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

Coherently Complete Algebraic Stacks in Positive Characteristic

Jarod Alper Jack Hall  and  David Benjamin Lim The University of Washington, Department of Mathematics, Box 354350, Seattle, WA 98195-4350, USA [email protected] School of Mathematics & Statistics
The University of Melbourne
Parkville, VIC, 3010
Australia
[email protected] [email protected]
Abstract.

With the long-term goal of proving local structure theorems of algebraic stacks in positive characteristic near points with reductive (but possibly non-linearly reductive) stabilizer, we conjecture that quotient stacks of the form [SpecA/G][\operatorname{Spec}A/G], with GG reductive and AGA^{G} complete local, are coherently complete along the unique closed point. We establish this conjecture in two interesting cases: (1) AGA^{G} is artinian and (2) GG acts trivially on SpecA\operatorname{Spec}A. We also establish coherent completeness results for graded unipotent group actions. In order to establish these results, we prove a number of foundational statements concerning cohomological and completeness properties of algebraic stacks—including on how these properties ascend and descend along morphisms.

1. Introduction

Let XX be a proper scheme over a complete noetherian local ring (R,𝔪R)(R,\mathfrak{m}_{R}). There are three important theorems that govern the formal geometry of XX, analogous to Grauert’s finiteness theorem and Serre’s GAGA theorems in the setting of complex analytic geometry:

  1. (1)

    Finiteness of Cohomology: For any coherent sheaf FF on XX, the cohomology group Hi(X,F)\mathrm{H}^{i}(X,F) is a finitely generated RR-module for all ii.

  2. (2)

    Formal Functions: Let X0:=XRR/𝔪RX_{0}:=X\otimes_{R}R/\mathfrak{m}_{R} to be the central fiber, XnX_{n} the nnth nilpotent thickening of X0X_{0}, and X^\widehat{X} the formal completion of XX along X0X_{0}. Then for any coherent sheaf FF on XX, there are natural isomorphisms:

    Hi(X,F)Hi(X^,F^)limnHi(Xn,F|Xn).\mathrm{H}^{i}(X,F)\simeq\mathrm{H}^{i}(\widehat{X},\widehat{F})\simeq\varprojlim_{n}\mathrm{H}^{i}(X_{n},F|_{X_{n}}).
  3. (3)

    Formal GAGA: There are equivalences of categories

    Coh(X)Coh(X^)limnCoh(Xn).\operatorname{Coh}(X)\stackrel{{\scriptstyle\sim}}{{\to}}\operatorname{Coh}(\widehat{X})\stackrel{{\scriptstyle\sim}}{{\to}}\varprojlim_{n}\operatorname{Coh}(X_{n}).

These properties and their extension to proper algebraic stacks [Con05, Ols05] are powerful tools in modern algebraic geometry. For example, they can be used to answer lifting questions about schemes to characteristic zero, or to prove the proper base change theorem in étale cohomology. In the context of moduli theory, Artin’s Criteria [Art74] implies that the effectivization of formal objects (e.g. curves, sheaves) is a necessary condition for a moduli stack to be algebraic. In practice, effectivization reduces to checking some incarnation of formal GAGA.

This paper investigates to what extent these three properties hold for (non-separated) algebraic stacks. Of primary interest are quotient stacks of the form [SpecA/G][\operatorname{Spec}A/G], where the invariant ring AGA^{G} is a complete local ring. To formulate our results, we introduce the following three definitions mirroring the three properties above:

  1. (1)

    A noetherian algebraic stack 𝒳\mathcal{X} over a noetherian ring RR is cohomologically proper if for every coherent sheaf \mathcal{F} on 𝒳\mathcal{X}, the cohomology group Hi(𝒳,)\mathrm{H}^{i}(\mathcal{X},\mathcal{F}) is a finitely generated RR-module for all ii.

Let 𝒳\mathcal{X} be a noetherian algebraic stack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} a closed substack. Let 𝒳^\widehat{\mathcal{X}} be the completion of 𝒳\mathcal{X} along 𝒳0\mathcal{X}_{0}; i.e. the ringed site (𝒳lis-ét,𝒪^𝒳,𝒳0)(\mathcal{X}_{\text{lis-\'{e}t}},\widehat{\mathcal{O}}_{\mathcal{X},\mathcal{X}_{0}}), where 𝒪^𝒳,𝒳0lim𝒪𝒳/n+1\widehat{\mathcal{O}}_{\mathcal{X},\mathcal{X}_{0}}\coloneqq\varprojlim\mathcal{O}_{\mathcal{X}}/\mathcal{I}^{n+1} (see Definition 3.1).

  1. (2)

    We say that the pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions if for every coherent sheaf \mathcal{F} on 𝒳\mathcal{X}, the natural map

    Hi(𝒳,)Hi(𝒳^,^)\mathrm{H}^{i}(\mathcal{X},\mathcal{F})\to\mathrm{H}^{i}(\widehat{\mathcal{X}},\widehat{\mathcal{F}})

    is an isomorphism for all ii.

  2. (3)

    We say that the pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) is coherently complete if the natural functor

    Coh(𝒳)Coh(𝒳^)\operatorname{Coh}(\mathcal{X})\to\operatorname{Coh}(\widehat{\mathcal{X}})

    is an equivalence of categories.

See Sections 2 and 3 for a further discussion of these definitions as well as examples.

1.1. Main conjecture and results

Conjecture 1.1.

Let GG be a smooth geometrically reductive group scheme over a complete noetherian local ring RR. Let X=SpecAX=\operatorname{Spec}A be an affine scheme of finite type over RR with an action of GG such that AG=RA^{G}=R. Let x[X/G]x\in[X/G] be the unique closed point and 𝒢x\mathcal{G}_{x} be its residual gerbe. Then [X/G][X/G] is cohomologically proper over AGA^{G}, and the pair ([X/G],𝒢x)([X/G],\mathcal{G}_{x}) is coherently complete and satisfies formal functions.

This conjecture differs from formal GAGA in two ways. First, when GG is not finite, the quotient stack [SpecA/G][\operatorname{Spec}A/G] is not separated, let alone proper, over SpecAG\operatorname{Spec}A^{G}. Second, in formal GAGA, the completion is taken with respect to the central fiber (i.e. the pullback of the maximal ideal from the base), whereas the completion in Theorem 1.2 is taken with respect to the smaller closed substack defined by the residual gerbe at the unique closed point. For instance, under the standard action of SL2\mathrm{SL}_{2} on 𝐀2=Speck[x,y]\mathbf{A}^{2}=\operatorname{Spec}k[x,y], the origin is the unique closed orbit and the invariant ring is k[x,y]SL2=kk[x,y]^{\mathrm{SL}_{2}}=k. Since the central fiber of [𝐀2/SL2]Speck[\mathbf{A}^{2}/\mathrm{SL}_{2}]\to\operatorname{Spec}k is everything, the coherent completeness of [𝐀2/SL2][\mathbf{A}^{2}/\mathrm{SL}_{2}] along the central fiber has no mathematical content. On the other hand, the coherent completeness of [𝐀2/SL2][\mathbf{A}^{2}/\mathrm{SL}_{2}] along the origin BSL2B\mathrm{SL}_{2} is a non-trivial statement.

This conjecture was established in [AHR19, Thm. 1.6] in the case that GG is linearly reductive. While in characteristic 0 the notions of linearly reductivity and reductivity agree, in positive characteristic linear reductivity is a very strong notion: a smooth affine group scheme is linearly reductive if it is an extension of a torus by a finite group whose order is prime to the characteristic [Nag62]. For example GLn\mathrm{GL}_{n} is reductive in any characteristic but linearly reductive only in characteristic 0.

Theorem 1.2.

1.1 holds in the following cases

  1. (1)

    RR is an artinian local ring, or

  2. (2)

    GG acts trivially on XX, i.e. A=RA=R and [X/G]=BG[X/G]=BG.

We expect that these two cases will assist in establishing 1.1 in general. Indeed, (1) reduces the statement that [X/G][X/G] is coherently complete and satisfies formal functions with respect to the residual gerbe 𝒢x\mathcal{G}_{x} to seemingly simpler statement that [X/G][X/G] satisfies these properties with respect to the central fiber of [SpecA/G]SpecAG[\operatorname{Spec}A/G]\to\operatorname{Spec}A^{G}. Our methods essentially reduce 1.1 to the ‘generically toric’ situation (Remark 8.2).

In fact, we establish a stronger result than (2): if GG is a geometrically reductive group scheme over an II-adically complete noetherian ring RR, then BGBG is cohomologically proper over RR, and the pair (BG,BGR/I)(BG,BG_{R/I}) is coherently complete and satisfies formal functions (Theorem 8.3). When GG is the base change of a reductive group defined over a field, this had been proved in [HLP23, Prop. 4.3.4].

It turns out that our results hold for a wider class than reductive group schemes: they hold for graded unipotent groups, which are precisely the type of groups that arise in non-reductive GIT [BDHK18].

Theorem 1.3.

Let GSpecRG\to\operatorname{Spec}R be a smooth affine group scheme over a complete local noetherian ring RR with residue field kk. Let GuGG_{u}\subset G be its unipotent radical with reductive quotient Gr=G/GuG_{r}=G/G_{u}. Let λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G be a one-parameter subgroup such that λ\lambda is central in GrG_{r} and acts positively on GuG_{u}. Then

  1. (1)

    BGBG is cohomologically proper over RR, and (BG,BGk)(BG,BG_{k}) is coherently complete and satisfies formal functions; and

  2. (2)

    if GG acts on an affine scheme X=SpecAX=\operatorname{Spec}A of finite type over RR such that AG=RA^{G}=R, [X/Gr][X/G_{r}] satisfies 1.1 (e.g., GrG_{r} is linearly reductive or RR is artinian) and λ\lambda acts semipositively on AA, then [X/G][X/G] is cohomologically proper, and there is a unique closed point x[X/G]x\in[X/G] such that ([X/G],𝒢x)([X/G],\mathcal{G}_{x}) is coherently complete and satisfies formal functions.

For example, if BGB\subset G is a Borel subgroup of a split reductive group scheme over RR, then B=PλB=P_{\lambda} for a regular one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G. The reductive quotient Br=TB_{r}=T of BB is a maximal torus (thus linearly reductive) such that λ\lambda is central. The theorem implies that BBBB is cohomologically proper and that (BB,BBk)(BB,BB_{k}) is coherently complete, and moreover that the same holds for a quotient stack [SpecA/G][\operatorname{Spec}A/G] if AB=RA^{B}=R and λ\lambda acts semipositively on AA.

On the other hand, these three cohomological properties do not hold for every algebraic group GG, e.g. unipotent groups (see Example 5.11).

1.2. Methods and other results

Our strategy to establish Theorem 1.2 is via descent. We prove that if f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} is a universally submersive morphism and 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} is a closed substack such that 𝒳\mathcal{X} and the higher base changes 𝒳×𝒴×𝒴𝒳\mathcal{X}\times_{\mathcal{Y}}\cdots\times_{\mathcal{Y}}\mathcal{X} satisfy the three cohomological properties—cohomologically properness, formal functions, and coherent completeness—along the preimage of 𝒴0\mathcal{Y}_{0}, then 𝒴\mathcal{Y} satisfies these properties along 𝒴0\mathcal{Y}_{0} (Theorem 6.1). The two most important cases are either when ff is proper and surjective, in which case one only needs to check that 𝒳\mathcal{X} has the properties along the preimage of 𝒴0\mathcal{Y}_{0}, and when ff is faithfully flat.

To apply the descent result, we construct suitable covers of [SpecA/G][\operatorname{Spec}A/G]. For the case of BGRBG_{R} for a reductive group GG over a complete local noetherian ring RR (i.e. Theorem 1.2(2)), we first reduce to the case that GG is split reductive. We consider subgroups TBGT\subset B\subset G where TT is a maximal torus and BB is a Borel subgroup, and consider the induced morphisms

BTBBBG.BT\to BB\to BG.

Since G/BG/B is projective, the map BBBGBB\to BG is proper and surjective, and hence Theorem 6.1 reduces the theorem to BBBB. On the other hand, BTBBBT\to BB is faithfully flat and affine. Since TT is a torus and hence linearly reductive, we know the cohomological properties for BTBT. Moreover, the higher base changes of BTBBBT\to BB are identified with quotient stacks [U×(n1)/T][U^{\times(n-1)}/T] where UU is the quotient B/TB/T of the right action, and TT acts on UU on the left. The TT-invariants of the action on the (n1)(n-1)-fold product U×(n1)U^{\times(n-1)} are identified with UU, which implies that [U×(n1)/T]SpecR[U^{\times(n-1)}/T]\to\operatorname{Spec}R is a good moduli space and thus satisfies the three cohomological properties. Theorem 6.1 then implies that BBBB satisfies the desired properties.

The construction of the covers in the case of [SpecA/G][\operatorname{Spec}A/G] with AGA^{G} artinian (i.e., Theorem 1.2(1)) is more involved. The general strategy is to find a one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G and consider the pair of morphisms

[Xλ+/Cλ][Xλ+/Pλ][X/G],[X^{+}_{\lambda}/C_{\lambda}]\to[X^{+}_{\lambda}/P_{\lambda}]\to[X/G],

where Xλ+=Hom𝐆m(𝐀1,X)X^{+}_{\lambda}={\rm Hom}^{\mathbf{G}_{m}}(\mathbf{A}^{1},X) is the ‘attractor’ locus parameterizing points of XX that have a limit under 𝐆m\mathbf{G}_{m}-action induced by λ\lambda (see §7.1), and where CλC_{\lambda} and PλP_{\lambda} are the centralizer and parabolic of λ\lambda (see §7.2). For any λ\lambda, the map [Xλ+/Pλ][X/G][X^{+}_{\lambda}/P_{\lambda}]\to[X/G] is proper since Xλ+XX^{+}_{\lambda}\subset X is a closed subscheme and G/PλG/P_{\lambda} is projective, while the map [Xλ+/Cλ][Xλ+/Pλ][X^{+}_{\lambda}/C_{\lambda}]\to[X^{+}_{\lambda}/P_{\lambda}] is faithfully flat and affine. A version of the Hilbert–Mumford criterion (Proposition 7.11) implies that there is one-parameter subgroup λ\lambda destabilizing at least one point in every GG-orbit, which gives that GXλ+=XG\cdot X^{+}_{\lambda}=X or in other words that [Xλ+/Pλ][X/G][X^{+}_{\lambda}/P_{\lambda}]\to[X/G] is surjective. The descent result (Theorem 6.1) therefore reduces to the claim to [Xλ+/Pλ][X^{+}_{\lambda}/P_{\lambda}]. However, [Xλ+/Cλ][X^{+}_{\lambda}/C_{\lambda}] (in addition to the higher base changes of [Xλ+/Cλ][Xλ+/Pλ][X^{+}_{\lambda}/C_{\lambda}]\to[X^{+}_{\lambda}/P_{\lambda}]) do not clearly satisfy the cohomological properties as CλC_{\lambda}, while reductive, is not linearly reductive. However, we establish a refined version of the Hilbert–Mumford criterion (Proposition 7.14) that yields a one-parameter subgroup λ\lambda such that every GG-orbit contains a point xx and such that x0:=limt0λ(t)xx_{0}:=\lim_{t\to 0}\lambda(t)\cdot x exists and lies in the unique closed GG-orbit, and moreover such that the induced one-parameter subgroup is regular in the stabilizer Gx0G_{x_{0}}. This implies that the connected component of Gx0CλG_{x_{0}}\cap C_{\lambda} is linearly reductive, which suffices to show that [Xλ+/Cλ][X^{+}_{\lambda}/C_{\lambda}] and the higher base changes satisfy the desired cohomological properties.

We also establish a number of other foundational results, such as:

  • every strongly cohomologically proper morphism (see Definition 4.1) satisfies formal functions (Theorem 4.6); and

  • coherent completeness ascends along proper and representable morphisms: if f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} is a proper representable morphism and (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}_{0}) is coherently complete, then (𝒳,f1(𝒴0))(\mathcal{X},f^{-1}(\mathcal{Y}_{0})) is coherently complete (Theorem 5.4).

1.3. Motivation and applications

The motivation for this paper is to extend the local structure theorem of [AHR20, Thm. 1.1] to apply more widely in positive characteristic. Specifically, [AHR20, Thm. 1.1] implies that a quasi-separated algebraic stack of finite type over an algebraically closed field with affine stabilizers is étale locally a quotient stack of the form [SpecA/GLn][\operatorname{Spec}A/\mathrm{GL}_{n}] near a point xx with linearly reductive stabilizer GxG_{x}; various generalizations of this were established in [AHR19] and [AHHR23]. The long-term goal of this project is to prove an analogous result near points with reductive stabilizer. Such a local structure theorem would be a powerful foundational result in stack theory with wide-ranging applications. Most notably, it would immediately extend the existence theorems for moduli spaces proved in [AHLH23, Thm. A] to positive characteristic. Namely, it would imply that in any characteristic, an algebraic stack 𝒳\mathcal{X} of finite type with affine diagonal admits a separated adequate moduli space if and only 𝒳\mathcal{X} is Θ\Theta-complete and 𝖲\mathsf{S}-complete.

Coherent completeness plays an essential role in the proof of [AHR20, Thm. 1.1]. The main idea of the proof in the case that x𝒳x\in\mathcal{X} is a smooth point is to consider the quotient 𝒯=[T𝒳,x/Gx]\mathcal{T}=[T_{\mathcal{X},x}/G_{x}] of the Zariski tangent space, and the base change

𝒯^:=𝒯×T𝒳,x//GxSpec𝒪^T𝒳,x//Gx,0\widehat{\mathcal{T}}:=\mathcal{T}\times_{T_{\mathcal{X},x}//G_{x}}\operatorname{Spec}\widehat{\mathcal{O}}_{T_{\mathcal{X},x}//G_{x},0}

along the completion of the GIT quotient at the image of the origin. Since 𝒯^\widehat{\mathcal{T}} is coherently complete along BGxBG_{x} by the linearly reductive case of 1.1 (i.e. [AHR19, Thm. 1.6]), Tannaka duality [HR19] implies that

Hom(𝒯^,𝒳)limHom(𝒯n,𝒳).\operatorname{Hom}(\widehat{\mathcal{T}},\mathcal{X})\cong\varprojlim\operatorname{Hom}(\mathcal{T}_{n},\mathcal{X}).

Using again that GxG_{x} is linearly reductive, deformation theory yields isomorphisms 𝒯n𝒳n\mathcal{T}_{n}\cong\mathcal{X}_{n} of the nnth nilpotent thickening of 0𝒯0\in\mathcal{T} and x𝒳x\in\mathcal{X}. The identification above implies that the maps 𝒯n𝒳n𝒳\mathcal{T}_{n}\stackrel{{\scriptstyle\sim}}{{\to}}\mathcal{X}_{n}\hookrightarrow\mathcal{X} extend to a map 𝒯^𝒳\widehat{\mathcal{T}}\to\mathcal{X}, which can then be approximated to produce the desired étale neighborhoods. In the reductive case, 1.1 would yield the coherent completeness of 𝒯^\widehat{\mathcal{T}} and be a big step in the direction of a local structure theorem. However, even with 1.1 resolved, deformation theory would remain an additional obstruction to establish a local structure theorem in the reductive case.

1.4. Comparison to the literature

The paper [HLP23] discusses related cohomological properties of stacks and establishes a number of related results. The authors introduce the notion of formal properness (see Remark 3.4) and establish for instance that cohomologically projective morphisms [HLP23, Def. 4.1.3] are formally proper [HLP23, Thm. 4.2.1], which is a variant of Theorem 5.4 (see Remark 5.5). They also prove Theorem 1.2(2) in the case that GG is the base change of a reductive group over a field [HLP23, Prop. 4.3.4].

One major difference between our paper and [HLP23] is that we are concerned with establishing coherent completeness of [SpecA/G][\operatorname{Spec}A/G] along the residual gerbe of the unique closed point. In contrast, [HLP23] focuses on coherent completeness results of stacks such as [SpecA/G][\operatorname{Spec}A/G] along the central fiber of a morphism [SpecA/G]SpecAG[\operatorname{Spec}A/G]\to\operatorname{Spec}A^{G}. In this sense, our results are substantially stronger than [HLP23] and are necessary for the intended applications to local structure theorems. Another feature of our paper is that we utilize only classical methods and avoid derived algebraic geometry.

1.5. Acknowledgements

We would like to thank Dmitry Kubrak, Zev Rosengarten, Sean Cotner, Bogdan Zavyalov, Brian Conrad, Jochen Heinloth, Dan Halpern-Leistner, and Elden Elmanto for very helpful conversations.

1.6. Conventions

We follow the conventions of [Sta21]. In particular, if 𝒳\mathcal{X} is an algebraic stack, the lisse-étale site 𝒳lisse,étale\mathcal{X}_{\textrm{lisse,\'{e}tale}} denotes the category with objects the morphisms U𝑝𝒳U\xrightarrow{p}\mathcal{X}, where UU is a scheme and pp is smooth; coverings of U𝒳U\to\mathcal{X} are given by étale coverings of UU [Sta21, Tag 0787].

2. Cohomological properness

For a morphism f:XYf\colon X\to Y of noetherian schemes (even algebraic spaces or algebraic stacks with finite diagonal), it is known that ff is proper if and only if the higher pushforwards 𝖱if\mathsf{R}^{i}f_{*} send coherent sheaves to coherent sheaves for all i0i\geq 0 [Ryd]. For algebraic stacks with infinite stabilizers, there does not exist such a comparatively simple characterization. It was noted, however, in [HLP23, Def. 2.4.1] that the higher pushforwards of a morphism preserving coherent sheaves was a useful condition and satisfied for various interesting types of algebraic stacks. In this vein, we make the following definition.

Definition 2.1.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a finite type morphism of noetherian algebraic stacks. We say that ff is:

  1. (1)

    cohomologically proper if for every coherent sheaf \mathcal{F} on 𝒳\mathcal{X}, 𝖱fDcoh+(𝒴)\mathsf{R}f_{\ast}\mathcal{F}\in D^{+}_{\operatorname{coh}}(\mathcal{Y}); and

  2. (2)

    universally cohomologically proper if for every morphism 𝒴𝒴\mathcal{Y}^{\prime}\to\mathcal{Y} of noetherian algebraic stacks, the base change 𝒳×𝒴𝒴𝒴\mathcal{X}\times_{\mathcal{Y}}\mathcal{Y}^{\prime}\to\mathcal{Y}^{\prime} is cohomologically proper.

Remark 2.2.

Universal cohomological properness is equivalent to the coherent pushforward property introduced in [HLP23, Def. 2.4.1].

Remark 2.3.

In Definition 2.1, by descent, it suffices to check cohomological properness on a smooth cover of the target. In particular, universal cohomological properness can be verified by only base changing to affine schemes.

Remark 2.4.

If f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} is cohomologically proper and Dcoh+(𝒳)\mathcal{F}\in D^{+}_{\operatorname{coh}}(\mathcal{X}), then 𝖱fDcoh+(𝒴)\mathsf{R}f_{*}\mathcal{F}\in D^{+}_{\operatorname{coh}}(\mathcal{Y}). This follows from the convergent hypercohomology spectral sequence:

𝖱ifj()i+j(𝖱f).\mathsf{R}^{i}f_{\ast}\mathscr{H}^{j}(\mathcal{F})\Rightarrow\mathscr{H}^{i+j}(\mathsf{R}f_{\ast}\mathcal{F}).
Remark 2.5.

If f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} is universally cohomologically proper, then it is universally closed (e.g., [Ryd] and [HLP23, Prop. 2.4.5]). The basic idea is to apply to the valuative criterion for universal closedness [Sta21, Tag 0CLV], so we are reduced to the situation where 𝒴=SpecD\mathcal{Y}=\operatorname{Spec}D, where (D,𝔪)(D,\mathfrak{m}) is a DVR with fraction field KK such that fK:𝒳KSpecKf_{K}\colon\mathcal{X}_{K}\to\operatorname{Spec}K has a section ss. We must show that, potentially after extending DD, that ss extends to a section over SpecD\operatorname{Spec}D. If pf1(𝔪)p\in f^{-1}(\mathfrak{m})\neq\emptyset, then we can find a DVR DD^{\prime} over DD such that SpecD𝒳\operatorname{Spec}D^{\prime}\to\mathcal{X} realizes the specialization from the image of ss to pp. In particular, we must show that f1(𝔪)=f^{-1}(\mathfrak{m})=\emptyset cannot occur if ff is cohomologically proper. If f1(𝔪)=f^{-1}(\mathfrak{m})=\emptyset, then ff factors through j:SpecKSpecRj\colon\operatorname{Spec}K\to\operatorname{Spec}R. It follows that H0(𝒳,𝒪𝒳)\mathrm{H}^{0}(\mathcal{X},\mathcal{O}_{\mathcal{X}}) is a non-zero finite dimensional KK-vector space, so it cannot be a finitely generated RR-module, which contradicts the cohomological properness of ff. This proves the claim.

The following two examples will be key in this article.

Example 2.6.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a proper morphism of noetherian algebraic stacks. Then ff is universally cohomologically proper. Since properness is stable under arbitrary base change, it suffices to prove that proper morphisms are cohomologically proper, which follows from [Ols05].

Example 2.7.

Let RR be a noetherian ring. Let AA be a finitely generated RR-algebra with an action of a reductive group scheme GG over SpecR\operatorname{Spec}R. Then π:[SpecA/G]Spec(AG)\pi\colon[\operatorname{Spec}A/G]\to\operatorname{Spec}(A^{G}) is universally cohomologically proper. Indeed, by Remark 2.3, it suffices to check that if BB is a noetherian AGA^{G}-algebra, then πB:[Spec(AAGB)/GB]Spec(B)\pi_{B}\colon[\operatorname{Spec}(A\otimes_{A^{G}}B)/G_{B}]\to\operatorname{Spec}(B) is cohomologically proper. But AAGBA\otimes_{A^{G}}B is a finitely generated BB-algebra and the result follows from [vdK15, Thm. 10.5].

3. Formal functions and coherent completeness

Throughout we use the following notation: if 𝒳\mathcal{X} is a noetherian algebraic stack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} is a closed substack defined by a coherent sheaf of ideals 𝒪𝒳\mathcal{I}\subset\mathcal{O}_{\mathcal{X}}, then we denote by 𝒳n𝒳\mathcal{X}_{n}\subset\mathcal{X} the closed substack defined by n+1\mathcal{I}^{n+1}. If \mathcal{F} is a quasi-coherent sheaf on 𝒳\mathcal{X}, we denote by n\mathcal{F}_{n} its pullback to 𝒳n\mathcal{X}_{n}.

Definition 3.1.

Let 𝒳\mathcal{X} be a noetherian algebraic stack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} be a closed substack defined by a coherent sheaf of ideals 𝒪𝒳\mathcal{I}\subset\mathcal{O}_{\mathcal{X}}. The completion of the pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) is the ringed site 𝒳^(𝒳lis-ét,𝒪^𝒳,𝒳0)\widehat{\mathcal{X}}\coloneqq(\mathcal{X}_{\text{lis-\'{e}t}},\widehat{\mathcal{O}}_{\mathcal{X},\mathcal{X}_{0}}), where the sheaf of rings is defined by the limit

𝒪^𝒳,𝒳0lim𝒪𝒳/n+1\widehat{\mathcal{O}}_{\mathcal{X},\mathcal{X}_{0}}\coloneqq\varprojlim\mathcal{O}_{\mathcal{X}}/\mathcal{I}^{n+1}

in the category of lisse-étale modules.

There is a canonical morphism of ringed sites c𝒳,𝒳0:𝒳^𝒳c_{\mathcal{X},\mathcal{X}_{0}}\colon\widehat{\mathcal{X}}\to\mathcal{X}, which is flat [GZB15, Lem. 3.3] and whose pullback preserves coherence; we will often suppress the subscript by writing c𝒳:𝒳^𝒳c_{\mathcal{X}}\colon\widehat{\mathcal{X}}\to\mathcal{X} or simply c:𝒳^𝒳c\colon\widehat{\mathcal{X}}\to\mathcal{X} if there is little possibility for confusion. For a coherent 𝒪𝒳\mathcal{O}_{\mathcal{X}}-module \mathcal{F}, we denote by ^\widehat{\mathcal{F}} the pullback of \mathcal{F} to 𝒳^\widehat{\mathcal{X}}. There is a natural identification ^lim/n+1\widehat{\mathcal{F}}\cong\varprojlim\mathcal{F}/\mathcal{I}^{n+1}\mathcal{F}. There is also an exact equivalence of abelian categories Coh(𝒳^)limCoh(𝒳^n)\operatorname{Coh}(\widehat{\mathcal{X}})\simeq\varprojlim\operatorname{Coh}(\widehat{\mathcal{X}}_{n}) [Con05, Thm. 2.3].

Definition 3.2.

Let 𝒳\mathcal{X} be a noetherian algebraic stack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} be a closed substack.

  1. (1)

    (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions if for any coherent sheaf \mathcal{F} on 𝒳\mathcal{X}, the natural map:

    Hi(𝒳,)Hi(𝒳^,^)\mathrm{H}^{i}(\mathcal{X},\mathcal{F})\to\mathrm{H}^{i}(\widehat{\mathcal{X}},\widehat{\mathcal{F}})

    is an isomorphism for all i0i\geq 0;

  2. (2)

    (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) is coherently complete if the functor:

    Coh(𝒳)Coh(𝒳^)limCoh(𝒳^n)\operatorname{Coh}(\mathcal{X})\to\operatorname{Coh}(\widehat{\mathcal{X}})\simeq\varprojlim\operatorname{Coh}(\widehat{\mathcal{X}}_{n})

    is an equivalence.

  3. (3)

    (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies derived formal functions if the functor:

    c:Dcoh(𝒳)Dcoh(𝒳^)c^{*}\colon D_{\operatorname{coh}}(\mathcal{X})\to D_{\operatorname{coh}}(\widehat{\mathcal{X}})

    is fully faithful.

  4. (4)

    (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) is derived coherently complete if the functor:

    c:Dcoh(𝒳)Dcoh(𝒳^)c^{*}\colon D_{\operatorname{coh}}(\mathcal{X})\to D_{\operatorname{coh}}(\widehat{\mathcal{X}})

    is an equivalence.

Examples are littered throughout the next few sections, with several in Section 5.

Remark 3.3.

The traditional formulation of “formal functions” in the literature is that Hi(𝒳,)limnHi(𝒳n,n)\mathrm{H}^{i}(\mathcal{X},\mathcal{F})\simeq\varprojlim_{n}\mathrm{H}^{i}(\mathcal{X}_{n},\mathcal{F}_{n}) for all i0i\geq 0 and coherent sheaves \mathcal{F} on 𝒳\mathcal{X}. We will refer to this as Zariski formal functions. If Coh(𝒳)\mathcal{F}\in\operatorname{Coh}(\mathcal{X}), then ^=holimnn\widehat{\mathcal{F}}=\mathrm{holim}_{{n}}\mathcal{F}_{n} in D(𝒳)D(\mathcal{X}) [Sta21, Tag 0A0K]. In particular, 𝖱Γ(𝒳^,^)holimn𝖱Γ(𝒳n,n)\mathsf{R}\Gamma(\widehat{\mathcal{X}},\widehat{\mathcal{F}})\simeq\mathrm{holim}_{{n}}\mathsf{R}\Gamma(\mathcal{X}_{n},\mathcal{F}_{n}). The Milnor exact sequence [Sta21, Tag 07KZ] implies that for each ii there is a short exact sequence:

(3.1) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}limn1Hi1(𝒳n,n)\textstyle{\varprojlim^{1}_{n}\mathrm{H}^{i-1}(\mathcal{X}_{n},\mathcal{F}_{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hi(𝒳^,^)\textstyle{\mathrm{H}^{i}(\widehat{\mathcal{X}},\widehat{\mathcal{F}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}limnHi(𝒳n,n)\textstyle{\varprojlim_{n}\mathrm{H}^{i}(\mathcal{X}_{n},\mathcal{F}_{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

It follows that formal functions implies that Hi(𝒳,)limnHi(𝒳n,n)\mathrm{H}^{i}(\mathcal{X},\mathcal{F})\twoheadrightarrow\varprojlim_{n}\mathrm{H}^{i}(\mathcal{X}_{n},\mathcal{F}_{n}) for all ii and \mathcal{F}. Conversely, Zariski formal functions implies that Hi(𝒳,)Hi(𝒳^,^)\mathrm{H}^{i}(\mathcal{X},\mathcal{F})\subseteq\mathrm{H}^{i}(\widehat{\mathcal{X}},\widehat{\mathcal{F}}). Hence, these conditions are not obviously equivalent in the generality that we are working. They are when i=0i=0 or when 𝒳\mathcal{X} has affine diagonal, however. This follows from an identical argument to that provided in [Knu71, Cor. V.2.20] for separated algebraic spaces.

Remark 3.4 (Formal properness).

In [HLP23, Defn. 1.1.3], the property of formal properness is introduced, which is related to the above definition of derived coherent completeness. A pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) consisting of an algebraic stack 𝒳\mathcal{X} and a cocompact closed substack 𝒳0\mathcal{X}_{0} is called complete if APerf(𝒳)APerf(𝒳^)\operatorname{APerf}(\mathcal{X})\stackrel{{\scriptstyle\sim}}{{\to}}\operatorname{APerf}(\widehat{\mathcal{X}}) is an equivalence. A morphism f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} of algebraic stacks is formally proper if for every complete pair (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}^{\prime}_{0}), the base change 𝒳×𝒴𝒴\mathcal{X}\times_{\mathcal{Y}}\mathcal{Y}^{\prime} is complete with respect to 𝒳×𝒴𝒴0\mathcal{X}\times_{\mathcal{Y}}\mathcal{Y}^{\prime}_{0}.

If 𝒳\mathcal{X} is a noetherian algebraic stack over a complete noetherian local ring SpecR\operatorname{Spec}R and x𝒳x\in\mathcal{X} is a closed point, then the conditions of (𝒳,𝒢x)(\mathcal{X},\mathcal{G}_{x}) being (derived) coherently complete and the morphism 𝒳SpecR\mathcal{X}\to\operatorname{Spec}R being formally proper are incomparable. On one hand, coherent completeness concerns the closed substack 𝒢x\mathcal{G}_{x} rather than the central fiber 𝒳RR/𝔪R\mathcal{X}\otimes_{R}R/\mathfrak{m}_{R} while formal properness includes a completeness property with respect to all base changes.

If the functor Coh(𝒳)Coh(𝒳^)\operatorname{Coh}(\mathcal{X})\to\operatorname{Coh}(\widehat{\mathcal{X}}) is fully faithful (e.g. (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) is coherently complete), then for any coherent sheaf \mathcal{F} on 𝒳\mathcal{X}, the natural map

H0(𝒳,)H0(𝒳^,^)\mathrm{H}^{0}(\mathcal{X},\mathcal{F})\to\mathrm{H}^{0}(\widehat{\mathcal{X}},\widehat{\mathcal{F}})

is an isomorphism. Additional hypotheses are needed to imply formal functions, i.e. that the comparison maps on higher cohomology are isomorphisms. On the other hand, if Dcoh(𝒳)Dcoh(𝒳^)D_{\operatorname{coh}}(\mathcal{X})\to D_{\operatorname{coh}}(\widehat{\mathcal{X}}) is fully faithful (i.e. (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies derived formal functions), then

Hi(𝒳,)Hi(𝒳^,^)\mathrm{H}^{i}(\mathcal{X},\mathcal{F})\to\mathrm{H}^{i}(\widehat{\mathcal{X}},\widehat{\mathcal{F}})

is an isomorphism for all ii (i.e. (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions). The proposition below implies in fact that formal functions is equivalent to derived formal functions, and that derived coherent completeness is equivalent to coherent completeness and formal functions.

Let 𝒳\mathcal{X} be a noetherian algebraic stack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} a closed substack. The morphism of ringed sites c:𝒳^𝒳c\colon\widehat{\mathcal{X}}\to\mathcal{X} induces an adjoint pair on unbounded derived categories:

c𝒳,𝒳0:Dqcoh(𝒳)D(𝒳^):c𝒳,𝒳0,qcoh,.c^{*}_{\mathcal{X},\mathcal{X}_{0}}\colon D_{\operatorname{qcoh}}(\mathcal{X})\leftrightarrows D(\widehat{\mathcal{X}})\colon c_{\mathcal{X},\mathcal{X}_{0},{\operatorname{qcoh}},*}.

The functor c𝒳,𝒳0,qcoh,c_{\mathcal{X},\mathcal{X}_{0},\operatorname{qcoh},*} coincides with the composition of the forgetful functor

c𝒳,𝒳0,:D(𝒳^)D(𝒳){c_{\mathcal{X},\mathcal{X}_{0},*}\colon D(\widehat{\mathcal{X}})\to D(\mathcal{X})}

and the quasi-coherator, which is the right adjoint to the inclusion Dqcoh(𝒳)D(𝒳)D_{\operatorname{qcoh}}(\mathcal{X})\to D(\mathcal{X}). See [Hal23, §4] for more details. We have the following simple proposition.

Proposition 3.5.

Let 𝒳\mathcal{X} be a noetherian algebraic stack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} a closed substack.

  1. (1)

    (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions if and only if it satisfies derived formal functions. Moreover, if 𝒢DCoh(𝒳)\mathcal{G}\in D_{\operatorname{Coh}}(\mathcal{X}), then 𝒢cqcoh,c𝒢\mathcal{G}\to c_{\operatorname{qcoh},*}c^{*}\mathcal{G} is an isomorphism.

  2. (2)

    If (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions, then it is coherently complete if and only if it is derived coherently complete. In this case,

    1. (a)

      cqcoh,c_{{\operatorname{qcoh}},*} is tt-exact on Dcoh(𝒳^)D_{\operatorname{coh}}(\widehat{\mathcal{X}}); and

    2. (b)

      if 𝔉Dcoh(𝒳^)\mathfrak{F}\in D_{\operatorname{coh}}(\widehat{\mathcal{X}}), then cqcoh,𝔉Dcoh(𝒳)c_{{\operatorname{qcoh}},*}\mathfrak{F}\in D_{\operatorname{coh}}(\mathcal{X}) and ccqcoh,𝔉𝔉c^{*}c_{{\operatorname{qcoh}},*}\mathfrak{F}\simeq\mathfrak{F}.

Proof.

Assume that (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies derived formal functions. Let Coh(𝒳)\mathcal{F}\in\operatorname{Coh}(\mathcal{X}); then

Hi(𝒳,)=Hom𝒪𝒳(𝒪𝒳,[i])Hom𝒪𝒳(𝒪𝒳^,^[i])=Hi(𝒳^,^).\mathrm{H}^{i}(\mathcal{X},\mathcal{F})=\operatorname{Hom}_{\mathcal{O}_{\mathcal{X}}}(\mathcal{O}_{\mathcal{X}},\mathcal{F}[i])\simeq\operatorname{Hom}_{\mathcal{O}_{\mathcal{X}}}(\mathcal{O}_{\widehat{\mathcal{X}}},\widehat{\mathcal{F}}[i])=\mathrm{H}^{i}(\widehat{\mathcal{X}},\widehat{\mathcal{F}}).

That is, (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions. Conversely, we first note that if (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions and 𝒬Dcoh+(𝒳)\mathcal{Q}\in D^{+}_{\operatorname{coh}}(\mathcal{X}), then 𝖱Γ(𝒳,𝒬)𝖱Γ(𝒳^,c𝒬)\mathsf{R}\Gamma(\mathcal{X},\mathcal{Q})\simeq\mathsf{R}\Gamma(\widehat{\mathcal{X}},c^{*}\mathcal{Q}). Indeed, by the hypercohomology spectral sequence, we have a diagram

Hi(𝒳,j(𝒬)){\mathrm{H}^{i}(\mathcal{X},\mathscr{H}^{j}(\mathcal{Q}))}Hi+j(𝒳,𝒬){\mathrm{H}^{i+j}(\mathcal{X},\mathcal{Q})}Hi(𝒳^,j(c𝒬)){\mathrm{H}^{i}(\widehat{\mathcal{X}},\mathscr{H}^{j}(c^{*}\mathcal{Q}))}Hi+j(𝒳^,c𝒬){\mathrm{H}^{i+j}(\widehat{\mathcal{X}},c^{*}\mathcal{Q})}(a)\scriptstyle{(a)}(b)\scriptstyle{(b)}

where (a) is a morphism of spectral sequences that is an isomorphism by formal functions. Hence (b) is an isomorphism as well. Now let Dcoh(𝒳)\mathcal{M}\in D^{-}_{\operatorname{coh}}(\mathcal{X}) and 𝒩Dcoh+(𝒳)\mathcal{N}\in D^{+}_{\operatorname{coh}}(\mathcal{X}). Then 𝖱om𝒪𝒳(,𝒩)Dcoh+(𝒳)\mathsf{R}\mathscr{H}om_{\mathcal{O}_{\mathcal{X}}}(\mathcal{M},\mathcal{N})\in D^{+}_{\operatorname{coh}}(\mathcal{X}) and so we have

𝖱Hom𝒪𝒳(,𝒩)\displaystyle\mathsf{R}\operatorname{Hom}_{\mathcal{O}_{\mathcal{X}}}(\mathcal{M},\mathcal{N}) 𝖱Γ(𝒳,𝖱om𝒪𝒳(M,N))\displaystyle\simeq\mathsf{R}\Gamma(\mathcal{X},\mathsf{R}\mathscr{H}om_{\mathcal{O}_{\mathcal{X}}}(M,N))
𝖱Γ(𝒳^,c𝖱om𝒪𝒳(,𝒩))\displaystyle\simeq\mathsf{R}\Gamma(\widehat{\mathcal{X}},c^{*}\mathsf{R}\mathscr{H}om_{\mathcal{O}_{\mathcal{X}}}(\mathcal{M},\mathcal{N})) (formal functions)\displaystyle(\mbox{formal functions})
𝖱Γ(𝒳^,𝖱om𝒪𝒳^(c,c𝒩))\displaystyle\simeq\mathsf{R}\Gamma(\widehat{\mathcal{X}},\mathsf{R}\mathscr{H}om_{\mathcal{O}_{\widehat{\mathcal{X}}}}(c^{*}\mathcal{M},c^{*}\mathcal{N})) (c is flat and Dcoh(𝒳))\displaystyle(\mbox{$c$ is flat and $\mathcal{M}\in D^{-}_{\operatorname{coh}}(\mathcal{X})$})
𝖱Hom𝒪𝒳^(c,c𝒩)\displaystyle\simeq\mathsf{R}\operatorname{Hom}_{\mathcal{O}_{\widehat{\mathcal{X}}}}(c^{*}\mathcal{M},c^{*}\mathcal{N})
𝖱Hom𝒪𝒳(,cqcoh,c𝒩).\displaystyle\simeq\mathsf{R}\operatorname{Hom}_{\mathcal{O}_{\mathcal{X}}}(\mathcal{M},c_{{\operatorname{qcoh}},*}c^{*}\mathcal{N}).

Let 𝒢Dcoh(𝒳)\mathcal{G}\in D_{\operatorname{coh}}(\mathcal{X}). To prove that formal functions implies derived formal functions, it suffices to prove that the morphism 𝒢cqcoh,c𝒢\mathcal{G}\to c_{{\operatorname{qcoh}},*}c^{*}\mathcal{G} is an isomorphism in D(𝒳)D(\mathcal{X}). By Lemma 3.7 and the above, 𝒢cqcoh,c𝒢\mathcal{\mathcal{G}}\to c_{{\operatorname{qcoh}},*}c^{*}\mathcal{G} is an isomorphism whenever 𝒢Dcoh+(𝒳)\mathcal{G}\in D^{+}_{\operatorname{coh}}(\mathcal{X}). In general, Lemma 3.6 provides

cqcoh,c𝒢\displaystyle c_{{\operatorname{qcoh}},*}c^{*}\mathcal{G} cqcoh,(holimnτnc𝒢)holimncqcoh,(τnc𝒢)\displaystyle\simeq c_{{\operatorname{qcoh}},*}(\mathrm{holim}_{{n}}\tau^{\geq-n}c^{*}\mathcal{G})\simeq\mathrm{holim}_{{n}}c_{{\operatorname{qcoh}},*}(\tau^{\geq-n}c^{*}\mathcal{G})
holimncqcoh,c(τn𝒢)holimnτn𝒢𝒢.\displaystyle\simeq\mathrm{holim}_{{n}}c_{{\operatorname{qcoh}},*}c^{*}(\tau^{\geq-n}\mathcal{G})\simeq\mathrm{holim}_{{n}}\tau^{\geq-n}\mathcal{G}\simeq\mathcal{G}.

This proves (1). For (2), by (1) it suffices to prove that the functor is essentially surjective. Equivalently, that the natural map ccqcoh,𝔉𝔉c^{*}c_{{\operatorname{qcoh}},*}\mathfrak{F}\to\mathfrak{F} is an isomorphism in Dcoh(𝒳^)D_{\operatorname{coh}}(\widehat{\mathcal{X}}). A standard induction argument on the length of the complex together with the derived formal functions assumption and coherent completeness, shows that for every 𝔉Dcohb(𝒳^)\mathfrak{F}\in D^{b}_{\operatorname{coh}}(\widehat{\mathcal{X}}) there is a unique Dcohb(𝒳)\mathcal{F}\in D^{b}_{\operatorname{coh}}(\mathcal{X}) such that c𝔉c^{*}\mathcal{F}\simeq\mathfrak{F}. If 𝒫Coh(𝒳)\mathcal{P}\in\operatorname{Coh}(\mathcal{X}), it follows from derived formal functions that the map cqcoh,𝔉\mathcal{F}\to c_{{\operatorname{qcoh}},*}\mathfrak{F} induces isomorphisms:

𝖱Hom𝒪𝒳(𝒫,)\displaystyle\mathsf{R}\operatorname{Hom}_{\mathcal{O}_{\mathcal{X}}}(\mathcal{P},\mathcal{F}) 𝖱Hom𝒪𝒳^(c𝒫,c)𝖱Hom𝒪𝒳^(c𝒫,𝔉)\displaystyle\simeq\mathsf{R}\operatorname{Hom}_{\mathcal{O}_{\widehat{\mathcal{X}}}}(c^{*}\mathcal{P},c^{*}\mathcal{F})\simeq\mathsf{R}\operatorname{Hom}_{\mathcal{O}_{\widehat{\mathcal{X}}}}(c^{*}\mathcal{P},\mathfrak{F})
𝖱Hom𝒪𝒳(𝒫,cqcoh,𝔉).\displaystyle\simeq\mathsf{R}\operatorname{Hom}_{\mathcal{O}_{{\mathcal{X}}}}(\mathcal{P},c_{{\operatorname{qcoh}},*}\mathfrak{F}).

It follows from Lemma 3.7 that the map cqcoh,𝔉\mathcal{F}\to c_{{\operatorname{qcoh}},*}\mathfrak{F} is an isomorphism, so ccqcoh,𝔉𝔉c^{*}c_{{\operatorname{qcoh}},*}\mathfrak{F}\to\mathfrak{F} is an isomorphism. Now let 𝔉Dcoh(𝒳^)\mathfrak{F}\in D^{-}_{\operatorname{coh}}(\widehat{\mathcal{X}}). Let n=cqcoh,(τn𝔉)Dcohb(𝒳)\mathcal{F}_{n}=c_{{\operatorname{qcoh}},*}(\tau^{\geq-n}\mathfrak{F})\in D^{b}_{\operatorname{coh}}(\mathcal{X}) and set =holimnn\mathcal{F}=\mathrm{holim}_{{n}}\mathcal{F}_{n}. Then [Sta21, Tag 0D6T] (also see Lemma 3.6) shows that q()q(n)Coh(𝒳)\mathscr{H}^{q}(\mathcal{F})\simeq\mathscr{H}^{q}(\mathcal{F}_{n})\in\operatorname{Coh}(\mathcal{X}) whenever qnq\geq-n. Hence, Dcoh(𝒳)\mathcal{F}\in D^{-}_{\operatorname{coh}}(\mathcal{X}). Lemma 3.6 gives a morphism:

c\displaystyle c^{*}\mathcal{F} =c(holimnn)holimncnholimnτn𝔉𝔉.\displaystyle=c^{*}(\mathrm{holim}_{{n}}\mathcal{F}_{n})\to\mathrm{holim}_{{n}}c^{*}\mathcal{F}_{n}\simeq\mathrm{holim}_{{n}}\tau^{\geq-n}\mathfrak{F}\simeq\mathfrak{F}.

Let q𝐙q\in\mathbf{Z}; then the above map induces isomorphisms:

q(c)cq()cq(q)q(cq)q(τq𝔉)q(𝔉).\mathscr{H}^{q}(c^{*}\mathcal{F})\simeq c^{*}\mathscr{H}^{q}(\mathcal{F})\simeq c^{*}\mathscr{H}^{q}(\mathcal{F}_{-q})\simeq\mathscr{H}^{q}(c^{*}\mathcal{F}_{-q})\simeq\mathscr{H}^{q}(\tau^{\geq q}\mathfrak{F})\simeq\mathscr{H}^{q}(\mathfrak{F}).

Hence, cF𝔉c^{*}F\simeq\mathfrak{F}. In particular, arguing as above using derived formal functions and Lemma 3.7, we see that ccqcoh,𝔉𝔉c^{*}c_{{\operatorname{qcoh}},*}\mathfrak{F}\to\mathfrak{F} is an isomorphism whenever 𝔉Dcoh(𝒳^)\mathfrak{F}\in D^{-}_{\operatorname{coh}}(\widehat{\mathcal{X}}). In general, let 𝔉Dcoh(𝒳^)\mathfrak{F}\in D_{\operatorname{coh}}(\widehat{\mathcal{X}}). Set 𝒢n=cqcoh,(τn𝔉)DCohn(𝒳)\mathcal{G}_{n}=c_{\operatorname{qcoh},*}(\tau^{\leq n}\mathfrak{F})\in D^{\leq n}_{\operatorname{Coh}}(\mathcal{X}). Let 𝒢=hocolimn𝒢n\mathcal{G}=\mathrm{hocolim}_{n}\mathcal{G}_{n}; then q(𝒢)q(𝒢n)Coh(𝒳)\mathscr{H}^{q}(\mathcal{G})\simeq\mathscr{H}^{q}(\mathcal{G}_{n})\in\operatorname{Coh}(\mathcal{X}) whenever qnq\leq n. In particular, 𝒢DCoh(𝒳)\mathcal{G}\in D_{\operatorname{Coh}}(\mathcal{X}). But [Sta21, Tag 0949] implies that

c𝒢=chocolimn𝒢nhocolimnc𝒢nhocolimnτn𝔉𝔉.c^{*}\mathcal{G}=c^{*}\mathrm{hocolim}_{n}\mathcal{G}_{n}\simeq\mathrm{hocolim}_{n}c^{*}\mathcal{G}_{n}\simeq\mathrm{hocolim}_{n}\tau^{\leq n}\mathfrak{F}\simeq\mathfrak{F}.

Arguing as before, the result follows. ∎

The following two lemmas featured in the proof of Proposition 3.5.

Lemma 3.6.

Let 𝒳\mathcal{X} be a noetherian algebraic stack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} a closed substack. If 𝔉Dcoh(𝒳^)\mathfrak{F}\in D_{\operatorname{coh}}(\widehat{\mathcal{X}}), then 𝔉holimnτn𝔉\mathfrak{F}\simeq\mathrm{holim}_{{n}}\tau^{\geq-n}\mathfrak{F}.

Proof.

This follows from [Sta21, Tag 0D6S] and Serre vanishing for coherent sheaves on affine noetherian formal schemes. ∎

Lemma 3.7.

Let 𝒳\mathcal{X} be a noetherian algebraic stack. Let f:𝒢f\colon\mathcal{F}\to\mathcal{G} be a morphism in Dqcoh+(𝒳)D^{+}_{\operatorname{qcoh}}(\mathcal{X}). If 𝖱Hom𝒪𝒳(𝒫,f)\mathsf{R}\operatorname{Hom}_{\mathcal{O}_{\mathcal{X}}}(\mathcal{P},f) is an isomorphism for all 𝒫Coh(𝒳)\mathcal{P}\in\operatorname{Coh}(\mathcal{X}), then ff is an isomorphism.

Proof.

Let 𝒞\mathcal{C} be a cone for ff. If 𝒞0\mathcal{C}\neq 0, choose d𝐙d\in\mathbf{Z} least such that d(𝒞)0\mathscr{H}^{d}(\mathcal{C})\neq 0. Then there is a non-zero coherent sheaf 𝒫\mathcal{P} and an injection 𝒫d(𝒞)\mathcal{P}\subseteq\mathscr{H}^{d}(\mathcal{C}) [LMB00, Prop. 15.4]. Hence, the induced map 𝒫[d]𝒞\mathcal{P}[-d]\to\mathcal{C} is non-zero, which is a contradiction. ∎

As a result, we can conclude that formal functions implies that Coh(𝒳)Coh(𝒳^)\operatorname{Coh}(\mathcal{X})\to\operatorname{Coh}(\widehat{\mathcal{X}}) is fully faithful.

Corollary 3.8.

Let 𝒳\mathcal{X} be a noetherian algebraic stack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} a closed substack. If (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions, then the functor

Coh(𝒳)Coh(𝒳^),^\operatorname{Coh}(\mathcal{X})\to\operatorname{Coh}(\widehat{\mathcal{X}}),\qquad\mathcal{F}\mapsto\widehat{\mathcal{F}}

is fully faithful, with essential image stable under kernels, cokernels and extensions. Moreover, it is an equivalence if for each non-zero 𝔉Coh(𝒳^)\mathfrak{F}\in\operatorname{Coh}(\widehat{\mathcal{X}}) there exist Coh(𝒳)\mathcal{F}\in\operatorname{Coh}(\mathcal{X}) and a non-zero map c𝔉c^{*}\mathcal{F}\to\mathfrak{F}.

Proof.

Since (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions, it satisfies derived formal functions (Proposition 3.5(1)). Hence, Coh(𝒳)Coh(𝒳^)\operatorname{Coh}(\mathcal{X})\to\operatorname{Coh}(\widehat{\mathcal{X}}) is exact and fully faithful. It follows immediately that the image of Coh(𝒳)\operatorname{Coh}(\mathcal{X}) in Coh(𝒳^)\operatorname{Coh}(\widehat{\mathcal{X}}) is stable under kernels and cokernels. For the extensions, we know that derived formal functions implies that if \mathcal{F}, 𝒢Coh(𝒳)\mathcal{G}\in\operatorname{Coh}(\mathcal{X}), then

Ext𝒪𝒳1(,𝒢)=Hom𝒪𝒳(,𝒢[1])Hom𝒪𝒳^(^,𝒢^[1])=Ext𝒪𝒳^1(^,𝒢^).\operatorname{Ext}^{1}_{\mathcal{O}_{\mathcal{X}}}(\mathcal{F},\mathcal{G})=\operatorname{Hom}_{\mathcal{O}_{\mathcal{X}}}(\mathcal{F},\mathcal{G}[1])\simeq\operatorname{Hom}_{\mathcal{O}_{\widehat{\mathcal{X}}}}(\widehat{\mathcal{F}},\widehat{\mathcal{G}}[1])=\operatorname{Ext}^{1}_{\mathcal{O}_{\widehat{\mathcal{X}}}}(\widehat{\mathcal{F}},\widehat{\mathcal{G}}).

Hence, the image is also stable under extensions.

If 𝔓Mod(𝒳^)\mathfrak{P}\in\operatorname{Mod}(\widehat{\mathcal{X}}), let cqcoh,0(𝔓)=0(cqcoh,𝔓)c_{\operatorname{qcoh},*}^{0}(\mathfrak{P})=\mathscr{H}^{0}(c_{\operatorname{qcoh},*}\mathfrak{P}), which is a left-exact functor. For the equivalence, the condition implies that cqcoh,0(𝔓)0c^{0}_{\operatorname{qcoh},*}(\mathfrak{P})\neq 0 whenever 𝔓Coh(𝒳^)\mathfrak{P}\in\operatorname{Coh}(\widehat{\mathcal{X}}). Now fix 𝔉Coh(𝔉^)\mathfrak{F}\in\operatorname{Coh}(\widehat{\mathfrak{F}}). If 𝒢cqcoh,0(𝔉)\mathcal{G}\subseteq c^{0}_{\operatorname{qcoh},*}(\mathfrak{F}) is a coherent subsheaf, let 𝔎=ker(c𝒢𝔉)\mathfrak{K}=\ker(c^{*}\mathcal{G}\to\mathfrak{F}). Then 𝔎Coh(𝒳^)\mathfrak{K}\in\operatorname{Coh}(\widehat{\mathcal{X}}) and there is an exact sequence:

0cqcoh,0(𝔎)cqcoh,0c𝒢cqcoh,0𝔉.0\to c^{0}_{\operatorname{qcoh},*}(\mathfrak{K})\to c^{0}_{\operatorname{qcoh},*}c^{*}\mathcal{G}\to c^{0}_{\operatorname{qcoh},*}\mathfrak{F}.

By Proposition 3.5(1), 𝒢cqcoh,0c𝒢\mathcal{G}\to c_{\operatorname{qcoh},*}^{0}c^{*}\mathcal{G} is an isomorphism. Hence, cqcoh,0(𝔎)=0c^{0}_{\operatorname{qcoh},*}(\mathfrak{K})=0 and so 𝔎=0\mathfrak{K}=0. Now write cqcoh,0𝔉=λλc_{\operatorname{qcoh},*}^{0}\mathfrak{F}=\cup_{\lambda}\mathcal{F}_{\lambda} as a filtered union of coherent subsheaves. Then {cλ}λ\{c^{*}\mathcal{F}_{\lambda}\}_{\lambda} defines an increasing union of coherent subsheaves of 𝔉\mathfrak{F}. By coherence of 𝔉\mathfrak{F}, there must be a λ0\lambda_{0} such that cλ=cλ0c^{*}\mathcal{F}_{\lambda}=c^{*}\mathcal{F}_{\lambda_{0}} for all λλ0\lambda\geq\lambda_{0}. By full faithfulness, λ=λ0\mathcal{F}_{\lambda}=\mathcal{F}_{\lambda_{0}} for all λλ0\lambda\geq\lambda_{0} and so cqcoh,0𝔉c_{\operatorname{qcoh},*}^{0}\mathfrak{F} is coherent. Finally, consider the exact sequence:

0ccqcoh,0𝔉𝔉0.0\to c^{*}c_{\operatorname{qcoh},*}^{0}\mathfrak{F}\to\mathfrak{F}\to\mathfrak{C}\to 0.

Applying cqcoh,0c_{\operatorname{qcoh},*}^{0}, we see that from Proposition 3.5(1) again that

0cqcoh,0𝔉idcqcoh,0𝔉cqcoh,000\to c_{\operatorname{qcoh},*}^{0}\mathfrak{F}\xrightarrow{\mathrm{id}}c_{\operatorname{qcoh},*}^{0}\mathfrak{F}\to c_{\operatorname{qcoh},*}^{0}\mathfrak{C}\to 0

is exact. It follows that cqcoh,0=0c_{\operatorname{qcoh},*}^{0}\mathfrak{C}=0 and so =0\mathfrak{C}=0, which proves the desired equivalence. ∎

4. Strong cohomological properness and formal functions

The main result of this section establishes a relationship between formal functions and a strengthening of cohomological properness. While something similar to Corollary 4.8 appeared in [HLP23], Corollary 4.9 is new in positive and mixed characteristic (it was established in the linearly reductive case in [AHR19, Thm. 1.6]). The arguments we use are simple generalizations of the related results of [GD67, III] and [Gro68, IX]. To this end, we make the following definition.

Definition 4.1.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a morphism of noetherian algebraic stacks, 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} a closed substack, and 𝒳0f1(𝒴0)\mathcal{X}_{0}\subset f^{-1}(\mathcal{Y}_{0}) a closed substack. We say that the triple (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) is strongly cohomologically proper if for sheaves of ideals 𝒥\mathcal{J} and \mathcal{I} defining 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X}, the morphism Spec¯𝒪𝒳(n0n)Spec¯𝒪𝒴(n0𝒥n)\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{I}^{n})\to\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}\mathcal{J}^{n}) is cohomologically proper.

Remark 4.2.

In Definition 4.1, from the commutative diagram:

𝒳\textstyle{\mathcal{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}Spec¯𝒪𝒳(n0n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{I}^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒴\textstyle{\mathcal{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒴(n0𝒥n),\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}\mathcal{J}^{n}),}

where the horizontal morphisms are closed immersions, it follows that ff is cohomologically proper. It is also easy to see that the definition only depends on the closed subsets |𝒳0||𝒳||\mathcal{X}_{0}|\subseteq|\mathcal{X}| and |𝒴0||𝒴||\mathcal{Y}_{0}|\subseteq|\mathcal{Y}| (not the ideals \mathcal{I} and 𝒥\mathcal{J}). Indeed, suppose that we have other ideals \mathcal{I}^{\prime}, 𝒥\mathcal{J}^{\prime} with vanishing locus |𝒳0||\mathcal{X}_{0}| and |𝒴0||\mathcal{Y}_{0}|. Now form the commutative diagram:

Spec¯𝒪𝒳(n0n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{I}^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒳(n0()n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}(\mathcal{I}\cap\mathcal{I}^{\prime})^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒳(n0n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{I}^{\prime n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒴(n0𝒥n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}\mathcal{J}^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒴(n0(𝒥𝒥)n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}(\mathcal{J}\cap\mathcal{J}^{\prime})^{n})}Spec¯𝒪𝒴(n0𝒥n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}\mathcal{J}^{\prime n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

Since 𝒳\mathcal{X} and 𝒴\mathcal{Y} are noetherian, it is easy to see that the horizontal morphisms are all finite and surjective. In particular, the cohomological properness of any of the vertical morphisms implies that of the others.

Remark 4.3.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a morphism of noetherian algebraic stacks with affine diagonal, 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} a closed substack, and 𝒳0f1(𝒴0)\mathcal{X}_{0}\subset f^{-1}(\mathcal{Y}_{0}) a closed substack. Assume that 𝒪𝒴f𝒪𝒳\mathcal{O}_{\mathcal{Y}}\to f_{*}\mathcal{O}_{\mathcal{X}} is an isomorphism (i.e., ff is Stein) and that 𝒥f\mathcal{J}\simeq f_{*}\mathcal{I}, where 𝒥\mathcal{J} defines 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} and \mathcal{I} defines 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X}. Then the triple (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) is strongly cohomologically proper if and only if the 𝒪𝒴\mathcal{O}_{\mathcal{Y}}-algebra n0f(n)\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n}) is finitely generated and the induced morphism Spec¯𝒪𝒳(n0n)Spec¯𝒪𝒴(n0f(n))\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{I}^{n})\to\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n})) is cohomologically proper. Indeed, we have the composition

Spec¯𝒪𝒳(n0n)Spec¯𝒪𝒴(n0f(n))Spec¯𝒪𝒴(n0𝒥n)𝒴.\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{I}^{n})\to\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n}))\to\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}\mathcal{J}^{n})\to\mathcal{Y}.

If (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) is strongly cohomologically proper, then n0f(n)\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n}) is a finite n0𝒥n\oplus_{n\geq 0}\mathcal{J}^{n}-algebra and so the induced morphism Spec¯𝒪𝒳(n0n)Spec¯𝒪𝒴(n0f(n))\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{I}^{n})\to\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n})) is cohomologically proper. Conversely, if n0f(n)\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n}) is a finitely generated 𝒪𝒴\mathcal{O}_{\mathcal{Y}}-algebra, then there is an N0N\gg 0 such that f(N)k=f(Nk)f_{*}(\mathcal{I}^{N})^{k}=f_{*}(\mathcal{I}^{Nk}) as ideals of f𝒪𝒳𝒪𝒴f_{*}\mathcal{O}_{\mathcal{X}}\simeq\mathcal{O}_{\mathcal{Y}} for all k0k\geq 0. But we have 𝒥Nf(N)f()=𝒥\mathcal{J}^{N}\subseteq f_{*}(\mathcal{I}^{N})\subseteq f_{*}(\mathcal{I})=\mathcal{J}. In particular, we have a commutative diagram:

Spec¯𝒪𝒳(n0n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{I}^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒳(k0Nk)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{k\geq 0}\mathcal{I}^{Nk})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒴(n0f(n))\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒴(k0f(N)k)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{k\geq 0}f_{*}(\mathcal{I}^{N})^{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒴(n0𝒥n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}\mathcal{J}^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Spec¯𝒪𝒴(k0𝒥Nk),\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{k\geq 0}\mathcal{J}^{Nk}),}

where the horizontal arrows are all finite and surjective. The top vertical arrow on the right is cohomologically proper and the bottom vertical arrow on the right is finite, whence the composition is cohomologically proper. It follows that the composition of the vertical arrows on the left is cohomologically proper, which proves the claim.

We have the following two key examples.

Example 4.4.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a universally cohomologically proper morphism of noetherian algebraic stacks. Let 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} be a closed substack. Then the triple (f,f1(𝒴0),𝒴0)(f,f^{-1}(\mathcal{Y}_{0}),\mathcal{Y}_{0}) is strongly cohomologically proper. To see this: let 𝒥𝒪𝒴\mathcal{J}\subset\mathcal{O}_{\mathcal{Y}} be a coherent sheaf of ideals definining 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y}. Since 𝒴\mathcal{Y} is noetherian, n0𝒥n\oplus_{n\geq 0}\mathcal{J}^{n} is a finitely generated 𝒪𝒴\mathcal{O}_{\mathcal{Y}}-algebra. Now form the cartesian square:

Spec¯𝒪𝒳(n0𝒥n𝒪𝒳)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{J}^{n}\mathcal{O}_{\mathcal{X}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f~\scriptstyle{\tilde{f}}𝒳\textstyle{\mathcal{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}Spec¯𝒪𝒴(n0𝒥n)\textstyle{\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}\mathcal{J}^{n})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒴.\textstyle{\mathcal{Y}.}

Since ff is universally cohomologically proper, f~\tilde{f} is cohomologically proper.

Example 4.5.

Let RR be a noetherian ring. Let AA be a finitely generated RR-algebra with an action of a reductive group scheme GG over SpecR\operatorname{Spec}R. Let IAI\subseteq A be a GG-equivariant ideal. Let 𝒳=[SpecA/G]\mathcal{X}=[\operatorname{Spec}A/G], 𝒳0=[Spec(A/I)/G]\mathcal{X}_{0}=[\operatorname{Spec}(A/I)/G], 𝒴0=Spec(AG/IG)\mathcal{Y}_{0}=\operatorname{Spec}(A^{G}/I^{G}) and f:𝒳Spec(AG)f\colon\mathcal{X}\to\operatorname{Spec}(A^{G}). Then the triple (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) is strongly cohomologically proper. Indeed, by Example 2.7, ff is universally cohomologically proper. Further, the GG-equivariant AA-algebra n0In\oplus_{n\geq 0}I^{n} is finitely generated and so (n0In)G=n0H0([SpecA/G],I~n)(\oplus_{n\geq 0}I^{n})^{G}=\oplus_{n\geq 0}\mathrm{H}^{0}([\operatorname{Spec}A/G],\tilde{I}^{n}) is a finitely generated AGA^{G}-algebra, where I~\tilde{I} denotes the coherent sheaf of ideals on [SpecA/G][\operatorname{Spec}A/G] associated to II. By Example 2.7 again, the induced morphism [Spec(n0In)/G]Spec((n0In)G)[\operatorname{Spec}(\oplus_{n\geq 0}I^{n})/G]\to\operatorname{Spec}((\oplus_{n\geq 0}I^{n})^{G}) is cohomologically proper and the claim follows from Remark 4.3. If AGA^{G} is artinian, this also follows from Theorem 1.2.

We now have the main result of this section.

Theorem 4.6.

Let f:𝒳SpecRf\colon\mathcal{X}\to\operatorname{Spec}R be a finite type morphism of noetherian algebraic stacks, JRJ\subset R an ideal, and 𝒳0f1(Spec(R/J))\mathcal{X}_{0}\subset f^{-1}(\operatorname{Spec}(R/J)) a closed substack. Assume that the triple (f,𝒳0,Spec(R/J))(f,\mathcal{X}_{0},\operatorname{Spec}(R/J)) is strongly cohomologically proper and RR is JJ-adically complete.

  1. (1)

    Let \mathcal{I} be a coherent 𝒪𝒳\mathcal{O}_{\mathcal{X}}-ideal defining the closed immersion 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X}. Then Γ(𝒳,𝒪𝒳)\Gamma(\mathcal{X},\mathcal{O}_{\mathcal{X}}) is Γ(𝒳,)\Gamma(\mathcal{X},\mathcal{I})-adically complete.

  2. (2)

    Formal functions holds for the pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}). Moreover, if \mathcal{F} is a coherent 𝒪𝒳\mathcal{O}_{\mathcal{X}}-module, then the natural morphisms:

    Hq(𝒳,)Hq(𝒳^,^)limnHq(𝒳n,n)\mathrm{H}^{q}(\mathcal{X},\mathcal{F})\to\mathrm{H}^{q}(\widehat{\mathcal{X}},\widehat{\mathcal{F}})\to\varprojlim_{n}\mathrm{H}^{q}(\mathcal{X}_{n},\mathcal{F}_{n})

    are isomorphisms for all q0q\geq 0.

  3. (3)

    If 𝔉\mathfrak{F} is a coherent 𝒪𝒳^\mathcal{O}_{\widehat{\mathcal{X}}}-module, then Hq(𝒳^,𝔉)\mathrm{H}^{q}(\widehat{\mathcal{X}},\mathfrak{F}) is a finitely generated RR-module and the natural morphism:

    Hq(𝒳^,𝔉)limnHq(𝒳n,𝔉n)\mathrm{H}^{q}(\widehat{\mathcal{X}},\mathfrak{F})\to\varprojlim_{n}\mathrm{H}^{q}(\mathcal{X}_{n},\mathfrak{F}_{n})

    is an isomorphism for all q0q\geq 0.

Before we prove Theorem 4.6, we provide several corollaries.

Corollary 4.7.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a universally cohomologically proper morphism of noetherian algebraic stacks. Then ff admits a Stein factorization:

𝒳f¯Spec¯𝒪𝒴(f𝒪𝒳)𝑎𝒴,\mathcal{X}\xrightarrow{\bar{f}}\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(f_{*}\mathcal{O}_{\mathcal{X}})\xrightarrow{a}\mathcal{Y},

where aa is finite and f¯\bar{f} is universally cohomologically proper and universally closed with geometrically connected fibers.

Proof.

By Remark 2.5, it suffices to prove that if 𝒴=SpecR\mathcal{Y}=\operatorname{Spec}R, where (R,𝔪)(R,\mathfrak{m}) is an 𝔪\mathfrak{m}-adically complete noetherian local ring and RH0(𝒳,𝒪𝒳)R\simeq\mathrm{H}^{0}(\mathcal{X},\mathcal{O}_{\mathcal{X}}); then the closed fiber 𝒳0=𝒳RR/𝔪Spec(R/𝔪)\mathcal{X}_{0}=\mathcal{X}\otimes_{R}R/\mathfrak{m}\to\operatorname{Spec}(R/\mathfrak{m}) is connected. By Example 4.4, the triple (𝒳,𝒳0,f)(\mathcal{X},\mathcal{X}_{0},f) is strongly cohomologically proper. It follows from Theorem 4.6(2) that RH0(𝒳,𝒪𝒳)limnH0(𝒳n,𝒪𝒳n)R\simeq\mathrm{H}^{0}(\mathcal{X},\mathcal{O}_{\mathcal{X}})\simeq\varprojlim_{n}\mathrm{H}^{0}(\mathcal{X}_{n},\mathcal{O}_{\mathcal{X}_{n}}). If 𝒳0\mathcal{X}_{0} is disconnected, then the corresponding idempotent e0e_{0} lifts to a non-trivial idempotent ee of RR, which contradicts RR being local [Sta21, Tag 0G7X]. ∎

Corollary 4.8.

Let π:𝒳SpecR\pi\colon\mathcal{X}\to\operatorname{Spec}R be a universally cohomologically proper morphism of noetherian algebraic stacks. Let JRJ\subseteq R be an ideal. If RR is JJ-adically complete, then formal functions holds for the pair (𝒳,π1(SpecR/J))(\mathcal{X},\pi^{-1}(\operatorname{Spec}R/J)).

Proof.

Combine Example 4.4 with Theorem 4.6. ∎

Corollary 4.9.

Let RR be a noetherian ring. Let AA be a finitely generated RR-algebra with an action of a reductive group scheme GG over SpecR\operatorname{Spec}R. Let IAI\subseteq A be a GG-equivariant ideal. Let 𝒳=[SpecA/G]\mathcal{X}=[\operatorname{Spec}A/G], 𝒳0=[Spec(A/I)/G]\mathcal{X}_{0}=[\operatorname{Spec}(A/I)/G], 𝒴0=Spec(AG/IG)\mathcal{Y}_{0}=\operatorname{Spec}(A^{G}/I^{G}), 𝒴=Spec(AG)\mathcal{Y}=\operatorname{Spec}(A^{G}) and f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} the induced morphism. If AGA^{G} is IGI^{G}-adically complete, then the triple (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) is strongly cohomologically proper and the pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions.

Proof.

Combine Example 4.5 and Theorem 4.6. ∎

We recall some background on filtrations that will be important for the proof of Theorem 4.6. Let AA be a ring, IAI\subset A an ideal, and MM an AA-module. A filtration (Mn)n𝐙(M_{n})_{n\in\mathbf{Z}} of MM is II-good (or II-stable) if the following three conditions are satisfied:

  1. (1)

    M=MkM=M_{k} for some k𝐙k\in\mathbf{Z};

  2. (2)

    IMnMn+1IM_{n}\subset M_{n+1} for all n𝐙n\in\mathbf{Z}; and

  3. (3)

    IMn=Mn+1IM_{n}=M_{n+1} for all n0n\gg 0.

Obviously, the filtration (In+1M)n1(I^{n+1}M)_{n\geq-1} is II-good; the topology that this filtration defines on MM is called the II-adic topology on MM. A key observation is that the topology on MM defined by any II-good filtration is equivalent to the II-adic topology on MM [AM69, Lem. 10.6]. A much deeper fact is that if AA is noetherian and MM is finitely generated, then a filtration (Mn)(M_{n}) is II-stable if and only if M=n𝐙MnM_{\ast}=\oplus_{n\in\mathbf{Z}}M_{n} is a finitely generated A=n0InA_{\ast}=\oplus_{n\geq 0}I^{n}-module [AM69, Lem. 10.8]. A key consequence of this whole theory is that if MM is a finitely generated AA-module and AA is II-adically complete, then MM is II-adically complete.

Assume now that we are in the situation of Theorem 4.6. Let \mathcal{M} be a coherent 𝒪𝒳\mathcal{O}_{\mathcal{X}}-module. Let =n0n\mathcal{I}_{\ast}=\oplus_{n\geq 0}\mathcal{I}^{n} and let =n0n\mathcal{M}_{\ast}=\oplus_{n\geq 0}\mathcal{I}^{n}\mathcal{M}. The quasi-coherent 𝒪𝒳\mathcal{O}_{\mathcal{X}}-algebra \mathcal{I}_{\ast} is of finite type and \mathcal{M}_{\ast} is a coherent \mathcal{I}^{\ast}-module [AM69, Lem. 10.8].

Let I=Γ(𝒳,)I_{\ast}=\Gamma(\mathcal{X},\mathcal{I}_{\ast}) and Mq=Hq(𝒳,)M^{q}_{\ast}=\mathrm{H}^{q}(\mathcal{X},\mathcal{M}_{\ast}). Let us briefly remark on the graded structure of MqM^{q}_{\ast}. If xIs=Γ(𝒳,s)x\in I_{s}=\Gamma(\mathcal{X},\mathcal{I}^{s}), then for all t0t\geq 0 there is an induced homomorphism of 𝒪𝒳\mathcal{O}_{\mathcal{X}}-modules that ts+t\mathcal{I}^{t}\mathcal{M}\to\mathcal{I}^{s+t}\mathcal{M} that is multiplication by xx. It follows that we obtain an induced morphism:

μx,,tq:Hq(𝒳,t)Hq(𝒳,t+s).\mu_{x,\mathcal{M},t}^{q}\colon\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t}\mathcal{M})\to\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t+s}\mathcal{M}).

This is how MqM^{q}_{\ast} becomes a graded II_{\ast}-module. In particular, IsMtqMt+sqI_{s}M^{q}_{t}\subset M^{q}_{t+s} denotes the image of the natural R=I0R=I_{0}-module homomorphism IsRMtqMt+sqI_{s}\otimes_{R}M^{q}_{t}\to M^{q}_{t+s}.

Further, the canonical inclusions tt\mathcal{I}^{t}\mathcal{M}\subset\mathcal{I}^{t^{\prime}}\mathcal{M} for ttt\geq t^{\prime} give rise to an inverse system (Hq(𝒳,t))t0(\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t}\mathcal{M}))_{t\geq 0} with transition map ν,t,tq:Hq(𝒳,t)Hq(𝒳,t)\nu_{\mathcal{M},t,t^{\prime}}^{q}\colon\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t}\mathcal{M})\to\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t^{\prime}}\mathcal{M}) when ttt\geq t^{\prime}. It follows that the composition:

Hq(𝒳,t)μx,,tqHq(𝒳,t+s)ν,t+s,tqHq(𝒳,t)\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t}\mathcal{M})\xrightarrow{\mu_{x,\mathcal{M},t}^{q}}\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t+s}\mathcal{M})\xrightarrow{\nu_{\mathcal{M},t+s,t}^{q}}\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t}\mathcal{M})

coincides with multiplication by xx on Hq(𝒳,t)\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t}\mathcal{M}) as an RR-module. In particular, if PHq(𝒳,t)P\subset\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t}\mathcal{M}) is an RR-submodule, then we have the equality of RR-submodules of Hq(𝒳,t)\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{t}\mathcal{M}):

(4.1) Γ(𝒳,s)P=ν,t+s,tq(IsP).\Gamma(\mathcal{X},\mathcal{I}^{s})P=\nu^{q}_{\mathcal{M},t+s,t}(I_{s}P).

By assumption, II_{\ast} is a finitely generated RR-algebra and MqM^{q}_{\ast} is a finitely generated and graded II_{\ast}-module. It follows that for some sufficiently large NN, that as ideals of the finite RR-algebra Γ(𝒳,𝒪𝒳)\Gamma(\mathcal{X},\mathcal{O}_{\mathcal{X}}), we have Γ(𝒳,N)k=Γ(𝒳,Nk)\Gamma(\mathcal{X},\mathcal{I}^{N})^{k}=\Gamma(\mathcal{X},\mathcal{I}^{Nk}) for all k0k\geq 0. In particular, replacing II and \mathcal{I} by INI^{N} and N\mathcal{I}^{N}, respectively, we may assume that Γ(𝒳,)k=Γ(𝒳,k)\Gamma(\mathcal{X},\mathcal{I})^{k}=\Gamma(\mathcal{X},\mathcal{I}^{k}) for all k0k\geq 0.

Proof of Theorem 4.6.

We first prove (1). The discussion above showed that n0In\oplus_{n\geq 0}I^{n} is a finite n0Jn\oplus_{n\geq 0}J^{n}-module. It follows that the II-adic filtration on SS is JJ-stable and so the JJ-adic topology is equivalent to the II-adic topology. Since RR is JJ-adically complete and SS is a finite RR-module, SS is II-adically complete too.

The proof of (2) follows Serre’s argument in [GD67, III.4.1.5] (cf. [FGI+05, §8.2] and [AHR19, §4]). By (1), we may assume that R=Γ(𝒳,𝒪𝒳)R=\Gamma(\mathcal{X},\mathcal{O}_{\mathcal{X}}) and Γ(𝒳,k)=Γ(𝒳,)k\Gamma(\mathcal{X},\mathcal{I}^{k})=\Gamma(\mathcal{X},\mathcal{I})^{k} for all k0k\geq 0 and so I=k0Ik=RI_{\ast}=\oplus_{k\geq 0}I^{k}=R_{\ast}. By Remark 3.3, it suffices to prove that Hq(𝒳,)limnHq(𝒳n,n)\mathrm{H}^{q}(\mathcal{X},\mathcal{F})\to\varprojlim_{n}\mathrm{H}^{q}(\mathcal{X}_{n},\mathcal{F}_{n}) is an isomorphism for all coherent sheaves \mathcal{F} on 𝒳\mathcal{X} and q0q\geq 0 and {Hq1(𝒳n,n)}n\{H^{q-1}(\mathcal{X}_{n},\mathcal{F}_{n})\}_{n} satisfies the Artin–Rees condition (this implies that the lim1\lim^{1} term in (3.1) vanishes).

We now let \mathcal{F} be a coherent 𝒪𝒳\mathcal{O}_{\mathcal{X}}-module. Let q0q\geq 0 and n1n\geq-1 and consider the exact sequence of RR-modules:

(4.2) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Rnq\textstyle{R^{q}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hq\textstyle{H^{q}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hnq\textstyle{H^{q}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Qnq\textstyle{Q^{q}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0,\textstyle{0,}

where Hq=Hq(𝒳,)H^{q}=\mathrm{H}^{q}(\mathcal{X},\mathcal{F}), Hnq=Hq(𝒳,/n+1F)H^{q}_{n}=\mathrm{H}^{q}(\mathcal{X},\mathcal{F}/\mathcal{I}^{n+1}F), Lnq=Hq(𝒳,n+1)L^{q}_{n}=\mathrm{H}^{q}(\mathcal{X},\mathcal{I}^{n+1}\mathcal{F}),

Rnq\displaystyle R^{q}_{n} =ker(HqHnq)=im(LnqHq),and\displaystyle=\ker(H^{q}\to H^{q}_{n})=\mathrm{im}(L^{q}_{n}\to H^{q}),\,\mbox{and}
Qnq\displaystyle Q^{q}_{n} =im(HnqLnq+1)=ker(Lnq+1Hq+1).\displaystyle=\mathrm{im}(H^{q}_{n}\to{L}^{q+1}_{n})=\ker({L}^{q+1}_{n}\to H^{q+1}).

The result follows from the following three claims:

  1. (1)

    the filtration (Rnq)n1(R^{q}_{n})_{n\geq-1} on HqH^{q} is II-good;

  2. (2)

    the inverse system (Qnq)(Q^{q}_{n}) is Artin–Rees zero (i.e., there exists an ss such that Qn+sqQnqQ_{n+s}^{q}\to Q^{q}_{n} is 0 for all nn);

  3. (3)

    the inverse system (Hnq1)(H^{q-1}_{n}) satisfies the uniform Artin–Rees condition (i.e., there is an ss such that the images of the morphisms Hnq1Hnq1H^{q-1}_{n^{\prime}}\to H^{q-1}_{n} agree for all nn+sn^{\prime}\geq n+s).

Indeed, the exact sequence (4.2) induces the following short exact sequence:

(4.3) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hq/Rnq\textstyle{H^{q}/R^{q}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hnq\textstyle{H^{q}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Qnq\textstyle{Q^{q}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

We now take inverse limits, and obtain the following exact sequence:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}limnHq/Rnq\textstyle{\varprojlim_{n}H^{q}/R^{q}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}limnHnq\textstyle{\varprojlim_{n}H^{q}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}limnQnq.\textstyle{\varprojlim_{n}Q^{q}_{n}.}

Since the system (Qnq)(Q^{q}_{n}) is Artin–Rees zero, it follows immediately that limnQnq=0\varprojlim_{n}Q^{q}_{n}=0. Moreover, the filtration (Rnq)(R^{q}_{n}) on HqH^{q} is II-good and since HqH^{q} is a finitely generated RR-module, it follows that the natural map HqlimnHq/RnqH^{q}\to\varprojlim_{n}H^{q}/R^{q}_{n} is an isomorphism. What results from all of this is an isomorphism:

Hq(𝒳,)limnHq(𝒳,/n+1).\mathrm{H}^{q}(\mathcal{X},\mathcal{F})\simeq\varprojlim_{n}\mathrm{H}^{q}(\mathcal{X},\mathcal{F}/\mathcal{I}^{n+1}\mathcal{F}).

That is, Zariski formal functions holds for the pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}).

We first establish that the filtration (Rnq)n1(R^{q}_{n})_{n\geq-1} on HqH^{q} is II-good. To see this, we first note that R1=HqR_{-1}=H^{q}. We now apply the previous discussion to the 𝒪X\mathcal{O}_{X}-module \mathcal{I}\mathcal{F}. It follows that Lq=n0Lnq{L}^{q}_{\ast}=\oplus_{n\geq 0}{L}^{q}_{n} is a finitely generated RR_{\ast}-module. But the graded RR_{\ast}-module n0Rnq\oplus_{n\geq 0}R^{q}_{n} is the image of the graded RR_{\ast}-module homomorphism

n0Lnqn0Hq,\oplus_{n\geq 0}{L}^{q}_{n}\to\oplus_{n\geq 0}H^{q},

and so n1Rnq\oplus_{n\geq-1}R^{q}_{n} is also a finitely generated graded II_{\ast}-module. By [AM69, Lem. 10.8], it follows that the filtration (Rnq)n1(R^{q}_{n})_{n\geq-1} is II-good.

We next prove that the inverse system (Qnq)(Q^{q}_{n}) is Artin–Rees zero. First observe that Qq=n0QnqQ_{\ast}^{q}=\oplus_{n\geq 0}Q_{n}^{q} is a RR_{\ast}-submodule of Lq+1=n0Lnq+1L_{\ast}^{q+1}=\oplus_{n\geq 0}L_{n}^{q+1}, which is a finitely generated RR_{\ast}-module. Hence, QqQ_{\ast}^{q} is a finitely generated RR_{\ast}-module. In particular, there exist integers hh, l0l\geq 0 such that IhQkq=Qh+kqI_{h}Q_{k}^{q}=Q_{h+k}^{q} for all klk\geq l. But QnqQ_{n}^{q} is always a quotient of HnqH^{q}_{n} and HnqH^{q}_{n} is annihilated by In+1I^{n+1} and so if ml+hm\geq l+h, then write m=th+r+lm=th+r+l, where 0r<h0\leq r<h. Then

Il+h+1Qmq=Il+h+1IthQl+rIl+r+1IthQl+r=0.I^{l+h+1}Q_{m}^{q}=I^{l+h+1}I_{th}Q_{l+r}\subset I^{l+r+1}I_{th}Q_{l+r}=0.

It follows from (4.1) and the above that if s=(h+2)(l+h)l+h+1s=(h+2)(l+h)\geq l+h+1, then for t0t\geq 0 we have

ν,t+s,tq+1(Qt+sq)=ν,t+s,tq+1(I(h+1)(l+h)Qt+h+lq)=I(h+1)(l+h)Qt+l+hqIl+h+1Qt+l+hq=0.\nu_{\mathcal{I}\mathcal{F},t+s,t}^{q+1}(Q^{q}_{t+s})=\nu_{\mathcal{I}\mathcal{F},t+s,t}^{q+1}(I_{(h+1)(l+h)}Q^{q}_{t+h+l})=I^{(h+1)(l+h)}Q^{q}_{t+l+h}\subset I^{l+h+1}Q^{q}_{t+l+h}=0.

Finally, the exact sequence of (4.3) and basic properties of the Artin–Rees condition shows that it suffices to prove that the inverse systems (Hq/Rn)(H^{q}/R_{n}) and (Qn)(Q_{n}) satisfy the uniform Artin–Rees condition. Since (Qn)(Q_{n}) is Artin–Rees zero, it satisfies the uniform Artin–Rees condition. Further, since (Hq/Rn)(H^{q}/R_{n}) is a surjective system it trivially satisfies the uniform Artin–Rees condition. This proves (2)

We now prove (3). Note that n0n/n+1\oplus_{n\geq 0}\mathcal{I}^{n}/\mathcal{I}^{n+1} is a coherent n0n\oplus_{n\geq 0}\mathcal{I}^{n}-algebra. In particular, the cohomological properness of Spec¯𝒪𝒳(n0n)Spec¯𝒪𝒴(n0f(n))\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}}}(\oplus_{n\geq 0}\mathcal{I}^{n})\to\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{Y}}}(\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n})) implies that n0f(n/n+1)\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n}/\mathcal{I}^{n+1}) is a coherent n0f(n)\oplus_{n\geq 0}f_{*}(\mathcal{I}^{n})-algebra. It follows that:

  1. (1)

    n0H0(𝒳0,n/n+1)\oplus_{n\geq 0}\mathrm{H}^{0}(\mathcal{X}_{0},\mathcal{I}^{n}/\mathcal{I}^{n+1}) is a finitely generated Γ(𝒳,𝒪𝒳)\Gamma(\mathcal{X},\mathcal{O}_{\mathcal{X}})-algebra; and

  2. (2)

    Spec¯𝒪𝒳0(n0n/n+1)Spec(n0H0(𝒳0,n/n+1))\underline{\operatorname{Spec}}_{\mathcal{O}_{\mathcal{X}_{0}}}(\oplus_{n\geq 0}\mathcal{I}^{n}/\mathcal{I}^{n+1})\to\operatorname{Spec}(\oplus_{n\geq 0}\mathrm{H}^{0}(\mathcal{X}_{0},\mathcal{I}^{n}/\mathcal{I}^{n+1})) is cohomologically proper.

The arguments of [GD67, III.3.4.4] now apply verbatim to prove (3). ∎

5. Permanence of properties

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a morphism of noetherian algebraic stacks, 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} a closed substack, and 𝒳0f1(𝒴0)\mathcal{X}_{0}\subset f^{-1}(\mathcal{Y}_{0}) a closed substack. Then there is an induced diagram:

Mod(𝒳)\textstyle{\operatorname{Mod}(\mathcal{X})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c𝒳\scriptstyle{c^{*}_{\mathcal{X}}}f\scriptstyle{f_{*}}Mod(𝒳^)\textstyle{\operatorname{Mod}(\widehat{\mathcal{X}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c𝒳,\scriptstyle{c_{\mathcal{X},*}}f^\scriptstyle{\widehat{f}_{*}}Mod(𝒴)\textstyle{\operatorname{Mod}(\mathcal{Y})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c𝒴\scriptstyle{c^{*}_{\mathcal{Y}}}f\scriptstyle{f^{*}}Mod(𝒴^),\textstyle{\operatorname{Mod}(\widehat{\mathcal{Y}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces,}f^\scriptstyle{\widehat{f}^{*}}c𝒴,\scriptstyle{c_{\mathcal{Y},*}}

as well as natural isomorphisms of functors c𝒳ff^c𝒴c_{\mathcal{X}}^{*}f^{*}\simeq\widehat{f}^{*}c_{\mathcal{Y}}^{*} and fc𝒳,c𝒴,f^f_{*}c_{\mathcal{X},*}\simeq c_{\mathcal{Y},*}\widehat{f}_{*}. Because of the lack of functoriality of the lisse-étale topos, the left derived functors of ff^{*} and f^\widehat{f}^{*} are somewhat subtle if ff is not smooth. The derived functors on the level of unbounded derived categories 𝖱f:D(𝒳)D(𝒴)\mathsf{R}f_{*}\colon D(\mathcal{X})\to D(\mathcal{Y}) and 𝖱f^:D(𝒳^)D(𝒴^)\mathsf{R}\widehat{f}_{*}\colon D(\widehat{\mathcal{X}})\to D(\widehat{\mathcal{Y}}) always exist, however. As c𝒳,c_{\mathcal{X},*} and c𝒴,c_{\mathcal{Y},*} are exact (they are just forgetful functors), it follows that we also have natural isomorphisms (𝖱f)c𝒳,c𝒴,𝖱f^(\mathsf{R}f_{*})\circ c_{\mathcal{X},*}\simeq c_{\mathcal{Y},*}\circ\mathsf{R}\widehat{f}_{*}. If D(𝒳)\mathcal{F}\in D(\mathcal{X}), then there is thus always a comparison morphism:

(5.1) (𝖱f)=c𝒴(𝖱f)𝖱f^c𝒳=𝖱f^^,(\mathsf{R}f_{*}\mathcal{F})^{\wedge}=c_{\mathcal{Y}}^{*}(\mathsf{R}f_{*}\mathcal{F})\to\mathsf{R}\widehat{f}_{*}c_{\mathcal{X}}^{*}\mathcal{F}=\mathsf{R}\widehat{f}_{*}\widehat{\mathcal{F}},

which comes from the adjoint of the morphism:

𝖱f(𝖱f)c𝒳,c𝒳c𝒴,𝖱f^c𝒳.\mathsf{R}f_{*}\mathcal{F}\to(\mathsf{R}f_{*})\circ c_{\mathcal{X},*}\circ c_{\mathcal{X}}^{*}\mathcal{F}\simeq c_{\mathcal{Y},*}\circ\mathsf{R}\widehat{f}_{*}\circ c_{\mathcal{X}}^{*}\mathcal{F}.

We now introduce a relative version of Definition 3.2 (cf. [HLP23, Defn. 1.1.3]).

Definition 5.1.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a morphism of noetherian algebraic stacks, 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} is a closed substack, and 𝒳0f1(𝒴0)\mathcal{X}_{0}\subset f^{-1}(\mathcal{Y}_{0}) is closed substack. We say that the triple (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) satisfies relative formal functions (resp. relatively (derived) coherently complete) if (𝒳R^,(𝒳0)R^)(\mathcal{X}_{\widehat{R}},(\mathcal{X}_{0})_{\widehat{R}}) satisfies formal functions (resp. (derived) coherent completeness) for all noetherian rings RR and smooth morphisms ρ:SpecR𝒴\rho\colon\operatorname{Spec}R\to\mathcal{Y}, where R^\widehat{R} denotes the adic completion of RR with respect to the ideal defining ρ1(𝒴0)SpecR\rho^{-1}(\mathcal{Y}_{0})\subset\operatorname{Spec}R, 𝒳R^=𝒳×𝒴SpecR^\mathcal{X}_{\widehat{R}}=\mathcal{X}\times_{\mathcal{Y}}\operatorname{Spec}\widehat{R}, and (𝒳0)R^(\mathcal{X}_{0})_{\widehat{R}} is the preimage of 𝒳0\mathcal{X}_{0} under 𝒳R^𝒳\mathcal{X}_{\widehat{R}}\to\mathcal{X}.

We have the following simple lemma.

Lemma 5.2.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a morphism of noetherian algebraic stacks, 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} a closed substack, and 𝒳0f1(𝒴0)\mathcal{X}_{0}\subset f^{-1}(\mathcal{Y}_{0}) a closed substack. If the triple (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) satisfies relative formal functions, then for every Dcoh+(𝒳)\mathcal{F}\in D^{+}_{\operatorname{coh}}(\mathcal{X}) the natural comparison morphism:

(𝖱f)𝖱f^^(\mathsf{R}f_{\ast}\mathcal{F})^{\wedge}\to\mathsf{R}\widehat{f}_{\ast}\widehat{\mathcal{F}}

is an isomorphism. If, in addition, the pair (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}_{0}) satisfies formal functions, then so does (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}).

Proof.

The second claim follows from the first since for any coherent sheaf \mathcal{F} on 𝒳\mathcal{X}, we have isomorphisms:

𝖱Γ(𝒳,)𝖱Γ(𝒴,𝖱f)𝖱Γ(𝒴^,(𝖱f))𝖱Γ(𝒴^,𝖱f^^)𝖱Γ(𝒳^,^).\mathsf{R}\Gamma(\mathcal{X},\mathcal{F})\simeq\mathsf{R}\Gamma(\mathcal{Y},\mathsf{R}f_{\ast}\mathcal{F})\simeq\mathsf{R}\Gamma(\widehat{\mathcal{Y}},(\mathsf{R}f_{\ast}\mathcal{F})^{\wedge})\simeq\mathsf{R}\Gamma(\widehat{\mathcal{Y}},\mathsf{R}\widehat{f}_{\ast}\widehat{\mathcal{F}})\simeq\mathsf{R}\Gamma(\widehat{\mathcal{X}},\widehat{\mathcal{F}}).

To treat the first claim, let SpecR𝒴\operatorname{Spec}R\to\mathcal{Y} be a smooth morphism, IRI\subset R the ideal defined by the preimage of 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y}, and R^\widehat{R} the II-adic completion of RR. Let fR:𝒳RSpecRf_{R}\colon\mathcal{X}_{R}\to\operatorname{Spec}R be the base change morphism, where 𝒳R=𝒳×𝒴SpecR\mathcal{X}_{R}=\mathcal{X}\times_{\mathcal{Y}}\operatorname{Spec}R. For an integer ii, i((𝖱f))\mathscr{H}^{i}((\mathsf{R}f_{\ast}\mathcal{F})^{\wedge}) is the sheafification of the presheaf:

(SpecR𝒴)Hi(𝒳R,R)RR^Hi(𝒳R^,R^),(\operatorname{Spec}R\to\mathcal{Y})\mapsto\mathrm{H}^{i}(\mathcal{X}_{R},\mathcal{F}_{R})\otimes_{R}\widehat{R}\simeq\mathrm{H}^{i}(\mathcal{X}_{\widehat{R}},\mathcal{F}_{\widehat{R}}),

with the last isomorphism by flat base change. One similarly finds that i(𝖱f^^)\mathscr{H}^{i}(\mathsf{R}\widehat{f}_{\ast}\widehat{\mathcal{F}}) is the sheafification of the presheaf:

(SpecR𝒴)Hi(𝒳^R^,^R^).(\operatorname{Spec}R\to\mathcal{Y})\mapsto\mathrm{H}^{i}(\widehat{\mathcal{X}}_{\widehat{R}},\widehat{\mathcal{F}}_{{\widehat{R}}}).

By assumption, the pair (𝒳R^,(𝒳0)R^)(\mathcal{X}_{\widehat{R}},(\mathcal{X}_{0})_{\widehat{R}}) satisfies formal functions, and the result follows. ∎

The following is a reformulation of Theorem 4.6 in the relative and derived situation.

Theorem 5.3.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a finite type morphism of noetherian algebraic stacks, 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} a closed substack, and 𝒳0f1(𝒴0)\mathcal{X}_{0}\subset f^{-1}(\mathcal{Y}_{0}) a closed substack. Assume that the triple (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) is strongly cohomologically proper.

  1. (1)

    The triple (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) satisfies relative formal functions.

  2. (2)

    The comparison morphism (𝖱f)𝖱f^^(\mathsf{R}f_{*}\mathcal{F})^{\wedge}\to\mathsf{R}\widehat{f}_{*}\widehat{\mathcal{F}} is an isomorphism for all DCoh+(𝒳)\mathcal{F}\in D^{+}_{\operatorname{Coh}}(\mathcal{X}).

  3. (3)

    The functor 𝖱f^\mathsf{R}\widehat{f}_{*} sends DCoh+(𝒳^)D^{+}_{\operatorname{Coh}}(\widehat{\mathcal{X}}) to DCoh+(𝒴^)D^{+}_{\operatorname{Coh}}(\widehat{\mathcal{Y}}).

Proof.

By Lemma 5.2, (1)\Rightarrow(2). The other claims follow from Theorem 4.6. ∎

Cohomologically proper morphisms and relative formal functions are stable under composition. The analogous question for coherent completeness seems more subtle. The following theorem asserts that they ascend under proper representable morphisms, however.

Theorem 5.4.

Let 𝒴\mathcal{Y} be a noetherian algebraic stack with affine stabilizers and 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} a closed substack. Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a proper and representable morphism. If (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}_{0}) satisfies formal functions and is coherently complete, then so does (𝒳,f1(𝒴0))(\mathcal{X},f^{-1}(\mathcal{Y}_{0})).

Remark 5.5.

This theorem is similar in spirit to the statement of [HLP23, Thm. 4.2.1] that cohomologically projective morphisms (see [HLP23, Def.4.1.3]) are formally proper. Projective morphisms are cohomologically projective and thus formally proper. Using Rydh’s Chow Lemma [Ryd16a] and the fact that formal properness descends under proper surjective morphisms (Remark 6.2), it follows that proper morphisms are also formally proper.

Proof of Theorem 5.4.

Let 𝒳0=f1(𝒴0)\mathcal{X}_{0}=f^{-1}(\mathcal{Y}_{0}). We have the following diagram:

Dqcoh+(𝒳)\textstyle{D^{+}_{\operatorname{qcoh}}(\mathcal{X})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝖱f\scriptstyle{\mathsf{R}f_{*}}Dqcoh(𝒳)\textstyle{D_{\operatorname{qcoh}}(\mathcal{X})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}c\scriptstyle{c^{*}}D(𝒳^)\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces D(\widehat{\mathcal{X}})}cqcoh,\scriptstyle{c_{{\operatorname{qcoh}},*}}D+(𝒳^)\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces D^{+}(\widehat{\mathcal{X}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝖱f^\scriptstyle{\mathsf{R}\widehat{f}_{*}}Dqcoh+(𝒴)\textstyle{D^{+}_{\operatorname{qcoh}}(\mathcal{Y})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Dqcoh(𝒴)\textstyle{D_{\operatorname{qcoh}}(\mathcal{Y})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d^{*}}D(𝒴^)\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces D(\widehat{\mathcal{Y}})}dqcoh,\scriptstyle{d_{{\operatorname{qcoh}},*}}D+(𝒴^).\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces D^{+}(\widehat{\mathcal{Y}}).}

Since ff is proper and representable, it is universally cohomologically proper [Ols05, Thm. 1.2], so the triple (f,𝒳0,𝒴0)(f,\mathcal{X}_{0},\mathcal{Y}_{0}) is strongly cohomologically proper (Example 4.4). In particular, Theorem 5.3 implies that ff satisfies relative formal functions and 𝖱f^\mathsf{R}\widehat{f}_{*} sends Dcoh+(𝒳^)D^{+}_{\operatorname{coh}}(\widehat{\mathcal{X}}) to Dcoh+(𝒴^)D^{+}_{\operatorname{coh}}(\widehat{\mathcal{Y}}). A simple calculation using homotopy limits shows that because ff is representable, 𝖱f^\mathsf{R}\widehat{f}_{*} even sends Dcohb(𝒳^)D^{b}_{\operatorname{coh}}(\widehat{\mathcal{X}}) to Dcohb(𝒴^)D^{b}_{\operatorname{coh}}(\widehat{\mathcal{Y}}). In particular, Lemma 5.2 implies that (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions. By Proposition 3.5(2) it remains to prove that

Coh(𝒳)Coh(𝒳^)\operatorname{Coh}(\mathcal{X})\to\operatorname{Coh}(\widehat{\mathcal{X}})

is essentially surjective. We first prove this under the following assumptions:

  1. (1)

    ff is projective (so comes with a relatively ample line bundle \mathcal{L}); and

  2. (2)

    𝒳0\mathcal{X}_{0} is a Cartier divisor.

In this case, let 𝔉Coh(𝒳^)\mathfrak{F}\in\operatorname{Coh}(\widehat{\mathcal{X}}) and let n𝐙n\in\mathbf{Z}. Then by the projection formula [Hal23, Thm. A.12],

𝖱f(n𝒪𝒳𝖫cqcoh,𝔉)\displaystyle\mathsf{R}f_{*}(\mathcal{L}^{\otimes n}\otimes^{\mathsf{L}}_{\mathcal{O}_{\mathcal{X}}}c_{{\operatorname{qcoh}},*}\mathfrak{F}) 𝖱fcqcoh,(^n𝒪𝒳^𝖫𝔉)\displaystyle\simeq\mathsf{R}f_{*}c_{{\operatorname{qcoh}},*}({\widehat{\mathcal{L}}}^{\otimes n}\otimes^{\mathsf{L}}_{\mathcal{O}_{\widehat{\mathcal{X}}}}\mathfrak{F})
dqcoh,𝖱f^(^n𝒪𝒳^𝖫𝔉).\displaystyle\simeq d_{{\operatorname{qcoh}},*}\mathsf{R}\widehat{f}_{*}({\widehat{\mathcal{L}}}^{\otimes n}\otimes^{\mathsf{L}}_{\mathcal{O}_{\widehat{\mathcal{X}}}}\mathfrak{F}).

But 𝖱f^(^n𝒪𝒳^𝖫𝔉)Dcohb(𝒴^)\mathsf{R}\widehat{f}_{*}({\widehat{\mathcal{L}}}^{\otimes n}\otimes^{\mathsf{L}}_{\mathcal{O}_{\widehat{\mathcal{X}}}}\mathfrak{F})\in D^{b}_{\operatorname{coh}}(\widehat{\mathcal{Y}}) and since (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}_{0}) is derived coherently complete, by (Proposition 3.5(2)) dqcoh,𝖱f^(^n𝒪𝒳^𝖫𝔉)Dcohb(𝒴).d_{{\operatorname{qcoh}},*}\mathsf{R}\widehat{f}_{*}({\widehat{\mathcal{L}}}^{\otimes n}\otimes^{\mathsf{L}}_{\mathcal{O}_{\widehat{\mathcal{X}}}}\mathfrak{F})\in D^{b}_{\operatorname{coh}}(\mathcal{Y}).

Let p:SpecA𝒴p\colon\operatorname{Spec}A\to\mathcal{Y} be a smooth cover by an affine scheme. Let 𝒳A=𝒳×𝒴SpecA\mathcal{X}_{A}=\mathcal{X}\times_{\mathcal{Y}}\operatorname{Spec}A and let fA:𝒳ASpecAf_{A}\colon\mathcal{X}_{A}\to\operatorname{Spec}A and q:𝒳A𝒳q\colon\mathcal{X}_{A}\to\mathcal{X} be the induced morphisms. We may choose pp so that qq^{*}\mathcal{L} is ample relative to fAf_{A}. Then

𝖱Γ(𝒳A,(q)n𝒪𝒳A𝖫qcqcoh,𝔉)\displaystyle\mathsf{R}\Gamma(\mathcal{X}_{A},(q^{*}\mathcal{L})^{\otimes n}\otimes^{\mathsf{L}}_{\mathcal{O}_{\mathcal{X}_{A}}}q^{*}c_{{\operatorname{qcoh}},*}\mathfrak{F}) 𝖱Γ(𝒳A,q(n𝒪𝒳𝖫cqcoh,𝔉))\displaystyle\simeq\mathsf{R}\Gamma(\mathcal{X}_{A},q^{*}(\mathcal{L}^{\otimes n}\otimes^{\mathsf{L}}_{\mathcal{O}_{\mathcal{X}}}c_{{\operatorname{qcoh}},*}\mathfrak{F}))
𝖱Γ(SpecA,𝖱fA,q(n𝒪𝒳𝖫cqcoh,𝔉))\displaystyle\simeq\mathsf{R}\Gamma(\operatorname{Spec}A,\mathsf{R}f_{A,*}q^{*}(\mathcal{L}^{\otimes n}\otimes^{\mathsf{L}}_{\mathcal{O}_{\mathcal{X}}}c_{{\operatorname{qcoh}},*}\mathfrak{F}))
𝖱Γ(SpecA,p𝖱f(n𝒪𝒳𝖫cqcoh,𝔉))\displaystyle\simeq\mathsf{R}\Gamma(\operatorname{Spec}A,p^{*}\mathsf{R}f_{*}(\mathcal{L}^{\otimes n}\otimes^{\mathsf{L}}_{\mathcal{O}_{\mathcal{X}}}c_{{\operatorname{qcoh}},*}\mathfrak{F}))
Dcohb(A).\displaystyle\in D^{b}_{\operatorname{coh}}(A).

Since this is true for all n𝐙n\in\mathbf{Z} and qq^{*}\mathcal{L} is ample relative to fAf_{A}, it follows from [Hal23, Thm. 3.8] that qcqcoh,𝔉Dcohb(𝒳A)q^{*}c_{{\operatorname{qcoh}},*}\mathfrak{F}\in D^{b}_{\operatorname{coh}}(\mathcal{X}_{A}). By smooth descent, cqcoh,𝔉Dcohb(𝒳)c_{{\operatorname{qcoh}},*}\mathfrak{F}\in D^{b}_{\operatorname{coh}}(\mathcal{X}). It remains to prove that the adjunction morphism

(5.2) ccqcoh,𝔉𝔉\displaystyle c^{*}c_{{\operatorname{qcoh}},*}\mathfrak{F}\to\mathfrak{F}

is an isomorphism.

To prove this, observe that 𝒪𝒳0\mathcal{O}_{\mathcal{X}_{0}} is a perfect complex on 𝒳\mathcal{X} since 𝒳0𝒳\mathcal{X}_{0}\subseteq\mathcal{X} is Cartier. It follows from the projection formula that

(ccqcoh,𝔉)𝒪𝒳^𝖫c𝒪𝒳0c(cqcoh,𝔉𝒪𝒳𝖫𝒪𝒳0)ccqcoh,(𝔉𝒪𝒳^𝖫c𝒪𝒳0)𝔉𝒪𝒳^𝖫c𝒪𝒳0.(c^{*}c_{{\operatorname{qcoh}},*}\mathfrak{F})\otimes^{\mathsf{L}}_{\mathcal{O}_{\widehat{\mathcal{X}}}}c^{*}\mathcal{O}_{\mathcal{X}_{0}}\simeq c^{*}(c_{{\operatorname{qcoh}},*}\mathfrak{F}\otimes^{\mathsf{L}}_{\mathcal{O}_{{\mathcal{X}}}}\mathcal{O}_{\mathcal{X}_{0}})\simeq c^{*}c_{{\operatorname{qcoh}},*}(\mathfrak{F}\otimes^{\mathsf{L}}_{\mathcal{O}_{\widehat{\mathcal{X}}}}c^{*}\mathcal{O}_{\mathcal{X}_{0}})\simeq\mathfrak{F}\otimes^{\mathsf{L}}_{\mathcal{O}_{\widehat{\mathcal{X}}}}c^{*}\mathcal{O}_{\mathcal{X}_{0}}.

In summary, if 𝔎\mathfrak{K} denotes the cone of (5.2), we have shown that 𝔎𝒪𝒳^𝖫c𝒪𝒳00\mathfrak{K}\otimes^{\mathsf{L}}_{\mathcal{O}_{\widehat{\mathcal{X}}}}c^{*}\mathcal{O}_{\mathcal{X}_{0}}\simeq 0. To finish the proof, suppose that 𝔎0\mathfrak{K}\neq 0. Then since 𝔎Dcohb(𝒳^)\mathfrak{K}\in D^{b}_{\operatorname{coh}}(\widehat{\mathcal{X}}), we may choose a largest integer qq such that q(𝔎)0\mathscr{H}^{q}(\mathfrak{K})\neq 0. Then

q(𝔎)𝒪𝒳^c𝒪𝒳0q(𝔎𝒪𝒳^𝖫c𝒪𝒳0)0\mathscr{H}^{q}(\mathfrak{K})\otimes_{\mathcal{O}_{\widehat{\mathcal{X}}}}c^{*}\mathcal{O}_{\mathcal{X}_{0}}\simeq\mathscr{H}^{q}(\mathfrak{K}\otimes^{\mathsf{L}}_{\mathcal{O}_{\widehat{\mathcal{X}}}}c^{*}\mathcal{O}_{\mathcal{X}_{0}})\simeq 0

and so by Nakayama’s Lemma, q(𝔎)0\mathscr{H}^{q}(\mathfrak{K})\simeq 0, which is a contradiction.

In general, Rydh’s Chow Lemma [Ryd16a] provides a blow-up 𝒳𝜋𝒳\mathcal{X}^{\prime}\xrightarrow{\pi}\mathcal{X} such that the composition g=fπg=f\circ\pi is projective. We may replace 𝒳\mathcal{X}^{\prime} by an additional blow-up so that 𝒳0=g1(𝒴0)𝒳\mathcal{X}_{0}^{\prime}=g^{-1}(\mathcal{Y}_{0})\subseteq\mathcal{X}^{\prime} becomes Cartier. It follows from the case already considered that (𝒳,𝒳0)(\mathcal{X}^{\prime},\mathcal{X}^{\prime}_{0}) is derived coherently complete. By Theorem 6.1(3), it follows that (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) is derived coherently complete. ∎

We also have the following easy result.

Lemma 5.6.

Let 𝒳\mathcal{X} be a noetherian algebraic stack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} a closed substack. Suppose that 𝒵𝒳\mathcal{Z}\subset\mathcal{X} is a closed substack such that |𝒳0||𝒵||\mathcal{X}_{0}|\subseteq|\mathcal{Z}|. If (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions (resp. is coherently complete), then the same holds for (𝒳,𝒵)(\mathcal{X},\mathcal{Z}).

We collect here some examples of algebraic stacks 𝒳\mathcal{X} that satisfy each of the three properties: cohomological properness, formal functions and coherent completeness.

Example 5.7 (Noetherian affine schemes).

Let AA be a noetherian ring and let IAI\subseteq A be an ideal. Then (SpecA,SpecA/I)(\operatorname{Spec}A,\operatorname{Spec}A/I) is coherently complete if and only if AA is II-adically complete [AHR19, Ex. 3.9].

Example 5.8 (Proper algebraic stacks).

A proper morphism 𝒳𝒴\mathcal{X}\to\mathcal{Y} of noetherian algebraic stacks is universally cohomologically proper [Ols05, Thm. 1.2]. In the case that 𝒴=SpecR\mathcal{Y}=\operatorname{Spec}R and IRI\subset R is an ideal such that RR is II-adically complete, then setting 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} to be the preimage of SpecR/ISpecR\operatorname{Spec}R/I\subset\operatorname{Spec}R, the pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) satisfies formal functions and is coherently complete [Ols05, Con05].

Example 5.9 (Good moduli spaces).

Let 𝒳SpecR\mathcal{X}\to\operatorname{Spec}R be a good moduli space [Alp13], where 𝒳\mathcal{X} is a noetherian algebraic stack with affine diagonal. Then 𝒳SpecR\mathcal{X}\to\operatorname{Spec}R is of finite type [AHR20, Thm. A.1], and RR is noetherian and 𝒳\mathcal{X} is cohomologically proper over RR [Alp13, Thm. 4.16 (x)]. Since good moduli spaces are compatible with arbitrary base change, 𝒳\mathcal{X} is even universally cohomologically proper over RR.

Let 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} be a closed substack defined by a coherent sheaf of ideals 𝒪𝒳\mathcal{I}\subset\mathcal{O}_{\mathcal{X}} and let I=Γ(𝒳,)I=\Gamma(\mathcal{X},\mathcal{I}) to be the corresponding ideal of RR. If 𝒳0\mathcal{X}_{0} has the resolution property and RR is II-adically complete, then the pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) is coherently complete and satisfies formal functions [AHR19, Thm. 1.6]. When 𝒳\mathcal{X} has the resolution property and 𝒳0\mathcal{X}_{0} is the preimage of SpecR/I\operatorname{Spec}R/I, then this was the main result of [GZB15]. If RR is equicharacteristic, 𝒳\mathcal{X} has the resolution property and 𝒳0\mathcal{X}_{0} is supported at a closed point, then this is [AHR20, Thm. 1.3].

We also point out the following converse. Let 𝒳\mathcal{X} be a noetherian algebraic stack with noetherian adequate moduli space π:𝒳X\pi\colon\mathcal{X}\to X. Let 𝒳0𝒳\mathcal{X}_{0}\subseteq\mathcal{X} be a closed substack with scheme-theoretic image X0XX_{0}\subset X. If X=SpecAX=\operatorname{Spec}A is affine and the pair (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) is coherently complete, then the pair (X,X0)(X,X_{0}) is coherently complete [AHR19, Lem. 3.5(1)].

Example 5.10 (Adequate moduli spaces).

Let 𝒳=[SpecA/G]\mathcal{X}=[\operatorname{Spec}A/G], where GG is a reductive algebraic group over a noetherian ring RR acting on an affine scheme SpecA\operatorname{Spec}A of finite type over RR. Then the cohomology ring H(𝒳,𝒪𝒳)=H(G,A)H^{*}(\mathcal{X},\mathcal{O}_{\mathcal{X}})=H^{*}(G,A) is a finitely generated RR-algebra. See [TvdK10, Thm. 1.1] for the case when RR is a field and [vdK15] in general. It follows that 𝒳Spec(AG)\mathcal{X}\to\operatorname{Spec}(A^{G}) is cohomologically proper. If JAJ\subseteq A is a GG-equivariant ideal AGA^{G} is JGJ^{G}-adically complete, then (𝒳,[Spec(A/J)/G])(\mathcal{X},[\operatorname{Spec}(A/J)/G]) satisfies formal functions (Corollary 4.9).

Example 5.11 (Unipotent groups).

If UU is a unipotent (affine) algebraic group over an algebraically closed field kk, then BUBU is cohomlogically proper if and only if char(k)=0{\rm char}(k)=0. Indeed, if char(k)=0{\rm char}(k)=0, then UU admits a filtration by 𝐆a\mathbf{G}_{a}’s and thus it suffices to show that B𝐆aB\mathbf{G}_{a} is cohomologically proper. Any finite dimensional representation of 𝐆a\mathbf{G}_{a} has a filtration by trivial representations and so it is enough to compute the cohomology Hi(B𝐆a,k)\mathrm{H}^{i}(B\mathbf{G}_{a},k) of the trivial representation. Using the Čech complex corresponding to the cover SpeckB𝐆a\operatorname{Spec}k\to B\mathbf{G}_{a}, one can compute that Hi(B𝐆a,k)=k\mathrm{H}^{i}(B\mathbf{G}_{a},k)=k if i=0,1i=0,1 and zero otherwise.

On the other hand, if char(k)=p{\rm char}(k)=p, then we can compute (again using the Čech complex) that H1(B𝐆a,k)H^{1}(B\mathbf{G}_{a},k) is infinite dimensional. Thus B𝐆aB\mathbf{G}_{a} is not cohomologically proper and it follows that neither is BUBU for any unipotent group UU.

Supposing again that char(k)=0{\rm char}(k)=0, let (R,𝔪)(R,\mathfrak{m}) be a complete local noetherian ring with residue field kk such that RR is not artinian. Then (BUR,BUk)(BU_{R},BU_{k}) is not coherently complete. It suffices to assume that U=𝐆aU=\mathbf{G}_{a} and that R=k[[ϵ]]R=k[[\epsilon]]. Consider the exponential map 𝐆a,R/𝔪n𝐆m,R/𝔪n\mathbf{G}_{a,R/\mathfrak{m}^{n}}\to\mathbf{G}_{m,R/\mathfrak{m}^{n}} defined by texp(ϵx)=1+(ϵx)+12(ϵx)2+t\mapsto\exp(\epsilon x)=1+(\epsilon x)+\frac{1}{2}(\epsilon x)^{2}+\cdots, where tt and xx are coordinates of 𝐆m\mathbf{G}_{m} and 𝐆a\mathbf{G}_{a}, respectively. This defines a compatible system of non-trivial line bundles on B𝐆a,R/𝔪nB\mathbf{G}_{a,R/\mathfrak{m}^{n}} that does not algebraize to a line bundle on B𝐆a,RB\mathbf{G}_{a,R}. Indeed, every line bundle on B𝐆a,RB\mathbf{G}_{a,R} is trivial.

Example 5.12 (Universal cohomological properness of BGBG in characteristic 0).

If GG is an affine group scheme of finite type over an algebraically closed field kk of characteristic 0, then BGBG is cohomologically proper. To see this, let GuG_{u} be the unipotent radical of GG so that Gr:=G/GuG_{r}:=G/G_{u} is linearly reductive. Since BGrBG_{r} has vanishing higher coherent cohomology, it is cohomologically proper. Therefore, it suffices to prove that BGBGrBG\to BG_{r} is cohomologically proper. By flat base change and descent, it remains to prove that the base change BGuSpeckBG_{u}\to\operatorname{Spec}k (of BGBGrBG\to BG_{r} by SpeckBGr\operatorname{Spec}k\to BG_{r}) is cohomologically proper but this follows from Example 5.11.

More generally, BGSpeckBG\to\operatorname{Spec}k is universally cohomologically proper. In particular, it follows that if TT is an JJ-adically complete noetherian kk-algebra, then (BGT,BGT/J)(BG_{T},BG_{T/J}) satisfies formal functions (Corollary 4.8). To see this, let RR be a noetherian kk-algebra. We first observe that BGRBG_{R} has finite cohomological dimension. Since GG is affine, we may choose an embedding GGLnG\subset\mathrm{GL}_{n}. This induces a morphism BGRBGLn,RBG_{R}\to B\mathrm{GL}_{n,R} of algebraic stacks such that the base change by SpecRBGLn,R\operatorname{Spec}R\to B\mathrm{GL}_{n,R} is the quotient GLn,R/GR\mathrm{GL}_{n,R}/G_{R}, which we know has finite cohomological dimension. Let dd be the cohomological dimension. If MM is a coherent BGRBG_{R}-module, then there is a BGkBG_{k}-module NN and a surjection NR=NkRMN_{R}=N\otimes_{k}R\to M. Indeed, the morphism r:BGRBGkr\colon BG_{R}\to BG_{k} is affine, so the adjunction rrMMr^{*}r_{*}M\to M is surjective. We may write rM=limλNλr_{*}M=\varinjlim_{\lambda}N_{\lambda}, where each MλM_{\lambda} is a vector bundles on BGkBG_{k}. Since MM is coherent on BGRBG_{R}, there is a λ0\lambda_{0} sufficiently large such that rrNλMr^{*}r_{*}N_{\lambda}\twoheadrightarrow M for all λλ0\lambda\geq\lambda_{0}. Now take N=Nλ0N=N_{\lambda_{0}}. Letting LL be its kernel, there is a long exact sequence

Hd(BGR,L)Hd(BGR,NR)Hd(BGR,M)0\cdots\to\mathrm{H}^{d}(BG_{R},L)\to\mathrm{H}^{d}(BG_{R},N_{R})\to\mathrm{H}^{d}(BG_{R},M)\to 0

By flat base change, we know that Hd(BGR,NR)=Hd(BG,N)kR\mathrm{H}^{d}(BG_{R},N_{R})=\mathrm{H}^{d}(BG,N)\otimes_{k}R, which is a finite RR-module, so Hd(BGR,M)\mathrm{H}^{d}(BG_{R},M) is a finite RR-module. Since this holds for any coherent BGRBG_{R}-module MM, we may perform descending induction to conclude that Hi(BGR,M)\mathrm{H}^{i}(BG_{R},M) is a finite RR-module for all ii.

6. Descent of properties

The main result of this section is a criterion for descending cohomological properness, formal functions and coherent completeness along universally submersive morphisms. Recall that a finite type morphism 𝒳𝒴\mathcal{X}\to\mathcal{Y} of noetherian algebraic stacks is universally submersive if for every morphism V𝒴V\to\mathcal{Y} from a scheme, the base change 𝒳×𝒴VV\mathcal{X}\times_{\mathcal{Y}}V\to V is surjective and VV has the quotient topology. This is equivalent to requiring that for every morphism SpecR𝒴\operatorname{Spec}R\to\mathcal{Y} from a DVR RR there is an extension of DVRs RRR\to R^{\prime} and a lift SpecR𝒳\operatorname{Spec}R^{\prime}\to\mathcal{X} of the composition SpecRSpecR𝒳\operatorname{Spec}R^{\prime}\to\operatorname{Spec}R\to\mathcal{X}.

The most important examples for us are when 𝒳𝒴\mathcal{X}\to\mathcal{Y} is representable, proper, and surjective, or when 𝒳𝒴\mathcal{X}\to\mathcal{Y} is representable, faithfully flat, and locally of finite presentation. For proper morphisms, these descent results are rather standard. The novelty here is that they hold more generally as long as you require the corresponding property on not only 𝒳\mathcal{X} but the higher fiber products

(𝒳/𝒴)n𝒳×𝒴×𝒴𝒳n(\mathcal{X}/\mathcal{Y})^{n}\coloneqq\underbrace{\mathcal{X}\times_{\mathcal{Y}}\cdots\times_{\mathcal{Y}}\mathcal{X}}_{n}

for all n1n\geq 1.

Theorem 6.1.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a representable, universally submersive and finite type morphism of noetherian algebraic stacks.

  1. (1)

    Suppose that 𝒴\mathcal{Y} is of finite type over a noetherian ring RR. If (𝒳/𝒴)n(\mathcal{X}/\mathcal{Y})^{n} is cohomologically proper over SpecR\operatorname{Spec}R for n1n\geq 1, then so is 𝒴\mathcal{Y}.

  2. (2)

    Let 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} be a closed substack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} be its preimage. If ((𝒳/𝒴)n,(𝒳0/𝒴0)n)((\mathcal{X}/\mathcal{Y})^{n},(\mathcal{X}_{0}/\mathcal{Y}_{0})^{n}) satisfies formal functions for n1n\geq 1, then so does (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}_{0}).

  3. (3)

    Let 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} be a closed substack defined by a coherent sheaf of ideals 𝒪𝒴\mathcal{I}\subset\mathcal{O}_{\mathcal{Y}} and let 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} be its preimage. If (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}) is coherently complete and ((𝒳/𝒴)n,(𝒳0/𝒴0)n)((\mathcal{X}/\mathcal{Y})^{n},(\mathcal{X}_{0}/\mathcal{Y}_{0})^{n}) satisfies formal functions for n1n\geq 1, then (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}_{0}) is coherently complete and satisfies formal functions.

If f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} is proper and surjective, then in each case one only needs to require the properties for n=1n=1.

Remark 6.2.

This theorem is similar in spirit to the statement [HLP23, Thm. 3.3.1] that an h-covering 𝒳𝒳\mathcal{X}^{\prime}\to\mathcal{X} induces an equivalence of \infty-categories of almost perfect complexes on 𝒳\mathcal{X} and on the Čech nerve of 𝒳𝒳\mathcal{X}^{\prime}\to\mathcal{X} and its consequence [HLP23, Thm. 4.3.1] that formal properness descends under proper surjective morphisms.

Proof.

We will argue (1) and (2) by noetherian induction on the abelian category Coh(𝒴)\operatorname{Coh}(\mathcal{Y}) and we may assume that every proper closed substack 𝒵𝒴\mathcal{Z}\subsetneq\mathcal{Y} is cohomologically proper over RR and that the pair (𝒵,𝒵𝒳0)(\mathcal{Z},\mathcal{Z}\cap\mathcal{X}_{0}) satisfies formal functions. For (3), note that (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}_{0}) satisfies formal functions by (2) and the functor Coh(𝒴)Coh(𝒴^)\operatorname{Coh}(\mathcal{Y})\to\operatorname{Coh}(\widehat{\mathcal{Y}}) is fully faithful, with image stable under kernels, cokernels and extensions, by Corollary 3.8. Thus it suffices to show that Coh(𝒴)Coh(𝒴^)\operatorname{Coh}(\mathcal{Y})\to\operatorname{Coh}(\widehat{\mathcal{Y}}) is essentially surjective. To accomplish this, we will also argue via dévissage on the abelian category Coh(𝒴^)\operatorname{Coh}(\widehat{\mathcal{Y}}) and we may assume that if 𝔊\mathfrak{G} is a coherent sheaf on 𝒴^\widehat{\mathcal{Y}} annihilated by 𝒥𝒪𝒴^\mathcal{J}\mathcal{O}_{\widehat{\mathcal{Y}}} for some non-zero coherent sheaf of ideals 𝒥𝒪𝒴\mathcal{J}\subset\mathcal{O}_{\mathcal{Y}}, then 𝔊\mathfrak{G} is in the essential image.

In each case, we may further reduce to the situation where 𝒴\mathcal{Y} is reduced. By generic flatness, there is a dense open substack 𝒰𝒴\mathcal{U}\subset\mathcal{Y} such that f1(𝒰)𝒰f^{-1}(\mathcal{U})\to\mathcal{U} is flat. By Rydh’s extension [Ryd16b] of Raynaud-Gruson’s theorem [RG71] on flatification by blow-ups, there is a commutative diagram:

𝒳{\mathcal{X}^{\prime}}𝒳{\mathcal{X}}𝒴{\mathcal{Y}^{\prime}}𝒴,{\mathcal{Y},}g\scriptstyle{g^{\prime}}f\scriptstyle{f^{\prime}}f\scriptstyle{f}g\scriptstyle{g}

where g:𝒴𝒴g\colon\mathcal{Y}^{\prime}\to\mathcal{Y} is a blow-up along a closed substack contained in the complement of 𝒰\mathcal{U}, 𝒳\mathcal{X}^{\prime} is the strict transform of 𝒳\mathcal{X} along g:𝒴𝒴g\colon\mathcal{Y}^{\prime}\to\mathcal{Y}, and f:𝒳𝒴f^{\prime}\colon\mathcal{X}^{\prime}\to\mathcal{Y}^{\prime} is flat. Observe that g:𝒴𝒴g\colon\mathcal{Y}^{\prime}\to\mathcal{Y} is a surjective, proper and representable morphism. Moreover, since f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} is universally submersive, f:𝒳𝒴f^{\prime}\colon\mathcal{X}^{\prime}\to\mathcal{Y}^{\prime} is surjective. Indeed, for every point y|𝒴|y^{\prime}\in|\mathcal{Y}^{\prime}| there is a morphism SpecR𝒴\operatorname{Spec}R\to\mathcal{Y}^{\prime} from a DVR whose generic point maps to g1(𝒰)g^{-1}(\mathcal{U}). As 𝒳𝒴\mathcal{X}\to\mathcal{Y} is universally submersive, there exists an extension of DVRs RRR\to R^{\prime} and a lift SpecR𝒳\operatorname{Spec}R^{\prime}\to\mathcal{X} of SpecRSpecR𝒴𝒴\operatorname{Spec}R^{\prime}\to\operatorname{Spec}R\to\mathcal{Y}^{\prime}\to\mathcal{Y}. The induced map SpecR𝒳×𝒴𝒴\operatorname{Spec}R^{\prime}\to\mathcal{X}\times_{\mathcal{Y}}\mathcal{Y}^{\prime} factors through the strict transform 𝒳\mathcal{X}^{\prime}. The image of the closed point under SpecR𝒳\operatorname{Spec}R^{\prime}\to\mathcal{X}^{\prime} is a preimage of yy^{\prime} and we conclude that f:𝒳𝒴f^{\prime}\colon\mathcal{X}^{\prime}\to\mathcal{Y}^{\prime} is a faithfully flat morphism of finite type.

Since g:𝒴𝒴g\colon\mathcal{Y}^{\prime}\to\mathcal{Y} is separated and representable, its diagonal is a closed immersion and it follows that (𝒳/𝒴)n(𝒳/𝒴)n(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{n}\to(\mathcal{X}^{\prime}/\mathcal{Y})^{n} is also closed immersion for each n1n\geq 1. Since g:𝒳𝒳g^{\prime}\colon\mathcal{X}^{\prime}\to\mathcal{X} is proper, (𝒳/𝒴)n(𝒳/𝒴)n(\mathcal{X}^{\prime}/\mathcal{Y})^{n}\to(\mathcal{X}/\mathcal{Y})^{n} is proper and it follows that (𝒳/𝒴)n(𝒳/𝒴)n(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{n}\to(\mathcal{X}/\mathcal{Y})^{n} is proper and thus universally cohomologically proper (Example 5.8).

We now establish (1). The hypotheses that each (𝒳/𝒴)n(\mathcal{X}/\mathcal{Y})^{n} is cohomologically proper over RR implies that (𝒳/𝒴)n(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{n} is also cohomologically proper over RR. Since f:𝒳𝒴f^{\prime}:\mathcal{X}^{\prime}\to\mathcal{Y}^{\prime} is faithfully flat, cohomological descent implies that for a coherent sheaf \mathcal{M} on 𝒴\mathcal{Y}^{\prime}, there is a convergent spectral sequence:

Hi((𝒳/𝒴)j,(𝒳/𝒴)j)Hi+j(𝒴,)\mathrm{H}^{i}((\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j},\mathcal{M}_{(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j}})\Rightarrow\mathrm{H}^{i+j}(\mathcal{Y}^{\prime},\mathcal{M})

where (𝒳/𝒴)j\mathcal{M}_{(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j}} denotes the pullback of \mathcal{M} under the projection (𝒳/𝒴)j𝒴(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j}\to\mathcal{Y}^{\prime}. It follows immediately that 𝒴\mathcal{Y}^{\prime} is cohomologically proper over RR.

For a coherent sheaf \mathcal{F} on 𝒴\mathcal{Y}, consider the exact triangle of complexes on 𝒴\mathcal{Y}:

𝖱gg𝒞[1].\mathcal{F}\to\mathsf{R}g_{\ast}g^{\ast}\mathcal{F}\to\mathcal{C}\to\mathcal{F}[1].

Since gg is proper and representable, Rgg\mathrm{R}g_{\ast}g^{\ast}\mathcal{F} and also 𝒞\mathcal{C} have bounded coherent cohomology. Since 𝒴\mathcal{Y}^{\prime} is cohomologically proper over RR, we conclude (using Remark 2.4) that

𝖱Γ(𝒴,𝖱gg)𝖱Γ(𝒴,g)Dcoh+(R).\mathsf{R}\Gamma(\mathcal{Y},\mathsf{R}g_{\ast}g^{\ast}\mathcal{F})\simeq\mathsf{R}\Gamma(\mathcal{Y}^{\prime},g^{\ast}\mathcal{F})\in D^{+}_{\operatorname{coh}}(R).

Since g:𝒴𝒴g\colon\mathcal{Y}^{\prime}\to\mathcal{Y} is an isomorphism over 𝒰𝒴\mathcal{U}\subset\mathcal{Y}, 𝒞|U0\mathcal{C}|_{U}\simeq 0. Therefore in order to show that 𝖱Γ(𝒴,)Dcoh+(R)\mathsf{R}\Gamma(\mathcal{Y},\mathcal{F})\in D^{+}_{\operatorname{coh}}(R), it suffices to show that 𝖱Γ(𝒴,𝒞)Dcoh+(R)\mathsf{R}\Gamma(\mathcal{Y},\mathcal{C})\in D^{+}_{\operatorname{coh}}(R). By using the spectral sequence 𝖱iπj(𝒞)i+j(𝖱π𝒞)\mathsf{R}^{i}\pi_{\ast}\mathscr{H}^{j}(\mathcal{C})\Rightarrow\mathscr{H}^{i+j}(\mathsf{R}\pi_{\ast}\mathcal{C}) as in Remark 2.4, it suffices to show that if \mathcal{E} is a coherent sheaf on 𝒴\mathcal{Y} such that |U=0\mathcal{E}|_{U}=0, then 𝖱Γ(𝒴,)Dcoh+(R)\mathsf{R}\Gamma(\mathcal{Y},\mathcal{E})\in D^{+}_{\operatorname{coh}}(R). But in this case \mathcal{E} is supported on a proper closed substack 𝒵𝒳\mathcal{Z}\subsetneq\mathcal{X} and the statement follows from the dévissage hypothesis.

For (2) and (3), we also establish the notation that the closed substacks 𝒴0𝒴\mathcal{Y}^{\prime}_{0}\subset\mathcal{Y}^{\prime} and 𝒳0𝒳\mathcal{X}^{\prime}_{0}\subset\mathcal{X}^{\prime} denote the preimages of 𝒴0\mathcal{Y}_{0}. For (2), since (𝒳/𝒴)n(𝒳/𝒴)n(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{n}\to(\mathcal{X}/\mathcal{Y})^{n} is universally cohomologically proper and ((𝒳/𝒴)n,(𝒳0/𝒴0)n)((\mathcal{X}/\mathcal{Y})^{n},(\mathcal{X}_{0}/\mathcal{Y}_{0})^{n}) satisfies formal functions for each n1n\geq 1, it follows from Lemma 5.2, Theorem 5.3(1) and Example 4.4 that ((𝒳/𝒴)n,(𝒳0/𝒴0)n)((\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{n},(\mathcal{X}^{\prime}_{0}/\mathcal{Y}^{\prime}_{0})^{n}) also satisfies formal functions.

Since f:𝒳𝒴f^{\prime}\colon\mathcal{X}^{\prime}\to\mathcal{Y}^{\prime} is faithfully flat, cohomological descent implies that for each coherent sheaf \mathcal{M}^{\prime} on 𝒴\mathcal{Y}^{\prime}, we have a morphism of convergent spectral sequences:

Hi((𝒳/𝒴)j,(𝒳/𝒴)j){\mathrm{H}^{i}((\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j},\mathcal{M}^{\prime}_{(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j}})}Hi+j(𝒴,){\mathrm{H}^{i+j}(\mathcal{Y}^{\prime},\mathcal{M}^{\prime})}Hi((𝒳/𝒴)j,^(𝒳/𝒴)j){\mathrm{H}^{i}((\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j\wedge{\hskip 2.84526pt}},\widehat{\mathcal{M}^{\prime}}_{(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j\wedge{\hskip 2.84526pt}}})}Hi+j(𝒴^,^),{\mathrm{H}^{i+j}(\widehat{\mathcal{Y}^{\prime}},\widehat{\mathcal{M}^{\prime}}),}

where (𝒳/𝒴)j(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j\wedge{\hskip 2.84526pt}} is the completion of (𝒳/𝒴)j(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{j} along (𝒳0/𝒴0)j(\mathcal{X}^{\prime}_{0}/\mathcal{Y}^{\prime}_{0})^{j} and 𝒴^\widehat{\mathcal{Y}^{\prime}} is the completion of 𝒴\mathcal{Y}^{\prime} along 𝒴0\mathcal{Y}^{\prime}_{0}. Since ((𝒳/𝒴)n,(𝒳0/𝒴0)n)((\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{n},(\mathcal{X}^{\prime}_{0}/\mathcal{Y}^{\prime}_{0})^{n}) satisfies formal functions, so does (𝒴,𝒴0)(\mathcal{Y}^{\prime},\mathcal{Y}^{\prime}_{0}).

For a coherent sheaf \mathcal{F} on 𝒴\mathcal{Y}, we again consider the exact triangle:

𝖱gg𝒞[1],\mathcal{F}\to\mathsf{R}g_{\ast}g^{\ast}\mathcal{F}\to\mathcal{C}\to\mathcal{F}[1],

which induces a morphism of exact triangles

𝖱Γ(𝒴,){\mathsf{R}\Gamma(\mathcal{Y},\mathcal{F})}𝖱Γ(𝒴,𝖱gg){\mathsf{R}\Gamma(\mathcal{Y},\mathsf{R}g_{\ast}g^{\ast}\mathcal{F})}𝖱Γ(𝒴,𝒞){\mathsf{R}\Gamma(\mathcal{Y},\mathcal{C})}𝖱Γ(𝒴,)[1]{\mathsf{R}\Gamma(\mathcal{Y},\mathcal{F})[1]}𝖱Γ(𝒴^,^){\mathsf{R}\Gamma(\widehat{\mathcal{Y}},\widehat{\mathcal{F}})}𝖱Γ(𝒴^,(𝖱gg)){\mathsf{R}\Gamma(\widehat{\mathcal{Y}},(\mathsf{R}g_{\ast}g^{\ast}\mathcal{F})^{\wedge{\hskip 2.84526pt}})}𝖱Γ(𝒴^,𝒞^){\mathsf{R}\Gamma(\widehat{\mathcal{Y}},\widehat{\mathcal{C}})}𝖱Γ(𝒴^,^)[1].{\mathsf{R}\Gamma(\widehat{\mathcal{Y}},\widehat{\mathcal{F}})[1].}

Since (𝒴,𝒴0)(\mathcal{Y}^{\prime},\mathcal{Y}^{\prime}_{0}) satisfies formal functions, the second vertical map is an isomorphism. Since 𝒞|U0\mathcal{C}|_{U}\simeq 0, the third vertical map is also a isomorphism. Indeed, we can reduce as above to the case that 𝒞\mathcal{C} is a coherent sheaf supported in cohomological degree 0 in which case 𝒞\mathcal{C} is supported on a proper closed substack 𝒵𝒴\mathcal{Z}\subsetneq\mathcal{Y} and we may apply the dévissage hypothesis. Thus 𝖱Γ(𝒴,)𝖱Γ(𝒴^,^)\mathsf{R}\Gamma(\mathcal{Y},\mathcal{F})\to\mathsf{R}\Gamma(\widehat{\mathcal{Y}},\widehat{\mathcal{F}}) is a an isomorphism and (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}_{0}) satisfies formal functions.

For (3), let 𝔊\mathfrak{G} be a coherent sheaf on 𝒴^\widehat{\mathcal{Y}}. By coherent completeness of (𝒳,𝒳0)(\mathcal{X},\mathcal{X}_{0}), there is a coherent sheaf \mathcal{F} on 𝒳\mathcal{X} and an isomorphism ^f^𝔊\widehat{\mathcal{F}}\cong\widehat{f}^{*}\mathfrak{G}. Denoting by \mathcal{F}^{\prime} the pullback gg^{\prime*}\mathcal{F}, we have that completion ^\widehat{\mathcal{F}^{\prime}} of \mathcal{F}^{\prime} along 𝒳0\mathcal{X}^{\prime}_{0} is identified with g^f^𝔊f^g^𝔊\widehat{g^{\prime}}^{*}\widehat{f}^{*}\mathfrak{G}\cong\widehat{f^{\prime}}^{*}\widehat{g}^{\,*}\mathfrak{G}. Letting p1p_{1} and p2p_{2} denote the two projections 𝒳×𝒴𝒳\mathcal{X}^{\prime}\times_{\mathcal{Y}^{\prime}}\mathcal{X}^{\prime}, then because ^\widehat{\mathcal{F}^{\prime}} is the pullback of a coherent sheaf on 𝒴^\widehat{\mathcal{Y}^{\prime}} we have an isomorphism

ϑ:p1^p2^\vartheta:\widehat{p^{*}_{1}\mathcal{F}^{\prime}}\stackrel{{\scriptstyle\sim}}{{\to}}\widehat{p^{*}_{2}\mathcal{F}^{\prime}}

of coherent sheaves on the completion of (𝒳/𝒴)2(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{2} along the preimage of 𝒴0\mathcal{Y}_{0} satisfying the cocycle condition on the completion of (𝒳/𝒴)3(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{3} along the preimage of 𝒴0\mathcal{Y}_{0}. As the morphisms (𝒳/𝒴)n(𝒳/𝒴)n(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{n}\to(\mathcal{X}/\mathcal{Y})^{n} are universally cohomologically proper for n1n\geq 1, ((𝒳/𝒴)n,(𝒳0/𝒴0)n)((\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{n},(\mathcal{X}^{\prime}_{0}/\mathcal{Y}^{\prime}_{0})^{n}) satisfy formal functions. It follows that there is a unique isomorphism θ:p1p2\theta\colon p_{1}^{*}\mathcal{F}^{\prime}\stackrel{{\scriptstyle\sim}}{{\to}}p_{2}^{*}\mathcal{F}^{\prime} of coherent sheaves on (𝒳/𝒴)2(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{2} satisfying the cocycle condition on (𝒳/𝒴)3(\mathcal{X}^{\prime}/\mathcal{Y}^{\prime})^{3} such that θ^=ϑ\widehat{\theta}=\vartheta. By faithfully flat descent, there is a coherent sheaf 𝒢\mathcal{G}^{\prime} on 𝒴\mathcal{Y}^{\prime} such that there is an isomorphism f𝒢f^{\prime*}\mathcal{G}^{\prime}\cong\mathcal{F}^{\prime} compatible with θ\theta. Since the coherent sheaves 𝒢^\widehat{\mathcal{G}^{\prime}} and g^𝔊\widehat{g}^{*}\mathfrak{G} are described via the same formal descent data with respect to 𝒳^𝒴^\widehat{\mathcal{X}^{\prime}}\to\widehat{\mathcal{Y}^{\prime}}, there is an isomorphism 𝒢^g^𝔊\widehat{\mathcal{G}^{\prime}}\cong\widehat{g}^{*}\mathfrak{G}. Theorem 5.3(2) now gives equivalences g^g^𝔊g^𝒢^g𝒢^\widehat{g}_{*}\widehat{g}^{*}\mathfrak{G}\cong\widehat{g}_{*}\widehat{\mathcal{G}^{\prime}}\cong\widehat{g_{*}\mathcal{G}^{\prime}} and thus g^g^𝔊\widehat{g}_{*}\widehat{g}^{*}\mathfrak{G} is in the essential image of Coh(𝒴)Coh(𝒴^)\operatorname{Coh}(\mathcal{Y})\to\operatorname{Coh}(\widehat{\mathcal{Y}}). As g:𝒴𝒴g\colon\mathcal{Y}^{\prime}\to\mathcal{Y} is an isomorphism over 𝒰\mathcal{U}, the adjunction morphism

𝔊g^g^𝔊\mathfrak{G}\to\widehat{g}_{*}\widehat{g}^{*}\mathfrak{G}

is an isomorphism over 𝒰\mathcal{U}. We claim that the kernel and cokernel of this adjunction morphism are annihilated by some power of sheaf ideals 𝒥𝒪𝒴^\mathcal{J}\subset\mathcal{O}_{\widehat{\mathcal{Y}}} defining the reduced complement of 𝒰\mathcal{U}. Indeed, this can be checked on a smooth presentation SpecB𝒴\operatorname{Spec}B\to\mathcal{Y}. Let IBI\subset B be the ideal defined by the preimage of 𝒴0\mathcal{Y}_{0} in SpecB\operatorname{Spec}B, let B^\widehat{B} be the II-adic completion of BB, and let 𝒴B=𝒴×𝒴SpecB\mathcal{Y}^{\prime}_{B}=\mathcal{Y}^{\prime}\times_{\mathcal{Y}}\operatorname{Spec}B be endowed with the (proper) projection morphism gB:𝒴BSpecBg_{B}\colon\mathcal{Y}^{\prime}_{B}\to\operatorname{Spec}B. The module of sections of 𝔊\mathfrak{G} over SpecB𝒴\operatorname{Spec}B\to\mathcal{Y} is a finite B^\widehat{B}-module NN while the module of sections of g^g^𝔊\widehat{g}_{*}\widehat{g}^{*}\mathfrak{G} over SpecB𝒴\operatorname{Spec}B\to\mathcal{Y} is H0(𝒴^B,gB^N)\mathrm{H}^{0}(\widehat{\mathcal{Y}^{\prime}}_{B},\widehat{g_{B}}^{*}N), where gB^\widehat{g_{B}} is the completion of gg along the preimage of 𝒴0\mathcal{Y}_{0}. Moreover, the adjunction morphism 𝔊g^g^𝔊\mathfrak{G}\to\widehat{g}_{*}\widehat{g}^{*}\mathfrak{G} corresponds to the morphism

NH0(𝒴^B,gB^N)H0(𝒴B^,gB^N)N\to\mathrm{H}^{0}(\widehat{\mathcal{Y}^{\prime}}_{B},\widehat{g_{B}}^{*}N)\simeq\mathrm{H}^{0}(\mathcal{Y}^{\prime}_{\widehat{B}},g_{\widehat{B}}^{*}N)

and both the kernel and cokernel are annihilated by some power of the ideal of sections of 𝒥\mathcal{J} over SpecB𝒴\operatorname{Spec}B\to\mathcal{Y} because 𝒴B^SpecB^\mathcal{Y}^{\prime}_{\widehat{B}}\to\operatorname{Spec}\widehat{B} is an isomorphism over the open subset SpecB^Spec(B^/𝒥(SpecB)B^)\operatorname{Spec}\widehat{B}-\operatorname{Spec}(\widehat{B}/\mathscr{J}(\operatorname{Spec}B)\widehat{B}). We know that the essential image of Coh(𝒴)Coh(𝒴^)\operatorname{Coh}(\mathcal{Y})\to\operatorname{Coh}(\widehat{\mathcal{Y}}) is stable under kernels, cokernels and extensions. Since g^g^𝔊\widehat{g}_{*}\widehat{g}^{*}\mathfrak{G} is in the essential image and both the kernel and cokernel of 𝔊g^g^𝔊\mathfrak{G}\to\widehat{g}_{*}\widehat{g}^{*}\mathfrak{G} are in the essential image by the dévissage hypothesis, we conclude that 𝔊\mathfrak{G} is in the essential image. ∎

Corollary 6.3.

Let f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} be a universally submersive and representable morphism of noetherian algebraic stacks of finite type over a noetherian II-adically complete ring RR. Let 𝒴0𝒴\mathcal{Y}_{0}\subset\mathcal{Y} be a closed substack and 𝒳0𝒳\mathcal{X}_{0}\subset\mathcal{X} be its preimage. Assume that Γ(𝒴,𝒪𝒴)=R\Gamma(\mathcal{Y},\mathcal{O}_{\mathcal{Y}})=R and that Γ(𝒴0,𝒪𝒴0)=R/I\Gamma(\mathcal{Y}_{0},\mathcal{O}_{\mathcal{Y}_{0}})=R/I. Suppose that for each n1n\geq 1, (𝒳/𝒴)n(\mathcal{X}/\mathcal{Y})^{n} is cohomologically proper over SpecR\operatorname{Spec}R and that there is a closed substack 𝒵n(𝒳/𝒴)n\mathcal{Z}_{n}\subset(\mathcal{X}/\mathcal{Y})^{n} such that |𝒵n||(𝒳0/𝒴0)n)||\mathcal{Z}_{n}|\subset|(\mathcal{X}_{0}/\mathcal{Y}_{0})^{n})| and such that the pair ((𝒳/𝒴)n,𝒵n)((\mathcal{X}/\mathcal{Y})^{n},\mathcal{Z}_{n}) is coherently complete and satisfies formal functions. Then 𝒴\mathcal{Y} is cohomologically proper over SpecR\operatorname{Spec}R and the pair (𝒴,𝒴0)(\mathcal{Y},\mathcal{Y}_{0}) is coherently complete and satisfies formal functions.

If f:𝒳𝒴f\colon\mathcal{X}\to\mathcal{Y} is surjective, proper, and representable, then in each case one only needs to require the properties for n=1n=1.

Proof.

The corollary follows directly from Theorem 6.1 using Lemma 5.6 to deduce that the pair ((𝒳/𝒴)n,(𝒳0/𝒴0)n)((\mathcal{X}/\mathcal{Y})^{n},(\mathcal{X}_{0}/\mathcal{Y}_{0})^{n}) is coherently complete and satisfies formal functions. ∎

7. 𝐆m\mathbf{G}_{m}-actions and destabilizing one-parameter subgroups

After reviewing properties of 𝐆m\mathbf{G}_{m}-actions and one parameter subgroups, we establish a refinement of the stabilization theorem in GIT (sometimes called the Hilbert–Mumford criterion) establishing the existence of destabilizing one-parameter subgroups that are regular in the stabilizer of the limit (Proposition 7.14).

7.1. The fixed and attractor subschemes

Let GG be a connected, smooth, and affine group scheme of finite type over a noetherian ring RR. Let X=SpecAX=\operatorname{Spec}A be an affine scheme over RR with an action of GG. For any one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G, we define the subfunctors

Xλ0\displaystyle X^{0}_{\lambda} \displaystyle\coloneqq Hom¯R𝐆m(SpecR,X)(fixed locus)\displaystyle\underline{\operatorname{Hom}}_{R}^{\mathbf{G}_{m}}(\operatorname{Spec}R,X)\hskip 12.80365pt\text{(fixed locus)}
Xλ+\displaystyle X^{+}_{\lambda} \displaystyle\coloneqq Hom¯R𝐆m(𝐀1,X)(attractor)\displaystyle\underline{\operatorname{Hom}}_{R}^{\mathbf{G}_{m}}(\mathbf{A}^{1},X)\hskip 29.87547pt\text{(attractor)}

of XX as introduced in [Dri13]. These functors are represented by closed subschemes of XX. Indeed, the λ\lambda-action on XX induces a 𝐙\mathbf{Z}-grading A=nAnA=\bigoplus_{n}A_{n}. Let Iλ=n<0AnI^{-}_{\lambda}=\sum_{n<0}A_{n} and Iλ+=n>0AnI^{+}_{\lambda}=\sum_{n>0}A_{n} denote the ideals generated by homogeneous elements in strictly negative and strictly positive degree; then one can check that Xλ0X^{0}_{\lambda} and Xλ+X^{+}_{\lambda} are represented by the closed subschemes SpecA/(Iλ+Iλ+)\operatorname{Spec}A/(I^{-}_{\lambda}+I^{+}_{\lambda}) and SpecA/Iλ\operatorname{Spec}A/I^{-}_{\lambda}, respectively. If kk is an algebraically closed field over RR, then the kk-points of Xλ0X^{0}_{\lambda} are the 𝐆m\mathbf{G}_{m}-fixed points the kk-points of Xλ+X^{+}_{\lambda} can be described as

Xλ+(k)={xX(k)|limt0λ(t)xexists}.X^{+}_{\lambda}(k)=\{x\in X(k)\,\,|\,\,\lim_{t\to 0}\lambda(t)\cdot x\,\,\text{exists}\}.

The inclusion Xλ0Xλ+X^{0}_{\lambda}\hookrightarrow X^{+}_{\lambda} has a natural retraction given by

ev0:Xλ+Xλ0,xlimt0λ(t)x.\operatorname{ev}_{0}\colon X^{+}_{\lambda}\to X^{0}_{\lambda},\qquad x\mapsto\lim_{t\to 0}\lambda(t)\cdot x.
Remark 7.1.

If X=SpecAX=\operatorname{Spec}A is integral, then Xλ0X^{0}_{\lambda} and Xλ+X^{+}_{\lambda} may be non-reduced and reducible. For example, if 𝐆m\mathbf{G}_{m} acts on Speck[x,y,z,w]/(xwy2z2)\operatorname{Spec}k[x,y,z,w]/(xw-y^{2}z^{2}) with weights (1,0,0,1)(1,0,0,-1), then Xλ0=Speck[y,z]/(y2z2)X^{0}_{\lambda}=\operatorname{Spec}k[y,z]/(y^{2}z^{2}) and Xλ+=Speck[x,y,z]/(y2z2)X^{+}_{\lambda}=\operatorname{Spec}k[x,y,z]/(y^{2}z^{2}). If XX is integral, it can also happen that Xλ0(k)X^{0}_{\lambda}(k) is a finite set with more than one point; e.g., if X=Speck[x,y,z]/(x(x1)+yz)X=\operatorname{Spec}k[x,y,z]/(x(x-1)+yz) with weights (0,1,1)(0,1,-1), then Xλ0=Speck[x]/(x(x1))X^{0}_{\lambda}=\operatorname{Spec}k[x]/(x(x-1)).

7.2. Centralizer, parabolic and unipotent subgroups

In this subsection, we recall the dynamic approach to algebraic groups with respect to one-parameter subgroup as discussed in [CGP15, Chapter 2.1] (also see [Con14, §§4-5]). Let GG be a connected, smooth, and affine group scheme of finite type over a noetherian ring RR. Let GG act on itself via conjugation. Then a one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G induces an action of 𝐆m\mathbf{G}_{m} on GG (again via conjugation). Consider the following subgroups of GG:

Zλ\displaystyle Z_{\lambda} ={gG|λ(t)gλ(t)1=g for all t𝐆m}\displaystyle=\{g\in G\,|\,\,\text{$\lambda(t)g\lambda(t)^{-1}=g$ for all $t\in\mathbf{G}_{m}$}\} (centralizer of λ),\displaystyle\text{(centralizer of $\lambda$)},
Pλ\displaystyle P_{\lambda} ={gG|limt0λ(t)gλ(t)1exists}\displaystyle=\{g\in G\,|\,\lim_{t\to 0}\lambda(t)g\lambda(t)^{-1}\,\,\text{exists}\} (parabolic of λ\lambda),
Uλ\displaystyle U_{\lambda} ={gG|limt0λ(t)gλ(t)1=1}\displaystyle=\{g\in G\,|\,\lim_{t\to 0}\lambda(t)g\lambda(t)^{-1}=1\} (unipotent of λ),\displaystyle\text{(unipotent of $\lambda$)},\

which extend naturally to subgroup functors of GG. When it is necessary to record the group GG, we write ZλGZ^{G}_{\lambda}, PλGP^{G}_{\lambda}, and UλGU^{G}_{\lambda}. These functors are represented by closed subgroup schemes of GG. Indeed, observe that with the notation introduced in §7.1, we have that Zλ=Gλ0Z_{\lambda}=G^{0}_{\lambda} and Pλ=Gλ+P_{\lambda}=G^{+}_{\lambda}, and that UλU_{\lambda} is identified with the kernel of Pλ=Gλ+Gλ0=ZλP_{\lambda}=G^{+}_{\lambda}\to G^{0}_{\lambda}=Z_{\lambda} defined by glimt0λ(t)gλ(t)1g\mapsto\lim_{t\to 0}\lambda(t)g\lambda(t)^{-1}. See also [CGP15, Lems. 2.1.4 and 2.1.5].

There is a split exact sequence

(7.1) 1UλPλZλ1.1\to U_{\lambda}\to P_{\lambda}\to Z_{\lambda}\to 1.

Over a field, the groups ZλZ_{\lambda}, PλP_{\lambda} and UλU_{\lambda} are well-known to satisfy several nice properties. If GG is smooth (resp. connected), so is Zλ,PλZ_{\lambda},P_{\lambda} and UλU_{\lambda}. The group UλU_{\lambda} is unipotent. If GG is connected and reductive then so is ZλZ_{\lambda} [Con20, Thm. C.2.1]. Moreover, in this case, UλU_{\lambda} is the unipotent radical of PλP_{\lambda} and G/PλG/P_{\lambda} is projective [CGP15, Prop. 2.2.9]. In general, if GG is a split reductive group scheme over a noetherian ring RR and λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G is a one-parameter subgroup, then ZλZ_{\lambda}, PλP_{\lambda}, and UλU_{\lambda} are closed subgroup schemes of GG smooth over RR, G/PλG/P_{\lambda} is projective over RR and ZλZ_{\lambda} is reductive [Con14, Thm. 4.1.7, Ex. 4.1.9, Cor. 5.2.8].

If GG acts on an affine scheme XX of finite type over RR, there are natural actions of ZλZ_{\lambda} and PλP_{\lambda} on Xλ0X^{0}_{\lambda} and Xλ+X^{+}_{\lambda} respectively. The evaluation map ev0:Xλ+Xλ0\operatorname{ev}_{0}\colon X^{+}_{\lambda}\to X^{0}_{\lambda} is equivariant with respect to PλZλP_{\lambda}\to Z_{\lambda}, and induces a morphism ev0:[Xλ+/Pλ][Xλ0/Zλ]\operatorname{ev}_{0}\colon[X^{+}_{\lambda}/P_{\lambda}]\to[X^{0}_{\lambda}/Z_{\lambda}] on quotient stacks.

7.3. Positively graded actions

Here we consider a distinguished class of actions that appear frequently in practice (see Example 7.4).

Definition 7.2.

Recall that an action of 𝐆m\mathbf{G}_{m} on an affine scheme SpecA\operatorname{Spec}A over a noetherian ring RR is given by a grading A=dAdA=\bigoplus_{d}A_{d}. The action is positively graded (resp. semipositively graded) if Ad=0A_{d}=0 for d<0d<0 and A0A_{0} is finite over RR (resp., if Ad=0A_{d}=0 for d<0d<0). A representation V=dVdV=\bigoplus_{d}V_{d} of 𝐆m\mathbf{G}_{m} is positively graded (resp., semipositively graded) if Vd=0V_{d}=0 for d0d\leq 0 (resp., if Vd=0V_{d}=0 for d<0d<0).

For the action of an algebraic group GG on an affine scheme SpecA\operatorname{Spec}A over RR and a one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G, we say that SpecA\operatorname{Spec}A is λ\lambda-positive (resp. λ\lambda-semipositive) if the 𝐆m\mathbf{G}_{m}-action on SpecA\operatorname{Spec}A induced by λ\lambda is positively graded (resp. semipositively graded).

Lemma 7.3.

Let 𝐆m\mathbf{G}_{m} act on an affine scheme X=SpecAX=\operatorname{Spec}A of finite type over a field kk.

  1. (1)

    The action is semipositively graded if and only if Xid+=XX^{+}_{\operatorname{id}}=X.

  2. (2)

    The action is positively graded if and only if X𝐆mX^{\mathbf{G}_{m}} is finite over Speck\operatorname{Spec}k.

Proof.

The first part is clear from the definitions. For the second, observe that the condition that Ad=0A_{d}=0 for d<0d<0 implies that X𝐆m=Spec(A/I+)=SpecA0X^{\mathbf{G}_{m}}=\operatorname{Spec}(A/I^{+})=\operatorname{Spec}A_{0}. ∎

The following is our key example.

Example 7.4.

If GG is an algebraic group acting on an affine scheme X=SpecAX=\operatorname{Spec}A of finite type over kk and λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G is a one-parameter subgroup, then Xλ+X^{+}_{\lambda} is always λ\lambda-semipositive. Moreover, Xλ+X^{+}_{\lambda} is λ\lambda-positive if and only if Xλ0X^{0}_{\lambda} is finite over Speck\operatorname{Spec}k.

Lemma 7.5.

Let GG be an affine group scheme of finite type over a noetherian ring RR. Let SpecA\operatorname{Spec}A and SpecB\operatorname{Spec}B be affine schemes of finite type over RR with actions of GG. Let λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G be a one-parameter subgroup. If AA is λ\lambda-positive with Aλ=RA^{\lambda}=R and BB is λ\lambda-positive (resp. λ\lambda-semipositive), then ARBA\otimes_{R}B is λ\lambda-positive (resp. λ\lambda-semipositive) and (ARB)G=BG(A\otimes_{R}B)^{G}=B^{G}.

Proof.

Let A=d0AdA=\bigoplus_{d\geq 0}A_{d} and B=d0BdB=\bigoplus_{d\geq 0}B_{d} be the gradings induced by the λ\lambda-action. Then (ARB)d=i+j=dAiBj(A\otimes_{R}B)_{d}=\bigoplus_{i+j=d}A_{i}\otimes B_{j} is λ\lambda-semipositive with

(ARB)λ=AλRBλ=Bλ.(A\otimes_{R}B)^{\lambda}=A^{\lambda}\otimes_{R}B^{\lambda}=B^{\lambda}.

It follows that (ARB)G=BG(A\otimes_{R}B)^{G}=B^{G}. ∎

Example 7.6.

Let GG be a reductive algebraic group over an algebraically closed field kk. If λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G is a one-parameter subgroup, then the λ\lambda-grading on Γ(Uλ)=Γ(Uλ,𝒪Uλ)\Gamma(U_{\lambda})=\Gamma(U_{\lambda},\mathcal{O}_{U_{\lambda}}) is positively graded and Γ(Uλ)λ=k\Gamma(U_{\lambda})^{\lambda}=k. To see this, choose a maximal torus TT containing λ\lambda. Since the root system Φ(G,T)\Phi(G,T) is reduced [Con20, Cor. 2.2.1], there is a decomposition as schemes UλaUaU_{\lambda}\simeq\prod_{a}U_{a} [CGP15, Cor. 3.3.12], where the product is over all root groups UaU_{a} with aΦ(G,T)a\in\Phi(G,T) satisfying a,λ>0\langle a,\lambda\rangle>0. Each Ua𝐆aU_{a}\simeq\mathbf{G}_{a} as group schemes over kk, hence Γ(Uλ,𝒪Uλ)\Gamma(U_{\lambda},\mathcal{O}_{U_{\lambda}}) is a polynomial ring in the variables xax_{a} indexed by aΦ(G,T)a\in\Phi(G,T), where the λ\lambda-weight of xax_{a} is a,λ\langle a,\lambda\rangle [CGP15, Prop. 2.1.8]. Therefore, the λ\lambda-weight of a monomial xa1m1xakmkx_{a_{1}}^{m_{1}}\cdots x_{a_{k}}^{m_{k}} is strictly positive unless m1==mk=0m_{1}=\cdots=m_{k}=0. It then follows from Lemma 7.5 that the λ\lambda-grading on Γ(Uλ)n\Gamma(U_{\lambda})^{\otimes n} is also positively graded for n1n\geq 1 and that (Γ(Uλ)n)λ=k(\Gamma(U_{\lambda})^{\otimes n})^{\lambda}=k.

More generally, if GG is a split reductive group scheme over a noetherian ring RR and λ\lambda is a one-parameter subgroup contained in a maximal torus TGT\subset G, then Γ(Uλ)\Gamma(U_{\lambda}) is λ\lambda-positive with Γ(Uλ)λ=R\Gamma(U_{\lambda})^{\lambda}=R.

7.4. Regular one-parameter subgroups

If TT is a torus over a noetherian ring RR, there is a perfect bilinear pairing between the character lattice 𝕏(T)Hom(T,𝐆m)\mathbb{X}^{\ast}(T)\coloneqq\operatorname{Hom}(T,\mathbf{G}_{m}) and the lattice of one-parameter subgroups (or cocharacter lattice) 𝕏(T)Hom(𝐆m,T)\mathbb{X}_{\ast}(T)\coloneqq\operatorname{Hom}(\mathbf{G}_{m},T)

(7.2) ,:𝕏(T)×𝕏(T)End(𝐆m)𝐙\langle\cdot,\cdot\rangle\colon\mathbb{X}^{\ast}(T)\times\mathbb{X}_{\ast}(T)\to\operatorname{End}(\mathbf{G}_{m})\simeq\mathbf{Z}

given by a,λ=aλ\langle a,\lambda\rangle=a\circ\lambda.

Recall that if TT is a maximal torus of GG, the root system of the pair (G,T)(G,T) is the subset Φ(G,T)𝕏(T){0}\Phi(G,T)\subset\mathbb{X}^{\ast}(T)-\{0\} consisting of non-trivial weights for the adjoint action of TT on Lie(G)\operatorname{Lie}(G).

Definition 7.7.

A one-parameter subgroup λ:𝐆mT\lambda\colon\mathbf{G}_{m}\to T is regular with respect to TT if a,λ0\langle a,\lambda\rangle\neq 0 for all aΦ(G,T)a\in\Phi(G,T).

Remark 7.8.

When G=GLnG=\mathrm{GL}_{n} and TT is the diagonal torus in GG, then the Lie algebra Lie(G)=Matn,n{\rm Lie}(G)={\rm Mat}_{n,n} is the vector space of n×nn\times n matrices. The basis element Ei,jE_{i,j} with a 1 in position (i,j)(i,j) and 0 elsewhere has weight χiχj\chi_{i}-\chi_{j}, where χi:T𝐆m\chi_{i}\colon T\to\mathbf{G}_{m} denotes the character defined by (tk)ti(t_{k})\mapsto t_{i}. Therefore a one-parameter subgroup λ=(λk):𝐆mT\lambda=(\lambda_{k})\colon\mathbf{G}_{m}\to T is regular if and only if the λi\lambda_{i}’s are distinct.

Lemma 7.9.

Let GG be a connected, smooth, and affine group scheme over a noetherian ring RR. Let TGT\subset G be a maximal torus and λ\lambda a one-parameter subgroup of TT.

  1. (1)

    λ\lambda is regular with respect to TT if and only if CentG(T)=Zλ\operatorname{Cent}^{G}(T)=Z_{\lambda}.

  2. (2)

    If GG is reductive, then λ\lambda is regular with respect to TT if and only if ZλZ_{\lambda} is a maximal torus in GG.

  3. (3)

    If TGT^{\prime}\subset G is another maximal torus containing the image of λ\lambda, then λ\lambda is regular with respect to TT if and only the same is true with respect to TT^{\prime}. In other words, the definition of regularity is independent of the maximal torus TT chosen.

Proof.

Since GG is smooth connected, so too are CentG(T)\operatorname{Cent}^{G}(T) [Con14, Lem. 2.2.4] and ZλZ_{\lambda}. Therefore, the containment CentG(T)Zλ\operatorname{Cent}^{G}(T)\subset Z_{\lambda} is an equality if and only if it is true on geometric fibers. In particular, we may assume that RR is an algebraically closed field. In this case, it suffices to prove that Lie(CentG(T))=Lie(Zλ)\operatorname{Lie}(\operatorname{Cent}^{G}(T))=\operatorname{Lie}(Z_{\lambda}). Now

Lie(G)0=Lie(G)T=Lie(CentG(T))Lie(Zλ)=Lie(G)λ=aΦ(G,T){0}:a,λ=0Lie(G)a.\operatorname{Lie}(G)_{0}=\operatorname{Lie}(G)^{T}=\operatorname{Lie}(\operatorname{Cent}^{G}(T))\subset\operatorname{Lie}(Z_{\lambda})=\operatorname{Lie}(G)^{\lambda}=\oplus_{a\in\Phi(G,T)\cup\{0\}\,:\,\langle a,\lambda\rangle=0}\operatorname{Lie}(G)_{a}.

This proves that λ\lambda is regular if and only if CentG(T)=Zλ\operatorname{Cent}^{G}(T)=Z_{\lambda}. Moreover if GG is reductive then CentG(T)=T\operatorname{Cent}^{G}(T)=T [Con20, Prop. 2.3.1], proving (1) and (2).

To prove (3), suppose that the image of λ\lambda is contained in some other maximal torus TT^{\prime}. By the preceding discussion, it suffices to show that dimCentG(T)=dimZλ\dim\operatorname{Cent}^{G}(T^{\prime})=\dim Z_{\lambda} when RR is an algebraically closed field. We have dimZλ=dimCentG(T)\dim Z_{\lambda}=\dim\operatorname{Cent}^{G}(T) since λ\lambda is regular with respect to TT. On the other hand, all maximal tori of GG are RR-conjugate, so dimCentG(T)=dimCentG(T)\dim\operatorname{Cent}^{G}(T^{\prime})=\dim\operatorname{Cent}^{G}(T) and hence λ\lambda is also regular with respect to TT^{\prime}. ∎

7.5. Destabilizing one-parameter subgroups

A degeneration from a non-closed orbit to a closed orbit can be realized by a one-parameter subgroup. This is the classical destabilization theorem in GIT, which implies the Hilbert–Mumford criterion [Mum65, p. 53].

Proposition 7.10.

Let GG be a reductive algebraic group over an algebraically closed field kk. Let XX be an affine scheme of finite type over kk with an action of GG. For any point xX(k)x\in X(k), there is a one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G such that limt0λ(t)x\lim_{t\to 0}\lambda(t)\cdot x exists and has closed orbit.

When XX is integral and has a unique closed orbit, it is possible to find a one-parameter subgroup destabilizing every orbit.

Proposition 7.11.

Let GG be a reductive algebraic group over an algebraically closed field kk. Let X=SpecAX=\operatorname{Spec}A be an integral affine scheme of finite type over kk with an action of GG. Assume that there is a unique closed orbit GG-orbit Gx0XGx_{0}\subset X. Then there exists a one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G such that for any point xX(k)x\in X(k), there exists gG(k)g\in G(k) such that limt0λ(t)gxGx0\lim_{t\to 0}\lambda(t)\cdot gx\in Gx_{0}. In particular, GXλ+=XG\cdot X^{+}_{\lambda}=X and Xλ0Gx0\emptyset\neq X^{0}_{\lambda}\subset Gx_{0}.

Proof.

Let K=K(X)K=K(X) be the function field of XX and let ξX(K)\xi\in X(K) be the generic point. Letting KK¯K\to\overline{K} be an algebraic closure, we consider the base change XkK¯X\otimes_{k}\overline{K} with the point ξ¯(XkK¯)(K¯)\overline{\xi}\in(X\otimes_{k}\overline{K})(\overline{K}) induced from ξ\xi. By Proposition 7.10, there is a one-parameter subgroup λ¯:𝐆m,K¯GK¯\overline{\lambda}\colon\mathbf{G}_{m,\overline{K}}\to G_{\overline{K}} such that limt0λ¯(t)ξ¯\lim_{t\to 0}\overline{\lambda}(t)\cdot\overline{\xi} exists and has closed GK¯G_{\overline{K}}-orbit. If TGT\subset G is a maximal torus, then there exists an element gG(K¯)g\in G(\overline{K}) such that gλ¯g1TK¯g\overline{\lambda}g^{-1}\in T_{\overline{K}}. Since extension of scalars gives an isomorphism

𝐙nHomk(𝐆m,T)HomK¯(𝐆m,K¯,TK¯)𝐙n,\mathbf{Z}^{n}\cong\operatorname{Hom}_{k}(\mathbf{G}_{m},T)\stackrel{{\scriptstyle\sim}}{{\to}}\operatorname{Hom}_{\overline{K}}(\mathbf{G}_{m,\overline{K}},T_{\overline{K}})\cong\mathbf{Z}^{n},

there is a one-parameter subgroup λ:𝐆mT\lambda\colon\mathbf{G}_{m}\to T with λK¯=gλ¯g1\lambda_{\overline{K}}=g\overline{\lambda}g^{-1}. Since the generic point ξ[X/G](K)\xi\in[X/G](K) is in the image of the proper map [Xλ+/Pλ][X/G][X^{+}_{\lambda}/P_{\lambda}]\to[X/G], this map is surjective. Moreover, every point x[X/G](k)x\in[X/G](k) has a representative in X(k)X(k) which specializes via λ\lambda to the unique closed orbit.

Alternatively, we can appeal to the Kempf–Hesselink stratification. Kempf proved in [Kem78, Thm. 3.4] that for any point xX(k)x\in X(k), there is a destabilizing one-parameter subgroup λx:𝐆mG\lambda_{x}\colon\mathbf{G}_{m}\to G (unique up to conjugation by the unipotent radical UλxU_{\lambda_{x}}) with limt0λx(t)xGx0\lim_{t\to 0}\lambda_{x}(t)x\in Gx_{0} minimizing the normalized Hilbert–Mumford index μ(x,)/||||\mu(x,-)/||-||, where ||||||-|| is a fixed conjugation-invariant norm on 𝕏(G)\mathbb{X}_{*}(G). In [Hes79], Hesselink used Kempf’s optimal destabilizing one-parameter subgroup to show that XGx0=Sλi,MiX\setminus Gx_{0}=\coprod S_{\lambda_{i},M_{i}} admits a stratification into locally closed GG-invariant subschemes where each λi\lambda_{i} is a one-parameter subgroup and Mi<0M_{i}<0. For any kk-point xSλi,Mix\in S_{\lambda_{i},M_{i}} there is an element gG(k)g\in G(k) such that λi\lambda_{i} is an optimal one-parameter subgroup for gxgx with μ(gx,λi)/λi=Mi\mu(gx,\lambda_{i})/||\lambda_{i}||=M_{i}. Since XX is integral, there is a generic strata Sλ,MS_{\lambda,M} for some (λ,M){(λi,Mi)}(\lambda,M)\in\{(\lambda_{i},M_{i})\}, which is dense. Since [Xλ+/Pλ][X/G][X^{+}_{\lambda}/P_{\lambda}]\to[X/G] is proper and its image contains the dense set [Sλ,M/G][S_{\lambda,M}/G], it must be surjective. The statement follows. ∎

Example 7.12.

Under the SLn\mathrm{SL}_{n} action on 𝐀n\mathbf{A}^{n} given by the standard representation, the origin is the unique closed orbit and the complement 𝐀n0\mathbf{A}^{n}\setminus 0 is a dense orbit. Kempf’s optimal one-parameter subgroup for (1,0,,0)(1,0,\ldots,0) is λ(t)=diag(tn1,t1,,t1)\lambda(t)=\mathrm{diag}(t^{n-1},t^{-1},\ldots,t^{-1}). Observe that this subgroup is unfortunately not regular, which is a property we desire for our application. However, nearby deformations are regular: for distinct positive integers d1,,dn1d_{1},\ldots,d_{n-1} with sum dd, the one-parameter subgroup λ(t)=diag(td,td1,,tdn1)\lambda^{\prime}(t)=\mathrm{diag}(t^{d},t^{-d_{1}},\ldots,t^{-d_{n-1}}) is regular and also destabilizes the generic orbit.

In fact, one can always find destabilizing one-parameter subgroups λ\lambda of a point xx such that λ\lambda is regular in the stabilizer of the limit x0=limt0λ(t)xx_{0}=\lim_{t\to 0}\lambda(t)\cdot x. This is proven in Proposition 7.14, which relies on the following lemma.

Lemma 7.13.

Let GG be a reductive algebraic group over an algebraically closed field kk. Let XX be an affine scheme of finite type over kk with an action of GG and let xX(k)x\in X(k). Let λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G be a one-parameter subgroup such that x0:=limt0λ(t)xx_{0}:=\lim_{t\to 0}\lambda(t)x has closed orbit. If ρ:𝐆mGx0\rho\colon\mathbf{G}_{m}\to G_{x_{0}} is a one-parameter subgroup commuting with λ\lambda (e.g., ρ\rho and λ\lambda are contained in the same maximal torus of Gx0G_{x_{0}}), then for n0n\gg 0

limt0(λnρ)(t)xGx0.\lim_{t\to 0}(\lambda^{n}\rho)(t)\cdot x\in Gx_{0}.
Proof.

We first determine the pushout of the diagram:

(7.3)
B𝐆m[𝐀1/𝐆m][𝐀1/𝐆m]×B𝐆m,
\begin{split}\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 37.73526pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\\crcr}}}\ignorespaces{\hbox{\kern-13.42545pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{B\mathbf{G}_{m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 13.42546pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@hook{1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 61.73526pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 0.0pt\raise-4.20555pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@hook{1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 0.0pt\raise-29.5pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 61.73526pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[\mathbf{A}^{1}/\mathbf{G}_{m}]}$}}}}}}}{\hbox{\kern-37.73526pt\raise-41.13776pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{[\mathbf{A}^{1}/\mathbf{G}_{m}]\times B\mathbf{G}_{m},}$}}}}}}}\ignorespaces}}}}\ignorespaces\end{split}

where the top map B𝐆m[𝐀1/𝐆m]B\mathbf{G}_{m}\hookrightarrow[\mathbf{A}^{1}/\mathbf{G}_{m}] is the closed immersion defining the origin and the left map B𝐆m[𝐀1/𝐆m]×B𝐆mB\mathbf{G}_{m}\hookrightarrow[\mathbf{A}^{1}/\mathbf{G}_{m}]\times B\mathbf{G}_{m} is the open immersion corresponding to the product of B𝐆mB\mathbf{G}_{m} and the inclusion of the open point 1:Speck[𝐀1/𝐆m]1\colon\operatorname{Spec}k\hookrightarrow[\mathbf{A}^{1}/\mathbf{G}_{m}]. Expressing the diagram 𝐆m2\mathbf{G}_{m}^{2}-equivariantly, we can write B𝐆m=[Speck[y]y/𝐆m2]B\mathbf{G}_{m}=[\operatorname{Spec}k[y]_{y}/\mathbf{G}_{m}^{2}], [𝐀1/𝐆m]×B𝐆m=[Speck[y]/𝐆m2][\mathbf{A}^{1}/\mathbf{G}_{m}]\times B\mathbf{G}_{m}=[\operatorname{Spec}k[y]/\mathbf{G}_{m}^{2}], and [𝐀1/𝐆m]=[Speck[y,z]y/𝐆m2][\mathbf{A}^{1}/\mathbf{G}_{m}]=[\operatorname{Spec}k[y,z]_{y}/\mathbf{G}_{m}^{2}] under the diagonal action (t1,t2)(y,z)=(t1y,t2z)(t_{1},t_{2})\cdot(y,z)=(t_{1}y,t_{2}z). The 𝐆m2\mathbf{G}_{m}^{2}-equivariant pushout is determined by the fiber product of rings

k[y]y\textstyle{k[y]_{y}}k[y,z]y\textstyle{k[y,z]_{y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0z\scriptstyle{0\mapsfrom z}k[y]\textstyle{k[y]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k[y,zy,zy2,].\textstyle{k[y,\frac{z}{y},\frac{z}{y^{2}},\ldots].\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

We can write k[y,zy,zy2,]=nk[y,zn]k[y,\frac{z}{y},\frac{z}{y^{2}},\ldots]=\bigcup_{n}k[y,z_{n}] with zn=zynz_{n}=\frac{z}{y^{n}}, where 𝐆m2\mathbf{G}_{m}^{2} acts via (t1,t2)(y,zn)=(t1y,t2t1nzn)(t_{1},t_{2})\cdot(y,z_{n})=(t_{1}y,\frac{t_{2}}{t_{1}^{n}}z_{n}). Identifying 𝐀2=Speck[y,zn]\mathbf{A}^{2}=\operatorname{Spec}k[y,z_{n}], then 𝐆m2\mathbf{G}_{m}^{2} acts via the degree matrix Dn=(10n1)D_{n}=\begin{pmatrix}1&0\\ -n&1\end{pmatrix}. By generalities of pushouts of stacks (see [AHHR23, §4]), the pushout of (7.3) is a limit

limn[𝐀2/Dn𝐆m2]\varprojlim_{n}\,[\mathbf{A}^{2}/_{D_{n}}\mathbf{G}_{m}^{2}]

of quotient stacks with affine transition morphisms.

The point xX(k)x\in X(k) and one-parameter subgroup λ\lambda define a morphism λx:[𝐀1/𝐆m][X/G]\lambda_{x}\colon[\mathbf{A}^{1}/\mathbf{G}_{m}]\to[X/G] with λx(1)x\lambda_{x}(1)\simeq x and λx(0)x0\lambda_{x}(0)\simeq x_{0}. Since ρ\rho commutes with λ\lambda, we also have a morphism (λx,ρ):[𝐀1/𝐆m]×B𝐆m[X/G](\lambda_{x},\rho)\colon[\mathbf{A}^{1}/\mathbf{G}_{m}]\times B\mathbf{G}_{m}\to[X/G] giving a commutative diagram of solid arrows

B𝐆m\textstyle{B\mathbf{G}_{m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}z=0\scriptstyle{z=0}y0\scriptstyle{y\neq 0}[𝐀1/𝐆m]\textstyle{[\mathbf{A}^{1}/\mathbf{G}_{m}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λx\scriptstyle{\lambda_{x}}[𝐀1/𝐆m]×B𝐆m\textstyle{[\mathbf{A}^{1}/\mathbf{G}_{m}]\times B\mathbf{G}_{m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(λx,ρ)\scriptstyle{(\lambda_{x},\rho)}[𝐀2/Dn𝐆m2]\textstyle{[\mathbf{A}^{2}/_{D_{n}}\mathbf{G}_{m}^{2}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ\scriptstyle{\Psi}[X/G].\textstyle{[X/G].}

Since [X/G][X/G] is of finite type, the induced morphism limn[𝐀2/Dn𝐆m2][X/G]\varprojlim_{n}[\mathbf{A}^{2}/_{D_{n}}\mathbf{G}_{m}^{2}]\to[X/G] factors through [𝐀2/Dn𝐆m2][\mathbf{A}^{2}/_{D_{n}}\mathbf{G}_{m}^{2}] for n0n\gg 0, and thus defines the dotted arrow Ψ:[𝐀2/Dn𝐆m2][X/G]\Psi\colon[\mathbf{A}^{2}/_{D_{n}}\mathbf{G}_{m}^{2}]\to[X/G].

Now consider the map [𝐀1/𝐆m][𝐀2/Dn𝐆m2][\mathbf{A}^{1}/\mathbf{G}_{m}]\to[\mathbf{A}^{2}/_{D_{n}}\mathbf{G}_{m}^{2}] induced by the diagonal 𝐀1𝐀2\mathbf{A}^{1}\to\mathbf{A}^{2} and the group homomorphism 𝐆m𝐆m2\mathbf{G}_{m}\to\mathbf{G}_{m}^{2} defined by t(t,tn+1)t\mapsto(t,t^{n+1}). The composition [𝐀1/𝐆m][𝐀2/Dn𝐆m2]Ψ[X/G][\mathbf{A}^{1}/\mathbf{G}_{m}]\to[\mathbf{A}^{2}/_{D_{n}}\mathbf{G}_{m}^{2}]\xrightarrow{\Psi}[X/G] defines a morphism such that 0 maps to x0x_{0} and such that the induced map 𝐆mGx0\mathbf{G}_{m}\to G_{x_{0}} on stabilizers is given by λn+1ρ\lambda^{n+1}\rho. The statement follows. ∎

Proposition 7.14.

Let GG be a reductive algebraic group over an algebraically closed field kk, and let XX be an affine scheme of finite type over kk with an action of GG. For any point xX(k)x\in X(k), there is a one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G such that x0:=limt0λ(t)xx_{0}:=\lim_{t\to 0}\lambda(t)\cdot x has closed orbit and such that the induced one-parameter subgroup λ:𝐆mGx0\lambda\colon\mathbf{G}_{m}\to G_{x_{0}} is regular.

Moreover, if there is a unique closed orbit Gx0Gx_{0} and XX is integral, then there is a one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G such that for any point yX(k)y\in X(k), there exists gG(k)g\in G(k) such that y:=limt0λ(t)gyGx0y^{\prime}:=\lim_{t\to 0}\lambda(t)\cdot gy\in Gx_{0} and such that the induced map λ:𝐆mGy\lambda\colon\mathbf{G}_{m}\to G_{y^{\prime}} is a regular one-parameter subgroup.

Proof.

Let λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G be a one-parameter subgroup such that x0:=limt0λ(t)xx_{0}:=\lim_{t\to 0}\lambda(t)\cdot x has closed orbit (Proposition 7.10). Let TGx0T\subset G_{x_{0}} be a maximal torus containing the image of λ\lambda. Choose a one-parameter subgroup ρ:𝐆mT\rho\colon\mathbf{G}_{m}\to T such that the composition λnρ:𝐆mTGx0\lambda^{n}\rho\colon\mathbf{G}_{m}\to T\to G_{x_{0}} is a regular one-parameter subgroup for n0n\gg 0. Lemma 7.13 implies that x0=limt0(λnρ)(t)xx_{0}=\lim_{t\to 0}(\lambda^{n}\rho)(t)\cdot x for n0n\gg 0.

The addendum follows from applying the main statement to the generic point ξX\xi\in X as in the proof of Proposition 7.11. There exist a one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G, a finite extension K=K(X)KK=K(X)\to K^{\prime}, and gG(K)g\in G(K^{\prime}) such that gξXλ+(K)g\xi^{\prime}\in X^{+}_{\lambda}(K^{\prime}) and ξ0:=limt0λ(t)gξ\xi^{\prime}_{0}:=\lim_{t\to 0}\lambda(t)\cdot g\xi^{\prime} has closed orbit in XKX_{K^{\prime}}, where ξX(K)\xi^{\prime}\in X(K^{\prime}) is the lift of ξX(K)\xi\in X(K). Choose a maximal torus TGT\subset G containing λ\lambda and a one-parameter subgroup ρ:𝐆mT\rho\colon\mathbf{G}_{m}\to T such that the base change of λnρ\lambda^{n}\rho to KK^{\prime} is regular in Gξ0G_{\xi^{\prime}_{0}} for n0n\gg 0. Lemma 7.13 implies that λnρ\lambda^{n}\rho also destabilizes gξg\xi^{\prime}. It follows that any GG-orbit of a kk-point in XX contains a kk-point yy such that y:=limt0(λnρ)(t)yGx0y^{\prime}:=\lim_{t\to 0}(\lambda^{n}\rho)(t)\cdot y\in Gx_{0}. Since the base change of Gy=StabG(y)G_{y^{\prime}}=\operatorname{Stab}^{G}(y^{\prime}) to KK^{\prime} is conjugate to (GK)ξ0=StabGK(ξ0)(G_{K^{\prime}})_{\xi^{\prime}_{0}}=\operatorname{Stab}^{G_{K^{\prime}}}(\xi^{\prime}_{0}), the one-parameter subgroup λnρ:𝐆mGy\lambda^{n}\rho\colon\mathbf{G}_{m}\to G_{y^{\prime}} is also regular. ∎

Remark 7.15.

One can also formulate this result as saying that the degeneration fan 𝒟eg([SpecA/G],x)\mathcal{D}eg([\operatorname{Spec}A/G],x), introduced in [Hal14], has dimension at least the reductive rank of Gx0G_{x_{0}}.

8. Proofs of coherent completeness

In this section, we prove the main coherent completeness theorem (Theorem 1.2) as well as the coherent completeness result for positively graded actions (Theorem 1.3).

8.1. Proof of the main theorem

Proof of Theorem 1.2(1).

Let 𝒳=[X/G]\mathcal{X}=[X/G]. By Theorem 6.1, we can replace 𝒳\mathcal{X} with 𝒳red\mathcal{X}_{{\operatorname{red}}} and RR with RredR_{{\operatorname{red}}}, and we can thus assume that R=kR=k is a field and 𝒳\mathcal{X} is reduced. If G0G^{0} denotes the connected component of the identity, then [X/G0][X/G][X/G^{0}]\to[X/G] is finite and étale. Applying Theorem 6.1 reduces us to the case that GG is connected. Letting [SpecA/G]=i𝒳i[\operatorname{Spec}A/G]=\coprod_{i}\mathcal{X}_{i} be the irreducible decomposition, then each 𝒳i\mathcal{X}_{i} is isomorphic to [SpecAi/G][\operatorname{Spec}A_{i}/G] with AiG=kA_{i}^{G}=k. The morphism i𝒳i[X/G]\coprod_{i}\mathcal{X}_{i}\to[X/G] is a proper and surjective, and by applying Theorem 6.1 again, we are further reduced to the case that AA is an integral domain.

Since [SpecA/G][\operatorname{Spec}A/G] has a unique closed point, Proposition 7.14 and limit methods imply that there is a one-parameter subgroup λ:𝐆m,kGk\lambda\colon\mathbf{G}_{m,k^{\prime}}\to G_{k^{\prime}} where kkk\to k^{\prime} is a finite field extension such that G(Xk)λ+=XkG\cdot(X_{k^{\prime}})^{+}_{\lambda}=X_{k^{\prime}} and such that for any closed point y(Xk)λ+y\in(X_{k^{\prime}})^{+}_{\lambda}, the limit y:=limt0λ(t)gyy^{\prime}:=\lim_{t\to 0}\lambda(t)\cdot gy has closed orbit and the induced map λ:𝐆m,kGy\lambda\colon\mathbf{G}_{m,k^{\prime}}\to G_{y^{\prime}} is a regular one-parameter subgroup. By Theorem 6.1, we may replace kk with kk^{\prime} and thus we can assume that λ\lambda is defined over kk. We can further assume that the unique closed point x[X/G]x\in[X/G] lifts to a kk-point of XX.

We consider the composition

[Xλ+/Zλ]𝑓[Xλ+/Pλ]𝑔[X/G].[X^{+}_{\lambda}/Z_{\lambda}]\xrightarrow{f}[X^{+}_{\lambda}/P_{\lambda}]\xrightarrow{g}[X/G].

The morphism gg is proper (since Xλ+XX^{+}_{\lambda}\subset X is a closed subscheme and G/PλG/P_{\lambda} is projective) and surjective (since GXλ+=XG\cdot X^{+}_{\lambda}=X). By applying Theorem 6.1, we are reduced to showing that ([Xλ+/Pλ],g1(𝒢x))([X^{+}_{\lambda}/P_{\lambda}],g^{-1}(\mathcal{G}_{x})) satisfies formal functions and is coherent complete. Since gg is proper, the pushforward g𝒪[Xλ+/Pλ]g_{*}\mathcal{O}_{[X^{+}_{\lambda}/P_{\lambda}]} is coherent and it follows that Γ(Xλ+)Pλ\Gamma(X^{+}_{\lambda})^{P_{\lambda}} is finite over kk (where Γ(Xλ+)\Gamma(X^{+}_{\lambda}) is shorthand for Γ(Xλ+,𝒪Xλ+)\Gamma(X^{+}_{\lambda},\mathcal{O}_{X^{+}_{\lambda}})).

On the other hand, the morphism ff is faithfully flat and of finite type. The higher base changes of ff can be computed as

(8.1) ([Xλ+/Zλ]/[Xλ+/Pλ])n[(Xλ+×(Uλ)n1)/Zλ].([X^{+}_{\lambda}/Z_{\lambda}]\,/\,[X^{+}_{\lambda}/P_{\lambda}])^{n}\cong[(X^{+}_{\lambda}\times(U_{\lambda})^{n-1})/Z_{\lambda}].

Since the composition [X0/Zλ][X+/Pλ]ev0[X0/Zλ][X^{0}/Z_{\lambda}]\to[X^{+}/P_{\lambda}]\xrightarrow{\operatorname{ev}_{0}}[X^{0}/Z_{\lambda}] is the identity, the map Γ(Xλ0)ZλΓ(Xλ+)Pλ\Gamma(X^{0}_{\lambda})^{Z_{\lambda}}\to\Gamma(X^{+}_{\lambda})^{P_{\lambda}} is injective and thus Γ(Xλ0)Zλ\Gamma(X^{0}_{\lambda})^{Z_{\lambda}} is also finite over kk. But since Γ(Xλ+)λ=Γ(Xλ0)\Gamma(X^{+}_{\lambda})^{\lambda}=\Gamma(X^{0}_{\lambda}), we can conclude that Γ(Xλ+)Zλ=Γ(Xλ0)Zλ\Gamma(X^{+}_{\lambda})^{Z_{\lambda}}=\Gamma(X^{0}_{\lambda})^{Z_{\lambda}} is finite over kk. Since ZλZ_{\lambda} is reductive, the map [Xλ+/Zλ]SpecΓ(Xλ+)Zλ[X^{+}_{\lambda}/Z_{\lambda}]\to\operatorname{Spec}\Gamma(X^{+}_{\lambda})^{Z_{\lambda}} is an adequate moduli space. The stack [Xλ+/Zλ][X^{+}_{\lambda}/Z_{\lambda}] has finitely many closed points y1,,ynXλ0y_{1},\ldots,y_{n}\in X_{\lambda}^{0} with each yiy_{i} contained in (gf)1(𝒢x)(g\circ f)^{-1}(\mathcal{G}_{x}). By construction, each one-parameter subgroup λ:𝐆mStabG(yi)\lambda\colon\mathbf{G}_{m}\to\operatorname{Stab}^{G}(y_{i}) is regular. It follows that the stabilizer StabZλ(yi)=ZλStabG(yi)\operatorname{Stab}^{Z_{\lambda}}(y_{i})=Z_{\lambda}\cap\operatorname{Stab}^{G}(y_{i}) is a maximal torus of StabG(yi)\operatorname{Stab}^{G}(y_{i}), and in particular linearly reductive. By [AHR19, Thm. 4.21], [Xλ+/Zλ]SpecΓ(Xλ+)Zλ[X^{+}_{\lambda}/Z_{\lambda}]\to\operatorname{Spec}\Gamma(X^{+}_{\lambda})^{Z_{\lambda}} is a good moduli space. Thus [Xλ+/Zλ][X^{+}_{\lambda}/Z_{\lambda}] is coherently complete and satisfies formal functions with respect to 𝒢y1𝒢yn\mathcal{G}_{y_{1}}\cup\cdots\cup\mathcal{G}_{y_{n}}, and therefore also coherently complete and satisfies formal functions with respect to (gf)1(𝒢x)(g\circ f)^{-1}(\mathcal{G}_{x}) (Lemma 5.6).

Similarly, we claim that the higher base changes (8.1) are coherently complete and satisfied formal functions with respect to the preimage of 𝒢x\mathcal{G}_{x}. Since Γ(Uλ)λ=k\Gamma(U_{\lambda})^{\lambda}=k (Example 7.6), the global sections of ([Xλ+/Zλ]/[Xλ+/Pλ])n([X^{+}_{\lambda}/Z_{\lambda}]\,/\,[X^{+}_{\lambda}/P_{\lambda}])^{n} are identified with the global sections of [Xλ+/Zλ][X^{+}_{\lambda}/Z_{\lambda}] (Lemma 7.5). Thus, ([Xλ+/Zλ]/[Xλ+/Pλ])nSpecΓ(Xλ+)Zλ([X^{+}_{\lambda}/Z_{\lambda}]\,/\,[X^{+}_{\lambda}/P_{\lambda}])^{n}\to\operatorname{Spec}\Gamma(X^{+}_{\lambda})^{Z_{\lambda}} is an adequate moduli space. Since the closed points have linearly reductive stabilizers, it is in fact a good moduli space and the claim follows.

By Theorem 6.1, [Xλ+/Pλ][X^{+}_{\lambda}/P_{\lambda}] is coherently complete and satisfies formal functions with respect to the preimage of 𝒢x\mathcal{G}_{x}, and as we’ve already observed this implies that [X/G][X/G] is coherently complete and satisfies formal functions with respect to 𝒢x\mathcal{G}_{x}. ∎

Remark 8.1.

When XX has a GG-fixed point xXGx\in X^{G}, then there is a more direct argument relying on the generic stabilization theorem (Proposition 7.11) but not on its refinement (Proposition 7.14). Indeed, as in the proof above, we can reduce to the case that R=kR=k is a field, GG is connected, and XX is integral. Proposition 7.11 implies that after replacing kk with a finite extension, there exists a one-parameter subgroup λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G such that GXλ+=XG\cdot X^{+}_{\lambda}=X and Xλ0={x}X^{0}_{\lambda}=\{x\} (set-theoretically). Let TGT\subset G be a maximal torus containing λ\lambda and BPλB\subset P_{\lambda} be a Borel containing TT. In the composition

[Xλ+/T]𝑓[Xλ+/B]𝑔[X/G],[X^{+}_{\lambda}/T]\xrightarrow{f}[X^{+}_{\lambda}/B]\xrightarrow{g}[X/G],

the morphism ff is proper and surjective, and gg is faithfully flat. The higher base changes of ff can be computed as

([Xλ+/T]/[Xλ+/B])n[(Xλ+×Un1)/T]([X^{+}_{\lambda}/T]\,/\,[X^{+}_{\lambda}/B])^{n}\cong[(X^{+}_{\lambda}\times U^{n-1})/T]

where U=B/TU=B/T.

Since Xλ+X^{+}_{\lambda} is λ\lambda-positive with Γ(Xλ+)\Gamma(X^{+}_{\lambda}) finite over kk (Lemma 7.3) and Γ(Uλ)λ=k\Gamma(U_{\lambda})^{\lambda}=k (Lemma 7.3), the global sections of ([Xλ+/T]/[Xλ+/B])n([X^{+}_{\lambda}/T]\,/\,[X^{+}_{\lambda}/B])^{n} are also finite over kk (c.f. (Lemma 7.5)). It follows that for each n1n\geq 1, the higher base changes ([Xλ+/T]/[Xλ+/B])n\big{(}[X^{+}_{\lambda}/T]\,/\,[X^{+}_{\lambda}/B]\big{)}^{n} admit good moduli spaces that are finite schemes over kk, and that the higher base changes are coherently complete and satisfy formal functions with respect to the preimage of 𝒢x\mathcal{G}_{x}. The coherent completeness and the formal functions of ([X/G],𝒢x)([X/G],\mathcal{G}_{x}) follow from applying Theorem 6.1 subsequently to gg and ff.

Remark 8.2.

Standard reductions show that 1.1 for all RR and GG follows from the case where RR is an integral domain and G=GLn,RG=\mathrm{GL}_{n,R}. Applying the method of Theorem 1.2(1) to the fraction field of RR and replacing RR by a finite extension, it follows that it suffices to prove 1.1 when [X/G][X/G] has a saturated and dense open with “nice” stabilizers [AHR19, §2]—in particular, they are linearly reductive.

The following theorem generalizes Theorem 1.2(2).

Theorem 8.3.

Let GG be a smooth geometrically reductive group scheme over a noetherian ring RR.

  1. (1)

    The morphism BGSpecRBG\to\operatorname{Spec}R is cohomologically proper.

  2. (2)

    If in addition RR is II-adically complete for an ideal IRI\subset R, then the pair (BG,BGR/I)(BG,BG_{R/I}) satisfies formal functions and is coherently complete.

Proof.

Since GSpecRG\to\operatorname{Spec}R is smooth and geometrically reductive, the connected component of the identity G0G^{0} is a reductive group scheme over RR and G/G0G/G^{0} is finite over RR. By applying Theorem 6.1 to the finite cover BG0BGBG^{0}\to BG, we are reduced to the case that GSpecRG\to\operatorname{Spec}R is reductive. After an étale cover of SpecR\operatorname{Spec}R, GG becomes split reductive. Applying the faithfully flat version of Theorem 6.1, we are further reduced to showing that if GG is split reductive with maximal torus TT and Borel subgroup BB. Now consider the composition

BTBBBG.BT\to BB\to BG.

Since G/BG/B is projective over RR, the map BBBGBB\to BG is proper. On the other hand, BTBBBT\to BB is faithfully flat and the higher base changes can be described as

(BT/BB)n=[Un1/T](BT/BB)^{n}=[U^{n-1}/T]

where U=B/TU=B/T is the unipotent radical arising as the quotient of the left action by TT and TT acts diagonally on Un1U^{n-1} via the right action. Since Γ(U)T=R\Gamma(U)^{T}=R (Example 7.6) and thus Γ(Un1)T=R\Gamma(U^{n-1})^{T}=R (Lemma 7.5), the map (BT/BB)nSpecR(BT/BB)^{n}\to\operatorname{Spec}R is a good moduli space for each nn. Both statements follow by applying Theorem 6.1 to BTBBBT\to BB and then BBBGBB\to BG. ∎

8.2. Positively graded group schemes

Let GG be a smooth affine group scheme over a complete noetherian local ring RR with residue field kk. Let GuG_{u} be the unipotent radical of GG and Gr=G/GuG_{r}=G/G_{u} its reductive quotient. Let λ:𝐆mG\lambda\colon\mathbf{G}_{m}\to G be a one-parameter subgroup.

Definition 8.4.

We say that GG is positively graded with respect to λ\lambda if the exact sequence 1GuGGr11\to G_{u}\to G\to G_{r}\to 1 splits, the conjugation action on Γ(Gu,𝒪Gu)\Gamma(G_{u},\mathcal{O}_{G_{u}}) is λ\lambda-positive, and λ\lambda is central in GrG_{r}.

Remark 8.5.

If GG is defined over a field, then the conjugation action on Γ(Gu,𝒪Gu)\Gamma(G_{u},\mathcal{O}_{G_{u}}) is λ\lambda-positive if and only if the conjugation action of λ\lambda on Lie(Gu)\operatorname{Lie}(G_{u}) is positive, and λ\lambda is central in GrG_{r} if and only if the adjoint action of λ\lambda on Lie(Gr)=Lie(G)/Lie(Gu)\operatorname{Lie}(G_{r})=\operatorname{Lie}(G)/\operatorname{Lie}(G_{u}) is trivial.

Since the λ\lambda-action on Γ(G,𝒪Gu)\Gamma(G,\mathcal{O}_{G_{u}}) is positive and λ\lambda is central in GG, the parabolic PλGP_{\lambda}^{G} must be all of GG, and the short exact sequence 1GuGGr11\to G_{u}\to G\to G_{r}\to 1 is identified with the sequence 1UλPλZλ11\to U_{\lambda}\to P_{\lambda}\to Z_{\lambda}\to 1.

We can now prove Theorem 1.3: if GG is positively graded with respect to λ\lambda, then

  1. (1)

    BGBG is cohomologically proper over kk, and (BG,BGk)(BG,BG_{k}) is coherently complete and satisfies formal functions, and

  2. (2)

    if GG acts on an affine scheme X=SpecAX=\operatorname{Spec}A of finite type over RR such that AG=RA^{G}=R, λ\lambda acts semipositively on AA and [X/Gr][X/G_{r}] satisfies 1.1, then [X/G][X/G] is cohomologically proper over AGA^{G} and [X/G][X/G] satisfies formal functions and is coherently complete along its unique closed point.

Proof of Theorem 1.3.

Let GrGG_{r}\to G be a group homomorphism splitting the surjection GGrG\to G_{r}. For (1), we consider the the morphism

BGrBGBG_{r}\to BG

induced from the splitting. The map BGrBGBG_{r}\to BG is faithfully flat and there is an isomorphism

(BGr/BG)n[Gun1/Gr](BG_{r}/BG)^{n}\cong[G_{u}^{n-1}/G_{r}]

where GrG_{r} acts diagonally via the conjugation action. We know that λ\lambda acts positively on Gu=UλG_{u}=U_{\lambda} with Γ(Gu)λ=R\Gamma(G_{u})^{\lambda}=R (Example 7.6), and that the same holds for the λ\lambda-action on the higher fiber products (Lemma 7.5). In particular, (BGr/BG)nSpecR(BG_{r}/BG)^{n}\to\operatorname{Spec}R are good moduli spaces. Since BGrBG_{r} is cohomologically proper (Theorem 1.2(2)) and (BGr,B(Gr)k)(BG_{r},B(G_{r})_{k}) is coherently complete and satisfies formal functions, Theorem 6.1 implies that the same holds for (BG,BGk)(BG,BG_{k}).

For (2), we first claim that [SpecA/G][\operatorname{Spec}A/G] has a unique closed point xx. We will use essentially the same argument as in the case that GG is linearly reductive. Let π:[SpecA/G]SpecAG\pi\colon[\operatorname{Spec}A/G]\to\operatorname{Spec}A^{G}. It suffices to show that if 𝒵1,𝒵2[SpecA/G]\mathcal{Z}_{1},\mathcal{Z}_{2}\subset[\operatorname{Spec}A/G] are closed substacks, then each π(𝒵i)SpecAG\pi(\mathcal{Z}_{i})\subset\operatorname{Spec}A^{G} is closed and π(𝒵1)π(𝒵2)=π(𝒵1𝒵2)\pi(\mathcal{Z}_{1})\cap\pi(\mathcal{Z}_{2})=\pi(\mathcal{Z}_{1}\cap\mathcal{Z}_{2}). Let I1I_{1}, I2AI_{2}\subset A be the GG-invariant ideals defining 𝒵1\mathcal{Z}_{1} and 𝒵2\mathcal{Z}_{2}. Since AA is λ\lambda-positive, we have AG=AGrA^{G}=A^{G_{r}}. Thus, the image of 𝒵i\mathcal{Z}_{i} is identified with the image of the closed GrG_{r}-invariant ideal V(Ii)V(I_{i}) under [SpecA/Gr]SpecAGr[\operatorname{Spec}A/G_{r}]\to\operatorname{Spec}A^{G_{r}}, and it follows that π(𝒵i)\pi(\mathcal{Z}_{i}) is closed by the reductive case. Similarly, since since I1I_{1}, I2I_{2}, and I1+I2I_{1}+I_{2} are λ\lambda-positive, we have

π(𝒵1𝒵2)=V((I1+I2)G)=V((I1+I2)Gr)=V(I1Gr)V(I2Gr)=V(I1G)V(I2G)=π(𝒵1)π(𝒵2),{\pi(\mathcal{Z}_{1}\cap\mathcal{Z}_{2})=V((I_{1}+I_{2})^{G})=V((I_{1}+I_{2})^{G_{r}})=V(I_{1}^{G_{r}})\cap V(I_{2}^{G_{r}})=V(I_{1}^{G})\cap V(I_{2}^{G})=\pi(\mathcal{Z}_{1})\cap\pi(\mathcal{Z}_{2}),}

where the middle equality used the reductive case. The claim follows.

We now consider the morphism [SpecA/Gr][SpecA/G][\operatorname{Spec}A/G_{r}]\to[\operatorname{Spec}A/G] and use the following identification of the higher fiber products

([SpecA/Gr]/[SpecA/G])n=[(SpecA×Gun1)/Gr].([\operatorname{Spec}A/G_{r}]/[\operatorname{Spec}A/G])^{n}=[(\operatorname{Spec}A\times G_{u}^{n-1})/G_{r}].

Since λ\lambda acts positively on GuG_{u} with Γ(Gu)λ=R=AG\Gamma(G_{u})^{\lambda}=R=A^{G} and λ\lambda acts semipositively on AA, the group of global sections of each higher fiber product is identified with RR by Lemma 7.5. As GrG_{r} is reductive, each higher fiber product is an adequate moduli space over SpecR\operatorname{Spec}R. It follows from Theorem 6.1, 1.1 and Corollary 4.9 that [SpecA/G][\operatorname{Spec}A/G] is cohomologically proper, and is coherently complete and satisfies formal functions with respect to 𝒢x\mathcal{G}_{x}. ∎

References

  • [AHHR23] Jarod Alper, Jack Hall, Daniel Halpern-Leistner, and David Rydh. Artin algebraization for pairs with applications to the local structure of stacks and Ferrand pushouts. to appear in Forum Math. Sigma, 2023.
  • [AHLH23] Jarod Alper, Daniel Halpern-Leistner, and Jochen Heinloth. Existence of moduli spaces for algebraic stacks. Invent. Math., Aug 2023.
  • [AHR19] Jarod Alper, Jack Hall, and David Rydh. The étale local structure of algebraic stacks, 2019.
  • [AHR20] Jarod Alper, Jack Hall, and David Rydh. A Luna étale slice theorem for algebraic stacks. Ann. of Math. (2), 191(3):675–738, 2020.
  • [Alp13] Jarod Alper. Good moduli spaces for Artin stacks. Ann. Inst. Fourier (Grenoble), 63(6):2349–2402, 2013.
  • [AM69] Michael F. Atiyah and Ian G. Macdonald. Introduction to commutative algebra. Addison-Wesley Publishing Co., Reading, Mass.-London-Don Mills, Ont., 1969.
  • [Art74] Michael Artin. Versal deformations and algebraic stacks. Invent. Math., 27:165–189, 1974.
  • [BDHK18] Gergely Bérczi, Brent Doran, Thomas Hawes, and Frances Kirwan. Geometric invariant theory for graded unipotent groups and applications. J. Topol., 11(3):826–855, 2018.
  • [CGP15] Brian Conrad, Ofer Gabber, and Gopal Prasad. Pseudo-reductive groups, volume 26 of New Mathematical Monographs. Cambridge University Press, Cambridge, second edition, 2015.
  • [Con05] Brian Conrad. Formal GAGA for Artin stacks. https://math.stanford.edu/~conrad/papers/formalgaga.pdf, 2005.
  • [Con14] Brian Conrad. Reductive group schemes. In Autour des schémas en groupes. Vol. I, volume 42/43 of Panor. Synthèses, pages 93–444. Soc. Math. France, Paris, 2014.
  • [Con20] Brian Conrad. Algebraic groups II. https://www.ams.org/open-math-notes/omn-view-listing?listingId=110663, 2020.
  • [Dri13] Vladimir Drinfeld. On algebraic spaces with an action of 𝔾m\mathbb{G}_{m}, 2013.
  • [FGI+05] Barbara Fantechi, Lothar Göttsche, Luc Illusie, Steven L. Kleiman, Nitin Nitsure, and Angelo Vistoli. Fundamental algebraic geometry, volume 123 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2005. Grothendieck’s FGA explained.
  • [GD67] Alexander Grothendieck and Jean Dieudonné. Éléments de géométrie algébrique. I.H.E.S. Publ. Math, 4, 8, 11, 17, 20, 24, 28, 32, 1960, 1961, 1961, 1963, 1964, 1965, 1966, 1967.
  • [Gro68] Alexander Grothendieck. Cohomologie locale des faisceaux cohérents et théorèmes de Lefschetz locaux et globaux (SGA(SGA 2)2). Advanced Studies in Pure Mathematics, Vol. 2. North-Holland Publishing Co., Amsterdam; Masson & Cie, Editeur, Paris, 1968. Augmenté d’un exposé par Michèle Raynaud, Séminaire de Géométrie Algébrique du Bois-Marie, 1962.
  • [GZB15] Anton Geraschenko and David Zureick-Brown. Formal GAGA for good moduli spaces. Algebr. Geom., 2(2):214–230, 2015.
  • [Hal14] Daniel Halpern-Leistner. On the structure of instability in moduli theory, 2014.
  • [Hal23] Jack Hall. GAGA theorems. J. Math. Pures Appl. (9), 175:109–142, 2023.
  • [Hes79] Wim H. Hesselink. Desingularizations of varieties of nullforms. Invent. Math., 55(2):141–163, 1979.
  • [HLP23] Daniel Halpern-Leistner and Anatoly Preygel. Mapping stacks and categorical notions of properness. Compos. Math., 159(3):530–589, 2023.
  • [HR19] Jack Hall and David Rydh. Coherent Tannaka duality and algebraicity of Hom-stacks. Algebra Number Theory, 13(7):1633–1675, 2019.
  • [Kem78] George R. Kempf. Instability in invariant theory. Ann. of Math. (2), 108(2):299–316, 1978.
  • [Knu71] Donald Knutson. Algebraic spaces. Lecture Notes in Mathematics, Vol. 203. Springer-Verlag, Berlin-New York, 1971.
  • [LMB00] Gérard Laumon and Laurent Moret-Bailly. Champs algébriques, volume 39 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, 2000.
  • [Mum65] David Mumford. Geometric invariant theory. Ergebnisse der Mathematik und ihrer Grenzgebiete, Neue Folge, Band 34. Springer-Verlag, Berlin-New York, 1965.
  • [Nag62] Masayoshi Nagata. Complete reducibility of rational representations of a matric group. J. Math. Kyoto Univ., 1:87–99, 1961/62.
  • [Ols05] Martin C. Olsson. On proper coverings of Artin stacks. Adv. Math., 198(1):93–106, 2005.
  • [RG71] Michel Raynaud and Laurent Gruson. Critères de platitude et de projectivité. Techniques de “platification” d’un module. Invent. Math., 13:1–89, 1971.
  • [Ryd] David Rydh. If the direct image of f preserves coherent sheaves on noetherian schemes, how to show f is proper? MathOverflow. URL:https://mathoverflow.net/q/182902 (version: 2014-10-14).
  • [Ryd16a] David Rydh. Equivariant flatification, étalification and compactification, 2016. in preparation.
  • [Ryd16b] David Rydh. Functorial flatification of proper morphisms, 2016. in preparation.
  • [Sta21] The Stacks Project Authors. Stacks Project. https://stacks.math.columbia.edu, 2021.
  • [TvdK10] Antoine Touzé and Wilberd van der Kallen. Bifunctor cohomology and cohomological finite generation for reductive groups. Duke Math. J., 151(2):251–278, 2010.
  • [vdK15] Wilberd van der Kallen. Good Grosshans filtration in a family. In Autour des schémas en groupes. Vol. III, volume 47 of Panor. Synthèses, pages 111–129. Soc. Math. France, Paris, 2015.