1 Introduction and Main Results

Local-to-global properties are ubiquitous in topology, geometry, and quantum field theory. The prototypical example of a local-to-global, or descent, property is the gluing of local sections of a sheaf: given a manifold M with an open covering \(\{U_i\}_{i \in I}\), global sections of a sheaf F on M are in bijection with families \(\{f_i \in F(U_i)\}_{i \in I}\) such that \(f_{i|U_{ij}} = f_{j|U_{ij}}\) for all \(i,j \in I\), with \(U_{ij} {:}{=}U_i \cap U_j\).

Geometric structures on manifolds are, in general, not described by sheaves of sets; instead, one needs to pass to sheaves of higher categories. For example, complex vector bundles form a sheaf on the site of manifolds and open coverings which is valued in categories rather than sets. In the works of Schreiber [\(\underline{{\mathscr {H}}}_{\infty ,0}(Q\underline{M},F)\). Here, \(\underline{M}\) is the presheaf on \({{\mathscr {C}}\text {art} }\) consisting of smooth maps to M, the functor Q is a cofibrant replacementFootnote 2 in \({\mathscr {H}}_{\infty ,0}\), and \(\underline{{\mathscr {H}}}_{\infty ,0}(-,-)\) denotes the -enriched hom functor in \({\mathscr {H}}_{\infty ,0}\). If F is the simplicial presheaf on cartesian spaces which describes G-bundles or n-gerbes, for instance, then the Kan complex \(\underline{{\mathscr {H}}}_{\infty ,0}(Q\underline{M},F)\) is the \(\infty \)-groupoid of G-bundles or n-gerbes on M, respectively. In fact, we could use any presheaf X on \({{\mathscr {C}}\text {art} }\) in place of \(\underline{M}\) and thus study geometries on any such X. (Later we will even allow X to be a presheaf of \(\infty \)-groupoids itself.)

We write \({{\widehat{{\mathscr {C}}}}}= \textrm{Fun}({\mathscr {C}}^{\text {op}},{{\mathscr {S}}\text {et} })\) for the category of (small) presheaves of sets on \({\mathscr {C}}\). Motivated by the above discussion, we pose the following two questions for sheaves on a generic site \(({\mathscr {C}},\tau )\):

  1. (1)

    If F satisfies descent with respect to the pretopology \(\tau \), does the assignment \(X \mapsto \underline{{\mathscr {H}}}_{\infty ,0}(QX,F)\) satisfy descent with respect to a related pretopology \({{\widehat{\tau }}}\) on \({{\widehat{{\mathscr {C}}}}}\)?

  2. (2)

    If so, can we understand which \({{\widehat{\tau }}}\)-sheaves on \({{\widehat{{\mathscr {C}}}}}\) arise as extensions \(\underline{{\mathscr {H}}}_{\infty ,0}(Q(-),F)\) of some ordinary sheaf F on \(({\mathscr {C}},\tau )\)?

We answer both of these questions in this paper, working over general Grothendieck sites \(({\mathscr {C}},\tau )\): any pretopology \(\tau \) on \({\mathscr {C}}\) induces a pretopology \({{\widehat{\tau }}}\) of generalised coverings on \({{\widehat{{\mathscr {C}}}}}\). We prove that with respect to such a pair of pretopologies, the answer to the first question is always ‘yes’. One way of saying this is that the process of extending any geometric structure from objects of \({\mathscr {C}}\) to objects of \({{\widehat{{\mathscr {C}}}}}\) does not destroy the descent property; we can still think of the extended structure as having geometric flavour and perform local-to-global constructions.

In concrete terms, the induced Grothendieck pretopology \({{\widehat{\tau }}}\) on \({{\widehat{{\mathscr {C}}}}}\) consists of the \(\tau \)-local epimorphisms, also called generalised coverings [17], which are defined as follows. Let \({{\mathscr {Y}}}:{\mathscr {C}}\rightarrow {{\widehat{{\mathscr {C}}}}}\) denote the Yoneda embedding of \({\mathscr {C}}\). A morphism \(\pi :Y \rightarrow X\) in \({{\widehat{{\mathscr {C}}}}}\) is a \(\tau \)-local epimorphism if, for every \(c \in {\mathscr {C}}\) and every map \({{\mathscr {Y}}}_c \rightarrow X\), there exists a covering \(\{c_i \rightarrow c\}_{i \in I}\) in the site \(({\mathscr {C}},\tau )\) such that each composition \({{\mathscr {Y}}}_{c_i} \rightarrow {{\mathscr {Y}}}_c \rightarrow X\) factors through \(\pi \); this is an abstract way of saying that the morphism \(\pi \) has local sections.

Let \({{\mathscr {M}}}\) be a left proper cellular or combinatorial simplicial model category. We consider the projective model structures on \(\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})\) and \(\textrm{Fun}({{\widehat{{\mathscr {C}}}}}{}^{\text {op}}, {{\mathscr {M}}})\), and their left Bousfield localisations \(\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})^{loc}\) and \(\textrm{Fun}({{\widehat{{\mathscr {C}}}}}{}^{\text {op}}, {{\mathscr {M}}})^{loc}\) at the \(\tau \)-coverings and \({{\widehat{\tau }}}\)-coverings, respectively. That is, the fibrant objects in these localised model categories are \({{\mathscr {M}}}\)-valued homotopy sheaves on \(({\mathscr {C}}, \tau )\) and \(({{\widehat{{\mathscr {C}}}}}, {{\widehat{\tau }}})\), respectively. We show the following result (see Theorems 3.303.37 and Proposition 3.38 below):

Theorem 1.1

Let \({{\mathscr {M}}}\) be a left proper simplicial model category which is cellular or combinatorial and \({\mathscr {C}}\) a category. Let \(\tau \) be a Grothendieck coverage on \({\mathscr {C}}\), and \({{\widehat{\tau }}}\) the induced Grothendieck coverage of \(\tau \)-local epimorphisms on \({{\widehat{{\mathscr {C}}}}}\). There is a Quillen adjunction

figure a

whose right adjoint restricts on fibrant objects to the homotopy right Kan extension along the Yoneda embedding \({{\mathscr {Y}}}:{\mathscr {C}}\hookrightarrow {{\widehat{{\mathscr {C}}}}}\),

$$\begin{aligned} S_{{{\mathscr {M}}}|{{\mathscr {M}}}_{fib}} \cong {\textrm{hoRan}}_{{\mathscr {Y}}}\,. \end{aligned}$$

A prime application is the case of homotopy sheaves of higher categories: denoting by \(\mathscr {C}\mathscr {S}\mathscr {S}_n\) the model category of n-fold complete spaces [2, 9, 41], let \({\mathscr {H}}_{\infty ,n}^{loc}\) be the left Bousfield localisation of the projective model structure on \(\textrm{Fun}({\mathscr {C}}^{\text {op}}, \mathscr {C}\mathscr {S}\mathscr {S}_n)\) at the Čech nerves of coverings in \(({\mathscr {C}},\tau )\). Similarly, we let \({{\widehat{{\mathscr {H}}}}}_{\infty ,n}^{loc}\) denote the left Bousfield localisation of \(\textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, \mathscr {C}\mathscr {S}\mathscr {S}_n)\) at the Čech nerves of the \(\tau \)-local epimorphisms. As a direct corollary to Theorem 1.1 we obtain:

Corollary 1.2

Let \(n \in {\mathbb {N}}_0\), and let \({\mathscr {C}}\) be a small category.

  1. (1)

    For any fibrant \({\mathscr {F}}\in {\mathscr {H}}_{\infty ,n}\), the presheaf \(\underline{{\mathscr {H}}}_{\infty ,n} \big ( Q(-), {\mathscr {F}}\big )\) is equivalent to the homotopy right Kan extension \({\textrm{hoRan}}_{{{\mathscr {Y}}}} {\mathscr {F}}\) of \({\mathscr {F}}\) along the Yoneda embedding \({{\mathscr {Y}}}:{\mathscr {C}}\rightarrow {{\widehat{{\mathscr {C}}}}}\). In particular, \(\underline{{\mathscr {H}}}_{\infty ,n} \big ( Q(-), {\mathscr {F}}\big )\) presents the \(\infty \)-categorical right Kan extension of presheaves of \((\infty ,n)\)-categories along \({{\mathscr {Y}}}\).

  2. (2)

    If \(({\mathscr {C}},\tau )\) is a Grothendieck site, there is a Quillen adjunction

    figure b

This answers our first question affirmatively since \(\underline{{\mathscr {H}}}_{\infty ,n} (Q(-), -)\) preserves fibrant objects.

In Sect. 4.1 we deduce the following strictification and descent result (Theorem 4.18 and Theorem 4.12, respectively), thereby filling significant gaps in the literature on diffeological spaces:

Theorem 1.3

Let \({{\mathscr {V}}\hspace{-0.02cm}{\mathscr {B}}\hspace{-0.02cm}\text {un} }_{{\mathscr {D}}{\textrm{fg}}}:{{\mathscr {D}}{\textrm{fg}}}^{\text {op}}\rightarrow {{\mathscr {C}}\text {at} }\) be the pseudo-functor which assigns to a diffeological space its category of diffeological vector bundles. The following statements hold true:

  1. (1)

    There is a strict functor \(\iota ^* {{\mathscr {V}}\hspace{-0.02cm}{\mathscr {B}}\hspace{-0.02cm}\text {un} }_{str}\) (which we construct explicitly) and an objectwise equivalence of pseudo-functors \({{\mathscr {D}}{\textrm{fg}}}^{\text {op}}\rightarrow {{\mathscr {C}}\text {at} }_\text {L} \),

    $$\begin{aligned} {\mathcal {A}}:\iota ^* {{\mathscr {V}}\hspace{-0.02cm}{\mathscr {B}}\hspace{-0.02cm}\text {un} }_{str} \overset{\sim }{\longrightarrow }{{\mathscr {V}}\hspace{-0.02cm}{\mathscr {B}}\hspace{-0.02cm}\text {un} }_{{\mathscr {D}}{\textrm{fg}}}\,. \end{aligned}$$
  2. (2)

    Diffeological vector bundles satisfy descent along subductions of diffeological spaces.

Here we use Rezk’s classifying diagram functor to deduce a 1-categorical analogue of Corollary 1.2 for \(n=1\). This readily provides the desired descent result; the main remaining work in proving Theorem 1.3 goes into the strictification of \({{\mathscr {V}}\hspace{-0.02cm}{\mathscr {B}}\hspace{-0.02cm}\text {un} }_{{\mathscr {D}}{\textrm{fg}}}\).

Further, we deduce from Corollary 1.2 that various flavours of the 2-category of bundle gerbes with connection as introduced by Waldorf [4.34 and 4.35).

In Sect. 5 we answer the second question above. To do so, we pass to quasi-categorical language: let \({\textsf {C}}\) be a quasi-category with a Grothendieck (pre)topology \(\tau \). Let \({\widehat{{\textsf {C}}}_\infty }\) denote the quasi-category of (small) presheaves of spaces on \({\textsf {C}}\). We write \({\textsf {Sh}}({\textsf {C}},\tau )\) for the quasi-category of \(\tau \)-sheaves on \({\textsf {C}}\) and \({\textsf {Sh}}_!({\widehat{{\textsf {C}}}_\infty },{{\widehat{\tau }}})\) for the quasi-category \({{\widehat{\tau }}}\)-sheaves on \({\widehat{{\textsf {C}}}_\infty }\) whose underlying functor \({\widehat{{\textsf {C}}}_\infty }\rightarrow {\textsf {S}}^{\text {op}}\) preserves colimits.

Theorem 1.4

Let \(({\textsf {C}},\tau )\) be as above, and let \({{\mathscr {Y}}}_*\) denote the \(\infty \)-categorical right Kan extension of presheaves along the Yoneda embedding \({{\mathscr {Y}}}:{\textsf {C}}^{\text {op}}\rightarrow {\widehat{{\textsf {C}}}_\infty }^{\text {op}}\). The adjunction \({{\mathscr {Y}}}^* \dashv {{\mathscr {Y}}}_*\) restricts to an equivalence of quasi-categories

figure c

Finally, to make contact with the first part of this paper, we prove a similar result where \({\textsf {C}}= {\textsf {N}}{\mathscr {C}}\) is an ordinary category; this relates quasi-categorical \(\tau \)-sheaves of spaces on \({\textsf {N}}{\mathscr {C}}\) to \({{\widehat{\tau }}}\)-sheaves of spaces on the 1-category of set-valued presheaves \({{\widehat{{\mathscr {C}}}}}= \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {S}}\text {et} }_\text {L} )\) (Theorem 5.14).

1.1 Conventions

Sizes and universes Throughout, we choose and fix a nested triple of Grothendieck universes \(\text {S} \in \text {L} \in \text {XL} \),Footnote 3 and we assume that \(\text {S} \) contains the natural numbers. We write \({{\mathscr {S}}\text {et} }_\text {S} \) and for the categories of \(\text {S} \)-small sets and \(\text {S} \)-small simplicial sets, respectively, and analogously for the universes \(\text {L} \) and \(\text {XL} \). As is common in the literature, we refer to the elements in \({{\mathscr {S}}\text {et} }_\text {S} \) simply as small sets, those of \({{\mathscr {S}}\text {et} }_\text {L} \) as large sets and those of \({{\mathscr {S}}\text {et} }_\text {XL} \) as extra large sets. All indexing sets will be assumed to be \(\text {S} \)-small.

Let \({\mathscr {C}}\) be a small category. Consider the category \({{\widehat{{\mathscr {C}}}}}{:}{=}\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {S}}\text {et} }_\text {S} )\) of \({{\mathscr {S}}\text {et} }_\text {S} \)-valued presheaves on \({\mathscr {C}}\). Observe that \({{\widehat{{\mathscr {C}}}}}\) is no longer small, since \({{\mathscr {S}}\text {et} }_\text {S} \) is not small. However, since \({{\mathscr {S}}\text {et} }_\text {S} \) is a large category (its objects and morphisms each form sets in \(\text {L} \)), it follows that \({{\widehat{{\mathscr {C}}}}}\) is a large category.

The Yoneda embedding \({{\widehat{{{\mathscr {Y}}}}}}:{{\widehat{{\mathscr {C}}}}}\rightarrow \textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, {{\mathscr {S}}\text {et} }_\text {L} )\) is fully faithful; observing that an object of \({{\mathscr {S}}\text {et} }_\text {S} \) is also an object of \({{\mathscr {S}}\text {et} }_\text {L} \), the standard proof applies. Further, the Yoneda Lemma holds true for both \({{\mathscr {Y}}}\) and \({{\widehat{{{\mathscr {Y}}}}}}\) (again by the usual method of proof). Finally, observe for any \(c \in {\mathscr {C}}\) and for any \(X \in {{\widehat{{\mathscr {C}}}}}\), by the Yoneda Lemma, there are canonical isomorphisms

$$\begin{aligned} {{\widehat{{{\mathscr {Y}}}}}}_X ({{\mathscr {Y}}}_c) = {{\widehat{{\mathscr {C}}}}}({{\mathscr {Y}}}_c, X) \cong X(c) \in {{\mathscr {S}}\text {et} }_\text {S} \subset {{\mathscr {S}}\text {et} }_\text {L} \,. \end{aligned}$$

Categories and higher categories In this article we use two different models for higher categories. Section 2 and Appendix A do not make use of higher categories. In Sects. 3 and 4 we use n-fold complete Segal spaces (introduced in [2]) as a model for \((\infty ,n)\)-categories. In particular, the term ‘\(\infty \)-groupoid’ shall always refer to a Kan complex, and ‘\((\infty ,n)\)-category’ shall always mean a n-fold complete Segal space. In Sect. 5, and only there, all higher categories are modelled as quasi-categories in the sense of [12, 27, 34]. In order to make clear the distinction between ordinary categories (possibly endowed with model structures), n-fold complete Segal spaces, and quasi-categories, we denote ordinary categories by script letters \({\mathscr {C}}, {\mathscr {D}}, \ldots \), n-fold complete Segal spaces (or presheaves thereof) by ordinary capitals \(F, G, \ldots \), and quasi-categories by sans-serif capitals \({\textsf {C}}, {\textsf {D}}, \ldots \). In particular, the quasi-categories of small spaces is denoted \({\textsf {S}}_\text {S} \), and analogously for the universes \(\text {L} \) and \(\text {XL} \). The term ‘small space’ (used on its own, as opposed to ‘Segal space’, for instance) shall always mean an object in the quasi-category \({\textsf {S}}_\text {S} \) of small spaces, and analogously for the terms ‘large space’ and ‘extra large space’ and the universes \(\text {L} \) and \(\text {XL} \), respectively.

Enriched categories If \({{\mathscr {V}}}\) is a monoidal category and \({\mathscr {C}}\) is a \({{\mathscr {V}}}\)-enriched category (or \({{\mathscr {V}}}\)-category), we denote the \({{\mathscr {V}}}\)-enriched hom-objects of \({\mathscr {C}}\) by \(\underline{{\mathscr {C}}}^{{\mathscr {V}}}(-,-)\). If is the category of simplicial sets, we also write . If \({{\mathscr {V}}}\) is a symmetric monoidal model category and \({\mathscr {C}}\) is a \({{\mathscr {V}}}\)-enriched model category (in the sense of [23, Sec. 4]), we will equivalently say that \({\mathscr {C}}\) is a model \({{\mathscr {V}}}\)-category.

Tractability We recall from [3, Def. 1.21] that a large model category \({{\mathscr {M}}}\) is tractable if there exists a regular \(\text {L} \)-small cardinal \(\lambda \) such that the category underlying \({{\mathscr {M}}}\) is locally \(\lambda \)-presentable, and there exist \(\text {L} \)-small sets IJ of morphisms with the following properties: the source and target of the morphisms in I and J are \(\lambda \)-presentable and cofibrant, the fibrations (resp. trivial fibrations) in \({{\mathscr {M}}}\) are precisely those morphisms satisfying the right lifting property with respect to J (resp. I). In particular, \({{\mathscr {M}}}\) is combinatorial. For further details and background on the formalism of enriched Bousfield localisation, we refer the reader to [3].

Diagrams For \({\mathscr {J}}\) a large category and \({\mathscr {C}}\) a \(\text {L} \)-tractable or \(\text {L} \)-combinatorial model category (cf. [2, 3]), the projective and the injective model structures on \(\textrm{Fun}({\mathscr {J}}, {\mathscr {C}})\) exist and are again \(\text {L} \)-tractable (resp. \(\text {L} \)-combinatorial); we denote them by \(\textrm{Fun}({\mathscr {J}}, {\mathscr {C}})_{proj}\) and \(\textrm{Fun}({\mathscr {J}}, {\mathscr {C}})_{inj}\), respectively. If \({\mathscr {J}}\) is a Reedy category, we denote the Reedy model structure on \(\textrm{Fun}({\mathscr {J}}, {\mathscr {C}})\) by \(\textrm{Fun}({\mathscr {J}}, {\mathscr {C}})_{Reedy}\).

Left Bousfield localisation. We recall that if \({{\mathscr {M}}}\) is a large simplicial model category with a chosen collection of morphisms A, then

  • \(z \in {{\mathscr {M}}}\) is A-local if it is fibrant in \({{\mathscr {M}}}\) and for every morphism \(f :x \rightarrow y\) in A the induced morphism

    $$\begin{aligned} \underline{{{\mathscr {M}}}}(Qy, z) \xrightarrow {(Qf)^*} \underline{{{\mathscr {M}}}}(Qx,z) \end{aligned}$$

    is a weak homotopy equivalence in , and

  • \(f \in {{\mathscr {M}}}(x,y)\) is an A-local weak equivalence if for every A-local object \(z \in {{\mathscr {M}}}\) the morphism

    $$\begin{aligned} \underline{{{\mathscr {M}}}}(Qy, z) \xrightarrow {(Qf)^*} \underline{{{\mathscr {M}}}}(Qx,z) \end{aligned}$$

    is a weak equivalence in .

In that case, the left Bousfield localisation \(L_A {{\mathscr {M}}}\), provided it exists, is a large simplicial model category with a simplicial left Quillen functor \({{\mathscr {M}}}\rightarrow L_A {{\mathscr {M}}}\) which is universal among simplicial left Quillen functors out of \({{\mathscr {M}}}\) that send A-local weak equivalences to weak equivalences (cf. for instance [3, Def. 4.42, Def. 4.45]).

2 Grothendieck Sites and Local Epimorphisms

We recall the definition of a Grothendieck pretopology and a site. Throughout this article, we assume that \({\mathscr {C}}\) is a small category.

Definition 2.1

We write \({{\widehat{{\mathscr {C}}}}}{:}{=}\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {S}}\text {et} }_\text {S} )\) for the large category of small presheaves on \({\mathscr {C}}\).

Let \({{\mathscr {Y}}}:{\mathscr {C}}\rightarrow {{\widehat{{\mathscr {C}}}}}\), \(c \mapsto {{\mathscr {Y}}}_c\), denote the Yoneda embedding of \({\mathscr {C}}\).

Definition 2.2

([26, Def. 2.1.9]) Let \({\mathscr {C}}\) be a small category.

  1. (1)

    A coverage, or Grothendieck pretopology, on \({\mathscr {C}}\) is given by assigning to every object \(c \in {\mathscr {C}}\) a small set \(\tau (c)\) of families of morphisms \(\{f_i :c_i \rightarrow c\}_{i \in I}\) (with \(I \in {{\mathscr {S}}\text {et} }_\text {S} \)) satisfying the following properties: \(\{1_c\} \in \tau (c)\) for each \(c \in {\mathscr {C}}\), and for every family \(\{f_i :c_i \rightarrow c\}_{i \in I} \in \tau (c)\) and any morphism \(g :c' \rightarrow c\) in \({\mathscr {C}}\) there exists a family \(\{f'_j :c'_j \rightarrow c'\}_{j \in J} \in \tau (c')\) such that for every \(j \in J\) we find some \(i \in I\) and a commutative diagram

    figure d
  2. (2)

    The families in \(\{f_i :c_i \rightarrow c\}_{i \in I} \in \tau (c)\) are called covering families for c.

  3. (3)

    A (Grothendieck) site is a category \({\mathscr {C}}\) equipped with a coverage \(\tau \).

Later we will use the following technical condition:

Definition 2.3

We call a site \(({\mathscr {C}},\tau )\) closed if it satisfies the following condition: let \(\{c_i \rightarrow c\}_{i \in I}\) be any covering family in \(({\mathscr {C}}, \tau )\). Further, for each \(i \in I\), let \(\{c_{i,j} \rightarrow c_i\}_{j \in J_i}\) be a covering family in \(({\mathscr {C}},\tau )\). Then, there exists a covering family \(\{d_k \rightarrow c\}_{k \in K}\) such that every morphism \(d_k \rightarrow c\) factors through one of the composites \(c_{i,j} \rightarrow c_i \rightarrow c\).

Example 2.4

Let \({{\mathscr {C}}\text {art} }\) be the category of cartesian spaces, i.e. of sub-manifolds of \({\mathbb {R}}^\infty \) that are diffeomorphic to some \({\mathbb {R}}^n\), with smooth maps between these manifolds as morphisms. This category is small and has finite products. A coverage on \({{\mathscr {C}}\text {art} }\) is defined by calling a family \(\{\iota _i :c_i \rightarrow c\}_{i \in I}\) a covering family if it satisfies

  1. (1)

    all \(\iota _i\) are open embeddings (in particular \(\dim (c_i) = \dim (c)\) for all \(i \in I\)),

  2. (2)

    the \(\iota _i\) cover c, i.e. \(c = \bigcup _{i \in I} \iota _i(c_i)\), and

  3. (3)

    every finite intersection \(\iota _{i_0}(c_{i_0}) \cap \cdots \cap \iota _{i_m}(c_{i_m})\), with \(i_0, \ldots , i_m \in I\), is a cartesian space.

Coverings \(\{\iota _i :c_i \rightarrow c\}_{i \in I}\) with these properties are called differentiably good open coverings of c. They induce a coverage \(\tau _{dgop}\) on \({{\mathscr {C}}\text {art} }\), which turns \(({{\mathscr {C}}\text {art} },\tau _{dgop})\) into a site. This site is even closed, because every open cover of \(c \in {{\mathscr {C}}\text {art} }\) admits a differentiably good refinement [19, Cor. A.1]. There is a canonical embedding \({{\mathscr {C}}\text {art} }\hookrightarrow {{\mathscr {M}}{\text {fd}}}\) of \({{\mathscr {C}}\text {art} }\) into the category of all smooth manifolds. Thereby, any manifold M defines a presheaf \(\underline{M}\) on \({{\mathscr {C}}\text {art} }\), defined by \(\underline{M}(c) = {{\mathscr {M}}{\text {fd}}}(c,M)\). \(\triangleleft \)

Example 2.5

Let \({{\mathscr {O}}\text {p} }\) be the small category whose objects are open subsets of \({\mathbb {R}}^n\) for any (varying) \(n \in {\mathbb {N}}_0\) and whose morphisms are all smooth maps between such open subsets. The category \({{\mathscr {O}}\text {p} }\) has finite products and carries a coverage whose covering families are the open coverings. The resulting site is denoted \(({{\mathscr {O}}\text {p} },\tau _{op})\); observe that it is closed. \(\triangleleft \)

Definition 2.6

Let \(({\mathscr {C}},\tau )\) be a site. A morphism \(\pi :Y \rightarrow X\) in \({{\widehat{{\mathscr {C}}}}}\) is a \(\tau \)-local epimorphism if for every \(c \in {\mathscr {C}}\) and each morphism \(\varphi :{{\mathscr {Y}}}_c \rightarrow X\) there exists a covering \(\{f_j :c_j \rightarrow c\}_{j \in J} \in \tau (c)\) and morphisms \(\{ \varphi _j :{{\mathscr {Y}}}_{c_j} \rightarrow Y \}_{j \in J}\) with \(\pi \circ \varphi _j = \varphi \circ f_j\) for all \(j \in J\).

One directly checks the following:

Lemma 2.7

For any site \(({\mathscr {C}},\tau )\), the class of \(\tau \)-local epimorphisms is stable under pullback. In particular, the collection of \(\tau \)-local epimorphisms defines a coverage \({{\widehat{\tau }}}\) on \({{\widehat{{\mathscr {C}}}}}\).

Note that in the Grothendieck site \(({{\widehat{{\mathscr {C}}}}}, {{\widehat{\tau }}})\) every covering family consists of a single morphism. We list some general properties of \(\tau \)-local epimorphisms:

Lemma 2.8

Let \(({\mathscr {C}},\tau )\) be a site.

  1. (1)

    Consider morphisms \(p \in {{\widehat{{\mathscr {C}}}}}(Y,X)\), \(q \in {{\widehat{{\mathscr {C}}}}}(Z,Y)\). If \(p \circ q\) is a \(\tau \)-local epimorphism, then so is p.

  2. (2)

    For any covering family \(\{c_i \rightarrow c\}_{i \in I}\), the induced morphism \(\coprod _{i \in I} {{\mathscr {Y}}}_{c_i} \rightarrow {{\mathscr {Y}}}_c\) is a \(\tau \)-local epimorphism.

  3. (3)

    \(\tau \)-local epimorphisms are stable under colimits: let \({\mathscr {J}}\) be a small category, let \(D', D :{\mathscr {J}}\rightarrow {{\widehat{{\mathscr {C}}}}}\) be diagrams in \({{\widehat{{\mathscr {C}}}}}\), and let \(\pi :D' \rightarrow D\) be a morphism of diagrams such that each component \(\pi _j :D'_j \rightarrow D_j\) is a \(\tau \)-local epimorphism, for any \(j \in {\mathscr {J}}\). Then, the induced morphism \({\text {colim}}\, \pi :{\text {colim}}\, D' \rightarrow {\text {colim}}\, D\) is a \(\tau \)-local epimorphism.

  4. (4)

    If \(X(c) \ne \emptyset \) for each \(c \in {\mathscr {C}}\), then the projection \(X \times Z \rightarrow Z\) is a \(\tau \)-local epimorphism, for any \(Z \in {{\widehat{{\mathscr {C}}}}}\).

  5. (5)

    The site \(({\mathscr {C}},\tau )\) is closed if and only if \(\tau \)-local epimorphisms are stable under composition, if and only if \(({{\widehat{{\mathscr {C}}}}}, {{\widehat{\tau }}})\) is closed.

Proof

Claims (1) appears as property LE2 in [31, p. 390]. Claim (2) follows straightforwardly from the definition of a \(\tau \)-local epimorphism. Claim (3) can be found as [31, Prop. 16.1.12].

Claim (4) is immediate since by assumption the projection \(X(c) \times Y(c) \rightarrow Y(c)\) is surjective for each \(c \in {\mathscr {C}}\); we can hence use the identity covering of c to obtain the desired lift.

For (5), it readily follows from the closedness of \(({\mathscr {C}},\tau )\) that \(\tau \)-local epimorphisms are stable under composition, and thus also that \(({{\widehat{{\mathscr {C}}}}},{{\widehat{\tau }}})\) is closed. On the other hand, assume that \(\tau \)-local epimorphisms are stable under composition. Consider an object \(c \in {\mathscr {C}}\), a covering family \(\{f_i :c_i \rightarrow c\}_{i \in I}\), and for each \(i \in I\) a covering family \(\{c_{i,k} \rightarrow c_i\}_{k \in K_i}\). By claims (2) and (3), we thus obtain \(\tau \)-local epimorphisms

$$\begin{aligned} \coprod _{i \in I} \coprod _{k \in K_i} {{\mathscr {Y}}}_{c_{i,k}} \longrightarrow \coprod _{i \in I} {{\mathscr {Y}}}_{c_i} \longrightarrow {{\mathscr {Y}}}_c\,. \end{aligned}$$

By assumption, their composition is a \(\tau \)-local epimorphism again, and hence it follows (using that \({{\widehat{{\mathscr {C}}}}}({{\mathscr {Y}}}_c,-)\) preserves colimits) that \(({\mathscr {C}},\tau )\) is closed. Finally, assume that \(({{\widehat{{\mathscr {C}}}}},{{\widehat{\tau }}})\) is closed. Let \(f :Z \rightarrow Y\) and \(g :Y \rightarrow X\) be \(\tau \)-local epimorphisms, and let \(p {:}{=}g \circ f\). By assumption, there exists a \(\tau \)-local epimorphism \(Z' \rightarrow X\) which factors through g; that is, there exists a morphism \(q :Z' \rightarrow Z\) such that \(p \circ q\) is a \(\tau \)-local epimorphism. But then p is a \(\tau \)-local epimorphism by claim (1). \(\square \)

Let \(({\mathscr {C}},\tau )\) be a site, and consider a covering \({{\mathcal {U}}}= \{c_i \rightarrow c\}_{i \in I}\). We can form its Čech nerve, which is the simplicial object in \({{\widehat{{\mathscr {C}}}}}\) whose level-n object reads as

$$\begin{aligned} {\check{C}}{{\mathcal {U}}}_n {:}{=}\coprod _{i_0, \ldots , i_n \in I} C_{i_0 \ldots i_n}\,, \quad \text {with} \quad C_{i_0 \ldots i_n} {:}{=}{{\mathscr {Y}}}_{c_{i_0}} \underset{{{\mathscr {Y}}}_c}{\times }\ \cdots \underset{{{\mathscr {Y}}}_c}{\times }\ {{\mathscr {Y}}}_{c_{i_n}} \quad \in {{\widehat{{\mathscr {C}}}}}\,. \end{aligned}$$
(2.9)

Note that, in general, \(C_{i_0 \ldots i_n} \in {{\widehat{{\mathscr {C}}}}}\) is not representable as soon as \(n \ne 0\). The simplicial structure morphisms are given by projecting out or doubling the i-th factor, respectively. Depending on the context we will view the Čech nerve \({\check{C}}{{\mathcal {U}}}\) either as a simplicial object \({\check{C}}{{\mathcal {U}}}_\bullet \) in \({{\widehat{{\mathscr {C}}}}}\) or as an augmented simplicial object \({\check{C}}{{\mathcal {U}}}_\bullet \rightarrow {{\mathscr {Y}}}_c\) in \({{\widehat{{\mathscr {C}}}}}\). For convenience, we recall the following definitions:

Definition 2.10

Let \(({\mathscr {C}},\tau )\) be a site, and let \(X \in {{\widehat{{\mathscr {C}}}}}\) be a presheaf on \({\mathscr {C}}\). Then, X is called a sheaf on \(({\mathscr {C}},\tau )\) if for every object \(c \in {\mathscr {C}}\) and for every covering family \(\{f_i :c_i \rightarrow c\}_{i \in I} \in \tau (c)\) the diagram

figure e

is an equaliser diagram in \({{\mathscr {S}}\text {et} }_\text {L} \). The category \({{\mathscr {S}}\text {h} }({\mathscr {C}},\tau )\) of sheaves on \(({\mathscr {C}},\tau )\) is the full subcategory of \({{\widehat{{\mathscr {C}}}}}\) on the sheaves.

3 Homotopy Sheaves and Descent Along Local Epimorphisms

Let \(({\mathscr {C}},\tau )\) be a small site, and let \({{\mathscr {Y}}}:{\mathscr {C}}\rightarrow {{\widehat{{\mathscr {C}}}}}\) denote its Yoneda embedding. In this section we show that the homotopy right Kan extension along \({{\mathscr {Y}}}_*\) maps large higher \(\tau \)-sheaves on \({\mathscr {C}}\) to large higher \({{\widehat{\tau }}}\)-sheaves on \({{\widehat{{\mathscr {C}}}}}\). Here, \({{\widehat{\tau }}}\) is the pretopology of \(\tau \)-local epimorphisms on \({{\widehat{{\mathscr {C}}}}}\) (see Definition 2.6 and Lemma 2.7).

3.1 Sheaves of \(\infty \)-Groupoids

Let \({\mathscr {C}}\) be a small category. Let denote the extra-large category of large simplicial presheaves on \({\mathscr {C}}\). It is enriched, tensored and cotensored over . Composing the Yoneda embedding with the functor , we obtain a fully faithful functor \({{\mathscr {Y}}}:{\mathscr {C}}\rightarrow {\mathscr {H}}_{\infty ,0}\). From now on, we view the category \({\mathscr {H}}_{\infty ,0}\) as endowed with the projective model structure.

Definition 3.1

A presheaf of \(\infty \)-groupoids on \({\mathscr {C}}\) is a fibrant object in \({\mathscr {H}}_{\infty ,0}\).

Proposition 3.2

\({\mathscr {H}}_{\infty ,0}\) is a left proper, tractable, model -category. If \({\mathscr {C}}\) has finite products, then \({\mathscr {H}}_{\infty ,0}\) is additionally a symmetric monoidal model category.

Proof

The existence and tractability of the projective model structure follow from [3, Thm. 2.14] (since \({\mathscr {C}}\) is small). \({\mathscr {H}}_{\infty ,0}\) is left proper since its pushouts and weak equivalences are defined objectwise, its cofibrations are (in particular) objectwise cofibrations, and is left proper. Its simplicial enrichment is a consequence of [3, Prop. 4.50]. If \({\mathscr {C}}\) has finite products, then \({\mathscr {H}}_{\infty ,0}\) is symmetric monoidal by [3, Prop. 4.52], using that the Yoneda embedding preserves limits and that \({\mathscr {H}}_{\infty ,0}\) is a simplicial model category. \(\square \)

Note that [3, Prop. 4.52] gives a more general criterion for when a projective model structure inherits a symmetric monoidal structure; however, for us the main case of interest will be where \({\mathscr {C}}\) admits finite products.

Projective model categories of simplicial presheaves have an explicit cofibrant replacement functor Q [18]. In order to write it down, it is convenient to introduce the two-sided bar construction; here, we follow [44, Sec. 4.2]:

Definition 3.3

Let \({{\mathscr {M}}}\) be a large cocomplete category tensored over . Consider a large category \({\mathscr {I}}\) and functors \(H :{\mathscr {I}}\rightarrow {{\mathscr {M}}}\) and .

  1. (1)

    The two-sided simplicial bar construction of (GH) is the simplicial object

  2. (2)

    The two-sided bar construction of (FG) is the realisation of \(B_\bullet (G,{\mathscr {I}},H)\), i.e.

For later reference, we record the following direct consequences of this definition:

Lemma 3.4

In the setting of Definition 3.3, the following statements hold true:

  1. (1)

    Bar constructions commute: let \({\mathscr {J}}\) be a second large category, and let and \(D :{\mathscr {I}}\times {\mathscr {J}}\rightarrow {{\mathscr {M}}}\) be diagrams. Then, there is a canonical natural isomorphism

    $$\begin{aligned} B \big ( G', {\mathscr {J}}, B(G, {\mathscr {I}}, D) \big ) \cong B \big ( G, {\mathscr {I}}, B(G', {\mathscr {J}}, D) \big )\,. \end{aligned}$$
  2. (2)

    Bar constructions are pointwise: if is the category of large simplicial presheaves on a large category \({\mathscr {C}}\) and \(c \in {\mathscr {C}}\) is any object, then evaluation at c commutes with bar constructions: there is a natural isomorphism

    $$\begin{aligned} B(G,{\mathscr {I}},H)(c) \cong B \big ( G, {\mathscr {I}}, H(c) \big )\,. \end{aligned}$$
  3. (3)

    In the above setting, let additionally . Then there is a natural morphism

    where \({\text {diag}}\) denotes the functor taking the diagonal of a bisimplicial set. This morphism is a weak equivalence of simplicial sets.

  4. (4)

    If \({{\mathscr {M}}}\) is a large simplicial model category with cofibrant replacement functor \(Q^{{\mathscr {M}}}\), the bar construction is a model for the homotopy colimit, i.e.

    $$\begin{aligned} \underset{j \in {\mathscr {J}}}{{\text {hocolim}}}^{{\mathscr {M}}}(Dj) \simeq B(*, {\mathscr {J}}, Q^{{\mathscr {M}}}\circ D)\,. \end{aligned}$$

    In particular, if Dj is already cofibrant in \({{\mathscr {M}}}\) for each \(j \in {\mathscr {J}}\), then we may compute the homotopy colimit of D as

    $$\begin{aligned} \underset{j \in {\mathscr {J}}}{{\text {hocolim}}}^{{\mathscr {M}}}(Dj) \simeq B(*, {\mathscr {J}}, D)\,. \end{aligned}$$

Proof

Claim (1) follows readily from the fact that colimits commute. Claim (2) is a direct consequence of the fact that colimits in presheaf categories are computed pointwise. Finally, Claim (3) is an application of the Bousfield-Kan map (see [22, Sec. 18.7] for details), and Claim (4) is [44, Cor. 5.1.3]. \(\square \)

Remark 3.5

Analogous statements hold true in the Grothendieck universes \(\text {S} \) and \(\text {XL} \) of small and extra-large sets, respectively. \(\triangleleft \)

We now define the cofibrant replacement functor in \({\mathscr {H}}_{\infty ,0}\) which we will be using throughout this article; it was introduced in [18].

Definition 3.6

We let \(Q :{\mathscr {H}}_{\infty ,0} \rightarrow {\mathscr {H}}_{\infty ,0}\) denote the functor whose action on \(F \in {\mathscr {H}}_{\infty ,0}\) is given by

$$\begin{aligned} (QF)_n = B(F,{\mathscr {C}},{{\mathscr {Y}}})_n = \coprod _{c_0, \ldots , c_n \in {\mathscr {C}}} {{\mathscr {Y}}}_{c_0} \times {\mathscr {C}}(c_0, c_1) \times \cdots \times {\mathscr {C}}(c_{n-1}, c_n) \times F_n(c_n)\,. \end{aligned}$$
(3.7)

Observe that for X a simplicially constant simplicial presheaf on \({\mathscr {C}}\), we have a canonical isomorphism

$$\begin{aligned} QX = B(*, {\mathscr {C}}_{/X}, {{\mathscr {Y}}})\,. \end{aligned}$$

Definition 3.8

Let \(({\mathscr {C}},\tau )\) be a small Grothendieck site, and let \({{\check{\tau }}}\) denote the class of morphisms in \({\mathscr {H}}_{\infty ,0}\) consisting of Čech nerves of coverings in \(({\mathscr {C}},\tau )\) (see (2.9)). Since \({\mathscr {C}}\) is small, this is a small set. We let \({\mathscr {H}}_{\infty ,0}^{loc}\) denote the left Bousfield localisation \({\mathscr {H}}_{\infty ,0}^{loc} {:}{=}L_{{\check{\tau }}}{\mathscr {H}}_{\infty ,0}\) of the large category \({\mathscr {H}}_{\infty ,0}\). A sheaf of \(\infty \)-groupoids on \(({\mathscr {C}},\tau )\) is a fibrant object in \({\mathscr {H}}_{\infty ,0}^{loc}\).

The model structure on \({\mathscr {H}}_{\infty ,0}^{loc}\) is also called the local projective model structure on . Explicitly, an object F in is fibrant if

  1. (1)

    F(c) is a Kan complex for every \(c \in {\mathscr {C}}\), and

  2. (2)

    for every covering family \({{\mathcal {U}}}= \{c_i \rightarrow c\}_{i \in I}\) in \(({\mathscr {C}},\tau )\), the morphism

    is a weak equivalence in (compare (2.9)). Here we have used that Q is a left adjoint.

Proposition 3.9

\({\mathscr {H}}_{\infty ,0}^{loc}\) is a proper, tractable, -model category. If \({\mathscr {H}}_{\infty ,0}\) is symmetric monoidal as a model category, then so is \({\mathscr {H}}_{\infty ,0}^{loc}\).

Proof

The first claim follows form [3, Thm. 2.14, Thm 4.56] and the fact that is a large, proper, tractable model category. The second claim is an application of [3, Thm. 4.58]. \(\square \)

Remark 3.10

If \({\mathscr {C}}\) has finite products, then the conditions of [3, Prop. 4.52] are satisfied, so that in this case \({\mathscr {H}}_{\infty ,0}\), and hence also \({\mathscr {H}}_{\infty ,0}^{loc}\), are symmetric monoidal model categories. \(\triangleleft \)

The -enriched hom in both \({\mathscr {H}}_{\infty ,0}\) and \({\mathscr {H}}_{\infty ,0}^{loc}\) is given by

where the argument of the end is given by the internal hom in . In particular, there is a natural isomorphism \(\underline{{\mathscr {H}}}_{\infty ,0}({{\mathscr {Y}}}_c, F) \cong F(c)\) for any \(c \in {\mathscr {C}}\) and any \(F \in {\mathscr {H}}_{\infty ,0}\).

Since \({\mathscr {C}}\) is a small category (hence also large, since \(\text {S} \in \text {L} \)) and \({{\mathscr {S}}\text {et} }_\text {S} \) is a large category, \({{\widehat{{\mathscr {C}}}}}= \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {S}}\text {et} }_\text {S} )\) is a large category. We define the extra-large category

Since is \(\text {L} \)-tractable, the projective model structure on \({{\widehat{{\mathscr {H}}}}}_{\infty ,0}\) exists and is itself \(\text {L} \)-tractable [3, Thm. 2.14]. We will always view \({{\widehat{{\mathscr {H}}}}}_{\infty ,0}\) as endowed with this projective model structure.

By Proposition 2.7, if \(({\mathscr {C}},\tau )\) is a small site, then \(({{\widehat{{\mathscr {C}}}}}, {{\widehat{\tau }}})\) is a large site. Let \(({{\widehat{\tau }}}){}\check{}\) denote the large set of Čech nerves of coverings in \(({{\widehat{{\mathscr {C}}}}},{{\widehat{\tau }}})\) (this is a large set since \({{\widehat{{\mathscr {C}}}}}\) is large). Since \({{\widehat{{\mathscr {H}}}}}_{\infty ,0}\) is left proper and \(\text {L} \)-tractable, and \(({{\widehat{\tau }}}){}\check{}\) is \(\text {L} \)-small, we can form the left Bousfield localisation

$$\begin{aligned} {{\widehat{{\mathscr {H}}}}}_{\infty ,0}^{loc} {:}{=}L_{({{\widehat{\tau }}}){}\check{}} {{\widehat{{\mathscr {H}}}}}_{\infty ,0}\,. \end{aligned}$$

Both \({{\widehat{{\mathscr {H}}}}}_{\infty ,0}\) and \({{\widehat{{\mathscr {H}}}}}_{\infty ,0}^{loc}\) are symmetric monoidal since \({{\widehat{{\mathscr {C}}}}}\) has finite products (by Propositions 3.2 and 3.9, respectively).

The Yoneda embedding \({{\mathscr {Y}}}:{\mathscr {C}}\rightarrow {{\widehat{{\mathscr {C}}}}}\) induces a pullback functor \({{\mathscr {Y}}}^* :{{\widehat{{\mathscr {H}}}}}_{\infty ,0} \rightarrow {\mathscr {H}}_{\infty ,0}\). Since both \({\mathscr {C}}\) and \({{\widehat{{\mathscr {C}}}}}\) are large, and since both \({{\widehat{{\mathscr {H}}}}}_{\infty ,0}\) and \({\mathscr {H}}_{\infty ,0}\) have all large limits and colimits, there is a triple of adjunctions

figure f

We can compute the right adjoint as

$$\begin{aligned} ({{\mathscr {Y}}}_* F)(X) = \int _{c \in {\mathscr {C}}^{\text {op}}} F(c)^{({{\widehat{{\mathscr {C}}}}})^{\text {op}}(X,{{\mathscr {Y}}}_c)} \cong \int _{c \in {\mathscr {C}}^{\text {op}}} F(c)^{X(c)} \cong \underline{{\mathscr {H}}}_{\infty ,0}(X,F)\,. \end{aligned}$$

(Note that here we view \(X \in {{\widehat{{\mathscr {C}}}}}\) as an object in \({\mathscr {H}}_{\infty ,0}\) via the canonical inclusions .)

We now ask whether one of the functors \({{\mathscr {Y}}}_!\) or \({{\mathscr {Y}}}_*\) maps sheaves of \(\infty \)-groupoids on \(({\mathscr {C}},\tau )\) to sheaves of \(\infty \)-groupoids on \(({{\widehat{{\mathscr {C}}}}}, {{\widehat{\tau }}})\), i.e. preserves fibrant objects as a functor \({\mathscr {H}}_{\infty ,0}^{loc} \rightarrow {{\widehat{{\mathscr {H}}}}}_{\infty ,0}^{loc}\). Since this indicates that we are looking for a right Quillen functor, we focus on the right adjoint \({{\mathscr {Y}}}_*\). In general, \({{\mathscr {Y}}}_*\) will not even preserve fibrant objects as a functor \({\mathscr {H}}_{\infty ,0} \rightarrow {{\widehat{{\mathscr {H}}}}}_{\infty ,0}\). However, if \({\tilde{Q}} :{\mathscr {H}}_{\infty ,0} \rightarrow {\mathscr {H}}_{\infty ,0}\) is a functorial cofibrant replacement, then the functor

$$\begin{aligned} S^{{\tilde{Q}}}_{\infty ,0} {:}{=}{\tilde{Q}}^* \circ {{\mathscr {Y}}}_* :{\mathscr {H}}_{\infty ,0} \longrightarrow {{\widehat{{\mathscr {H}}}}}_{\infty ,0}\,, \qquad \big ( S^{{\tilde{Q}}}_{\infty ,0}(F) \big )(X) {:}{=}\underline{{\mathscr {H}}}({\tilde{Q}}X,F) \end{aligned}$$

preserves fibrations and trivial fibrations: \({\mathscr {H}}_{\infty ,0}\) is a model -category, and so the simplicially enriched hom

is a right Quillen functor whenever \(A \in \underline{{\mathscr {H}}}_{\infty ,0}\) is cofibrant.

Remark 3.11

The above holds true for any cofibrant replacement functor \({\tilde{Q}}\) in \({\mathscr {H}}_{\infty ,0}\). However, for technical reasons we always use Dugger’s functor Q from (3.7) for cofibrant replacement in the projective model structure of simplicial presheaves in the remainder of this article. We abbreviate \(S_{\infty ,0} {:}{=}S^Q_{\infty ,0}\). \(\triangleleft \)

Let \(F \in {\mathscr {H}}_{\infty ,0}\) and \(X \in {{\widehat{{\mathscr {C}}}}}\). Using that QX is a bar construction we compute

where and \(B^{{\mathscr {H}}_{\infty ,0}}\) are the cobar and bar constructions in and in \({\mathscr {H}}_{\infty ,0}\), respectively. If F is projectively fibrant, then the last expression is a model for the homotopy limit (see also recalled as Proposition 3.31 below),

Now, using the Yoneda Lemma and [44, Ex. 9.2.11] we conclude:

Proposition 3.12

If \(F \in {\mathscr {H}}_{\infty ,0}\) is fibrant, \(S_{\infty ,0} F\) is the homotopy right Kan extension of F along the (opposite of the) Yoneda embedding of \({\mathscr {C}}\), i.e.

$$\begin{aligned} S_{\infty ,0} F \simeq {\textrm{hoRan}}_{{\mathscr {Y}}}(F)\,. \end{aligned}$$

Remark 3.13

In view of Sect. 5, we emphasize that \(S_{\infty ,0}\) is a presentation of the \(\infty \)-categorical right Kan extension of presheaves of spaces on \({\mathscr {C}}\) to presheaves of spaces on \({{\widehat{{\mathscr {C}}}}}\). \(\triangleleft \)

In order to show that \(S_{\infty ,0}\) is a right Quillen functor, it remains to show that it is a right adjoint. Since \({{\widehat{{\mathscr {C}}}}}\) is a large category, \({{\widehat{{\mathscr {C}}}}}\)-indexed (co)ends exist in , the extra-large category of large simplicial sets. Therefore, a left adjoint to \(S_{\infty ,0}\) is given by

$$\begin{aligned} {\widetilde{Re}}_{\infty ,0} :{{\widehat{{\mathscr {H}}}}}\rightarrow {\mathscr {H}}\,, \qquad {\widetilde{Re}}_{\infty ,0}({\widehat{G}}) {:}{=}\int ^{X \in {{\widehat{{\mathscr {C}}}}}^{\text {op}}} {\widehat{G}}(X) \otimes QX\,. \end{aligned}$$

We also set

$$\begin{aligned} Re_{\infty ,0}({\widehat{G}}) {:}{=}Q \circ {{\mathscr {Y}}}^*({\widehat{G}}) \end{aligned}$$

Lemma 3.14

There is a canonical natural isomorphism

$$\begin{aligned} {\widetilde{Re}}_{\infty ,0} \cong Re_{\infty ,0}\,. \end{aligned}$$

Thus, \(Re_{\infty ,0}\) is also a left adjoint to \(S_{\infty ,0}\).

Proof

Using (3.7), we have the following natural isomorphisms

$$\begin{aligned} \big ( {\widetilde{Re}}_{\infty ,0}({\widehat{G}}) \big )_n&= \int ^{X \in {{\widehat{{\mathscr {C}}}}}^{\text {op}}} \coprod _{c_0, \ldots , c_n \in {\mathscr {C}}} {{\mathscr {Y}}}_{c_0} \times {\mathscr {C}}(c_0, c_1) \times \cdots \times {\mathscr {C}}(c_{n-1}, c_n) \times {{\widehat{{\mathscr {C}}}}}({{\mathscr {Y}}}_{c_n},X) \times {\widehat{G}}_n(X) \\&\cong \coprod _{c_0, \ldots , c_n \in {\mathscr {C}}} {{\mathscr {Y}}}_{c_0} \times {\mathscr {C}}(c_0, c_1) \times \cdots \times {\mathscr {C}}(c_{n-1}, c_n) \times {\widehat{G}}_n({{\mathscr {Y}}}_{c_n}) \\&= \big ( (Q \circ {{\mathscr {Y}}}^*)({\widehat{G}}) \big )_n\,, \end{aligned}$$

where the isomorphism is an application of the Yoneda Lemma for \({{\widehat{{\mathscr {C}}}}}\). \(\square \)

Proposition 3.15

The functors \(Re_{\infty ,0}\) and \(S_{\infty ,0}\) give rise to a simplicial Quillen adjunction which sits inside the following non-commutative diagram

figure g

Further, there exists a natural isomorphism \(\eta :Re_{\infty ,0} \circ {\widehat{{{\mathscr {Y}}}}} \overset{\cong }{\longrightarrow }Q\).

Proof

We have already shown above that \(Re_{\infty ,0} \dashv S_{\infty ,0}\) and that \(S_{\infty ,0}\) preserves fibration and trivial fibrations. The rest is then a direct application of [18, Prop. 2.3]. In the present case, \(\eta \) as written down there turns out to be an isomorphism as a consequence of the Yoneda Lemma: recall that \(Re_{\infty ,0} = Q \circ {{\mathscr {Y}}}^*\) and that there is a canonical isomorphism

$$\begin{aligned} {{\mathscr {Y}}}^*({{\widehat{{{\mathscr {Y}}}}}}_X) = {{\widehat{{\mathscr {C}}}}}({{\mathscr {Y}}}_{(-)}, X) \cong X\,, \end{aligned}$$

natural in \(X \in {{\widehat{{\mathscr {C}}}}}\). \(\square \)

Proposition 3.16

The functors \(Re_{\infty ,0}\) and \(S_{\infty ,0}\) have the following properties:

  1. (1)

    The functor \(Re_{\infty ,0} :{{\widehat{{\mathscr {H}}}}}_{\infty ,0} \rightarrow {\mathscr {H}}_\infty \) is homotopical.

  2. (2)

    Let \({{\widehat{Q}}}:{{\widehat{{\mathscr {H}}}}}_{\infty ,0} \rightarrow {{\widehat{{\mathscr {H}}}}}_{\infty ,0}\) denote the cofibrant replacement functor on \({{\widehat{{\mathscr {H}}}}}_{\infty ,0}\) defined in analogy with (3.7). For each fibrant object \(F \in {\mathscr {H}}_{\infty ,0}\), there is a zig-zag of weak equivalences

    $$\begin{aligned} F \overset{\sim }{\longleftarrow }\ Q F \overset{\sim }{\longrightarrow }Re_{\infty ,0} \circ S_{\infty ,0}(F) \overset{\sim }{\longleftarrow }\ Re_{\infty ,0} \circ {{\widehat{Q}}}\circ S_{\infty ,0} (F) \,, \end{aligned}$$

    which is natural in F.

Proof

By Lemma 3.14, \(Re_{\infty ,0} \cong Q \circ {{\mathscr {Y}}}^*\). The functor \({{\mathscr {Y}}}^*\) is homotopical since weak equivalences in both \({\mathscr {H}}_{\infty ,0}\) and in \({{\widehat{{\mathscr {H}}}}}_{\infty ,0}\) are defined objectwise. Thus, claim (1) follows since Q is homotopical. As a consequence, the morphism

$$\begin{aligned} Re_{\infty ,0} \circ {{\widehat{Q}}}\circ S_{\infty ,0} (F) \longrightarrow Re_{\infty ,0} \circ S_{\infty ,0}(F) \end{aligned}$$

is a weak equivalence. For \(F \in {\mathscr {H}}_{\infty ,0}\), we find that

$$\begin{aligned} \big ( {{\mathscr {Y}}}^* \circ S_{\infty ,0}(F) \big ) (c) = S_{\infty ,0}(F)({{\mathscr {Y}}}_c) = \underline{{\mathscr {H}}}_{\infty ,0}(Q{{\mathscr {Y}}}_c, F)\,, \end{aligned}$$

and since F is fibrant and \({{\mathscr {Y}}}_c\) is cofibrant, the functor \(\underline{{\mathscr {H}}}_{\infty ,0}(-,F)\) preserves the weak equivalence \(Q{{\mathscr {Y}}}_c \rightarrow {{\mathscr {Y}}}_c\). Hence, there is a natural weak equivalence \(F \overset{\sim }{\longrightarrow }{{\mathscr {Y}}}^* \circ S_{\infty ,0}(F)\). To complete the proof, we apply the homotopical functor Q to this morphism. \(\square \)

We now investigate whether the Quillen adjunction \(Re_{\infty ,0} \dashv S_{\infty ,0}\) between the projective model structures descends to a Quillen adjunction between the local projective model structures. We recall the following standard result:

Proposition 3.17

[22, Prop. 3.1.6, Prop. 3.3.18] Let \({{\mathscr {M}}}\) and \({{\mathscr {N}}}\) be two large simplicial model categories, and let \(F \dashv G\) be a simplicial Quillen adjunction from \({{\mathscr {M}}}\) to \({{\mathscr {N}}}\). Suppose that AB are large sets of morphisms in \({{\mathscr {M}}}\) and in \({{\mathscr {N}}}\), respectively, such that the left Bousfield localisations \(L_A {{\mathscr {M}}}\) and \(L_B {{\mathscr {N}}}\) exist. If G maps B-local objects to A-local objects, then \(F \dashv G\) descends to a Quillen adjunction between the localised model categories.

Theorem 3.18

([17, Cor. A.3]) Let \({\mathscr {C}}\) be a small category endowed with a coverage \(\tau \), and let \(\pi :Y \rightarrow X\) be a \(\tau \)-local epimorphism in \({{\widehat{{\mathscr {C}}}}}\) with Čech nerve \({\check{C}}\pi \). The augmentation map \(\pi ^{[\bullet ]} :{\check{C}}\pi \rightarrow X\) is a weak equivalence in \({\mathscr {H}}_{\infty ,0}^{loc}\).

The proof of Theorem 3.18 is rather technical; it requires a “wrestling match with the small object argument” ([17] Subsection A.12). We refer the reader to that reference for details.

Remark 3.19

In [17], Theorem 3.18 is proven for the Čech localisation of the injective model structure \({\mathscr {H}}^i\) on \({\mathscr {H}}\). However, it also holds true in \({\mathscr {H}}_{\infty ,0}\) because the injective and projective model structures on \({\mathscr {H}}\) (and \({{\widehat{{\mathscr {H}}}}}\)) have the same weak equivalences, and so do the associated local model structures (see also [11, Prop. 2.6] for details). \(\triangleleft \)

Lemma 3.20

Let \(\pi :Y \rightarrow X\) be a \(\tau \)-local epimorphism in \({{\widehat{{\mathscr {C}}}}}\). Let denote the simplicial diagram which sends to the simplicially constant presheaf \({\check{C}}\pi _n\). There is a commutative triangle in \({\mathscr {H}}_{\infty ,0}\):

figure h

The top morphism is an objectwise weak equivalence, and the diagonal morphisms are weak equivalences in \({\mathscr {H}}_{\infty ,0}^{loc}\).

Proof

The top morphism is induced by the Bousfield-Kan, or last-vertex map. It is an objectwise weak equivalence by Lemma 3.4(3). The right-hand side is a local weak equivalence by Theorem 3.18. Thus, it remains to show that the triangle commutes. We check this explicitly: for each object \(c \in {\mathscr {C}}\), there are canonical natural isomorphisms

where \({\textsf {N}}\) denotes the nerve functor. For , recall the last-vertex map

of small simplicial sets (see, for instance, [14, Lemma 7.3.11, Par. 7.3.14]). We describe this explicitly in our case, i.e. for \(X = {\check{C}}\pi (c)\). An n-simplex in this simplicial set is explicitly given by a sequence of morphisms

figure i

of small simplicial sets. We denote these data by \((\alpha , \kappa )\). Consider the map \(\varphi _\alpha :[n] \rightarrow [k_n]\) in which acts as

$$\begin{aligned} i \longmapsto \alpha _{n-1,n} \circ \cdots \circ \alpha _{i,i+1}(k_i)\,. \end{aligned}$$

The top morphism in diagram (3.21) acts by sending the n-simplex \((\alpha , \kappa )\) to the n-simplex of \({\check{C}}\pi (c)\) which is given by the composition

figure j

where the first map is the one constructed above from the data \((\alpha , \kappa )\). Explicitly, for a given n-simplex , the resulting map

$$\begin{aligned} {\check{C}}\pi _{k_n} \cong \underbrace{Y \times _X \cdots \times _X Y}_{k_n\text { copies of }Y} \longrightarrow {\check{C}}\pi _n \cong \underbrace{Y \times _X \cdots \times _X Y}_{n\text { copies of }Y} \end{aligned}$$

is the projection onto the Y-factors in positions \(\varphi _\alpha (0), \ldots , \varphi _\alpha (n)\). This map commutes with the maps to X. \(\square \)

Proposition 3.22

The functor \(S_{\infty ,0} :{\mathscr {H}}_{\infty ,0}^{loc} \rightarrow {{\widehat{{\mathscr {H}}}}}_{\infty ,0}^{loc}\) preserves local objects.

Proof

Let \(\pi :Y \rightarrow X\) be a \(\tau \)-local epimorphism in \({{\widehat{{\mathscr {C}}}}}\), and let \(F \in {\mathscr {H}}_{\infty ,0}^{loc}\) be fibrant. We need to show that the canonical morphism

is a weak equivalence in . Substituting the definition of \(S_{\infty ,0}\), this is the same as the canonical map

Observe that there is a canonical isomorphism

figure k

in the homotopy category (in fact, this can even be modelled as an isomorphism in in light of Lemma 3.4(1) and Proposition 3.31 below). Since the functor Q is a bar construction (see (3.7)), it commutes with the homotopy colimit (which is also a bar construction) by Lemma 3.4(1). Then, the morphism

figure l

is the image of the left-hand diagonal morphism in the commutative triangle (3.21) under the functor \(\underline{{\mathscr {H}}}_{\infty ,0}(Q(-),F)\). Observe that the functor

is homotopical since F is fibrant in \({\mathscr {H}}_{\infty ,0}^{loc}\) and \({\mathscr {H}}_{\infty ,0}^{loc}\) is enriched over . Combining this with Lemma 3.20, we obtain that the above morphism is a weak equivalence in . \(\square \)

Combining Proposition 3.17 and Proposition 3.22, we obtain

Theorem 3.23

The Quillen adjunction \(Re_{\infty ,0}: {{\widehat{{\mathscr {H}}}}}_{\infty ,0} \rightleftarrows {\mathscr {H}}_{\infty ,0}:S_{\infty ,0}\) induces a Quillen adjunction

figure m

In Sect. 5 we provide a version of Theorem 3.23 for fully coherent diagrams of spaces in a quasi-categorical language; on the underlying quasi-categories, we will show that \(S_{\infty ,0}\) is fully faithful and we identify its essential image.

3.2 Homotopy Sheaves with Values in Simplicial Model Categories

We now extend the results from Sect. 3.1 to homotopy sheaves on a small site \(({\mathscr {C}}, \tau )\), valued not just in , but in any large, combinatorial or cellular simplicial model category. We begin by recalling some technology for enriched model categories.

Recall the notion of a \({{\mathscr {V}}}\) be a symmetric monoidal model category. A \({{\mathscr {V}}}\)-enriched model category, or model \({{\mathscr {V}}}\)-category \({{\mathscr {M}}}\) (see, for instance, [23, Def. 4.2.18]). We denote the three functors which are part of the enrichment by

$$\begin{aligned} (-) \otimes (-) :{{\mathscr {V}}}\times {{\mathscr {M}}}\rightarrow {{\mathscr {M}}}\,, \qquad \underline{{{\mathscr {M}}}}^{{\mathscr {V}}}(-,-) :{{\mathscr {M}}}^{\text {op}}\times {{\mathscr {M}}}\rightarrow {{\mathscr {V}}}\,, \qquad \{-,-\} :{{\mathscr {V}}}^{\text {op}}\times {{\mathscr {M}}}\rightarrow {{\mathscr {M}}}\,. \end{aligned}$$

We import the following definition (using [44, Thms. 7.6.3, 11.5.1]):

Definition 3.24

Let \({\mathscr {I}}\) be a small category.

  1. (1)

    The weighted \({\mathscr {I}}\)-colimit functor of \({{\mathscr {M}}}\) is the functor

    $$\begin{aligned} (-) \otimes _{\mathscr {I}}(-) :\textrm{Fun}({\mathscr {I}}^{\text {op}}, {{\mathscr {V}}}) \times \textrm{Fun}({\mathscr {I}}, {{\mathscr {M}}}) \longrightarrow {{\mathscr {M}}}\,, \qquad (D, F) \longmapsto \int ^{i \in {\mathscr {I}}} Di \otimes Fi\,. \end{aligned}$$
    (3.25)
  2. (2)

    The weighted \({\mathscr {I}}\)-limit functor of \({{\mathscr {M}}}\) is the functor

    $$\begin{aligned} \{-,-\}^{\mathscr {I}}:\textrm{Fun}({\mathscr {I}}, {{\mathscr {V}}})^{\text {op}}\times \textrm{Fun}({\mathscr {I}}, {{\mathscr {M}}}) \longrightarrow {{\mathscr {M}}}\,, \qquad (D, F) \longmapsto \int _{i \in {\mathscr {I}}} \{Di, Fi\}\,. \end{aligned}$$
    (3.26)

From now on we consider with the Kan-Quillen model structure, and thus \({{\mathscr {M}}}\) a large simplicial model category. The following result is [20, Thms. 3.2, 3.3] (see also [44, Thm. 11.5.1]).

Theorem 3.27

Let \({{\mathscr {M}}}\) be a large simplicial model category and \({\mathscr {I}}\) a large category. Suppose that the projective and injective model structures on \(\textrm{Fun}({\mathscr {I}}, {{\mathscr {M}}})\) both exist.

  1. (1)

    The weighted \({\mathscr {I}}\)-colimit functor (3.25) is a left Quillen functor of two variables [23, Def. 4.2.1] if its source and target are either endowed with the projective and injective model structures, respectively, or with the injective and projective model structures, respectively.

  2. (2)

    The weighted \({\mathscr {I}}\)-limit functor (3.26) is a right Quillen functor of two variables if its source and target are either both endowed with the projective model structures, or are both endowed with the injective model structures.

Recall that if \({{\mathscr {M}}}\) is combinatorial (in \(\text {L} \)), then the projective and injective model structures on \(\textrm{Fun}({\mathscr {I}}, {{\mathscr {M}}})\) exist and are again combinatorial (see, for instance, [3, Thms. 2.14, 2.16]). The projective model structure also exists when \({{\mathscr {M}}}\) is cellular, and is itself cellular [22, Prop. 12.1.5].

We also define functors

In the original reference [20] the following observation is the crucial ingredient in the proof of Theorem 3.27:

Proposition 3.28

In the situation of Theorem 3.27(2), with either the projective or injective model structure on both and \(\textrm{Fun}({\mathscr {I}}, {{\mathscr {M}}})\), the functors

form a Quillen adjunction of two variables.

Proof

In the projective case the functor

applies the -enrichment of \({{\mathscr {M}}}\) objectwise. Since pullbacks in functor categories are computed objectwise, it follows from the properties of this enrichment that \(\underline{{{\mathscr {M}}}}^{ptw}_{\mathscr {I}}(-,-)\) satisfies the pullback-product axiom (see [23, Lemma 4.2.2]). One proves the injective case analogously, using the functor \((-) \otimes ^{ptw}_{\mathscr {I}}(-)\) in place of \(\underline{{{\mathscr {M}}}}^{ptw}_{\mathscr {I}}\). \(\square \)

As before, we let \(Q :{\mathscr {H}}_{\infty ,0} \rightarrow {\mathscr {H}}_{\infty ,0}\) denote Dugger’s cofibrant replacement functor (see Definition 3.6). Consider the functors

$$\begin{aligned} \begin{array}{llll} Re_{{\mathscr {M}}}:\textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, {{\mathscr {M}}}) &{}\longrightarrow \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})\,, &{} \quad Re_{{\mathscr {M}}}({\widehat{G}})(c) &{}{:}{=}Q(-)(c) \otimes _{{{\widehat{{\mathscr {C}}}}}^{\text {op}}} {\widehat{G}} = \int ^{X \in {{\widehat{{\mathscr {C}}}}}} QX(c) \otimes {\widehat{G}}(X)\,, \\ S_{{\mathscr {M}}}:\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}}) &{}\longrightarrow \textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, {{\mathscr {M}}})\,, &{} \quad S_{{\mathscr {M}}}(F)(X) &{}{:}{=}\{QX, F\}^{{\mathscr {C}}^{\text {op}}} = \int _{c \in {\mathscr {C}}} \big \{ QX(c), Fc \big \}\,. \end{array} \end{aligned}$$

They form an adjoint pair

figure n

Proposition 3.30

The adjunction (3.29) is a Quillen adjunction with respect to the projective model structures.

Proof

Let \(X \in {{\widehat{{\mathscr {C}}}}}\), and consider the functor

$$\begin{aligned} \text {ev} _X \circ S_{{\mathscr {M}}}:\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}}) \longrightarrow {{\mathscr {M}}}\,, \qquad F \longmapsto \int _{c \in {\mathscr {C}}} \big \{ QX(c), Fc \big \} = \{QX, F\}^{{\mathscr {C}}^{\text {op}}}\,. \end{aligned}$$

Here we have used the notation from (3.26). By Theorem 3.27 and [23, Rmk. 4.2.3] this functor is right Quillen, for each \(X \in {{\widehat{{\mathscr {C}}}}}\). Since fibrations and weak equivalences in the projective model structure on \(\textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, {{\mathscr {M}}})\) are defined objectwise, it follows that the functor \(S_{{\mathscr {M}}}\) preserves projective fibrations and trivial fibrations. \(\square \)

We recall some useful technology for computing homotopy limits from [44] (see, in particular, Chapters 4–6). Let \({{\mathscr {M}}}\) be a large simplicial model category (enriched, tensored and cotensored over ). Let \({\mathscr {I}}\) be a large category and consider functors and \(F :{\mathscr {I}}\rightarrow {{\mathscr {M}}}\). The (two-sided) cosimplicial cobar construction associated to these data is the cosimplicial object in \({{\mathscr {M}}}\) whose l-th level reads as

$$\begin{aligned} C^l(G, {\mathscr {I}}, F)\! =\! \prod _{i \in ({\textsf {N}}{\mathscr {J}})_l} \{G i_0, F i_l \} \cong \prod _{i_0, \ldots , i_l \in {\mathscr {I}}} \big \{ G i_0 \otimes ({\mathscr {I}}(i_0, i_1) \times \cdots \times {\mathscr {I}}(i_{l-1}, i_l)),\, F i_l \big \} \quad \in {{\mathscr {M}}}\,, \end{aligned}$$

where denotes the nerve of \({\mathscr {I}}\). The (two-sided) cobar construction \(C(G, {\mathscr {I}}, F)\) in \({{\mathscr {M}}}\) is the totalisation

formed in \({{\mathscr {M}}}\). If the ambient model category \({{\mathscr {M}}}\) is not clear from context, we will write \(C_{{\mathscr {M}}}\) for the cobar construction in \({{\mathscr {M}}}\).

Proposition 3.31

[44, Cor. 5.1.3] The cobar construction is a model for the homotopy limit: let \({{\mathscr {M}}}\) be a large simplicial model category with fibrant replacement functor \(R^{{\mathscr {M}}}\), and let \(F :{\mathscr {I}}\rightarrow {{\mathscr {M}}}\) a functor as above. Then,

$$\begin{aligned} \underset{i \in {\mathscr {I}}}{{\text {holim}}}^{{\mathscr {M}}}(Fi) \simeq C(*, {\mathscr {I}}, R^{{\mathscr {M}}}F)\,. \end{aligned}$$

In particular, if \(Fi \in {{\mathscr {M}}}\) is fibrant for each \(i \in {\mathscr {I}}\), then

$$\begin{aligned} \underset{i \in {\mathscr {I}}}{{\text {holim}}}^{{\mathscr {M}}}(Fi) \simeq C(*, {\mathscr {I}}, F)\,. \end{aligned}$$

Lemma 3.32

Let \({{\mathscr {M}}}\) be a large simplicial model category, \({\mathscr {I}}\) a large category and \(F :{\mathscr {I}}\rightarrow {{\mathscr {M}}}\) a functor. Suppose that the projective model structure on \(\textrm{Fun}({\mathscr {I}}, {{\mathscr {M}}})\) exists and that F is projectively fibrant. Then, the cobar construction \(C(*, {\mathscr {I}}, F) \in {{\mathscr {M}}}\) is fibrant.

Proof

Recall the notation from Definition 3.24. By [44, Thm. 6.6.1] there is a canonical natural isomorphism

$$\begin{aligned} C(*, {\mathscr {I}}, D) \cong \int _{i \in {\mathscr {I}}} \big \{ N({\mathscr {I}}_{/i}),\, Fi \big \} = \{N({\mathscr {I}}_{/(-)}), F\}^{\mathscr {I}}\,. \end{aligned}$$

By Theorem 3.27, the functor is a right Quillen functor in two variables if we endow and \(\textrm{Fun}({\mathscr {I}}, {{\mathscr {M}}})\) with their projective model structures. The claim then follows from the fact that the functor is a projectively cofibrant object in \(\textrm{Fun}({\mathscr {I}}, {{\mathscr {M}}})\) [22, Cor. 14.8.8]. \(\square \)

Definition 3.33

Suppose \({{\mathscr {M}}}\) is left proper, as well as cellular or combinatorial. Let \(\tau \) be a Grothendieck coverage on \({\mathscr {C}}\), and \({{\widehat{\tau }}}\) the induced Grothendieck coverage of \(\tau \)-local epimorphisms on \({{\widehat{{\mathscr {C}}}}}\).

  1. (1)

    Let \(S({{\mathscr {M}}}, \tau )\) be the collection of morphisms in \(\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})\) of the form

    $$\begin{aligned} Q ({\check{C}}{{\mathcal {U}}}\rightarrow {{\mathscr {Y}}}_c) \otimes _{{\mathscr {C}}^{\text {op}}}^{ptw} m\,, \end{aligned}$$

    where \({{\mathcal {U}}}= \{c_i \rightarrow c\}_{i \in I}\) is a \(\tau \)-covering of \(c \in {\mathscr {C}}\), with associated Čech nerve \({\check{C}}{{\mathcal {U}}}\in {\mathscr {H}}_{\infty ,0}\), and \(m \in {{\mathscr {M}}}\) is a cofibrant object in \({{\mathscr {M}}}\). Define the left Bousfield localisation

    $$\begin{aligned} \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})^{loc} {:}{=}L_{S({{\mathscr {M}}}, \tau )} \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})\,. \end{aligned}$$
  2. (2)

    Let \(S({{\mathscr {M}}}, {{\widehat{\tau }}})\) be the collection of morphisms in \(\textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, {{\mathscr {M}}})\) of the form

    $$\begin{aligned} ({\check{C}}\pi \rightarrow {{\widehat{{{\mathscr {Y}}}}}}_X) \otimes _{{\mathscr {C}}^{\text {op}}}^{ptw} m\,, \end{aligned}$$

    where \(\pi :Y \rightarrow X\) is a \(\tau \)-local epimorphism in \({{\widehat{{\mathscr {C}}}}}\), with associated Čech nerve \({\check{C}}\pi \in {{\widehat{{\mathscr {H}}}}}_{\infty ,0}\), and \(m \in {{\mathscr {M}}}\) is a cofibrant object in \({{\mathscr {M}}}\). Define the left Bousfield localisation

    $$\begin{aligned} \textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, {{\mathscr {M}}})^{loc} {:}{=}L_{S({{\mathscr {M}}}, {{\widehat{\tau }}})} \textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, {{\mathscr {M}}})\,. \end{aligned}$$

These Bousfield localisations exist by [22, Thm. 4.4.1] and [3, Thm. 4.7]. The following is a generalisation of [3, Thm. 4.56] to the case where the model category \({{\mathscr {M}}}\) is not monoidal:

Proposition 3.34

In the situation of Definition 3.33 the following statements hold true:

  1. (1)

    An object \(F \in \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})\) is \(S({{\mathscr {M}}}, \tau )\)-local if and only if it is projectively fibrant and, for each \(\tau \)-covering \({{\mathcal {U}}}= \{c_i \rightarrow c\}_{i \in I}\), the induced morphism

    in \({{\mathscr {M}}}\) is a weak equivalence (where \({\check{C}}_k {{\mathcal {U}}}\) is viewed as a simplicially constant presheaf on \({\mathscr {C}}\)).

  2. (2)

    An object \({\widehat{G}} \in \textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, {{\mathscr {M}}})\) is \(S({{\mathscr {M}}}, {{\widehat{\tau }}})\)-local if and only if it is projectively fibrant and, for each \({{\widehat{\tau }}}\)-covering \(\pi :Y \rightarrow X\) in \({{\widehat{{\mathscr {C}}}}}\), the induced morphism

    in \({{\mathscr {M}}}\) is a weak equivalence.

Remark 3.35

Suppose that \(({\mathscr {C}}, \tau )\) is such that, for each \(\tau \)-covering \({{\mathcal {U}}}= \{c_i \rightarrow c\}_{i \in I}\), each presheaf \(C_{i_0 \cdots i_k}\) from (2.9) is again representable. In that case, the Čech nerve \({\check{C}}{{\mathcal {U}}}\) is projectively cofibrant (it is then objectwise a coproduct of representables), and so we can omit the cofibrant replacement functor Q from Definition 3.33(1) and Proposition 3.34. Furthermore, the enrichment and the Yoneda Lemma provide a canonical isomorphism

$$\begin{aligned} \{ {\check{C}}_k {{\mathcal {U}}}, F \}^{{\mathscr {C}}^{\text {op}}} = \prod _{i_0, \ldots , i_n \in I} \{ C_{i_0 \cdots i_n}, F \}^{{\mathscr {C}}^{\text {op}}} \cong \prod _{i_0, \ldots , i_n \in I} F(C_{i_0 \cdots i_n})\,, \end{aligned}$$

thus reproducing the more familiar form of the homotopy descent condition. \(\triangleleft \)

Proof of Proposition 3.34

We show Claim (1); the proof of Claim (2) is analogous. By definition, an object \(F \in \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})\) is \(S({{\mathscr {M}}}, \tau )\)-local if and only if it is projectively fibrant and, for each cofibrant \(m \in {{\mathscr {M}}}\) and each \(\tau \)-covering \({{\mathcal {U}}}= \{c_i \rightarrow c\}_{i \in I}\) in \({\mathscr {C}}\), the morphism

$$\begin{aligned} \underline{\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})} \big ( Q ({\check{C}}{{\mathcal {U}}}) \otimes _{{\mathscr {C}}^{\text {op}}}^{ptw} m, F \big ) \longrightarrow \underline{\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})} \big ( Q{{\mathscr {Y}}}_c \otimes _{{\mathscr {C}}^{\text {op}}}^{ptw} m, F \big ) \end{aligned}$$

induced by the augmentation \({\check{C}}{{\mathcal {U}}}\rightarrow {{\mathscr {Y}}}_c\) is a weak homotopy equivalence in (note that the left-hand arguments in both functors above are cofibrant in \(\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})\), and so the above simplicially enriched hom spaces indeed compute map** spaces). By adjointness, this morphism is isomorphic to the morphism

$$\begin{aligned} \underline{{{\mathscr {M}}}} \big ( m, \{ Q ({\check{C}}{{\mathcal {U}}}), F \}^{{\mathscr {C}}^{\text {op}}} \big ) \longrightarrow \underline{{{\mathscr {M}}}} \big ( m, \{ Q {{\mathscr {Y}}}_c, F \}^{{\mathscr {C}}^{\text {op}}} \big ) \simeq \underline{{{\mathscr {M}}}} \big ( m, F(c) \big )\,. \end{aligned}$$

Next, we use the isomorphisms in the homotopy category ,

Here we have used that we can commute Q and the homotopy colimit, since both are bar constructions. The claim now follows from the fact that a morphism \(f :a \rightarrow b\) in \({{\mathscr {M}}}\) between fibrant objects is a weak equivalence if and only if, for each cofibrant \(m \in {{\mathscr {M}}}\), the induced morphism

$$\begin{aligned} f^* :\underline{{{\mathscr {M}}}}(m, b) \longrightarrow \underline{{{\mathscr {M}}}}(m, a) \end{aligned}$$

is a weak homotopy equivalence; in other words, we use that the functors \(\underline{{{\mathscr {M}}}}(m, -)\), for \(m \in {{\mathscr {M}}}\) cofibrant, jointly detect weak equivalences between fibrant objects [22, Prop. 9.7.1]. \(\square \)

Definition 3.36

If \(({\mathscr {D}}, \tau _{\mathscr {D}})\) is an \(\text {L} \)-small category with a Grothendieck coverage \(\tau _{\mathscr {D}}\) and \({{\mathscr {M}}}\) a left proper, \(\text {L} \)-small, combinatorial or cellular model category, then we call the fibrant objects of \(L_{S({{\mathscr {M}}}, \tau _{\mathscr {D}})}\textrm{Fun}({\mathscr {D}}^{\text {op}}, {{\mathscr {M}}})\) the \({{\mathscr {M}}}\)-valued (homotopy) sheaves on \({\mathscr {D}}\).

Theorem 3.37

Let \({{\mathscr {M}}}\) be a left proper cellular or combinatorial simplicial model category in the universe \(\text {L} \) and \({\mathscr {C}}\) a small category. Let \(\tau \) be a Grothendieck coverage on \({\mathscr {C}}\), and \({{\widehat{\tau }}}\) the induced Grothendieck coverage of \(\tau \)-local epimorphisms on \({{\widehat{{\mathscr {C}}}}}\). The Quillen adjunction (3.29) descends to a Quillen adjunction

figure o

Proof

By Proposition 3.17, it suffices to show that the right adjoint \(S_{{\mathscr {M}}}\) preserves local objects. To that end, let \(F \in \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})\) be \(S({\mathscr {C}}, \tau )\)-local, and let \(\pi :Y \rightarrow X\) be a \(\tau \)-local epimorphism in \({{\widehat{{\mathscr {C}}}}}\). By Proposition 3.34 we thus need to show that the canonical morphism

is a weak equivalence in \({{\mathscr {M}}}\). Since \({\check{C}}_k \pi \) is representable in , the Yoneda Lemma for \({{\widehat{{\mathscr {C}}}}}\) implies that

$$\begin{aligned} \{ {\check{C}}_k \pi , S_{{\mathscr {M}}}F \}^{{{\widehat{{\mathscr {C}}}}}^{\text {op}}} \cong (S_{{\mathscr {M}}}F)({\check{C}}_k \pi ) = \{Q({\check{C}}_k \pi ), F\}^{{\mathscr {C}}^{\text {op}}}\,. \end{aligned}$$

The problem is thus equivalent to showing that the canonical morphism

is a weak equivalence in \({{\mathscr {M}}}\). Since the functors jointly detect weak equivalences, where m ranges over all cofibrant objects in \({{\mathscr {M}}}\) (and where \({{\mathscr {M}}}_{fib} \subset {{\mathscr {M}}}\) is the full subcategory on the fibrant objects), it suffices to show that, for each cofibrant \(m \in {{\mathscr {M}}}\), the morphism

is a weak homotopy equivalence. Here we have used that F is projectively fibrant, so that the functor \(\{Q(-), F\}^{{\mathscr {C}}^{\text {op}}}\) takes values in \({{\mathscr {M}}}_{fib}\), and that the homotopy limit of an objectwise fibrant diagram is again a fibrant object. Using the simplicial enrichment and the two-variable Quillen adjunction from Proposition 3.28, the above problem is further equivalent to showing that, for each cofibrant \(m \in {{\mathscr {M}}}\), the morphism

is a weak homotopy equivalence. Note that this morphism is the same as the canonical morphism

in the homotopy descent condition for the simplicial presheaf

By Theorem 3.23 it now suffices to show that the simplicial presheaf

is \(\tau \)-local, for each cofibrant object \(m \in {{\mathscr {M}}}\).

To that end, let \({{\mathcal {U}}}= \{c_i \rightarrow c\}_{i \in I}\) be a \(\tau \)-covering in \({\mathscr {C}}\). We have to check that the canonical morphism

is a weak homotopy equivalence. Again using the two-variable Quillen adjunctions, this is equivalent to checking that

is a weak homotopy equivalence, for each cofibrant \(m \in {{\mathscr {M}}}\). Again since the functors \(\underline{{{\mathscr {M}}}}(m,-)\) jointly detect weak equivalences between fibrant objects in \({{\mathscr {M}}}\), this is the case if and only if

is a weak equivalence in \({{\mathscr {M}}}\). However, that is true by the assumption that F is \(S({{\mathscr {M}}}, \tau )\)-local. \(\square \)

Finally, we record a generalisation of Proposition 3.12:

Proposition 3.38

In the setting of Theorem 3.30, for each projectively fibrant \(F \in \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {M}}})\) and any \(X \in {{\widehat{{\mathscr {C}}}}}\), there is a canonical natural weak equivalence

$$\begin{aligned} S_{{\mathscr {M}}}(F)(X) \simeq ({\textrm{hoRan}}_{{\mathscr {Y}}}F)(X)\,. \end{aligned}$$

Proof

We have canonical weak equivalences

In the second-to-last line we have used the enriched cobar construction for the simplicially enriched category \({{\mathscr {M}}}\). The final weak equivalence follows from Proposition 3.31. \(\square \)

3.3 Sheaves of \((\infty ,n)\)-Categories

The specific instances of Theorem 3.37 in which we are interested most arise from choosing the target model category \({{\mathscr {M}}}\) to be a model for \((\infty ,n)\)-categories. Various such model categories exist in the literature, including the model structures for n-fold complete Segal spaces [2,

Further, the category is cartesian closed, with internal hom given by

where is the Yoneda embedding of . For each \(n \in {\mathbb {N}}\), there is a monoidal adjunction

figure p

where \({\textsf {c}}_\bullet \) is the constant-diagram functor, i.e. \(({\textsf {c}}_\bullet X)_{k_0, \ldots , k_{n-1}} = X_{k_1, \ldots , k_{n-1}}\). Therefore, the category is enriched, tensored, and cotensored over for any \(0 \le k \le n\).

Moreover, for each \(k,l \in {\mathbb {N}}\) there is a functor

Iterating this, we have a functor

which satisfies that and \({\textsf {c}}_\bullet X = \Delta ^0 \boxtimes X\), for each .

The category of large bisimplicial sets carries a Reedy model structure, and it is a well-known fact that this coincides with the injective model structure [22, Thm. 15.8.7]. It is a less well-known fact that an analogous identity of the injective and Reedy model structures holds true for the categories [7, Prop. 3.15]. Whenever we refer to as a model category, we will mean the Reedy (or, equivalently, the injective) model structure. The model category is cellular, left proper, combinatorial and cartesian closed, i.e. it is a symmetric monoidal model structure in the sense of [23, Def. 4.2.6] whose monoidal product agrees with the categorical binary product: this follows readily from the fact that with the Kan-Quillen model structure is cartesian closed [23, Prop. 4.2.8] and the Reedy model structure on agrees with the injective model structure.

For the reader’s convenience, we briefly recall the presentation of a model category for n-fold complete Segal spaces, following [25, App. A] (see, in particular, the ‘Lurie-presentation’ in loc. cit.).

For \(k \in {\mathbb {N}}\), \(k \ge 2\), define the k-th spine inclusion

figure q

in . Localising at maps of this type will implement the Segal conditions.

Let denote the pushout

figure r

where the top morphism consists of the inclusions of \(\Delta ^1\) into \(\Delta ^3\) as the edge \(0 \rightarrow 2\) and \(1 \rightarrow 3\), respectively. Let \(\kappa \) denote the collapse map

$$\begin{aligned} \kappa :K \rightarrow \Delta ^0\,. \end{aligned}$$

Localising at maps of the type \(\kappa \) will implement the completeness conditions below.

Definition 3.39

Let \(n \in {\mathbb {N}}\). The model category for (large) n-fold complete Segal spaces is obtained by left Bousfield localisation of at the following morphisms:

  1. (1)

    (Segal conditions) For each \(k \in {\mathbb {N}}\), \(i \in \{0, \ldots , n-1\}\) and each \(k_0, \ldots , k_{i-2}, k_i, \ldots , k_{n-1} \in {\mathbb {N}}_0\), we localise at the morphism

    $$\begin{aligned} \Delta ^{k_0} \boxtimes \cdots \boxtimes \Delta ^{k_{i-2}} \boxtimes \sigma _k \boxtimes \Delta ^{k_i} \boxtimes \cdots \boxtimes \Delta ^{k_{n-1}}\,. \end{aligned}$$
  2. (2)

    (Completeness) For each \(i \in \{0, \ldots , n-1\}\) and each \(k_0, \ldots , k_{i-2}, k_i, \ldots , k_{n-1} \in {\mathbb {N}}_0\), we localise at the morphism

    $$\begin{aligned} \Delta ^{k_0} \boxtimes \cdots \boxtimes \Delta ^{k_{i-2}} \boxtimes \kappa \boxtimes \Delta ^{k_i} \boxtimes \cdots \boxtimes \Delta ^{k_{n-1}}\,. \end{aligned}$$
  3. (3)

    (Essential constancy) For each \(i \in \{0, \ldots , n-1\}\) and each \(k_0, \ldots , k_{i-2}, k_i, \ldots , k_{n-1} \in {\mathbb {N}}_0\), we localise at the morphism

    $$\begin{aligned} \Delta ^{k_0} \boxtimes \cdots \boxtimes \Delta ^{k_{i-2}} \boxtimes \Delta ^0 \boxtimes \Delta ^{k_i} \boxtimes \cdots \boxtimes \Delta ^{k_{n-1}} \longrightarrow \Delta ^{k_0} \boxtimes \cdots \boxtimes \Delta ^{k_{i-2}} \boxtimes \Delta ^0 \boxtimes \Delta ^0 \boxtimes \cdots \boxtimes \Delta ^0\,. \end{aligned}$$

We denote the resulting model category by \(\mathscr {C}\mathscr {S}\mathscr {S}_n\). Fibrant objects in \(\mathscr {C}\mathscr {S}\mathscr {S}_n\) are called n-fold complete Segal spaces, or \((\infty ,n)\)-categories.

Remark 3.40

The model categories \(\mathscr {C}\mathscr {S}\mathscr {S}_n\) exist and are left proper, cellular and simplicial [25, Lemma A.6]. In particular, we can apply Theorem 3.37 to presheaves with values in \(\mathscr {C}\mathscr {S}\mathscr {S}_n\), for each \(n \in {\mathbb {N}}\). \(\triangleleft \)

Lemma 3.41

Let \(n \in {\mathbb {N}}\), let \({\mathscr {I}}\) be a large category, and let \(D :{\mathscr {I}}\rightarrow \mathscr {C}\mathscr {S}\mathscr {S}_n\) be a projectively fibrant diagram, i.e. Di is an n-fold complete Segal space for every \(i \in {\mathscr {J}}\).

  1. (1)

    The homotopy limit of D can be computed levelwise: for each \(k \in {\mathbb {N}}_0\), there exist canonical natural isomorphisms

    $$\begin{aligned} \big ( {\text {holim}}^{\mathscr {C}\mathscr {S}\mathscr {S}_n}_{\mathscr {J}}(D) \big )_k \cong {\text {holim}}^{\mathscr {C}\mathscr {S}\mathscr {S}_{n-1}}_{\mathscr {J}}(D_k)\,, \end{aligned}$$

    where for \(i \in {\mathscr {J}}\) the object \(D_k i = (Di)_k\) is the k-th simplicial level of .

  2. (2)

    The homotopy limit of D can be computed objectwise when viewing the diagram as a functor . That is, for any there is an isomorphism (natural in )

Proof

We use the explicit model for homotopy limits in simplicially enriched categories from Proposition 3.31. For the sake of brevity, for \(l \in {\mathbb {N}}_0\), we set

Since D is objectwise fibrant, a model for the homotopy limit of D is given by the two-sided cobar construction \(C(*,{\mathscr {I}},D) \in \mathscr {C}\mathscr {S}\mathscr {S}_n\) [44, Cor. 5.1.3].

For any \(n \in {\mathbb {N}}_0\), there are canonical natural isomorphisms of simplicial sets

The last identity relies on [44, Cor. 5.1.3] (see also Proposition 3.31), together with the following facts: the functor \(D :{\mathscr {I}}\rightarrow \mathscr {C}\mathscr {S}\mathscr {S}_n\) is objectwise fibrant. In particular, for each \(i \in {\mathscr {J}}\), the object is fibrant in the Reedy model structure (equivalently, the injective model structure), where carries the Reedy (equivalently, injective) model structure. Forgetting the localisations in Definition 3.39 at morphisms which are non-trivial in the first -factor, we further obtain that \((Di)_m\) is fibrant in \(\mathscr {C}\mathscr {S}\mathscr {S}_{n-1}\), for each \(i \in {\mathscr {I}}\) and . Consequently, the functor , \(i \mapsto D_m(i) = (Di)_m\) is also fibrant in the projective model structure on . Thus, the cobar construction in the second-to-last line indeed models its homotopy limit. Part (2) follows in analogy with part (1), replacing \({{\mathscr {Y}}}^n\) with \({{\mathscr {Y}}}\) (i.e. replacing \(\Delta ^k \boxtimes \Delta ^0 \boxtimes \cdots \boxtimes \Delta ^0\) by \(\Delta ^{k_0} \boxtimes \cdots \boxtimes \Delta ^{k_{n-1}}\)) and using that injectively fibrant diagrams are, in particular, projectively fibrant. \(\square \)

Definition 3.42

Let \({\mathscr {C}}\) be a small category.

  1. (1)

    We write

    $$\begin{aligned} {\mathscr {H}}_{\infty ,n} {:}{=}\textrm{Fun}({\mathscr {C}}^{\text {op}}, \mathscr {C}\mathscr {S}\mathscr {S}_n) \end{aligned}$$

    and view this as endowed with the projective model structure. A fibrant object in \({\mathscr {H}}_{\infty ,n}\) is called a presheaf of \((\infty ,n)\)-categories on \({\mathscr {C}}\).

  2. (2)

    In the notation of Definition 3.33, we define the left Bousfield localisation

    $$\begin{aligned} {\mathscr {H}}_{\infty ,n}^{loc} {:}{=}L_{S(\mathscr {C}\mathscr {S}\mathscr {S}_n,\tau )} {\mathscr {H}}_{\infty ,n}\,. \end{aligned}$$

    A fibrant object in \({\mathscr {H}}_{\infty ,n}^{loc}\) is called a sheaf of \((\infty ,n)\)-categories on \(({\mathscr {C}},\tau )\).

Proposition 3.43

Let \({\mathscr {C}}\) be a small category. Then \({\mathscr {H}}_{\infty ,n}\) is a left proper, tractable (large) simplicial model category. Further, if \({\mathscr {C}}\) is endowed with a Grothendieck pretopology \(\tau \), then \({\mathscr {H}}_{\infty ,n}^{loc}\) is a \(\mathscr {C}\mathscr {S}\mathscr {S}_n\)-model category, and it is symmetric monoidal if \({\mathscr {H}}_{\infty ,n}\) is so.

Proof

The statement about \({\mathscr {H}}_{\infty ,n}\) is a consequence of [3, Prop. 4.50]. For \({\mathscr {H}}_{\infty ,n}^{loc}\), the claim follows from [3, Prop. 4.47]. The statement about symmetric monoidal structures is an application of [3, Thm. 4.58]. \(\square \)

Analogously to Sects. 3.1, \({\mathscr {H}}_{\infty ,n}\) is symmetric monoidal whenever \({\mathscr {C}}\) has finite products.

Definition 3.44

Let \({\mathscr {C}}\) be a small category. As before, we write \({{\widehat{{\mathscr {C}}}}}= \textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {S}}\text {et} }_\text {S} )\).

  1. (1)

    We write

    $$\begin{aligned} {{\widehat{{\mathscr {H}}}}}_{\infty ,n} {:}{=}\textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, \mathscr {C}\mathscr {S}\mathscr {S}_n) \end{aligned}$$

    for the category \(\textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, \mathscr {C}\mathscr {S}\mathscr {S}_n)\), endowed with the projective model structure. A fibrant object in \({{\widehat{{\mathscr {H}}}}}_{\infty ,n}\) is called a presheaf of \((\infty ,n)\)-categories on \({{\widehat{{\mathscr {C}}}}}\).

  2. (2)

    In the notation of Definition 3.33, we define the left Bousfield localisation

    $$\begin{aligned} {{\widehat{{\mathscr {H}}}}}_{\infty ,n}^{loc} {:}{=}L_{S(\mathscr {C}\mathscr {S}\mathscr {S}_n,{{\widehat{\tau }}})} {{\widehat{{\mathscr {H}}}}}_{\infty ,n}\,. \end{aligned}$$

    A fibrant object in \({{\widehat{{\mathscr {H}}}}}_{\infty ,n}^{loc}\) is called a sheaf of \((\infty ,n)\)-categories on \(({{\widehat{{\mathscr {C}}}}},{{\widehat{\tau }}})\).

Remark 3.45

Lemma 3.41 implies that an object \({\mathscr {F}}\in \textrm{Fun}({\mathscr {C}}^{\text {op}}, \mathscr {C}\mathscr {S}\mathscr {S}_n)\) is \(S(\mathscr {C}\mathscr {S}\mathscr {S}_n,\tau )\)-local if and only if, for each the object is \(\tau \)-local, and analogously for \(({{\widehat{{\mathscr {C}}}}}, {{\widehat{\tau }}})\) in place of \(({\mathscr {C}}, \tau )\). \(\triangleleft \)

In the notation of (3.29), we further write

$$\begin{aligned} S_{\infty ,n} {:}{=}S_{\mathscr {C}\mathscr {S}\mathscr {S}_n} :{\mathscr {H}}_{\infty ,n} \longrightarrow {{\widehat{{\mathscr {H}}}}}_{\infty ,n}\,, \qquad (S_{\infty ,n} {\mathscr {F}})(X) = \{ QX, {\mathscr {F}}\}^{{\mathscr {C}}^{\text {op}}}\,. \end{aligned}$$

We can compute this functor more explicitly as follows: for , we have canonical isomorphisms

We have the following direct corollary of Proposition 3.38:

Proposition 3.46

On any fibrant object \({\mathscr {F}}\in {\mathscr {H}}_{\infty ,n}\), the functor \(S_{\infty ,n}\) agrees with the homotopy right Kan extension of \({\mathscr {F}}:{\mathscr {C}}^{\text {op}}\rightarrow \mathscr {C}\mathscr {S}\mathscr {S}_n\) along the Yoneda embedding,

$$\begin{aligned} S_{\infty ,n}{\mathscr {F}}\cong {\textrm{hoRan}}_{{\mathscr {Y}}}({\mathscr {F}})\,. \end{aligned}$$

As a direct application of Proposition 3.30, we have:

Proposition 3.47

The functor \(S_{\infty ,n}\) is the right adjoint in a Quillen adjunction

figure s

Moreover, Theorem 3.37 implies the following result:

Corollary 3.48

Let \(({\mathscr {C}},\tau )\) be a small site. For any \(n \in {\mathbb {N}}\), the Quillen adjunction \(Re_{\infty ,n} \dashv S_{\infty ,n}\) descends to a Quillen adjunction

figure t

Remark 3.49

The descent results in this subsection also imply corresponding results for strict presheaves of 1-categories, i.e. functors

$$\begin{aligned} F :{\mathscr {C}}^{\text {op}}\rightarrow {{\mathscr {C}}\text {at} }_\text {L} \,, \end{aligned}$$

from \({\mathscr {C}}^{\text {op}}\) to the 1-category of large categories. Rezk’s classifying diagram functor \(N_{rel} :{{\mathscr {C}}\text {at} }_\text {L} \rightarrow \mathscr {C}\mathscr {S}\mathscr {S}_1\) from large categories to large complete Segal spaces is a fully faithful embedding of categories which preserves and detects weak equivalences (weak equivalences in \({{\mathscr {C}}\text {at} }_\text {L} \) being equivalences of categories) and which preserves exponentials; this is the content of [41, Thm. 3.7]. Explicitly, if \({\mathscr {D}}^\sim \) denotes the maximal subgroupoid of a category \({\mathscr {D}}\), we have, for each \(n \in {\mathbb {N}}_0\),

$$\begin{aligned} (N_{rel} {\mathscr {D}})_n = N \big ( ({\mathscr {D}}^{[n]})^\sim \big )\,. \end{aligned}$$

Now let \(({\mathscr {C}}, \tau )\) be a site. One checks that, given a presheaf \(F :{\mathscr {C}}^{\text {op}}\rightarrow {{\mathscr {C}}\text {at} }_\text {L} \) and a \(\tau \)-covering family \({{\mathcal {U}}}= \{c_i \rightarrow c\}_{i \in \Lambda }\) such that all \(C_{i_0 \ldots i_n}\) are representable (see (2.9) for the notation), there is an equivalence

$$\begin{aligned} \underline{{\mathscr {H}}}_{\infty ,1}(Q {\check{C}}{{\mathcal {U}}}, N_{rel} F) \simeq \underline{{\mathscr {H}}}_{\infty ,1}({\check{C}}{{\mathcal {U}}}, N_{rel} F) \simeq N_{rel} {{\mathscr {D}}\text {esc} }(F; {{\mathcal {U}}})\,, \end{aligned}$$

where \({{\mathscr {D}}\text {esc} }(F; {{\mathcal {U}}})\) is the usual descent category associated to F and the covering family \({{\mathcal {U}}}\) (see, for instance, [37, Def. 2.5] or [10, Def. 2.9]). It follows that F is a sheaf of categories precisely if \(N_{rel} F\) is fibrant in \({\mathscr {H}}_{\infty ,1}^{loc}\). Given a small presheaf \(X \in {{\widehat{{\mathscr {C}}}}}\), one can check that

$$\begin{aligned} \underline{{\mathscr {H}}}_{\infty ,1}(QX, N_{rel} F) \cong N_{rel} \Big ( {\text {holim}}^{{{\mathscr {C}}\text {at} }_\text {L} } \big ( ({\mathscr {C}}_{/X})^{\text {op}}\longrightarrow {\mathscr {C}}^{\text {op}}\overset{F}{\longrightarrow }\ {{\mathscr {C}}\text {at} }_\text {L} \big ) \Big )\,, \end{aligned}$$

in \(\mathscr {C}\mathscr {S}\mathscr {S}_1\), where the homotopy limit is computed in the canonical model structure on \({{\mathscr {C}}\text {at} }_\text {L} \) (see, for instance, [40]). We define a functor

$$\begin{aligned}&S_{1,1} :\textrm{Fun}({\mathscr {C}}^{\text {op}}, {{\mathscr {C}}\text {at} }_\text {L} ) \longrightarrow \textrm{Fun}({{\widehat{{\mathscr {C}}}}}^{\text {op}}, {{\mathscr {C}}\text {at} }_\text {L} )\,, \\&S_{1,1}(F)(X) {:}{=}{\text {holim}}^{{{\mathscr {C}}\text {at} }_\text {L} } \big ( ({\mathscr {C}}_{/X})^{\text {op}}\longrightarrow {\mathscr {C}}^{\text {op}}\overset{F}{\longrightarrow }\ {{\mathscr {C}}\text {at} }_\text {L} \big )\,. \end{aligned}$$

Given a \(\tau \)-local epimorphism \(\pi :Y \rightarrow X\), we thus have a commutative diagram

figure u

Since \(N_{rel}\) reflects weak equivalences, it now follows from Corollary 3.48 that \(S_{1,1}F\) satisfies \({{\widehat{\tau }}}\)-descent whenever F satisfies \(\tau \)-descent. \(\triangleleft \)