Coherently Complete Algebraic Stacks in Positive Characteristic
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 , with reductive and complete local, are coherently complete along the unique closed point. We establish this conjecture in two interesting cases: (1) is artinian and (2) acts trivially on . 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 be a proper scheme over a complete noetherian local ring . There are three important theorems that govern the formal geometry of , analogous to Grauert’s finiteness theorem and Serre’s GAGA theorems in the setting of complex analytic geometry:
-
(1)
Finiteness of Cohomology: For any coherent sheaf on , the cohomology group is a finitely generated -module for all .
-
(2)
Formal Functions: Let to be the central fiber, the th nilpotent thickening of , and the formal completion of along . Then for any coherent sheaf on , there are natural isomorphisms:
-
(3)
Formal GAGA: There are equivalences of categories
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 , where the invariant ring is a complete local ring. To formulate our results, we introduce the following three definitions mirroring the three properties above:
-
(1)
A noetherian algebraic stack over a noetherian ring is cohomologically proper if for every coherent sheaf on , the cohomology group is a finitely generated -module for all .
Let be a noetherian algebraic stack and a closed substack. Let be the completion of along ; i.e. the ringed site , where (see Definition 3.1).
-
(2)
We say that the pair satisfies formal functions if for every coherent sheaf on , the natural map
is an isomorphism for all .
-
(3)
We say that the pair is coherently complete if the natural functor
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 be a smooth geometrically reductive group scheme over a complete noetherian local ring . Let be an affine scheme of finite type over with an action of such that . Let be the unique closed point and be its residual gerbe. Then is cohomologically proper over , and the pair is coherently complete and satisfies formal functions.
This conjecture differs from formal GAGA in two ways. First, when is not finite, the quotient stack is not separated, let alone proper, over . 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 on , the origin is the unique closed orbit and the invariant ring is . Since the central fiber of is everything, the coherent completeness of along the central fiber has no mathematical content. On the other hand, the coherent completeness of along the origin is a non-trivial statement.
This conjecture was established in [AHR19, Thm. 1.6] in the case that is linearly reductive. While in characteristic 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 is reductive in any characteristic but linearly reductive only in characteristic 0.
Theorem 1.2.
1.1 holds in the following cases
-
(1)
is an artinian local ring, or
-
(2)
acts trivially on , i.e. and .
We expect that these two cases will assist in establishing 1.1 in general. Indeed, (1) reduces the statement that is coherently complete and satisfies formal functions with respect to the residual gerbe to seemingly simpler statement that satisfies these properties with respect to the central fiber of . Our methods essentially reduce 1.1 to the ‘generically toric’ situation (Remark 8.2).
In fact, we establish a stronger result than (2): if is a geometrically reductive group scheme over an -adically complete noetherian ring , then is cohomologically proper over , and the pair is coherently complete and satisfies formal functions (Theorem 8.3). When 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 be a smooth affine group scheme over a complete local noetherian ring with residue field . Let be its unipotent radical with reductive quotient . Let be a one-parameter subgroup such that is central in and acts positively on . Then
-
(1)
is cohomologically proper over , and is coherently complete and satisfies formal functions; and
-
(2)
if acts on an affine scheme of finite type over such that , satisfies 1.1 (e.g., is linearly reductive or is artinian) and acts semipositively on , then is cohomologically proper, and there is a unique closed point such that is coherently complete and satisfies formal functions.
For example, if is a Borel subgroup of a split reductive group scheme over , then for a regular one-parameter subgroup . The reductive quotient of is a maximal torus (thus linearly reductive) such that is central. The theorem implies that is cohomologically proper and that is coherently complete, and moreover that the same holds for a quotient stack if and acts semipositively on .
On the other hand, these three cohomological properties do not hold for every algebraic group , 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 is a universally submersive morphism and is a closed substack such that and the higher base changes satisfy the three cohomological properties—cohomologically properness, formal functions, and coherent completeness—along the preimage of , then satisfies these properties along (Theorem 6.1). The two most important cases are either when is proper and surjective, in which case one only needs to check that has the properties along the preimage of , and when is faithfully flat.
To apply the descent result, we construct suitable covers of . For the case of for a reductive group over a complete local noetherian ring (i.e. Theorem 1.2(2)), we first reduce to the case that is split reductive. We consider subgroups where is a maximal torus and is a Borel subgroup, and consider the induced morphisms
Since is projective, the map is proper and surjective, and hence Theorem 6.1 reduces the theorem to . On the other hand, is faithfully flat and affine. Since is a torus and hence linearly reductive, we know the cohomological properties for . Moreover, the higher base changes of are identified with quotient stacks where is the quotient of the right action, and acts on on the left. The -invariants of the action on the -fold product are identified with , which implies that is a good moduli space and thus satisfies the three cohomological properties. Theorem 6.1 then implies that satisfies the desired properties.
The construction of the covers in the case of with artinian (i.e., Theorem 1.2(1)) is more involved. The general strategy is to find a one-parameter subgroup and consider the pair of morphisms
where is the ‘attractor’ locus parameterizing points of that have a limit under -action induced by (see §7.1), and where and are the centralizer and parabolic of (see §7.2). For any , the map is proper since is a closed subscheme and is projective, while the map is faithfully flat and affine. A version of the Hilbert–Mumford criterion (Proposition 7.11) implies that there is one-parameter subgroup destabilizing at least one point in every -orbit, which gives that or in other words that is surjective. The descent result (Theorem 6.1) therefore reduces to the claim to . However, (in addition to the higher base changes of ) do not clearly satisfy the cohomological properties as , 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 such that every -orbit contains a point and such that exists and lies in the unique closed -orbit, and moreover such that the induced one-parameter subgroup is regular in the stabilizer . This implies that the connected component of is linearly reductive, which suffices to show that 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 is a proper representable morphism and is coherently complete, then 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 near a point with linearly reductive stabilizer ; 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 of finite type with affine diagonal admits a separated adequate moduli space if and only is -complete and -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 is a smooth point is to consider the quotient of the Zariski tangent space, and the base change
along the completion of the GIT quotient at the image of the origin. Since is coherently complete along by the linearly reductive case of 1.1 (i.e. [AHR19, Thm. 1.6]), Tannaka duality [HR19] implies that
Using again that is linearly reductive, deformation theory yields isomorphisms of the th nilpotent thickening of and . The identification above implies that the maps extend to a map , which can then be approximated to produce the desired étale neighborhoods. In the reductive case, 1.1 would yield the coherent completeness of 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 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 along the residual gerbe of the unique closed point. In contrast, [HLP23] focuses on coherent completeness results of stacks such as along the central fiber of a morphism . 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
2. Cohomological properness
For a morphism of noetherian schemes (even algebraic spaces or algebraic stacks with finite diagonal), it is known that is proper if and only if the higher pushforwards send coherent sheaves to coherent sheaves for all [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 be a finite type morphism of noetherian algebraic stacks. We say that is:
-
(1)
cohomologically proper if for every coherent sheaf on , ; and
-
(2)
universally cohomologically proper if for every morphism of noetherian algebraic stacks, the base change 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 is cohomologically proper and , then . This follows from the convergent hypercohomology spectral sequence:
Remark 2.5.
If 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 , where is a DVR with fraction field such that has a section . We must show that, potentially after extending , that extends to a section over . If , then we can find a DVR over such that realizes the specialization from the image of to . In particular, we must show that cannot occur if is cohomologically proper. If , then factors through . It follows that is a non-zero finite dimensional -vector space, so it cannot be a finitely generated -module, which contradicts the cohomological properness of . This proves the claim.
The following two examples will be key in this article.
Example 2.6.
Let be a proper morphism of noetherian algebraic stacks. Then 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 be a noetherian ring. Let be a finitely generated -algebra with an action of a reductive group scheme over . Then is universally cohomologically proper. Indeed, by Remark 2.3, it suffices to check that if is a noetherian -algebra, then is cohomologically proper. But is a finitely generated -algebra and the result follows from [vdK15, Thm. 10.5].
3. Formal functions and coherent completeness
Throughout we use the following notation: if is a noetherian algebraic stack and is a closed substack defined by a coherent sheaf of ideals , then we denote by the closed substack defined by . If is a quasi-coherent sheaf on , we denote by its pullback to .
Definition 3.1.
Let be a noetherian algebraic stack and be a closed substack defined by a coherent sheaf of ideals . The completion of the pair is the ringed site , where the sheaf of rings is defined by the limit
in the category of lisse-étale modules.
There is a canonical morphism of ringed sites , which is flat [GZB15, Lem. 3.3] and whose pullback preserves coherence; we will often suppress the subscript by writing or simply if there is little possibility for confusion. For a coherent -module , we denote by the pullback of to . There is a natural identification . There is also an exact equivalence of abelian categories [Con05, Thm. 2.3].
Definition 3.2.
Let be a noetherian algebraic stack and be a closed substack.
-
(1)
satisfies formal functions if for any coherent sheaf on , the natural map:
is an isomorphism for all ;
-
(2)
is coherently complete if the functor:
is an equivalence.
-
(3)
satisfies derived formal functions if the functor:
is fully faithful.
-
(4)
is derived coherently complete if the functor:
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 for all and coherent sheaves on . We will refer to this as Zariski formal functions. If , then in [Sta21, Tag 0A0K]. In particular, . The Milnor exact sequence [Sta21, Tag 07KZ] implies that for each there is a short exact sequence:
(3.1) |
It follows that formal functions implies that for all and . Conversely, Zariski formal functions implies that . Hence, these conditions are not obviously equivalent in the generality that we are working. They are when or when 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 consisting of an algebraic stack and a cocompact closed substack is called complete if is an equivalence. A morphism of algebraic stacks is formally proper if for every complete pair , the base change is complete with respect to .
If is a noetherian algebraic stack over a complete noetherian local ring and is a closed point, then the conditions of being (derived) coherently complete and the morphism being formally proper are incomparable. On one hand, coherent completeness concerns the closed substack rather than the central fiber while formal properness includes a completeness property with respect to all base changes.
If the functor is fully faithful (e.g. is coherently complete), then for any coherent sheaf on , the natural map
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 is fully faithful (i.e. satisfies derived formal functions), then
is an isomorphism for all (i.e. 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 be a noetherian algebraic stack and a closed substack. The morphism of ringed sites induces an adjoint pair on unbounded derived categories:
The functor coincides with the composition of the forgetful functor
and the quasi-coherator, which is the right adjoint to the inclusion . See [Hal23, §4] for more details. We have the following simple proposition.
Proposition 3.5.
Let be a noetherian algebraic stack and a closed substack.
-
(1)
satisfies formal functions if and only if it satisfies derived formal functions. Moreover, if , then is an isomorphism.
-
(2)
If satisfies formal functions, then it is coherently complete if and only if it is derived coherently complete. In this case,
-
(a)
is -exact on ; and
-
(b)
if , then and .
-
(a)
Proof.
Assume that satisfies derived formal functions. Let ; then
That is, satisfies formal functions. Conversely, we first note that if satisfies formal functions and , then . Indeed, by the hypercohomology spectral sequence, we have a diagram
where (a) is a morphism of spectral sequences that is an isomorphism by formal functions. Hence (b) is an isomorphism as well. Now let and . Then and so we have
Let . To prove that formal functions implies derived formal functions, it suffices to prove that the morphism is an isomorphism in . By Lemma 3.7 and the above, is an isomorphism whenever . In general, Lemma 3.6 provides
This proves (1). For (2), by (1) it suffices to prove that the functor is essentially surjective. Equivalently, that the natural map is an isomorphism in . A standard induction argument on the length of the complex together with the derived formal functions assumption and coherent completeness, shows that for every there is a unique such that . If , it follows from derived formal functions that the map induces isomorphisms:
It follows from Lemma 3.7 that the map is an isomorphism, so is an isomorphism. Now let . Let and set . Then [Sta21, Tag 0D6T] (also see Lemma 3.6) shows that whenever . Hence, . Lemma 3.6 gives a morphism:
Let ; then the above map induces isomorphisms:
Hence, . In particular, arguing as above using derived formal functions and Lemma 3.7, we see that is an isomorphism whenever . In general, let . Set . Let ; then whenever . In particular, . But [Sta21, Tag 0949] implies that
Arguing as before, the result follows. ∎
The following two lemmas featured in the proof of Proposition 3.5.
Lemma 3.6.
Let be a noetherian algebraic stack and a closed substack. If , then .
Proof.
Lemma 3.7.
Let be a noetherian algebraic stack. Let be a morphism in . If is an isomorphism for all , then is an isomorphism.
Proof.
Let be a cone for . If , choose least such that . Then there is a non-zero coherent sheaf and an injection [LMB00, Prop. 15.4]. Hence, the induced map is non-zero, which is a contradiction. ∎
As a result, we can conclude that formal functions implies that is fully faithful.
Corollary 3.8.
Let be a noetherian algebraic stack and a closed substack. If satisfies formal functions, then the functor
is fully faithful, with essential image stable under kernels, cokernels and extensions. Moreover, it is an equivalence if for each non-zero there exist and a non-zero map .
Proof.
Since satisfies formal functions, it satisfies derived formal functions (Proposition 3.5(1)). Hence, is exact and fully faithful. It follows immediately that the image of in is stable under kernels and cokernels. For the extensions, we know that derived formal functions implies that if , , then
Hence, the image is also stable under extensions.
If , let , which is a left-exact functor. For the equivalence, the condition implies that whenever . Now fix . If is a coherent subsheaf, let . Then and there is an exact sequence:
By Proposition 3.5(1), is an isomorphism. Hence, and so . Now write as a filtered union of coherent subsheaves. Then defines an increasing union of coherent subsheaves of . By coherence of , there must be a such that for all . By full faithfulness, for all and so is coherent. Finally, consider the exact sequence:
Applying , we see that from Proposition 3.5(1) again that
is exact. It follows that and so , 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 be a morphism of noetherian algebraic stacks, a closed substack, and a closed substack. We say that the triple is strongly cohomologically proper if for sheaves of ideals and defining and , the morphism is cohomologically proper.
Remark 4.2.
In Definition 4.1, from the commutative diagram:
where the horizontal morphisms are closed immersions, it follows that is cohomologically proper. It is also easy to see that the definition only depends on the closed subsets and (not the ideals and ). Indeed, suppose that we have other ideals , with vanishing locus and . Now form the commutative diagram:
Since and 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 be a morphism of noetherian algebraic stacks with affine diagonal, a closed substack, and a closed substack. Assume that is an isomorphism (i.e., is Stein) and that , where defines and defines . Then the triple is strongly cohomologically proper if and only if the -algebra is finitely generated and the induced morphism is cohomologically proper. Indeed, we have the composition
If is strongly cohomologically proper, then is a finite -algebra and so the induced morphism is cohomologically proper. Conversely, if is a finitely generated -algebra, then there is an such that as ideals of for all . But we have . In particular, we have a commutative diagram:
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 be a universally cohomologically proper morphism of noetherian algebraic stacks. Let be a closed substack. Then the triple is strongly cohomologically proper. To see this: let be a coherent sheaf of ideals definining . Since is noetherian, is a finitely generated -algebra. Now form the cartesian square:
Since is universally cohomologically proper, is cohomologically proper.
Example 4.5.
Let be a noetherian ring. Let be a finitely generated -algebra with an action of a reductive group scheme over . Let be a -equivariant ideal. Let , , and . Then the triple is strongly cohomologically proper. Indeed, by Example 2.7, is universally cohomologically proper. Further, the -equivariant -algebra is finitely generated and so is a finitely generated -algebra, where denotes the coherent sheaf of ideals on associated to . By Example 2.7 again, the induced morphism is cohomologically proper and the claim follows from Remark 4.3. If is artinian, this also follows from Theorem 1.2.
We now have the main result of this section.
Theorem 4.6.
Let be a finite type morphism of noetherian algebraic stacks, an ideal, and a closed substack. Assume that the triple is strongly cohomologically proper and is -adically complete.
-
(1)
Let be a coherent -ideal defining the closed immersion . Then is -adically complete.
-
(2)
Formal functions holds for the pair . Moreover, if is a coherent -module, then the natural morphisms:
are isomorphisms for all .
-
(3)
If is a coherent -module, then is a finitely generated -module and the natural morphism:
is an isomorphism for all .
Before we prove Theorem 4.6, we provide several corollaries.
Corollary 4.7.
Let be a universally cohomologically proper morphism of noetherian algebraic stacks. Then admits a Stein factorization:
where is finite and is universally cohomologically proper and universally closed with geometrically connected fibers.
Proof.
By Remark 2.5, it suffices to prove that if , where is an -adically complete noetherian local ring and ; then the closed fiber is connected. By Example 4.4, the triple is strongly cohomologically proper. It follows from Theorem 4.6(2) that . If is disconnected, then the corresponding idempotent lifts to a non-trivial idempotent of , which contradicts being local [Sta21, Tag 0G7X]. ∎
Corollary 4.8.
Let be a universally cohomologically proper morphism of noetherian algebraic stacks. Let be an ideal. If is -adically complete, then formal functions holds for the pair .
Proof.
Combine Example 4.4 with Theorem 4.6. ∎
Corollary 4.9.
Let be a noetherian ring. Let be a finitely generated -algebra with an action of a reductive group scheme over . Let be a -equivariant ideal. Let , , , and the induced morphism. If is -adically complete, then the triple is strongly cohomologically proper and the pair 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 be a ring, an ideal, and an -module. A filtration of is -good (or -stable) if the following three conditions are satisfied:
-
(1)
for some ;
-
(2)
for all ; and
-
(3)
for all .
Obviously, the filtration is -good; the topology that this filtration defines on is called the -adic topology on . A key observation is that the topology on defined by any -good filtration is equivalent to the -adic topology on [AM69, Lem. 10.6]. A much deeper fact is that if is noetherian and is finitely generated, then a filtration is -stable if and only if is a finitely generated -module [AM69, Lem. 10.8]. A key consequence of this whole theory is that if is a finitely generated -module and is -adically complete, then is -adically complete.
Assume now that we are in the situation of Theorem 4.6. Let be a coherent -module. Let and let . The quasi-coherent -algebra is of finite type and is a coherent -module [AM69, Lem. 10.8].
Let and . Let us briefly remark on the graded structure of . If , then for all there is an induced homomorphism of -modules that that is multiplication by . It follows that we obtain an induced morphism:
This is how becomes a graded -module. In particular, denotes the image of the natural -module homomorphism .
Further, the canonical inclusions for give rise to an inverse system with transition map when . It follows that the composition:
coincides with multiplication by on as an -module. In particular, if is an -submodule, then we have the equality of -submodules of :
(4.1) |
By assumption, is a finitely generated -algebra and is a finitely generated and graded -module. It follows that for some sufficiently large , that as ideals of the finite -algebra , we have for all . In particular, replacing and by and , respectively, we may assume that for all .
Proof of Theorem 4.6.
We first prove (1). The discussion above showed that is a finite -module. It follows that the -adic filtration on is -stable and so the -adic topology is equivalent to the -adic topology. Since is -adically complete and is a finite -module, is -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 and for all and so . By Remark 3.3, it suffices to prove that is an isomorphism for all coherent sheaves on and and satisfies the Artin–Rees condition (this implies that the term in (3.1) vanishes).
We now let be a coherent -module. Let and and consider the exact sequence of -modules:
(4.2) |
where , , ,
The result follows from the following three claims:
-
(1)
the filtration on is -good;
-
(2)
the inverse system is Artin–Rees zero (i.e., there exists an such that is for all );
-
(3)
the inverse system satisfies the uniform Artin–Rees condition (i.e., there is an such that the images of the morphisms agree for all ).
Indeed, the exact sequence (4.2) induces the following short exact sequence:
(4.3) |
We now take inverse limits, and obtain the following exact sequence:
Since the system is Artin–Rees zero, it follows immediately that . Moreover, the filtration on is -good and since is a finitely generated -module, it follows that the natural map is an isomorphism. What results from all of this is an isomorphism:
That is, Zariski formal functions holds for the pair .
We first establish that the filtration on is -good. To see this, we first note that . We now apply the previous discussion to the -module . It follows that is a finitely generated -module. But the graded -module is the image of the graded -module homomorphism
and so is also a finitely generated graded -module. By [AM69, Lem. 10.8], it follows that the filtration is -good.
We next prove that the inverse system is Artin–Rees zero. First observe that is a -submodule of , which is a finitely generated -module. Hence, is a finitely generated -module. In particular, there exist integers , such that for all . But is always a quotient of and is annihilated by and so if , then write , where . Then
It follows from (4.1) and the above that if , then for we have
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 and satisfy the uniform Artin–Rees condition. Since is Artin–Rees zero, it satisfies the uniform Artin–Rees condition. Further, since is a surjective system it trivially satisfies the uniform Artin–Rees condition. This proves (2)
5. Permanence of properties
Let be a morphism of noetherian algebraic stacks, a closed substack, and a closed substack. Then there is an induced diagram:
as well as natural isomorphisms of functors and . Because of the lack of functoriality of the lisse-étale topos, the left derived functors of and are somewhat subtle if is not smooth. The derived functors on the level of unbounded derived categories and always exist, however. As and are exact (they are just forgetful functors), it follows that we also have natural isomorphisms . If , then there is thus always a comparison morphism:
(5.1) |
which comes from the adjoint of the morphism:
We now introduce a relative version of Definition 3.2 (cf. [HLP23, Defn. 1.1.3]).
Definition 5.1.
Let be a morphism of noetherian algebraic stacks, is a closed substack, and is closed substack. We say that the triple satisfies relative formal functions (resp. relatively (derived) coherently complete) if satisfies formal functions (resp. (derived) coherent completeness) for all noetherian rings and smooth morphisms , where denotes the adic completion of with respect to the ideal defining , , and is the preimage of under .
We have the following simple lemma.
Lemma 5.2.
Let be a morphism of noetherian algebraic stacks, a closed substack, and a closed substack. If the triple satisfies relative formal functions, then for every the natural comparison morphism:
is an isomorphism. If, in addition, the pair satisfies formal functions, then so does .
Proof.
The second claim follows from the first since for any coherent sheaf on , we have isomorphisms:
To treat the first claim, let be a smooth morphism, the ideal defined by the preimage of , and the -adic completion of . Let be the base change morphism, where . For an integer , is the sheafification of the presheaf:
with the last isomorphism by flat base change. One similarly finds that is the sheafification of the presheaf:
By assumption, the pair 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 be a finite type morphism of noetherian algebraic stacks, a closed substack, and a closed substack. Assume that the triple is strongly cohomologically proper.
-
(1)
The triple satisfies relative formal functions.
-
(2)
The comparison morphism is an isomorphism for all .
-
(3)
The functor sends to .
Proof.
By Lemma 5.2, (1)(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 be a noetherian algebraic stack with affine stabilizers and a closed substack. Let be a proper and representable morphism. If satisfies formal functions and is coherently complete, then so does .
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 . We have the following diagram:
Since is proper and representable, it is universally cohomologically proper [Ols05, Thm. 1.2], so the triple is strongly cohomologically proper (Example 4.4). In particular, Theorem 5.3 implies that satisfies relative formal functions and sends to . A simple calculation using homotopy limits shows that because is representable, even sends to . In particular, Lemma 5.2 implies that satisfies formal functions. By Proposition 3.5(2) it remains to prove that
is essentially surjective. We first prove this under the following assumptions:
-
(1)
is projective (so comes with a relatively ample line bundle ); and
-
(2)
is a Cartier divisor.
In this case, let and let . Then by the projection formula [Hal23, Thm. A.12],
But and since is derived coherently complete, by (Proposition 3.5(2))
Let be a smooth cover by an affine scheme. Let and let and be the induced morphisms. We may choose so that is ample relative to . Then
Since this is true for all and is ample relative to , it follows from [Hal23, Thm. 3.8] that . By smooth descent, . It remains to prove that the adjunction morphism
(5.2) |
is an isomorphism.
To prove this, observe that is a perfect complex on since is Cartier. It follows from the projection formula that
In summary, if denotes the cone of (5.2), we have shown that . To finish the proof, suppose that . Then since , we may choose a largest integer such that . Then
and so by Nakayama’s Lemma, , which is a contradiction.
In general, Rydh’s Chow Lemma [Ryd16a] provides a blow-up such that the composition is projective. We may replace by an additional blow-up so that becomes Cartier. It follows from the case already considered that is derived coherently complete. By Theorem 6.1(3), it follows that is derived coherently complete. ∎
We also have the following easy result.
Lemma 5.6.
Let be a noetherian algebraic stack and a closed substack. Suppose that is a closed substack such that . If satisfies formal functions (resp. is coherently complete), then the same holds for .
We collect here some examples of algebraic stacks that satisfy each of the three properties: cohomological properness, formal functions and coherent completeness.
Example 5.7 (Noetherian affine schemes).
Let be a noetherian ring and let be an ideal. Then is coherently complete if and only if is -adically complete [AHR19, Ex. 3.9].
Example 5.8 (Proper algebraic stacks).
Example 5.9 (Good moduli spaces).
Let be a good moduli space [Alp13], where is a noetherian algebraic stack with affine diagonal. Then is of finite type [AHR20, Thm. A.1], and is noetherian and is cohomologically proper over [Alp13, Thm. 4.16 (x)]. Since good moduli spaces are compatible with arbitrary base change, is even universally cohomologically proper over .
Let be a closed substack defined by a coherent sheaf of ideals and let to be the corresponding ideal of . If has the resolution property and is -adically complete, then the pair is coherently complete and satisfies formal functions [AHR19, Thm. 1.6]. When has the resolution property and is the preimage of , then this was the main result of [GZB15]. If is equicharacteristic, has the resolution property and is supported at a closed point, then this is [AHR20, Thm. 1.3].
We also point out the following converse. Let be a noetherian algebraic stack with noetherian adequate moduli space . Let be a closed substack with scheme-theoretic image . If is affine and the pair is coherently complete, then the pair is coherently complete [AHR19, Lem. 3.5(1)].
Example 5.10 (Adequate moduli spaces).
Let , where is a reductive algebraic group over a noetherian ring acting on an affine scheme of finite type over . Then the cohomology ring is a finitely generated -algebra. See [TvdK10, Thm. 1.1] for the case when is a field and [vdK15] in general. It follows that is cohomologically proper. If is a -equivariant ideal is -adically complete, then satisfies formal functions (Corollary 4.9).
Example 5.11 (Unipotent groups).
If is a unipotent (affine) algebraic group over an algebraically closed field , then is cohomlogically proper if and only if . Indeed, if , then admits a filtration by ’s and thus it suffices to show that is cohomologically proper. Any finite dimensional representation of has a filtration by trivial representations and so it is enough to compute the cohomology of the trivial representation. Using the Čech complex corresponding to the cover , one can compute that if and zero otherwise.
On the other hand, if , then we can compute (again using the Čech complex) that is infinite dimensional. Thus is not cohomologically proper and it follows that neither is for any unipotent group .
Supposing again that , let be a complete local noetherian ring with residue field such that is not artinian. Then is not coherently complete. It suffices to assume that and that . Consider the exponential map defined by , where and are coordinates of and , respectively. This defines a compatible system of non-trivial line bundles on that does not algebraize to a line bundle on . Indeed, every line bundle on is trivial.
Example 5.12 (Universal cohomological properness of in characteristic 0).
If is an affine group scheme of finite type over an algebraically closed field of characteristic , then is cohomologically proper. To see this, let be the unipotent radical of so that is linearly reductive. Since has vanishing higher coherent cohomology, it is cohomologically proper. Therefore, it suffices to prove that is cohomologically proper. By flat base change and descent, it remains to prove that the base change (of by ) is cohomologically proper but this follows from Example 5.11.
More generally, is universally cohomologically proper. In particular, it follows that if is an -adically complete noetherian -algebra, then satisfies formal functions (Corollary 4.8). To see this, let be a noetherian -algebra. We first observe that has finite cohomological dimension. Since is affine, we may choose an embedding . This induces a morphism of algebraic stacks such that the base change by is the quotient , which we know has finite cohomological dimension. Let be the cohomological dimension. If is a coherent -module, then there is a -module and a surjection . Indeed, the morphism is affine, so the adjunction is surjective. We may write , where each is a vector bundles on . Since is coherent on , there is a sufficiently large such that for all . Now take . Letting be its kernel, there is a long exact sequence
By flat base change, we know that , which is a finite -module, so is a finite -module. Since this holds for any coherent -module , we may perform descending induction to conclude that is a finite -module for all .
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 of noetherian algebraic stacks is universally submersive if for every morphism from a scheme, the base change is surjective and has the quotient topology. This is equivalent to requiring that for every morphism from a DVR there is an extension of DVRs and a lift of the composition .
The most important examples for us are when is representable, proper, and surjective, or when 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 but the higher fiber products
for all .
Theorem 6.1.
Let be a representable, universally submersive and finite type morphism of noetherian algebraic stacks.
-
(1)
Suppose that is of finite type over a noetherian ring . If is cohomologically proper over for , then so is .
-
(2)
Let be a closed substack and be its preimage. If satisfies formal functions for , then so does .
-
(3)
Let be a closed substack defined by a coherent sheaf of ideals and let be its preimage. If is coherently complete and satisfies formal functions for , then is coherently complete and satisfies formal functions.
If is proper and surjective, then in each case one only needs to require the properties for .
Remark 6.2.
Proof.
We will argue (1) and (2) by noetherian induction on the abelian category and we may assume that every proper closed substack is cohomologically proper over and that the pair satisfies formal functions. For (3), note that satisfies formal functions by (2) and the functor is fully faithful, with image stable under kernels, cokernels and extensions, by Corollary 3.8. Thus it suffices to show that is essentially surjective. To accomplish this, we will also argue via dévissage on the abelian category and we may assume that if is a coherent sheaf on annihilated by for some non-zero coherent sheaf of ideals , then is in the essential image.
In each case, we may further reduce to the situation where is reduced. By generic flatness, there is a dense open substack such that is flat. By Rydh’s extension [Ryd16b] of Raynaud-Gruson’s theorem [RG71] on flatification by blow-ups, there is a commutative diagram:
where is a blow-up along a closed substack contained in the complement of , is the strict transform of along , and is flat. Observe that is a surjective, proper and representable morphism. Moreover, since is universally submersive, is surjective. Indeed, for every point there is a morphism from a DVR whose generic point maps to . As is universally submersive, there exists an extension of DVRs and a lift of . The induced map factors through the strict transform . The image of the closed point under is a preimage of and we conclude that is a faithfully flat morphism of finite type.
Since is separated and representable, its diagonal is a closed immersion and it follows that is also closed immersion for each . Since is proper, is proper and it follows that is proper and thus universally cohomologically proper (Example 5.8).
We now establish (1). The hypotheses that each is cohomologically proper over implies that is also cohomologically proper over . Since is faithfully flat, cohomological descent implies that for a coherent sheaf on , there is a convergent spectral sequence:
where denotes the pullback of under the projection . It follows immediately that is cohomologically proper over .
For a coherent sheaf on , consider the exact triangle of complexes on :
Since is proper and representable, and also have bounded coherent cohomology. Since is cohomologically proper over , we conclude (using Remark 2.4) that
Since is an isomorphism over , . Therefore in order to show that , it suffices to show that . By using the spectral sequence as in Remark 2.4, it suffices to show that if is a coherent sheaf on such that , then . But in this case is supported on a proper closed substack and the statement follows from the dévissage hypothesis.
For (2) and (3), we also establish the notation that the closed substacks and denote the preimages of . For (2), since is universally cohomologically proper and satisfies formal functions for each , it follows from Lemma 5.2, Theorem 5.3(1) and Example 4.4 that also satisfies formal functions.
Since is faithfully flat, cohomological descent implies that for each coherent sheaf on , we have a morphism of convergent spectral sequences:
where is the completion of along and is the completion of along . Since satisfies formal functions, so does .
For a coherent sheaf on , we again consider the exact triangle:
which induces a morphism of exact triangles
Since satisfies formal functions, the second vertical map is an isomorphism. Since , the third vertical map is also a isomorphism. Indeed, we can reduce as above to the case that is a coherent sheaf supported in cohomological degree 0 in which case is supported on a proper closed substack and we may apply the dévissage hypothesis. Thus is a an isomorphism and satisfies formal functions.
For (3), let be a coherent sheaf on . By coherent completeness of , there is a coherent sheaf on and an isomorphism . Denoting by the pullback , we have that completion of along is identified with . Letting and denote the two projections , then because is the pullback of a coherent sheaf on we have an isomorphism
of coherent sheaves on the completion of along the preimage of satisfying the cocycle condition on the completion of along the preimage of . As the morphisms are universally cohomologically proper for , satisfy formal functions. It follows that there is a unique isomorphism of coherent sheaves on satisfying the cocycle condition on such that . By faithfully flat descent, there is a coherent sheaf on such that there is an isomorphism compatible with . Since the coherent sheaves and are described via the same formal descent data with respect to , there is an isomorphism . Theorem 5.3(2) now gives equivalences and thus is in the essential image of . As is an isomorphism over , the adjunction morphism
is an isomorphism over . We claim that the kernel and cokernel of this adjunction morphism are annihilated by some power of sheaf ideals defining the reduced complement of . Indeed, this can be checked on a smooth presentation . Let be the ideal defined by the preimage of in , let be the -adic completion of , and let be endowed with the (proper) projection morphism . The module of sections of over is a finite -module while the module of sections of over is , where is the completion of along the preimage of . Moreover, the adjunction morphism corresponds to the morphism
and both the kernel and cokernel are annihilated by some power of the ideal of sections of over because is an isomorphism over the open subset . We know that the essential image of is stable under kernels, cokernels and extensions. Since is in the essential image and both the kernel and cokernel of are in the essential image by the dévissage hypothesis, we conclude that is in the essential image. ∎
Corollary 6.3.
Let be a universally submersive and representable morphism of noetherian algebraic stacks of finite type over a noetherian -adically complete ring . Let be a closed substack and be its preimage. Assume that and that . Suppose that for each , is cohomologically proper over and that there is a closed substack such that and such that the pair is coherently complete and satisfies formal functions. Then is cohomologically proper over and the pair is coherently complete and satisfies formal functions.
If is surjective, proper, and representable, then in each case one only needs to require the properties for .
Proof.
The corollary follows directly from Theorem 6.1 using Lemma 5.6 to deduce that the pair is coherently complete and satisfies formal functions. ∎
7. -actions and destabilizing one-parameter subgroups
After reviewing properties of -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 be a connected, smooth, and affine group scheme of finite type over a noetherian ring . Let be an affine scheme over with an action of . For any one-parameter subgroup , we define the subfunctors
of as introduced in [Dri13]. These functors are represented by closed subschemes of . Indeed, the -action on induces a -grading . Let and denote the ideals generated by homogeneous elements in strictly negative and strictly positive degree; then one can check that and are represented by the closed subschemes and , respectively. If is an algebraically closed field over , then the -points of are the -fixed points the -points of can be described as
The inclusion has a natural retraction given by
Remark 7.1.
If is integral, then and may be non-reduced and reducible. For example, if acts on with weights , then and . If is integral, it can also happen that is a finite set with more than one point; e.g., if with weights , then .
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 be a connected, smooth, and affine group scheme of finite type over a noetherian ring . Let act on itself via conjugation. Then a one-parameter subgroup induces an action of on (again via conjugation). Consider the following subgroups of :
(parabolic of ), | ||||
which extend naturally to subgroup functors of . When it is necessary to record the group , we write , , and . These functors are represented by closed subgroup schemes of . Indeed, observe that with the notation introduced in §7.1, we have that and , and that is identified with the kernel of defined by . See also [CGP15, Lems. 2.1.4 and 2.1.5].
There is a split exact sequence
(7.1) |
Over a field, the groups , and are well-known to satisfy several nice properties. If is smooth (resp. connected), so is and . The group is unipotent. If is connected and reductive then so is [Con20, Thm. C.2.1]. Moreover, in this case, is the unipotent radical of and is projective [CGP15, Prop. 2.2.9]. In general, if is a split reductive group scheme over a noetherian ring and is a one-parameter subgroup, then , , and are closed subgroup schemes of smooth over , is projective over and is reductive [Con14, Thm. 4.1.7, Ex. 4.1.9, Cor. 5.2.8].
If acts on an affine scheme of finite type over , there are natural actions of and on and respectively. The evaluation map is equivariant with respect to , and induces a morphism 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 on an affine scheme over a noetherian ring is given by a grading . The action is positively graded (resp. semipositively graded) if for and is finite over (resp., if for ). A representation of is positively graded (resp., semipositively graded) if for (resp., if for ).
For the action of an algebraic group on an affine scheme over and a one-parameter subgroup , we say that is -positive (resp. -semipositive) if the -action on induced by is positively graded (resp. semipositively graded).
Lemma 7.3.
Let act on an affine scheme of finite type over a field .
-
(1)
The action is semipositively graded if and only if .
-
(2)
The action is positively graded if and only if is finite over .
Proof.
The first part is clear from the definitions. For the second, observe that the condition that for implies that . ∎
The following is our key example.
Example 7.4.
If is an algebraic group acting on an affine scheme of finite type over and is a one-parameter subgroup, then is always -semipositive. Moreover, is -positive if and only if is finite over .
Lemma 7.5.
Let be an affine group scheme of finite type over a noetherian ring . Let and be affine schemes of finite type over with actions of . Let be a one-parameter subgroup. If is -positive with and is -positive (resp. -semipositive), then is -positive (resp. -semipositive) and .
Proof.
Let and be the gradings induced by the -action. Then is -semipositive with
It follows that . ∎
Example 7.6.
Let be a reductive algebraic group over an algebraically closed field . If is a one-parameter subgroup, then the -grading on is positively graded and . To see this, choose a maximal torus containing . Since the root system is reduced [Con20, Cor. 2.2.1], there is a decomposition as schemes [CGP15, Cor. 3.3.12], where the product is over all root groups with satisfying . Each as group schemes over , hence is a polynomial ring in the variables indexed by , where the -weight of is [CGP15, Prop. 2.1.8]. Therefore, the -weight of a monomial is strictly positive unless . It then follows from Lemma 7.5 that the -grading on is also positively graded for and that .
More generally, if is a split reductive group scheme over a noetherian ring and is a one-parameter subgroup contained in a maximal torus , then is -positive with .
7.4. Regular one-parameter subgroups
If is a torus over a noetherian ring , there is a perfect bilinear pairing between the character lattice and the lattice of one-parameter subgroups (or cocharacter lattice)
(7.2) |
given by .
Recall that if is a maximal torus of , the root system of the pair is the subset consisting of non-trivial weights for the adjoint action of on .
Definition 7.7.
A one-parameter subgroup is regular with respect to if for all .
Remark 7.8.
When and is the diagonal torus in , then the Lie algebra is the vector space of matrices. The basis element with a 1 in position and 0 elsewhere has weight , where denotes the character defined by . Therefore a one-parameter subgroup is regular if and only if the ’s are distinct.
Lemma 7.9.
Let be a connected, smooth, and affine group scheme over a noetherian ring . Let be a maximal torus and a one-parameter subgroup of .
-
(1)
is regular with respect to if and only if .
-
(2)
If is reductive, then is regular with respect to if and only if is a maximal torus in .
-
(3)
If is another maximal torus containing the image of , then is regular with respect to if and only the same is true with respect to . In other words, the definition of regularity is independent of the maximal torus chosen.
Proof.
Since is smooth connected, so too are [Con14, Lem. 2.2.4] and . Therefore, the containment is an equality if and only if it is true on geometric fibers. In particular, we may assume that is an algebraically closed field. In this case, it suffices to prove that . Now
This proves that is regular if and only if . Moreover if is reductive then [Con20, Prop. 2.3.1], proving (1) and (2).
To prove (3), suppose that the image of is contained in some other maximal torus . By the preceding discussion, it suffices to show that when is an algebraically closed field. We have since is regular with respect to . On the other hand, all maximal tori of are -conjugate, so and hence is also regular with respect to . ∎
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 be a reductive algebraic group over an algebraically closed field . Let be an affine scheme of finite type over with an action of . For any point , there is a one-parameter subgroup such that exists and has closed orbit.
When is integral and has a unique closed orbit, it is possible to find a one-parameter subgroup destabilizing every orbit.
Proposition 7.11.
Let be a reductive algebraic group over an algebraically closed field . Let be an integral affine scheme of finite type over with an action of . Assume that there is a unique closed orbit -orbit . Then there exists a one-parameter subgroup such that for any point , there exists such that . In particular, and .
Proof.
Let be the function field of and let be the generic point. Letting be an algebraic closure, we consider the base change with the point induced from . By Proposition 7.10, there is a one-parameter subgroup such that exists and has closed -orbit. If is a maximal torus, then there exists an element such that . Since extension of scalars gives an isomorphism
there is a one-parameter subgroup with . Since the generic point is in the image of the proper map , this map is surjective. Moreover, every point has a representative in which specializes via 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 , there is a destabilizing one-parameter subgroup (unique up to conjugation by the unipotent radical ) with minimizing the normalized Hilbert–Mumford index , where is a fixed conjugation-invariant norm on . In [Hes79], Hesselink used Kempf’s optimal destabilizing one-parameter subgroup to show that admits a stratification into locally closed -invariant subschemes where each is a one-parameter subgroup and . For any -point there is an element such that is an optimal one-parameter subgroup for with . Since is integral, there is a generic strata for some , which is dense. Since is proper and its image contains the dense set , it must be surjective. The statement follows. ∎
Example 7.12.
Under the action on given by the standard representation, the origin is the unique closed orbit and the complement is a dense orbit. Kempf’s optimal one-parameter subgroup for is . 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 with sum , the one-parameter subgroup is regular and also destabilizes the generic orbit.
In fact, one can always find destabilizing one-parameter subgroups of a point such that is regular in the stabilizer of the limit . This is proven in Proposition 7.14, which relies on the following lemma.
Lemma 7.13.
Let be a reductive algebraic group over an algebraically closed field . Let be an affine scheme of finite type over with an action of and let . Let be a one-parameter subgroup such that has closed orbit. If is a one-parameter subgroup commuting with (e.g., and are contained in the same maximal torus of ), then for
Proof.
We first determine the pushout of the diagram:
(7.3) |
where the top map is the closed immersion defining the origin and the left map is the open immersion corresponding to the product of and the inclusion of the open point . Expressing the diagram -equivariantly, we can write , , and under the diagonal action . The -equivariant pushout is determined by the fiber product of rings
We can write with , where acts via . Identifying , then acts via the degree matrix . By generalities of pushouts of stacks (see [AHHR23, §4]), the pushout of (7.3) is a limit
of quotient stacks with affine transition morphisms.
The point and one-parameter subgroup define a morphism with and . Since commutes with , we also have a morphism giving a commutative diagram of solid arrows
Since is of finite type, the induced morphism factors through for , and thus defines the dotted arrow .
Now consider the map induced by the diagonal and the group homomorphism defined by . The composition defines a morphism such that 0 maps to and such that the induced map on stabilizers is given by . The statement follows. ∎
Proposition 7.14.
Let be a reductive algebraic group over an algebraically closed field , and let be an affine scheme of finite type over with an action of . For any point , there is a one-parameter subgroup such that has closed orbit and such that the induced one-parameter subgroup is regular.
Moreover, if there is a unique closed orbit and is integral, then there is a one-parameter subgroup such that for any point , there exists such that and such that the induced map is a regular one-parameter subgroup.
Proof.
Let be a one-parameter subgroup such that has closed orbit (Proposition 7.10). Let be a maximal torus containing the image of . Choose a one-parameter subgroup such that the composition is a regular one-parameter subgroup for . Lemma 7.13 implies that for .
The addendum follows from applying the main statement to the generic point as in the proof of Proposition 7.11. There exist a one-parameter subgroup , a finite extension , and such that and has closed orbit in , where is the lift of . Choose a maximal torus containing and a one-parameter subgroup such that the base change of to is regular in for . Lemma 7.13 implies that also destabilizes . It follows that any -orbit of a -point in contains a -point such that . Since the base change of to is conjugate to , the one-parameter subgroup is also regular. ∎
Remark 7.15.
One can also formulate this result as saying that the degeneration fan , introduced in [Hal14], has dimension at least the reductive rank of .
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 . By Theorem 6.1, we can replace with and with , and we can thus assume that is a field and is reduced. If denotes the connected component of the identity, then is finite and étale. Applying Theorem 6.1 reduces us to the case that is connected. Letting be the irreducible decomposition, then each is isomorphic to with . The morphism is a proper and surjective, and by applying Theorem 6.1 again, we are further reduced to the case that is an integral domain.
Since has a unique closed point, Proposition 7.14 and limit methods imply that there is a one-parameter subgroup where is a finite field extension such that and such that for any closed point , the limit has closed orbit and the induced map is a regular one-parameter subgroup. By Theorem 6.1, we may replace with and thus we can assume that is defined over . We can further assume that the unique closed point lifts to a -point of .
We consider the composition
The morphism is proper (since is a closed subscheme and is projective) and surjective (since ). By applying Theorem 6.1, we are reduced to showing that satisfies formal functions and is coherent complete. Since is proper, the pushforward is coherent and it follows that is finite over (where is shorthand for ).
On the other hand, the morphism is faithfully flat and of finite type. The higher base changes of can be computed as
(8.1) |
Since the composition is the identity, the map is injective and thus is also finite over . But since , we can conclude that is finite over . Since is reductive, the map is an adequate moduli space. The stack has finitely many closed points with each contained in . By construction, each one-parameter subgroup is regular. It follows that the stabilizer is a maximal torus of , and in particular linearly reductive. By [AHR19, Thm. 4.21], is a good moduli space. Thus is coherently complete and satisfies formal functions with respect to , and therefore also coherently complete and satisfies formal functions with respect to (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 . Since (Example 7.6), the global sections of are identified with the global sections of (Lemma 7.5). Thus, 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, is coherently complete and satisfies formal functions with respect to the preimage of , and as we’ve already observed this implies that is coherently complete and satisfies formal functions with respect to . ∎
Remark 8.1.
When has a -fixed point , 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 is a field, is connected, and is integral. Proposition 7.11 implies that after replacing with a finite extension, there exists a one-parameter subgroup such that and (set-theoretically). Let be a maximal torus containing and be a Borel containing . In the composition
the morphism is proper and surjective, and is faithfully flat. The higher base changes of can be computed as
where .
Since is -positive with finite over (Lemma 7.3) and (Lemma 7.3), the global sections of are also finite over (c.f. (Lemma 7.5)). It follows that for each , the higher base changes admit good moduli spaces that are finite schemes over , and that the higher base changes are coherently complete and satisfy formal functions with respect to the preimage of . The coherent completeness and the formal functions of follow from applying Theorem 6.1 subsequently to and .
Remark 8.2.
Standard reductions show that 1.1 for all and follows from the case where is an integral domain and . Applying the method of Theorem 1.2(1) to the fraction field of and replacing by a finite extension, it follows that it suffices to prove 1.1 when 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 be a smooth geometrically reductive group scheme over a noetherian ring .
-
(1)
The morphism is cohomologically proper.
-
(2)
If in addition is -adically complete for an ideal , then the pair satisfies formal functions and is coherently complete.
Proof.
Since is smooth and geometrically reductive, the connected component of the identity is a reductive group scheme over and is finite over . By applying Theorem 6.1 to the finite cover , we are reduced to the case that is reductive. After an étale cover of , becomes split reductive. Applying the faithfully flat version of Theorem 6.1, we are further reduced to showing that if is split reductive with maximal torus and Borel subgroup . Now consider the composition
Since is projective over , the map is proper. On the other hand, is faithfully flat and the higher base changes can be described as
where is the unipotent radical arising as the quotient of the left action by and acts diagonally on via the right action. Since (Example 7.6) and thus (Lemma 7.5), the map is a good moduli space for each . Both statements follow by applying Theorem 6.1 to and then . ∎
8.2. Positively graded group schemes
Let be a smooth affine group scheme over a complete noetherian local ring with residue field . Let be the unipotent radical of and its reductive quotient. Let be a one-parameter subgroup.
Definition 8.4.
We say that is positively graded with respect to if the exact sequence splits, the conjugation action on is -positive, and is central in .
Remark 8.5.
If is defined over a field, then the conjugation action on is -positive if and only if the conjugation action of on is positive, and is central in if and only if the adjoint action of on is trivial.
Since the -action on is positive and is central in , the parabolic must be all of , and the short exact sequence is identified with the sequence .
We can now prove Theorem 1.3: if is positively graded with respect to , then
-
(1)
is cohomologically proper over , and is coherently complete and satisfies formal functions, and
-
(2)
if acts on an affine scheme of finite type over such that , acts semipositively on and satisfies 1.1, then is cohomologically proper over and satisfies formal functions and is coherently complete along its unique closed point.
Proof of Theorem 1.3.
Let be a group homomorphism splitting the surjection . For (1), we consider the the morphism
induced from the splitting. The map is faithfully flat and there is an isomorphism
where acts diagonally via the conjugation action. We know that acts positively on with (Example 7.6), and that the same holds for the -action on the higher fiber products (Lemma 7.5). In particular, are good moduli spaces. Since is cohomologically proper (Theorem 1.2(2)) and is coherently complete and satisfies formal functions, Theorem 6.1 implies that the same holds for .
For (2), we first claim that has a unique closed point . We will use essentially the same argument as in the case that is linearly reductive. Let . It suffices to show that if are closed substacks, then each is closed and . Let , be the -invariant ideals defining and . Since is -positive, we have . Thus, the image of is identified with the image of the closed -invariant ideal under , and it follows that is closed by the reductive case. Similarly, since since , , and are -positive, we have
where the middle equality used the reductive case. The claim follows.
We now consider the morphism and use the following identification of the higher fiber products
Since acts positively on with and acts semipositively on , the group of global sections of each higher fiber product is identified with by Lemma 7.5. As is reductive, each higher fiber product is an adequate moduli space over . It follows from Theorem 6.1, 1.1 and Corollary 4.9 that is cohomologically proper, and is coherently complete and satisfies formal functions with respect to . ∎
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 , 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 . 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.