisse. The restriction
$i^*L$
to Z of L is
$g^*g_*M_{\mid Z}$
which has
$M_{\mid Z}$
as a direct factor because g is étale.
Because the support of
$M_{\mid U}$
is contained in Z we have that
$i_*M_{\mid Z}= i_*i^*M_{\mid U}\simeq M_{\mid U}$
. By the above, this shows that
$M_{\mid U}$
is a direct factor of
$i_*L_{\mid Z}$
. Let j be the inclusion of the open complement of Z in U. By localisation we have that
$i_*L_{\mid Z} \simeq L/j_!j^*L$
. By Proposition 4.3 the objects L and
$j^*L$
are admissible, thus, as by Corollary 4.14, the object
$j_!j^*L$
is then admissible, we deduce by Corollary 4.11 that the object
$i_*L_{\mid Z}$
is admissible. Thus, the restriction
$M_{\mid U}$
of M to U is admissible as a direct factor of an admissible object.
If
$\iota\colon U\to X$
is the open immersion then
$\iota_!M_{\mid U}$
is admissible by Corollary 4.14. We have an exact sequence

in which the support of
$M/\iota_!M_{\mid U}$
is strictly smaller than the support of M. By Noetherian induction
$M/\iota_!M_{\mid U}$
is admissible, thus by Lemma 4.10 we finally obtain that M is admissible.
Corollary 4.19. Let X be a variety. The natural functor
$\mathrm{D}^b({\mathcal{M}_{\mathrm{ct}}}(X))\to\mathrm{D}(X)$
is an equivalence.
Proof. We apply [Reference Beilinson, Bernstein and DeligneBBD82, Proposition 3.1.16]. The conditions are verified because of the previous Theorem 4.18.
5. Higher categorical enhancements
5.1 Categorical preliminaries
Recall the following fundamental definition due to Lurie.
Definition 5.1. An
$\infty$
-category
$\mathcal{C}$
is stable if it has all finite limits and colimits, if initial objects coincide with final objects, and if any commutative square in
$\mathcal{C}$
is cocartesian if and only if it is cartesian. An exact functor between stable
$\infty$
-categories is an
$\infty$
-functor preserving finite limits and colimits.
The homotopy category
$\mathrm{ho}(\mathcal{C})$
of a stable
$\infty$
-category is always a triangulated category [Reference LurieLur17, Theorem 1.1.2.14], and exact functors induce triangulated functors. By definition, a t-structure on a stable
$\infty$
-category is a t-structure on its homotopy category.
Definition 5.2. The bounded derived
$\infty$
-category of an abelian category
$\mathcal{A}$
is the
$\infty$
-category
$\mathcal{D}^b(\mathcal{A})$
obtained by inverting quasi-isomorphisms in the category
$\mathrm{Ch}^b(\mathcal{A})$
of bounded complexes of objects of
$\mathcal{A}$
. It is a stable
$\infty$
-category with a t-structure, and its homotopy category is the usual derived category
$\mathrm{D}^b(\mathcal{A})$
.
If
$F\colon \mathcal{A}\to \mathcal{C}$
is a functor between an abelian category and a stable
$\infty$
-category, we will say that F is exact if it preserves finite coproducts and it sends short exact sequences to exact triangles.
We will use the following result from Bunke, Cisinski, Kasprowski and Winges.
Theorem 5.3 [Reference Bunke, Cisinski, Kasprowski and WingesBCKW24, Corollary 7.4.12]. Let
$\mathcal{C}$
be a stable
$\infty$
-category and let
$\mathcal{A}$
be an abelian category. Then restriction to the heart gives an equivalence of
$\infty$
-categories

between
$\infty$
-categories of exact functors.
There is an easy generalisation with multiple variables.
Corollary 5.4. Let
$\mathcal{C}$
be a stable
$\infty$
-category and let
$\mathcal{A}_1,\dots,\mathcal{A}_n$
be abelian categories. Denote by
$\mathrm{Fun}^{\mathrm{nex}}(\prod_i\mathcal{D}^b(\mathcal{A}_i),\mathcal{C})$
the
$\infty$
-category of n-multi-exact functors
$\prod_i \mathcal{D}^b(\mathcal{A}_i)\to\mathcal{C}$
, that is, functors that are exact in each variable. Denote also similarly by
$\mathrm{Fun}^{\mathrm{nex}}(\prod_i\mathcal{A}_i,\mathcal{C})$
the
$\infty$
-category of n-multi-exact functors. Then the restriction to the hearts functor

is an equivalence of
$\infty$
-categories.
Proof. We give the proof for
$n=2$
; the general case is similar and can be reduced to the case
$n=2$
by induction. We have canonical equivalence of
$\infty$
-categories
$\mathrm{Fun}(\mathcal{D}^b(\mathcal{A}_1)\times\mathcal{D}^b(\mathcal{A}_2),\mathcal{C})\simeq \mathrm{Fun}(\mathcal{D}^b(\mathcal{A}_1),\mathrm{Fun}(\mathcal{D}^b(\mathcal{A}_2),\mathcal{C}))$
. This equivalences induces

By Theorem 5.3 the latter category is exactly
$\mathrm{Fun}^{\mathrm{ex}}(\mathcal{A}_1,\mathrm{Fun}^{\mathrm{ex}}(\mathcal{A}_2,\mathcal{C}))$
, which again is equivalent to
$ \mathrm{Fun}^{\mathrm{2ex}}(\mathcal{A}_1\times\mathcal{A}_2,\mathcal{C})$
.
We will also need a result on the monoidal structure of the bounded derived category. We denote by
$\mathrm{SymMono}_1$
the 2-category of symmetric monoidal 1-categories and symmetric monoidal functors.
Proposition 5.5. Let
$\mathcal{B}$
be a symmetric monoidal abelian category such that the tensor product is exact in each variable. Then there is a canonical symmetric monoidal structure on the stable
$\infty$
-category
$\mathcal{D}^b(\mathcal{B})$
such that the inclusion functor
$\mathcal{B}\to \mathcal{D}^b(\mathcal{B})$
is symmetric monoidal. Moreover, for any category I and any functor
$\mathcal{A}: I\to \mathrm{SymMono}_1$
such that for each
$i\in I$
the category
$\mathcal{A}(i)$
is a symmetric monoidal abelian category with exact tensor product as above, and such that the transition functors are exact, there is a lift of the functor
$\mathcal{D}^b\circ \mathcal{A}\colon I\to \mathrm{Cat}_\infty$
to a functor
$I\to \mathrm{CAlg}(\mathrm{Cat}_\infty).$
Proof. The first case follows from the second with
$I=*$
. Let
$\mathcal{A}:I\to \mathrm{SymMono}_1$
be a diagram of symmetric monoidal abelian categories and symmetric monoidal exact functors, and such that for each
$i\in I$
the tensor product on
$\mathcal{A}(i)$
is exact in both variables. We can compose this functor with
$\mathrm{Ch}^b(-)$
to obtain a diagram of symmetric monoidal additive categories (the monoidal structure on
$\mathrm{Ch}^b(\mathcal{A}(i))$
is induced by that of
$\mathcal{A}(i)$
, using the sign trick). By [Reference Mac LaneML98, Chapter XI, Section 1, Theorem 1] we know that any symmetric monoidal 1-category has all higher coherences. Written differently as in Lurie [Reference LurieLur17, Corollary 5.1.1.7], the forgetful functor induces an equivalence of 2-categories

Thus, we can see
$\mathrm{Ch}^b\circ\mathcal{A}$
as a functor

Now by [Reference LurieLur17, Theorem 2.4.3.18 and Proposition 2.4.2.5] the data of this functor is equivalent to the data of a
$I^\amalg$
-monoid, which is classified by a cocartesian fibration
$p\colon \mathfrak{C}\to I^\amalg$
(we denote by
$I^\amalg$
the
$\infty$
-operad constructed in [Reference LurieLur17, Construction 2.4.3.1]). By [Reference HinichHin16, Proposition 2.1.4] we can fiberwise invert quasi-isomorphisms to obtain a cocartesian fibration
$q\colon \mathfrak{D}\to I^\amalg$
which classifies (using unstraighteningFootnote
1
and [Reference LurieLur17, Theorem 2.4.3.18 and Proposition 2.4.2.5] again) a functor
$I\to\mathrm{CAlg}(\mathrm{Cat}_\infty)$
sending an object
$i\in I$
to
$\mathcal{D}^b(\mathcal{A}(i))$
. The fact that the functor
$\mathcal{A}(i)\to\mathcal{D}^b(\mathcal{A}(i))$
is symmetric monoidal comes from the exactness of the tensor product on
$\mathcal{A}(i)$
.
We also have a version of the universal property of
$\mathcal{D}^b(\mathcal{A})$
(see Theorem 5.3) that works in families.
Proposition 5.6. Let
$\mathcal{C}: I\to \mathrm{CAlg}(\mathrm{Cat}_\infty)$
be a diagram of stably symmetric monoidal
$\infty$
-categories such that for each
$i\in I$
the
$\infty$
-category
$\mathcal{C}(i)$
has a t-structure such that the tensor product is t-exact in each variable and every arrow
$i\to j$
in I induces a t-exact functor. Then the canonical functors
$\mathcal{D}^b(\mathcal{C}(i)^\heartsuit)\to \mathcal{C}(i)$
assemble to give a natural transformation
$\mathcal{D}^b(\mathcal{C}^\heartsuit)\Rightarrow \mathcal{C}$
of functors
$I\to \mathrm{CAlg}(\mathrm{Cat}_\infty)$
.
Proof. For
$\mathcal{B}$
an abelian category, denote by
$\mathcal{K}^b(\mathcal{B})$
the
$\infty$
-category of bounded complexes in
$\mathcal{B}$
, that is, the
$\infty$
-categorical localisation of the additive category of bounded chain complexes
$\mathrm{Ch}^b(\mathcal{B})$
with respect to chain homotopies. By the work of Bunke, Cisinski, Kasprowski and Winges, we know [Reference Bunke, Cisinski, Kasprowski and WingesBCKW24, Theorem 7.4.0] that
$\mathcal{K}^b(\mathcal{B})$
is the value at
$\mathcal{B}$
of a functor

from the
$\infty$
-category of additive
$\infty$
-categories and coproduct-preserving functors to the
$\infty$
-category
$\mathrm{Cat}_\infty^\mathrm{ex}$
of small stable
$\infty$
-categories and exact functors. Moreover, this functor is left adjoint to the forgetful functor

By Cisinski [Reference CisinskiCis19, Theorem 6.1.22] this induces an adjunction

where the category
$I^\amalg$
is as introduced in the proof of the above proposition. Our assumptions on
$\mathcal{C}$
imply that if we denote by
$\mathcal{A}(i)$
the heart of the t-structure on
$\mathcal{C}(i)$
, we have a natural transformation
$\mathcal{A}\Rightarrow\mathcal{C}$
of functors
$I\to\mathrm{CAlg}(\mathrm{Cat}_\infty^\mathrm{add}$
). Now, as in the proof of Proposition 5.5, we have from Lurie [Reference LurieLur17, Theorem 2.4.3.18 and Proposition 2.4.2.5] that the data of this natural transformation is equivalent to the data of a natural transformation
$\mathfrak{A}\Rightarrow \mathfrak{C}$
of
$I^\amalg$
-monoids
$I^\amalg\to \mathrm{Cat}_\infty^\mathrm{add}$
(see [Reference LurieLur17, Definition 2.4.2.1] for a definition). By (5.2), this gives a natural transformation

of functors
$I^\amalg\to \mathrm{Cat}_\infty^\mathrm{ex}$
. We claim that
$\mathcal{K}^b\circ\mathfrak{A}$
is still a
$I^\amalg$
-monoid. Indeed, it suffices to prove that
$\mathcal{K}^b$
preserves finite products, but both in
$\mathrm{Cat}_\infty^\mathrm{ex}$
and in
$\mathrm{Cat}_\infty^\mathrm{add}$
finite product and coproduct agree (this can be proven exactly as in [Reference Bunke, Cisinski, Kasprowski and WingesBCKW24, Lemma 2.1.38]), so that this follows from
$\mathcal{K}^b$
being a left adjoint which thus preserves colimits.
Unravelling the definitions, we have proven that the natural transformation
$\mathcal{K}^b(\mathcal{A})\Rightarrow \mathcal{C}$
is well defined and is symmetric monoidal. We will now show that this factors through
$\mathcal{D}^b(\mathcal{A})$
, and that everything stays symmetric monoidal. For this, we will see our map of
$I^\amalg$
-monoid
$\mathcal{K}^b\circ\mathfrak{A}\Rightarrow \mathfrak{C}$
through straightening, thus as a commutative triangle

where
$\kappa$
and
$\gamma$
are the cocartesian fibrations classifying the functors
$\mathcal{K}^b\circ\mathfrak{A}$
and
$\mathfrak{C}$
. We consider a marking W on
$\int_{I^\amalg}\mathcal{K}^b\circ\mathfrak{A}=:\mathscr{K}$
to be the set of arrows that are products of quasi-isomorphisms. That is, an arrow f in
$\mathscr{K}$
is marked if there exist an integer n and an object
$\underline{i}=(i_1,\dots,i_n)\in I^\amalg_{\langle n\rangle}$
such that f is a map in the fiber
$\prod_{j=1}^n\mathcal{K}^b(\mathcal{A}(i_j))$
above
$\underline{i}$
and is of the form
$f_1\times\cdots\times f_n$
with each
$f_j$
being a quasi-isomorphism. Then the cocartesian fibration
$\kappa$
, together with the data of W and the marking of isomorphisms in
$I^\amalg$
, is a marked cocartesian fibration in the sense of Hinich [Reference HinichHin16, Definition 2.1.1]. Because each
$\mathcal{A}(i)$
is the heart of a t-structure on
$\mathcal{C}(i)$
, all maps in W are sent to isomorphisms in
$\int_{I^\amalg}\mathfrak{C}=:\mathscr{C}$
. Indeed, over a multi-index
$(i_1,\dots,i_n)$
in
$I^\amalg$
, the functor
$\mathfrak{r}$
becomes the functor

which sends marked maps to isomorphisms by [Reference Bunke, Cisinski, Kasprowski and WingesBCKW24, Proposition 7.4.11]. Thus, the map

factors through

By [Reference HinichHin16, Proposition 2.1.4], the map
$\mathscr{K}[W^{-1}]\to I^\amalg$
is a cocartesian fibration, and moreover, the fiber above an
$\underline{i}=(i_1,\dots,i_n)$
is the localisation of
$\prod_{j= 1}^n\mathcal{K}^b(\mathcal{A}(i_j))$
with respect to products of quasi-isomorphisms, which is
$\prod_{j=1}^n\mathcal{D}^b(\mathcal{A}(i_j))$
. Because
$I^\amalg$
is a 1-category, to check that the diagram

commutes it suffices by adjunction to check that
$\mathrm{ho}(-)$
of it commutes, which is clear because the horizontal map preserves fibres. Thus, after unstraightening we obtain a map of
$I^\amalg$
-monoids

which express exactly the symmetric monoidality of a natural transformation
$\mathcal{D}^b\circ\mathcal{A}\to\mathcal{C}$
, finishing the proof.
Remark 5.7. The proof of Proposition 5.6 also gives the same result with
$\mathcal{K}^b(-)$
instead of
$\mathcal{D}^b(-)$
, and the fact that
$\mathcal{K}^b(-)\Rightarrow\mathcal{D}^b(-)$
is symmetric monoidal on symmetric monoidal abelian categories and symmetric monoidal exact functors.
We will use the following two theorems.
Theorem 5.8 [Reference Nguyen, Raptis and SchradeNRS20, Theorem 3.3.1]. Let
$\mathcal{D}$
be an
$\infty$
-category which admits finite limits and let
$\mathcal{C}$
be an
$\infty$
-category. Let
$G:\mathcal{D}\to\mathcal{C}$
be a functor which preserves finite limits. Then G admits a left adjoint if and only if
$\mathrm{ho}(G):\mathrm{ho}(\mathcal{D})\to\mathrm{ho}(\mathcal{C})$
does.
Corollary 5.9. Let
$\mathcal{C}$
and
$\mathcal{D}$
be stable
$\infty$
-categories. Let
$G:\mathcal{D}\to\mathcal{C}$
be an exact functor. Then G admits a left (respectively, a right) adjoint if and only if
$\mathrm{ho}(G):\mathrm{ho}(\mathcal{D})\to\mathrm{ho}(\mathcal{C})$
does.
Proof. The case of the left adjoint is the above theorem, and the case of the right adjoint follows by passing to the opposite
$\infty$
-categories.
Theorem 5.10 [Reference CisinskiCis19, Theorem 7.6.10]. Let
$F:\mathcal{D}\to\mathcal{C}$
be a functor between
$\infty$
-categories having finite limits. Assume that F preserves finite limits. Then F is an equivalence of
$\infty$
-categories if and only if
$\mathrm{ho}(F):\mathrm{ho}(\mathcal{D})\to\mathrm{ho}(\mathcal{C})$
is an equivalence of categories.
Recall that by [Reference LurieLur18a, Tag 01F1], we have the following proposition.
Proposition 5.11. Let
$F:\mathcal{C}\to\mathcal{D}$
be a functor of
$\infty$
-categories. Assume that for a collection
$A\subset\mathcal{C}$
of objects of
$\mathcal{C}$
is given the data, for each
$a\in A$
of an isomorphism
$f_a:F(a)\to d_a$
with
$d_a\in\mathcal{D}$
. Then there exist a functor
$G:\mathcal{C}\to\mathcal{D}$
and a natural equivalence
$g:F\Rightarrow G$
such that for every
$a\in A$
, we have
$g_a = f_a$
(and this is really an equality).
Proof. Indeed, denote by
$\iota : A\to\mathcal{C}$
the canonical monomorphism of simplicial sets. Then by [Reference LurieLur18a, Tag 01F1], the restriction

is an isofibration of simplicial sets (cf. [Reference CisinskiCis19, 3.3.15]). The data of
$f_a$
defines an isomorphism
$\gamma$
in
$\mathrm{Fun}(A,\mathcal{D})$
between
$F_{\mid A}$
and the map of simplicial sets
$A\to\mathcal{D}$
defined by
$a\mapsto d_a$
. By the lifting property of isofibrations, there exist a functor
$G:\mathcal{C}\to\mathcal{D}$
and an isomorphism
$F\simeq G$
, that is a natural equivalence
$g:F\Rightarrow G$
such that
$\iota^*g = \gamma$
, which is precisely the statement of the proposition.
To prove descent statements for categories of perverse Nori motives and mixed Hodge modules, we will use the following proposition and its convenient corollary. Robin Carlier really helped the author in the writing of this proposition.
Proposition 5.12. Let
$\chi : (\mathcal{C}^\mathrm{op})^{\lhd}\to \mathrm{Cat}_\infty$
be a functor. For any
$f:c\to c'$
in
$(\mathcal{C}^\mathrm{op})^{\lhd}$
, denote
$f^*=\chi(f)$
, and assume that any such
$f^*$
has a right adjoint
$f_*$
. For each
$c\in\mathcal{C}$
, denote by
$f_c:c\to v$
the unique map from c to the end point v (we have that
$(\mathcal{C}^\mathrm{op})^\lhd \simeq (\mathcal{C}^\rhd)^\mathrm{op}$
). Let
$\overline{\pi}:\overline{\mathcal{D}}\to (\mathcal{C}^\mathrm{op})^{\lhd}$
be the cocartesian fibration that classifies
$\chi$
, and let
$\pi:\mathcal{D}\to \mathcal{C}^\mathrm{op}$
be its pullback by the inclusion
$\mathcal{C}^\mathrm{op}\to(\mathcal{C}^\mathrm{op})^{\lhd}$
.
Then
$\chi$
is a limit diagram if and only if the following two conditions are satisfied.
-
(i) The family
$(f_c^*)_{c\in\mathcal{C}}$ is conservative.
-
(ii) For any cocartesian section
$X:\mathcal{C}^\mathrm{op}\to \mathcal{D} $ of
$\pi$ , the limit
$X(v):=\lim_{c\in\mathcal{C}}(f_c)_*X(c)$ exists in
$\chi(v)$ , and the map
$f_c^*X(v)\to X(c)$ adjoint to the canonical map
$$\lim_c (f_c)_*X(c)\to (f_c)_*X(c)$$
Proof. By [Reference LurieLur18a, Construction 02T0], there is a diffraction functor

that fits in a commutative triangle.

By the diffraction criterion [Reference LurieLur18a, Theorem 02T8],
$\chi$
is a limit diagram if and only if the restriction morphism in (5.5) is an equivalence. In that case, by [Reference LurieLur18a, Remark 02TF], the functor Df is an equivalence.
Assume that for any cocartesian section
$X:\mathcal{C}^\mathrm{op}\to \mathcal{D}$
of
$\pi$
, the functor
$c\mapsto (f_c)_*X(c)$
admits a limit in
$\overline{D}_v$
. By the preservation of limits of any right adjoint in
$(\mathcal{C}^\mathrm{op})^\lhd$
and [Reference LurieLur18a, Corollary 0311 and Corollary 02KY], this is equivalent to saying that for any cocartesian section
$X:\mathcal{C}^\mathrm{op}\to\mathcal{D}$
of
$\pi$
, the lifting problem

admits a solution
$\overline{X}$
which is a
$\overline{\pi}$
-limit [Reference LurieLur18a, Definition 02KG]. By [Reference LurieLur18a, Example 02ZA], in that case
$\overline{X}$
is the right Kan extension functor along the inclusion
$i:\mathcal{C}^\mathrm{op}\to(\mathcal{C}^\mathrm{op})^\lhd$
. This easily implies that the restriction functor
$\mathrm{res}:\mathrm{Fun}_{/(\mathcal{C}^\mathrm{op})^\lhd}^\mathrm{Cocart}((\mathcal{C}^\mathrm{op})^\lhd,\overline{\mathcal{D}})\to \mathrm{Fun}_{/\mathcal{C}^\mathrm{op}}^\mathrm{Cocart}(\mathcal{C}^\mathrm{op},\mathcal{D})$
is fully faithful as the left adjoint of the right Kan extension along
$i : \mathcal{C}^\mathrm{op}\to(\mathcal{C}^\mathrm{op})^\lhd$
.
Therefore, assuming point (ii) of the proposition,
$\chi$
is a limit diagram if and only if the restriction functor is conservative by [Reference LurieLur18a, Corollary 03UZ]. But as the family of evaluations
$\mathrm{ev}_c : \mathrm{Fun}(\mathcal{C}^\mathrm{op},\mathcal{D})\to\mathcal{D}$
is conservative, this is equivalent to the family of functors
$\mathrm{ev}_c\circ\mathrm{res}=\mathrm{ev}_c\circ\mathrm{Df}\circ\mathrm{ev}_v = f_c^*$
being conservative, which is point (i) of the proposition.
By (5.5), if
$\chi$
is a limit diagram, Df is an equivalence which implies that (i) holds. To finish the proof of the proposition, we therefore only have to show that if
$\chi$
is a limit diagram, then (ii) holds. Let
$\overline{X}$
be a cocartesian section of
$\overline{\pi}$
. Assume first that
$\overline{X}$
is a
$\overline{\pi}$
-limit diagram. By [Reference LurieLur18a, Example 03ZA] this is equivalent to supposing that
$\overline{X}$
is the
$\overline{\pi}$
-right Kan extension of its restriction X to
$\mathcal{C}^\mathrm{op}$
. Then, by [Reference LurieLur18a, Corollary 0307], the limit of
$c\mapsto (f_c)_*X(c)$
exists and the fact that
$\overline{X}$
is cocartesian is equivalent to the fact that for every
$c\in\mathcal{C}$
, the canonical map

is an equivalence.
Therefore, to show that (ii) holds, we only have to show that any cocartesian section
$\overline{X}$
of
$\overline{\pi}$
is a
$\overline{\pi}$
-limit. This is exactly what is proved in the second paragraph of the proof of [Reference LurieLur11, Theorem 5.17].
Corollary 5.13 (See also [Reference LurieLur17, Corollary 5.2.2.37]). Let
$\chi,\chi':(\mathcal{C}^\mathrm{op})^{\lhd}\to\mathrm{Cat}_\infty$
be two functors as in Proposition 5.12, and assume one has a natural transformation
$B:\chi\to\chi'$
which commutes with the right adjoints. Suppose also that the limits of point (ii) in Proposition 5.12 exist for
$\chi$
, that
$B_v:\chi(v)\to\chi'(v)$
preserves them and that each
$B_c$
for
$c\in \mathcal{C}$
is conservative. Then if
$\chi'$
is a limit diagram, so is
$\chi$
.
5.2 Enhancement of the six-functor formalism
In this subsection, we will work with perverse Nori motives and mixed Hodge modules interchangeably. Therefore, we denote by
$\mathrm{Var}_k$
the category of quasi-projective k-schemes when dealing with perverse Nori motives or the category of separated and reduced finite-type
$\mathbb{C}$
-schemes when dealing with mixed Hodge structures. We will call elements of
$\mathrm{Var}_k$
‘varieties’ or ‘k-varieties’. For
$X\in\mathrm{Var}_k$
, we will denote by
${\mathcal{M}_{\mathrm{perv}}}(X)$
the category of perverse Nori motives constructed in [Reference Ivorra and MorelIM22] or the category of Saito’s mixed Hodge modules constructed in [Reference SaitoSai90]. We will have to use realisation functors, hence we denote by
$R: \mathrm{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to \mathrm{D}^b_c(X)$
either the Betti realisation of perverse Nori motives when
$k\subset\mathbb{C}$
, the underlying
$\mathbb{Q}$
-structure functor of mixed Hodge modules, or the
$\ell$
-adic realisation of perverse Nori motives. Here
$\mathrm{D}^b_c(X)$
is the derived category of bounded cohomologically constructible sheaves on
$X^\mathrm{an}$
in the two first cases, and the derived category of constructible
$\ell$
-adic étale sheaves on X. The triangulated category
$\mathrm{D}^b_c(X)$
is endowed with a perverse and a constructible t-structures and the functor R is t-exact (for both t-structures) and conservative.
We begin with a recollection on the
$\infty$
-categorical enhancement of derived categories of constructible sheaves.
Proposition 5.14. There exists a functor

such that for each map
$f:Y\to X$
in
$\mathrm{Var}_k$
, the symmetric monoidal induced by
$\mathcal{D}_c^b(f)$
on the homotopy categories is the classical pullback of constructible sheaves.
Proof. In both analytic sheaves and étale sheaves, for
$X\in\mathrm{Var}_k$
, the category
$\mathrm{D}^b_c(X)$
is embedded in a bigger category
$\mathcal{D}(X)$
which is of the form
$\mathcal{D}(\mathrm{Shv}(\mathfrak{X},\mathscr{O}))$
with
$\mathfrak{X}$
a topos and
$\mathscr{O}$
a sheaf of rings on
$\mathfrak{X}$
and is naturally a symmetric monoidal stable
$\infty$
-category. Indeed, in the case of analytic sheaves one can take
$\mathfrak{X}$
to be the categories of analytic opens of the complex points of X, and
$\mathscr{O}$
is the constant sheaf
$\mathbb{Q}_X$
. In the case of
$\ell$
-adic sheaves, following [Reference Hemo, Richarz and ScholbachHRS23], one can take
$\mathfrak{X}$
to be the proétale topos of X, and
$\mathscr{O}$
to be the sheaf of rings induced by the condensed ring
$\mathbb{Q}_\ell = \mathbb{Z}_\ell[1/\ell]$
. This construction has a canonical lift to the world of symmetric monoidal
$\infty$
-categories thanks to [Reference LurieLur18b, Remark 2.1.0.5]. As the constructions
$X\mapsto \mathfrak{X}$
and
$\mathfrak{X}\mapsto \mathcal{D}(\mathrm{Shv}(\mathfrak{X},\mathscr{O}))$
are functorial, we obtain a functor

Now as for each f in
$\mathrm{Var}_k$
, the functor
$f^*\colon \mathcal{D}(Y)\to\mathcal{D}(X)$
preserves constructible objects, this induces a functor

which gives the usual pullback on homotopy categories by definition in the case of analytic sheaves, and by [Reference Hemo, Richarz and ScholbachHRS23, Theorem 7.7] for
$\ell$
-adic sheaves.
Construction 5.15. We have a symmetric monoidal homotopy 2-functor

By restricting to the constructible hearts, this induces a functor

such that the tensor product is exact in each variable and the pullback functors are exact (this can be checked after realisation). By applying Proposition 5.5 we obtain a functor

Moreover, the equivalence in Corollary 4.19 lifts to an equivalence of
$\infty$
-categories thanks to Theorem 5.10. Proposition 5.11 then gives a functor

canonically equivalent to
$\mathcal{D}^b({\mathcal{M}_{\mathrm{ct}}}(-))$
.
Proposition 5.16. Let
$f:Y\to X$
be a map in
$\mathrm{Var}_k$
. Then the functor induced by
$\mathcal{D}^b({\mathcal{M}_{\mathrm{ct}}}(f))$
on the homotopy categories is the pullback functor constructed by Ivorra and Morel or by Saito, depending on which case we are in.
Proof. We claim that if we know that the pullback functors constructed by Ivorra, Morel and Saito admit
$\infty$
-categorical lifts
$f^*$
, then they will be canonically equivalent to
$f^*_\mathrm{new} := \mathcal{D}^b({\mathcal{M}_{\mathrm{ct}}}(f))$
. Indeed, in that case the
$\infty$
-functors
$f^*$
and
$f^*_\mathrm{new}$
coincide on the constructible heart by definition, hence they are equivalent by Theorem 5.3.
We prove this first for mixed Hodge modules. For any map
$f:Y\to X$
of varieties, there is the factorisation

where i is the closed immersion given by the inclusion of the graph of f (our varieties are separated) and p is the projection. By definition, we have
$f^*\simeq p^*\circ i^*$
. The functor
$p^*$
is given by the external product with the constant object,
$p^*(M)=M\boxtimes \mathbb{Q}_Y$
, hence is an
$\infty$
-functor because the external product is defined as the derived functor on the perverse hearts. The functor
$i^*$
is the left adjoint of the functor
$i_*$
, which is obtained [Reference SaitoSai90, (4.2.4) and (4.2.10)] by deriving the functor
$i_*$
on the perverse heart and is therefore an
$\infty$
-functor. Thus, by Corollary 5.9 the functor
$i^*$
is an
$\infty$
-functor. Finally,
$f^*$
admits an
$\infty$
-categorical lift.
Now for perverse Nori motives, any map
$f:Y\to X$
, being a map of quasi-projective varieties, factors as

with i a closed immersion and g a smooth map. Hence,
$f^*\simeq i^*g^*$
. The case of
$i^*$
is the same as for mixed Hodge modules, and thus we have an
$\infty$
-functor. For the smooth map g, we can assume that X is connected (as a direct sum of
$\infty$
-functors is an
$\infty$
-functor), so that g is smooth of pure relative dimension d for some
$d\in\mathbb{N}$
. Now by definition
$f^* = (f^*[d])[-d]$
, where
$f^*[d]$
is the derived functor of the exact functor
$f^*[d]$
on the perverse heart, hence we have an
$\infty$
-functor.
Corollary 5.17. Let
$f:Y\to X$
be a map in
$\mathrm{Var}_k^\mathrm{op}$
. Then the
$\infty$
-functor

admits a right adjoint
$f_*$
. Moreover, if f is smooth, then
$f^*$
admits a left adjoint
$f_\sharp$
.
Proof. These adjoints exist on the homotopy categories, thus they have
$\infty$
-categorical lifts by Corollary 5.9.
Proposition 5.18. For each
$X\in\mathrm{Var}_k$
, the monoidal structure obtained in Construction 5.15 induces on the homotopy categories the same monoidal structure as that constructed by Saito for mixed Hodge modules and by Terenzi [Reference TerenziTer24] for perverse Nori motives.
Moreover, the projection formula for a smooth morphism holds in
$ \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
.
Proof. Note that in both [Reference TerenziTer24] and [Reference SaitoSai90] the tensor product is built in the following way. There is an external tensor product
$\boxtimes :{\mathcal{M}_{\mathrm{perv}}}(X)\times{\mathcal{M}_{\mathrm{perv}}}(X)\to{\mathcal{M}_{\mathrm{perv}}}(X \times X)$
that is exact in both variables and hence induces by Corollary 5.4 an
$\infty$
-functor
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\times \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X\times X))$
which we compose with the pullback
$\Delta^*$
under the diagonal
$\Delta :X\to X\times X$
to obtain the tensor product, thus given by
$M\otimes N = \Delta^*(M\boxtimes N)$
. The other structural morphisms of the monoidal structure are obtained similarly. Denote by
$\otimes^T : \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\times \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
Saito’s or Terenzi’s tensor product and denote by
$\otimes^\infty : \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\times \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
the one we just constructed. We want to check that
$\mathrm{ho}(\otimes^\infty)=\mathrm{ho}(\otimes^T)$
. For this there are several isomorphisms to check (see [Reference Mac LaneML98, Section XI.1]):
-
(i) A functorial isomorphism
$\otimes^T\Rightarrow \otimes^\infty$ .
-
(ii) Modulo point (i) isomorphism, an identification of the unit isomorphism
$\rho:M\otimes 1\to M$ and
$\lambda:1\otimes M\to M$ .
-
(iii) Modulo point (i) isomorphism, an identification of the associativity isomorphism
$c : M\otimes(N\otimes P)\to (M\otimes N)\otimes P$ .
-
(iv) Modulo point (i) isomorphism, an identification of the commutativity isomorphism
$\gamma : M\otimes N\to N\otimes M$ .
By definition,
$\otimes^\infty$
coincides with
$\otimes^T$
when restricted to
${\mathcal{M}_{\mathrm{ct}}}(X)\times{\mathcal{M}_{\mathrm{ct}}}(X)$
. By Corollary 5.4 this gives (i). We explain how to obtain (iii) and the other checks will be left to the reader as they are very similar. Denote by

the functor sending M,N,P to
$M\otimes^\infty (N\otimes^\infty P)$
and by

the functor sending M,N,P to
$(M\otimes^\infty N)\otimes^\infty P$
. The associativity isomorphism is a natural equivalence
$c^\infty:t_1^\infty\Rightarrow t_2^\infty$
. We give the same definition for
$t_1^T$
,
$t_2^T$
and
$c^T$
. Again by definition, the images by the restriction functor of the morphisms
$c^\infty$
and
$c^T$
are the same, hence, up to the equivalence of (i), we obtain (iii).
For the projection formula, the arrow exists by a play of adjunctions, and is an equivalence on the homotopy categories.
Corollary 5.19. Let X be a variety. The tensor structure on
$ \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
is closed.
Proof. Again this is true on the homotopy categories and we can apply Corollary 5.9.
Proposition 5.20. Let X be a variety. The realisation functor

admits a canonical lift to a map in
$\mathrm{CAlg}(\mathrm{Cat}_\infty)$
.
Moreover, it induces a natural transformation

on functors

Proof. In the case of the Betti realisation this is easy as we have an equivalence

thanks to Nori’s theorem. This enables us to consider the functor
$\Delta^1\times\mathrm{Var}_k^\mathrm{op}\to\mathrm{CAlg}(\mathrm{Cat}_1)$
that sends (0,f) to
$f^*:{\mathcal{M}_{\mathrm{ct}}}(Y)\to {\mathcal{M}_{\mathrm{ct}}}(X)$
, the edge (1,f) to
$f^*:\mathrm{Cons}(Y)\to\mathrm{Cons}(X)$
and
$(0\to 1,X)$
to the realisation functor. Therefore, by Proposition 5.5 we obtain the result.
For the
$\ell$
-adic realisation it is not true that the canonical functor

is an equivalence. What we did in the case of the Betti realisation gives a symmetric monoidal natural transformation

Hence, we have to check that the
$\mathrm{real}_X$
assemble to give a symmetric monoidal natural transformation

This is Proposition 5.6.
Corollary 5.21. The realisation functor

commutes with all the operations constructed above. This includes pullbacks
$f^*$
, pushforwards
$f_*$
, left adjoints
$f_\sharp$
when f is a smooth map, the tensor product
$-\otimes-$
and the internal homomorphism
$\mathscr{{H om}}\,(-,-)$
.
Proof. Indeed, this is true on the homotopy categories, and the exchange morphisms exist because
$\mathcal{R}$
commutes with pullbacks and tensor products by definition.
Verdier duality also lifts.
Proposition 5.22. Let X be a k-variety. Then the Verdier duality functor
$\mathbb{D}_X$
has a
$\infty$
-categorical lift
$\mathbb{D}_X\colon \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))^\mathrm{op}$
which commutes with the Betti realisation. Moreover, there is a canonical isomorphism
$\mathbb{D}_X\circ\mathbb{D}_X\simeq \mathrm{Id}_{\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))}$
.
Proof. Indeed, the functor
$\mathbb{D}_X$
, both in Ivorra and Morel and in Saito, is constructed as the trivial derived functor of the exact functor
$\mathbb{D}_X\colon {\mathcal{M}_{\mathrm{perv}}}(X)\to{\mathcal{M}_{\mathrm{perv}}}(X)^\mathrm{op}$
, thus it is an
$\infty$
-functor. Thus, the functor also commutes with the Betti realisation as an
$\infty$
-functor (as in fact we apply the functor
$\mathcal{D}^b(-)$
to the commutative square involving Verdier duality of
${\mathcal{M}_{\mathrm{perv}}}(X)$
,
$\mathrm{Perv}(X)$
and R). The fact that
$\mathbb{D}_X\circ\mathbb{D}_X\simeq \mathrm{Id}_{\mathcal{D}^b_\mathcal{M}(X)}$
holds can be checked on the homotopy category as well.
For free, we also obtain the exceptional functoriality.
Corollary 5.23. The exceptional functors
$f^!$
and
$f_!$
constructed by Ivorra, Morel and Saito have
$\infty$
-categorical lifts that assemble to give functors

Moreover, Verdier duality is a natural isomorphism between the star functoriality (
$f_*$
or
$f^*$
) and the shriek functoriality of the opposite category (
$f_!$
or
$f^!$
).
Proof. The last statement of this corollary gives a construction. Indeed, the formula
$\mathbb{D}\circ f_* \simeq f_!\circ\mathbb{D}$
tells us that if we apply Proposition 5.11 to the isomorphisms
$\mathbb{D}_X : \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))^\mathrm{op}$
, the functor
$\mathrm{Sch}^{(\mathrm{op})}\to\mathrm{Cat}_\infty$
obtained encodes exactly the exceptional functoriality.
Theorem 5.24. The functor
$X\mapsto \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
, with respect to the
$(-)^*$
-functoriality, is a hypersheaf with respect to h-topology.
Proof. Denote
${\mathcal{D}^b_\mathcal{M}}(X):=\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
. By [Reference Cisinski and DégliseCD19, Corollary 3.3.38] (or see the proof of [Reference Bhatt and MathewBM21, Proposition 5.11]), it suffices to prove that this functor is an étale hypersheaf because of localisation and proper base change hold for
${\mathcal{D}^b_\mathcal{M}}$
. We will use Corollary 5.13 with a realisation functor (either the
$\ell$
-adic realisation for Nori motives, or the underlying
$\mathbb{Q}$
-structure functor for mixed Hodge modules) that we will denote by

Note that
$\mathrm{D}^b_c$
has étale hyperdescent. Now, to apply Corollary 5.13 we need to make sure that the limits ensuring descent exist in
${\mathcal{D}^b_\mathcal{M}}$
. This is not obvious. However, for any variety X and
$m\in\mathbb{N}$
, if we denote by
$\mathcal{D}^{[-m,m]}_\mathcal{M}(X)$
the full subcategory of
${\mathcal{D}^b_\mathcal{M}}(X)$
of objects concentrated, for the constructible t-structure, in degrees
$[-m,m]$
, we have

Moreover, the functor
${\mathcal{D}^b_\mathcal{M}}$
is a hypersheaf if and only if for each m, the functor
$\mathcal{D}^{[-m,m]}_\mathcal{M}$
is a hypersheaf, and the same holds for
$\mathrm{D}^b_c$
. Indeed, if
${\mathcal{D}^b_\mathcal{M}}$
is a hypersheaf, then as being concentrated in degrees
$[-m,m]$
can be tested after pulling back by any surjective map of schemes, it turns out that each
$\mathcal{D}^{[-m,m]}_\mathcal{M}$
is also a hypersheaf. Conversely, if each
$\mathcal{D}^{[-m,m]}_\mathcal{M}$
is a hypersheaf, then because pullback functors
$f^*$
are t-exact functors, in turns out that the conditions to test that
${\mathcal{D}^b_\mathcal{M}}$
is a hypersheaf only occur in
$\mathcal{D}^{[-m,m]}_\mathcal{M}$
: if
$U_\bullet\to X$
is a hypercovering, full faithfulness of the sheaf condition is computed in
$\mathcal{D}^{[-m,m]}_\mathcal{M}(X)$
for some m, and then essential surjectivity holds because, given an element
$K=(K_n)_n$
of the limit, there exists a fixed m such that K has its
$K_0$
, hence also all its factors
$K_n$
in
$\mathcal{D}^{[-m,m]}_\mathcal{M}(U_n)$
.
Thus, because
$\mathrm{D}^{[-m,m]}_c$
is an étale hypersheaf, it suffices to check the conditions of Corollary 5.13 for the realisation functor

Let
$f_\bullet\colon Y_\bullet\to X$
be an étale hypercovering. As the realisation functor is conservative, it suffices to check the following assertions.
-
(i) For each complex M in
$\mathcal{D}^{[-m,m]}_\mathcal{M}(X)$ , the limit
\[\lim_\Delta \tau^{\leqslant m}(f_n)_*f_n^*M\]
$\Delta$ is the simplicial category and
$\tau^{\leqslant m}$ is the truncation functor for the t-structure.
-
(ii) The realisation functor
\[\mathcal{D}_\mathcal{M}^{[-m,m]}(X)\to\mathrm{D}_c^{[-m,m]}(X)\]
In fact as
$\mathcal{D}^{[-m,m]}_\mathcal{M}(X)$
is a
$(2m+1)$
-category, the above totalisation in (i) is finite and thus exists (because the finite limit
$\lim (f_n)_*f_n^*M$
exists in
$\mathcal{D}_\mathcal{M}^{\geqslant -m}$
because it exists in
$\mathcal{D}^b_\mathcal{M}$
, and then we apply the right adjoint
$\tau^{\leqslant m}$
), and the realisation functor clearly commutes with such a finite limit because the realisation functor

commutes with finite limits (the inclusion
$\mathrm{D}^{\geqslant -m}\to\mathrm{D}^b$
commutes with limits) and the image by the realisation functor of the above limit in (i) is the truncation by
$\tau^{\leqslant m}$
, which is a right adjoint, of a limit in
$\mathrm{D}^{\geqslant -m}_c(X)$
.
6. The realisation functor and applications
6.1 Construction of the realisation functor
In this section, we construct realisation functors from the category of étale motives to the derived categories of mixed Hodge modules and perverse Nori motives that commute with the six operations. Denote by
$\mathrm{Sch}_k$
the category of finite-type k-schemes. We use Drew and Gallauer’s result in [Reference Drew and GallauerDG22] in which they prove that the stable motivic
$\mathbb{A}^1$
-invariant
$\infty$
-category
$\mathcal{SH}$
has a universal property among systems of
$\infty$
-categories called coefficient systems: these are functors

that have the minimal functoriality needed for the six operations to exist thanks to Ayoub’s thesis (see Appendix A for a precise definition and some results on coefficient systems). Any coefficient system
$C\in\mathrm{CoSys}_k$
with values in idempotent complete
$\infty$
-categories receives a unique realisation functor

from the compact objects of the motivic homotopy category that commutes with enough operations to ensure that it commutes with the six operations (if C has some good properties, see below).
Theorem 6.1. Denote by
$\mathcal{D}_\mathcal{M}^b$
the functor
$\mathrm{Var}_k^\mathrm{op}\to\mathrm{CAlg}(\mathrm{\mathrm{Cat}_\infty})$
of Construction 5.15. The functor
$\mathcal{D}_\mathcal{M}^b$
defines an object of
$\mathrm{CoSys}_k$
.
Proof. This is the content of Section 5.2 and Proposition A.4. See Definition A.1 for the precise definition of coefficient systems.
The above theorem contains the statement that perverse Nori motives and mixed Hodge modules extend to every finite-type k-scheme, together with all the operations and h-hyperdescent (the only non-trivial statement here is the h-hyperdescent, but this follows from the fact that the Kan extension of a hypersheaf on a basis of a site to the full site is still a hypersheaf; see, for example, [Reference AokiAok23, Theorem A.6]).
Remark 6.2. Using that the mapping spectra in
$\mathcal{D}^b_\mathcal{M}$
,
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}})$
and
$\mathcal{D}^b({\mathcal{M}_{\mathrm{ct}}})$
are étale sheaves (for the perverse heart, one has to use affine étale morphisms) and the fact that a limit of a diagram of stable
$\infty$
-categories with t-structures and t-exact transition functors is endowed with a t-structure (see [Reference Richarz and ScholbachRS20, Lemma 3.2.18]), one can show that after extension to finite-type k-schemes, we still have

and in the case of perverse Nori motives, the category
${\mathcal{M}_{\mathrm{perv}}}(X)$
is the universal category as for quasi-projective schemes. The weights also extend to that case.
Notation 6.3. For X a finite-type k-scheme, we will denote by
$\mathcal{D}^b(\mathrm{MHM}(X))$
(respectively, by
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
) the value at X of the functor
${\mathcal{D}^b_\mathcal{M}}$
in the mixed Hodge modules case (respectively, in the Nori motives case).
Before going into the proof of the main theorem, we recall a construction originally due to Fabien Morel, and we will use [Reference Cisinski and DégliseCD19, Section 16.2] as a reference. Let X be a finite-type k-scheme. There is a canonical decomposition

of the rational part of the motivic homotopy category
$\mathcal{SH}(X)$
over X. It comes from the decomposition in
$\mathcal{SH}_\mathbb{Q}(X)$
of the unit
$\mathbb{Q}_X$
as a direct sum

Moreover, the object
$\mathbb{Q}_X^+$
has a commutative algebra structure (because it is idempotent) in
$\mathcal{SH}(X)_\mathbb{Q}$
such that the projection map
$\mathbb{Q}_X\to\mathbb{Q}_X^+$
is a map of algebras. By [Reference Cisinski and DégliseCD19, Theorem 16.2.13], there is a canonical identification of
$\mathcal{SH}_\mathbb{Q}^+(X)\simeq \mathrm{Mod}_{\mathbb{Q}^+}(\mathcal{SH}_\mathbb{Q}(X))$
with étale motives
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
. In particular, a symmetric monoidal colimit-preserving
$\infty$
-functor

to a stable presentably symmetric monoidal
$\infty$
-category
$\mathcal{D}$
factors through
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
if and only if
$F(\mathbb{Q}^-_X)=0$
. By [Reference Cisinski and DégliseCD19, Corollary 16.2.14] the object
$\mathbb{Q}^-_X\in\mathcal{SH}_\mathbb{Q}(X)$
vanishes whenever
$-1$
is a sum of squares in the residual fields of X. We will need the following lemma.
Lemma 6.4. The maps
$\mathcal{SH}_\mathbb{Q}(X)\to\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
and
$\mathcal{SH}(X)\to\mathcal{SH}_\mathbb{Q}(X)$
are part of morphisms of coefficient systems
$\mathcal{SH}_\mathbb{Q}\to \mathcal{DM}^{\mathrm{\acute{e}t}}$
and
$\mathcal{SH}\to\mathcal{SH}_\mathbb{Q}$
.
Proof. This is a particular case of a theorem by Drew [Reference DrewDre18, Theorem 8.10] (although in the second case of rationalisation a simpler proof could have been given; see [Reference AyoubAyo14a, Lemma A.11]), where for example in the first case we take (in his notation)
$\mathcal{V} = \mathcal{SH}_\mathbb{Q}(\mathrm{Spec}(k))$
,
$\mathcal{M}^*=\mathcal{SH}_\mathbb{Q}$
and
$A = \mathbb{Q}^+_{\mathrm{Spec}(k)}$
.
Theorem 6.5. There exists an (essentially unique) morphism of coefficients systems
$\mathcal{DM}^{\mathrm{\acute{e}t}}\to \mathrm{Ind} {\mathcal{D}^b_\mathcal{M}}$
on the category
$\mathrm{Sch}_k$
. It restricts to a functor

The functor (6.1) commutes with the six operations, that is, all the exchange transformations are isomorphisms.
In the case of mixed Hodge modules we will denote the realisation functor by
$\mathrm{Hdg}^*$
, and in the case of Nori motives we will denote it by
$\mathrm{Nor}^*$
.
Proof. Thanks to Theorem 6.1 we know that
$\mathcal{D}^b_\mathcal{M}$
is an object of
$\mathrm{CoSys}_k$
, hence its indization
$\mathrm{Ind}\mathcal{D}^b_\mathcal{M}$
is an object of
$\mathrm{CoSys}_k^\mathrm{c}$
. Together with Theorem A.2, this gives a morphism of coefficient systems

Note that the functor
$\mathrm{Ind}\mathcal{D}^b_\mathcal{M}$
naturally takes values in
$\mathrm{CAlg}(\mathrm{Pr}^L)_{\backslash \mathrm{Ind}\mathcal{D}^b_\mathcal{M}(\mathrm{Spec}(k))}$
, the coslice
$\infty$
-category, because
$\mathrm{Spec}(k)$
is the initial object of
$\mathrm{Sch}_k^\mathrm{op}$
. Using the functor
$\rho_{\mathcal{SH}}$
over
$\mathrm{Spec}(k)$
, we see that
$\mathrm{Ind}\mathcal{D}^b_\mathcal{M}$
takes values in
$\mathrm{CAlg}(\mathrm{Pr}^L)_{\backslash \mathcal{SH}(\mathrm{Spec}(k))}$
, the
$\infty$
-category of
$\mathcal{SH}(\mathrm{Spec}(k))$
-algebras in
$\mathrm{Pr}^L$
, where we endowed
$\mathrm{Pr}^L$
with the Lurie tensor product constructed in [Reference LurieLur17, Section 4.8.1]. We claim that
$\rho_{\mathcal{SH},k}\colon\mathcal{SH}(\mathrm{Spec}(k))\to\mathrm{Ind}{\mathcal{D}^b_\mathcal{M}}(\mathrm{Spec}(k))$
factors through
$\mathcal{DM}^\mathrm{\acute{e}t}(\mathrm{Spec}(k))$
.
First, note that because
$\mathrm{Ind}{\mathcal{D}^b_\mathcal{M}}(\mathrm{Spec}(k))$
is
$\mathbb{Q}$
-linear, the functor
$\rho_{\mathcal{SH},k}$
factors through
$\mathcal{SH}_\mathbb{Q}(\mathrm{Spec}(k))$
to give a symmetric monoidal functor

Moreover, by the universal property of the localisation, if K is a finite extension of k in which we added a square root of
$-1$
, the square

commutes, where
$g\colon \mathrm{Spec}(K)\to\mathrm{Spec}(k)$
is the structural morphism. Now because
$\mathbb{Q}^-_{\mathrm{Spec}(k)}$
is a direct factor of
$\mathbb{Q}_{\mathrm{Spec}(k)}$
, its image under
$\rho_{\mathcal{SH}_\mathbb{Q},k}$
lands in
${\mathcal{D}^b_\mathcal{M}}(\mathrm{Spec}(k))$
. By étale descent of
${\mathcal{D}^b_\mathcal{M}}$
proven in Theorem 5.24, the functor
$g^*$
is conservative on
${\mathcal{D}^b_\mathcal{M}}(\mathrm{Spec}(k))$
, and as
$\mathbb{Q}^-_{\mathrm{Spec}(K)}=0$
, we see that the image of
$\mathbb{Q}^-_{\mathrm{Spec}(k)}$
in
$\mathrm{Ind}{\mathcal{D}^b_\mathcal{M}}(\mathrm{Spec}(k))$
vanishes. By the previous discussion Lemma 6.4, this proves that the functor
$\rho_{\mathcal{SH}_\mathbb{Q},k}$
factors through
$\mathcal{DM}^\mathrm{\acute{e}t}(\mathrm{Spec}(k))$
to give a functor that we will denote by
$\widetilde{\nu_k}$
. Through
$\widetilde{\nu_k}$
we see that our functor
$\mathrm{Ind}{\mathcal{D}^b_\mathcal{M}}$
takes values in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(\mathrm{Spec}(k))$
-algebras in
$\mathrm{Pr}^L$
. By [Reference LurieLur17, Remark 7.3.2.13] applied to the left adjoint functor

the forgetful functor

is right adjoint to the functor

Applying this adjunction termwise, we obtain by [Reference CisinskiCis19, Theorem 6.1.22] an adjunction between functors on
$\mathrm{Sch}^\mathrm{op}_k$
. In particular, the natural transformation
$\rho_{\mathcal{SH}}$
may be seen as a map
$\mathcal{SH}\to \mathrm{ff}(\mathrm{Ind}{\mathcal{D}^b_\mathcal{M}})$
in the
$\infty$
-category

By adjunction, we obtain a natural transformation

of functors on
$\mathrm{Sch}_k^\mathrm{op}$
and with values in
$\mathrm{CAlg}(\mathrm{Pr}^L)_{\backslash \mathcal{DM}^\mathrm{\acute{e}t}(\mathrm{Spec}(k))}$
. The left-hand sides send a scheme X to the
$\infty$
-category

By using that

and [Reference LurieLur17, Theorem 4.8.4.6], we see that

This proves that
$\rho_\mathcal{SH}$
factors through the functor
$\mathcal{DM}^\mathrm{\acute{e}t}$
, giving us a natural transformation

of functors on
$\mathrm{Sch}^\mathrm{op}_k$
with values in
$\mathrm{CAlg}(\mathrm{Pr}^L)$
. The factorisation is given by the unit of the adjunction between functor categories described above.
We claim that we obtained a morphism of coefficient systems. Let
$f\colon X\to Y$
be a smooth map. Because
$\mathcal{DM}^\mathrm{\acute{e}t}(X)$
is generated under colimits by objects of the form
$a_\mathrm{\acute{e}t}^\mathbb{Q}(M)$
with
$M\in\mathcal{SH}(X)$
and
$a_\mathrm{\acute{e}t}^\mathbb{Q}$
the morphism
$\mathcal{SH}\to\mathcal{DM}^\mathrm{\acute{e}t}$
, it suffices to prove that the natural exchange morphism

is an equivalence. This follows from the following sequence of equivalences:

where we used in (1) the fact that
$\rho_{\mathcal{SH}}$
is a morphism of coefficient systems and in (2) the fact that
$a_\mathrm{\acute{e}t}^\mathbb{Q}$
is a morphism of coefficient systems by Lemma 6.4.
Denote by
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
(respectively, by
$\mathcal{D}^{\mathrm{gm}}_\mathcal{M}(X)$
) the thick full subcategory of
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
(respectively, of
$\mathrm{Ind}{\mathcal{D}^b_\mathcal{M}}(X)$
) spanned by the
$f_\sharp \mathbb{Q}_Y(i)$
for
$f:Y\to X$
smooth and
$i\in \mathbb{Z}$
. As
${\mathcal{D}^b_\mathcal{M}}(X)\subset \mathrm{Ind}{\mathcal{D}^b_\mathcal{M}}(X)$
is thick and each
$f_\sharp \mathbb{Q}_Y(i)$
is in
${\mathcal{D}^b_\mathcal{M}}(X)$
, we have
$\mathcal{D}^{\mathrm{gm}}_\mathcal{M}(X)\subset {\mathcal{D}^b_\mathcal{M}}(X)$
. Because
$\widetilde{\nu}$
is a morphism of coefficient systems, it induces a morphism of coefficient systems

In particular, at the level of triangulated categories, we obtain a premotivic morphism that satisfies all the hypotheses of [Reference Cisinski and DégliseCD19, Theorem 4.4.25]:
$\mathcal{D}^{\mathrm{gm}}_\mathcal{M}$
is
$\tau$
-generated for
$\tau$
the Tate twist,
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c$
is
$\tau$
-dualisable by absolute purity,
$\mathcal{D}^{\mathrm{gm}}_\mathcal{M}$
is separated because it is a sub-pullback formalism of
${\mathcal{D}^b_\mathcal{M}}$
which satisfies étale descent by Theorem 5.24 (which is equivalent to separateness in the
$\mathbb{Q}$
-linear case by [Reference Cisinski and DégliseCD19, Corollairy 3.3.34]) and all Tate twists are invertible in
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c$
by construction.
Therefore, the functor
$\nu^{\mathrm{gm}}$
commutes with the six operations. One has to be a little careful here: a priori, [Reference Cisinski and DégliseCD19, Theorem 4.4.25] implies only that
$\nu^{\mathrm{gm}}$
commutes with all the operations when
$\mathcal{D}^{\mathrm{gm}}_\mathcal{M}$
is endowed with the operations they construct in their book, and not necessarily with the constructions of Saito, Ivorra and Morel as in Corollary 5.23. We know that for
$f:Y\to X$
a morphism,
$f^*$
is the same for both constructions, hence by adjunction this is also the case for
$f_*$
. For
$j:U\to X$
an open immersion the
$j_\sharp$
are the same, but
$j_\sharp = j_!$
, hence by Nagata compactification (and the fact that for a proper p, we have
$p_*\simeq p_!$
) and composition both
$f_!$
are isomorphic for any
$f:Y\to X$
separated. By adjunction we also have an isomorphism between the
$f^!$
. This defines a isomorphism of étale sheaves of
$\infty$
-categories, hence we also have compatibility of operations for non-separated morphisms (when they are defined for
$\mathcal{DM}^{\mathrm{\acute{e}t}}$
). The tensor product is the same, hence also the internal
$\mathscr{H} \mathscr {om}\,$
by adjunction.
We have obtained that the functor

commutes with the six operations at the triangulated categories level. But all exchange morphisms exist at the level of
$\infty$
-categories, and the functor from an
$\infty$
-category to its homotopy category is conservative, hence
$\nu$
commutes with all the operations.
Corollary 6.6. The realisation functor

commutes with the six operations.
Proof. Indeed, it suffices to check that the functor
$\mathcal{SH}_c\to\mathcal{DM}^{\mathrm{\acute{e}t}}_c$
commutes with the operations. This is the content of the next lemma.
This lemma is probably well known to experts, but the author could not find a reference.
Lemma 6.7. The morphisms
$\rho_\mathbb{Q}\colon\mathcal{SH}_c\to\mathcal{SH}_{\mathbb{Q},c}$
and
$a_{\mathrm{\acute{e}t}}\colon\mathcal{SH}_{\mathbb{Q},c}\to\mathcal{DM}^\mathrm{\acute{e}t}_c$
commute with the six operations.
Proof. The case of
$\rho_\mathbb{Q}$
follows from [Reference AyoubAyo14a, Théorème A.15], thus we focus on
$a_{\mathrm{\acute{e}t}}$
. We know that
$a_{\mathrm{\acute{e}t}}$
is a morphism of coefficient systems generated by objects of the form
$g_\sharp\mathbf{1}_Y(n)$
for
$g\colon Y\to X$
a smooth map and
$n\in\mathbb{Z}$
, thus to prove that it commutes with the six operations it suffices to check that it commutes with pushforwards (see the proof of [Reference Cisinski and DégliseCD19, Theorem 4.4.25]). Let
$f\colon Y\to X$
be a morphism of schemes. As the functor
$a_{\mathrm{\acute{e}t}}$
commutes with
$f^*$
its right adjoint
$\mathrm{ff}^{\mathrm{\acute{e}t}}$
commutes with
$f_*$
, and because
$\mathrm{ff}^{\mathrm{\acute{e}t}}$
is conservative it suffices to check that the natural transformation

is an equivalence. Moreover, by [Reference LurieLur17, Corollary 4.2.4.8] there is a canonical identification of functors

Thus, we have to prove that the natural map

is an equivalence. This is the case because as
$\mathbb{Q}^+$
is a direct factor of
$\mathbb{Q}$
, it is a dualisable object of
$\mathcal{SH}_\mathbb{Q}$
so that we can apply [Reference AyoubAyo14b, Lemma 2.8], using that
$\mathbb{Q}^+_Y\simeq f^*\mathbb{Q}^+_X$
.
Corollary 6.8. There exists a Hodge realisation functor
$R_H:\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}})\to\mathcal{D}^b(\mathrm{MHM})$
compatible with the six operations. It factors the Hodge realisation that we have constructed in Theorem 6.5.
Proof. By Theorem 6.5, there exists a Hodge realisation from
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(-,\mathbb{Q})$
to the derived category of mixed Hodge modules. Therefore, taking the underlying
$\mathbb{Q}$
-structure and then perverse
${\ ^{\mathrm{p}}\mathrm{H}}^0$
, we obtain a factorisation of the perverse
${\ ^{\mathrm{p}}\mathrm{H}}^0$
of the Betti realisation through
$\mathrm{MHM}(X)$
. The universal property of perverse Nori motives then gives a faithful exact functor
${\mathcal{M}_{\mathrm{perv}}}(X)\to \mathrm{MHM}(X)$
, which induces a functor
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to \mathcal{D}^b(\mathrm{MHM}(X))$
. Now all the constructions of [Reference Ivorra and MorelIM22] are the constructions of the operations for mixed Hodge modules, hence this functor is compatible with the operations. The fact that this functor factors the Hodge realisation we constructed above follows from the universal property of
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c$
and the fact that
$R_H$
can be promoted to a natural transformation of
$\infty$
-functors on
$\mathrm{Sch}_k^\mathrm{op}$
, using that it is also the derived functor of its restriction to the constructible heart.
Recall that by [Reference Ivorra and MorelIM22, Corollary 6.18], [Reference BondarkoBon11, Proposition 2.3.1] and [Reference HébertHéb11, Theorem 3.3] the categories
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
,
$\mathcal{D}^b(\mathrm{MHM}(X))$
and
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
admit weight structures (for the last one, note that Beilinson motives and étale motives are equivalent by [Reference Cisinski and DégliseCD19, Theorem 16.1.2]).
Proposition 6.9. Let X be a finite-type k-scheme. Then the functors


and

are compatible with weights.
Proof. First note that by definition of the weights on
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
, the
$\ell$
-adic realisation
$R_\ell :\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to\mathcal{D}^b_c(X,\mathbb{Q}_\ell)$
is compatible with weights and reflects weights (i.e., if a complex
$K\in\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
satisfies that
$R_\ell(K)$
is of weights less than or equal to w, then K is of weights less than or equal to w). As all weight filtrations and structures considered are bounded, it suffices to show that our functors map weight-zero objects to weight-zero objects. Let us first consider functors with source
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
. For these, is suffices to check that the image of the relative motive of a smooth proper X-scheme is pure of weight zero. For the
$\ell$
-adic realisation this is [Reference DeligneDel80] and [Reference MorelMor19]. Therefore, the functor
$\mathrm{Nor}^* : \mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
is also weight exact. The fact that
$\mathrm{Hdg}^*$
is exact is the fact that the mixed Hodge module is associated to a smooth proper X-scheme is pure of weight zero; see [Reference SaitoSai06, Theorem 6.7 and Proposition 6.6].
Now all that remains is
$R_H : \mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to\mathcal{D}^b(\mathrm{MHM}(X))$
. By [Reference Ivorra and MorelIM22, Corollary 6.27], any pure complex of perverse Nori motives is the direct sum of shifts of its
${\ ^{\mathrm{p}}\mathrm{H}}^i$
, hence to check that
$R_H$
is weight exact it suffices to show that it sends pure objects of
${\mathcal{M}_{\mathrm{perv}}}(X)$
to pure objects of
$\mathrm{MHM}(X)$
. By construction of
$R_H$
, we have a commutative diagram

with
$H_u$
the universal functor defining
${\mathcal{M}_{\mathrm{perv}}}(X)$
. But by [Reference Ivorra and MorelIM22, Corollary 6.27],
$H_u$
sends pure objects to pure objects, and any motive
$M\in{\mathcal{M}_{\mathrm{perv}}}(X)$
is a quotient of some
$\mathrm{H}_u(N)$
for
$N\in \mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
. Assume that M is pure of some weight w. Then M is a direct factor of a quotient of the weight filtration of
$\mathrm{H}^0(N)$
. As
$\mathrm{Hdg}^*$
is weight exact, it follows that
$R_H(M)$
is a direct factor of a mixed Hodge module of weight w, hence
$R_H$
is weight exact.
Remark 6.10. Our construction of the realisation functor
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
is very similar to that of Ivorra in [Reference IvorraIvo16] (which works for smooth varieties). Indeed, given a smooth variety X, in both definitions one first gives the image of the smooth X-schemes in
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
. To be more precise, Ivorra only gives the image of smooth affine X-schemes and then extends to smooth X-schemes by colimits, but as the functor
$(f:Y\to X)\mapsto f_!f^!\mathbb{Q}_X$
is a Nisnevich cosheaf, its value on smooth X-schemes is indeed determined by its value on smooth affine X-schemes. Once one has the value of the functor on smooth X-schemes, the rest of the construction is just a matter of applying the universal property of
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
: one has to check that the functor on smooth X-schemes is an
$\mathbb{A}^1$
-invariant étale sheaf and
$\mathbb{P}^1$
-stable. Therefore, it is easy to see that if both definitions on smooth affine schemes (ours with the six operations and Ivorra’s using cellular complexes and Beilinson’s basic lemma) coincide, then our functor
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
and Ivorra’s are equivalent. In particular, this would prove that Ivorra’s construction is compatible with the Betti realisation constructed by Ayoub on Voevodsky motives, hence that the category of perverse Nori motives as in [Reference Ivorra and MorelIM22] is equivalent to the diagram category of relative pairs studied in [Reference IvorraIvo17] and [Reference Ivorra and MorelIM22, Section 2.10].
This can be reduced to checking a simple fact, which the author did not manage to do: in [Reference IvorraIvo16, Proposition 4.15] Ivorra proves that for a given smooth affine map
$f:Y\to X$
with X a smooth variety, there is a specific cellular stratification
$Y^\bullet$
of Y such that the cellular complex induced on Y is an
$\mathrm{H}^0f_!$
-acyclic resolution of
$f^!\mathbb{Q}_X$
, hence that there is an isomorphism in
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
between Ivorra’s functor at Y and the homology of Y, that is,
$\mathrm{ra}^{\mathcal{M}_{\mathrm{perv}}}_X(Y,Y_\bullet)\simeq f_!f^!\mathbb{Q}_X$
. This isomorphism induces a canonical and functorial isomorphism on each
${\ ^{\mathrm{p}}\mathrm{H}}^i$
. The question is whether one can construct such an isomorphism in a functorial way, as this would give the desired isomorphism of functors. This does not seem easy, and it is unlikely that one can arrange those specific stratifications so that they are compatible with a given morphism of smooth affine X-schemes.
6.2 Nori motives as modules in étale motives
Notation 6.11. Recall that to be coherent with the notation of [Reference AyoubAyo22], we will denote by
$\mathrm{Nor}^*$
the realisation functor

where we have denoted by
$\mathcal{DN}$
the indization of the bounded derived category of perverse motives. We will denote by
$\mathcal{DN}_c$
its full subcategory of compact objects.
By construction, the functor
$\mathrm{Nor}^*$
preserves colimits, hence it has a right adjoint
$\mathrm{Nor}_*:\mathcal{DN}\to \mathcal{DM}^{\mathrm{\acute{e}t}}$
. Denote

This is an
$\mathbb{E}_\infty$
-ring object in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(k)$
. For X a finite-type k-scheme, denote by
$\mathscr{N}_{\mid X}\in\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
the restriction of
$\mathscr{N}_k$
to X. As in [Reference AyoubAyo22, Construction 1.91],
$\mathrm{Nor}^*$
factors canonically through the functor

as

Lemma 6.12. Let k be a field of characteristic zero and let
$\pi_X:X\to\mathrm{Spec} k$
be a finite-type k-scheme. The functor

is fully faithful. Moreover, the natural map
$\pi_X^*\mathrm{Nor}_*\mathbb{Q}_{\mathrm{Spec} k}\to\mathrm{Nor}_*\mathbb{Q}_X$
is an equivalence.
Proof. The proof adds nothing new to [Reference AyoubAyo22, Remark 1.97]. Denote by
$\widetilde{\mathrm{Nor}_*}$
the right adjoint of
$\widetilde{\mathrm{Nor}^*}$
. Assume first that
$X = \mathrm{Spec} k$
. We have to show that the unit map

is an equivalence. By [Reference IwanariIwa18, Lemma 2.6], the
$\infty$
-category
$\mathrm{Mor}_{\mathscr{N}_k}(\mathcal{DM}^{\mathrm{\acute{e}t}}(k))$
is compactly generated by
$f_\sharp\mathbb{Q}_Y(i)\otimes\mathscr{N}_k$
for
$f:Y\to \mathrm{Spec} k$
smooth and
$i\in \mathbb{Z}$
. As
$\widetilde{\mathrm{Nor}^*}$
of those compact generators are compact objects of
$\mathcal{DN}(k)$
, the right adjoint
$\widetilde{\mathrm{Nor}_*}$
preserves colimits. Thus, it suffices to check that for any compact object
$C\otimes\mathscr{N}_k$
of the above form, the map

is an equivalence. As the forgetful functor
$\mathrm{Mor}_{\mathscr{N}_k}(\mathcal{DM}^{\mathrm{\acute{e}t}}(k))\to \mathcal{DM}^{\mathrm{\acute{e}t}}(k)$
is conservative it is even sufficient to prove that

is an equivalence in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(k)$
. This is true because over a field, compact objects in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(k)$
are dualisable, and for dualisable objects this map is an equivalence by [Reference AyoubAyo14b, Lemma 2.8].
Now for a general X we use that the functor
$\widetilde{\mathrm{Nor}^*}$
commutes with the six operations (this follows from the fact that
$\mathscr{N}_X$
is a filtered colimit of dualisable objects, thus the projection formulae of the form
$f_*(M)\otimes \mathscr{N} \simeq f_*(M\otimes \mathscr{N})$
are true by [Reference AyoubAyo14b, Lemma 2.8], and this implies commutation with all the operations as in the proof of [Reference Cisinski and DégliseCD19, Theorem 4.4.2.5]). For compact
$M,N\in \mathrm{Mod}_{\pi_X^*\mathscr{N}_k}(\mathcal{DM}^{\mathrm{\acute{e}t}}(X))$
we have

which finishes the proof of full faithfulness. Now that we know that the unit

is an equivalence in
$\mathrm{Mod}_{\pi_X^*\mathscr{N}_k}(\mathcal{DM}^{\mathrm{\acute{e}t}}(X))$
, by plugging in
$\pi_X^*\mathscr{N}_k$
we obtain

and we have

Therefore, the canonical map

is an equivalence, thus it stays an equivalence after applying the forgetful functor, leaving us with the equivalence

in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
.
From the previous lemma there is no more ambiguity for
$\mathscr{N}_X=\pi_X^*\mathscr{N}_k\simeq \mathrm{Nor}_*\mathbb{Q}_X\in\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
.
Proposition 6.13. Let
$k=\mathrm{colim}_i k_i$
be a filtered colimit of fields of characteristic zero. Then the natural functors

and

are equivalences in
$\mathrm{Cat}_\infty$
and
$\mathrm{Pr}^L$
, respectively.
The same is true for
$\eta = \lim_i U_i$
with
$\eta\in X$
the generic point of an integral finite-type k-scheme and the
$U_i$
are the non-empty open subsets of X:

Proof. As the functor Ind, from small idempotent complete stable
$\infty$
-categories to compactly generated presentable stable categories, is an equivalence and the inclusion functor
$\mathrm{Pr}^L_\omega\to\mathrm{Pr}^L$
preserves colimits, the second statement follows from the first. By [Reference Ivorra and MorelIM22, Proposition 6.8] (or [Reference Ivorra and MorelIM22, Corollary 6.10] for the generic point case), the diagram
$\mathcal{M}:I^{\rhd}\to \mathrm{AbCAt}$
from I to the (2,1)-category of small abelian categories with exact functors sending
$i\in I$
to
${\mathcal{M}_{\mathrm{perv}}}(\mathrm{Spec} k_i)$
and the cone point v to
${\mathcal{M}_{\mathrm{perv}}}(\mathrm{Spec} k)$
is a colimit diagram. By [Reference Bunke, Cisinski, Kasprowski and WingesBCKW24, Theorem 7.4.9] the functor sending an Abelian category to the
$\infty$
-category of bounded complexes up to homotopy is a left adjoint, hence commutes with colimits, so that we obtain a colimit diagram

Now as
$\mathcal{D}^b(\mathcal{A})$
is a localisation of
$\mathcal{K}^b(\mathcal{A})$
for any abelian category
$\mathcal{A}$
, and as the transitions of
$\mathcal{K}^b\circ\mathcal{M}$
preserve quasi-isomorphisms,
$\mathcal{D}^b\circ \mathcal{M}$
is a colimit diagram.
Corollary 6.14. Let
$f:\mathrm{Spec} k\to\mathrm{Spec} \mathbb{Q}$
be a map of fields. Then the natural map
$f^*\mathscr{N}_\mathbb{Q}\to\mathscr{N}_k$
is an equivalence. Consequently, for any scheme X of finite type over k, if we denote by
$p_X:X\to\mathrm{Spec} \mathbb{Q}$
the unique map (which has to factor through f), then
$\mathscr{N}_X\simeq p_X^*\mathscr{N}_\mathbb{Q}$
.
Proof. The second claim follows from the first. Write
$k=\mathrm{colim}_i k_i$
with
$k_i/\mathbb{Q}$
of finite type. Then the same proof as in [Reference Ayoub, Gallauer and VezzaniAGV22, Lemma 3.5.7], using continuity of both
$\mathcal{DM}^{\mathrm{\acute{e}t}}$
and
$\mathcal{DN}$
for fields (Proposition 6.13) and that these categories are compactly generated, gives that the natural map

is an equivalence so that this reduces to the case where
$k/\mathbb{Q}$
is of finite type. Then k is the generic point of an integral
$\mathbb{Q}$
-scheme, and again the same proof as in [Reference Ayoub, Gallauer and VezzaniAGV22, Lemma 3.5.7], using continuity for the generic point of perverse Nori motives, reduces to showing that
$\pi_U^*\mathscr{N}_\mathbb{Q} \to \mathscr{N}_U$
is an equivalence for U of finite type over
$\mathbb{Q}$
, which is Lemma 6.12.
Therefore, there is now even less ambiguity for
$\mathscr{N}_X = (X\to \mathrm{Spec} \mathbb{Q})^*\mathscr{N}_\mathbb{Q}$
.
Proposition 6.15. Let
$X=\lim_i X_i$
be a limit of
$\mathbb{Q}$
-schemes with affine transitions. The natural morphism

is an equivalence where the colimit is computed in
$\mathrm{Pr}^{L}$
.
Proof. By [Reference Elmanto, Hoyois, Iwasa and KellyEHIK21, Lemma 5.1 (ii)], the functor

is an equivalence. Therefore, the assignment
$i\mapsto (\mathcal{DM}^{\mathrm{\acute{e}t}}(X_i,\mathbb{Q}),\mathscr{N}_{\mid X_i})$
is a colimit diagram in the category
$\mathrm{Pr}^{\mathrm{Alg}}$
of [Reference LurieLur17, Notation 4.8.5.10]. By [Reference LurieLur17, Theorem 4.8.5.11], the functor
$(\mathcal{C},A)$
that sends a presentable symmetric monoidal category
$\mathcal{C}$
together with an algebra object
$A\in\mathcal{C}$
to the category
$\mathrm{Mod}_A(\mathcal{C})$
has a right adjoint, hence preserves colimits. This is why the assignment
$i\mapsto \mathrm{Mod}_{\mathscr{N}_{X_i}}(\mathcal{DM}^{\mathrm{\acute{e}t}}(X_i,\mathbb{Q}))$
is a colimit diagram. We found a precise statement giving the symmetric monoidal structure in the paper by Ayoub, Gallauer and Vezzani. The precise statement is [Reference Ayoub, Gallauer and VezzaniAGV22, Lemma 3.5.6].
Proposition 6.16. The functor

is an equivalence for every finite-type k-scheme X.
Proof. The functor is fully faithful by Lemma 6.12. We first prove that it is essential surjective for
$X=\mathrm{Spec} K$
the spectrum of a field, not necessarily finite over k. As
$\mathcal{DN}(\mathrm{Spec} K)$
is compactly generated it suffices to show that
$\widetilde{\mathrm{Nor}^*}$
reaches all compact objects
$M\in\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(K))$
. As the functor is already fully faithful, by dévissage it suffices to check that any object of the heart
$M\in{\mathcal{M}_{\mathrm{perv}}}(K)$
is in the image of
$\widetilde{\mathrm{Nor}^*}$
. Any such object has a weight filtration whose graded pieces are direct factors of
$\mathrm{H}^{n+2j}(\pi_X)_*\mathbb{Q}_X(i)$
with
$\pi_X:X\to\mathrm{Spec} K$
a smooth projective morphism by Arapura [Reference Huber and Müller-StachHMS17, Theorem 10.2.5]. Note that here we used that if K is a finite-type extension of k, then the category
${\mathcal{M}_{\mathrm{perv}}}(\mathrm{Spec} K)$
obtained by colimit as motives over the generic point of a variety over k coincides with Nori motives over K. This holds with a beautiful proof in [Reference TerenziTer24, Proposition 1.15] or can be proven by using that the
$\ell$
-adic realisation is absolute, that is, does not depend on the base field. Therefore, by dévissage again it suffices to check that any such
$\mathrm{H}^{n+2j}(\pi_X)_*\mathbb{Q}_X(i)$
is in the image. But as
$(\pi_X)_*\mathbb{Q}_X(i)$
is pure of weight
$-2i$
, it is the direct sum of its
$\mathrm{H}^n[-n]$
, thus it suffices to show that each
$(\pi_X)_*\mathbb{Q}_X(i)$
is in the image, which is obvious by compatibility with the six operations.
Now for a finite-type k-scheme, the proposition is proven by Noetherian induction. Assume that X is such that the statement is true for any proper closed subset of X. We take an object
$M\in\mathcal{DN}(X)$
that we may assume to be compact. Let
$\eta$
be a generic point of X. By the case of a field, there exists
$N\in\mathrm{Mod}_\mathscr{N}(\mathcal{DM}^{\mathrm{\acute{e}t}}(\eta))$
, compact, such that
$\widetilde{\mathrm{Nor}^*}(N)=M_\eta$
. By Proposition 6.15 there exist an open subset U of X and an object
$N'\in\mathrm{Mod}_\mathscr{N}(\mathcal{DM}^{\mathrm{\acute{e}t}}(U))$
, compact, such that
$(N')_\eta = N$
. The functor
$\mathrm{colim}_{\eta\in U}\mathcal{DN}(U)\to\mathcal{DN}(\eta)$
is fully faithful by Proposition 6.13. Thus, the isomorphism
$\widetilde{\mathrm{Nor}^*}(N)\to M_\eta$
lifts to a smaller open subset V of U : we have
$\widetilde{\mathrm{Nor}^*}((N')_{\mid V})\simeq M_{\mid V}$
. Denote by Z the reduced closed complement of V in X. By induction hypothesis, there exists a
$N''\in\mathrm{Mod}_\mathscr{N}(\mathcal{DM}^{\mathrm{\acute{e}t}}(Z))$
such that
$\widetilde{\mathrm{Nor}^*}(N'')=M_{\mid Z}$
. The cofiber sequence
$\widetilde{\mathrm{Nor}^*}(j_!(N')_{\mid V})\to M \to \widetilde{\mathrm{Nor}^*}(i_*N'')$
with
$j:U\to X$
and
$i:Z\to X$
the immersion then ensures that M is the image of
$\widetilde{\mathrm{Nor}^*}$
, which finishes the proof.
Remark 6.17. The same arguments would prove that the
$\infty$
-category

of geometric origin objects in the indization of the derived category of mixed Hodge modules is also the category of modules over some algebra
$\mathscr{H}_X$
in
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
. Note that the lack of continuity of mixed Hodge modules would prevent things like
$p_X^*\mathscr{H}_{\mathrm{Spec} \mathbb{Q}}\simeq \mathscr{H}_{X}$
happening, so one would have to be careful of which algebra is used and an additional argument over the generic point will be needed, similarly to the case of the Betti realisation. Similar ideas are developed in [Reference DrewDre18] and the fact that one would obtain
$(X\to \mathrm{Spec} k)^*\mathscr{H}_{\mathrm{Spec} k}\simeq \mathscr{H}_X$
identifies Drew’s category as the full subcategory
$\mathcal{DH}^\mathrm{geo}(X)$
of geometric origin mixed Hodge modules. His considerations with enriched categories of motives also have an application here: his results, together with the same proof as above, would show that his category
$\mathbf{DH}$
of motivic enriched Hodge modules is the smallest full subcategory of
$\mathrm{Ind}\mathcal{D}^b(\mathrm{MHM}(-))$
that is stable under truncation and filtered colimts, and that contains the
$f_*g^*H$
for
$f:Y\to X$
proper,
$g:Y\to\mathrm{Spec} \mathbb{C}$
the structural morphism, and
$H\in\mathcal{D}(\mathrm{Ind}\mathrm{MHS}^p)$
. See [Reference TubachTub24] for a detailed discussion.
Corollary 6.18. The
$\infty$
-functor
$X\mapsto\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
has a unique extension
$\mathcal{DN}_c(X)$
to quasi-compact and quasi-separated schemes of characteristic zero such that for all limits
$X=\lim_i X_i$
of such schemes with affine transitions the natural functor

is an equivalence of
$\infty$
-categories, where we compute the colimit in
$\mathrm{Pr}^{L}$
.
Proof. Let
$\mathcal{C}$
be category of schemes that are of finite type over a field of characteristic zero. Let
$\mathrm{Sch}_\mathbb{Q}$
be the category of quasi-compact and quasi-separated
$\mathbb{Q}$
-schemes, and
$\iota : \mathcal{C}\to\mathrm{Sch}_\mathbb{Q}$
the inclusion. Nori motives, together with the pullback functoriality, define a functor
$\mathcal{C}^\mathrm{op} \to\mathrm{CAlg}(\mathrm{Cat}_\infty)$
which sends limits of schemes with affine transitions to colimits of
$\infty$
-categories by Proposition 6.15 together with the fact that the colimit is taken in
$\mathrm{Pr}^L$
and hence restricts to compact objects. We can denote by
$\mathcal{DN}_c$
its left Kan extension along
$\iota$
.
Corollary 6.19. Let X be a finite-type k-scheme. The
$\infty$
-category
$\mathcal{DN}(X)$
is compactly generated by the
$f_\sharp\mathbb{Q}_Y(i)$
for
$f:Y\to X$
smooth and
$i\in \mathbb{Z}$
.
Proof. Indeed, this the case for
$\mathcal{DM}^{\mathrm{\acute{e}t}}(X)$
, thus for
$\mathrm{Mod}_{\mathscr{N}_X}(\mathcal{DM}^{\mathrm{\acute{e}t}}(X))\simeq \mathcal{DN}(X)$
by [Reference IwanariIwa18, Lemma 2.6].
We finish with the proof of the following natural corollary, whose proof was surprisingly harder than expected for a result that the author took for granted for a long time.
Corollary 6.20. Let X be a quasi-projective k-scheme. The two functors

given by
${\ ^{\mathrm{p}}\mathrm{H}}^0\circ\mathrm{Nor}^*$
and the universal functor
$\mathrm{H}_\mathrm{univ}$
, are canonically equivalent.
Proof. First, note that the functor
${\ ^{\mathrm{p}}\mathrm{H}}^0\circ\mathrm{Nor}^*$
gives a factorisation of the
${\ ^{\mathrm{p}}\mathrm{H}}^0$
of the Betti realisation of étale motives

By the universal property of perverse Nori motives, this induces a canonical functor

such that
$R_B\circ F \simeq R_B$
and
${\ ^{\mathrm{p}}\mathrm{H}}^0\circ\mathrm{Nor}^*\simeq F\circ \mathrm{H}_\mathrm{univ}$
. Now the compatibility of
$\mathrm{Nor}^*$
with the operations ensures that for any perverse t-exact operation h the functor F commutes with the operation h. Hence, F commutes with
$i_*$
for i a closed immersion and with
$f^*[d]$
for f a smooth map of relative dimension d. We will still denote by F the functor induced by F on
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
. As F commutes with
$i_*$
, there exists an exchange morphism
$i^*\circ F\to F \circ i^*$
. Applying the Betti realisation (or the
$\ell$
-adic realisation if the field k is too big) where this exchange morphism is an equivalence, we see that F commutes with
$i^*$
. The same argument works to show that F commutes with
$f_\sharp$
for smooth f: because F commutes with
$f^*$
there is a natural transformation
$f_\sharp \circ F \to F\circ f_\sharp$
which is an equivalence after realisation, hence F also commutes with the
$f_\sharp$
functors. It also commutes with external tensor product because is it perverse t-exact, hence with the internal tensor product which is obtained from the external one by pulling back along the diagonal.
The functor F induces a compact and colimit-preserving functor

which commutes with all pullbacks, tensor products and
$f_\sharp$
for f-1pt, a smooth map. Denote by G its right adjoint. We claim that the natural transformation

is an equivalence. Indeed, by Corollary 6.19 we have that
$\mathrm{Ind}\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
is generated by colimits by the
$f_\sharp\mathbb{Q}_Y(i)$
for
$f:Y\to X$
a smooth map and
$i\in \mathbb{Z}$
. This means that it suffices to check that for
$M\in\mathrm{Ind}\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
, the map

is an equivalence. As G and F commute with filtered colimits and
$f_\sharp\mathbb{Q}_Y(i)$
is compact, we can assume that M is as well compact. Moreover, compact objects are constructed by extensions, finite limits, finite colimits and direct factors from the same
$f_\sharp\mathbb{Q}_Y(i)$
, thus by dévissage we can assume that
$M = g_\sharp\mathbb{Q}_Z(j)$
for some smooth map
$g:Z\to X$
and
$j\in\mathbb{Z}$
. Using that G is a right adjoint and that F commutes with the operations, the above map is equivalent to the map

and, as
$F(\mathbb{Q}) = \mathbb{Q}$
, this finishes the proof.
Now that F is fully faithful, it is automatically essentially surjective because we know it reaches the generators
$f_\sharp\mathbb{Q}(i)$
of
$\mathrm{Ind}\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
. Now, as F commutes with the operations, we have by universality of
$\mathcal{DM}^{\mathrm{\acute{e}t}}$
that
$F\circ \mathrm{Nor}^*\simeq \mathrm{Nor}^*$
. Indeed, because we proved that F commutes with the tensor product and because it is t-exact for the constructible t-structure we can see F as the derived functor of a symmetric monoidal functor on the constructible heart, apply Proposition 5.5 and deduce that F is a morphism of coefficient systems. This proves that the compositions of the morphisms
$\mathrm{Nor}^*$
and
$F\circ\mathrm{Nor}^*$
with the morphism
$\mathcal{SH}\to\mathcal{DM}^\mathrm{\acute{e}t}$
agree by universality of
$\mathcal{SH}$
, thus also their factorisation through the localisation
$\mathcal{DM}^\mathrm{\acute{e}t}$
of
$\mathcal{SH}$
. In particular over X,
$F\circ {\ ^{\mathrm{p}}\mathrm{H}}^0\circ \mathrm{Nor}^* \simeq {\ ^{\mathrm{p}}\mathrm{H}}^0\circ\mathrm{Nor}^*$
, hence
$F\circ F\circ \mathrm{H}_\mathrm{univ} \simeq {\ ^{\mathrm{p}}\mathrm{H}}^0\circ \mathrm{Nor}^*$
, and the universal property of
${\mathcal{M}_{\mathrm{perv}}}(X)$
tells us that
$F\circ F\simeq F$
, so that
$F\simeq \mathrm{Id}$
.
6.3 Relations with the t-structure conjecture
Let k be a field of characteristic zero. Recall the following conjecture ([Reference AndréAnd04, Section 21.1.7] and [Reference BeilinsonBei12]).
Conjecture 6.21. There exists a non-degenerate t-structure on
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(k)$
compatible with the tensor product and such that the
$\ell$
-adic realisation functor is t-exact.
In [Reference BeilinsonBei12] Beilinson proves that if the characteristic of k is zero the conjecture implies that the realisation functors of
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(k)$
are conservative. In [Reference BondarkoBon15, Theorem 3.1.4], Bondarko proves that if such a t-structure exists for all fields of characteristic zero, then there exists a perverse t-structure on
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
for each finite-type k-scheme X. The joint conservativity of the family of
$x^*$
for
$x\in X$
and the conservativity of the realisation functor over fields implies that in that case, the
$\ell$
-adic realisation over X is conservative and t-exact, when its target is endowed with the perverse t-structure.
Theorem 6.22. Assume Conjecture 6.21 for all fields of characteristic zero. Let X be a finite-type k-scheme. Then the heart of the perverse t-structure of
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
is canonically equivalent to the category
${\mathcal{M}_{\mathrm{perv}}}(X)$
of perverse Nori motives. Moreover, the functor
$\mathrm{Nor}^* : \mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
is an equivalence of stable
$\infty$
-categories. This implies that
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
is the derived category of the abelian category of both its perverse and constructible heart.
From now on, we assume that Conjecture 6.21 is true. We begin with two lemmas, in which we denote by
$\mathrm{Var}_k$
the category of quasi-projective k-varieties.
Lemma 6.23. Let X be quasi-projective scheme. Denote by
$\mathcal{M}(X)$
the heart of the perverse t-structure on
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
. There is a faithful exact functor
$F_X:{\mathcal{M}_{\mathrm{perv}}}(X)\to \mathcal{M}(X)$
that commutes with the
$\ell$
-adic realisation. Denote by

the canonical functor induced by the universal property of the bounded derived category applied to the functor

The composition
$\mathrm{Nor}^*_X\circ \gamma_X$
is equivalent to the identity functor
$\mathrm{Id}_{\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))}$
. Moreover, the
$\mathrm{ho}(\gamma_X)$
fit in a morphism
$\gamma$
of functors
$\mathrm{Var}^\mathrm{op}_k\to\mathrm{SymMono}_1$
with values in symmetric monoidal categories, using the
$(-)^*$
-functoriality.
Proof. By assumption the restriction to the heart of the
$\ell$
-adic realisation functor

is faithful exact and the composition
$\rho_\ell \circ \, \rm ^P{H^0} $
with the perverse
$$\rm ^P{H^0}$$
on
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
is isomorphic to the perverse
$\mathrm{H}^0$
of the
$\ell$
-adic realisation. By the universal property of perverse Nori motives [Reference Ivorra and MorelIM22, Proposition 6.10] this provides a faithful exact functor
$F_X\colon{\mathcal{M}_{\mathrm{perv}}}(X)\to\mathcal{M}(X)$
such that the composition
$\rho_\ell\circ F_X$
is isomorphic to the
$\ell$
-adic realisation
$R_\ell$
of perverse Nori motives and the composition
$F_X\circ\mathrm{H}_\mathrm{univ}$
with the universal functor
$\mathrm{H}_\mathrm{univ}\colon\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to{\mathcal{M}_{\mathrm{perv}}}(X)$
is the perverse
$$\rm ^P{H^0}$$
on
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
: we have a commutative (up to natural isomorphism) diagram.

Now consider the endofunctor

of
${\mathcal{M}_{\mathrm{perv}}}(X)$
. Because
$\mathrm{Nor}^*$
is perverse t-exact (this can be checked after
$\ell$
-adic realisation), if we compose
$F_X$
with the universal
$\mathrm{H}_\mathrm{univ}$
we obtain

as functors
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\to{\mathcal{M}_{\mathrm{perv}}}(X)$
. But by Corollary 6.20 the functor
${\ ^{\mathrm{p}}\mathrm{H}}^0\circ\mathrm{Nor}^*$
can canonically be identified with
$\mathrm{H}_\mathrm{univ}$
. Thus, we have a commutative diagram.

By the universal property of
${\mathcal{M}_{\mathrm{perv}}}(X)$
, the only vertical faithful exact functor fitting in the above diagram and making the two outer triangles commute is (up to natural isomorphism) the identity functor: this implies that there exists a natural equivalence
$\mathrm{Nor}^*\circ F_X\simeq\mathrm{Id}_{{\mathcal{M}_{\mathrm{perv}}}(X)}$
.
Now by the universal property of the bounded derived category we have a functor
$\mathcal{D}^b(\mathcal{M}(X))\to\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
, thus by composition with the (trivially) derived functor of
$F_X$
there is a functor

By definition, the restriction to
${\mathcal{M}_{\mathrm{perv}}}(X)$
of
$\gamma_X$
is precisely the functor
$F_X$
composed with the inclusion
$\mathcal{M}(X)\to\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
. In particular, the composition
$\mathrm{Nor}^*_X\circ\gamma_X$
induces the functor
$\mathrm{Nor}^*_X\circ F_X$
between the hearts. By the universal property of
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
and the above paragraph, this implies that there exists a natural equivalence
$\mathrm{Nor}^*_X\circ\gamma_X\simeq \mathrm{Id}_{\mathcal{D}({\mathcal{M}_{\mathrm{perv}}}(X))}$
.
Using [Reference Ivorra and MorelIM22, Section 2.2] and the t-exactness of the
$\ell$
-adic realisation of étale motives we see that there exist natural isomorphisms
$i_*\circ \gamma\simeq \gamma\circ i_*$
for i a closed immersion,
$f^*[d]\circ\gamma\simeq \gamma\circ f^*[d]$
for f a smooth morphism of relative dimension d, and
$\gamma(-)\boxtimes\gamma(-)\simeq \gamma(-\boxtimes -)$
, with
$\boxtimes$
the external tensor product. Indeed, for the pullback by a smooth map
$f\colon Y\to X$
we have a commutative diagram

obtained by the universal property of
${\mathcal{M}_{\mathrm{perv}}}(X)$
, which induces a commutative square

that gives the desired natural isomorphism by applying the universal property of the bounded derived category. The cases of
$i_*$
and
$\boxtimes$
are similar. By adjunction there is a natural transformation
$i^*\circ \gamma\Rightarrow \gamma\circ i^*$
, which is an isomorphism because
$\mathrm{Nor}^*$
is conservative, commutes with
$i^*$
and is a retraction of
$\gamma$
. Using the fact that any morphism of quasi-projective varieties can be written as a composition of a closed immersion and a smooth map, together with the formula for the tensor product using the external tensor product and pullbacks, we see that
$\mathrm{ho}(\gamma)$
commutes with pullbacks and tensor products, giving the last claim of the lemma.
By the lemma above we know that we have a natural transformation

of contravariant functors on quasi-projective with values in symmetric monoidal 1-categories. We need an
$\infty$
-categorical enhancement of this functor. This is the next lemma. Note that as for perverse Nori motives, the existence of a perverse t-structure on étale motives
$\mathcal{DM}^\mathrm{\acute{e}t}_c(X)$
implies that there is also a constructible t-structure, whose heart we denote by
$\mathcal{CM}(X)$
. The functors
$\gamma_X$
and
$\mathrm{Nor}^*_X$
are t-exact for the constructible t-structure, as this can be checked after
$\ell$
-adic realisation.
Lemma 6.24. Let X be a quasi-projective scheme. The functor
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\to\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
induces a symmetric monoidal functor
${\mathcal{M}_{\mathrm{ct}}}(X)\to \mathcal{CM}(X)$
. This gives a natural transformation

in
$\mathrm{Fun}(\mathrm{Var}_k^\mathrm{op},\mathrm{CAlg}(\mathrm{Cat}_\infty))$
, where we use the
$(-)^*$
-functoriality. Moreover,
$\delta$
is a morphism of coefficient systems and satisfies
$\mathrm{Nor}^*\circ\delta\simeq\mathrm{Id}_{\mathcal{D}^b({\mathcal{M}_{\mathrm{ct}}}(-))}$
. In fact, we have
$\mathrm{ho}(\delta)\simeq \gamma$
.
Proof. Because
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))\xrightarrow{\gamma_X}\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
and
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)\xrightarrow{\mathrm{Nor}^*}\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))$
are t-exact for the constructible t-structure, we have faithful exact functors
${\mathcal{M}_{\mathrm{ct}}}(X)\to\mathcal{CM}(X)$
and
$\mathcal{CM}(X)\to {\mathcal{M}_{\mathrm{ct}}}(X)$
whose composition is the identity of
${\mathcal{M}_{\mathrm{ct}}}(X)$
by Lemma 6.23.
Now by Proposition 5.6 there is a natural transformation

of functors
$\mathrm{Sch}_k^\mathrm{op}\to\mathrm{CAlg}(\mathrm{Cat}_\infty)$
. By the last part of Lemma 6.23 we have a functor
$\Delta^1\times\mathrm{Var}_k^\mathrm{op}\to\mathrm{SymMono}_1$
sending, for
$f\colon Y\to X$
a map of quasi-projective varieties, the map
$(0<1,f)$
to
${\mathcal{M}_{\mathrm{ct}}}(Y)\xrightarrow{f^*}{\mathcal{M}_{\mathrm{ct}}}(X)\xrightarrow{\gamma_X}\mathcal{CM}(X)$
, which is also isomorphic to
${\mathcal{M}_{\mathrm{ct}}}(Y)\xrightarrow{\gamma_Y}\mathcal{CM}(Y)\xrightarrow{f^*}\mathcal{CM}(X)$
. Thus, by Proposition 5.6 there is a natural transformation

of functors
$\mathrm{Var}_k^\mathrm{op}\to \mathrm{CAlg}(\mathrm{Cat}_\infty)$
. Composing the two, we obtain a natural transformation

of functors with values in symmetric monoidal stable
$\infty$
-categories. The restriction to
${\mathcal{M}_{\mathrm{ct}}}(X)$
of
$\delta_X$
is, by definition, the restriction to
${\mathcal{M}_{\mathrm{ct}}}(X)$
of
$\gamma_X$
. Thus,
$\delta_X\simeq \gamma_X$
by Theorem 5.3.
To prove that
$\delta$
is a morphism of coefficient systems, it suffices to prove that if
$f\colon Y\to X$
is a smooth morphism of varieties, the natural transformation
$f_\sharp\circ\delta_Y\to\delta_X\circ f_\sharp$
is an equivalence. For this, we may apply the conservative functor

so that the exchange morphism above becomes

Now remark that for S a variety, the composition
$\mathrm{Nor}^*_S\circ\delta_S$
is the identity functor and the functor
$\mathrm{Nor}^*$
commutes with the
$\ell$
-adic realisation, thus we have an isomorphism

As the functor
$R_\ell$
commutes with
$f_\sharp$
, the proof is finished.
Proof of Theorem 6.22. Because both
$\mathcal{DM}^\mathrm{\acute{e}t}_c$
and
$\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(-))$
are Zariski hypersheaves and
$\mathrm{Nor}^*$
is a morphism of sheaves, we may assume that X is quasi-projective.
Let
$\delta_X$
be the functor of Lemma 6.24. We know that
$\mathrm{Nor}^*_X\circ\delta_X\simeq\mathrm{Id}_{\mathcal{D}^b({\mathcal{M}_{\mathrm{perv}}}(X))}.$
We can now show that
$\delta_X\circ\mathrm{Nor}^*_X\simeq\mathrm{Id}_{\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)}$
. By Lemma 6.24 and Theorem 6.5 the functor
$\delta_X\circ\mathrm{Nor}^*_X$
is a morphism of coefficient systems. In particular, its composition with the canonical symmetric monoidal functor

is isomorphic to the canonical symmetric monoidal functor
$M_X$
. Using the universal property of
$\mathcal{DM}^{\mathrm{\acute{e}t}}_c(X)$
proved by Robalo in [Reference RobaloRob15, Corollary 2.29] (or rather, its étale and
$\mathbb{Q}$
-linear counterpart which can be deduced from it) this implies that

This finishes the proof as now we have that
$\mathrm{Nor}^*_X$
and
$\delta_X$
are inverses of each other.
Appendix A. Coefficient systems
Recall the following definition due to Drew and Gallauer in [Reference Drew and GallauerDG22].
Definition A.1 A functor
$C:\mathrm{Sch}^\mathrm{op}_k\to\mathrm{CAlg}(\mathrm{Cat}_\infty^\mathrm{st})$
taking values in symmetric monoidal stable
$\infty$
-categories and exact symmetric monoidal functors is called a coefficient system if it satisfies the following properties.
-
(i) Pushforwards. For every
$f:Y\to X$ in
$\mathrm{Sch}_k$ , the pullback functor
$f^*$ admits a right adjoint
$f_*:C(Y)\to C(X)$ .
-
(ii) Internal homs. For every
$X\in\mathrm{Sch}_k$ , the symmetric monoidal structure on C(X) is closed.
-
(iii). For each smooth morphism
$p:Y\to X\in\mathrm{Sch}_k$ , the functor
$p^*:C(X)\to C(Y)$ admits a left adjoint
$p_\sharp$ , and the following:
-
(a) Smooth base change. For each cartesian square

in
$\mathrm{Sch}_k$
, the exchange transformation
$p'_\sharp (f')^*\to f^*p_\sharp$
is an equivalence.
-
(b) Smooth projection formula. The exchange transformation
\[p_\sharp(p^*(-)\otimes -)\to -\otimes p_\sharp(-)\]
$C(X)\times C(Y)\to C(Y)$ .
-
(iv) Localisation.
$C(\emptyset)\simeq 0$ and for each closed immersion
$Z\to X$ in
$\mathrm{Sch}_k$ with complementary open immersion
$j:U\to X$ , the square

is cartesian in
$\mathrm{Cat}_\infty^\mathrm{st}$
.
-
(v)
$\mathbb{A}^1$ -homotopy invariance. For each
$X\in\mathrm{Sch}_k$ , if
$\pi_{\mathbb{A}^1}:\mathbb{A}^1_X\to X$ denotes the canonical projection then the functor
$\pi_{\mathbb{A}^1}^*:C(X)\to C(\mathbb{A}^1_X)$ is fully faithful.
-
(vi) T-stability. The composite
$\pi_{\mathbb{A}^1,\sharp} s_*:C(X)\to C(X)$ is an equivalence.
A morphism of coefficient systems is a natural transformation
$\phi:C\to C'$
such that for each smooth morphism
$p:Y\to X$
in
$\mathrm{Sch}_k$
, the exchange transformation

is an equivalence. This defines a sub-
$\infty$
-category
$\mathrm{CoSy}_k\subset\mathrm{Fun}(\mathrm{Sch}^\mathrm{op}_k,\mathrm{CAlg}(\mathrm{Cat}_\infty^\mathrm{st}))$
. We say that a coefficient system C is cocomplete if it takes values in symmetric monoidal cocomplete
$\infty$
-categories and colimit-preserving functors. We denote by
$\mathrm{CoSy}_k^c$
their
$\infty$
-category.
Recall the main theorem of [Reference Drew and GallauerDG22].
Theorem A.2 (Drew and Gallauer). The stable
$\mathbb{A}^1$
-homotopy
$\infty$
-category defines an object
$\mathcal{SH}\in \mathrm{CoSy}_k^c$
. It is the initial object.
Definition A.3. We let
$\mathrm{QPCoSy}_k$
be the
$\infty$
-category defined in the same terms as in Definition A.1 but with functors

with source quasi-projective varieties instead of finite-type k-schemes.
Proposition A.4 The restriction functor

is an equivalence of
$\infty$
-categories with quasi-inverse the right Kan extension.
Proof. By [Reference GallauerGal22, Corollary 2.19] or [Reference Drew and GallauerDG22, Proposition 7.13], any coefficient system (on all finite-type k-schemes or varieties) has Nisnevich descent, thus Zariski descent. Restriction and right Kan extension give an equivalence of categories between Zariski sheaves

because any finite-type k-scheme admits a Zariski covering by affine schemes. Therefore, it suffices to check that if
$C\in \mathrm{QPCoSy}_k$
, its right Kan extension to
$\mathrm{Sch}^\mathrm{op}$
is an object of
$\mathrm{CoSy}_k$
, and that a morphism in
$\mathrm{QPCoSy}_k$
gives a morphism in
$\mathrm{CoSy}_k$
after Kan extension.
We denote by
$\mathcal{D}$
the right Kan extension of the functor

along the inclusion
$\mathrm{Var}_k^\mathrm{op} \to \mathrm{Sch}_k^\mathrm{op}$
. Because C is a Zariski sheaf, for
$X\in\mathrm{Sch}_k$
we may choose a Zariski covering
$p:U\to X$
with U affine, and as p is quasi-projective, all
$U_n := U^{\times_X {n+1}}$
are quasi-projective over k and we have

in
$\mathrm{CAlg}(\mathrm{Cat}_\infty)$
. But then if
$f: Y\to X$
is a smooth morphism, base change induces a map of simplicial schemes
$f_\bullet:V_\bullet\to U_\bullet$
with
$V_n = U_n\times_X Y$
and we also have

in
$\mathrm{CAlg}(\mathrm{Cat}_\infty)$
. In fact, the functor

is the limit of the functors

in the
$\infty$
-category
$\mathrm{Fun}(\Delta^1,\mathrm{Cat}_\infty)$
. Because each
$f_n^*$
has a left adjoint
$(f_n)_\sharp$
, and for every map
$[n]\to [m]$
in
$\Delta$
the square

commutes thanks to the canonical Beck–Chevalley morphism and the smooth base change property, the diagram
$(f_n^*)_{[n]\in \Delta}$
in fact takes values in
$\mathrm{Fun}^{\mathrm{LAd}}(\Delta^1,\mathrm{Cat}_\infty)$
, the
$\infty$
-category of left adjointable functors (see [Reference LurieLur17, Definition 4.7.4.16]). Thus, by [Reference LurieLur17, Corollary 4.7.4.18] the limit
$f^*$
is also a left adjointable functor: the left adjoint
$f_\sharp$
exists and has a base change property with respect to all the functors
$\mathcal{D}(X)\to C(U_n)$
. Moreover, because
$f^*$
is symmetric monoidal for
$M\in \mathcal{D}(X)$
and
$N\in\mathcal{D}(Y)$
there exists a canonical map

This map is an equivalence. Indeed, if
$M_n = M_{\mid U_n}$
, we have a map of diagrams
$(f_n^*)_n\to (f_n^*)_n$
given by the commutative squares

which, because of the projection formula for each
$f_n$
and the base change for
$(f_m)_\sharp$
, is a map of diagrams with values in
$\mathrm{Fun}^{\mathrm{LAd}}(\Delta^1,\mathrm{Cat}_\infty)$
. Thus, at the limit the map
$f^*\to f^*$
in
$\mathrm{Fun}(\Delta^1,\mathrm{Cat}_\infty)$
given by

is left adjointable, giving the projection formula.
Moreover, if
$g\colon X'\to X$
is a map in
$\mathrm{Sch}_k$
we can form the pullback

and, doing a base change to
$U_\bullet$
, we obtain a pullback square of simplicial objects in
$\mathrm{Var}_k$
of the form

Then, as above, the commutative square

can be seen as a map
$f^*\to (f')^*$
in
$\mathrm{Fun}(\Delta^1,\mathrm{Cat}_\infty)$
, which itself is a limit of maps
$b'_n\to b_n$
in
$\mathrm{Fun}(\Delta^1,\mathrm{Cat}_\infty)$
. As this last simplicial diagram of maps takes values in
$\mathrm{Fun}^\mathrm{LAd}(\Delta^1,\mathrm{Cat}_\infty)$
because the smooth base change holds for maps in
$\mathrm{Var}_k$
, the limit is left adjointable, giving the smooth base for the pair (f,g).
The existence of pushforwards and internal
$\mathscr{{H}}\,$
is proven in a similar way to the existence of the
$f_\sharp$
(using right adjointable squares instead of left adjointable). The only thing to check in order to obtain those functors as functors to a limit of
$\infty$
-categories is that over quasi-projective schemes these functors commute with pullbacks by Zariski covering so that they define a compatible system as in (A1). This is the case because these pullbacks
$\pi^*$
are right adjoints to
$\pi_\sharp$
and the smooth base change and the projection formula ensure the commutativity of
$\pi_\sharp$
with the left adjoints of internal
$\mathscr{{H om}}\,$
and pushforwards. Localisation follows from the fact that the category
$\mathrm{Sq}^\mathrm{cart}(\mathrm{Cat}_\infty)$
of cartesian squares in
$\mathrm{Cat}_\infty$
is closed under limits in the category of commutative squares of
$\infty$
-categories. The properties of
$\mathbb{A}^1$
-invariance and T-stability can be checked Zariski locally thanks to the smooth projection formula and the smooth base change, hence they hold. The same is true for extensions of morphisms of coefficient systems.
Remark A.5. The above proposition on extension of coefficient system from quasi-projective to finite-type objects works verbatim if one replaces
$\mathrm{Spec} k$
with any Noetherian finite-dimensional scheme B. Moreover, the proposition also holds with quasi-projective varieties replaced by separated reduced finite-type k-schemes.
Finally, recall the following theorem.
Theorem A.6 (Ayoub, Cisinski, Dàlise, Rönsdings and Voevodsky). Any coefficient system has the six functors.
Acknowledgements
This work was written during my PhD supervised by Sophie Morel at the UMPA in Lyon. I would like to thank her for the faith and constant attention she has for my work and for the immense amount of time she is spending with me on this thesis. I am grateful to the anonymous referee for their valuable comments and suggestions that helped make this paper more readable. I am much obliged to Joseph Ayoub for finding a mistake during a talk I gave, whose correction simplified the third part of this paper. I also thank Frédéric Déglise for sharing some ideas with me, and Robin Carlier for taking the time to explain to me so many facts about
$\infty$
-categories. I had valuable email exchanges with Marc Hoyois. I also had useful conversations with Georg Lehner, Raphaël Ruimy and Luca Terenzi.
Conflicts of interest
None.
Financial support
The author was funded by the École Normale Supérieure de Lyon.
Journal information
Compositio Mathematica is owned by the Foundation Compositio Mathematica and published by the London Mathematical Society in partnership with Cambridge University Press. All surplus income from the publication of Compositio Mathematica is returned to mathematics and higher education through the charitable activities of the Foundation, the London Mathematical Society and Cambridge University Press.