aligntableaux=center, boxsize=0.5cm
On Schützenberger modules of the cactus group
Abstract.
The cactus group acts on the set of standard Young tableau of a given shape by (partial) Schützenberger involutions. It is natural to extend this action to the corresponding Specht module by identifying standard Young tableau with the Kazhdan-Lusztig basis. We term these representations of the cactus group “Schützenberger modules”, denoted , and in this paper we investigate their decomposition into irreducible components. We prove that when is a hook shape, the cactus group action on factors through and the resulting multiplicities are given by Kostka coefficients. Our proof relies on results of Berenstein and Kirillov and Chmutov, Glick, and Pylyavskyy.
1. Introduction
1.1. Background
Let be a reductive complex Lie algebra. In Kashiwara’s theory of -crystals, the cactus group plays a role analogous to that of the braid group in representations of the quantum group . Indeed just as the -strand braid group acts on -fold tensor products of representations of (resulting in a braided category), the cactus group acts on -fold tensor products of crystals (resulting in a coboundary category) [HK06]. And just as the type braid group acts on any integrable representation of , the type cactus group acts on any -crystal [HKRW]. This latter “internal” action is our focus.
Before describing our results, we highlight the appearance of the internal cactus group action in several recent theorems.
Losev construced an action of the cactus group on the Weyl group of , and showed that it interacts nicely with Kazhdan-Lusztig cells [LosCacti]. For , this recovers the external action of the cactus group corresponding to the zero weight space of the -fold tensor product of the standard representation.
Losev constructs his action by showing that certain wall-crossing functors are perverse equivalences in the sense of Chuang and Rouquier [CRperv]. This was recently extended in work of Halacheva, Losev, Licata and the second author [HLLY]. We show that for any categorical representation of , the Rickard complexes corresponding to the half-twist are perverse equivalences. From this we obtain the internal cactus group action on any integrable representation. In [GY, GY2] Gossow and the second author explain how to recover this cactus group action in type A directly from the representation without appealing to categorical techniques.
In a different direction, Halacheva, Kamnitzer, Rybnikov and Weekes study the action of Gaudin algebras on tensor product multiplicity spaces [HKRW]. Their main tool is a crystal structure on eigenvectors for shift of argument subalgebras, which are a family of commutative algebras acting on irreducible representations. In particular they show that the internal action of the cactus group controls the monodromy of these eigenvectors.
In this paper we initiate a study of representations of the cactus group which arise as permutation modules from the internal action on crystals. We’ll now describe our work in detail.
1.2. Our work
We specialise to the case of . The corresponding cactus group is an infinite group generated by , for subintervals , subject to the relations in Definition 2.2. It is isomorphic to the orbifold fundamental group of the real locus of the wonderful compactification of , where is the regular locus in the reflection representation of the symmetric group [DJS03].
Let be a partition of . The cactus group acts on , the set of standard Young tableau of shape , where the generator acts by a partial Schützenberger involution. Letting denote the Specht module of , it is natural to view the -action on as an action on the Kazhdan-Lusztig basis of . We thus obtain a -action on , which we term the “Schützenberger module”, and denote .
Our main problem, which to our knowledge has not been studied, is the following:
Problem 1.1.
Determine the irreducible constituents of .
An obvious obstruction to solving this problem is that we do not have a classification of the finite dimensional irreducible representation of (it is of wild representation type). Nevertheless, there are naturally occurring families of irreducible representations of obtained by inflation from symmetric groups.
Indeed, there is a natural homomorphism , and this can be generalised to a surjective map , for (cf. Lemma 2.3). For we let be the irreducible -module on obtained via pullback by .
Our main theorem solves Problem 1.1 in the case when is a hook partition. To a hook partition, we associate a composition of given by .
Theorem 1.2.
Let be a hook partition. We have an isomorphism of -modules
(1.3) |
where are the Kostka numbers, unless .
Note that in the outlying case, is simply the two-dimensional module with basis elements interchanged by ( and act trivially).
Our main tools for proving the theorem come from work of Berenstein and Kirillov [BeKi] and Chmutov, Glick, and Pylyavskyy [CGP]. The former define a group of symmetries of Gelfand-Tsetlin patterns (i.e. semistandard Young tableau), which the latter show is a quotient of the cactus group. These results allow us to show that in the case of a hook shape, the -action on factors through , and to identify resulting permutation module.
Acknowledgement 1.4.
This work was undertaken for an Honours Thesis by the first author at the University of Sydney in 2020, under the supervision of the second author. The second author is partially supported by the ARC grant DP180102563.
2. Background
2.1. Young tableau
In this section we briefly recall the basic combinatorics of Young tableau. For more details see [Sagan]. Let . A partition of , written , is a weakly decreasing sequence of nonnegative integers that sum to :
If we drop the weakly decreasing condition, we get the notion of a composition of .
We use Young diagrams to represent partitions and compositions. A Young diagram for a composition is a finite collection of cells, arranged in left-justified rows, where the -th row length is the -th entry of .
Let . A Young tableau of shape is a filling of the corresponding Young diagram with positive integers. For example, here is a Young diagram and tableau of shape :
The Young tableau is semistandard (respectively standard) if the entries are weakly increasing (respectively strictly increasing) along rows, and strictly increasing down columns. The content of a tableau of shape is the composition of , , where is the number of ’s appearing in .
Given and , we let denote the set of semistandard Young tableau of shape and cells filled in with numbers . We let denote the set of standard Young tableau of shape and cells filled in with the numbers .
The Kostka number is defined equivalently as: the number of of shape and content , the dimension of the weight space in the irreducible representation of of highest weight , or as the multiplicity of in the permutation module (Equation (2.1)).
Given partitions we write if for every . Let be two partitions such that . The skew-diagram of shape is given by removing the boxes of in . A skew tableau is a labelling of these boxes with positive integers. Here is a skew diagram and tableau of shape for and :
Similar to above, semistandard tableau on skew shapes are skew-tableau with weakly increasing labels along the rows and strictly increasing labels down the columns. Standard tableau on skew shapes are semistandard tableau whose entries strictly increase along the rows.
Let be a composition of . Let be two diagrams of shape with entries . We write if and are row-equivalent, i.e. they have the same entries in each row. An equivalence class for this relation is a -tabloid. We let be the set of -tabloids. A tabloid can be pictured in a manner similar to tableau. For example, here is a -tabloid:
representing the equivalence class of the diagram with entries in the first row and in the second.
Set . Given a tableau we let be the tableau obtain by deleting all cells with entries not in .
2.2. The symmetric group
Let . Let denote the symmetric group on . Let denote the simple transposition swapping and . Finite dimensional irreducible complex representations of are indexed by partitions of . The irreducible representation corresponding to is the Specht module . For instance, is the trivial representation, is the sign representation, and is the standard representation.
The Specht module has a remarkable basis indexed by called the Kazhdan-Lusztig basis, which we denote . To construct this basis one needs to pass to the Iwahori-Hecke algebra associated to . Kazhdan and Lusztig constructed a canonical basis of the Hecke algebra, which gives rise also to bases of its cell modules. In type A, these cell modules are the irreducible Specht modules and the specialisation leads to the basis . For more details, see e.g. [KL79, GaMc, Rhoad].
Given a composition of , let be the corresponding parabolic subgroup. Let denote the induced module from the trivial representation. The module has a basis indexed by the set of row tabloids , where the action is given by permutation of entries. Kostka numbers encode the decomposition of into Specht modules:
(2.1) |
2.3. The cactus group
Given an interval , let be the subgroup of permutations which fix . In the notation of the previous section, is the parabolic subgroup , where . Let be the longest element, that is “flips” the interval via for .
Definition 2.2.
Let be an integer. The cactus group is generated by generators , indexed by the intervals , subject to the following relations: , if and, if .
The cactus group is an infinite group, which has its origins in (a) the study of symmetry groups of universal covers of blow-ups of projective hyperplane arrangements [DJS03], and (b) the study of commutators in the category of crystals for a semisimple Lie algebra [HK06].
Note that there is also a slightly different presentation of the cactus group that’s often used, where generators are indexed by subdiagrams of the Dynkin diagram of a semisimple Lie algebra [HKRW]. The cactus group defined above corresponds to type .
Symmetric groups are naturally quotients of the cactus group. Indeed, we have a map , which is a surjective group homomorphism since the defining relations of are satisfied also by the elements . This map can be generalised as follows
Lemma 2.3.
For any the assignment:
defines a surjective group homomorphism .
Proof.
The third defining relation of is only non-obvious relation to check. Suppose we have intervals . We need to show that .
If then so the equation holds. Otherwise, we have that . Let . Then a quick calculation shows that , which proves the desired relation. ∎
By inflation we obtain irreducible representations of on , for and , which we denote .
Remark 2.4.
It’s possible to generalise the maps above further. Given nonnegative numbers such that , we have a map given by
It’s straightforward to check that this satisfies the defining relations of . We recover defined in Lemma 2.3 as .
2.4. Operations on Young tableau
In order to construct the Schützenberger modules, we need to first recall some operations on Young tableau. For more details see [Sagan].
2.4.1. Jeu De Taquin
The Jeu de Taquin is a map taking a semistandard skew tableau to a rectified semistandard tableau, which we recall now. Let . Call a removable box of a movable box of . Then is defined as follows:
-
(1)
Choose a movable box of . Move this box with the following rules:
-
(a)
If it is adjacent to a box to its east and south, let them be and respectively.
If , then swap with
Otherwise, swap with
-
(b)
If it is adjacent to exactly one box to its east or south, swap it with that box.
-
(c)
Repeat this process until it is not adjacent to any boxes to its east or south.
-
(a)
-
(2)
Repeat this process with another movable box until there are no movable boxes left.
For example,
The rectification of a skew semistandard tableau via is independent of the choice of the removable boxes at each iteration.
2.4.2. Promotion
The promotion operation is a map defined as follows.
-
(1)
Turn every box labelled 1 to a dummy box.
-
(2)
Apply to the dummy boxes.
-
(3)
Reduce every non-dummy box’s label by 1.
-
(4)
Relabel the dummy boxes to
2.4.3. Schützenberger Involution
The Schützenberger Involution is a map defined by
Where , while leaving constant.
The Schützenberger involution can be shown to be an involution. Moreover and are of the same shape, and if then .
Definition 2.5.
Let . For , define the partial Schützenberger involution
to be the Schützenberger involution on (where the relabelling is ), while leaving constant. Let for .
Proposition 2.6.
[BeKi] The operators satisfy the following relations:
-
(1)
If , then .
-
(2)
For , we have .
Define a map by for an interval .
Proposition 2.7.
[HKRW] The cactus group acts on the set via , and is an invariant subset.
Proof.
First, we have
as we have established that the Schützenberger involution is an involution. Then we have in general
and the last two relations follow from Proposition 2.6. This shows that acts on .
Note that if and only if . Since as well, we have . Thus and the claim follows. ∎
Remark 2.8.
The set naturally carries a -crystal structure, isomorphic to the crystal of the irreducible representation of of highest weight [HoKa]. The subset is the weight zero elements of the crystal. The -action described in the above proposition coincides with the internal cactus group action on this crystal. For more details see [HKRW], where the internal cactus group action is constructed for any semisimple Lie algebra.
3. The Schützenberger modules
3.1. Definition and preliminary results
Let . Recall the Kazhdan-Lusztig basis of . We define a action on using (partial) Schützenberger involutions:
We term the resulting representation the Schützenberger module of , and denote it by .
Let and define the -module by the decomposition . We begin with some preliminary observations about .
Proposition 3.1.
Let and set . Then is an irreducible -module.
Proof.
It suffices to show that is surjective. We proceed by induction on . The base case is trivial. Define to be the standard tableau with the box in the second row. Then
Consider the restriction of to , where we regard as the subgroup generated by . Notice that the image of does not change the position of the box. Thus, fixes .
On the other hand, if , we have a bijection between and given by appending the box to the end of the first row for each . This bijection commutes with the action as disregards . By the induction hypothesis,
As is generated by and ,
as . (Applying the first promotion step shows that the box is in the first row for . In fact, .) ∎
In general, is not irreducible. Indeed, let be the dual map, where the tableau is reflected by the diagonal from northwest to southeast. Here, is the dual shape of .
The maps and commute on standard Young tableaux (but not in general). It follows that commutes with the promotion map, and hence the Schützenberger involution. Thus commutes with the action on standard Young tableau.
We therefore obtain an isomorphism . In particular, for a self-dual shape , we have a non-trivial automorphism . As is an involution, we have an eigenspace decomposition corresponding to the eigenvalues of :
Notice that , and hence there exists a submodule such that . Thus
and hence . Consequently is not irreducible for self-dual .
3.2. The Berenstein-Kirillov group
In order to undertake a more detailed study of Schützenberger modules we will utilise Gelfand-Tsetlin patterns and their symmetries, as developed by Berenstein and Kirillov.
Let and . A Gelfand-Tsetlin pattern with rows and top row is a triangular arrangement of nonnegative integers such that and the top row is . Denote the set of such patterns as .
Define a map as follows. Let . For , as is semistandard, cannot have more than rows. Let the shape of be . As the notation suggests, set equal to . The following is immediate:
Proposition 3.2.
is a bijection.
Bereinstein and Kirillov defined operators acting on as follows. Let . Define for by , where
For the edge cases we let and .
Proposition 3.3.
[BeKi] The operators satisfy the following relations.
where .
It is conjectured by Berenstein and Kirillov that these generate all relations among the operators .
Definition 3.4.
The Berenstein-Kirillov group is the group generated by with relations as in Proposition 3.3.
By transport of structure via , we have an action of on . Following [CGP], we will describe this explicitly. Let . Recall that is a disjoint union of rectangles and strips of the form
We say a strip is of type if it contains many -boxes and many -boxes.
Define such that acts on by replacing each strip of type with a strip of type , and leaving the rectangles unchanged:
In the example above, strips of type were swapped with strips of type .
We define to be the tableau obtained by replacing with , and leaving the other boxes unchanged.
Lemma 3.5.
Let and . Recall
Then the strip of in row is of type starting at column .
Proof.
Recall that corresponds to the number of boxes in row of labelled from .
Assume there is no rectangles with its first row in row . This means that every box above in the -th row has a label less than or equal to . This is precisely when , and in this case, the strip indeed starts at column . On the other hand, if there is such a rectangle, then we have . This rectangle ends at column , hence the strip starts at column after it as claimed.
The proof for is similar. The strip is indeed of type as the boxes labelled in row span columns to by the correspondence given by . ∎
Proposition 3.6.
For we have that .
Proof.
It suffices to show that . Notice that does not affect the shape of for all and . Furthermore, in , each strip in row spans column to column and is of type . Thus in , this strip is replaced by a strip of type .
Let and set . Let
Recall that is an operation on which affects only the -th row, hence for all . Thus, as follows:
Similarly, we have . Thus the strip at row for also starts and ends at the same column, but is of type . However, as ,
Thus . ∎
This proposition implies that for standard Young tableaux, the action of is particularly easy to describe:
Corollary 3.7.
Let for . Then swaps the two boxes and if they are not adjacent, otherwise .
Proof.
As is standard, consists of two boxes, which can be non-adjacent, horizontally adjacent, or vertically adjacent:
The non-adjacent case is essentially two disjoint strips of type and each. Thus swaps the two boxes. The vertically adjacent case has no strips, while the horizontally adjacent case is of type , which stays constant under . ∎
Consider the elements defined by
Although we won’t use the following theorem of Berenstein and Kirillov, we include a (new) proof since it provides important context for what follows.
Theorem 3.8.
[BeKi, Section 2] The action of and on are equivalent to and , the promotion and Schützenberger involution operations on .
Proof.
Let . We first prove that by induction on .
For the base case, and act by identity on and . Thus, .
For the inductive case, notice that the step of and are identical until the dummy box becomes adjacent to the box.
Case 1: If the dummy box is never adjacent to in the step of ,
Then the step of and are identical, and they only differ in the relabelling step. For , the dummy box is labelled as , and the box is kept constant, while every other box’s label is reduced by 1. On the other hand, for , the dummy box is labelled as , while every other box, including , has its label reduced by 1. Furthermore, in both cases, and are not adjacent due to our assumption. Thus, acts by swapping and on and we have .
Case 2: If the dummy box comes adjacent to in the step of ,
Then the step of must have an extra step of swapping the dummy box with . Then after the relabelling steps of and , we have . Furthermore, by assumption, we have and adjacent, thus acts by identity on .
Thus we have overall . Hence by induction, . Then we have by definition
and the result follows. ∎
We now recall a theorem of Chmutov, Glick, and Pylyavskyy, which identifies the Berenstein-Kirillov group with a quotient of the cactus group.
Definition 3.9.
The reduced cactus group is the quotient of by the relations
(C3) |
where for .
Remark 3.10.
Since and the relations defining the reduced cactus group are conjugates of a single relation, that is, .
The following is the main result of [CGP].
Theorem 3.11.
There is a group isomorphism given by
Corollary 3.12.
Let . The action of on factors through .
Proof.
Let . Notice that is a non-identity element. Using we have
By Theorem 3.8, the action of is equivalent to the action of . However,
It follows that acts by identity on , and the action of on factors through the projection map . ∎
Remark 3.13.
The corollary is a special case of a more general result of Kashiwara [Kash94, Theorem 7.2.2], which implies that the internal cactus group action on any normal -crystal factors over the reduced cactus group of type (see also [HKRW, Remark 5.21]).
3.3. The case of a hook shape
In this section we will prove our main result, which describes the Schützenberger modules in the case when is a hook shape, i.e. for some and some number of s. For this we make crucial use of the connection between the cactus group and the Berenstein-Kirillov group explained in Theorem 3.11.
Recall that acts on for . This gives rise to a representation
Recall that acts by swapping and if they are not adjacent, and otherwise does nothing. As and are always adjacent for standard tableaux, always acts by identity, and thus .
The remaining generators satisfy the relations of . In particular,
Assume for the purposes of discussion that for , we have
() |
This would give a surjective group homomorphism
Since (Theorem 3.11), and the action of on factors through (Corollary 3.12), this will allow us to use the representation theory of to study . The following lemma describes when this approach is feasible.
Lemma 3.14.
Let . Then relation holds for all if and only if or is a hook shape.
Proof.
As and act on depending on how , , and are adjacent, we consider the ways the three boxes can be adjacent.
-
Case 1:
If the three boxes are all non-adjacent, is true. For example:
-
Case 2:
If two boxes are adjacent and one is not adjacent to either, then is true. For example:
-
Case 3:
If all three are adjacent in a single row or single column, then is true. In this case, and are always adjacent, so always acts by identity. The same is true for and , so also always acts by identity.
-
Case 4:
If all three are adjacent in the following shape, then is not true.
In this case, and are always adjacent while and are always not adjacent. Thus acts by identity while acts by swapping and . Hence for with this formation,
Hence if is a shape that does not allow the Case 4 configuration for (since we ignore ), then will hold true. This is exactly when or is a hook shape. ∎
Remark 3.15.
In general for all shapes , we have . To see that we don’t necessarily have , consider and the following tableau
A quick calculation shows:
Theorem 3.16.
For a hook shape not of the form or , the map is an isomorphism.
Proof.
We have already shown that is surjective. Notice that for , has two distinct non-identity elements, namely and . These are nontrivial because there is a such that
and indeed are all distinct.
Assume for contradiction . If , the only other normal subgroups of are , the alternating group, and . In either case, we have , hence
which gives a contradiction. Thus the kernel must be the trivial group for .
For , the only possible hook shapes are . For the case, using the notation from the proof of Proposition 3.1, we have that interchanges and for . Thus is isomorphic the subgroup of generated by the simple transposition matrices, which is clearly isomorphic to . The case is dual to the case.
Let us examine the case for . There are six tableaux in .
Hence we can view as elements in acting on the subscript: . Then we have:
The subgroup generated by these elements has more than six elements. Since every nontrivial normal subgroup of has index at most 6 ( where is the Klein four group), we conclude as above that the map is indeed injective. ∎
Corollary 3.17.
Let be a hook partition not of the form or . Then the representation factors over as follows:
Proof.
Definition 3.18.
Let be a hook shape. The boxes in the first row (excluding the first box) are the arm of , and the boxes in the first column (excluding the first box) are the leg of . We let be the two-part composition of formed by the arm and leg of .
In the example below, has arm length and leg length and .
Proposition 3.19.
Let be a hook shape. Then the representation is isomorphic to the permutation module .
Proof.
Set . We have shown that by the dual map, so we can assume without loss of generality that .
Since and are both permutation modules of , it suffices to prove that there exists a bijection between the standard bases of each module that commutes with the action of .
Define the operation given by the following illustration:
Notice that we lose the hinge, i.e. , so the entries in are now from . Thus we subtract 1 from each label, and define a map given by .
We first show that is a bijection. Let . Any of cardinality uniquely determines , where is the set of numbers in the arm of . Similarly for any of cardinality determines , where is the set of numbers in the first row. The bijection is then .
Next we show that commutes with . Recall that acts via on , which swaps and if the two boxes are not adjacent in the tableau, and if they are it acts by the identity.
Notice that for , and , are adjacent in if and only if and are in the same row in . Thus, acts on trivially if and only if it acts trivially on . Otherwise, the boxes swap. In , the boxes swap from the arm to the leg and vice versa. In , the boxes swap rows. Since the arm maps to the first row and the leg maps to the second row, this shows that commutes with every transposition. ∎
We are now ready to prove our main result.
Proof of Theorem 1.2.
In the setting of the theorem, is a hook partition not equal to . Consider first the case when is not of the form . By the above proposition and Equation (2.1), there is an isomorphism of -modules
By Corollary 3.17 this implies the isomorphism of Equation (1.3).
The remaining cases are easily dealt with by direct computation. If or ) the Kostka number is zero unless , in which case it is equal to . Thus both sides of (1.3) are isomorphic to the trivial -module. ∎