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

Categorical braid group actions
and cactus groups

Iva Halacheva I. Halacheva: Department of Mathematics, Northeastern University, USA [email protected] Anthony Licata A. Licata: Mathematical Sciences Institute, Australian National University, Australia [email protected] Ivan Losev I. Losev: Department of Mathematics, Yale University, and School of Mathematics, IAS, USA [email protected]  and  Oded Yacobi O. Yacobi: School of Mathematics and Statistics, University of Sydney, Australia [email protected]
Abstract.

Let 𝔤\mathfrak{g} be a semisimple simply-laced Lie algebra of finite type. Let 𝒞\mathcal{C} be an abelian categorical representation of the quantum group Uq(𝔤)U_{q}(\mathfrak{g}) categorifying an integrable representation VV. The Artin braid group BB of 𝔤\mathfrak{g} acts on Db(𝒞)D^{b}(\mathcal{C}) by Rickard complexes, providing a triangulated equivalence

Θw0:Db(𝒞μ)Db(𝒞w0(μ))\displaystyle\Theta_{w_{0}}:D^{b}(\mathcal{C}_{\mu})\to D^{b}(\mathcal{C}_{w_{0}(\mu)})

where μ\mu is a weight of VV, and Θw0\Theta_{w_{0}} is a positive lift of the longest element of the Weyl group.

We prove that this equivalence is t-exact up to shift when VV is isotypic, generalising a fundamental result of Chuang and Rouquier in the case 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2}. For general VV, we prove that Θw0\Theta_{w_{0}} is a perverse equivalence with respect to a Jordan-Hölder filtration of 𝒞\mathcal{C}.

Using these results we construct, from the action of BB on VV, an action of the cactus group on the crystal of VV. This recovers the cactus group action on VV defined via generalised Schützenberger involutions, and provides a new connection between categorical representation theory and crystal bases. We also use these results to give new proofs of theorems of Berenstein-Zelevinsky, Rhoades, and Stembridge regarding the action of symmetric group on the Kazhdan-Lusztig basis of its Specht modules.

1. Introduction

In their seminal work, Chuang and Rouquier introduced 𝔰𝔩2\mathfrak{sl}_{2} categorifications on abelian categories [CR]. Their definition mirrors the notion of an 𝔰𝔩2\mathfrak{sl}_{2} representation on a vector space: weight spaces are replaced by weight categories, Chevalley generators acting on them are replaced by Chevalley functors, and Lie algebra relations are replaced by isomorphisms of functors. But, crucially, these isomorphisms are part of the “higher data” of categorification.

The richness of this theory was immediately evident. As a corollary of an 𝔰𝔩2\mathfrak{sl}_{2} categorification on representations of symmetric groups in positive characteristic, Chuang and Rouquier proved Broue’s abelian defect conjecture in that case. The essential tool allowing them to do this is the Rickard complex, which is a categorical lifting of the reflection matrix in SL2SL_{2}, and provides a derived equivalence between opposite weight categories.

Subsequently, Rouquier and Khovanov-Lauda vastly generalised this theory to quantum symmetrisable Kac-Moody algebras Uq(𝔤)U_{q}(\mathfrak{g}) [Rou2KM, KLI, KLII]. Let 𝕜\mathbbm{k} be any field. A graded abelian 𝕜\mathbbm{k}-linear category 𝒞\mathcal{C} endowed with a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}) possesses a family of Rickard complexes Θi\Theta_{i}, indexed by the simple roots of 𝔤\mathfrak{g}, acting on the derived category Db(𝒞)D^{b}(\mathcal{C}).

Henceforth let 𝔤\mathfrak{g} be a semisimple simply-laced Lie algebra of finite type with Dynkin diagram II, WW its Weyl group, and BB its Artin braid group. Let 𝒞\mathcal{C} be a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}) as in the previous paragraph. Cautis and Kamnitzer proved that Rickard complexes satisfy the braid relations, as conjectured by Rouquier [CK3]. This defines an action of BB on Db(𝒞)D^{b}(\mathcal{C}), and is our main object of study.

Categorical braid group actions defined via Rickard complexes have many significant applications. For example, in low dimensional topology, the type A link homology theories (in particular Khovanov homology) emerge as a byproduct of these types of categorical braid group actions [CK, CKII, LQR]. In mirror symmetry, the theory of spherical twists plays an important role, and these all arise from categorical 𝔰𝔩2\mathfrak{sl}_{2} representations [ST01].

To describe our first theorem, recall that minimal categorifications are certain distinguished categorifications of simple representations. On these the Rickard complex Θi\Theta_{i} is t-exact up to shift [CR, Theorem 6.6]. Notice that this is a result about 𝔰𝔩2\mathfrak{sl}_{2} categorifications, and in fact, this is one of Chuang-Rouquier’s key technical results which they use to prove the derived equivalence.

We generalise this result to Uq(𝔤)U_{q}(\mathfrak{g}), where we show that the composition of Rickard complexes corresponding to a positive lift of the longest element w0Ww_{0}\in W is t-exact up to shift on any isotypic categorification. More precisely:

Theorem A.

[Theorem 6.4 & Corollary 6.7] Let 𝒞\mathcal{C} be a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}) categorifying an isotypic representation of type λ\lambda, where λ\lambda is a dominant integral weight. Let μ\mu be any weight, and let nn be the height of μw0(λ)\mu-w_{0}(\lambda). Then the derived equivalence

Θw0𝟙μ[n]:Db(𝒞μ)Db(𝒞w0(μ))\displaystyle\Theta_{w_{0}}\mathbbm{1}_{\mu}[n]:D^{b}(\mathcal{C}_{\mu})\to D^{b}(\mathcal{C}_{w_{0}(\mu)})

is t-exact.

This theorem is the technical heart of the paper. In order to prove it we introduce a new combinatorial notion of “marked words” (Section 5). This allows us to use relations between Θi\Theta_{i} and Chevalley functors established by Cautis and Kamnitzer to deduce the commutation relations involving Θw0\Theta_{w_{0}} (Proposition 5.9). We then use these relations to prove the theorem by induction on nn.

Our second theorem describes Θw0\Theta_{w_{0}} on an arbitrary categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}), also generalising a result of Chuang-Rouquier in the case 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2}. Indeed, their study of the Rickard complex on an 𝔰𝔩2\mathfrak{sl}_{2} categorification led them to define the notion of a “perverse equivalence” [CRperv].

Consider an equivalence of triangulated categories 𝖥:𝒯𝒯\mathsf{F}:\mathcal{T}\to\mathcal{T}^{\prime} with t-structures [BBD]. Suppose further that 𝒯\mathcal{T} (respectively 𝒯\mathcal{T}^{\prime}) is filtered by thick triangulated subcategories

0𝒯0𝒯r=𝒯,0𝒯0𝒯r=𝒯,\displaystyle 0\subset\mathcal{T}_{0}\subset\cdots\subset\mathcal{T}_{r}=\mathcal{T},\quad 0\subset\mathcal{T}_{0}^{\prime}\subset\cdots\subset\mathcal{T}_{r}^{\prime}=\mathcal{T}^{\prime},

and 𝖥\mathsf{F} is compatible with these filtrations (cf. Section 4.1 for precise definitions). Then, roughly speaking, 𝖥\mathsf{F} is a perverse equivalence if on each subquotient 𝖥:𝒯i/𝒯i1𝒯i/𝒯i1\mathsf{F}:\mathcal{T}_{i}/\mathcal{T}_{i-1}\to\mathcal{T}_{i}^{\prime}/\mathcal{T}_{i-1}^{\prime} is t-exact up to shift.

Since their introduction, perverse equivalences have proven useful in various contexts (e.g. representations of finite groups [Cra-Rou], geometric representation theory and mirror symmetry [Agan], and algebraic combinatorics [GJN]). Our second theorem shows that perverse equivalences are ubiquitous in categorical representation theory:

Theorem B.

[Theorem 6.8] Let 𝒞\mathcal{C} be a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}), and let μ\mu be any weight. The derived equivalence Θw0𝟙μ:Db(𝒞μ)Db(𝒞w0(μ))\Theta_{w_{0}}\mathbbm{1}_{\mu}:D^{b}(\mathcal{C}_{\mu})\to D^{b}(\mathcal{C}_{w_{0}(\mu)}) is a perverse equivalence with respect to a Jordan-Hölder or isotypic filtration of 𝒞\mathcal{C}.

Note that if JIJ\subseteq I is a subdiagram, and w0Jw_{0}^{J} is corresponding longest element, then this theorem implies that Θw0J𝟙μ\Theta_{w_{0}^{J}}\mathbbm{1}_{\mu} is a perverse equivalence for any JJ. We also remark that our argument go through in the ungraded setting, where 𝒞\mathcal{C} is a categorical representation of 𝔤\mathfrak{g}.

Let us explain the filtration arising in Theorem B more precisely. We apply Rouquier’s Jordan-Hölder theory for representations of 22-Kac-Moody algebras to our setting [Rou2KM]. We thus obtain a filtration of 𝒞\mathcal{C} by Serre subcategories,

0𝒞0𝒞r=𝒞,\displaystyle 0\subset\mathcal{C}_{0}\subset\cdots\subset\mathcal{C}_{r}=\mathcal{C},

such that each factor 𝒞i\mathcal{C}_{i} is a subrepresentation, and each subquotient 𝒞i/𝒞i1\mathcal{C}_{i}/\mathcal{C}_{i-1} categorifies either a simple module (Theorem 3.6) or an isotypic component (Remark 3.9). Then Θw0𝟙μ\Theta_{w_{0}}\mathbbm{1}_{\mu} is a perverse equivalence with respect to the filtration whose ii-th filtered component consists of complexes in Db(𝒞μ)D^{b}(\mathcal{C}_{\mu}) with cohomology supported in 𝒞i\mathcal{C}_{i}.

We remark that in the case 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2} this gives a more conceptual proof of a result of Chuang-Rouquier [CRperv, Proposition 8.4]. If 𝒞\mathcal{C} is the tensor product categorification of the nn-fold tensor product of the standard representation of 𝔰𝔩n\mathfrak{sl}_{n}, we recover a theorem of the third author [LosCacti]. We explain this in Example 8.3, where we show how to interpret the filtration on the principal block of the BGG category 𝒪\mathcal{O} using the Robinson-Schensted correspondence.

In fact, the third author and Bezrukavnikov formulated a principle that suitable categorical braid group representations should have a “crystal limit” [BLet, Section 9]. As an application of our results we can make this precise in the setting of categorical representations of Uq(𝔤)U_{q}(\mathfrak{g}).

Recall that to an integrable representation VV of Uq(𝔤)U_{q}(\mathfrak{g}), Kashiwara associated its crystal basis 𝐁V\mathbf{B}_{V} [Kash90], which is closely related to Lusztig’s canonical basis [GL92]. If VV is categorified by 𝒞\mathcal{C} then there is a natural identification 𝐁V=𝖨𝗋𝗋(𝒞)\mathbf{B}_{V}=\mathsf{Irr}(\mathcal{C}), the set of isomorphism classes of simple objects in 𝒞\mathcal{C} up to shift (cf. Proposition 3.4) .

One of the most important features of the theory is the existence of a tensor product, endowing the category of crystals with a monoidal structure. The commutator of crystals is controlled by a group called the cactus group, just as BB controls the commutator in the category of representations of Uq(𝔤)U_{q}(\mathfrak{g}) [HK06]. There is also an internal cactus group action, mirroring Lusztig’s internal braid group action on VV. Indeed, there is a cactus group CC associated to 𝔤\mathfrak{g} (or rather to its Dynkin diagram II), which can be presented by generators cJc_{J} indexed by connected subdiagrams JIJ\subseteq I (cf. Section 7.1). Then CC acts on 𝐁V\mathbf{B}_{V} via the so-called Schützenberger involutions (cf. Theorem 7.3).

So, starting with an integrable representation VV of the quantum group we obtain: an action of BB on V, a 𝔤\mathfrak{g}-crystal 𝐁V\mathbf{B}_{V}, and an action of CC on 𝐁V\mathbf{B}_{V}. We schematically picture this situation as follows:

Uq(𝔤)V{U_{q}(\mathfrak{g})\curvearrowright V}BV{B\curvearrowright V}𝔤-crystal 𝐁V{\mathfrak{g}\text{-crystal }\mathbf{B}_{V}}C𝐁V{C\curvearrowright\mathbf{B}_{V}}?\scriptstyle{?}

Naturally one asks: can we “crystallise” the braid group action on VV directly to obtain the cactus group action on 𝐁V\mathbf{B}_{V}? Our results allow us to answer this in the affirmative.

The key point is that a perverse equivalence 𝖥:𝒯𝒯\mathsf{F}:\mathcal{T}\to\mathcal{T}^{\prime} induces a bijection 𝖨𝗋𝗋(𝒯)𝖨𝗋𝗋((𝒯))\mathsf{Irr}(\mathcal{T}^{\heartsuit})\leftrightarrow\mathsf{Irr}((\mathcal{T}^{\prime})^{\heartsuit}), where 𝒯\mathcal{T}^{\heartsuit} denotes the heart of the t-structure. In the setting of Theorem B, we obtain a bijection φI:𝖨𝗋𝗋(𝒞)𝖨𝗋𝗋(𝒞)\varphi_{I}:\mathsf{Irr}(\mathcal{C})\to\mathsf{Irr}(\mathcal{C}). In fact, if JIJ\subseteq I is a subdiagram and 𝔤J𝔤I\mathfrak{g}_{J}\subseteq\mathfrak{g}_{I} is the corresponding Lie subalgebra, we can regard 𝒞\mathcal{C} as a categorical representation of Uq(𝔤J)U_{q}(\mathfrak{g}_{J}) by restriction. By Theorem B we also obtain a bijection φJ:𝖨𝗋𝗋(𝒞)𝖨𝗋𝗋(𝒞)\varphi_{J}:\mathsf{Irr}(\mathcal{C})\to\mathsf{Irr}(\mathcal{C}).

Theorem C.

[Theorem 7.7 & Theorem 7.12] Let 𝒞\mathcal{C} be a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}), categorifying the integrable representation VV. The assignment cJφJc_{J}\mapsto\varphi_{J} defines an action of CC on 𝐁V=𝖨𝗋𝗋(𝒞)\mathbf{B}_{V}=\mathsf{Irr}(\mathcal{C}), and this agrees with the combinatorial action arising from Schützenberger involutions.

We thus obtain the sought-after crystalisation process for braid groups:

BVBDb(𝒞)C𝖨𝗋𝗋(𝒞),B\curvearrowright V\quad\rightsquigarrow\quad B\curvearrowright D^{b}(\mathcal{C})\quad\rightsquigarrow\quad C\curvearrowright\mathsf{Irr}(\mathcal{C}),

which associates a cactus group set 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}) to the braid group representation of BB on VV. The first appearance of such a crystalisation process is in the work of the third author, where a cactus group action on WW is constructed [LosCacti]. It’s an interesting question to crystallise the braid group action without appealing to categorical representation theory.

Finally we remark that perversity of Rickard complexes, and more specifically the t-exactness of Θw0\Theta_{w_{0}} on isotypic categorifications as in Theorem A, is a fruitful vantage from which to view results in algebraic combinatorics.

For example, we show in Section 8.2 how to use this to easily recover theorems of Berenstein-Zelevinsky [BZ96] and Stembridge [Stem96], namely that the action of w0Snw_{0}\in S_{n} on the Kazhdan-Lusztig basis of a Specht module of SnS_{n} is governed by the evacuation operator on standard Young tableaux. We note that this theorem was earlier proven by Mathas in slightly different form (without explicit reference to the evacuation operator, and credited to J.J. Graham) [Mathas, Theorem 3.1], and a similar result was shown even earlier by Lusztig in 1990 [LusCBII, Corollary 5.9].

As another example, we use our methods to also recover Rhoades’ Theorem that the Coxeter element (1,2,,n)Sn(1,2,\ldots,n)\in S_{n} acts on the Kazhdan-Lusztig basis of a Specht module associated to a rectangular partition by the promotion operator. This point of view led us to generalise Rhoades’ result to arbitrary partitions [GY1], and isolate the class of permutations (the separable permutations) for which such results can hold [GY2].

Acknowledgement

We would like to thank Sabin Cautis, Ian Grojnowski, Joel Kamnitzer, Aaron Lauda, Andrew Mathas, Peter McNamara, Bregje Pauwels, Raphäel Rouquier, and Geordie Williamson for insightful discussions. We are grateful to two anonymous referees for their helpful comments. I.L. is partially supported by the NSF under grant DMS-2001139. O.Y. is supported by the Australian Research Council Grants DP180102563 and DP230100654.

2. Background on quantum groups

2.1. The quantum group

In this article we work with a simply-laced quantum group Uq(𝔤)U_{q}(\mathfrak{g}) of finite type. Recall that we have an associated Cartan datum and a root datum, which consists of:

  • A finite set II,

  • a symmetric bilinear form (,)(\cdot,\cdot) on I\mathbb{Z}I satisfying (i,i)=2(i,i)=2 and (i,j){0,1}(i,j)\in\{0,-1\} for all ijIi\neq j\in I,

  • a free \mathbb{Z}-module XX, called the weight lattice, and

  • a choice of simple roots {αi}iIX\{\alpha_{i}\}_{i\in I}\subset X and simple coroots {hi}iIX=Hom(X,)\{h_{i}\}_{i\in I}\subset X^{\vee}=\operatorname{Hom}(X,\mathbb{Z}) satisfying hi,αj=(i,j)\langle h_{i},\alpha_{j}\rangle=(i,j), where ,:X×X\langle\cdot,\cdot\rangle:X^{\vee}\times X\to\mathbb{Z} is the natural pairing.

The quantum group Uq(𝔤)U_{q}(\mathfrak{g}) is the unital, associative, (q)\mathbb{C}(q) algebra generated by Ei,Fi,Kh,(iI,hX)E_{i},F_{i},K_{h},(i\in I,h\in X^{\vee}) subject to relations:

  1. (1)

    K0=1K_{0}=1 and KhKh=Kh+hK_{h}K_{h^{\prime}}=K_{h+h^{\prime}} for any h,hXh,h^{\prime}\in X^{\vee},

  2. (2)

    KhEi=qh,αiEiKhK_{h}E_{i}=q^{\langle h,\alpha_{i}\rangle}E_{i}K_{h} for any iI,hXi\in I,h\in X^{\vee},

  3. (3)

    KhFi=qh,αiFiKhK_{h}F_{i}=q^{-\langle h,\alpha_{i}\rangle}F_{i}K_{h} for any iI,hXi\in I,h\in X^{\vee},

  4. (4)

    EiFjFjEi=δijKiKi1qq1E_{i}F_{j}-F_{j}E_{i}=\delta_{ij}\frac{K_{i}-K_{i}^{-1}}{q-q^{-1}}, where we set Ki=KhiK_{i}=K_{h_{i}}, and

  5. (5)

    for all iji\neq j,

    a+b=hi,αj+1(1)aEi(a)EjEi(b)=0 and a+b=hi,αj+1(1)aFi(a)FjFi(b)=0,\sum_{a+b=-\langle h_{i},\alpha_{j}\rangle+1}(-1)^{a}E_{i}^{(a)}E_{j}E_{i}^{(b)}=0\quad\text{ and }\quad\sum_{a+b=-\langle h_{i},\alpha_{j}\rangle+1}(-1)^{a}F_{i}^{(a)}F_{j}F_{i}^{(b)}=0,

    where Ei(a)=Eia/[a]!,Fi(a)=Fia/[a]!E_{i}^{(a)}=E_{i}^{a}/[a]!,F_{i}^{(a)}=F_{i}^{a}/[a]!, and [a]!=i=1aqiqiqq1[a]!=\prod_{i=1}^{a}\frac{q^{i}-q^{-i}}{q-q^{-1}}.

We let aij=(i,j)a_{ij}=(i,j), so that (aij)i,jI(a_{ij})_{i,j\in I} is a Cartan matrix. Given λX\lambda\in X we abbreviate λi=hi,λ\lambda_{i}=\langle h_{i},\lambda\rangle, and let

X+={λX:λi0 for all iI}X_{+}=\{\lambda\in X:\lambda_{i}\geq 0\text{ for all }i\in I\}

be the set of dominant weights.

Let RXR\subset X be the root lattice, defined as the \mathbb{Z}-span of the simple roots, and let R+RR_{+}\subset R be the \mathbb{N}-span of the simple roots. We define the usual preorder \succ on XX by λμ\lambda\succeq\mu if λμR+\lambda-\mu\in R_{+}. For μR\mu\in R let ht(μ)ht(\mu) denote the height of μ\mu, i.e. ht(iaiαi)=iaiht(\sum_{i}a_{i}\alpha_{i})=\sum_{i}a_{i}.

When convenient, we also view II as the Dynkin diagram of 𝔤\mathfrak{g}, and make reference to subdiagrams or diagram automorphisms of II.

2.2. Braid group actions on integrable representations of Uq(𝔤)U_{q}(\mathfrak{g}).

Given a Uq(𝔤)U_{q}(\mathfrak{g})-module VV and μX\mu\in X we let VμV_{\mu} denote the μ\mu weight space of VV. For λX+\lambda\in X_{+} we let L(λ)L(\lambda) be the irreducible representation of Uq(𝔤)U_{q}(\mathfrak{g}) of highest weight λ\lambda. Let 𝖨𝗌𝗈λ(V)\mathsf{Iso}_{\lambda}(V) denote the λ\lambda-isotypic component of VV. We say that VV is isotypic if there exists λX+\lambda\in X_{+} such that V=𝖨𝗌𝗈λ(V)V=\mathsf{Iso}_{\lambda}(V).

The representation L(λ)L(\lambda) has a canonical basis, which we denote by 𝐁(λ)\mathbf{B}(\lambda) [Lusbook]. We let vλv_{\lambda} (respectively vλlowv_{\lambda}^{low}) denote the unique highest weight (respectively lowest weight) element of 𝐁(λ)\mathbf{B}(\lambda).

Let B=BIB=B_{I} denote the braid group of type II, which is generated by θi(iI)\theta_{i}\;(i\in I) subject to the braid relations:

θiθj\displaystyle\theta_{i}\theta_{j} =θjθi, if (i,j)=0, and\displaystyle=\theta_{j}\theta_{i},\text{ if }(i,j)=0,\text{ and }
θiθjθi\displaystyle\theta_{i}\theta_{j}\theta_{i} =θjθiθj, if (i,j)=1.\displaystyle=\theta_{j}\theta_{i}\theta_{j},\text{ if }(i,j)=-1.

Let W=WIW=W_{I} be the Weyl group of type II, which has generators si(iI)s_{i}\;(i\in I) subject to the braid relations, and in addition the quadratic relation si2=1s_{i}^{2}=1. Let w0Ww_{0}\in W be the longest element. Recall that WW acts on XX via siλ=λhi,λαis_{i}\cdot\lambda=\lambda-\langle h_{i},\lambda\rangle\alpha_{i}. We define τ:II\tau:I\to I by the equality ατ(i)=w0(αi)\alpha_{\tau(i)}=-w_{0}(\alpha_{i}) for any iIi\in I.

To JIJ\subset I a subdiagram, we associate WJWW_{J}\subset W the parabolic subgroup, w0JWJw_{0}^{J}\in W_{J} its longest element, and τJ:II\tau_{J}:I\to I the bijection given by

ατJ(i)={w0J(αi) if iJ,αiotherwise.\displaystyle\alpha_{\tau_{J}(i)}=\begin{cases}-w_{0}^{J}(\alpha_{i})&\text{ if }i\in J,\\ \alpha_{i}&\text{otherwise.}\end{cases}

For any wWw\in W we can consider its positive lift θwB\theta_{w}\in B, where θw=θi1θi\theta_{w}=\theta_{i_{1}}\cdots\theta_{i_{\ell}} and w=si1siw=s_{i_{1}}\cdots s_{i_{\ell}} is any reduced decomposition.

Let VV be an integrable representation of Uq(𝔤)U_{q}(\mathfrak{g}). A fundamental structure of VV, discovered by Lusztig, is that it admits (several) braid group symmetries, sometimes referred to as the “quantum Weyl group actions”. To recall this, let 𝟣μ\mathsf{1}_{\mu} denote the projection onto the μ\mu weight space. For each iIi\in I we define 𝗍i:VV\mathsf{t}_{i}:V\to V by:

𝗍i𝟣μ=ba=(μ,αi)(q)bEi(a)Fi(b)𝟣μ.\displaystyle\mathsf{t}_{i}\mathsf{1}_{\mu}=\sum_{b-a=(\mu,\alpha_{i})}(-q)^{-b}E_{i}^{(a)}F_{i}^{(b)}\mathsf{1}_{\mu}. (2.1)

Note that the indexing set of this sum is infinite, but the sum itself is finite on VV. In the notation of [Lusbook], 𝗍i=𝗍i,1′′\mathsf{t}_{i}=\mathsf{t}^{\prime\prime}_{i,-1}. (Note that the formula given in [Lusbook] is more complicated. This simpler form was initially observed in [CR] in the non-quantum setting, and generalised in [CKM] to the quantum setting.). The assignment θi𝗍i\theta_{i}\mapsto\mathsf{t}_{i} defines an action of BB on VV, and so we can unambiguously write 𝗍w\mathsf{t}_{w} for any wWw\in W.

3. Categorical representations of Uq(𝔤)U_{q}(\mathfrak{g})

3.1. Notation

Fix a field 𝕜\mathbbm{k} of any characteristic. In this paper we will be concerned mostly with abelian 𝕜\mathbbm{k}-linear categories 𝒞\mathcal{C}. We will assume throughout that each block of 𝒞\mathcal{C} is a finite abelian category [EGNO, Definition 1.8.6].

Recall that a category 𝒜\mathcal{A} is graded if it is equipped with an auto-equivalence 1:𝒜𝒜\langle 1\rangle:\mathcal{A}\to\mathcal{A} called the “shift functor”. We let \langle\ell\rangle be the auto-equivalence obtained by applying the shift functor \ell times. We denote by 𝖨𝗋𝗋(𝒜)\mathsf{Irr}(\mathcal{A}) the set of equivalence classes of simple objects of 𝒜\mathcal{A} up to shift.

A functor F:𝒜𝒜F:\mathcal{A}\to\mathcal{A}^{\prime} between graded categories is graded if it commutes with the shift functors. We denote by [𝒜][\mathcal{A}]_{\mathbb{Z}} the Grothendieck group of an abelian category 𝒜\mathcal{A}. If 𝒜\mathcal{A} is graded we denote by [𝒜][q,q1][\mathcal{A}]_{\mathbb{Z}[q,q^{-1}]} the quotient of [𝒜][q,q1][\mathcal{A}]_{\mathbb{Z}}\otimes_{\mathbb{Z}}\mathbb{Z}[q,q^{-1}] by the additional relation q[M]=[M1]q[M]=[M\langle-1\rangle]. This quotient is naturally a [q,q1]\mathbb{Z}[q,q^{-1}]-module. Set

[𝒜]\displaystyle[\mathcal{A}]_{\mathbb{C}} =[𝒜],\displaystyle=[\mathcal{A}]_{\mathbb{Z}}\otimes_{\mathbb{Z}}\mathbb{C},
[𝒜](q)\displaystyle[\mathcal{A}]_{\mathbb{C}(q)} =[𝒜][q,q1][q,q1](q).\displaystyle=[\mathcal{A}]_{\mathbb{Z}[q,q^{-1}]}\otimes_{\mathbb{Z}[q,q^{-1}]}\mathbb{C}(q).

For a 𝕜\mathbbm{k}-algebra AA, we let AmodA\mathrm{-mod} be the category of finitely generated AA-modules, and if AA is \mathbb{Z}-graded, we let AmodA\mathrm{-mod}_{\mathbb{Z}} be the category of finitely generated \mathbb{Z}-graded modules. These are naturally 𝕜\mathbbm{k}-linear abelian categories.

3.2. Definition

In this section we introduce our main objects of study: representations of the 2-quantum group on abelian categories, which we refer to as “categorical representations” of Uq(𝔤)U_{q}(\mathfrak{g}). This definition of the categorified quantum groups and their categorical actions is originally due to Rouquier and Khovanov-Lauda [Rou2KM, KLI, KLII].

In the literature, there are a number of slightly different-looking examples of 2-representations, depending not just on whether or not one is working with the Khovanov-Lauda or Rouquier 2-categories, but also depending on whether or not one is interested in categorifications of representations of Uq(𝔤)U_{q}(\mathfrak{g}) or of U(𝔤)U(\mathfrak{g}). At the categorical level, the difference between Uq(𝔤)U_{q}(\mathfrak{g}) or of U(𝔤)U(\mathfrak{g}) arises from grading considerations; in categorifications of representations of Uq(𝔤)U_{q}(\mathfrak{g}), one works with categories enriched in graded vector spaces, whereas in categorifications of representations of U(𝔤)U(\mathfrak{g}) no such grading is needed.

Fortunately, the different notions of categorical representations - and in particular relationships between the Rouquier and Khovanov-Lauda frameworks in both graded and ungraded settings, have been brought in line by work of Cautis-Lauda, who proved that in the graded setting, integrable 2-representations of the Rouquier 2-category induce 2-representations of the Khovanov-Lauda 2-category [CaLa], and by work of Brundan [Brundandef], who proved that the underlying ungraded 2-categories of Rouquier and Khovanov-Lauda are equivalent.

The important point for us is that all our theorems, including our main results (Theorem 6.4 and Theorem 6.8), remain true in any 2-representation of the Khovanov-Lauda-Rouqier 2-category, in either the graded or ungraded setting. Since the compatibility of the internal grading with the braid group is of some independent combinatorial interest, we elect to keep track of the gradings in the rest of the paper, and leave it to the reader to verify that the arguments go through while ignoring the gradings and working in the setup of, e.g. Brundan [Brundandef]. To that end, we have chosen to follow the notation and conventions of Cautis-Lauda below.

A categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}) consists of the following data:

  • A family of graded abelian 𝕜\mathbbm{k}-linear categories 𝒞μ\mathcal{C}_{\mu} indexed by μX\mu\in X. We refer to each 𝒞μ\mathcal{C}_{\mu} as a weight category.

  • Exact graded functors 𝖤i𝟙μ:𝒞μ𝒞μ+αi\mathsf{E}_{i}\mathbbm{1}_{\mu}:\mathcal{C}_{\mu}\to\mathcal{C}_{\mu+\alpha_{i}} and 𝖥i𝟙μ:𝒞μ𝒞μαi\mathsf{F}_{i}\mathbbm{1}_{\mu}:\mathcal{C}_{\mu}\to\mathcal{C}_{\mu-\alpha_{i}}, for iIi\in I and μX\mu\in X. We refer to 𝖤i,𝖥i\mathsf{E}_{i},\mathsf{F}_{i} as Chevalley functors.

  • A collection of natural transformations between these functors. We won’t be using directly these natural transformations in this work, so refer the reader to [CaLa, Definition 1.1] for their definition.

This data is subject to the conditions spelled out in items (1)-(5) of [CaLa, Definition 1.1]. We only record those that are relevant for us:

  1. (1)

    The functors 𝖤i𝟙μ\mathsf{E}_{i}\mathbbm{1}_{\mu} and 𝖥i𝟙μ\mathsf{F}_{i}\mathbbm{1}_{\mu} are biadjoint up to a specified degree shift (see (3.1) below).

  2. (2)

    The powers of 𝖤i\mathsf{E}_{i} carry an action of the KLR algebra associated to QQ, where QQ denotes a choice of units (tij)ijI(t_{ij})_{i\neq j\in I} in 𝕜×\mathbbm{k}^{\times}. These units satisfy some restrictions which are not relevant for us.

  3. (3)

    We have the following isomorphisms:

    𝖥i𝖤i𝟙μ\displaystyle\mathsf{F}_{i}\mathsf{E}_{i}\mathbbm{1}_{\mu} 𝖤i𝖥i𝟙μ[hi,μ]𝟙μ, if hi,μ0,\displaystyle\cong\mathsf{E}_{i}\mathsf{F}_{i}\mathbbm{1}_{\mu}\oplus_{[-\langle h_{i},\mu\rangle]}\mathbbm{1}_{\mu},\text{ if }\langle h_{i},\mu\rangle\leq 0,
    𝖤i𝖥i𝟙μ\displaystyle\mathsf{E}_{i}\mathsf{F}_{i}\mathbbm{1}_{\mu} 𝖥i𝖤i𝟙μ[hi,μ]𝟙μ, if hi,μ0,\displaystyle\cong\mathsf{F}_{i}\mathsf{E}_{i}\mathbbm{1}_{\mu}\oplus_{[\langle h_{i},\mu\rangle]}\mathbbm{1}_{\mu},\text{ if }\langle h_{i},\mu\rangle\geq 0,
    𝖤i𝖥j𝟙μ\displaystyle\mathsf{E}_{i}\mathsf{F}_{j}\mathbbm{1}_{\mu} 𝖥j𝖤i𝟙μ.\displaystyle\cong\mathsf{F}_{j}\mathsf{E}_{i}\mathbbm{1}_{\mu}.

    This notation is explained as follows: for a Laurent polynomial f=faqaf=\sum f_{a}q^{a}, fA\oplus_{f}A is a direct sum over aa\in\mathbb{Z} of faf_{a} copies of AaA\langle a\rangle, and [n]:=qn1+qn3++q1n[n]:=q^{n-1}+q^{n-3}+\cdots+q^{1-n}.

Usually we just say that 𝒞=μ𝒞μ\mathcal{C}=\bigoplus_{\mu}\mathcal{C}_{\mu} is a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}) (the remaining data is implicit). Given an integrable Uq(𝔤)U_{q}(\mathfrak{g})-module VV, we say that 𝒞\mathcal{C} is a categorification of VV if 𝒞\mathcal{C} is a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}) such that [𝒞](q)V[\mathcal{C}]_{\mathbb{C}(q)}\cong V as Uq(𝔤)U_{q}(\mathfrak{g})-modules.

An additive categorification is a categorical representation on a graded additive 𝕜\mathbbm{k}-linear category 𝒱\mathcal{V} satisfying the same conditions as above, except the Chevalley functors are of course only required to be additive. We let 𝒱i\mathcal{V}^{i} be the idempotent completion of 𝒱\mathcal{V}.

Given an abelian category 𝒞\mathcal{C}, we consider the additive category 𝒞proj\mathcal{C}\mathrm{-proj} defined as the full subcategory of projective objects in 𝒞\mathcal{C}. Note that if 𝒞\mathcal{C} is a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}), then 𝒞proj\mathcal{C}\mathrm{-proj} naturally inherits the structure of an additive categorification.

As a consequence of this definition, and in particular condition (2), there exist divided power functors 𝖤i(r)𝟙λ𝖤ir𝟙λ,𝖥i(r)𝟙λ𝖥ir𝟙λ\mathsf{E}_{i}^{(r)}\mathbbm{1}_{\lambda}\subset\mathsf{E}_{i}^{r}\mathbbm{1}_{\lambda},\mathsf{F}_{i}^{(r)}\mathbbm{1}_{\lambda}\subset\mathsf{F}_{i}^{r}\mathbbm{1}_{\lambda} which categorify the usual divided powers on the level of the quantum group. Again, we refer the reader to [CaLa] and references therein for further details. We note that their adjoints are related as follows:

(𝖤i(r)𝟙λ)R\displaystyle(\mathsf{E}_{i}^{(r)}\mathbbm{1}_{\lambda})_{R} 𝖥i(r)𝟙λ+rαir(λi+r),\displaystyle\cong\mathsf{F}_{i}^{(r)}\mathbbm{1}_{\lambda+r\alpha_{i}}\langle r(\lambda_{i}+r)\rangle, (3.1)
(𝖤i(r)𝟙λ)L\displaystyle(\mathsf{E}_{i}^{(r)}\mathbbm{1}_{\lambda})_{L} 𝖥i(r)𝟙λ+rαir(λi+r).\displaystyle\cong\mathsf{F}_{i}^{(r)}\mathbbm{1}_{\lambda+r\alpha_{i}}\langle-r(\lambda_{i}+r)\rangle. (3.2)

3.3. Crystals

We recall the definition of a crystal.

Definition 3.3 ([Kash90]).

A 𝔤\mathfrak{g}-crystal is a finite set 𝐁\mathbf{B} together with maps:

e~i,f~i:𝐁𝐁{0},εi,φi:𝐁,wt:𝐁X\widetilde{e}_{i},\widetilde{f}_{i}:\mathbf{B}\rightarrow\mathbf{B}\sqcup\{0\},\quad\varepsilon_{i},\varphi_{i}:\mathbf{B}\rightarrow\mathbb{Z},\quad\text{wt}:\mathbf{B}\rightarrow X

for all iIi\in I, such that:

  1. (1)

    for any b,b𝐁b,b^{\prime}\in\mathbf{B}, e~i(b)=b\widetilde{e}_{i}(b)=b^{\prime} if and only if b=f~i(b)b=\widetilde{f}_{i}(b^{\prime}) ,

  2. (2)

    for all b𝐁b\in\mathbf{B}, if e~i(b)𝐁\widetilde{e}_{i}(b)\in\mathbf{B} then wt(e~i(b))=wt(b)+αi\text{wt}(\widetilde{e}_{i}(b))=\text{wt}(b)+\alpha_{i}, and if f~i(b)𝐁\widetilde{f}_{i}(b)\in\mathbf{B} then wt(f~i(b))=wt(b)αi\text{wt}(\widetilde{f}_{i}(b))=\text{wt}(b)-\alpha_{i},

  3. (3)

    for all b𝐁b\in\mathbf{B}, εi(b)=max{n:e~in(b)0}\varepsilon_{i}(b)=\max\{n\in\mathbb{Z}~{}:~{}\widetilde{e}^{n}_{i}(b)\neq 0\}, φi(b)=max{n:f~in(b)0}\varphi_{i}(b)=\max\{n\in\mathbb{Z}~{}:~{}\widetilde{f}^{n}_{i}(b)\neq 0\},

  4. (4)

    for all b𝐁b\in\mathbf{B}, φi(b)εi(b)=wt(b),hi\varphi_{i}(b)-\varepsilon_{i}(b)=\left\langle\text{wt}(b),h_{i}\right\rangle.

Any Uq(𝔤)U_{q}(\mathfrak{g})-representation VV has a corresponding 𝔤\mathfrak{g}-crystal 𝐁=𝐁V\mathbf{B}=\mathbf{B}_{V}. The underlying set of 𝐁\mathbf{B} is in natural bijection with a particular basis of VV (the “global crystal basis” or “canonical basis”). The maps e~i,f~i\widetilde{e}_{i},\widetilde{f}_{i} are related to the raising and lowering Chevalley operators; vaguely speaking they encode information about the leading terms of the Chevalley operators acting on this basis. In particular, for an integral dominant weight λ\lambda, the canonical basis 𝐁(λ)\mathbf{B}(\lambda) of L(λ)L(\lambda) carries a natural crystal structure [GL92].

The crystal of VV naturally arises via categorical representation theory. Namely, as we describe in the next proposition, if 𝒞\mathcal{C} is a categorification of VV, then 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}) carries a crystal structure isomorphic to 𝐁V\mathbf{B}_{V}. This follows from [CR, Proposition 5.20] and [LV], and is explained in detail in [BDcryst].

Proposition 3.4.

([BDcryst, Theorem 4.31]) The set 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}) together with:

  • Kashiwara operators defined as 𝖤~i(X)=soc𝖤i(X)\widetilde{\mathsf{E}}_{i}(X)=\operatorname{soc}{\mathsf{E}_{i}(X)}, 𝖥~i(X)=soc𝖥i(X)\widetilde{\mathsf{F}}_{i}(X)=\operatorname{soc}{\mathsf{F}_{i}(X)} for X𝖨𝗋𝗋(𝒞)X\in\mathsf{Irr}(\mathcal{C}),

  • wt(X)=μ\text{wt}(X)=\mu for X𝒞μX\in\mathcal{C}_{\mu}, and

  • ε(X)=max{n|𝖤in(X)0}\varepsilon(X)=max\{n\;|\;\mathsf{E}_{i}^{n}(X)\neq 0\}, and φ(X)=max{n|𝖥in(X)0}\varphi(X)=max\{n\;|\;\mathsf{F}_{i}^{n}(X)\neq 0\},

is a 𝔤\mathfrak{g}-crystal isomorphic to the crystal 𝐁=𝐁V\mathbf{B}=\mathbf{B}_{V}.

3.4. Jordan-Hölder series

A categorification of a simple representation (respectively an isotypic representation) is called a simple categorification (respectively an isotypic categorification). There is a distinguished categorification of L(λ)L(\lambda) called the minimal categorification and denoted (λ)\mathcal{L}(\lambda) [CR, Rou2KM, KLI, Webmerged, KK]. It is characterized by the fact that (λ)λ(λ)w0(λ)𝕜mod\mathcal{L}(\lambda)_{\lambda}\cong\mathcal{L}(\lambda)_{w_{0}(\lambda)}\cong\mathbbm{k}\mathrm{-mod}_{\mathbb{Z}}. We let 𝕜low(λ)w0(λ),𝕜high(λ)λ\mathbbm{k}_{low}\in\mathcal{L}(\lambda)_{w_{0}(\lambda)},\mathbbm{k}_{high}\in\mathcal{L}(\lambda)_{\lambda} be the generators.

The Jordan-Hölder Theorem for categorical representations will play an important role in our work. This was originally developed by Rouquier for additive categorifications [Rou2KM], and in this section we transfer these results to the abelian setting. To set this up, recall that given finite abelian 𝕜\mathbbm{k}-linear categories 𝒜,\mathcal{A},\mathcal{B} the Deligne tensor product 𝒜𝕜\mathcal{A}\otimes_{\mathbbm{k}}\mathcal{B} is universal for the functor assigning to every such abelian category 𝒞\mathcal{C} the category of bilinear bifunctors 𝒜×𝒞\mathcal{A}\times\mathcal{B}\to\mathcal{C} right exact in both variables [EGNO, Definition 1.11.1]. The tensor product is again a finite abelian 𝕜\mathbbm{k}-linear category, and there is a bifunctor 𝒜×𝒜𝕜,(X,Y)XY\mathcal{A}\times\mathcal{B}\to\mathcal{A}\otimes_{\mathbbm{k}}\mathcal{B},(X,Y)\mapsto X\otimes Y. This construction enjoys the following properties:

  1. (1)

    The tensor product is unique up to unique equivalence,

  2. (2)

    for finite 𝕜\mathbbm{k}-algebras A,BA,B we have that (Amod)𝕜(Bmod)(A𝕜B)mod(A\mathrm{-mod})\otimes_{\mathbbm{k}}(B\mathrm{-mod})\cong(A\otimes_{\mathbbm{k}}B)\mathrm{-mod}, and

  3. (3)

    Hom𝒜𝕜(X1Y1,X2Y2)Hom𝒜(X1,Y1)Hom(X2,Y2)\operatorname{Hom}_{\mathcal{A}\otimes_{\mathbbm{k}}\mathcal{B}}(X_{1}\otimes Y_{1},X_{2}\otimes Y_{2})\cong\operatorname{Hom}_{\mathcal{A}}(X_{1},Y_{1})\otimes\operatorname{Hom}_{\mathcal{B}}(X_{2},Y_{2}).

Let 𝒞\mathcal{C} be a categorical representation, and let 𝒜\mathcal{A} be a finite 𝕜\mathbbm{k}-linear abelian category. We can endow 𝒞𝕜𝒜\mathcal{C}\otimes_{\mathbbm{k}}\mathcal{A} with a structure of a categorical representation, by setting (𝒞𝕜𝒜)μ=𝒞μ𝕜𝒜(\mathcal{C}\otimes_{\mathbbm{k}}\mathcal{A})_{\mu}=\mathcal{C}_{\mu}\otimes_{\mathbbm{k}}\mathcal{A}, defining Chevalley functors 𝖤i𝟙μ𝟙𝒜\mathsf{E}_{i}\mathbbm{1}_{\mu}\otimes\mathbbm{1}_{\mathcal{A}}, etc. If 𝒞\mathcal{C} is a simple categorification, then clearly 𝒞𝕜𝒜\mathcal{C}\otimes_{\mathbbm{k}}\mathcal{A} is an isotypic categorification. Conversely, we have:

Lemma 3.5.

Let 𝒞\mathcal{C} be an isotypic categorification of type λX+\lambda\in X_{+}. Then there exists an abelian category 𝒜\mathcal{A} such that 𝒞(λ)𝕜𝒜\mathcal{C}\cong\mathcal{L}(\lambda)\otimes_{\mathbbm{k}}\mathcal{A}.

Proof.

Since 𝒞\mathcal{C} is an isotypic categorification, so is 𝒞proj\mathcal{C}\mathrm{-proj}. By Rouquier’s Jordan-Hölder series for additive categorifications [Rou2KM, Theorem 5.8], there exists an additive 𝕜\mathbbm{k}-linear category \mathcal{M} such that

𝒞proj((λ)proj𝕜)i.\displaystyle\mathcal{C}\mathrm{-proj}\cong(\mathcal{L}(\lambda)\mathrm{-proj}\otimes_{\mathbbm{k}}\mathcal{M})^{i}.

Note that no filtration appears here since, in the notation of [Rou2KM], (𝒞proj)λlw=(𝒞proj)λ(\mathcal{C}\mathrm{-proj})_{-\lambda}^{lw}=(\mathcal{C}\mathrm{-proj})_{-\lambda}.

Let {Pi}\{P_{i}\} (respectively {Qj}\{Q_{j}\}) be a complete list of the projective indecomposable objects of (λ)\mathcal{L}(\lambda) (respectively \mathcal{M}). Let P=i,jPiQjP=\bigoplus_{i,j}P_{i}\otimes Q_{j}, and let B=End𝒞(P)opB=\operatorname{End}_{\mathcal{C}}(P)^{op}. By Morita theory, Bmod𝒞B\mathrm{-mod}\cong\mathcal{C}. On the other hand,

BEnd(λ)(iPi)opEnd(jQj)op.\displaystyle B\cong\operatorname{End}_{\mathcal{L}(\lambda)}(\bigoplus_{i}P_{i})^{op}\otimes\operatorname{End}_{\mathcal{M}}(\bigoplus_{j}Q_{j})^{op}.

Since (λ)End(λ)(iPi)opmod\mathcal{L}(\lambda)\cong\operatorname{End}_{\mathcal{L}(\lambda)}(\bigoplus_{i}P_{i})^{op}\mathrm{-mod} we have the desired result with 𝒜=End(jQj)opmod\mathcal{A}=\operatorname{End}_{\mathcal{M}}(\bigoplus_{j}Q_{j})^{op}\mathrm{-mod}. ∎

Theorem 3.6.

Let 𝒞\mathcal{C} be a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}). Then there exists a filtration by Serre subcategories

0=𝒞0𝒞1𝒞n=𝒞,\displaystyle 0=\mathcal{C}_{0}\subset\mathcal{C}_{1}\subset\cdots\subset\mathcal{C}_{n}=\mathcal{C}, (3.7)

such that for each ii: 𝒞i\mathcal{C}_{i} is a subrepresentation of 𝒞\mathcal{C}, 𝒞i/𝒞i1\mathcal{C}_{i}/\mathcal{C}_{i-1} is a simple categorification of type λiX+\lambda_{i}\in X_{+}, and the list of highest weights is weakly increasing so that λiλji<j\lambda_{i}\prec\lambda_{j}\Longrightarrow i<j.

Proof.

The 𝔤\mathfrak{g}-crystal 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}) is isomorphic to a finite direct sum of irreducible crystals 𝐁(λ)\mathbf{B}(\lambda) for various λ\lambda, i.e. we have an isomorphism

𝖨𝗋𝗋(𝒞)λX+𝐁(λ)mλ,\mathsf{Irr}(\mathcal{C})\cong\bigoplus_{\lambda\in X_{+}}\mathbf{B}(\lambda)^{\oplus m_{\lambda}},

where mλ0m_{\lambda}\geq 0 and only finitely many are nonzero.

Define M={λX+|mλ0}M=\{\lambda\in X_{+}\;|\;m_{\lambda}\neq 0\}, and let λM\lambda\in M. We claim that there exists a highest weight simple object L𝒞λL\in\mathcal{C}_{\lambda}. Indeed, otherwise for any simple object L𝒞λL\in\mathcal{C}_{\lambda} there exists iIi\in I such that 𝖤i(X)0\mathsf{E}_{i}(X)\neq 0. This implies that soc(𝖤i(X))0\mathrm{soc}(\mathsf{E}_{i}(X))\neq 0, and by Proposition 3.4 we conclude that 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}) has no highest weight elements of weight λ\lambda, a contradiction.

Now take λM\lambda\in M which is minimal with respect to \preceq, and let L𝒞λL\in\mathcal{C}_{\lambda} be a highest weight object. Let 𝒞1\mathcal{C}_{1} be the Serre subcategory of 𝒞\mathcal{C} generated by objects

{𝖥i1𝖥i(L)|ijI,0}.\displaystyle\{\mathsf{F}_{i_{1}}\cdots\mathsf{F}_{i_{\ell}}(L)\;|\;i_{j}\in I,\ell\geq 0\}. (3.8)

By the exactness and bi-adjunction of the Chevalley functors, 𝒞1\mathcal{C}_{1} is a subrepresentation of 𝒞\mathcal{C}. Moreover it categorifies L(λ)L(\lambda). Indeed, by our choice of λ\lambda there cannot be any highest weight objects with weight λ\prec\lambda occuring in 𝒞1\mathcal{C}_{1}, and by construction the only simple object in (𝒞1)λ(\mathcal{C}_{1})_{\lambda} is LL.

Next consider the categorical representation 𝒞/𝒞1\mathcal{C}/\mathcal{C}_{1} and repeat this construction. This produces a Serre subcategory 𝒞2𝒞/𝒞1\mathcal{C}_{2}^{\prime}\subset\mathcal{C}/\mathcal{C}_{1} which is again a simple categorification. Let π:𝒞𝒞/𝒞1\pi:\mathcal{C}\to\mathcal{C}/\mathcal{C}_{1} be the natural quotient functor, and define 𝒞2=π1(𝒞2)\mathcal{C}_{2}=\pi^{-1}(\mathcal{C}_{2}^{\prime}). Clearly, we have that 𝒞1𝒞2\mathcal{C}_{1}\subset\mathcal{C}_{2}, 𝒞2\mathcal{C}_{2} is Serre, it is a subrepresentation, and 𝒞2/𝒞1𝒞2\mathcal{C}_{2}/\mathcal{C}_{1}\cong\mathcal{C}_{2}^{\prime} is a simple categorification.

Iterating this process produces a filtration of 𝒞\mathcal{C} such that each composition factor is a simple categorification, and the highest weights of the subquotients are weakly increasing. ∎

Remark 3.9.

Note that the construction in the proof of Theorem 3.6 can produce also an isotypic filtration with similar properties. Namely, if λ1,,λN\lambda_{1},\ldots,\lambda_{N} is a list of the distinct isotypic types appearing in [𝒞](q)[\mathcal{C}]_{\mathbb{C}(q)}, and we choose any ordering of this list so that λiλji<j\lambda_{i}\prec\lambda_{j}\Longrightarrow i<j, then there is a filtration 0=𝒞0𝒞1𝒞N=𝒞0=\mathcal{C}^{\prime}_{0}\subset\mathcal{C}^{\prime}_{1}\subset\cdots\subset\mathcal{C}^{\prime}_{N}=\mathcal{C} such that 𝒞k/𝒞k1\mathcal{C}^{\prime}_{k}/\mathcal{C}^{\prime}_{k-1} categorifies the isotypic component of [𝒞](q)[\mathcal{C}]_{\mathbb{C}(q)} of highest weight λk\lambda_{k}. To construct this isotypic filtration, consider the Jordan-Hölder filtration from the theorem. From the proof of Theorem 3.6, it’s easy to see that one can ensure that the subquotients which categorify the same simple representations appear in sequence. Assuming then that our Jordan-Hölder filtration satisfies this property, a coarsening of it is the desired isotypic filtration.

Remark 3.10.

Note that one can read off the isotypic filtration of 𝒞\mathcal{C} from the crystal structure on 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}). Indeed, suppose that 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}) decomposes into components

𝖨𝗋𝗋(𝒞)=𝐗(λ1)𝐗(λN),\mathsf{Irr}(\mathcal{C})=\mathbf{X}(\lambda_{1})\sqcup\cdots\sqcup\mathbf{X}(\lambda_{N}),

where λ1,,λNX+\lambda_{1},\ldots,\lambda_{N}\in X_{+} are distinct dominant integral weights, and 𝐗(λi)\mathbf{X}(\lambda_{i}) is a disjoint union of copies of 𝐁(λi)\mathbf{B}(\lambda_{i}). Further, we arrange the weights as above so that λiλji<j\lambda_{i}\prec\lambda_{j}\Longrightarrow i<j. Let 𝒞i\mathcal{C}_{i} be the Serre subcategory of 𝒞\mathcal{C} generated by simple objects LL such that [L]𝐗(λj)[L]\in\mathbf{X}(\lambda_{j}), where jij\leq i. Then it follows from Remark 3.9 that {0}𝒞1𝒞2𝒞N\{0\}\subset\mathcal{C}_{1}\subset\mathcal{C}_{2}\subset\cdots\subset\mathcal{C}_{N} is an isotypic filtration of 𝒞\mathcal{C}.

3.5. The categorical braid group action

Let 𝒞\mathcal{C} be a categorical representation of Uq(𝔤),μXU_{q}(\mathfrak{g}),\mu\in X and iIi\in I. We define a complex of functors Θi𝟙μ\Theta_{i}\mathbbm{1}_{\mu}, supported in nonpositive cohomological degrees, where for r0r\geq 0 the r-r component is

(Θi𝟙μ)r={𝖤i(μi+r)𝖥i(r)𝟙μs if μi0,𝖥i(μi+r)𝖤i(r)𝟙μr if μi0.\displaystyle(\Theta_{i}\mathbbm{1}_{\mu})^{-r}=\begin{cases}\mathsf{E}_{i}^{(-\mu_{i}+r)}\mathsf{F}_{i}^{(r)}\mathbbm{1}_{\mu}\langle-s\rangle&\text{ if }\mu_{i}\leq 0,\\ \mathsf{F}_{i}^{(\mu_{i}+r)}\mathsf{E}_{i}^{(r)}\mathbbm{1}_{\mu}\langle-r\rangle&\text{ if }\mu_{i}\geq 0.\end{cases}

The differential dr:(Θi𝟙μ)r(Θi𝟙μ)r+1d^{r}:(\Theta_{i}\mathbbm{1}_{\mu})^{-r}\to(\Theta_{i}\mathbbm{1}_{\mu})^{-r+1} is defined using the counits of the bi-adjunctions relating 𝖤i\mathsf{E}_{i} and 𝖥i\mathsf{F}_{i} (see [Cauclasp, Section 4] for details). This produces a functor Θi𝟙μ:Db(𝒞μ)Db(𝒞si(μ))\Theta_{i}\mathbbm{1}_{\mu}:D^{b}(\mathcal{C}_{\mu})\to D^{b}(\mathcal{C}_{s_{i}(\mu)}), which following Chuang and Rouquier we call the Rickard complex.

It’s straightforward to verify that the Rickard complex Θi𝟙μ\Theta_{i}\mathbbm{1}_{\mu} categorifies Lusztig’s braid group operators 𝗍i𝟣μ\mathsf{t}_{i}\mathsf{1}_{\mu} ([Cauclasp, Section 2]). On the level of categories we have the following two theorems of Chuang-Rouquier and Cautis-Kamnitzer, which are the fundamental results about Rickard complexes. Note that the latter theorem was conjectured in [Rou2KM, Conjecture 5.19].

Theorem 3.11.

[CR, Theorem 6.4] For any μ,i\mu,i, Θi𝟙μ:Db(𝒞μ)Db(𝒞si(μ))\Theta_{i}\mathbbm{1}_{\mu}:D^{b}(\mathcal{C}_{\mu})\to D^{b}(\mathcal{C}_{s_{i}(\mu)}) is an equivalence of triangulated categories.

Theorem 3.12.

[CK3, Theorem 6.3] The Rickard complexes satisfy the braid relations:

ΘiΘj𝟙μ\displaystyle\Theta_{i}\Theta_{j}\mathbbm{1}_{\mu} ΘjΘi𝟙μ if (i,j)=0,\displaystyle\cong\Theta_{j}\Theta_{i}\mathbbm{1}_{\mu}\text{ if }(i,j)=0,
ΘiΘjΘi𝟙μ\displaystyle\Theta_{i}\Theta_{j}\Theta_{i}\mathbbm{1}_{\mu} ΘjΘiΘj𝟙μ if (i,j)=1,\displaystyle\cong\Theta_{j}\Theta_{i}\Theta_{j}\mathbbm{1}_{\mu}\text{ if }(i,j)=-1,

thereby defining a weak action of BB on Db(𝒞)D^{b}(\mathcal{C}).

This action is “weak” since we don’t make any claim on the canonicity of the functorial isomorphisms. Nevertheless, for wWw\in W we define Θw𝟙μ:=Θi1Θi𝟙μ\Theta_{w}\mathbbm{1}_{\mu}:=\Theta_{i_{1}}\circ\cdots\circ\Theta_{i_{\ell}}\mathbbm{1}_{\mu}, where w=si1siw=s_{i_{1}}\cdots s_{i_{\ell}} is a reduced expression. Thus Θw𝟙μ\Theta_{w}\mathbbm{1}_{\mu} is defined up to isomorphism, but not canonical isomorphism. Luckily, everything we do in this paper only requires Θw𝟙μ\Theta_{w}\mathbbm{1}_{\mu} to be defined up to isomorphism.

As a consequence of their proof of Theorem 3.11, Chuang and Rouquier show that the inverse of Θi𝟙μ\Theta_{i}\mathbbm{1}_{\mu} is its right adjoint. We denote this functor by Θi𝟙μ:Db(𝒞μ)Db(𝒞si(μ))\Theta_{i}^{\prime}\mathbbm{1}_{\mu}:D^{b}(\mathcal{C}_{\mu})\to D^{b}(\mathcal{C}_{s_{i}(\mu)}), so that ΘiΘi𝟙μΘiΘi𝟙μ𝟙μ\Theta_{i}\Theta_{i}^{\prime}\mathbbm{1}_{\mu}\cong\Theta^{\prime}_{i}\Theta_{i}\mathbbm{1}_{\mu}\cong\mathbbm{1}_{\mu}. As a complex of functors, Θi𝟙μ\Theta_{i}^{\prime}\mathbbm{1}_{\mu} is supported in nonnegative cohomological degrees, where for s0s\geq 0 the ss component is

(Θi𝟙μ)s={𝖤i(s)𝖥i(μi+s)𝟙μs(2μi2s+1) if μi0,𝖥i(s)𝖤i(μi+s)𝟙μs(2μi+2s+1) if μi0.\displaystyle(\Theta_{i}^{\prime}\mathbbm{1}_{\mu})^{s}=\begin{cases}\mathsf{E}_{i}^{(s)}\mathsf{F}_{i}^{(\mu_{i}+s)}\mathbbm{1}_{\mu}\langle s(-2\mu_{i}-2s+1)\rangle&\text{ if }\mu_{i}\geq 0,\\ \mathsf{F}_{i}^{(s)}\mathsf{E}_{i}^{(-\mu_{i}+s)}\mathbbm{1}_{\mu}\langle s(-2\mu_{i}+2s+1)\rangle&\text{ if }\mu_{i}\leq 0.\end{cases}

4. Perverse equivalences

4.1. General definition

Let 𝒯\mathcal{T} be a triangulated category with shift functor [1]:𝒯𝒯[1]:\mathcal{T}\rightarrow\mathcal{T}. In the cases of most interest to us, 𝒯\mathcal{T} is a subcategory of a derived category, in which case [1][1] is the homological shift functor. Suppose 𝒯\mathcal{T} has a t-structure t=(𝒯0,𝒯0)t=(\mathcal{T}^{\leq 0},\mathcal{T}^{\geq 0}), with heart 𝒯=𝒯0𝒯0\mathcal{T}^{\heartsuit}=\mathcal{T}^{\leq 0}\cap\mathcal{T}^{\geq 0} [BBD]. Recall that a triangulated functor F:𝒯𝒮F:\mathcal{T}\to\mathcal{S} between triangulated categories with tt-structure is tt-exact if F(𝒯0)𝒮0F(\mathcal{T}^{\leq 0})\subseteq\mathcal{S}^{\leq 0} and F(𝒯0)𝒮0F(\mathcal{T}^{\geq 0})\subseteq\mathcal{S}^{\geq 0}. We let F[p]:𝒯𝒮F[p]:\mathcal{T}\to\mathcal{S} denote the pre-composition of FF with the pp-shift [p][p].

Now let 𝒮𝒯\mathcal{S}\subset\mathcal{T} be a thick triangulated subcategory, and consider the quotient functor Q:𝒯𝒯/𝒮Q:\mathcal{T}\to\mathcal{T}/\mathcal{S}. Following [CRperv], we say that tt is compatible with 𝒮\mathcal{S} if t𝒯/𝒮=(Q(𝒯0),Q(𝒯0))t_{\mathcal{T}/\mathcal{S}}=(Q(\mathcal{T}^{\leq 0}),Q(\mathcal{T}^{\geq 0})) is a t-structure on 𝒯/𝒮\mathcal{T}/\mathcal{S}. By [CRperv, Lemmas 3.3 & 3.9], if tt is compatible with 𝒮\mathcal{S} then (𝒯/𝒮)=𝒯/𝒯𝒮(\mathcal{T}/\mathcal{S})^{\heartsuit}=\mathcal{T}^{\heartsuit}/\mathcal{T}^{\heartsuit}\cap\mathcal{S}, and t𝒮=(𝒮𝒯0,𝒮𝒯0)t_{\mathcal{S}}=(\mathcal{S}\cap\mathcal{T}^{\leq 0},\mathcal{S}\cap\mathcal{T}^{\geq 0}) is a t-structure on 𝒮\mathcal{S} such that 𝒮=𝒯𝒮\mathcal{S}^{\heartsuit}=\mathcal{T}^{\heartsuit}\cap\mathcal{S}.

Now suppose that 𝒯,𝒯\mathcal{T},\mathcal{T}^{\prime} are two triangulated categories with t-structures t,tt,t^{\prime}. Suppose further that we have filtrations by thick triangulated subcategories:

0𝒯0𝒯1𝒯r=𝒯,0𝒯0𝒯1𝒯r=𝒯,\displaystyle 0\subset\mathcal{T}_{0}\subset\mathcal{T}_{1}\subset\cdots\subset\mathcal{T}_{r}=\mathcal{T},\quad 0\subset\mathcal{T}_{0}^{\prime}\subset\mathcal{T}_{1}^{\prime}\subset\cdots\subset\mathcal{T}_{r}^{\prime}=\mathcal{T}^{\prime},

such that for every ii, tt is compatible with 𝒯i\mathcal{T}_{i} and tt^{\prime} is compatible with 𝒯i\mathcal{T}^{\prime}_{i}. By [CRperv, Lemma 3.11], t𝒯it_{\mathcal{T}_{i}} is also compatible with 𝒯i1\mathcal{T}_{i-1}, and hence 𝒯i/𝒯i1\mathcal{T}_{i}/\mathcal{T}_{i-1} inherits a natural t-structure. Let p:{0,,r}p:\{0,\ldots,r\}\rightarrow\mathbb{Z}. The data (𝒯,𝒯,p)(\mathcal{T}_{\bullet},\mathcal{T}^{\prime}_{\bullet},p) is termed a perversity triple.

Although Chuang and Rouquier didn’t formulate perverse equivalences for graded categories, it is straightforward to extend their definitions to this setting.

Definition 4.1.

A graded equivalence of graded triangulated categories 𝖥:𝒯𝒯\mathsf{F}:\mathcal{T}\to\mathcal{T}^{\prime} is a (graded) perverse equivalence with respect to (𝒯,𝒯,p)(\mathcal{T}_{\bullet},\mathcal{T}^{\prime}_{\bullet},p) if for every ii,

  1. (1)

    𝖥(𝒯i)=𝒯i\mathsf{F}(\mathcal{T}_{i})=\mathcal{T}_{i}^{\prime}, and

  2. (2)

    the induced equivalence 𝖥[p(i)]:𝒯i/𝒯i1𝒯i/𝒯i1\mathsf{F}[-p(i)]:\mathcal{T}_{i}/\mathcal{T}_{i-1}\to\mathcal{T}_{i}^{\prime}/\mathcal{T}_{i-1}^{\prime} is t-exact.

For brevity, we say 𝖥\mathsf{F} is perverse if it is a graded perverse equivalence with respect to some perversity datum. Since we will be working exclusively in the graded setting, a perverse equivalence for us will always mean a graded perverse equivalence.

A perverse equivalence 𝖥:𝒯𝒯\mathsf{F}:\mathcal{T}\to\mathcal{T}^{\prime} induces a bijection φ𝖥:𝖨𝗋𝗋(𝒯)𝖨𝗋𝗋(𝒯)\varphi_{\mathsf{F}}:\mathsf{Irr}(\mathcal{T}^{\heartsuit})\to\mathsf{Irr}(\mathcal{T}^{\prime\heartsuit}). Indeed, by (2) 𝖥[p(i)]\mathsf{F}[-p(i)] induces a bijection 𝖨𝗋𝗋(𝒯i)𝖨𝗋𝗋(𝒯i1)𝖨𝗋𝗋(𝒯i)𝖨𝗋𝗋(𝒯i1)\mathsf{Irr}(\mathcal{T}_{i}^{\heartsuit})\setminus\mathsf{Irr}(\mathcal{T}_{i-1}^{\heartsuit})\to\mathsf{Irr}(\mathcal{T}_{i}^{\prime\heartsuit})\setminus\mathsf{Irr}(\mathcal{T}_{i-1}^{\prime\heartsuit}), and these yield φ𝖥\varphi_{\mathsf{F}}.

Although the construction of φ𝖥\varphi_{\mathsf{F}} depends on a choice of perversity triple, the resulting bijection does not when 𝒯,𝒯\mathcal{T}^{\heartsuit},\mathcal{T}^{\prime\heartsuit} have finitely many simple objects. This follows from the following lemma.

Lemma 4.2 ([LosCacti], Lemma 2.4).

Suppose that 𝒯,𝒯\mathcal{T}^{\heartsuit},\mathcal{T}^{\prime\heartsuit} have finitely many simple objects, and for i=1,2i=1,2 let 𝖥i:𝒯𝒯\mathsf{F}_{i}:\mathcal{T}\to\mathcal{T}^{\prime} be a perverse equivalence with respect to the perversity datum (𝒯i,,𝒯i,,pi)(\mathcal{T}_{i,\bullet},\mathcal{T}^{\prime}_{i,\bullet},p_{i}). If the induced maps [𝖥1],[𝖥2]:[𝒯][𝒯][\mathsf{F}_{1}],[\mathsf{F}_{2}]:[\mathcal{T}]_{\mathbb{Z}}\to[\mathcal{T}^{\prime}]_{\mathbb{Z}} coincide then φ𝖥1=φ𝖥2\varphi_{\mathsf{F}_{1}}=\varphi_{\mathsf{F}_{2}}.

Corollary 4.3.

Suppose that 𝒯,𝒯\mathcal{T}^{\heartsuit},\mathcal{T}^{\prime\heartsuit} have finitely many simple objects, and 𝖥:𝒯𝒯\mathsf{F}:\mathcal{T}\to\mathcal{T}^{\prime} is perverse. Then φ𝖥\varphi_{\mathsf{F}} is independent of the choice of perversity triple.

Proof.

Suppose 𝖥\mathsf{F} is a graded perverse equivalence with respect to two choices of perversity triples (𝒯i,,𝒯i,,pi)(\mathcal{T}_{i,\bullet},\mathcal{T}^{\prime}_{i,\bullet},p_{i}), i=1,2i=1,2. Now apply the lemma. ∎

The proofs of the following lemmas are straightforward.

Lemma 4.4.

Suppose 𝖥:𝒯𝒯\mathsf{F}:\mathcal{T}\rightarrow\mathcal{T}^{\prime} is perverse. Then for any \ell\in\mathbb{Z}, 𝖥[]\mathsf{F}[\ell] is also perverse and φ𝖥[]=φ𝖥\varphi_{\mathsf{F}[\ell]}=\varphi_{\mathsf{F}}.

Lemma 4.5.

Suppose 𝖥:𝒯𝒯\mathsf{F}:\mathcal{T}\rightarrow\mathcal{T}^{\prime} is a perverse equivalence with respect to (𝒯,𝒯,p)(\mathcal{T}_{\bullet},\mathcal{T}^{\prime}_{\bullet},p), and 𝖦:𝒯𝒯′′\mathsf{G}:\mathcal{T}^{\prime}\rightarrow\mathcal{T}^{\prime\prime} is a perverse equivalence with respect to (𝒯,𝒯′′,q)(\mathcal{T}^{\prime}_{\bullet},\mathcal{T}^{\prime\prime}_{\bullet},q). Then 𝖦𝖥\mathsf{G}\circ\mathsf{F} is a perverse equivalence with respect to (𝒯,𝒯′′,p+q)(\mathcal{T}_{\bullet},\mathcal{T}^{\prime\prime}_{\bullet},p+q), and φ𝖦𝖥=φ𝖦φ𝖥\varphi_{\mathsf{G}\circ\mathsf{F}}=\varphi_{\mathsf{G}}\circ\varphi_{\mathsf{F}}.

Lemma 4.6.

Let 𝒯,𝒯\mathcal{T},\mathcal{T}^{\prime} be triangulated with t-structures t,tt,t^{\prime}, and let 𝒮𝒯,𝒮𝒯\mathcal{S}\subset\mathcal{T},\mathcal{S}^{\prime}\subset\mathcal{T}^{\prime} be thick triangulated subcategories such that tt is compatible with 𝒮\mathcal{S} and tt^{\prime} is compatible with 𝒮\mathcal{S}^{\prime}. Suppose further that 𝖥:𝒯𝒯\mathsf{F}:\mathcal{T}\to\mathcal{T}^{\prime} is a perverse equivalence with respect to (𝒯,𝒯,p)(\mathcal{T}_{\bullet},\mathcal{T}^{\prime}_{\bullet},p), and 𝖥(𝒮)=𝒮\mathsf{F}(\mathcal{S})=\mathcal{S}^{\prime}.

Define 𝒮i=𝒮𝒯i,𝒮i=𝒮𝒯i\mathcal{S}_{i}=\mathcal{S}\cap\mathcal{T}_{i},\mathcal{S}_{i}^{\prime}=\mathcal{S}^{\prime}\cap\mathcal{T}_{i}^{\prime} and (𝒯/𝒮)i=Q(𝒯i),(𝒯/𝒮)i=Q(𝒯i)(\mathcal{T}/\mathcal{S})_{i}=Q(\mathcal{T}_{i}),(\mathcal{T}^{\prime}/\mathcal{S}^{\prime})_{i}=Q^{\prime}(\mathcal{T}_{i}^{\prime}). Let 𝖦:𝒮𝒮\mathsf{G}:\mathcal{S}\to\mathcal{S}^{\prime} and 𝖧:𝒯/𝒮𝒯/𝒮\mathsf{H}:\mathcal{T}/\mathcal{S}\to\mathcal{T}^{\prime}/\mathcal{S}^{\prime} be the induced equivalences. Then:

  1. (1)

    𝖦\mathsf{G} is a perverse equivalence with respect to (𝒮,𝒮,p)(\mathcal{S}_{\bullet},\mathcal{S}^{\prime}_{\bullet},p), and φ𝖦=φ𝖥|𝖨𝗋𝗋𝒮\varphi_{\mathsf{G}}=\varphi_{\mathsf{F}}|_{\mathsf{Irr}{\mathcal{S}^{\heartsuit}}}.

  2. (2)

    𝖧\mathsf{H} is a perverse equivalence with respect to ((𝒯/𝒮),(𝒯/𝒮),p)((\mathcal{T}/\mathcal{S})_{\bullet},(\mathcal{T}^{\prime}/\mathcal{S}^{\prime})_{\bullet},p), and φ𝖧=φ𝖥|𝖨𝗋𝗋𝒯𝖨𝗋𝗋𝒮.\varphi_{\mathsf{H}}=\varphi_{\mathsf{F}}|_{\mathsf{Irr}{\mathcal{T}^{\heartsuit}}\setminus\mathsf{Irr}{\mathcal{S}^{\heartsuit}}}.

Lemma 4.7 ([LosCacti], Lemma 2.4).

Suppose 𝖥:𝒯𝒯\mathsf{F}:\mathcal{T}\rightarrow\mathcal{T}^{\prime} is perverse, and 𝖦\mathsf{G} (respectively 𝖦\mathsf{G}^{\prime}) is an autoequivalence of 𝒯\mathcal{T} (respectively 𝒯\mathcal{T}^{\prime}) which is t-exact up to shift. Then 𝖦𝖥𝖦\mathsf{G}^{\prime}\circ\mathsf{F}\circ\mathsf{G} is perverse, and φ𝖦𝖥𝖦=φ𝖦φ𝖥φ𝖦\varphi_{\mathsf{G}^{\prime}\circ\mathsf{F}\circ\mathsf{G}}=\varphi_{\mathsf{G}^{\prime}}\circ\varphi_{\mathsf{F}}\circ\varphi_{\mathsf{G}^{\prime}}.

Note that in [LosCacti, Lemma 2.4] is stated for functors which are t-exact. Our formulation for functors which are t-exact up to shift follows by Lemma 4.4.

4.2. Derived categories of graded abelian categories

We now specialise to the case of derived categories. We recall that if 𝒜\mathcal{A} is an abelian category, then the bounded derived category Db(𝒜)D^{b}(\mathcal{A}) has a standard t-structure (Db(𝒜)0,Db(𝒜)0)(D^{b}(\mathcal{A})^{\leq 0},D^{b}(\mathcal{A})^{\geq 0}) whose heart is 𝒜\mathcal{A}.

Given a 𝒜\mathcal{B}\subset\mathcal{A} a Serre subcategory, we let Db(𝒜)Db(𝒜)D_{\mathcal{B}}^{b}(\mathcal{A})\subset D^{b}(\mathcal{A}) denote the thick subcategory consisting of complexes with cohomology supported in \mathcal{B}. The category Db(𝒜)D_{\mathcal{B}}^{b}(\mathcal{A}) inherits a natural t-structure from the standard t-structure on Db(𝒜)D^{b}(\mathcal{A}): Db(𝒜)0=Db(𝒜)Db(𝒜)0D_{\mathcal{B}}^{b}(\mathcal{A})^{\leq 0}=D_{\mathcal{B}}^{b}(\mathcal{A})\cap D^{b}(\mathcal{A})^{\leq 0} and Db(𝒜)0=Db(𝒜)Db(𝒜)0D_{\mathcal{B}}^{b}(\mathcal{A})^{\geq 0}=D_{\mathcal{B}}^{b}(\mathcal{A})\cap D^{b}(\mathcal{A})^{\geq 0}. The heart of the t-structure on Db(𝒜)D_{\mathcal{B}}^{b}(\mathcal{A}) is \mathcal{B}. Moreover, if 𝒞\mathcal{C}\subset\mathcal{B} is another Serre subcategory then the tt-structure on Db(𝒜)D_{\mathcal{B}}^{b}(\mathcal{A}) is compatible with D𝒞b(𝒜)D_{\mathcal{C}}^{b}(\mathcal{A}). In particular, the quotient Db(𝒜)/D𝒞b(𝒜)D_{\mathcal{B}}^{b}(\mathcal{A})/D_{\mathcal{C}}^{b}(\mathcal{A}) inherits a natural t-structure whose heart is /𝒞\mathcal{B}/\mathcal{C}.

For the remainder of this section let 𝒜,𝒜\mathcal{A},\mathcal{A}^{\prime} be graded abelian categories. In the setting of derived categories of graded abelian categories, a perverse equivalence can be packaged as follows. We can encode a perversity triple (𝒜,𝒜,p)(\mathcal{A}_{\bullet},\mathcal{A}^{\prime}_{\bullet},p) using filtrations on the abelian categories: 𝒜\mathcal{A}_{\bullet} and 𝒜\mathcal{A}^{\prime}_{\bullet} are filtrations by shift-invariant Serre subcategories:

0=𝒜1𝒜0𝒜1𝒜r=𝒜,0=𝒜1𝒜0𝒜1𝒜r=𝒜.\displaystyle 0=\mathcal{A}_{-1}\subset\mathcal{A}_{0}\subset\mathcal{A}_{1}\subset\ldots\subset\mathcal{A}_{r}=\mathcal{A},\quad 0=\mathcal{A}^{\prime}_{-1}\subset\mathcal{A}^{\prime}_{0}\subset\mathcal{A}^{\prime}_{1}\subset\ldots\subset\mathcal{A}^{\prime}_{r}=\mathcal{A}^{\prime}.

Then a graded equivalence 𝖥:Db(𝒜)Db(𝒜)\mathsf{F}:D^{b}(\mathcal{A})\to D^{b}(\mathcal{A}^{\prime}) is a perverse with respect to (𝒜,𝒜,p)(\mathcal{A}_{\bullet},\mathcal{A}^{\prime}_{\bullet},p) if conditions (1) and (2) of Definition 4.1 hold for 𝒯i=D𝒜ib(𝒜)\mathcal{T}_{i}=D^{b}_{\mathcal{A}_{i}}(\mathcal{A}) and 𝒯i=D𝒜ib(𝒜)\mathcal{T}_{i}^{\prime}=D^{b}_{\mathcal{A}^{\prime}_{i}}(\mathcal{A}^{\prime})

As above, a graded perverse equivalence 𝖥:Db(𝒜)Db(𝒜)\mathsf{F}:D^{b}(\mathcal{A})\to D^{b}(\mathcal{A}^{\prime}) induces a bijection φ𝖥:𝖨𝗋𝗋(𝒜)𝖨𝗋𝗋(𝒜)\varphi_{\mathsf{F}}:\mathsf{Irr}(\mathcal{A})\to\mathsf{Irr}(\mathcal{A}^{\prime}).

The following standard lemma will be useful in the proof of our main result.

Lemma 4.8.

Let 𝒜,𝒜\mathcal{A},\mathcal{A}^{\prime} be abelian categories, and ,\mathcal{B},\mathcal{B}^{\prime} Serre subcategories. Let aba\leq b be integers, and 𝖥i:𝒜𝒜\mathsf{F}_{i}:\mathcal{A}\to\mathcal{A}^{\prime} be exact functors for aiba\leq i\leq b. Suppose these functors fit into a complex 𝖥=(𝖥a𝖥a+1𝖥b)\mathsf{F}=(\mathsf{F}_{a}\to\mathsf{F}_{a+1}\to\cdots\to\mathsf{F}_{b}), defining a functor

𝖥:Db(𝒜)Db(𝒜).\mathsf{F}:D^{b}(\mathcal{A})\to D^{b}(\mathcal{A}^{\prime}).

If 𝖥i()\mathsf{F}_{i}(\mathcal{B})\subset\mathcal{B}^{\prime} for all aiba\leq i\leq b, then 𝖥(Db(𝒜))Db(𝒜)\mathsf{F}(D_{\mathcal{B}}^{b}(\mathcal{A}))\subseteq D_{\mathcal{B}^{\prime}}^{b}(\mathcal{A}^{\prime}).

5. Some commutation relations

We fix throughout a categorical representation 𝒞\mathcal{C} of Uq(𝔤)U_{q}(\mathfrak{g}). Recall that w0Ww_{0}\in W is the longest word, and let w0=si1si2sirw_{0}=s_{i_{1}}s_{i_{2}}\cdots s_{i_{r}} be a reduced expression. We consider the composition of Rickard complexes which categorifies the positive lift in BB of w0w_{0}:

Θw0𝟙λ=Θi1Θir𝟙μ:Db(𝒞λ)Db(𝒞w0(λ)).\Theta_{w_{0}}\mathbbm{1}_{\lambda}=\Theta_{i_{1}}\cdots\Theta_{i_{r}}\mathbbm{1}_{\mu}:D^{b}(\mathcal{C}_{\lambda})\to D^{b}(\mathcal{C}_{w_{0}(\lambda)}).

In preparation for the proofs our main results in the next section, we prove some commutation relations between Θw0\Theta_{w_{0}} and the Chevalley functors.

5.1. Cautis’ relations

To begin, we recall some relations of Cautis (building on work with Kamnitzer [CK3]). Although they are stated only for type A, their proofs apply to any simply-laced Lie algebra.

Lemma 5.1 (Lemma 4.6, [Cauclasp]).

For any iI,λXi\in I,\lambda\in X we have the following relations:

Θi𝖤i𝟙λ\displaystyle\Theta_{i}\mathsf{E}_{i}\mathbbm{1}_{\lambda} 𝖥iΘi𝟙λ[1]λi,\displaystyle\cong\mathsf{F}_{i}\Theta_{i}\mathbbm{1}_{\lambda}[1]\langle\lambda_{i}\rangle,
Θi𝖥i𝟙λ\displaystyle\Theta_{i}\mathsf{F}_{i}\mathbbm{1}_{\lambda} 𝖤iΘi𝟙λ[1]λi.\displaystyle\cong\mathsf{E}_{i}\Theta_{i}\mathbbm{1}_{\lambda}[1]\langle-\lambda_{i}\rangle.
Remark 5.2.

The careful reader will notice that actually Cautis proves Lemma 5.1 under certain conditions on λ\lambda. For instance, the first relation is only proven in the case when λi0\lambda_{i}\leq 0. To deduce the general case from this, one can rewrite the relation as 𝖤iΘi1Θi1𝖥i[1]λi\mathsf{E}_{i}\Theta_{i}^{-1}\cong\Theta_{i}^{-1}\mathsf{F}_{i}[1]\langle\lambda_{i}\rangle. Now recall that there is an anti-automorphism σ~\widetilde{\sigma} on the 𝔰𝔩2\mathfrak{sl}_{2} 2-category which on objects maps nnn\mapsto-n [Laucq, Section 5.6]. This anti-automorphism maps Θi1\Theta_{i}^{-1} to Θi\Theta_{i}, and hence applying it to the relation above we deduce the desired relation in the case when λi0\lambda_{i}\geq 0.

Alternatively, in a recent preprint Vera proves a version of the relation between the Rickard complex and Chevalley functors in the (bounded homotopy category of the) 𝔰𝔩2\mathfrak{sl}_{2} 2-category, which of course implies it also in any 2-representation [Vera].

Next we recall the categorical analogues of commutators [Ei,Ej][E_{i},E_{j}] acting on representations of Uq(𝔤)U_{q}(\mathfrak{g}). Given nodes i,jIi,j\in I such that (i,j)=1(i,j)=-1 and λX\lambda\in X, define complexes of functors

𝖤ij𝟙λ:Db(𝒞λ)Db(𝒞λ+αi+αj),𝖤ij𝟙λ\displaystyle\mathsf{E}_{ij}\mathbbm{1}_{\lambda}:D^{b}(\mathcal{C}_{\lambda})\to D^{b}(\mathcal{C}_{\lambda+\alpha_{i}+\alpha_{j}}),\quad\mathsf{E}_{ij}\mathbbm{1}_{\lambda} =𝖤i𝖤j𝟙λ1𝖤j𝖤i𝟙λ,\displaystyle=\mathsf{E}_{i}\mathsf{E}_{j}\mathbbm{1}_{\lambda}\langle-1\rangle\to\mathsf{E}_{j}\mathsf{E}_{i}\mathbbm{1}_{\lambda},
𝖥ij𝟙λ:Db(𝒞λ)Db(𝒞λαiαj),𝖥ij𝟙λ\displaystyle\mathsf{F}_{ij}\mathbbm{1}_{\lambda}:D^{b}(\mathcal{C}_{\lambda})\to D^{b}(\mathcal{C}_{\lambda-\alpha_{i}-\alpha_{j}}),\quad\mathsf{F}_{ij}\mathbbm{1}_{\lambda} =𝖥i𝖥j𝟙λ𝖥j𝖥i𝟙λ1.\displaystyle=\mathsf{F}_{i}\mathsf{F}_{j}\mathbbm{1}_{\lambda}\to\mathsf{F}_{j}\mathsf{F}_{i}\mathbbm{1}_{\lambda}\langle 1\rangle.

In both instances the differential is given by the element TijT_{ij} arising from the KLR algebra, and the left term of the complex is in homological degree zero [Cauclasp].

Lemma 5.3 (Lemma 5.2, [Cauclasp]).

Let i,jI,λXi,j\in I,\lambda\in X and suppose (i,j)=1(i,j)=-1. We have the following isomorphisms:

𝖤ijΘi𝟙λ\displaystyle\mathsf{E}_{ij}\Theta_{i}\mathbbm{1}_{\lambda} {Θi𝖤j if λi>0,Θi𝖤j[1]1 if λi0\displaystyle\cong\begin{cases}\Theta_{i}\mathsf{E}_{j}&\text{ if }\lambda_{i}>0,\\ \Theta_{i}\mathsf{E}_{j}[1]\langle-1\rangle&\text{ if }\lambda_{i}\leq 0\end{cases}
𝖥ijΘi𝟙λ\displaystyle\mathsf{F}_{ij}\Theta_{i}\mathbbm{1}_{\lambda} {Θi𝖥j if λi0,Θi𝖥j[1]1 if λi<0\displaystyle\cong\begin{cases}\Theta_{i}\mathsf{F}_{j}&\text{ if }\lambda_{i}\geq 0,\\ \Theta_{i}\mathsf{F}_{j}[-1]\langle 1\rangle&\text{ if }\lambda_{i}<0\end{cases}
𝟙λΘj𝖤ij\displaystyle\mathbbm{1}_{\lambda}\Theta_{j}\mathsf{E}_{ij} {𝖤iΘj if λj<0,𝖤iΘj[1]1 if λj0\displaystyle\cong\begin{cases}\mathsf{E}_{i}\Theta_{j}&\text{ if }\lambda_{j}<0,\\ \mathsf{E}_{i}\Theta_{j}[1]\langle-1\rangle&\text{ if }\lambda_{j}\geq 0\end{cases}
𝟙λΘj𝖥ij\displaystyle\mathbbm{1}_{\lambda}\Theta_{j}\mathsf{F}_{ij} {𝖥iΘj if λj0,𝖥iΘj[1]1 if λj>0\displaystyle\cong\begin{cases}\mathsf{F}_{i}\Theta_{j}&\text{ if }\lambda_{j}\leq 0,\\ \mathsf{F}_{i}\Theta_{j}[-1]\langle 1\rangle&\text{ if }\lambda_{j}>0\end{cases}

5.2. Marked words

We now introduce a combinatorial set-up which we’ll use to prove Proposition 5.9 below. A marked word is a word in the elements of II with one letter marked: 𝐚=(i1,i2,,i¯,,in)\mathbf{a}=(i_{1},i_{2},...,\underline{i_{\ell}},...,i_{n}). From 𝐚\mathbf{a} we can define a functor and an element of WW:

Φ(𝐚)\displaystyle\Phi(\mathbf{a}) =Θi1Θi1𝖥iΘi1Θin𝟙λ,\displaystyle=\Theta_{i_{1}}\cdots\Theta_{i_{\ell-1}}\mathsf{F}_{i_{\ell}}\Theta_{i_{\ell 1}}\cdots\Theta_{i_{n}}\mathbbm{1}_{\lambda},
w(𝐚)\displaystyle w(\mathbf{a}) =si1sin.\displaystyle=s_{i_{1}}\cdots s_{i_{n}}.

Note that unlike Φ(𝐚)\Phi(\mathbf{a}), w(𝐚)w(\mathbf{a}) forgets the location of the marked letter. We say that 𝐚\mathbf{a} is reduced, if the corresponding unmarked word is a reduced expression for w(𝐚)w(\mathbf{a}).

We will apply braid relations to marked words. Away from the marked letter these operate as usual, and at the marked letter we have:

(,k,¯,)\displaystyle(\ldots,k,\underline{\ell},\ldots) (,¯,k,), if (k,)=0 and,\displaystyle\leftrightarrow(\ldots,\underline{\ell},k,\ldots),\quad\text{ if }(k,\ell)=0\text{ and,} (5.4)
(,,k,¯,)\displaystyle(\ldots,\ell,k,\underline{\ell},\ldots) (,k¯,,k,), if (k,)=1.\displaystyle\leftrightarrow(\ldots,\underline{k},\ell,k,\ldots),\quad\text{ if }(k,\ell)=-1. (5.5)

For marked words 𝐚,𝐛\mathbf{a},\mathbf{b} we write 𝐚𝐛\mathbf{a}\sim\mathbf{b} if they are related by a sequence of braid relations.

Lemma 5.6.

Let 𝐚,𝐛\mathbf{a},\mathbf{b} be marked words which differ by a single braid relation. Then there exists k{0,±1}k\in\{0,\pm 1\} such that Φ(𝐚)Φ(𝐛)[k]k\Phi(\mathbf{a})\cong\Phi(\mathbf{b})[k]\langle-k\rangle.

Proof.

If the relation doesn’t involve the marked letter then Φ(𝐚)Φ(𝐛)\Phi(\mathbf{a})\cong\Phi(\mathbf{b}) since Rickard complexes satisfy the braid relations [CK3, Theorem 2.10]. Suppose then that the relation does involve the marked letter. If (j,)=0(j,\ell)=0 the result follows from the fact that Θj𝖥𝖥Θj\Theta_{j}\mathsf{F}_{\ell}\cong\mathsf{F}_{\ell}\Theta_{j}. Otherwise (j,)=1(j,\ell)=-1. Set μ=ssj(λ)αj\mu=s_{\ell}s_{j}(\lambda)-\alpha_{j}. Applying Lemma 5.3 (twice) we deduce that

𝟙μΘΘj𝖥𝟙λ=ΘΘj𝖥𝟙λ\displaystyle\mathbbm{1}_{\mu}\Theta_{\ell}\Theta_{j}\mathsf{F}_{\ell}\mathbbm{1}_{\lambda}=\Theta_{\ell}\Theta_{j}\mathsf{F}_{\ell}\mathbbm{1}_{\lambda} {Θ𝖥jΘj𝟙λ, if λj0,Θ𝖥jΘj𝟙λ[1]1, if λj<0.\displaystyle\cong\begin{cases}\Theta_{\ell}\mathsf{F}_{j\ell}\Theta_{j}\mathbbm{1}_{\lambda},&\text{ if }\lambda_{j}\geq 0,\\ \Theta_{\ell}\mathsf{F}_{j\ell}\Theta_{j}\mathbbm{1}_{\lambda}[1]\langle-1\rangle,&\text{ if }\lambda_{j}<0.\end{cases}
{𝖥jΘΘj𝟙λ[1]1 if λj0,μ>0,𝖥jΘΘj𝟙λ[1]1 if λj<0,μ0,𝖥jΘΘj𝟙λ otherwise.\displaystyle\cong\begin{cases}\mathsf{F}_{j}\Theta_{\ell}\Theta_{j}\mathbbm{1}_{\lambda}[-1]\langle 1\rangle&\text{ if }\lambda_{j}\geq 0,\mu_{\ell}>0,\\ \mathsf{F}_{j}\Theta_{\ell}\Theta_{j}\mathbbm{1}_{\lambda}[1]\langle-1\rangle&\text{ if }\lambda_{j}<0,\mu_{\ell}\leq 0,\\ \mathsf{F}_{j}\Theta_{\ell}\Theta_{j}\mathbbm{1}_{\lambda}&\text{ otherwise.}\end{cases}

Hence the result follows. ∎

Corollary 5.7.

Let 𝐚,𝐛\mathbf{a},\mathbf{b} be marked words such that 𝐚𝐛\mathbf{a}\sim\mathbf{b}. Then there exists an integer kk such that Φ(𝐚)Φ(𝐛)[k]k\Phi(\mathbf{a})\cong\Phi(\mathbf{b})[k]\langle-k\rangle.

Lemma 5.8.

Let 𝐚=(i1,,in,¯)\mathbf{a}=(i_{1},...,i_{n},\underline{\ell}) and 𝐛=(¯,i1,,in)\mathbf{b}=(\underline{\ell^{\prime}},i_{1},...,i_{n}) be reduced marked words such that w(𝐚)=w(𝐛)w(\mathbf{a})=w(\mathbf{b}). Then 𝐚𝐛\mathbf{a}\sim\mathbf{b}.

Proof.

We prove the claim by induction on nn. Since w(𝐚)=w(𝐛)w(\mathbf{a})=w(\mathbf{b}) there is a Matsumoto sequence of length MM relating the unmarked words 𝔞=(i1,,in,)\mathfrak{a}=(i_{1},...,i_{n},\ell) and 𝔟=(,i1,,in)\mathfrak{b}=(\ell^{\prime},i_{1},...,i_{n}). Let 𝔞0=𝔞𝔞1𝔞M=𝔟\mathfrak{a}_{0}=\mathfrak{a}\to\mathfrak{a}_{1}\to\cdots\to\mathfrak{a}_{M}=\mathfrak{b} denote the resulting sequence of unmarked words, starting at 𝔞\mathfrak{a} and ending 𝔟\mathfrak{b}. Let 𝔞r+1\mathfrak{a}_{r+1} be the first word in this sequence whose last entry is not \ell. In other words, the first rr steps in the sequence do not involve the last entry of 𝔞\mathfrak{a}, but the (r+1)st(r+1)^{st} step does.

We will now consider the same sequence of steps, but applied to the marked words. We then obtain a sequence of marked words 𝐚0=𝐚𝐚1\mathbf{a}_{0}=\mathbf{a}\to\mathbf{a}_{1}\to\cdots\;. Further, we know that 𝐚r=(j1,,jn,¯)\mathbf{a}_{r}=(j_{1},\dots,j_{n},\underline{\ell}) for some j1,,jnj_{1},\ldots,j_{n}, and the next step involves the marked letter.

If the (r+1)st(r+1)^{st} step is an application of (5.5) then jn1=j_{n-1}=\ell and the step is:

𝐚r=(j1,,jn2,,jn,¯)𝐚r+1=(j1,,jn2,jn¯,,jn).\displaystyle\mathbf{a}_{r}=(j_{1},\ldots,j_{n-2},\ell,j_{n},\underline{\ell})\to\mathbf{a}_{r+1}=(j_{1},\ldots,j_{n-2},\underline{j_{n}},\ell,j_{n}).

Let 𝐛:=(¯,j1,,jn2,,jn)\mathbf{b}^{\prime}:=(\underline{\ell^{\prime}},j_{1},...,j_{n-2},\ell,j_{n}), i.e. 𝐛\mathbf{b}^{\prime} is obtained from 𝐛\mathbf{b} by applying the first rr steps to its (last nn) entries as were used for the (first nn) entries of 𝐚\mathbf{a}. Note that 𝐚r+1\mathbf{a}_{r+1} and 𝐛\mathbf{b}^{\prime} are reduced marked words such that w(𝐚r+1)=w(𝐛)w(\mathbf{a}_{r+1})=w(\mathbf{b}^{\prime}), and both end in (,jn)(\ell,j_{n}). Hence, if we delete these last two entries we obtain two reduced marked words 𝐚′′=(j1,,jn2,jn¯)\mathbf{a}^{\prime\prime}=(j_{1},...,j_{n-2},\underline{j_{n}}) and 𝐛′′=(¯,j1,,jn2)\mathbf{b}^{\prime\prime}=(\underline{\ell^{\prime}},j_{1},...,j_{n-2}) such that w(𝐚′′)=w(𝐛′′)w(\mathbf{a}^{\prime\prime})=w(\mathbf{b}^{\prime\prime}). By induction 𝐚′′𝐛′′\mathbf{a}^{\prime\prime}\sim\mathbf{b}^{\prime\prime} and therefore 𝐚r+1𝐛\mathbf{a}_{r+1}\sim\mathbf{b}^{\prime}. Since clearly 𝐛𝐛\mathbf{b}^{\prime}\sim\mathbf{b}, we have our desired result:

𝐚𝐚r𝐚r+1𝐛𝐛.\mathbf{a}\sim\mathbf{a}_{r}\sim\mathbf{a}_{r+1}\sim\mathbf{b}^{\prime}\sim\mathbf{b}.

If the (r+1)st(r+1)^{st} step is an application of (5.4) then the step is:

𝐚r=(j1,,jn1,jn,¯)𝐚r+1=(j1,,jn1,¯,jn).\displaystyle\mathbf{a}_{r}=(j_{1},\ldots,j_{n-1},j_{n},\underline{\ell})\to\mathbf{a}_{r+1}=(j_{1},\ldots,j_{n-1},\underline{\ell},j_{n}).

Note that 𝐚r+1\mathbf{a}_{r+1} and 𝐛:=(¯,j1,,jn1,jn)\mathbf{b}^{\prime}:=(\underline{\ell^{\prime}},j_{1},...,j_{n-1},j_{n}) are reduced marked words such that w(𝐚r+1)=w(𝐛)w(\mathbf{a}_{r+1})=w(\mathbf{b}^{\prime}), and both end in jnj_{n}. Hence we can delete this last entry and apply a similar analysis as above.

5.3. The relation between Θw0\Theta_{w_{0}} and Chevalley functors

We now have the machinery in place to prove our main relation.

Proposition 5.9.

For any iI,λXi\in I,\lambda\in X we have the following relations:

Θw0𝖤i𝟙λ\displaystyle\Theta_{w_{0}}\mathsf{E}_{i}\mathbbm{1}_{\lambda} 𝖥τ(i)Θw0𝟙λ[1]λi,\displaystyle\cong\mathsf{F}_{\tau(i)}\Theta_{w_{0}}\mathbbm{1}_{\lambda}[1]\langle\lambda_{i}\rangle,
Θw0𝖥i𝟙λ\displaystyle\Theta_{w_{0}}\mathsf{F}_{i}\mathbbm{1}_{\lambda} 𝖤τ(i)Θw0𝟙λ[1]λi.\displaystyle\cong\mathsf{E}_{\tau(i)}\Theta_{w_{0}}\mathbbm{1}_{\lambda}[1]\langle-\lambda_{i}\rangle.
Proof.

We’ll prove the first relation, the second being entirely analogous. For two functors 𝖥,𝖦\mathsf{F},\mathsf{G} we write 𝖥𝖦\mathsf{F}\equiv\mathsf{G} if there exist integers ,k\ell,k such that 𝖥𝖦[]k\mathsf{F}\cong\mathsf{G}[\ell]\langle k\rangle.

We first show that Θw0𝖤i𝖥τ(i)Θw0\Theta_{w_{0}}\mathsf{E}_{i}\equiv\mathsf{F}_{\tau(i)}\Theta_{w_{0}} by induction on the rank of 𝔤\mathfrak{g}. The base case, when 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2}, follows from Lemma 5.1. For the inductive step let JIJ\subset I be a strict subdiagram containing ii. Recall the bijection τJ:II\tau_{J}:I\to I induced by the longest element w0JWJw_{0}^{J}\in W_{J}. Let u=w0(w0J)1u=w_{0}(w_{0}^{J})^{-1} and let u=si1sinu=s_{i_{1}}\cdots s_{i_{n}} be a reduced expression. Define two marked words:

𝐚\displaystyle\mathbf{a} =(i1,,in,τJ(i)¯),\displaystyle=(i_{1},\ldots,i_{n},\underline{\tau_{J}(i)}),
𝐛\displaystyle\mathbf{b} =(τ(i)¯,i1,,in).\displaystyle=(\underline{\tau(i)},i_{1},\ldots,i_{n}).

Note that w(𝐚)=w(𝐛)w(\mathbf{a})=w(\mathbf{b}). By the inductive hypothesis we have Θw0J𝖤i𝖥τJ(i)Θw0J\Theta_{w_{0}^{J}}\mathsf{E}_{i}\equiv\mathsf{F}_{\tau_{J}(i)}\Theta_{w_{0}^{J}}, and therefore

Θw0𝖤iΘuΘw0J𝖤iΘu𝖥τJ(i)Θw0JΦ(𝐚)Θw0J.\displaystyle\Theta_{w_{0}}\mathsf{E}_{i}\equiv\Theta_{u}\Theta_{w_{0}^{J}}\mathsf{E}_{i}\equiv\Theta_{u}\mathsf{F}_{\tau_{J}(i)}\Theta_{w_{0}^{J}}\equiv\Phi(\mathbf{a})\Theta_{w_{0}^{J}}.

Note that 𝐚,𝐛\mathbf{a},\mathbf{b} satisfy the hypothesis of Lemma 5.8, so by Lemmas 5.7 and 5.8 we have that Φ(𝐚)Φ(𝐛)\Phi(\mathbf{a})\equiv\Phi(\mathbf{b}), and hence Θw0𝖤iΦ(𝐛)Θw0J𝖥τ(i)Θw0.\Theta_{w_{0}}\mathsf{E}_{i}\equiv\Phi(\mathbf{b})\Theta_{w_{0}^{J}}\equiv\mathsf{F}_{\tau(i)}\Theta_{w_{0}}.

We now know there exist integers k,k,\ell such that Θw0𝖤i𝟙λ𝖥τ(i)Θw0𝟙λ[]k\Theta_{w_{0}}\mathsf{E}_{i}\mathbbm{1}_{\lambda}\cong\mathsf{F}_{\tau(i)}\Theta_{w_{0}}\mathbbm{1}_{\lambda}[\ell]\langle k\rangle, and it remains to show that =1,k=λi\ell=1,k=\lambda_{i}. Let u=w0siu=w_{0}s_{i} and let u=si1sinu=s_{i_{1}}\cdots s_{i_{n}} be a reduced expression. Define two marked words:

𝐚\displaystyle\mathbf{a} =(i1,,in,i¯),\displaystyle=(i_{1},\ldots,i_{n},\underline{i}),
𝐛\displaystyle\mathbf{b} =(τ(i)¯,i1,,in).\displaystyle=(\underline{\tau(i)},i_{1},\ldots,i_{n}).

Note that 𝐚,𝐛\mathbf{a},\mathbf{b} satisfy the hypothesis of Lemma 5.8, so by Lemmas 5.7 and 5.8 there exists an integer mm such that Φ(𝐚)Φ(𝐛)[m]m\Phi(\mathbf{a})\cong\Phi(\mathbf{b})[m]\langle-m\rangle. Hence we have that

Θw0𝖤i𝟙λ\displaystyle\Theta_{w_{0}}\mathsf{E}_{i}\mathbbm{1}_{\lambda} ΘuΘi𝖤i𝟙λ\displaystyle\cong\Theta_{u}\Theta_{i}\mathsf{E}_{i}\mathbbm{1}_{\lambda}
Θu𝖥iΘi𝟙λ[1]λi\displaystyle\cong\Theta_{u}\mathsf{F}_{i}\Theta_{i}\mathbbm{1}_{\lambda}[1]\langle\lambda_{i}\rangle
𝖥τ(i)ΘuΘi𝟙λ[m+1]m+λi\displaystyle\cong\mathsf{F}_{\tau(i)}\Theta_{u}\Theta_{i}\mathbbm{1}_{\lambda}[m+1]\langle-m+\lambda_{i}\rangle
𝖥τ(i)Θw0𝟙λ[m+1]m+λi\displaystyle\cong\mathsf{F}_{\tau(i)}\Theta_{w_{0}}\mathbbm{1}_{\lambda}[m+1]\langle-m+\lambda_{i}\rangle

showing that +k=1+λi\ell+k=1+\lambda_{i}.

On the other hand, we can deduce kk by inspecting the relation on the level of Grothendieck groups. Namely, by [KT, Lemma 5.4], we have that

𝗍w0Ei𝟣λ=qλiFτ(i)𝗍w0𝟣λ,\mathsf{t}_{w_{0}}E_{i}\mathsf{1}_{\lambda}=-q^{-\lambda_{i}}F_{\tau(i)}\mathsf{t}_{w_{0}}\mathsf{1}_{\lambda},

showing that k=λik=\lambda_{i}. ∎

6. On t-exactness and perversity of Θw0\Theta_{w_{0}}

In this section we will state and prove the central results of the paper. We fix throughout a categorical representation 𝒞\mathcal{C} of Uq(𝔤)U_{q}(\mathfrak{g}), and let w0=si1si2sinw_{0}=s_{i_{1}}s_{i_{2}}\cdots s_{i_{n}} be a reduced expression.

6.1. Θw0\Theta_{w_{0}} on isotypic categorifications

In this section we prove that Θw0\Theta_{w_{0}} is t-exact on any isotypic categorification. Fix λX+\lambda\in X_{+}. We write Θ=Θw0,L=L(λ)\Theta=\Theta_{w_{0}},L=L(\lambda) and =(λ)\mathcal{L}=\mathcal{L}(\lambda).

Lemma 6.1.

Let k{2,,n}k\in\{2,\ldots,n\} and set μ=siksin(w0(λ))\mu=s_{i_{k}}\cdots s_{i_{n}}(w_{0}(\lambda)). The weight space L(λ)μαik1L(\lambda)_{\mu-\alpha_{i_{k-1}}} is zero.

Proof.

By [Humph, Proposition 21.3], it suffices to find uWu\in W such that u(μαik1)w0(λ)u(\mu-\alpha_{i_{k-1}})\nsucc w_{0}(\lambda). Take u=sik1u=s_{i_{k-1}} and, noting that (μ,αik1)<0(\mu,\alpha_{i_{k-1}})<0, the result follows. ∎

Recall that vλv_{\lambda} and vλlowv_{\lambda}^{low} are the highest and lowest weight elements of the canonical basis 𝐁(λ)\mathbf{B}(\lambda).

Lemma 6.2.

[KT, Comment 5.10] We have 𝗍w0(vλlow)=vλ\mathsf{t}_{w_{0}}(v_{\lambda}^{low})=v_{\lambda}.

Proposition 6.3.

The equivalence Θ𝟙w0(λ):Db(w0(λ))Db(λ)\Theta\mathbbm{1}_{w_{0}(\lambda)}:D^{b}(\mathcal{L}_{w_{0}(\lambda)})\to D^{b}(\mathcal{L}_{\lambda}) satisfies 𝕜low𝕜high\mathbbm{k}_{low}\mapsto\mathbbm{k}_{high}, where both are considered as complexes concentrated in degree zero. In particular, under the equivalences λw0(λ)𝕜mod\mathcal{L}_{\lambda}\cong\mathcal{L}_{w_{0}(\lambda)}\cong\mathbbm{k}\mathrm{-mod}_{\mathbb{Z}}, Θ𝟙w0(λ)\Theta\mathbbm{1}_{w_{0}(\lambda)} is isomorphic to the identity autofunctor of Db(𝕜mod)D^{b}(\mathbbm{k}\mathrm{-mod}_{\mathbb{Z}}).

Proof.

Consider first the case 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2}. On the minimal categorification of highest weight mm, we have that Θ𝟙m(𝕜low)=𝕜high\Theta\mathbbm{1}_{-m}(\mathbbm{k}_{low})=\mathbbm{k}_{high}\langle\ell\rangle for some \ell by [CR, Theorem 6.6]. Since [Θ𝟙m]=𝗍1𝟣m[\Theta\mathbbm{1}_{-m}]=\mathsf{t}_{1}\mathsf{1}_{-m}, by Lemma 6.2 we conclude that =0\ell=0, and hence Θ𝟙m(𝕜low)=𝕜high\Theta\mathbbm{1}_{-m}(\mathbbm{k}_{low})=\mathbbm{k}_{high}.

For general 𝔤\mathfrak{g}, suppose XνX\in\mathcal{L}_{\nu} is simple and 𝖥iX=0\mathsf{F}_{i}X=0 for some iIi\in I. Consider \mathcal{L} as a categorical representation of 𝔰𝔩2\mathfrak{sl}_{2} by restriction to the ii-th root subalgebra. Then for some mm we have a morphism of categorical 𝔰𝔩2\mathfrak{sl}_{2} representations RX:(m),R_{X}:\mathcal{L}(m)\to\mathcal{L}, such that RX(𝕜low)=XR_{X}(\mathbbm{k}_{low})=X [CR, Theorem 5.24].

The functor RXR_{X} is equivariant for the categorical 𝔰𝔩2\mathfrak{sl}_{2} action on 𝒞\mathcal{C} determined by 𝖤i,𝖥i\mathsf{E}_{i},\mathsf{F}_{i} (in fact it is strongly equivariant in the sense of [LoWe, Definition 3.1]), and hence commutes with Θi𝟙ν\Theta_{i}\mathbbm{1}_{\nu}. Therefore we have that

Θi𝟙ν(X)\displaystyle\Theta_{i}\mathbbm{1}_{\nu}(X) Θi𝟙ν(RX(𝕜low))\displaystyle\cong\Theta_{i}\mathbbm{1}_{\nu}(R_{X}(\mathbbm{k}_{low}))
RX(Θi𝟙ν(𝕜low))\displaystyle\cong R_{X}(\Theta_{i}\mathbbm{1}_{\nu}(\mathbbm{k}_{low}))
RX(𝕜high),\displaystyle\cong R_{X}(\mathbbm{k}_{high})\in\mathcal{L},

and so Θi𝟙ν(X)\Theta_{i}\mathbbm{1}_{\nu}(X) is in homological degree zero. It follows that in the case when XX is not necessarily simple (but still assume that 𝖥iX=0\mathsf{F}_{i}X=0), Θi𝟙ν(X)\Theta_{i}\mathbbm{1}_{\nu}(X) is still in homological degree zero. Indeed, by induction on the length of a Jordan-Hölder filtration of XX one deduces this since Db()\mathcal{L}\subset D^{b}(\mathcal{L}) is extension closed.

Now we study Θ𝟙w0(λ)\Theta\mathbbm{1}_{w_{0}(\lambda)} applied to 𝕜low\mathbbm{k}_{low}. For k=2,,nk=2,\ldots,n, by Lemma 6.1,

𝖥ik1(ΘikΘin𝟙w0(λ)(𝕜low))=0.\mathsf{F}_{i_{k-1}}(\Theta_{i_{k}}\cdots\Theta_{i_{n}}\mathbbm{1}_{w_{0}(\lambda)}(\mathbbm{k}_{low}))=0.

By the previous paragraph, it follows that Θik1Θin𝟙w0(λ)(𝕜low)\Theta_{i_{k-1}}\cdots\Theta_{i_{n}}\mathbbm{1}_{w_{0}(\lambda)}(\mathbbm{k}_{low}) is in homological degree zero, and in particular, Θ𝟙w0(λ)(𝕜low)\Theta\mathbbm{1}_{w_{0}(\lambda)}(\mathbbm{k}_{low}) is supported in homological degree zero. Since in addition

[Θ𝟙w0(λ)]=𝗍w0𝟣w0(λ),[\Theta\mathbbm{1}_{w_{0}(\lambda)}]=\mathsf{t}_{w_{0}}\mathsf{1}_{w_{0}(\lambda)},

by Lemma 6.2 we conclude that Θ𝟙w0(λ)(𝕜low)𝕜high\Theta\mathbbm{1}_{w_{0}(\lambda)}(\mathbbm{k}_{low})\cong\mathbbm{k}_{high}. ∎

Theorem 6.4.

Let λX+\lambda\in X_{+} and set =(λ)\mathcal{L}=\mathcal{L}(\lambda). For any μX\mu\in X,

Θ𝟙μ[n]:Db(μ)Db(w0(μ))\Theta\mathbbm{1}_{\mu}[-n]:D^{b}(\mathcal{L}_{\mu})\to D^{b}(\mathcal{L}_{w_{0}(\mu)})

is t-exact, where n=ht(μw0(λ))n=ht(\mu-w_{0}(\lambda)).

Proof.

Consider P=𝖤i1𝖤i(𝕜low)μP=\mathsf{E}_{i_{1}}\cdots\mathsf{E}_{i_{\ell}}(\mathbbm{k}_{low})\in\mathcal{L}_{\mu}. We will first prove by induction on nn that there exists an integer kk such that

Θ(P)[n]𝖥j1𝖥j(𝕜high)kw0(μ),\displaystyle\Theta(P)[-n]\cong\mathsf{F}_{j_{1}}\cdots\mathsf{F}_{j_{\ell}}(\mathbbm{k}_{high})\langle k\rangle\in\mathcal{L}_{w_{0}(\mu)}, (6.5)

where jr=τ(ir)j_{r}=\tau(i_{r}).

The base case when n=0n=0 follows by Proposition 6.3. For the inductive step write P=𝖤i1(Q)P=\mathsf{E}_{i_{1}}(Q). Note that Qμαi1Q\in\mathcal{L}_{\mu-\alpha_{i_{1}}}. By Proposition 5.9 we have that

Θ𝟙μ(P)[n]\displaystyle\Theta\mathbbm{1}_{\mu}(P)[-n] =Θ𝖤i1𝟙μαi1(Q)[n]\displaystyle=\Theta\mathsf{E}_{i_{1}}\mathbbm{1}_{\mu-\alpha_{i_{1}}}(Q)[-n]
𝖥τ(i1)Θ𝟙μαi1(Q)[n+1](μαi1,αi1).\displaystyle\cong\mathsf{F}_{\tau(i_{1})}\Theta\mathbbm{1}_{\mu-\alpha_{i_{1}}}(Q)[-n+1]\langle(\mu-\alpha_{i_{1}},\alpha_{i_{1}})\rangle.

By hypothesis

Θ𝟙μαi1(Q)[n+1]𝖥j2𝖥j(𝕜high)kw0(μαi1)\Theta\mathbbm{1}_{\mu-\alpha_{i_{1}}}(Q)[-n+1]\cong\mathsf{F}_{j_{2}}\cdots\mathsf{F}_{j_{\ell}}(\mathbbm{k}_{high})\langle k\rangle\in\mathcal{L}_{w_{0}(\mu-\alpha_{i_{1}})}

for some kk, and hence Equation (6.5) follows.

Since up to grading shift, any projective indecomposable object in μ\mathcal{L}_{\mu}, respectively w0(μ)\mathcal{L}_{w_{0}(\mu)}, is a summand of an object of the form 𝖤i1𝖤i(𝕜low)\mathsf{E}_{i_{1}}\cdots\mathsf{E}_{i_{\ell}}(\mathbbm{k}_{low}) (respectively 𝖥j1𝖥j(𝕜high)\mathsf{F}_{j_{1}}\cdots\mathsf{F}_{j_{\ell}}(\mathbbm{k}_{high})), it follows that Θ𝟙μ[n]\Theta\mathbbm{1}_{\mu}[-n] takes projective objects in μ\mathcal{L}_{\mu} to projective objects in w0(μ)\mathcal{L}_{w_{0}(\mu)}. Since Θ𝟙μ[n]\Theta\mathbbm{1}_{\mu}[-n] is a derived equivalence it follows that it is t-exact.

Remark 6.6.

Theorem 6.4 is a generalisation of [CR, Theorem 6.6], which covers the 𝔰𝔩2\mathfrak{sl}_{2} case. Note that [CR, Theorem 6.6] is crucial in the work of Chuang and Rouquier, since it’s one of the main technical results needed to prove that Rickard complexes are invertible. Our proof in the general case follows a completely different approach, but it does not give a new proof in the case of 𝔰𝔩2\mathfrak{sl}_{2}. Indeed we use [CR, Theorem 6.6] explicitly in the proof of Proposition 6.3, and more generally we use the fact the Θi\Theta_{i} is invertible throughout.

Corollary 6.7.

Suppose 𝒞\mathcal{C} is an isotypic categorification of type λ\lambda, for some λX+\lambda\in X_{+}, and let μX\mu\in X. Then Θ𝟙μ[n]:Db(𝒞μ)Db(𝒞w0(μ))\Theta\mathbbm{1}_{\mu}[-n]:D^{b}(\mathcal{C}_{\mu})\to D^{b}(\mathcal{C}_{w_{0}(\mu)}) is a t-exact equivalence, where n=ht(μw0(λ))n=ht(\mu-w_{0}(\lambda)).

Proof.

By Lemma 3.5, there exists an abelian 𝕜\mathbbm{k}-linear category 𝒜\mathcal{A} such that 𝒞(λ)𝕜𝒜\mathcal{C}\cong\mathcal{L}(\lambda)\otimes_{\mathbbm{k}}\mathcal{A} as categorical representations. We have that

Θ𝟙μ[n]((λ)μ𝕜𝒜)Θ𝟙μ[n]((λ)μ)𝕜𝒜(λ)w0(μ)𝕜𝒜,\Theta\mathbbm{1}_{\mu}[-n](\mathcal{L}(\lambda)_{\mu}\otimes_{\mathbbm{k}}\mathcal{A})\cong\Theta\mathbbm{1}_{\mu}[-n](\mathcal{L}(\lambda)_{\mu})\otimes_{\mathbbm{k}}\mathcal{A}\cong\mathcal{L}(\lambda)_{w_{0}(\mu)}\otimes_{\mathbbm{k}}\mathcal{A},

proving that Θ𝟙μ[n]:Db(𝒞μ)Db(𝒞w0(μ))\Theta\mathbbm{1}_{\mu}[-n]:D^{b}(\mathcal{C}_{\mu})\to D^{b}(\mathcal{C}_{w_{0}(\mu)}) is a t-exact equivalence. ∎

6.2. Θw0\Theta_{w_{0}} on general categorical representations

In this section we prove that Θw0\Theta_{w_{0}} is a perverse equivalence on an arbitrary categorical representation. Fix μX\mu\in X such that 𝒞μ\mathcal{C}_{\mu} is nonzero. For ease of notation, set 𝒜=𝒞μ\mathcal{A}=\mathcal{C}_{\mu} and 𝒜=𝒞w0(μ)\mathcal{A}^{\prime}=\mathcal{C}_{w_{0}(\mu)}.

Consider a filtration by Serre subcategories

0=𝒞0𝒞1𝒞r=𝒞,\displaystyle 0=\mathcal{C}_{0}\subset\mathcal{C}_{1}\subset\cdots\subset\mathcal{C}_{r}=\mathcal{C},

which can be either the Jordan-Hölder filtration (Theorem 3.6) or the isotypic filtration (Remark 3.9). So for every ii, 𝒞i\mathcal{C}_{i} is a subrepresentation of 𝒞\mathcal{C}, and 𝒞i/𝒞i1\mathcal{C}_{i}/\mathcal{C}_{i-1} is either a simple categorification or an isotypic one. Define λiX+\lambda_{i}\in X_{+} by requiring that [𝒞i/𝒞i1](q)[\mathcal{C}_{i}/\mathcal{C}_{i-1}]_{\mathbb{C}(q)} is a representation of type λi\lambda_{i}.

Construct filtrations of 𝒜\mathcal{A} and 𝒜\mathcal{A}^{\prime} by 𝒜i=𝒞i𝒜,𝒜i=𝒞i𝒜\mathcal{A}_{i}=\mathcal{C}_{i}\cap\mathcal{A},\mathcal{A}_{i}^{\prime}=\mathcal{C}_{i}\cap\mathcal{A}^{\prime}. These are Serre subcategories of 𝒜\mathcal{A} and 𝒜\mathcal{A}^{\prime} respectively. Let p:{0,,r}p:\{0,...,r\}\to\mathbb{Z} be given by p(i)=ht(μw0(λi))p(i)=ht(\mu-w_{0}(\lambda_{i})).

Theorem 6.8.

Θw0𝟙μ:Db(𝒜)Db(𝒜)\Theta_{w_{0}}\mathbbm{1}_{\mu}:D^{b}(\mathcal{A})\to D^{b}(\mathcal{A}^{\prime}) is a perverse equivalence with respect to (𝒜,𝒜,p)(\mathcal{A}_{\bullet},\mathcal{A}_{\bullet}^{\prime},p) for either the Jordan-Hölder or the isotypic filtration.

Proof.

Since 𝒞i𝒞\mathcal{C}_{i}\subset\mathcal{C} is a categorical subrepresentation, the terms of the functor Θw0𝟙μ\Theta_{w_{0}}\mathbbm{1}_{\mu} leave 𝒞i\mathcal{C}_{i} invariant, and in particular take objects in 𝒜i\mathcal{A}_{i} to 𝒜i\mathcal{A}_{i}^{\prime}. By Lemma 4.8 this implies that Θw0𝟙μ(D𝒜ib(𝒜))D𝒜ib(𝒜)\Theta_{w_{0}}\mathbbm{1}_{\mu}(D^{b}_{\mathcal{A}_{i}}(\mathcal{A}))\subseteq D^{b}_{\mathcal{A}_{i}^{\prime}}(\mathcal{A}^{\prime}).

Now, 𝒞i/𝒞i1\mathcal{C}_{i}/\mathcal{C}_{i-1} is a simple or isotypic categorification (of type λi\lambda_{i}). By Corollary 6.7, Θw0𝟙μ[p(i)]\Theta_{w_{0}}\mathbbm{1}_{\mu}[-p(i)] restricts to an abelian equivalence 𝒜i/𝒜i1𝒜i/𝒜i1\mathcal{A}_{i}/\mathcal{A}_{i-1}\to\mathcal{A}_{i}^{\prime}/\mathcal{A}_{i-1}^{\prime}, i.e. the functor

Θw0𝟙μ[p(i)]:D𝒜ib(𝒜)/D𝒜i1b(𝒜)D𝒜ib(𝒜)/D𝒜i1b(𝒜)\Theta_{w_{0}}\mathbbm{1}_{\mu}[-p(i)]:D^{b}_{\mathcal{A}_{i}}(\mathcal{A})/D^{b}_{\mathcal{A}_{i-1}}(\mathcal{A})\to D^{b}_{\mathcal{A}^{\prime}_{i}}(\mathcal{A}^{\prime})/D^{b}_{\mathcal{A}^{\prime}_{i-1}}(\mathcal{A}^{\prime})

is a t-exact equivalence. This shows that Θw0𝟙μ\Theta_{w_{0}}\mathbbm{1}_{\mu} is a perverse equivalence with respect to (𝒜,𝒜,p)(\mathcal{A}_{\bullet},\mathcal{A}^{\prime}_{\bullet},p). ∎

Remark 6.9.

The 𝔰𝔩2\mathfrak{sl}_{2} case of Theorem 6.8 appears as [CRperv, Proposition 8.4], by a different argument relying on a technical lemma [CRperv, Lemma 4.12].

7. Crystalising the braid group action

Already in the work of Chuang and Rouquier, a close connection is established between categorical representation theory and the theory of crystals (although it is not phrased in this language, cf. Proposition 3.4 below). In this section we describe a new component of this theory. More precisely, let VV be an integrable representation of Uq(𝔤)U_{q}(\mathfrak{g}). Recall that Lusztig has defined a braid group action on VV [Lusbook]. In this section we explain how to use our results to “crystalise” this braid group action to obtain a cactus group action on the crystal of VV, recovering the recently discovered action by generalised Schützenberger involutions.

7.1. Cactus groups

The cactus group associated to the Dynkin diagram II has several incarnations. Geometrically, it appears as the fundamental group of a space associated to the Cartan subalgebra 𝔥\mathfrak{h} of 𝔤\mathfrak{g}. Namely, let 𝔥reg𝔤\mathfrak{h}^{\text{reg}}\subseteq\mathfrak{g} denote the regular elements of 𝔥\mathfrak{h}. The cactus group C=CIC=C_{I} is the WW-equivariant fundamental group of the real locus of the de Concini-Procesi wonderful compactification of 𝔥reg\mathfrak{h}^{\text{reg}} (see [DJS03], [HKRW, Section 2] for further details):

C=π1W((𝔥reg)¯()).C=\pi^{W}_{1}(\overline{\mathbb{P}(\mathfrak{h}^{\text{reg}})}(\mathbb{R})).

There is a surjective map CWC\to W, and the kernel of this map is called the pure cactus group. In type A it is the fundamental group of the Deligne-Mumford compactification of the moduli space of real genus 0 curves with n+1n+1 marked points [HK06].

The cactus group has a presentation using Dynkin diagram combinatorics. For any subdiagram JIJ\subseteq I, recall that τJ:JJ\tau_{J}:J\to J is the diagram automorphism induced by the longest element w0JWJw_{0}^{J}\in W_{J}.

Definition 7.1.

The cactus group C=CIC=C_{I} is generated by cJc_{J}, where JIJ\subseteq I is a connected subdiagram, subject to the following relations:

  • (i)

    cJ2=1c_{J}^{2}=1 for all JIJ\subseteq I,

  • (ii)

    cJcK=cKcJc_{J}c_{K}=c_{K}c_{J}, if JK=J\cap K=\varnothing and there are no edges connecting any jJj\in J to any kKk\in K, and

  • (iii)

    cJcK=cKcτK(J)c_{J}c_{K}=c_{K}c_{\tau_{K}(J)} if JKJ\subseteq K.

The surjective map CWC\to W mentioned above is given by cJw0Jc_{J}\mapsto w_{0}^{J}. We are interested in the cactus group in connection to the theory of crystals.

A 𝔤\mathfrak{g}-crystal 𝐁\mathbf{B} is called normal if it is isomorphic to a disjoint union λ𝐁(λ)\sqcup_{\lambda}\mathbf{B}(\lambda) for some collection of highest weights λ\lambda. The category of normal 𝔤\mathfrak{g}-crystals has the structure of a coboundary category analogous to the braided tensor category structure on Uq(𝔤)U_{q}(\mathfrak{g})-representations. It is realized through an “external” cactus group action of CAn1C_{A_{n-1}} on nn-tensor products of 𝔤\mathfrak{g}-crystals, described by Henriques and Kamnitzer [HK06, Theorems 6,7].

We are interested in the “internal” cactus group action of CC on any 𝔤\mathfrak{g}-crystal 𝐁\mathbf{B}. Both the internal and external actions rely on the following combinatorially defined maps, which are generalisations of the partial Schützenberger involutions in type AA.

Definition 7.2.

The generalised Schützenberger involution ξλ\xi_{\lambda} on 𝐁(λ)\mathbf{B}(\lambda) is the set map defined uniquely by the following properties. For all b𝐁(λ)b\in\mathbf{B}(\lambda) and iIi\in I:

  1. (1)

    wt(ξλ(b))=w0wt(b)\text{wt}(\xi_{\lambda}(b))=w_{0}\text{wt}(b),

  2. (2)

    ξλe~i(b)=f~τ(i)ξλ(b)\xi_{\lambda}\widetilde{e}_{i}(b)=\widetilde{f}_{\tau(i)}\xi_{\lambda}(b),

  3. (3)

    ξλf~i(b)=e~τ(i)ξλ(b)\xi_{\lambda}\widetilde{f}_{i}(b)=\widetilde{e}_{\tau(i)}\xi_{\lambda}(b).

Note that from (1), ξλ\xi_{\lambda} swaps the lowest and highest weight element of 𝐁(λ)\mathbf{B}(\lambda), and the rest of its behavior is then determined from (2) and (3). The generalised Schützenberger involution ξ\xi on 𝐁=λ𝐁(λ)\mathbf{B}=\sqcup_{\lambda}\mathbf{B}(\lambda) is the set map which acts as ξλ\xi_{\lambda} on each irreducible component 𝐁(λ)\mathbf{B}(\lambda).

Note that (1) implies that ξλ\xi_{\lambda} maps the lowest weight element to the highest weight element (and vice-versa), and then (2) and (3) ensure that it is uniquely defined.

For JIJ\subseteq I, denote by 𝐁J\mathbf{B}_{J} the crystal 𝐁\mathbf{B} restricted to the subdiagram JJ. We denote the corresponding Schützenberger involution by ξJ\xi_{J}.

Theorem 7.3.

([HKRW, Theorem 5.19]) For any 𝔤\mathfrak{g}-crystal 𝐁\mathbf{B}, the assignment cJξJc_{J}\mapsto\xi_{J} defines a (set-theoretic) action of CC on 𝐁\mathbf{B}.

7.2. The cactus group action arising from Rickard complexes

We now explain how cactus group actions arise from categorical representations, analogous to the construction of the crystal on 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}) in Proposition 3.4.

Let 𝒞\mathcal{C} be a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}). For any weight μX\mu\in X, by Theorem 6.8 Θw0𝟙μ:Db(𝒞μ)Db(𝒞w0(μ))\Theta_{w_{0}}\mathbbm{1}_{\mu}:D^{b}(\mathcal{C}_{\mu})\to D^{b}(\mathcal{C}_{w_{0}(\mu)}) is a perverse equivalence, and hence it induces a bijection φI1μ:𝖨𝗋𝗋(𝒞μ)𝖨𝗋𝗋(𝒞w0(μ))\varphi_{I}1_{\mu}:\mathsf{Irr}(\mathcal{C}_{\mu})\to\mathsf{Irr}(\mathcal{C}_{w_{0}(\mu)}). By varying μ\mu we obtain a bijection φI:𝖨𝗋𝗋(𝒞)𝖨𝗋𝗋(𝒞).\varphi_{I}:\mathsf{Irr}(\mathcal{C})\to\mathsf{Irr}(\mathcal{C}).

Now let JIJ\subseteq I be a connected subdiagram, and let 𝔤J𝔤\mathfrak{g}_{J}\subset\mathfrak{g} be the corresponding subalgebra. By restriction, 𝒞\mathcal{C} is also a categorical representation of Uq(𝔤J)U_{q}(\mathfrak{g}_{J}), and hence by the above discussion we also obtain a bijection φJ:𝖨𝗋𝗋(𝒞)𝖨𝗋𝗋(𝒞).\varphi_{J}:\mathsf{Irr}(\mathcal{C})\to\mathsf{Irr}(\mathcal{C}).

We will prove that this family of bijections defines an action of the cactus group 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}). First we need the following technical result. The important point here is just that there exists an integer nn such that 𝗍w02𝟣μ=±qn𝟣μ\mathsf{t}^{2}_{w_{0}}\mathsf{1}_{\mu}=\pm q^{n}\mathsf{1}_{\mu}.

Lemma 7.4.

Let λX+,μX\lambda\in X_{+},\mu\in X, and let μw0(λ)=r=1αir\mu-w_{0}(\lambda)=\sum_{r=1}^{\ell}\alpha_{i_{r}}, where irIi_{r}\in I. Set jr=τ(ir)j_{r}=\tau(i_{r}) and define n(λ,μ)n(\lambda,\mu)\in\mathbb{Z} by

n(λ,μ)=2(r=1λjr+11rsajrjs+(λ,ρ))n(\lambda,\mu)=2\left(\sum_{r=1}^{\ell}\lambda_{j_{r}}+1-\sum_{1\leq r\leq s\leq\ell}a_{j_{r}j_{s}}+(\lambda,\rho)\right)

Then on L(λ)μL(\lambda)_{\mu} we have

𝗍w02𝟣μ=(1)2λ,ρqn(λ,μ)𝟣μ.\displaystyle\mathsf{t}^{2}_{w_{0}}\mathsf{1}_{\mu}=(-1)^{\langle 2\lambda,\rho^{\vee}\rangle}q^{n(\lambda,\mu)}\mathsf{1}_{\mu}. (7.5)
Proof.

We will prove the claim by induction on ht(μw0(λ))ht(\mu-w_{0}(\lambda)). For μ=w0(λ)\mu=w_{0}(\lambda), we have that =0\ell=0 so n(λ,μ)=(λ,ρ)n(\lambda,\mu)=(\lambda,\rho). By [KT, Equation (7)] 𝗍w0(vλ)=(1)λ,ρq(λ,ρ)vλlow\mathsf{t}_{w_{0}}(v_{\lambda})=(-1)^{\langle\lambda,\rho^{\vee}\rangle}q^{(\lambda,\rho)}v_{\lambda}^{low}, which, combined with Lemma 6.2, implies that 𝗍w02(vλlow)=(1)λ,ρq(λ,ρ)vλlow\mathsf{t}_{w_{0}}^{2}(v_{\lambda}^{low})=(-1)^{\langle\lambda,\rho^{\vee}\rangle}q^{(\lambda,\rho)}v_{\lambda}^{low}. Since dim(L(λ)w0(λ))=1\dim(L(\lambda)_{w_{0}(\lambda)})=1, this proves the base case.

Now choose any μ\mu and suppose (7.5) holds for any weight μ\mu^{\prime} such that ht(μw0(λ))<ht(μw0(λ))ht(\mu^{\prime}-w_{0}(\lambda))<ht(\mu-w_{0}(\lambda)). Consider v=Ej1EjvλlowL(λ)μv=E_{j_{1}}\cdots E_{j_{\ell}}v_{\lambda}^{low}\in L(\lambda)_{\mu}. First note that by [KT, Lemma 5.4] we have that

𝗍w02Ei=q2Ki2Ei𝗍w02.\displaystyle\mathsf{t}_{w_{0}}^{2}E_{i}=q^{2}K_{i}^{-2}E_{i}\mathsf{t}_{w_{0}}^{2}. (7.6)

Setting v=Ej2Ejvλlowv^{\prime}=E_{j_{2}}\cdots E_{j_{\ell}}v_{\lambda}^{low}, by induction we have

𝗍w02v\displaystyle\mathsf{t}_{w_{0}}^{2}v =q2Kj12Ej1𝗍w02v\displaystyle=q^{2}K_{j_{1}}^{-2}E_{j_{1}}\mathsf{t}_{w_{0}}^{2}v^{\prime}
=(q2Kj12Ej1)(1)λ,ρqn(λ,μαj1)v\displaystyle=(q^{2}K_{j_{1}}^{-2}E_{j_{1}})(-1)^{\langle\lambda,\rho^{\vee}\rangle}q^{n(\lambda,\mu-\alpha_{j_{1}})}v^{\prime}
=(1)λ,ρq2+n(λ,μαj1)2(μ,αj1)v\displaystyle=(-1)^{\langle\lambda,\rho^{\vee}\rangle}q^{2+n(\lambda,\mu-\alpha_{j_{1}})-2(\mu,\alpha_{j_{1}})}v

One checks easily that n(λ,μ)=2+n(λ,μαj1)2(μ,αj1)n(\lambda,\mu)=2+n(\lambda,\mu-\alpha_{j_{1}})-2(\mu,\alpha_{j_{1}}), proving that 𝗍w02v=(1)λ,ρqn(λ,μ)v\mathsf{t}_{w_{0}}^{2}v=(-1)^{\langle\lambda,\rho^{\vee}\rangle}q^{n(\lambda,\mu)}v. Since this holds for any vector of the form Ej1EjvλlowE_{j_{1}}\cdots E_{j_{\ell}}v_{\lambda}^{low} in L(λ)μL(\lambda)_{\mu}, this completes the inductive step. ∎

Theorem 7.7.

The assignment cJφJc_{J}\mapsto\varphi_{J} defines an action of CC on 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}).

Proof.

We need to show that the bijections φJ\varphi_{J} satisfy the cactus group relations.

Relation (i): Without loss of generality we may assume J=IJ=I. Fix a weight μ\mu. Our aim is to show that

φIφI1μ=Id𝖨𝗋𝗋(𝒞μ).\varphi_{I}\varphi_{I}1_{\mu}=\mathrm{Id}_{\mathsf{Irr}(\mathcal{C}_{\mu})}. (7.8)

Since the filtration of 𝒞w0(μ)\mathcal{C}_{w_{0}(\mu)} which we use in the perversity data of Θw0𝟙μ\Theta_{w_{0}}\mathbbm{1}_{\mu}, agrees with the filtration of 𝒞w0(μ)\mathcal{C}_{w_{0}(\mu)} which we use in the perversity data of Θw0𝟙w0(μ)\Theta_{w_{0}}\mathbbm{1}_{w_{0}(\mu)}, by Lemma 4.5 the composition Θw0Θw0𝟙μ\Theta_{w_{0}}\Theta_{w_{0}}\mathbbm{1}_{\mu} is a perverse autoequivalence of Db(𝒞μ)D^{b}(\mathcal{C}_{\mu}).

The functor [2λ,ρ]n(λ,μ)[\langle 2\lambda,\rho^{\vee}\rangle]\langle n(\lambda,\mu)\rangle is also a perverse autoequivalence of Db(𝒞μ)D^{b}(\mathcal{C}_{\mu}). By Lemma 7.4 these two perverse equivalences induce the same map on Grothendieck groups, and hence by Lemma 4.2 they also induce the same bijection. Since the bijection induced by [2λ,ρ]n(λ,μ)[\langle 2\lambda,\rho^{\vee}\rangle]\langle n(\lambda,\mu)\rangle is the identity, this proves relation (i).

Relation (ii): Let J,KIJ,K\subset I be disjoint subdiagrams with no connecting edges. Our aim is to show that

φJφK1μ=φKφJ1μ.\displaystyle\varphi_{J}\varphi_{K}1_{\mu}=\varphi_{K}\varphi_{J}1_{\mu}. (7.9)

We prove a slightly more general statement, namely that for any categorical representation 𝒞\mathcal{C} of Uq(𝔤J×𝔤K)U_{q}(\mathfrak{g}_{J}\times\mathfrak{g}_{K}), relation (7.9) holds.

Note that ΘJΘK𝟙μΘKΘJ𝟙μ\Theta_{J}\Theta_{K}\mathbbm{1}_{\mu}\cong\Theta_{K}\Theta_{J}\mathbbm{1}_{\mu} are isomorphic perverse equivalences, so they induce the same bijections. It remains to show that

φΘJΘK𝟙μ=φJφK1μ.\displaystyle\varphi_{\Theta_{J}\Theta_{K}\mathbbm{1}_{\mu}}=\varphi_{J}\circ\varphi_{K}1_{\mu}. (7.10)

Consider first the case when 𝒞\mathcal{C} categorifies an simple representation of Uq(𝔤J×𝔤K)U_{q}(\mathfrak{g}_{J}\times\mathfrak{g}_{K}). A minimal categorification of Uq(𝔤J×𝔤K)U_{q}(\mathfrak{g}_{J}\times\mathfrak{g}_{K}) is of the form (λ)𝕜(μ)\mathcal{L}(\lambda)\otimes_{\mathbbm{k}}\mathcal{L}(\mu), where λ\lambda is a highest weight for 𝔤J\mathfrak{g}_{J} and μ\mu is a highest weight for 𝔤K\mathfrak{g}_{K}. Hence by Lemma 3.5, a simple categorification of Uq(𝔤J×𝔤K)U_{q}(\mathfrak{g}_{J}\times\mathfrak{g}_{K}) is of the form (λ)𝕜(μ)𝕜𝒜\mathcal{L}(\lambda)\otimes_{\mathbbm{k}}\mathcal{L}(\mu)\otimes_{\mathbbm{k}}\mathcal{A} for some abelian category 𝒜\mathcal{A}.

This implies that as a categorical representation of Uq(𝔤J)U_{q}(\mathfrak{g}_{J}) (respectively Uq(𝔤K)U_{q}(\mathfrak{g}_{K})), 𝒞\mathcal{C} categorifies an isotypic representation. By Corollary 6.7 ΘJ𝟙w0K(μ)\Theta_{J}\mathbbm{1}_{w_{0}^{K}(\mu)} and ΘK𝟙μ\Theta_{K}\mathbbm{1}_{\mu} are t-exact up shift on isotypics categorifications. Hence Equation (7.10) follows by Lemma 4.5.

Now consider a Jordan-Hölder filtration (Theorem 3.6):

0=𝒞0𝒞n=𝒞,\displaystyle 0=\mathcal{C}_{0}\subset\cdots\subset\mathcal{C}_{n}=\mathcal{C},

where for every ii, 𝒞i\mathcal{C}_{i} is a subrepresentation of 𝒞\mathcal{C}, and 𝒞i/𝒞i1\mathcal{C}_{i}/\mathcal{C}_{i-1} is a simple categorification of Uq(𝔤J×𝔤K)U_{q}(\mathfrak{g}_{J}\times\mathfrak{g}_{K}). Equation (7.10) now follows by an easy induction on ii. Indeed the base case when i=1i=1 holds by the paragraph above, and the inductive step by Lemma 4.6.

Relation (iii): We need to show that φJφK1μ=φKφτK(J)1μ\varphi_{J}\varphi_{K}1_{\mu}=\varphi_{K}\varphi_{\tau_{K}(J)}1_{\mu}, where JKJ\subset K. Again, we may assume that K=IK=I. Note that we have an isomorphism at the level of functors:

Θw01Θw0JΘw0𝟙μΘw0τK(J)𝟙μ,\displaystyle\Theta_{w_{0}}^{-1}\Theta_{w_{0}^{J}}\Theta_{w_{0}}\mathbbm{1}_{\mu}\cong\Theta_{w_{0}^{\tau_{K}(J)}}\mathbbm{1}_{\mu},

which lifts the corresponding relation in BB. Since this is an isomorphism of perverse equivalences, they must induce the same bijections by Lemma 4.2.

It remains to show that

φΘw01Θw0JΘw0𝟙μ=φI1φJφI.\displaystyle\varphi_{\Theta_{w_{0}}^{-1}\Theta_{w_{0}^{J}}\Theta_{w_{0}}\mathbbm{1}_{\mu}}=\varphi_{I}^{-1}\circ\varphi_{J}\circ\varphi_{I}. (7.11)

When 𝒞\mathcal{C} is a simple categorification, by Corollary 6.7 Θw0𝟙μ\Theta_{w_{0}}\mathbbm{1}_{\mu} is t-exact (up to shift). Hence Equation (7.11) follows by Lemma 4.7. Now apply the same reasoning as in the proof of Relation (ii) to deduce equation (7.11) in the general case.

7.3. Reconciling the two cactus group actions

Let 𝒞\mathcal{C} be a categorical representation of Uq(𝔤)U_{q}(\mathfrak{g}), and consider the 𝔤\mathfrak{g}-crystal 𝐁=𝖨𝗋𝗋(𝒞)\mathbf{B}=\mathsf{Irr}(\mathcal{C}). There are two actions of the cactus group on 𝐁\mathbf{B}, the first arising combinatorially via Schützenberger involutions (Theorem 7.3) and the other categorically via Theorem 7.7.

Theorem 7.12.

The two actions of the cactus group on 𝐁\mathbf{B} agree.

Proof.

It suffices to show that φI=ξI\varphi_{I}=\xi_{I}. First, suppose 𝒞\mathcal{C} is a simple categorification of type λX+\lambda\in X_{+}. In this case 𝐁=𝐁(λ)\mathbf{B}=\mathbf{B}(\lambda), and ξ=ξI\xi=\xi_{I} is determined by:

ξ(vλ)\displaystyle\xi(v_{\lambda}) =vλlow, and\displaystyle=v_{\lambda}^{low},\text{ and }
ξ(e~i(v))\displaystyle\xi(\widetilde{e}_{i}(v)) =f~τ(i)(ξ(v)) for all v𝐁,\displaystyle=\widetilde{f}_{\tau(i)}(\xi(v))\text{ for all }v\in\mathbf{B},

so we need to show that φI\varphi_{I} satisfies these properties as well.

The first is an immediate consequence of Corollary 6.7. To show that φI\varphi_{I} satisfies the second property, fix μX\mu\in X and iIi\in I. We set n=ht(μw0(λ)),j=τ(i)n=ht(\mu-w_{0}(\lambda)),j=\tau(i), and write Θ=Θw0\Theta=\Theta_{w_{0}}.

Consider the following diagram:

Db(𝒞μ){D^{b}(\mathcal{C}_{\mu})}Db(𝒞w0(μ)){D^{b}(\mathcal{C}_{w_{0}(\mu)})}Db(𝒞μ+αi){D^{b}(\mathcal{C}_{\mu+\alpha_{i}})}Db(𝒞w0(μ)αj){D^{b}(\mathcal{C}_{w_{0}(\mu)-\alpha_{j}})}Θ𝟙μ[n]μi\scriptstyle{\Theta\mathbbm{1}_{\mu}[-n]\langle\mu_{i}\rangle}𝖤i𝟙μ\scriptstyle{\mathsf{E}_{i}\mathbbm{1}_{\mu}}𝖥j𝟙w0(μ)\scriptstyle{\mathsf{F}_{j}\mathbbm{1}_{w_{0}(\mu)}}Θ𝟙μ+αi[n1]\scriptstyle{\Theta\mathbbm{1}_{\mu+\alpha_{i}}[-n-1]} (7.13)

By Proposition 5.9 this diagram commutes (note that we shifted both sides of the equation by n1-n-1). By Theorem 6.4 both horizontal arrows are in fact t-exact equivalences so this restricts to a diagram of abelian categories:

𝒞μ{\mathcal{C}_{\mu}}𝒞w0(μ){\mathcal{C}_{w_{0}(\mu)}}𝒞μ+αi{\mathcal{C}_{\mu+\alpha_{i}}}𝒞w0(μ)αj{\mathcal{C}_{w_{0}(\mu)-\alpha_{j}}}Θ𝟙μ[n]μi\scriptstyle{\Theta\mathbbm{1}_{\mu}[-n]\langle\mu_{i}\rangle}𝖤i𝟙μ\scriptstyle{\mathsf{E}_{i}\mathbbm{1}_{\mu}}𝖥j𝟙w0(μ)\scriptstyle{\mathsf{F}_{j}\mathbbm{1}_{w_{0}(\mu)}}Θ𝟙μ+αi[n1]\scriptstyle{\Theta\mathbbm{1}_{\mu+\alpha_{i}}[-n-1]} (7.14)

Let L𝒞μL\in\mathcal{C}_{\mu} be a simple object, and let L=Θ𝟙μ(L)[n]μiL^{\prime}=\Theta\mathbbm{1}_{\mu}(L)[-n]\langle\mu_{i}\rangle. Note that L𝒞w0(μ)L^{\prime}\in\mathcal{C}_{w_{0}(\mu)} is simple and φI(L)=L\varphi_{I}(L)=L^{\prime}. By the above diagram we have an isomorphism

Θ𝟙μ+αi(𝖤i(L))[n1]𝖥j(L).\displaystyle\Theta\mathbbm{1}_{\mu+\alpha_{i}}(\mathsf{E}_{i}(L))[-n-1]\cong\mathsf{F}_{j}(L^{\prime}).

Now, 𝖥~j(L)𝖥j(L)\widetilde{\mathsf{F}}_{j}(L^{\prime})\subset\mathsf{F}_{j}(L^{\prime}) is the unique simple subobject. On the other hand, since Θ𝟙μ+αi[n1]\Theta\mathbbm{1}_{\mu+\alpha_{i}}[-n-1] is an abelian equivalence, Θ𝟙μ+αi(𝖤~i(L))[n1]Θ𝟙μ+αi(𝖤i(L))[n1]\Theta\mathbbm{1}_{\mu+\alpha_{i}}(\widetilde{\mathsf{E}}_{i}(L))[-n-1]\subset\Theta\mathbbm{1}_{\mu+\alpha_{i}}(\mathsf{E}_{i}(L))[-n-1] is a simple subobject. Therefore

Θ𝟙μ+αi(𝖤~i(L))[n1]𝖥~j(L).\displaystyle\Theta\mathbbm{1}_{\mu+\alpha_{i}}(\widetilde{\mathsf{E}}_{i}(L))[-n-1]\cong\widetilde{\mathsf{F}}_{j}(L^{\prime}).

Since the equivalence class of the left hand side is φIe~i(L)\varphi_{I}\circ\widetilde{e}_{i}(L), this shows that φI\varphi_{I} satisfies the second defining property, and hence the two cactus group actions agree in the case of a simple categorification.

The general case when 𝒞\mathcal{C} is not necessarily a simple categorification follows easily using the Jordan-Hölder filtration (Theorem 3.6) and Lemma 4.6. ∎

8. Examples and Applications

8.1. Examples

We now examine three examples. The first two consider minimal categorifications of the adjoint representation, while the third studies the categorification of the nn-fold tensor product of the standard representation of 𝔰𝔩n\mathfrak{sl}_{n}. For ease of presentation, we ignore gradings and consider non-quantum categorical representations.

Example 8.1.

Let’s consider the first non-trivial example of Theorem 6.4: the minimal categorification of the adjoint representation of 𝔰𝔩2\mathfrak{sl}_{2}. We can model this as follows:

𝕜mod{\mathbbm{k}\mathrm{-mod}}Rmod{R\mathrm{-mod}}𝕜mod{\mathbbm{k}\mathrm{-mod}}ind\scriptstyle{ind}res\scriptstyle{res}res\scriptstyle{res}ind\scriptstyle{ind}

where R=𝕜[x]/(x2)R=\mathbbm{k}[x]/(x^{2}) [CR, Example 5.17]. Here 𝕜mod\mathbbm{k}\mathrm{-mod} is the ±2\pm 2 weight category, and RmodR\mathrm{-mod} is the zero weight category. The arrows describe the 𝖤,𝖥\mathsf{E},\mathsf{F} functors (we omit the higher structure).

Consider the Rickard complex Θ=Θ𝟙0:Db(Rmod)Db(Rmod)\Theta=\Theta\mathbbm{1}_{0}:D^{b}(R\mathrm{-mod})\to D^{b}(R\mathrm{-mod}). For MRmodM\in R\mathrm{-mod}, we have:

Θ(M)=R𝕜MM,\displaystyle\Theta(M)=R\otimes_{\mathbbm{k}}M\to M,

where the differential is given by the action map, and MM is in cohomological degree 0. It’s an exercise to verify that Θ(M)\Theta(M) is quasi-isomorphic to M[1]M^{\prime}[1], where MM^{\prime} is the twist of MM by the automorphism of RR given by a+bxabxa+bx\mapsto a-bx. This shows that Θ[1]\Theta[-1] is the t-exact equivalence MMM\mapsto M^{\prime}. ∎

Example 8.2.

More generally, one can consider the minimal categorification of the adjoint representation of a simple simply-laced Lie algebra 𝔤\mathfrak{g}. This was studied by Khovanov and Huerfano in [HuerKh], who used zigzag algebras to model this category.

For a weight α\alpha of the adjoint representation of 𝔤\mathfrak{g}, the weight category 𝒞α\mathcal{C}_{\alpha} is taken to be 𝕜mod\mathbbm{k}\mathrm{-mod} as long as α0\alpha\neq 0. However, the zero weight category is more interesting: 𝒞0:=Amod\mathcal{C}_{0}:=A\mathrm{-mod}, where AA is the zigzag algebra associated to the Dynkin diagram II of 𝔤\mathfrak{g} (cf. [HuerKh] for the precise definition).

The isomorphism classes of indecomposable projective left AA-modules {Pi}\{P_{i}\} and the isomorphism classes of indecomposable projective right AA-modules {Qi}\{Q_{i}\} are both indexed by iIi\in I. Tensoring with these modules defines functors

𝖤i:𝒞0𝒞αi,MQiAM,\displaystyle\mathsf{E}_{i}:\mathcal{C}_{0}\longrightarrow\mathcal{C}_{\alpha_{i}},\quad M\mapsto Q_{i}\otimes_{A}M,
𝖤i:𝒞αi𝒞0,VPi𝕜V.\displaystyle\mathsf{E}_{i}:\mathcal{C}_{-\alpha_{i}}\longrightarrow\mathcal{C}_{0},\quad V\mapsto P_{i}\otimes_{\mathbbm{k}}V.

The functors 𝖥i\mathsf{F}_{i} are defined analogously and are biadjoint to the 𝖤i\mathsf{E}_{i}.

The Rickard complex Θi𝟙0\Theta_{i}\mathbbm{1}_{0} is given by tensoring with the complex

Pi𝕜QiAP_{i}\otimes_{\mathbbm{k}}Q_{i}\rightarrow A

of (A,A)(A,A)-bimodules, where the differential is given by the multiplication in AA, and AA sits in cohomological degree zero. These functors are autoequivalences of the derived category Db(𝒞0)D^{b}(\mathcal{C}_{0}).

By Theorem 6.4 Θw0𝟙0[n]\Theta_{w_{0}}\mathbbm{1}_{0}[-n] is t-exact, where n+1n+1 is the Coxeter number of 𝔤\mathfrak{g}. This auto-equivalence can be explicitly described as follows: the automorphism of the Dynkin diagram τ:II\tau:I\to I induces an automorphism ψ\psi of AA. Then we claim that Θw0𝟙0[n]\Theta_{w_{0}}\mathbbm{1}_{0}[-n] is the abelian autoequivalence of AmodA\mathrm{-mod} defined by twisting with ψ\psi. Indeed, since Θw0𝟙0[n]\Theta_{w_{0}}\mathbbm{1}_{0}[-n] is an abelian autoequivalence, it is determined up to isomorphism by its action on simple objects. Moreover, the action on simple objects can be read off from the action of WW on the Grothendieck group of AmodA\mathrm{-mod} as follows: [Amod][A\mathrm{-mod}]_{\mathbb{Z}} is isomorphic to the Cartain subalgebra by mapping [Li][L_{i}] (LiL_{i} is the simple head of PiP_{i}) to the simple root vector HiH_{i}. And on the root vectors we have that w0Hi=Hτ(i)w_{0}\cdot H_{i}=-H_{\tau(i)}

Example 8.3.

Let 𝔤=𝔰𝔩n\mathfrak{g}=\mathfrak{sl}_{n} and consider the nn-fold tensor power of the standard representation VnV^{\otimes n}. Categorifications of VnV^{\otimes n} have been well-studied, and a model 𝒞\mathcal{C} for this categorical representation can be constructed using the BGG category 𝒪\mathcal{O} of 𝔤\mathfrak{g} [Sussan2007, MS09]. In this model, the principal block 𝒪0𝒪\mathcal{O}_{0}\subset\mathcal{O} appears as the zero weight category of 𝒞\mathcal{C}, and the Rickard complexes acting on Db(𝒪0)D^{b}(\mathcal{O}_{0}) are the well-known shuffling functors.

By Theorem 6.8 Θw0𝟙0:Db(𝒪0)Db(𝒪0)\Theta_{w_{0}}\mathbbm{1}_{0}:D^{b}(\mathcal{O}_{0})\to D^{b}(\mathcal{O}_{0}) is a perverse equivalence with respect to an isotypic filtration. In fact, this recovers the type A case of a theorem of the third named author [LosCacti], using completely different methods (in [LosCacti] the perversity of Θw0\Theta_{w_{0}} is proved using the theory of W-algebras). We can interpret the filtration of 𝒪0\mathcal{O}_{0} arising from our perspective concretely using the Robinson-Schensted correspondence .

Recall that the simple objects in 𝒪0\mathcal{O}_{0} are the irreducible highest weight representations L(w),wSnL(w),w\in S_{n}, where L(w)L(w) has highest weight wρρw\rho-\rho (ρ\rho is the half-sum of positive roots of 𝔰𝔩n\mathfrak{sl}_{n}).

We view a partition λ\lambda of nn simultaneously as a dominant integral weight for 𝔰𝔩n\mathfrak{sl}_{n}, and as an index for the irreducible Specht module SλS^{\lambda} of the symmetric group SnS_{n}. Let SYT(λ)\operatorname{SYT}(\lambda) denote the set of standard Young tableau of shape λ\lambda, and let d(λ)=|SYT(λ)|d(\lambda)=|\operatorname{SYT}(\lambda)|. Recall that d(λ)=dimSλd(\lambda)=\dim S^{\lambda}.

Choose an ordering of the partitions of nn, λ1,,λr\lambda_{1},\ldots,\lambda_{r}, so that if λiλj\lambda_{i}\prec\lambda_{j} in the dominance order, then i<ji<j. Note that the dominance order on partitions of nn is equivalent to the positive root ordering on partitions (thought of as weights for 𝔰𝔩n\mathfrak{sl}_{n}). By Remark 3.9 there is an isotypic filtration on 𝒞\mathcal{C}, 0𝒞1𝒞N=𝒞,0\subset\mathcal{C}_{1}\subset\cdots\subset\mathcal{C}_{N}=\mathcal{C}, where 𝒞i/𝒞i1\mathcal{C}_{i}/\mathcal{C}_{i-1} is an isotypic categorification of type λi\lambda_{i}. Then Θw0\Theta_{w_{0}} is a perverse equivalence with respect to the filtration 𝒪0,i=𝒪0𝒞i\mathcal{O}_{0,i}=\mathcal{O}_{0}\cap\mathcal{C}_{i} and the perversity function p(i)=ht(λi)p(i)=ht(\lambda_{i}).

We would like now to define the categories 𝒪0,i\mathcal{O}_{0,i} more explicitly using the Robinson-Schensted correspondence, which recall is a bijection [Sagan]:

𝖱𝖲:Snj=1rSYT(λj)×SYT(λj),w(P(w),Q(w)).\displaystyle\mathsf{RS}:S_{n}\longrightarrow\bigsqcup_{j=1}^{r}\operatorname{SYT}(\lambda_{j})\times\operatorname{SYT}(\lambda_{j}),\;w\mapsto(P(w),Q(w)).

We will use a crystal analogue of classical Schur-Weyl duality. This is given by a crystal isomorphism:

𝐁(ϖ1)nj=1r𝐁(λj)×SYT(λj).\displaystyle\mathbf{B}(\varpi_{1})^{\otimes n}\longrightarrow\bigsqcup_{j=1}^{r}\mathbf{B}(\lambda_{j})\times\operatorname{SYT}(\lambda_{j}). (8.4)

Now, recall that for a partition λ\lambda of nn, the underlying set of the crystal 𝐁(λ)\mathbf{B}(\lambda) can be chosen to be the set of semistandard Young tableaux of shape λ\lambda with entries 1,,n1,\ldots,n, and the weight zero subset of 𝐁(λ)\mathbf{B}(\lambda) is precisely SYT(λ)\operatorname{SYT}(\lambda). The essential point is that the isomorphism (8.4) can be chosen so that it restricts to the map 𝖱𝖲\mathsf{RS} on the elements of weight zero ([Shimozono, Theorem 3.5]). (Note that the elements of weight zero in 𝐁(ϖ1)n\mathbf{B}(\varpi_{1})^{\otimes n} are naturally identified with permutations of 1,,n1,\ldots,n.)

This shows that as an element of the crystal 𝐁(ϖ1)n\mathbf{B}(\varpi_{1})^{\otimes n}, [L(w)][L(w)] is in a connected component whose highest weight is the shape of Q(w)Q(w) (or equivalently P(w)P(w)). Therefore, following Remark 3.10, we can construct the isotypic filtration be defining 𝒪0,i\mathcal{O}_{0,i} to be the Serre subcategory of 𝒪0\mathcal{O}_{0} generated by L(w)L(w) such that the shape of Q(w)Q(w) is among λ1,,λi\lambda_{1},\ldots,\lambda_{i}.

8.2. Type A combinatorics

In this final section, we specialise to type A and discuss the combinatorics of Kazhdan-Lusztig bases and standard Young tableaux from the vantage of perverse equivalences.

Set I=An1={1,,n1}I=A_{n-1}=\{1,\ldots,n-1\}. We continue with the notation in Example 8.3 and view a partition λn\lambda\vdash n simultaneously as a dominant integral highest weight for 𝔰𝔩n\mathfrak{sl}_{n}, and as an index of the Specht module SλS^{\lambda}. Recall that by Schur-Weyl duality, L(λ)0L(\lambda)_{0} is isomorphic to SλS^{\lambda}. The Kazhdan-Lusztig basis of the Hecke algebra naturally descends to a basis of SλS^{\lambda}, which we denote {CT}\{C_{T}\} indexed by TSYT(λ)T\in\operatorname{SYT}(\lambda). For further details we recommend the exposition in [Rhoad].

Consider the minimal categorification (λ)\mathcal{L}(\lambda), and in particular its zero weight category (λ)0\mathcal{L}(\lambda)_{0}. For convenience, we forget the grading and work in the non-quantum setting. The simple objects L(T)(λ)0L(T)\in\mathcal{L}(\lambda)_{0} are indexed by TSYT(λ)T\in\operatorname{SYT}(\lambda), and hence a perverse equivalence 𝖥:Db((λ)0)Db((λ)0)\mathsf{F}:D^{b}(\mathcal{L}(\lambda)_{0})\to D^{b}(\mathcal{L}(\lambda)_{0}) induces a bijection φ𝖥:SYT(λ)SYT(λ)\varphi_{\mathsf{F}}:\operatorname{SYT}(\lambda)\to\operatorname{SYT}(\lambda).

The bijection φI\varphi_{I} studied in the previous section specialises to the well-known Schützenberger involution on standard Young tableau, otherwise known as the “evacuation operator” ee [Sagan]. Indeed, by Theorem 7.12 φI\varphi_{I} recovers the cactus group action on the crystal 𝖨𝗋𝗋(𝒞)\mathsf{Irr}(\mathcal{C}) by generalised Schützenberger involutions. The Schützenberger involution is well-known to agree with the evacuation operator [StanleyII, Theorem A1.2.10]). Note that this is elementary: it follows directly from the fact that the evacuation operator satisfies the properties of Definition 7.2 on standard Young tableaux.

The promotion operator j:SYT(λ)SYT(λ)j:\operatorname{SYT}(\lambda)\to\operatorname{SYT}(\lambda) is another important function in algebraic combinatorics, which is closely connected to the RSK correspondence and related ideas such as jeu de taquin. Letting J={1,,n2}IJ=\{1,\ldots,n-2\}\subset I, we can express promotion in terms of the Schützenberger involution: j=φIφJj=\varphi_{I}\varphi_{J}. We refer the reader to [Sagan] for a detailed exposition. We can now see easily that promotion also arises from a perverse equivalence:

Proposition 8.5.

Let cn=(1,2,,n)Snc_{n}=(1,2,\ldots,n)\in S_{n} be the long cycle. Then

Θcn:Db((λ)0)Db((λ)0)\Theta_{c_{n}}:D^{b}(\mathcal{L}(\lambda)_{0})\to D^{b}(\mathcal{L}(\lambda)_{0})

is a perverse equivalence whose associated bijection is the promotion operator: φΘcn=j\varphi_{\Theta_{c_{n}}}=j.

Proof.

Notice that cn=w0w0Jc_{n}=w_{0}w_{0}^{J} for JJ as above. Now recall that Θw0\Theta_{w_{0}} is (up to shift) a t-exact autoequivalence of Db((λ)0)D^{b}(\mathcal{L}(\lambda)_{0}) (Theorem 6.4). Since Θw0J\Theta_{w_{0}^{J}} is a perverse equivalence (Theorem 6.8), its inverse is too. Therefore ΘcnΘw0Θw0J1\Theta_{c_{n}}\cong\Theta_{w_{0}}\Theta_{w_{0}^{J}}^{-1} is also a perverse autoequivalence. By Lemma 4.7 we have that φΘcn=φIφJ\varphi_{\Theta_{c_{n}}}=\varphi_{I}\varphi_{J}, and hence we recover the promotion operator. ∎

We can also use this set-up to extract information about the action of SnS_{n} on the Kazhdan-Lusztig basis of S(λ)S(\lambda). This is based on the following elementary lemma:

Lemma 8.6.

Let wSn,λnw\in S_{n},\lambda\vdash n and suppose Θw:Db((λ)0)Db((λ)0)\Theta_{w}:D^{b}(\mathcal{L}(\lambda)_{0})\to D^{b}(\mathcal{L}(\lambda)_{0}) is t-exact up to shift. Then for any TSYT(λ)T\in\operatorname{SYT}(\lambda), wCT=±CSw\cdot C_{T}=\pm C_{S}, where S=φΘw(T)S=\varphi_{\Theta_{w}}(T).

Proof.

Since Θw\Theta_{w} is t-exact up to shift, we have:

[Θw(L(T))]=±[L(S)].\displaystyle[\Theta_{w}(L(T))]=\pm[L(S)].

The result now follows since the isomorphism [(λ)0]Sλ,L(T)CT[\mathcal{L}(\lambda)_{0}]_{\mathbb{C}}\cong S^{\lambda},L(T)\mapsto C_{T}, is SnS_{n}-equivariant, and by [BauqWeyl, Proposition 10], the action of the braid group B=BnB=B_{n} on L(λ)0L(\lambda)_{0} factors through SnS_{n}. ∎

Applying this lemma to Theorem 6.4 we obtain a result of Berenstein-Zelevinsky and Stembridge:

Corollary 8.7.

[BZ96, Stem96] The action of the longest element on the Kazhdan-Lusztig basis recovers the Schützenberger evacuation operator, i.e. for w0Sn,λnw_{0}\in S_{n},\lambda\vdash n and TSYT(λ)T\in\operatorname{SYT}(\lambda), we have that w0CT=±Ce(T)w_{0}\cdot C_{T}=\pm C_{e(T)}.

Similarly we can prove a result of Rhoades regarding the action of the long cycle cn=(1,2,,n)c_{n}=(1,2,\ldots,n) on the Kazhdan-Lusztig basis. Note that in the statement below the significance of λ\lambda being rectangular is that the restriction of S(λ)S(\lambda) to Sn1S_{n-1} in this case remains irreducible.

Proposition 8.8.

(cf. [Rhoad, Proposition 3.5]) Let λ=(ab)n\lambda=(a^{b})\vdash n be a rectangular partition. Then for any TSYT(λ)T\in\operatorname{SYT}(\lambda), the action of the long cycle on the Kazhdan-Lusztig basis element CTC_{T} recovers the promotion operator:

cnCT=±Cj(T).c_{n}\cdot C_{T}=\pm C_{j(T)}.
Proof.

As above, set J={1,,n2}IJ=\{1,\ldots,n-2\}\subset I. Recall that, since (λ)\mathcal{L}(\lambda) is a simple categorification, we know Θw0:Db((λ)0)Db((λ)0)\Theta_{w_{0}}:D^{b}(\mathcal{L}(\lambda)_{0})\to D^{b}(\mathcal{L}(\lambda)_{0}) is t-exact up to shift by Theorem 6.4. However, L(λ)|𝔰𝔩n1L(\lambda)|_{\mathfrak{sl}_{n-1}} is no longer simple, so a priori we only know that Θw0J:Db((λ)0)Db((λ)0)\Theta_{w_{0}^{J}}:D^{b}(\mathcal{L}(\lambda)_{0})\to D^{b}(\mathcal{L}(\lambda)_{0}) is perverse, but not necessarily t-exact up to shift. We first prove that Θw0J\Theta_{w_{0}^{J}} is indeed t-exact up to shift.

Consider the set of functors which are monomials in the Chevalley functors indexed by JJ:

𝐌={𝖦j1𝖦js|𝖦{𝖤,𝖥},jJ}.\displaystyle\mathbf{M}=\{\mathsf{G}_{j_{1}}\cdots\mathsf{G}_{j_{s}}\;|\;\mathsf{G}\in\{\mathsf{E},\mathsf{F}\},\;j_{\ell}\in J\}.

Let 𝒞\mathcal{C} be the abelian category generated by {𝖬(X)|X(λ)0,𝖬𝐌}\{\mathsf{M}(X)\;|\;X\in\mathcal{L}(\lambda)_{0},\mathsf{M}\in\mathbf{M}\}, that is, 𝒞\mathcal{C} is the category closed under subobjects and quotients of objects of the form 𝖬(X)\mathsf{M}(X). Since the Chevalley functors are exact, it is easy to see that 𝒞\mathcal{C} is a categorical representation of Uq(𝔰𝔩n1)U_{q}(\mathfrak{sl}_{n-1}).

Let ν(n1)\nu\vdash(n-1) be the (unique) partition obtained from λ\lambda by removing a box. We claim that 𝒞\mathcal{C} is a categorification of L(ν)L(\nu). Note that this is a categorification of the fact that L(ν)Uq(𝔰𝔩n1)L(λ)0L(\nu)\cong U_{q}(\mathfrak{sl}_{n-1})\cdot L(\lambda)_{0}. Indeed, it is clear that [𝒞][\mathcal{C}]_{\mathbb{C}} contains L(ν)L(\nu). On the other hand, if 𝒞μ0\mathcal{C}_{\mu}\neq 0 then μ\mu is in the root lattice of 𝔰𝔩n1\mathfrak{sl}_{n-1}, and L(ν)L(\nu) is the unique constituent of L(λ)|𝔰𝔩n1L(\lambda)|_{\mathfrak{sl}_{n-1}} whose weights are in the root lattice.

Now observe that 𝒞0=(λ)0\mathcal{C}_{0}=\mathcal{L}(\lambda)_{0}. Hence, by Corollary 6.7 it follows that Θw0J:Db((λ)0)Db((λ)0)\Theta_{w_{0}^{J}}:D^{b}(\mathcal{L}(\lambda)_{0})\to D^{b}(\mathcal{L}(\lambda)_{0}) is t-exact up to shift. Since Θw0\Theta_{w_{0}} is also t-exact up to shift, it follows that Θcn\Theta_{c_{n}} is too. The result now follows by Proposition 8.5 and Lemma 8.6. ∎

References