<<

arXiv:2006.02315v3 [math.AG] 23 Feb 2021 nutvl,tefunctor the Intuitively, yashm,a rvni [ in proven as scheme, a by loehr eoti h diagram the obtain we Altogether, etr rjc 072//T/05 n nttt fMath of Institute and 2017/26/D/ST1/00755 project Center, proper smooth a its atfiaino the of pactification xlcty o a for Explicitly, emtial nerlafieagbacmni ihzr a zero with monoid algebraic affine integral geometrically a scheme hntligaota about talking When opcicto.Teeautosa eoadoeetn om to extend one and zero at evaluations The compactification. rcesi he steps three in proceeds { f ∞ 7→ iue1 Figure G IŁNCIBRL EOPSTO O EUTV RUSI P IN GROUPS REDUCTIVE FOR DECOMPOSITION BIAŁYNICKI-BIRULA Let ooti h lsia oiie(ep eaie Białynic negative) (resp. positive classical the obtain To nttt fMteais nvriyo Warsaw, of University , of Institute Warsaw, of University Mathematics, of Institute } k obt,wietetaseslbakcreis curve black transversal the while -orbits, pit are -points ) • • f epciey o once ieryrdciegroup reductive linearly connected a For respectively. k ( h map the the s 0 eafil.B a By field. a be X omalrecasta otisalmniswt eutv u reductive with monoids [ from all statements contains existence the that extends class large a form A X ) BSTRACT over htsends that ∈ ii map limit h aua asascae to associated maps natural The – X G k k G t iłnciBrl eopsto sastvle func -valued a is decomposition Białynicki-Birula its , i -scheme epoeteeitneo iłnciBrl decompositio Białynicki-Birula of existence the prove We . X iemaps give X euvratmorphisms -equivariant : π n pairs and G X π X ϕ G + obtof -orbit ( X iergroup linear 0 ϕ X X ato eawy enaleft a always we -action : → : ) + + S S X OCI EIIJWADŁKS SIENKIEWICZ ŁUKASZ AND JELISIEJEW JOACHIM ( X + hc eedwwt h trivial the with endow we which → parameterizes JS19 S X ) htsends that → π s ( + X : X G X = f and ] ( G ( X k , 1  ) G G otecntn family constant the to G ) ϕ → X ( = ) and htsends that ema mohafiegopshm ffiietp over type finite of scheme affine smooth a mean we : CHARACTERISTIC G X G .I 1. X JS19 + ϕ × ( [email protected] f G ssot o smooth for smooth is [email protected] π k s to NTRODUCTION G ( , X ) S X 0 m AHR20 f obt in -orbits ) and , ϕ → : A X | sa“ii”o h otdgnrt on fthis of point degenerate most the or “limit” a is G 1 X ϕ S 1 × + 1 X mtc,Pls cdm fSciences. of Academy Polish ematics, + S X to n h hc lecrei h ii of limit the is curve blue thick the and h lecre ntemdl eoethe denote middle the in curves blue The . ]. = → | : + ϕ S ϕ ( G i | k G X is X → 0 X m ) × ato.Let -action. h restrictions The . G S { ∪ → ϕ hc opciyto compactify which iBrl eopsto [ decomposition ki-Birula X : X -equivariant i ru nalcaatrsis This characteristics. all in group nit . = S dwt ntgroup unit with nd . 0 X G . } → ϕ G p ffunctors: of aps ) ato,tefntri ie by given is the -action, G 0 ( and i X X k ◦ h functor the , X upre yPls ainlScience National Polish by supported , ) for n h ro frepresentability of proof The . G pr h map The . G h ii a a section a has map limit The . 2 ( G : ealna ru and group linear a be ep monoids Kempf

G , . G tor × ( = ) f S X 7→ → + X f G G + : saprilcom- partial a is G = obt.Indeed, -orbits. X i o fixed a For . X f m which , srepresented is . BB73 ( ( Map , ϕ 1 A OSITIVE ) ) 1 ∈ etake we ] ( S = G . X , G X G and m ) G be G ∪ k - . . 2 JOACHIM JELISIEJEW AND ŁUKASZ SIENKIEWICZ

(I) introduce a formal version Xˆ of the functor X+ and prove its representability. The stage for this part is the formal neighbourhood of XG, hence the question becomes essentially affine and the of G plays a central role, (II) prove that the formalization map X+ → Xˆ is an for affine schemes, (III) for a G-scheme X find an affine G-equivariant étale cover of fixed points of X and use an easy descent argument to show that the natural map X+ → Xˆ is an isomorphism. The existence of such a cover is proven in [AHR20], which crucially depends on the linear reductivity of G, see[AHR20, Prop 3.1]. In positive characteristic the only connected linearly reductive groups are tori. Thus it is a natural question whether one could extend those existence results to reductive groups in positive char- acteristic. This seems also interesting from the point of view of geometric representation theory, similarly to how the Gm-case is used in [DG14]. The aim of the present article is to prove that X+ is representable for a large class of algebraic monoids G with zero, so called Kempf monoids. An algebraic monoid G with zero is a Kempf G monoid if there exists a central one-parameter subgroup m,k → Z(G) such that the induced map G → G extends to a map A1 → G that sends 0 to 0 . We prove that every monoid G with m,k k G zero and with reductive unit group is a Kempf monoid. We stress that we make no assumptions on the characteristic. Reductive as usual that the unipotent radical of G is trivial. There are plenty of Kempf monoids with non-reductive unit group as well, such as the monoid of upper- triangular matrices. The main result of this paper is the following representability result. Theorem 1.1. Let G be a Kempf monoid with zero and with unit group G and X be a Noetherian G- scheme over k. Then the functor X+ is representable and affine of finite type over XG. This directly extends the previous results for the one-dimensional torus [Dri13, AHR20] and linearly reductive groups [JS19, AHR20]. This extension is particularly far-reaching in positive characteristic, where tori are the only geometrically connected linear groups. It also clarifies the situation in general in that the complicated representation theory for G poses no obstructions to representability. In its ideas, the proof proceeds along the steps (I)-(III) above. However, there are two funda- mental problems along the way: • the representation theory for G is complicated and thus step (I) requires much more care. We introduce finitely generated Serre subcategories of representations and heavily em- ploy the Kempf torus (G ) inside G and the corresponding A1 inside G . m k k k k • the analogue of (III) is not known and it is clear that the current ideas are insufficient (the problems arising are similar to the ones with smoothness discussed below). We overcome ˆ this by employing Tannakian formalism of Hall-Rydh [HR19] to get a map iXˆ : X → X mimicking the unit map iX. In this way, Xˆ becomes a G-scheme with a map to X. This induces a section of the formalization map X+ → Xˆ and implies that it is an isomorphism. The language of stacks is most appropriate for (III). We delegate this part to the appendix in order to make the paper more accessible. The Tannaka duality in the required generality follows from the results of [HR19]. However [HR19] needs to be applied with care: the main results of that article require additional assumptions, so to deduce our claim we go into details of their argument. In the smooth case, we can say a little more about the πX. G + G + Proposition 1.2. Let x ∈ X be such that πX : X → X is smooth at sX(x) ∈ X . Then locally near the point x, the map πX is an affine space fiber bundle with an action of G fiberwise. It would be desirable to have Proposition 1.2 for every smooth X, without any assumptions + m m m2 on X . However, this is still open. The main problem is that the x → x/ x may not have a G-equivariant splitting, so the regularity of x ∈ X does not immediately imply the regularity of x ∈ X+, see Example 2.16. This is the same issue which implies that XG may not be smooth for smooth G-schemes X. Indeed, for example SLp acting on itself by conjugation has fixed points µp, which is non-reduced. In general [FN77] shows that smoothness of fixed points BIAŁYNICKI-BIRULA DECOMPOSITION FOR REDUCTIVE GROUPS IN POSITIVECHARACTERISTIC 3 for actions on smooth varieties characterizes linearly reductive groups. Very curiously, it seems that very few examples of non-smooth fixed points are known and all known examples seem to have smooth underlying reduced schemes. Also, by Lemma 2.1 below and affineness of πX we G have XG = (X+)G = (X+) m , so we cannot hope for X+ to be smooth in general, as G -fixed k k k k m points would be smooth while XG can be singular. However, we do not have examples of smooth k + G X such that πX : X → X is not smooth.

ACKNOWLEDGEMENTS We thank Torsten Wedhorn and Andrew Salmon for enquiring about extending the Białynicki- Birula decomposition to positive characteristic that encouraged us to write this paper. We also thank Michel Brion and Andrew Salmon for helpful remarks on the preliminary version of this paper. We thank the anonymous referee for careful reading and pointing out several imprecise points.

2. MONOIDS AND AFFINE SCHEMES Throughout, we fix a base field k. We do not impose any characteristic, algebraically closed, perfect or other assumptions on k. An algebraic monoid is a geometrically integral affine variety G together with an associative multiplication µ : G × G → G that has an identity element 1 ∈ G(k). × The unit group of G is the open subset consisting of all invertible elements. We denote it by G or simply by G. The unit group G is open in G, so it is dense and connected. The monoid is reductive if G is reductive, i.e., Gk contains no nonzero normal unipotent subgroups. Every G-representation is assumed to be rational, that is, a union of finite dimensional subrepresenta- tions on which G acts regularly. A canonical reference for reductive groups in all characteristics is [SGA3], for a summary see [Dem65]. Great introductions to algebraic monoids are for exam- ple [Bri14, Ren05]. We remark that the affineness assumption on G is redundant for reductive G: it follows by descent from [Rit98, Lemma 2] that every geometrically integral variety over k with an associative multiplication and with dense, reductive unit group is affine.

G 2.1. Kempf monoids. Let G be a Kempf monoid with unit group G and m,k → G be the Kempf torus. The Kempf line is the corresponding map A1 → G. Let T be the image of G in G. This k m,k is a one-dimensional algebraic group and it is geometrically connected, since it is connected and has a rational point 1T, see [sta17, Tag 04KV]. Therefore, T is a torus; in particular it is linearly reductive.

Lemma 2.1. Let X = Spec (A) be a Gk-scheme such that the action of the Kempf torus on X is trivial. Then the Gk-action on X is trivial.

Proof. Since 1G,0G are in the closure of Kempf’s torus, the assumption implies in particular that the actions of 1G and 0G are equal. For every k-scheme S and S-points x ∈ X(S) and g ∈ G(S) we have gx = (g · 1G)x = g · (1G · x) = g · (0G · x)=(g · 0G) · x = 0G · x = 1G · x, whence the action is trivial. 

N Corollary 2.2. Let X = Spec (A) be a G-scheme. Consider the -grading on Ak associated to the Kempf torus. Then (A )> is the ideal of XG in X . k 0 k k

Proof. Since G acts trivially on XG, also the Kempf line acts trivially, so I(XG) ⊃ (A )> . Con- k k k 0 versely, the ideal (Ak)>0 is Gk-stable since the Kempf torus is central, so it is Gk-stable and so Z = V((Ak)>0) ⊂ Xk is a Gk-scheme with a trivial action of the Kempf torus. By Lemma 2.1 also the G-action on Z is trivial, so Z ⊂ XG and so equality holds.  4 JOACHIM JELISIEJEW AND ŁUKASZ SIENKIEWICZ

2.2. Reductive monoids are Kempf. To prove the result in the title, we need a notion slightly weaker than an algebraic monoid. In this subsection a monoid M is an affine scheme of finite type with an identity and an associative multiplication (this notion is not used elsewhere in the article). Lemma 2.3. Let G be a semisimple group and G be a monoid with G contained in the unit group of G. Then G is closed in G. Proof. Since G is affine of finite type, it has a faithful finite-dimensional representation. We fix such a representation V and the associated embedding i : G ⊂ Endk(V). The group i(G) is an image of semisimple group, hence it has no non-zero characters so in particular det|i(G) is trivial and so we have i(G) ⊂ SL(V). The map i : G → SL(V) is a group , so its image is closed. Also SL(V)=(det = 1) is closed in End(V). Summing up, the group i(G) is closed in End(V) so also in G.  Corollary 2.4. In the setting of Lemma 2.3 assume moreover that G has a zero and that G ⊂ G is dense. Then G is the zero monoid.

Proof. Lemma 2.3 shows that G is closed in G. As it is dense, we have G = G. But then 0G ∈ G  is an invertible element and this happens only if G = {0G}. Proposition 2.5. Let G be a reductive monoid with zero. Let G be its unit group and let Z be the connected component of the identity in the center Z(G) with its reduced structure. Then for every point g ∈ G(k), the point 0G lies in the closure of Z · g. Proof. Since G is reductive, Z is an algebraic torus [SGA3, XII.4.11]. Consider the quotient variety

0 Z Q := G // Z = Spec H (G, OG)  and the quotient map π : G → Q. Since Z is a torus, this is a categorical quotient, so the map π ◦ µ : G × G → G → Q descends to a map µ¯ : Q × Q → Q which makes Q a monoid with zero. The map π and the inclusion G → G are both dominant, hence so is their composition, which implies that the subgroup π(G) ⊂ Q is dense. The group G/Z is semisimple [SGA3, XXII,4.3.5], so also its image π(G) is semisimple. Now the monoid Q satisfies assumptions of Corollary 2.4, so Q is a point. The points of Q = G // Z correspond to closed Z-orbits in G, so there is only one closed orbit and it is equal to {0G}. By general theory, the closure of every Z-orbit in G contains this closed orbit. This concludes the proof.  Corollary 2.6 (Reductive monoids are Kempf). Let G be a reductive monoid with zero. Then G is a Kempf monoid. Proof. Let Z := Z(G) be the connected component of identity in the center of the unit group G of G. Since G is reductive, Z is an algebraic torus. Let Z be the closure of Z in G. By Proposition 2.5 it contains the point 0G, which is a fixed point of the torus Z. Then the cone corresponding, in toric varieties sense, to the normalization of Zk is pointed, so any general one-parameter subgroup G → Z extends to A1 → Z that sends 0 to 0 .  m,k k k k G Remark 2.7. Upon completion of this paper, we learned that over an algebraically closed field the existence of Kempf torus is proven in [Rit98, Proposition 4]. We thank Michel Brion for the reference. 2.3. Representability for affine schemes: general setup. In this section we make no reductivity assumptions on G. Fix an algebraic monoid G with unit group G. Let V be a G-representation. We say that V is a G-representation if the G-action extends to a map G × V → V. If V is finite dimensional, then by the coaction map it is a G-representation if and only if there exists an em- 0 ⊕ dim V bedding of V into H (G, OG) .A G-representation that is a subrepresentation or quotient BIAŁYNICKI-BIRULA DECOMPOSITION FOR REDUCTIVE GROUPS IN POSITIVECHARACTERISTIC 5 of G-representation is a G-representation itself. Also the product of G-representations is a G-representation, but the dual of a G-representation is not necessarily a G-representation. Lemma 2.8. Let V be a G-representation. Then there exists a quotient V ։ W of G-representations such that (1) W isa G-representation, (2) for every other G-equivariant map V → W′ with W′ a G-representation there exists a unique factorization V → W → W′ where W → W′ is a G-equivariant map.

In other words, the quotient V → W represents the functor RepG → Set given by HomG(V, −). Proof. Consider first the case when V is finite-dimensional and view it as V = Spec Sym(V∨). ∨ 0 ∨ ∨ The coaction map becomes σV : V → H (G, OG) ⊗ V . Let U−1 := V and let U0 be the pullback defined by the diagram

0 ∨ U0 H (G, OG) ⊗k V

∨ σV 0 ∨ V H (G, OG) ⊗k V

Let Un for n = 1, 2, . . . be constructed inductively as the pullback

0 Un H (G, OG) ⊗k Un−1 (2.9)

0 Un−1 H (G, OG) ⊗k Un−2

Finally let U := Un. Diagram (2.9) implies that the coaction map on U factors as Tn 0 U → H (G, OG) ⊗k U. Therefore the G-representation W := Spec Sym(U) is in fact a G-representation. The inclu- -sion U ֒→ V∨ induces a surjective map V → W. Consider any other map V → W′ to a G representation; then replacing W′ by the image of V we may assume W′ is finite-dimensional, so W′ = Spec Sym(U′) for some i : U′ → V∨ and we obtain a of coactions σ ′ W′ 0 ′ U H (G, OG) ⊗k U

i id ⊗i

∨ σV 0 ∨ V H (G, OG) ⊗ V ′ ∨ It follows from the construction of {Un} that i factors as U → U → V . Consider now the case of a general (but, as always, rational) G-representation V. For every finite-dimensional subrepresentation V′ consider the corresponding quotient V′ → W(V′) and its kernel K(V′). For an inclusion of finite-dimensional subrepresentations V′ → V′′, the universal property of W(−) ′ ′′ ′ ′′ ′ gives a map W(V ) → W(V ) whence an inclusion K(V ) → K(V ). The union K = ′ K(V ) SV is a G-subrepresentation of V and V/K has the required universal property.  Proposition 2.10. Let X = Spec (A) be an affine G-scheme. Then X+ is represented by a closed sub- scheme.

Proof. Let A → A/G be the universal quotient G-representation as in Lemma 2.8. Let I ⊂ A be the ideal generated by G. The closed scheme Spec (A/I) is G-stable and in fact its G-action extends to G. We claim that X+ = V(I) ≃ Spec (A/I). Pick a G-scheme Y with a G-equivariant 0 map ϕ : Y → X. The map ϕ factors through the G-scheme Spec H (Y, OY), so we may assume 0 that Y is affine. The pullback A → H (Y, OY) maps A to a G-representation, hence maps G to zero, and thus factors through X+.  6 JOACHIM JELISIEJEW AND ŁUKASZ SIENKIEWICZ

Now we slightly generalize Proposition 2.10 taking into account open immersions. We say that a G-scheme X is locally G-linear if it is covered by G-stable open affine subschemes. We say that X is locally G-linear if it is covered by G-stable open affine subschemes. .Lemma 2.11 (open immersions, [JS19, Prop 5.2]). Let U ֒→ X be an open immersion of G-schemes Then the diagram π U+ U UG

π X+ X XG is cartesian. Sketch of proof. Arguing as in the proof of Proposition 2.5 using the Kempf torus, we see that the point 0 ∈ G has no G-stable open neighbourhoods except the whole G, so for every scheme S, the locus 0 × S ⊂ G × S has no G-stable open neighbourhoods except the whole G × S. Hence G G  ϕ : G × S → X factors through U if and only if ϕ|0×S : S → X factors through U . + G + Proposition 2.12. Let X be a locally G-linear scheme. Then πX : X → X is affine, so X is repre- sented by a locally linear G-scheme.

Proof. Let {Ui} be a G-stable open affine cover of X. By Lemma 2.11 the pullback of πX via G G + G Ui → X is the map πU : U → U . This map is affine by Proposition 2.10. It follows that the map π becomes affine after the pullback to the cover U → X, so π is affine.  X Fi i X Remark 2.13. Proposition 2.12 can be stated with weaker assumptions: we do not really need {U } to cover X, just the inclusion U ⊃ XG. i Si i 2.4. Representabilityfor affine schemes: reductive case. Proposition 2.10, while satisfactory for our general purposes, says little about the resulting scheme. In this section we prove that one can say more in the reductive case. Fix a reductive monoid G with unit group G. We additionally assume that the variety G is normal, however we do not need to assume that G has a zero. All maximal tori of G are conjugate. We fix one such torus T and its closure T ⊂ G. We say that a T-representation is a T-representation if the action of T extends to an action of T. A simple T-representation is an outsider representation if it is not a T-representation. Lemma 2.14 (Extension principle). Let V be a G-representation. Then the following are equivalent (1) the representation V is a G-representation, (2) the representation V is a T-representation. Proof. The implication 1 =⇒ 2 is trivial, and the converse is [Ren05, Theorem 5.2]. (the referenced theorem requires G to be normal and this is the main point where we use normality of G).  We now prove the representability of X+ for X affine. Proposition 2.15. Let X = Spec (A) be an affine G-scheme. Then X+ is represented by a closed sub- scheme whose ideal is the smallest G-ideal containing all outsider representations of T in A. Proof. Let us discuss the easiest case: X a G-scheme. Let µ : G × X → X denote the action. Then every family ϕ1 : S → X extends uniquely to a G-equivariant family ϕ = µ ◦ (id ×ϕ1) : G × S → + X, hence iX : X → X is an isomorphism. Let us return to the general case. Consider X as a T-variety and let X˜ be the Białynicki-Birula decomposition for T. As T is a torus, it is linearly reductive, so by [JS19, Proposition 4.5] the scheme X˜ is a closed subscheme of X. Let J := I(X˜ ) ⊂ A be its ideal and let I ⊂ A be the ideal generated by G · J. By[JS19, Proposition 4.5] the ideal J is generated by all irreducible T- subrepresentations of A that are not T-representations. From linear reductivity of T we deduce that A/I is a T-representation. BIAŁYNICKI-BIRULA DECOMPOSITION FOR REDUCTIVE GROUPS IN POSITIVECHARACTERISTIC 7

Let X′ = Spec (A/I). By construction, X′ is the largest G-scheme contained in X˜ . The co- ordinate ring A/I of X′ is a G-representation that is also a T-representation, hence by Exten- sion Principle 2.14 it is a G-representation, so X′ is a G-scheme. Every G-equivariant family ϕ : G × Spec (C) → X induces a G-equivariant pullback map # 0 ϕ : A → H (G, OG) ⊗k C. The right hand side is a G-representation, so a T-representation, so ker(ϕ#) contains J. The →֒ ′pullback is G-equivariant, so ker(ϕ#) contains I as well. This shows that ϕ factors through X X and consequently that X+ = (X′)+. But X′ is an affine G-scheme, hence the “easiest case” from the beginning of the present proof applies and gives X′+ = X′. 

Example 2.16 (Issues with regularity). Let x ∈ X = Spec (A) be a G-fixed k-point which we + identify with sX(x) ∈ X . Applying Proposition 2.15 we get that for every n ∈ N the truncated n,+ mn+1 n mn+1 local ring A := OX+,x/ x is the quotient of A := OX,x/ x by the smallest G-ideal containing all outsider representations of An with respect to T. m ։ m m2 If the group G was linearly reductive, we could take a G-equivariant section s of x x/ x and conclude that An,+ is the quotient of An by the image under s of the G-representation gen- m m2 ˆ erated by the T-outsider representations of x/ x. Consequently, the complete local ring OX+,x would be the quotient of Oˆ X,x by a sequence of elements whose images in the cotangent space are linearly independent. When x ∈ X is regular, we would conclude that x ∈ X+ is regular. This would be essentially the classical Iversen’s [Ive72] argument on the smoothness of fixed points of G for G acting on smooth X. However, for groups G that are not linearly reductive Iversen’s argument fails and so we cannot hope for the existence of section s in our case.

3. FORMAL G-SCHEMES In this section we prove the main results for the formal Białynicki-Birula functor Xˆ that we will introduce in the next section. The main advantage of the formal functor over X+ is that it is defined on the affine level; correspondingly in this section we speak the language of algebra rather than geometry. Throughout, we assume that G is a Kempf monoid and that 0 ∈ N.

3.1. Setup. We begin with basic definitions. A formal abelian group is a sequence (An)n∈N of ։ abelian groups together with surjections πn : An+1 An. A morphism of formal abelian groups ′ ′ (An) → (An) is a family of fn : An → An such that the diagram

πn An+1 An

fn+1 fn ′ ′ πn ′ An+1 An commutes for every n. We obtain an abelian of formal abelian groups. Using abstract- nonsense for this category, we can define formal algebras (as algebra objects), G-actions etc. Be- low we gather some specific cases. For a k-algebra B, we say that a formal abelian group is a formal B-algebra if all An’s are B-algebras and πn are maps of B-algebras. A over a formal B-algebra (An) is a formal abelian group (Mn) where additionally each Mn is an An-module and πn is a homomorphism of An+1-modules. We say that a formal B-algebra (An) is a G-algebra if all algebras An are G-algebras, the maps πn are G-equivariant of algebras and moreover the structure maps B → An are G-equivariant for the trivial G-action on B. Sim- ilarly we define G-algebras and G- or G-modules. In particular, whenever we speak about a G-action on a module (Mn) over a G-algebra (An), we assume that the multiplication maps An ⊗ Mn → Mn are G-equivariant.

Definition 3.1 (Formalization). Let A be an abelian group and (In ⊂ A)n∈N an increasing se- quence of subgroups. The associated formal abelian group is (A/In)n with the natural surjections. n+1 When A is a ring and In := I for an ideal I ⊂ A then the associated formal abelian group is a 8 JOACHIM JELISIEJEW AND ŁUKASZ SIENKIEWICZ

n+1 ring which we call the formalization of (A, I). When M is an A-module, then (M/I M)n is an n+1 (A/I )n-module called the formalization of (M, I).

Definition 3.2 (adic algebras and algebraizations). We say that a formal algebra A =(An) is adic if for every m ≥ n the map Am → An is surjective and n+1 ker(Am → An) = ker(Am → A0) . m+1 This implies in particular that ker(Am → A0) = 0 for every m. For a formal adic algebra (An) its algebraization is an algebra A with an ideal I such that the formalization of (A, I) is isomorphic to (An). A module (Mn)n over an adic algebra (An)n is adic if the maps Mn → Mn−1 are surjective and for every m ≥ n we have ker(Mm → Mn) = ker(Am → An)Mm so that the maps induce Mm ⊗Am An → Mn. Formalizations of algebras are adic. In the next sections we will prove that in certain situations “standard” adic algebras admit algebraizations.

3.2. Serre subcategories of G-linearized sheaves. We will be interested in constructing alge- braizations of a given formal algebra in an G-equivariant way (see Section 3.3). If G was linearly reductive, this would mean that we just look separately at each λ-isotypic component for a sim- ple G-representation λ. But the category of G-representations is far from semisimple, so working with irreducible representations makes little sense. We need a generalization of them. While in the linearly reductive case every finite dimensional representation is a direct sum of simple representations, here every representation has a filtration with simple subquotients. Lemma 3.3 (Jordan-Hölder). Let V be a finite dimensional G-representation. Then there exists a filtra- tion 0 = V0 ⊂ V1 ⊂ ... ⊂ Vr+1 = V with simple subquotients Ei = Vi+1/Vi. Theset {E0,..., Er} does not depend on the filtration chosen. The representations E0,..., Er are called the composition factors of V.

Proof. This follows from the Jordan-Hölder theorem. 

In the linearly reductive case, for a simple representation λ and a finite dimensional represen- tation V, one has the isotypic component V[λ] ⊂ V associated to λ: the largest subrepresentation that is a direct sum of λ’s. In our situation, for a set of simple G-representations λ = {λ1,..., λr} and a representation V we define V[λ] ⊂ V to be the largest subrepresentation whose composi- tion factors belong to {λ1,..., λr}. To put this idea on a solid footing we use Serre subcategories. Definition 3.4. Let C be an abelian category. A Serre subcategory S is a full subcategory closed under direct sums and such that for any short exact sequence

0 → V1 → V2 → V3 → 0 we have V2 ∈ S if and only if V1, V3 ∈ S. Lemma 3.5. Let S ⊂ C be a Serre subcategory and V ∈ C be an object. Then there exists a unique largest subobject of V that lies in S. We denote it by V[S]. For a morphism f : V1 → V2 in C the image f (V1[S]) lies in V2[S], so we get an induced morphism f [S] : V[S] → W[S]. For an exact sequence

0 V1 V2 V3 we get an exact sequence 0 → V1[S] → V2[S] → V3[S], so the functor (−)[S] is left exact. Proof. Consider the family {W } of all the subobjects of V that lie in S. Then W lies in S, i L i hence its quotient ∑ Wi ⊂ V lies in S. Take V[S] := ∑ Wi. For f : V1 → V2 the object f (V1[S]) is a quotient of V1[S], hence f (V1[S]) ∈ S, so f (V1[S]) ⊂ V2[S] by definition of V2[S]. Finally, ker(V2[S] → V3[S]) is in S as a subobject of V2[S] and is a subrepresentation of V1 since the  sequence is exact. Thus ker(V2[S] → V3[S]) ⊂ V1[S] which proves exactness. It is in general not true, even in our setup, that (−)[S] is right exact. In the following, we will use the two main examples of Serre subcategories. BIAŁYNICKI-BIRULA DECOMPOSITION FOR REDUCTIVE GROUPS IN POSITIVECHARACTERISTIC 9

Example 3.6. Fix a set λ = {λ1,..., λr} of simple G-representations and consider the full subcat- egory of RepG consisting of representations V that are unions of finite dimensional subrepresen- tations with composition factors in λ. This is a Serre category, which we denote by RepG[λ] and call the Serre category generated by λ. We denote by (−)[λ] the associated functor. For a G-representation V the inclusion of sets induces a partial order on the set of all λ’s which makes the set {V[λ]}λ a direct system and V its colimit. Remark 3.7. For a representation V, the subspace V[λ] is a subrepresentation, in particular it is rational. But also a stronger “global” rationality condition is satisfied: there exists a finite 0 dimensional k-linear subspace W ⊂ H (G, OG) such that the coaction map for any V[λ] has 0 image in W ⊗ V[λ] ⊂ H (G, OG) ⊗k V[λ]. Indeed, let W be the minimal subspace such that the coactions of all (finitely many!) simple representations from λ have image in W ⊗ (−). Then W satisfies the assumptions. Example 3.8. Let T ⊂ Z(G) be a central torus (which means that T ≃ Gm for some m, in k m,k particular T is linearly reductive). Let X be a finite set of simple T-representations and consider the full subcategory of RepG consisting of G-representations V which as a T-representation are direct sums of elements of X . This is a Serre category that we denote by RepG[X ]. We denote by (−)[X ] the associated functor. Remark 3.9. Since T is central, for every its simple representation χ and every G-representation V the isotypic component V[χ] is a G-subrepresentation, therefore so is V[X ] = V[χ]. (we Lχ∈X see here a clash of notation between isotypic components and Serre categories; fortunately both notations agree). The functor (−)[X ] is exact. The constructions (−)[λ] and (−)[X ] are connected as follows. For λ, let X (λ) be the set of all T-weights that appear in λ . For a G-representation V, we have V[λ] ⊂ V[X (λ)]. On the Lλi∈λ i other hand, if w ∈ V[X (λ)], then we may form µ(w) = {µ1,..., µs} where µi are composition factors of the representation generated by w. Since G · w ⊂ V[X (λ)], we have X (µ(w)) ⊂X (λ). By definition w lies in V[µ(w)], therefore we obtain

(3.10) V[X ] = [ V[µ]. µ : X (µ)⊂X

3.3. G-equivariant algebraizations. In this subsection we fix a torus T ⊂ Z(G), not necessarily G split. We assume that the Kempf torus ( m)k → G factors through T. Below we use the conven- tion that application of (−)[λ] precedes taking the limit, so that limn Mn[λ] := limn(Mn[λ]) and similarly for X .

Definition 3.11. Let (Mn) be a formal abelian group with a G-action. Its λ-algebraization is

lim(Mn) := colimλ lim Mn[λ], G n ։ ։ By Remark 3.7, limG(Mn) is a G-representation. It is the limit of the diagram . . . Mn+1 ։ ։ Mn ... M0 in the category of abelian groups with G-action; hence the notation. If A = (An)n is a formal algebra, then limG(A) is an algebra as well. Indeed for λ and µ let λ ⊗ µ denote the Serre subcategory generated by composition factors of all {λi ⊗ µj}λi∈λ,µj∈µ. Then the multiplication on an An restricts to An[λ] ⊗ An[µ] → An[λ ⊗ µ] and induces multiplication (limn An[λ]) ⊗ (limn An[µ]) → limn An[λ ⊗ µ] and thus a multiplication on limG(A). Remark 3.12. The reader might wonder what is the connection between algebraizations from Definition 3.1 and λ-algebraizations. We will see in Theorem 3.22 that in favorable conditions a λ-algebraization is an algebraization.

Definition 3.13. Let (Mn) be a formal abelian group with a T-action. Its X -algebraization is

lim(Mn) := colimX lim Mn[X ], T n where the colimit is taken over all finitely generated Serre subcategories of RepT. 10 JOACHIM JELISIEJEW AND ŁUKASZ SIENKIEWICZ

Lemma 3.14. Let 0 →M→N →P be an exact sequence of formal abelian groups with G-action (resp. T-action). Then the induced sequence 0 → limG(M) → limG(N ) → limG(P) is exact (resp., the induced sequence of limT(−) is exact). Proof. This follows from left exactness of (−)[X ] and (−)[λ] and the fact that colimits in both algebraizations are taken over filtered sets. 

Since RepT is semisimple we have a canonical isomorphism

lim(Mn) = lim Mn[χ], M n T χ where χ runs through all simple T-representations. It follows that

(3.15) (lim(Mn))[X ] = lim Mn[X ] T n ։ ։ for every set of simple T-representation X . The T-module limT(Mn) is the limit of . . . Mn+1 ։ ։ Mn ... M0 in the category of abelian groups with T-action. For a formal abelian group (Mn) ֒ with G-action, the restriction RepG → RepT induces a injective λ-X -comparison map:

(lim(Mn) ֒→ lim(Mn (3.16) G T that embeds Mn[λ] into Mn[X (λ)] and so limn Mn[λ] into limn Mn[X (λ)]. The reason to intro- duce both algebraizations is that λ-algebraization is more canonical — it comes with a G-action — but also more challenging to work with, mostly because [−](λ) is not right-exact. We will prove that the λ-X -comparison map is an isomorphism in situations of interest, see Corollary 3.21. For this we employ various stabilization results that we now prove. 3.4. Stabilization. We keep the setup from the previous subsection (Subsection 3.3). We say that a formal abelian group (Mn) with a G-action stabilizes G-equivariantly if for every λ the map Mn+1[λ] → Mn[λ] is an isomorphism for all n large enough. Similarly, we say that a formal abelian group (Mn) with a T-action stabilizes T-equivariantly if for every X the map Mn+1[X ] → Mn[X ] is an isomorphism for all n large enough. If (Mn) has a G-action and stabilizes T-equivariantly then it also stabilizes G-equivariantly. Indeed, for every λ the map Mn+1[X (λ)] → Mn[X (λ)] is an isomorphism for n ≫ 0 and so using (3.10) we deduce that also the map Mn+1[λ] = Mn+1[X (λ)][λ] → Mn[X (λ)][λ] = Mn[λ] is an isomorphism. We now give an instance where those stabilizations hold. Definition 3.17 (standard adic formal algebra). We say that a formal G-algebra A is a standard adic formal G-algebra if all of the following hold (1) A is an adic algebra (as in Definition 3.2), (2) A is an A0-algebra (which means that every An is an A0-algebra and An → An−1 are surjections of A0-algebras), G (3) Spec (An) → Spec (A0) is an isomorphism for every n. A formal algebra obtained from a G-scheme X = Spec (A) with I = I(XG) is standard adic. We will see in Section 4.2 that under good conditions standard adic algebras come from G-schemes. Let M =(Mn)n be a module over a standard adic algebra (An)n. If the module M is equipped with a T-action then we say that M is grounded if there exists a finite set of T-characters X such that M0 = M0[X ]. By Definition 3.17(3) for n = 0 the G-action on A0 is trivial, so the module M is grounded whenever M0 is a finitely generated A0-module. n Example 3.18. Let T be Gm and let A =(k[t]/t )n be equipped with a T-action coming from the 1 n n −1 2 n −1 standard grading. The A-modules M =(k[t]/t ⊕ (k[t]/t )t )n and M =( i∈N(k[t]/t )t )n 3 n −i L 3 are grounded, while M = N(k[t]/t )t is not grounded. The difference is that M has Li∈ n generators of arbitrarily negative weights, while the generators M2 are infinite, but their weights are bounded from below. BIAŁYNICKI-BIRULA DECOMPOSITION FOR REDUCTIVE GROUPS IN POSITIVECHARACTERISTIC 11

Definition 3.19 (stabilization indices). For a simple T-representation χ, we define nχ ∈ Z as the maximal weight of the Kempf torus in χk. We set nX = max(nχ : χ ∈X ) and nλ := nX (λ). For example, if χ is a trivial T-representation then nχ = 0.

Lemma 3.20 (T-equivariant stabilization lemma). Let A =(An) be a standard adic formal G-algebra. Let M =(Mn) be an adic A-module with a T-action. Assume that M is grounded. (For example, this is the case when M0 is a finitely generated A0-module.) Fix a finite set of characters X . Then there exists a natural number nX ,M such that for every n > nX ,M the surjection ։ Mn[X ] Mn−1[X ] is an isomorphism. If M = A then we may take nX ,M = nX . Proof. The claim is under field extensions, so we pass to the algebraic closure of k. We also immediately reduce to the case X = {χ}. Let m be the minimal weight of the Kempf torus on M0; such a weight exists since the set of weights is finite because M is grounded. We claim that nχ,M = nχ + |m| satisfies the assumptions. By Corollary 2.2 for every n the ideal ker(An → A0) consists entirely of positive weights. Since M is adic, we have M0 = Mn/(ker(An → A0)Mn). By the graded Nakayama lemma, the minimal weight of the Kempf torus on Mn is equal to m. We have a short exact sequence

0 → ker(Mn → Mn−1)[χ] → Mn[χ] → Mn−1[χ] → 0 so it is enough to prove that ker(Mn → Mn−1)[χ] = 0 for all n > nχ,M. Since A and M are adic, n we have ker(Mn → Mn−1) = ker(An → A0) Mn. If n > nχ + |m| then the minimal weight of n the Kempf torus in ker(An → A0) Mn is greater than nχ + |m| + m which in turn is greater than n or equal nχ. The isotypic component (ker(An → A0) Mn)[χ] is thus equal to zero, as claimed. If  M = A then m = 0 and so nχ,M = nχ. As discussed at the beginning of this subsection, for a G-module M its T-equivariant stabi- lization implies G-equivariant stabilization. Corollary 3.21. Let A be a standard adic formal G-algebra and let M be an adic G-module. Assume M is grounded. Then the λ-X -comparison map (3.16) for M is an isomorphism. Consequently, the map limG(Mn) → Mn is surjective for every n. Proof. The comparison map is always injective, so it is enough to prove surjectivity. Let m ∈ ′ ′ limT(Mn) and choose X such that m ∈ limn Mn[X ]. Let n := nX ,M. For n > n the map Mn[X ] → Mn−1[X ] is an isomorphism, so limn Mn[X ] → Mn′ [X ] is an isomorphism. Since T is central in G, the subgroup Mn′ [X ] is G-stable and so there exists a finite set of G-representations µ such that m ∈ Mn′ [µ]. But then X (µ) ⊂ X and by the discussion above, also limn Mn[µ] → Mn′ [µ] is an isomorphism, so m lifts to an element of limn Mn[µ] and so m is in the image of the λ-X -comparison map. The final claim follows because limT(Mn) → Mn is surjective by right- exactness of (−)[X ].  Theorem 3.22 (T-equivariant algebraization exists). Let A and M be as in Lemma 3.20. Let M = n+1 limT M and A = limG A. Then ker(M → Mn) = ker(A → A0) M for all n. Therefore M admits an algebraization (M,ker(A → A0)). In particular, A admits an algebraization (limG(A),ker(A → A0)).

Proof. Corollary 3.21 applies to A and shows that A = limT A. Let πn : A → An be the canonical M M maps and let In = ker(πn). Let πn : M → Mn be the canonical maps. The maps πn and πn are surjective by exactness of (−)[X ]. This implies that M n+1 n+1 n+1 πn (I0 M) = πn(I0 )Mn = ker(An → A0) Mn = 0 M n+1 because A and M areadic. We have proven ker(πn ) ⊃ I0 M. It remains to prove the other con- M tainment. Take an element m ∈ ker(πn ) and fix X such that m ∈ limn Mn[X ]. By T-stabilization ′ > for the module M, for any n nX ,M the map limn Mn[X ] → Mn′ [X ] is an isomorphism. This 12 JOACHIM JELISIEJEW AND ŁUKASZ SIENKIEWICZ

n+1 isomorphism maps the element m to an element of ker(Mn′ → Mn) = ker(An′ → A0) Mn′ . M Since πn′ and πn′ are surjective, we have M n+1 n+1 n+1 πn′ (I0 M) = πn′ (I0) πn′ (M) = ker(An′ → A0) Mn′ . M n+1 Since (−)[X ] is right-exact, we see that πn′ ((I0 M)[X ]) contains m. So we pick an element n+1 M M M j ∈ (I0 M)[X ] such that πn′ (j) = πn′ (i). Then πn′ (i − j) = 0 and by (3.15) we have i − j ∈ M  limn Mn[X ]. But πn′ restricted to limn Mn[X ] is an isomorphism, so i − j = 0. Proposition 3.23 (generating sets for grounded modules). Let A, M, A, M be as in Theorem 3.22. Let F be an A-module with a T-linearization and a T-equivariant homomorphismof A-modules p : F → M such that the associated map p : F0 → M0 is surjective. Then p is surjective.

n+1 Proof. Consider the map pn : Fn → Mn which is just pn = p ⊗A (A/I ). Let Gn = pn(Fn). Since n+1 p0 is surjective and M is adic, we have Mn = Gn + IMn. But I Mn = 0 so 2 n+1 Mn = Gn + IMn = Gn + I(Gn + IMn) = Gn + I Mn = ... = Gn + I Mn = Gn. Let G = p(F) ⊂ M. It is enough to prove that G[X ] = M[X ] for all X . By exactness, we have M[X ] = limn Mn[X ] for all X . By Stabilization Lemma 3.20, the map M[X ] → Mn[X ] is an isomorphism for n large enough and we obtain the following diagram

G[X ] M[X ]

≃ ≃ Gn[X ] Mn[X ] so G[X ] → M[X ] is surjective as well as injective. 

Corollary 3.24 (finitely generated modules). Let A be a standard adic formal G-algebra with an alge- braization (A, I) and let M =(Mn)n be an adic module over A with a T-action. Then, the following are equivalent

(1) the A0-module M0 is finitely generated, (2) every An-module Mn is finitely generated, (3) limT(M) is a finitely generated A-module. If these hold, we say that M is a finitely generated A-module.

Proof. 3. =⇒ 2. Since (−)[X ] is right-exact for every X , the map limT(M) → Mn is surjective n+1 so Mn is a finitely generated A-module. But M is adic, so I annihilates Mn and so Mn is a finitely generated An-module. The implication 2. =⇒ 1. is trivial and 1. =⇒ 3. follows from Proposition 3.23 since M is grounded by assumption and T is linearly reductive, so we can find a T-equivariant surjection  from a free A-module F with a T-linearization onto M0 and its T-equivariant lift to limT(M).

We are now ready to prove the main equivalence statement for grounded modules. Consider a standard adic formal G-algebra A = (An)n with algebraization (A, I = ker(A → A0)) as in Theorem 3.22, so that the algebra A is a formalization of A. Consider the two

adic,G G G adic,G Alg := lim(−) : Mod → ModA and Formalize: ModA → Mod , G A A n+1 G defined by Formalize(M)=(M/I M)n∈N and Alg(M) = limG(M), where ModA denotes adic,G the category of A-modules with a G-action and ModA of adic G-modules over A. Theorem 3.25 (Algebraization for grounded G-modules). The functors Alg and Formalize restrict to an equivalence between the category of finitely generated A-modules with a G-action and finitely gen- erated adic A-modules (in the sense of Corollary 3.24) with G-action. BIAŁYNICKI-BIRULA DECOMPOSITION FOR REDUCTIVE GROUPS IN POSITIVECHARACTERISTIC 13

Proof. Using Corollary 3.21 we identity λ- and X -algebraizations of finitely generated A-modules. By Corollary 3.24 the functors Alg and Formalize map finitely generated modules to finitely generated ones. It remains to prove that they give an equivalence. By Theorem 3.22, the composition Formalize ◦ Alg is isomorphic to identity. Consider Alg ◦ Formalize. Choose an A-module M with a G-linearization, let M = Formalize(M) and M′ = Alg(M) and consider the natural map M → M′. By Stabilization Lemma 3.20, the map M[X ] → M′[X ] is surjective for every X . Since every element of M′ sits inside some M′[X ], we get that M → M′ is surjective. By the same argument as in Stabilization Lemma, for every X the map M[X ] → (M/In+1M)[X ] is an isomorphism for n large enough, so M → M′ is injective. 

4. FORMAL BIAŁYNICKI-BIRULA FUNCTORS In this section we introduce the formal version of Białynicki-Birula functors. The notation and Z mn+1 general outline are consistent with [JS19, Section 6]. For an n ∈ ≥0 let Gn = V( 0 ) ⊂ G be the n-th thickening of the k-point 0 ∈ G. Consider the set-valued functor

Xˆ (S) = (ϕ ) Z | ϕ : G × S → X, ϕ is G-equivariant and (ϕ ) = ϕ for all n . n n n∈ ≥0 n n n n+1 |Gn×S n o

We have a natural formalization map X+ → Xˆ given by ϕ 7→ (ϕ ) . Eventually, we will prove |Gn×S n that it is an isomorphism. The crucial technical advantage of Xˆ over X+ is that it avoids the G topological issues altogether: the set-theoretic image of each ϕn lies in X .

4.1. Formal G-schemes and G-linearized sheaves. We introduce some notation for the formal neighbourhoods of fixed points in X. It is convenient to make this in a very general way.

Definition 4.1. A formal G-scheme consists of a sequence of G-schemes Z =(Zn)n∈Z≥0 and equi- variant closed immersions

Z0 Z1 ... Zn Zn+1 ... such that G ֒ (1) the G-action on Z0 is trivial and Z0 → Zn is an isomorphism, m+1 (2) if I ⊂OZn is the ideal sheaf defining Z0 ⊂ Zn, then for all m ≤ n the ideal sheaf I defines Zm ⊂ Zn.

A morphism of formal G-schemes f : (Wn) → (Zn) is a family of G-equivariant morphisms ( fn : Wn → Zn)n compatible with closed immersions Wn ֒→ Wn+1 and Zn ֒→ Zn+1. If every Zn is a G-scheme .and the closed immersions Zn ֒→ Zn+1 are G-equivariant then we say that Z is a formal G-scheme Similarly, a morphism of formal G-schemes is a morphism of formal G-schemes such that every fn is G-equivariant. If every Zn is locally Noetherian, we say that Z is locally Noetherian.

Remark 4.2. For a formal G-scheme Z = (Zn) and any m > n the radicals of ideal sheaves .of Zn ⊂ Zm and Z0 ⊂ Zm are equal, so Z0 ֒→ Zn is a homeomorphism of topological spaces Therefore for every n, the G-action on the |Zn| is trivial, so every Zn is locally G-linear.

Definition 4.3 (Quasi-coherent sheaves on Z). Let Z =(Zn)n be a formal G-scheme. The closed inclusions Zn ֒→ Zn+1 induce a sequence of restriction maps

(4.4) . . . QcohG(Zn+1) QcohG(Zn) . . . QcohG(Z0)

A quasi-coherent G-linearized sheaf on Z is a sequence (Fn ∈ QcohG(Zn))n together with isomor- phisms in : Fn+1|Zn → Fn.A morphism of quasi-coherent G-linearized sheaves ϕ : (F•, i•) → (G•, j•) 14 JOACHIM JELISIEJEW AND ŁUKASZ SIENKIEWICZ is a sequence of morphisms ϕn : Fn → Gn such that for every n the following diagram commutes

in Fn+1|Z = Fn+1 ⊗O OZ Fn n Zn+1 n

ϕn ϕn+1⊗id jn Gn+1|Z = Gn+1 ⊗O OZ Gn n Zn+1 n

These objects and morphisms form the category QcohG(Z) of G-linearized quasi-coherent sheaves on Z which is a symmetric monoidal category, that is a 2-categorical limit of the diagram (4.4). We say that a G-linearized sheaf F of Z is of finite type if there exists an affine open cover of |Z0| such that the restrictions of F to each member of this open cover are finitely generated modules in the ft sense of Corollary 3.24. We denote by QcohG(Z) the full subcategory of QcohG(Z) consisting of finite type sheaves. We have a natural formalization and — much subtler — algebraization in the G-equivariant setting, that are the geometric analogues of the ones from Section 3. Eventually we will use them to show that the formalization map X+ → Xˆ is an isomorphism. G Definition 4.5 (Formalization). Let X be a G-scheme and I⊂OX be the ideal defining X . The n+1 + formal G-scheme Zˆ associated to X is the sequence (Zn)n where Zn =(V(I )) is the Białynicki- Birula decomposition of the (n + 1)-th thickening of the fixed point set; the decomposition exists because the action of G on the topological space of V(In+1) is trivial, so this scheme is locally linear and Proposition 2.12 applies. The inclusions Zn → Z induce restrictions QcohG(Z) → QcohG(Zn) which together give a natural comparison functor QcohG(Z) → QcohG(Z) that is cocontinuous and tensor (see appendix for definitions) and preserves finite type sheaves. Definition 4.6 (Algebraization). Let Z be a formal G-scheme. An algebraization of Z is a G- scheme Z such that Z is isomorphic to the formal G-scheme associated to Z and the associated ft ft restriction functor QcohG(Z) → QcohG(Z) is an equivalence.

4.2. Algebraization, geometric counterpart. To a formal G-scheme Z =(Zn) we may associate the family of maps πn : Zn → Z0 which are multiplications by 0 ∈ G. In particular, πn+1|Zn = πn for all n. By Remark 4.2 and Proposition 2.12, the maps πn are affine, so we have a sequence of sheaves of quasi-coherent G-algebras on Z0:

An :=(πn)∗OZn ։ and surjections An+1 An. We call (An) the sheaf of G-algebras associated to Z. The category RepG[λ] behaves well when we pass from RepG to RepG and then to G-equivariant sheaves on a G-scheme Z0 with a trivial G-action. Indeed, fix a quasi-coherent sheaf A with a G-action and define A[λ] ⊂ A by setting H0(U, A[λ]) := H0(U, A)[λ] for every open U. Since (−)[λ] 0 is left exact, we obtain a subsheaf. For every open U the algebra H (U, OZ0 ) is a trivial G- 0 0 0 representation, so the structural map H (U, OZ0 ) ⊗k H (U, A) → H (U, A) descends to 0 0 0 H (U, OZ0 ) ⊗k H (U, A)[λ] → H (U, A)[λ]. and so A[λ] ⊂ A is a subsheaf of G-modules and OZ0 -modules; since the action of G on Z0 is trivial those structures commute. We can thus form λ-algebraizations of sheaves. For a sheaf of algebras A its λ-algebraization is a sheaf of algebras as well. If we fix a central torus T the same holds for (−)[X ] instead of (−)[λ].

Theorem 4.7 (algebraization of a formal G-scheme). Let Z = (Zn) be a locally Noetherian formal

G-scheme with associated sheaf of algebras (An). The scheme Z = Spec Z0 (limG(An)) is the colimit of Zn in the category of locally linear G-schemes and it is an algebraization of Z. Moreover, the scheme Z is G locally linear and πZ : Z → Z is affine of finite type, so Z is locally Noetherian.

Proof. Let A = limG(An). We also fix central Kempf torus Gm,k → Z(G) and its image T ⊂ Z(G). This is a one-dimensional torus, possibly non-split, see Subsection 2.1. By Theorem 3.22 applied BIAŁYNICKI-BIRULA DECOMPOSITION FOR REDUCTIVE GROUPS IN POSITIVECHARACTERISTIC 15

n+1 to sections of A, the map A → An is surjective and its kernel is equal to ker(A → A0) section- wise. Therefore indeed the formalization of Z is Z. Now we prove that Z = colim Zn in the category of locally linear G-schemes. Fix a locally G linear G-scheme W with coherent maps fn : Zn → W. In particular, we get a map f0 : Z0 → W . G ′ By Proposition 2.12 the multiplication by 0G map πW : W → W is affine. Let W := W ×WG ′ ′ Z0. The map W → Z0 is affine and G-equivariant. We have induced maps fn = fn × πn : Zn → ′ W over Z0. To sum up, we have the following situation and we want to find the dashed arrow Z ′ fn

f ′ n+1 ′ Zn Zn+1 W W

πn+ 1 aff aff πn f0 G Z0 W ′ All schemes Zn, Z and W are affine over Z0. Let B be the sheaf of G-algebras on Z0 such that ′ ′ W = Spec Z0 (B). Note that B = B[λ]. The maps fn induce compatible G-equivariant mor- # S # phisms fn : B → An. By G-equivariance, for every λ, they restrict to fn : B[λ] → An[λ] and give morphisms F[λ] : B[λ] → A[λ]. Composing those with A[λ] → A, we obtain F[λ] : B[λ] → A which in turn glue to F : B → A. # (Compatibility of any two F[λ1] and F[λ2] follows from compatibility of fn, because of stabiliza- tion.) This is a G-equivariant homomorphism of OZ0 -algebras and it gives the required map Z → W′ and in turn Z → W. By assumption the scheme Z is locally Noetherian; we will use it to prove that Z → Z0 is of finite type. The argument of [JS19, Theorem 6.8] generalizes verbatim; we repeat it for com- pleteness. By assumption the scheme Z0 is locally Noetherian. For every n, in the OZ0 -module An we have a coherent ideal sheaf Kn := ker(An → A0) and by definition of formal schemes we n n n−1 have Kn = ker(An → An) = 0, so 0 = Kn ⊂ Kn ⊂ ... ⊂ K ⊂ An is a finite filtration whose quotients are coherent OZn -modules that are in fact OZ0 -modules. Thus the OZ0 -module An is coherent as well. Section-wise application Stabilization Lemma 3.20 implies that for every λ the 2 OZ0 -module A[λ] is coherent. The A-module J0/J0 ⊂ A2 is coherent, so we can fix λ and a ⊕r ⊕r 2 map A[λ] →J0 that restricts to a surjection A[λ] →J0/J0 . We claim that A is generated as ⊕r an OZ0 -algebra by the image of A[λ] . To prove this, we pass to k and use the grading induced by the Kempf torus. By Stabilization Lemma for the trivial representation, the degree zero part of A is indeed A0 = OZ0 . Therefore J0 is positively graded and the result follows from graded Nakayama lemma. ft ft It remains to prove that QcohG(Z) → QcohG(Z) is an equivalence, but this follows from Theorem 3.25.  4.3. Main proofs. In this short section we apply the results of the previous one to prove Theo- rem 1.1. Proof of Theorem 1.1. We begin by constructing a formal G-scheme and its algebraization. Let X n+1 G be a Noetherian G-scheme and Xn = V(I ) where I ⊂OX is the sheaf defining X , so G + X0 = X . Let Zn = Xn be the Białynicki-Birula decomposition, it is a closed subscheme of Xn, see Definition 4.5. Let Z = (Zn) and let Xˆ be the algebraization of Z as in Theorem 4.7. The scheme Xˆ has a natural map Xˆ → XG, but a priori lacks the “inclusion of cells” map Xˆ → X. In fact, the construction of this inclusion map is subtle topologically, as Xˆ is obtained from a formal neighbourhood of XG and does not see directly the topology of X. We now construct this map using the formalism of Tannaka duality, which is described in the appendix. 16 JOACHIM JELISIEJEW AND ŁUKASZ SIENKIEWICZ

ft ˆ ft By Theorem 4.7, we have QcohG(X) = limn QcohG(Zn). The pullback via closed inclu- ft ft ֒ ֒ sions Zn → Xn → X gives a functor F : QcohG(X) → limn QcohG(Zn). The composition ft ft ˆ F : QcohG(X) → QcohG(X) is a cocontinuous tensor functor. Since X is Noetherian, each G- linearized quasi-coherent sheaf on X is a filtered limit of coherent G-linearized sheaves, see for example [sta17, Tag 07TU]. The category of quasi-coherent sheaves of Xˆ with G-linearization ad- mits all limits, so the functor F extends to a functor QcohG(X) → QcohG(Xˆ ) that we also denote by F. Let p : X → Spec (k) and p : Xˆ → Spec (k) be the structure maps. Then F ◦ p∗ and p∗ are X Xˆ X Xˆ isomorphic as this holds coherently for all Zn. By Theorem A.1, we obtain a G-equivariant map I : Xˆ → X ˆ such that for every n the map I|Zn is the inclusion of Zn. The G-scheme X is a scheme over X via the equivariant map I, so we have a family Φ : G × Xˆ → X. It remains to prove that it is universal. For every other morphism ϕ : G × S → X we have restrictions ϕn : Gn × S → Xn, which factor uniquely through Zn and hence give maps S → Zn. Since Xˆ is a colimit of Zn’s by Proposition 4.7, we obtain a map S → Xˆ . Since ϕn is a pullback of the canonical family on Zn, the family ϕ agrees to any finite order with the pullback of the universal family Φ, hence ϕ is equal to this pullback. This proves that (Xˆ , Φ) represents the functor X+. 

Proposition 1.2 now follows very similarly as the proof in the linearly reductive case [JS19, §7]. + G The key point is the mere existence and affineness of πX : X → X .

Proof of Proposition 1.2. Since πX is affine and smooth at x by assumption, we can write it locally on XG as π : Spec (B) → Spec (A), where B is a G-algebra and π is G-invariant. Since π has a section, the map π# : A → B is injective. Suppose first k = k. Then a Kempf line sits inside G and induces an N-grading on B such that A ⊂ B0. By Corollary 2.2, we in fact have A = B0. In this setup, the claim follows from [JS19, Lemma 7.2]. For general k, we may assume x is closed. Now the morphism X+ → (XG) is an affine k k space fibration near x by the previous case, so in fact near x it a trivial affine space fibration, i.e., a projection from a trivial vector bundle, so X+ → XG is a GL-torsor. But being a GL-torsor bundle  is Zariski-local, so we get that also πX is such locally near x and that concludes the proof.

APPENDIX A. TANNAKIAN FORMALISM Let G be a linear group over k and X bea G-scheme. In this appendix we recover G-equivariant morphisms from pullback-like maps. This is used crucially in the proof of Theorem 1.1 to obtain the “embedding of cells” map Xˆ → X. We need some preliminary notions. A cocontinuous functor F between tensor categories is a functor that preserves all small colimits; in particular it is right-exact. A tensor functor is a strong symmetric monoidal functor, which means that F(M ⊗ N) ≃ F(M) ⊗ F(N) and F(1) ≃ 1 and these isomorphisms are subject to compatibility conditions[SR72, I.4.1.1-4.2.4]. For a G-scheme T ∗ the equivariant map pT : T → Spec (k) induces a pullback pT : RepG → QcohG(T) which maps a G-representation V to V ⊗k OT with the natural linearization. For a G-equivariant morphism ∗ ∗ ∗ ϕ : Y → X of G-schemes, we have a natural isomorphism αϕ : ϕ ◦ pX → pY of functors. Our main result is the following.

Theorem A.1. Let X be a quasi-compact quasi-separated G-scheme. Let Y be a G-scheme. Let F : QcohG(X) → ∗ ∗ QcohG(Y) be a cocontinuous tensor functor and let α : F ◦ pX → pY be an isomorphism of func- tors RepG → QcohG(Y). Then there exists a unique G-equivariant morphism f : Y → X such that ∗ (F, α) ≃ ( f , α f ). The most natural language of proving (and even formulating) Theorem A.1 is the theory of algebraic stacks. We recall below the bare minimum necessary for its proof. For an introduction BIAŁYNICKI-BIRULA DECOMPOSITION FOR REDUCTIVE GROUPS IN POSITIVECHARACTERISTIC 17 to stacks, see [Ols16]. For a sketch of the proof without using the language of stacks, see the first arXiv version of the present article. Let ⋆ = Spec (k). Let T be a G-scheme over Spec (k). The quotient stack [T/G] is the category fibred in groupoids, whose S-points are pairs (π, γ), where π : P → S is a principal G-bundle over S and γ : P → T is a G-equivariant morphism. Since G is affine, this stack has affine diago- nal. There is a forgetful map pT : [T/G] → [⋆/G] that maps the pair (π, γ) to the pair (π,pr ◦γ), where pr: T → Spec (k) is the structural map. The category QcohG(T) is equivalent to the cate- gory Qcoh([T/G]) of quasi-coherent sheaves on the stack [T/G],see [Ols16, Example 9.1.19]. Let now X, Y be two G-schemes over Spec (k). We now recall the definition of the groupoid Mor[⋆/G] ([Y/G] , [X/G]). Its objects are pairs ( f , α), where f : [Y/G] → [X/G] is a morphism ′ ′ of stacks and α : pX ◦ f → pY is an isomorphism of functors. Its morphisms ( f , α) → ( f , α ) are ′ ′ isomorphisms of functors β : f → f that satisfy α ◦ (pXβ) = α. We now define the groupoid Homr⊗,≃,[⋆/G](QcohG(X), QcohG(Y)). Its objects are pairs (F, A) where F : QcohG(X) → QcohG(Y) ∗ ∗ is a cocontinuous tensor functor and A : F ◦ pX → pY is an isomorphism of functors. Its mor- ′ ′ ′ ′ ∗ phisms (F, A) → (F , A ) are isomorphisms of functors B : F → F that satisfy A ◦ (BpX) = A. Both groupoids defined above are the usual mapping spaces of two categories fibered in groupoids over a third such category. Let MorG(Y, X) be the set of G-equivariant morphisms ϕ : Y → X. There is a natural functor

[−/G] : MorG(Y, X) → Mor[⋆/G]([Y/G] , [X/G]) that maps ϕ to ([ϕ/G], id). Asin[HR19] there is also a natural functor

ω[X/G]([Y/G]) : Mor[⋆/G]([Y/G] , [X/G]) → Homr⊗,≃,[⋆/G](QcohG(X), QcohG(Y)). Theorem A.1 can be restated as: the composition of the two functors above is an equivalence of categories. We will show that both functors are equivalences. For the first, this follows from general properties. For the second, this follows from certain results of [HR19], however not directly from their main theorem, which requires the stack [Y/G] to be locally excellent. Lemma A.2. The functor [−/G] is an equivalence of categories.

Proof. Call a functor ( f , α) ∈ Mor[⋆/G]([Y/G] , [X/G]) strict if α = id. We claim that for every ( f , α) there is a uniquely defined strict functor ( fstrict, id) isomorphic to ( f , α). Indeed, for a scheme S and a bundle π : P → S with a map γ : P → Y we have f ((π, γ)) = (π′, δ) and α((π, γ)) is an isomorphism α(π,γ) X P′ P δ π′ π S −1 so we define fstrict((π, γ)) as the bundle (π, δ ◦ α(π,γ)). This proves existence. Uniqueness follows directly the definition: the only isomorphism of strict functors is the identity. The above strictifi- cation shows that we only need to prove that [−/G] induces a onto the set of strict func- tors. Consider the trivial bundle (pr, σY). Then [ϕ/G](pr, σY)=(pr, τ) where τ(g, y) = gϕ(y), in particular ϕ = τ(1G, −). This shows that [−/G] is injective. Note that we use strictness here, to have a canonical section (1G, −). It remains to show that [−/G] is onto the set of strict func- tors. Consider a strict functor ( f , id), let (pr, τ) = f (pr, σY) and define ϕ := τ(1G, −). Consider a scheme S, a trivial bundle pr2 : G × S → S and an equivariant map γ : G × S → Y. We have commutative diagrams

µ×id γ id ×γ σ G × G × S S G × S Y G × G × S G G × YY Y

pr23 pr2 pr23 pr2 pr γ G × S2 S G × S Y

The bundles in the leftmost columns are equal and their maps to Y agree. Let f (pr2, γ)=(pr2, δ). Applying f to both diagrams above, we get that δ(gh, s) = τ(g, γ(h, s)) for all g, h ∈ G and s ∈ S. 18 JOACHIM JELISIEJEW AND ŁUKASZ SIENKIEWICZ

For g = 1G we get δ = ϕ ◦ γ and thus f maps (pr2, γ) to (pr2, ϕ ◦ γ). Applying this for γ = σY we deduce that ϕ is equivariant. Therefore, the functors f and [ϕ/G] agree on all trivial bundles. By descent, they agree everywhere.  Proposition A.3 (tensoriality). Suppose that X is quasi-compact, quasi-separated. Then the functor ω[X/G] is an equivalence. Proof. The proof closely follows the argument in [HR19, Theorem 4.10]. The group G is affine, so every quotient stack by G has affine diagonal and then [HR19, Proposition 4.8(i)] implies that ω[X/G]([Y/G]) and ω[⋆/G]([Y/G]) are fully faithful. To prove that ω[X/G]([Y/G]) is essentially surjective, pick an element

(F, A) ∈ Homr⊗,≃,[⋆/G](QcohG(X), QcohG(Y)). The group G is smooth and we have the smooth affine atlas s : X → [X/G]. Let Λ be the coor- dinate ring of G with its natural linearization. The element s∗OX corresponds in QcohG(X) to Λ ∗ Λ ∗ ∗ the sheaf ⊗k OX ≃ pX( ). Since FpX is isomorphic to pY, the sheaf of algebras F(s∗OX) is iso- morphic to Λ ⊗k OY ∈ QcohG(Y). The stack Spec [Y/G]F(s∗OX) is isomorphic to Spec [Y/G](Λ ⊗k OY) = Y. By general nonsense and as explained in [HR19, Corollary 3.6] the functor F induces a cocontinuous tensor functor ′ F : Qcoh(X) ≃ Qcoh(Spec [X/G](s∗OX)) → Qcoh(Spec [Y/G]F(s∗OX)) ≃ Qcoh(Y). Now by [BC14], since X is quasi-compact, quasi-separated, it follows that F′ is isomorphic to a pullback functor. The map Y → [Y/G] is smooth, so by [HR19, Lemma 4.2(iii)] the functor F is isomorphic to a pullback functor of some f : [Y/G] → [X/G]. The isomorphism F ≃ f ∗ ∗ ∗ ∗ composed with A gives an isomorphism α : f pX → pY. The pair ( f , α) is isomorphic to (F, A), which concludes the proof. 

REFERENCES [AHR20] J. Alper, J. Hall, and D. Rydh. A Luna étale slice theorem for algebraic stacks. Ann. of Math. (2), 191(3):675–738, 2020. doi:10.4007/annals.2020.191.3.1. [BB73] A. Białynicki-Birula. Some theorems on actions of algebraic groups. Ann. of Math. (2), 98:480–497, 1973. doi:10.2307/1970915. [BC14] M. Brandenburg and A. Chirvasitu. Tensor functors between categories of quasi-coherent sheaves. J. Algebra, 399:675–692, 2014. doi:10.1016/j.jalgebra.2013.09.050. [Bri14] M. Brion. On algebraic semigroups and monoids. In Algebraic monoids, group embeddings, and alge- braic combinatorics, volume 71 of Fields Inst. Commun., pages 1–54. Springer, New York, 2014. URL https://doi.org/10.1007/978-1-4939-0938-4_1. [Dem65] M. Demazure. Schémas en groupes réductifs. Bull. Soc. Math. France, 93:369–413, 1965. URL http://www.numdam.org/item?id=BSMF_1965__93__369_0. [DG14] V. Drinfeld and D. Gaitsgory. On a theorem of Braden. Transform. Groups, 19(2):313–358, 2014. doi:10.1007/s00031-014-9267-8. [Dri13] V. Drinfeld. On algebraic spaces with an action of Gm. arXiv:1308.2604., 2013. [FN77] J. Fogarty and P. Norman. A fixed-point characterization of linearly reductive groups. In Contributions to algebra (collection of papers dedicated to Ellis Kolchin), pages 151–155. 1977. [HR19] J. Hall and D. Rydh. Coherent Tannaka duality and algebraicity of Hom-stacks. Algebra Number Theory, 13(7):1633–1675, 2019. doi:10.2140/ant.2019.13.1633. [Ive72] B. Iversen. A fixed point formula for action of tori on algebraic varieties. Invent. Math., 16:229–236, 1972. doi:10.1007/BF01425495. [JS19] J. Jelisiejew and Ł. Sienkiewicz. Białynicki-Birula decomposition for reductive groups. Journal de Mathématiques Pures et Appliquées, 2019. doi:https://doi.org/10.1016/j.matpur.2019.04.006. [Ols16] M. Olsson. Algebraic spaces and stacks, volume 62 of American Mathematical Society Colloquium Publications. Ameri- can Mathematical Society, Providence, RI, 2016. URL https://doi.org/10.1090/coll/062. [Ren05] L. E. Renner. Linear algebraic monoids, volume 134 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, 2005. Invariant Theory and Algebraic Transformation Groups, V. [Rit98] A. Rittatore. Algebraic monoids and group embeddings. Transformation Groups, 3(4):375–396, 1998. [SGA3] Schémas en groupes. III: Structure des schémas en groupes réductifs. Séminaire de Géométrie Algébrique du Bois Marie 1962/64 (SGA 3). Dirigé par M. Demazure et A. Grothendieck. Lecture Notes in Mathematics, Vol. 153. Springer-Verlag, Berlin-New York, 1970. [SR72] N. Saavedra Rivano. Catégories Tannakiennes. Lecture Notes in Mathematics, Vol. 265. Springer-Verlag, Berlin- New York, 1972. BIAŁYNICKI-BIRULA DECOMPOSITION FOR REDUCTIVE GROUPS IN POSITIVECHARACTERISTIC 19

[sta17] Stacks Project. http://math.columbia.edu/algebraic_geometry/stacks-git, 2017.