Home On signed Young permutation modules and signed p-Kostka numbers
Article Publicly Available

On signed Young permutation modules and signed p-Kostka numbers

  • Eugenio Giannelli EMAIL logo , Kay Jin Lim , William O’Donovan and Mark Wildon
Published/Copyright: February 25, 2017

Abstract

We prove the existence and main properties of signed Young modules for the symmetric group, using only basic facts about symmetric group representations and the Broué correspondence. We then prove new reduction theorems for the signed p-Kostka numbers, defined to be the multiplicities of signed Young modules as direct summands of signed Young permutation modules. We end by classifying the indecomposable signed Young permutation modules and determining their endomorphism algebras.

1 Introduction

Let F be a field of odd prime characteristic p and let 𝔖n denote the symmetric group of degree n. In this article we investigate the modular structure of the p-permutation F𝔖n-modules defined by inducing a linear representation of a Young subgroup of 𝔖n to 𝔖n.

Let 𝒫2(n) be the set of all pairs of partitions (α|β) such that |α|+|β|=n. For (α|β)𝒫2(n), the signed Young permutation moduleM(α|β) is the F𝔖n-module defined by

(1.1)M(α|β)=Ind𝔖α×𝔖β𝔖n(F(𝔖α)sgn(𝔖β)).

In [7, p. 651], Donkin defines a signed Young module to be an indecomposable direct summand of a signed Young permutation module and proves the following theorem.

Theorem 1.1 (Donkin [7]).

There exist indecomposable FSn-modules Y(λ|pμ) for (λ|pμ)P2(n) with the following properties:

  1. if (α|β)𝒫2(n), then M(α|β) is isomorphic to a direct sum of modules Y(λ|pμ) for (λ|pμ)𝒫2(n) such that (λ|pμ)(α|β),

  2. [M(λ|pμ):Y(λ|pμ)]=1,

  3. if λ=i=0rpiλ(i) and μ=i=0r-1piμ(i) are the p-adic expansions of λ and μ , as defined in (2.1), then Y(λ|pμ) has as a vertex a Sylow p-subgroup of the Young subgroup 𝔖ρ, where ρ is the partition of n having exactly |λ(i)|+|μ(i-1)| parts of size pi for each i{0,,r}.

Here (λ|pμ)(α|β) refers to the dominance order on 𝒫2(n), as defined in Section 2.3 below and, in (iii), μ(-1) should be interpreted as the partition of 0.

Donkin’s definition of signed Young modules and his proof of his theorem use the Schur superalgebra. In Section 4 we give an independent proof using only basic facts about symmetric group representations and the Broué correspondence for p-permutation modules; our proof shows that the Y(λ|pμ) may be defined by Definition 4.11. (Theorem 1.1 characterizes the signed Young module Y(λ|pμ) as the unique summand of M(λ|pμ) appearing in M(α|β) only if (λ|pμ)(α|β), so the two definitions are equivalent.) As a special case we obtain the existence and main properties of the Young modules, which we define by Yλ=Y(λ|). These are precisely the indecomposable summands of the Young permutation modulesMα=M(α|). We state this result, and discuss the connection with [10], and with the original definition of Young modules via the Schur algebra [18], in Section 5.1.

In [14], Hemmer conjectured, motivated by known results on tilting modules for Schur algebras, that the signed Young modules are exactly the self-dual modules for symmetric groups with Specht filtrations. This was shown to be false in [23]; the fourth author later proved in [26] that if n66 and G is a subgroup of 𝔖n such that the ordinary character of M=IndG𝔖nF is multiplicity free, then every indecomposable summand of M is a self-dual module with a Specht filtration. Despite the failure of Hemmer’s conjecture, it is clear that signed Young modules are of considerable interest. In particular, a strong connection between simple Specht modules and signed Young modules has been established by Hemmer [14] and by Danz and the second author [6]. More precisely, Hemmer showed that every simple Specht module is isomorphic to a signed Young module, and Danz and the second author established their labels.

In Section 6 we study signed p-Kostka numbers, defined to be the multiplicities of signed Young modules as direct summands of signed Young permutation modules. These generalize the p-Kostka numbers considered in [11, 12, 15, 16]. Given the p-Kostka numbers for 𝔖n it is routine to calculate the decomposition matrix of 𝔖n in characteristic p (see [12, Section 3]). It is therefore no surprise that a complete understanding of the p-Kostka numbers seems to be out of reach. However, as the references above demonstrate, many partial results and significant advances have been obtained. Our first main theorem is a relation between signed p-Kostka numbers. We refer the reader to Notation 3.8 for the definitions of the composition 𝜹(0) and the set Λ((α|β),ρ).

Theorem 1.2.

Let (α|β),(λ|pμ)P2(n). Then

[M(pα|pβ):Y(pλ|p2μ)][M(α|β):Y(λ|pμ)].

Furthermore, if 𝛅(0)= for all (𝛄|𝛅)Λ((α|β),ρ), then equality holds.

Example 6.4 shows that strict inequality may hold in Theorem 1.2. This is an important fact, since it appears to rule out a routine proof of Theorem 1.2 using the theory of weights for the Schur superalgebra: we explain this obstacle later in the introduction. However, in Corollary 6.3, we obtain the following asymptotic stability of the signed p-Kostka numbers:

[M(α|β):Y(λ|pμ)][M(pα|pβ):Y(pλ|p2μ)]=[M(p2α|p2β):Y(p2λ|p3μ)]=.

If β=, then the condition on 𝜹(0) holds for all (𝜸|𝜹)Λ((α|β),ρ) and Theorem 1.2 specializes to Gill’s result [12, Theorem 1] that [Mpα:Ypλ]=[Mα:Yλ] for all partitions α and λ of n.

Our second main theorem describes the relation between signed p-Kostka numbers for partitions differing by a p-power of a partition. Let 𝒞2(m) be the set consisting of all pairs of compositions (α|β) such that |α|+|β|=m. We refer the reader to equation (5.1) in Section 5.2 for the definition of p(λ|pμ).

Theorem 1.3.

Let m, n and k be natural numbers. Let (π|π~)C2(m), (λ|pμ)P2(m), (ϕ|ϕ~)C2(n) and (α|pβ)P2(n). If k>p(λ|pμ), then

[M(π+pkϕ|π~+pkϕ~):Y(λ+pkα|p(μ+pkβ))][M(π|π~):Y(λ|pμ)][M(pϕ|pϕ~):Y(pα|p2β)].

Moreover, if pk>max{π1,π~1}, then equality holds.

In particular, taking ϕ=α=(r) and ϕ~=β=, we see that

[M(π+pk(r)|π~):Y(λ+pk(r)|pμ)]M[(π|π~):Y(λ|pμ)]

with equality whenever pk>max{π1,π~1}.

Our third main theorem classifies the indecomposable signed Young permutation modules.

Theorem 1.4.

Let (α|β)P2(n). The signed Young permutation module M(α|β) is indecomposable if and only if one of the following conditions holds.

  1. (α|β)=((m)|(n)) for some non-negative integers m,n such that either

    1. m=0,

    2. n=0, or

    3. m+n is divisible by p.

  2. (α|β) is either ((kp-1,1)|) or (|(kp-1,1)) for some k.

In cases (i) (a) and (i) (b), we have EndFSnM(α|β)F. In the remaining cases we have EndFSnM(α|β)F[x]/x2.

In particular, Theorem 1.4 classifies all indecomposable Young permutation modules up to isomorphism, recovering [12, Theorem 2] for fields of odd characteristic. Note that the Young permutation module M(n-1,1)=M((n-1,1)|)=M((n-1)|(1)) appears in both parts (i) and (ii). If M(α|β) is indecomposable, then there exist unique partitions λ and μ such that M(α|β)Y(λ|pμ). These partitions are determined in Proposition 7.1.

Schur algebras

Our results may be applied to obtain corollaries on modules for the Schur algebra. Fix n,d with dn and let GLd(F) be the general linear group of d×d matrices over F. Let ρ:GLd(F)GLm(F) be a representation of GLd(F) of dimension m. We say that ρ is a polynomial representation of degree n if the matrix coefficients ρ(X)ij for each i,j{1,,m} are polynomials of degree n in the coefficients of the matrix X. Given a polynomial representation ρ:GLd(F)GL(V) of degree n, the image of V under the Schur functorf is the subspace of V on which the diagonal matrices diag(a1,,ad)GLd(F) act as a1an. It is easily seen that f(V) is preserved by the permutation matrices in GLd(F) that fix the final d-n vectors in the standard basis of Fd. Thus f(V) is a module for F𝔖n.

The category of polynomial representations of GLd(F) of degree n is equivalent to the category of modules for the Schur algebra SF(d,n). We refer the reader to [13] for the definition of SF(d,n) and further background. In this setting, the Schur functor may be defined by VeV, where eSF(d,n) is an idempotent such that eSF(d,n)eF𝔖n. It follows that f is an exact functor from the category of polynomial representations of GLd(F) of degree n to the category of F𝔖n-modules.

Let E denote the natural GLd(F)-module. Given α𝒫(n), let Symα(E) and β(E) denote the corresponding divided symmetric and exterior powers of E, defined as quotient modules of En. The mixed powersSymαEβE for (α|β)𝒫2(n) generate the category of GLd(F)-modules of degree n. In [7], Donkin defines a listing module to be an indecomposable direct summand of a mixed power. (As the nautical parlance suggests, listing modules generalize tilting modules). By [7, Proposition 3.1 c], for each (λ|μ)𝒫2(n) there exists a unique listing module List(λ|pμ) such that f(List(λ|pμ))Y(λ|pμ). By [7, Proposition 3.1a], we have

f(SymαEβE)M(α|β).

Moreover, by [7, Proposition 3.1b], the Schur functor induces an isomorphism

EndGLd(F)(SymαEβE)End𝔖n(M(α|β)).

Thus each of our three main theorems has an immediate translation to a result on multiplicities of listing modules in certain mixed powers. For example Theorem 1.4 classifies the indecomposable GLd(F)-mixed powers and shows that each has an endomorphism algebra, as a GLd(F)-module, of dimension at most 2. It is also worth noting that many of Gill’s results from [12] are reproved in greater generality in the Schur algebra setting in a recent paper of Donkin [8].

Steinberg tensor product formula

As Gill remarks in [12], some of his results can be obtained using weight spaces and the Steinberg Tensor Product Theorem for irreducible representations of the group GLd(F). We explain the connection here, since this remark is also relevant to this work. Let α be a composition of n where dn and let ξαSF(d,n) be the idempotent defined in [13, Section 3.2] such that ξαV is the α-weight space, denoted Vα, of the SF(d,n)-module V; the idempotent e defining the Schur functor is ξ(1n). For λ a partition of n, let L(λ) denote the irreducible representation of GLd(F) with highest weight λ, thought of as a module for SF(d,n). Let Proj(λ) be the projective cover of L(λ). By James’ original definition of Young modules (this is shown to be equivalent to ours in Section 5.1), we have Yλ=f(Proj(λ)); moreover,

[Mα:Yλ]=[Symα(E):Proj(λ)]=dimFHom(S(d,n)ξα,L(λ))=dimFξαL(λ).

(Here Symα(E)En is the contravariant dual, as defined in [13, 2.7a], of the quotient module Symα(E) of En.) Thus

(1.2)[Mα:Yλ]=dimFL(λ)α.

As an example of this relationship between p-Kostka numbers and dimensions of weight spaces, we use (1.2) to deduce Theorem 1 in [12]. By the Steinberg Tensor Product Theorem, L(pλ)=L(λ), where is the Frobenius map, acting on representing matrices by sending each entry to its pth power. Clearly there is a canonical vector space isomorphism (L(λ))pαL(λ)α. Therefore

[Mpα:Ypλ]=dimFL(pλ)pα=dimFL(λ)α=[Mα:Yλ]

as required.

Schur superalgebras

Our Theorem 1.2 generalizes the result just proved, so it is natural to ask if it can be proved in a similar way, replacing the Schur algebra with the Schur superalgebra defined in [7]. Let a,b. Given (λ|pμ)𝒫2(n) where λ has at most a parts and μ has at most b parts, let L(λ|pμ) denote the irreducible module of highest weight (λ|pμ) for the Schur superalgebra S(a|b,n), defined in [7, p. 661]. By [7, Section 2.3], we have

(1.3)[M(α|β):Y(λ|pμ)]=dimFL(λ|pμ)(α|β)

generalizing (1.2).

Let GL(a|b) denote the super general linear group defined in [4, Section 2]. As En is a generator for the category of polynomial representations of GL(a|b) of degree n, it follows from [7, p. 660, (1)] that the category of such modules is equivalent to the module category of S(a|b,n). Taking the even degree part of GL(a|b) recovers GLa(F)×GLb(F). (More precisely, the even degree part is isomorphic to the product of the affine group schemes corresponding to these two general linear groups.) The Frobenius map is identically zero on the odd degree part of GL(a|b), so induces a map :GL(a|b)GLa(F)×GLb(F). Let be the corresponding inflation functor, sending modules for GLa(F)×GLb(F) to modules for GL(a|b). By [4, Remark 4.6 (iii)] we have

L(pλ|pμ)=(L(λ)L(μ)),

where denotes an outer tensor product. Taking weight spaces we get

L(pλ|pμ)(pα|pβ)L(λ)αL(μ)β.

By (1.3) we have [M(pα|pβ):Y(pλ|pμ)]=dimFL(λ)αdimFL(μ)β. Replacing μ with pμ and applying the Steinberg Tensor Product Formula, this implies the asymptotic stability of signed p-Kostka numbers mentioned after Theorem 1.2.

Stated for GL(a|b)-modules, the remaining part of Theorem 1.2 becomes

dimFL(pλ|pμ2)(pα|pβ)dimFL(λ|pμ)(α|β).

This does not follow from the results mentioned so far, or from the version of the Steinberg Tensor Product Theorem for GL(a|b)-modules proved in [22], because the module on the right-hand side is not an inflation. Moreover, translated into this setting, a special case of Example 6.4 shows that dimFL((1)|)(|(1))=1 whereas dimFL((p)|)(|(p))=dimFL((p))dimFL()(p)=0, so it is certainly not the case that equality always holds. (Further examples of this type are given by the general case of Example 6.4.) Whether or not a proof using supergroups is possible, the authors believe that since Theorem 1.2 can be stated within the context of symmetric groups, it deserves a proof in this setting.

Klyachko’s multiplicity formula

Klyachko’s multiplicity formula [21, Corollary 9.2] expresses the p-Kostka number [Mα:Yλ] in terms of p-Kostka numbers for p-restricted partitions. Our Corollary 5.2 gives a generalization to signed Young modules. Specializing this result we obtain a symmetric group proof of Klyachko’s formula in the form

(1.4)[Mα:Yλ]=(𝜸|)Λ((α|),ρ)i=0r[M𝜸(i):Yλ(i)],

where λ=i=0rpiλ(i) is the p-adic expansion of λ, ρ is the partition defined in Theorem 1.1 (iii) and the set Λ((α|),ρ) is as defined in Notation 3.8.

Outline

In Section 2 we recall the main ideas concerning the Brauer construction for p-permutation modules and set up our notation for symmetric group modules and modules for wreath products. In Section 3 we find the Broué quotients of signed Young permutation modules. In Section 4 we use these results, together with James’ Submodule Theorem, to define Young modules and signed Young modules in the symmetric group setting. We then prove Donkin’s Theorem 1.1. We give some immediate corollaries of this theorem in Section 5. In Sections 6 and 7, we prove Theorems 1.2, 1.3 and 1.4.

2 Preliminaries

We work with left modules throughout. For background on vertices and sources and other results from modular representation theory we refer the reader to [1]. For an account of the representation theory of the symmetric group we refer the reader to [17] or [19], or for more recent developments, to [20].

2.1 Indecomposable summands

Let G be a finite group. Let M and N be FG-modules. We write NM if N is isomorphic to a direct summand of M. We have already used the notation [M:N] for the number of summands in a direct sum decomposition of M that are isomorphic to the indecomposable module N. This multiplicity is well defined by the Krull–Schmidt Theorem (see [1, Section 4, Lemma 3]). The proof of the following lemma is easy.

Lemma 2.1.

Let M and N be FG-modules, and let N be indecomposable. Suppose that H is a normal subgroup of G acting trivially on both the modules M and N. Let M¯ and N¯ be the corresponding F[G/H]-modules. Then

[M:N]=[M¯:N¯].

2.2 Broué correspondence

Let G be a finite group. An FG-module V is said to be a p-permutation module if for every Sylow p-subgroup P of G there exists a linear basis of V that is permuted by P. A useful characterization of p-permutation modules is given by the following theorem (see [3, (0.4)]).

Theorem 2.2.

An indecomposable FG-module V is a p-permutation module if and only if there exists a p-subgroup P of G such that VIndPGF; equivalently, V has trivial source.

It easily follows that the class of p-permutation modules is closed under restriction and induction and under taking direct sums, direct summands and tensor products.

We now recall the definition and the basic properties of Brauer quotients. Given an FG-module V and P a p-subgroup of G, the set of fixed points of P on V is denoted by

VP={vV:gv=v for all gP}.

It is easy to see that VP is an FNG(P)-module on which P acts trivially. For Q a proper subgroup of P, the relative trace map TrQP:VQVP is the linear map defined by

TrQP(v)=gP/Qgv,

where the sum is over a complete set of left coset representatives for Q in P. The definition of this map does not depend on the choice of the set of representatives. We observe that

TrP(V)=Q<PTrQP(VQ)

is an FNG(P)-module on which P acts trivially. We define the Brauer quotient of V with respect to P to be the F[NG(P)/P]-module

V(P)=VP/TrP(V).

If V is an indecomposable FG-module and P is a p-subgroup of G such that V(P)0, then P is contained in a vertex of V. Broué proved in [3] that the converse holds for p-permutation modules.

Theorem 2.3 ([3, Theorem 3.2]).

Let V be an indecomposable p-permutation module and let P be a vertex of V. Let Q be a p-subgroup of G. Then V(Q)0 if and only if QPg for some gG.

Here Pg denotes the conjugate gPg-1 of P. If V is an FG-module with p-permutation basis with respect to a Sylow p-subgroup P~ of G and PP~, then, taking for each orbit of P on the sum of the basis elements in that orbit, we obtain a basis for VP. Each sum over an orbit of size p or more is a relative trace from a proper subgroup of P. Hence V(P) is isomorphic to the F-span of

P={v:gv=v for all gP}.

Thus Theorem 2.3 has the following corollary.

Corollary 2.4.

Let V be a p-permutation FG-module with p-permutation basis B with respect to a Sylow p-subgroup of G containing a subgroup P. The Brauer quotient V(P) has BP as a basis. Moreover, V has an indecomposable summand with a vertex containing P if and only if BP.

The next result states what is now known as the Broué correspondence.

Theorem 2.5 ([3, Theorems 3.2 and 3.4]).

An indecomposable p-permutation FG-module V has vertex P if and only if V(P) is a projective F[NG(P)/P]-module. Furthermore, we have the following statements.

  1. The Brauer map sending V to V(P) is a bijection between the isomorphism classes of indecomposable p-permutation FG-modules with vertex P and the isomorphism classes of indecomposable projective modules for F[NG(P)/P]. Regarded as an FNG(P)-module, V(P) is the Green correspondent of V.

  2. Suppose that V has vertex P. If M is a p-permutation FG-module, then V is a direct summand of M if and only if V(P) is a direct summand of M(P). Moreover, [M:V]=[M(P):V(P)].

The following lemma allows the Broué correspondence to be applied to monomial modules such as signed Young permutation modules.

Lemma 2.6.

Let A be a subset of F×. Let M be an FG-module with an F-basis B={m1,,mr} such that, if gG and miB, then gmi=amj for some aA and some mjB. Then, for any p-subgroup P of G, there exist coefficients a1,,arA such that {a1m1,,armr} is a p-permutation basis of M with respect to P.

Proof.

Let {i1,,is} be a subset of {1,,r} such that is the disjoint union of 1,,s, where, for each 1js,

j={mk:gmij=agmk for some gP and agA}.

Suppose that gmij=amk and gmij=amk for some g,gP and a,aA. Then we have g-1gmij=aa-1mij and, consequently, Fmij is a 1-dimensional Fg-1g-module. Since P is a p-subgroup, it follows that Fmij is the trivial Fg-1g-module. Hence a=a. Thus the coefficient ag is independent of the choice of g, and depends only on mij and mk.

For each 1js, let

Λj={akmk:gmij=akmk for some gP and akA}.

By the previous paragraph, j=1sΛj is a basis of M. It is sufficient to prove that each Λj is permuted by P. Let xP, and let akmk, akmkΛj. Suppose that x(akmk)=b(akmk) for some bF. We have gmij=akmk and gmij=akmk for some g, gP. Thus g-1xgmij=bmij. Repeating the argument in the first paragraph, we see that Fmij is the trivial Fg-1xg-module and so b=1. ∎

The Brauer quotient of an outer tensor product of p-permutation modules is easily described.

Lemma 2.7.

Let G1 and G2 be finite groups, let M1,M2 be p-permutation FG1- and FG2-modules, and let P1 and P2 be p-subgroups of G1 and G2, respectively. Then

(M1M2)(P1×P2)M1(P1)M2(P2)

as a representation of

NG1×G2(P1×P2)/(P1×P2)(NG1(P1)/P1)×(NG2(P2)/P2).

Proof.

The statement follows from an easy application of Theorem 2.5. ∎

2.3 Partitions and compositions

Let n0. A composition of n is a sequence of non-negative integers α=(α1,,αr) such that αr0 and α1++αr=n. In this case, we write (α)=r and |α|=n. The unique composition of 0 is denoted by ; we have ()=0. The Young subgroup𝔖α is the subgroup 𝔖α1××𝔖αr of 𝔖n, where the ith factor 𝔖αi acts on the set {α1++αi-1+1,,α1++αi-1+αi}. Let α=(α1,,αr) and β=(β1,,βs) be compositions and let q. We denote by qα and αβ the compositions of q|α| and |α|+|β| defined by

qα=(qα1,,qαr),
αβ=(α1,,αr,β1,,βs),

respectively. We set 0α=. We denote by α+β the composition of |α|+|β| defined by

α+β=(α1+β1,,αs+βs,αs+1,,αr),

where we have assumed, without loss of generality, that sr. We define α-β similarly, in the case when βiαi for each is.

A composition α is a partition if it is non-increasing. A partition α is called p-restricted if αi-αi+1<p for all i1. We denote the set of compositions, partitions and p-restricted partitions of n by 𝒞(n), 𝒫(n) and 𝒫(n), respectively. A partition α is p-regular if its conjugate α, defined by αj=|{i:αij}|, is p-restricted. It is well known that if λ is a partition, then there exist unique p-restricted partitions λ(i) for i0 such that

(2.1)λ=i0piλ(i).

We call this expression the p-adic expansion of λ.

Let 𝒫2(n), 𝒞2(n) and 𝒫2(n) be the sets consisting of all pairs (λ|ν) of partitions, compositions and p-restricted partitions, respectively, such that |λ|+|ν|=n. Here λ or ν may be the empty composition . For (λ|ν),(α|β)𝒫2(n), we say that (λ|ν)dominates(α|β), and write (λ|ν)(α|β), if, for all k1, we have

  1. i=1kλii=1kαi, and

  2. |λ|+i=1kνi|α|+i=1kβi.

(As a standing convention we declare that λi=0 whenever λ is a partition and i>(λ).) This defines a partial order on the set 𝒫2(n) called the dominance order. This order becomes the usual dominance order on partitions when restricted to the subsets {(λ|)𝒫2(n)} or {(|ν)𝒫2(n)} of 𝒫2(n).

2.4 Modules for symmetric groups

Let n0, let 𝔖n be the symmetric group on the set {1,,n} and let 𝔄n be its alternating subgroup. Given a subgroup H of 𝔖n, we denote the trivial representation of H by F(H), and the restriction of the sign representation of 𝔖n to H by sgn(H). In the case when H=𝔖γ for some composition γ of n we write F(γ) and sgn(γ) for F(H) and sgn(H), respectively. If γ=(n), we reduce the number of parentheses by writing F(n) and sgn(n), respectively.

For λ a p-regular partition of n, let Dλ be the F𝔖n-module defined by

Dλ=Sλ/rad(Sλ),

where Sλ is the Specht module labelled by λ (see [17, Chapter 4]). By [17, Theorem 11.5] each Dλ is simple, and each simple F𝔖n-module is isomorphic to a unique Dλ. The simple F𝔖n-modules can also be labelled by p-restricted partitions. For λ𝒫(n) we set Dλ=soc(Sλ). The connection between the two labellings is given by DλDλsgn(n). For λ𝒫(n), let Pλ denote the projective cover of the simple F𝔖n-module Dλ.

Finally, for γ𝒫(n), let χγ denote the ordinary irreducible character of Sγ, defined over the rational field.

2.5 Modules for wreath products

Let m and let G be a finite group. Recall that the multiplication in the group G𝔖m is given by

(g1,,gm;σ)(g1,,gm;σ)=(g1gσ-1(1),,gmgσ-1(m);σσ),

for (g1,,gm;σ),(g1,,gm;σ)G𝔖m. (Our notation for wreath products is taken from [19, Section 4.1].) Let M be an FG-module. The m-fold tensor product of M becomes an F[G𝔖m]-module with the action given by

(g1,,gm;σ)(v1vm)=sgn(σ)g1vσ-1(1)gmvσ-1(m)

for (g1,,gm;σ)G𝔖m, v1,,vmM. We denote this module by M^m. Note that we have twisted the action of the top group 𝔖m by the sign representation. Thus, in the notation of [19, 4.3.14], we have

M^m=(#mM)~Inf𝔖mG𝔖m(sgn(m)).

The 1-dimensional module sgn(k)^n will be important to us. In our applications k will be a p-power, and so odd. Since a transposition in the top group 𝔖n acts on {1,,kn} as a product of k disjoint transpositions, and so has odd sign, there is a simpler definition of this module, as Res𝔖k𝔖n𝔖knsgn(kn). More generally, given α𝒞(n) and an odd number k, we define

(2.2)sgn(k)^α=Res(𝔖k𝔖α1)××(𝔖k𝔖αr)𝔖knsgn(kn).

For use in the proof of Proposition 4.5 we briefly recall the character theory of the group C2𝔖n. Let χλ denote the irreducible character of 𝔖n labelled by λ𝔖n. For (λ|μ)𝒫2(n), with |λ|=m1 and |μ|=m2, we define χ(λ|μ) to be the ordinary character of the following module for C2𝔖n:

IndC2(𝔖m1×𝔖m2)C2𝔖n(Inf𝔖m1C2𝔖m1(χλ)(Inf𝔖m2C2𝔖m2(χμ)sgn(2)^m2)).

A standard Clifford theory argument (see for instance [19, Theorem 4.34]) shows that the characters χ(λ|μ) for (λ|μ)𝒫2(n) are precisely the irreducible characters of C2𝔖n.

2.6 Sylow p-subgroups of 𝔖n

Let Pp be the cyclic group (1,2,,p)𝔖p of order p. Let P1={1} and, for d1, set

Ppd+1=PpdPp={(σ1,,σp;π):σ1,,σpPpd,πPp}.

By [19, 4.1.22, 4.1.24], Ppd is a Sylow p-subgroup of 𝔖pd.

Let n. Let n=i=0rnipi, where 0ni<p for i{0,,r}, and let nr0 be the p-adic expansion of n. By [19, 4.1.22, 4.1.24], the Sylow p-subgroups of 𝔖n are each conjugate to the direct product i=0r(Ppi)ni. Hence if we define Pn to be a Sylow p-subgroup of the Young subgroup i=0r(𝔖pi)ni, then Pn is a Sylow p-subgroup of 𝔖n. The normalizer N𝔖n(Pn) of Pn in 𝔖n is denoted by Nn.

Whenever ρ=(ρ1,,ρr)𝒞(n), we denote by Pρ a Sylow p-subgroup of 𝔖ρ, defined so that Pρ=i=1rPρi. In the special case when

ρ=(1m0,pm1,,(ps)ms)=(1,,1m0 copies,p,,pm1 copies,,ps,,psms copies),

where mi0 for each i, we have Pρ=i=0s(Ppi)mi; in particular, the group Pρ has precisely mi orbits of size pi on the set {1,2,,n} for each i. We write Nρ=N𝔖n(Pρ).

3 The Brauer quotients of signed Young permutation modules

In this section, we determine the Brauer quotients of signed Young permutation modules with respect to Sylow subgroups of Young subgroups. Our main result is Proposition 3.12; this generalizes [9, Proposition 1]. The description of the Brauer quotients is combinatorial, using the (α|β)-tableaux defined below.

Fix n and (α|β)𝒞2(n). Let α=(α1,,αr) and β=(β1,,βs). The diagram[α][β] is the set consisting of the boxes(i,j)2 for i and j such that either1ir and 1jαiorr+1ir+s and 1jβi-r. A box (i,j) is said to be in rowi. The subset of [α][β] consisting of the boxes belonging to the first r rows (respectively, the last s rows) is denoted by [α] (respectively, [β]).

Definition 3.1.

An (α|β)-tableauT is a bijective function

T:[α][β]{1,,n}.

For (i,j)[α][β], the (i,j)-entry of T is T(i,j).

We represent an (α|β)-tableauT by putting the (i,j)-entry of T in the box (i,j) of the diagram [α][β]. Considering [α] as the Young diagram [α], we denote the α-tableau T([α]) by T+. Similarly, we denote the β-tableau T([β]) by T-. It will sometimes be useful to write

T=(T+|T-).

The (α|β)-tableau T is row standard if the entries in each row of T are increasing from left to right, i.e. both T+ and T- are row standard in the usual sense. We denote by Tα|β the unique row standard (α|β)-tableau such that for all i, j{1,,n}, if i is in row a of Tα|β and j is in row b of Tα|β and ij, then ab. For example,

where the thicker line separates the two parts of the tableau.

Let 𝒯(α|β) be the set of all (α|β)-tableaux. If T𝒯(α|β) and g𝔖n, then we define gT to be the (α|β)-tableau obtained by applying g to each entry of T, i.e. (gT)(i,j)=g(T(i,j)). This defines an action of 𝔖n on the set 𝒯(α|β). The vector space F𝒯(α|β) over F with basis 𝒯(α|β) is therefore a permutation F𝔖n-module.

For each T𝒯(α|β), let R(T)𝔖n be the row stabilizer of T in 𝔖n, consisting of those g𝔖n such that the rows of T and gT coincide as sets. Then R(T)=R(T+)×R(T-), where R(T+) and R(T-) are the row stabilizers of T+ and T-, respectively, in the usual sense. Denote by U(α|β) the subspace of F𝒯(α|β) spanned by

{T-sgn(g2)g1g2T:T𝒯(α|β),(g1,g2)R(T+)×R(T-)}.

In fact, U(α|β) is an F𝔖n-submodule of F𝒯(α|β), since for all h𝔖n and for any (g1,g2)R(T+)×R(T-) and T𝒯(α|β) we have

h(T-sgn(g2)gT)=hT-sgn(g2h)gh(hT)U(α|β),

where g=g1g2, since ghRh(T)=R(hT) and g2hR((hT)-).

Definition 3.2.

For each T𝒯(α|β), we write

{T}={(T+|T-)}

for the element T+U(α|β)F𝒯(α|β)/U(α|β) and call it an (α|β)-tabloid.

Note that g{T}={gT} for all g𝔖n and T𝒯(α|β). If T,T𝒯(α|β) are such that T-=T- and T+ is obtained by swapping two entries in the same row of T+, then {T}={T}. On the other hand, if T+=T+ and T- is obtained by swapping two entries in the same row of T-, then {T}=-{T}. The graphical representation of (α|β)-tableaux is shown in Example 3.5 below.

Let

Ω(α|β)={{T}:T is a row standard (α|β)-tableau}F𝒯(α|β)/U(α|β).

It is clear that Ω(α|β) is an F-basis of F𝒯(α|β)/U(α|β). We write FΩ(α|β) for the F𝔖n-module F𝒯(α|β)/U(α|β).

Lemma 3.3.

Let (α|β)C2(n).

  1. The F𝔖n-module FΩ(α|β) is isomorphic to the signed Young permutation M(α|β).

  2. For any p-subgroup P of 𝔖n, there exist coefficients a{T}{±1} for each {T}Ω(α|β) such that

    {a{T}{T}:{T}Ω(α|β)}

    is a p-permutation basis for FΩ(α|β)M(α|β) with respect to P.

Proof.

By the remarks after Definition 3.2 there is an isomorphism

F(α)sgn(β)F{Tα|β}

of F[𝔖α×𝔖β]-modules. Since |Ω(α|β)|=dimFM(α|β), part (i) follows from the characterization of induced modules in [1, Section 8, Corollary 3]. Part (ii) follows from Lemma 2.6, since, for all {T}Ω(α|β) and σ𝔖n, we have σ{T}=±{T} for some {T}Ω(α|β). ∎

In view of Lemma 3.3 (i), we shall identify M(α|β) with FΩ(α|β), so that M(α|β) has the set of (α|β)-tabloids as a basis.

The next corollary follows from Lemma 3.3 and Corollary 2.4.

Corollary 3.4.

Let (α|β)C2(n).

  1. Let P be a p-subgroup of 𝔖n. The F[N𝔖n(P)/P]-module M(α|β)(P) has a linear basis consisting of all the (α|β)-tabloids {T} that are fixed by P.

  2. Let ρ=(1n0,pn1,,(pr)nr) be a partition of n. The group

    Nρ/Pρ𝔖n0×((Np/Pp)𝔖n1)××((Npr/Ppr)𝔖nr)

    acts on the set of Pρ-fixed (α|β)-tabloids by transitively permuting the entries in Pρ-orbits of size pi according to 𝔖ni and, within each Pρ-orbit of size pi, permuting its entries according to Npi/Ppi, for all i{0,1,,r}.

More explicitly, the basis in Corollary 3.4 (i) consists of all (α|β)-tabloids {T} such that T is row standard and each row of T is a union of orbits of P on {1,,n}. This can be seen in the following example.

Example 3.5.

Let p=3. Consider the 3-subgroups

Q1=(1,2,3),(4,5,6),(7,8,9)andQ2=(4,5,6),(7,8,9)

of 𝔖9. By Corollary 3.4 (i), since there are no ((2,1)|(6))-tabloids fixed by Q1, we have M((2,1)|(6))(Q1)=0. On the other hand, M(Q2) has a basis consisting of the ((2,1)|(6))-tabloids

where the bold line separates each T+ from T-. Taking ρ=(1,1,1,3,3), we have Pρ=Q2 and

N𝔖9(Q2)=𝔖3×(N𝔖3(P3)𝔖2)=𝔖3×(𝔖3𝔖2).

The first factor 𝔖3 permutes the entries 1,2,3 of each tabloid without sign, and the second factor 𝔖3𝔖2 permutes the entries 4,5,6,7,8,9 with sign. The subgroup Q2 acts trivially on the tabloids. Thus if {U} and {V} are the first two ((2,1)|(6))-tabloids above, then

Res𝔖3𝔖2N𝔖9(Q2)(F{U})sgn(3)^2

and (23)(45){V}=-{U}. Note that the isomorphism above requires the sign twist in the definition of M^ for M an F𝔖m-module that we commented on in Section 2.5.

Given m, we define a 1-dimensional F[Nk𝔖m]-module by

sgn(Nk)^m=ResNk𝔖m𝔖k𝔖msgn(k)^m.

Using this, we may now define three key families of modules. Let denote the bifunctor sending a pair (U|V) where U is an F𝔖m1-module and V is an F𝔖m2-module to the F[C2𝔖m]-module

IndC2(𝔖m1×𝔖m2)C2𝔖m(Inf𝔖m1C2𝔖m1(U)(Inf𝔖m2C2𝔖m2(V)sgn(2)^m2)).

Definition 3.6.

Let k, let m0 and let (γ|δ)𝒞2(m). Let |γ|=m1 and |δ|=m2.

  1. We define Vk(γ|δ) to be the F[𝔖k𝔖m]-module

    Ind(𝔖k𝔖m1)×(𝔖k𝔖m2)𝔖k𝔖m(Inf𝔖m1𝔖k𝔖m1(Mγ)(Inf𝔖m2𝔖k𝔖m2(Mδ)sgn(k)^m2))
  2. We define Wk(γ|δ) to be the F[(Nk/Pk)𝔖m]-module obtained from

    ResNk𝔖m𝔖k𝔖mVk(γ|δ)IndNk(𝔖m1×𝔖m2)Nk𝔖m(Inf𝔖m1Nk𝔖m1(Mγ)(Inf𝔖m2Nk𝔖m2(Mδ)sgn(Nk)^m2))

    via the canonical surjection

    Nk𝔖m(Nk𝔖m)/(Pk)m(Nk/Pk)𝔖m.
  3. For k2, we define W¯k(γ|δ) to be the F[C2𝔖m]-module (Mγ|Mδ). We define

    W¯1(γ|δ)=Inf𝔖mC2𝔖mW1(γ|δ).

Note that Wk(γ|δ) may equivalently be defined to be the F[(Nk/Pk)𝔖m]-module obtained from

(3.1)IndNk(𝔖γ×𝔖δ)Nk𝔖m(F(Nk𝔖γ)sgn(Nk)^δ)

via the canonical surjection as in Definition 3.6 (ii). We have

W1(γ|δ)=V1(γ|δ)M(γ|δ)

as F𝔖m-modules. When k2, the F[C2𝔖m]-module W¯k(α|β) is isomorphic to Vk(α|β), considered as an F[C2𝔖m]-module via the canonical surjection

𝔖k𝔖m(𝔖k𝔖m)/𝔄km(𝔖k/𝔄k)𝔖mC2𝔖m.

Similarly, we have that W¯k(α|β) is isomorphic to ResNk𝔖m𝔖k𝔖mVk(α|β), considered as an F[C2𝔖m]-module via the canonical surjection

Nk𝔖m(Nk𝔖m)/(N𝔄k(Pk))mC2𝔖m.

Lemma 3.7.

For all k2 and all (γ|δ)C2(m), we have

W¯k(γ|δ)V2(γ|δ),

as F[C2Sm]-modules.

Proof.

It suffices to show that sgn(k)^m2sgn(2)^m2 as F[C2𝔖m2]-modules, where sgn(k) is regarded as an FC2-module via the canonical surjection

𝔖k𝔖k/𝔄kC2.

This is clear since sgn(k)sgn(2) as FC2-modules in this regard. ∎

The following notation will be used to describe the direct summands of the Brauer quotients of the signed Young permutation modules M(α|β).

Notation 3.8.

Let (α|β)𝒞2(n) and ρ=(1n0,pn1,(p2)n2,,(pr)nr)𝒞(n). We write Λ((α|β),ρ) for the set consisting of all pairs of tuples of compositions (𝜸|𝜹)=(𝜸(0),𝜸(1),,𝜸(r)|𝜹(0),𝜹(1),,𝜹(r)) such that:

  1. α=i=0rpi𝜸(i), β=i=0rpi𝜹(i), and

  2. |𝜸(i)|+|𝜹(i)|=ni for each i{0,,r}.

Let (α|β)𝒞2(n). Recall that Ω(α|β) is the basis of M(α|β) consisting of all (α|β)-tabloids. As remarked after Corollary 3.4, an F-basis of M(α|β)(Pρ) is obtained by taking those (α|β)-tabloids {(T+|T-)}Ω(α|β) such that the rows of T+ and T- are unions of the orbits of Pρ. Given such a basis element {(T+|T-)} and i{0,,r}, let 𝜸j(i) and 𝜹k(i) be the numbers of Pρ-orbits of length pi in rows j and k of T+ and T-, respectively. For each i{0,,r}, let

𝜸(i)=(𝜸1(i),𝜸2(i),),
𝜹(i)=(𝜹1(i),𝜹2(i),).

Note that |𝜸(i)|+|𝜹(i)|=ni for each i, and so

(𝜸(0),𝜸(1),,𝜸(r)|𝜹(0),𝜹(1),,𝜹(r))Λ((α|β),ρ).

We say that the (α|β)-tabloid {(T+|T-)} is of ρ-type (𝜸|𝜹). For example, if p=3, n=9 and ρ=(13,32), so Pρ=(4,5,6),(7,8,9), then the ((3,3)|(3))-tabloid

has ρ-type ((3),(0,1)|,(1)).

We denote the set of all (α|β)-tabloids of ρ-type (𝜸|𝜹) by Ω((α|β),ρ)(𝜸|𝜹). Then the disjoint union

(3.2)Ω((α|β),ρ)=(𝜸|𝜹)Λ((α|β),ρ)Ω((α|β),ρ)(𝜸|𝜹)

is an F-basis of M(α|β)(Pρ). Thus, as F-vector spaces, we have

(3.3)M(α|β)(Pρ)=FΩ((α|β),ρ)=(𝜸|𝜹)Λ((α|β),ρ)FΩ((α|β),ρ)(𝜸|𝜹).

It is clear that (3.3) is in fact a decomposition of FNρ-modules, since Nρ permutes orbits of Pρ of the same size as blocks for its action, and therefore preserves the ρ-type in its action on (α|β)-tabloids. Furthermore, Pρ fixes all (α|β)-tabloids having a specified ρ-type. Therefore we obtain the following lemma.

Lemma 3.9.

Let (α|β)C2(n) and let ρ=(1n0,pn1,,(pr)nr) be a partition of n. The Brauer quotient of M(α|β) with respect to the subgroup Pρ has the following direct sum decomposition into F[Nρ/Pρ]-modules:

M(α|β)(Pρ)=(𝜸|𝜹)Λ((α|β),ρ)FΩ((α|β),ρ)(𝜸|𝜹).

In view of Lemma 3.9, to understand the Brauer quotient M(α|β)(Pρ) of the signed Young permutation module M(α|β), it suffices to understand each of the F[Nρ/Pρ]-modules FΩ((α|β),ρ)(𝜸|𝜹).

Definition 3.10.

Suppose that (α|β)𝒞2(n) and that ρ=(1n0,pn1,,(pr)nr) is a partition of n. Let the orbits of Pρ of size pi be 𝒪i,1,,𝒪i,ni. Let

Θ:Ω((α|β),ρ)(𝜸|𝜹)Λ((α|β),ρ)i=0rΩ(𝜸(i)|𝜹(i))

be the bijective function defined as follows. Suppose that {T}Ω(α|β) is of ρ-type (𝜸|𝜹). For each 0ir, let {Ti} be the (𝜸(i)|𝜹(i))-tabloid such that Ti is row standard, and row k of (Ti)+ (respectively, (Ti)-) contains j if and only if row k of T+ (respectively, T-) contains the orbit 𝒪i,j. Define

Θ({T})=({Ti})i=0,1,,r.

We note that, by the definition of Pρ,

𝒪i,j={(j-1)pi+1+=0i-1np,,jpi+=0i-1np}

for i{0,,r} and j{1,,ni}. Clearly, the bijection Θ in Definition 3.10 restricts to a bijection, also denoted Θ,

Θ:Ω((α|β),ρ)(𝜸|𝜹)i=0rΩ(𝜸(i)|𝜹(i)).

Since |Ω(𝜸(i)|𝜹(i))|=dimFM(𝜸(i)|𝜹(i))=[𝔖ni:(𝔖𝜸(i)×𝔖𝜹(i))], we obtain the following lemma.

Lemma 3.11.

Let (α|β)C2(n), ρ=(1n0,pn1,,(pr)nr) such that |ρ|=n, and let (𝛄|𝛅)Λ((α|β),ρ). Set

H=i=0rNpi(𝔖𝜸(i)×𝔖𝜹(i))=i=0r(Npi𝔖𝜸(i))×(Npi𝔖𝜹(i))Nρ.

Then |Ω((α|β),ρ)(𝛄|𝛅)|=[Nρ:H].

We have reached the main result of this section.

Proposition 3.12.

Suppose that (α|β)C2(n) and that

ρ=(1n0,pn1,,(pr)nr)𝒞(n).

Regarded as an F[Nρ/Pρ]-module, the Brauer quotient M(α|β)(Pρ) of the signed Young permutation module M(α|β) with respect to Pρ satisfies

M(α|β)(Pρ)(𝜸|𝜹)Λ((α|β),ρ)i=0rWpi(𝜸(i)|𝜹(i)).

Proof.

Recall that for each pair (λ|μ)𝒞2(n), we have defined a row-standard (λ|μ)-tableau Tλ|μ immediately after Definition 3.1. Fix (𝜸|𝜹)Λ((α|β),ρ) and let Z=FΩ((α|β),ρ)(𝜸|𝜹). By Lemma 3.9, it suffices to show that

Zi=0rWpi(𝜸(i)|𝜹(i))

as FNρ-modules with Pρ acting trivially, or equivalently, by (3.1), that

(3.4)Zi=0rIndNpi(𝔖𝜸(i)×𝔖𝜹(i))Npi𝔖ni(F(Npi𝔖𝜸(i))sgn(Npi)^𝜹(i)).

Let {S}Ω((α|β),ρ)(𝜸|𝜹) be the unique (α|β)-tabloid such that

Θ({S})=(T𝜸(0)|𝜹(0),T𝜸(1)|𝜹(1),,T𝜸(r)|𝜹(r))i=0rΩ(𝜸(i)|𝜹(i)).

Using the Nρ-action on Z, we observe that Z is a cyclic FNρ-module generated by {S}. Let X be the subspace of Z linearly spanned by {S}. By the definition of {S}, the subspace X is an FH-module, where

H=i=0rNpi(𝔖𝜸(i)×𝔖𝜹(i))=i=0r((Npi𝔖𝜸(i))×(Npi𝔖𝜹(i)))Nρ,

and there is an isomorphism

X(F(N1𝔖𝜸(0))sgn(N1)^𝜹(0))(F(Npr𝔖𝜸(r))sgn(Npr)^𝜹(r))

of FH-modules. Since dimFZ=[Nρ:H]dimFX by Lemma 3.11, we have ZIndHNρX by the characterization of induced modules in [1, Section 8, Corollary 3]. Hence we obtain the isomorphism (3.4) as desired. ∎

4 Young modules and signed Young modules

In this section we define Young modules and signed Young modules in the setting of the symmetric group and prove Theorem 1.1.

4.1 Vertices

As a first step we identify the possible vertices of summands of signed Young modules. Recall from Section 2.6 that Pk denotes a Sylow subgroup of 𝔖k and, if ρ is a partition of n, then Pρ denotes a Sylow subgroup of the Young subgroup 𝔖ρ of 𝔖n. We require the following lemma from [9]; a proof, slightly shorter than the one in [9], is included to make the article self-contained.

Lemma 4.1 (Erdmann [9, Lemma 1]).

Let G be a finite group and let M be a p-permutation FG-module. If P and P~ are p-subgroups of G such that P<P~ and dimFM(P)=dimFM(P~), then no indecomposable summand of M has vertex P.

Proof.

Suppose, for a contradiction, that U is such a summand. Let M=UV where V is a complementary FG-module. By Corollary 2.4, we have U(P)0 and U(P~)=0. Thus

M(P~)=U(P~)V(P~)=V(P~)

and

M(P)=U(P)V(P).

This is a contradiction, since taking a p-permutation basis for V and applying Corollary 2.4 shows that dimFV(P)dimFV(P~). ∎

Proposition 4.2.

Let (α|β)C2(n). If P is a vertex of an indecomposable summand of M(α|β), then there exists ρ=(1n0,pn1,,(pr)nr)C(n) such that P is conjugate in Sn to Pρ.

Proof.

Let H be the Young subgroup of 𝔖n having the same orbits as P on {1,,n} and let P~ be a Sylow p-subgroup of H. Note that P~ has the same orbits on {1,,n} as P: suppose that each subgroup has exactly ni orbits of size pi for each i{0,,r}, so P~ is conjugate in 𝔖n to Pρ. It suffices to prove that P=P~.

Let {T} be an (α|β)-tabloid fixed by P. As remarked following Corollary 3.4, each row of T is a union of orbits of P. Therefore each row is a union of orbits of P~, and so if gP~, then g{T}=±{T}. Since g has p-power order, we see that g{T}={T}. It now follows from Corollary 3.4 that

dimFM(α|β)(P)=dimFM(α|β)(P~).

By Lemma 4.1 we have P=P~, as required. ∎

Combining Proposition 3.12 and Proposition 4.2, we see that the Broué correspondents of the non-projective indecomposable summands of M(α|β) are certain outer tensor products of the projective indecomposable summands of the F[(Npi/Ppi)𝔖m]-modules Wpi(γ|δ) in Definition 3.6. In fact, it is most convenient to factor out a further subgroup that acts trivially, and consider projective summands of the F[C2𝔖m]-modules W¯pi(γ|δ).

4.2 Projective summands of W¯k(γ|δ)

Fix k and m1,m20. Let m=m1+m2. Recall from Section 2.5 that if α𝒫(n), that is, α is a p-restricted partition of n, then Pα denotes the projective cover of the simple F𝔖n-module Dα.

We remind the reader that the bifunctor was defined just before Definition 3.6.

Definition 4.3.

Let (α|β)𝒫2(m). We define Q¯(α|β)=(Pα|Pβ).

Example 4.10 gives an example of these modules. Note that each tensor factor is projective, so each Q¯(α|β) is projective.

Lemma 4.4.

The F[C2Sm]-modules F(Dα|Dβ) for every (α|β)RP2(m) form a complete set of non-isomorphic simple F[C2Sm]-modules. Moreover, the F[C2Sm]-module Q¯(α|β) is the projective cover of F(Dα|Dβ) and the modules Q¯(α|β) for (α|β)RP2(m) form a complete set of non-isomorphic indecomposable projective modules for F[C2Sm].

Proof.

The first claim follows from the construction of simple modules for wreath products stated in [19, Theorem 4.34]. For the second, note that by functoriality, there is a surjection Q¯(α|β)=(Pα|Pβ)(Dα|Dβ). Hence the projective F[C2𝔖m]-module Q¯(α|β) has the projective cover of (Dα|Dβ) as a summand. Since the inertial group of

(Pα|Pβ)=Inf𝔖m1C2𝔖m1(Pα)(Inf𝔖m2C2𝔖m2(Pβ)sgn(2)^m2)

is (C2𝔖m1)×(C2𝔖m2), it follows from [2, Proposition 3.13.2] that Q¯(α|β) is indecomposable. Therefore Q¯(α|β) is the projective cover of (Dα|Dβ). ∎

Let G be a finite group. By [2, Section 3.11], we may associate a character to a p-permutation FG-module M by taking a p-modular system (K,𝒪,F) compatible with F and an 𝒪G-module M𝒪 whose p-modular reduction is M. The ordinary character of M is then the character of the KG-module K𝒪M𝒪. If M is projective and indecomposable, the ordinary character of M may equivalently be defined by Brauer reciprocity (see for instance [25, Section 15.4]).

Proposition 4.5.

Let (γ|δ)P2(m), where |γ|=m1 and |δ|=m2. Each indecomposable projective summand of W¯k(γ|δ) is isomorphic to some Q¯(α|β), where (α|β)RP2(m) satisfies

  1. |α|=m1 and |β|=m2,

  2. αγ and βδ.

Proof.

By Lemma 4.4, each indecomposable projective summand of W¯k(γ|δ) is isomorphic to some Q¯(α|β). By the ‘wedge’ shape of the decomposition matrix of 𝔖n with columns labelled by p-restricted partitions (see for instance [20, Theorem 5.2]) and Brauer reciprocity, the ordinary character of Pα contains the irreducible character χα exactly once. Hence the ordinary character of Q¯(α|β) contains the character

χ(α|β)=IndC2(𝔖m1×𝔖m2)C2𝔖m(Inf𝔖m1C2𝔖m1(χα)×(Inf𝔖m2C2𝔖m2(χβ)sgn(2)^m2))

defined in Section 2.5 exactly once.

We now consider when the ordinary character of W¯k(γ|δ) contains χ(α|β). The restriction of W¯k(γ|δ) to the base group in the wreath product C2𝔖m is a direct sum of 1-dimensional submodules. In each such submodule, m1 of the factors in the product C2m act trivially and m2 of the factors act as sgn(2). It follows by basic Clifford theory that the ordinary character of W¯k(γ|δ) contains the character χ(α|β) only if |α|=m1 and |β|=m2. By Young’s rule (see for instance [17, Theorem 13.13]), the ordinary character of Mγ contains χα only if αγ, and similarly the ordinary character of Mδ contains χβ only if βδ. It follows that if Q¯(α|β) is a summand of W¯k(γ|δ) then α𝒫(m1), β𝒫(m2), αγ and βδ. ∎

4.3 Definition of signed Young modules

We define signed Young modules as the Broué correspondents of tensor products of suitable inflations of the modules Q¯(α|β). To make this precise, we need the three further families of modules defined below: their definition follows the same pattern as the p-permutation modules Vk(γ|δ), Wk(γ|δ) and W¯k(γ|δ) in Definition 3.6.

Definition 4.6.

Let k, let m0, and let (α|β)𝒫2(m). Let m1=|α| and m2=|β|. The F[𝔖k𝔖m]-module Rk(α|β) is defined by

Rk(α|β)=(Pα|Pβ).

By convention,

Rk(|β)=Inf𝔖m2𝔖k𝔖m2(Pβ)sgn(k)^m2,

and similarly for Rk(α|). Furthermore if m=0, then Rk(|) is the trivial F𝔖0-module. If k=1, then we identify 𝔖k𝔖m with 𝔖m and get

R1(α|β)=Ind𝔖m1×𝔖m2𝔖m(Pα(Pβsgn(m2))).

Recall from Section 2.6 that Pk is a fixed Sylow p-subgroup of 𝔖k and that Nk=N𝔖k(Pk).

Definition 4.7.

Let k, let m0, and let (α|β)𝒫2(m). Let Qk(α|β) be the F[(Nk/Pk)𝔖m]-module defined by

Qk(α|β)=ResNk𝔖m𝔖k𝔖mRk(α|β)

considered as an F[(Nk/Pk)𝔖m]-module via the canonical surjection

Nk𝔖m(Nk𝔖m)/(Pk)m(Nk/Pk)𝔖m.

Again if k=1, we identify N1𝔖m with 𝔖m and we have

Q1(α|β)=R1(α|β).

Since (𝔄k)m acts trivially on Rk(α|β), we see that (N𝔄k(Pk)/Pk)m acts trivially on Qk(α|β). It is clear that

(4.1)Qk(α|β)IndNk(𝔖m1×𝔖m2)Nk𝔖m(Inf𝔖m1Nk𝔖m1(Pα)(Inf𝔖m2Nk𝔖m2(Pβ)sgn(Nk)^m2)),

again regarded as an F[(Nk/Pk)𝔖m]-module by this canonical surjection.

Definition 4.8.

Let k, let m0, and let (α|β)𝒫2(m). For k2, let Q¯k(α|β) be the F[C2𝔖m]-module obtained from Qk(α|β) via the canonical surjection

(Nk/Pk)𝔖m((Nk/Pk)𝔖m)/(N𝔄k(Pk)/Pk)mC2𝔖m.

We define the F[C2𝔖m]-module Q¯1(α|β) by

Q¯1(α|β)=Inf𝔖mC2𝔖mQ1(α|β).

The following lemma justifies the notation Q¯k(α|β) for the projective modules just defined.

Lemma 4.9.

Let k2 and let (α|β)RP2(m), where mN0. Then

Q¯k(α|β)Q¯(α|β)R2(α|β)

as F[C2Sm]-modules.

Proof.

The first isomorphism is clear from the definitions and the second follows as in Lemma 3.7. ∎

We pause to give a small example showing the exceptional behaviour when k=1.

Example 4.10.

Let p=3 and let k2. Let ε=Inf𝔖3𝔖k𝔖3(sgn(3)). There are four mutually non-isomorphic 1-dimensional simple F[𝔖k𝔖3]-modules, namely

F(k)^3,sgn(k)^3,F(k)^3ε,sgn(k)^3ε,

where the trivial module appears as F(k)^3ε. The projective covers of these modules are

Rk((1,1,1)|),Rk(|(2,1)),Rk((2,1)|),Rk(|(1,1,1)),

respectively. Quotienting out by the trivial action of the group 𝔄k, the corresponding modules Q¯(α|β) for F[C2𝔖3] are precisely the projective covers of the four 1-dimensional simple modules for F[C2𝔖3]. The four remaining simple modules for F[C2𝔖3], each projective; by Lemma 4.4, they are isomorphic to the modules Q¯(α|β) where both α and β are non-empty. By contrast, when k=1, identifying 𝔖1𝔖3 with 𝔖3 as described after Definition 4.6, we have

Q¯1((1,1,1)|)Q¯1(|(2,1))P(1,1,1)M(2,1)sgn

and

Q¯1((2,1)|)Q¯1(|(1,1,1))P(2,1)M(2,1).

We are finally ready to define signed Young modules.

Definition 4.11.

Let (λ|pμ)𝒫2(n). Let

λ=i0piλ(i)andμ=i0piμ(i)

be the p-adic expansions of λ and μ, as defined in (2.1). Let n0=|λ(0)| and let ni=|λ(i)|+|μ(i-1)| for each i. Let r be maximal such that nr0 and let ρ=(1n0,pn1,,(pr)nr). We define the signed Young moduleY(λ|pμ) to be the unique (up to isomorphism) F𝔖n-module V such that

V(Pρ)Q1(λ(0)|)Qp(λ(1)|μ(0))Qpr(λ(r)|μ(r-1)).

We define a Young module to be a signed Young module of the form Y(λ|).

The isomorphism above is an isomorphism of projective F[𝔖n0×((Np/Pp)𝔖n1)××((Npr/Ppr)𝔖nr)]-modules. Observe that Pρ is trivial if and only if λ is p-restricted and μ=; in this case Q1(λ(0)|) is regarded as a 𝔖n-module by identifying N1𝔖n with 𝔖n, and since λ=λ(0) we have

Y(λ|)=Q1(λ|)=Pλ.

The following proposition gives part of Theorem 1.1 (i).

Proposition 4.12.

The following statements hold.

  1. If (α|β)𝒫2(n), then M(α|β) is a direct sum of signed Young modules.

  2. If α𝒫(n), then Mα is a direct sum of Young modules.

Proof.

Let (α|β)𝒫2(n) and let V be an indecomposable summand of M(α|β). By Proposition 4.2 there exists ρ=(1m0,pm1,,(pr)mr)𝒞(n) such that Pρ is a vertex of V. Recall that

Nρ/Pρ𝔖n0×((Np/Pp)𝔖n1)××((Npr/Ppr)𝔖nr).

By Proposition 3.12, there exists (𝜸|𝜹)Λ((α|β),ρ) such that the projective F[Nρ/Pρ]-module V(Pρ) is a direct summand of

W1(𝜸(0)|𝜹(0))Wp(𝜸(1)|𝜹(1))Wpr(𝜸(r)|𝜹(r)).

By Lemmas 4.4 and 4.9 there exist partitions λ(0),,λ(r) and μ(0),,μ(r-1) such that

V(Pρ)=Q1(λ(0))Qp(λ(1)|μ(0))Qpr(λ(r)|μ(r-1)).

By Theorem 2.5, VY(λ|pμ), where λ=i=0rpiλ(i) and μ=i=0r-1piμ(i). This proves part (i). For (ii), observe that if β=, we have 𝜹(i)= for each i, and so μ(i)= for each i. ∎

4.4 Column symmetrization of (α|β)-tabloids

To deal with the projective summands of signed Young permutation modules, we require the following corollary of the key lemma used by James to prove his Submodule Theorem in [17]. Given a tableau 𝔱 with entries from a set 𝒪, let C𝔱𝔖𝒪 be the group of permutations which fix the columns of 𝔱 setwise. Set κ𝔱=gC𝔱sgn(g)g.

Proposition 4.13.

Let λP(n) and let t be a λ-tableau. In any direct sum decomposition of Mλ into indecomposable modules there is a unique summand Uλ such that κtUλ0. Moreover, if αP(n), then κtUα=0 unless λα.

Proof.

This follows immediately from [17, Lemma 4.6]. ∎

By the Krull–Schmidt Theorem, the Uλ are well-defined up to isomorphism. It is clear that UαUβ if and only if α=β.

We also need the following generalization of part of James’ lemma.

Lemma 4.14.

Let (α|β)C2(n) and let T=(T+|T-) be an (α|β)-tableau. Let λP(n) and let t be a λ-tableau. If κt{T}0, then (λ|)(α|β).

Proof.

Let 𝒪 be the set of entries of T+. Let H=𝔖𝒪𝔖λ and let 𝒪1,,𝒪s be the orbits of H on 𝒪, ordered so that |𝒪1||𝒪s|. Let

ν=(|𝒪1|,,|𝒪s|)𝒫(|α|).

The jth largest orbit of H has size at most λj. Therefore we have νjλj for each j{1,,s}, and so ν is a subpartition of λ. It immediately follows that

(4.2)i=1kλii=1kνi

for all k. (By our standing convention, νi=0 if i>(ν).)

Let 𝔱 be a ν-tableau having the entries of 𝒪j in its jth column. Observe that C𝔱C𝔱. Choose g1,,gsC𝔱 such that C𝔱=g1C𝔱gsC𝔱, where the union is disjoint. We have

κ𝔱=(sgn(g1)g1++sgn(gs)gs)κ𝔱.

Since κ𝔱{T}0, we have κ𝔱{T}0. Since C𝔱 fixes the entries in T-, it follows that κ𝔱{T+}0. The argument used to prove [17, Lemma 4.6] now shows that any two entries in the same row of {T+} lie in different columns of 𝔱, and so να. Hence, by (4.2), we have (λ|)(α|β), as required. ∎

4.5 Proof of Theorem 1.1

For convenience we repeat the statement of this theorem below.

Theorem 1.1 (Donkin [7]).

There exist indecomposable FSn-modules Y(λ|pμ) for (λ|pμ)P2(n) with the following properties:

  1. if (α|β)𝒫2(n), then M(α|β) is isomorphic to a direct sum of modules Y(λ|pμ) for (λ|pμ)𝒫2(n) such that (λ|pμ)(α|β),

  2. [M(λ|pμ):Y(λ|pμ)]=1,

  3. if λ=i=0rpiλ(i) and μ=i=0r-1piμ(i) are the p-adic expansions of λ and μ , as defined in (2.1), then Y(λ|pμ) has as a vertex a Sylow p-subgroup of the Young subgroup 𝔖ρ, where ρ is the partition of n having exactly |λ(i)|+|μ(i-1)| parts of size pi for each i{0,,r}.

We shall prove the theorem by showing that parts (i), (ii) and (iii) of Theorem 1.1 hold when Y(λ|pμ) is as defined in Definition 4.11. In fact, part (iii) holds by definition, so we may concentrate on parts (i) and (ii).

Proof of Theorem 1.1.

We work by induction on n0. If n<p, then F𝔖n is semisimple and the modules Y(λ|) for λ𝒫(n) form a complete set of simple F𝔖n-modules. Hence parts (i) and (ii) follow from Proposition 4.5. Now let np.

We first deal with non-projective summands. Let (λ|pμ)𝒫2(n) and suppose that either λ is not p-restricted or μ. Let n0=|λ(0)| and let

ni=|λ(i)|+|μ(i-1)|for i.

Let ρ=(1n0,pn1,,(pr)nr).

By Theorem 2.5 and Proposition 3.12, [M(α|β):Y(λ|pμ)] is equal to the sum of the following products over all (𝜸|𝜹)Λ((α|β),ρ):

[W1(𝜸(0)|𝜹(0)):Pλ(0)]i=1r[Wpi(𝜸(i)|𝜹(i)):Qpi(λ(i)|μ(i-1))].

Suppose the product is non-zero for (𝜸|𝜹)Λ((α|β),ρ). Then Pλ(0) is a direct summand of W1(𝜸(0)|𝜹(0))M(𝜸(0)|𝜹(0)). Since Pλ(0)=Y(λ(0)|), it follows from the inductive hypothesis that (λ(0)|)(𝜸(0)|𝜹(0)). Similarly we have that Qpi(λ(i)|μ(i-1)) is a direct summand of Wpi(𝜸(i)|𝜹(i)) for each i{1,,r}. By Proposition 4.5, we have λ(i)𝜸(i) and μ(i-1)𝜹(i) for each such i. Hence

(4.3)λ-λ(0)=i=1rpiλ(i)i=1rpi𝜸(i)=α-𝜸(0)

and

(4.4)pμ=i=1rpiμ(i-1)i=1rpi𝜹(i)=β-𝜹(0).

Hence λα-𝜸(0)+λ(0)α and

|λ|+i=1jpμi=|α|+|𝜹(0)|+i=1jpμi
|α|+|𝜹(0)|+i=1j(β-𝜹(0))i|α|+i=1jβi

for all j. Therefore (λ|pμ)(α|β). By Proposition 4.12, every summand of M(α|β) is isomorphic to some Y(λ|pμ), so this proves part (i) in the non-projective case. If (α|β)=(λ|pμ), then, by divisibility considerations, 𝜸(0)=λ(0) and 𝜹(0)=. Moreover, equality holds in both (4.3) and (4.4), so we have 𝜸(i)=λ(i) and 𝜹(i)=μ(i-1) for each i{1,,r}. Conversely, if 𝜸 and 𝜹 are defined in this way, then the product is 1. This proves part (ii) in the non-projective case.

We now deal with the projective summands. By Proposition 4.12, if α𝒫(n), then Mα is a direct sum of modules Y(λ|) for λ𝒫(n). The argument so far shows that if α is not p-restricted, then Y(α|) is a summand of Mα, and Y(α|) is a summand of Mγ only if αγ. Therefore, inductively working down the dominance order on partitions, we see that, for each such α, the submodule Uα in Proposition 4.13 is Y(α|). By counting, the remaining Uα for α𝒫(n) are the modules Y(λ|) for λ𝒫(n). Again working inductively down the dominance order of partitions, it follows from Proposition 4.5 that Uα=Y(α|) for each α𝒫(n). This proves part (i) in the projective case when β=, and also proves part (ii) in the projective case.

Finally, suppose that λ is p-restricted and Y(λ|) is a direct summand of M(α|β). Let 𝔱 be a λ-tableau. By Proposition 4.12, we have κ𝔱M(α|β)0. Hence there exists an (α|β)-tabloid {T} such that κ𝔱{T}0. By Lemma 4.14 we have (λ|)(α|β). This completes the proof of part (i) in the projective case. ∎

5 Applications of Theorem 1.1

5.1 Equivalent definitions

We observed in the introduction that since signed Young modules are characterized by Theorem 1.1, our definition of signed Young modules agrees with Donkin’s in [7]. Similarly Theorem 1.1 characterizes the Young module Y(λ|) as the unique summand of Mλ appearing in Mμ only if λμ. By [18, Theorem 3.1 (i)], James’ Young modules admit the same characterisation. The two definitions therefore agree. In [10], Erdmann and Schroll consider Young modules for finite general linear groups. Adapting their proof to symmetric groups (this is mentioned as a possibility in [10], as a way to correct [9]), their definition of the Young modules uses the characterization in Proposition 4.12. Our proof of Theorem 1.1 shows these definitions agree; of course this also follows from the alternative characterization just mentioned.

Remark 5.1.

(i) The counting argument used in our proof of the projective case of Theorem 1.1 is motivated by similar counting arguments used in [10]; the authors of [10] thank Burkhard Külshammer for suggesting this approach.

(ii) We have assumed throughout that F has odd prime characteristic p. It is possible to construct Young modules when p=2 and to prove the analogue of Theorem 1.1 by adapting (and simplifying) the approach herein.

(iii) The analogue of signed Young modules for the finite general linear group GLn(𝔽q) are the linear source modules induced from powers of the determinant representation of parabolic subgroups of GLn(𝔽q). These modules seem worthy of study, especially given the difficulty of working directly with Specht modules for GLn(𝔽q).

5.2 Klyachko’s formula and other applications

The following corollary generalizes Klyachko’s formula to signed Young modules. It is proved in the first step of our proof of Theorem 1.1; alternatively it follows from this theorem by taking Broué correspondents.

Corollary 5.2.

If (α|β) and (λ|pμ)P2(n), then

[M(α|β):Y(λ|pμ)]=(𝜸|𝜹)Λ((α|β),ρ)[W1(𝜸(0)|𝜹(0)):Y(λ(0)|)]
×i=1r[Wpi(𝜸(i)|𝜹(i)):Qpi(λ(i)|μ(i-1))].

We remark that the reduction formula for signed p Kostka numbers in Corollary 5.2 has previously been obtained by Danz, the first and the second authors in [5].

The proof of the following lemma is very easy and is left to the reader. Recall that the notation for the concatenation of two compositions was defined in Section 2.3.

Lemma 5.3.

Let

ρ=(1m0,pm1,,(pr)mr),
γ=(1n0,pn1,,(ps)ns),

be partitions of m and n, respectively, and let k>r. Then

Pρpkγ=Pρ×Ppkγ,
N𝔖m+pkn(Pρ×Ppkγ)=N𝔖m(Pρ)×N𝔖pkn(Ppkγ),
Nρpkγ/Pρpkγ=(Nρ/Pρ)×(Npkγ/Ppkγ).

Let (λ|pμ)𝒫2(n). Suppose that the p-adic expansions of λ and μ are

λ=i0piλ(i)andμ=i0piμ(i),

respectively. Let μ(-1)=. If r is maximal such that |λ(r)|+|μ(r-1)|0, then we set

(5.1)p(λ|pμ)=r.

Lemma 5.4.

Let (λ|pμ)P2(n) and let Pρ be a vertex of the signed Young module Y(λ|pμ).

  1. The signed Young module Y(pλ|p2μ) has vertex Ppρ.

  2. Suppose that k>p(λ|pμ) and let (α|β)𝒫2(m) for some m. Then Y(λ+pkα|p(μ+pkβ)) has vertex Pρ×Ppkγ, where Pγ is a vertex of Y(α|pβ). Moreover, Y(λ|pμ)(Pρ)Y(pkα|pk+1β)(Ppkγ) is isomorphic to the Broué correspondent Y(λ+pkα|p(μ+pkβ))(Pρ×Ppkγ).

Proof.

Suppose that λ, μ have p-adic expansions i0piλ(i), i0piμ(i), respectively. It is clear that the partitions pλ and pμ have p-adic expansions

pλ=i1piλ(i-1)andpμ=i1piμ(i-1),

respectively. So |(pλ)(0)|=0, and

|(pλ)(i)|+|(pμ)(i-1)|=|λ(i-1)|+|μ(i-2)|for all i1,

where we set μ(-1)=. By Definition 4.11, Y(pλ|p2μ) has vertex Ppρ, proving part (i).

Let r=p(λ|pμ). For part (ii), since k>r, the p-adic expansions of λ+pkα and μ+pkβ are

λ+pkα=0irpiλ(i)+ikpiα(i-k),
μ+pkβ=0irpiμ(i)+ikpiβ(i-k),

respectively. By Definition 4.11, Y(λ+pkα|p(μ+pkβ)) has vertex Pη, where

η=(1|λ(0)|,p|λ(1)|+|μ(0)|,,(pr)|λ(r)|+|μ(r-1)|,(pk)|α(0)|,(pk+1)|α(1)|+|β(0)|,)=ρpkγ.

Thus Pη=Pρpkγ=Pρ×Ppkγ. By Definition 4.11 and Lemma 5.3, we have

Y(λ+pkα|p(μ+pkβ))(Pρpkγ)
=Q1(λ(0)|)Qp(λ(1)|μ(0))Qpr(λ(r)|μ(r-1))
Qpk(α(0)|)Qpk+1(α(1)|β(0))
Y(λ|pμ)(Pρ)Y(pkα|pk+1β)(Ppkγ),

as required. ∎

The following result is an interesting special case of [6, Theorem 3.18]. It is included to illustrate a technique used again in the proof of Proposition 7.1.

Lemma 5.5.

Let nN. If n=mp+c where mN0 and 0c<p, then

sgn(n)Y((1c)|(mp)).

Proof.

Let n=i=0rpini be the p-adic expansion of n, and let

ρ=(1n0,pn1,,(pr)nr).

By Definition 4.11, the signed Young module Y((1c)|(mp)) has Pρ as a vertex and

Y((1c)|(mp))(Pρ)Q1((1c)|)Qp(|(n1))Qpr(|(nr))

as a module for F[Nρ/Pρ]. Since ni<p, we have

Qpi(|(ni))Inf𝔖ni(Npi/Ppi)𝔖ni(F(ni))sgn(Npi)^ni
ResNpi𝔖ni𝔖pini(sgn(pini)),

where the second isomorphism follows from (2.2), regarding the right-hand side as a representation of (Npi/Ppi)𝔖ni. Hence there is the following isomorphism of FNρ-modules:

Y((1c)|(mp))(Pρ)ResNρ𝔖n(sgn(n)).

On the other hand, since Pρ is a Sylow p-subgroup of 𝔖n, it is a vertex of sgn(n), and clearly

sgn(n)(Pρ)ResNρ𝔖n(sgn(n))

as an FNρ-module. The Broué correspondence is bijective (see Theorem 2.5), so we have Y((1c)|(mp))sgn(n). ∎

6 Signed p-Kostka numbers

In this section we prove Theorem 1.2 and Theorem 1.3. We work mainly with the F[(Nk/Pk)𝔖m]-modules Wk(γ|δ) and Qk(α|β) defined in Definitions 3.6 and 4.7, and the F[C2𝔖m]-modules W¯k(γ|δ) and Q¯k(α|β) obtained from them by factoring out the trivial action of the even permutations in the base group of the wreath product.

We begin with a key lemma for the proof of Theorem 1.2.

Lemma 6.1.

Let nN. For any (γ|δ)C2(n) and (λ|μ)RP2(n) we have

  1. [Wpi+1(γ|δ):Qpi+1(λ|μ)]=[Wpi(γ|δ):Qpi(λ|μ)] for all i1,

  2. [Wp(γ|):Qp(λ|)]=[W1(γ|):Q1(λ|)],

  3. [Wp(γ|δ):Qp(λ|)]=0 if δ.

Proof.

By Lemma 3.7 and Lemma 4.9 we have

Q¯pj(λ|μ)R2(λ|μ)andW¯pj(γ|δ)V2(γ|δ)

for all j1. Part (i) now follows by applying Lemma 2.1. For (ii), if δ=μ=, then

W¯p(γ|)V2(γ|)=Inf𝔖mC2𝔖m(Mγ)=W¯1(γ|),
Q¯p(λ|)R2(λ|)=Inf𝔖mC2𝔖m(Pλ)=Q¯1(λ|).

So

[W¯p(γ|):Q¯p(λ|)]=[W¯1(γ|):Q¯1(λ|)].

Now apply Lemma 2.1. Finally, the third part follows from Proposition 4.5 and Lemma 2.1. ∎

We are now ready to prove Theorem 1.2.

Proof of Theorem 1.2.

Let Pρ be a vertex of Y(λ|pμ). By Definition 4.11 we have

ρ=(1n0,pn1,(p2)n2,,(pr)nr),

where n0=|λ(0)| and ni=|λ(i)|+|μ(i-1)| for all i{1,,r}. By the Broué correspondence (see Theorem 2.5) and the description of the Broué correspondents of signed Young modules in Lemma 5.4, it is equivalent to show that

[M(pα|pβ)(Ppρ):Y(pλ|p2μ)(Ppρ)][M(α|β)(Pρ):Y(λ|pμ)(Pρ)].

Let Λ=Λ((α|β),ρ) and Λ=Λ((pα|pβ),pρ) be as defined in Notation 3.8. Observe that Λ consists of all compositions

(,𝜸(0),𝜸(1),,𝜸(r)|,𝜹(0),𝜹(1),,𝜹(r))

where (𝜸(0),𝜸(1),,𝜸(r)|𝜹(0),𝜹(1),,𝜹(r))Λ. By Lemma 3.12 applied to M(pα|pβ)(Ppρ), we have

M(pα|pβ)(Ppρ)(𝜸|𝜹)Λi=0r+1Wpi(𝜸(i)|𝜹(i))
=(𝜸|𝜹)ΛW1(|)i=0rWpi+1(𝜸(i)|𝜹(i)).

By Definition 4.11 and Lemma 5.4 (i), we obtain both

[M(pα|pβ)(Ppρ):Y(pλ|p2μ)(Ppρ)]
=(𝜸|𝜹)Λi=0r[Wpi+1(𝜸(i)|𝜹(i)):Qpi+1(λ(i)|μ(i-1))],
[M(α|β)(Pρ):Y(λ|pμ)(Pρ)]
=(𝜸|𝜹)Λi=0r[Wpi(𝜸(i)|𝜹(i)):Qpi(λ(i)|μ(i-1))],

where, as usual, μ(-1)=. By Lemma 6.1, we have

[Wpi+1(𝜸(i)|𝜹(i)):Qpi+1(λ(i)|μ(i-1))]=[Wpi(𝜸(i)|𝜹(i)):Qpi(λ(i)|μ(i-1))]

for all i1, and for i=0 whenever 𝜹(0)=. Otherwise, when i=0 and 𝜹(0), we have

0=[Wp(𝜸(0)|𝜹(0)):Qp(λ(0)|)][W1(𝜸(0)|𝜹(0)):Q1(λ(0)|)].

This completes the proof. ∎

Corollary 6.2.

Let (α|β),(λ|pμ)P2(n). Suppose that λ(0)=. Then

[M(pα|pβ):Y(pλ|p2μ)]=[M(α|β):Y(λ|pμ)].

Proof.

Let ρ𝒞(n) be defined by

ρ=(1|λ(0)|,p|λ(1)|+|μ(0)|,,(pr)|λ(r)|+|μ(r-1)|).

The vertex Pρ of Y(λ|pμ) has no fixed points in {1,2,,n}. Hence 𝜹(0)= for any (𝜸|𝜹)Λ((α|β),ρ). The result now follows from Theorem 1.2. ∎

It is now very easy to deduce the asymptotic stability of signed p-Kostka numbers mentioned in the introduction.

Corollary 6.3.

Let (α|β),(λ|pμ)P2(n). Then, for every natural number k2, we have

[M(pkα|pkβ):Y(pkλ|pk+1μ)]=[M(pα|pβ):Y(pλ|p2μ)][M(α|β):Y(λ|pμ)].

Proof.

This follows immediately from Corollary 6.2 and Theorem 1.2. ∎

Example 6.4.

We present a family of examples where the inequality in Theorem 1.2 is strict. Let 0<c<p, let m and let n=mp+c. Since 𝔖mp×𝔖c has index coprime to p in 𝔖n, the trivial module Y((n)|) is a direct summand of M((mp,c)|); the multiplicity is 1 since M((mp,c)|) comes from a transitive action of 𝔖n. By Lemma 5.5 we have sgn(n)Y((1c)|(mp)). Thus

[M(|(mp,c)):Y((1c)|(mp))]
=[M(|(mp,c))sgn(n):Y((1c)|(mp))sgn(n)]
=[M((mp,c)|):Y((n)|)]
=1.

On the other hand,

[M((mp2,cp)|):Y((mp2)|(p(1c)))]=0

because, by [7, 2.3 (6)], the signed Young modules are pairwise non-isomorphic and so the signed Young module Y((mp2)|p(1c)) is not isomorphic to a Young module. Thus we have

[M(|p(mp,c)):Y(p(1c)|p(mp))]
=[M(|(mp2,cp))sgn(np):Y(p(1c)|p(mp))sgn(np)]
=[M((mp2,cp)|):Y((mp2)|p(1c))]
=0,

where the penultimate equation is obtained using [6, Theorem 3.18]. This shows that

[M(|p(mp,c)):Y(p(1c)|p(mp))]<[M(|(mp,c)):Y((1c)|(mp))].

We now turn to the proof of Theorem 1.3. We need a further result on the Brauer quotients of signed Young permutation modules.

Proposition 6.5.

Let m,nN and let (π|π~)C2(m) and (ϕ|ϕ~)C2(n). Let ρC(m) and γC(n) be compositions of the form

ρ=(1m0,pm1,,(pr)mr),
γ=(1n0,pn1,,(ps)ns).

For all kN such that k>r, we have that M(π|π~)(Pρ)M(pkϕ|pkϕ~)(Ppkγ) is isomorphic to a direct summand of M(π+pkϕ|π~+pkϕ~)(Pρpkγ). Furthermore, if pk>max{π1,π~1}, then

M(π|π~)(Pρ)M(pkϕ|pkϕ~)(Ppkγ)M(π+pkϕ|π~+pkϕ~)(Pρpkγ)

as F[NSm+pkn(Pρpkγ)/Pρpkγ]-modules.

Note that, in Proposition 6.5, while i=0rmipi=m and i=0rnipi=n, these need not be the base p expressions for either m or n.

Proof.

Since k>r, by Lemma 5.3, we have

Pρpkγ=Pρ×Ppkγ.

To ease the notation, we denote by M, M1, M2 the modules M(π+pkϕ|π~+pkϕ~), M(π|π~), M(pkϕ|pkϕ~), respectively. Further, let P=Pρpkγ. By Corollary 3.4, we know that M(P) has as a basis the subset of Ω(π+pkϕ|π~+pkϕ~) consisting of all {R} such that R is a row standard (π+pkϕ|π~+pkϕ~)-tableau whose rows are unions of P-orbits. Similarly, we define bases 1 and 2 of Ω(π|π~) and Ω(pkϕ|pkϕ~) for M1(Pρ) and M2(Ppkγ), respectively; here each (pkϕ|pkϕ~)-tableau S of 2 is filled with the numbers m+1,m+2,,m+pkn.

For {T}1 and {S}2, let

ψ:1×2

be the map defined by

ψ({T},{S})={(R+|R-)},

where R+ is the row standard (π+pkϕ)-tableau such that row i of R+ is the union of row i of T+ and row i of S+, and R- is the row standard (π~+pkϕ~)-tableau such that row i of R- is the union of row i of T- and row i of S-. Here we have used the convention row i of T+ is empty if i>(π), and so on. The map ψ is well defined since the rows of R=(R+|R-) are union of orbits of P=Pρ×Ppkγ on {1,2,,m+pkn}.

Clearly ψ is injective and so it induces an injection of vector spaces

θ:M1(Pρ)M2(Ppkγ)M(P)

defined by θ({T}{S})=ψ({T},{S}). By Lemma 5.3, we may regard the domain and codomain of θ as FN𝔖m+pkn(P)-modules with trivial P-action. It is not difficult to check that

θ(g({T}{S}))=gθ({T}{S})

for all gN𝔖m+pkn(P), {T}1 and {S}2. Therefore θ is an injective homomorphism of FN𝔖m+pkn(P)-modules, and hence an injective homomorphism of F[N𝔖m+pkn(P)/P]-modules. Since both M1(Pρ) and M2(Ppkγ) are projective and hence injective, their outer tensor product is also injective. Therefore, the map θ splits and we obtain that M1(Pρ)M2(Ppkγ) is a direct summand of M(P).

The second assertion follows easily by observing that, if pk>max{π1,π~1}, then the map ψ defined above is a bijection. ∎

We are now ready to prove Theorem 1.3.

Proof of Theorem 1.3.

Let ρ𝒞(m) and γ𝒞(n) be defined by

ρ=(1|λ(0)|,p|λ(1)|+|μ(0)|,,(pr)|λ(r)|+|μ(r-1)|),
γ=(1|α(0)|,p|α(1)|+|β(0)|,,(ps)|α(s)|+|β(s-1)|),

where r=p(λ|pμ) and s=p(α|pβ), respectively. By Definition 4.11, Pρ is a vertex of Y(λ|pμ) and Pγ is a vertex of Y(α|pβ). Since k>r, by Lemma 5.3, we have Pρpkγ=Pρ×Ppkγ and

N𝔖m+pkn(Pρ×Ppkγ)=N𝔖m(Pρ)×N𝔖pkn(Ppkγ).

By Lemma 5.4, Y(pkα|pk+1β) has vertex Ppkγ and Y(λ+pkα|p(μ+pkβ)) has vertex Pρpkγ. Moreover, the Broué correspondent of Y(λ+pkα|p(μ+pkβ)) is

Y(λ|pμ)(Pρ)Y(pkα|pk+1β)(Ppkγ).

By Proposition 6.5, we have

M(π|π~)(Pρ)M(pkϕ|pkϕ~)(Ppkγ)M(π+pkϕ|π~+pkϕ~)(Pρpkγ).

Therefore, using Theorem 2.5 (ii), we deduce that

[M(π+pkϕ|π~+pkϕ~):Y(λ+pkα|p(μ+pkβ))]
=[M(π+pkϕ|π~+pkϕ~)(Pρpkγ):Y(λ+pkα|p(μ+pkβ))(Pρpkγ)]
[M(π|π~)(Pρ)M(pkϕ|pkϕ~)(Ppkγ):
Y(λ|pμ)(Pρ)Y(pkα|pk+1β)(Ppkγ)]
=[M(π|π~)(Pρ):Y(λ|pμ)(Pρ)]
[M(pkϕ|pkϕ~)(Ppkγ):Y(pkα|pk+1β)(Ppkγ)]
=[M(π|π~):Y(λ|pμ)][M(pkϕ|pkϕ~):Y(pkα|pk+1β)]
=[M(π|π~):Y(λ|pμ)][M(pϕ|pϕ~):Y(pα|p2β)],

where the final equality follows from Corollary 6.3. If pk>max{π1,π~1}, then Proposition 6.5 implies that we have equalities throughout. ∎

7 Indecomposable signed Young permutation modules

In this section, in the spirit of Gill’s result [12, Theorem 2], we classify all indecomposable signed Young permutation modules over the field F and determine their endomorphism algebras and their labels as signed Young modules. By [12], we know that any indecomposable Young permutation module is of the form M(m) or M(kp-1,1). It is immediate from the definition of signed Young permutation modules in (1.1) that

M(α|β)Ind𝔖|α|×𝔖|β|𝔖|α|+|β|(Mα(Mβsgn(|β|)).

As such, by Gill’s result, any indecomposable signed Young permutation module is of one of the forms M((m)|(n)), M((m)|(kp-1,1)), M((kp-1,1)|(m)) or M((kp-1,1)|(p-1,1)). Since

M((m)|(kp-1,1))sgn(m+kp)M((kp-1,1)|(m)),

there are essentially three different forms to consider.

Proof of Theorem 1.4.

Let M1=M((m)|(n)). If m=0, then M1 is the sign representation, and if n=0, then M1 is the trivial representation. In these cases, M1 is simple with 1-dimensional endomorphism ring. Suppose that both m and n are non-zero. By the Littlewood–Richardson rule, the module M1 has a Specht series with top Specht factor S(m+1,1n-1) and bottom Specht factor S(m,1n). If m+n is not divisible by p, then the p-cores of (m+1,1n-1) and (m,1n) are non-empty and distinct and so S(m+1,1n-1) and S(m,1n) lie in different blocks. Consequently, M1 is decomposable. Now suppose that m+n is divisible by p. In this case, by Peel’s result [24],

S(m+1,1n-1)={F,n=1,[DλDγ],n2,S(m,1n)={sgn(m+n),m=1,[DμDλ],m2,

where μ, λ and γ are the p-regularization of the partitions (m,1n), (m+1,1n-1) and (m+2,1n-2), respectively (see [19, 6.3.48]). If m=1, then M((1)|(n))M(|(n,1)) is indecomposable. Similarly, if n=1, we have that M((m)|(1))M((m,1)|) is indecomposable. Moreover, since M((m,1)|) has a Loewy series with factors F,D(m,1),F, the endomorphism algebra EndF𝔖m+1M((m,1)|) is 2-dimensional. Tensoring by the sign representation we obtain the same result for EndF𝔖n+1(|(n,1)).

We now study the case when m,n2. In this case, both the head and socle of M1 contain the simple module Dλ. Also, as a signed Young permutation module, M1 is self-dual. Suppose that Dγ is not isomorphic to a composition factor of any direct summand of M1 containing Dλ in its head (and hence in its socle). Then Dγ is necessarily isomorphic to a direct summand of M1. From the Specht series, there is a surjection ψ from M1 onto the Specht module S=S(m+1,1n-1). Since S has composition factors Dγ and Dλ, we have ψ(Dγ)0 and so ψ(Dγ)Dγ. Let Y be an indecomposable direct summand of M1 such that ψ(Y) contains a composition factor Dλ. This shows that ψ(Y)Dλ and hence

S=ψ(DγY)DγY/(Ykerψ)DγDλ.

This is absurd since S is indecomposable. Hence there exists an indecomposable direct summand of M1 containing Dλ in its head and that does not contain Dγ in its head or in its socle. Dually, there exists an indecomposable direct summand of M1 containing Dλ in its head, that does not contain Dμ in its head or in its socle. Thus the only possibility is that M1 is indecomposable with the Loewy structure

[DλDμDγDλ]

and has 2-dimensional endomorphism ring.

Let M2=M((kp-1,1)|(m)). By Gill’s result, if m=0, then M2 is indecomposable and if m=1, then M2M((kp-1,12)|) is decomposable. Suppose that m2. By the Young and Littlewood–Richardson rules, M2 has a Specht series with Specht factors

S1=S(kp+1,1m-1),S2=S(kp,2,1m-2),S3=S(kp,1m),
S4=S(kp-1,2,1m-1),S5=S(kp-1,1m+1),

with S3 occurring twice. If m0 mod p, then S1 and S3 lie in different blocks. If m0 mod p, then S3 and S4 belong to different blocks. Thus we conclude that M2 is decomposable whenever m2.

Let M3=M((kp-1,1)|(p-1,1)). Then M3M((kp-1,12)|(p-1)). By Gill’s result, since M(kp-1,12) is decomposable, we have that

M((kp-1,12)|(p-1))
=Ind𝔖kp+1×𝔖p-1𝔖kp+p(M(kp-1,12)(M(p-1)sgn(p-1)))

is decomposable. ∎

We end by determining the labels of the indecomposable signed Young permutation modules. By the remark immediately following the statement of Theorem 1.4, it suffices to consider the modules M((m)|(n)) where either m=0, n=0 or m+n is divisible by p.

Proposition 7.1.

Let m,nN. Let n=n0+pn, where 0n0<p. There are isomorphisms M((m)|)Y((m)|), M(|(n))Y((1n0)|(pn)) and, provided m+n is divisible by p, M((m)|(n))Y((m,1n0)|(pn)).

Proof.

Clearly M((n)|)Y((n)|)F(n). The second isomorphism follows from Lemma 5.5. In the remaining case, m, n>0 and m+n is divisible by p. Let m=i0mipi and let n=i0nipi be the p-adic expansions. Let r be the greatest integer such that mr+nr0. Let P be a Sylow p-subgroup of 𝔖m×𝔖n. By Proposition 3.12 we have an isomorphism of F[N𝔖m+n(P)/P]-modules

M((m)|(n))(P)W1((m0)|(n0))Wp((m1)|(n1))Wpr((mr)|(nr)).

By Definition 4.11, the signed Young module Y((m,1n0)|(pn)) satisfies

Y((m,1n0)|(pn))(P)=Y((m0,1n0)|)i=1rQpi((mi)|(ni)),

where Qpi((mi)|(ni)) is the F[(Nk/Pk)𝔖m]-module defined in Definition 4.7. The Broué correspondence is bijective (see Theorem 2.5), so it suffices to prove that the tensor factors in these two modules agree.

Observe that m0+n0 is a multiple of p and m0+n0<2p. If m0=n0=0, we have

W1(|)=Y(|).

Next, we assume that m0+n0=p. The F𝔖p-module

W1((m0)|(n0))M((m0)|(n0))

is indecomposable by Theorem 1.4. The only signed Young module for F𝔖p that is not a Young module is the sign representation. Since n0<p, we see that M((m0)|(n0)) is a Young module. The proof of Theorem 1.4 shows that it has a Specht filtration with S(m0,1n0) at the bottom and S(m0+1,1n0-1) at the top. Therefore W1((m0)|(n0))=M((m0)|(n0))Y((m0,1n0)|), as required.

Finally, suppose that i1. By Definition 3.6 (ii), we have that Wpi((mi)|(ni)) is the F[(Npi/Ppi)𝔖mi+ni]-module obtained from

IndNpi(𝔖mi×𝔖ni)Npi𝔖mi+ni(Inf𝔖miNpi𝔖mi(F(mi))((Inf𝔖niNpi𝔖ni(F(ni))sgn(Npi)^ni)).

by the canonical surjection (Npi𝔖mi+ni)/(Ppi)mi+ni(Npi/Ppi)𝔖mi+ni. Since mi,ni<p the projective covers P(mi) and P(ni) are the trivial F𝔖mi- and F𝔖ni-modules, respectively. Therefore, by (4.1), we have

Wpi((mi)|(ni))Qpi((mi)|(ni)),

again as required. ∎


Communicated by Radha Kessar


Funding statement: The first author is supported by the ERC advanced grant 291512. The second author is supported by Singapore Ministry of Education AcRF Tier 1 grant RG13/14. Part of this work was done while the first and fourth authors visited National University of Singapore in December 2014; this visit was supported by London Mathematical Society grant 41406 and funding from National University of Singapore and Royal Holloway, University of London.

Acknowledgements

We thank Kai Meng Tan for his valuable suggestions made at the start of this work. We thank an anonymous referee for a careful reading of this paper.

References

[1] J. L. Alperin, Local Representation Theory, Cambridge Stud. Adv. Math. 11, Cambridge University Press, Cambridge, 1986. 10.1017/CBO9780511623592Search in Google Scholar

[2] D. J. Benson, Representations and Cohomology I, Cambridge Stud. Adv. Math. 30, Cambridge University Press, Cambridge, 1995. Search in Google Scholar

[3] M. Broué, On Scott modules and p-permutation modules: An approach through the Brauer homomorphism, Proc. Amer. Math. Soc. 93 (1985), no. 3, 401–408. 10.2307/2045600Search in Google Scholar

[4] J. Brundan and J. Kujawa, A new proof of the Mullineux conjecture, J. Algebraic Combin. 18 (2003), 13–39. 10.1023/A:1025113308552Search in Google Scholar

[5] S. Danz, E. Giannelli and K. J. Lim, On signed p-Kostka matrices, in preparation. Search in Google Scholar

[6] S. Danz and K. J. Lim, Signed Young modules and simple Specht modules, Adv. Math. 307 (2017), 369–416. 10.1016/j.aim.2016.11.016Search in Google Scholar

[7] S. Donkin, Symmetric and exterior powers, linear source modules and representations of Schur superalgebras, Proc. London Math. Soc. (3) 83 (2001), 647–680. 10.1112/plms/83.3.647Search in Google Scholar

[8] S. Donkin, Some remarks on Gill’s theorems on Young modules, J. Algebra 430 (2015), 1–14. 10.1016/j.jalgebra.2014.12.045Search in Google Scholar

[9] K. Erdmann, Young modules for symmetric groups, J. Aust. Math. Soc. 71 (2001), no. 2, 201–210. 10.1017/S1446788700002846Search in Google Scholar

[10] K. Erdmann and S. Schroll, On Young modules of general linear groups, J. Algebra 310 (2007), no. 1, 434–451. 10.1016/j.jalgebra.2006.11.004Search in Google Scholar

[11] M. Fang, A. Henke and S. Koenig, Comparing GL(n)-representations by characteristic-free isomorphisms between generalized Schur algebras, Forum Math. 20 (2008), no. 1, 45–79. 10.1515/FORUM.2008.003Search in Google Scholar

[12] C. Gill, Young module multiplicities, decomposition numbers and the indecomposable Young permutation modules, J. Algebra Appl. 13 (2014), no. 5, Article ID 1350147. 10.1142/S0219498813501478Search in Google Scholar

[13] J. A. Green, Polynomial Representations of GLn, Lecture Notes in Math. 830, Springer, Berlin, 1980. 10.1007/BFb0092296Search in Google Scholar

[14] D. J. Hemmer, Irreducible Specht modules are signed Young modules, J. Algebra 305 (2006), 433–441. 10.1016/j.jalgebra.2006.01.046Search in Google Scholar

[15] A. Henke, On p-Kostka numbers and Young modules, European J. Combin. 26 (2005), 923–942. 10.1016/j.ejc.2004.06.006Search in Google Scholar

[16] A. Henke and S. Koenig, Relating polynomials GL(n)-representations in different degrees, J. Reine Angew. Math. 551 (2002), 219–235. 10.1515/crll.2002.083Search in Google Scholar

[17] G. D. James, The Representation Theory of the Symmetric Groups, Lecture Notes in Math. 682, Springer, Berlin, 1978. 10.1007/BFb0067708Search in Google Scholar

[18] G. D. James, Trivial source modules for symmetric groups, Arch. Math. (Basel) 41 (1983), 294–300. 10.1007/BF01371400Search in Google Scholar

[19] G. D. James and A. Kerber, The Representation Theory of the Symmetric Group, Addison-Wesley, Boston, 1981. Search in Google Scholar

[20] A. Kleshchev, Representation theory of symmetric groups and related Hecke algebras, Bull. Amer. Math. Soc. 47 (2009), 419–481. 10.1090/S0273-0979-09-01277-4Search in Google Scholar

[21] A. A. Klyachko, Direct summand of permutation modules, Sel. Math. Sov. 3 (1984), 45–55. Search in Google Scholar

[22] J. Kujawa, The Steinberg tensor product theorem for GL(m|n), Contemp. Math. 413 (2006), 123–132. 10.1090/conm/413/07843Search in Google Scholar

[23] R. Paget, A family of modules with Specht and dual Specht filtrations, J. Algebra 312 (2007), 880–890. 10.1016/j.jalgebra.2007.03.022Search in Google Scholar

[24] M. H. Peel, Hook representations of the symmetric groups, Glasg. Math. J. 12 (1971), 136–149. 10.1017/S0017089500001245Search in Google Scholar

[25] J. P. Serre, Linear Representations of Finite Groups, Grad. Texts in Math. 42, Springer, New York, 1977. 10.1007/978-1-4684-9458-7Search in Google Scholar

[26] M. Wildon, Multiplicity-free representations of symmetric groups, J. Pure Appl. Algebra 213 (2009), 1464–1477. 10.1016/j.jpaa.2008.12.009Search in Google Scholar

Received: 2015-12-21
Revised: 2016-9-26
Published Online: 2017-2-25
Published in Print: 2017-7-1

© 2017 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 29.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/jgth-2017-0007/html
Scroll to top button