Hostname: page-component-7f64f4797f-pmsz9 Total loading time: 0 Render date: 2025-11-08T18:23:01.287Z Has data issue: false hasContentIssue false

A flag version of Beilinson–Drinfeld Grassmannian for surfaces

Published online by Cambridge University Press:  06 November 2025

Benjamin Hennion
Affiliation:
Laboratoire de Mathématiques d’Orsay, Université Paris-Saclay, Orsay, France benjamin.hennion@universite-paris-saclay.fr
Valerio Melani
Affiliation:
DIMAI, Università degli Studi di Firenze, Firenze, Italy valerio.melani@unifi.it
Gabriele Vezzosi
Affiliation:
DIMAI, Università degli Studi di Firenze, Firenze, Italy gabriele.vezzosi@unifi.it
Rights & Permissions [Opens in a new window]

Abstract

We define and study a generalization of the Beilinson–Drinfeld Grassmannian to the case where the curve is replaced by a smooth projective surface $X$, and the trivialization data are given on loci associated with a nonlinear flag of closed subschemes. We first establish some general formal gluing results for moduli of (almost) perfect complexes and torsors. We construct a simplicial object $\operatorname {\underline {\mathsf{Fl}}}_X$ of flags of closed subschemes of a smooth projective surface $X$, associated with the operation of taking union of flags. We prove that this simplicial object has the $2$-Segal property. For an affine complex algebraic group $G$, we define a derived, flag analogue $\mathcal{G}r_X$ of the Beilinson–Drinfeld Grassmannian of $G$-bundles on the surface $X$, and show that most of the properties of the Beilinson–Drinfeld Grassmannian for curves can be extended to our flag generalization: we prove a factorization formula, the existence of a canonical flat connection and define a chiral product on suitable sheaves on $\operatorname {\underline {\mathsf{Fl}}}_X$ and on $\mathcal{G}r_X$. We sketch the construction of actions of flags analogues of the loop group and of the positive loop group on $\mathcal{G}r_X$. To fixed ‘large’ flags on $X$, we associate ‘exotic’ derived structures on the stack of $G$-bundles on $X$.

MSC classification

Information

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of Foundation Compositio Mathematica

1. Introduction

Given a connected smooth complex projective curve $C$ and ${\bf G}$ a smooth affine $\mathbb{C}$ -group scheme, the corresponding (global) affine Grassmannian is the functor

\begin{align*}\underline {{\rm Gr}}_{C,{\bf G}}: (\mathsf{Aff}_{\mathbb{C}})^{{\rm op}} \to {\bf Sets} \, : \, S \longmapsto \left \{ (x, \mathcal{E}, \varphi ) \, |\, x \in C(S), \mathcal{E}\in {\rm Bun}^{{\bf G}}(C \times S), \varphi \;:\; \mathcal{E}_{| (C\times S)\smallsetminus \Gamma _{x}} \simeq \mathcal{E}_0 \right \},\end{align*}

where ${\rm Bun}^{{\bf G}}(C \times S)$ is the groupoid of ${\bf G}$ -torsors on $C \times S$ , $\Gamma _x \subset C \times S$ denotes the graph of $x \in C(S)$ and $\mathcal{E}_0$ denotes the trivial ${\bf G}$ -bundle (see, e.g., [Reference ZhuZhu17, 1.4]). A local version of $\underline {{\rm Gr}}_{C,{\bf G}}$ can also be defined using ${\bf G}$ -bundles only defined on a formal neighbourhood of $\Gamma _x$ and trivializations defined on the corresponding affine punctured neighbourhood [Reference ZhuZhu17, (1.2.1)], and Beauville–Laszlo theorem [Reference Beauville and LaszloBL95] ensures that such a local version of the affine Grassmannian is in fact equivalent to the global version recalled above (see, e.g., [Reference ZhuZhu17, Theorem 1.4.2]).

A brilliant idea of A. Beilinson and V. Drinfeld was to generalize the construction of the affine Grassmannian by allowing an arbitrary, finite number of points $x$ . This can be done by first introducing the so called Ran space of the curve $C$ as the presheaf

\begin{align*}\mathsf{Ran}_C: (\mathsf{Aff}_{\mathbb{C}})^{{\rm op}} \to {\bf Sets} \, : \, S \longmapsto \left \{ \textrm {finite non-empty subsets of } C(S)\right \},\end{align*}

and then defining

\begin{align*}\underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}}: (\mathsf{Aff}_{\mathbb{C}})^{{\rm op}} \to {\bf Sets} : S \longmapsto \left \{ (\underline {x}, \mathcal{E}, \varphi ) |\, \underline {x} \in \mathsf{Ran}_C (S), \mathcal{E}\in {\rm Bun}^{{\bf G}}(C \times S), \varphi : \mathcal{E}_{| (C\times S)\smallsetminus \Gamma _{\underline {x}}} \simeq \mathcal{E}_0 \right \}\!,\end{align*}

where $\Gamma _{\underline {x}} \subset C \times S$ denotes the joint graph of $\underline {x} \in \mathsf{Ran}_C(S)$ . The functor $\underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}}$ is called the (Ran version of the) Beilinson–Drinfeld affine Grassmannian of the pair $(C, {\bf G})$ . There is an obvious morphism of presheaves $p: \underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}} \to \mathsf{Ran}_C$ , and the trivial ${\bf G}$ -bundle endows it with a canonical section $\mathsf{triv}: \mathsf{Ran}_C \to \underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}}$ . Although $\mathsf{Ran}_C$ is not representable,Footnote 1 not even by an ind-scheme, the map $p$ is itself ind-representable and it is even ind-proper when ${\bf G}$ is reductive [Reference ZhuZhu17, 3.3]. The presheaf $\underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}}$ is not represented by an ind-scheme but it is the presheaf (non-filtered) colimit of ind-schemes under proper maps (i.e. a pseudo-ind-scheme as in [Reference GaitsgoryGai13, 1.2]). Although passing from $\underline {{\rm Gr}}_{C,{\bf G}}$ to $\underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}}$ might seem an easy and obvious step, it does make it possible for $\underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}}$ to exhibit an important property called factorization. First of all notice that $\mathsf{Ran}_C$ carries a (non-unital) semigroup structure given by the union

\begin{align*}\cup \;:\; \mathsf{Ran}_C \times \mathsf{Ran}_C \to \mathsf{Ran}_C: (\{\underline {x}\}, \{\underline {y}\}) \mapsto \{\underline {x}, \underline {y}\}.\end{align*}

Let $(\mathsf{Ran}_C \times \mathsf{Ran}_C)_{{\rm disj}} \subset \mathsf{Ran}_C \times \mathsf{Ran}_C$ be the sub-presheaf defined by those pairs $(\{\underline {x}\}, \{\underline {y}\})$ such that $\{\underline {x}\} \cap \{\underline {y}\}= \emptyset$ . We define similarly $(\mathsf{Ran}_C \times \mathsf{Ran}_C \times \mathsf{Ran}_C )_{{\rm disj}} \subset \mathsf{Ran}_C \times \mathsf{Ran}_C \times \mathsf{Ran}_C$ . Then, the factorization property of $\underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}}$ can be stated as the existence of a natural isomorphism (called factorization isomorphism) between the pullbacks of the two diagrams

(1.1)

which is compatible, in an obvious way, with the map $\mathsf{triv}: \mathsf{Ran}_C \to \underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}}$ and satisfies a natural cocycle condition on $(\mathsf{Ran}_C \times \mathsf{Ran}_C \times \mathsf{Ran}_C )_{{\rm disj}}$ [Reference ZhuZhu17, Theorem 3.3.3]. We may denote $\mathsf{Gr}_{C,{\bf G}; 2}^{{\rm BD}}$ the common (up to canonical isomorphism) pullback of these two diagrams. The map $p:\mathsf{Gr}_{C,{\bf G}; 2}^{{\rm BD}} \to \mathsf{Ran}_C$ carries a flat connection ([Reference ZhuZhu17, Theorem 3.1.20] and [Reference Beilinson and DrinfeldBD99, 2.8]) and actions of suitable Ran-versions ${\rm L}_{\mathsf{Ran}}{\bf G}$ and ${\rm L}^+_{\mathsf{Ran}}{\bf G}$ of the loop group ${\rm L}{\bf G}$ and of the positive loop group ${\rm L}^+{\bf G}$ , respectively ([Reference GaitsgoryGai22, 1.1.2] and [Reference Mirkovic and VilonenMV07] § 5, (5.5)).

In this paper we propose and study an analogue of Beilinson-Drinfeld Grassmannian, called the flag Grassmannian, where the curve $C$ is replaced by a smooth complex projective surface $X$ , and points in the Ran space of $C$ are replaced with (families) of flags of subschemes in $X$ . Our investigation on the flag Grassmannian has its origins in some conversations of the third author with Mauro PortaFootnote 2 triggered by § 3.4.6 of [Reference Beilinson and DrinfeldBD04] where an equivalent approach to factorization algebras sheaves is based on effective Cartier divisors (possibly considered up to reduced equivalence) on a curve, rather than on the Ran space of the curve. This remark immediately gave us the idea of a possible replacement of the Ran space in the higher dimensional case of a surface, with the further possibility of including arbitrary flags consisting of an effective Cartier divisor together with a codimension 1 subscheme in this divisor.

Let us summarize the results of this paper.

Formal glueing. In proving several properties of our flag Grassmannian, we will need a higher-dimensional version of Beauville–Laszlo gluing, therefore we devote the first three sections to establish a general version of formal gluing. Let $k$ be field. For a $k$ -scheme $Y$ and a closed subscheme $Z$ , with open complement $U$ , roughly speaking, formal gluing (whenever it exists) is a way of describing geometric objects of some kind defined on $Y$ , as geometric objects of the same kind defined on $U$ and on the formal completion $\widehat {Z}$ of $Z$ in $Y$ , that are suitably compatible. The notion of compatibility here is a delicate one. Morally, guided by a purely topological intuition, one would like to say that a geometric object $\mathcal{E}_U$ and a geometric object $\mathcal{E}_{{\widehat {Z}}}$ are compatible if they restrict to the same object (or to an equivalent one) on the intersection $U \cap {\widehat {Z}}= {\widehat {Z}} \smallsetminus Z$ . However, strictly speaking, in algebraic geometry this intersection is empty, so the notion of compatibility cannot be the naive one, at least if we want to remain within the realm of algebraic geometry, i.e. we want to keep viewing $Y$ , $U$ and $\widehat {Z}$ as algebro-geometric objects. Many formal glueing results have been proved, and we give here only a probably non-exhaustive list: [Reference WeilWei82], [Reference ArtinArt70], [Reference Ferrand and RaynaudFR70], [Reference Beauville and LaszloBL95], [Reference Moret-BaillyMB96], [Reference Ben-Bassat and TemkinBBT13], [Reference BhattBha16], [Reference SchäppiSch15], [Reference Hall and RydhHR16] and [Reference MathewMat22]. In the first sections, based on some of these previous works (mainly [Reference BhattBha16] and [Reference MathewMat22]), we address the general formal glueing problem when the geometric objects above are

  • almost perfect modules (also known as pseudo-coherent complexes) on $Y$ , whose derived moduli stack is denoted by ${\bf Coh}^-_Y$ in the paper,

  • perfect modules on $Y$ , whose derived moduli stack is denoted by ${\bf Perf}_Y$ in the paper,

  • ${\bf G}$ -bundles on $Y$ , for ${\bf G}$ an affine $k$ -group scheme, whose derived moduli stack is denoted by ${\bf Bun}^{{\bf G}}_Y$ in the paper,

and $Y$ is a locally Noetherian stack over $k$ . If ${\bf F} = {\bf Coh}^-, {\bf Perf}$ or ${\bf Bun}^{{\bf G}}$ , and $Z\subset Y$ is a closed substack, restriction induces an equivalence of derived stacks over $Y$ (i.e. defined on the category $\mathsf{dAff}_{Y}$ of derived affine Noetherian schemes ${\rm Spec}\, R$ with a map ${\rm Spec}\, R \to Y$ )

(1.2)

This is proved in Proposition 3.3 and Theorem 3.5. In this formula, $\widehat {Z}$ denotes the formal completion of $Z$ inside $Y$ , while ${\widehat {Z}}^{{\rm aff}} \smallsetminus Z$ is defined using the fibre functor Footnote 3 of affinization (see Definitions 2.10 and 2.16). In particular, when $R$ is an underived commutative $k$ -algebra together with a map ${\rm Spec}\, R \to Y$ , the value of ${\widehat {Z}}^{{\rm aff}} \smallsetminus Z$ on $({\rm Spec}\, R \to Y)$ is the expected one

\begin{align*}({\widehat {Z}}^{{\rm aff}} \smallsetminus Z)({\rm Spec}\, R \to Y) = {\rm Spec}(\widehat {R}_{{\rm I}_Z})\smallsetminus {\rm Spec}(R/{\rm I}_Z),\end{align*}

where ${\rm I}_Z\subset R$ denotes the ideal of the pullback of $Z$ to ${\rm Spec}\, R$ , and $\widehat {R}_{{\rm I}_Z}$ is the ring completion of $R$ at the ideal ${\rm I}_Z$ . Note that, by algebraization (Proposition 3.1), the canonical map ${\bf F}_{{\widehat {Z}}} \to {\bf F}_{{\widehat {Z}}^{{\rm aff}}}$ is an equivalence of derived stacks over $Y$ (where ${\widehat {Z}}^{{\rm aff}}({\rm Spec}\, R \to Y)= {\rm Spec}(\widehat {R}_{{\rm I}_Z})$ , when ${\rm Spec}\, R$ is a Noetherian underived affine $k$ -scheme endowed with a map to $Y$ ). In the presence of a longer nonlinear flag $(Z_k\subset Z_{k-1} \subset \cdots \subset Z_1 \subset Y)$ of closed substacks in $Y$ , the formal glueing formula (1.2) can be iterated yielding what we call a flag decomposition of ${\bf F}$ (see § $\,$ 3.4).

Using the map $Y \to {\rm Spec}\, k$ , and the induced functor

(1.3) \begin{align} \mathsf{dSt}_Y \to \mathsf{dSt}_k \;:\; {\bf F} \mapsto \underline {\mathsf{F}}, \end{align}

from derived stacks defined over $\mathsf{dAff}_{Y}$ to derived stacks defined over $\mathsf{dAff}_{k}$ , formula (1.2) translates into the corresponding obvious formal glueing equivalence between derived stacks over $k$

(1.4)

(see Construction 3.7).

In our analysis, the case of ${\bf G}$ -bundles follows from the case of perfect modules (since ${\bf Bun}^{{\bf G}}_Y$ is Perf-local in the variable $Y$ , see Lemma 2.24), while the case of almost perfect modules is similar to that of perfect modules. Therefore, we have decided to present these three formal gluing results together, although in the rest of the paper we will mainly concentrate on the study of ${\bf G}$ -bundles. Moreover, for the sake of completeness, and since this does not require an excessive effort, we actually establish formal gluing and flag decomposition results, in the more general case where $Y$ a locally Noetherian derived stack over $k$ , $Z \subset Y$ is a derived closed substack.

Section 3.5 explains how to deduce from the previous results formal gluing for the underived versions of ${\bf F} = {\bf Coh}^-, {\bf Perf}$ or ${\bf Bun}^{{\bf G}}$ .

Nonlinear flags. Starting with § 4, we will let $k=\mathbb{C}$ , and fix a smooth complex projective surface $X$ . In § 4 we introduce the simplicial object $\underline {\mathsf{Fl}}_{X,\bullet }$ of families of nonlinear flags on $X$ . Here is the basic idea of the construction of $\underline {\mathsf{Fl}}_{X,\bullet }$ . We first define $\operatorname {\underline {\mathsf{Fl}}}_{X,1}$ (which we also denote simply by $\operatorname {\underline {\mathsf{Fl}}}_X$ ) as the presheaf of posets on $\mathsf{Aff}_{\mathbb{C}}$ assigning to $S$ the poset whose elements are pairs $(D, Z)$ consisting of a relative effective Cartier divisor $D$ on $X \times S/ S$ , and an element $Z$ in the Hilbert scheme of $0$ -dimensional subschemes of $X \times S /S$ , subject to the nesting condition $Z \subset D$ . The partial order on $\operatorname {\underline {\mathsf{Fl}}}_{X,1}(S)$ is defined as follows: we say that $(D,Z) \leq (D',Z')$ (also denoted as a morphism $(D,Z) \to (D',Z')$ ) if there exists a commutative square of closed subschemes of $X \times S$

such that the canonical induced map $Z \to D\times _{D'} Z'$ yields an isomorphism on the induced reduced scheme structures. Note that by forgetting the partial order structure, $\underline {\mathsf{Fl}}_{X,1}$ is just a variant of the so-called flag Hilbert scheme of $X$ (see, e.g., [Reference SernesiSer06, 4.5]). The main step in the construction of $\underline {\mathsf{Fl}}_{X,\bullet }$ is to define $\underline {\mathsf{Fl}}_{X,2}$ in such a way that it supports a union map $\bigcup \;:\; \underline {\mathsf{Fl}}_{X,2} \to \underline {\mathsf{Fl}}_{X,1}$ of the form $(D_1, Z_1) \cup (D_2, Z_2)= (D_1 +D_2, Z_1 \cup Z_2)$ . In order to ensure that such a map is well defined (e.g. that $Z_1 \cup Z_2$ is flat over the test scheme $S$ ), we define $\underline {\mathsf{Fl}}_{X,2}(S)$ as the sub-poset of pairs $((D_1, Z_1), (D_2, Z_2))$ in $\underline {\mathsf{Fl}}_{X,1}(S) \times \underline {\mathsf{Fl}}_{X,1}(S)$ satisfying the following good position (or transversality) condition: $D_1 \cap D_2$ is a relative effective Cartier divisor inside both $D_1 /S$ and $D_2 / S$ , and $Z_1 \cap Z_2 = D_1 \cap D_2$ (Definition 4.9). Note that, if $X$ is a curve, then $\underline {\mathsf{Fl}}_{X,2}$ classifies pairs of disjoint Cartier divisors. Finally, we build our full simplicial object $\underline {\mathsf{Fl}}_{X,\bullet }$ as the classifying simplicial object associated with the partially defined unital monoid structure $\cup$ on $\underline {\mathsf{Fl}}_{X,1}$ (see § 4.4). Note that $\operatorname {\underline {\mathsf{Fl}}}_{X, 0}= \left \{ * \right \}$ , $\partial _0, \partial _2: \operatorname {\underline {\mathsf{Fl}}}_{X,2} \to \operatorname {\underline {\mathsf{Fl}}}_{X,1}$ are, respectively, the projection on the second and first flag, while $\partial _1: \operatorname {\underline {\mathsf{Fl}}}_{X,2} \to \operatorname {\underline {\mathsf{Fl}}}_{X,1}$ is the union map. The main result of § 4 (Theorem 4.21) is that the simplicial object $\underline {\mathsf{Fl}}_{X,\bullet }$ has the so-called $2$ -Segal property [Reference Dyckerhoff and KapranovDK19], i.e. the diagram

defines an associative algebra structure in correspondences. This will induce, later in the paper, the same property on the simplicial flag Grassmannian $\mathcal{G}r_{X, \bullet }$ .

Flag Grassmannian. Fix, once and for all, a smooth affine group scheme ${\bf G}$ over $\mathbb{C}$ , that will be most often omitted from the notation. In §§ 5 and 6 we build the simplicial flag Grassmannian $\mathcal{G}r_{X,\bullet }$ for ${\bf G}$ -bundles, together with a simplicial map $\mathcal{G}r_{X,\bullet } \to \underline {\mathsf{Fl}}_{X,\bullet }$ . We sketch here, in a slightly simplified way, the basic ideas of the constructions, where furthermore the schemes parametrizing families of flags will be underived. The actual rigorous constructions in the main body of the paper follow a slightly different and more technical path (mainly in order to take care of higher coherences and derived structures), and giving here all such details would obscure the intuition. Moreover, in this Introduction, we will focus on the flag Grassmannian for ${\bf G}$ -bundles; variations involving perfect complexes are considered in the main text.

For any affine (underived) Noetherian test scheme $S$ , and any flag $(D,Z)\in \underline {\mathsf{Fl}}_{X,1}(S)$ , so that $(D,Z)$ is a flag inside $X \times S$ , we consider (Definition 5.7) the derived stack over $X\times S$ (notation being the one introduced in the previous paragraph)

(1.5) \begin{align} {\bf Gr}_X(S)(D,Z) := {\bf Bun}_{X \times S} \times _{{\bf Bun}_{(X \times S) \smallsetminus D} \times _{{\bf Bun}_{(X \times S)\smallsetminus D \times _{X\times S} {\widehat {Z}}^{{\rm aff}}}} {\bf Bun}_{{\widehat {Z}}}} \left \{ * \right \} \,\,\, \in \, \mathsf{dSt}_{X \times S}, \end{align}

where the map from $\left \{ * \right \}$ represents the trivial ${\bf G}$ -bundle $\mathcal{E}_0$ . This derived stack over $X \times S$ is called the flag Grassmannian of $X\times S \,/ S$ at the fixed flag $(D,Z)$ , and it classifies triples $(\mathcal{E}, \varphi , \psi )$ where $\mathcal{E}$ is a family of ${\bf G}$ -bundles on $X \times S$ , $\varphi$ a family of trivializations of $\mathcal{E}$ on $(X\times S)\smallsetminus D$ and $\psi$ a family of trivializations of $\mathcal{E}$ on $\widehat {Z}$ , such that $\varphi$ and $\psi$ are compatible on the ‘intersection’ $(X \times S)\smallsetminus D \times _{X\times S} {\widehat {Z}}^{{\rm aff}}$ . Like in the classical case where $X$ is a curve, there is also a corresponding local version (Definition 5.8)

(1.6) \begin{align} {\bf Gr}^{{\rm loc}}_X(S)(D,Z) := {\bf Bun}_{\widehat {D}} \times _{{\bf Bun}_{\widehat {D} \smallsetminus D} \times _{{\bf Bun}_{\widehat {D}\smallsetminus D \times _{\widehat {D}} {\widehat {Z}}^{{\rm aff}}}} {\bf Bun}_{{\widehat {Z}}}} \left \{ * \right \} \,\,\, \in \, \mathsf{dSt}_{X \times S}, \end{align}

and formal gluing implies that the obvious restriction

\begin{align*}{\bf Gr}_X(S)(D,Z) \longrightarrow {\bf Gr}^{{\rm loc}}_X(S)(D,Z)\end{align*}

is an equivalence of derived stacks over $X \times S$ (Lemma 5.9). Translating these constructions into absolute derived stacks (i.e. stacks over $k=\mathbb{C}$ ) by using the functor (1.3) defined above, we get derived stacks $\underline {\mathsf{Gr}}_X(S)(D,Z)$ and $\underline {\mathsf{Gr}}^{{\rm loc}}_X(S)(D,Z)$ over $k$ , together with a restriction equivalence $\underline {\mathsf{Gr}}_X(S)(D,Z) \simeq \underline {\mathsf{Gr}}^{{\rm loc}}_X(S)(D,Z)$ . Since $D$ is a relative effective Cartier divisor in $X \times S \, / S$ , one can prove that the truncation ${\rm t}_0(\underline {\mathsf{Gr}}_X(S)(D,Z))$ , hence ${\rm t}_0(\underline {\mathsf{Gr}}^{{\rm loc}}_X(S)(D,Z))$ , is actually a sheaf of sets (Lemma 5.12). In particular the global sections ${\rm Gr}_X(S)(D,Z)$ of both $\underline {\mathsf{Gr}}_X(S)(D,Z)$ and $\underline {\mathsf{Gr}}^{{\rm loc}}_X(S)(D,Z)$ , form a set. Since $(X\times S)\smallsetminus D$ , $\widehat {D}$ , $\widehat {D}^{{\rm aff}}$ , $\widehat {Z}$ , ${\widehat {Z}}^{{\rm aff}}$ and $\widehat {D}\smallsetminus D$ are all insensitive to the scheme structure on $D$ and $Z$ , ${\bf Gr}_X(S)(D,Z)$ , $ {\bf Gr}^{{\rm loc}}_X(S)(D,Z)$ , $\underline {\mathsf{Gr}}_X(S)(D,Z)$ and $\underline {\mathsf{Gr}}^{{\rm loc}}_X(S)(D,Z)$ are topologically invariant $\,$ 5.4). With the test scheme $S$ fixed, by using again formal gluing, one can prove that the assignment $(D,Z) \mapsto {\rm Gr}_X(S)(D,Z)$ defines, for each test scheme $S$ , a functor ${\rm Gr}_X(S): \operatorname {\underline {\mathsf{Fl}}}_{X}(S) \to {\bf Sets},$ and that the assignment $S\mapsto {\rm Gr}_X(S)$ is suitably functorial in the argument $S$ (see § $\,$ 5.3).

The Grothendieck construction applied to ${\rm Gr}_X(S): \operatorname {\underline {\mathsf{Fl}}}_{X}(S) \to {\bf Sets}$ defines $p_1(S): \mathcal{G}r_X(S) \to \operatorname {\underline {\mathsf{Fl}}}_X(S)$ , and the fuctoriality in $S$ of ${\rm Gr}_X(S)$ (established in § $\,$ 5.3), builds for us $\mathcal{G}r_X$ as a functor $\mathsf{dAff}^{{\rm op}} \to {\rm Cat}_\infty$ , together with a morphism of functorsFootnote 4

(1.7) \begin{align} p_1: \mathcal{G}r_X \longrightarrow \operatorname {\underline {\mathsf{Fl}}}_X. \end{align}

In other words, when $S$ is an underived test affine scheme, the objects of $\mathcal{G}r_X(S)$ consist of pairs $(F, \underline {\mathcal{E}})$ where $F$ is a flag in $X \times S$ , and $\underline {\mathcal{E}} \in {\rm Gr}_X(S)$ . Note that the trivial bundle induces an obvious section $\mathsf{triv}: \operatorname {\underline {\mathsf{Fl}}}_X \to \mathcal{G}r_X$ of $p_1$ .

One can show that the map $p_1: \mathcal{G}r_X \to \operatorname {\underline {\mathsf{Fl}}}_X$ carries a canonical flat connection $\,$ 5.3). This is conveniently and compactly rephrased by asserting that the map $p_1: \mathcal{G}r_X \to \operatorname {\underline {\mathsf{Fl}}}_X$ is actually the pullback along the canonical map $\operatorname {\underline {\mathsf{Fl}}}_X \to \operatorname {\underline {\mathsf{Fl}}}^{DR}_X$ of a suitable derived stack $\mathcal{G}r'_X$ mapping to the de Rham stack $\operatorname {\underline {\mathsf{Fl}}}^{DR}_X$ of $\operatorname {\underline {\mathsf{Fl}}}_X$ . This should be seen as the analogue for surfaces of the existence of a canonical flat connection on Beilinson–Drinfeld Grassmannian (see [Reference ZhuZhu17, Proposition 3.1.20]). A further important fact (Theorem 6.20) is that $\mathcal{G}r_X \to \operatorname {\underline {\mathsf{Fl}}}^{DR}_X$ (and also $p_1: \mathcal{G}r_X \to \operatorname {\underline {\mathsf{Fl}}}_X$ ) is, in an appropriate sense, represented by a derived ind-scheme (not of finite typeFootnote 5 ).

We are now able to construct the simplicial flag Grassmannian $\mathcal{G}r_{X, \bullet }$ together with a simplicial map $p_{\bullet }: \mathcal{G}r_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }$ (see § $\,$ 6.4). First of all we put $\mathcal{G}r_{X, 0}= \left \{ * \right \}$ , $\mathcal{G}r_{X, 1}=\mathcal{G}r_X$ , and define $\mathcal{G}r_{X, 2}$ via the following pullback diagram.

(1.8)

A non-trivial result (see § 6.1), obtained through formal gluing inside punctured formal neighbourhoods, allows us to defineFootnote 6 projection maps

(1.9) \begin{align} q_i (S; F_1, F_2): {\rm Gr}_X(S)(F_1 \cup F_2) \longrightarrow {\rm Gr}_X(S)(F_i), \quad i=1,2, \end{align}

for arbitrary flags $(F_1, F_2) \in \operatorname {\underline {\mathsf{Fl}}}_{X,2}(S)$ that are suitably functorial both with respect to maps $(F_1, F_2) \to (F'_1, F'_2)$ in $\operatorname {\underline {\mathsf{Fl}}}_{X,2}(S)$ , and with respect to maps $S' \to S$ between test schemes. Thus, the projection maps of (1.9) induce morphisms $\partial _0, \partial _2: \mathcal{G}r_{X,2} \to \mathcal{G}r_{X,1}$ such that the diagram

commutes, for $j=0, 2$ . Using the empty flag to build degeneracy maps, we get a 2-truncated simplicial object $\mathcal{G}r_{X,\leq 2}$ . Now, via a general extension result (§ Appendix A), we complete $\mathcal{G}r_{X,\leq 2}$ to a full simplicial object $\mathcal{G}r_{X,\bullet }$ , called the simplicial flag Grassmannian of the pair $(X, {\bf G})$ , equipped with a canonical simplicial map $p_{\bullet }: \mathcal{G}r_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }$ , extending $p_1$ and $p_2$ . The $2$ -Segal property of $\operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }$ , and the properties of $p_{\bullet }$ , imply that the simplicial flag Grassmannian $\mathcal{G}r_{X, \bullet }$ is again $2$ -Segal. One can furthermore prove that the simplicial map $p_{\bullet }$ is endowed with a canonical flat connection (extending the one already established on $p_1$ ), or equivalently, that the previous construction yields in fact a map of simplicial objects $\mathcal{G}r'_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}^{DR}_{X, \bullet }$ inducing $p_{\bullet }: \mathcal{G}r_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }$ via pullback along the canonical map $\operatorname {\underline {\mathsf{Fl}}}_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}^{DR}_{X, \bullet }$ .

Flag-factorization. In § 6.2 we prove the flag-factorization property for the flag Grassmannian $\mathcal{G}r_X$ . More precisely, and in complete analogy with (1.1), we prove (see Remark 6.17) that the map $(\partial _0, \partial _2): \mathcal{G}r_{X,2} \to \mathcal{G}r_{X, 1} \times \mathcal{G}r_{X,1}$ establishes an isomorphismFootnote 7 (called the factorization isomorphism) between the pullbacks of the following diagrams.

(1.10)

The proof boils down to showing that the projection maps (1.9) induce an isomorphism

(1.11)

for any pair $(F_1,F_2) \in \operatorname {\underline {\mathsf{Fl}}}_{X, 2}(S)$ , and for any $S$ (Theorem 6.2). One important and non-trivial ingredient of the proof (Lemma 6.7) consists in the equivalence

for arbitrary (derived) closed subschemes $T_1, T_2$ of an arbitrary affine (derived) scheme.

Flag-dependent exotic structures on the stack of ${\bf G}$ -bundles. For a (underived) Noetherian scheme $S$ and a smooth morphism $Y \to S$ of relative dimension $\geq 2$ , we say that a flag $(D,Z)$ in $Y/S$ is large if $Z_s$ meets any irreducible component of $D_s$ , for any $s \in S$ . As A. Beilinson observed (private communication), the set of global sections

\begin{align*}{\rm Gr}(Y; D,Z)= {\rm Bun}^{{\bf G}}(Y) \times _{{\rm Bun}^{{\rm G}}(Y\setminus D) \times _{{\rm Bun}^{{\bf G}}(\widehat {Z} \setminus D)} {\rm Bun}^{{\bf G}}(\widehat {Z}) } \left \{ \overline {\mathcal{E}} \right \}\end{align*}

is a singleton,Footnote 8 for any fixed ${\bf G}$ -bundle $\overline {\mathcal{E}}$ on $Y$ (Proposition 7.2). This observation allows us, in § 7, to associate with any large flag on a smooth projective surface $X$ a derived structure on the classical stack of ${\bf G}$ -bundles on $X$ (Proposition 7.6). These derived structures (given by the corresponding derived Hecke stacks, see § 7.2) are, in general, different from the usual one, i.e. $\underline {\mathsf{Bun}}_{X}$ (see Examples 7.7 and 7.8), and therefore we call them exotic derived structures on the truncation ${\rm t}_0(\underline {\mathsf{Bun}}_{X})$ , i.e. on the classical stack of ${\bf G}$ -bundles on $X$ . We remark that different choices of large flags in $X$ yield different exotic structures.

Flag-chiral structures. In § 8 we define flag analogues of chiral algebras and chiral categories. In the classical case, i.e. when $C$ is a smooth projective curve, one can define the so called chiral tensor product on appropriate categories $\mathsf{Shv}(\mathsf{Ran}_C)$ of sheaves on the Ran space of $C$ , essentially by performing the pull–push construction (of the exterior tensor product of such sheaves) along the diagram

(1.12)

where $q$ is the obvious inclusion (see [Reference Francis and GaitsgoryFG12] for details). When $X$ is a smooth projective surface, in order to get a flag analogue of this chiral product, one might naively perform the same construction by replacing (1.12) by the diagram

(1.13)

where recall that $\partial _1$ is the union map of good flags. This can certainly be done if we forget maps between flags (i.e. considering $\operatorname {\underline {\mathsf{Fl}}}_{X}$ and $\operatorname {\underline {\mathsf{Fl}}}_{X,2}$ as sets-valued rather than posets-valued), see § 8.1. In order to deal with the general case, we use the tools developed by S. RaskinFootnote 9 [Reference RaskinRas19], together with the Beilinson–Drinfeld idea in [Reference Beilinson and DrinfeldBD04, 3.4.6], to get a definition of flag-chiral category and of quasi-coherent flag-factorization algebra, together with their crystal and $\mathcal{D}$ -Modules counterparts (Definition 8.2). Some examples of these structures are given.

Why this flag Grassmannian? There are various reasons for considering our flag Grassmannian $\mathcal{G}r_X \to \underline {\mathsf{Fl}}_X$ , for $X$ a smooth projective algebraic surface, as a reasonable generalization of the Beilinson–Drinfeld Ran Grassmannian $\underline {{\rm Gr}}^{{\rm BD}}_{C,{\bf G}} \to \mathsf{Ran}_C$ , defined for a smooth projective algebraic curve $C$ . First of all, if we take $X$ to be a curve, and we only keep the divisorial part $D$ of the flag (the lowest-dimensional part $Z$ being then pushed into dimension $-1$ , so reasonably considered to be empty), then ${\bf Gr}_X(S)(D,Z=\emptyset )$ , out of which $\mathcal{G}r_X$ is constructed by varying the flags, reduces to the usual affine Grassmannian of the curve with respect to the divisor $D$ . Secondly, and most importantly, as we have tried to show above in this Introduction, our flag Grassmannian $\mathcal{G}r_X \to \underline {\mathsf{Fl}}_X$ enjoys most of the nice properties ofthe Beilinson–Drinfeld Ran Grassmannian. Namely, it has (a flags’ version of) the factorization property, it carries a natural flat connection and it naturally accommodates (flags’ versions of) chiral algebra structures. It is also worth mentioning that our general construction can be performed by replacing the stack ${\bf Bun}^{{\bf G}}$ by the stack ${\bf Perf}$ (see Remark 5.6 and 5.11).

On the other hand, as already observed, for ‘large’ flags on the surface (Definition 7.1), the corresponding flag Grassmannian is a purely derived stackFootnote 10 (Proposition 7.2). This is a consequence of imposing compatibility between the two trivializations in formula (1.5), expressed by the lowest term ${\bf Bun}_{(X \times S)\smallsetminus D \times _{X\times S} {\widehat {Z}}^{{\rm aff}}}$ in the fibre product. Although this property has the effect of producing interesting exotic derived structures on the stack of ${\bf G}$ -bundles on the surface (as explained earlier in this Introduction), it may well be considered as a weakness of our construction. We believe that moving from a curve to a higher-dimensional variety, and assuming that in such higher-dimensional cases flags of subvarieties should enter in the construction, naturally forces the derived geometrical part of any corresponding flag version of the affine Grassmannian to be predominant in some sense. This is basically due to the existence of Hartogs-like phenomena, like the example of ${\bf G}$ -bundles on $\mathbb{A}^2$ and on $\mathbb{A}^2 \setminus 0$ shows (the corresponding derived stacks are different while their truncations agree, see Remark 7.4). Nonetheless, there are variations on our definition of the flag Grassmannian that behave better from this specific point of view. Some of them are discussed in the next paragraph, and they form the object of a forthcoming paper [Reference Hennion, Melani and VezzosiHMV].

Leftovers and future directions. We list here some possible further lines of research naturally suggested by this paper.

(i) Actions. In Appendix B we sketch the construction of flag versions of the loop and positive loop group (both fibred over $\operatorname {\underline {\mathsf{Fl}}}_X$ ), well known in the curve case, together with their actions on the flag Grassmannian. This construction should be rigorously enhanced to the derived setting, and one should then be able to make sense, at least for ${\bf Shv}$ denoting $\ell$ -adic sheaves or $\mathcal{D}$ -Modules, of the equivariant $\infty$ -category ${\bf Shv}^{\mathcal{L}^{+}G_X}(\mathcal{G}r_{X})$ of $\mathcal{L}^{+}G_X$ -equivariant sheaves on $\mathcal{G}r_{X}$ , endowed with an appropriate convolution monoidal structure. This should also be extended to the other flag Grassmannians introduced in (ii) and (iii) below.

(ii) Mapping down and other flag Grassmannians. Fix a smooth divisor $D_0$ in a smooth projective surface $X$ . For each test scheme $S$ , and any flag on $X\times S /S$ of the form $(D_0 \times S, Z)$ , we may consider

\begin{align*}\underline {\mathsf{Gr}}_{D_0}(S;Z) := \underline {\mathsf{Bun}}_{D_0 \times S} \times _{\underline {\mathsf{Bun}}_{(D_0 \times S)\setminus Z}} \{ {\rm triv} \}\end{align*}

i.e. the affine Grassmannian of the curve $D_0$ at the ‘points’ $Z$ . We may try to construct a map $\underline {\mathsf{Gr}}(X\times S)(D_0 \times S, Z) \to \underline {\mathsf{Gr}}_{D_0}(S;Z)$ , and varying both $S$ and $Z \subset D_0 \times S$ , a map

(1.14) \begin{align} \mathcal{G}r_{X;D_0} \longrightarrow \mathcal{G}r_{D_0} ,\end{align}

where $\mathcal{G}r_{D_0}$ is (a Cartier-divisors version of) the Beilinson–Drinfeld Ran Grassmannian of the curve $D_0 \subset X$ . A drawback of the definition of flag Grassmannian in this paper is that this map is necessarily trivial (i.e. factors through the trivial bundle). This is essentially due to the compatibility required between the two trivializations in (1.5). Thus, in order to get a non-trivial map, one may modify our definition of the flag Grassmannian by replacing (1.5) with either of the following:Footnote 11

\begin{align*}\underline {\mathsf{Gr}}_X^{{\rm glob}}(S)(D,Z):= \underline {\mathsf{Bun}}_{X \times S} \times _{\underline {\mathsf{Bun}}_{\widehat {D}\setminus Z} \times \underline {\mathsf{Bun}}_{\widehat {Z}}} \{ {\rm triv} \},\quad \underline {\mathsf{Gr}}^{\rm glob,\ res}_{X}(S)(D,Z):= \underline {\mathsf{Bun}}_{X \times S} \times _{\underline {\mathsf{Bun}}_{\widehat {D}\setminus Z}} \{ {\rm triv} \},\end{align*}
\begin{align*}\underline {\mathsf{Gr}}_X^{{\rm fat}}(S)(D,Z):= \underline {\mathsf{Bun}}_{\widehat {D}} \times _{\underline {\mathsf{Bun}}_{\widehat {D}\setminus Z} \times \underline {\mathsf{Bun}}_{\widehat {Z}}} \{ {\rm triv} \},\qquad \underline {\mathsf{Gr}}^{\rm fat,\ res}_{X}(S)(D,Z):= \underline {\mathsf{Bun}}_{\widehat {D}} \times _{\underline {\mathsf{Bun}}_{\widehat {D}\setminus Z}} \{ {\rm triv} \}.\end{align*}

As done with the flag Grassmannian of this paper, one can show [Reference Hennion, Melani and VezzosiHMV] that there are corresponding $2$ -Segal simplicial objects $\mathcal{G}r^{{\rm glob}}_{X, \bullet }$ , $\mathcal{G}r^{\rm glob,\ res}_{X, \bullet }$ in $\mathsf{dSt}^{{\rm Cat}_{\infty }}/\underline {\mathsf{Bun}}_{X}$ (with $\mathcal{G}r^{{\rm glob}}_{X, 0}= \mathcal{G}r^{\rm glob,\ res}_{X, 0} := \underline {\mathsf{Bun}}_{X}$ ),Footnote 12 and $2$ -Segal objects $\mathcal{G}r^{{\rm fat}}_{X, \bullet }$ , $\mathcal{G}r^{\rm fat,\ res}_{X, \bullet }$ in $\mathsf{dSt}^{{\rm Cat}_{\infty }}$ , together with simplicial maps $\mathcal{G}r^{{\rm glob}}_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}^{DR, {\rm op}}_{X, \bullet } \times \underline {\mathsf{Bun}}_{X}$ , $\mathcal{G}r^{\rm glob,\ res}_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}^{DR, {\rm op}}_{X, \bullet } \times \underline {\mathsf{Bun}}_{X}$ , $\mathcal{G}r^{{\rm fat}}_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}^{DR, {\rm op}}_{X, \bullet }$ and $\mathcal{G}r^{\rm fat,\ res}_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}^{DR, {\rm op}}_{X, \bullet }$ . Moreover, one can show that $\mathcal{G}r^{{\rm glob}}_{X, \bullet }$ , $\mathcal{G}r^{\rm glob,\ res}_{X, \bullet }$ , $\mathcal{G}r^{{\rm fat}}_{X, \bullet }$ and $\mathcal{G}r^{\rm fat,\ res}_{X, \bullet }$ all satisfy an appropriate flag-factorization property.Footnote 13 For this modified flag Grassmannians, the corresponding maps (1.14) are non-trivial. This also suggests the possibility that suitably defined Hecke operators on $X$ induce Hecke operators on the curve $D_0$ , perhaps generating interesting subalgebras of Hecke operators on $D_0$ . Investigations in this directions are ongoing [Reference Hennion, Melani and VezzosiHMV].

(iii) More local flag Grassmannians and the double affine Grassmannian. By considering even more local Grassmannians with respect to those considered above in (ii), one can try to give an algebro-geometrical interpretation of the (usual) affine Grassmannian for the loop group ${\rm L}{\bf G}$ , as discussedFootnote 14 in [Reference Braverman and FinkelbergBF10, § 3]. Indeed one can run the flag Grassmannian construction of this paper by replacing (1.5) with

\begin{align*}\underline {\mathsf{Gr}}_X^{{\rm double}}(S)(D,Z)=\underline {\mathsf{Bun}}_{(\widehat {Z}^{\widehat {D}})^{{\rm aff}} \setminus D} \times _{\underline {\mathsf{Bun}}_{(\widehat {Z}^{\rm aff,\ \widehat{D}}\setminus Z)^{{\rm aff}} \setminus D}} \{ {\rm triv} \},\end{align*}

where notation need some explanation. First of all, $(\widehat {Z}^{\widehat {D}})^{{\rm aff}}$ is the affinization of the ind-scheme $\widehat {Z}^{\widehat {D}}= \{ \widehat {Z}^{D_n}\}_{n\geq 0}$ , where $\widehat {Z}^{D_n}$ is the formal completion of $Z$ inside the $n$ th formal neighbourhood $D_n$ of $D$ inside the ambient scheme $X \times S$ . Secondly, $\widehat {Z}^{\rm aff,\ \widehat{D}}$ is the ind-scheme $\{ \widehat {Z}^{{\rm aff}, D_n}\}_{n\geq 0}$ (where $\widehat {Z}^{{\rm aff}, D_n}$ is the affinization of the completion of $Z$ inside $D_n$ ); then we subtract $Z$ from this ind-scheme, and we affinize it before finally subtracting $D$ : this is the meaningFootnote 15 of $(\widehat {Z}^{\rm aff,\ \widehat{D}}\setminus Z)^{{\rm aff}} \setminus D$ . Note that, with this definition, for $S= {\rm Spec}\, k$ , $X= \mathbb{A}_k^2$ , $D=\{ y=0 \}$ and $Z=\{ x=y=0 \}$ , we get $(\widehat {Z}^{\widehat {D}})^{{\rm aff}} \setminus D = {\rm Spec} (k[[x]]((y)))$ , and $(\widehat {Z}^{\rm aff,\ \widehat{D}}\setminus Z)^{{\rm aff}} \setminus D = {\rm Spec} (k((x))((y)))$ , and this explains the relation with [Reference Braverman and FinkelbergBF10, § 3]. Further relations between [Reference Braverman and FinkelbergBF10, Reference Braverman and FinkelbergBF12, Reference Braverman and KazhdanBK11] and this very local version of the flag Grassmannian are being investigated [Reference Maffei, Melani and VezzosiMMV25].

(iv) Replacing flags. One could run the construction of this paper (also for the flag Grassmannians listed in (ii) and (iii) above) by replacing $\operatorname {\underline {\mathsf{Fl}}}_{X,\bullet }$ with either the Ran space of nested Cartier divisors in $X$ , the Ran space of $\operatorname {\underline {\mathsf{Fl}}}_X$ , or even the Ran space of families of nested closed subsets of $X$ . This approach might also indicate how to extend our constructions to smooth projective varieties $X$ of dimension $\gt 2$ . This is currently being investigated in [Reference Hennion, Melani and VezzosiHMV].

(v) Open questions. Finally, here are a few questions (inspired by [Reference GaitsgoryGai13] and [Reference Mirkovic and VilonenMV07]) we are considering but have, so far, few clues about.

Let $\mathcal{G}r_X^{{\rm any}}$ be either $\mathcal{G}r_X$ or $\mathcal{G}r^{{\rm glob}}$ or $\mathcal{Gr}^{\rm glob,\ res}$ , and consider the obvious map $\mathcal{G}r^{{\rm any}}_X \to \underline {\mathsf{Bun}}_X$ . Are the fibres of this map de Rham-contractible? If not, does this map induce at least an interesting relation between the de Rham cohomology of $\mathcal{G}r^{{\rm any}}_X$ and the de Rham cohomology of $\underline {\mathsf{Bun}}_X\, ?$

Using (i) above, given any flag Grassmannian $\mathcal{G}r_X$ (i.e. the one in the paper or the ones listed in (ii) and (iii) above), can we interpret in a Tannakian-like way the categoryFootnote 16 of $\mathcal{L}^{+}G_X$ -equivariant perverse sheaves on $\mathcal{G}r_X$ ? To start with, one might fix a flag, and pose the same question for the flag Grassmannian at this fixed flag.

Related works. It is very much unclear what a topic named ‘Geometric Langlands program for surfaces’ should precisely mean or whether it exists at all, at least in the broad structured sense it has in the case of curves. With no claim of exhaustivity, we will merely list, among the ones we are aware of, those sources that actually inspired and motivated our research project. The paper [Reference Ginzburg, Kapranov and VasserotGKV95] was an inspiring and pioneering work in this direction but, at least to our knowledge, its lines of research have not been pursued since. The article [Reference Kapranov, Gelfand, Lepowsky and WilsonKap95] also contains very interesting suggestions about the possible existence of a Langlands program for higher-dimensional varieties, although here the emphasis is more on higher local fields and formal analogies with topological quantum field theories. As we already mentioned, A. Beilinson and V. Drinfeld’s [Reference Beilinson and DrinfeldBD04] (especially § 3.4.6), although dealing with the curve case, was the actual starting point of our project.

Another question about a possible Geometric Langlands program for surfaces, passed on to the third author by M. Kapranov, is whether it should concern ${\bf G}$ -bundles or ${\bf G}$ -gerbes. In any case, the set-up of formal gluing, nonlinear flags and the flag Grassmannian investigated in our paper for ${\bf G}$ -bundles makes sense, mutatis mutandis, for ${\bf G}$ -gerbes, too. However, algebraization is likely to be false for ${\bf G}$ -gerbes,Footnote 17 and this is a problem in developing this direction of research.

While finishing the paper, S. Raskin informed us that, in a 2010 conversation with him, I. Mirkovic proposed to use flags of the sort we consider here, in order to extend Belinson-Drinfeld’s Grassmannian to higher-dimensional varieties. We are not aware of the specific definition of the flag Grassmannian he proposed nor of subsequent publicly available work along these lines.

On the other hand, all of the interesting literature on the double affine Grassmannian (e.g. [Reference Braverman and FinkelbergBF10, Reference Braverman and FinkelbergBF12, Reference Braverman and KazhdanBK11] to list a few papers) seems to go in a different but related direction with respect to our paper. An ongoing collaboration between the second and third authors and Andrea Maffei (Pisa) is aiming at establishing a link between the two approaches [Reference Maffei, Melani and VezzosiMMV25].

Notation. Our base field will be $k=\mathbb{C}$ . All our schemes will, unless otherwise stated, be $k$ -schemes. For a scheme $T$ , ${\bf Sh}_T$ will denote the category of flat sheaves of sets on $T$ . Analogously, $\mathsf{dSt}_T$ will denote the $(\infty ,1)$ -category of flat derived stacks in $\infty$ -groupoids on $T$ . We write ${\bf Sh}$ and $\mathsf{dSt}$ when $T={\rm Spec} \, k$ . We will denote by ${\bf PoSets}$ the $1$ -category of (small) posetal categories, i.e. (small) categories whose $\textrm {Hom}$ -sets are either empty or a singleton. In a 1-category $C$ , if $: u: Z\to X$ , and $u': Z'\to X$ are monos, and there exists an isomorphism $u \simeq u'$ in $C/X$ , then this isomorphism is unique, and we will sometimes simply write $u=u'$ . The other relevant notation will be introduced and explained in the main text.

2. Sheaves and bundles on the punctured formal neighbourhood

In this first section, we define and construct the stacks of bundles (or (pseudo-)perfect complexes) on punctured formal neighbourhoods in a very general setting. This will allow us to state and prove, in the next section, a statement à la Beauville and Laszlo for general (derived) stacks.

2.1 Recollection on derived stacks

We start by fixing a handful of notation from derived algebraic geometry. For a comprehensive review of the subject, we refer the reader to [Reference ToënToë14].

Definition 2.1. Let us recall a few definitions, and fix our notation.

  • We will denote by ${\rm CAlg}_k$ the usual $1$ -category of commutative $k$ -algebras. In the derived geometry literature, such algebras are sometimes called discrete, underived or $0$ -truncated.

  • We will denote by ${\bf sCAlg}_k$ the $\infty$ -category of simplicial commutative $k$ -algebras up to homotopy equivalence.

  • For any $A \in {\bf sCAlg}_k$ , we will denote by $\pi _i A$ its $i^{{\rm th}}$ homotopy group. Recall that $\pi _0 A$ is endowed with a commutative $k$ -algebra structure, and that for any $i$ , the abelian group $\pi _i A$ is endowed with a structure of $\pi _0 A$ -module.

  • A simplicial commutative algebra $A$ is Noetherian if $\pi _0 A$ is Noetherian and $\pi _i A$ is of finite type over $\pi _0 A$ . We denote by ${\bf sCAlg}^{{\rm Noeth}}_k$ the $\infty$ -category of Noetherian simplicial commutative algebras.

  • A (locally Noetherian) derived prestack is a functor ${\bf sCAlg}^{{\rm Noeth}}_k \to {\rm Gpd}_\infty$ with values in the category of $\infty$ -groupoids. A derived stack is a derived prestack satisfying the étale hyperdescent condition. We will denote by $\mathsf{dSt}_k$ the $\infty$ -category of derived stacks. For $X \in \mathsf{dSt}_k$ , we denote by $\mathsf{dSt}_X := \mathsf{dSt}_k/X$ the category of stacks over $X$ .

  • We denote by $\mathsf{dAff}_k \subset \mathsf{dSt}_k$ the full subcategory of (Noetherian) derived affine schemes (i.e. functors represented by some $A \in {\bf sCAlg}^{{\rm Noeth}}_k$ ).

  • For any derived stack $X$ , we will denote by ${\rm t}_0 X$ its restriction to (discrete) commutative $k$ -algebras. The functor ${\rm t}_0 X$ is then a (non-derived) stack. It comes with a canonical morphism ${\rm t}_0 X \to X$ . We say that $X$ is a derived enhancement of ${\rm t}_0 X$ .

  • For $X \to Y$ , $X' \to Y$ morphisms between derived stacks, we will denote by $X \times _Y X'$ their derived fibre product. Sometimes, although not systematically, we will emphasize this by writing $X \times _Y^{{\rm d}} X'$ .

Given a derived affine scheme $X$ and a closed subset $Z$ of the scheme ${\rm t}_0 X$ , the open complement ${\rm t}_0 X \smallsetminus Z$ has a canonical derived enhancement as a derived subscheme of $X$ . We call it the open complement of $Z$ in $X$ and denote it by $X \smallsetminus Z$ .

Definition 2.2. Let $X \in \mathsf{dSt}_k$ .

  • A closed substack $Z \subset X$ is the datum of a non-derived stack $Z$ and a map $Z \to X$ such that for any $U \in \mathsf{Aff}_k$ and any $U \to X$ , the projection ${\rm t}_0(Z \times _{ X} U) \to U$ is a closed immersion of (affine) schemes.

  • Given a closed substack $Z \subset X$ , its open complement is the derived stack

    \begin{align*} X \smallsetminus Z := \mathop {{\rm colim}}\limits _{U \in \mathsf{dAff} /X} U \smallsetminus Z_U, \end{align*}
    where $Z_U$ stands for the pullback $Z \times _X U$ .

Definition 2.3. Let $X$ be a derived stack. We define its de Rham stack $X_{{\rm dR}}$ as the derived prestack whose $A$ -points are

\begin{align*} X_{{\rm dR}}(A) = X((\pi _0 A)_{{\rm red}}). \end{align*}

It is endowed with a canonical morphism $X \to X_{{\rm dR}}$ .

Definition 2.4. Let $Z \to X$ be a map of derived stacks. The (derived) formal completion of $Z$ in $X$ (or of $X$ along $Z$ ) is defined as the pullback

\begin{align*} {\widehat {Z}} = X \mathop{\times}^{{\rm d}}_{X_{{\rm dR}}} Z_{{\rm dR}} \in \mathsf{dSt}_k. \end{align*}

Remark 2.5. The notation $\widehat {Z}$ is ambiguous as the formal completion of $Z$ in $X$ strongly depends on the map $Z\to X$ . On the other hand, the de Rham stack associated with $Z$ only depends on its truncation ${\rm t}_0 Z$ . In particular, the formal completion $\widehat {Z}$ does not depend on the derived structure on $Z$ .

Lemma 2.6. Let $Z \to X \leftarrow Y \in \mathsf{dSt}_k$ and let $Z_Y$ be the fibre product $Z_Y = Z \times ^{{\rm d}}_X Y$ . Let $\widehat {Z}$ be the formal completion of $Z$ in $X$ and ${\widehat {Z}}_Y$ be the formal completion of $Z_Y$ in $Y$ (i.e. the formal completion along the projection map $Z_Y\to Y\!$ ). The canonical morphism

\begin{align*} {\widehat {Z}}_Y \to Y \mathop{\times}^{{\rm d}}_X {\widehat {Z}} \end{align*}

is an equivalence.

Proof. This is immediate, since the functor $X \mapsto X_{{\rm dR}}$ preserves fibre products.

The following statement justifies the above definition of the derived formal completion.

Theorem 2.7. Let $X$ be a derived scheme and $Z \subset X$ a closed substack (subscheme). The following assertions hold.

  1. (i) (Gaitsgory and Rozenblyum) The formal completion $\widehat {Z}$ is a stack and is representable by a derived ind-scheme.

  2. (ii) The truncation ${\rm t}_0({\widehat {Z}})$ is canonically isomorphic to the (usual) formal completion of ${\rm t}_0(X)$ along ${\rm t}_0(Z)$ .

Proof. Assertion (i) is [Reference Gaitsgory and RozenblyumGR14, Proposition 6.3.1]. Let us prove assertion (ii). Using Lemma 2.6, we can assume that $X$ is affine. Let $A$ denote the algebra of functions on ${\rm t}_0 X$ and let $I \subset A$ be an ideal defining ${\rm t}_0 Z$ in ${\rm t}_0 X$ .

The truncation functor ${\rm t}_0$ preserves fibre products. It follows that a $B$ -point of ${\rm t}_0({\widehat {Z}})$ is a commutative diagram

where $\sqrt {B} \subset B$ is the nilradical of $B$ . Since $B$ is Noetherian, it amounts to a map $f \;:\; A \to B$ such that $f(I)$ is nilpotent in $B$ . The functor ${\rm t}_0({\widehat {Z}})$ is therefore represented by the ind-scheme $\textrm {colim Spec}(A/I^p)$ .

2.2 Categories of modules

Definition 2.8. Let $A \in {\bf sCAlg}_k$ be a simplicial commutative algebra. Let ${\rm Perf}(A)$ denote the full stable $\infty$ -subcategory of ${\rm QCoh}(A)$ spanned by its compact objects. Let ${{\rm Coh}}^-(A)$ denote the full stable $\infty$ -subcategory of ${\rm QCoh}(A)$ spanned by almost perfect $A$ -modules [Reference LurieLur17, Definition 7.2.4.10].

Remark 2.9. If $A \in {\rm CAlg}_k$ , then a complex $\mathcal F$ belongs to ${{\rm Coh}}^-(A)$ if and only if:

  1. (i) for every integer $i \in \mathbb Z$ , the cohomology group $\mathrm H^i(\mathcal F)$ is coherent over $A$ ; and

  2. (ii) $\mathrm H^i(\mathcal F)=0$ for $i \gg 0$ .

The assignment $\textrm {Spec}(A) \mapsto {{\rm Coh}}^-(A)$ can be promoted to an $\infty$ -functor

\begin{align*} {\bf Coh}^{\otimes ,-} \;:\; \mathsf{dAff}_k^{{\rm op}} \to {{\rm Cat}}_\infty ^{{\rm Ex},\otimes } , \end{align*}

using the $(-)^*$ functoriality. It comes with a pointwise fully faithful natural transformation ${\bf Coh}^{\otimes ,-} \to {\bf QCoh}^\otimes$ .

Since ${\bf QCoh}^{\otimes }$ is an hypercomplete sheaf for the étale topology, and since being coherent is a local property, the functor ${\bf Coh}^{\otimes ,-}$ is an hypercomplete sheaf for the étale topology. In particular, both ${\bf QCoh}^\otimes$ and ${\bf Coh}^{\otimes ,-}$ extend uniquely into limit-preserving functors

\begin{align*} {\bf QCoh}^{\otimes },\,{\bf Coh}^{\otimes ,-} \;:\; \mathsf{dSt}_k^{{\rm op}} \to {{\rm Cat}}_\infty ^{{\rm Ex},\otimes }. \end{align*}

2.3 Fibre functors and punctured neighbourhoods

The first issue arising in the study of punctured formal neighbourhoods is that whenever $Z$ is a closed subset of a scheme $X$ , the formal scheme ${\widehat {Z}}\smallsetminus Z$ is empty. To circumvent this, we will see that any (formal) scheme is determined by what we call its fibre functor, and that the punctured formal neighbourhood admits a somewhat natural (and non-trivial) fibre functor (which is not representable).

Definition 2.10. We fix $X\in \mathsf{dSt}_k$ . Recall that $\mathsf{dAff}_{/X}$ is the category of Noetherian affine derived $k$ -schemes equipped with a morphism to $X$ . Denote by ${\mathfrak{F}}_X$ the forgetful functor $\mathsf{dAff}_{/X} \to \mathsf{dSt}_k$ mapping $S \to X$ to $S$ . A fibre functor over $X$ is a functor

\begin{align*} {\mathfrak{F}} \;:\; \mathsf{dAff}_{/X} \to \mathsf{dSt}_k ,\end{align*}

together with a natural transformation ${\mathfrak{F}} \to {\mathfrak{F}}_X$ . We denote by $\mathsf{dFibF}_X = {\rm Fun}(\mathsf{dAff}_{/X}, \mathsf{dSt}_k)/{\mathfrak{F}}_X$ the $\infty$ -category of fibre functors over $X$ .

For any stack $Y$ over $X$ , we denote by ${\mathfrak{F}}_Y$ the functor

\begin{align*} {\mathfrak{F}}_Y \;:\; \begin{aligned} \mathsf{dAff}_{/X} & \to \mathsf{dSt}_k \\ U & \mapsto U \mathop{\times}_{X} Y \end{aligned} \end{align*}

This construction being functorial, any $Y \in \mathsf{dSt}_k$ yields a fibre functor ${\mathfrak{F}}_Y \to {\mathfrak{F}}_X$ . We call it the fibre functor of $Y$ (over $X$ ).

Remark 2.11. All in all, a fibre functor associates with every $U \to X$ a stack over $U$ , with only weak compatibilities. They can be thought as ‘non-quasi-coherent stacks’ over $X$ .

Remark also that the terminal fibre functor ${\mathfrak{F}}_X$ maps to the constant functor with value $X$ , so that fibre functors can be seen as having values in $\mathsf{dSt}_X$ .

Example 2.12. Let $X \in \mathsf{dSt}_k$ and $Z \subset X$ a closed substack. The formal completion ${\widehat {Z}} := \widehat X_Z$ (Definition 2.4) yields a fibre functor ${\mathfrak{F}}_{\widehat {Z}}$ . Lemma 2.6 implies that for any $U \in \mathsf{dAff}_{/X}$ , the stack ${\mathfrak{F}}_{\widehat {Z}}(U)$ is canonically equivalent to the completion ${\widehat {Z}}_U$ of $Z_U := Z \times _X U$ inside $U$ .

Lemma 2.13. Assume $X$ is locally Noetherian. The functor ${\mathfrak{F}}_\bullet \;:\; \mathsf{dSt}_X \to \mathsf{dFibF}_X$ mapping $Y \in \mathsf{dSt}_X$ to ${\mathfrak{F}}_Y$ is fully faithful. Its essential image consists of fibre functors ${\mathfrak{F}} \to {\mathfrak{F}}_X$ such that for any $U \to V \in \mathsf{dAff}_{/X}$ , the induced ${\mathfrak{F}}(U) \to {\mathfrak{F}}(V) \times _V U$ is an equivalence.

Proof. The functor ${\mathfrak{F}}_\bullet$ admits a left adjoint computing the colimit of the underlying functor $\mathsf{dAff}_{/X} \to \mathsf{dSt}_k$ . Since colimits in $\mathsf{dSt}_k$ are universal, the adjunction morphism

\begin{align*} \mathop {{\rm colim}}_{U \in \mathsf{dAff}_{/X}} U \mathop{\times} _X Y \to Y ,\end{align*}

is an equivalence. In particular ${\mathfrak{F}}_\bullet$ is fully faithful.

To compute its essential image, we interpret the notion of fibre functors in terms of Grothendieck constructions. Let $t \;:\; \mathsf{dSt}_k^{\Delta^1} := {\rm Fun}(\Delta^1, \mathsf{dSt}_k) \to \mathsf{dSt}_k$ be the target functor (i.e. evaluation at $1$ ). It is a Cartesian fibration classifying the functor $U \mapsto \mathsf{dSt}_U$ sending maps to pullbacks (see [Reference LurieLur09, Corollaries 2.4.7.12 and 5.2.2.5]). A fibre functor $\mathfrak{F}$ over $X$ is tantamount to a section $s_{\mathfrak{F}}$ of $t$ over $\mathsf{dAff}_{/X}$ .

Given a fibre functor $\mathfrak{F}$ , the condition that for any $U \to V \in \mathsf{dAff}_{/X}$ , the map ${\mathfrak{F}}(U) \to {\mathfrak{F}}(V) \times _V U$ is an equivalence is equivalent to the condition that the section $s_{\mathfrak{F}}$ is Cartesian. By [Reference LurieLur09, Corollary 3.3.3.2], the category of Cartesian sections is equivalent to the limit category $\lim _{U \in \mathsf{dAff}_{/X}} \mathsf{dSt}_U$ . Since the functor $U \mapsto \mathsf{dSt}_U$ is a stack (the stack of stacks, see for instance [Reference LurieLur09, Theorem 6.1.3.9]), this limit is equivalent (through the base-change functors) to $\mathsf{dSt}_X$ . It follows that the essential image of ${\mathfrak{F}}_\bullet$ is indeed as announced.

Definition 2.14. A fibre functor $\mathfrak{F}$ is called representable if it belongs to the essential image of ${\mathfrak{F}}_\bullet$ .

Definition 2.15. Let $X \in \mathsf{dSt}_k$ and let $Z \subset X$ be a closed substack. The map ${\mathfrak{F}}_Z \to {\mathfrak{F}}_X$ of fibre functors is a pointwise closed immersion of derived affine schemes. Any morphism $U \to V \in \mathsf{dAff}_{/X}$ maps the open complement of $Z_U = Z \times _X U$ to the open complement of $Z_V = Z \times _X V$ . In particular, we have a fibre functor

\begin{align*} {\mathfrak{F}}_{X \smallsetminus Z} \;:\; U \mapsto U \smallsetminus Z_U. \end{align*}

We easily check using Lemma 2.13 that ${\mathfrak{F}}_{X \smallsetminus Z}$ is representable. We denote by $X \smallsetminus Z$ its representative (so that $X \smallsetminus Z \simeq {\rm colim}_U U \smallsetminus Z_U$ ) and call it the open complement of $Z$ in $X$ .

Definition 2.16. Let $X \in \mathsf{dSt}_k$ and let $Z \subset X$ be a closed substack.

  • The affinized formal neighbourhood of $Z$ in $X$ is the fibre functor ${\mathfrak{F}}_{{\widehat {Z}}}^{{\rm aff}}$ obtained by applying a pointwise affinization to ${\mathfrak{F}}_{{\widehat {Z}}}$ . More explicitly

    \begin{align*} {\mathfrak{F}}_{{\widehat {Z}}}^{{\rm aff}} \;:\; U \mapsto ({\mathfrak{F}}_{{\widehat {Z}}}(U))^{{\rm aff}} := {\rm Spec}(\Gamma({\widehat {Z}}_U, \mathcal O_{{\widehat {Z}}_U})). \end{align*}
    Since ${\mathfrak{F}}_X$ has values in derived affine schemes, the structure map ${\mathfrak{F}}_{\widehat {Z}} \to {\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}}$ factors as ${\mathfrak{F}}_{\widehat {Z}} \to {\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}} \to {\mathfrak{F}}_X$ , giving ${\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}}$ the structure of a fibre functor over $X$ .
  • The punctured formal neighbourhood of $Z$ in $X$ is the fibre functor ${\mathfrak{F}}_{{\widehat {Z}}\smallsetminus Z}$

    \begin{align*} {\mathfrak{F}}_{{\widehat {Z}}\smallsetminus Z} \;:\; U \mapsto {\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}}(U) \mathop{\times} _X (X \smallsetminus Z) \simeq {\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}}(U) \mathop{\times} _U (U \smallsetminus Z_U). \end{align*}

Example 2.17. Assume $A \in {\rm CAlg}_k$ (i.e. $A$ is a usual, discrete commutative $k$ -algebra), $U = {\rm Spec} A$ and $Z_U = Z \times _X U$ is the scheme ${\rm Spec}(A/I)$ , for $I \subset A$ an ideal. We have

\begin{align*} {\mathfrak{F}}_{\widehat {Z}}(U) = Spf({\widehat A}) \hspace {0.3em} {\rm ,} \hspace {1em} {\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}}(U) = {\rm Spec}({\widehat A}) \hspace {1em} {\rm and} \hspace {1em} {\mathfrak{F}}_{{\widehat {Z}}\smallsetminus Z}(U) = {\rm Spec}({\widehat A}) \smallsetminus {\rm Spec}(A/I), \end{align*}

where $\widehat A$ is the $I$ -completion of $A$ .

Remark 2.18. The affinization process forgets all about the $I$ -adic topology that was contained in the formal neighbourhood. For this reason, the fibre functors ${\mathfrak{F}}^{{\rm aff}}_{\widehat {Z}}$ and ${\mathfrak{F}}_{{\widehat {Z}} \smallsetminus Z}$ are (typically) not representable.

2.4 Restriction of fibre functors

Any morphism of locally Noetherian derived stacks $f \;:\; X \to Y$ induces a projection functor $p_f \;:\; \mathsf{dAff}_{/X} \to \mathsf{dAff}_{/Y}$ . Since we have ${\mathfrak{F}}_X = {\mathfrak{F}}_Y \circ p_f$ , this in turn induces a restriction functor $f^{-1} \;:\; \mathsf{dFibF}_Y \to \mathsf{dFibF}_X$ given by

\begin{align*} f^{-1} {\mathfrak{F}} \;:\; (U \to X) \mapsto {\mathfrak{F}}(U \to X \to Y). \end{align*}

This is trivially functorial, leading to a contravariant $\infty$ -functor $X \mapsto \mathsf{dFibF}_X$ . It satisfies the following properties (whose straightforward proofs are left to the reader).

Lemma 2.19. Let $f \;:\; X \to Y$ be a morphism of locally Noetherian derived stacks.

  1. (i) If $T \to Y$ is a derived stack, then $f^{-1}({\mathfrak{F}}_T) \simeq {\mathfrak{F}}_{T \times _Y X}$ .

  2. (ii) The functor $f^{-1}$ preserves small colimits as well as pullbacks.

  3. (iii) If $Z \to Y$ is a closed substack, then

    \begin{align*} f^{-1}({\mathfrak{F}}_{\widehat {Z}}) \simeq {\mathfrak{F}}_{{\widehat {Z}}_X}, \hspace {1em} f^{-1}({\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}}) \simeq {\mathfrak{F}}_{{\widehat {Z}}_X}^{{\rm aff}} \hspace {1em}{\rm and}\hspace {1em} f^{-1}({\mathfrak{F}}_{{\widehat {Z}}\smallsetminus Z}) \simeq {\mathfrak{F}}_{{\widehat {Z}}_X \smallsetminus Z_X}, \end{align*}
    where $Z_X := Z \times _Y X$ .

2.5 Sheaves on punctured formal neighbourhoods

Definition 2.20. Let $X \in \mathsf{dSt}_k$ and let ${\mathfrak{F}} \;:\; \mathsf{dAff}_{/X} \to \mathsf{dSt}_X$ be a fibre functor over $X$ . The stacks of (right bounded) coherent sheaves and perfect complexes of $\mathfrak{F}$ are respectively the functors $ \left ( \mathsf{dAff}_{/X} \right )^{{\rm op}} \to {{\rm Cat}}_\infty ^{{\rm Ex},\otimes }$ given by

\begin{align*} {\bf Coh}^-_{{\mathfrak{F}}} \;:\; U & \mapsto {\bf Coh}^{\otimes ,-}({\mathfrak{F}}(U)),\\ {\bf Perf}_{{\mathfrak{F}}} \;:\; U &\mapsto {\bf Perf}^{\otimes }({\mathfrak{F}}(U)). \end{align*}

When $\mathfrak{F}$ is either ${\mathfrak{F}}_Y$ for $Y \in \mathsf{dSt}_X$ or ${\mathfrak{F}}^{{\rm aff}}_{{\widehat {Z}}\smallsetminus Z}$ for $Z \subset X$ a closed substack, we will simply write

\begin{align*} {\bf Coh}^-_{Y} := {\bf Coh}^-_{{\mathfrak{F}}_Y} \hspace {0.3em},\hspace {1em} {\bf Coh}^-_{{\widehat {Z}}\smallsetminus Z} := {\bf Coh}^-_{{\mathfrak{F}}_{{\widehat {Z}}\smallsetminus Z}} \hspace {0.3em},\hspace {1em} {\bf Perf}_{Y} := {\bf Perf}_{{\mathfrak{F}}_Y} \hspace {1em}{\rm and} \hspace {1em} {\bf Perf}_{{\widehat {Z}}\smallsetminus Z} := {\bf Perf}_{{\mathfrak{F}}_{{\widehat {Z}}\smallsetminus Z}}. \end{align*}

Theorem 2.21. (Mathew [Reference MathewMat22]). Let $X$ be a derived stack and $Z \subset X$ a closed substack. The functors

\begin{align*} {\bf Coh}^-_{{\widehat {Z}} \smallsetminus Z}\hspace {0.7em}{\rm and} \hspace {0.7em} {\bf Perf}_{{\widehat {Z}} \smallsetminus Z} \;:\; \left ( \mathsf{dAff}_{/X} \right )^{{\rm op}} \to {{\rm Cat}}_\infty ^{{\rm Ex},\otimes } \end{align*}

are hypercomplete sheaves with respect to the flat topology.

Proof. The case where $X$ is affine is dealt with in [Reference MathewMat22, Theorem 7.8].Footnote 18 The question being local in $X$ , the general case follows.

2.6 The $\boldsymbol{G}$ -bundles on punctured formal neighbourhoods

In this paragraph, we will extend Theorem 2.21 to the case of ${\bf G}$ -bundles, for ${\bf G}$ an affine group scheme over $k$ .

Definition 2.22. Let $X \in \mathsf{dSt}_k$ and ${\mathfrak{F}} \in \mathsf{dFibF}_X$ . We denote by ${\bf Bun}^{{\bf G}}_{\mathfrak{F}} \;:\; (\mathsf{dAff}_{/X})^{{\rm op}} \to {\rm Gpd}_\infty$ the derived prestack

\begin{align*} {\bf Bun}^{{\bf G}}_{\mathfrak{F}} \;:\; U \mapsto {\rm Bun}^{{\bf G}}({\mathfrak{F}}(U)). \end{align*}

Whenever ${\mathfrak{F}} = {\mathfrak{F}}_Y$ for $Y \in \mathsf{dSt}_X$ or ${\mathfrak{F}} = {\mathfrak{F}}_{{\widehat {Z}} \smallsetminus Z}$ for $Z \subset X$ , we will write

\begin{align*} {\bf Bun}^{{\bf G}}_Y := {\bf Bun}^{{\bf G}}_{{\mathfrak{F}}_Y} \hspace {1em} {\rm and} \hspace {1em} {\bf Bun}^{{\bf G}}_{{\widehat {Z}} \smallsetminus Z} := {\bf Bun}^{{\bf G}}_{{\mathfrak{F}}_{{\widehat {Z}} \smallsetminus Z}}. \end{align*}

Proposition 2.23. For $X \in \mathsf{dSt}_k$ and $Z \subset X$ a closed substack, the prestack

\begin{align*} {\bf Bun}^{{\bf G}}_{{\widehat {Z}} \smallsetminus Z} \;:\; (\mathsf{dAff}_{/X})^{{\rm op}} \to {\rm Gpd}_\infty ,\end{align*}

is a hypercomplete stack for the flat topology.

The next lemma will be used in the proof of Proposition 2.23.

Lemma 2.24. Let $f \;:\; Y_1 \to Y_2$ be a map of derived stacks. We denote by ${\rm Perf} (f) \;:\; {\rm Perf}(Y_2) \to {\rm Perf}(Y_1)$ and ${\rm Bun} (f) \;:\; {\rm Bun}^{{\bf G}}(Y_2) \to {\rm Bun}^{{\bf G}}(Y_1)$ the pullback functors. Then:

  1. (a) if ${\rm Perf} (f)$ is fully faithful, then ${\rm Bun} (f)$ is fully faithful;

  2. (b) if ${\rm Perf} (f)$ is an equivalence, then ${\rm Bun} (f)$ is an equivalence.

Proof. To prove this lemma, we will use Tannaka duality as proven for instance in [Reference LurieLur04, Theorem 3.4.2] (see also [Reference Bhatt and Halpern-LeistnerBHL17, Theorem 4.1]). It gives us a description of the groupoid of ${\bf G}$ -bundles in terms of monoidal functors. Namely, for any stack $Y\!$ , the map

\begin{align*} \alpha _Y \;:\; {\rm Bun}^{{\bf G}}(Y) \simeq {\rm Map}(Y, {\mathrm B} {\bf G}) \to {\rm Fun}^{\otimes }({\rm QCoh}({\mathrm B} {\bf G}),{\rm QCoh}(Y)) ,\end{align*}

is fully faithful. Note also that we have a canonical embedding

\begin{align*} {\rm Fun}^{\otimes }({\rm Perf}({\mathrm B} {\bf G}),{\rm Perf}(Y)) \to {\rm Fun}^{\otimes }({\rm QCoh}({\mathrm B} {\bf G}),{\rm QCoh}(Y)) ,\end{align*}

given by the inclusion ${\rm Perf}(Y) \to {\rm QCoh}(Y)$ and the left Kan extension functor along ${\rm Perf}({\mathrm B} {\bf G}) \to {\rm QCoh}({\mathrm B} {\bf G})$ (recall that ${\bf G}$ being affine, ${\rm QCoh}({\mathrm B} {\bf G})$ is compactly generated by ${\rm Perf}({\mathrm B} {\bf G})$ ). The functor $\alpha _Y$ actually factors through ${\rm Fun}^{\otimes }({\rm Perf}({\mathrm B} {\bf G}),{\rm Perf}(Y))$ . Let now $f \;:\; Y_1 \to Y_2$ be a morphism. We get a commutative diagram

(2.1)

where $F_{{\rm Perf}}$ denotes composition with the pullback ${\rm Perf} (f)$ , and $F_{{\rm QCoh}}$ composition with the pullback ${\rm QCoh} (f)$ . From what precedes, we get that ${\rm Bun} (f)$ is fully faithful if $F_{{\rm Perf}}$ is. This proves (a).

From now on, we assume ${\rm Perf} (f)$ (hence $F_{{\rm Perf}}$ ) is an equivalence. Recall the description of the essential images of the functors $\alpha _{Y_i}$ given in [Reference LurieLur04, Theorem 3.4.2]. A monoidal functor $\beta \;:\; {\rm Perf}({\mathrm B} {\bf G}) \to {\rm Perf}(Y_1)$ then lies in the essential image of $\alpha _{Y_1}$ if and only if its preimage $F_{{\rm Perf}}^{-1}(\beta )$ lies in the essential image of $\alpha _{Y_2}$ . It follows that ${\rm Bun} (f)$ is an equivalence.

Remark 2.25. In the above lemma, one could replace the classifying stack of ${\bf G}$ -bundles ${\mathrm B} {\bf G}$ by any geometric stack with affine diagonal (so that [Reference LurieLur04] Theorem 3.4.2] applies). Stacks satisfying the conclusion of Lemma 2.24(b) are often called ${\rm Perf}$ -local stacks.

Proof of Proposition 2.23. Given $U \in \mathsf{dAff}_{/X}$ and $U_\bullet \to U$ a hypercovering, we consider the morphism in $\mathsf{dSt}_k$

\begin{align*} f \;:\; \mathop{\rm colim}_{[n] \in \Delta } {\mathfrak{F}}_{{\widehat {Z}} \smallsetminus Z}(U_n) \to {\mathfrak{F}}_{{\widehat {Z}} \smallsetminus Z}(U). \end{align*}

By Theorem 2.21, the corresponding pullback functor ${\rm Perf} (f)$ is an equivalence. Using Lemma 2.24(b), we get that also ${\rm Bun} (f)$ is an equivalence.

3. Formal glueing and flag decomposition

The goal of this section is to prove a very general version of the Beauville–Laszlo theorem [Reference Beauville and LaszloBL95]. The Beauville–Laszlo theorem states that a vector bundle on a curve $C$ amount to the data of a bundle on the complement of a point $x \in C$ , a bundle on the formal neighbourhood of said point $x$ , and some glueing datum on the punctured formal neighbourhood.

3.1 Algebraization

We start with an algebraization result, to explain how sheaves or bundles on a formal neighbourhood are to be considered. The main part of its proof is due to Lurie.

Proposition 3.1 (Lurie). Let $X \in \mathsf{dSt}_k$ and $Z \subset X$ a closed substack. The morphism of fibre functors ${\mathfrak{F}}_{{\widehat {Z}}} \to {\mathfrak{F}}_{{\widehat {Z}}}^{{\rm aff}}$ induces equivalences

\begin{align*} {\bf Coh}^-_{{\mathfrak{F}}_{{\widehat {Z}}}^{{\rm aff}}} & \overset \sim \longrightarrow {\bf Coh}^-_{{\mathfrak{F}}_{\widehat {Z}}} = {\bf Coh}^-_{\widehat {Z}},\\ {\bf Perf}_{{\mathfrak{F}}_{{\widehat {Z}}}^{{\rm aff}}} & \overset \sim \longrightarrow {\bf Perf}_{{\mathfrak{F}}_{\widehat {Z}}} = {\bf Perf}_{\widehat {Z}},\\ \hspace {1em} {\bf Bun}^{{\bf G}}_{{\mathfrak{F}}_{{\widehat {Z}}}^{{\rm aff}}} & \overset \sim \longrightarrow {\bf Bun}^{{\bf G}}_{{\mathfrak{F}}_{\widehat {Z}}} = {\bf Bun}^{{\bf G}}_{\widehat {Z}}. \end{align*}

Proof. The case of

\begin{align*} {\bf Coh}^-_{{\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}}} \longrightarrow {\bf Coh}^-_{\widehat {Z}} \end{align*}

is dealt with in [Reference LurieLur11, Theorem 5.3.2]. This equivalence is symmetric monoidal and therefore induces an equivalence between the dualizable objects, so that we get the equivalence

\begin{align*} {\bf Perf}_{{\mathfrak{F}}_{{\widehat {Z}}}^{{\rm aff}}} \overset \sim \longrightarrow {\bf Perf}_{\widehat {Z}}. \end{align*}

We deal with the last case using Lemma 2.24(b) on the morphism ${\mathfrak{F}}_{\widehat {Z}}(U) \to {\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}}(U)$ for every $U \in \mathsf{dAff}_{/X}$ .

Lemma 3.2. The functors ${\bf Coh}^-_{\widehat {Z}}$ , ${\bf Perf}_{\widehat {Z}}$ and ${\bf Bun}^{{\bf G}}_{\widehat {Z}}$ are hypercomplete sheaves for the flat topology.

Proof. The fibre functor ${\mathfrak{F}}_{\widehat {Z}}$ is representable. In particular, it preserves hypercoverings. The result hence follows from usual descent for coherent or perfect complexes, or ${\bf G}$ -bundles.

3.2 The affine case

We start with the affine case, due to Bhatt in the non-derived setting, and extended by Lurie to derived geometry.

We fix $X \in \mathsf{dSt}_k$ and $Z \subset X$ a closed substack. We have the following diagram of fibre functors.

(3.1)

Proposition 3.3. Let $X \in \mathsf{dSt}_k$ and $Z \subset X$ a closed substack. Let ${\bf G}$ be an affine $k$ -group scheme. The above diagram of fibre functors and Proposition 3.1 induce natural equivalences of derived stacks

\begin{align*} {\bf Perf}_X & \overset \sim \longrightarrow {\bf Perf}_{X \smallsetminus Z} \mathop{\times}_{{\bf Perf}_{{\widehat {Z}} \smallsetminus Z}} {\bf Perf}_{\widehat {Z}}, \\ {\bf Coh}^-_X & \overset \sim \longrightarrow {\bf Coh}^-_{X \smallsetminus Z} \mathop{\times} _{{\bf Coh}^-_{{\widehat {Z}} \smallsetminus Z}} {\bf Coh}^-_{\widehat {Z}} ,\\ {\bf Bun}^{{\bf G}}_X & \overset \sim \longrightarrow {\bf Bun}^{{\bf G}}_{X \smallsetminus Z} \mathop{\times}_{{\bf Bun}^{{\bf G}}_{{\widehat {Z}} \smallsetminus Z}} {\bf Bun}^{{\bf G}}_{\widehat {Z}}. \end{align*}

Proof. This can be checked pointwise. We therefore fix $U \in \mathsf{dAff}_{/X}$ . Using [Reference LurieLur18, Proposition 7.4.2.1], together with Proposition 3.1, imply that the base-change functor

\begin{align*} {\bf Perf}_X(U) \to {\bf Perf}_{X \smallsetminus Z}(U) \mathop{\times}_{{\bf Perf}_{{\widehat {Z}} \smallsetminus Z}(U)} {\bf Perf}_{\widehat {Z}}(U) ,\end{align*}

is an equivalence. Consider now the morphism in $\mathsf{dSt}_k$

\begin{align*} f \;:\; {\mathfrak{F}}_{X \smallsetminus Z}(U) \mathop{\amalg} _{{\mathfrak{F}}_{{\widehat {Z}} \smallsetminus Z}(U)} {\mathfrak{F}}_{{\widehat {Z}}}(U) \to {\mathfrak{F}}_X(U) = U. \end{align*}

By the above, ${\rm Perf}(f)$ is an equivalence. Lemma 2.24(b) shows that so is ${\rm Bun}^{{\bf G}}(f)$ .

In the case of ${\bf Coh}^-$ , it follows from [Reference LurieLur18, Remark 7.4.2.2], using that $U$ is Noetherian.

3.3 The global case

We can now focus on the global statement.

Construction 3.4. Recall that for any presentable $\infty$ -category $\mathcal C$ , any hypercomplete étale sheaf $(\mathsf{dAff}_{/X})^{{\rm op}} \to {\mathcal C}$ extends canonically to a limit-preserving functor $(\mathsf{dSt}_X)^{{\rm op}} \to {\mathcal C}$ . In particular, using Theorem 2.21 and Proposition 2.23 , we can extend ${\bf Coh}^-_{{\widehat {Z}} \smallsetminus Z}$ , ${\bf Perf}_{{\widehat {Z}} \smallsetminus Z}$ and ${\bf Bun}^{{\bf G}}_{{\widehat {Z}} \smallsetminus Z}$ to every locally Noetherian derived stack over $X$ . Note that descent for coherent/perfect complexes and bundles implies that ${\bf Coh}^-_{T}$ , ${\bf Perf}_{T}$ and ${\bf Bun}^{{\bf G}}_{T}$ (for $T$ being either $X$ , $X \smallsetminus Z$ or $\widehat {Z}$ – see Lemma 3.2 ) are also hypercomplete sheaves and can similarly be extended to all of $\mathsf{dSt}_X$ .

In the case where $X$ itself is locally Noetherian, we can now take global sections and get the following global version of Proposition 3.3.

Theorem 3.5. Let $X \in \mathsf{dSt}_k$ and let $Z \subset X$ be a closed substack. We have three equivalences of $\infty$ -categories

\begin{align*} {\rm Perf}(X) & \overset \sim \longrightarrow {\rm Perf}(X \smallsetminus Z) \mathop{\times} _{{\rm Perf}({\widehat {Z}} \smallsetminus Z)} {\rm Perf}({\widehat {Z}}), \\ {{\rm Coh}}^-(X) & \overset \sim \longrightarrow {{\rm Coh}}^-(X \smallsetminus Z) \mathop{\times}_{{{\rm Coh}}^-({\widehat {Z}} \smallsetminus Z)} {{\rm Coh}}^-({\widehat {Z}}), \\ {\rm Bun}^{{\bf G}}(X) & \overset \sim \longrightarrow {\rm Bun}^{{\bf G}}(X \smallsetminus Z) \mathop{\times}_{{\rm Bun}^{{\bf G}}({\widehat {Z}} \smallsetminus Z)} {\rm Bun}^{{\bf G}}({\widehat {Z}}), \end{align*}

where ${\rm Perf}({\widehat {Z}} \smallsetminus Z)$ , ${{\rm Coh}}({\widehat {Z}} \smallsetminus Z)$ and ${\rm Bun}^{{\bf G}}({\widehat {Z}} \smallsetminus Z)$ denote respectively ${\bf Perf}_{{\widehat {Z}} \smallsetminus Z}(X)$ , ${\bf Coh}^-_{{\widehat {Z}} \smallsetminus Z}(X)$ and ${\bf Bun}^{{\bf G}}_{{\widehat {Z}} \smallsetminus Z}(X)$ . The first two equivalences are equivalences of symmetric monoidal $\infty$ -categories.

3.4 Flag decomposition

In Theorem 3.5, the derived stack $X$ can be very general: locally Noetherian schemes are allowed, as well as formal completions of locally Noetherian schemes along closed subschemes. More generally, if $X \in \mathsf{dSt}_k$ and $Z \subset X$ is a closed substack, then the formal completion $\widehat {Z}$ is also locally Noetherian. Indeed, this statement is local in $X$ so we can assume $X$ to be affine and Noetherian. In this case, we can see using [Reference Gaitsgory and RozenblyumGR14, Proposition 6.7.4] that $\widehat {Z}$ is a colimit of Noetherian derived affine schemes and thus do belong to $\mathsf{dSt}_k$ .

In particular, both Proposition 3.3 and Theorem 3.5 can be iterated as follows.

Corollary 3.6. Let $X \in \mathsf{dSt}_k$ , $Z_1 \subset Z_2 \subset X$ be a flag of closed substacks, and ${\bf G}$ be an affine $k$ -group scheme. We have natural equivalences of derived stacks

\begin{align*} {\bf Perf}_X & \overset \sim \longrightarrow {\bf Perf}_{X \smallsetminus Z_2} \mathop{\times}_{{\bf Perf}_{{\widehat {Z}}_2 \smallsetminus Z_2}}\Big({\bf Perf}_{{\widehat {Z}}_2 \smallsetminus Z_{1}}\mathop{\times}_{{\bf Perf}_{{\widehat {Z}}_{1} \smallsetminus Z_{1}}} {\bf Perf}_{{\widehat {Z}}_1} \Big),\\ {\bf Coh}^-_X & \overset \sim \longrightarrow {\bf Coh}^-_{X \smallsetminus Z_2} \mathop{\times}_{{\bf Coh}^-_{{\widehat {Z}}_2 \smallsetminus Z_2}}\Big( {\bf Coh}^-_{{\widehat {Z}}_2 \smallsetminus Z_{1}}\mathop{\times}_{{\bf Coh}^-_{{\widehat {Z}}_{1} \smallsetminus Z_{1}}} {\bf Coh}^-_{{\widehat {Z}}_1} \Big),\\ {\bf Bun}^{{\bf G}}_X & \overset \sim \longrightarrow {\bf Bun}^{{\bf G}}_{X \smallsetminus Z_2} \mathop{\times}_{{\bf Bun}^{{\bf G}}_{{\widehat {Z}}_2 \smallsetminus Z_2}}\Big( {\bf Bun}^{{\bf G}}_{{\widehat {Z}}_2 \smallsetminus Z_{1}}\mathop{\times}_{{\bf Bun}^{{\bf G}}_{{\widehat {Z}}_{1} \smallsetminus Z_{1}}} {\bf Bun}^{{\bf G}}_{{\widehat {Z}}_1} \Big). \end{align*}

If we suppose, moreover, that $X\in \mathsf{dSt}_k$ , taking global sections of the previous equivalences yield equivalences of $\infty$ -categories

\begin{align*} {\rm Perf}(X) & \simeq {\rm Perf}(X \smallsetminus Z_2) \mathop{\times} _{{\rm Perf}({\widehat {Z}}_2 \smallsetminus Z_2)} \Big({\rm Perf}({\widehat {Z}}_2 \smallsetminus Z_{1}) \mathop{\times} _{{\rm Perf}({\widehat {Z}}_{1} \smallsetminus Z_{1})} {\rm Perf} ({\widehat {Z}}_1) \Big),\\ {{\rm Coh}}^-(X) & \simeq {{\rm Coh}}^-(X \smallsetminus Z_2) \mathop{\times} _{{{\rm Coh}}^-({\widehat {Z}}_2 \smallsetminus Z_2)} \Big({{\rm Coh}}^-({\widehat {Z}}_2 \smallsetminus Z_{1}) \mathop{\times} _{{{\rm Coh}}^-({\widehat {Z}}_{1} \smallsetminus Z_{1})} {{\rm Coh}}^-({\widehat {Z}}_1)\Big),\\ {\rm Bun}^{{\bf G}}(X) & \simeq {\rm Bun}^{{\bf G}}(X \smallsetminus Z_2) \mathop{\times} _{{\rm Bun}^{{\bf G}}({\widehat {Z}}_2 \smallsetminus Z_2)} \Big({\rm Bun}^{{\bf G}}({\widehat {Z}}_2 \smallsetminus Z_{1}) \mathop{\times} _{{\rm Bun}^{{\bf G}}({\widehat {Z}}_{1} \smallsetminus Z_{1})} {\rm Bun}^{{\bf G}}({\widehat {Z}}_1)\Big). \end{align*}

The first two equivalences are equivalences of symmetric monoidal $\infty$ -categories.

Similar statements obviously hold for flags $Z_1 \subset \cdots \subset Z_n \subset X$ of arbitrary length; we leave the interested reader to write them down explicitly.

Construction 3.7. Let $\mathcal{C}$ denote either ${{\rm Cat}}_\infty ^{{\rm Ex},\otimes }$ or ${\rm Gpd}_\infty$ , and let $X\in \mathsf{dSt}_k$ . Observe that there is an equivalence between the category of $\mathcal{C}$ -valued hypercomplete sheaves ${\rm St}_{{\rm fl}}(\mathsf{dAff}_{/X}, \mathcal{C})$ for the flat topology on $\mathsf{dAff}_{/X}$ , and the category ${\rm St}_{{\rm fl}}(X_{{\rm fl}}, \mathcal{C})$ of $\mathcal{C}$ -valued hypercomplete sheaves for the flat topology on the big Noetherian flat site $X_{{\rm fl}} := (\mathsf{dSt}_k/X, {\rm fl})$ . Now, the map $X \to {\rm Spec}\, k$ induces a functor $\mathsf{dSt}_k/X \to \mathsf{dSt}_k$ , which is continuous for the flat topologies and induces a morphism of sites $u \;:\; (\mathsf{dSt}_k, {\rm fl}) \to (\mathsf{dSt}_k/X, {\rm fl})$ . Therefore we get a pullback-preserving functor

To ease notation, we still denote this composition by $u^{-1}$ . For $Z\subset X$ a closed substack, we will put

\begin{align*} \underline {\mathsf{Perf}}_X := u^{-1}{\bf Perf}_X , \,\, \underline {\mathsf{Perf}}_{X \smallsetminus Z}:= u^{-1}{\bf Perf}_{X \smallsetminus Z}, \,\, \underline {\mathsf{Perf}}_{{\widehat {Z}} \smallsetminus Z}:=u^{-1}{\bf Perf}_{{\widehat {Z}} \smallsetminus Z}, \,\, \underline {\mathsf{Perf}}_{\widehat {Z}}:=u^{-1}{\bf Perf}_{\widehat {Z}}, \\ \underline {\mathsf{Coh}}^-_X := u^{-1}{\bf Coh}^-_X , \,\, \underline {\mathsf{Coh}}^-_{X \smallsetminus Z}:= u^{-1}{\bf Coh}^-_{X \smallsetminus Z}, \,\, \underline {\mathsf{Coh}}^-_{{\widehat {Z}} \smallsetminus Z}:=u^{-1}{\bf Coh}^-_{{\widehat {Z}} \smallsetminus Z}, \,\, \underline {\mathsf{Coh}}^-_{\widehat {Z}}:=u^{-1}{\bf Perf}_{\widehat {Z}}, \\ \underline {\mathsf{Bun}}^{{\bf G}}_X := u^{-1}{\bf Bun}^{{\bf G}}_X , \,\, \underline {\mathsf{Bun}}^{{\bf G}}_{X \smallsetminus Z}:= u^{-1}{\bf Bun}^{{\bf G}}_{X \smallsetminus Z}, \,\, \underline {\mathsf{Bun}}^{{\bf G}}_{{\widehat {Z}} \smallsetminus Z}:=u^{-1}{\bf Bun}^{{\bf G}}_{{\widehat {Z}} \smallsetminus Z}, \,\, \underline {\mathsf{Bun}}^{{\bf G}}_{\widehat {Z}}:=u^{-1}{\bf Bun}^{{\bf G}}_{\widehat {Z}} ,\end{align*}

and similar definitions for longer flags $Z_1 \subset \cdots \subset Z_n \subset X$ . These are $\mathcal{C}$ -valued hypercomplete sheaves on $\mathsf{dAff}_k$ for the flat topology, and both formal gluing (Proposition 3.3 ) and flag decomposition (Corollary 3.6 ) hold for them (since $u^{-1}$ preserves fibre products). In particular, for $X\in \mathsf{dSt}_k$ and $Z \subset X$ a closed substack, restriction induces equivalences in $\mathsf{dSt}_k$

\begin{align*} \underline {\mathsf{Perf}}_X \simeq \underline {\mathsf{Perf}}_{X\smallsetminus Z} \mathop{\times} _{\underline {\mathsf{Perf}}_{{\widehat {Z}}\smallsetminus Z}} \underline {\mathsf{Perf}}_{{\widehat {Z}}} \quad {\rm and}\quad \underline {\mathsf{Bun}}^{{\bf G}}_X \simeq \underline {\mathsf{Bun}}^{{\bf G}}_{X\smallsetminus Z} \mathop{\times} _{\underline {\mathsf{Bun}}^{{\bf G}}_{{\widehat {Z}}\smallsetminus Z}} \underline {\mathsf{Bun}}^{{\bf G}}_{{\widehat {Z}}}. \end{align*}

3.5 Non-derived version

The results presented so far all involve derived geometry. They imply similar results in classical geometry, which we will make explicit in this section, for the reader’s convenience.

(a) Fibre functors. The notion of fibre functors introduced in Definition 2.10 has an obvious non-derived analogue: a fibre functor over a stack $X$ is a functor $\mathsf{Aff}_X \to {\rm St}_X$ equipped with a natural transformation to the inclusion ${\mathfrak{F}}_X \;:\; \mathsf{Aff}_X \subset {\rm St}_X$ . We denote by $\mathsf{FibF}_X$ the $(2,1)$ -category of fibre functors.Footnote 19 Computing pointwise fibres defines a fully faithful functor $\tilde {\mathfrak{F}}_\bullet \;:\; {\rm St}_X \to \mathsf{FibF}_X$ .

Classical and derived fibre functors are related by two functors:

  • the inclusion $i \;:\; \mathsf{FibF}_X \to \mathsf{dFibF}_X^{\leq 1} \subset \mathsf{dFibF}_X$ induced by the inclusion ${\rm St}_X \subset \mathsf{dSt}_X$ and a Kan extension;

  • The truncation functor $\tau \;:\; \mathsf{dFibF}_X^{\leq 1} \to \mathsf{FibF}_X$ (restrict at the source and truncate the values);

where $\mathsf{dFibF}_X^{\leq 1}$ denotes the full subcategory of $\mathsf{dFibF}_X$ spanned by fibre functors $\mathfrak{F}$ such that for any affine (classical) scheme $S$ over $X$ , the classical truncation of ${\mathfrak{F}}(S)$ is a $1$ -stack (i.e. lies in ${\rm St}$ ). There is a canonical equivalence $\tau \circ i \simeq {{\rm id}}$ . Note that $i$ does not commute with the functors $\tilde {\mathfrak{F}}_\bullet \;:\; {\rm St}_X \to \mathsf{FibF}_X$ and ${\mathfrak{F}}_\bullet \;:\; {\rm St}_X \subset \mathsf{dSt}_X \to \mathsf{dFibF}_X$ in general. It does, however, in every case of interest to us: if the structural morphism $Y \to X$ is flat, then $i(\tilde {\mathfrak{F}}_Y\!) = {\mathfrak{F}}_Y$ .

(b) Formal glueing. These relations between derived and non-derived fibre functors will allow to deduce a non-derived version of the algebraization and formal glueing theorems. This is only possible due to the following observation.

Fix $X \in {\rm St}$ . Using the inclusions $\mathsf{Aff}_X \hookrightarrow \mathsf{dAff}_X$ and ${\rm St} \hookrightarrow \mathsf{dSt}$ , we get canonical functors

Lemma 3.8. Let $X \in {\rm St}$ and $Z \subset X$ a closed substack. The restrictions of the fibre functors

\begin{align*} {\mathfrak{F}}_X, \hspace {1em} {\mathfrak{F}}_{{\widehat {Z}}}, \hspace {1em} {\mathfrak{F}}^{{\rm aff}}_{{\widehat {Z}}}\quad {\rm and} \quad {\mathfrak{F}}_{{\widehat {Z}} \smallsetminus Z} \end{align*}

to $\mathsf{Aff}_X$ all belong to the essential image of $\mathsf{FibF}_X$ , with pre-image given respectively by

\begin{align*} S \mapsto S, \hspace {1em} S \mapsto {\widehat {Z}}_S, \hspace {1em} S \mapsto {\widehat {Z}}_S^{{\rm aff}}\quad {\rm and} \quad S \mapsto {\widehat {Z}}_S^{{\rm aff}} \smallsetminus Z_S, \end{align*}

where $Z_S$ is the closed subscheme $Z \times _X S \subset S$ .

Proof. The first two cases follow from the flatness of $S$ (respectively ${\widehat {Z}}_S$ ) over $S$ . The other two cases are trivial consequences.

Corollary 3.9. Both algebraization Proposition 3.1 and formal glueing Proposition 3.3 and Theorem 3.5 apply mutatis mutandis to the underived setting.

4. The 2-Segal object of good flags on a surface

In the previous section we worked by fixing a flag (in some fixed scheme or stack). In this section we will allow the flags to move in flat families (over a fixed scheme), and then move on to consider tuples of families of such flags.

Throughout this section, we will fix a smooth projective complex surface $X$ . Everything will depend on $X$ , so we allow ourselves to sometimes suppress $X$ from the notation.

We denote by $\operatorname {Hilb}_X$ the Hilbert scheme functor of (families of) $0$ -dimensional subschemes in $X$ (i.e. $\operatorname {Hilb}_X$ is the disjoint union of all the $\operatorname {Hilb}^P_X$ for Hilbert polynomials $P$ of degree $0$ ). We will write $\operatorname {Hilb}^{\leq 0}_X$ for the functor sending a $\mathbb{C}$ -scheme $S$ to $\operatorname {Hilb}_X(S) \cup \{ \emptyset _S \}$ , where $\emptyset _S$ denotes the empty subscheme $\emptyset \hookrightarrow X\times S$ (note that $\emptyset$ is flat over $S$ ). We denote by ${\rm Car}_{X}$ the set-valued functor of relative effective Cartier divisors on $X/\mathbb{C}$ , sending a $\mathbb{C}$ -scheme $S$ to the set of relative effective Cartier divisors on $(X\times S) /S$ (as in [Reference Katz and MazurKM, 1.2.3]).

4.1 Flags

Definition 4.1.

  • Let $S\in \mathsf{Aff}$ . We define ${\rm Fl}_X(S)$ as the subset of ${\rm Car}_{X}(S)\times \operatorname {Hilb}^{\leq 0}_X(S)$ consisting of pairs $(D,Z)$ with $Z$ a closed $(X\times S)-$ subscheme of $D$ (in which case, we will simply write $Z\subset D$ )Footnote 20 . Such pairs will be called families of flags on $X$ relative to $S$ .

  • The category $\operatorname {\underline {\mathsf{Fl}}}_X (S)$ is the category (poset) with set of objects ${\rm Fl}_X(S)$ and morphisms sets

    \begin{align*} \textrm {Hom}_{\operatorname {\underline {\mathsf{Fl}}}_X(S)}((D', Z'), (D, Z)) \end{align*}
    being a singleton if there is a commutative diagram of closed $(X\times S)$ -immersions

    (4.1)

    such that $(Z' \to D' \times _D Z)_{{\rm red}}$ is an isomorphism; $\textrm {Hom}_{\operatorname {\underline {\mathsf{Fl}}}_X(S)}((D', Z'), (D, Z))$ is otherwise empty. The composition is the unique one.

Since all maps are closed immersions, if a diagram like (4.1) exists, then it is unique, so that $\operatorname {\underline {\mathsf{Fl}}}_X (S)$ is indeed a poset.

Remark 4.2. According to Definition 4.1, the empty flag $(\emptyset , \emptyset )$ is an initial object in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ . However, if $(D,Z) \in \operatorname {\underline {\mathsf{Fl}}}_X(S)$ , then there is no morphism $(D, \emptyset ) \to (D,Z)$ in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ unless $Z=\emptyset$ . On the other hand, for any closed immersion $D'\hookrightarrow D$ , $(D', \emptyset ) \to (D,\emptyset )$ is in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ . In particular, the category ${\bf Car}_X(S)$ of relative effective Cartier divisors on $X\times S/S$ (with maps given by inclusions) fully embeds into $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ . A typical example of a morphism in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ is obtained from a chain of closed immersion $Z' \hookrightarrow D' \hookrightarrow D$ : $(D', Z') \to (D,Z')$ is in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ .

The inverse image functor preserves inclusions of closed subschemes, and for a map of schemes $Y'\to Y$ and a closed subscheme $T \hookrightarrow Y$ , the canonical closed immersion

\begin{align*} \left (Y'\mathop{\times} _Y T\right )_{\!{\rm red}} \longrightarrow Y'_{{\rm red}} \mathop{\times} _{Y_{{\rm red}}} T_{{\rm red}} \end{align*}

induces an isomorphism on the underlying reduced subschemes. Therefore, the inverse image functor along $S'\to S$ preserves morphisms between flags as defined in Definition 4.1, so that the following definition is well posed.

Definition 4.3. We denote by $\operatorname {\underline {\mathsf{Fl}}}_X: \mathsf{Aff}^{{\rm op}} \to {\bf PoSets}$ the induced functor, and call it the functor of relative flags on $X$ .

Remark 4.4. Our definitions 4.1 and 4.3 are flag analogues of [Reference Beilinson and DrinfeldBD04, 3.4.6] where the authors give an alternative definition of factorization algebras using effective Cartier divisors (on a curve) modulo reduced equivalence, instead of using the Ran space. It was exactly this remark in [Reference Beilinson and DrinfeldBD04] that helped us in crystallizing our first ideas on the topics of the current section of this paper.

Proposition 4.5. $\operatorname {\underline {\mathsf{Fl}}}_X$ is a stack of posets on the Zariski site of $\mathsf{Aff}$ .

Let ${\rm forget}: {\bf PoSets} \to {\bf Sets}$ denote the forgetful functor, sending a poset $(S, \leq )$ to the set $S$ .

Lemma 4.6. The composite functor

is represented by a scheme.

Proof. Since $X$ is smooth projective, ${\rm Car}_{X} \times \operatorname {Hilb}^{\leq 0}_{X}$ is represented by a scheme, and the nesting condition $Z\subset D$ for flags is a closed subscheme condition (see also [Reference SernesiSer06, 4.5]).

4.2 Good pairs of flags

In order to be short, we will write that $D$ is a reCd on $X\times S/S$ to mean that $D$ is a relative effective Cartier divisor on $X \times S/S$ .

Definition 4.7. Let $S \in \mathsf{Aff}$ , and $D, D' \in {\rm Car}_{X}(S)$ . We say that the pair of reCd’s $(D,D')$ on $X\times S/S$ is good if $D\cap D' \in {\rm Car}_{D/S}(S) \cap {\rm Car}_{D'/S}(S)$ (i.e. $D\cap D'$ is an effective Cartier divisor inside both $D$ and $D'$ , and moreover $D\cap D'$ is flat over $S$ ).

It is obvious that, for $D, D' \in {\rm Car}_{X}(S)$ , $(D,D')$ is good if and only if $(D',D)$ is.

Lemma 4.8. Let $B$ be a commutative ring and let $f, f'$ be non-zero divisors in $B$ . The following assertions are equivalent:

  1. (i) $f'$ is not a zero divisor in $B/f$ ;

  2. (ii) $f$ is not a zero divisor in $B/f'$ .

As a consequence, for any $S \in \mathsf{Aff}$ and $D, D' \in {\rm Car}_{X}(S)$ , the following are equivalent:

  1. (a) the pair $(D, D')$ is good;

  2. (b) $D \cap D'$ is a relative effective Cartier divisor in $D$ ;

  3. (c) $D \cap D'$ is a relative effective Cartier divisor in $D'$ .

Proof. By symmetry, it is enough to prove that (i) implies (ii). Let $x, y \in B$ be such that $xf' = yf$ . By assumption, the image of $f'$ in $B/f$ is a non-zero divisor, so there exists $t \in B$ such that $x = tf$ . Since $f$ is not a zero divisor in $B$ , the equality $yf = xf' = tff'$ implies $y = tf'$ . As this holds for any $x$ and any $y$ as above, we get that $f$ is not a zero divisor in $B/f'$ .

Definition 4.9. Let $S \in \mathsf{Aff}$ .

  • We define ${\rm Fl}_{X,2}(S)$ as the subset of ${\rm Fl}_X(S)\times {\rm Fl}_X(S)$ whose elements are good pairs $((D,Z), (D',Z'))$ of families of flags of $X$ relative to $S$ , i.e. those pairs satisfying the following conditions.

    1. (i) The pair of reCd’s $(D, D')$ is good (Definition 4.7).

    2. (ii) $D\cap D' = Z\cap Z'$ ;

  • The category $\operatorname {\underline {\mathsf{Fl}}}_{X,2}$ is the full subcategory (which is again a poset) of $\operatorname {\underline {\mathsf{Fl}}}_X(S)^{\times 2}$ spanned by ${\rm Fl}_{X,2}(S)$ .

Remark 4.10.

  1. (a) Note that condition (i) Definition 4.9 easily implies (see, e.g., [Sta13, Lemma 0C4R]) that for a good pair $((D,Z), (D',Z'))$ , we have $D + D'= D\cup D'$ , so that, in particular, $D\cup D'$ is flat over $S$ .

  2. (b) Let $S \in \mathsf{Aff}$ . If we denote by $(\emptyset , \emptyset )\in {\rm Fl}(S)$ the empty flag, then the pair $((D,Z), (\emptyset , \emptyset ))$ is a good pair, for any $(D,Z)\in {\rm Fl}(S)$ (indeed, note that the empty scheme is an effective Cartier divisor in itself, since $0$ is not a zero divisor in the zero ring). Moreover, the pair $((D,Z), (D',Z')) \in {\rm Fl}_X(S) \times {\rm Fl}_X(S)$ is good if and only if $((D',Z'), (D,Z))$ is good.

  3. (c) If $X$ is a smooth projective curve (instead of our surface), and for obvious reasons we limit ourselves to flags of type $(D, \emptyset )$ , then a pair $((D,\emptyset ), (D', \emptyset )) \in \operatorname {\underline {\mathsf{Fl}}}_{X,2}(S)$ is good if and only if $D$ and $D'$ are disjoint divisors. If $X$ is a smooth surface, then any pair $((D',Z'),(D,Z))$ of flags with $D\cap D' =\emptyset$ is good (but the converse is clearly false).

Note that conditions (i) and (ii) in Definition 4.9 are stable under arbitrary base change: this is obvious for (ii), while it follows for instance from [Sta13, Lemma 063U] for (i). In other words, for every morphism $S' \to S$ of affine schemes, the induced functor

\begin{align*} \operatorname {\underline {\mathsf{Fl}}}_X(S) \times \operatorname {\underline {\mathsf{Fl}}}_X(S) \to \operatorname {\underline {\mathsf{Fl}}}_X(S') \times \operatorname {\underline {\mathsf{Fl}}}_X(S') ,\end{align*}

restricts to a functor

\begin{align*} \operatorname {\underline {\mathsf{Fl}}}_{X,2}(S) \to \operatorname {\underline {\mathsf{Fl}}}_{X,2}(S'). \end{align*}

So we actually get, using inverse images, a functor

\begin{align*}\operatorname {\underline {\mathsf{Fl}}}_{X, 2}: \mathsf{Aff}^{{\rm op}} \longrightarrow {\bf PoSets}, \,\, S \longmapsto \operatorname {\underline {\mathsf{Fl}}}_{X, 2}(S).\end{align*}

Lemma 4.11. Let $S \in \mathsf{Aff}$ .

  1. (i) If $Z$ and $Z'$ are closed subschemes of $X \times S$ , both flat over $S$ , and $Z\cap Z'$ is flat over $S$ , then $Z\cup Z'$ is flat over $S$ .

  2. (ii) Let $Z$ and $Z'$ be closed subschemes of $X \times S$ such that $Z\cap Z'$ is flat over $S$ . Let $f:S' \to S$ an arbitrary morphism, and $F:= {\rm id}_X \times f: X \times S' \to X\times S$ . Then, the canonical map

    \begin{align*} F^{-1}(Z) \cup F^{-1}(Z') \to F^{-1}(Z \cup Z')\end{align*}
    is an isomorphism of closed subschemes of $X\times S'$ (or, equivalently, the $\mathcal{O}_{X\times S}$ -ideals of $F^{-1}(Z) \cup F^{-1}(Z')$ and of $F^{-1}(Z \cup Z')$ coincide).
  3. (iii) Let $D_1, D_2 \in {\rm Car}_X(S)$ , $f:S' \to S$ is an arbitrary morphism, and $F:= {\rm id}_X \times f: X \times S' \to X\times S$ . Then, the canonical map

    \begin{align*} {\rm Car}_X (f)(D_1) +{\rm Car}_X (f)(D_2)&= F^{-1}(D_1)+ F^{-1}(D_2) \longrightarrow F^{-1}(D_1 +D_2)\\&={\rm Car}_X (f)(D_1 + D_2 )\end{align*}
    is an isomorphism of relative effective Cartier divisors on $X\times S'/S'$ (or equivalently, the $\mathcal{O}_{X\times S}$ -ideals of $F^{-1}(D_1)+ F^{-1}(D_2)$ and of $F^{-1}(D_1 +D_2)$ coincide).

Proof. To prove (i), consider the pushout (of schemes)

where we know that $Z, Z'$ and $Z\cap Z'$ are flat over $S$ . To conclude that also the pushout $Z\cup Z'$ is flat over $S$ , we may we may work locally on $X \times S$ , and thus replace it by an affine scheme, so that both $Z$ and $Z'$ are affine, and thus so is $Z\cap Z'$ . In particular, if we put $X \times S={\rm Spec} A$ , $Z\cap Z'={\rm Spec} B$ , we are therefore in the following situation: $B, C, D$ are flat $A$ -algebras and $\pi _C: C\to B$ , $\pi _D: D\to B$ are surjective morphisms of $A$ -algebras and we want to conclude that the fibre product $C\times _B D$ is flat over $A$ . This follows from the exact sequence of $A$ -modules $0\to C\times _B D \to C\times D \to B \to 0$ , since both $C\times D$ and $B$ are then flat over $A$ . To prove (ii) we adopt the same reduction to the affine case and same notation as in the proof of (i), and let $\varphi \;:\; A \to A'$ such that $f={\rm Spec} (\varphi )$ . We need to prove that the canonical map of $A'$ -algebras $\rho \;:\; (C\times _B D)\otimes _A A' \to (C\otimes _A A')\times _{B\otimes _A A'} (D\otimes _A A')$ , sending $((c,d)\otimes a')$ to $ (c\otimes a', d\otimes a')$ is an isomorphism. It will be enough to prove that $\rho$ is an isomorphism of $A'$ -modules. Apply the base-change functor $(-)\otimes _A A'$ to the exact sequence

of $A$ -modules, and observe that the canonical map $(C\times D)\otimes _A A' \to (C\otimes _A A') \times (D\otimes _A A')$ is an isomorphism (distributivity of tensor product on direct sums in $\mathsf{Mod}_{A'}$ ). By hypothesis, $B$ is flat over $A$ , so that ${\rm Tor}^A_1(B, A')=0$ , and we get an exact sequence of $A'$ -modules

(4.2)

Now, exactness of (4.2) is equivalent to $\rho$ being an isomorphism of $A'$ -modules.

Finally, (iii) is a well-known result that can be proved again by reducing to the affine case: $S'={\rm Spec} R' \to {\rm Spec} R=S$ , $X\times S= {\rm Spec} A$ , $X\times S' = {\rm Spec} (A\otimes _R R')$ , $D_i ={\rm Spec} (A/f_i)$ , $i=1,2$ . Then $D_1+D_2 = {\rm Spec} (A/f_1f_2)$ and $F^{-1}(D_1 + D_2)= {\rm Spec} (A'/(f_1f_2)')$ where $(-)': A \to A'=A\otimes _R R'$ is the canonical map. Since $(-)'$ is a morphism of algebras, and $F^{-1}(D_1) + F^{-1}(D_2)= {\rm Spec} (A'/(f_1)'(f_2)')$ , we conclude.

Lemma 4.11 allows for the following.

Definition 4.12. For $S\in \mathsf{Aff}$ and $((D,Z), (D',Z')) \in \operatorname {\underline {\mathsf{Fl}}}_{X,2}(S)$ , define

\begin{align*}\cup _S((D,Z), (D',Z')):=(D+D', Z\cup Z') \in \operatorname {\underline {\mathsf{Fl}}}_X(S),\end{align*}

called a union of good pairs. We will most often write $(D,Z) \cup (D',Z')$ for $\cup _S((D,Z), (D',Z'))$ .

Note that Definition 4.12 is well posed since $D+D'$ is a relative effective Cartier divisor on $X$ over $S$ (this is classical, see, e.g., [Reference Katz and MazurKM, Ch. 1]), $Z\cup Z'$ lies in $\operatorname {Hilb}_{X}^0(S)$ (since by Lemma 4.11(i), Definition 4.9(i) and (ii) ensure that if $((D,Z), (D',Z')) \in {\rm Fl}_2(S)$ , then $Z \cup Z'$ is again flat over $S$ ), and obviously $Z\cup Z' \subset D+D'$ .

Remark 4.13. The union of good pairs is obviously commutative and unital, with unit given by the empty flag $(\emptyset , \emptyset )$ . Moreover, by Remark 4.10(a), we have

\begin{align*} (D,Z)\cup (D',Z') := (D+D', Z\cup Z')= (D \cup D', Z\cup Z') \end{align*}

for any good pair $((D,Z), (D',Z'))$ .

Lemma 4.14. The union $\cup \;:\; \operatorname {\underline {\mathsf{Fl}}}_{X, 2} \to \operatorname {\underline {\mathsf{Fl}}}_{X}$ is a morphism of prestacks.

Proof. For each fixed $S \in \mathsf{Aff}$ , we have seem that $\cup _S$ is well defined on objects. On morphisms, the only point that needs some argument is to show that if $(j_1, j_2): (F_1, F_2) \to (F'_1, F'_2)$ is a morphism in $\operatorname {\underline {\mathsf{Fl}}}_{X, 2}(S)$ ( $F_i=(D_i, Z_i), F'_i=(D'_i, Z'_i), \, i=1,2$ ), then the induced map $j_1 \cup j_2: F_1 \cup F_2 \to F'_1 \cup F'_2$ has the property that the canonical map

\begin{align*}Z_1 \cup Z_2 \to (D_1+D_2) \cap (Z'_1 \cup Z'_2)\end{align*}

is an isomorphism on the underlying reduced schemes. This follows from the fact that both $Z_1 \to D_1 \cap Z'_1$ , and $Z_2 \to D_2 \cap Z'_2$ are isomorphisms on the underlying reduced schemes, and from the fact that it is enough to prove the equality $Z_1 \cup Z_2 = (D_1+D_2) \cap (Z'_1 \cup Z'_2)$ at the level of the underlying topological spaces (which is an elementary verification, using that the pairs $(F_1, F_2)$ and $(F'_1, F'_2)$ are good). Finally, naturality in $S$ is an easy consequence of Lemma 4.11 (ii) and (iii).

4.3 Good tuples of flags

Proposition 4.15. Fix $S = {\rm Spec}(A)$ an affine test scheme and $(D_1, \ldots , D_n) \in {\rm Car}_{X}(S)^n$ . The following conditions are equivalent:

  1. (i) for any $i \neq j$ , the pair of divisors $(D_i, D_j)$ is good (in the sense of Definition 4.7 );

  2. (ii) for any $i$ , the pair $(D_1 + \cdots + D_i ,D_{i+1})$ is good;

  3. (iii) for any disjoint subsets $I$ and $J$ of $\{1, \ldots , n\}$ , the pair $(\sum _{i\in I} D_i ,\sum _{j \in J} D_j)$ is good.

We shall say that a family of such divisors is good if it satisfies the above conditions.

Proof. Let us first prove (i) and (ii) are equivalent. This straightforwardly reduces to the case of three divisors $(D_1, D_2, D_3)$ .

Assume first that they are pairwise good and let us prove that $D_1 + D_2$ and $D_3$ are good. The statement being local, we choose an affine chart ${\rm Spec} B \subset X \times S$ and pick non-zero divisors $f_i \in B$ cutting out $D_i$ . As $D_i$ and $D_j$ are good (for $i \neq j$ ), the image of $f_i$ in $B/f_j$ is a non-zero divisor as well.

Let $g,h \in B$ such that $f_3 g=f_1f_2h$ . Since $f_3$ is a non-zero divisor in $B/f_1$ , there exists $g' \in B$ such that $g = f_1 g'$ . As $f_1$ is a non-zero divisor in $A$ , we get $f_3 g' = f_2 h$ . Since $f_3$ is a non-zero divisor in $B/f_2$ , we get $g''$ such that $g' = f_2 g''$ . This in turn implies $f_3 g'' = h$ . This shows that $f_1f_2$ is a non-zero divisor in $B/f_3$ .

In order to prove that $(D_1 + D_2) \cap D_3$ is a relative effective Cartier divisor in $D_3$ , it remains to show it is flat over $A$ . Consider then the following sequence:

(4.3)

The morphism $\pi$ is obviously surjective, and its kernel coincides with the image of $\alpha$ . To prove that $\alpha$ is injective, consider $b \in B$ such that $f_2b=0 \in B/(f_1f_2, f_3)$ . Pick $x, y \in B$ such that $f_2b = f_1f_2x + f_3y$ . Since $f_3$ is a non-zero divisor in $B/f_2$ , we find $y'$ satisfying $f_2y' = y$ . As $f_2$ is a non-zero divisor in $B$ , we get $b - f_1x = f_3 y'$ . In particular $b=0 \in B/(f_1, f_3)$ and $\ker (\alpha ) = 0$ . The above sequence is therefore exact. Since both $B/(f_1, f_3)$ and $B/(f_2, f_3)$ are flat over $A$ , so is $B/(f_1f_2, f_3)$ . This shows that $(D_1 + D_2) \cap D_3$ is a relative effective Cartier divisor in $D_3$ . Using Lemma 4.8, we deduce that (i) implies (ii).

We now assume that $(D_1, D_2)$ and $(D_1 + D_2, D_3)$ are good pairs of divisors. We fix as above non-zero divisors $f_i \in B$ generating the ideal corresponding to $D_i$ . Let $g, h \in B$ such that $gf_1 = hf_3$ . Multiplying by $f_2$ and using that $f_1f_2$ is not a zero divisor in $B/f_3$ , we get $t \in B$ such that $g = tf_3$ . In particular, $f_1$ is a non-zero divisor in $B/f_3$ . By Lemma 4.8, $f_3$ is not a zero divisor in $B/f_1$ either. The roles of $f_1$ and $f_2$ being symmetric, the same holds for $f_2$ in place of $f_1$ .

The sequence (4.3) above is still exact. Since $B/(f_1f_2, f_3)$ is flat over $A$ , the associated long Tor exact sequence yields

\begin{align*} \forall n \geq 1,\quad {\rm Tor}_n^A(B/(f_1, f_3), -) \simeq {\rm Tor}_{n+1}^A(B/(f_2, f_3), -). \end{align*}

Again, the role of $f_1$ and $f_2$ being symmetric, we also get ${\rm Tor}_n^A(B/(f_2, f_3), -) \simeq {\rm Tor}_{n+1}^A(B/(f_1, f_3), -)$ . As a consequence, the Tor-functors ${\rm Tor}_n^A(B/(f_1, f_3), -)$ are $2$ -periodic

\begin{align*} \forall n \geq 1,\quad {\rm Tor}_n^A(B/(f_1, f_3), -) \simeq {\rm Tor}_{n+2}^A(B/(f_1, f_3), -). \end{align*}

However, the exact sequence $0 \to B/f_1 \overset {f_3}{\to } B/f_1 \to B/(f_1, f_3) \to 0$ and the assumption that $B/f_1$ is flat over $A$ imply that $B/(f_1, f_3)$ has Tor-dimension at most $1$ . Together with the above $2$ -periodicity, we deduce that $B/(f_1, f_3)$ is flat over $A$ . We proved that the pair $(D_1, D_3)$ is good. By symmetric, so is $(D_2, D_3)$ and thus (ii) implies (i).

Clearly, (iii) implies (ii). The converse implication is proven by induction, up to renumbering the divisors.

An analogue of Proposition 4.15 for flags does not quite hold: the right notion of good families of flags cannot be defined in terms of pairs of elements.

Definition 4.16. Fix $S$ an affine test scheme. We define the notion of good families of flags inductively on the cardinal:

  • a pair $(F_1, F_2) \in \operatorname {\underline {\mathsf{Fl}}}_X(S)^2$ is good if it is in the sense of Definition 4.9;

  • a family $(F_1, \ldots , F_n) \in \operatorname {\underline {\mathsf{Fl}}}_X(S)^n$ is good if for any $i \lt j$ , the pair $(F_i, F_j)$ is good, and the family $(F_i \cup F_j, F_1, \ldots , F_{i-1}, F_{i+1}, \ldots , F_{j-1}, F_{j+1}, \ldots , F_n)$ is good.

Proposition 4.17. Fix $S$ an affine test scheme and $(F_1, \ldots , F_n) \in \operatorname {\underline {\mathsf{Fl}}}_X(S)^n$ . The following conditions are equivalent:

  1. (i) the family $(F_1, \ldots , F_n)$ is good;

  2. (ii) the families $(F_2, \ldots , F_n)$ , $(F_1, \ldots , F_{n-1})$ and $(F_1, \ldots , F_i \cup F_{i+1}, \ldots , F_n)$ for $0 \lt i \lt n$ , are all good;

  3. (iii) for any $i$ , the pair $(F_1 \cup \cdots \cup F_i, F_{i+1})$ is good;

  4. (iv) the family $(F_1, \ldots , F_{n-1})$ is good and there exists $1 \leq i \leq n-2$ such that the family $(F_1, \ldots , F_i \cup F_{i+1}, \ldots , F_n)$ is good;

  5. (v) the family $(F_2, \ldots , F_n)$ is good and there exists $2 \leq i \leq n-1$ such that the family $(F_1, \ldots , F_i \cup F_{i+1}, \ldots , F_n)$ is good.

Remark 4.18. The definition of good families of flags is independent of the order of the flags. As a consequence, there also are characterizations similar to (ii) through (v) but where the role of the flags are permuted.

In particular, a family $(F_1, \ldots , F_n)$ is good if and only if there exists a permutation $\sigma$ such that each pair $(F_{\sigma (1)} \cup \ldots \cup F_{\sigma (i)}, F_{\sigma (i+1)})$ is good. In this case, this actually holds for any permutation $\sigma$ .

We can interpret this fact as: ‘A family of flags is good if there exists a way of computing its union by adding flags one by one. In this case, every such way of computing the union is valid’.

Lemma 4.19. Let $S$ be an affine scheme and $F_1$ , $F_2$ and $F_3$ be three flags in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ . The following statements are equivalent:

  1. (i) the pairs $(F_1, F_2)$ and $(F_1 \cup F_2, F_3)$ are good;

  2. (ii) the pairs $(F_1, F_3)$ and $(F_1 \cup F_3, F_2)$ are good;

  3. (iii) the pairs $(F_2, F_3)$ and $(F_2 \cup F_3, F_1)$ are good.

Proof. By symmetry, it suffices to show the first assertion implies the second one. Assume thus that $(F_1, F_2)$ and $(F_1 \cup F_2, F_3)$ are good pairs of flags. For each $i$ , we denote by $(D_i, Z_i)$ the flag $F_i$ . By Proposition 4.15, the pairs of divisors $(D_1, D_3)$ and $(D_1 + D_3, D_2)$ are good. Remark that this implies $D_1 + D_3 = D_1 \cup D_3$ . It is therefore enough to verify $D_1 \cap D_3 = Z_1 \cap Z_3$ and $(D_1 \cup D_3) \cap D_2 = (Z_1 \cup Z_3) \cap Z_2$ . This can be checked locally in some Zariski chart ${\rm Spec} B$ of $X \times S$ . We denote by $f_k$ a non-zero divisor of $B$ of whom $D_k$ is the zero-locus, by $I_k$ the ideal $(f_k)$ and by $J_k$ the ideal of functions vanishing on $Z_k$ (so that by definition $I_k \subset J_k$ ). By assumption $I_1 + I_2 = J_1 + J_2$ and $I_3 + (I_1 \cap I_2) = J_3 + (J_1 \cap J_2)$ . We have to prove

\begin{align*} \left\{\begin{array}{l@{\quad}l} I_1 + I_3 = J_1 + J_3, & {\rm (a)} \\ I_2 + (I_1 \cap I_3)=J_2 + (J_1 \cap J_3). & {\rm (b)} \end{array} \right. \end{align*}

This is elementary. Here are the details.

  1. (a) Let $ \xi _1 + \xi _3 \in J_1 + J_3$ , $\xi _i \in J_i$ . Since $J_1 \subset I_1 + I_2$ , we get $\xi _1= x_1 + x_2 \in I_1 + I_2$ , so that $x_2 =\xi _1 - x_1 \in I_2 \cap J_1 \subset J_2 \cap J_1$ . Therefore $x_2+ \xi _3= x_3 + y_{12} \in I_3 + (I_1 \cap I_2)$ for some $x_3 \in I_3$ and $y_{12} \in I_1 \cap I_2$ . This gives $\xi _1 + \xi _3= (x_1 + y_{12}) + x_3 \in I_1 + I_3$ .

  2. (b) Let $\xi _2 + \xi _{13} \in J_2 + (J_1 \cap J_3)$ , with $\xi _2 \in J_2$ and $\xi _{13} \in J_1 \cap J_3$ . We have $\xi _{13} \in J_1 \subset I_1 + I_2$ so there exist $y_1 \in I_1$ and $y_2 \ni I_2$ such that $\xi _{13} = y_1 + y_2$ . We get $y_2 = \xi _{13} - y_1 \in I_2 \cap J_1 \subset J_2 \cap J_1$ . On the other hand, $y_1 = \xi _{13} - y_2 \in J_3 + (J_2 \cap J_1) = I_3 + (I_1 \cap I_2)$ , so that we have $y_1 = t_3 + t_{12}$ for some $t_3 \in I_3$ and $t_{13}$ in $(I_1 \cap I_2)$ . We find $t_3 \in I_1 \cap I_3$ , since $y_1 \in I_1$ . Therefore, $\xi _{13}= (t_3 + t_{12}) + y_2 \in I_2 + (I_1 \cap I_3)$ . Hence, we are left to prove that $\xi _2 \in I_2 + (I_1 \cap I_3)$ as well. Decomposing $\xi _2 = x_1 + x_2 \in I_1 + I_2$ , we find $x_1 \in I_1 \cap J_2 \subset J_1 \cap J_2 \subset I_3 + (I_1 \cap I_2)$ . Let $\alpha _3 \in I_3$ and $\alpha _{12} \in I_1 \cap I_2$ such that $x_1 = \alpha _3 + \alpha _{12}$ . This implies $\alpha _3 \in I_1\cap I_3$ and in turn $\xi _2 = x_2 + x_1 = x_2 + \alpha _3 + \alpha _{12} \in I_2 + (I_1 \cap I_3)$ .

Proof of Proposition 4.17. Let us first observe that any subfamily of a good family is good. In particular (i) implies (ii). Moreover, a rapid induction on $i$ shows that (ii) implies (iii).

Let us now assume (iii). We will prove (iii) $\implies$ (i) by induction on $n$ . For $n=2$ , the implication is tautological. Assume now that it holds for $n \geq 2$ and pick a family $(F_1, \ldots , F_{n+1}) \in \operatorname {\underline {\mathsf{Fl}}}_X(S)^{n+1}$ satisfying (iii). For any $1 \leq i \lt j \leq n+1$ , we denote by $E_{ij}$ the $n$ -tuple

\begin{align*} E_{ij} = (F_i \cup F_j, F_k, k \in \{1, \ldots , n+1\} \smallsetminus \{i,j\}). \end{align*}

We have to show that each pair $(F_i, F_j)$ and each $E_{ij}$ is good.

The truncated family $(F_1, \ldots , F_n)$ satisfies (iii) and thus, by induction, is good. In particular, for any $1 \leq i \lt j \leq n$ , the pair $(F_i, F_j)$ and the family $(F_i \cup F_j, F_k, k \in \{1, \ldots , n\} \smallsetminus \{i,j\})$ are good. With the original assumption that $(F_1 \cup \cdots \cup F_n, F_{n+1})$ is good, it follows that the $n$ -tuple $E_{ij}$ satisfies (iii) and is thus good.

It remains to handle the pairs $(F_i, F_{n+1})$ and the families $E_{i,n+1}$ . The $n$ -tuple $(F_1 \cup F_2, F_3, \ldots , F_{n+1})$ satisfies (iii) and is thus good. In particular, the pair $(F_i, F_{n+1})$ is good for $i \geq 3$ . The pairs $(F_1, F_2)$ and $(F_1 \cup F_2, F_{n+1})$ are also good, and thus so are the pairs $(F_1, F_{n+1})$ and $(F_2, F_{n+1})$ by Lemma 4.19.

By induction, to prove that $E_{i,n+1}$ is a good $n$ -tuple, it suffices to show it satisfies (iii): i.e. that each pair $P_k^i := (F_1^i \cup \cdots \cup F_k^i, F_{k+1}^i)$ is good, where $F_m^i = F_m$ if $m \neq i$ and $F_i^i = F_i \cup F_{n+1}$ .

Assume for now $i \geq 2$ . By assumption, the pair $P_k^i$ is good if $k \lt i - 1$ . The $(n+2-i)$ -tuple $(F_1 \cup \cdots \cup F_i, F_{i+1}, \ldots , F_{n+1})$ satisfies (iii) and is thus good. In particular, the pair $(F_1 \cup \cdots \cup F_i, F_{n+1})$ is good. Since $(F_1 \cup \ldots \cup F_{i-1}, F_i)$ is a good pair, Lemma 4.19 implies that $P_{i-1}^i = (F_1 \cup \ldots \cup F_{i-1}, F_i \cup F_{n+1})$ is a good pair as well. Moreover $P_k^i = P_k^1$ as soon as $k \geq i$ , so the case $i=1$ induces the others.

We are left to prove that for any $1 \leq k \lt n$ , the pair $P_k^1 = (F_{n+1} \cup F_1 \cup \cdots \cup F_k, F_{k+1})$ is good. This also follows from Lemma 4.19, since both $(F_1 \cup \ldots \cup F_k, F_{k+1})$ and $(F_1 \cup \ldots \cup F_{k+1}, F_{n+1})$ are good pairs of flags. The family $(F_1, \ldots , F_{n+1})$ is therefore good. This concludes the proof of the implication (iii) $\implies$ (i) by induction on $n$ .

It remains to prove (iv) $\iff$ (ii) $\iff$ (v). By symmetry, dealing with the first equivalence is enough. Moreover, we tautologically have (ii) $\implies$ (iv). Conversely, assume (iv) holds for a family $(F_1, \ldots , F_n)$ . Since $(F_1, \ldots , F_{n-1})$ is good, it satisfies (iii) and each pair $(F_1 \cup \ldots \cup F_{p-1}, F_p)$ is good for $p \lt n$ .

Moreover, there exists $i \leq n-1$ such that $(F_1, \ldots , F_i \cup F_{i+1}, \ldots , F_n)$ is good. This family thus also satisfies (iii) and thus the pair $(F_1 \cup \ldots \cup F_{n-1}, F_n)$ is good as well. In particular, the family $(F_1, \ldots , F_n)$ satisfies (iii) and is thus good.

4.4 The $2$ -Segal object of good flags

The union operation of Lemma 4.14 is actually a partially defined commutative monoid structure on $\operatorname {\underline {\mathsf{Fl}}}_{X}$ : the operation is only defined for some pairs of elements. This sort of structure requires extra care when defining (and checking) the various constraints (here, associativity and commutativity). For instance, we need to ensure that every operation is actually well defined when writing down the associativity constraint.

A concise way of writing those details consists in using $2$ -Segal objects. The notion of $2$ -Segal object was first introduced in [Reference Dyckerhoff and KapranovDK19]. Recall from [Reference Feller, Garner, Joachim, Proulx and WeberFGJ+21] that a $2$ -Segal object in a category $\mathcal{C}$ with fibre products, is a simplicial object $T_{\bullet }$ in $\mathcal{C}$ satisfying the following $2$ -Segal condition: for any $(i,n)$ with $0\lt i\lt n$ , the following squares are Cartesian in $\mathcal{C}$ .

The category ${\bf 2\mbox{-}Seg}_{\Delta }(\mathcal{C})$ of $2$ -Segal objects in $\mathcal{C}$ is the corresponding full sub-category of the category of simplicial objects in $\mathcal{C}$ .

We will thus define a $2$ -Segal object $\operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }$ in prestacks that we will later use to construct a factorization structure on our flag version of the Beilinson–Drinfeld affine Grassmannian. Another important feature of the category ${\bf 2\mbox{-}Seg}_{\Delta }(\mathcal{C})$ for an arbitrary category $\mathcal{C}$ with fibre products, is its equivalence [Reference SternSte21, Theorem 2.25] with the category ${\bf Alg}(\mathcal{C}_{{\rm corr}})$ of associative algebras in the category $\mathcal{C}_{{\rm corr}}$ of correspondences (or spans) in $\mathcal{C}$ . This equivalence will be exploited later, in § 8.

Let us start by fixing a test scheme $S\in \mathsf{Aff}$ . We will construct a simplicial object in $\mathsf{PoSet}$ , which we denote by $\operatorname {\underline {\mathsf{Fl}}}_{X,\bullet }(S)$ .

Definition 4.20. For any $n \geq 0$ , we define $\operatorname {\underline {\mathsf{Fl}}}_{X,n}(S)$ as the full sub-poset of $\operatorname {\underline {\mathsf{Fl}}}_X(S)^n$ of good families of flags (in the sense of Definition 4.16). By convention, we set $\operatorname {\underline {\mathsf{Fl}}}_{X,0}(S): = *$ , $\operatorname {\underline {\mathsf{Fl}}}_{X,1} (S):=\operatorname {\underline {\mathsf{Fl}}}_X(S)$ .

By Proposition 4.17, for $0\leq i \leq n$ we have (order-preserving) face maps $\partial _i^{n} \;:\; \operatorname {\underline {\mathsf{Fl}}}_{X, n} (S) \to \operatorname {\underline {\mathsf{Fl}}}_{X, n-1} (S)$ whose values on objects is defined as follows:

\begin{align*} \partial _i^n: (F_1,\ldots , F_n) \longmapsto\left\{\begin{array}{l@{\quad}l@{\quad}l}(F_2, \ldots , F_n) & if & i=0 ,\\ (F_1 , \ldots , F_i \cup F_{i+1}, \ldots F_n) & if & 0\lt i\lt n ,\\ (F_1 ,\ldots , F_{n-1}) & if & i=n . \end{array}\right. \end{align*}

Similarly, we have degeneracy maps $\sigma _i^{n} \;:\; \operatorname {\underline {\mathsf{Fl}}}_{X, n} (S) \to \operatorname {\underline {\mathsf{Fl}}}_{X, n+1} (S)$ defined as

\begin{align*} \sigma _i^n: (F_1,F_2,\ldots , F_n) \longmapsto (F_1,\ldots , F_{i}, \emptyset , F_{i+1}, \ldots F_n). \end{align*}

It is immediate to verify that these give $\operatorname {\underline {\mathsf{Fl}}}_{X, \bullet } (S)$ the structure of a simplicial object in $\mathsf{PoSet}$ .

These constructions are functorial in $S$ : every $\operatorname {\underline {\mathsf{Fl}}}_{X, n} (S)$ is a functor $\mathsf{Aff}^{{\rm op}} \to \mathsf{PoSet}$ , as they are subfunctors of $\operatorname {\underline {\mathsf{Fl}}}_{X} (S)^{\times n}$ . Moreover, functoriality of the degeneracy maps $\sigma _i^n$ with respect to maps in $\mathsf{Aff}^{{\rm op}}$ is obvious, while the fact that also the face maps $\partial _{i}^n$ are functorial is a consequence of Lemma 4.14. Therefore, $S \mapsto \operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }(S)$ defines a functor

\begin{align*} \operatorname {\underline {\mathsf{Fl}}}_{X,\bullet } \;:\; \mathsf{Aff}^{{\rm op}} \to s\mathsf{PoSet} ,\end{align*}

or, equivalently, a simplicial object in ${\rm Fun}(\mathsf{Aff}^{{\rm op}}, \mathsf{PoSet})$ , which we still denote $\operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }$ .

Theorem 4.21. The simplicial object $\operatorname {\underline {\mathsf{Fl}}}_{X,\bullet }$ is a $2$ -Segal object.

Proof. First of all, recall (see [Reference Feller, Garner, Joachim, Proulx and WeberFGJ+21]) that we have to prove that $\operatorname {\underline {\mathsf{Fl}}}_{X,\bullet }$ satisfies the $2$ -Segal conditions, i.e. that for any $(i,n)$ with $0\lt i\lt n$ , the following squares are Cartesian.

We focus on the square $(\sigma )$ , the case $(\tau )$ being deduced by symmetry. We will show that the natural map of posets

\begin{align*} \operatorname {\underline {\mathsf{Fl}}}_{X, n+1} (S) \to H := \operatorname {\underline {\mathsf{Fl}}}_{X, n} (S) \mathop{\times} _{\partial _0,\, \operatorname {\underline {\mathsf{Fl}}}_{X, n-1} (S), \, \partial _i} \operatorname {\underline {\mathsf{Fl}}}_{X, n} (S) \end{align*}

is an isomorphism. To begin with, observe that the composition

is the canonical inclusion. Moreover, the restriction of $({{\rm id}}, p_1) \;:\; \operatorname {\underline {\mathsf{Fl}}}_X^n \times \operatorname {\underline {\mathsf{Fl}}}_X^n \to \operatorname {\underline {\mathsf{Fl}}}_X^{n+1}$ to $H$ is injective (as a map of posets). It is then enough to prove that $\operatorname {\underline {\mathsf{Fl}}}_{X, n+1}$ and $H$ have the same points. This is exactly the content of the equivalence (i) $\iff$ (v) from Proposition 4.17.

Remark 4.22. Since the union of flags is obviously strictly commutative, one can easily verify that the associative algebra in correspondences induced (via [Reference SternSte21]) by the 2-Segal object $\operatorname {\underline {\mathsf{Fl}}}_{X,\bullet }$ is, in fact, commutative.

5. The flag Grassmannian of a surface

5.1 Set-up and notation

We will fix once and for all a smooth reductive affine group scheme over $\mathbb{C}$ , and therefore we will simply write ${\bf Bun}$ and $\underline {\mathsf{Bun}}$ for ${\bf Bun}^{{\bf G}}$ or $\underline {\mathsf{Bun}}^{{\bf G}}$ , respectively (see Construction 3.7 for the difference between the bold and the underlined notation). We let ${\rm Gpd}_\infty$ denote the $\infty$ -category of $\infty$ -groupoids (i.e. spaces or, equivalently, simplicial sets), and its objects will be often simply called groupoids (rather than $\infty$ -groupoids).

For a fibre functor $\mathfrak{F}$ over a locally Noetherian (derived) scheme $Y$ (Definition 2.10), and a (derived) $Y$ -scheme $V \to Y$ , we write

(5.1) \begin{align} \begin{aligned} {\bf Bun}_{{\mathfrak{F}}}(V)= {\rm Bun}({\mathfrak{F}}(V)):= \mathop{\rm holim}_{\substack {{\rm Spec} R \to V \\ R \ {\rm Noetherian}}} {\rm Bun}({\mathfrak{F}}(R)). \end{aligned} \end{align}

In particular, we call ${\rm Bun}({\mathfrak{F}}(Y))$ the global sections of the stack ${\bf Bun}_{{\mathfrak{F}}}$ and, in order to simplify our notation, we will often write simply ${\rm Bun}({\mathfrak{F}})$ instead of ${\rm Bun}({\mathfrak{F}}(Y))$ for these global sections.

We identify, even notationally, (Lemma 2.13) a derived scheme $V \to Y$ over $Y$ with its fibre functor ${\mathfrak{F}}_V$ sending ${\rm Spec} R \to Y$ to ${\rm Spec} R \times _Y V$ .

Definition 5.1. Let ${\mathfrak{F}}_1$ and ${\mathfrak{F}}_2$ be fibre functors over $Y$ . We denote by ${\mathfrak{F}}_1 \cap {\mathfrak{F}}_2$ (and call it intersection) the fibre functor

\begin{align*} {\mathfrak{F}}_1 \cap {\mathfrak{F}}_2 := {\mathfrak{F}}_1 \mathop{\times} _{{\mathfrak{F}}_Y} {\mathfrak{F}}_2 \;:\; S \mapsto {\mathfrak{F}}_1(S) \mathop{\times} _S {\mathfrak{F}}_2(S). \end{align*}

We denote by ${\mathfrak{F}}_1 \cup {\mathfrak{F}}_2$ (and call it the union) the fibre functor

\begin{align*} {\mathfrak{F}}_1 \cup {\mathfrak{F}}_2 := {\mathfrak{F}}_1 \mathop{\amalg} _{{\mathfrak{F}}_1 \cap {\mathfrak{F}}_2} {\mathfrak{F}}_2 \;:\; S \mapsto {\mathfrak{F}}_1(S) \mathop{\amalg}_{{\mathfrak{F}}_1(S) \cap {\mathfrak{F}}_2(S)} {\mathfrak{F}}_2(S). \end{align*}

These notation are abusive, as we do not require the structural morphisms ${\mathfrak{F}}_i(S) \to S$ to be immersions. We will thus restrict their use to situations where said structural morphisms are wannabe immersions (even though they might not formally be).

Let $Y$ be a locally Noetherian derived scheme and $Z\subset T$ be closed (derived) subschemes in $Y$ . Recall from §§ 2 and 3, the derived stacks in $\mathsf{dSt}_Y$

\begin{align*} & & \mathsf{dAff}_{/Y} & \longrightarrow {\rm Gpd}_\infty \\ & {\bf Bun}_Y \;:\; & (U \to Y) & \longmapsto {\rm Bun}(U),\\ & {\bf Bun}_{Y \smallsetminus T} \;:\; & (U \to Y) & \longmapsto {\rm Bun}(U \smallsetminus T_U) ,\\ & {\bf Bun}_{\widehat {T}} \;:\; & (U \to Y) & \longmapsto {\rm Bun}(\widehat T_U),\\ & {\bf Bun}_{\widehat T^{{\rm aff}}} \;:\; & (U \to Y) & \longmapsto {\rm Bun}({\mathfrak{F}}^{{\rm aff}}_{\widehat {T}}(U))= {\rm Bun}(\widehat T_U^{{\rm aff}}) ,\\ & {\bf Bun}_{\widehat {T}\smallsetminus T} \;:\; & (U \to Y) & \longmapsto {\rm Bun}({\mathfrak{F}}_{\widehat {T}\smallsetminus T}(U))= {\rm Bun}( \widehat T_U^{{\rm aff}} \smallsetminus T_U ),\\ & {\bf Bun}_{{\widehat {Z}}} \;:\; & (U \to Y) & \longmapsto {\rm Bun}({\widehat {Z}}_U),\\ & {\bf Bun}_{{\widehat {Z}}^{{\rm aff}}} \;:\; & (U \to Y) & \longmapsto {\rm Bun}({\mathfrak{F}}^{{\rm aff}}_{{\widehat {Z}}}(U))= {\rm Bun}({\widehat {Z}}_U^{{\rm aff}} ),\\ & {\bf Bun}_{(Y\smallsetminus T) \cap {\widehat {Z}}^{{\rm aff}} } \;:\; & (U \to Y) & \longmapsto {\rm Bun}((U\smallsetminus T_U) \times _U {\widehat {Z}}_U^{{\rm aff}}),\\ & {\bf Bun}_{(\widehat {T}\smallsetminus T) \cap {\widehat {Z}}^{{\rm aff}} } \;:\; & (U \to Y) & \longmapsto {\rm Bun}((\widehat T_U^{{\rm aff}} \smallsetminus T_U) \times _U {\widehat {Z}}_U^{{\rm aff}}), \end{align*}

where $T_U := T \times _Y U$ , $Z_U := Z \times _Y U$ .

We have obvious restriction maps in $\mathsf{dSt}_Y$

(5.2) \begin{align} \begin{gathered} \begin{aligned} {\bf Bun}_{Y} & \longrightarrow {\bf Bun}_{Y \smallsetminus T}, & \hspace {4em} {\bf Bun}_{Y} & \longrightarrow {\bf Bun}_{{\widehat {Z}}^{{\rm aff}} } \\ {\bf Bun}_{Y\smallsetminus T} & \longrightarrow {\bf Bun}_{(Y\smallsetminus T) \cap {\widehat {Z}}^{{\rm aff}} }, & {\bf Bun}_{{\widehat {Z}}^{{\rm aff}}} & \longrightarrow {\bf Bun}_{(Y\smallsetminus T) \cap {\widehat {Z}}^{{\rm aff}} } \\ {\bf Bun}_{{\widehat {Z}}^{{\rm aff}}} & \longrightarrow {\bf Bun}_{(\widehat {T}\smallsetminus T) \cap {\widehat {Z}}^{{\rm aff}} }, & {\bf Bun}_{\widehat T^{{\rm aff}}} & \longrightarrow {\bf Bun}_{\widehat {T}\smallsetminus T} \end{aligned} \\ {\bf Bun}_{\widehat {T}\smallsetminus T} \longrightarrow {\bf Bun}_{(\widehat {T}\smallsetminus T) \cap {\widehat {Z}}^{{\rm aff}} }. \end{gathered} \end{align}

Moreover, algebraization (Proposition 3.1) gives equivalences

\begin{align*} {\bf Bun}_{{\widehat {Z}}^{{\rm aff}} } \stackrel {\sim }{\longrightarrow } {\bf Bun}_{{\widehat {Z}}}, \quad {\bf Bun}_{\widehat T^{{\rm aff}} } \stackrel {\sim }{\longrightarrow } {\bf Bun}_{\widehat T}. \end{align*}

5.2 The flag Grassmannian of a surface

In the next definition we use the notation established in § 5.1.

Definition 5.2. Let $Y$ be a locally Noetherian scheme and $Z\subset T$ be closed subschemes in $Y$ :

  • we will call Grassmannian of $Y$ relative to $(T, Z)$ both

    and

  • we put ${\rm Gr}(Y;T,Z):= \Gamma (Y , {\bf Gr}(Y; T, Z)) = \Gamma ({\rm Spec} \mathbb{C} , \underline {\mathsf{Gr}} (Y; T, Z)) \in {\rm Gpd}_\infty .$

Remark 5.3. Since the functor $u^{-1}: \mathsf{dSt}_Y \to \mathsf{dSt}_k$ is exact, we have $u({\bf Gr}(Y; T, Z))\simeq \underline {\mathsf{Gr}}(Y; T, Z)$ .

Remark 5.4. It is useful to notice that the definition of ${\bf Gr}(Y; T, Z)$ , and hence of $\underline {\mathsf{Gr}}(Y; T, Z)$ does not require that $Z$ is a closed subscheme of $T$ , but just that $T$ and $Z$ are, possibly unrelated, closed subschemes of $Y$ .

Remark 5.5. Let $U:={\rm Spec} R \to Y$ , with $R$ Noetherian (and not derived), and consider the following pullback diagrams.

Denoting by $\mathcal E_0$ the trivial bundle (whatever the base), then ${\bf Gr}(Y; T, Z)(U)$ has:

  • objects $(\mathcal{E}, \varphi , \psi )$ , where $\mathcal{E}\in {\rm Bun}(U)$ , $\varphi \;:\; \mathcal{E}_{| U\smallsetminus T_U} \simeq \mathcal{E}_0$ is an isomorphism in ${\rm Bun}(U\smallsetminus T_U)$ , and $\psi \;:\; \mathcal{E}_{| {\rm Spec} (\widehat {R}_J)} \simeq \mathcal{E}_0$ an isomorphism in ${\rm Bun}({\rm Spec} (\widehat {R}_J))$ , such that $\varphi _{| (U\smallsetminus T_U) \times _U {\rm Spec} (\widehat {R}_J) } = \psi _{| (U\smallsetminus T_U) \times _U {\rm Spec} (\widehat {R}_J) }$ ;

  • morphisms $(\mathcal{E}, \varphi , \psi ) \to (\mathcal{E}', \varphi ', \psi ')$ are (iso)morphisms $\alpha \;:\; \mathcal{E} \to \mathcal{E}'$ such that the following diagrams commute:

Remark 5.6. One can replace, in Definition 5.2, the stack of ${\bf G}$ -bundles with the stack of perfect complexes (respectively, with the stack of almost perfect complexes), and the trivial ${\bf G}$ -bundle on $Y$ with a fixed perfect complex (respectively, a fixed almost perfect complex) $\mathcal{E}_0$ on $Y$ ; we thus obtain Perf-versions (respectively, ${\rm Coh}^{-}$ -versions) of the flag Grassmannian that may be denoted by ${\bf Gr}_{\mathcal{E}_0}^{{\rm Perf}}(Y; T, Z)$ and $\underline {\mathsf{Gr}}_{\mathcal{E}_0}^{{\rm Perf}}(Y; T, Z)$ (respectively, ${\bf Gr}_{\mathcal{E}_0}^{{\rm Coh}^{-}}(Y; T, Z)$ and $\underline {\mathsf{Gr}}_{\mathcal{E}_0}^{{\rm Coh}^{-}}(Y; T, Z)$ ).

As in the previous section, let $X$ be a complex projective smooth algebraic surface, $S$ an arbitrary Noetherian affine test scheme over $\mathbb{C}$ . In this context, Definition 5.2 specializes to the following.

Definition 5.7. For an arbitrary flag $(D,Z) \in \operatorname {\underline {\mathsf{Fl}}}_{X}(S)$ :

  • we will call the Grassmannian relative to the pair $(S,(D,Z))$ both

    \begin{align*} {\bf Gr}_X (S)(D, Z) &:= {\bf Gr}(X\times S; D, Z)\\& = {\bf Bun}_{X \times S} \mathop{\times} _{{\bf Bun}_{(X \times S \smallsetminus D) \cup {\widehat {Z}}^{{\rm aff}}}} \{{\rm trivial}\} \in \mathsf{dSt}_{X \times S} \subset {\rm Fun}\big(\mathsf{dAff}_{X \times S}^{{\rm op}}, {\rm Gpd}_\infty \big), \end{align*}

    and

    \begin{align*} \underline {\mathsf{Gr}}_X (S)(D, Z) &:= \underline {\mathsf{Gr}}(X\times S; D, Z)\\& = \underline {\mathsf{Bun}}_{X \times S} \mathop{\times} _{\underline {\mathsf{Bun}}_{(X \times S\smallsetminus D) \cup {\widehat {Z}}^{{\rm aff}}}} \{{\rm trivial}\} \in \mathsf{dSt}_k \subset {\rm Fun}\big(\mathsf{dAff}_k^{{\rm op}}, {\rm Gpd}_\infty \big); \end{align*}
  • we define

    \begin{align*}{\rm Gr}_X(S)(D,Z):= \Gamma (X\times S , {\bf Gr}_X (S)(D, Z))= \Gamma ({\rm Spec} \mathbb{C}, \underline {\mathsf{Gr}}_X (S)(D, Z)) \in {\rm Gpd}_\infty .\end{align*}

We also give a local version of Definitions 5.2 and 5.7.

Definition 5.8 (Local Grassmannian). Let $Y$ be a locally Noetherian scheme, and $(Z\subset T \subset Y)$ be closed subschemes. Denote by ${\widehat {Z}} \smallsetminus T$ (or ${\mathfrak{F}}_{{\widehat {Z}} \smallsetminus T}$ ) the fibre functor

\begin{align*} {\widehat {Z}} \smallsetminus T := (Y \smallsetminus T) \cap {\widehat {Z}}^{{\rm aff}} \simeq \widehat T \smallsetminus T \mathop{\times} _{\widehat T^{{\rm aff}}} {\widehat {Z}}^{{\rm aff}}, \end{align*}

and by ${\widehat {Z}} \cup _{\widehat T} \widehat T \smallsetminus T$ the fibre functor

\begin{align*} {\widehat {Z}} \cup _{\widehat T} \widehat T \smallsetminus T := {\widehat {Z}}^{{\rm aff}} \mathop{\amalg} _{{\widehat {Z}} \smallsetminus T} \widehat T \smallsetminus T. \end{align*}

We call the local Grassmannian associated with $Z \subset T \subset Y$ both the following derived stacks:

If $S$ is a test scheme and $(D, Z) \in \operatorname {\underline {\mathsf{Fl}}}_X(S)$ , we also set

\begin{align*} {\bf Gr}^{{\rm loc}}_X(S)(D,Z) := {\bf Gr}^{{\rm loc}}(X \times S; D, Z) \hspace {2em} {\rm and} \hspace {2em} \underline {\mathsf{Gr}}^{{\rm loc}}_X(S)(D,Z) := \underline {\mathsf{Gr}}^{{\rm loc}}(X \times S; D, Z). \end{align*}

The following result is a flag Grassmannian version of the well-known equivalence between the local and the global affine Grassmannian of a curve (see, e.g., [Reference ZhuZhu17, Theorem 1.4.2]).

Lemma 5.9. Let $Y$ be a locally Noetherian scheme, and let $Z\subset T \subset Y$ be closed subschemes. There are canonical restriction equivalences

If $(D, Z) \in \operatorname {\underline {\mathsf{Fl}}}_X(S)$ , they specialize to equivalences

\begin{align*} {\bf Gr}^{{\rm loc}}_X(S)(D,Z) \simeq {\bf Gr}_X(S)(D,Z) \hspace {2em}{\it and}\hspace {2em} \underline {\mathsf{Gr}}^{{\rm loc}}_X(S)(D,Z) \simeq \underline {\mathsf{Gr}}_X(S)(D,Z). \end{align*}

Proof. It is enough to prove the statement for ${\bf Gr} (Y; T, Z)$ , since then the statement for $\underline {\mathsf{Gr}} (Y; T, Z)$ follows by applying the holim-preserving functor $\mathsf{dSt}_Y \to \mathsf{dSt}_k$ . The special statement for $\underline {\mathsf{Gr}}_X(S)(D, Z)$ then follows.

Consider now the following commutative diagram.

By definition of $Y \smallsetminus T \cup {\widehat {Z}}^{{\rm aff}}$ and ${\widehat {Z}} \cup _{\widehat T} \widehat T \smallsetminus T$ , the square $(\delta )$ and the bottom rectangle (formed by $(\gamma )$ and $(\delta )$ ) are Cartesian. It follows that $(\gamma )$ is Cartesian as well. The vertical rectangle formed by $(\beta )$ and $(\gamma )$ is also Cartesian by formal glueing Theorem 3.5. We deduce that $(\beta )$ itself is Cartesian. The square $(\alpha )$ is Cartesian as well, by definition of the Grassmannian. We conclude that the upper rectangle formed by $(\alpha )$ and $(\beta )$ is also Cartesian, so that

\begin{align*} {\bf Gr}(Y; T, Z) \simeq {\bf Bun}_{\widehat T^{{\rm aff}}} \mathop{\times} _{{\bf Bun}_{{\widehat {Z}} \cup _{\widehat T} \widehat T \smallsetminus T}} \{{\rm trivial}\} =: {\bf Gr}^{{\rm loc}}(Y; T, Z). \end{align*}

By taking global sections over $Y$ (or $X \times S$ ), we immediately get the following.Footnote 21

Corollary 5.10. Let $Y$ be a locally Noetherian scheme, and $(Z \subset T \subset Y)$ be closed subschemes. There is a natural equivalence of groupoids

\begin{align*}{\rm Gr}(Y; T, Z) \simeq {\rm Bun}(\widehat T) \times _{{\rm Bun}(\widehat T \smallsetminus T) \times _{{\rm Bun}((\widehat T \smallsetminus T) \times _{\widehat T^{{\rm aff}}} {\widehat {Z}}^{{\rm aff}})} {\rm Bun}({\widehat {Z}})} \,\, \{{\rm trivial}\} =: {\rm Gr}^{{\rm loc}}(Y; T,Z) .\end{align*}

In particular, for $(D,Z) \in \operatorname {\underline {\mathsf{Fl}}}_X(S)$ , there is a natural bijection

\begin{align*}{\rm Gr}_X(S)(D, Z) \simeq {\rm Bun}(\widehat {D}) \times _{{\rm Bun}(\widehat {D} \smallsetminus D) \times _{{\rm Bun}((\widehat {D} \smallsetminus D) \times _{{\hat {{D}}}^{{\rm aff}}} {\widehat {Z}}^{{\rm aff}})} {\rm Bun}({\widehat {Z}})} \,\, \{{\rm trivial}\} =: {\rm Gr}^{{\rm loc}}_X(S)(D,Z).\end{align*}

Remark 5.11. Continuing Remark 5.6, there are obvious local versions of ${\bf Gr}_{\mathcal{E}_0}^{{\rm Perf}}(Y; T, Z)$ , $\underline {\mathsf{Gr}}_{\mathcal{E}_0}^{{\rm Perf}}(Y; T, Z)$ , ${\bf Gr}_{\mathcal{E}_0}^{{\rm Coh}^{-}}(Y; T, Z)$ and $\underline {\mathsf{Gr}}_{\mathcal{E}_0}^{{\rm Coh}^{-}}(Y; T, Z)$ (analogues of Definition 5.8). Moreover, since only formal gluing (Theorem 3.5) has been used in the proof of Lemma 5.9, a similar ‘local=global’ statement holds for these Perf and ${\rm Coh}^{-}$ versions.

Lemma 5.12. Let $Y$ be a locally Noetherian scheme and $Z\subset T$ be closed subschemes in $Y$ . If $Y\setminus T$ is quasi-compact and schematically dense in $Y$ ,Footnote 22 then the truncation ${\rm t}_0(\underline {\mathsf{Gr}}(Y; T, Z))$ is actually a sheaf of sets. In particular, its global sections ${\rm Gr}(Y;T,Z)$ form a set.

Proof. Let $S={\rm Spec} R$ be a Noetherian affine scheme. We will use the notation of Remark 5.5. We have to prove that $\underline {\mathsf{Gr}}(Y; T, Z)(S)$ is a set (i.e. equivalent to a discrete groupoid). Now the obvious functor $\underline {\mathsf{Gr}}(Y; T, Z)(S) \to \underline {\mathsf{Gr}}(Y; T, \emptyset )(S)$ sending $(\mathcal{E}, \varphi , \psi )$ to $(\mathcal{E}, \varphi )$ (and identical on morphisms) is faithful, therefore it will be enough to show that $\underline {\mathsf{Gr}}(Y; T, \emptyset )(S)$ is a set. Define $T_S:= T \times _Y (S\times Y)$ . Observe that, since being quasi-compact open and scheme-theoretically dense is stable under flat pullbacks [Sta13, TAG081I] and $S\times Y \to Y$ is flat, we have that $(Y\setminus T) \times _Y (S \times Y)= (S \times Y)\setminus T_S$ is quasi-compact and schematically dense inside $S \times Y$ .

Let $(\mathcal{E}, \varphi ), (\mathcal{E'}, \varphi ') \in \underline {\mathsf{Gr}}(Y; T, \emptyset )(S)$ , where $\mathcal{E}\to S \times Y$ is a ${\bf G}$ -bundle, $\varphi$ a trivialization of $\mathcal{E}$ on $(S \times Y) \setminus T_S$ (and similarly for $(\mathcal{E'}, \varphi ')$ ). Let $\alpha , \beta \;:\; (\mathcal{E}, \varphi ) \to (\mathcal{E'}, \varphi ')$ be morphisms (hence isomorphisms) in the groupoid $\underline {\mathsf{Gr}}(Y; T, \emptyset )(S)$ . Since a groupoid is equivalent to a set if and only if its Hom sets are either empty or consist of a singleton, it will be enough to show that $\alpha =\beta$ . Consider $\beta ^{-1}\circ \alpha \;:\; (\mathcal{E}, \varphi ) \to (\mathcal{E}, \varphi )$ and ${\rm id}: (\mathcal{E}, \varphi ) \to (\mathcal{E}, \varphi )$ . We want to prove $\beta ^{-1}\circ \alpha = {\rm id}$ . Observe that ${\rm Aut}_{{\rm Bun}(S\times Y)}(\mathcal{E})$ injects into ${\rm Hom}_{{\bf Sch}_{\mathbb{C}}}(\mathcal{E}, {\bf G})$ (the image being ${\bf G}$ -equivariant maps). Let us denote by $\rho _{\alpha , \beta }: \mathcal{E} \to {\bf G}$ , and $\rho _{{\rm id}}: \mathcal{E} \to {\bf G}$ the images of $\beta ^{-1}\circ \alpha$ , and of ${\rm id}$ inside ${\rm Hom}_{{\bf Sch}_{\mathbb{C}}}(\mathcal{E}, {\bf G})$ . It will be enough to show that $\rho _{\alpha , \beta }= \rho _{{\rm id}}$ .

We know, by definition of morphisms in the groupoid $\underline {\mathsf{Gr}}(Y; T, \emptyset )(S)$ , that $\beta ^{-1}\circ \alpha$ and ${\rm id}$ agree when restricted to $\mathcal{E}_{|\,(S\times Y) \setminus T_S}$ . Therefore, $\rho _{\alpha , \beta }$ and $\rho _{{\rm id}}$ coincide on

\begin{align*} \mathcal{E} \times _{S\times Y} ((S \times Y)\setminus T_S) = \mathcal{E} \setminus (\mathcal{E}\times _{S\times Y} T_S). \end{align*}

Now, since any ${\bf G}$ -bundle is (faithfully) flat over its base, $\mathcal{E} \to S\times Y$ is flat, $\mathcal{E} \setminus (\mathcal{E}\times _{S\times Y} T_S)$ is quasi-compact open and schematically dense in $\mathcal{E}$ (again by [Sta13, TAG081I]). Now, ${\bf G}$ is a separated scheme, and the two maps $\rho _{\alpha , \beta }, \,\rho _{{\rm id}} \;:\; \mathcal{E} \to {\bf G}$ coincide on the quasi-compact open and schematically dense $\mathcal{E} \setminus (\mathcal{E}\times _{S\times Y} T_S)$ , therefore they coincide on all of $\mathcal{E}$ [Sta13, TAG01RH].

Here is an immediate (recalling Definition 5.7) corollary of the previous lemma.

Corollary 5.13. Let $X$ be a complex projective smooth algebraic surface, $S$ an arbitrary Noetherian affine test scheme over $\mathbb{C}$ . Then, for an arbitrary flag $(D,Z) \in \operatorname {\underline {\mathsf{Fl}}}_{X}(S)$ , the truncation ${\rm t}_0(\underline {\mathsf{Gr}}_X (S)(D, Z))$ is actually a sheaf of sets. In particular, its global sections ${\rm Gr}_X(S)(D,Z)$ form a set.

5.3 Functorialities and flat connection

We prove here the functorialities of ${\rm Gr}_X(S)(F)$ both in the flag $F$ , and in the test scheme $S$ . See Corollary 5.18 below. We fix a smooth projective surface $X$ over $k$ .

To any affine scheme $S$ over $k$ and any closed subschemes $D, Z \subset X \times S_{{\rm red}}$ , we associate two (ind-)schemes flat over $Y := X \times S$

\begin{align*} Y \smallsetminus D \to Y \hspace {1em}{\rm and}\hspace {1em} {\widehat {Z}} \to Y. \end{align*}

They induce underived fibre functorsFootnote 23 $\tilde {\mathfrak{F}}_{Y \smallsetminus D}$ and $\tilde {\mathfrak{F}}_{{\widehat {Z}}}$ over $Y$ . We denote by $\tilde {\mathfrak{F}}_{Y \smallsetminus D \cap {\widehat {Z}}}$ the intersection

\begin{align*} \tilde {\mathfrak{F}}_{Y \smallsetminus D \cap {\widehat {Z}}} := \tilde {\mathfrak{F}}_{Y \smallsetminus D} \mathop{\times} _{Y} \tilde {\mathfrak{F}}_{{\widehat {Z}}}^{{\rm aff}}. \end{align*}

Let $\Lambda$ be the category . We denote by $\Lambda ^{D,Z}_Y$ the diagram of fibre functors

Notice that for any morphism of Noetherian affine schemes $S' \to S$ , pulling back along the induced morphism $f \;:\; Y' = X \times S' \to X \times S=Y$ gives, by (an underived version of) Lemma 2.19, a canonical isomorphism of diagrams

(5.3)

where $D' = D \times _{S_{{\rm red}}} S'_{{\rm red}}$ and $Z' = Z \times _{S_{{\rm red}}} S'_{{\rm red}}$ .

Finally, the assignment $(D, Z) \mapsto \Lambda ^{D,Z}_Y$ is covariant in $Z$ and contravariant in $D$ . In particular, given a commutative diagram (of closed subschemes)

we get a co-correspondence (of diagrams)

Let $\underline {\mathsf{Cl}}_{X,{\rm dR}}(S)$ denoteFootnote 24 the category of closed immersions $Z \hookrightarrow D \hookrightarrow X \times S_{{\rm red}}$ (and morphisms being commutative diagrams as above). It is cumbersome but easy to check that this construction gives a laxFootnote 25 functor to the $2$ -category of correspondences in diagrams of underived fibre functors

Using (5.3) and Lemma 2.19(ii), those assemble into a natural transformation (by lax functors)

Composition with the inclusion $i \;:\; \mathsf{FibF}_{X \times -} \to \mathsf{dFibF}_{X \times -}$ of fibre functors into derived fibre functors, and using the canonical morphisms $i{\mathfrak{F}}_0 \mathop{\amalg} _{i{\mathfrak{F}}_1} i{\mathfrak{F}}_2 \to i({\mathfrak{F}}_0 \mathop{\amalg} _{{\mathfrak{F}}_1} {\mathfrak{F}}_2)$ , we get a natural transformation (by lax $\infty$ -functors, see [Reference Gaitsgory and RozenblyumGR17a, Chapter 10 § 3.])

Taking pointwise pushouts of our $\Lambda$ -indexed diagrams (of derived fibre functors) yields a natural transformation (by lax $\infty$ -functors)

which evaluates, for $S \in \mathsf{Aff}$ and $Y := X \times S$ , to

(5.4)

Lemma 5.14. Fix $S$ a test scheme and $Y := X \times S$ . Let $(D_1, Z_1) \to (D_2, Z_2) \in \underline {\mathsf{Cl}}_{X,{\rm dR}}(S)$ . If the canonical morphism $Z_1 \to Z_2 \times _{D_2} D_1$ is a homeomorphism (i.e. induces an isomorphism between the associated reduced schemes), then the canonical morphism

\begin{align*} {\mathfrak{F}}_{Y \smallsetminus D_1 \cup {\widehat {Z}}_1} \to {\mathfrak{F}}_{Y \smallsetminus D_1 \cup {\widehat {Z}}_2} \end{align*}

induces a equivalences of derived stacks over $Y$

\begin{align*} {\bf Perf}_{Y \smallsetminus D_1 \cup {\widehat {Z}}_2} & \overset \sim \longrightarrow {\bf Perf}_{Y \smallsetminus D_1 \cup {\widehat {Z}}_1}, \\ {\bf Bun}^{{\bf G}}_{Y \smallsetminus D_1 \cup {\widehat {Z}}_2} & \overset \sim \longrightarrow {\bf Bun}^{{\bf G}}_{Y \smallsetminus D_1 \cup {\widehat {Z}}_1} \\ {\it and}\hspace {1em} {\bf Coh}^-_{Y \smallsetminus D_1 \cup {\widehat {Z}}_2} & \overset \sim \longrightarrow {\bf Coh}^-_{Y \smallsetminus D_1 \cup {\widehat {Z}}_1}. \end{align*}

Proof. Consider the following commutative diagram (of categorical derived stacks over $Y$ ).

The square $(\sigma )$ is Cartesian by definition. Fixing a derived test scheme $T$ over $Y$ , we can apply Proposition 3.3 to $D_1 \times _Y {\mathfrak{F}}_{{\widehat {Z}}_2}^{{\rm aff}}(T) \subset {\mathfrak{F}}_{{\widehat {Z}}_2}^{{\rm aff}}(T)$ . We deduce that $(\tau )$ is Cartesian as well. The large rectangle is thus Cartesian and the result follows (with a similar argument for the cases of $\mathbf G$ -bundles or of complexes of coherent sheaves).

Let $\mathcal{M} \;:\; Y \mapsto \mathcal{M}_Y$ denote either ${\bf Perf}$ , ${\bf Bun}^{{\bf G}}$ or ${\bf Coh}^-$ . Denote by ${\mathcal{M}}{/}\underline {\mathsf{dSt}}^{{\rm Cat}_\infty }_X$ the stack (in $\infty$ -categories) of derived stacks in $\infty$ -categories under $\mathcal{M}$

\begin{align*} {\mathcal{M}}{/}\underline {\mathsf{dSt}}^{{\rm Cat}_\infty }_X \;:\; S \mapsto {\mathcal{M}_{X \times S}}{/}\mathsf{dSt}^{{\rm Cat}_\infty }_{X \times S} .\end{align*}

The stack $\mathcal{M}$ induces a morphism categorical prestacks

Lemma 5.15. Composing the lax $\infty$ -functor ${\mathfrak{F}}^\cup \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}} \to \mathsf{dFibF}_{X \times -}^{{\rm cocorr}}$ with ${\rm M}$ gives a morphism (still denote by $\mathrm M$ ) of prestacks in $\infty$ -categories

Proof. This follows from the fact that for any $Y$ , the functor $\mathcal{M}_Y \;:\; \mathsf{dFibF}_Y^{{\rm op}} \to \mathsf{dSt}^{{\rm Cat}_\infty }_Y$ maps pushout squares to pullback squares.

Definition 5.16. For any test affine scheme $S$ , let $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}}(S)$ the (non-full) subcategory of $\underline {\mathsf{Cl}}_{X,{\rm dR}}(S)$ containing every object, but keeping only the morphisms $(D_1, Z_1) \to (D_2, Z_2)$ such that $Z_1 \to Z_2 \times _{D_2} D_1$ is a homeomorphism (i.e. an isomorphism on the associated reduced schemes). This condition being stable under base change, this defines a sub-prestack $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}}$ of $\underline {\mathsf{Cl}}_{X,{\rm dR}}$ .

Theorem 5.17. The composition

factors through $\underline {\mathsf{dSt}}^{{\rm Cat}_\infty }_X \subset \left (\left (\underline {\mathsf{dSt}}^{{\rm Cat}_\infty }_X\right )^{{\rm op}}\right )^{{\rm cocorr}}$ , therefore inducing

Proof. Fix a test scheme $S \in \mathsf{Aff}_k$ . From Lemma 5.14 and (5.4), we get:

  1. (i) for each $S \in \mathsf{Aff}_k$ , the lax $\infty$ -functor $\mathrm M_S \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}}(S) \to \left (\left ({\mathcal{M}_{X \times S}}{/}\mathsf{dSt}^{{\rm Cat}_\infty }_{X \times S}\right )^{{\rm op}}\right )^{{\rm cocorr}}$ is strict (i.e. the morphisms making the lax structure are equivalences);

  2. (ii) For each $S \in \mathsf{Aff}_k$ , the $\infty$ -functor $\mathrm M_S$ has values in the (non-full) sub- $(\infty ,1)$ -category ${\mathcal{M}_{X \times S}}{/}\mathsf{dSt}^{{\rm Cat}_\infty }_{X \times S}$ (embedded the standard way into the category of cocorrespondences).

In particular, the map $\mathrm M$ factors as $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}} \to {\underline {\mathcal{M}}_X}{/}\underline {\mathsf{dSt}}^{{\rm Cat}_\infty }_X$ giving the announced functor.

Corollary 5.18. The flag Grassmannians ${\bf Gr}(X \times S; D, Z)$ (respectively $\underline {\mathsf{Gr}}(X \times S; D, Z)$ ) assemble into a morphism

\begin{align*} {\bf Gr}_X \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}} \to \underline {\mathsf{dSt}}_X \subset \underline {\mathsf{dSt}}^{{\rm Cat}_\infty }_X \hspace {1cm}(\rm respectively\ \underline {\mathsf{Gr}}_X \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}} \to \underline {\mathsf{dSt}}_k) \end{align*}

of stacks $(\mathsf{Aff}_k)^{{\rm op}} \to {\rm Cat}_\infty$ .

Proof. Consider the morphism

Taking pointwise the homotopy fibre of the map ${\bf Bun}^{{\bf G}}_{X \times S}\left (X \times S\right ) \to {\bf Bun}^{{\bf G}}_{X \times S}({\mathfrak{F}}_{X \times S \smallsetminus D \cup {\widehat {Z}}})$ gives the announced functoriality of the Grassmannian $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}} \to \underline {\mathsf{dSt}}^{{\rm Cat}_\infty }_X$ . This functor has values in the substack of derived stacks in $\infty$ -groupoids $\underline {\mathsf{dSt}}_X$ .

The morphism $\underline {\mathsf{Gr}}_X$ is obtained from ${\bf Gr}_X$ be composing with the natural morphism $\underline {\mathsf{dSt}}_X \to \underline {\mathsf{dSt}}_k$ mapping a derived stack over $X \times S$ to its derived stack of relative global sections.

Remark 5.19.

  1. (a) The result of Corollary 5.18, and in particular the fact that the flag Grassmannian can be seen as a functor out of the de Rham stack of closed pairs $(D,Z)$ , means our Grassmannian has a crystal structure (which, by analogy with the linear set-up, intuitively amounts to a flat connection; see also [Reference Beilinson and DrinfeldBD99, 2.8.2]). This fact should be interpreted as an analogue of the fact the usual Beilinson–Drinfeld Grassmannian over some curve $X$ is a (factorizing) $\mathcal{D}_X$ -scheme in [Reference Beilinson and DrinfeldBD04].

  2. (b) Using Theorem 5.17, together with remarks 5.6 and 5.11, one can straightforwardly define functorial flags Grassmannians ${\bf Gr}_{X, \mathcal{E}_0}^{{\rm Perf}}$ , $\underline {\mathsf{Gr}}_{X,\mathcal{E}_0}^{{\rm Perf}}$ , ${\bf Gr}_{X,\mathcal{E}_0}^{{\rm Coh}^{-}}$ and $\underline {\mathsf{Gr}}_{X,\mathcal{E}_0}^{{\rm Coh}^{-}}$ for perfect or coherent complexes. Although our focus is on $\mathbf G$ -bundles in this section, we will keep this level of generality in the upcoming technical lemmas (as it does not impact the complexity of the arguments and may be useful for further work in this direction).

5.4 Topological invariance of the flag Grassmannian

We prove here that the flag Grassmannian is, in an appropriate sense, insensitive to non-reduced structures on flags. We already know it carries a flat connection. With the following lemma, we add that it is invariant under reduced equivalence.

Lemma 5.20. Let $S$ be a test scheme and $Z \subset D \subset X \times S_{{\rm red}}$ be closed subschemes. The canonical diagram

defines a morphism $\alpha$ in $\underline {\mathsf{Cl}}_{X,{\rm dR}}(S)$ . The cocorrespondence (of diagrams of fibre functors over $Y := X \times S$ )

is an isomorphism (i.e. both $\overrightarrow \tau _\alpha$ and $\overleftarrow \tau _\alpha$ are invertible).

Proof. The (ind-)schemes $Y \smallsetminus D$ and $\widehat {Z}$ only depend on the underlying closed subsets of $D$ and $Z$ , respectively. As a consequence, the same is true of the fibre functors $\tilde {\mathfrak{F}}_{Y \smallsetminus D}$ and $\tilde {\mathfrak{F}}_{\widehat {Z}}^{{\rm aff}}$ . The result follows.

Corollary 5.21. For any $(D,Z) \in \underline {\mathsf{Cl}}_{X,{\rm dR}}(S)$ , there is a canonical equivalence

\begin{align*} {\bf Gr}(X \times S; D, Z) \simeq {\bf Gr}(X \times S; D_{{\rm red}}, Z_{{\rm red}}) \in \mathsf{dSt}_{X \times S}. \end{align*}

The following Corollary establishes the topological invariance of the flag Grassmannian: the values of $\underline {\mathsf{Gr}}_X (S)$ on two reduced equivalent flags agree (up to a canonical isomorphism).

Corollary 5.22. Let $X$ be a complex projective smooth algebraic surface, $S$ an arbitrary Noetherian affine scheme over $\mathbb{C}$ , and $Y=X \times S$ . If $(D,Z), (D', Z') \in \operatorname {\underline {\mathsf{Fl}}}_{X}(S)$ are such that $D_{{\rm red}}= D'_{{\rm red}}$ and $Z_{{\rm red}}= Z'_{{\rm red}}$ , then there is a canonical equivalence

\begin{align*} {\bf Gr}_X(S)(D, Z) &:= {\bf Gr}(Y; D, Z) \simeq {\bf Gr}(Y; D_{{\rm red}}, Z_{{\rm red}})\\& = {\bf Gr}(Y; D'_{{\rm red}}, Z'_{{\rm red}}) \simeq {\bf Gr}(Y; D', Z') =: {\bf Gr}_X(S)(D', Z'). \end{align*}

Remark 5.23.

  1. (a) It would be tempting, by Corollary 5.22, to consider the quotient of $\operatorname {\underline {\mathsf{Fl}}}_{X}(S)$ modulo the reduced equivalence relation, and then define our Grassmannian as a functor out of this quotient. However, this would not be so helpful for our purposes, since unfortunately the corresponding version of $\operatorname {\underline {\mathsf{Fl}}}^{\bullet , {\rm good}}_{X}(S)$ modulo reduced equivalence will no longer be $2$ -Segal.

  2. (b) As in Remark 5.19(b), the above results (corollaries 5.21 and 5.22) can be trivially generalized to perfect or coherent complexes, i.e. to ${\bf Gr}_{\mathcal{E}_0}^{{\rm Perf}}(X\times S; -,-)$ and ${\bf Gr}_{\mathcal{E}_0}^{{\rm Coh}^{-}}(X \times S; -, -)$ .

5.5 Functorialities of the local flag Grassmannian

Since by Lemma 5.9, the local Grassmannian ${\bf Gr}_X^{{\rm loc}}$ is equivalent to the global one ${\bf Gr}_X$ , we get covariant functorialities

\begin{align*} {\bf Gr}_X^{{\rm loc}} \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}} \to \underline {\mathsf{dSt}}_X \hspace {1em} {\rm and} \hspace {1em} \underline {\mathsf{Gr}}_X^{{\rm loc}} \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}} \to \underline {\mathsf{dSt}}_k. \end{align*}

There is, however, a more interesting and less intuitive contravariant functoriality, mapping a morphism in $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}}(S)$ corresponding to a reduced-Cartesian commutative square

to a pullback morphism $\alpha ^* \;:\; {\bf Gr}^{{\rm loc}}_X(Y; D_2, Z_2) \to {\bf Gr}^{{\rm loc}}_X(Y; D_1, Z_1)$

(5.5)

One could prove the above pullback morphisms are functorial enough to get a morphism

\begin{align*} ({\bf Gr}_X^{{\rm loc}}, (-)^*) \;:\; (\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}})^{{\rm op}} \to \underline {\mathsf{dSt}}_X. \end{align*}

We will, however, be more interested in the compatibility between the pullback along one morphism of flags and the covariant functoriality we established so far. This compatibility can be thought as a base-change formula, as usual only valid under additional assumptions – see Remark 5.25(a) below.

Let us construct this pullback formally, and prove it is partially compatible with the covariant functoriality. Denote by ${\boxminus } := \Lambda^2_0 \times \Delta^1$ the category depicted as

Let $\mathsf{dFibF}_{X \times -}^{\boxminus }$ denote the stack of $\boxminus$ -diagrams of fibre functors, and by $(\mathsf{dFibF}_{X \times -}^{\boxminus })^{{\rm cocorr}}$ the $(\infty , 2)$ -categorical stack of cocorrespondences therein. Informally, objects of $(\mathsf{dFibF}_{X \times S}^{\boxminus })^{{\rm cocorr}}$ are $\boxminus$ -diagrams of fibre functors, and morphisms are cocorrespondences of diagrams, thus commutative diagrams of the following shape.

Finally, $2$ -morphisms are maps of $\boxminus$ -diagrams between the hats of the cocorrespondences involved (together with the obvious commutativity constraints).

Denote by $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{\Delta^1}$ the stack of arrows in $\underline {\mathsf{Cl}}_{X,{\rm dR}}$ . Using the same method as in the construction of ${\mathfrak{F}}^\cup$ (see (5.4)), we obtain a natural transformation by lax $\infty$ -functors ${\mathfrak{F}}_{{\rm loc}}^\cup \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{\Delta^1} \to (\mathsf{dFibF}_{X \times S}^{\boxminus })^{{\rm cocorr}}$ evaluating, on $S \in \mathsf{Aff}_k$ , to

Similarly to Lemma 5.15, composing ${\mathfrak{F}}^\cup _{{\rm loc}}$ with ${\bf Bun}^{{\bf G}}$ yields a natural transformation by lax $\infty$ -functors

\begin{align*} \underline {\mathsf{Cl}}_{X,{\rm dR}}^{\Delta^1} \to (\mathsf{dFibF}_{X \times S}^{\boxminus })^{{\rm cocorr}} \to (((\underline {\mathsf{dSt}}_X)^{{\rm op}})^{\boxminus })^{{\rm cocorr}} ,\end{align*}

mapping $(S,\alpha \;:\; (D_1, Z_1) \to (D_2, Z_2))$ to the following commutative diagram.

Taking the homotopy fibres of the horizontal morphisms over the trivial bundle, and setting $Y := X \times S$ and ${\bf Gr}_X^{{\rm loc}}(\alpha ) := {\bf Bun}^{{\bf G}}_{\widehat D_1} \times _{{\bf Bun}^{{\bf G}}_{\widehat D_1 \smallsetminus D_2 \cup {\widehat {Z}}_1}} \{{\rm trivial}\}$ , we get the cocorrespondence in $\mathsf{dSt}_Y$ depicted in (5.5),

By functoriality of homotopy fibres, we get a natural transformation by lax $\infty$ -functors

\begin{align*} G \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{\Delta^1} \to (\mathsf{dFibF}_{X \times S}^{\boxminus })^{{\rm cocorr}} \to (((\underline {\mathsf{dSt}}_X)^{{\rm op}})^{\boxminus })^{{\rm cocorr}} \to (((\underline {\mathsf{dSt}}_X)^{\Lambda^2_2})^{{\rm op}})^{{\rm cocorr}} .\end{align*}

Consider now the stack in $(\infty ,1)$ -categories $(\underline {\mathsf{dSt}}_X)^{\Delta^1}$ . It embeds naturally in the stack in $(\infty ,2)$ -categories $(((\underline {\mathsf{dSt}}_X)^{\Lambda^2_2})^{{\rm op}})^{{\rm cocorr}}$ , via the assignment

Theorem 5.24. Denote by $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1_{{\rm t}}}$ the categorical prestack mapping $S \in \mathsf{Aff}_k$ to the (non-full) sub-category of the category $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}}(S)^{\Delta^1}$ containing every object, but only morphisms corresponding to commutative squares

such that Footnote 26 $\overline {D_2 \smallsetminus D_1} \cap D_1' = \overline {D_2 \smallsetminus D_1} \cap Z_1'$ (as closed topological subspaces of $X \times S$ ). Denote by $G^{{\rm h}}$ the restriction of $G$ to the substack $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1_{{\rm t}}}$ of $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{\Delta^1}$ .

  1. (a) For any $S \in \mathsf{Aff}_k$ , the lax structure on $G^{{\rm h}}(S)$ is strict, and $G^{{\rm h}}$ is thus an honest morphism of $\infty$ -categorical stacks.

  2. (b) The morphism $G^{{\rm h}}$ has values in the substack $(\underline {\mathsf{dSt}}_X)^{\Delta^1}$ of $(((\underline {\mathsf{dSt}}_X)^{\Lambda^2_2})^{{\rm op}})^{{\rm cocorr}}$ , thus inducing a morphism

    \begin{align*} {\bf Gr}^{{\rm loc},\Delta^1}_X \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1_{{\rm t}}} \to (\underline {\mathsf{dSt}}_X)^{\Delta^1}. \end{align*}

Remark 5.25.

  1. (a) The main consequence of the above theorem is the existence of a pullback as announced at the beginning of this section: for any morphism $(D_1, Z_1) \to (D_2, Z_2)$ , seen as an object of $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1}(S)$ , we get the pullback morphism ${\bf Gr}_X^{{\rm loc}}(X \times S; D_2, Z_2) \to {\bf Gr}_X^{{\rm loc}}(X \times S; D_1, Z_1)$ announced in (5.5). Its functoriality implies the following base-change formula:

  2. (b) When restricting to degenerated arrows $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}} \subset \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1_{{\rm t}}}$ , we find back the covariant functoriality ${\bf Gr}_X^{{\rm loc}} \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h}} \to \underline {\mathsf{dSt}}_X$ without relying on the equivalence with the global Grassmannian.

  3. (c) As in Remarks 5.19(b) and 5.23(b), the above theorem extends mutatis mutandis to a version of the flag Grassmannian for perfect complexes and almost perfect complexes.

The proof of Theorem 5.24 is very similar to that of Theorem 5.17, and will rely on two crucial Lemmas 5.30 and 5.31, the proof of which requires some preliminary results.

5.5.1 $\mathcal M$ -coCartesian squares

Definition 5.26. A commutative square of $\infty$ -categories

is $1$ -Cartesian if the induced $\infty$ -functor $A \to B \times _D C$ is fully faithful.

Lemma 5.27. Consider three commutative squares $(\sigma )$ , $(\sigma ')$ and $(\sigma '')$ of $\infty$ -categories and morphisms $(\sigma ) \to (\sigma ') \leftarrow (\sigma '')$ . Denote by $(\tau )$ the square obtained as the pullback $(\sigma ) \times _{(\sigma ')} (\sigma '')$ . We represent these data as commutative diagrams.

If both $(\sigma )$ and $(\sigma '')$ are Cartesian, and if $(\sigma ')$ is $1$ -Cartesian, then $(\tau )$ is Cartesian. If $(\sigma )$ , $(\sigma ')$ and $(\sigma '')$ are all $1$ -Cartesian, so is $(\tau )$ .

Proof. Denote by $f$ the canonical morphism $f \;:\; A \to B \times _D C$ . We also set $f'$ and $f''$ similarly, and

\begin{align*} g \;:\; A \mathbin {\underset {A'}{\times }} A'' \to B \mathbin {\underset {B'}{\times }} B'' \underset {D \mathbin {\underset {D'}{\times }} D''}{\times } C \mathbin {\underset {C'}{\times }} C''. \end{align*}

The result follows from the assumptions together with the trivial observation that $g = f \times _{f'} f''$ .

Definition 5.28. Fix a commutative square $(\sigma )$ of derived fibre functors over a base $Y$ and $\mathcal M$ a derived stack over $Y$ . Denote by $(\sigma )_{\mathcal M}$ the associated commutative square of derived pre-stacks over $Y$

  1. (i) We say that $(\sigma )$ is $\mathcal M$ -coCartesian if $(\sigma )_{\mathcal M}$ is Cartesian (i.e. is pointwise Cartesian).

  2. (ii) We say that $(\sigma )$ is $\mathcal M$ - $1$ -coCartesian if $(\sigma )_{\mathcal M}$ is pointwise $1$ -Cartesian.

A number of our previous results can be formulated using the notion of ${\bf Perf}$ -, ${\bf Coh}^-$ - or ${\rm B}{\bf G}$ -coCartesian squares (e.g. Proposition 3.3). We will need this consequence of Lemma 2.24(a).

Lemma 5.29. Let $(\sigma )$ be a commutative square in $\mathsf{FibF}_Y$ for some base $Y$

If $(\sigma )$ is ${\bf Perf}$ - $1$ -coCartesian, then $(\sigma )$ is ${\rm B}{\bf G}$ - $1$ -coCartesian.

Proof. Denote by $\mathfrak{F}$ the pushout ${\mathfrak{F}}_2 \mathop{\amalg} _{{\mathfrak{F}}_1} {\mathfrak{F}}_3$ computed in $\mathsf{dFibF}_{Y}$ , and by $F \;:\; {\mathfrak{F}} \to {\mathfrak{F}}_4$ the induced morphism. By assumption, for any $S$ over $Y$ , the $\infty$ -functor ${\rm Perf}({\mathfrak{F}}_4(S)) \to {\rm Perf}({\mathfrak{F}}(S))$ is fully faithful. Using Lemma 2.24(a), we deduce that ${\bf Bun}^{{\bf G}}_{{\mathfrak{F}}_4}(S) = {\rm Bun}^{{\bf G}}({\mathfrak{F}}_4(S)) \to {\rm Bun}^{{\bf G}}({\mathfrak{F}}(S)) = {\bf Bun}^{{\bf G}}_{{\mathfrak{F}}}(S)$ is also fully faithful.

5.5.2 Two keys lemmas After these preliminaries, we are ready to prove the key lemmas allowing for the proof of Theorem 5.24.

Lemma 5.30. Let $S \in \mathsf{Aff}_k$ and $Y := X \times S$ . Consider a diagram of closed subschemes of $Y$

and assume $\overline {D_2 \smallsetminus D_1} \cap D_1' = \overline {D_2 \smallsetminus D_1} \cap Z_1'$ (as topological spaces). Then the natural morphism of derived fibre functors

induces equivalences of derived stacks over $Y$

Proof. For a derived test scheme $T$ over $X \times S$ , we fix the notation

\begin{align*} \widehat D_{1,T}^{\prime {\rm aff}} := {\mathfrak{F}}^{{\rm aff}}_{\widehat D_1'}(T), \hspace {2em} {\widehat {Z}}_{1,T}^{\prime {\rm aff}} := {\mathfrak{F}}^{{\rm aff}}_{{\widehat {Z}}_1'}(T). \end{align*}

We also by $D_1^T$ , $D_2^T$ and $C$ the closed subschemes of $\widehat D_{1,T}^{\prime {\rm aff}}$ induced respectively by the closed subschemes $D_1$ , $D_2$ andFootnote 27 $\overline {D_2 \smallsetminus D_1}$ of $X \times S$ . Because $\overline {D_2 \smallsetminus D_1} \cap D_1' = \overline {D_2 \smallsetminus D_1} \cap Z_1'$ , the formal completion of $\widehat D_{1,T}^{\prime {\rm aff}}$ at $C$ is canonically equivalent to ${\widehat {Z}}_{1,T}^{\prime {\rm aff}}$ . Applying [Reference LurieLur18, 7.4.1.1] to $A = {\mathbb{R}}\Gamma (\widehat D_{1,T}^{\prime {\rm aff}}, \mathcal O)$ , $B = {\mathbb{R}}\Gamma ({\widehat {Z}}_{1,T}^{\prime {\rm aff}}, \mathcal O)$ and $\mathcal C = {\rm QCoh}(\widehat D_{1,T}^{\prime {\rm aff}} \smallsetminus D_1^T)$ , we deduce after passing to dualizable objects, that the natural square

is Cartesian. It follows that the square

is ${\bf Perf}$ -coCartesian. By definition of the fibre functors ${\mathfrak{F}}_{\widehat D_1' \smallsetminus D_1 \cup {\widehat {Z}}_1'}$ and ${\mathfrak{F}}_{\widehat D_1' \smallsetminus D_2 \cup {\widehat {Z}}_1'}$ , we get a commutative square

in which:

  • both lateral faces are coCartesian in $\mathsf{dFibF}_{X \times S}$ ; and

  • the back square is ${\bf Perf}$ -coCartesian.

This implies that the front square is ${\bf Perf}$ -coCartesian as well, so that ${\bf Perf}_{\widehat D_1' \smallsetminus D_2 \cup {\widehat {Z}}_1'} \to {\bf Perf}_{\widehat D_1' \smallsetminus D_1 \cup {\widehat {Z}}_1'}$ is an equivalence as claimed. The case of ${\bf Bun}^{{\bf G}}$ follows by using Lemma 2.24(b).

Lemma 5.31. Let $S \in \mathsf{Aff}_k$ and $Y := X \times S$ . Consider a diagram of closed subschemes of $Y$

such that $Z_1 = D_1 \cap Z_1'$ and $\overline {D_2 \smallsetminus D_1} \cap D_1' = \overline {D_2 \smallsetminus D_1} \cap Z_1'$ as topological subspaces of $Y$ . The commutative diagram of fibre functors

is ${\bf Perf}$ -coCartesian (and thus also ${\rm B}{\bf G}$ -coCartesian).

Proof. The square $(\sigma )$ fits in the commutative diagram below.

Using Lemma 5.30, we see that $g$ induces an equivalence once composed with ${\bf Perf}$ . Moreover, using $Z_1 = D_1 \cap Z_1'$ and $\overline {D_2 \smallsetminus D_1} \cap D_1' = \overline {D_2 \smallsetminus D_1} \cap Z_1'$ , we deduce the equality (of topological subspaces) $\overline {D_2 \smallsetminus D_1} \cap D_1 = \overline {D_2 \smallsetminus D_1} \cap Z_1$ . Applying Lemma 5.30 once more, to the case $D_1' = D_1$ , implies that $f$ induces an equivalence as well, once compose with ${\bf Perf}$ .

In particular, the square $(\sigma )$ is ${\bf Perf}$ -coCartesian if and only if the outer square is. We may thus assume $D_2 = D_1$ without lose of generality.

Under this new assumption, the square $(\sigma )$ decomposes as follows.

It thus suffices to prove that $(\tau )$ is ${\bf Perf}$ -coCartesian and that $\psi$ induces an equivalence ${\bf Perf}(\psi )$ . Consider the following commutative diagram.

For formal reasons, the upper square is coCartesian. Moreover, the outer square is ${\bf Perf}$ -coCartesian by Proposition 3.3. It follows that $(\tau )$ is ${\bf Perf}$ -coCartesian. We now focus on the morphism $\psi$ . We have a commutative diagram

in which the lateral squares are coCartesian by definition. In particular, the morphism ${\bf Perf}(\psi )$ is an equivalence if and only if the upper face is ${\bf Perf}$ -coCartesian. Since $Z_1 = D_1 \cap Z_1'$ , for any derived affine scheme $T$ over $Y = X \times S$ , the affinization of the formal completion of ${\mathfrak{F}}_{\widehat Z_1'}^{{\rm aff}}(T)$ at the closed subscheme induced by $D_1 \times _Y T$ is canonically equivalent to ${\mathfrak{F}}_{{\widehat {Z}}_1}^{{\rm aff}}(T)$ . In particular, Proposition 3.3 implies this upper square is indeed ${\bf Perf}$ -coCartesian.

5.5.3 Proof of Theorem 5.24

Proof of Theorem 5.24. Let $S \in \mathsf{Aff}_k$ . We start with the lax $\infty$ -functor

\begin{align*} G^{{\rm h}} \;:\; \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1_{{\rm t}}}(S) \to (((\underline {\mathsf{dSt}}_X)^{\Lambda^2_2})^{{\rm op}})^{{\rm cocorr}}(S) = (((\mathsf{dSt}_{X \times S})^{\Lambda^2_2})^{{\rm op}})^{{\rm cocorr}}. \end{align*}

Fix a morphism of $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1_{{\rm t}}}(S)$ , corresponding to a commutative diagram

such that $\overline {D_2 \smallsetminus D_1} \cap D_1' = \overline {D_2 \smallsetminus D_1} \cap Z_1'$ . Its image by ${\mathfrak{F}}^\cup _{{\rm loc}}(S)$ is the diagram.

By Lemma 5.31, the three squares $(\alpha )$ , $(\beta )$ and $(\gamma )$ are ${\rm B}{\bf G}$ -coCartesian. As a consequence, the image of any morphism of $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1_{{\rm t}}}(S)$ is a cocorrespondence (of $\Lambda^2_2$ -diagrams) where the morphism from ‘source to hat’ is an equivalence. It follows that the lax $\infty$ -functor $G^{{\rm h}}(S)$ is in fact a strict $\infty$ -functor with values in the (non-full) sub- $(\infty ,1)$ -category

\begin{align*} (\mathsf{dSt}_{X \times S})^{\Lambda^2_2} \subset (((\mathsf{dSt}_{X \times S})^{\Lambda^2_2})^{{\rm op}})^{{\rm cocorr}}. \end{align*}

Consider now an object $(D_1, Z_1) \to (D_2, Z_2)$ of $\underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1_{{\rm t}}}(S)$ . Its image by $G^{{\rm h}}(S)$ is the diagram (depicted in (5.5))

Lemma 5.30 implies that $u$ is an equivalence. In particular, the functor $G^{{\rm h}}(S)$ has values in the full subcategory $\mathsf{dSt}_{X \times S} \subset (\mathsf{dSt}_{X\times S})^{\Lambda^2_2}$ . This concludes the proof of Theorem 5.24.

6. The flag Grassmannian as a $2$ -Segal object

By Beilinson and Drinfeld, the classical affine Grassmannian (in dimension 1) carries an important factorization structure (see, e.g., [Reference ZhuZhu17]). In essence, it consists of the way a $\mathbf G$ -bundle on a curve trivialized outside two points is tantamount to two bundles, each one trivialized outside one of the points.

In this section, we will construct a similar factorization structure, where disjoint points are to be replaced with pairs of flags in good position. Part of the difficulties arising in this construction is to handle the higher coherences and homotopies the structure involves.

6.1 The projection maps ${\bf Gr}_X(S)(F_1 \cup F_2) \to {\bf Gr}_X(S)(F_i)$

We will construct a $2$ -Segal structure on the flag Grassmannian. The first step is to define functorial projection maps ${\bf Gr}_X(S)(F_1 \cup F_2) \to {\bf Gr}_X(S)(F_i)$ , $i=1,2$ , for any pair $(F_1, F_2) \in \operatorname {\underline {\mathsf{Fl}}}_{X,2}(S_{{\rm red}})$ . This starts with the simple observation: the natural inclusions $F_i \to F_1 \cup F_2$ are morphism in $\operatorname {\underline {\mathsf{Fl}}}_X(S_{{\rm red}})$ .

In particular, we get natural transformations

(6.1)

The natural transformations $a \;:\; \partial _0 \Rightarrow \partial _1$ and $b \;:\; \partial _2 \Rightarrow \partial _1$ each correspond to a morphism $\operatorname {\underline {\mathsf{Fl}}}_{X,2} \to \operatorname {\underline {\mathsf{Fl}}}_X^{\Delta^1}$ to the stack of arrows of flags. Taking the de Rham stack on each side, we denote by $\underline {a}$ and $\underline {b}$ the induced morphisms

Lemma 6.1. The inclusion $\operatorname {\underline {\mathsf{Fl}}}_X^{{\rm dR}} \to \underline {\mathsf{Cl}}_{X,{\rm dR}}$ induces an inclusion $\operatorname {\underline {\mathsf{Fl}}}_{X,{\rm dR}}^{\Delta^1} \to \underline {\mathsf{Cl}}_{X,{\rm dR}}^{{\rm h},\Delta^1_t}$ (recall the notation from Theorem 5.24 ).

Proof. This is straightforward point-set topology.

Using the morphism ${\bf Gr}^{{\rm loc},\Delta^1}_X$ from Theorem 5.24(b), we get

Remembering that ${\bf Gr}^{{\rm loc},\Delta^1}_X$ swaps source and target, each composition ${\bf Gr}^{{\rm loc},\Delta^1}_X \circ \underline {a}$ or ${\bf Gr}^{{\rm loc},\Delta^1}_X \circ \underline {b}$ amounts to one of the following natural transformations.

(6.2)

Evaluating at $S \in \mathsf{Aff}_k$ and at a pair of flags $(F_1, F_2) \in \operatorname {\underline {\mathsf{Fl}}}_{X,2}^{{\rm dR}}(S) := \operatorname {\underline {\mathsf{Fl}}}_{X,2}(S_{{\rm red}})$ , we get the announced functorial projection morphisms (in $\mathsf{dSt}_{X \times S}$ ):

\begin{align*} {\bf Gr}_X^{{\rm loc}}(S)(F_1 \cup F_2) \to {\bf Gr}_X^{{\rm loc}}(S)(F_i). \end{align*}

6.2 Factorization property

The goal of this section is to prove the following factorization formula.

Theorem 6.2 (Factorization). The natural transformations from (6.2) induce an equivalence

Remark 6.3.

  1. (a) Theorem 6.2 can be seen as a factorization structure on our flag Grassmannian. Indeed, after evaluating at a Noetherian derived affine scheme $S$ and at a good pair of flags $(F_1, F_2)$ in $X \times S_{{\rm red}}$ , the statement becomes that the constructed map

    is an equivalence.

  2. (b) Similarly to Remarks 5.19(b), 5.23(b) and 5.25(c), the factorization property of Theorem 6.2 extends to flag Grassmannians of perfect complexes, and the proof provided below also applies in this case.

The proof of Theorem 6.2 has prerequisites that we will address first. It is thus postponed until the end of this section.

Let $A \in {\bf sCAlg}_k$ be a simplicial commutative algebra over $k$ . For $f \in \pi _0 A$ , denote by $A^f$ the following (homotopy) pushout.

(6.3)

For any family $f_\bullet = (f_1, \ldots , f_n)$ of elements of $ \pi _0 A$ , we denote by $A^{f_\bullet }$ or $A^{f_1, \ldots , f_n}$ the tensor product $\bigotimes _A A^{f_i}$ . Note that $A^f$ is equivalent to the mapping cone of the morphism $A \to A$ given by the multiplication $f$ . More generally $A^{f_\bullet }$ is nothing but the Koszul complex associated with the sequence $(f_1, \ldots , f_n)$ .

Moreover, if $f_\bullet = ( f_1, \ldots , f_n ) \subset \pi _0 A$ and $p$ is a positive integer, we will use the notation $f_\bullet ^p = ( f_1^p, \ldots , f_n^p )$ . Notice that if $f,g \in \pi _0 A$ , there is a natural map $A^{fg} \to A^{f}$ (induced, in terms of pushout squares as in (6.3), by $x \mapsto gx$ ). In particular, for any $p \geq q$ and any family $f_\bullet$ as above, we get a canonical morphism $A^{f^p_\bullet } \to A^{f^q_\bullet }$ . The pro-diagram $(A^{f_\bullet ^p})_p$ corresponds, by [Reference Gaitsgory and RozenblyumGR14, Proposition 6.7.4], to an explicit derived affine ind-scheme computing the formal neighbourhood of $\{f_\bullet=0\}$ in ${\rm Spec} A$ .

Given two such families $f_\bullet = (f_1,\ldots , f_n)$ and $g_\bullet =(g_1,\ldots ,g_m)$ , we denote by $f_\bullet g_\bullet$ the family in $\pi _0 A$ consisting of all possible products between an element of $f_\bullet$ and an element of $g_\bullet$ , while $f_\bullet \cup g_\bullet$ denotes $(f_1,\ldots ,f_n,g_1,\ldots ,g_m)$ .

Lemma 6.4. For any $A \in {\bf sCAlg}_k$ and for any families $f_\bullet = (f_1, \ldots , f_n)$ and $g_\bullet = (g_1, \ldots , g_m)$ of elements of $\pi _0 A$ , there is a natural equivalence

\begin{align*} \lim _p A^{f_\bullet ^p g_\bullet ^p} \simeq \lim _p \left ( A^{f_\bullet ^p} \mathop{\times} _{ A^{f_\bullet ^p} \otimes _A A^{g_\bullet ^p}} A^{g_\bullet ^p} \right ). \end{align*}

Sublemma 6.5. Let $f, g \in \pi _0 A$ . There is a natural equivalence $A^{f,g} \simeq (A^f)^g$ (where in the right-hand side, $g$ is seen as an element in $\pi _0(A^f)$ ).

Proof. It follows from contemplating the following pushout diagrams.

Sublemma 6.6. Let $f, g \in \pi _0 A$ . There is a natural equivalence

(6.4) \begin{align} A^{fg} \simeq A^f \mathop{\times} _{A^{f,g}} A^g. \end{align}

Proof. To prove this sublemma, we will rely on explicit models in the (equivalent model) category of connective commutative dg-algebras. We thus assume that $A$ is now an explicit commutative dg-algebra $\cdots \to A^{-1} \to A^0 \to 0$ , and we fix lifts of $f$ and $g$ to $A^0$ . We will abusively denote those lifts by $f$ and $g$ as well.

The pushout (6.3) yields explicit models for $A^f$ as the cdga (quasi-free over $A$ )

\begin{align*} A[\varepsilon ] := \left (\cdots \to A^{-2} \oplus \varepsilon A^{-1} \to A^{-1} \oplus \varepsilon A^{0} \to A^0\right ) ,\end{align*}

with differential mapping $\varepsilon$ to $f$ . We get similar models $A[\tau ]$ for $A^g$ and $A[\varepsilon ,\tau ]$ for $A^{f,g}$ . Then the pullback on the right-hand side of (6.4) is the (homotopy) pullback follows.

(6.5) \begin{align} A[\varepsilon ] \mathop{\times} _{A[\varepsilon ,\tau ]} A[\tau ]. \end{align}

In order to compute this homotopy pullback, it suffices to resolve $A[\varepsilon ] \to A[\varepsilon ,\tau ]$ with a fibration. For example, consider the algebra $A[\varepsilon ,\bar {\varepsilon }, d\bar {\varepsilon }]$ , where the added generators $\bar {\varepsilon }, d\bar {\varepsilon }$ are in degree $-1$ and $0$ respectively, and the differential sends $\bar {\varepsilon }$ to $d\bar {\varepsilon }$ . It is straightforward to check that the assignments

\begin{align*} \varepsilon \mapsto \varepsilon , \ \ \ \bar {\varepsilon } \mapsto \tau , \ \ \ d\bar {\varepsilon } \mapsto g \end{align*}

define a fibration $A[\varepsilon ,\bar {\varepsilon }, d\bar {\varepsilon }] \to A[\varepsilon ,\tau ]$ , replacing $A[\varepsilon ] \to A[\varepsilon ,\tau ]$ .

It follows that an explicit model for the homotopy pullback (6.5) is given by the strict pullback

\begin{align*} B=A[\varepsilon ,\bar {\varepsilon }, d\bar {\varepsilon }] \mathop{\times} _{A[\varepsilon ,\tau ]} A[\tau ]. \end{align*}

Explicitly, the elements of $B$ are polynomials in $A[\varepsilon ,\bar {\varepsilon }, d\bar {\varepsilon }]$ such that, if evaluated in $\bar {\varepsilon }=\tau , d\bar {\varepsilon }=g$ , reduce to polynomials involving $\tau$ as unique variable. Said differently, the elements of $B$ are the polynomials in $A[\varepsilon ,\bar {\varepsilon }, d\bar {\varepsilon }]$ of the form

\begin{align*} \varepsilon (g-d\bar {\varepsilon }) q_1(\bar {\varepsilon }, d\bar {\varepsilon }) + q_2 (\bar {\varepsilon }, d\bar {\varepsilon }) ,\end{align*}

where $q_1,q_2 \in A[\bar {\varepsilon }, d\bar {\varepsilon }]$ . We now claim that $B$ is quasi-isomorphic to $A[z]$ , where $z$ is a degree $-1$ generator with differential $fg$ . In fact, setting $t=\varepsilon (g-d\bar {\varepsilon })+f\bar {\varepsilon }$ , we have that $B=A[t,\bar {\varepsilon },d\bar {\varepsilon }]$ , where now the differential of the generator $t$ is $fg$ . It is now easy to verify that the natural map $A[z] \to B$ sending $z$ to $t$ is a quasi-isomorphism.

Noticing that $A[z]$ is a model for $A^{fg}$ concludes the proof of the sublemma.

Proof of Lemma 6.4. We proceed by induction on $m + n$ . The case $m + n=2$ is either trivial (if either $m$ or $n$ vanish) or follows directly from Sublemma 6.6 in the case $m = n=1$ .

We now prove the inductive step. Let $f_\bullet = (f_1, \ldots , f_n)$ and $g_\bullet = (g_1, \ldots , g_m)$ be families of elements of $H^0( A)$ . Suppose without loss of generality that $m\gt 1$ , and denote by $\bar {g}_\bullet$ the family $(g_1, \ldots , g_{m-1})$ . Then Sublemma 6.5 yields

\begin{align*} \lim _p A^{f_\bullet ^p q_\bullet ^p } \simeq \lim _p ( A^{(f_\bullet ^pg_m^p)} )^{(f_\bullet ^p \bar {g}_\bullet ^p)}. \end{align*}

By finality of the diagonal, the last limit can also be written as

\begin{align*} \lim _p ( A^{(f_\bullet ^pg_m^p)} )^{(f_\bullet ^p \bar {g}_\bullet ^p)} \simeq \lim _{p,q} ( A^{(f_\bullet ^q g_m^q)} )^{(f_\bullet ^p \bar {g}_\bullet ^p)}.\end{align*}

Setting $B_q = A^{f_\bullet ^q g_m^q}$ , we can now use the inductive hypothesis to obtain an equivalence

\begin{align*} \lim _p A^{f_\bullet ^p q_\bullet ^p } \simeq \lim _{p,q} \left ( B_q \right )^{f_\bullet ^p \bar {g}_\bullet ^p} \simeq \lim _{p,q} \left ( B_q^{f_\bullet ^p} \mathop{\times} _{ B_q^{f_\bullet ^p} \otimes _{B_q} B_q^{\bar {g}_\bullet ^p}} B_q^{\bar {g}_\bullet ^p} \right ) \simeq \lim _{p} \left ( B_p^{f_\bullet ^p} \mathop{\times} _{ B_p^{f_\bullet ^p} \otimes _{B_p} B_p^{\bar {g}_\bullet ^p}} B_p^{\bar {g}_\bullet ^p} \right ) ,\end{align*}

where in the last identification we used once again a finality argument. In other words, invoking subLemma 6.5, we obtained a Cartesian square.

(6.6)

Notice, however, [Reference Gaitsgory and RozenblyumGR14, Proposition 6.7.4] implies that if $h_\bullet$ is a family in $\pi _0(A)$ , then $\lim _p A^{h_\bullet ^p}$ only depends on the ideal of $\pi _0(A)$ generated by the elements of $h_\bullet$ . Therefore, the Cartesian square (6.6) can be equivalently rewritten as follows.

(6.7)

But repeating the same arguments as before, we obtain that the bottom map of diagram (6.7) fits in a Cartesian square

(6.8)

and composing the two Cartesian squares (6.7) and (6.8) yields the desired statement.

Lemma 6.7. Let $Y$ be a Noetherian derived affine scheme and $Z_1, Z_2 \subset Y$ be closed subschemes. Let $Z_{12} := Z_1 \cap Z_2$ and $Z = Z_1 \cup Z_2$ . The canonical functor

\begin{align*} {\rm Perf}_{\widehat {Z}} \to {\rm Perf}_{\widehat Z_1} \mathop{\times}_{{\rm Perf}_{\widehat Z_{12}}} {\rm Perf}_{\widehat Z_2} \end{align*}

is an equivalence.

Proof. Let $A$ denote (the homotopy type of) the simplicial algebra of functions on $Y$ . Since $A$ is Noetherian, we can and do fix generators $f_\bullet = (f_1, \ldots , f_n)$ and $g_\bullet =(g_1, \ldots , g_m)$ of the ideals $I$ and $J$ of $\pi _0 A$ of functions vanishing on $Z_1$ and $Z_2$ respectively.

For any positive integer $p$ , we write $f_\bullet ^p$ for the family $(f_1^p, \ldots , f_n^p)$ (and similarly for $g_\bullet ^p$ ). By [Reference Gaitsgory and RozenblyumGR14, Proposition 6.7.4], we have

\begin{align*} \widehat Z_1 \simeq {\rm colim}_p {\rm Spec}(A^{f_\bullet ^p}), \hspace {2em} \widehat Z_2 \simeq {\rm colim}_p {\rm Spec}(A^{g_\bullet ^p}) \hspace {2em}{\rm and} \hspace {2em} \widehat Z_{12} \simeq {\rm colim}_p {\rm Spec}\left (A^{f_\bullet ^p} \otimes _A A^{g_\bullet ^p}\right ). \end{align*}

Notice that for every $p$ the morphisms

\begin{align*} A^{f^p_\bullet } \to A^{f_\bullet ^p} \otimes _A A^{g_\bullet ^p}, \ \ \ \ \ A^{g^p_\bullet } \to A^{f_\bullet ^p} \otimes _A A^{g_\bullet ^p}\end{align*}

are surjective on $H^0$ . Therefore, it follows from Theorems 16.2.0.1 and 16.2.3.1 of [Reference LurieLur18] that, for every positive integer $p$ , the square

is Cartesian. Taking the limit over $p$ yields an equivalence

\begin{align*} \lim _p {\rm Perf}_{A^{f_\bullet ^p} \mathop{\times} _{ A^{f_\bullet ^p} \otimes _A A^{g_\bullet ^p}} A^{g_\bullet ^p}} \simeq {\rm Perf}_{\widehat Z_1} \mathop{\times} _{{\rm Perf}_{\widehat Z_{12}}} {\rm Perf}_{\widehat Z_2},\end{align*}

and algebraization implies that

\begin{align*} \lim _p {\rm Perf}_{A^{f_\bullet ^p} \mathop{\times} _{ A^{f_\bullet ^p} \otimes _A A^{g_\bullet ^p}} A^{g_\bullet ^p}} \simeq {\rm Perf}_{\lim _{p} A^{f_\bullet ^p} \mathop{\times} _{ A^{f_\bullet ^p} \otimes _A A^{g_\bullet ^p}} A^{g_\bullet ^p}}. \end{align*}

Using Lemma 6.4, we obtain therefore that

(6.9) \begin{align} {\rm Perf}_{\lim _{p} A^{f^p_\bullet g_\bullet ^p}} \simeq {\rm Perf}_{\widehat Z_1} \mathop{\times} _{{\rm Perf}_{\widehat Z_{12}}} {\rm Perf}_{\widehat Z_2}. \end{align}

In virtue of [Reference Gaitsgory and RozenblyumGR14, Proposition 6.7.4], the limit $\lim _p A^{f^p_\bullet g_\bullet ^p}$ only depends on the radical of the ideal of $H^0(A)$ generated by the family $f_\bullet g_\bullet$ . But the elements of the family $f_\bullet g_\bullet$ generate the ideal $IJ$ , whose radical coincides with $I \cap J$ . Therefore,

\begin{align*} \widehat {Z} \simeq {\rm colim}_p {\rm Spec}(A^{f_\bullet ^p g_\bullet ^p}) ,\end{align*}

which, together with equivalence (6.9), concludes the proof.

From now on and up until the end of this section, we fix a test Noetherian affine scheme $S$ , and $(F_1, F_2)$ a good pair of flags in $X \times S$ . We write $Y = X \times S$ and $F_i = (D_i, Z_i)$ .

Lemma 6.8. The following squares of fibre functors are both ${\bf Perf}$ -coCartesian and ${\rm B}{\bf G}$ -coCartesian (see Definition 5.28 ).

Proof. After evaluating the fibre functors involved on a derived affine scheme, this becomes a straightforward application of Lemma 6.7 and Lemma 2.24(b).

Lemma 6.9. The following commutative squares of fibre functors are ${\bf Perf}$ - $1$ -coCartesian and ${\rm B}{\bf G}$ - $1$ -coCartesian (see Definition 5.28 ).

Proof. We fix a test scheme $T \in \mathsf{Aff}_Y$ . We denote by $A$ the ring of functions of the affine ind-scheme ${\mathfrak{F}}_{\widehat {D_1 + D_2}}(T)$ , by $A_i$ that of ${\mathfrak{F}}_{\widehat {D_i}}(T)$ for $i=1,2$ and by $A_{12}$ that of ${\mathfrak{F}}_{\widehat {D_1 \cap D_2}}(T)$ . By Lemma 6.7, we have $A \simeq A_1 \times _{A_{12}} A_2$ . Denote by $U$ the open subscheme $U := {\mathfrak{F}}_{\widehat {D_1 + D_2} \smallsetminus D_1 + D_2}(T)$ of ${\mathfrak{F}}_{\widehat {D_1 + D_2}}^{{\rm aff}}(T) = {\rm Spec}(A)$ . Finally, we denote by $U_i$ the intersection $U_i := U \times _{{\rm Spec} A} {\rm Spec} A_i$ , for $i=1, 2$ or $12$ . Because $U \to {\rm Spec} A$ is flat, we get a (homotopy) pullback square

in the derived category of $U$ . This implies (by [Reference LurieLur18]) that the canonical functor

is fully faithful. By Lemma 5.29, we deduce that $(\sigma _D^\circ )$ is ${\rm B}{\bf G}$ - $1$ -coCartesian. The case of $(\sigma ^\circ _Z)$ is isomorphic, up to replacing $A$ with the ring of functions on $\widehat {Z_1 \cup Z_2}$ , and changing $U$ accordingly.

Corollary 6.10. The following commutative square of fibre functors is ${\bf Perf}$ - $1$ -coCartesian (and thus ${\rm B}{\bf G}$ - $1$ -coCartesian as well).

Proof. The commutative squares of fibre functors from Lemmas 6.8 and 6.9 are related by canonical morphisms $(\sigma _Z) \leftarrow (\sigma _Z^\circ ) \to (\sigma _D^\circ )$ . The pushout of this diagram of squares is nothing but the square $(\tau )$ . Applying Lemma 5.27 (pointwise) to the induced diagram $(\sigma _Z)_{{\bf Perf}} \to (\sigma _Z^\circ )_{{\bf Perf}} \leftarrow (\sigma _D^\circ )_{{\bf Perf}}$ , using Lemmas 6.8 and 6.9, implies that $(\tau )_{{\bf Perf}}$ is $1$ -Cartesian as announced.

We can now conclude our proof of Theorem 6.2.

Proof of Theorem 6.2. Consider the morphisms of commutative squares $(\sigma _D)_{{\rm B}{\bf G}} \to (\tau )_{{\rm B}{\bf G}} \leftarrow (*)$ , where $(*)$ is the constant square with value $*$ , and the map $(*) \to (\tau )_{{\rm B}{\bf G}}$ selects the trivial bundles. Denote by $(\alpha )_{\textrm {Gr}}$ the pullback $(\sigma _D)_{{\rm B}{\bf G}} \times _{(\tau )_{{\rm B}{\bf G}}} (*)$ . Explicitly, $(\alpha )_{\textrm {Gr}}$ is the upper square in the commutative diagram

where the four vertical sequences are fibre sequences. Indeed, the leftmost sequence is a fibre sequence by Lemma 5.9; the two middle ones by Lemma 5.30. As for the rightmost sequence, it is a fibre sequence since the pair of flags is assumed to be good, so that $\widehat {D_1 \cap D_2} = \widehat {Z_1 \cap Z_2}$ .

Finally, we know that $(\sigma _D)_{\mathrm B {\bf G}}$ is a Cartesian square (Lemma 6.8) and that $(\tau )_{\mathrm B {\bf G}}$ is $1$ -Cartesian (Corollary 6.10). Since $(*)$ is obviously Cartesian, we get by Lemma 5.27 that $(\alpha )_{{\rm Gr}}$ is Cartesian. In particular, we have

6.3 The Grassmannian as a fibred stack

In order to better handle the homotopy coherent factorization structure the Grassmannian naturally carries, we will see the Grassmannian as a fibred (derived) stack. As before, we fix $X$ a smooth projective surface over $k$ .

Recall that the flag Grassmannian is a natural transformation $\underline {\mathsf{Gr}}_X \simeq \underline {\mathsf{Gr}}_X^{{\rm loc}} \;:\; \operatorname {\underline {\mathsf{Fl}}}_X^{{\rm dR}} \to \underline {\mathsf{dSt}}_k$ between functors $\mathsf{Aff}_k^{{\rm op}} \to {\rm Cat}_\infty$ . The stack of derived stacks $\underline {\mathsf{dSt}}_k$ is endowed with a canonical derived structure, and thus comes from a functor $\underline {\mathsf{dSt}}_k \;:\; \mathsf{dAff}_k^{{\rm op}} \to {\rm Cat}_\infty$ . Consider the left Kan extension of $\operatorname {\underline {\mathsf{Fl}}}_X$ to derived affine schemes, also denoted by $\operatorname {\underline {\mathsf{Fl}}}_X$ (so equipped with a trivial derived structure), and consider its de Rham stack $\operatorname {\underline {\mathsf{Fl}}}_X^{{\rm dR}}$ . The flag Grassmannian becomes a natural transformation $\underline {\mathsf{Gr}}_X \simeq \underline {\mathsf{Gr}}_X^{{\rm loc}} \;:\; \operatorname {\underline {\mathsf{Fl}}}_X^{{\rm dR}} \to \underline {\mathsf{dSt}}_k$ between functors $\mathsf{dAff}_k \to {\rm Cat}_\infty$ .

Definition 6.11. By the universal property of the (categorical) stack of stacks, we get a pointwise coCartesian fibration

\begin{align*} p \;:\; \mathcal{G}r_X \to \operatorname {\underline {\mathsf{Fl}}}_X^{{\rm dR}} \in \mathsf{dSt}^{{\rm Cat}_\infty }_k ,\end{align*}

whose fibre $p_S^{-1}(D,Z)$ at $S \in \mathsf{dAff}_k$ and $(D, Z) \in \operatorname {\underline {\mathsf{Fl}}}_X(S_{{\rm red}})$ is the $\infty$ -groupoid $\Gamma (S, {\bf Gr}(S; D, Z))$ .

Remark 6.12. Let $S\in \mathsf{Aff}$ (i.e. a Noetherian underived affine scheme). Then, $\mathcal{G}r_X(S)$ is the category whose objects are pairs $(F=(D,Z), \underline {\mathcal{E}})$ , where $F \in \operatorname {\underline {\mathsf{Fl}}}_{X}(S_{{\rm red}})$ and $\underline {\mathcal{E}} \in {\rm Gr}_X(S)(F)=\Gamma (S, {\bf Gr}(S; D, Z))$ , with morphisms $(F, \underline {\mathcal{E}}) \to (F', \underline {\mathcal{E}}')$ given by pairs $(i, \alpha )$ of morphisms $i: F \to F'$ in $\operatorname {\underline {\mathsf{Fl}}}_{X}(S_{{\rm red}})$ , and $\alpha \;:\; {\rm Gr}_X(S)(i)(\underline {\mathcal{E}}) \simeq \underline {\mathcal{E}}'$ . The composition is straightforward. Moreover, $p_S\;:\; \mathcal{G}r_X(S) \to \operatorname {\underline {\mathsf{Fl}}}_{X}(S_{{\rm red}})$ sends $(F, \underline {\mathcal{E}})$ to $F$ , and $(i, \alpha ): (F, \underline {\mathcal{E}}) \to (F', \underline {\mathcal{E}}')$ to $i: F\to F'$ . We leave to the reader the further explicit description of functoriality of $\mathcal{G}r_X(S)$ with respect to morphisms $S' \to S$ in $\mathsf{Aff}$ .

6.4 The simplicial flag Grassmannian

Definition 6.13. We denote by $\mathcal{G}r_{X,2} \to \operatorname {\underline {\mathsf{Fl}}}_{X,2}^{{\rm dR}}$ the fibred derived stack obtained by pulling back $\mathcal{G}r_X \to \operatorname {\underline {\mathsf{Fl}}}_X^{{\rm dR}}$ along $\partial _1 \;:\; \operatorname {\underline {\mathsf{Fl}}}_{X,2}^{{\rm dR}} \to \operatorname {\underline {\mathsf{Fl}}}_X^{{\rm dR}}$ .

Remark 6.14. Under the equivalence between fibred stacks and morphisms to the stack of stacks, the fibred stack $\mathcal{G}r_{X,2} \to \operatorname {\underline {\mathsf{Fl}}}_{X,2}^{{\rm dR}}$ corresponds to the composite

\begin{align*} \underline {\mathsf{Gr}}_X \circ \partial _1 \;:\; \operatorname {\underline {\mathsf{Fl}}}_{X,2}^{{\rm dR}} \to \underline {\mathsf{dSt}}_k. \end{align*}

Together with Lemma 5.9, the natural transformations of (6.2) induce a commutative diagram.

Since both degeneracy morphisms $\sigma _0,\sigma _1 \;:\; \operatorname {\underline {\mathsf{Fl}}}_X \to \operatorname {\underline {\mathsf{Fl}}}_{X,2}$ satisfy $\partial _1 \circ \sigma _i = {{\rm id}}$ , they trivially induce degeneracy morphisms $\sigma _0,\sigma _1 \;:\; \mathcal{G}r_X \to \mathcal{G}r_{X,2}$ . All in all, we get a truncated simplicial object.

Note moreover that since $\operatorname {\underline {\mathsf{Fl}}}_{X,\bullet }$ is a $2$ -Segal object, and since the de Rham functor preserves fibre products, the simplicial object $\operatorname {\underline {\mathsf{Fl}}}_{X,\bullet }^{{\rm dR}}$ is also $2$ -Segal. Applying Proposition A.1 to $X_\bullet = \operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }^{{\rm dR}}$ and $Y^{\leq 2}_\bullet = \mathcal{G}r_{X, \leq 2}$ , we get a 2-Segal simplicial object $\mathcal{G}r_{X, \bullet }$ in ${\rm \mathrm{Fun}}(\mathsf{dAff}_k^{{\rm op}}, {\rm Cat}_\infty )$ together with a simplicial map $p_\bullet \;:\; \mathcal{G}r_{X, \bullet } \to \operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }^{{\rm dR}}$ . Notice that by construction, the derived stack $\mathcal{G}r_{X,n}$ can be obtained as the following pullback.

Definition 6.15. We will refer to $\mathcal{G}r_{X, \bullet }$ as the $2$ -Segal simplicial flag Grassmannian.

Remark 6.16. Using Remark 4.22 one can easily verify that the associative algebra in correspondences induced (via [Reference SternSte21]) by the 2-Segal object $\mathcal{G}r_{X, \bullet }$ is, in fact, commutative (i.e. $E_{\infty }$ ).

Remark 6.17. With this simplicial point of view, factorization (Theorem 6.2) can be reinterpreted as: the following square

is Cartesian.

Remark 6.18. Factorization (Theorem 6.2 or Remark 6.17) allows for the following, equivalent, description of the simplicial flag Grassmannian. Define $\widetilde {\mathcal{G}r}_{X,2}$ as the pullback

then use Theorem 6.2 to show that there is a further induced map $\tilde {\partial }_1: \widetilde {\mathcal{G}r}_{X,2} \to \mathcal{G}r_{X}$ satisfying the hypothesis of Proposition A.1, so that Proposition A.1 produces for us a simplicial and 2-Segal object $\widetilde {\mathcal{G}r}_{X,\bullet }$ together with a map to $\operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }$ . Using again Proposition A.1 for $X_\bullet = \mathcal{G}r_{X,\bullet }$ and $Y_\bullet ^{\leq 2}= \widetilde {\mathcal{G}r}_{X,\bullet }$ (and vice versa), we get induced maps, in both directions, between $\widetilde {\mathcal{G}r}_{X,\bullet }$ and $\mathcal{G}r_{X,\bullet }$ which are mutually inverse isomorphisms over $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_{X, \bullet }$ , by factorization.

Remark 6.19. A completely analogue construction to the one leading to Definition 6.15 yields a simplicial flag Grassmannian $\mathcal{G}r^{{\rm Perf}}_{X, \bullet }$ , for the stack of perfect complexes on $X$ (see Remark 6.2(b)).

6.5 Representability

Theorem 6.20. The projection $p \;:\; \mathcal{G}r_X \to \operatorname {\underline {\mathsf{Fl}}}_X^{{\rm dR}}$ is representable by a derived ind-scheme, in the following sense: for any derived affine scheme $S$ and any morphism $F \;:\; S \to \operatorname {\underline {\mathsf{Fl}}}_X^{{\rm dR}}$ (in $\mathsf{dSt}_k^{{\rm Cat}_\infty }$ ), the fibre $\mathcal{G}r_X^F := p^{-1}(F)$ lies in $\mathsf{dSt}_k \subset \mathsf{dSt}_k^{{\rm Cat}_\infty }$ and is representable by a derived ind-scheme.

Remark 6.21. The fibre $\mathcal{G}r_X^F$ tautologically identifies with the derived stack over $S$ classified by the morphism

It immediately follows that $\mathcal{G}r_X^F$ belongs to $\mathsf{dSt}_k$ .

Proof. The proof of Theorem 6.20 will rely on the representability criterion [Reference Gaitsgory and RozenblyumGR17b, Chapter 2, Corollary 1.3.13]. We will thus:

  1. (i) show the underlying underived stack ${\rm t}_0(\mathcal{G}r_X^{F})$ is representable by an ind-scheme; and

  2. (ii) show the derived stack $\mathcal{G}r_X^{F}$ admits a connective deformation theory.

Notice first that the stack ${\rm t}_0(\mathcal{G}r_X^{F})$ is actually a sheaf (on $\mathsf{Aff}_S$ ), by Lemma 5.12.

We start by treating the case ${\bf G}=GL_n$ . If $F$ is of the form $(D, Z = \emptyset )$ , since being a vector bundle is an open condition inside the moduli of torsion-free sheaves, [Reference Halpern-Leinster, Herrero and JonesHLHJ24, Proposition 3.13 and Remark. 3.8] implies (by taking $X$ in loc. cit. to be our $X \times S$ , while $S$ is the same, so that $X\times S \to S$ is indeed projective and of finite presentation) that ${\rm t}_0(\mathcal{G}r_X^{F})$ is represented by an ind-scheme which is ind-quasi-projective over $S$ . For a general flag $F=(D, Z)$ , we consider the obvious forgetful map $q \;:\; {\rm t}_0(\mathcal{G}r_X^{F}) \to {\rm t}_0(\mathcal{G}r_X^{(D, \emptyset )})$ between sheaves on $\mathsf{Aff}_S$ . We claim that this map is representable and affine.

Indeed, first notice that $\underline {\mathsf{Bun}}_{{\widehat {Z}}^{{\rm aff}}} \simeq \underline {\mathsf{Bun}}_{{\widehat {Z}}}$ has an affine diagonal (not of finite type, in general): this follows from [Reference WangWan11, Theorem 1.0.1] applied to $\underline {\mathsf{Bun}}_{Z_{n}}$ , where $Z_{n}$ is the $n$ th thickening of $Z$ in $Y:=X\times S$ . Now, the fibre of $q \;:\; {\rm t}_0(\mathcal{G}r_X^{F}) \to {\rm t}_0(\mathcal{G}r_X^{(D, \emptyset )})$ at $T={\rm Spec} A \to {\rm t}_0(\mathcal{G}r_X^{(D, \emptyset )})$ , corresponding to some $(\underline {\mathcal{E}}, \varphi )$ , is the closed subscheme of the affine scheme

\begin{align*} \underline {{\rm Isom}}_{\underline {\mathsf{Bun}}_{{\widehat {Z}}^{{\rm aff}},\, T}}(\underline {\mathcal{E}}_{|\widehat Z_T^{{\rm aff}}}, \mathcal{E}_0) ,\end{align*}

defined by imposing the restriction to a fixed element on the open subscheme $Y_T\smallsetminus D_T \cap \widehat Z_T^{{\rm aff}}$ . Hence, the map $q$ is indeed representable by affines. Therefore (see, e.g., [Reference RicharzRic19, Lemma 1.7]), ${\rm t}_0(\mathcal{G}r_X^{F})$ is represented by an ind-scheme, since ${\rm t}_0(\mathcal{G}r_X^{(D, \emptyset )})$ is.

Now, the same technique as in [Reference ZhuZhu17, § 1.2], allows us to deduce from the ${\bf G}=GL_n$ case the general ${\bf G}$ case. More precisely, embed ${\bf G}$ into $GL_n$ so that $GL_n/{\bf G}$ is affine; this gives, for ${\bf G}$ , a closed sub-ind-scheme of the corresponding ${\rm t}_0(\mathcal{G}r_X^{F})$ for $GL_n$ . This concludes step (i).

We now focus on step (ii): proving that the derived stack $\mathcal{G}r_X^F$ admits a connective deformation theory. Let $q \;:\; {\rm Spec} B \to \mathcal{G}r_X^F$ be a point. Set $X_B := X \times {\rm Spec} B$ and denote by $D_B$ and $Z_B$ the corresponding closed subschemes. The point $q$ induces a ${\bf G}$ -bundle $\mathcal P$ on $X_B$ . We denote by $\mathcal P_{{\rm ad}} \in {\rm Vect}(X_B)$ the associated adjoint (vector) bundle.

Using that $X$ is proper, the derived stack $\mathcal{G}r_X^F$ then admits an ind-pro-perfect tangent $B$ -complex at $q$ , described as

\begin{align*} \mathbb T_{\mathcal{G}r_X^F,q}[-1] = \operatorname {\mathsf{R}\Gamma }\left (\widehat D_B, \mathcal P_{{\rm ad}}\right ) \mathop{\times} _{\operatorname {\mathsf{R}\Gamma }\left (\widehat D_B \smallsetminus D_B \cup \widehat Z_B, \mathcal P_{{\rm ad}}\right )} 0. \end{align*}

It is stable under base change, and the stack $\mathcal{G}r_X^F$ thus admits a global deformation theory. Moreover, as $D_B$ is a divisor, the homotopy fibre of the morphism $\operatorname {\mathsf{R}\Gamma }(\widehat D_B, \mathcal P_{{\rm ad}}) \to \operatorname {\mathsf{R}\Gamma }(\widehat D_B \smallsetminus D_B, \mathcal P_{{\rm ad}})$ has negative Tor-amplitude. It follows that $\mathcal{G}r_X^F$ has a connective deformation theory.

Remark 6.22. It is useful to record that the first part of the proof of Theorem 6.20 shows that the restriction of $\mathcal{G}r_X$ to purely divisorial flags is represented by derived ind-quasi-projective ind-scheme.

Remark 6.23. If one knows that the union map $\cup \;:\; \operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_{X,2} \to \operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X$ is (representable, in the sense of Theorem 6.20, and) proper, one can conclude, via the pullback

that $\partial _1: \mathcal{G}r_{X, 2} \to \mathcal{G}r_{X}$ is (representable and) proper, too. We can prove that $\cup \;:\; \operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_{X,2} \to \operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X$ is representable (in the sense of Theorem 6.20): this is essentially a consequence of the fact that, if $F\in \operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X(S)$ , and $f: S \to T$ is a map of derived affine schemes, then $F_1 \cup F_2 = F'_1 \cup F'_2 = f_{{\rm red}}^{*} F$ implies $(F_1, F_2)= (F'_1, F'_2)$ , for any $(F_1, F_2)$ and $(F'_1, F'_2)$ in $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_{X,2}(T)$ . On the other hand, although we believe that $\cup \;:\; \operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_{X,2} \to \operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X$ is furthermore proper, we have, at the moment, a complete proof only for $X$ a (smooth) cubic surface in $\mathbb{P}_k^{3}$ .

7. Hecke stacks and exotic derived structures on ${\rm Bun}_X$

7.1 Grassmannian at a fixed ‘large’ flag

Definition 7.1. Let $S$ be a Noetherian $\mathbb{C}$ -scheme, and $Y \to S$ a smooth morphism of relative dimension $\geq 2$ . A flag $(D,Z)$ in $Y/S$ will be called large if for any $s \in S$ , the fibre $Z_s$ meets any irreducible component of the fibre $D_s$ .

Proposition 7.2 (A. Beilinson). Let $S$ be a Noetherian $\mathbb{C}$ -scheme, $Y \to S$ a smooth morphism of relative dimension $\geq 2$ , and $(D,Z)$ a large flag in $Y/S$ . Then, for any fixed ${\bf G}$ -bundle $\overline {\mathcal{E}}$ on $Y$ , the set of global sections

\begin{align*}{\rm Bun}^{{\bf G}}(Y) \times _{{\rm Bun}^{{\rm G}}(Y\setminus D) \times _{{\rm Bun}^{{\bf G}}(\widehat {Z} \setminus D)} {\rm Bun}^{{\bf G}}(\widehat {Z}) } \left \{ \overline {\mathcal{E}} \right \}\end{align*}

is a singleton. In particular, for $\overline {\mathcal{E}}$ being the trivial bundle, we have that the set ${\rm Gr}(Y;D,Z)$ (see Definition 5.2 ) Footnote 28 is a singleton.

Proof. We thank A. Beilinson for providing the following detailed proof. First of all, the argument will be local in $Y$ , so it will be enough to consider the case where $\overline {\mathcal{E}}$ is the trivial bundle.

Consider a function $f \in \Gamma (Y\setminus D, \mathcal{O}_Y)$ , and $U \supset (X\setminus D)$ the largest open subset of $Y$ to which $f$ extends as a regular function (such a $U$ exists since $D$ is an effective Cartier divisor). Define $T: Y \setminus U$ . For any point $y \in Y$ , let $Y^{\wedge }_y= {\rm Spec} (\widehat {\mathcal{O}_{Y,y}})$ . The map $\pi ^{\wedge }_y: Y_y^{\wedge } \to Y$ is flat, so the base change $D_y^{\wedge }$ of $D$ along $\pi ^{\wedge }_y$ is still an effective Cartier divisor on $Y_y^{\wedge }$ .

Claim (a) If $y \in T$ , then the restriction of $f$ to $Y_y^{\wedge } \setminus D_y^{\wedge }$ does not extend to $Y_y^{\wedge }$ .

(b) For any point $s \in S$ , the fibre $T_s$ is either empty or the union of some irreducible components of the fibre divisor $D_s$ .

Proof of claim. (a) By replacing $D$ with $nD$ for $n$ sufficiently large, we may assume that $f \in \Gamma (Y, \mathcal{O}_Y (D))$ . Let $\mathcal{L}:= \mathcal{O}_Y (D)/ \mathcal{O}_Y$ , and $\overline {f}$ be the image of $f$ in $\Gamma (Y, \mathcal{L})$ . Note that $\mathcal{L}$ is an invertible $\mathcal{O}_Y$ -Module. Suppose that claim (a) is wrong, and there exists $y \in T$ such that $f$ does extend to $Y_y^{\wedge }$ . Then, the section $(\pi ^{\wedge }_y)^*\overline {f} \in \Gamma (Y_y^{\wedge }, (\pi ^{\wedge }_y)^*(\mathcal{L}) )= \widehat {\mathcal{O}_{Y,y}}(D_y^{\wedge })/ \widehat {\mathcal{O}_{Y,y}}$ vanishes. Since $ \mathcal{O}_{Y,y} \to \widehat {\mathcal{O}_{Y,y}}$ is faithfully flat, this implies that also the image of $\overline {f}$ in $\mathcal{O}_{Y,y}(D)/\mathcal{O}_{Y,y}$ vanishes, so that $y \notin T$ , which is a contradiction.

(b) $T_s$ is a closed subset of $D_s$ , so it will be enough to show that $T_s$ does not contain a closed point of $D_s$ as a connected component. Indeed, if $y$ is such a point then, replacing $Y$ by a neighbourhood of $y$ , we can assume (by semi-continuity of the fibres applied to $T/S$ [Reference GrothendieckGro66, 13.1.3]) that $T$ is quasi-finite over $S$ . Thus, all the fibres $T_{s'}$ for $s'\in S$ , are closed points in $Y_{s'}$ , and so ${\rm depth}(\mathcal{O}_x)\geq 2$ for any $x \in T$ . Thus $f$ is regular at $T$ , which is a contradiction. The proof of the claim is complete.

The claim remains true if we replace $f$ by an element in $\Gamma (Y\setminus D, \mathcal{O}^n_Y)$ , for any $n$ , i.e. by a section of $\mathbb{A}_Y^n$ on $Y \setminus D$ . Therefore, the claim still holds if $f$ is replaced by any section on $Y\setminus D$ of any affine $Y$ -scheme of finite type $E/X$ (since, locally on $Y$ , such an $E/Y$ is a closed subscheme of some $\mathbb{A}_Y^n$ ). In particular, we may take $\mathcal{E}/Y$ to be a ${\bf G}$ -bundle on $Y$ .

Consider

\begin{align*}(\mathcal{E}, \varphi , \psi ) \in {\rm Gr}(Y;D,Z) = {\rm Bun}^{{\bf G}}(Y) \times _{{\rm Bun}^{{\rm G}}(Y\setminus D) \times _{{\rm Bun}^{{\bf G}}(\widehat {Z} \setminus D)} {\rm Bun}^{{\bf G}}(\widehat {Z}) } \left \{ {\rm triv} \right \},\end{align*}

where $\mathcal{E}$ is a ${\bf G}$ -bundle on $Y$ , $\varphi$ (respectively $\psi$ ) is a trivialization of $\mathcal{E}$ on $Y\setminus D$ (respectively on $\widehat {Z}$ ), and $\varphi$ and $\psi$ agree on $\widehat {Z} \setminus D$ . The trivialization $\varphi$ can be identified, as usual, with a section of $\mathcal{E}/Y$ on $Y\setminus D$ . Let $z\in Z$ . Since

\begin{align*}Y_z^{\wedge } \setminus D_z^{\wedge } = (\widehat {Z}^{{\rm aff}}\setminus D)_z \hookrightarrow (\widehat {Z}^{{\rm aff}})_z =Y_z^{\wedge },\end{align*}

the compatibility between $\varphi$ and $\psi$ on $\widehat {Z} \setminus D$ implies that the restriction of $\varphi$ to $Y_z^{\wedge } \setminus D_z^{\wedge }$ does extend (via $\psi$ ) to $Y_z^{\wedge }$ . Hence, by claim (a), $z \notin Y$ . Since $z \in Z$ is arbitrary, we get $Z \cap T= \emptyset$ . Therefore, for any $s \in S$ , we have $Z_s \cap T_s = \emptyset$ . But $(D,Z)$ is a large flag in $Y/S$ , so $Z_s$ meets any irreducible component of $D_s$ , and (claim (b)) $T_s$ is either empty or a union of irreducible components of $D_s$ . Therefore $T_s$ must be empty, for any $s \in S$ . Thus $T=\emptyset$ , and this means (by definition of $T$ ) that the trivialization $\varphi$ of $\mathcal{E}$ on $Y\setminus D$ extends to a trivialization $\overline {\varphi }$ on the whole $Y$ , whose restriction to $\widehat {Z}^{{\rm aff}}$ (hence to $\widehat {Z}$ ), coincides with $\psi$ . Moreover, since $D$ is an effective Cartier divisor in $Y$ , and $\mathcal{E}$ is separated, two such trivializations $\overline {\varphi }$ and $\overline {\varphi }'$ extending the same $\varphi$ , can be identified with maps $Y \to \mathcal{E}$ that coincide on the quasi-compact schematically dense open subset $(Y\setminus D)\subset Y$ , hence $\overline {\varphi }= \overline {\varphi }'$ . In other words the extension, $\overline {\varphi }$ exists and is unique. This ends the proof of the proposition.

Remark 7.3. Note that the analogue of Proposition 7.2 is obviously false when the stack of ${\bf G}$ -bundles is replaced with the stack of perfect complexes.

Remark 7.4. Proposition 7.2 does not imply that (with the same notation as in Proposition 7.2) the derived stack

\begin{align*}\underline {\mathsf{Bun}}^{{\bf G}}(Y) \times _{\underline {\mathsf{Bun}}^{{\rm G}}(Y\setminus D) \times _{\underline {\mathsf{Bun}}^{{\bf G}}(\widehat {Z} \setminus D)} \underline {\mathsf{Bun}}^{{\bf G}}(\widehat {Z}) } \left \{ \overline {\mathcal{E}} \right \}\end{align*}

is trivial. This should be considered as analogous to the fact that while at the level of classical (truncated) stacks the restriction map ${\rm t}(\underline {\mathsf{Bun}}_{\mathbb{A}^2}^{{\rm G}}) \to {\rm t}(\underline {\mathsf{Bun}}_{\mathbb{A}^2 \setminus 0}^{{\rm G}})$ is an equivalence, this is no more true at the level of derived stacks (as can be checked on cotangent complexes).

Remark 7.5. Proposition 7.2 does not imply that the flag-factorization property (see, e.g., Remark 6.3) is trivial, i.e. it is of the form singleton=singleton (not even for the truncations), for arbitrary pair of good flags. Take, for example, $X= \mathbb{P}^1 \times \mathbb{P}^1$ , fix $x_1, x_2, y \in \mathbb{P}^1$ with $x_1 \neq x_2$ , and consider the flags $F_1 =(D_1= \mathbb{P}^1 \times \{ x_1 \}, Z_1=\{ (y, x_1)\})$ , and $F_2 =(D_2= \mathbb{P}^1 \times \{ x_2 \}, Z_2=\emptyset )$ . Then, $(F_1, F_2) \in \operatorname {\underline {\mathsf{Fl}}}_{X,2}(\mathbb{C})$ , so flag factorization holds, but neither $F_2$ nor $F_1 \cup F_2$ are large flags.

7.2 Exotic derived structures on $\mathsf{Bun}_X$

Let $F=(D,Z) \in \underline {\mathsf{Fl}}_X(\mathbb{C})$ . For simplicity, let us denote by $\mathcal{H}_{{\bf G}, X,F}$ the derived stack

\begin{align*}\mathcal{H}_{{\bf G}, X,F}:= \underline {\mathsf{Bun}}^{{\bf G}}_{X} \times _{\underline {\mathsf{Bun}}^{{\bf G}}_{X\setminus D} \times _{\underline {\mathsf{Bun}}^{{\bf G}}_{\widehat {Z} \setminus D}} \underline {\mathsf{Bun}}^{{\bf G}}_{\widehat {Z}} } \underline {\mathsf{Bun}}^{{\bf G}}_{X}.\end{align*}

This will be called the derived Hecke stack of $X$ at the fixed flag $F$ .

Proposition 7.6. If $F=(D, Z)$ is a large flag in $X/\mathbb{C}$ (Definition 7.1 ), i.e. $Z$ meets any irreducible component of $D$ , then the truncation ${\rm t}(p_i)$ of either of the canonical projections $p_i$ , $i=1,2$ ,

\begin{align*}{\rm t}(p_1), \, {\rm t}(p_2) \;:\; {\rm t}(\mathcal{H}_{{\bf G}, X,F}) \longrightarrow {\rm t}(\underline {\mathsf{Bun}}^{{\bf G}}_X),\end{align*}

is an equivalence of underived stacks, with quasi-inverse given by the diagonal morphism

\begin{align*}\Delta \;:\; {\rm t}(\underline {\mathsf{Bun}}^{{\bf G}}_X) \to {\rm t}(\mathcal{H}_{{\bf G}, X,F})\,\,\,\,\, \mathcal{E} \longmapsto (\mathcal{E}, \mathcal{E}; {\rm id}, {\rm id}).\end{align*}

Proof. We will prove the statement for the second projection, the argument being the same for the first projection. It is obvious that $\Delta$ is a section of ${\rm t}(p_2)$ . So, it is enough to prove that there exists a pseudo-natural transformationFootnote 29 $\alpha \;:\; \Delta \circ {\rm t}(p_2) \to {\rm Id}_{{\rm t}(\mathcal{H}_{{\bf G}, X,F})}$ , such that $\alpha (S)$ is a morphism (then, necessarily an isomorphism) between endofunctors of the groupoid ${\rm t}(\mathcal{H}_{{\bf G}, X,F})(S)$ , for any (classical) Noetherian affine scheme $S$ .

We first observe that, for any Noetherian $S$ , $(D \times S, Z\times S)$ is still a large flag in $X \times S / S$ . Indeed, by applying the same argument to any irreducible component of $D$ , we may clearly suppose that $D$ is itself irreducible. But $D$ is defined over $\mathbb{C}$ , so it is actually geometrically irreducible. So, if $s\;:\;{\rm Spec}(k(s)) \to S$ is a point in $S$ , the fibre $D_s$ is irreducible ( $k(s)$ being a field extension of $\mathbb{C}$ ) and the closed subscheme $Z_s \hookrightarrow D_s$ is non-empty, since ${\rm Spec}(k(s)) \to {\rm Spec}\, \mathbb{C}$ is faithfully flat and $Z\neq \emptyset$ by hypothesis. Therefore $(D \times S, Z\times S)$ is still a large flag in $X \times S / S$ , as claimed. For $\underline {\mathcal{E}}:=(\mathcal{E}_1, \mathcal{E}_2; \varphi , \psi ) \in {\rm t}(\mathcal{H}_{{\bf G}, X,F})(S)$ (where $\varphi \;:\; \mathcal{E}_1 | X_S\setminus D_S \simeq \mathcal{E}_2 | X_S\setminus D_S$ , and $\psi \;:\; \mathcal{E}_1 | \widehat {Z_S} \simeq \mathcal{E}_2 | \widehat {Z_S}$ ), we note that $(\Delta \circ {\rm t}(p_2))(S)(\underline {\mathcal{E}}) =(\mathcal{E}_2, \mathcal{E}_2; {\rm id}, {\rm id} )$ , so that both $(\Delta \circ {\rm t}(p_2))(S)(\underline {\mathcal{E}})$ and ${\rm Id}_{{\rm t}(\mathcal{H}_{{\bf G}, X,F})}(S)(\underline {\mathcal{E}})= (\mathcal{E}_1, \mathcal{E}_2; \varphi , \psi )$ belong to

\begin{align*}{\rm Bun}^{{\bf G}}(X_S) \times _{{\rm Bun}^{{\rm G}}(X_S\setminus D_S) \times _{{\rm Bun}^{{\bf G}}(\widehat {Z_S} \setminus D_S)} {\rm Bun}^{{\bf G}}(\widehat {Z_S}) } \left \{\mathcal{E}_2 \right \},\end{align*}

which is a singleton by Proposition 7.2 (since $(D \times S, Z\times S)$ is still a large flag in $X \times S / S$ ). Therefore, there exists a (unique) isomorphism

\begin{align*}\alpha (S)(\underline {\mathcal{E}}): (\mathcal{E}_2, \mathcal{E}_2; {\rm id}, {\rm id} )=(\Delta \circ {\rm t}(p_2))(S)(\underline {\mathcal{E}})\simeq {\rm Id}_{{\rm t}(\mathcal{H}_{{\bf G}, X,F})}(S)(\underline {\mathcal{E}})= (\mathcal{E}_1, \mathcal{E}_2; \varphi , \psi ),\end{align*}

which, we recall, is of the form $(\overline {\varphi }^{-1}, {\rm id}_{\mathcal{E}_2})$ where $\overline {\varphi }: \mathcal{E}_1 \simeq \mathcal{E}_2$ is the unique common extension of $\varphi$ and $\psi$ to all of $X_S$ .

Finally, the fact that $\alpha (S)$ is a natural transformation between endofunctors of ${\rm t}(\mathcal{H}_{{\bf G}, X,F})(S)$ can be shown as follows. Let

\begin{align*}(f_1,f_2): \underline {\mathcal{E}}:=(\mathcal{E}_1, \mathcal{E}_2; \varphi , \psi )\longrightarrow \underline {\mathcal{E}}':=(\mathcal{E}'_1, \mathcal{E}'_2; \varphi ', \psi ' )\end{align*}

be a morphism in ${\rm t}(\mathcal{H}_{{\bf G}, X,F})(S)$ . We need to prove that the diagram

commutes. Recalling the definition of $\alpha (S)$ , this commutativity easily boils down to showing that the two compositions

(7.1)

agree.

Note that, since $(f_1,f_2)$ is a morphism in ${\rm t}(\mathcal{H}_{{\bf G}, X,F})(S)$ , we have in particular a commutative diagram of isomorphisms.

In particular, by definition of $\alpha (S)$ , we deduce that the two composite maps in (7.1) do agree on $\mathcal{E}_2 |_{X_S \setminus D_S}$ . Since $D$ is an effective Cartier divisor in $X$ , the same is true for $D_S$ in $X_S$ , hence $X_S\setminus D_S$ is quasi-compact and schematically dense in $X_S$ . Now,

\begin{align*} \mathcal{E}_2|_{X_S\setminus D_S} =\mathcal{E}_2 \times _{X_S} (X_S\setminus D_S) = \mathcal{E}_2 \setminus (\mathcal{E}_2 \times _{X_S} D_S) ,\end{align*}

and $\mathcal{E}_2 \to X_S$ is flat (any ${\bf G}$ -bundle is (faithfully) flat over its base), so that $\mathcal{E}_2 \setminus (\mathcal{E}_2 \times _{X_S} D_S)$ is quasi-compact open and schematically dense in $\mathcal{E}_2$ [Sta13, TAG081I]. Now, $\mathcal{E}'_1$ is a separated scheme, and the two maps $\mathcal{E}_2 \to \mathcal{E}'_1$ in ((7.1) coincide on the quasi-compact open and schematically dense $\mathcal{E}_2 \setminus \mathcal{E}_2\times _{X_S} D_S$ , therefore they coincide on all of $\mathcal{E}_2$ [Sta13, TAG01RH]. Therefore, $\alpha (S)$ is indeed a natural transformation, for any $S$ . We leave to the reader the verification of the pseudo-naturality behaviour of $\alpha (S)$ with respect to maps $S' \to S$ . We simply note that this pseudo-functoriality follows from the uniqueness of the extensions $\overline {\varphi }$ and $\overline {\varphi _{S'}}$ of the isomorphisms $\varphi :\mathcal{E}_1 |_{X_S \setminus D_S} \simeq \mathcal{E}_2 |_{X_S \setminus D_S}$ and $\varphi _{S'}: \mathcal{E}_1 |_{X_{S'} \setminus D_{S'}} \simeq \mathcal{E}_2 |_{X_{S'} \setminus D_{S'}}$ , and is simplified by the fact that the diagrams defining the pseudo-naturality of the pseudo-natural transformations $\Delta \circ {\rm t}(p_2)$ and ${\rm Id}_{{\rm t}(\mathcal{H}_{{\bf G}, X,F})}$ are actually strictly commutative.

As a consequence of Proposition 7.6, for each large flag $F$ in $X/\mathbb{C}$ , we may view $\mathcal{H}_{{\bf G}, X,F}$ as a derived enhancement of the underived stack ${\rm t}(\underline {\mathsf{Bun}}^{{\bf G}}_X)$ , that will be denoted as $\underline {\mathsf{Bun}}^{{\bf G}}_{F,X}$ . As the following computations show, in general $\underline {\mathsf{Bun}}^{{\bf G}}_{F,X}$ is an exotic derived structure on ${\rm t}(\underline {\mathsf{Bun}}^{{\bf G}}_{X})$ , i.e. $\underline {\mathsf{Bun}}^{{\bf G}}_{F,X} \neq \underline {\mathsf{Bun}}^{{\bf G}}_{X}$ , or more precisely, neither of the canonical projection maps $\underline {\mathsf{Bun}}^{{\bf G}}_{F,X} \to \underline {\mathsf{Bun}}^{{\bf G}}_{X}$ is an equivalence of derived stacks.

Example 7.7. Let $X= \mathbb{P}^2$ , $D= \mathbb{P}^1$ , and $Z$ be any point in $D$ , and define $F=(D,Z)$ . Consider the (rigid) vector bundle $\mathcal{E}:= \mathcal{O} \oplus \mathcal{O}(3)$ on $X$ , that we identify, notationally with the $\mathbb{C}$ -point $(\mathcal{E}, \mathcal{E}, {\rm id})$ of $\mathcal{H}_{\mathsf{GL}_2, \mathbb{P}^2,F}$ . We will denote by $\mathbb{T}_{F; \mathcal{E}} := \mathbb{T}_{\mathcal{E}}\mathcal{H}_{\mathsf{GL}_2, \mathbb{P}^2,F}$ , and by $\mathbb{T}_{\mathcal{E}} := \mathbb{T}_{\mathcal{E}}\underline {\mathsf{Bun}}^{{\bf G}}_X$ . We wish to compare $\mathbb{T}_{F; \mathcal{E}}$ with $\mathbb{T}_{\mathcal{E}}$ . In order to do this, we use the fibre sequences of complexes of $\mathbb{C}$ -vector spaces

(7.2) \begin{align} \mathbb{T}_{F;\; \mathcal{E}} \longrightarrow \mathbb{T}_{\mathcal{E}} \times \mathbb{T}_{\mathcal{E}} \longrightarrow \mathbb{T}_{\mathcal{E}}':= \mathbb{T}_{(\mathcal{E}, \mathcal{E}, {\rm id})}\big(\underline {\mathsf{Bun}}^{{\bf G}}_{X\setminus D} \times _{\underline {\mathsf{Bun}}^{{\bf G}}_{\widehat {Z} \setminus D}} \underline {\mathsf{Bun}}^{{\bf G}}_{\widehat {Z}} \big) ,\end{align}
(7.3) \begin{align} \mathbb{T}_{\mathcal{E}}' \longrightarrow \mathbb{T}_{\mathcal{E}}\big(\underline {\mathsf{Bun}}^{{\bf G}}_{X\setminus D}\big) \times \mathbb{T}_{\mathcal{E}}\big(\underline {\mathsf{Bun}}^{{\bf G}}_{\widehat {Z}}\big) \longrightarrow \mathbb{T}_{\mathcal{E}}\big(\underline {\mathsf{Bun}}^{{\bf G}}_{\widehat {Z} \setminus D}\big). \end{align}

If we use homogeneous coordinates $[x_0,x_1,x_2]$ on $X= \mathbb{P}^2$ , and identify $D=\left \{ x_0=0\right \}$ , $Z=\left \{ x_0 = x_1=0\right \}$ , we have

\begin{align*}X\setminus D=U_0 \simeq \mathbb{A}_{t_1,t_2}^2={\rm Spec}(\mathbb{C}[t_1, t_2]) ,\quad \,t_i := x_i/x_0, i=1,2 ,\end{align*}
\begin{align*}\widehat {Z}^{{\rm aff}} \simeq (U_{2})_{Z}^{\wedge } \simeq (\mathbb{A}_{s_0,s_1}^2)_{(0,0)}^{\wedge }= {\rm Spec}(\mathbb{C}[[s_0,s_1]]),\quad \,s_j = x_j / x_0, j=0,1,\end{align*}
\begin{align*} \widehat {Z} \setminus D \simeq {\rm Spec}(\mathbb{C}[[s_0,s_1]][s_0^{-1}]).\end{align*}

We have used above the usual notation $U_i :=\left \{ x_i \neq 0\right \}$ for the standard open subsets in $\mathbb{P}^2$ . Since $X\setminus D$ , $\widehat {Z}^{{\rm aff}}$ , and $\widehat {Z} \setminus D$ are affine schemes, ${\rm H^i}(\mathbb{T}_{\mathcal{E}})= 0$ for $i\neq -1, 0, 1$ , and

\begin{align*}{\rm H^{-1}}(\mathbb{T}_{\mathcal{E}})= {\rm Ext}_{\mathbb{P}^2}^0(\mathcal{E},\mathcal{E}) \simeq \mathbb{C}^{12}, \qquad {\rm H^{0}}(\mathbb{T}_{\mathcal{E}})= {\rm Ext}_{\mathbb{P}^2}^1(\mathcal{E}, \mathcal{E}) =0, \qquad {\rm H^{1}}(\mathbb{T}_{\mathcal{E}})= {\rm Ext}_{\mathbb{P}^2}^2(\mathcal{E},\mathcal{E}) \simeq \mathbb{C},\end{align*}

the fibre sequences (7.2) and (7.3) yield exact sequences

(7.4)

(7.5)

(7.6)

In ((7.4), we can compute that ${\rm H}^{-1}(\mathbb{T}'_{\mathcal{E}}) \simeq {\rm ker} (\rho ) \simeq \mathbb{C}^{12}$ , and that ${\rm H}^{0}(\mathbb{T}'_{\mathcal{E}}) \simeq {\rm coker}(\rho )$ is infinite-dimensional since, e.g. the linearly independent family

\begin{align*} \left \{\left(\begin{array}{c@{\quad}c} \frac {1}{s_0^N}\sum _{n\geq 0} s^n_1 & 0\\ 0 & 0 \end{array} \right) \right \}_{N\geq 1} \end{align*}

of endomorphisms of $\mathbb{C}[[s_0,s_1]][s_0^{-1}]^{\oplus \,2}$ does not belong to the image of $\rho$ .

Moreover, in ((7.5) we have that ${\rm ker}(g) \simeq \mathbb{C}^{12}$ embedded as the diagonal inside $\mathbb{C}^{12} \oplus \mathbb{C}^{12}$ . In fact, obviously the diagonal sits inside ${\rm ker}(g)$ ; on the other hand, if $(a,b) \in {\rm ker}(g)$ , and we interpret $a$ and $b$ as maps $\overline {a}, \, \overline {b}: X \to \mathcal{E}^{\vee } \otimes \mathcal{E}$ , we have that $\overline {a}_{|X\setminus D}= \overline {b}_{|X\setminus D}$ (since $(a,b) \in {\rm ker}(g)$ ). Now, $X\setminus D$ is schematically dense in $X$ ( $D$ being an effective Cartier divisor), and $\mathcal{E}^{\vee } \otimes \mathcal{E}$ is separated, hence $\overline {a}= \overline {b}$ , i.e. $a=b$ .

By putting all these results together, we finally get that

\begin{align*}&{\rm H}^{-1}(\mathbb{T}_{F; \mathcal{E}}) \simeq {\rm H}^{-1}(\mathbb{T}_{ \mathcal{E}}) \simeq \mathbb{C}^{12}, \qquad {\rm H}^{0}(\mathbb{T}_{F; \mathcal{E}}) \simeq {\rm H}^{0}(\mathbb{T}_{ \mathcal{E}}) = 0 \\ &\quad \dim _{\mathbb{C}}\, {\rm H}^{1}(\mathbb{T}_{F; \mathcal{E}})= \infty \quad {\rm while} \,\, {\rm H}^{1}(\mathbb{T}_{\mathcal{E}})\simeq \mathbb{C}.\end{align*}

In particular, $\underline {\mathsf{Bun}}^{{\bf G}}_{F,\mathbb{P}^2}$ is an exotic derived structure on ${\rm t}(\underline {\mathsf{Bun}}^{{\bf G}}_{\mathbb{P}^2})$ .

Example 7.8. Keeping the same notation of Example 7.7, we give here an example of a non-rigid bundle $\mathcal{E}$ on $X=\mathbb{P}^2$ for which $\mathbb{T}_{F; \mathcal{E}}$ and $\mathbb{T}_{\mathcal{E}}$ are again different, where $F=(D,Z)$ is the same flag of Example 7.7. We omit the detailed computations, that, however, follow the same lines of the previous Example. We consider $\mathcal{E}:= \mathcal{O}_{\mathbb{P}^2} \oplus {\rm T}_{\mathbb{P}^2}$ (where ${\rm T}_{\mathbb{P}^2}$ is the tangent bundle). Again ${\rm H^i}(\mathbb{T}_{\mathcal{E}})= 0$ for $i\neq -1, 0, 1$ , and in this case we have

\begin{align*}{\rm H^{-1}}(\mathbb{T}_{\mathcal{E}})= {\rm Ext}_{\mathbb{P}^2}^0(\mathcal{E},\mathcal{E}) \simeq \mathbb{C}^{10}, \qquad {\rm H^{0}}(\mathbb{T}_{\mathcal{E}})= {\rm Ext}_{\mathbb{P}^2}^1(\mathcal{E}, \mathcal{E}) \simeq \mathbb{C}, \qquad {\rm H^{1}}(\mathbb{T}_{\mathcal{E}})= {\rm Ext}_{\mathbb{P}^2}^2(\mathcal{E},\mathcal{E}) =0.\end{align*}

The sequence ((7.4) is the same but ${\rm H}^{-1}(\mathbb{T}'_{\mathcal{E}}) \simeq {\rm ker} (\rho ) \simeq \mathbb{C}^{14}$ , and ${\rm H}^{0}(\mathbb{T}'_{\mathcal{E}}) \simeq {\rm coker}(\rho )$ is infinite-dimensional since, e.g. the linearly independent family

\begin{align*}\left\{\left(\begin{array}{c@{\quad}c@{\quad}c}\frac {1}{s_0^N}\sum _{n\geq 0} s^n_1 & 0 & 0\\ 0 & 0 & 0\\ 0 & 0 & 0 \end{array}\right) \right\}_{N\geq 1}\end{align*}

of endomorphisms of $\mathbb{C}[[s_0,s_1]][s_0^{-1}]^{\oplus \, 3}$ does not belong to the image of $\rho$ .

The sequences ((7.5) and ((7.6) are replaced by the sequence

(7.7)

and the same argument used in Example 7.7 shows that $ {\rm H}^{-1}(\mathbb{T}_{F; \mathcal{E}}) \simeq \mathbb{C}^{10}$ . Moreover, ((7.7) shows that ${\rm H}^{1}(\mathbb{T}_{F; \mathcal{E}})$ is infinite-dimensional, since ${\rm H}^{0}(\mathbb{T}'_{\mathcal{E}})$ is. The upshot is that

\begin{align*}&{\rm H}^{-1}(\mathbb{T}_{F; \mathcal{E}}) \simeq {\rm H}^{-1}(\mathbb{T}_{ \mathcal{E}}) \simeq \mathbb{C}^{10}, \qquad {\rm H}^{0}(\mathbb{T}_{F; \mathcal{E}}) \simeq \mathbb{C}^k\, \textrm {for some}\, 4\leq k\leq 6,\quad {\rm H}^{0}(\mathbb{T}_{ \mathcal{E}}) \simeq \mathbb{C} \\&\qquad \qquad\qquad \! \dim _{\mathbb{C}}\, {\rm H}^{1}(\mathbb{T}_{F; \mathcal{E}})= \infty ,\quad {\rm H}^{1}(\mathbb{T}_{\mathcal{E}})=0.\end{align*}

Remark 7.9. Example 7.8 shows that the deformation space associated with the exotic structure $\underline {\mathsf{Bun}}^{{\bf G}}_{F,X}$ is not in general a subspace of the non-exotic one, but rather the latter is a quotient of the former. This is a general phenomenon that can be explained heuristically by observing that the choice of a flag $F$ does not simply put constraints on the usual deformation space, thus selecting a subspace thereof. The deformation space of $\underline {\mathsf{Bun}}^{{\bf G}}_{F,X}$ rather describes deformations of more data (bundle and trivializations), so the map from this deformation space to the non-exotic deformation space is expected to have non-trivial fibres. In other words, the map of derived stacks $\underline {\mathsf{Bun}}^{{\bf G}}_{F,X} \to \underline {\mathsf{Bun}}^{{\bf G}}_{X}$ is rather a forgetful map, not an inclusion.

8. Flag-chiral product and flag-factorization objects

In this section we suggest a replacement of the chiral tensor product (defined for sheaves on the Ran space of a curve, see [Reference Beilinson and DrinfeldBD04, Reference Francis and GaitsgoryFG12]) for sheaves or stacks over $\mathsf{Fl}_X$ , $X$ being a surface. We call this the flag-chiral tensor product. As a consequence (and inspired by [Reference RaskinRas19]), we define flag analogues of factorization algebras (called flag-factorization algebras) and of factorization categories (called flag-factorization categories).

8.1 Naive flag-chiral product

As a warm up, we define a flag-chiral product on $\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ , where $\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ is defined as the composition

${\bf PoSets} \to {\bf Sets}$ being the forgetful functor (forgetting the partial order).

Let us denote by $\mathsf{Shv}(-)$ a 3-functors formalismFootnote 30 in the sense of [Reference ScholzeSch22, Definition 2.4], i.e. a lax symmetric monoidal functor $\mathsf{Corr}(\mathcal{C}_{{\bf Sch}_{\mathbb{C}}}, E) \to {\bf Cat}_{\infty }$ , where $\mathcal{C}_{{\bf Sch}_{\mathbb{C}}}$ is an appropriate subcategory of ${\bf Sch}_{\mathbb{C}}$ , and $E$ is an appropriate class of maps in ${\bf Sch}_{\mathbb{C}}$ . Here, the word appropriate depends on the specific 3-functors formalism chosen. An example of particular interest for us will be $\mathcal{C}_{{\bf Sch}_{\mathbb{C}}}$ consisting of separated schemes of finite type over $\mathbb{C}$ , $E$ all morphisms between them, and $\mathsf{Shv}(-)= \mathcal{D}\mbox{-}\mathsf{Mod}$ given by $\mathcal{D}$ -modules (see [Reference ScholzeSch22, Appendix to Lecture VIII:D-modules]). The reader will find other examples in the appendices of [Reference ScholzeSch22]. Since $\operatorname {\underline {\mathsf{Fl}}}^{\flat }_{X, \bullet }$ is a $2$ -Segal object (Theorem 4.21), $\operatorname {\underline {\mathsf{Fl}}}_X^\flat$ is an algebra in $\mathsf{Corr}(\mathcal{C}_{{\bf Sch}_{\mathbb{C}}}, E)$ (e.g. for the choices corresponding to $\mathsf{Shv}(-)= \mathcal{D}\mbox{-}\mathsf{Mod}$ ), hence by composing with $\mathsf{Shv}(-)$ , we get a monoidal structure $\otimes ^{\flat ,\, {\rm ch}}$ on $\mathsf{Shv}(\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X)$ , that will be called the naive flag-chiral monoidal structure. In particular, considering the correspondence

we have the binary productFootnote 31

\begin{align*}\mathcal{F}\otimes ^{\flat ,\, {\rm ch}} \mathcal{G} = p_{!} q^*(\mathcal{F}\boxtimes \mathcal{G}), \,\,\,\, \mathcal{F}, \mathcal{G} \in \mathsf{Shv}(\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X).\end{align*}

Though not needed, the union map $\partial _1^{\flat }$ is probably proper for general $X$ (it is, e.g. when $X$ is a smooth cubic surface, see Remark 6.23). One can prove that $\otimes ^{\flat ,\, {\rm ch}}$ is actually a symmetric monoidal structure on $\mathsf{Shv}(\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X)$ .

8.2 Flag-chiral product and flag-factorization algebras

We sketch here a version of the naive flag-chiral product for sheaves on $\operatorname {\underline {\mathsf{Fl}}}_X$ (so that maps between flags are also taken into account). The reason for this comes from [Reference Beilinson and DrinfeldBD04, 3.4.6] where factorization algebras over a given curve $C$ (defined using the Ran space of $C$ ) are given an equivalent formulation using the category fibred in posets over ${\bf Sch}_{\mathbb{C}}$ of relative effective Cartier divisors on $C/\mathbb{C}$ with maps given by inclusions of the support of the divisors. These maps are crucial for such a comparison. The drawback of the necessity of considering maps between divisors is that the corresponding fibred category is not a stack in groupoids anymore but rather a stack in categories (namely, in posets), and to our knowledge, no 3-functors formalism is available for sheaves on such objects. So, we have to take a different approach. Specifically, we follow the ideas, notation and constructions given by Sam Raskin in [Reference RaskinRas19], where he also shows ([Reference RaskinRas19, Proposition 6.19.4 and Remark 6.19.5]) that his definition recovers the definition of factorization algebras of [Reference Francis and GaitsgoryFG12] (the latter being an $\infty$ -categorical elaboration of the original approach in [Reference Beilinson and DrinfeldBD04]).

We denote by $\mathsf{PreStk}^{{\rm lax}}$ the category of lax-prestacks on $\mathsf{dAff}_{\mathbb{C}}$ (i.e. of admissible functors $\mathsf{dAff}_{\mathbb{C}}^{{\rm op}} \to {\bf Cat}$ [Reference RaskinRas19, 4.14]), and $\mathsf{PreStk}_{{\rm corr}}^{{\rm lax}}$ the categoryFootnote 32 of correspondences in $\mathsf{PreStk}^{{\rm lax}}$ [Reference RaskinRas19, 4.28]. Note that, for any lax-prestack $\mathcal{X}$ , there is a category $\mathsf{QCoh}(\mathcal{X})$ of quasi-coherent complexes on $\mathcal{X}$ , defined as the category of natural transformations $\mathcal{X} \to \mathsf{QCoh}$ [Reference RaskinRas19, s4.15]. As observed in Remark 4.22, $\operatorname {\underline {\mathsf{Fl}}}_{X}$ is actually a unital commutative (i.e. $E_{\infty }$ ) algebra in correspondences, and we thus view it as an object, denoted as $\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}$ , in $\mathsf{CAlg}(\mathsf{PreStk}_{{\rm corr}}^{{\rm lax}})$ . This notation is meant to parallel Raskin’s $\mathsf{Ran}_{\mathcal{X}}^{\rm un,\ ch}$ in [Reference RaskinRas19, 4.28]. The same is true for $\operatorname {\underline {\mathsf{Fl}}}_{X}^{{\rm dR}}$ , whose unital commutative algebra structure will be denoted as $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}, {\rm ch}}_{X} \in \mathsf{CAlg}(\mathsf{PreStk}_{{\rm corr}}^{{\rm lax}})$ . Having these two commutative algebras $\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}$ and $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}, {\rm ch}}_{X}$ at our disposal, we may consider multiplicative sheaves $\mathcal{C}$ of categories over any of them [Reference RaskinRas19, Definition 5.21.1], and, given such a multiplicative sheaf $\mathcal{C}$ of categories, we have a notion of multiplicative object in $\mathcal{C}$ [Reference RaskinRas19, Definition 5.27.1]. In particular, the sheaves of categories $\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}}$ and $\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}, {\rm ch}}_{X}}$ both carry a canonical multiplicative structure [Reference RaskinRas19, 5.21.2]. Therefore, we can give the following.

Definition 8.1. A flag pre-chiral category on $X$ is a unital multiplicative sheaf of categories on the unital commutative algebra $\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}$ .

A flag pre-chiral crystal category on $X$ is a unital multiplicative sheaf of categories on the unital commutative algebra $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}, {\rm ch}}_{X}$ .

A quasi-coherent flag pre-factorization algebra on $X$ is a multiplicative object in the flag pre-chiral category $\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}}$ on $X$ .

A $\mathcal{D}$ -module flag pre-factorization algebra on $X$ is a multiplicative object in the flag pre-chiral crystal category $\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}, {\rm ch}}_{X}}$ on $X$ .

We denote by $\mathsf{FlCats}^{{\rm ch},{\rm pre}}_{X}$ (respectively, $\mathsf{FlCrysCats}^{\rm ch,\ pre}_{X}$ ) the $2$ -category of flag pre-chiral categories (respectively of flag pre-chiral crystal categories) on $X$ . Analogously, we denote by $\mathsf{FlFact}^{{\rm pre}}_{X}$ (respectively, $\mathsf{FlFact}^{{\rm dR}, {\rm pre}}_{X}$ ) the category of quasi-coherent flag pre-factorization algebras (respectively $\mathcal{D}$ -module flag pre-factorization algebras) on $X$ .

Unzipping Definition 8.1, we get the following (partial but more concrete) descriptions. Let us consider the correspondence (i.e. morphism in $\mathsf{PreStk}_{{\rm corr}}^{{\rm lax}}$ )

and the canonical unit section (i.e. $0$ th degeneracy map of $\operatorname {\underline {\mathsf{Fl}}}_{X, \bullet }$ ) $\mathsf{e}:=\sigma _0: {\rm Spec}\, \mathbb{C} \to \operatorname {\underline {\mathsf{Fl}}}_X$ , given by the empty flag. Then, a flag pre-chiral category $\mathcal{C}$ on $X$ have, in particular, canonical binary product and unit maps [Reference RaskinRas19, 5.21.1]

(8.1) \begin{equation} \begin{aligned} &\mu _{\mathcal{C}}: q^*(\mathcal{C}\boxtimes \mathcal{C}) \longrightarrow p^*(\mathcal{C}) \,\,\, {\rm in}\,\,\, \mathsf{ShvCat}_{\operatorname {\underline {\mathsf{Fl}}}_{X,2}}, \\ &\quad \epsilon _{\mathcal{C}}: \mathsf{QCoh}_{{\rm Spec}\, \mathbb{C}}=\mathsf{Vect} \longrightarrow \mathsf{e}^*(\mathcal{C}) \,\, \,{\rm in}\,\,\, \mathsf{ShvCat}_{{\rm Spec}\, \mathbb{C}} = \mathsf{Cat}_{\mathbb{C}} \end{aligned}\end{equation}

(where $\mathsf{Vect}$ denotes the category of complexes of $\mathbb{C}$ -vector spaces) which are both equivalences, and satisfy appropriate commutativity and associativity conditions. And we have similar maps for $n$ -ary products, for any $n$ , satisfying appropriate commutativity and associativity conditions. An analogous description is available when $\mathcal{C}$ is a flag pre-chiral crystal category on $X$ .

In the special case of $\mathcal{C}= \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}}$ , we have

\begin{align*}&\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}}\boxtimes \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}} \simeq \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X} \times \operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}}, \qquad q^*(\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X} \times \operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}}) \simeq \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}_{X,2}},\\ & \qquad\qquad\qquad p^*(\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}}) \simeq \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}_{X,2}},\end{align*}

and the (binary) product and unit maps have the following obvious from

\begin{align*}\mu _{\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}}}= {\rm Id} \;:\; \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}_{X,2}} \longrightarrow \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}_{X,2}}, \qquad \epsilon _{\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}}}= {\rm Id}: \mathsf{Vect} \longrightarrow \mathsf{Vect}.\end{align*}

Therefore, if $\mathsf{A}$ is a quasi-coherent flag pre-factorization algebra on $X$ , by definition, [Reference RaskinRas19, 4.19.3, 5.5]

\begin{align*}\mathsf{A} \in \Gamma (\operatorname {\underline {\mathsf{Fl}}}_X, \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}_{X}})&= \mathsf{Hom}_{\mathsf{ShvCat}_{\operatorname {\underline {\mathsf{Fl}}}_X}}(\mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}_{X}}, \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}_{X}})\simeq \mathsf{QCoh}(\operatorname {\underline {\mathsf{Fl}}}_{X})\\&:=\mathsf{Nat}(\operatorname {\underline {\mathsf{Fl}}}_{X}, \mathsf{QCoh}) \in \mathsf{DGCat}_{{\rm cont}}, \end{align*}

and we have, in particular, canonical binary product and unit maps [Reference RaskinRas19, 5.27]

(8.2) \begin{align} \mu _{\mathsf{A}}: q^*(\mathsf{A}\boxtimes \mathsf{A}) \longrightarrow p^*(\mathsf{A}) \,\,\, {\rm in}\,\,\, \Gamma (\operatorname {\underline {\mathsf{Fl}}}_{X,2}, \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}_{X,2}})=\mathsf{QCoh}(\operatorname {\underline {\mathsf{Fl}}}_{X,2}) , \end{align}
(8.3) \begin{align} \epsilon _{\mathsf{A}}: \mathbb{C} \longrightarrow \mathsf{e}^*(\mathsf{A})= \mathsf{A}_{|(\emptyset ,\emptyset )} \,\, \,{\rm in}\,\,\, \mathsf{Vect} ,\end{align}

which are both equivalences, and satisfy appropriate commutativity and associativity conditions. And we have similar maps for $n$ -ary products, for any $n$ , satisfying appropriate commutativity and associativity conditions. An analogous description is available when $\mathsf{A}$ is $\mathcal{D}$ -module flag pre-factorization algebra on $X$ . Note that in this case, $\mathsf{A} \in \Gamma (\operatorname {\underline {\mathsf{Fl}}}_X, \mathsf{QCoh}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_{X}})\simeq \mathsf{Nat}(\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_{X}, \mathsf{QCoh})$ .

In order to omit the prefix ‘pre-’ in Definition 8.1, i.e. to pass from flag pre-chiral (crystal) categories (respectively, pre-factorization algebras) on $X$ to flag-chiral (crystal) categories (respectively, flag-factorization algebras) on $X$ , we use the intuition from [Reference Beilinson and DrinfeldBD04, 3.4.6] where sheaves of factorization algebras (in the Cartier approach discussed there) are required to be insensitive to the scheme structures of the divisors. This suggests the definition below.

Let $\mathcal{C}$ be either a flag pre-chiral category or a flag pre-chiral crystal category on $X$ , and $\mathsf{A}$ either a quasi-coherent flag pre-factorization algebra or a $\mathcal{D}$ -module flag pre-factorization algebra on $X$ . For $S \in \mathsf{Aff}_{\mathbb{C}}$ , (i.e. a classical underivedFootnote 33 affine Noetherian scheme), and a morphism $\alpha \;:\; F_1 =(D_1, Z_1) \to (D_2, Z_2)=F_2$ a morphism in either $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ or in $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X(S)$ , we have an induced functor $\mathcal{C}_S(\alpha ): \mathcal{C}_S(F_1) \to \mathcal{C}_S(F_2)$ and a morphism $\mathsf{A}_S(\alpha ):\mathsf{A}_S(F_1) \to \mathsf{A}_S(F_2)$ . We say that the morphism $\alpha$ is a reduced isomorphism if the induced maps on the underlying reduced subschemes $(D_1)_{{\rm red}} \to (D_2)_{{\rm red}}$ and $(Z_1)_{{\rm red}} \to (Z_2)_{{\rm red}}$ are both isomorphisms.

Definition 8.2. A flag pre-chiral category (respectively, a flag pre-chiral crystal category) $\mathcal{C}$ on $X$ is a flag-chiral category (respectively a flag-chiral crystal category) on $X$ , if for any (underived) $S \in \mathsf{Aff}_{\mathbb{C}}$ , and any reduced isomorphism $\alpha$ in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ (respectively in $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X(S)$ ), the induced functor $\mathcal{C}_S(\alpha )$ is an equivalence.

A quasi-coherent flag pre-factorization algebra (respectively, a $\mathcal{D}$ -module flag pre-factorization algebra) $\mathsf{A}$ on $X$ is a quasi-coherent flag-factorization algebra (respectively, a $\mathcal{D}$ -module flag-factorization algebra) if for any $S \in \mathsf{dAff}_{\mathbb{C}}$ , and any reduced isomorphism $\alpha$ in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ (respectively in $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X(S)$ ), the induced morphism $\mathsf{A}_S(\alpha )$ is an equivalence.

The full sub-2-category of $\mathsf{FlCats}^{{\rm ch},{\rm pre}}_{X}$ (respectively, of $\mathsf{FlCrysCats}^{\rm ch,\ pre}_{X}$ ) whose objects are flag-chiral categories (respectively, flag-chiral crystal categories) is denoted by $\mathsf{FlCats}^{{\rm ch}}_{X}$ (respectively, $\mathsf{FlCrysCats}^{{\rm ch}}_{X}$ ).

The full subcategory of $\mathsf{FlFact}^{{\rm pre}}_{X}$ (respectively, of $\mathsf{FlFact}^{{\rm dR}, {\rm pre}}_{X}$ ) whose objects are quasi-coherent flag-factorization algebras (respectively, $\mathcal{D}$ -module flag-factorization algebras) is denoted by $\mathsf{FlFact}_{X}$ (respectively, $\mathsf{FlFact}^{{\rm dR}}_{X}$ ).

Example 8.3. Obviously $\mathcal{O}_{\operatorname {\underline {\mathsf{Fl}}}_X}$ (respectively, $\mathcal{O}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X}$ ) is an object in $\mathsf{FlFact}_{X}$ (respectively, $\mathsf{FlFact}^{{\rm dR}}_{X}$ ). Note that for any test affine underived scheme $S$ , and any map $F_1 \to F_2$ in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ (respectively, in $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X(S)$ ), the induced map (which is part of the data defining $\mathcal{O}$ , see [Reference RaskinRas19, (4.16.1)]) $(\mathcal{O}_{\operatorname {\underline {\mathsf{Fl}}}_X})_S (F_1)= \mathcal{O}_S \to (\mathcal{O}_{\operatorname {\underline {\mathsf{Fl}}}_X})_S (F_2)= \mathcal{O}_S$ is the identity (respectively, the induced map $(\mathcal{O}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X})_S (F_1)= \mathcal{O}_{S_{{\rm red}}} \to (\mathcal{O}_{\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X})_S (F_2)= \mathcal{O}_{S_{{\rm red}}}$ is the identity.)

By Remark 6.16, the 2-Segal objects $\mathcal{G}r_{X, \bullet }$ and $\mathcal{G}r^{{\rm dR}}_{X, \bullet }$ give rise to commutative algebras in $\mathsf{PreStk}_{{\rm corr}}^{{\rm lax}}$ , that will be denoted, respectively, as $\mathcal{G}r^{{\rm ch}}_{X}$ and $\mathcal{G}r^{{\rm dR}, {\rm ch}}_{X}$ . So, the same arguments used above for $\operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_X$ and $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}, {\rm ch}}_X$ , allow for a definition of $\mathsf{Gr}$ -chiral categories and of $\mathsf{Gr}$ -chiral crystal categories.

Moreover, as the sheaves of categories $\mathsf{QCoh}_{\mathcal{G}r^{{\rm ch}}_{X}}$ and $\mathsf{QCoh}_{\mathcal{G}r^{{\rm dR}, {\rm ch}}_{X}}$ are, as always, multiplicative and unital sheaves of categories on, respectively, $\mathcal{G}r^{{\rm ch}}_{X}$ and $\mathcal{G}r^{{\rm dR}, {\rm ch}}_{X}$ , we also have the notions of quasi-coherent $\mathsf{Gr}$ -factorization algebra and of $\mathcal{D}$ -module $\mathsf{Gr}$ -factorization algebra. A more detailed investigations of these objects is deferred to a future paper.

Example 8.4. As in Example 8.3, $\mathcal{O}_{\mathcal{G}r^{{\rm ch}}_{X}}$ (respectively, $\mathcal{O}_{\mathcal{G}r^{{\rm dR}, {\rm ch}}_{X}}$ ) is a quasi-coherent $\mathsf{Gr}$ -factorization algebra (respectively, a $\mathcal{D}$ -module $\mathsf{Gr}$ -factorization algebra). Moreover, if $p: \mathcal{G}r^{{\rm ch}}_{X} \to \operatorname {\underline {\mathsf{Fl}}}^{{\rm ch}}_{X}$ (respectively $p^{{\rm dR}}: \mathcal{G}r^{{\rm dR}, {\rm ch}}_{X} \to \operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}, {\rm ch}}_{X}$ ) denotes the projection maps, both are maps of commutative algebras in $\mathsf{PreStk}_{{\rm corr}}^{{\rm lax}}$ , and $p_* (\mathcal{O}_{\mathcal{G}r^{{\rm ch}}_{X}})$ (respectively, $(p^{{\rm dR}})_*(\mathcal{O}_{\mathcal{G}r^{{\rm dR}, {\rm ch}}_{X}})$ ) is an object in $\mathsf{FlFact}_{X}$ (respectively, in $\mathsf{FlFact}^{{\rm dR}}_{X}$ ).

Remark 8.5. When $X$ is a curve, the definitions 8.1 and 8.2 admit obvious analogues, where $\operatorname {\underline {\mathsf{Fl}}}_X$ is replaced by the lax-prestack $\underline {\mathsf{Car}}_{X}$ (of relative effective Cartier divisors on $X$ , with morphisms given by inclusions of Cartier). We expect that, in this version, Definition 8.2 recovers Raskin’s definitions of chiral category and of factorization algebras [Reference RaskinRas19, Definition 6.2.1].

Remark 8.6. One can adapt the formalism of [Reference RaskinRas19] in order to define the notion of a flag-factorization (crystal) space over $X$ , in such a way that the flag-Grassmannian $\mathcal{G}r_X$ is such a factorization (crystal) space. We leave the details to the interested reader.

Appendix A. Extensions into simplicial objects

In this appendix, we will prove the technical Proposition A.1, used in § 6.4.

Proposition A.1. Let $\mathcal C$ be an $\infty$ -category with all countable limits. Let $X_\bullet \;:\; \Delta ^{{\rm op}} \to {\mathcal C}$ be a simplicial object. Denote by $\Delta _{\leq 2}$ the full subcategory of $\Delta$ spanned by $[0]$ , $[1]$ and $[2]$ . Let $Y^{\leq 2}_\bullet \;:\; \Delta _{\leq 0}^{{\rm op}} \to {\mathcal C}$ and let $p_{\bullet }^{\leq 2} \;:\; Y_{\bullet }^{\leq 2} \to X_{|\Delta _{\leq 2}}$ be a morphism (of truncated simplicial objects). Assume that:

  1. (a) the morphism $Y_0 \to X_0$ is an equivalence;

  2. (b) the square

    is Cartesian.

Then $Y_\bullet ^{\leq 2}$ and $p_\bullet ^{\leq 2}$ extend naturally into a morphism of simplicial objects $p_\bullet \;:\; Y_\bullet \to X_\bullet$ such that:

  1. (i) for any $n \geq 2$ , the commutative square

    is Cartesian, where $\delta _X^n$ and $\delta _Y^n$ are the images by $X$ and $Y$ of the map $\ell _n \;:\; [1] \to [n]$ sending $0$ to $0$ and $1$ to $n$ ;

  2. (ii) the morphism $p_\bullet$ is relatively $2$ -Segal. In particular, if $X_\bullet$ is $2$ -Segal, so is $Y_\bullet$ .

Proof. Denote by $\Delta ^{{\rm int}}$ the subcategory of $\Delta$ containing every object, but only morphisms $f \;:\; [n] \to [m]$ satisfying $f(0) = 0$ and $f(n) = m$ . We also set $\Delta ^{{\rm int}}_{\leq 2} := \Delta ^{{\rm int}} \cap \Delta _{\leq 2}$ . Note that $\ell _n$ is the only morphism $[1] \to [n]$ in $\Delta ^{{\rm int}}$ . The restriction of the morphism $p_\bullet ^{\leq 2}$ to $\Delta ^{{\rm int}}_{\leq 2}$ can be represented as a diagram indexed $p' \;:\; D \to {\mathcal C}$ by the pullback

where $I = \Delta ^1$ is the interval category. We define $p_\bullet ^{{\rm int}}$ as the right Kan extension of $p'$ along the inclusion $D \to (\Delta ^{{\rm int}})^I$ . By our assumptions (a) and (b), $p'$ is itself a right Kan extension its restriction to $\Delta ^{{\rm int}}_{\leq 1} \times _{\Delta ^{{\rm int}}} (\Delta ^{{\rm int}})^I$ . As a consequence, the morphism $p_\bullet ^{{\rm int}}$ satisfies condition (i).

Notice that $p_\bullet ^{{\rm int}}$ already defines all the degeneracy maps as well as the face morphisms $Y_n \to Y_{n_1}$ except $\partial _0$ and $\partial _n$ . In particular, in order to extend $p_\bullet ^{{\rm int}}$ to a fully fledge morphism of simplicial objects, we only need to construct in a canonical way the face morphisms $\partial _0$ and $\partial _n$ . To do so, we need the following trivial observation.

Observation: Let $\epsilon _i \;:\; [n-1] \to [n]$ be the morphism in $\Delta$ whose image is $[n] \smallsetminus \{i\}$ (so that $\partial _i$ of a simplicial object is the image of $\epsilon _i$ ) For $i = 0$ or $n$ , there exists a unique morphism $g_i \;:\; [2] \to [n]$ in $\Delta ^{{\rm int}}$ such that the following diagrams commute.

Indeed, the only possible choice is $g_i(0) = 0$ , $g_i(2) = n$ and $g_0(1) = 1$ or $g_n(1) = n-1$ . As a consequence, we can canonically form the plain diagrams

and there exist essentially unique dashed arrows completing the diagrams, as the front squares are Cartesian. It follows that $p_\bullet ^{{\rm int}}$ extends to a morphism of simplicial objects $p_\bullet \;:\; Y_\bullet \to X_\bullet$ .

Remains to prove that the projection $p_\bullet$ is relatively $2$ -Segal. For $Z$ a simplicial object (actually, either $X$ or $Y$ ), for any integer $n \geq 2$ , and any $0 \lt j \lt n$ , we consider the commutative diagrams

where $Z_n^{(i),0}$ , $Z_n^{(i),n}$ , $q_i^0$ and $q_i^n$ are defined by the squares being Cartesian.

Notice then that, for any $0 \lt j \lt n$ , in the commutative diagram

both the rightmost square and the outer rectangle are Cartesian. It follows that the leftmost square is also Cartesian. In particular, in the commutative diagram, for $i = 0, n$ and $0 \lt j \lt n$ ,

the upper, bottom and front squares are Cartesian. It follows that so is the back square. Finally, we can form the following commutative diagram.

By what precedes, both the rightmost square, and the outer rectangle are Cartesian. In particular, so is the leftmost square. This being true for $i = 0,n$ and $0 \lt j \lt n$ , this is precisely the relative $2$ -Segal condition on the projection $p_\bullet$ .

Appendix B. Actions

In this appendix we construct some natural actions of appropriate flags analogues of the loop and positive loop group on the flag Grassmannian. In order to simplify the exposition, we work in the underived setting: all test affine schemes and all stacks (namely the stack of ${\bf G}$ -bundles) will be implicitly underived. Generalizing the constructions to the derived setting would pose no problems but would increase the length (and possibly the readability) of this section. Although we focus on the stack of ${\bf G}$ -bundles, as in the main text, similar results hold for the stack of perfect complexes together with the choice of a fixed perfect complex $\mathcal{E}_0$ on $X$ (replacing the trivial ${\bf G}$ -bundle).

B.1 Flags version of the (positive) loop group

Let $X$ be our smooth projective surface and $S$ be an arbitrary test scheme in $\mathsf{Aff}$ . As usual, we will sometimes write $Y$ for $X \times S$ , and $\mathcal{E}_0$ will denote the trivial ${\bf G}$ -bundle (we will leave the context to specify which space $\mathcal{E}_0$ actually lives on).

Definition B.1. Let $F=(D,Z) \in \operatorname {\underline {\mathsf{Fl}}}_X(S)$ , $Y:=X \times S$ . We define the following.

  • the sheaf of groups over $Y$

    \begin{align*}\mathsf{LG}_X(S,F):= * \times _{{\bf Bun}_{\widehat {D}\smallsetminus D} \times _{{\bf Bun}_{(\widehat {D}\smallsetminus D) \times _{Y} {\hat {{Z}}}^{{\rm aff}}}} {\bf Bun}_{{\hat {{Z}}}}}\,\, *,\end{align*}
    where the two maps from $*$ coincide and are given by the trivial ${\bf G}$ -bundle $\mathcal{E}_0$ . In other words, $\mathsf{LG}_X(S,F)$ is the loop stack $\Omega _{\mathcal{E}_0}({\bf Bun}_{\widehat {D}\smallsetminus D} \times _{{\bf Bun}_{(\widehat {D}\smallsetminus D) \times _{Y} {\hat {{Z}}}^{{\rm aff}}}} {\bf Bun}_{{\hat {{Z}}}})$ at the point corresponding to the trivial ${\bf G}$ -bundle;
  • the group

    \begin{align*}LG_X(S,F):= {\rm Aut}_{{\rm Bun}(\widehat {D}\smallsetminus D)}(\mathcal{E}_0)\times _{{\rm Aut}_{{\rm Bun}((\widehat {D}\smallsetminus D) \times _{Y} {\widehat {Z}}^{{\rm aff}})}(\mathcal{E}_0)} {\rm Aut}_{{\rm Bun}({\widehat {Z}})}(\mathcal{E}_0),\end{align*}
    i.e. $LG_X(S,F)=\Gamma (X \times S, {\mathsf{LG}}_X(S,F))$ ;
  • the sheaf of groups over $Y$

    \begin{align*}\mathsf{L^{+}G}_X(S,F):= * \times _{{\bf Bun}_{\widehat {D} }} \,\, *,\end{align*}
    i.e. the loop stack $\Omega _{\mathcal{E}_0}({\bf Bun}_{\widehat {D}})$ of ${\bf Bun}_{\widehat {D}}$ at the point corresponding to the trivial ${\bf G}$ -bundle;
  • the group $L^{+}G_X(S,F)$ of global sections of $\mathsf{L^{+}G}_X(S,F)$ , i.e. $L^{+}G_X(S,F) = {\rm Aut}_{{\rm Bun}(\widehat {D})}(\mathcal{E}_0)$ .

Remark B.2. Consider the obvious projection ${\bf Gr}_X(S)(F) \to {\bf Bun}_{\widehat {D}}$ (see Lemma 5.9). The fibre of this map at the trivial bundle map $* \to {\bf Bun}_{\widehat {D}}$ coincides with the sheaf $\mathsf{LG}_X(S,F)$ , and the induced map $\mathsf{LG}_X(S)(F) \to {\bf Gr}_X(S)(F)$ of stacks over $X \times S$ is a section of the canonical projection ${\bf Gr}_X(S)(F) \to \mathsf{LG}_X(S)(F)$ . Moreover, there is an obvious morphism $\mathsf{L^{+}G}_X(S,F) \to \mathsf{LG}_X(S,F)$ of sheaves of groups, given by restricting the trivializations.

Remark B.3. Since we are dealing with ${\bf G}$ -principal bundles, we have

\begin{align*}\mathsf{LG}_X(S,F) \simeq {\bf G}_{\widehat {D}\smallsetminus D, X\times S} \times _{{\bf G}_{(\widehat {D}\smallsetminus D) \times _{X \times S} {\widehat {Z}}^{{\rm aff}}, X\times S}} {\bf G}_{{\widehat {Z}},X\times S},\end{align*}

and

\begin{align*}\mathsf{L^{+}G}_X(S,F) \simeq {\bf G}_{\widehat {D},X\times S},\end{align*}

where for a map $V \to W$ ( $V$ being allowed to be a fibre functor on $W$ ), we write ${\bf G}_{V,W}$ for the functor

\begin{align*}{\bf G}_{V,W}: \mathsf{Aff}_{W}^{{\rm op}} \to {\bf Grp}, \,\,\, (T \to W) \longmapsto {\rm Hom}_{{\bf Sch}_{\mathbb{C}}}(T\times _{W} V, {\bf G})= {\rm Hom}_{{\bf Sch}_{W}}(T\times _{W} V, {\bf G}\times W).\end{align*}

Remark B.4. Quite generally, given $\mathcal{X} \to \mathcal{Y}$ , and $y: * \to \mathcal{Y}$ in ${\rm St}_T$ , the loop stack $\Omega _y\mathcal{Y}$ acts naturally on the fibre $\mathcal{X}_y=\mathcal{X}\times _{f,\mathcal{Y},y} \, *$ . In particular, the sheaf in groups $\mathsf{LG}_X(S,F)$ actsFootnote 34 on the stack ${\bf Gr}_X(S)(F)$ by changing the trivializations. Namely, if $(T \to X\times S) \in \mathsf{Aff}_Y$ , we have

\begin{align*}\mathsf{LG}_X(S,F)(T)&= \{ (\varphi , \psi ) \, | \varphi \in {\rm Aut}_{{\rm Bun}((\widehat {D}\smallsetminus D) \times _Y T)}(\mathcal{E}_0), \, \psi \in {\rm Aut}_{{\rm Bun}({\widehat {Z}}\times _Y T)}(\mathcal{E}_0) ,\, \varphi |_{(\widehat {D}\smallsetminus D) \times _{Y} {\widehat {Z}} \times _Y T}\\&= \psi |_{(\widehat {D}\smallsetminus D) \times _{Y} {\widehat {Z}} \times _Y T} \},\end{align*}

while

\begin{align*}\underline {\mathsf{Gr}}_X(S)(F)(T)&=\{ (\mathcal{E}, \phi , \chi ) \, |\, \mathcal{E} \in {\rm Bun}(\widehat {D}\times _Y T), \phi \;:\; \mathcal{E}|_{(\widehat {D}\smallsetminus D) \times _Y T)} \simeq \mathcal{E}_0, \, \chi \;:\; \mathcal{E}|_{{\widehat {Z}} \times _Y T} \simeq \mathcal{E}_0 ,\\ &\quad \phi |_{(\widehat {D}\smallsetminus D) \times _Y {\widehat {Z}}\times _Y T} =\chi |_{(\widehat {D}\smallsetminus D) \times _Y {\widehat {Z}}\times _Y T} \} \end{align*}

(with obvious morphisms). The (left) action of $\mathsf{LG}_X(S,F)(T)$ on ${\bf Gr}_X(S)(F)(T)$ is then given, on objects, by

\begin{align*}(\varphi , \psi )\cdot (\mathcal{E}, \phi , \chi ) := (\mathcal{E}, \varphi \circ \phi , \psi \circ \chi ).\end{align*}

This action should be considered as a higher-dimensional analogue of the well-known action of ${\bf G}_{\mathcal{K}_x}$ on the local affine Grassmannian at a point $x$ inside a curve (see [Reference Beilinson and DrinfeldBD99, 4.5] or [Reference Mirkovic and VilonenMV07, 2]). As a consequence of the above action, passing to global sections, the group $G_X(S,F)$ acts on the groupoid ${\rm Gr}_X(S)(F)$ (again by change of trivializations, with similar formulas).

Remark B.5. $\mathsf{L^{+}G}_X(S,F)$ acts on the stack ${\bf Gr}_X(S)(F)$ , by changing trivializations, the action being induced by the $\mathsf{LG}_X(S,F)$ -action of Remark B.4 together with the map $\mathsf{L^{+}G}_X(S,F) \to \mathsf{LG}_X(S,F)$ of Remark B.2. This action should be considered as a higher-dimensional analogue of the well-known action of ${\bf G}_{\mathcal{O}_x}$ on the local affine Grassmannian at a point $x$ inside a curve (see [Reference Beilinson and DrinfeldBD99, 4.5] or [Reference Mirkovic and VilonenMV07, 2]).

As a consequence of the above action, passing to global sections, the group $L^{+}G_X(S,F)$ acts on the groupoid ${\rm Gr}_X(S)(F)$ .

Proposition B.6. For any $S \in \mathsf{Aff}$ , we have functors

  1. (i)

    \begin{align*}\mathsf{L^{+}G}_X(S,-): \operatorname {\underline {\mathsf{Fl}}}_X(S) \longrightarrow {\bf ShGrp}^{{\rm op}}_{X\times S},\end{align*}
  2. (ii)

    \begin{align*}\mathsf{LG}_X(S,-):\operatorname {\underline {\mathsf{Fl}}}_X(S) \longrightarrow {\bf ShGrp}^{{\rm corr}}_{X\times S}, \end{align*}
  3. and a natural transformation $\mathsf{L^{+}G}_X(S,-) \to \mathsf{LG}_X(S,-)$ (see Remark B.2 ).

Proof. To ease notation, we put $Y=X\times S$ . Let $i: F_1=(D_1,Z_1) \to (D_2 , Z_2)=F_2$ be a morphism in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ . Since $\widehat {D}_1 \hookrightarrow \widehat {D}_2$ , (i) is obvious. In order to prove (ii), we observe that restrictions along the maps $\widehat {D}_1 \smallsetminus D_1 \to \widehat {D}_2 \smallsetminus D_1$ , ${\widehat {Z}}_1 \to {\widehat {Z}}_2$ , $(\widehat {D}_2 \smallsetminus D_2) \cap {\widehat {Z}}_2 \to (\widehat {D}_2 \smallsetminus D_1) \cap {\widehat {Z}}_2$ , $(\widehat {D}_1 \smallsetminus D_1) \cap {\widehat {Z}}_1 \to (\widehat {D}_2 \smallsetminus D_1) \cap {\widehat {Z}}_2$ yield maps

\begin{align*}{\bf Bun}_{\widehat {D}_2 \smallsetminus D_1} \to {\bf Bun}_{\widehat {D}_1 \smallsetminus D_1}\,, \,\,\, {\bf Bun}_{{\widehat {Z}}_2 } \to {\bf Bun}_{{\widehat {Z}}_1 },\end{align*}
\begin{align*}{\bf Bun}_{(\widehat {D}_2 \smallsetminus D_1) \cap {\widehat {Z}}_2} \to {\bf Bun}_{(\widehat {D}_2 \smallsetminus D_2) \cap {\widehat {Z}}_2}\,, \,\,\, {\bf Bun}_{(\widehat {D}_2 \smallsetminus D_1) \cap {\widehat {Z}}_2} \to {\bf Bun}_{(\widehat {D}_1 \smallsetminus D_1) \cap {\widehat {Z}}_1}.\end{align*}

Taking loops at $\mathcal{E}_0$ of the previous maps (and observing that taking loops at $\mathcal{E}_0$ commutes with fibre products of stacks over $X\times S$ ), we get a correspondence in ${\bf ShGrp}_{X\times S}$ ,

(B.1)

i.e. a morphismFootnote 35 $\mathsf{LG}_X(S,i): \mathsf{LG}_X(S,F_1) \to \mathsf{LG}_X(S,F_2)$ in ${\bf ShGrp}^{{\rm corr}}_{X\times S}$ . We leave to the reader verifying the existence of canonical compatibility isomorphismsFootnote 36 $\mathsf{LG}_X(S,i\circ j) \simeq \mathsf{LG}_X(S,i) \circ \mathsf{LG}_X(S,j)$ for composable morphisms $i, j$ in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ , i.e. that if are morphisms in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ , the following diagram is Cartesian:

the maps being the obvious restrictions.

The following result is a long but straightforward verification. Recall (e.g. [Reference RomagnyRom05, Definition 1.3]) that for a morphism of stacks endowed with an action of the a (pre)sheaf of groups $H$ , being an $H$ -equivariant morphism between $H$ -stacks consists of data and is not just a property of the morphism.

Lemma B.7. Let $S \in \mathsf{Aff}$ .

  • There is a canonical action of $\mathsf{LG}_X(S,-):\operatorname {\underline {\mathsf{Fl}}}_X(S) \longrightarrow {\bf ShGrp}^{{\rm corr}}_{X\times S}$ on ${\bf Gr}_X(S)(-):\operatorname {\underline {\mathsf{Fl}}}_X(S) \longrightarrow {\rm St}_{X\times S}$ . Precisely, for any $i \;:\; F_1 \to F_2$ in $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ , and any $T \in \mathsf{Aff}_{X\times S}$ , the morphism

    \begin{align*}{\bf Gr}_X(S)(\alpha )(T): {\bf Gr}_X(S)(F_1)(T) \to {\bf Gr}_X(S)(F_2)(T),\end{align*}
    has an $\mathsf{H}_X(S)(F_1,F_2)$ -equivariant structure, where $\mathsf{H}_X(S)(F_1,F_2)$ is the sheaf of groups on $X\times S$ defining the correspondence $\mathsf{LG}_X(S,i): \mathsf{LG}_X(S,F_1) \to \mathsf{LG}_X(S,F_2)$ ,

    (see diagram (B.1)).

  • Analogously, there is a canonical action of $\mathsf{L^+ G}_X(S,-):\operatorname {\underline {\mathsf{Fl}}}_X(S) \longrightarrow {\bf ShGrp}_{X\times S}$ on ${\bf Gr}_X(S)(-):\operatorname {\underline {\mathsf{Fl}}}_X(S) \longrightarrow {\rm St}_{X\times S}$ (see Remark B.5 and Proposition B.6 (iii)).

By applying the (limit-preserving) functor ${\bf St}_{X\times S} \to {\bf St}_{\mathbb{C}} \, \;:\; \mathcal{X} \mapsto \underline {\mathcal{X}}$ to $\mathsf{LG}_X(S,-)$ , $\mathsf{L^+ G}_X(S,-)$ , and ${\bf Gr}_X(S)$ , we obtain analogous actions of $\underline {\mathsf{LG}}_X(S,-)$ , $\underline {\mathsf{L^+ G}}_X(S,-)$ on $\underline {\mathsf{Gr}}_X(S)$ . We leave to the reader to establish the functoriality in $S$ of $\underline {\mathsf{LG}}_X(S,-)$ , $\underline {\mathsf{L^+G}}_X(S,-)$ , and of their actions on $\underline {\mathsf{Gr}}_X(S)$ .

B.2 Action on $\mathcal{G}r^{\flat }_{X}$

Definition B.8. Let ${\bf PoSets} \to {\bf Sets}$ denote the forgetful functor, sending a poset $(S, \leq )$ to the set $S$ . We denote by $\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X,\, \mathcal{G}r^{\flat }_X$ the composite functors

By integrating $\underline {\mathsf{L^+G}}_X(S,-)$ over all flags and all $S$ ’s, we now define an object $\mathcal{L}^{+}G_X \to \operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ and its action on $\mathcal{G}r^{\flat }_X \to \operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ .

Namely, let us consider the functorFootnote 37 $\mathsf{Aff}^{{\rm op}} \to {\bf Sets}$

\begin{align*}\mathcal{L}^{+}G_X\, \;:\; \,S \longmapsto \{ (F, \alpha ) \, | \, F\in \operatorname {\underline {\mathsf{Fl}}}^{\flat }_X(S), \, \alpha \in L^{+}G_X(S)(F) \}.\end{align*}

There is an obvious morphism of functors $h: \mathcal{L}^{+}G_X \to \operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ , such that, for any $F: S \to \operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ , the fibre of $h$ at such $F \in \operatorname {\underline {\mathsf{Fl}}}^{\flat }_X(S)$ is exactly the group $L^{+}G_X(S, F)$ of Definition B.1. We will say that $h: \mathcal{L}^{+}G_X \to \operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ is fibred in groups. By Proposition B.6, $\mathcal{L}^{+}G_X$ acts on $\mathcal{G}r^{\flat }_{X}$ over $\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ , i.e. for any $F: S \to \operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ the fibre of $h$ at $F$ acts on the fibre of $p_1: \mathcal{G}r^{\flat }_{X} \to \operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ at $F$ . In other words, there is a morphism

(B.2) \begin{align} {\rm act}:\mathcal{L}^{+}G_X \times _{\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X} \mathcal{G}r^{\flat }_{X} \longrightarrow \mathcal{G}r^{\flat }_{X} \end{align}

satisfying the usual axioms of a group action over $\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ . Note that this action is analogous to the family of actions of $(L^+G)_{X^I}$ on the Beilinson–Drinfeld $Gr_{X^I}$ for variable non-empty finite sets $I$ (see, e.g., [Reference ZhuZhu17, Proposition 3.19]).

Remark B.9. The authors currently ignore whether the action of Lemma B.7 or the one in (B.2) are, in a suitable sense, transitive (as it happens in the classical case of curves for the Ran Grassmannian). This question, suggested by the referee, surely deserves further investigation.

B.3 Equivariant sheaves

Using the action morphism (B.2)

\begin{align*}{\rm act}: \mathcal{L}G^{+}_X \times _{\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X} \mathcal{G}r^{\flat }_{X} \longrightarrow \mathcal{G}r^{\flat }_{X}\end{align*}

defined in § B, one can define the $\infty$ -category ${\bf Shv}^{\mathcal{L}G^{+}_X}(\mathcal{G}r^{\flat }_{X}; \mathbb{Q}_{\ell })$ (respectively, $\mathcal{D}{\bf \mbox{-}Mod}^{\mathcal{L}G^{+}_X}(\mathcal{G}r^{\flat }_{X})$ ) of $\mathcal{L}G^{+}_X$ -equivariant $\ell$ -adic sheaves on $\mathcal{G}r^{\flat }_{X}$ (respectively, of $\mathcal{L}G^{+}_X$ -equivariant $\mathcal{D}$ -Modules on $\mathcal{G}r^{\flat }_{X}$ ), and then prove that the naive chiral product of § $\,$ 8.1 can be upgraded to monoidal structures on ${\bf Shv}^{\mathcal{L}G^{+}_X}(\mathcal{G}r^{\flat }_{X}; \mathbb{Q}_{\ell })$ and on $\mathcal{D}{\bf \mbox{-}Mod}^{\mathcal{L}G^{+}_X}(\mathcal{G}r^{\flat }_{X})$ . These should be flag analogues of the equivariant chiral monoidal structure considered in [Reference ButsonBut20, Chapter 2] for the Ran space of a curve. These $\infty$ -categories and their monoidal structures, together with their non-naive counterparts (obtained by developing an equivariant version of § $\,$ 8.2), in the derived setting, will be investigated in a future work.

B.4 Kac–Moody groups

We conclude with an informal account relating our flag Grassmannian (for a fixed flag) with some version of a Kac–Moody group. We thus fix a surface $X$ and a flag $Z \subset D \subset X$ . The local flag Grassmannian $\underline {\mathsf{Gr}}^{{\rm loc}}_X(X; D, Z)$ is by definition the fibre product

\begin{align*} \underline {\mathsf{Gr}}^{{\rm loc}}_X(X; D, Z) = \underline {\mathsf{Bun}}_{\widehat D} \mathop{\times} _{\underline {\mathsf{Bun}}_{\widehat D \smallsetminus D \cup \widehat Z}} \{{\rm trivial}\}. \end{align*}

Using [Reference HennionHen17] and methods from [Reference Faonte, Hennion and KapranovFHK19, § 5.5], the stack $\underline {\mathsf{Bun}}_{\widehat D \smallsetminus D}$ comes equipped with a so-called determinantal $\mathbb G_{\mathrm m}$ -gerbe $[\mathsf{Det}] \to \underline {\mathsf{Bun}}_{\widehat D \smallsetminus D}$ . In induces in particular a central extension of the group stack of automorphisms of the trivial bundle

\begin{align*} 1 \to \mathbb G_{\mathrm m} \to \widetilde {\Omega _{\xi _0}\underline {\mathsf{Bun}}_{\widehat D \smallsetminus D}} \to \Omega _{\xi _0}\underline {\mathsf{Bun}}_{\widehat D \smallsetminus D} \to 1. \end{align*}

This provides an analogue to the Kac–Moody group extension to the case of flags. The extension restricts to an extension of the loop group $\mathcal{L}G_X({\rm Spec} {\mathbb{C}}, D, Z)$ . This extension then splits once restricted to the positive loop group $\mathcal{L}G^+_X({\rm Spec} {\mathbb{C}}, D, Z)$ (by an argument similar to that of [Reference Faonte, Hennion and KapranovFHK19, Proposition 5.5.4]), similarly to the classical case of curves.

Acknowledgments

Several people helped in various ways the realization of this paper. M. Porta started this project together with the first and third authors, and he contributed some important ideas that shaped the current paper; in particular, §§ 2 and 3 should be considered as joint work with him. M. Kapranov shared with us some of his stimulating ideas and visions on the subject of a Geometric Langlands program on higher-dimensional varieties. D. Beraldo was great in carefully and patiently explaining us some tricky details in the Geometric Langlands program. G. Nocera followed closely the recent evolution of this paper, and generously shared his ideas about the curve case. A. F. Herrero made very interesting comments, and suggested Lemma 5.12. S. Raskin gave us extremely useful and knowledgeable feedbacks: we have taken into account some of them in the current version (e.g. about the chiral tensor product in § 8), the rest will be addressed in a subsequent paper. A. Beilinson sent us a detailed proof of Proposition 7.2: we thank him a lot both for this and for his interest in our work. N. Rozenblyum and L. Ramero helped us in the proof of a preliminary underived version of Lemma 6.4. We owe a lot to D. Beraldo’s and D. Gaitsgory’s brilliant papers on the Geometric Langlands program. C. Barwick, R. Donagi, E. Elmanto, T. Pantev, M. Pippi and B. Toën offered valuable comments and encouragement. We sincerely thank all of them.

Conflicts of Interest

None.

Journal Information

Moduli is published as a joint venture of the Foundation Compositio Mathematica and the London Mathematical Society. As not-for-profit organisations, the Foundation and Society reinvest $100\%$ of any surplus generated from their publications back into mathematics through their charitable activities.

Footnotes

1 Actually, $\mathsf{Ran}_C$ is not even an étale sheaf.

2 M. Porta, together with the first and third authors, is also responsible for our first version of formal and flag glueing results (see arXiv:1607.04503).

3 This notion of fibre functor should not be confused with the one in the theory of Tannakian categories.

4 Here, we use the same symbol both for $\operatorname {\underline {\mathsf{Fl}}}_X$ and for its canonical derived extension (i.e. its Kan extension to derived affine schemes).

5 A stronger finiteness is obtained by restricting to the sub-functor $\underline {\mathsf{Car}}_{X} \hookrightarrow \operatorname {\underline {\mathsf{Fl}}}_{X}$ of relative effective Cartier divisors on $X$ (i.e. restricting to flags of type $(D, \emptyset )$ ): such a restriction is, in fact, represented by an ind-quasi-projective ind-scheme, hence of finite type.

6 For simplicity, we only state here the results for the global version ${\rm Gr}_X(S)$ .

7 This isomorphism is obviously compatible with the section $\mathsf{triv}: \operatorname {\underline {\mathsf{Fl}}}_X \to \mathcal{G}r_X$ of $p_1$ , introduced above. Moreover, Remark 6.18 gives the analogue of the natural cocycle condition on $(\mathsf{Ran}_C \times \mathsf{Ran}_C \times \mathsf{Ran}_C )_{{\rm disj}}$ satisfied by the factorization isomorphism for Beilinson–Drinfeld Grassmannian [Reference ZhuZhu17, Theorem 3.3.3].

8 Note that this result is clearly false for the Perf-version of the flag Grassmannian.

9 Note that in [Reference RaskinRas19, Remark 6.19.5] Raskin proves the equivalence between his approach with the Francis and Gaitsgory one.

10 However, this is peculiar to ${\bf G}$ -bundles, and it is not true for the Perf-version of the flag Grassmannian.

11 We have written here derived stacks over $k$ , using notation (1.3).

12 Hence we have corresponding algebras in correspondences in $\mathsf{dSt}/\underline {\mathsf{Bun}}_{X}$ .

13 And the same is true for the obvious corresponding Hecke stacks.

14 At the very end of § 3, the authors write ‘It is not clear how to think about the affine Grassmannian of ${\bf G}_{{\rm aff}}$ in terms of algebraic geometry’.

15 Another interesting way of interpreting $(\widehat {Z}^{\rm aff,\ \widehat\ {D}}\setminus Z)^{{\rm aff}} \setminus D$ is as the (suitably defined) triple intersection of the three ‘basic adelic’ opens associated with the flag $D,Z$ , i.e. $(X\times S)\setminus D$ , $\widehat {D} \setminus Z$ and $\widehat {Z}$ .

16 Note that even the definition of this category is not straightforward.

17 B. Toën, private communication.

18 Note that Mathew works in the context of ring spectra, while we are using simplicial rings. However, since the statement only deals with categories of modules, the difference between those contexts is harmless here. Indeed, the category of modules over a simplicial algebra agrees with that of modules over the geometric realization of said algebra in ring spectra.

19 Recall that by convention, ${\rm St}_X$ denotes the $(2,1)$ -category of $1$ -stacks over $X$ .

20 Since both $D\hookrightarrow X\times S$ and $Z\hookrightarrow X\times S$ are closed immersions, if there exists a morphism $Z\to D$ over $X\times S$ , then it is unique and it is a closed immersion. This justifies our apparently sloppy notation $Z\subset D$ .

21 Recall that, for a fibre functor $\mathfrak{F}$ over $Y$ , we denote by ${\rm Bun}({\mathfrak{F}})$ or, equivalently, by ${\rm Bun}({\mathfrak{F}}(Y))$ , the groupoid of global sections of the stack over $Y$ ${\bf Bun}_{{\mathfrak{F}}}$ , or equivalently the groupoid of global (i.e. over ${\rm Spec} \mathbb{C}$ ) sections of the corresponding stack $\underline {\mathsf{Bun}}_{{\mathfrak{F}}}\simeq u({\bf Bun}_{{\mathfrak{F}}})$ (see beginning of § 5.1).

22 For example, if there exists an effective Cartier divisor $D \hookrightarrow Y$ such that $D$ and $T$ share the same underlying topological subspace of $|Y|$ [Sta13, TAG07ZU].

23 Recall § 3.5. We are only using underived fibre functors here to avoid the hassle of building $\infty$ -functors while honest functors can do.

24 The ${\rm dR}$ lower script highlights that $\underline {\mathsf{Cl}}_{X,{\rm dR}}$ really is the de Rham stack of a more general construction.

25 Recall that a lax functor is a functor $F$ where the composition is not quite preserved. This lack of compatibility is replaced by a natural transformation evaluating to a map $F(f) \circ F(g) \to F(f \circ g)$ for each pair of composable arrows $(f,g)$ .

26 One shows easily with a little point-set topology that this condition is stable under (vertical) composition as well as base change.

27 We endow $\overline {D_2 \smallsetminus D_1}$ with its reduced scheme structure.

28 We use here, and in the previous display, the fact that a derived stack $\mathcal{X}$ and its truncation ${\rm t}\mathcal{X}$ share the same points with values in classical (i.e. underived) schemes.

29 We are viewing here underived stacks, classically, as pseudo-functors $({\rm Aff}^{{\rm Noeth}})^{{\rm op}} \to {\rm Grpds}$ . For the classical notions of pseudo-functors and pseudo-natural transformation between them, the reader may refer, e.g., to [Reference StreetStr96, § 9].

30 A 6-functor formalism, i.e. a 3-functors formalism in which all the three functors involved have adjoints, in the same spirit was defined previously in [Reference Gaitsgory and RozenblyumGR17a], and the reader might well prefer to use this source.

31 For direct and inverse images, we keep the notation of [Reference ScholzeSch22, Definition 2.4].

32 Both $\mathsf{PreStk}_{{\rm corr}}^{{\rm lax}}$ and $\mathsf{PreStk}^{{\rm lax}}$ can be seen as categories (i.e. $(\infty , 1)$ -categories) or as 2-categories (i.e. $(\infty , 2)$ -categories); we will specify when the 2-category structure is used.

33 Note that if $\mathcal{C}$ is a flag pre-chiral crystal category on $X$ , or if $\mathsf{A}$ is a $\mathcal{D}$ -module flag pre-factorization algebra on $X$ , we may allow any $S \in \mathsf{dAff}_{\mathbb{C}}$ , since by definition $\operatorname {\underline {\mathsf{Fl}}}^{{\rm dR}}_X(S)= \operatorname {\underline {\mathsf{Fl}}}_X ({\rm t}_0(S)_{{\rm red}}).$

34 In other words, the data we give below make ${\bf Gr}_X(S)(F)$ into a lax presheaf of groupoids on the classifying prestack ${\rm B}\mathsf{LG}_X(S)(F)$ . See, e.g., [Reference RomagnyRom05] for generalities about actions of (pre)sheaves of groups on (pre)stacks

35 Recall that in the category of correspondences in a category with fibre products, a morphism from a given source to a given target is the same thing as a morphism from the given target to the given source.

36 Note that ${\bf ShGrp}^{{\rm corr}}_{X\times S}$ is a 2-category in the usual way.

37 Since we aim at constructing an object acting on $\mathcal{G}r^{\flat }_{X}$ over $\operatorname {\underline {\mathsf{Fl}}}^{\flat }_X$ , we have chosen to disregard morphisms between flags, i.e. the partial order structure on $\operatorname {\underline {\mathsf{Fl}}}_X(S)$ .

References

Artin, M., Algebraization of formal moduli: II – existence of modifications, Ann. Math. 91 (1970), 88135.10.2307/1970602CrossRefGoogle Scholar
Beauville, A. and Laszlo, Y., Un lemme de descente, Sci. Paris Sér. I Math. 320 (1995), 335340.Google Scholar
Beilinson, A. and Drinfeld, V., Quantization of Hitchin’s integrable system and Hecke eigensheaves. Unpublished book (1999), http://www.math.uchicago.edu/∼drinfeld/langlands/QuantizationHitchin.pdf.Google Scholar
Beilinson, A. and Drinfeld, V., Chiral algebras, Colloquium Publications, vol. 51 (American Mathematical Society, 2004).Google Scholar
Ben-Bassat, O. and Temkin, M., Berkovich spaces and tubular descent, Adv. Math. 234 (2013), 217238, MR 1786197.10.1016/j.aim.2012.10.016CrossRefGoogle Scholar
Bhatt, B., Algebraization and Tannaka duality, Camb. J. Math. 4 (2016), 403461.10.4310/CJM.2016.v4.n4.a1CrossRefGoogle Scholar
Bhatt, B. and Halpern-Leistner, D., Tannaka duality revisited, Adv. Math. 316 (2017), 576612.10.1016/j.aim.2016.08.040CrossRefGoogle Scholar
Braverman, A. and Finkelberg, M., Pursuing the double affine Grassmannian, I: Transversal slices via instantons on ${A}_k$ -singularities , Duke Math. J. 152 (2010), 175206.10.1215/00127094-2010-011CrossRefGoogle Scholar
Braverman, A. and Finkelberg, M., Pursuing the double affine Grassmannian II: Convolution, Adv. Math. 230 (2012), 414432.10.1016/j.aim.2011.12.007CrossRefGoogle Scholar
Braverman, A. and Kazhdan, D., The spherical Hecke algebra for affine Kac-Moody groups I, Ann. Math. 174 (2011), 16031642.10.4007/annals.2011.174.3.5CrossRefGoogle Scholar
Butson, D., Equivariant localization in factorization homology and applications in mathematical physics I: Foundations, Preprint (2020), arXiv:2011.14988.Google Scholar
Dyckerhoff, T. and Kapranov, M., Higher Segal Spaces, Lecture Notes in Mathematics, vol. 2244 (Springer International Publishing, 2019).Google Scholar
Faonte, G., Hennion, B. and Kapranov, M., Higher Kac–Moody algebras and moduli spaces of G-bundles, Adv. Math. 346 (2019), 389466.10.1016/j.aim.2019.01.040CrossRefGoogle Scholar
Feller, M., Garner, R., Joachim, K., Proulx, M. U. and Weber, M., Every 2-Segal space is unital, Commun. Contemp. Math. 23 (2021), 2050055.10.1142/S0219199720500558CrossRefGoogle Scholar
Ferrand, D. and Raynaud, M., Fibres formelles d’un anneau local noethérien, Ann. Sci. École Norm. Sup. (4) 3 (1970), 295311.10.24033/asens.1195CrossRefGoogle Scholar
Francis, J. and Gaitsgory, D., Chiral Koszul duality, Select. Math. 18 (2012), 2787.10.1007/s00029-011-0065-zCrossRefGoogle Scholar
Gaitsgory, D., Contractibility of the space of rational maps, Invent. Math. 191 (2013), 91196.10.1007/s00222-012-0392-5CrossRefGoogle Scholar
Gaitsgory, D., The semi-infinite intersection cohomology sheaf-II: the Ran space version (Springer International Publishing, Cham, 2022), 151265.Google Scholar
Gaitsgory, D. and Rozenblyum, N., Dg indschemes, in Perspectives in representation theory, Contemporary Mathematics, vol. 610 (American Mathematical Society, 2014), 139251.10.1090/conm/610/12080CrossRefGoogle Scholar
Gaitsgory, D. and Rozenblyum, N., A study in derived algebraic geometry. Vol. I. Correspondences and duality, Mathematical Surveys and Monographs, vol. 221 (American Mathematical Society, Providence, RI, 2017).Google Scholar
Gaitsgory, D. and Rozenblyum, N., A study in derived algebraic geometry. Vol. II. Deformations, Lie theory and formal geometry, Mathematical Surveys and Monographs, vol. 221 (American Mathematical Society, Providence, RI, 2017).Google Scholar
Ginzburg, V., Kapranov, M. and Vasserot, É, Langlands reciprocity for algebraic surfaces, Math. Res. Lett. 2 (1995), 147160; MR 1786197.10.4310/MRL.1995.v2.n2.a4CrossRefGoogle Scholar
Grothendieck, A., Éléments de géométrie algébrique : IV. étude locale des schémas et des morphismes de schémas, Troisième partie, Publ. Math. l’IHÉS 28 (1966), 5255.10.1007/BF02684343CrossRefGoogle Scholar
Hall, J. and Rydh, D., Mayer-Vietoris squares in algebraic geometry, Preprint (2016), arXiv:1606.08517.Google Scholar
Halpern-Leinster, D., Herrero, A. F. and Jones, T., Moduli spaces of sheaves via affine Grassmannians, Journal für die reine und angewandte Mathematik (Crelles Journal) 809 (2024), 159215.Google Scholar
Hennion, B., Tate objects in stable ( $\infty$ ,1)-categories , Homol. Homotopy Appl. 19 (2017), 373395.10.4310/HHA.2017.v19.n2.a18CrossRefGoogle Scholar
Hennion, B., Melani, V. and Vezzosi, G., Flag Grassmannians , in preparation.Google Scholar
Kapranov, M. M. Analogies between the langlands correspondence and topological quantum field theory , in Functional Analysis on the Eve of the 21st Century: Volume I: In Honor of the Eightieth Birthday of I, eds Gelfand, S., Lepowsky, J. and Wilson, R. L. (Birkhäuser, Boston, 1995), 119151.Google Scholar
Katz, N. M. and Mazur, B., Arithmetic moduli of elliptic curves, Annals of Mathematics Studies, vol. 108.Google Scholar
Lurie, J., Tannaka duality for geometric stacks, Preprint (2004), arXiv:math/0412266.Google Scholar
Lurie, J., Higher topos theory, Annals of Mathematics Studies, vol. 170 (Princeton University Press, Princeton, NJ, 2009); MR 1786197.Google Scholar
Lurie, J., DAG XII: Proper morphisms, completions and the Grothendieck existence theorem, Preprint (2011), https://www.math.ias.edu/∼lurie/papers/DAG-XII.pdf.Google Scholar
Lurie, J., Higher algebra, Preprint (2017), https://www.math.ias.edu/∼lurie/papers/HA.pdf.Google Scholar
Lurie, J., Spectral algebraic geometry, Preprint (2018), https://www.math.ias.edu/~lurie/papers/SAG-rootfile.pdf.Google Scholar
Maffei, A., Melani, V. and Vezzosi, G., Two dimensional versions of the affine Grassmannian and their geometric description. Preprint (2025), arXiv:2503.16353.Google Scholar
Mathew, A., Faithfully flat descent of almost perfect complexes in rigid geometry, J. Pure Appl. Algebra 226 (2022), 106938.10.1016/j.jpaa.2021.106938CrossRefGoogle Scholar
Mirkovic, I. and Vilonen, K., Geometric Langlands duality and representations of algebraic groups over commutative rings, Ann. Math. 166 (2007), 95143.10.4007/annals.2007.166.95CrossRefGoogle Scholar
Moret-Bailly, L., Un problème de descente, Bull. Soc. Math. France 124 (1996), 559585.10.24033/bsmf.2293CrossRefGoogle Scholar
Raskin, S., Chiral categories, Preprint (2019), https://gauss.math.yale.edu/∼sr2532/chiralcats.pdf.Google Scholar
Richarz, T., Basics on the affine Grassmannian , Summer School Luminy 2019 (2019), https://timo-richarz.com/wp-content/uploads/2020/02/BoAG_02.pdf.Google Scholar
Romagny, M., Group actions on stacks and applications, Michigan Math. J. 53 (2005), 209236.10.1307/mmj/1114021093CrossRefGoogle Scholar
Schäppi, D., Descent via Tannaka duality, Preprint (2015), arXiv:1505.05681.Google Scholar
Scholze, P., Six-functor formalism, Lecture Notes (2022), https://people.mpim-bonn.mpg.de/scholze/SixFunctors.pdf.Google Scholar
Sernesi, E., Deformations of algebraic schemes , Grundlehren der Mathematischen Wissenschaften, vol. 334 (Springer, Berlin–Heidelberg, 2006).Google Scholar
The Stacks Project Authors, Stacks Project (2013), http://stacks.math.columbia.edu.Google Scholar
Stern, W. H., 2-Segal objects and algebras in spans, J. Homotopy Relat. Str. 16 (2021), 297361.10.1007/s40062-021-00282-8CrossRefGoogle Scholar
Street, R., Categorical structures, in Handbook of Algebra, vol. 1, ed. M. Hazewinkel (North-Holland, 1996), 529577.10.1016/S1570-7954(96)80019-2CrossRefGoogle Scholar
Toën, B., Derived algebraic geometry, EMS Surv. Math. Sci. 1 (2014), 153240.10.4171/emss/4CrossRefGoogle Scholar
Wang, J., The moduli space of G-bundles, Preprint (2011), arXiv:1104.4828.Google Scholar
Weil, A., Adèles and algebraic groups (Birkäuser, 1982).10.1007/978-1-4684-9156-2CrossRefGoogle Scholar
Zhu, X., An introduction to affine Grassmannians and the geometric Satake equivalence, in Geometry of Moduli Spaces and Representation Theory, IAS/Park City Mathematics Series, vol. 24 (American Mathematical Society, Providence, RI, 2017), 59154.10.1090/pcms/024/02CrossRefGoogle Scholar