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

\ytableausetup

aligntableaux=center, boxsize=0.5cm

On Schützenberger modules of the cactus group

Jongmin Lim J. Lim: School of Mathematics and Statistics, University of Sydney, Australia [email protected]  and  Oded Yacobi O. Yacobi: School of Mathematics and Statistics, University of Sydney, Australia [email protected]
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 S𝖲𝖼𝗁λS^{\lambda}_{\mathsf{Sch}}, and in this paper we investigate their decomposition into irreducible components. We prove that when λ\lambda is a hook shape, the cactus group action on S𝖲𝖼𝗁λS^{\lambda}_{\mathsf{Sch}} factors through Sn1S_{n-1} 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 𝔤\mathfrak{g} be a reductive complex Lie algebra. In Kashiwara’s theory of 𝔤\mathfrak{g}-crystals, the cactus group plays a role analogous to that of the braid group in representations of the quantum group Uq(𝔤)U_{q}(\mathfrak{g}). Indeed just as the nn-strand braid group acts on nn-fold tensor products of representations of Uq(𝔤)U_{q}(\mathfrak{g}) (resulting in a braided category), the cactus group CnC_{n} acts on nn-fold tensor products of crystals (resulting in a coboundary category) [HK06]. And just as the type 𝔤\mathfrak{g} braid group acts on any integrable representation of Uq(𝔤)U_{q}(\mathfrak{g}), the type 𝔤\mathfrak{g} cactus group acts on any 𝔤\mathfrak{g}-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 𝔤\mathfrak{g}, and showed that it interacts nicely with Kazhdan-Lusztig cells [LosCacti]. For 𝔤=𝔰𝔩n\mathfrak{g}=\mathfrak{sl}_{n}, this recovers the external action of the cactus group corresponding to the zero weight space of the nn-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 Uq(𝔤)U_{q}(\mathfrak{g}), 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 𝔤\mathfrak{g} 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 𝔤=𝔰𝔩n\mathfrak{g}=\mathfrak{sl}_{n}. The corresponding cactus group CnC_{n} is an infinite group generated by cJc_{J}, for subintervals J[1,n]J\subset[1,n], 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 𝔥reg/Sn\mathfrak{h}^{reg}/S_{n}, where 𝔥reg\mathfrak{h}^{reg} is the regular locus in the reflection representation of the symmetric group SnS_{n} [DJS03].

Let λ\lambda be a partition of nn. The cactus group CnC_{n} acts on SYT(λ)\mathrm{SYT}(\lambda), the set of standard Young tableau of shape λ\lambda, where the generator cJc_{J} acts by a partial Schützenberger involution. Letting SλS^{\lambda} denote the Specht module of SnS_{n}, it is natural to view the CnC_{n}-action on SYT(λ)\mathrm{SYT}(\lambda) as an action on the Kazhdan-Lusztig basis of SλS^{\lambda}. We thus obtain a CnC_{n}-action on SλS^{\lambda}, which we term the “Schützenberger module”, and denote S𝖲𝖼𝗁λS^{\lambda}_{\mathsf{Sch}}.

Our main problem, which to our knowledge has not been studied, is the following:

Problem 1.1.

Determine the irreducible constituents of S𝖲𝖼𝗁λS^{\lambda}_{\mathsf{Sch}}.

An obvious obstruction to solving this problem is that we do not have a classification of the finite dimensional irreducible representation of CnC_{n} (it is of wild representation type). Nevertheless, there are naturally occurring families of irreducible representations of CnC_{n} obtained by inflation from symmetric groups.

Indeed, there is a natural homomorphism CnSnC_{n}\to S_{n}, and this can be generalised to a surjective map πk:CnSk\pi_{k}:C_{n}\to S_{k}, for 1kn1\leq k\leq n (cf. Lemma 2.3). For λk\lambda\vdash k we let SπkλS^{\lambda}_{\pi_{k}} be the irreducible CnC_{n}-module on SλS^{\lambda} obtained via pullback by πk\pi_{k}.

Our main theorem solves Problem 1.1 in the case when λ\lambda is a hook partition. To λ=(a,1b)\lambda=(a,1^{b}) a hook partition, we associate a composition of n1n-1 given by λ~=(a1,b)\widetilde{\lambda}=(a-1,b).

Theorem 1.2.

Let λn\lambda\vdash n be a hook partition. We have an isomorphism of CnC_{n}-modules

S𝖲𝖼𝗁λμn1Kμλ~Sπn1μS^{\lambda}_{\mathsf{Sch}}\cong\bigoplus_{\mu\;\vdash\;n-1}K_{\mu\widetilde{\lambda}}S^{\mu}_{\pi_{n-1}} (1.3)

where Kμλ~K_{\mu\widetilde{\lambda}} are the Kostka numbers, unless λ=(2,1)\lambda=(2,1).

Note that in the outlying case, S𝖲𝖼𝗁(2,1)S^{(2,1)}_{\mathsf{Sch}} is simply the two-dimensional module with basis elements interchanged by c[1,3]c_{[1,3]} (c[1,2]c_{[1,2]} and c[2,3]c_{[2,3]} 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 CnC_{n}-action on S𝖲𝖼𝗁λS^{\lambda}_{\mathsf{Sch}} factors through Sn1S_{n-1}, 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 n1n\geq 1. A partition of nn, written λn\lambda\vdash n, is a weakly decreasing sequence of nonnegative integers that sum to nn:

λ=(λ1,,λn),λ1λn0,iλi=n.\lambda=(\lambda_{1},\ldots,\lambda_{n}),\;\lambda_{1}\geq\cdots\geq\lambda_{n}\geq 0,\;\sum_{i}\lambda_{i}=n.

If we drop the weakly decreasing condition, we get the notion of a composition of nn.

We use Young diagrams to represent partitions and compositions. A Young diagram for a composition μ\mu is a finite collection of cells, arranged in left-justified rows, where the ii-th row length is the ii-th entry of μ\mu.

Let λn\lambda\vdash n. A Young tableau of shape λ\lambda is a filling of the corresponding Young diagram with positive integers. For example, here is a Young diagram and tableau of shape λ=(5,3,2)\lambda=(5,3,2):

\ytableaushort12746,318,29\ytableaushort{12746,318,29}

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 TT of shape λ\lambda is the composition of nn, μ(T)=(μ1,μ2,)\mu(T)=(\mu_{1},\mu_{2},\ldots), where μi\mu_{i} is the number of ii’s appearing in TT.

Given λn\lambda\vdash n and m1m\geq 1, we let SSYT(λ,m)\mathrm{SSYT}(\lambda,m) denote the set of semistandard Young tableau of shape λ\lambda and cells filled in with numbers 1,,m1,\ldots,m. We let SYT(λ)\mathrm{SYT}(\lambda) denote the set of standard Young tableau of shape λ\lambda and cells filled in with the numbers 1,,n1,\ldots,n.

The Kostka number KλμK_{\lambda\mu} is defined equivalently as: the number of TSSYT(λ,n)T\in\mathrm{SSYT}(\lambda,n) of shape λ\lambda and content μ\mu, the dimension of the μ\mu weight space in the irreducible representation of 𝔤𝔩n\mathfrak{gl}_{n} of highest weight λ\lambda, or as the multiplicity of SλS^{\lambda} in the permutation module MμM^{\mu} (Equation (2.1)).

Given partitions μ,λ\mu,\lambda we write μλ\mu\subseteq\lambda if μiλi\mu_{i}\leq\lambda_{i} for every ii. Let λ,μ\lambda,\mu be two partitions such that μλ\mu\subseteq\lambda. The skew-diagram of shape λ/μ\lambda/\mu is given by removing the boxes of μ\mu in λ\lambda. A skew tableau is a labelling of these boxes with positive integers. Here is a skew diagram and tableau of shape λ/μ\lambda/\mu for λ=(5,5,3)\lambda=(5,5,3) and μ=(3,2)\mu=(3,2):

\ytableaushort\none\none\none31,\none\none213,311\ytableaushort{\none\none\none 31,\none\none 213,311}

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 μ\mu be a composition of nn. Let T,TT,T^{\prime} be two diagrams of shape μ\mu with entries 1,,n1,\ldots,n. We write TTT\sim T^{\prime} if TT and TT^{\prime} are row-equivalent, i.e. they have the same entries in each row. An equivalence class for this relation is a μ\mu-tabloid. We let Tab(μ)\mathrm{Tab}(\mu) be the set of μ\mu-tabloids. A tabloid can be pictured in a manner similar to tableau. For example, here is a (3,4)(3,4)-tabloid:

\ytableausetuptabloids\ytableaushort123,4567\ytableausetup{tabloids}\ytableaushort{123,4567}

representing the equivalence class of the diagram with entries 1,2,31,2,3 in the first row and 4,5,6,74,5,6,7 in the second.

Set [a,b]={a,a+1,,b}[a,b]=\{a,a+1,\ldots,b\}. Given a tableau TT we let T|[a,b]T|_{[a,b]} be the tableau obtain by deleting all cells with entries not in [a,b][a,b].

2.2. The symmetric group

Let n1n\geq 1. Let SnS_{n} denote the symmetric group on {1,2,,n}\{1,2,\ldots,n\}. Let siSns_{i}\in S_{n} denote the simple transposition swapping ii and i+1i+1. Finite dimensional irreducible complex representations of SnS_{n} are indexed by partitions λ\lambda of nn. The irreducible representation corresponding to λn\lambda\vdash n is the Specht module SλS^{\lambda}. For instance, S(n)S^{(n)} is the trivial representation, S1nS^{1^{n}} is the sign representation, and S(n1,1)S^{(n-1,1)} is the standard representation.

The Specht module SλS^{\lambda} has a remarkable basis indexed by SYT(λ)\mathrm{SYT}(\lambda) called the Kazhdan-Lusztig basis, which we denote {bTTSYT(λ)}\{b_{T}\;\mid\;T\in\mathrm{SYT}(\lambda)\}. To construct this basis one needs to pass to the Iwahori-Hecke algebra Hn(q)H_{n}(q) associated to SnS_{n}. 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 q1q\mapsto 1 leads to the basis {bT}\{b_{T}\}. For more details, see e.g. [KL79, GaMc, Rhoad].

Given a composition μ\mu of nn, let SμSnS_{\mu}\subseteq S_{n} be the corresponding parabolic subgroup. Let MμM^{\mu} denote the induced module IndSμSn()\text{Ind}_{S_{\mu}}^{S_{n}}(\mathbb{C}) from the trivial representation. The module MμM^{\mu} has a basis indexed by the set of row tabloids Tab(μ)\mathrm{Tab}(\mu), where the action is given by permutation of entries. Kostka numbers encode the decomposition of MμM^{\mu} into Specht modules:

MμλKλμSλ.M^{\mu}\cong\bigoplus_{\lambda}K_{\lambda\mu}S^{\lambda}. (2.1)

2.3. The cactus group

Given an interval J=[a,b][1,n]J=[a,b]\subseteq[1,n], let SJSnS_{J}\subseteq S_{n} be the subgroup of permutations which fix iJi\notin J. In the notation of the previous section, SJS_{J} is the parabolic subgroup SμS_{\mu}, where μ=(1a1,ba+1,1nb)\mu=(1^{a-1},b-a+1,1^{n-b}). Let wJSJw_{J}\in S_{J} be the longest element, that is wJw_{J} “flips” the interval [a,b][a,b] via a+ibia+i\mapsto b-i for 0iba0\leq i\leq b-a.

Definition 2.2.

Let n2n\geq 2 be an integer. The cactus group CnC_{n} is generated by (n2){n\choose 2} generators cJc_{J}, indexed by the intervals J{[a,b]| 1a<bn}J\in\{[a,b]\ |\ 1\leq a<b\leq n\}, subject to the following relations: cJ2=1c_{J}^{2}=1, cJcK=cKcJc_{J}c_{K}=c_{K}c_{J} if JK=J\cap K=\varnothing and, cJcK=cwJ(K)cJc_{J}c_{K}=c_{w_{J}(K)}c_{J} if KJK\subseteq J.

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 𝔤\mathfrak{g} [HKRW]. The cactus group defined above corresponds to type An1A_{n-1}.

Symmetric groups are naturally quotients of the cactus group. Indeed, we have a map πn:CnSn,cJwJ\pi_{n}:C_{n}\to S_{n},\;c_{J}\mapsto w_{J}, which is a surjective group homomorphism since the defining relations of CnC_{n} are satisfied also by the elements wJSnw_{J}\in S_{n}. This map can be generalised as follows

Lemma 2.3.

For any 1kn1\leq k\leq n the assignment:

c[a,b]{w[a,bn+k] if nk<ba,1 otherwise,\displaystyle c_{[a,b]}\mapsto\begin{cases}w_{[a,b-n+k]}&\text{ if }n-k<b-a,\\ 1&\text{ otherwise,}\end{cases}

defines a surjective group homomorphism πk:CnSk\pi_{k}:C_{n}\to S_{k}.

Proof.

The third defining relation of CnC_{n} is only non-obvious relation to check. Suppose we have intervals K=[a,b]JIK=[a,b]\subseteq J\subseteq I. We need to show that πk(cJ)πk(cK)=πk(cwJ(K))πk(cJ)\pi_{k}(c_{J})\pi_{k}(c_{K})=\pi_{k}(c_{w_{J}(K)})\pi_{k}(c_{J}).

If nkban-k\geq b-a then πk(cK)=πk(cwJ(K))=1\pi_{k}(c_{K})=\pi_{k}(c_{w_{J}(K)})=1 so the equation holds. Otherwise, we have that nk<ban-k<b-a. Let K=[a,bn+k]K^{\prime}=[a,b-n+k]. Then a quick calculation shows that wJ(K)=wJ(K)w_{J}(K)^{\prime}=w_{J^{\prime}}(K^{\prime}), which proves the desired relation. ∎

By inflation we obtain irreducible representations of CnC_{n} on SλS^{\lambda}, for λk\lambda\vdash k and 1kn1\leq k\leq n, which we denote SπkλS^{\lambda}_{\pi_{k}}.

Remark 2.4.

It’s possible to generalise the maps above further. Given nonnegative numbers i,ji,j such that i+j<ni+j<n, we have a map π(i,j):CnS[1+i,nj]Snij\pi_{(i,j)}:C_{n}\to S_{[1+i,n-j]}\cong S_{n-i-j} given by

c[a,b]{w[a+i,bj] if i+j<ba,1 otherwise,\displaystyle c_{[a,b]}\mapsto\begin{cases}w_{[a+i,b-j]}&\text{ if }i+j<b-a,\\ 1&\text{ otherwise,}\end{cases}

It’s straightforward to check that this satisfies the defining relations of CnC_{n}. We recover πk\pi_{k} defined in Lemma 2.3 as π(0,nk)\pi_{(0,n-k)}.

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 𝗃𝖽𝗍\mathsf{jdt} taking a semistandard skew tableau to a rectified semistandard tableau, which we recall now. Let TSSYT(λ/μ,n)T\in\mathrm{SSYT}(\lambda/\mu,n). Call a removable box of μ\mu a movable box of TT. Then 𝗃𝖽𝗍(T)\mathsf{jdt}(T) is defined as follows:

  1. (1)

    Choose a movable box \ytableausetupnotabloids\ytableaushort(green)\ytableausetup{notabloids}\ytableaushort{{*(green)}} of TT. Move this box with the following rules:

    1. (a)

      If it is adjacent to a box to its east and south, let them be ii and jj respectively.

      \ytableaushort(green)i,j\ytableaushort{{*(green)}i,j}

      If i<ji<j, then swap with \ytableaushorti\ytableaushort{i}

      \ytableaushorti(green),j\ytableaushort{i{*(green)},j}

      Otherwise, swap with \ytableaushortj\ytableaushort{j}

      \ytableaushortji,(green)\ytableaushort{ji,*(green)}
    2. (b)

      If it is adjacent to exactly one box to its east or south, swap it with that box.

    3. (c)

      Repeat this process until it is not adjacent to any boxes to its east or south.

  2. (2)

    Repeat this process with another movable box until there are no movable boxes left.

For example,

𝗃𝖽𝗍(\ytableaushort\none\none122,\none2445,23)=\ytableaushort12225,244\none\none,3\none\mathsf{jdt}\left(\ytableaushort{\none\none 122,\none 2445,23}\right)=\ytableaushort{12225,244\none\none,3\none}

The rectification of a skew semistandard tableau via 𝗃𝖽𝗍\mathsf{jdt} is independent of the choice of the removable boxes at each iteration.

2.4.2. Promotion

The promotion operation is a map :SSYT(λ,n)SSYT(λ,n)\partial:\mathrm{SSYT}(\lambda,n)\to\mathrm{SSYT}(\lambda,n) defined as follows.

  1. (1)

    Turn every box labelled 1 to a dummy box.

  2. (2)

    Apply 𝗃𝖽𝗍\mathsf{jdt} to the dummy boxes.

  3. (3)

    Reduce every non-dummy box’s label by 1.

  4. (4)

    Relabel the dummy boxes to nn

\ytableaushort1123,223,44,5\ytableaushort(green)(green)23,223,44,5\ytableaushort2223,34(green),4(green),5\ytableaushort1112,235,35,4\ytableaushort{1123,223,44,5}\quad\longrightarrow\quad\ytableaushort{*(green)*(green)23,223,44,5}\quad\longrightarrow\quad\ytableaushort{2223,34*(green),4*(green),5}\quad\longrightarrow\quad\ytableaushort{1112,235,35,4}

2.4.3. Schützenberger Involution

The Schützenberger Involution is a map ξ:SSYT(λ,n)SSYT(λ,n)\xi:\mathrm{SSYT}(\lambda,n)\to\mathrm{SSYT}(\lambda,n) defined by

ξ=12n\xi=\partial_{1}\circ\partial_{2}\circ\cdots\circ\partial_{n}

Where k(T)|[1,k]:=(T|[1,k])\partial_{k}(T)|_{[1,k]}:=\partial\left(T|_{[1,k]}\right), while leaving T|[k+1,n]T|_{[k+1,n]} constant.

\ytableaushort1123,223,44,55\ytableaushort1112,235,35,44\ytableaushort1144,225,35,43\ytableaushort1144,235,35,4\ytableaushort{1123,223,44,5}\quad\xrightarrow{\quad\partial_{5}\quad}\quad\ytableaushort{1112,235,35,4}\quad\xrightarrow{\quad\partial_{4}\quad}\quad\ytableaushort{1144,225,35,4}\quad\xrightarrow{\quad\partial_{3}\quad}\quad\ytableaushort{1144,235,35,4}

The Schützenberger involution can be shown to be an involution. Moreover TT and ξ(T)\xi(T) are of the same shape, and if μ(T)=(μ1,,μn)\mu(T)=(\mu_{1},\ldots,\mu_{n}) then μ(ξ(T))=(μn,,μ1)\mu(\xi(T))=(\mu_{n},\ldots,\mu_{1}).

Definition 2.5.

Let λn\lambda\vdash n. For k=2,3,,nk=2,3,\cdots,n, define the partial Schützenberger involution

ξ[1,k]:SSYT(λ,n)SSYT(λ,n)\xi_{[1,k]}:\mathrm{SSYT}(\lambda,n)\to\mathrm{SSYT}(\lambda,n)

to be the Schützenberger involution on T|[1,k]T\big{|}_{[1,k]} (where the relabelling is ik+1ii\mapsto k+1-i), while leaving T|[k+1,n]T\big{|}_{[k+1,n]} constant. Let ξ[a,b]=ξ[1,b]ξ[1,ba+1]ξ[1,b]\xi_{[a,b]}=\xi_{[1,b]}\circ\xi_{[1,b-a+1]}\circ\xi_{[1,b]} for 1a<bn1\leq a<b\leq n.

Proposition 2.6.

[BeKi] The operators ξ[a,b]\xi_{[a,b]} satisfy the following relations:

  1. (1)

    If 1i<j<j+1<k<ln1\leq i<j<j+1<k<l\leq n, then ξ[i,j]ξ[k,l]=ξ[k,l]ξ[i,j]\xi_{[i,j]}\xi_{[k,l]}=\xi_{[k,l]}\xi_{[i,j]}.

  2. (2)

    For 1ik<ljn1\leq i\leq k<l\leq j\leq n, we have ξ[i,j]ξ[k,l]ξ[i,j]=ξ[i+jl,i+jk]\xi_{[i,j]}\xi_{[k,l]}\xi_{[i,j]}=\xi_{[i+j-l,i+j-k]}.

Define a map φ:CnAut(SYT(λ))\varphi:C_{n}\to\operatorname{Aut}(\mathrm{SYT}(\lambda)) by cJξJc_{J}\mapsto\xi_{J} for an interval J[1,n]J\subseteq[1,n].

Proposition 2.7.

[HKRW] The cactus group CnC_{n} acts on the set SSYT(λ,n)\mathrm{SSYT}(\lambda,n) via φ\varphi, and SYT(λ)\mathrm{SYT}(\lambda) is an invariant subset.

Proof.

First, we have

ξ[1,b]2=1\xi_{[1,b]}^{2}=1

as we have established that the Schützenberger involution is an involution. Then we have in general

ξ[a,b]2=ξ[1,b]ξ[1,ba+1]2ξ[1,b]=1,\xi_{[a,b]}^{2}=\xi_{[1,b]}\xi_{[1,b-a+1]}^{2}\xi_{[1,b]}=1,

and the last two relations follow from Proposition 2.6. This shows that CnC_{n} acts on SSYT(λ,n)\mathrm{SSYT}(\lambda,n).

Note that μ(T)=(1,1,,1)\mu(T)=(1,1,\ldots,1) if and only if TSYT(λ)T\in\mathrm{SYT}(\lambda). Since μ(ξ(T))=(1,1,,1)\mu(\xi(T))=(1,1,\ldots,1) as well, we have ξ(T)SYT(λ)\xi(T)\in\mathrm{SYT}(\lambda). Thus ξ[1,k](T)SYT(λ)\xi_{[1,k]}(T)\in\mathrm{SYT}(\lambda) and the claim follows. ∎

Remark 2.8.

The set SSYT(λ,n)\mathrm{SSYT}(\lambda,n) naturally carries a 𝔰𝔩n\mathfrak{sl}_{n}-crystal structure, isomorphic to the crystal of the irreducible representation of 𝔰𝔩n\mathfrak{sl}_{n} of highest weight λ\lambda [HoKa]. The subset SYT(λ)\mathrm{SYT}(\lambda) is the weight zero elements of the crystal. The CnC_{n}-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 λn\lambda\vdash n. Recall the Kazhdan-Lusztig basis {bTTSYT(λ)}\{b_{T}\;\mid\;T\in\mathrm{SYT}(\lambda)\} of SλS^{\lambda}. We define a CnC_{n} action on SλS^{\lambda} using (partial) Schützenberger involutions:

cJbT=bξJ(T).c_{J}\cdot b_{T}=b_{\xi_{J}(T)}.

We term the resulting representation the Schützenberger module of CnC_{n}, and denote it by (ρλ,S𝖲𝖼𝗁λ)(\rho_{\lambda},S^{\lambda}_{\mathsf{Sch}}).

Let vλ=TSYT(λ)bTv_{\lambda}=\sum_{T\in\mathrm{SYT}(\lambda)}b_{T} and define the CnC_{n}-module VλV^{\lambda} by the decomposition S𝖲𝖼𝗁λ=vλVλS^{\lambda}_{\mathsf{Sch}}=\mathbb{C}v_{\lambda}\oplus V^{\lambda}. We begin with some preliminary observations about VλV^{\lambda}.

Proposition 3.1.

Let n3n\geq 3 and set λ=(n1,1)\lambda=(n-1,1). Then VλV^{\lambda} is an irreducible CnC_{n}-module.

Proof.

It suffices to show that φ:CnAut(SYT(λ))\varphi:C_{n}\to\operatorname{Aut}(\mathrm{SYT}(\lambda)) is surjective. We proceed by induction on nn. The base case n=3n=3 is trivial. Define TkT_{k} to be the standard tableau with the box \ytableaushortk+1\ytableaushort{{\scriptstyle k+1}} in the second row. Then SYT(λ)={T1,,Tn1}.\mathrm{SYT}(\lambda)=\{T_{1},\ldots,T_{n-1}\}.

Consider the restriction of φ:CnAut(SYT(λ))\varphi:C_{n}\to\operatorname{Aut}(\mathrm{SYT}(\lambda)) to Cn1CnC_{n-1}\subset C_{n}, where we regard Cn1C_{n-1} as the subgroup generated by {c[a,b]| 1a<bn1}\{c_{[a,b]}\ |\ 1\leq a<b\leq n-1\}. Notice that the image of Cn1C_{n-1} does not change the position of the \ytableaushortn\ytableaushort{n} box. Thus, φ(Cn1)\varphi(C_{n-1}) fixes Tn1\mathrm{T}_{n-1}.

On the other hand, if μ=(n2,1)\mu=(n-2,1), we have a bijection between SYT(μ)\mathrm{SYT}(\mu) and SYT(λ){Tn1}\mathrm{SYT}(\lambda)\setminus\{T_{n-1}\} given by appending the box \ytableaushortn\ytableaushort{n} to the end of the first row for each TSYT(μ)T\in\mathrm{SYT}(\mu). This bijection commutes with the φ(Cn1)\varphi(C_{n-1}) action as φ(Cn1)\varphi(C_{n-1}) disregards \ytableaushortn\ytableaushort{n}. By the induction hypothesis,

φ(Cn1)Aut(SYT(μ)){σAut(SYT(λ))|σ(Tn1)=Tn1}.\varphi(C_{n-1})\cong\operatorname{Aut}(\mathrm{SYT}(\mu))\cong\{\sigma\in\operatorname{Aut}(\mathrm{SYT}(\lambda))\ |\ \sigma(T_{n-1})=T_{n-1}\}.

As φ(Cn)\varphi(C_{n}) is generated by φ(Cn1)\varphi(C_{n-1}) and φ(c[1,n])\varphi(c_{[1,n]}),

φ(Cn){σAut(SYT(λ))|σ(Tn1)=Tn1},φ(c[1,n])Aut(SYT(λ))\varphi(C_{n})\cong\langle\{\sigma\in\operatorname{Aut}(\mathrm{SYT}(\lambda))\ |\ \sigma(T_{n-1})=T_{n-1}\},\varphi(c_{[1,n]})\rangle\cong\operatorname{Aut}(\mathrm{SYT}(\lambda))

as c[1,n](Tn1)Tn1c_{[1,n]}(T_{n-1})\neq T_{n-1}. (Applying the first promotion step shows that the \ytableaushortn\ytableaushort{n} box is in the first row for c[1,n](Tn1)c_{[1,n]}(T_{n-1}). In fact, c[1,n](Tn1)=T1c_{[1,n]}(T_{n-1})=T_{1}.) ∎

In general, VλV^{\lambda} is not irreducible. Indeed, let δ:SYT(λ)SYT(λ)\delta:\mathrm{SYT}(\lambda)\to\mathrm{SYT}(\lambda^{\prime}) be the dual map, where the tableau is reflected by the diagonal from northwest to southeast. Here, λ\lambda^{\prime} is the dual shape of λ\lambda.

\ytableaushort1234,56,7δ\ytableaushort157,26,3,4\ytableaushort{1234,56,7}\xrightarrow{\quad\delta\quad}\ytableaushort{157,26,3,4}

The maps 𝗃𝖽𝗍\mathsf{jdt} and δ\delta commute on standard Young tableaux (but not in general). It follows that δ\delta commutes with the promotion map, and hence the Schützenberger involution. Thus δ\delta commutes with the CnC_{n} action on standard Young tableau.

We therefore obtain an isomorphism δ:S𝖲𝖼𝗁λS𝖲𝖼𝗁λ\delta:S^{\lambda}_{\mathsf{Sch}}\to S^{\lambda^{\prime}}_{\mathsf{Sch}}. In particular, for a self-dual shape λ=λ\lambda=\lambda^{\prime}, we have a non-trivial automorphism δ:S𝖲𝖼𝗁λS𝖲𝖼𝗁λ\delta:S^{\lambda}_{\mathsf{Sch}}\to S^{\lambda}_{\mathsf{Sch}}. As δ\delta is an involution, we have an eigenspace decomposition S𝖲𝖼𝗁λ=S+λSλS^{\lambda}_{\mathsf{Sch}}=S_{+}^{\lambda}\oplus S_{-}^{\lambda} corresponding to the eigenvalues ±1\pm 1 of δ\delta:

S+λ\displaystyle S_{+}^{\lambda} =span{bT+bδ(T)|TSYT(λ)}\displaystyle=span\{b_{T}+b_{\delta(T)}\ |\ T\in\mathrm{SYT}(\lambda)\}
Sλ\displaystyle S_{-}^{\lambda} =span{bTbδ(T)|TSYT(λ)}\displaystyle=span\{b_{T}-b_{\delta(T)}\ |\ T\in\mathrm{SYT}(\lambda)\}

Notice that vλS+λv_{\lambda}\in S_{+}^{\lambda}, and hence there exists a submodule WλW^{\lambda} such that S+λ=vλWλS_{+}^{\lambda}=\mathbb{C}v_{\lambda}\oplus W^{\lambda}. Thus

S𝖲𝖼𝗁λ=vλWλSλ,S^{\lambda}_{\mathsf{Sch}}=\mathbb{C}v_{\lambda}\oplus W^{\lambda}\oplus S_{-}^{\lambda},

and hence Vλ=WλSλV^{\lambda}=W^{\lambda}\oplus S_{-}^{\lambda}. Consequently VλV^{\lambda} is not irreducible for self-dual λ\lambda.

Theorem 1.2 generalises Proposition 3.1 to arbitrary hook-shaped partitions.

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 nn\in\mathbb{N} and λn\lambda\vdash n. A Gelfand-Tsetlin pattern with nn rows and top row λ\lambda is a triangular arrangement of nonnegative integers {λi,j}1ijn\{\lambda_{i,j}\}_{1\leq i\leq j\leq n} such that λi,j+1λi,jλi+1,j+1\lambda_{i,j+1}\geq\lambda_{i,j}\geq\lambda_{i+1,j+1} and the top row is λ\lambda. Denote the set of such patterns as GTP(λ,n)\mathrm{GTP}(\lambda,n).

[λ1,nλ2,nλ3,nλn,nλ1,n1λ2,n1λn1,n1λ1,n2λn2,n2λ1,1]\begin{bmatrix}\lambda_{1,n}&&\lambda_{2,n}&&\lambda_{3,n}&&\cdots&&\lambda_{n,n}\\ &\lambda_{1,n-1}&&\lambda_{2,n-1}&&\cdots&&\lambda_{n-1,n-1}\\ &&\lambda_{1,n-2}&&\cdots&&\lambda_{n-2,n-2}\\ &&&\ddots&&\reflectbox{$\ddots$}\\ &&&&\lambda_{1,1}\end{bmatrix}

Define a map Φ:SSYT(λ,n)GTP(λ,n)\Phi:\mathrm{SSYT}(\lambda,n)\to\mathrm{GTP}(\lambda,n) as follows. Let TSSYT(λ,n)T\in\mathrm{SSYT}(\lambda,n). For 1kn1\leq k\leq n, as TT is semistandard, T|[1,k]T\big{|}_{[1,k]} cannot have more than kk rows. Let the shape of T|[1,k]T\big{|}_{[1,k]} be (λ1,k,λ2,k,,λk,k)(\lambda_{1,k},\lambda_{2,k},\cdots,\lambda_{k,k}). As the notation suggests, set Φ(T)\Phi(T) equal to 𝒯={λi,j}1ijn\mathcal{T}=\{\lambda_{i,j}\}_{1\leq i\leq j\leq n}. The following is immediate:

Proposition 3.2.

Φ:SSYT(λ,n)GTP(λ,n)\Phi:\mathrm{SSYT}(\lambda,n)\to\mathrm{GTP}(\lambda,n) is a bijection.

Bereinstein and Kirillov defined operators acting on GTP(λ,n)\mathrm{GTP}(\lambda,n) as follows. Let 𝒯={λi,j}1ijnGTP(λ,n)\mathcal{T}=\{\lambda_{i,j}\}_{1\leq i\leq j\leq n}\in\mathrm{GTP}(\lambda,n). Define τk:GTP(λ,n)GTP(λ,n)\tau_{k}:\mathrm{GTP}(\lambda,n)\to\mathrm{GTP}(\lambda,n) for 1kn11\leq k\leq n-1 by τk(𝒯)={λ~i,j}1ijn\tau_{k}(\mathcal{T})=\{\widetilde{\lambda}_{i,j}\}_{1\leq i\leq j\leq n}, where

ai,j\displaystyle a_{i,j} :=min(λi,j+1,λi1,j1)\displaystyle:=\min(\lambda_{i,j+1},\lambda_{i-1,j-1})
bi,j\displaystyle b_{i,j} :=max(λi,j1,λi+1,j+1)\displaystyle:=\max(\lambda_{i,j-1},\lambda_{i+1,j+1})
λ~i,j\displaystyle\widetilde{\lambda}_{i,j} :=λi,j(jk)\displaystyle:=\lambda_{i,j}\qquad(j\neq k)
λ~i,k\displaystyle\widetilde{\lambda}_{i,k} :=ai,k+bi,kλi,k\displaystyle:=a_{i,k}+b_{i,k}-\lambda_{i,k}

For the edge cases we let a1,j=λ1,j+1a_{1,j}=\lambda_{1,j+1} and bj,j=λj+1,j+1b_{j,j}=\lambda_{j+1,j+1}.

Proposition 3.3.

[BeKi] The operators τ1,τ2,,τn1\tau_{1},\tau_{2},\cdots,\tau_{n-1} satisfy the following relations.

τk2\displaystyle\tau_{k}^{2} =11kn1\displaystyle=1\quad\quad 1\leq k\leq n-1
τkτl\displaystyle\tau_{k}\tau_{l} =τlτk|kl|2\displaystyle=\tau_{l}\tau_{k}\quad|k-l|\geq 2
(τ1τ2)6\displaystyle(\tau_{1}\tau_{2})^{6} =1\displaystyle=1
(τ1qk)4\displaystyle(\tau_{1}q_{k})^{4} =1k3\displaystyle=1\quad\quad k\geq 3

where qk:=(τ1)(τ2τ1)(τ3τ2τ1)(τkτk1τ1)q_{k}:=(\tau_{1})(\tau_{2}\tau_{1})(\tau_{3}\tau_{2}\tau_{1})\cdots(\tau_{k}\tau_{k-1}\cdots\tau_{1}).

It is conjectured by Berenstein and Kirillov that these generate all relations among the operators τ1,τ2,,τn1\tau_{1},\tau_{2},\ldots,\tau_{n-1}.

Definition 3.4.

The Berenstein-Kirillov group BKnBK_{n} is the group generated by t1,t2,,tn1t_{1},t_{2},\ldots,t_{n-1} with relations as in Proposition 3.3.

By transport of structure via Φ\Phi, we have an action of BKnBK_{n} on SSYT(λ,n)\mathrm{SSYT}(\lambda,n). Following [CGP], we will describe this explicitly. Let TSSYT(λ,n)T\in\mathrm{SSYT}(\lambda,n). Recall that T|[k,k+1]T\big{|}_{[k,k+1]} is a disjoint union of rectangles and strips of the form

\ytableaushortkk\none[]k,k+1k+1\none[]k+1or\ytableaushortkk\none[]kk+1k+1\none[]k+1\ytableaushort{kk{\none[\cdots]}k,\scriptstyle k+1\scriptstyle k+1{\none[\cdots]}\scriptstyle k+1}\qquad\text{or}\qquad\ytableaushort{kk{\none[\cdots]}k\scriptstyle k+1\scriptstyle k+1{\none[\cdots]}\scriptstyle k+1}

We say a strip is of type (a,b)(a,b) if it contains aa many \ytableaushortk\ytableaushort{k}-boxes and bb many \ytableaushortk+1\ytableaushort{\scriptstyle k+1}-boxes.

Define τ~k:SSYT(λ,n)SSYT(λ,n)\widetilde{\tau}_{k}:\mathrm{SSYT}(\lambda,n)\to\mathrm{SSYT}(\lambda,n) such that τ~k(T)\widetilde{\tau}_{k}(T) acts on T|[k,k+1]T\big{|}_{[k,k+1]} by replacing each strip of type (a,b)(a,b) with a strip of type (b,a)(b,a), and leaving the rectangles unchanged:

{ytableau}\none&\none\none\none\none\nonekkkk+1k+1k+1\none\none\none\nonekk+1k+1k+1\nonekkkk+1k+1k+1τ~k{ytableau}\none&\none\none\none\none\nonekkkkkk+1\none\none\none\nonekk+1k+1k+1\nonekkk+1kk+1k+1\ytableau\none&\none\none\none\none\none kkk\scriptstyle k+1\scriptstyle k+1\scriptstyle k+1\\ \none\none\none\none k\scriptstyle k+1\scriptstyle k+1\scriptstyle k+1\\ \none kkk\\ \scriptstyle k+1\scriptstyle k+1\scriptstyle k+1\xrightarrow{\quad\widetilde{\tau}_{k}\quad}\ytableau\none&\none\none\none\none\none kkkkk\scriptstyle k+1\\ \none\none\none\none k\scriptstyle k+1\scriptstyle k+1\scriptstyle k+1\\ \none kk\scriptstyle k+1\\ k\scriptstyle k+1\scriptstyle k+1

In the example above, strips of type (0,1),(1,0),(1,1),(1,3)(0,1),(1,0),(1,1),(1,3) were swapped with strips of type (1,0),(0,1),(1,1),(3,1)(1,0),(0,1),(1,1),(3,1).

We define τ~k(T)\widetilde{\tau}_{k}(T) to be the tableau obtained by replacing T|[k,k+1]T\big{|}_{[k,k+1]} with τ~k(T|[k,k+1])\widetilde{\tau}_{k}(T\big{|}_{[k,k+1]}), and leaving the other boxes unchanged.

Lemma 3.5.

Let TSSYT(λ,n)T\in\mathrm{SSYT}(\lambda,n) and Φ(T)=𝒯={λi,j}1ijnGTP(λ,n)\Phi(T)=\mathcal{T}=\{\lambda_{i,j}\}_{1\leq i\leq j\leq n}\in\mathrm{GTP}(\lambda,n). Recall

ai,k=min(λi,k+1,λi1,k1)bi,k=max(λi+1,k+1,λi,k1)a_{i,k}=\min(\lambda_{i,k+1},\lambda_{i-1,k-1})\qquad b_{i,k}=\max(\lambda_{i+1,k+1},\lambda_{i,k-1})

Then the strip of T|[k,k+1]T\big{|}_{[k,k+1]} in row ii is of type (λi,kbi,k,ai,kλi,k)(\lambda_{i,k}-b_{i,k},a_{i,k}-\lambda_{i,k}) starting at column bi,k+1b_{i,k}+1.

Proof.

Recall that λi,k\lambda_{i,k} corresponds to the number of boxes in row ii of TT labelled from 1,2,,k1,2,\cdots,k.
Assume there is no rectangles with its first row in row ii. This means that every box above \ytableaushortk+1\ytableaushort{\scriptstyle k+1} in the i+1i+1-th row has a label less than or equal to k1k-1. This is precisely when λi+1,k+1λi,k1\lambda_{i+1,k+1}\leq\lambda_{i,k-1}, and in this case, the strip indeed starts at column bi,k+1=λi,k1+1b_{i,k}+1=\lambda_{i,k-1}+1. On the other hand, if there is such a rectangle, then we have λi+1,k+1>λi,k1\lambda_{i+1,k+1}>\lambda_{i,k-1}. This rectangle ends at column bi,k=λi+1,k+1b_{i,k}=\lambda_{i+1,k+1}, hence the strip starts at column bi,k+1b_{i,k}+1 after it as claimed.

\ytableaushort\none\nonekk\none[],k+1\none[]\ytableaushort\none\nonekk\none[],k+1k+1\none[]\ytableaushort\none\nonekk\none[],k+1k+1k+1\none[]\ytableaushort{\none\none kk{\none[\cdots]},\scriptstyle k+1{\none[\cdots]}}\qquad\ytableaushort{\none\none kk{\none[\cdots]},\scriptstyle k+1\scriptstyle k+1{\none[\cdots]}}\qquad\ytableaushort{\none\none kk{\none[\cdots]},\scriptstyle k+1\scriptstyle k+1\scriptstyle k+1{\none[\cdots]}}
λi+1,k+1<λi,k1λi+1,k+1=λi,k1λi+1,k+1>λi,k1\lambda_{i+1,k+1}<\lambda_{i,k-1}\qquad\lambda_{i+1,k+1}=\lambda_{i,k-1}\qquad\lambda_{i+1,k+1}>\lambda_{i,k-1}

The proof for ai,ka_{i,k} is similar. The strip is indeed of type (λi,kbi,k,ai,kλi,k)(\lambda_{i,k}-b_{i,k},a_{i,k}-\lambda_{i,k}) as the boxes labelled kk in row ii span columns bi,k+1b_{i,k}+1 to λi,k\lambda_{i,k} by the correspondence given by Φ\Phi. ∎

Proposition 3.6.

For TSSYT(λ,n)T\in\mathrm{SSYT}(\lambda,n) we have that tkT=τ~k(T)t_{k}\cdot T=\widetilde{\tau}_{k}(T).

Proof.

It suffices to show that Φτ~k=τkΦ\Phi\widetilde{\tau}_{k}=\tau_{k}\Phi. Notice that τ~k\widetilde{\tau}_{k} does not affect the shape of T|[1,j]T\big{|}_{[1,j]} for all 1jn1\leq j\leq n and jkj\neq k. Furthermore, in TT, each strip in row ii spans column bi,k+1b_{i,k}+1 to column ai,ka_{i,k} and is of type (λi,kbi,k,ai,kλi,k)(\lambda_{i,k}-b_{i,k},a_{i,k}-\lambda_{i,k}). Thus in τ~k(T)\widetilde{\tau}_{k}(T), this strip is replaced by a strip of type (ai,kλi,k,λi,kbi,k)(a_{i,k}-\lambda_{i,k},\lambda_{i,k}-b_{i,k}).

Let 𝒯=Φ(T)\mathcal{T}=\Phi(T) and set τk(𝒯)={λ~i,j}1ijn\tau_{k}(\mathcal{T})=\{\widetilde{\lambda}_{i,j}\}_{1\leq i\leq j\leq n}. Let

a~i,k=min(λ~i,k+1,λ~i1,k1)b~i,k=max(λ~i+1,k+1,λ~i,k1)\widetilde{a}_{i,k}=\min(\widetilde{\lambda}_{i,k+1},\widetilde{\lambda}_{i-1,k-1})\qquad\widetilde{b}_{i,k}=\max(\widetilde{\lambda}_{i+1,k+1},\widetilde{\lambda}_{i,k-1})

Recall that τk\tau_{k} is an operation on GTP(λ,n)\mathrm{GTP}(\lambda,n) which affects only the kk-th row, hence λi,j=λ~i,j\lambda_{i,j}=\widetilde{\lambda}_{i,j} for all jkj\neq k. Thus, a~i,k=ai,k\widetilde{a}_{i,k}=a_{i,k} as follows:

a~i,k\displaystyle\widetilde{a}_{i,k} =min(λ~i,k+1,λ~i1,k1)\displaystyle=\min(\widetilde{\lambda}_{i,k+1},\widetilde{\lambda}_{i-1,k-1})
=min(λi,k+1,λi1,k1)=ai,k\displaystyle=\min(\lambda_{i,k+1},\lambda_{i-1,k-1})=a_{i,k}

Similarly, we have b~i,k=bi,k\widetilde{b}_{i,k}=b_{i,k}. Thus the strip at row ii for Φ1(τk(𝒯))\Phi^{-1}(\tau_{k}(\mathcal{T})) also starts and ends at the same column, but is of type (λ~i,kbi,k,ai,kλ~i,k)(\widetilde{\lambda}_{i,k}-b_{i,k},a_{i,k}-\widetilde{\lambda}_{i,k}). However, as λ~i,k=ai,k+bi,kλi,k\widetilde{\lambda}_{i,k}=a_{i,k}+b_{i,k}-\lambda_{i,k},

(λ~i,kbi,k,ai,kλ~i,k)=(ai,kλi,k,λi,kbi,k)(\widetilde{\lambda}_{i,k}-b_{i,k},a_{i,k}-\widetilde{\lambda}_{i,k})=(a_{i,k}-\lambda_{i,k},\lambda_{i,k}-b_{i,k})

Thus τ~k(T)=Φ1(τk(𝒯))\widetilde{\tau}_{k}(T)=\Phi^{-1}(\tau_{k}(\mathcal{T})). ∎

This proposition implies that for standard Young tableaux, the action of BKnBK_{n} is particularly easy to describe:

Corollary 3.7.

Let TSYT(λ)T\in\mathrm{SYT}(\lambda) for λn\lambda\vdash n. Then tkt_{k} swaps the two boxes \ytableaushortk\ytableaushort{k} and \ytableaushortk+1\ytableaushort{\scriptstyle k+1} if they are not adjacent, otherwise tkT=Tt_{k}\cdot T=T.

Proof.

As TT is standard, T|[k,k+1]T\big{|}_{[k,k+1]} consists of two boxes, which can be non-adjacent, horizontally adjacent, or vertically adjacent:

\ytableaushort\nonek,k+1\ytableaushortkk+1\ytableaushortk,k+1\ytableaushort{\none k,\scriptstyle k+1}\qquad\ytableaushort{k\scriptstyle k+1}\qquad\ytableaushort{k,\scriptstyle k+1}

The non-adjacent case is essentially two disjoint strips of type (1,0)(1,0) and (0,1)(0,1) each. Thus τ~k\widetilde{\tau}_{k} swaps the two boxes. The vertically adjacent case has no strips, while the horizontally adjacent case is of type (1,1)(1,1), which stays constant under τ~k\widetilde{\tau}_{k}. ∎

Consider the elements pk,qkBKnp_{k},q_{k}\in BK_{n} defined by

pk:=tktk1t1qk=p1p2pkp_{k}:=t_{k}t_{k-1}\cdots t_{1}\qquad q_{k}=p_{1}p_{2}\cdots p_{k}

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 pkp_{k} and qkq_{k} on SYT(λ,n)\mathrm{SYT}(\lambda,n) are equivalent to k+1\partial_{k+1} and c[1,k+1]c_{[1,k+1]}, the promotion and Schützenberger involution operations on T|[1,k+1]T\big{|}_{[1,k+1]}.

Proof.

Let TSSYT(λ)T\in\mathrm{SSYT}(\lambda). We first prove that pk=k+1p_{k}=\partial_{k+1} by induction on kk.
For the base case, 2\partial_{2} and t1t_{1} act by identity on \ytableaushort12\ytableaushort{12} and \ytableaushort1,2\ytableaushort{1,2}. Thus, 2=t1=p1\partial_{2}=t_{1}=p_{1}.
For the inductive case, notice that the 𝗃𝖽𝗍\mathsf{jdt} step of k\partial_{k} and k+1\partial_{k+1} are identical until the dummy box becomes adjacent to the \ytableaushortk+1\ytableaushort{\scriptstyle k+1} box.

Case 1: If the dummy box is never adjacent to \ytableaushortk+1\ytableaushort{\scriptstyle k+1} in the 𝗃𝖽𝗍\mathsf{jdt} step of k\partial_{k},
Then the 𝗃𝖽𝗍\mathsf{jdt} step of k\partial_{k} and k+1\partial_{k+1} are identical, and they only differ in the relabelling step. For k\partial_{k}, the dummy box is labelled as kk, and the \ytableaushortk+1\ytableaushort{\scriptstyle k+1} box is kept constant, while every other box’s label is reduced by 1. On the other hand, for k+1\partial_{k+1}, the dummy box is labelled as k+1k+1, while every other box, including \ytableaushortk+1\ytableaushort{\scriptstyle k+1}, has its label reduced by 1. Furthermore, in both cases, \ytableaushortk\ytableaushort{k} and \ytableaushortk+1\ytableaushort{\scriptstyle k+1} are not adjacent due to our assumption. Thus, tkt_{k} acts by swapping \ytableaushortk\ytableaushort{k} and \ytableaushortk+1\ytableaushort{\scriptstyle k+1} on k(T)\partial_{k}(T) and we have tkk(T)=k+1(T)t_{k}\partial_{k}(T)=\partial_{k+1}(T).

Case 2: If the dummy box comes adjacent to \ytableaushortk+1\ytableaushort{\scriptstyle k+1} in the 𝗃𝖽𝗍\mathsf{jdt} step of k\partial{k},
Then the 𝗃𝖽𝗍\mathsf{jdt} step of k+1\partial_{k+1} must have an extra step of swapping the dummy box with \ytableaushortk+1\ytableaushort{\scriptstyle k+1}. Then after the relabelling steps of k\partial_{k} and k+1\partial_{k+1}, we have k(T)=k+1(T)\partial_{k}(T)=\partial_{k+1}(T). Furthermore, by assumption, we have \ytableaushortk\ytableaushort{k} and \ytableaushortk+1\ytableaushort{\scriptstyle k+1} adjacent, thus tkt_{k} acts by identity on k(T)\partial_{k}(T).

Thus we have overall tkk=k+1t_{k}\circ\partial_{k}=\partial_{k+1}. Hence by induction, k+1=tkk=tkpk1=pk\partial_{k+1}=t_{k}\circ\partial_{k}=t_{k}\circ p_{k-1}=p_{k}. Then we have by definition

c[1,k+1]=ξk=12k=p1p2pk=qkc_{[1,k+1]}=\xi_{k}=\partial_{1}\partial_{2}\cdots\partial_{k}=p_{1}p_{2}\cdots p_{k}=q_{k}

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 Cn0C^{0}_{n} is the quotient of CnC_{n} by the relations

cici+1ci=ci+1cici+1\displaystyle c_{i}c_{i+1}c_{i}=c_{i+1}c_{i}c_{i+1} (C3)

where ci=c[i,i+1]c_{i}=c_{[i,i+1]} for 1in11\leq i\leq n-1.

Remark 3.10.

Since ci=c[1,i+2]c2c[1,i+2]c_{i}=c_{[1,i+2]}c_{2}c_{[1,i+2]} and ci+1=c[1,i+2]c1c[1,i+2]c_{i+1}=c_{[1,i+2]}c_{1}c_{[1,i+2]} the relations defining the reduced cactus group are conjugates of a single relation, that is, Cn0=Cn/(c[1,2]c[2,3])3C^{0}_{n}=C_{n}/\langle(c_{[1,2]}c_{[2,3]})^{3}\rangle.

The following is the main result of [CGP].

Theorem 3.11.

There is a group isomorphism χ:Cn0BKn\chi:C^{0}_{n}\to BK_{n} given by

c[1,i]qi12in.c_{[1,i]}\mapsto q_{i-1}\qquad 2\leq i\leq n.
Corollary 3.12.

Let λn\lambda\vdash n. The action of CnC_{n} on SYT(λ)\mathrm{SYT}(\lambda) factors through Cn0C_{n}^{0}.

Proof.

Let x=(c[1,2]c[2,3])3Cnx=(c_{[1,2]}c_{[2,3]})^{3}\in C_{n}. Notice that xx is a non-identity element. Using c[1,2]c[2,3]=c[1,2]c[1,3]c[1,2]c[1,3]c_{[1,2]}c_{[2,3]}=c_{[1,2]}c_{[1,3]}c_{[1,2]}c_{[1,3]} we have

x=(c[1,2]c[2,3])3=(c[1,2]c[1,3])6.x=(c_{[1,2]}c_{[2,3]})^{3}=(c_{[1,2]}c_{[1,3]})^{6}.

By Theorem 3.8, the action of xx is equivalent to the action of y=(t1(t1t2t1))6BKny=(t_{1}(t_{1}t_{2}t_{1}))^{6}\in BK_{n}. However,

y=(t1(t1t2t1))6=(t2t1)6=1y=(t_{1}(t_{1}t_{2}t_{1}))^{6}=(t_{2}t_{1})^{6}=1

It follows that xx acts by identity on SYT(λ)\mathrm{SYT}(\lambda), and the action of CnC_{n} on SYT(λ)\mathrm{SYT}(\lambda) factors through the projection map π:CnCn0\pi:C_{n}\to C^{0}_{n}. ∎

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 𝔤\mathfrak{g}-crystal factors over the reduced cactus group of type 𝔤\mathfrak{g} (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 λ\lambda is a hook shape, i.e. λ=(m,1,,1)\lambda=(m,1,\ldots,1) for some mm and some number of 11s. 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 BKn=t1,t2,,tn1BK_{n}=\langle t_{1},t_{2},\cdots,t_{n-1}\rangle acts on SYT(λ)\mathrm{SYT}(\lambda) for λn\lambda\vdash n. This gives rise to a representation

ψ:BKnGL(S𝖲𝖼𝗁λ).\psi:BK_{n}\to GL(S^{\lambda}_{\mathsf{Sch}}).

Recall that tkt_{k} acts by swapping \ytableaushortk\ytableaushort{k} and \ytableaushortk+1\ytableaushort{\scriptstyle k+1} if they are not adjacent, and otherwise does nothing. As \ytableaushort1\ytableaushort{1} and \ytableaushort2\ytableaushort{2} are always adjacent for standard tableaux, t1t_{1} always acts by identity, and thus ψ(t1)=1\psi(t_{1})=1.

The remaining generators ψ(t2),,ψ(tn1)\psi(t_{2}),\cdots,\psi(t_{n-1}) satisfy the relations of BKnBK_{n}. In particular,

ψ(tk)2=1andψ(tk)ψ(tl)=ψ(tl)ψ(tk)|kl|2\psi(t_{k})^{2}=1\qquad\text{and}\qquad\psi(t_{k})\psi(t_{l})=\psi(t_{l})\psi(t_{k})\qquad|k-l|\geq 2

Assume for the purposes of discussion that for k2k\geq 2, we have

(ψ(tk)ψ(tk+1))3=1\displaystyle(\psi(t_{k})\psi(t_{k+1}))^{3}=1 (\star)

This would give a surjective group homomorphism

η:Sn1im(ψ)skψ(tk+1)\eta:S_{n-1}\to\operatorname{im}(\psi)\qquad s_{k}\mapsto\psi(t_{k+1})

Since BKnCn0BK_{n}\cong C_{n}^{0} (Theorem 3.11), and the action of CnC_{n} on SYT(λ)\mathrm{SYT}(\lambda) factors through Cn0C_{n}^{0} (Corollary 3.12), this will allow us to use the representation theory of Sn1S_{n-1} to study S𝖲𝖼𝗁λS^{\lambda}_{\mathsf{Sch}}. The following lemma describes when this approach is feasible.

Lemma 3.14.

Let λn\lambda\vdash n. Then relation ()(\star) holds for all TSYT(λ)T\in\mathrm{SYT}(\lambda) if and only if λ=(2,2)\lambda=(2,2) or λ\lambda is a hook shape.

Proof.

As tkt_{k} and tk+1t_{k+1} act on TSYT(λ)T\in\mathrm{SYT}(\lambda) depending on how \ytableaushortk\ytableaushort{k}, \ytableaushortk+1\ytableaushort{\scriptstyle k+1}, and \ytableaushortk+2\ytableaushort{\scriptstyle k+2} are adjacent, we consider the ways the three boxes can be adjacent.

  1. Case 1:

    If the three boxes are all non-adjacent, ()(\star) is true. For example:

    \ytableaushort\none\nonek,\nonek+1,k+2tk+1\ytableaushort\none\nonek,\nonek+2,k+1tk\ytableaushort\none\nonek+1,\nonek+2,ktk+1\ytableaushort\none\nonek+2,\nonek+1,k\ytableaushort{\none\none k,\none\scriptstyle k+1,\scriptstyle k+2}\xrightarrow{t_{k+1}}\ytableaushort{\none\none k,\none\scriptstyle k+2,\scriptstyle k+1}\xrightarrow{t_{k}}\ytableaushort{\none\none\scriptstyle k+1,\none\scriptstyle k+2,k}\xrightarrow{t_{k+1}}\ytableaushort{\none\none\scriptstyle k+2,\none\scriptstyle k+1,k}
    tk\ytableaushort\none\nonek+2,\nonek,k+1tk+1\ytableaushort\none\nonek+1,\nonek,k+2tk\ytableaushort\none\nonek,\nonek+1,k+2\xrightarrow{t_{k}}\ytableaushort{\none\none\scriptstyle k+2,\none k,\scriptstyle k+1}\xrightarrow{t_{k+1}}\ytableaushort{\none\none\scriptstyle k+1,\none k,\scriptstyle k+2}\xrightarrow{t_{k}}\ytableaushort{\none\none k,\none\scriptstyle k+1,\scriptstyle k+2}
  2. Case 2:

    If two boxes are adjacent and one is not adjacent to either, then ()(\star) is true. For example:

    \ytableaushort\none\nonek,k+1k+2tk+1\ytableaushort\none\nonek,k+1k+2tk\ytableaushort\none\nonek+1,kk+2tk+1\ytableaushort\none\nonek+2,kk+1\ytableaushort{\none\none k,\scriptstyle k+1\scriptstyle k+2}\xrightarrow{t_{k+1}}\ytableaushort{\none\none k,\scriptstyle k+1\scriptstyle k+2}\xrightarrow{t_{k}}\ytableaushort{\none\none\scriptstyle k+1,k\scriptstyle k+2}\xrightarrow{t_{k+1}}\ytableaushort{\none\none\scriptstyle k+2,k\scriptstyle k+1}
    tk\ytableaushort\none\nonek+2,kk+1tk+1\ytableaushort\none\nonek+1,kk+2tk\ytableaushort\none\nonek,k+1k+2\xrightarrow{t_{k}}\ytableaushort{\none\none\scriptstyle k+2,k\scriptstyle k+1}\xrightarrow{t_{k+1}}\ytableaushort{\none\none\scriptstyle k+1,k\scriptstyle k+2}\xrightarrow{t_{k}}\ytableaushort{\none\none k,\scriptstyle k+1\scriptstyle k+2}
  3. Case 3:

    If all three are adjacent in a single row or single column, then ()(\star) is true. In this case, \ytableaushortk\ytableaushort{k} and \ytableaushortk+1\ytableaushort{\scriptstyle k+1} are always adjacent, so tkt_{k} always acts by identity. The same is true for \ytableaushortk+1\ytableaushort{\scriptstyle k+1} and \ytableaushortk+2\ytableaushort{\scriptstyle k+2}, so tk+1t_{k+1} also always acts by identity.

  4. Case 4:

    If all three are adjacent in the following shape, then ()(\star) is not true.

    \ytableaushortkk+1,k+2\ytableaushort{k\scriptstyle k+1,\scriptstyle k+2}

    In this case, \ytableaushortk\ytableaushort{k} and \ytableaushortk+1\ytableaushort{\scriptstyle k+1} are always adjacent while \ytableaushortk+1\ytableaushort{\scriptstyle k+1} and \ytableaushortk+2\ytableaushort{\scriptstyle k+2} are always not adjacent. Thus tkt_{k} acts by identity while tk+1t_{k+1} acts by swapping \ytableaushortk+1\ytableaushort{\scriptstyle k+1} and \ytableaushortk+2\ytableaushort{\scriptstyle k+2}. Hence for TSYT(λ)T\in\mathrm{SYT}(\lambda) with this formation,

    (ψ(tk)ψ(tk+1))3(T)=ψ(tk+1)3(T)=ψ(tk+1)(T)T(\psi(t_{k})\psi(t_{k+1}))^{3}(T)=\psi(t_{k+1})^{3}(T)=\psi(t_{k+1})(T)\neq T

Hence if λ\lambda is a shape that does not allow the Case 4 configuration for k2k\geq 2 (since we ignore ψ(t1)\psi(t_{1})), then ()(\star) will hold true. This is exactly when λ=(2,2)\lambda=(2,2) or λ\lambda is a hook shape. ∎

Remark 3.15.

In general for all shapes λ\lambda, we have (ψ(tk)ψ(tk+1))6=1(\psi(t_{k})\psi(t_{k+1}))^{6}=1. To see that we don’t necessarily have (ψ(tk)ψ(tk+1))3=1(\psi(t_{k})\psi(t_{k+1}))^{3}=1, consider λ=(3,2)\lambda=(3,2) and the following tableau

T=\ytableaushort134,25T=\ytableaushort{1{3}{4},2{5}}

A quick calculation shows:

(t3t4)3(T)=\ytableaushort135,24(t_{3}t_{4})^{3}(T)=\ytableaushort{135,24}
Theorem 3.16.

For a hook shape λ\lambda not of the form (n),(1n)(n),(1^{n}) or (2,1)(2,1), the map η:Sn1im(ψ)\eta:S_{n-1}\to\operatorname{im}(\psi) is an isomorphism.

Proof.

We have already shown that η\eta is surjective. Notice that for n4n\geq 4, im(ψ)\operatorname{im}(\psi) has two distinct non-identity elements, namely ψ(t2)\psi(t_{2}) and ψ(t3)\psi(t_{3}). These are nontrivial because there is a TSYT(λ)T\in\mathrm{SYT}(\lambda) such that

T|[1,4]=\ytableaushort124,3or\ytableaushort13,2,4T\big{|}_{[1,4]}=\ytableaushort{124,3}\qquad\text{or}\qquad\ytableaushort{13,2,4}

and indeed T,ψ(t2)(T),ψ(t3)(T)T,\psi(t_{2})(T),\psi(t_{3})(T) are all distinct.

Assume for contradiction kerη{1}\ker\eta\neq\{1\}. If n5n\neq 5, the only other normal subgroups of Sn1S_{n-1} are An1A_{n-1}, the alternating group, and Sn1S_{n-1}. In either case, we have [Sn1:kerη]2[S_{n-1}:\ker\eta]\leq 2, hence

|Sn1/kerη|2<3|im(ψ)||S_{n-1}/\ker\eta|\leq 2<3\leq|\operatorname{im}(\psi)|

which gives a contradiction. Thus the kernel must be the trivial group for n5n\neq 5.

For n=5n=5, the only possible hook shapes are (4,1),(3,1,1),(2,1,1,1)(4,1),(3,1,1),(2,1,1,1). For the (4,1)(4,1) case, using the notation from the proof of Proposition 3.1, we have that tit_{i} interchanges Ti1T_{i-1} and TiT_{i} for i=2,3,4i=2,3,4. Thus im(ψ)\operatorname{im}(\psi) is isomorphic the subgroup of GL4()GL_{4}(\mathbb{C}) generated by the simple transposition matrices, which is clearly isomorphic to S4S_{4}. The (2,1,1,1)(2,1,1,1) case is dual to the (4,1)(4,1) case.

Let us examine the case for λ=(3,1,1)\lambda=(3,1,1). There are six tableaux in SYT(λ)\mathrm{SYT}(\lambda).

T1=\ytableaushort123,4,5T2=\ytableaushort124,3,5T3=\ytableaushort125,3,4T_{1}=\ytableaushort{123,4,5}\qquad T_{2}=\ytableaushort{124,3,5}\qquad T_{3}=\ytableaushort{125,3,4}
T4=\ytableaushort145,2,3T5=\ytableaushort135,2,4T6=\ytableaushort134,2,5T_{4}=\ytableaushort{145,2,3}\qquad T_{5}=\ytableaushort{135,2,4}\qquad T_{6}=\ytableaushort{134,2,5}

Hence we can view xim(ψ)x\in\operatorname{im}(\psi) as elements in S6S_{6} acting on the subscript: xTk=Tx(k)xT_{k}=T_{x(k)}. Then we have:

ψ(t2)=(2 6)(3 5)ψ(t3)=(1 2)(4 5)ψ(t4)=(2 3)(5 6)\psi(t_{2})=(2\ 6)(3\ 5)\qquad\psi(t_{3})=(1\ 2)(4\ 5)\qquad\psi(t_{4})=(2\ 3)(5\ 6)

The subgroup generated by these elements has more than six elements. Since every nontrivial normal subgroup of S4S_{4} has index at most 6 ([S4:K4]=6[S_{4}:K_{4}]=6 where K4K_{4} is the Klein four group), we conclude as above that the map is indeed injective. ∎

Corollary 3.17.

Let λn\lambda\vdash n be a hook partition not of the form (n),(1n)(n),(1^{n}) or (2,1)(2,1). Then the representation S𝖲𝖼𝗁λS^{\lambda}_{\mathsf{Sch}} factors over Sn1S_{n-1} as follows:

Cn{C_{n}}GL(S𝖲𝖼𝗁λ){GL(S^{\lambda}_{\mathsf{Sch}})}Sn1{S_{n-1}}ρλ\scriptstyle{\rho_{\lambda}}πn1\scriptstyle{\pi_{n-1}}η\scriptstyle{\eta}
Proof.

Recall we have the projection map π:CnCn0\pi:C_{n}\to C_{n}^{0} and the isomorphism χ:Cn0BKn\chi:C_{n}^{0}\to BK_{n}. By Theorem 3.11 and Corollary 3.12 we have the diagram:

Cn{C_{n}}GL(S𝖲𝖼𝗁λ){GL(S^{\lambda}_{\mathsf{Sch}})}BKn{BK_{n}}ρλ\scriptstyle{\rho_{\lambda}}χπ\scriptstyle{\chi\circ\pi}ψ\scriptstyle{\psi}

By Theorem 3.16, if we can extend this diagram with a map from BKnBK_{n} to Sn1S_{n-1}, obtaining the desired result. ∎

Definition 3.18.

Let λn\lambda\vdash n be a hook shape. The boxes in the first row (excluding the first box) are the arm of λ\lambda, and the boxes in the first column (excluding the first box) are the leg of λ\lambda. We let λ~=(a,b)\widetilde{\lambda}=(a,b) be the two-part composition of n1n-1 formed by the arm and leg of λ\lambda.

In the example below, λ=(5,1,1)\lambda=(5,1,1) has arm length 44 and leg length 22 and λ~=(4,2)\widetilde{\lambda}=(4,2).

{ytableau}&\ytableau{}&\\ \\ \\
Proposition 3.19.

Let λn\lambda\vdash n be a hook shape. Then the Sn1S_{n-1} representation (η,S𝖲𝖼𝗁λ)(\eta,S^{\lambda}_{\mathsf{Sch}}) is isomorphic to the permutation module Mλ~M^{\widetilde{\lambda}}.

Proof.

Set λ~=(a,b)\widetilde{\lambda}=(a,b). We have shown that S𝖲𝖼𝗁λS𝖲𝖼𝗁λS^{\lambda}_{\mathsf{Sch}}\cong S^{\lambda^{\prime}}_{\mathsf{Sch}} by the dual map, so we can assume without loss of generality that aba\geq b.

Since S𝖲𝖼𝗁λS^{\lambda}_{\mathsf{Sch}} and Mλ~M^{\widetilde{\lambda}} are both permutation modules of Sn1S_{n-1}, it suffices to prove that there exists a bijection between the standard bases of each module that commutes with the action of Sn1S_{n-1}.

Define the operation 𝖥𝗈𝗅𝖽\mathsf{Fold} given by the following illustration:

{ytableau}1&x1x2\none[]xay1y2\none[]yb𝖥𝗈𝗅𝖽\ytableausetuptabloids{ytableau}x1&x2x3xay1y2yb\ytableau 1&x_{1}x_{2}\none[\cdots]x_{a}\\ y_{1}\\ y_{2}\\ \none[\vdots]\\ y_{b}\xrightarrow{\quad\mathsf{Fold}\quad}\ytableausetup{tabloids}\ytableau x_{1}&x_{2}x_{3}\cdots x_{a}\\ y_{1}y_{2}\cdots y_{b}

Notice that we lose the hinge, i.e. \ytableausetupnotabloids\ytableaushort1\ytableausetup{notabloids}\ytableaushort{1}, so the entries in 𝖥𝗈𝗅𝖽(T)\mathsf{Fold}(T) are now from {2,3,,n}\{2,3,\cdots,n\}. Thus we subtract 1 from each label, and define a map F:SYT(λ)Tab(λ~)F:\mathrm{SYT}(\lambda)\to\mathrm{Tab}(\widetilde{\lambda}) given by F=(1)𝖥𝗈𝗅𝖽F=(-1)\circ\mathsf{Fold}.

\ytableaushort12345,6,7,8\ytableaushort{12345,6,7,8}\ytableausetuptabloids\ytableaushort2345,678\ytableausetup{tabloids}\ytableaushort{2345,678}\ytableaushort1234,567\ytableaushort{1234,567}𝖥𝗈𝗅𝖽\mathsf{Fold}1-1FF

We first show that F:SYT(λ)Tab(λ~)F:\mathrm{SYT}(\lambda)\to\mathrm{Tab}(\widetilde{\lambda}) is a bijection. Let Z={2,3,,a+b+1}Z=\{2,3,\cdots,a+b+1\}. Any XZX\subseteq Z of cardinality aa uniquely determines T(X)SYT(λ)T(X)\in\mathrm{SYT}(\lambda), where XX is the set of numbers in the arm of TT. Similarly for Z={1,2,,a+b}Z^{\prime}=\{1,2,\cdots,a+b\} any XZX^{\prime}\subseteq Z^{\prime} of cardinality aa determines P(X)Tab(λ~)P(X^{\prime})\in\mathrm{Tab}(\widetilde{\lambda}), where XX^{\prime} is the set of numbers in the first row. The bijection is then F:T(X)P(X1)F:T(X)\mapsto P(X-1).

Next we show that FF commutes with Sn1S_{n-1}. Recall that siSn1s_{i}\in S_{n-1} acts via ti+1BKnt_{i+1}\in BK_{n} on SYT(λ)\mathrm{SYT}(\lambda), which swaps \ytableausetupnotabloids\ytableaushorti+1\ytableausetup{notabloids}\ytableaushort{{\scriptstyle i+1}} and \ytableaushorti+2\ytableaushort{{\scriptstyle i+2}} if the two boxes are not adjacent in the tableau, and if they are it acts by the identity.

Notice that for i1i\geq 1, \ytableaushorti+1\ytableaushort{{\scriptstyle i+1}} and \ytableaushorti+2\ytableaushort{{\scriptstyle i+2}}, are adjacent in TSYT(λ)T\in\mathrm{SYT}(\lambda) if and only if \ytableaushorti\ytableaushort{i} and \ytableaushorti+1\ytableaushort{{\scriptstyle i+1}} are in the same row in F(T)Tab(λ~)F(T)\in\mathrm{Tab}(\widetilde{\lambda}). Thus, sis_{i} acts on TT trivially if and only if it acts trivially on F(T)F(T). Otherwise, the boxes swap. In TT, the boxes swap from the arm to the leg and vice versa. In F(T)F(T), the boxes swap rows. Since the arm maps to the first row and the leg maps to the second row, this shows that FF commutes with every transposition. ∎

We are now ready to prove our main result.

Proof of Theorem 1.2.

In the setting of the theorem, λ\lambda is a hook partition not equal to (2,1)(2,1). Consider first the case when λ\lambda is not of the form (n),(1n)(n),(1^{n}). By the above proposition and Equation (2.1), there is an isomorphism of Sn1S_{n-1}-modules

S𝖲𝖼𝗁λμn1Kμλ~Sμ.S^{\lambda}_{\mathsf{Sch}}\cong\bigoplus_{\mu\;\vdash\;n-1}K_{\mu\widetilde{\lambda}}S^{\mu}.

By Corollary 3.17 this implies the isomorphism of Equation (1.3).

The remaining cases are easily dealt with by direct computation. If λ=(n)\lambda=(n) or λ=(1n)\lambda=(1^{n})) the Kostka number Kμλ~K_{\mu\widetilde{\lambda}} is zero unless μ=(n1)\mu=(n-1), in which case it is equal to 11. Thus both sides of (1.3) are isomorphic to the trivial CnC_{n}-module. ∎

References