Categorical braid group actions
and cactus groups
Abstract.
Let be a semisimple simply-laced Lie algebra of finite type. Let be an abelian categorical representation of the quantum group categorifying an integrable representation . The Artin braid group of acts on by Rickard complexes, providing a triangulated equivalence
where is a weight of , and is a positive lift of the longest element of the Weyl group.
We prove that this equivalence is t-exact up to shift when is isotypic, generalising a fundamental result of Chuang and Rouquier in the case . For general , we prove that is a perverse equivalence with respect to a Jordan-Hölder filtration of .
Using these results we construct, from the action of on , an action of the cactus group on the crystal of . This recovers the cactus group action on 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 categorifications on abelian categories [CR]. Their definition mirrors the notion of an 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 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 , and provides a derived equivalence between opposite weight categories.
Subsequently, Rouquier and Khovanov-Lauda vastly generalised this theory to quantum symmetrisable Kac-Moody algebras [Rou2KM, KLI, KLII]. Let be any field. A graded abelian -linear category endowed with a categorical representation of possesses a family of Rickard complexes , indexed by the simple roots of , acting on the derived category .
Henceforth let be a semisimple simply-laced Lie algebra of finite type with Dynkin diagram , its Weyl group, and its Artin braid group. Let be a categorical representation of 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 on , 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 representations [ST01].
To describe our first theorem, recall that minimal categorifications are certain distinguished categorifications of simple representations. On these the Rickard complex is t-exact up to shift [CR, Theorem 6.6]. Notice that this is a result about 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 , where we show that the composition of Rickard complexes corresponding to a positive lift of the longest element is t-exact up to shift on any isotypic categorification. More precisely:
Theorem A.
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 and Chevalley functors established by Cautis and Kamnitzer to deduce the commutation relations involving (Proposition 5.9). We then use these relations to prove the theorem by induction on .
Our second theorem describes on an arbitrary categorical representation of , also generalising a result of Chuang-Rouquier in the case . Indeed, their study of the Rickard complex on an categorification led them to define the notion of a “perverse equivalence” [CRperv].
Consider an equivalence of triangulated categories with t-structures [BBD]. Suppose further that (respectively ) is filtered by thick triangulated subcategories
and is compatible with these filtrations (cf. Section 4.1 for precise definitions). Then, roughly speaking, is a perverse equivalence if on each subquotient 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 be a categorical representation of , and let be any weight. The derived equivalence is a perverse equivalence with respect to a Jordan-Hölder or isotypic filtration of .
Note that if is a subdiagram, and is corresponding longest element, then this theorem implies that is a perverse equivalence for any . We also remark that our argument go through in the ungraded setting, where is a categorical representation of .
Let us explain the filtration arising in Theorem B more precisely. We apply Rouquier’s Jordan-Hölder theory for representations of -Kac-Moody algebras to our setting [Rou2KM]. We thus obtain a filtration of by Serre subcategories,
such that each factor is a subrepresentation, and each subquotient categorifies either a simple module (Theorem 3.6) or an isotypic component (Remark 3.9). Then is a perverse equivalence with respect to the filtration whose -th filtered component consists of complexes in with cohomology supported in .
We remark that in the case this gives a more conceptual proof of a result of Chuang-Rouquier [CRperv, Proposition 8.4]. If is the tensor product categorification of the -fold tensor product of the standard representation of , 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 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 .
Recall that to an integrable representation of , Kashiwara associated its crystal basis [Kash90], which is closely related to Lusztig’s canonical basis [GL92]. If is categorified by then there is a natural identification , the set of isomorphism classes of simple objects in 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 controls the commutator in the category of representations of [HK06]. There is also an internal cactus group action, mirroring Lusztig’s internal braid group action on . Indeed, there is a cactus group associated to (or rather to its Dynkin diagram ), which can be presented by generators indexed by connected subdiagrams (cf. Section 7.1). Then acts on via the so-called Schützenberger involutions (cf. Theorem 7.3).
So, starting with an integrable representation of the quantum group we obtain: an action of on V, a -crystal , and an action of on . We schematically picture this situation as follows:
Naturally one asks: can we “crystallise” the braid group action on directly to obtain the cactus group action on ? Our results allow us to answer this in the affirmative.
The key point is that a perverse equivalence induces a bijection , where denotes the heart of the t-structure. In the setting of Theorem B, we obtain a bijection . In fact, if is a subdiagram and is the corresponding Lie subalgebra, we can regard as a categorical representation of by restriction. By Theorem B we also obtain a bijection .
Theorem C.
We thus obtain the sought-after crystalisation process for braid groups:
which associates a cactus group set to the braid group representation of on . The first appearance of such a crystalisation process is in the work of the third author, where a cactus group action on 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 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 on the Kazhdan-Lusztig basis of a Specht module of 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 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 of finite type. Recall that we have an associated Cartan datum and a root datum, which consists of:
-
•
A finite set ,
-
•
a symmetric bilinear form on satisfying and for all ,
-
•
a free -module , called the weight lattice, and
-
•
a choice of simple roots and simple coroots satisfying , where is the natural pairing.
The quantum group is the unital, associative, algebra generated by subject to relations:
-
(1)
and for any ,
-
(2)
for any ,
-
(3)
for any ,
-
(4)
, where we set , and
-
(5)
for all ,
where , and .
We let , so that is a Cartan matrix. Given we abbreviate , and let
be the set of dominant weights.
Let be the root lattice, defined as the -span of the simple roots, and let be the -span of the simple roots. We define the usual preorder on by if . For let denote the height of , i.e. .
When convenient, we also view as the Dynkin diagram of , and make reference to subdiagrams or diagram automorphisms of .
2.2. Braid group actions on integrable representations of .
Given a -module and we let denote the weight space of . For we let be the irreducible representation of of highest weight . Let denote the -isotypic component of . We say that is isotypic if there exists such that .
The representation has a canonical basis, which we denote by [Lusbook]. We let (respectively ) denote the unique highest weight (respectively lowest weight) element of .
Let denote the braid group of type , which is generated by subject to the braid relations:
Let be the Weyl group of type , which has generators subject to the braid relations, and in addition the quadratic relation . Let be the longest element. Recall that acts on via . We define by the equality for any .
To a subdiagram, we associate the parabolic subgroup, its longest element, and the bijection given by
For any we can consider its positive lift , where and is any reduced decomposition.
Let be an integrable representation of . A fundamental structure of , discovered by Lusztig, is that it admits (several) braid group symmetries, sometimes referred to as the “quantum Weyl group actions”. To recall this, let denote the projection onto the weight space. For each we define by:
(2.1) |
Note that the indexing set of this sum is infinite, but the sum itself is finite on . In the notation of [Lusbook], . (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 defines an action of on , and so we can unambiguously write for any .
3. Categorical representations of
3.1. Notation
Fix a field of any characteristic. In this paper we will be concerned mostly with abelian -linear categories . We will assume throughout that each block of is a finite abelian category [EGNO, Definition 1.8.6].
Recall that a category is graded if it is equipped with an auto-equivalence called the “shift functor”. We let be the auto-equivalence obtained by applying the shift functor times. We denote by the set of equivalence classes of simple objects of up to shift.
A functor between graded categories is graded if it commutes with the shift functors. We denote by the Grothendieck group of an abelian category . If is graded we denote by the quotient of by the additional relation . This quotient is naturally a -module. Set
For a -algebra , we let be the category of finitely generated -modules, and if is -graded, we let be the category of finitely generated -graded modules. These are naturally -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 . 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 or of . At the categorical level, the difference between or of arises from grading considerations; in categorifications of representations of , one works with categories enriched in graded vector spaces, whereas in categorifications of representations of 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 consists of the following data:
-
•
A family of graded abelian -linear categories indexed by . We refer to each as a weight category.
-
•
Exact graded functors and , for and . We refer to 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)
The functors and are biadjoint up to a specified degree shift (see (3.1) below).
-
(2)
The powers of carry an action of the KLR algebra associated to , where denotes a choice of units in . These units satisfy some restrictions which are not relevant for us.
-
(3)
We have the following isomorphisms:
This notation is explained as follows: for a Laurent polynomial , is a direct sum over of copies of , and .
Usually we just say that is a categorical representation of (the remaining data is implicit). Given an integrable -module , we say that is a categorification of if is a categorical representation of such that as -modules.
An additive categorification is a categorical representation on a graded additive -linear category satisfying the same conditions as above, except the Chevalley functors are of course only required to be additive. We let be the idempotent completion of .
Given an abelian category , we consider the additive category defined as the full subcategory of projective objects in . Note that if is a categorical representation of , then naturally inherits the structure of an additive categorification.
As a consequence of this definition, and in particular condition (2), there exist divided power functors 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:
(3.1) | ||||
(3.2) |
3.3. Crystals
We recall the definition of a crystal.
Definition 3.3 ([Kash90]).
A -crystal is a finite set together with maps:
for all , such that:
-
(1)
for any , if and only if ,
-
(2)
for all , if then , and if then ,
-
(3)
for all , , ,
-
(4)
for all , .
Any -representation has a corresponding -crystal . The underlying set of is in natural bijection with a particular basis of (the “global crystal basis” or “canonical basis”). The maps 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 , the canonical basis of carries a natural crystal structure [GL92].
The crystal of naturally arises via categorical representation theory. Namely, as we describe in the next proposition, if is a categorification of , then carries a crystal structure isomorphic to . 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 together with:
-
•
Kashiwara operators defined as , for ,
-
•
for , and
-
•
, and ,
is a -crystal isomorphic to the crystal .
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 called the minimal categorification and denoted [CR, Rou2KM, KLI, Webmerged, KK]. It is characterized by the fact that . We let 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 -linear categories the Deligne tensor product is universal for the functor assigning to every such abelian category the category of bilinear bifunctors right exact in both variables [EGNO, Definition 1.11.1]. The tensor product is again a finite abelian -linear category, and there is a bifunctor . This construction enjoys the following properties:
-
(1)
The tensor product is unique up to unique equivalence,
-
(2)
for finite -algebras we have that , and
-
(3)
.
Let be a categorical representation, and let be a finite -linear abelian category. We can endow with a structure of a categorical representation, by setting , defining Chevalley functors , etc. If is a simple categorification, then clearly is an isotypic categorification. Conversely, we have:
Lemma 3.5.
Let be an isotypic categorification of type . Then there exists an abelian category such that .
Proof.
Since is an isotypic categorification, so is . By Rouquier’s Jordan-Hölder series for additive categorifications [Rou2KM, Theorem 5.8], there exists an additive -linear category such that
Note that no filtration appears here since, in the notation of [Rou2KM], .
Let (respectively ) be a complete list of the projective indecomposable objects of (respectively ). Let , and let . By Morita theory, . On the other hand,
Since we have the desired result with . ∎
Theorem 3.6.
Let be a categorical representation of . Then there exists a filtration by Serre subcategories
(3.7) |
such that for each : is a subrepresentation of , is a simple categorification of type , and the list of highest weights is weakly increasing so that .
Proof.
The -crystal is isomorphic to a finite direct sum of irreducible crystals for various , i.e. we have an isomorphism
where and only finitely many are nonzero.
Define , and let . We claim that there exists a highest weight simple object . Indeed, otherwise for any simple object there exists such that . This implies that , and by Proposition 3.4 we conclude that has no highest weight elements of weight , a contradiction.
Now take which is minimal with respect to , and let be a highest weight object. Let be the Serre subcategory of generated by objects
(3.8) |
By the exactness and bi-adjunction of the Chevalley functors, is a subrepresentation of . Moreover it categorifies . Indeed, by our choice of there cannot be any highest weight objects with weight occuring in , and by construction the only simple object in is .
Next consider the categorical representation and repeat this construction. This produces a Serre subcategory which is again a simple categorification. Let be the natural quotient functor, and define . Clearly, we have that , is Serre, it is a subrepresentation, and is a simple categorification.
Iterating this process produces a filtration of 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 is a list of the distinct isotypic types appearing in , and we choose any ordering of this list so that , then there is a filtration such that categorifies the isotypic component of of highest weight . 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 from the crystal structure on . Indeed, suppose that decomposes into components
where are distinct dominant integral weights, and is a disjoint union of copies of . Further, we arrange the weights as above so that . Let be the Serre subcategory of generated by simple objects such that , where . Then it follows from Remark 3.9 that is an isotypic filtration of .
3.5. The categorical braid group action
Let be a categorical representation of and . We define a complex of functors , supported in nonpositive cohomological degrees, where for the component is
The differential is defined using the counits of the bi-adjunctions relating and (see [Cauclasp, Section 4] for details). This produces a functor , which following Chuang and Rouquier we call the Rickard complex.
It’s straightforward to verify that the Rickard complex categorifies Lusztig’s braid group operators ([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 , is an equivalence of triangulated categories.
Theorem 3.12.
[CK3, Theorem 6.3] The Rickard complexes satisfy the braid relations:
thereby defining a weak action of on .
This action is “weak” since we don’t make any claim on the canonicity of the functorial isomorphisms. Nevertheless, for we define , where is a reduced expression. Thus is defined up to isomorphism, but not canonical isomorphism. Luckily, everything we do in this paper only requires to be defined up to isomorphism.
As a consequence of their proof of Theorem 3.11, Chuang and Rouquier show that the inverse of is its right adjoint. We denote this functor by , so that . As a complex of functors, is supported in nonnegative cohomological degrees, where for the component is
4. Perverse equivalences
4.1. General definition
Let be a triangulated category with shift functor . In the cases of most interest to us, is a subcategory of a derived category, in which case is the homological shift functor. Suppose has a t-structure , with heart [BBD]. Recall that a triangulated functor between triangulated categories with -structure is -exact if and . We let denote the pre-composition of with the -shift .
Now let be a thick triangulated subcategory, and consider the quotient functor . Following [CRperv], we say that is compatible with if is a t-structure on . By [CRperv, Lemmas 3.3 & 3.9], if is compatible with then , and is a t-structure on such that .
Now suppose that are two triangulated categories with t-structures . Suppose further that we have filtrations by thick triangulated subcategories:
such that for every , is compatible with and is compatible with . By [CRperv, Lemma 3.11], is also compatible with , and hence inherits a natural t-structure. Let . The data 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 is a (graded) perverse equivalence with respect to if for every ,
-
(1)
, and
-
(2)
the induced equivalence is t-exact.
For brevity, we say 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 induces a bijection . Indeed, by (2) induces a bijection , and these yield .
Although the construction of depends on a choice of perversity triple, the resulting bijection does not when have finitely many simple objects. This follows from the following lemma.
Lemma 4.2 ([LosCacti], Lemma 2.4).
Suppose that have finitely many simple objects, and for let be a perverse equivalence with respect to the perversity datum . If the induced maps coincide then .
Corollary 4.3.
Suppose that have finitely many simple objects, and is perverse. Then is independent of the choice of perversity triple.
Proof.
Suppose is a graded perverse equivalence with respect to two choices of perversity triples , . Now apply the lemma. ∎
The proofs of the following lemmas are straightforward.
Lemma 4.4.
Suppose is perverse. Then for any , is also perverse and .
Lemma 4.5.
Suppose is a perverse equivalence with respect to , and is a perverse equivalence with respect to . Then is a perverse equivalence with respect to , and .
Lemma 4.6.
Let be triangulated with t-structures , and let be thick triangulated subcategories such that is compatible with and is compatible with . Suppose further that is a perverse equivalence with respect to , and .
Define and . Let and be the induced equivalences. Then:
-
(1)
is a perverse equivalence with respect to , and .
-
(2)
is a perverse equivalence with respect to , and
Lemma 4.7 ([LosCacti], Lemma 2.4).
Suppose is perverse, and (respectively ) is an autoequivalence of (respectively ) which is t-exact up to shift. Then is perverse, and .
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 is an abelian category, then the bounded derived category has a standard t-structure whose heart is .
Given a a Serre subcategory, we let denote the thick subcategory consisting of complexes with cohomology supported in . The category inherits a natural t-structure from the standard t-structure on : and . The heart of the t-structure on is . Moreover, if is another Serre subcategory then the -structure on is compatible with . In particular, the quotient inherits a natural t-structure whose heart is .
For the remainder of this section let 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 using filtrations on the abelian categories: and are filtrations by shift-invariant Serre subcategories:
Then a graded equivalence is a perverse with respect to if conditions (1) and (2) of Definition 4.1 hold for and
As above, a graded perverse equivalence induces a bijection .
The following standard lemma will be useful in the proof of our main result.
Lemma 4.8.
Let be abelian categories, and Serre subcategories. Let be integers, and be exact functors for . Suppose these functors fit into a complex , defining a functor
If for all , then .
5. Some commutation relations
We fix throughout a categorical representation of . Recall that is the longest word, and let be a reduced expression. We consider the composition of Rickard complexes which categorifies the positive lift in of :
In preparation for the proofs our main results in the next section, we prove some commutation relations between 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 we have the following relations:
Remark 5.2.
The careful reader will notice that actually Cautis proves Lemma 5.1 under certain conditions on . For instance, the first relation is only proven in the case when . To deduce the general case from this, one can rewrite the relation as . Now recall that there is an anti-automorphism on the 2-category which on objects maps [Laucq, Section 5.6]. This anti-automorphism maps to , and hence applying it to the relation above we deduce the desired relation in the case when .
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) 2-category, which of course implies it also in any 2-representation [Vera].
Next we recall the categorical analogues of commutators acting on representations of . Given nodes such that and , define complexes of functors
In both instances the differential is given by the element 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 and suppose . We have the following isomorphisms:
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 with one letter marked: . From we can define a functor and an element of :
Note that unlike , forgets the location of the marked letter. We say that is reduced, if the corresponding unmarked word is a reduced expression for .
We will apply braid relations to marked words. Away from the marked letter these operate as usual, and at the marked letter we have:
(5.4) | ||||
(5.5) |
For marked words we write if they are related by a sequence of braid relations.
Lemma 5.6.
Let be marked words which differ by a single braid relation. Then there exists such that .
Proof.
If the relation doesn’t involve the marked letter then since Rickard complexes satisfy the braid relations [CK3, Theorem 2.10]. Suppose then that the relation does involve the marked letter. If the result follows from the fact that . Otherwise . Set . Applying Lemma 5.3 (twice) we deduce that
Hence the result follows. ∎
Corollary 5.7.
Let be marked words such that . Then there exists an integer such that .
Lemma 5.8.
Let and be reduced marked words such that . Then .
Proof.
We prove the claim by induction on . Since there is a Matsumoto sequence of length relating the unmarked words and . Let denote the resulting sequence of unmarked words, starting at and ending . Let be the first word in this sequence whose last entry is not . In other words, the first steps in the sequence do not involve the last entry of , but the 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 . Further, we know that for some , and the next step involves the marked letter.
If the step is an application of (5.5) then and the step is:
Let , i.e. is obtained from by applying the first steps to its (last ) entries as were used for the (first ) entries of . Note that and are reduced marked words such that , and both end in . Hence, if we delete these last two entries we obtain two reduced marked words and such that . By induction and therefore . Since clearly , we have our desired result:
If the step is an application of (5.4) then the step is:
Note that and are reduced marked words such that , and both end in . Hence we can delete this last entry and apply a similar analysis as above.
∎
5.3. The relation between and Chevalley functors
We now have the machinery in place to prove our main relation.
Proposition 5.9.
For any we have the following relations:
Proof.
We’ll prove the first relation, the second being entirely analogous. For two functors we write if there exist integers such that .
We first show that by induction on the rank of . The base case, when , follows from Lemma 5.1. For the inductive step let be a strict subdiagram containing . Recall the bijection induced by the longest element . Let and let be a reduced expression. Define two marked words:
Note that . By the inductive hypothesis we have , and therefore
Note that satisfy the hypothesis of Lemma 5.8, so by Lemmas 5.7 and 5.8 we have that , and hence
We now know there exist integers such that , and it remains to show that . Let and let be a reduced expression. Define two marked words:
Note that satisfy the hypothesis of Lemma 5.8, so by Lemmas 5.7 and 5.8 there exists an integer such that . Hence we have that
showing that .
On the other hand, we can deduce by inspecting the relation on the level of Grothendieck groups. Namely, by [KT, Lemma 5.4], we have that
showing that . ∎
6. On t-exactness and perversity of
In this section we will state and prove the central results of the paper. We fix throughout a categorical representation of , and let be a reduced expression.
6.1. on isotypic categorifications
In this section we prove that is t-exact on any isotypic categorification. Fix . We write and .
Lemma 6.1.
Let and set . The weight space is zero.
Proof.
By [Humph, Proposition 21.3], it suffices to find such that . Take and, noting that , the result follows. ∎
Recall that and are the highest and lowest weight elements of the canonical basis .
Lemma 6.2.
[KT, Comment 5.10] We have .
Proposition 6.3.
The equivalence satisfies , where both are considered as complexes concentrated in degree zero. In particular, under the equivalences , is isomorphic to the identity autofunctor of .
Proof.
Consider first the case . On the minimal categorification of highest weight , we have that for some by [CR, Theorem 6.6]. Since , by Lemma 6.2 we conclude that , and hence .
For general , suppose is simple and for some . Consider as a categorical representation of by restriction to the -th root subalgebra. Then for some we have a morphism of categorical representations such that [CR, Theorem 5.24].
The functor is equivariant for the categorical action on determined by (in fact it is strongly equivariant in the sense of [LoWe, Definition 3.1]), and hence commutes with . Therefore we have that
and so is in homological degree zero. It follows that in the case when is not necessarily simple (but still assume that ), is still in homological degree zero. Indeed, by induction on the length of a Jordan-Hölder filtration of one deduces this since is extension closed.
Theorem 6.4.
Let and set . For any ,
is t-exact, where .
Proof.
Consider . We will first prove by induction on that there exists an integer such that
(6.5) |
where .
The base case when follows by Proposition 6.3. For the inductive step write . Note that . By Proposition 5.9 we have that
By hypothesis
for some , and hence Equation (6.5) follows.
Since up to grading shift, any projective indecomposable object in , respectively , is a summand of an object of the form (respectively ), it follows that takes projective objects in to projective objects in . Since 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 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 . Indeed we use [CR, Theorem 6.6] explicitly in the proof of Proposition 6.3, and more generally we use the fact the is invertible throughout.
Corollary 6.7.
Suppose is an isotypic categorification of type , for some , and let . Then is a t-exact equivalence, where .
Proof.
By Lemma 3.5, there exists an abelian -linear category such that as categorical representations. We have that
proving that is a t-exact equivalence. ∎
6.2. on general categorical representations
In this section we prove that is a perverse equivalence on an arbitrary categorical representation. Fix such that is nonzero. For ease of notation, set and .
Consider a filtration by Serre subcategories
which can be either the Jordan-Hölder filtration (Theorem 3.6) or the isotypic filtration (Remark 3.9). So for every , is a subrepresentation of , and is either a simple categorification or an isotypic one. Define by requiring that is a representation of type .
Construct filtrations of and by . These are Serre subcategories of and respectively. Let be given by .
Theorem 6.8.
is a perverse equivalence with respect to for either the Jordan-Hölder or the isotypic filtration.
Proof.
Since is a categorical subrepresentation, the terms of the functor leave invariant, and in particular take objects in to . By Lemma 4.8 this implies that .
Now, is a simple or isotypic categorification (of type ). By Corollary 6.7, restricts to an abelian equivalence , i.e. the functor
is a t-exact equivalence. This shows that is a perverse equivalence with respect to . ∎
Remark 6.9.
The 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 be an integrable representation of . Recall that Lusztig has defined a braid group action on [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 , recovering the recently discovered action by generalised Schützenberger involutions.
7.1. Cactus groups
The cactus group associated to the Dynkin diagram has several incarnations. Geometrically, it appears as the fundamental group of a space associated to the Cartan subalgebra of . Namely, let denote the regular elements of . The cactus group is the -equivariant fundamental group of the real locus of the de Concini-Procesi wonderful compactification of (see [DJS03], [HKRW, Section 2] for further details):
There is a surjective map , 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 curves with marked points [HK06].
The cactus group has a presentation using Dynkin diagram combinatorics. For any subdiagram , recall that is the diagram automorphism induced by the longest element .
Definition 7.1.
The cactus group is generated by , where is a connected subdiagram, subject to the following relations:
-
(i)
for all ,
-
(ii)
, if and there are no edges connecting any to any , and
-
(iii)
if .
The surjective map mentioned above is given by . We are interested in the cactus group in connection to the theory of crystals.
A -crystal is called normal if it is isomorphic to a disjoint union for some collection of highest weights . The category of normal -crystals has the structure of a coboundary category analogous to the braided tensor category structure on -representations. It is realized through an “external” cactus group action of on -tensor products of -crystals, described by Henriques and Kamnitzer [HK06, Theorems 6,7].
We are interested in the “internal” cactus group action of on any -crystal . Both the internal and external actions rely on the following combinatorially defined maps, which are generalisations of the partial Schützenberger involutions in type .
Definition 7.2.
The generalised Schützenberger involution on is the set map defined uniquely by the following properties. For all and :
-
(1)
,
-
(2)
,
-
(3)
.
Note that from (1), swaps the lowest and highest weight element of , and the rest of its behavior is then determined from (2) and (3). The generalised Schützenberger involution on is the set map which acts as on each irreducible component .
Note that (1) implies that 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 , denote by the crystal restricted to the subdiagram . We denote the corresponding Schützenberger involution by .
Theorem 7.3.
([HKRW, Theorem 5.19]) For any -crystal , the assignment defines a (set-theoretic) action of on .
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 in Proposition 3.4.
Let be a categorical representation of . For any weight , by Theorem 6.8 is a perverse equivalence, and hence it induces a bijection . By varying we obtain a bijection
Now let be a connected subdiagram, and let be the corresponding subalgebra. By restriction, is also a categorical representation of , and hence by the above discussion we also obtain a bijection
We will prove that this family of bijections defines an action of the cactus group . First we need the following technical result. The important point here is just that there exists an integer such that .
Lemma 7.4.
Let , and let , where . Set and define by
Then on we have
(7.5) |
Proof.
We will prove the claim by induction on . For , we have that so . By [KT, Equation (7)] , which, combined with Lemma 6.2, implies that . Since , this proves the base case.
Now choose any and suppose (7.5) holds for any weight such that . Consider . First note that by [KT, Lemma 5.4] we have that
(7.6) |
Setting , by induction we have
One checks easily that , proving that . Since this holds for any vector of the form in , this completes the inductive step. ∎
Theorem 7.7.
The assignment defines an action of on .
Proof.
We need to show that the bijections satisfy the cactus group relations.
Relation (i): Without loss of generality we may assume . Fix a weight . Our aim is to show that
(7.8) |
Since the filtration of which we use in the perversity data of , agrees with the filtration of which we use in the perversity data of , by Lemma 4.5 the composition is a perverse autoequivalence of .
The functor is also a perverse autoequivalence of . 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 is the identity, this proves relation (i).
Relation (ii): Let be disjoint subdiagrams with no connecting edges. Our aim is to show that
(7.9) |
We prove a slightly more general statement, namely that for any categorical representation of , relation (7.9) holds.
Note that are isomorphic perverse equivalences, so they induce the same bijections. It remains to show that
(7.10) |
Consider first the case when categorifies an simple representation of . A minimal categorification of is of the form , where is a highest weight for and is a highest weight for . Hence by Lemma 3.5, a simple categorification of is of the form for some abelian category .
This implies that as a categorical representation of (respectively ), categorifies an isotypic representation. By Corollary 6.7 and 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):
where for every , is a subrepresentation of , and is a simple categorification of . Equation (7.10) now follows by an easy induction on . Indeed the base case when holds by the paragraph above, and the inductive step by Lemma 4.6.
Relation (iii): We need to show that , where . Again, we may assume that . Note that we have an isomorphism at the level of functors:
which lifts the corresponding relation in . Since this is an isomorphism of perverse equivalences, they must induce the same bijections by Lemma 4.2.
It remains to show that
(7.11) |
When is a simple categorification, by Corollary 6.7 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 be a categorical representation of , and consider the -crystal . There are two actions of the cactus group on , 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 agree.
Proof.
It suffices to show that . First, suppose is a simple categorification of type . In this case , and is determined by:
so we need to show that satisfies these properties as well.
The first is an immediate consequence of Corollary 6.7. To show that satisfies the second property, fix and . We set , and write .
Consider the following diagram:
(7.13) |
By Proposition 5.9 this diagram commutes (note that we shifted both sides of the equation by ). By Theorem 6.4 both horizontal arrows are in fact t-exact equivalences so this restricts to a diagram of abelian categories:
(7.14) |
Let be a simple object, and let . Note that is simple and . By the above diagram we have an isomorphism
Now, is the unique simple subobject. On the other hand, since is an abelian equivalence, is a simple subobject. Therefore
Since the equivalence class of the left hand side is , this shows that satisfies the second defining property, and hence the two cactus group actions agree in the case of a simple categorification.
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 -fold tensor product of the standard representation of . 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 . We can model this as follows:
where [CR, Example 5.17]. Here is the weight category, and is the zero weight category. The arrows describe the functors (we omit the higher structure).
Consider the Rickard complex . For , we have:
where the differential is given by the action map, and is in cohomological degree . It’s an exercise to verify that is quasi-isomorphic to , where is the twist of by the automorphism of given by . This shows that is the t-exact equivalence . ∎
Example 8.2.
More generally, one can consider the minimal categorification of the adjoint representation of a simple simply-laced Lie algebra . This was studied by Khovanov and Huerfano in [HuerKh], who used zigzag algebras to model this category.
For a weight of the adjoint representation of , the weight category is taken to be as long as . However, the zero weight category is more interesting: , where is the zigzag algebra associated to the Dynkin diagram of (cf. [HuerKh] for the precise definition).
The isomorphism classes of indecomposable projective left -modules and the isomorphism classes of indecomposable projective right -modules are both indexed by . Tensoring with these modules defines functors
The functors are defined analogously and are biadjoint to the .
The Rickard complex is given by tensoring with the complex
of -bimodules, where the differential is given by the multiplication in , and sits in cohomological degree zero. These functors are autoequivalences of the derived category .
By Theorem 6.4 is t-exact, where is the Coxeter number of . This auto-equivalence can be explicitly described as follows: the automorphism of the Dynkin diagram induces an automorphism of . Then we claim that is the abelian autoequivalence of defined by twisting with . Indeed, since 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 on the Grothendieck group of as follows: is isomorphic to the Cartain subalgebra by mapping ( is the simple head of ) to the simple root vector . And on the root vectors we have that ∎
Example 8.3.
Let and consider the -fold tensor power of the standard representation . Categorifications of have been well-studied, and a model for this categorical representation can be constructed using the BGG category of [Sussan2007, MS09]. In this model, the principal block appears as the zero weight category of , and the Rickard complexes acting on are the well-known shuffling functors.
By Theorem 6.8 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 is proved using the theory of W-algebras). We can interpret the filtration of arising from our perspective concretely using the Robinson-Schensted correspondence .
Recall that the simple objects in are the irreducible highest weight representations , where has highest weight ( is the half-sum of positive roots of ).
We view a partition of simultaneously as a dominant integral weight for , and as an index for the irreducible Specht module of the symmetric group . Let denote the set of standard Young tableau of shape , and let . Recall that .
Choose an ordering of the partitions of , , so that if in the dominance order, then . Note that the dominance order on partitions of is equivalent to the positive root ordering on partitions (thought of as weights for ). By Remark 3.9 there is an isotypic filtration on , where is an isotypic categorification of type . Then is a perverse equivalence with respect to the filtration and the perversity function .
We would like now to define the categories more explicitly using the Robinson-Schensted correspondence, which recall is a bijection [Sagan]:
We will use a crystal analogue of classical Schur-Weyl duality. This is given by a crystal isomorphism:
(8.4) |
Now, recall that for a partition of , the underlying set of the crystal can be chosen to be the set of semistandard Young tableaux of shape with entries , and the weight zero subset of is precisely . The essential point is that the isomorphism (8.4) can be chosen so that it restricts to the map on the elements of weight zero ([Shimozono, Theorem 3.5]). (Note that the elements of weight zero in are naturally identified with permutations of .)
This shows that as an element of the crystal , is in a connected component whose highest weight is the shape of (or equivalently ). Therefore, following Remark 3.10, we can construct the isotypic filtration be defining to be the Serre subcategory of generated by such that the shape of is among .
∎
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 . We continue with the notation in Example 8.3 and view a partition simultaneously as a dominant integral highest weight for , and as an index of the Specht module . Recall that by Schur-Weyl duality, is isomorphic to . The Kazhdan-Lusztig basis of the Hecke algebra naturally descends to a basis of , which we denote indexed by . For further details we recommend the exposition in [Rhoad].
Consider the minimal categorification , and in particular its zero weight category . For convenience, we forget the grading and work in the non-quantum setting. The simple objects are indexed by , and hence a perverse equivalence induces a bijection .
The bijection studied in the previous section specialises to the well-known Schützenberger involution on standard Young tableau, otherwise known as the “evacuation operator” [Sagan]. Indeed, by Theorem 7.12 recovers the cactus group action on the crystal 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 is another important function in algebraic combinatorics, which is closely connected to the RSK correspondence and related ideas such as jeu de taquin. Letting , we can express promotion in terms of the Schützenberger involution: . 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 be the long cycle. Then
is a perverse equivalence whose associated bijection is the promotion operator: .
Proof.
We can also use this set-up to extract information about the action of on the Kazhdan-Lusztig basis of . This is based on the following elementary lemma:
Lemma 8.6.
Let and suppose is t-exact up to shift. Then for any , , where .
Proof.
Since is t-exact up to shift, we have:
The result now follows since the isomorphism , is -equivariant, and by [BauqWeyl, Proposition 10], the action of the braid group on factors through . ∎
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 and , we have that .
Similarly we can prove a result of Rhoades regarding the action of the long cycle on the Kazhdan-Lusztig basis. Note that in the statement below the significance of being rectangular is that the restriction of to in this case remains irreducible.
Proposition 8.8.
(cf. [Rhoad, Proposition 3.5]) Let be a rectangular partition. Then for any , the action of the long cycle on the Kazhdan-Lusztig basis element recovers the promotion operator:
Proof.
As above, set . Recall that, since is a simple categorification, we know is t-exact up to shift by Theorem 6.4. However, is no longer simple, so a priori we only know that is perverse, but not necessarily t-exact up to shift. We first prove that is indeed t-exact up to shift.
Consider the set of functors which are monomials in the Chevalley functors indexed by :
Let be the abelian category generated by , that is, is the category closed under subobjects and quotients of objects of the form . Since the Chevalley functors are exact, it is easy to see that is a categorical representation of .
Let be the (unique) partition obtained from by removing a box. We claim that is a categorification of . Note that this is a categorification of the fact that . Indeed, it is clear that contains . On the other hand, if then is in the root lattice of , and is the unique constituent of whose weights are in the root lattice.