On the action of the long cycle on the Kazhdan-Lusztig basis
Abstract
The complex irreducible representations of the symmetric group carry an important canonical basis called the Kazhdan-Lusztig basis. Although it is difficult to express how general permutations act on this basis, some distinguished permutations have beautiful descriptions. In 2010 Rhoades showed that the long cycle acts by the jeu-de-taquin promotion operator in the case when the irreducible representation is indexed by a rectangular partition. We prove a generalisation of this theorem in two directions: on the one hand we lift the restriction on the shape of the partition, and on the other hand we enlarge the result to the collection of all separable permutations.
keywords:
Specht modules, Kazhdan-Lusztig basis, promotion, evacuation1 Main Results
The complex irreducible representations, aka Specht modules, of the symmetric group are indexed by partitions of : for such a partition we denote by the associated representation. These carry a canonical basis called the Kazhdan-Lusztig basis , which is indexed by the set of standard Young tableaux of shape .
Although the Kazhdan-Lusztig basis enjoys many wonderful properties, it is quite complicated to express its transformation under the action of an arbitrary permutation. Nevertheless, owing to a close connection with the RSK correspondence, it is possible for certain distinguished elements.
In the mid 1990s Berenstein-Zelevinsky [1] and Stembridge [10] showed independently that the long element of acts on the Kazhdan-Lusztig basis (up to sign) by Schützenberger’s evacuation operator. More recently, Rhoades proved in the case of rectangular partitions that the long cycle acts on the Kazhdan-Lusztig basis (up to sign) by the jeu-de-taquin promotion operator, [8]. This was the essential new ingredient needed in Rhoades’ cyclic sieving phenomenon regarding the action of promotion on rectangular standard Young tableaux.
Our first main result is a generalisation of Rhoades’ Theorem to arbitrary partitions. Fix a partition and number its removable boxes moving downward. Define the index of to be the said number of the removable box containing , and denote it . Choose an ordering of the Kazhdan-Lusztig basis which is weakly increasing along the index. For let be the matrix of in this ordered basis.
Theorem 1.1.
Let be an arbitrary partition. Let be the QR decomposition of into the product of an orthogonal matrix and an upper triangular matrix . Then is a generalised permutation matrix for with entries , i.e. for any we have
The sign appearing above depends only on the index of .
Equivalently, this theorem states that acts on the Kazhdan-Lusztig basis by promotion on the leading term, i.e.
where the sum is over with index strictly smaller than the index of , and . In the case of a rectangular partition, the sum is empty and we recover Rhoades’ Theorem.
The key property of rectangular partitions that Rhoades leverages is that the restriction of the corresponding Specht module to remains irreducible, and the Kazhdan-Lusztig basis is unchanged if the space is regarded as an -module or an -module. It is therefore not surprising that the key step in our argument involves the interaction between the restriction functor and the Kazhdan-Lusztig basis in a more general setting, which may be interesting in its own right.
For define the subspace of . This gives rise to a filtration , where is the number of removable boxes of .
Theorem 1.2.
Each subspace is -invariant, and for every , there is an isomorphism of -modules , where is the partition obtained from by deleting the removable box.
Note that this theorem provides a new proof that the branching law for symmetric groups is multiplicity-free, as
Finally, we mention a generalisation of Theorem 1.1 to the class of all separable permutations, which includes the long element and the long cycle. By definition, separable permutations are those that can be obtained from by repeated applications of direct sum and skew sum. Alternatively, these are the permutations that avoid the patterns and [6].
Fix an arbitrary partition. In Section 5, we describe briefly how to associate to each separable permutation a bijection of such that an analogous result to Theorem 1.1 holds. That is, if is the matrix giving the action of on some given ordering of the KL basis, we have:
Theorem 1.3.
Let be any separable permutation, and let be the QR decomposition. Then is a (generalised) permutation matrix of .
In specific cases, we can prove this combinatorially but the general case requires results from categorical representation theory and in particular the action of braid groups on triangulated categories [4]. In our forthcoming work we will describe this connection, and use this to also derive variations of these theorems for arbitrary simply-laced semisimple Lie algebras.
2 Background
In this section we briefly recall some basic results we’ll need. We refer the reader to [9] for more details on notions from algebraic combinatorics, and to [2] for the necessary background on Kazhdan-Lusztig theory.
2.1 Combinatorics
Let be a partition of , which we depict using its associated Young diagram. Recall that a box in is called a removable box if its deletion results in another Young diagram. We label the removable boxes of by starting from the top, and moving down the tableau. For example, the partition has three removable boxes, with labelling given by:
Write for the set of standard Young tableaux of shape . We define the index of , denoted , to be the label of the removable box containing as above. If and have the same index, then their largest box occupies the same position. When ordering according to index, we can break ties by comparing the index values of the tableaux after removing their largest boxes. Repeating this procedure gives the total index ordering on .
Recall that the descent set of is the set of in such that appears strictly below in .
The promotion map is defined as follows: replace the box with value by a dot, and move this dot north-west by repeatedly swapping it with the box above or to the left. If both boxes exist, swap with the larger. Once the box has reached the north-west corner, increment all values in the tableau and replace the dot with 1.
The evacuation map is defined as follows: denote every box in the tableau ‘fixed’ or ‘unfixed’, with all boxes initially starting unfixed. At each step, perform the inverse of the promotion map on the unfixed boxes, and fix the largest remaining box. Repeat until all boxes are fixed. In the example below fixed boxes are in boldface:
Both promotion and evacuation are bijections on , and evacuation is in fact an involution. It will also be useful to define the partial evacuation maps. For , define the involution by fixing all but the smallest boxes, and performing evacuation on the remaining unfixed tableau.
Lemma 2.1.
For every , we have .
The RSK correspondence [7] gives a bijection between and ordered pairs of tableaux of some shape . We will denote the RSK correspondence by .
2.2 Representation theory
The symmetric group is generated by the simple transpositions for , satisfying the braid relations and . Given an element , we write for the length of in these generators.
Given permutations , we write in the Bruhat order if can be obtained from by successively swapping two elements in the permutation’s one-line notation, increasing the length of the intermediate permutations at each swap. In particular, implies . Denote by the (left) descent set of , which consists of the indices between and such that .
Let be an indeterminate. The Hecke algebra (of type A) is an algebra over generated by satisfying the braid relations and the quadratic relation for all . As a -module, has a basis indexed by . The quadratic relation shows that each , and hence each , is invertible. The bar involution of the Hecke algebra is the -linear extension of the maps and .
Theorem 2.2 (Kazhdan-Lusztig Basis, [5, Theorem 1.1]).
There is a unique basis of indexed by , with elements of the form
such that:
-
•
is an integer polynomial, and is non-zero only if ,
-
•
for all ,
-
•
if then , and
-
•
for all .
The are the well-known Kazhdan-Lusztig (KL) polynomials. The degree bound gives us a statistic on pairs of permutations. For , we write for the coefficient of degree in . If is even or is incomparable with in the Bruhat order, this coefficient will be zero. We also define when , so is symmetric.
Under the specialisation , we have the relation , and so . Set , so that is a basis for . Define a binary relation on by setting if, for some , appears with nonzero coefficient in the expansion of . Taking the transitive closure gives a preorder on , called the left KL preorder. By its construction, the span of for a fixed will be a submodule of . The equivalence classes induced by are the left cells of , and we write if and belong to the same left cell.
For a left cell of , we write for the linear span of elements . The following defines an irreducible action of on [2, (6.4)]:
(2.3) |
Left cells in are determined by the -tableau in the RSK correspondence. That is, if and only if and for some and . Therefore we can uniquely associate each left cell of to a tableau with boxes, and write . Appropriately restricting the RSK correspondence gives a bijection between the basis elements and , by identifying with the -tableau of . Hence, we can reindex the basis of as . We can rewrite the action on the KL basis in terms of tableaux only:
-
1.
Let and suppose . Then if and only if .
-
2.
For and tableaux we have:
where we identify permutations with their images under the RSK correspondence.
From this we obtain that for tableaux , the respective representations and are equal (not just isomorphic) up to a reindexing of basis elements.
These results have a number of implications. First, we can define the value between tableaux by for any . Furthermore, since depends only on the shape of , we can write for the representation , where is any tableau of shape . Finally, we can write the conditions on the descent set of in (2.3) in terms of the -tableau of under RSK. We incorporate these results into the following theorem.
Theorem 2.4 ([2, Theorem 6.5.3]).
Let be a partition. The module is irreducible, and has a basis with the following action:
(2.5) |
The irreducible representations of are called Specht modules. The above theorem provides a canonical basis for Specht modules, which we refer to as the KL basis.
3 The Branching Rule via the KL Basis
In this section we prove Theorem 1.2. Fix , and recall the filtration of by the subrepresentations , each defined as the subspace spanned by with . We will first show that always remains invariant under .
Let be the column super-strict tableau of shape , where the values are placed column-by-column, reading left-to-right. We take as an example:
Preimages of the RSK correspondence when are easy to describe. We simply write the elements of the -tableau in column order, reading from bottom-to-top.
Lemma 3.1.
Fix and , where is the number of removable boxes of . Then is an -submodule of .
Proof.
If , there is nothing to prove. Otherwise, choose arbitrary tableau in with and , or equivalently, . Instead of working in , we work in the module . To this end, define the unique permutations with images and under RSK. We are required to show that does not appear in (see (2.5) for the action) when . We consider cases depending on the Bruhat comparibility of and .
(a) and are not Bruhat compatible. Then by definition.
(b) . We have a chain of swaps with . At some point, the swap moves to the right. But this will decrease the length, which is a contradiction.
(c) and . Suppose appears in the expansion of . Then and , so by [2, Proposition 5.1.9].111Björner [2] uses the convention of right descent sets, but the result for left descent sets is analogous.
(d) and . Again, suppose appears in the expansion of for some . Then , and , which implies . But occurs at different positions in the one-line notation of and , so we must have . This is again a contradiction, since . ∎
Recall that is the partition obtained by deleting the removable box from . There is a bijection which removes the largest box of each tableau . By taking its linear extension, we have an isomorphism of vector spaces for each . Our goal is to show that this is in fact an isomorphism of -modules. For this, we need to show that the deletion map preserves the function .
We again define a specific recording tableau and work with the permutations explicitly. For a partition with removable boxes and , define the tableau by the following procedure:
-
•
Create a tableau of shape , and place the value in removable box .
-
•
Suppose is placed in the row. For , place the value at the end of row .
-
•
Place the remaining values into the tableau column by column.
For example, if and we have:
The proof that the function is preserved under the deletion map was originally given by Rhoades in the case of rectangular partitions. The following Lemma appears implicitly in the proof:
Lemma 3.2 ([8], Lemma 3.1).
Suppose and are elements of , written in their one-line notation. For a fixed , define the permutations
in . Then we have an equality of values .
Proposition 3.3.
Fix , and choose tableaux , both with index . Then
Proof.
Suppose the removable box of occurs in row . Define the following permutations by their images under the RSK correspondence, using the tableau .
By the above Lemma, it remains to show that and the result for follows mutatis mutandis.
The box containing the value in will be in the top row. Hence, is inserted directly into the top row when performing RSK on . In , we instead insert , which will also insert into the top row. Inserting will then push into the next row. Each following insertion for mimics , except the box is constantly pushed down the right-hand side of the insertion tableau. One can verify that this gives the correct tableaux-pair for , with an additional -box in row for each. ∎
For a tableau with index , write for the basis element of , where is the unique value such that , but .
By (2.5), for a tableau with index we have:
In the second case, we sum over with and . Then:
with the sum in the second case as before. If , then if and only if , and likewise for . Hence:
This is the action of on in . Thus, and commute on every basis element of for every generator of . This completes the proof of Theorem 1.2.
4 The long cycle action on the KL basis
In this section we prove Theorem 1.1. Let be the long element, and be the image of the long element of under the standard embedding into . Our goal is to relate the identity to from Lemma 2.1.
Note that will act on a tableau of size by (up to sign) [10]. Moreover, commutes with on . Therefore
Note that the sign depends only on the shape of . Since the index of matches the index of , we obtain:
Lemma 4.1.
Fix a partition . The action of on the KL basis of is
with the sum over satisfying , and are unknown coefficients. Furthermore, the sign of depends only on .
Theorem 1.1 follows immediately from:
Theorem 4.2.
Fix a partition . The action of on the KL basis of is
with the sum over satisfying , and are unknown coefficients. Furthermore, the sign of depends only on .
Proof.
Reindexing the sum from the Lemma with ,
We set and apply the long element to both sides:
Replacing with and with gives the result required. ∎
Example 4.3.
Fix . We arrange using the total index ordering:
Using MAGMA [3] we compute with respect to this ordered basis and calculate its QR decomposition:
One can verify that is precisely the permutation matrix .
Corollary 4.4.
Let and suppose has index . Then , with the sign depending only on . In particular, when is a rectangular partition, then acts by the promotion operator on the KL basis of up to sign.
5 Other permutations acting on the KL basis
In this section we discuss Theorem 1.3. Details will appear in forthcoming work.
Fix a partition . Let denote the set of generators of . For a non-empty subset , let denote the associated parabolic subgroup, and its longest element. To each such subset, we can associate a filtration of the Specht module such that each subspace is -invariant and spanned by KL basis elements, and the quotient representation is isomorphic to a simple -module.
This filtration induces a total preorder on , with when appears strictly before in this filtration. Using partial evacuation operators, we also associate a bijection on such that the action of on the KL basis is given by
for some constants , and a sign which depends only on the equivalence class of under .
We can extend this result in a natural way to chains of subsets of . Call a permutation descending if it is of the form for some increasing chain of subsets . It turns out that a permutation is descending if and only if it is separable.222This observation was communicated to the authors by Joel Gibson, who discovered this fact by enumerating the descending permutations computationally in MAGMA. Each descending permutation has an associated bijection and a total preorder on . The statement is as before, with the action of on the KL basis of given by
(5.1) |
for some constants , and a sign which depends only on the equivalence class of under . Theorem 1.3 follows from this formula.
When is a connected subset for , then is isomorphic to the symmetric group . Moreover, , where is the permutation reversing . Likewise, , where is the partial Schützenberger involution from Section 2.1. Finally, the preorder can be defined explicitly by: if and only if precedes in the total index ordering.
In the case of chains of connected subsets of , equation (5.1) can be proved combinatorially. However, in the case of general chains, we use results from categorical representation theory, which will appear in later work by the authors. This will also prove analogues of these results in other types concerning the action of an ADE Weyl group on the Lusztig’s dual canonical basis in the zero weight space of an irreducible representation of the corresponding Lie algebra.
Finally, we note that the result in (5.1) does not hold for arbitrary permutations. For example, take the non-separable permutation and . Choose any of the orderings of the KL basis for , and compute the decomposition of with respect to this basis. One can verify that is not a generalised permutation matrix under any of these orderings.
Acknowledgements.
The results in Sections 1-4 are part of the first author’s University of Sydney Honours Thesis (2021), supervised by the second author. The authors thank the anonymous referees for helpful comments, and Joel Gibson, Ed Heng, Tony Licata and Eloise Little for useful conversations that improved this work.References
- [1] A. Berenstein and A. Zelevinsky “Canonical bases for the quantum group of type and piecewise-linear combinatorics” In Duke Math. J. 82.3, 1996, pp. 473–502 DOI: 10.1215/S0012-7094-96-08221-6
- [2] A. Bjorner and F. Brenti “Combinatorics of Coxeter Groups”, Graduate Texts in Mathematics Springer Berlin Heidelberg, 2006 URL: https://books.google.com.au/books?id=1TBPz5sd8m0C
- [3] W. Bosma, J. Cannon and C. Playoust “The Magma algebra system. I. The user language” Computational algebra and number theory (London, 1993) In J. Symbolic Comput. 24.3-4, 1997, pp. 235–265 DOI: 10.1006/jsco.1996.0125
- [4] I. Halacheva, T. Licata, I. Losev and O. Yacobi “Categorical braid group actions and cactus groups”, 2021 arXiv:2101.05931
- [5] D. Kazhdan and G. Lusztig “Representations of Coxeter groups and Hecke algebras” In Invent. Math. 53.2, 1979, pp. 165–184 DOI: 10.1007/BF01390031
- [6] S. Kitaev “Patterns in permutations and words” With a foreword by Jeffrey B. Remmel, Monographs in Theoretical Computer Science. An EATCS Series Springer, Heidelberg, 2011, pp. xxii+494 DOI: 10.1007/978-3-642-17333-2
- [7] D. E. Knuth “Permutations, matrices, and generalized Young tableaux” In Pacific J. Math. 34, 1970, pp. 709–727 URL: http://projecteuclid.org.ezproxy.library.sydney.edu.au/euclid.pjm/1102971948
- [8] B. Rhoades “Cyclic sieving, promotion, and representation theory” In J. Combin. Theory Ser. A 117.1, 2010, pp. 38–76 DOI: 10.1016/j.jcta.2009.03.017
- [9] B. E. Sagan “The cyclic sieving phenomenon: a survey” In Surveys in combinatorics 2011 392, London Math. Soc. Lecture Note Ser. Cambridge Univ. Press, Cambridge, 2011, pp. 183–233
- [10] J. R. Stembridge “Canonical bases and self-evacuating tableaux” In Duke Math. J. 82.3, 1996, pp. 585–606 DOI: 10.1215/S0012-7094-96-08224-1