186
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Core blocks for Hecke algebras of type B and sign sequences

ORCID Icon
Pages 1965-1981 | Received 18 Aug 2023, Accepted 26 Oct 2023, Published online: 15 Nov 2023

Abstract

We consider the core blocks corresponding to the Hecke algebras of type B over a field of arbitrary characteristic. To each core block, we associate two non-negative integers which determine the indexing of the Specht modules and simple modules in the block, the weight of the block, the multicharge of the algebra (up to a shift) and the block decomposition matrix.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1 Introduction

Let H=Hr,n(q,Q) denote an Ariki-Koike algebra over a field F of characteristic p with quantum characteristic e{2,3,}{}. Each of these algebras decomposes into a direct sum of indecomposable two-sided ideals, its blocks, and so in order to try to understand the algebras one may study certain types of block in the hope that they are more manageable.

There is an important class of H-modules which are indexed by the r-multipartitions of n and which are known as Specht modules. The composition factors of each Specht module all lie in the same block and so we may think of partitioning the Specht modules into blocks. One may consider the block decomposition matrix, which records the composition factors of the Specht modules belonging to a given block. In general, computing block decomposition matrices is hard. There exist recursive algorithms which compute the transition coefficients for a highest weight module of the Fock space representation of Uq(sl̂e); by Ariki’s theorem [1], these coefficients coincide with the decomposition numbers for the Ariki-Koike algebras when p = 0. When p > 0, the transition coefficients give a lower bound for the decomposition numbers, and may be considered as a first approximation to them. It is reasonable to ask when this approximation is precise, or when the block decomposition matrix is independent of the characteristic of the field.

In this paper, we look at the core blocks for the Hecke algebras of type B, that is, the Ariki-Koike algebras H=H2,n(q,Q). These core blocks, which were initially introduced and studied by Fayers [Citation11], are blocks in which none of the bipartitions indexing the Specht modules have any removable e-rim hooks. If e= or e > n, all blocks are core blocks. To each core block B, we associate two non-negative integers nB and pB. We show that these two integers determine the indexing of the Specht modules and simple modules in the block, the weight of the block, the multicharge of H (up to a shift) and the block decomposition matrix. We present a result of George Witty [Citation27] that shows that in some situations they also determine the homomorphism space between pairs of Specht modules in the block. All these results are independent of the characteristic of the field.

When e= or e > n, decomposition numbers for the Hecke algebras of type B, or the corresponding cyclotomic q-Schur algbras, are not new. When e= and the parameters of the multicharge are weakly increasing, Leclerc and Miyachi [Citation19, Theorem 3] give a closed formula for the canonical basis elements of the irreducible highest weight representation V(Λ); by Ariki’s theorem [1], this gives the graded decomposition numbers for the Hecke algebras when e= and p = 0. When e= and p = 0, Brundan and Stroppel [Citation7, Section 9] apply the theory they have developed earlier in their paper to construct a basis for the radical of their cell module S(λ), thus determining the graded dimension of the simple modules. When e= or e > n, Hu and Mathas [Citation14, Appendix B] give a formula for the graded decomposition numbers of the quiver Schur algebras of level two in terms of tableaux combinatorics; this is independent of the characteristic of the field. In current work [Citation2, Citation3], the authors determine full submodule structures for the Specht modules of the core blocks of the Hecke algebras and the Weyl modules for their Schur algebras.

Each paper above uses a different notation. In this paper, we index the (bipartitions corresponding to) Specht modules in a core block by sign sequences, sequences containing elements from {0,,+} with a fixed number of entries of each type, where the 0 entries are essentially redundant. The integers nB and pB above count the numbers of –s and + s respectively. The structure of the core blocks means that our combinatorics feels extremely natural. When e=, it is straightforward to pass from our notation to that of [Citation19] or [Citation2].

The structure of this paper is as follows. In Section 2 we introduce the background material: Section 2.1 defines the relevant combinatorics and Section 2.2 introduces the Ariki-Koike algebras. For more information on the Ariki-Koike algebras and their combinatorics, we refer the reader to the survey paper of Mathas [Citation23] and for more information on their connections with the cyclotomic KLR algebras of type A, we refer them to the survey paper of Kleshchev [Citation17].

In Section 3, we move on to considering the core blocks, as introduced by Fayers [Citation11], when r = 2. In Section 3.1 we define the sign sequences and in Section 3.2 we show how they index the block decomposition matrices for core blocks. In Section 3.3 we briefly introduce the Fock space representation of Uv(sl̂e) and describe its connection with the Ariki-Koike algebras; more details can be found in the papers of Lascoux, Leclerc and Thibon [Citation18] and Ariki [Citation1]. In Section 3.4, we find the decomposition numbers for the core blocks. We start with the special case that the base tuple is flat, which includes the case that e=, and then use what is essentially a focused version of Scopes equivalence [Citation8, Citation25] to generalize it to arbitrary core blocks. Finally, in Section 3.5 we summarize our results, present Witty’s theorem and give some examples.

2 Background

Throughout this paper, we will use a parameter e{2,3,}{}. If e is finite, we define I={0,1,,e1} which we may identify with Z/eZ; otherwise we define I=Z. In both cases we take < to be the usual total order on I.

2.1 Multipartitions and abacus configurations

Suppose that n0. A partition of n is a sequence λ=(λ1,λ2,) of non-negative integers such that λ1λ2 and i1λi=n. We write |λ|=n. We let Λn denote the set of partitions of n and Λ=n0Λn denote the set of all partitions. Now suppose that r1. An r-multipartition, or multipartition, of n is an r-tuple of partitions λ=(λ(1),λ(2),,λ(r)) such that k=1r|λ(k)|=n. We write |λ|=n. We let Λnr denote the set of r-multipartitions of n and Λr=n0Λnr denote the set of all r-multipartitions.

Suppose that λΛr. The Young diagram of λ is the set [λ]={(x,y,k)Z>0×Z>0×{1,2,,r}|yλx(k)}.

We say that a node n[λ] is removable if [λ]{n} is the Young diagram of a multipartition and we say that n[λ] is addable if [λ]{n} is the Young diagram of a multipartition. The rim of [λ] is the set of nodes {(x,y,k)[λ]|(x+1,y+1,k)[λ]}. For h1, a removable h-rim hook is a set of h connected nodes in the rim such that removing those nodes gives the Young diagram of a multipartition.

Now fix e{2,3,}{}. Suppose that aZr. To each node n=(x,y,k)Z>0×Z>0×{1,2r} we associate its residue resa(n)I. If e is finite (resp. e=) we set resa(x,y,k)=ak+yxmode (resp. resa(x,y,k)=ak+yx). For λΛr we define the residue set of λ to be the multiset Resa(λ)={res(n)|n[λ]}. We define an equivalence relation a on Λr by saying that λaμ if and only if Resa(λ)=Resa(μ) and we refer to the a-equivalence classes of Λr as blocks. Clearly if e is finite and sZr with skakmode for 1kr then λaμ if and only if λsμ.

Given aZr, we define a subset ΛaΛr. For iI, we call an addable (resp. removable) node of residue i an addable (resp. removable) i-node. Given μΛr and iI, we define a total order C on the set of addable and removable nodes of [μ] by saying that (x1,y1,k1) C(x2,y2,k2) if k1<k2 or if k1 = k2 and x1<x2. We may then define the i-signature of μ by looking at all the addable and removable i-nodes of [μ] ordered according to C and writing a for an addable i-node and r for a removable i-node. We then construct the reduced i-signature of μ by repeatedly removing all adjacent (ra)-pairs until there are no such pairs left. If there are any r terms in the reduced i-signature of μ, the first r term corresponds to a removable i-node of [μ] which is called a good i-node.

It is usual to write + instead of a and – instead of r, however we want use those symbols for other purposes in this paper. The set Λa is defined recursively. Suppose that μΛr.

  • If |μ|=0 then μΛa.

  • Otherwise, if μ does not contain a good i-node for any iI then μΛa.

  • Otherwise, suppose that n is a good i-node of μ for some iI. Let μ¯ be the multipartition whose Young diagram is obtained from [μ] by removing the node n. Then μΛa if and only if μ¯Λa.

The multipartitions in Λa are known as Kleshchev multipartitions; they have the property that if μΛa then μ(k) is e-restricted for all 1kr. We set Λna=ΛaΛnr.

Given an r-multipartition, it is convenient to represent it as a r-tuple of abacus configurations. If e is finite, an e-abacus is an abacus with e vertical runners which are infinite in both directions and which are indexed from left to right by the elements of I. The possible bead positions are indexed by the elements of Z such that bead position b on the abacus is in row l of runner i where b=le+i and iI. If e=, the abacus has runners and bead positions indexed by the elements of I=Z, so that runner iZ contains either one bead or no beads.

For λΛ and aZ, define the β-set Ba(λ)={λxx+a|x1}.

We define the abacus configuration of λ with respect to a to be the abacus where we put a bead at position b for each bBa(λ). For λΛr and aZr we define the abacus configuration of λ with respect to a to be an r-tuple of abacus configurations, the kth one of which is the abacus configuration of λ(k) with respect to ak.

Operations on the Young diagram of λ can be translated into operations on the abacus configuration of λ.

Lemma 2.1.

Let λΛr and aZr. All residues below are with repect to a as are all abacus configurations.

  • Removing an i-node from component k of [λ] corresponds to moving a bead on runner i of the abacus of λ(k) back by one position.

  • Adding an i-node to component k of [λ] corresponds to moving a bead on runner i – 1 of the abacus of λ(k) forward by one position.

  • If e then adding (resp. removing) an e-rim hook to (resp. from) component k of [λ] corresponds to pushing a bead on the abacus of λ(k) down one position (resp. up one position).

Example.

Take e = 3. Let λ=((6,5,2,12),(6,1)) and take a=(8,6). Then the abacus configuration of λ with respect to a is given byand [λ] has three removable e-rim hooks. For example, [((42,2,11),(6,1))] is formed by [λ] by removing an e-rim hook from [λ(1)] and has abacus configuration (with respect to a)

.

When drawing abacuses we truncate the runners and assume that the positions above the drawn portion all contain beads while the positions below are empty.

We say that λΛr is a multicore if e= or if e< and no component of λ has any removable e-rim hooks; equivalently, λ is a multicore if no bead in the abacus of any component has an empty space above it. (This is independent of the choice of a used to draw the abacus configurations.) We say that a a-equivalence class B is a core block if every λB is a multicore. If λ is a multicore and aZr then for 1kr and iI we define lika(λ) as follows. If e is finite, set lika(λ)=max{lZ|le+iBak(λ(k))};in other words, lika(λ) is the lowest row of the abacus configuration for λ(k) with respect to ak which contains a bead on runner i. If e=, set lika(λ) to be 1 if iBak(λ(k)) and 0 otherwise.

Lemma 2.2.

[Citation11, Theorem 3.1] Let sZr and suppose that B is a s-equivalence class.

  1. Suppose e is finite. If B is a core block then there exists a=(a1,a2,,ar)Zr, with akskmode for all 1kr, and b=(b0,b1,,be1)Ze such that for each iI,1kr and λB,lika(λ) is equal to either bi or bi+1.

  2. Suppose e=. Then every block is a core block and if we set b=(,0,0,0,) and take a=s then for each iI,1kr and λB,lika(λ) is equal to either bi or bi+1.

Conversely, suppose that e is finite and that λΛr. If there exists a=(a1,a2,,ar)Zr, with akskmode for all 1kr, and b=(b0,b1,,be1)Ze such that for each iI and 1kr,lika(λ) is equal to either bi or bi+1, then the s-equivalence class of λ is a core block.

2.2 The Ariki-Koike algebra

Let r1 and n0 and let F be a field of characteristic p0. Choose qF  {0} and Q=(Q1,,Qr)Fr. The Ariki-Koike algebra H=Hr,n(q,Q) is the unital associative F-algebra with generators T0,,Tn1 and relations (Ti+q)(Ti1)=0, for 1in1,TiTj=TjTi, for 0i,jn1,|ij|>1,TiTi+1Ti=Ti+1TiTi+1, for 1in2,(T0Q1)(T0Qr)=0,T0T1T0T1=T1T0T1T0.

Define e2 to be minimal such that 1+q++qe1=0, or set e= if no such value exists. Two parameters Qk and Ql are q-connected if Qk=qaQl for some aI. Each Ariki-Koike algebra H is Morita equivalent to a direct sum of tensor products of smaller algebras whose parameters are all q-connected [Citation9] and so we will assume that all our parameters are q-connected, in fact, that they are all powers of q where q1. If a=(a1,a2,,ar)Zr satisfies Qk=qak for all 1kr then we call aa multicharge for H. If e is finite then qe=1 so there are infinitely many possible multicharges for H.

The algebra H is a cellular algebra [Citation10, Citation13] with the cell modules indexed by the r-multipartitions of n. The cell module Sλ indexed by the multipartition λ is called a Specht module. Due to the properties of cellular algebras, all the composition factors of Sλ lie in the same block, and so we can think of the Specht modules as being partitioned into blocks.

Proposition 2.3.

[Citation21, Theorem 2.11] Suppose that a is a multicharge for H and that λ,μΛnr. Then Sλ and Sμ lie in the same block of H if and only if Resa(λ)=Resa(μ).

This result explains why we called the a-equivalence classes blocks: two multipartitions of n are in the same class if and only if the corresponding Specht modules lie in the same block.

From the properties of cellular algebras, we know that there is a bilinear form on each cell module. If we define rad(Sμ) to be the radical of the Specht module Sμ with respect to this bilinear form then if a is a multicharge for H then rad(Sμ)Sμ if and only if μΛna and so {Dμ=Sμ/rad(Sμ)|μΛna}is a complete set of non-isomorphic irreducible H-modules. Given λΛnr and μΛna we define [Sλ:Dμ] to be the multiplicity of the simple module Dμ as a composition factor of the Specht module Sλ.

Brundan and Kleshchev [Citation4] have shown that the Ariki-Koike algebras are isomorphic to certain graded algebras defined by Khovanov and Lauda [Citation15, Citation16] and by Rouquier [Citation24]: the cyclotomic KLR algebras of type A. Through this isomorphism, we may think of H as being graded. There is a grading on the Specht modules [Citation6], thus we can define [Sλ:Dμ]vN[v,v1] to be the graded multiplicity of the simple module Dμ as a composition factor of the Specht module Sλ; we recover the original decomposition number by setting v = 1. For more details, we refer the reader to the survey paper [Citation17].

We define the (graded) decomposition matrix of H to be the matrix whose rows are indexed by the elements of Λnr and whose columns are indexed by the elements of Λna with entries equal to [Sλ:Dμ]v, for λΛnr and μΛna. If B is a a-equivalence class, where λB is such that |λ|=n, we define the block decomposition matrix of B to be the submatrix of the decomposition matrix of H whose rows and columns are indexed only by elements of B.

3 Core blocks when r = 2

For the rest of this paper, we assume r = 2 so that our Specht modules are indexed by bipartitions and we shall refer to multicores as bicores. We fix e{2,3,}{} and s=(s1,s2)I2. We shall take a block to mean an s-equivalence class of Λ2 and a core block to be a block B in which each λB is a bicore.

Let F be a field of characteristic p0. Fix qF where q is a primitive eth root of unity if e is finite and ql0,1 for any lZ otherwise. Let Q=(qs1,qs2). If B is a block then there exists n0 such that |λ|=n for all λB and so we associate to B the block of the Hecke algebra Hn=H2,n(q,Q) containing the Specht modules Sλ for λB. We denote this corresponding block of Hn by B̂ so that we have a correspondence BB̂ between s-equivalence classes and 2-sided indecomposable ideals of the Hecke algebras Hn; we refer to both as blocks.

3.1 Sign sequences

Set Δ={δ=(δi)iI|δi{,0,+} and if e= then δi=δi=0 for all i0}.

We say that δ,δΔ are essentially the same if the two sequences obtained by removing all the 0s from each of them are the same.

For δΔ, we define sets R(δ)Δ and S(δ)I2 as follows.

  1. Take R(δ)=δ and S(δ)=.

  2. If there do not exist i,jI with i < j and δi= and δj=+, end the process. Return R(δ) and S(δ).

  3. Otherwise, choose i,jI with i < j and δi= and δj=+ with the property that δm=0 for all i < m < j. Add (i, j) to S(δ) and set R(δ)i=R(δ)j=0. Go back to step (2).

Example.

Suppose that e = 19 and δ=(,0,,0,+,,0,0,+,,+,+,0,+,,0,+,0,+). We find it helpful to draw a diagram for δ as below.

Then S(δ)={(2,4),(5,8),(9,10),(0,11),(14,16)},R(δ)=(0,0,0,0,0,0,0,0,0,0,0,0,0,+,0,0,0,0,+).

Set Δ0={δΔ|R(δ) or +R(δ)}.

Suppose δΔ0. For each S={(i1,j1),,(it,jt)}S(δ), define δSΔ by setting δmS={+,m=il for some 1lt,,m=jl for some 1lt,δm,otherwise.

If δ=δS for some SS(δ), we write δ δ and set l(δ,δ)=|S|. We write if δ=δS for some SS(δ) where S has the property that if (i1,j1),(i2,j2)S(δ) with i1<i2<j2<j1 then (i1,j1)S(i2,j2)S.

Example.

Let e = 8 and δ=(,,+,,,+,+,)Δ0. We write δ asso that S(δ)={(1,2),(4,5),(3,6)}. Then we have SδSl(δS,δ)δS[graphic] δ(,,+,,,+,+,)0X{(1,2)}(,+,,,,+,+,)1X{(4,5)}(,,+,,+,,+,)1X{(3,6)}(,,+,+,,+,,)1{(1,2),(4,5)}(,+,,,+,,+,)2X{(1,2),(3,6)}(,+,,+,,+,,)2{(4,5),(3,6)}(,,+,+,+,,,)2X{(1,2),(4,5),(3,6)}(,+,,+,+,,,)3X

Informally, we can see that each element of the set {δS|SS(δ)} is obtained by swapping some pairs + that lie at either end of an arc in the diagram of δ.

3.2 Indexing the block decomposition matrices corresponding to core blocks

Let B be a core block. Suppose e is finite. By Lemma 2.2, we can find a=(a1,a2)Z2 and b=(b0,b1,,be1)Ze with the property that akskmode for k = 1, 2 and for any λB we have lika(λ)=bi or lika(λ)=bi+1, for iI and k = 1, 2. If a,b satisfy these conditions we call r=(a,b)a reduced pair for B. We define a total order (which depends on the choice of b) on I by saying that ij if bi < bj or if bi = bj and i < j. If I={i0,i1,,ie1} with i0i1ie1 we take π=πb(B) be the permutation that sends j to ij for jI. If e=, we define r=(s,b​​​​°) where b​​​​°=(,0,0,0,) to be the unique reduced pair for B. We then take to be the usual total order < on Z and π to be the identity permutation on Z.

For λB and r=(a,b)a reduced pair for B, define δλr=((δλr)i)iI by setting (δλr)i=lπ(i)2a(λ)lπ(i)1a(λ)for all iI. Our choice of r ensures that δλrΔ, where we abuse notation by identifying 1 with + and –1 with –.

If e is finite, the reduced pair r is not uniquely determined by the conditions above. However we shall see that unless |B|=1, δλr and δλr are essentially the same for any reduced pairs r and r.

Example.

Take e = 7 and s=(1,6). Let B be the core block containing the bipartition λ=((13,10,8,7,6,4,3,2,16),((13,12,10,9,8,6,53,3,23,16)).

If r=(a,b) is a reduced pair for B then there exists lZ such that a is of the form a=(8+7l,6+7l).

Taking a=(29,27) we have two choices for b, each of which gives a different ordering : b=(3,1,6,3,5,2,5)1503462δλr=(,,,+,+,,0),b=(3,1,5,3,5,2,5)1503246δλr=(,,,+,0,+,).

However, the only choice comes when we look at the third runner on the abacus configurations, that is, when l22a(λ)=l21a(λ)=6, and we note that the two choices for δλr are essentially the same.

Lemma 3.1.

[Citation11, Propn. 3.7] Suppose B is a core block and r=(a,b) is a reduced pair for B. Let λB and suppose i,jI with (δλr)i= and (δλr)j=+. Define sij(λ) to be the bicore μ with lmka(μ) ={lmka(λ)1,π(m)=i and k=1 or π(m)=j and k=2,lmka(λ)+1,π(m)=j and k=1 or π(m)=i and k=2,lmka(λ),otherwise.

Then μB and (δμr)m ={+,m=i,,m=j,(δλr)m,otherwise.

Moreover, if νB then we may form ν from λ by repeatedly applying operations of the form sij for some i, j as above.

Corollary 3.2.

Suppose that B is a core block and r is a reduced pair for B. Let λ,μB. Then δμr can be formed from δλr by permuting the entries equal to ±. Conversely, any sequence ϵΔ formed by permuting the entries equal to ± in δλr is equal to δνr for some νB.

Consequently, |B|=1 if and only if δλr or +δλr.

Lemma 3.3.

Suppose e is finite and that B is a core block with |B|>1. Suppose that r=(a,b) and r=(a,b) are reduced pairs for B with a=(a1,a2) and a=(a1,a2). Let λB.

  1. There exists dZ with ak=ak+de for k = 1, 2.

  2. The sequences δλr and δλr are essentially the same.

Proof.

  1. Since r and r are both reduced pairs for B, there exist d1,d2Z such that ak=ak+dke for k = 1, 2. As |B|>1, there exist i,jI such that

    li2a(λ)li1a(λ)=1li2a(λ)li1a(λ)=d2d11,lj2a(λ)lj1a(λ)=1lj2a(λ)lj1a(λ)=d2d1+1.

    so since |lm2a(λ)lm1a(λ)|1 for all mI, we must have d1 = d2.

  2. Using part (1) we may assume that ak=ak+de for k = 1, 2. If we set c=(b0+d,b1+d,,be1+d) then r̂=(a,c) is a reduced pair for B and δλr̂=δλr. Let

    I±={iI|li2a(λ)li1a(λ)},

    so that (δλr̂)i0 if and only if iI±. If iI± then we must have bi=min{li2a(λ),li1a(λ)}=ci. So restricting the order to the set I± we see that changing (a,c) to (a,b) permutes the elements of δλr̂ while keeping the non-zero entries in the same order. Hence δλr̂ and δλr are essentially the same.

If |B|=1, Lemma 3.3 does not hold. If e is finite and B={λ}, assume that we choose the reduced pair r such that 0δλr and 1δλr. This choice is arbitrary; it is simply made so that for every element of a core block we have an expression δλr which is essentially independent of the choice of the reduced pair r. Recall that if e=, there is a unique reduced pair for B. Unless we need to emphasize the reduced pair, we will henceforth write δλ instead of δλr.

Suppose that B is a core block. Take λB and hence define nB=#{iI|(δλ)i=},pB=#{iI|(δλ)i=+},mB=min{nB,pB}.

Given Corollary 3.2 and the discussion above, these parameters are well-defined and independent of the choice of λ.

Lemma 3.4.

There exists dZ such that if e is finite (resp. e=) then s1nB+dmode and s2pB+dmode (resp. s1=nB+d and s2=pB+d).

Proof.

Suppose that r=(a,b) is a reduced pair for B and let λB. Suppose e is finite. Then by [Citation20, Lemma 2.2], skakiIlika(λ)mode for k = 1, 2. Hence s2s1iI(li2a(λ)li1a(λ))iI(δλ)ipBnBmode.

Suppose e=. Choose t0. Then sk=ak=t+#{iI|it and lika(λ)=1} for k = 1, 2. Hence s2s1=it(li2a(λ)li1a(λ))=it(δλ)i=pBnB.

In [Citation12, Section 2.1], Fayers introduced the weight of a multipartition. Two multipartitions in the same block have the same weight and so for a block B we define wt(B) to be the weight of any multipartition belonging to B. We define the weight of the corresponding block B̂ as wt(B̂)=wt(B). We do not give Fayers’ definition here, but we note that, roughly speaking, the blocks B̂ of small weight tend to be easier to understand. We have |B|=1 if and only if wt(B)=0 which holds if and only if B̂ is simple.

Lemma 3.5.

[Citation12, Propn. 3.8] Suppose that B is a core block. Then wt(B)=mB.

Corollary 3.6.

Let B be a core block. If e is finite then 0wt(B)e2.

For a core block B, we have seen that the Specht modules in B̂ can be indexed by a subset of Δ. The next problem is to decide which μB index simple modules. We will see that this can be determined by looking at δμ. Let b​​​​°=(bi​​​​°)iI be the sequence where bi​​​​°=0 for all iI. We first show that it is sufficient to consider the case where r=(a,b​​​​°) is a reduced pair for B. Note that in this case, we have equal to the usual total order < on I.

Lemma 3.7.

[Citation22, Proposition 2.10] Suppose that B is a core block with (a,b) a reduced pair for B. Let μB. Then for iI and k = 1, 2 we have lika(μ)=bi+xik where xik{0,1}. We define μ¯Λ2 and a¯Z2 so that lika¯(μ¯)=xikfor iI and k = 1, 2; note that this abacus configuration does uniquely define both a¯ and μ¯ and that δμa=δμ¯a¯, so that if B¯ is the a¯-equivalence class containing μ¯ then (a¯,b​​​​°) is a reduced pair for B¯. Then μΛs if and only if μ¯Λa¯.

Lemma 3.8.

Suppose that B is a core block and r=(a,b​​​​°) is a reduced pair for B. Let μB. Suppose that i1,iI with i1<i and let ν be the bipartition whose Young diagram is formed by removing all removable i-nodes from [μ]. Then unless (δμ)i1=+ and (δμ)i= we have that μΛs if and only if νΛs.

Proof.

Since r=(a,b​​​​°), each component of [μ] has at most one removable i-node. If [μ] has no removable i-nodes, the lemma is trivial. So suppose that [μ] has one removable i-node. Then the i-residue sequence of μ is either r, ar or ra. In the first two cases, this removable node is good, and so the lemma holds. In the third case, the first component of [μ] has a removable i-node and the second has an addable i-node. Given r=(a,b​​​​°), this is only possible if (δμ)i1=+ and (δμ)i=.

Now suppose that [μ] has two removable i-nodes so the i-residue sequence of μ is rr and the removable i-node in the first component is good. Removing this node gives a bipartition, σ say, with μΛs if and only if σΛs. By the last paragraph, the one removable i-node in [σ] is good, so that σΛs if and only if νΛs. □

Lemma 3.9.

Suppose that B is a core block with |B|>1 and (a,b​​​​°) is a reduced pair for B. Let μB. There is a bipartition ν which satisfies the following conditions.

  • ν lies in a core block B and (a,b​​​​°) is a reduced pair for B.

  • δν is essentially the same as δμ.

  • There exist i1,i2I with i1i2 such that

    • For k = 1, 2, lmka(ν)=1 for all m<i1 and lmka(ν)=0 for all m>i2.

    • (δν)m{,+} for i1mi2.

Then μΛs if and only if νΛs.

Before proving this result, we give an example of bipartitions μ and ν which satisfy the lemma.

Example.

Below, we have δμ=(+,0,+,0,,,+,0,+,0,) and δν=(0,0,+,+,,,+,+,,0,0).

Proof of Lemma 3.9.

Suppose there exist i,jI with j < i and lika(μ)=1 for k = 1, 2 while ljka(μ)=0 for k = 1 or k = 2. Choose i minimal with this condition. Then following Lemma 3.8, we may remove all the removable i-nodes of [μ] to obtain the Young diagram of a bipartition σ˜ with σ˜Λs if and only if μΛs. Then the abacus configuration of σ˜ is formed from that of μ by swapping runners i – 1 and i, so that δσ˜ is formed from δμ by swapping the entries in the positions indexed by i – 1 and i; note that (δμ)i=0 so that δμ and δσ˜ are essentially the same. We continue in this way until we reach a bipartition σ where there do not exist i, j as above.

Now consider σ. Suppose there exist i,jI with i<j and lika(σ)=0 for k = 1, 2 while ljka(σ)=1 for k = 1 or k = 2. Choose i maximal with this property. Again following Lemma 3.8, we may remove all the removable (i+1)-nodes of [σ] to obtain the Young diagram of a bipartition ν˜ with ν˜Λs if and only if σΛs. Similarly to the case above, δν˜ is formed from δσ by swapping the entries in the positions indexed by i and i+1, and (δσ)i=0 so that δσ and δν˜ are essentially the same. We continue in this way until we reach a bipartition ν where there do not exist i,j as above. Then ν satisfies the conditions of the lemma and μΛs if and only if νΛs. □

Proposition 3.10.

Suppose that B is a core block and μB. Then μΛs if and only if δμΔ0.

Proof.

Suppose that (a,b) is a reduced pair for μ. By Lemma 3.7, we may assume that b=b​​​​°. If |B|=1 then the block B̂ is simple and so μΛs. So assume that |B|>1. By Lemma 3.9, we may assume that there exist i1,i2I with i1i2 such that

  • For k = 1, 2, lmka(μ)=1 for all m<i1 and lmka(μ)=0 for all m>i2.

  • (δμ)m{,+} for i1mi2.

We prove Proposition 3.10 by induction on |S(δμ)|. First suppose that |S(δμ)|=0. If δμΔ0 then |B|=1, which we have assumed is not true. So suppose that δμΔ0 so that there are entries equal to both – and + in δμ with the + terms occurring before the – terms. The abacus configuration of μ with respect to a looks like:

Then μ has only one removable node which is not good, hence μΛs.

Now suppose that |S(δμ)|>0 and Proposition 3.10 holds for any σ in a core block with |S(δσ)|<|S(δμ)|. Then there exist i,i+1I with i<i+1 and (δμ)i= and (δμ)i+1=+. By Lemma 3.8, we can remove the removable i + 1-node from [μ] to obtain the Young diagram of a bipartition σ with σΛs if and only if μΛs. Furthermore, we have (δσ)i=(δσ)i+1=0 and (δσ)m=(δμ)m for mi,i+1 so that δσΔ0 if and only if δμΔ0. Hence μΛsσΛsδσΔ0δμΔ0where the middle step follows from the inductive hypothesis. □

In this section, we have shown that if B is a core block then we can index the Specht modules corresponding to bipartitions in B by a subset of the set of sequences δΔ with nB entries equal to – and pB entries equal to +, with the simple modules indexed by sequences δΔ0. The next step is to determine the entries in the block decomposition matrix. To do this, we begin by introducing the Fock space representation of Uv(sl̂e).

3.3 The Fock space representation

Let U denote the quantized enveloping algebra U=Uv(sl̂e). This is a Q(v)-algebra with generators ei, fi for iI and vh for hP; the relations may be found in [Citation18]. Let Fs be the Q(v)-vector space with basis {sλ|λΛ2}. This becomes a U-module under the action described in [Citation18]; we call Fs the Fock space representation of U. The U-submodule Ms generated by s2 (where 2 is the unique bipartition of 0) is isomorphic to the irreducible highest weight module V(Υ) for some dominant integral weight Υ of U. This module has a canonical basis (in the sense of Lusztig and Kashiwara) which is indexed by the elements of Λs; we write Pμ for the canonical basis element indexed by μΛs. We define the transition coefficients dλμ(v) so that they satisfy the equation Pμ=λΛ2dλμ(v)sλ;then dλμ(v)=0 unless λsμ. Ariki’s Theorem and its graded analogue, below, relate the transition coefficients to the graded decomposition numbers.

Theorem 3.11.

[Citation1, Citation5] Suppose that Hn=H2,n(q,Q) is defined over a field of characteristic 0. Suppose that λ,μΛnr with μΛs. Then [Sλ:Dμ]v=dλμ(v).

The transition coefficients give us the decomposition numbers when p = 0 and can be considered as a first approximation to the decomposition numbers when the field is arbitrary; as we shall see below, there are situations when this first approximation is correct. In order to compute the transition coefficients, we now consider the action of fiU on a basis element sλFs.

If d > 0 and ν,λΛ2, write νd:iλ if [λ] is formed from [ν] by adding d nodes all of residue iI. If νd:iλ set N(ν,λ)=n[λ][ν]#{man addable i-node of ν with nCm}#{mare movable i-node of λ with nCm}.

For d > 0 and iI, define fi(d)=fid / [d]!U, the quantum divided power of fid. Then if νΛ2, fi(d)sν=νd:iλvN(ν,λ)sλ.

Lemma 3.12.

[Citation22, Proposition 2.3] Suppose νΛs is such that if σsν then [Sσ:Dν]v=dσν(v). Let i1,,itI and d1,,dt>0. Suppose fit(dt)fi1(d1)Pν=sμ+λμcλμ(v)sλwhere cλμ(v)vN[v] for all λμ. Then μΛs and if λμ then [Sλ:Dμ]v=cλμ(v).

The statement of Lemma 3.12 given in [Citation22] is only for the case when Pν=sν, but the proof for the slightly more general situation above is identical. Suppose i,jI. If ij, define fi,j=fjfj1fi+1.

If e is finite and i > j, define fi,j=fjfj1f0fe1fi+2fi+1.

Suppose that B is a core block and that (a,b) is a reduced pair for B. We say that b is flat if e= or if e is finite and there exists mI and bZ such that bm = b if mm and bm=b+1 otherwise. Note that if e is finite then in this case, mm+1e101m1.

Lemma 3.13.

Suppose that B is a core block and that r=(a,b) is a reduced pair for B with b flat. Let νB. Suppose that ij with lika(ν)=bi+1,ljka(ν)=bj,lm1a(ν)=lm2a(ν), for k=1,2 and imj.

Let I1={imj|lm1a(ν)=lm2a(ν)=bm+1}={i1,i2,,it}, where i1it. Set it+1=j and define f=fi1,i2fit1,itfit,it+1U.

Then fsν=sλ2+vsλ1where lmka(λ1) ={lmka(ν)1,m=i and k=1,lmka(ν)+1,m=j and k=1,lmka(ν),otherwise, lmka(λ2) ={lmka(ν)1,m=i and k=2,lmka(ν)+1,m=j and k=2,lmka(ν),otherwise, so that (δλ1)m ={+,m=i,,m=j,(δν)m,otherwise, (δλ2)m ={,m=i,+,m=j,(δν)m,otherwise.

Proof.

For 1st and k = 1, 2, define σks by lmka(σks)={lmka(ν)1,m=is and k=k,lmka(ν)+1,m=j and k=k,lmka(ν),otherwise,

so that λk=σk1 for k = 1, 2. We claim that if 1st then fis,is+1fit,it+1sν=sσ2s+vsσ1s.

Since b is flat, [ν] has exactly two addable (it+1)-nodes, one in the first component and one in the second, and no removable (it+1)-nodes. If we add an (it+1)-node to component k, for k = 1, 2, then the subsequent bipartition has exactly one addable (it+2)-node, in component k, and no removable (it+2)-nodes. Continuing in this way, we see that fit,it+1sν=sσ2t+vsσ1t.

Now suppose 1st1 and consider fis,is+1sσks+1, where k = 1, 2. If is+1=is+1 then [σks+1] has exactly one addable is+1-node, which is in component k, and no removable is+1-nodes, and so fis+1sσks+1=sσks. Otherwise, as above, fis,is+11sσks+1=sτ2+vsτ1,where [τk] is obtained by successively adding nodes to component k of [σks+1], for k=1,2. However, if kk,[τk] does not have any addable is+1-nodes so fis+1sτk=0. If k=k=1 (resp. k=k=2) then [τk] has an addable is+1-node on component 1 (resp. on component 2) and a removable is+1-node on component 2 (resp. on component 1) and so fis+1sτk=v1sσk (resp. fis+1sτk=sσk). So in both cases we have fis,is+1sσks+1=sσks.

The lemma follows. □

3.4 Decomposition matrices

Theorem 3.14.

Let B be a core block and suppose that r=(a,b) is a reduced pair for B with b flat. Let λ,μB with μΛa. Then [Sλ:Dμ]v=dλμ(v)={vl(δλ,δμ),δλ[graphic]δμ,0,otherwise.

Proof.

We work by induction on |S(μ)|=mB. If |S(μ)|=0 then B is simple and the result holds. So suppose |S(μ)|>0 and the theorem holds for all core blocks B with mB<mB. Choose (i,j)S(μ) such that there does not exist (i,j)S(μ) with i<i<j<j. Let ν be the bipartition defined by lmka(ν)={lmka(μ)+1,m=i and k=2,lmka(μ)1,m=j and k=2,lmka(μ),otherwise.

Then νB for some core block B such that (a,b) is a reduced pair for B. Also S(ν)=S(μ){(i,j)} so νΛs. Hence ν satisfies the conditions of the inductive hypothesis, so by Theorem 3.11, Pν=δσ[graphic]δνvl(σ,ν)sσ.

The bipartition ν with reduced pair (a,b), together with the pair (i, j) described above, satisfies the conditions of Lemma 3.13. We define f as in that lemma and compute each term fsσ that appears in the sum above, so that fPν=δλ[graphic]δμvl(λ,μ)sλ.

Since l(λ,μ)>0 for all λμ which appear in the sum, we have satisfied the conditions of Lemma 3.12 and so the theorem holds. □

Since there is always a reduced pair (s,b) with b flat when e= or n < e, this describes the block decomposition matrices for core blocks in these cases, recovering previous results; we note in particular the similarity between Theorem 3.14 and [Citation19, Theorem 3]. For the rest of this section, we assume that e is finite. Suppose that B and B are any two core blocks with nB=nB and pB=pB. Then we have a bijection Φ:BB such that δλ and δΦ(λ) are essentially the same for all λB. If μB then by Lemma 3.10, we have that μΛs if and only if Φ(μ)Λs.

Given two core blocks B and B, we say that B and B are δ-equivalent if nB=nB and pB=pB and if the map Φ defined above preserves the block decomposition matrix of B, that is, if λ,μB with μΛs then [Sλ:Dμ]v=[SΦ(λ):DΦ(μ)]v. In fact, we will show that any two core blocks B and B with nB=nB and pB=pB are δ-equivalent. Theorem 3.14 shows that this also holds when e=.

Lemma 3.15.

Let B be a core block and suppose that r=(a,b) is a reduced pair for B. Suppose that iI and bi+1>bi+xi where xi={1,i=e1,0,otherwise.

Define a map Φi:BΛ2 which sends λB to the bipartition whose Young diagram is obtained by removing all possible (i+1)-nodes from [λ]. Then Φi gives a bijection between B and a core block B such that if λB then δλ=δΦi(λ).

Proof.

Suppose λB. From our assumptions on b,[λ] has no addable (i+1)-nodes. If νB then by Lemma 3.1, lm1s(λ)+lm2s(λ)=lm1s(ν)+lm2s(ν) for all mI so that the number of removable (i+1)-nodes on both [λ] and [ν] is li+11s(λ)li1s(λ)xi+li+12s(λ)li2s(λ)xi=li+11s(ν)li1s(ν)xi+li+12s(ν)li2s(ν)xi.

Hence Φi(λ) and Φi(ν) lie in the same block, B say. If ie1,Φi acts on the abacus configurations by swapping the runners i and i + 1 on each component, so that δλ=δΦi(λ) and so by Corollary 3.2, Φi is a bijection between B and B. If i=e1, then (a,b) is a reduced pair for B, where bm={be1+1,m=0,b01,m=e1,bm,otherwise.

It is straightforward to check that if m0,e1 then bmbe1bmb0,be1bmb0b0bmbe1,b0bmbe1bj,so that again δλ=δΦi(λ) and Φi is a bijection between B and B. □

Lemma 3.16.

Suppose that we have the conditions of Lemma 3.15. Keep the notation of that lemma so that B=Φi(B). Suppose B is such that if σ,τB with τΛs then [Sσ:Dτ]v=dστ(v). Then B and B are δ-equivalent.

Proof.

Suppose that if λB then λ has t removable (i+1)-nodes; equivalently if σB then σ has t addable (i+1)-nodes and no removable (i+1)-nodes. Let Ψi=Φi1. If σB then fi+1(t)sσ=sΨi(σ).

Let τB be such that τΛs. Then fi+1(t)Pτ=fi+1(t)σdστ(v)sσ=σdστ(v)sΨi(σ).

Hence by repeated application of Lemma 3.12, [SΨi(σ):DΨi(τ)]v=dστ(v)=[Sσ:Dτ]v for σB. Since δσ=δΨi(σ) for all σB, the lemma follows. □

Lemma 3.17.

Suppose that B is a core block with reduced pair (a,b). Then there exists a core block B with reduced pair (a,b) such that B and B are δ-equivalent and b is flat.

Proof.

Following Lemma 3.15, if there exists iI  {e1} with bi<bi+1 then we may apply the map Φi to B. If b0be1>1 then we may apply the map Φe1 to B. By repeatedly applying these maps, we obtain a core block B=Φi1Φi2ΦizB where (a,b) is a reduced pair for B and b is flat. By Theorem 3.14, [Sσ:Dτ]v=dστ(v) for all σ,τB with τΛa. Hence by Lemma 3.16, we have that B and B are δ-equivalent. □

Corollary 3.18.

Suppose that B and B are core blocks. If nB=nB and pB=pB then B and B are δ-equivalent.

Theorem 3.19.

Let B be a core block and suppose λ,μB with μΛs. Then [Sλ:Dμ]v={vl(δλ,δμ),δλ[graphic]δμ,0,otherwise.

Proof.

Following Lemma 3.17, we know that B is δ-equivalent to a core block B with reduced pair (a,b) where b is flat. The theorem then follows from Theorem 3.14. □

3.5 Summary and examples

We collect together the main results of this paper.

Theorem 3.20.

Let B be a core block and let B0=BΛa. Let n=nB, p=pB and m=min{n,p}.

  1. We can identify a bipartition λB with a sequence δλΔ which contains n entries equal to – and p entries equal to +. If νB then δν is formed by permuting these n + p entries. Hence

    |B|=(n+pm).

  2. Suppose μB. Then μB0 if and only if δμΔ0. Hence

    |B0|=(n+pm)(n+pm1).

  3. We have wt(B)=m.

  4. There exists dZ such that if s=(s1,s2) then if e is finite (resp. e=) then s1n+dmode and s2p+dmode (resp. s1=n+d and s2=p+d).

  5. 5. Suppose that λ,μB with μB0. Then

    [Sλ:Dμ]v={vl(δλ,δμ),δλ[graphic]δμ,0,otherwise.

Proof.

All these results appear in the previous section, apart from the formula for |B0|. To see that this holds, first suppose that np. Identify each sequence with n entries equal to – and p entries equal to + with a (n, p)-tableau by putting the entries corresponding to – in the top row and to + in the bottom row. The tableau is standard if and only if the sequence lies in Δ0. Since the number of standard (n, p)-tableaux is given by the formula above, we are done. The case that n < p follows by symmetry. □

For λ,ν in a core block B, let dim(Hom(Sλ,Sν))v denote the graded dimension of the homomorphism space between Sλ and Sν. The next result appears in George Witty’s 2020 thesis [Citation27]. Although Witty only proves the case where (a,b​​​​°) is a reduced pair for B, he conjectures that the result holds for an arbitrary reduced pair [Citation27, Conjecture 4.32]. It is possible that one could prove this conjecture using Scopes equivalence [Citation26].

Theorem 3.21.

[Citation27, Theorem 4.27] Suppose that B is a core block and that (a,b​​​​°) is a reduced pair for B. Suppose that e2 or that nBpB. Let λ,νB. Then dim(Hom(Sλ,Sν))v={vl(δλ,δν),δλ[graphic]δν,0,otherwise.

If e = 2 or nB = pB then the formula above gives a lower bound for dim(Hom(Sλ,Sν))v.

Witty’s theorem is an application of results proved in his thesis about homomorphisms between Specht modules. His proofs involve very intricate calculations within the cyclotomic KLR algebras.

It is likely that a formula similar to that of Theorem 3.20 (5) holds for the graded decomposition numbers for the type B Schur algebras when μB0 and it seems to us that an argument via induction on |S(δμ)|, using the techniques developed in this paper, would prove the result. We thank the referee for asking this question, since our assumption when writing the paper was that our methods would not work for the Schur algebras.

We end this paper with some examples.

Example.

Suppose that e = 5 and s=(3,1). Let λ=((7,5,4,32,2,12),(14,10,6,4,3,22,1)) and let B be the core block containing λ. Then ((18,16),(2,4,3,1,5)) is a reduced pair for B. By drawing the abacus configurations, we can see that nB = 3 and pB = 1. The four bipartitions in the block are given by the abacus configurations below.

Then the decomposition matrix for B is λ1(,0,,,+)1λ2(,0,,+,)v1λ3(,0,+,,)v1λ4(+,0,,,)v

Example.

Suppose that B is a core block with nB = 2 and pB = 3. Then |B|=10 and by removing the 0s from the sequences {δλ|λB} we get Specht modules indexed by the 10 elements below. The block decomposition matrix for B is equal to (,,+,+,+)1(,+,,+,+)v1(+,,,+,+)v1(,+,+,,+)v1(+,,+,,+)vv2vv1(+,+,,,+)v2v(,+,+,+,)v(+,,+,+,)v2v(+,+,,+,)vv2(+,+,+,,)v2

Note that this is the matrix that appears in [Citation19, Example 6].

References

  • Ariki, S. (1996). On the decomposition numbers of the Hecke algebra of G(r,1,n). J. Math Kyoto Univ. 36:789–808.
  • Bowman, C., de Visscher, M., Hazi, A., Stroppel, C. Quiver presentations and isomorphisms of Hecke categories and Khovanov arc algebras. arXiv: 2309.13695.
  • Bowman, C., Dell’Arciprete, A., De Visscher, M., Hazi, A., Muth, R., Stroppel, C. Quiver presentations and Schur-Weyl dualities for Hecke categories and Khovanov arc algebras. in preparation.
  • Brundan, J., Kleshchev, A. (2009). Blocks of cyclotomic Hecke algebras and Khovanov-Lauda algebras. Invent. Math. 178:451–484. DOI: 10.1007/s00222-009-0204-8.
  • Brundan, J., Kleshchev, A. (2009). Graded decomposition numbers for cyclotomic Hecke algebras. Adv. Math. 222:1883–1942. DOI: 10.1016/j.aim.2009.06.018.
  • Brundan, J., Kleshchev, A., Wang, W. (2011). Graded Specht modules. J. Reine Angew. Math. 655:61–87.
  • Brundan, J., Stroppel, C. (2011). Highest weight categories arising from Khovanov’s diagram algebra III: category O. Rep. Theory 15:170–243.
  • Dell’Arciprete, A. Equivalence of decomposition matrices for blocks of Ariki-Koike algebras. arXiv: 2301.05153.
  • Dipper, R., Mathas, A. (2002). Morita equivalences of Ariki-Koike algebras. Math. Z. 240:579–610. DOI: 10.1007/s002090100371.
  • Dipper, R., James, G. D., Mathas, A. (1999). Cyclotomic q–Schur algebras. Math. Z. 229:385–416. DOI: 10.1007/PL00004665.
  • Fayers, M. (2007). Core blocks of Ariki-Koike algebras. J. Algebraic Comb. 26:47–81. DOI: 10.1007/s10801-006-0048-x.
  • Fayers, M. (2006). Weights of multipartitions and representations of Ariki-Koike algebras. Adv. Math. 06:112–133. DOI: 10.1016/j.aim.2005.07.017.
  • Graham, J., Lehrer, G. I. (1996). Cellular algebras. Invent. Math. 123:1–34. DOI: 10.1007/BF01232365.
  • Hu, J., Mathas, A. (2015). Quiver Schur algebras for linear quivers. Proc. LMS (3) 110:1315–1386. DOI: 10.1112/plms/pdv007.
  • Khovanov, M., Lauda, A. D. (2009). A diagrammatic approach to categorification of quantum groups I. Represent. Theory 13:309–347. DOI: 10.1090/S1088-4165-09-00346-X.
  • Khovanov, M., Lauda, A. D. (2011). A diagrammatic approach to categorification of quantum groups II. Trans. Amer. Math. Soc. 363:2685–2700. DOI: 10.1090/S0002-9947-2010-05210-9.
  • Kleshchev, A. (2010). Representation theory of symmetric groups and related Hecke algebras. Bull. Amer. Math. Soc. 47:419–481. DOI: 10.1090/S0273-0979-09-01277-4.
  • Lascoux, A., Leclerc, B., Thibon, J.-Y. (1996). Hecke algebras at roots of unity and crystal bases of quantum affine algebras. Commun. Math. Phys. 181:205–263. DOI: 10.1007/BF02101678.
  • Leclerc, B., Miyachi, H. (2004). Constructible characters and canonical bases. J. Algebra 277:298–317. DOI: 10.1016/j.jalgebra.2004.02.023.
  • Lyle, S. Rouquier blocks for Ariki-Koike algebras, arXiv: 2206.14720.
  • Lyle, S., Mathas, A. (2007). Blocks of cyclotomic Hecke algebras. Adv. Math. 216:854–878. DOI: 10.1016/j.aim.2007.06.008.
  • Lyle, S., Ruff, O. (2016). Graded decomposition numbers of Ariki–Koike algebras for blocks of small weight. J. Pure App. Algebra 220:2112–2142. DOI: 10.1016/j.jpaa.2015.10.019.
  • Mathas, A. (2004). The representation theory of the Ariki–Koike and cyclotomic q-Schur algebras. In: Representation Theory of Algebraic Groups and Quantum Groups, Advanced Studies in Pure Mathematics, Vol. 40. Tokyo: Mathematical Society of Japan, pp. 261–320.
  • Rouquier, R. (2008). 2-Kac-Moody algebras. arXiv:0812.5023.
  • Scopes, J. (1991). Cartan matrices and Morita equivalence for blocks of symmetric groups. J. Algebra 142:441–455. DOI: 10.1016/0021-8693(91)90319-4.
  • Webster, B. RoCK blocks for affine categorical representations. arXiv: 2301.01613.
  • Witty, G. (2020). Homomorphisms between Specht modules of KLR algebras. PhD Thesis, University of East Anglia.