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

Symmetric functions and Springer representations111MSC2010: 14N15,20G44

Syu Kato222Department of Mathematics, Kyoto University, Oiwake Kita-Shirakawa Sakyo Kyoto 606-8502 JAPAN E-mail:[email protected]
Abstract

The characters of the (total) Springer representations are identified with the Green functions by Kazhdan [Israel J. Math. 28 (1977)], and the latter are identified with Hall-Littlewood’s QQ-functions by Green [Trans. Amer. Math. Soc. (1955)]. In this paper, we present a purely algebraic proof that the (total) Springer representations of GL(n)\mathop{GL}(n) are Ext\mathrm{Ext}-orthogonal to each other, and show that it is compatible with the natural categorification of the ring of symmetric functions.

Dedicated to the memory of Tonny Albert Springer

Introduction

Let GG be a connected reductive algebraic group over an algebraically closed field with a Borel subgroup BB. Let WW be the Weyl groups of GG, and let 𝒩LieG\mathcal{N}\subset\mathrm{Lie}\,G denote the variety of nilpotent elements. The cohomology of a fiber of the Springer resolution

μ:T(G/B)𝒩,\mu:T^{*}(G/B)\longrightarrow\mathcal{N},

affords a representation of WW. This is widely recognized as the Springer representation [24], and it is proved to be an essential tool in representation theory of finite and pp-adic Chevalley groups [16, 13, 17, 18, 12]. Here and below, we understand that the Springer representation refers to the total cohomology of a Springer fiber instead of the top cohomology, commonly seen in the literature.

In [10], we found a module-theoretic realization of Springer representations that is axiomatized as Kostka systems. For G=GL(n)G=\mathop{GL}(n), it takes the following form: Let

A=An:=𝔖n[X1,,Xn]A=A_{n}:={\mathbb{C}}\mathfrak{S}_{n}\ltimes{\mathbb{C}}[X_{1},\ldots,X_{n}]

be a graded ring obtained by the smash product of the symmetric group 𝔖n\mathfrak{S}_{n} and a polynomial algebra [X1,,Xn]{\mathbb{C}}[X_{1},\ldots,X_{n}] such that deg𝔖n=0\deg\,\mathfrak{S}_{n}=0 and degXi=1\deg\,X_{i}=1 (1in1\leq i\leq n). Let A𝗀𝗆𝗈𝖽A\mathchar 45\relax\mathsf{gmod} be the category of finitely generated graded AA-modules. Let homA\mathrm{hom}_{A}, endA\mathrm{end}_{A}, and extA\mathrm{ext}_{A} denote the graded versions of HomA\mathrm{Hom}_{A}, EndA\mathrm{End}_{A}, and ExtA\mathrm{Ext}_{A}, respectively. The set of simple graded AA-modules is parametrized by 𝖨𝗋𝗋𝔖n\mathsf{Irr}\,\mathfrak{S}_{n} (up to grading shift), and is denoted as {Lλ}λ𝖨𝗋𝗋𝔖n\{L_{\lambda}\}_{\lambda\in\mathsf{Irr}\,\mathfrak{S}_{n}}. We have a projective cover PλLλP_{\lambda}\rightarrow L_{\lambda} as graded AA-modules.

Theorem A.

For each λ𝖨𝗋𝗋𝔖n\lambda\in\mathsf{Irr}\,\mathfrak{S}_{n}, we have two modules K~λ\widetilde{K}_{\lambda} and KλK_{\lambda} in An𝗀𝗆𝗈𝖽A_{n}\mathchar 45\relax\mathsf{gmod} with the following properties:

  1. 1.

    We have a sequence of AnA_{n}-module surjections PλK~λKλLλP_{\lambda}\rightarrow\!\!\!\!\!\rightarrow\widetilde{K}_{\lambda}\rightarrow\!\!\!\!\!\to K_{\lambda}\rightarrow\!\!\!\!\!\to L_{\lambda}, where the first map is obtained by annihilating all graded Jordan-Hölder components LμL_{\mu} such that μλ\mu\not\geq\lambda with respect to the dominance order on 𝖨𝗋𝗋𝔖n\mathsf{Irr}\,\mathfrak{S}_{n};

  2. 2.

    The graded ring endA(K~λ)\mathrm{end}_{A}(\widetilde{K}_{\lambda}) is a polynomial ring. The ((unique)) graded quotient endA(K~λ)0\mathrm{end}_{A}(\widetilde{K}_{\lambda})\to{\mathbb{C}}_{0}\cong{\mathbb{C}} yields Kλ0endA(K~λ)K~λK_{\lambda}\cong{\mathbb{C}}_{0}\otimes_{\mathrm{end}_{A}(\widetilde{K}_{\lambda})}\widetilde{K}_{\lambda};

  3. 3.

    We have the following ext\mathrm{ext}-orthogonality:

    extAi(K~λ,Kμ)δi,0δλ,μ.\mathrm{ext}^{i}_{A}(\widetilde{K}_{\lambda},K_{\mu}^{*})\cong{\mathbb{C}}^{\oplus\delta_{i,0}\delta_{\lambda,\mu}}.
Remark B.

If we identify λ𝖨𝗋𝗋𝔖n\lambda\in\mathsf{Irr}\,\mathfrak{S}_{n} with a partition, and hence with a nilpotent element xλ𝒩𝔤𝔩(n,)x_{\lambda}\in\mathcal{N}\subset\mathfrak{gl}(n,{\mathbb{C}}) via the theory of Jordan normal form, then we have

KλH(μ1(xλ),)and K~λHStabGL(n,)(xλ)(μ1(xλ),)K_{\lambda}\cong H^{\bullet}(\mu^{-1}(x_{\lambda}),{\mathbb{C}})\hskip 14.22636pt\text{and}\hskip 14.22636pt\widetilde{K}_{\lambda}\cong H^{\bullet}_{\mathrm{Stab}_{\mathop{GL}(n,{\mathbb{C}})}(x_{\lambda})}(\mu^{-1}(x_{\lambda}),{\mathbb{C}})

with a suitable adjustment of conventions ([10, 11]).

Theorem A follows from works of many people ([9, 8, 27, 15, 14, 2, 5]) in several different ways as well as an exact account ([10, 11]) that works for an arbitrary GG. All of these proofs utilize some structures (geometry, cells, or affine Lie algebras) that is hard to see in the category of graded AA-modules.

The main goal of this paper is to give a new proof of Theorem A based on a detailed analysis of KλK_{\lambda}^{*} due to Garsia-Procesi [6] and some algebraic results from [14, 10]. This completes author’s attempt [10, Appendix A] to give a proof of Theorem A inside the category of graded AA-modules.

As a byproduct, we obtain an interesting consequence: We call MA𝗀𝗆𝗈𝖽M\in A\mathchar 45\relax\mathsf{gmod} (resp. MAA𝗀𝗆𝗈𝖽M\in A\boxtimes A\mathchar 45\relax\mathsf{gmod}) to be Δ\Delta-filtered (resp. Δ¯\overline{\Delta}-filtered) if MM admits a decreasing separable filtration (resp. finite filtration) whose associated graded is isomorphic to the direct sum of {K~λ}λ\{\widetilde{K}_{\lambda}\}_{\lambda} (resp. direct sum of {LλKμ}λ,μ\{L_{\lambda}\boxtimes K_{\mu}\}_{\lambda,\mu}) up to grading shifts.

Theorem C (\doteq Theorem 2.37).

The induction of graded AA-modules sends the external tensor product of PλP_{\lambda} and a Δ\Delta-filtered module to a Δ\Delta-filtered module. Dually, the restriction of graded AnA_{n}-modules sends a Δ¯\overline{\Delta}-filtered module of AnA_{n} (=A0An)(=A_{0}\boxtimes A_{n}) to a Δ¯\overline{\Delta}-filtered module of ArAnrA_{r}\boxtimes A_{n-r} (0rn)(0\leq r\leq n).

Recall that the graded modules

n0K(An𝗀𝗆𝗈𝖽)((q))n0K(𝔖n𝗆𝗈𝖽),\bigoplus_{n\geq 0}K(A_{n}\mathchar 45\relax\mathsf{gmod})\subset\mathbb{Q}(\!(q)\!)\otimes_{\mathbb{Z}}\bigoplus_{n\geq 0}K(\mathfrak{S}_{n}\mathchar 45\relax\mathsf{mod}),

are Hopf algebras by Zelevinsky [28], that is identified with the ring Λ\Lambda of symmetric functions up to scalar extensions (Theorem 1.1). In particular, this ring is equipped with four bases {sλ}λ,{Qλ}λ,{Qλ}λ\{s_{\lambda}\}_{\lambda},\{Q^{\vee}_{\lambda}\}_{\lambda},\{Q_{\lambda}\}_{\lambda}, and {Sλ}λ\{S_{\lambda}\}_{\lambda}, usually referred to as the Schur functions, the Hall-Littlewood PP-functions, the Hall-Littlewood QQ-functions, and the big Schur functions, respectively ([19]). We exhibit a natural character identification (that we call the twisted Frobenius characteristic)

Modules of APλK~λKλLλBasis of ΛsλQλQλSλ\begin{array}[]{c|cccc}\textrm{Modules of }A&P_{\lambda}&\widetilde{K}_{\lambda}&K_{\lambda}&L_{\lambda}\\ \hline\cr\textrm{Basis of }\Lambda&s_{\lambda}&Q^{\vee}_{\lambda}&Q_{\lambda}&S_{\lambda}\end{array} (0.1)

that intertwines the products with inductions, and the coproducts with restrictions. (The complete symmetric functions and the elementary symmetric functions are expanded positively by the Schur functions, and hence corresponds to a direct sum of projective modules in this table).

Under this identification, Theorem C implies that the multiplication of a Schur function in Λ\Lambda exhibits positivity with respect to the Hall-Littlewood functions (Corollary 2.39). In addition, we deduce a homological interpretation of skew Hall-Littlewood functions (Corollary 2.40).

In a sense, our exposition here can be seen as a direct approach to an algebraic avatar of the Springer correspondence. We note that interpreting sheaves appearing in the Springer correspondence as constructible functions produces totally different algebraic avatar of the Springer correspondence via Hall algebras (as pursued in Shimoji-Yanagida [22]). Although our Hopf algebra structure is closely related to the Heisenberg categorification (cf. [1]), the author was not able to find a result of this kind in the literature. Nevertheless, he plans to write a follow-up paper that covers the relation with the Heisenberg categorification in an occasion.

Finally, the author was very grateful to find related [25] during the preparation of this paper.

1 Preliminaries

A vector space is always a {\mathbb{C}}-vector space, and a graded vector space refers to a \mathbb{Z}-graded vector space whose graded pieces are finite-dimensional and its grading is bounded from the below. Tensor products are taken over {\mathbb{C}} unless stated otherwise. We define the graded dimension of a graded vector space as

gdimM:=iqidimMi((q)).\mathrm{gdim}\,M:=\sum_{i\in\mathbb{Z}}q^{i}\dim_{{\mathbb{C}}}M_{i}\in\mathbb{Q}(\!(q)\!).

In case dimM<\dim\,M<\infty, we set M:=i(M)iM^{*}:=\bigoplus_{i\in\mathbb{Z}}(M^{*})_{i}, where (M)i:=(Mi)(M^{*})_{i}:=(M_{-i})^{*} for each ii\in\mathbb{Z}. We set [n]q:=1qn1q[n]_{q}:=\frac{1-q^{n}}{1-q} for each n0n\in\mathbb{Z}_{\geq 0}.

For a {\mathbb{C}}-algebra AA, let A𝗆𝗈𝖽A\mathchar 45\relax\mathsf{mod} denote the category of finitely generated left AA-modules. If AA is a graded algebra in the sense that A=i0AiA=\bigoplus_{i\in\mathbb{Z}_{\geq 0}}A_{i} and AiAjAi+jA_{i}A_{j}\subset A_{i+j} (i,j0i,j\in\mathbb{Z}_{\geq 0}), then we denote by A𝗀𝗆𝗈𝖽A\mathchar 45\relax\mathsf{gmod} the category of finitely generated graded AA-modules. We also have a full subcategory A𝖿𝗆𝗈𝖽A\mathchar 45\relax\mathsf{fmod} of A𝗀𝗆𝗈𝖽A\mathchar 45\relax\mathsf{gmod} consisting of finite-dimensional modules.

For a graded algebra AA, the category A𝗀𝗆𝗈𝖽A\mathchar 45\relax\mathsf{gmod} admits an autoequivalence n\left<n\right> for each nn\in\mathbb{Z} such that M=iMiM=\bigoplus_{i\in\mathbb{Z}}M_{i} is sent to Mn:=i(Mn)iM\left<n\right>:=\bigoplus_{i\in\mathbb{Z}}(M\left<n\right>)_{i}, where (Mn)i=Min(M\left<n\right>)_{i}=M_{i-n}. For M,NA𝗀𝗆𝗈𝖽M,N\in A\mathchar 45\relax\mathsf{gmod}, we set

homA(M,N)\displaystyle\mathrm{hom}_{A}(M,N) :=jhomA(M,N)j,homA(M,N)j:=HomA𝗀𝗆𝗈𝖽(Mj,N),\displaystyle\,:=\bigoplus_{j\in\mathbb{Z}}\mathrm{hom}_{A}(M,N)_{j},\hskip 8.53581pt\mathrm{hom}_{A}(M,N)_{j}:=\mathrm{Hom}_{A\mathchar 45\relax\mathsf{gmod}}(M\left<j\right>,N),
extAi(M,N)\displaystyle\mathrm{ext}^{i}_{A}(M,N) :=jextAi(M,N)j,extAi(M,N)j:=ExtA𝗀𝗆𝗈𝖽i(Mj,N).\displaystyle\,:=\bigoplus_{j\in\mathbb{Z}}\mathrm{ext}^{i}_{A}(M,N)_{j},\hskip 8.53581pt\mathrm{ext}^{i}_{A}(M,N)_{j}:=\mathrm{Ext}^{i}_{A\mathchar 45\relax\mathsf{gmod}}(M\left<j\right>,N).

In particular, homA(M,N)\mathrm{hom}_{A}(M,N) and extA(M,N)\mathrm{ext}^{\bullet}_{A}(M,N) are graded vector spaces if dimAi<\dim\,A_{i}<\infty for each i0i\in\mathbb{Z}_{\geq 0}. Moreover, homA(M,N)j\mathrm{hom}_{A}(M,N)_{j} consists of graded AA-module homomorphisms that raise the degree by jj.

For MA𝗀𝗆𝗈𝖽M\in A\mathchar 45\relax\mathsf{gmod}, the head of MM (that we denote by 𝗁𝖽M\mathsf{hd}\,M) is the maximal semisimple graded quotient of MM, and the socle of MM (that we denote by 𝗌𝗈𝖼M\mathsf{soc}\,M) is the maximal semisimple graded submodule of MM.

For a decreasing filtration

M=F0MF1MF2MM=F_{0}M\supset F_{1}M\supset F_{2}M\supset\cdots

of graded vector spaces, we define its kk-th associated graded piece as grkFM:=FkM/Fk+1M\mathrm{gr}^{F}_{k}M:=F_{k}M/F_{k+1}M (k0k\geq 0). We call such a filtration separable if k0FkM={0}\bigcap_{k\geq 0}F_{k}M=\{0\}.

For an exact category 𝒞\mathcal{C}, let [𝒞][\mathcal{C}] denote its Grothendieck group. For M𝒞M\in\mathcal{C}, we have its class [M][𝒞][M]\in[\mathcal{C}]. In case 𝒞\mathcal{C} admits the grading shift functor n\left<n\right> (nn\in\mathbb{Z}), an element f=nanqn[q±1]f=\sum_{n}a_{n}q^{n}\in\mathbb{Z}[q^{\pm 1}] (an0a_{n}\in\mathbb{Z}_{\geq 0}) defines the direct sum

Mf:=n(Mn)anM𝒞.M^{\oplus f}:=\bigoplus_{n\in\mathbb{Z}}\left(M\left<n\right>\right)^{\oplus a_{n}}\hskip 14.22636ptM\in\mathcal{C}.

We may represent a number that is not important by [q±1]\star\in\mathbb{Z}[q^{\pm 1}].

1.1 Partitions and the ring of symmetric functions

We employ [19] as the general reference about partitions and symmetric functions. We briefly recall some key notions there. The set of partitions is denoted by 𝒫{\mathcal{P}}, and the set of partitions of nn (0)(\in\mathbb{Z}_{\geq 0}) is denoted by 𝒫n{\mathcal{P}}_{n}. Each of 𝒫n{\mathcal{P}}_{n} is equipped with a partial order \leq such that (n)(n) is the largest element. We extend the order \leq to the whole 𝒫{\mathcal{P}} by declaring that elements of 𝒫n{\mathcal{P}}_{n} and 𝒫m{\mathcal{P}}_{m} are comparable only if n=mn=m. Let mi(λ)m_{i}(\lambda) be the multiplicity of ii, let (λ)\ell(\lambda) be the partition length, and let |λ||\lambda| be the partition size of λ𝒫\lambda\in{\mathcal{P}}. The conjugate partition of λ𝒫\lambda\in{\mathcal{P}} is denoted by λ\lambda^{\prime}. We set

n(λ):=i1(i1)λi=i1(λi2).n(\lambda):=\sum_{i\geq 1}(i-1)\lambda_{i}=\sum_{i\geq 1}\left(\begin{matrix}\lambda^{\prime}_{i}\\ 2\end{matrix}\right).

For λ𝒫n\lambda\in{\mathcal{P}}_{n} and 1j(λ)+11\leq j\leq\ell(\lambda)+1, let λ(j)𝒫n\lambda^{(j)}\in{\mathcal{P}}_{n} be the partition of (n+1)(n+1) obtained by rearranging {λi}ij{λj+1}\{\lambda_{i}\}_{i\neq j}\cup\{\lambda_{j}+1\}, and for 1j(λ)1\leq j\leq\ell(\lambda), we set λ(j)\lambda_{(j)} be the partition of (n1)(n-1) obtained by rearranging {λi}ij{λj1}\{\lambda_{i}\}_{i\neq j}\cup\{\lambda_{j}-1\}. We set

bλ(q)=j1((1q)(1qmj(λ))).b_{\lambda}(q)=\prod_{j\geq 1}\left((1-q)\cdots(1-q^{m_{j}(\lambda)})\right).

Let Λ\Lambda be the ring of symmetric functions with their coefficients in \mathbb{Z}. Let Λq\Lambda_{q} be its scalar extension to ((q))\mathbb{Q}(\!(q)\!). We have direct sum decompositions Λ=n0Λn\Lambda=\bigoplus_{n\geq 0}\Lambda_{n} and Λq=n0Λq,n\Lambda_{q}=\bigoplus_{n\geq 0}\Lambda_{q,n} into the graded components. The ring Λ\Lambda is equipped with four distinguished bases

{hλ}λ𝒫,{sλ}λ𝒫,{eλ}λ𝒫,and {mλ}λ𝒫,\{h_{\lambda}\}_{\lambda\in{\mathcal{P}}},\hskip 8.53581pt\{s_{\lambda}\}_{\lambda\in{\mathcal{P}}},\hskip 8.53581pt\{e_{\lambda}\}_{\lambda\in{\mathcal{P}}},\hskip 8.53581pt\text{and}\hskip 8.53581pt\{m_{\lambda}\}_{\lambda\in{\mathcal{P}}},

called (the sets of) complete symmetric functions, Schur functions, elementary symmetric functions, and monomial symmetric functions, respectively. We have equalities

h1=s(1)=e1=m(1),hn=s(n),and en=s(1n)n>0.h_{1}=s_{(1)}=e_{1}=m_{(1)},\hskip 8.53581pth_{n}=s_{(n)},\hskip 8.53581pt\text{and}\hskip 8.53581pte_{n}=s_{(1^{n})}\hskip 14.22636ptn\in\mathbb{Z}_{>0}.

We have a symmetric inner product (,)(\bullet,\bullet) on Λ\Lambda such that

(sλ,sμ)=(hλ,mμ)=δλ,μλ,μ𝒫.(s_{\lambda},s_{\mu})=(h_{\lambda},m_{\mu})=\delta_{\lambda,\mu}\hskip 14.22636pt\lambda,\mu\in{\mathcal{P}}.

The ring Λ\Lambda has a structure of a Hopf algebra with the coproduct Δ\Delta satisfying

Δ(hn)=i+j=nhihj,and Δ(en)=i+j=neiej\Delta(h_{n})=\sum_{i+j=n}h_{i}\otimes h_{j},\hskip 8.53581pt\text{and}\hskip 8.53581pt\Delta(e_{n})=\sum_{i+j=n}e_{i}\otimes e_{j}

and the antipode SS satisfying

S(hn)=(1)nen,and S(en)=(1)nhn.S(h_{n})=(-1)^{n}e_{n},\hskip 8.53581pt\text{and}\hskip 8.53581ptS(e_{n})=(-1)^{n}h_{n}.

The antipode SS preserves the inner product (,)(\bullet,\bullet).

1.2 Zelevinsky’s picture for symmetric groups

For a (not necessarily non-increasing) sequence λ=(λ1,λ2,)0\lambda=(\lambda_{1},\lambda_{2},\ldots)\in\mathbb{Z}_{\geq 0}^{\infty} such that jλj=n\sum_{j}\lambda_{j}=n, we define the subgroup

𝔖λ:=j1𝔖λj𝔖n.\mathfrak{S}_{\lambda}:=\prod_{j\geq 1}\mathfrak{S}_{\lambda_{j}}\subset\mathfrak{S}_{n}.

We usually omit 0 in λ=(λ1,λ2,)\lambda=(\lambda_{1},\lambda_{2},\ldots). Each λ𝒫n\lambda\in{\mathcal{P}}_{n} defines an irreducible representation of LλL_{\lambda} of 𝔖n\mathfrak{S}_{n}. We normalize LλL_{\lambda} such that

L(n)𝗍𝗋𝗂𝗏,and L(1n)𝗌𝗀𝗇.L_{(n)}\cong\mathsf{triv},\hskip 8.53581pt\text{and}\hskip 8.53581ptL_{(1^{n})}\cong\mathsf{sgn}.

For 0<r<n0<r<n, we have induction/restriction functors

Indr,nr:\displaystyle\mathrm{Ind}_{r,n-r}:\, 𝔖(r,nr)𝗆𝗈𝖽(M,N)𝔖n𝔖(r,nr)(MN)𝔖n𝗆𝗈𝖽\displaystyle{\mathbb{C}}\mathfrak{S}_{(r,n-r)}\mathchar 45\relax\mathsf{mod}\ni(M,N)\mapsto{\mathbb{C}}\mathfrak{S}_{n}\otimes_{{\mathbb{C}}\mathfrak{S}_{(r,n-r)}}(M\boxtimes N)\in\mathfrak{S}_{n}\mathchar 45\relax\mathsf{mod}
Resr,nr:\displaystyle\mathrm{Res}_{r,n-r}:\, 𝔖n𝗆𝗈𝖽𝔖(r,nr)𝗆𝗈𝖽,\displaystyle{\mathbb{C}}\mathfrak{S}_{n}\mathchar 45\relax\mathsf{mod}\longrightarrow{\mathbb{C}}\mathfrak{S}_{(r,n-r)}\mathchar 45\relax\mathsf{mod},

where the latter is the natural restriction. They induce corresponding maps between the Grothendieck groups that we denote by the same letter.

Theorem 1.1 (Zelevinsky [28]).

We have a \mathbb{Z}-module isomorphism

Ψ0:n0[𝔖n𝗆𝗈𝖽][Lλ]sλΛ.\Psi_{0}:\bigoplus_{n\geq 0}[{\mathbb{C}}\mathfrak{S}_{n}\mathchar 45\relax\mathsf{mod}]\ni[L_{\lambda}]\mapsto s_{\lambda}\in\Lambda.

with the following properties: For M𝔖r𝗆𝗈𝖽M\in{\mathbb{C}}\mathfrak{S}_{r}\mathchar 45\relax\mathsf{mod} and N𝔖n𝗆𝗈𝖽N\in{\mathbb{C}}\mathfrak{S}_{n}\mathchar 45\relax\mathsf{mod}, we have

Ψ0(Indr,n[MN])=Ψ0([M])Ψ0([N]),s=0nΨ0(Ress,ns[N])=Δ([N]).\Psi_{0}(\mathrm{Ind}_{r,n}\,[M\boxtimes N])=\Psi_{0}([M])\cdot\Psi_{0}([N]),\hskip 14.22636pt\sum_{s=0}^{n}\Psi_{0}(\mathrm{Res}_{s,n-s}\,[N])=\Delta([N]).

In particular, we have

hrΨ0([N])=Ψ0(Indr,n[L(r)N]),erΨ0([N])=Ψ0(Indr,n[L(1r)N]).h_{r}\cdot\Psi_{0}([N])=\Psi_{0}(\mathrm{Ind}_{r,n}\,[L_{(r)}\boxtimes N]),\hskip 14.22636pte_{r}\cdot\Psi_{0}([N])=\Psi_{0}(\mathrm{Ind}_{r,n}\,[L_{(1^{r})}\boxtimes N]).

1.3 The algebra AnA_{n} and its basic properties

We follow [10, §2] here. We set

An:=𝔖n[X1,,Xn],A_{n}:={\mathbb{C}}\mathfrak{S}_{n}\ltimes{\mathbb{C}}[X_{1},\ldots,X_{n}],

where 𝔖n\mathfrak{S}_{n} acts on the ring [X1,,Xn]{\mathbb{C}}[X_{1},\ldots,X_{n}] by

(w1)(1Xi)=(1Xw(i))(w1)w𝔖n,1in.(w\otimes 1)(1\otimes X_{i})=(1\otimes X_{w(i)})(w\otimes 1)\hskip 14.22636ptw\in\mathfrak{S}_{n},1\leq i\leq n.

We usually denote ww in place of w1w\otimes 1, and f[X1,,Xn]f\in{\mathbb{C}}[X_{1},\ldots,X_{n}] in place of 1f1\otimes f. The ring AnA_{n} acquires the structure of a graded ring by

degw=0,degXi=1w𝔖n,1in.\deg\,w=0,\hskip 8.53581pt\deg\,X_{i}=1\hskip 14.22636ptw\in\mathfrak{S}_{n},1\leq i\leq n.

The grading of the ring AnA_{n} is non-negative, and the positive degree part An+:=j>0AnjA_{n}^{+}:=\bigoplus_{j>0}A_{n}^{j} defines a graded ideal such that An/An+𝔖nAn0A_{n}/A_{n}^{+}\cong{\mathbb{C}}\mathfrak{S}_{n}\cong A_{n}^{0}. In particular, each LλL_{\lambda} can be understood to be a graded AnA_{n}-module concentrated in degree 0.

The assignments ww1w\mapsto w^{-1} (wWw\in W) and XiXiX_{i}\mapsto X_{i} (1in1\leq i\leq n) define an isomorphism AnAnopA_{n}\cong A_{n}^{op}. Therefore, if MAn𝖿𝗆𝗈𝖽M\in A_{n}\mathchar 45\relax\mathsf{fmod}, then MM^{*} acquires the structure of a graded AnA_{n}-module. We have (Lλ)Lλ(L_{\lambda})^{*}\cong L_{\lambda} for each λ𝒫n\lambda\in{\mathcal{P}}_{n} as 𝔖n\mathfrak{S}_{n} is a real reflection group.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, we have an idempotent eλ𝔖ne_{\lambda}\in{\mathbb{C}}\mathfrak{S}_{n} such that Lλ𝔖neλL_{\lambda}\cong{\mathbb{C}}\mathfrak{S}_{n}e_{\lambda}. We set Pλ:=AneλP_{\lambda}:=A_{n}e_{\lambda}.

Proposition 1.2 (see [10] §2).

The modules {Lλj}λ𝒫n,j\{L_{\lambda}\left<j\right>\}_{\lambda\in{\mathcal{P}}_{n},j\in\mathbb{Z}} is the complete collection of simple objects in An𝗀𝗆𝗈𝖽A_{n}\mathchar 45\relax\mathsf{gmod}. In addition, PλP_{\lambda} is the projective cover of LλL_{\lambda} in An𝗀𝗆𝗈𝖽A_{n}\mathchar 45\relax\mathsf{gmod} for each λ𝒫n\lambda\in{\mathcal{P}}_{n}. \Box

We define

K~λ:=Pλμλ,fhomA(Pμ,Pλ)Imfand Kλ:=K~λj>0,fhomA(Pλ,K~λ)jImf.\widetilde{K}_{\lambda}:=\frac{P_{\lambda}}{\sum_{\mu\not\geq\lambda,f\in\hom_{A}(P_{\mu},P_{\lambda})}\mathrm{Im}\,f}\hskip 8.53581pt\text{and}\hskip 8.53581ptK_{\lambda}:=\frac{\widetilde{K}_{\lambda}}{\sum_{j>0,f\in\hom_{A}(P_{\lambda},\widetilde{K}_{\lambda})_{j}}\mathrm{Im}\,f}.

For each MA𝗀𝗆𝗈𝖽M\in A\mathchar 45\relax\mathsf{gmod}, we set

[M:Lλ]q:=gdimhomA(Pλ,M)=iqidimHom𝔖n(Lλ,Mi)((q)).[M:L_{\lambda}]_{q}:=\mathrm{gdim}\,\mathrm{hom}_{A}(P_{\lambda},M)=\sum_{i\in\mathbb{Z}}q^{i}\dim\,\mathrm{Hom}_{\mathfrak{S}_{n}}(L_{\lambda},M_{i})\in\mathbb{Z}(\!(q)\!).

In case the q=1q=1 specialization of [M:Lλ]q[M:L_{\lambda}]_{q} makes sense, we denote it by [M:Lλ][M:L_{\lambda}].

Lemma 1.3 (see [10] §2).

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, we have

[Kλ:Lμ]q={0λμ1λ=μ,[K~λ:Lμ]q{0λμ1+q[[q]]λ=μ.[K_{\lambda}:L_{\mu}]_{q}=\begin{cases}0&\lambda\not\leq\mu\\ 1&\lambda=\mu\end{cases},\hskip 8.53581pt[\widetilde{K}_{\lambda}:L_{\mu}]_{q}\in\begin{cases}0&\lambda\not\leq\mu\\ 1+q\mathbb{Z}[\![q]\!]&\lambda=\mu\end{cases}.
Proof.

Immediate from the definition. ∎

For 0rn0\leq r\leq n, we consider the subalgebra

Ar,nr:=𝔖(r,nr)[X1,,Xn]ArAnrAn.A_{r,n-r}:={\mathbb{C}}\mathfrak{S}_{(r,n-r)}\ltimes{\mathbb{C}}[X_{1},\ldots,X_{n}]\cong A_{r}\boxtimes A_{n-r}\subset A_{n}.

We have induction/restriction functors

indr,nr\displaystyle\mathrm{ind}_{r,n-r} :Ar,nr𝗀𝗆𝗈𝖽MAnAr,nrMAn𝗀𝗆𝗈𝖽,\displaystyle\,:A_{r,n-r}\mathchar 45\relax\mathsf{gmod}\ni M\mapsto A_{n}\otimes_{A_{r,n-r}}M\in A_{n}\mathchar 45\relax\mathsf{gmod},
resr,nr\displaystyle\mathrm{res}_{r,n-r} :An𝗀𝗆𝗈𝖽Ar,nr𝗀𝗆𝗈𝖽.\displaystyle\,:A_{n}\mathchar 45\relax\mathsf{gmod}\longrightarrow A_{r,n-r}\mathchar 45\relax\mathsf{gmod}.

Since AnA_{n} is free of rank n!r!(nr)!\frac{n!}{r!(n-r)!} over Ar,nrA_{r,n-r}, we find that the both functors are exact, and preserves finite-dimensionality of the modules. We sometimes omit the functor resr,nr\mathrm{res}_{r,n-r} from notation in case there are no possible confusion.

We consider the category 𝒜:=n0An𝗀𝗆𝗈𝖽{\mathcal{A}}:=\bigoplus_{n\geq 0}A_{n}\mathchar 45\relax\mathsf{gmod}. We define

ind:=r,sindr,s:𝒜×𝒜𝒜,res:=r,sresr,s:𝒜𝒜𝒜.\mathrm{ind}:=\bigoplus_{r,s}\mathrm{ind}_{r,s}:{\mathcal{A}}\times{\mathcal{A}}\rightarrow{\mathcal{A}},\hskip 14.22636pt\mathrm{res}:=\bigoplus_{r,s}\mathrm{res}_{r,s}:{\mathcal{A}}\rightarrow{\mathcal{A}}\boxtimes{\mathcal{A}}.
Lemma 1.4.

We embed 𝔖n𝗆𝗈𝖽\mathfrak{S}_{n}\mathchar 45\relax\mathsf{mod} into An𝗀𝗆𝗈𝖽A_{n}\mathchar 45\relax\mathsf{gmod} by regarding M𝔖n𝗆𝗈𝖽M\in\mathfrak{S}_{n}\mathchar 45\relax\mathsf{mod} as a semisimple graded AnA_{n}-module concentrated in degree 0 for each n0n\in\mathbb{Z}_{\geq 0}. Then, we have

Indr,s=indr,sand Resr,s=resr,sr,s0\mathrm{Ind}_{r,s}=\mathrm{ind}_{r,s}\hskip 14.22636pt\text{and}\hskip 14.22636pt\mathrm{Res}_{r,s}=\mathrm{res}_{r,s}\hskip 14.22636ptr,s\in\mathbb{Z}_{\geq 0}

on n0𝔖n𝗆𝗈𝖽\bigoplus_{n\geq 0}\mathfrak{S}_{n}\mathchar 45\relax\mathsf{mod}. In particular, [𝒜][{\mathcal{A}}] can be understood as a ((Hopf)) subalgebra of ((q))Λ=Λq{\mathbb{C}}(\!(q)\!)\otimes\Lambda=\Lambda_{q} by extending the scalar in Theorem 1.1. \Box

The following three theorems are quite well-known to experts.

Theorem 1.5 (Frobenius-Nakayama reciprocity).

For MAr,nr𝗀𝗆𝗈𝖽M\in A_{r,n-r}\mathchar 45\relax\mathsf{gmod} and NAn𝗀𝗆𝗈𝖽N\in A_{n}\mathchar 45\relax\mathsf{gmod}, it holds

extAnk(indr,nrM,N)extAr,nrk(M,resr,nrN)k.\mathrm{ext}^{k}_{A_{n}}(\mathrm{ind}_{r,n-r}\,M,N)\cong\mathrm{ext}^{k}_{A_{r,n-r}}(M,\mathrm{res}_{r,n-r}\,N)\hskip 14.22636ptk\in\mathbb{Z}.
Proof.

This follows from the fact that AnA_{n} is a free Ar,nrA_{r,n-r}-module by the classical Frobenius reciprocity as indr,nr\mathrm{ind}_{r,n-r} sends a projective resolution of MM to a projective resolution of indr,nrM\mathrm{ind}_{r,n-r}\,M. ∎

Theorem 1.6.

For M,NAn𝖿𝗆𝗈𝖽M,N\in A_{n}\mathchar 45\relax\mathsf{fmod}, it holds

extAnk(M,N)extAnk(N,M)k.\mathrm{ext}^{k}_{A_{n}}(M,N)\cong\mathrm{ext}^{k}_{A_{n}}(N^{*},M^{*})\hskip 14.22636ptk\in\mathbb{Z}.
Proof.

We borrow terminology from [7, §2.2]. We have natural isomorphism

homAn(M,N)homAn(N,M).\mathrm{hom}_{A_{n}}(M,N)\cong\mathrm{hom}_{A_{n}}(N^{*},M^{*}).

Since the derived functors of the both sides (defined in an appropriate ambient categories) are δ\delta-functors in each variables, it suffices to see that they are universal δ\delta-functors. By approximating NN by its injective envelope (and hence NN^{*} by its projective cover), we find that the both sides are effacable on the second variables. Thus, they must coincide by [7, 2.2.1 Proposition]. ∎

Theorem 1.7.

The global dimension of AnA_{n} is finite. In particular, every MAn𝗀𝗆𝗈𝖽M\in A_{n}\mathchar 45\relax\mathsf{gmod} admits a graded projective resolution of finite length.

Proof.

See McConnell-Robson-Small [21, 7.5.6]. ∎

We have a [q±1]\mathbb{Z}[q^{\pm 1}]-bilinear inner product ,EP\left<\bullet,\bullet\right>_{EP} on [𝒜][{\mathcal{A}}] prolonging

An𝗀𝗆𝗈𝖽×An𝖿𝗆𝗈𝖽(M,N)i0(1)igdimextAni(M,N)((q)).A_{n}\mathchar 45\relax\mathsf{gmod}\times A_{n}\mathchar 45\relax\mathsf{fmod}\ni(M,N)\mapsto\sum_{i\geq 0}(-1)^{i}\mathrm{gdim}\,\mathrm{ext}^{i}_{A_{n}}(M,N^{*})^{*}\in\mathbb{Q}(\!(q)\!).
Lemma 1.8.

The pairing ,EP\left<\bullet,\bullet\right>_{EP} is a well-defined symmetric form on [𝒜][\mathcal{A}].

Proof.

Since the Euler-Poincaré form respects the short exact sequences, the form ,EP\left<\bullet,\bullet\right>_{EP} must be additive with respect to the both variables.

By the arrangement of duals in the definition of ,EP\left<\bullet,\bullet\right>_{EP}, we find that replacing MM with MnM\left<n\right> and replacing NN with NnN\left<n\right> both result in multiplying qnq^{n} (nn\in\mathbb{Z}). As the category 𝒜{\mathcal{A}} has finite direct sums, we conclude that ,EP\left<\bullet,\bullet\right>_{EP} must be [q±1]\mathbb{Z}[q^{\pm 1}]-bilinear.

We have

[An𝗀𝗆𝗈𝖽]=λ𝒫n[q±1][Pλ]λ𝒫n((q))[Lλ][A_{n}\mathchar 45\relax\mathsf{gmod}]=\bigoplus_{\lambda\in{\mathcal{P}}_{n}}\mathbb{Z}[q^{\pm 1}][P_{\lambda}]\subset\bigoplus_{\lambda\in{\mathcal{P}}_{n}}\mathbb{Q}(\!(q)\!)[L_{\lambda}]

by Proposition 1.2. In particular, Lλ,LμEP((q))\left<L_{\lambda},L_{\mu}\right>_{EP}\in\mathbb{Q}(\!(q)\!) (λ,μ𝒫n\lambda,\mu\in{\mathcal{P}}_{n}) uniquely determines a well-defined ((q))\mathbb{Q}(\!(q)\!)-bilinear form ,EP\left<\bullet,\bullet\right>_{EP} that restricts to [𝒜][{\mathcal{A}}]. It is symmetric by Theorem 1.6. ∎

2 Main results

Keep the setting of the previous section.

Definition 2.1.

Fix 0rn0\leq r\leq n. A Δ\Delta-filtration (resp. Δ¯\overline{\Delta}-filtration) of MAn𝗀𝗆𝗈𝖽M\in A_{n}\mathchar 45\relax\mathsf{gmod} is a decreasing separable filtration

M=F0MF1MF2MM=F_{0}M\supset F_{1}M\supset F_{2}M\supset\cdots

of graded AnA_{n}-modules (resp. graded Ar,nrA_{r,n-r}-modules) such that

grkFM{K~λm}λ𝒫n,m(resp. grkFM{LμKνm}μ𝒫r,ν𝒫nr,m)\mathrm{gr}^{F}_{k}M\in\{\widetilde{K}_{\lambda}\left<m\right>\}_{\lambda\in{\mathcal{P}}_{n},m\in\mathbb{Z}}\hskip 11.38109pt\text{(resp. }\mathrm{gr}^{F}_{k}M\in\{L_{\mu}\boxtimes K_{\nu}\left<m\right>\}_{\mu\in{\mathcal{P}}_{r},\nu\in{\mathcal{P}}_{n-r},m\in\mathbb{Z}}\text{)}

for each k0k\geq 0. In case MM admits a Δ\Delta-filtration, then we set

(M:K~λ)q:=k=0mqmχ(grkFMK~λm),(M:\widetilde{K}_{\lambda})_{q}:=\sum_{k=0}^{\infty}\sum_{m\in\mathbb{Z}}q^{m}\chi(\mathrm{gr}^{F}_{k}M\cong\widetilde{K}_{\lambda}\left<m\right>),

where χ(𝔛)\chi(\mathfrak{X}) takes value 11 if the proposition 𝔛\mathfrak{X} is true, and 0 otherwise.

Lemma 2.2 ([10] §2 or [14]).

The multiplicity (M:K~λ)q(M:\widetilde{K}_{\lambda})_{q} does not depend on the choice of Δ\Delta-filtration. \Box

The following theorem is not new (see Remark 2.4). Nevertheless, the author feels it might worth to report a yet another proof based on Garsia-Procesi [6], that differs significantly from other proofs and is carried out within the category of AnA_{n}-modules:

Theorem 2.3.

Let λ,μ𝒫n\lambda,\mu\in{\mathcal{P}}_{n}. We have the following:

  1. 1.

    For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, the graded ring endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}) is a polynomial ring generated by homogeneous polynomials of positive degrees;

  2. 2.

    The module K~λ\widetilde{K}_{\lambda} is free over endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}), and we have

    0endAn(K~λ)K~λKλ,{\mathbb{C}}_{0}\otimes_{\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda})}\widetilde{K}_{\lambda}\cong K_{\lambda},

    where 0{\mathbb{C}}_{0} is the unique graded one-dimensional quotient of endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda});

  3. 3.

    We have the Ext\mathrm{Ext}-orthogonality:

    extAni(K~λ,Kμ)δλ,μδi,0;\mathrm{ext}^{i}_{A_{n}}(\widetilde{K}_{\lambda},K_{\mu}^{*})\cong{\mathbb{C}}^{\oplus\delta_{\lambda,\mu}\delta_{i,0}};
  4. 4.

    Each PλP_{\lambda} admits a Δ\Delta-filtration, and we have (Pλ:K~μ)q=[Kμ:Lλ]q(P_{\lambda}:\widetilde{K}_{\mu})_{q}=[K_{\mu}:L_{\lambda}]_{q}.

Proof.

Postponed to §2.4. ∎

Remark 2.4.

Theorem 2.3 is originally proved in [10, 11] essentially in this form by using the geometry of Springer correspondence (that works for arbitrary Weyl groups with arbitrary cuspidal data). Theorem 2.3 also follows from results of Haiman [9, 8] that employ the geometry of Hilbert schemes of points on 2{\mathbb{C}}^{2}. We also have two algebraic proofs of Theorem 2.3, one is to use a detailed study of two-sided cells of affine Hecke algebras by Xi [27] together with König-Xi [15] and Kleshchev [14], and another is an analogous result for affine Lie algebras (Chari-Ion [2]) together with Feigin-Khoroshkin-Makedonskyi [5].

We exhibit applications of Theorem 2.3 in §2.5.

2.1 Garsia-Procesi’s theorem

For each 𝙸[1,n]\mathtt{I}\subset[1,n] and |𝙸|r1|\mathtt{I}|\geq r\geq 1, let er(𝙸)e_{r}(\mathtt{I}) be the rr-th elementary symmetric function with respect to the variables {Xi}i𝙸\{X_{i}\}_{i\in\mathtt{I}}. For λ𝒫n\lambda\in{\mathcal{P}}_{n}, we set

dr(λ):=λ1++λr(1rn).d_{r}(\lambda):=\lambda^{\prime}_{1}+\cdots+\lambda^{\prime}_{r}\hskip 14.22636pt(1\leq r\leq n).

We set

𝒞λ:={et(𝙸)rtrdr(λ),|𝙸|=r,𝙸[1,n]}.\mathcal{C}_{\lambda}:=\{e_{t}(\mathtt{I})\mid r\geq t\geq r-d_{r}(\lambda),|\mathtt{I}|=r,\mathtt{I}\subset[1,n]\}.

Let Iλ[X1,,Xn]I_{\lambda}\subset{\mathbb{C}}[X_{1},\ldots,X_{n}] be the ideal generated by 𝒞λ\mathcal{C}_{\lambda} (originally introduced in [26]).

Definition 2.5.

We set Rλ:=[X1,,Xn]/IλR_{\lambda}:={\mathbb{C}}[X_{1},\ldots,X_{n}]/I_{\lambda}, and call it the Garsia-Procesi module.

Lemma 2.6 ([6] §3).

The algebra RλR_{\lambda} admits a structure of graded AnA_{n}-module generated by L(n)L_{(n)}. In addition, [Rλ:L(n)]q=1[R_{\lambda}:L_{(n)}]_{q}=1.

Proof.

Since RλR_{\lambda} is the quotient of P(n)P_{(n)}, it suffices to see that the ideal IλI_{\lambda} is graded and 𝔖n\mathfrak{S}_{n}-stable. Since 𝒞λ\mathcal{C}_{\lambda} consists of homogeneous polynomials and it is stable under the 𝔖n\mathfrak{S}_{n}-action, we conclude the first assertion. For the second assertion, it suffices to notice that 𝒞λ\mathcal{C}_{\lambda} contains all the elementary symmetric polynomials in [X1,,Xn]{\mathbb{C}}[X_{1},\ldots,X_{n}], and hence IλI_{\lambda} contains all the positive degree part of [X1,,Xn]𝔖n{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}}. ∎

Theorem 2.7 (Garsia-Procesi [6] §1).

Let λ𝒫n\lambda\in{\mathcal{P}}_{n}. The [X1,,Xn]{\mathbb{C}}[X_{1},\ldots,X_{n}]-module RλR_{\lambda} admits a decreasing filtration

Rλ=F0RλF1RλF(λ)Rλ={0}R_{\lambda}=F_{0}R_{\lambda}\supset F_{1}R_{\lambda}\supset\cdots\supset F_{\ell(\lambda)}R_{\lambda}=\{0\} (2.1)

such that grjFRλRλ(j+1)j\mathrm{gr}^{F}_{j}R_{\lambda}\cong R_{\lambda_{(j+1)}}\left<j\right> for 0j<(λ)0\leq j<\ell(\lambda). In addition, this filtration respects the 𝔖n1\mathfrak{S}_{n-1}-action, and hence can be regarded as an A1,n1A_{1,n-1}-module filtration. \Box

Theorem 2.8 ([6] Theorem 3.1 and Theorem 3.2).

Let λ𝒫n\lambda\in{\mathcal{P}}_{n}. It holds:

  1. 1.

    We have (Rλ)n(λ)+1={0}(R_{\lambda})_{n(\lambda)+1}=\{0\};

  2. 2.

    We have a 𝔖n\mathfrak{S}_{n}-module isomorphism Rλind𝔖λ𝔖n𝗍𝗋𝗂𝗏R_{\lambda}\cong\mathrm{ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}\,\mathsf{triv}.

In particular, we have [Rλ:Lμ]0[R_{\lambda}:L_{\mu}]\neq 0 only if λμ\lambda\leq\mu. \Box

In view of [19, I​I​I (2.1)], we have the Hall-Littlewood PP- and QQ- functions in Λq\Lambda_{q} indexed by 𝒫{\mathcal{P}}, that we denote by QλQ^{\vee}_{\lambda} and QλQ_{\lambda}, respectively (we changed notation of PP-functions to QQ^{\vee} in order to avoid confusion with projective modules). They satisfy the following relation:

Qλ=bλ1QλΛq.Q^{\vee}_{\lambda}=b_{\lambda}^{-1}Q_{\lambda}\in\Lambda_{q}. (2.2)

We also have the big Schur function ([19, I​I​I (4.6)])

Sλ:=i<j(1qRij)Qλ,S_{\lambda}:=\prod_{i<j}(1-qR_{ij})Q_{\lambda},

where RijR_{ij} are the raising operators.

Theorem 2.9 ([6] §5, particularly (5.24)).

For each λ𝒫\lambda\in{\mathcal{P}}, the polynomial

Qλ:=μ[Kλ:Lμ]qSμΛqQ_{\lambda}:=\sum_{\mu}[K_{\lambda}:L_{\mu}]_{q}\cdot S_{\mu}\in\Lambda_{q}

is the Hall-Littlewood’s QQ-function. \Box

Theorem 2.10 ([19] I​I​I (4.9)).

There exists a (q)\mathbb{Q}(q)-linear bilinear form ,\left<\bullet,\bullet\right> on Λq\Lambda_{q} ((referred to as the Hall inner product)) characterized as

Qλ,Qμ=δλ,μ=Sλ,sμ\left<Q^{\vee}_{\lambda},Q_{\mu}\right>=\delta_{\lambda,\mu}=\left<S_{\lambda},s_{\mu}\right> (2.3)

for each λ,μ𝒫\lambda,\mu\in{\mathcal{P}}. \Box

Lemma 2.11.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, we have [Rλ:Lλ]q=qn(λ)[R_{\lambda}:L_{\lambda}]_{q}=q^{n(\lambda)}.

Proof.

By [19, p115] and the Frobenius reciprocity, LλL_{\lambda} contains a vector on which 𝔖λ\mathfrak{S}_{\lambda^{\prime}} acts by sign representation. Since the Vandermonde determinant offers the minimal degree realization of the sign representations of each 𝔖λj\mathfrak{S}_{\lambda^{\prime}_{j}} (1jλ11\leq j\leq\lambda_{1}), we find that Hom𝔖n(Lλ,(Rλ)m)0\mathrm{Hom}_{\mathfrak{S}_{n}}(L_{\lambda},(R_{\lambda})_{m})\neq 0 only if mn(λ)m\geq n(\lambda). It must be strict by Theorem 2.8 1). ∎

Proposition 2.12 ([10] Theorem A.4 and Corollary A.3).

We have

extAn1(Kλ,Lμ)=0λμ.\mathrm{ext}^{1}_{A_{n}}(K_{\lambda},L_{\mu})=0\hskip 14.22636pt\lambda\not\geq\mu.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, the head of KλK_{\lambda} is LλL_{\lambda}, and the socle of KλK_{\lambda} is L(n)n(λ)L_{(n)}\left<n(\lambda)\right>.

Proof.

By [10, Theorem A.4], the module KλK_{\lambda} is isomorphic to the module MλM_{\lambda} constructed there. They have the properties in the assertions by construction and [10, Theorem A.4]. ∎

Proposition 2.13 (De Concini-Procesi [4], Tanisaki [26]).

We have an isomorphism Rλn(λ)KλR_{\lambda}^{*}\left<n(\lambda)\right>\cong K_{\lambda} as graded AnA_{n}-modules.

Proof.

By Lemma 2.11, Rλn(λ)R_{\lambda}^{*}\left<n(\lambda)\right> is a graded AnA_{n}-module such that Lλ𝗁𝖽Rλn(λ)L_{\lambda}\subset\mathsf{hd}\,R_{\lambda}^{*}\left<n(\lambda)\right> and [Rλn(λ):Lμ]q=0[R_{\lambda}^{*}\left<n(\lambda)\right>:L_{\mu}]_{q}=0 if μλ\mu\not\geq\lambda and [Rλn(λ):Lλ]q=1[R_{\lambda}^{*}\left<n(\lambda)\right>:L_{\lambda}]_{q}=1. Thus, we obtain a map KλRλn(λ)K_{\lambda}\rightarrow R_{\lambda}^{*}\left<n(\lambda)\right> of graded AnA_{n}-modules. This map is injective as they share L(n)n(λ)L_{(n)}\left<n(\lambda)\right> as their socles.

We prove that KλRλn(λ)K_{\lambda}\subset R_{\lambda}^{*}\left<n(\lambda)\right> is an equality for every λ𝒫n\lambda\in{\mathcal{P}}_{n} by induction on nn. The case n=1n=1 is clear as the both are {\mathbb{C}}. Thanks to Theorem 2.7 and the induction hypothesis, we deduce that a (graded) direct summand of the head of Rλn(λ)R_{\lambda}^{*}\left<n(\lambda)\right> as A1,n1A_{1,n-1}-module must be of the shape Lλ(j)dL_{\lambda_{(j)}}\left<d\right> for 1j(λ)1\leq j\leq\ell(\lambda) and d0d\geq 0. The module Lλ(j)dL_{\lambda_{(j)}}\left<d\right> arises as the restriction of a (graded) 𝔖n\mathfrak{S}_{n}-module LμdL_{\mu}\left<d\right> (μ𝒫n\mu\in{\mathcal{P}}_{n}) such that λ(j)=μ(k)\lambda_{(j)}=\mu_{(k)} for 1k(μ)1\leq k\leq\ell(\mu). In case μ=λ\mu=\lambda, then [Rλn(λ):Lλ]q=1[R_{\lambda}^{*}\left<n(\lambda)\right>:L_{\lambda}]_{q}=1 forces Lλ(j)dLλ𝗁𝖽Kλ𝗁𝖽Rλn(λ)L_{\lambda_{(j)}}\left<d\right>\subset L_{\lambda}\subset\mathsf{hd}\,K_{\lambda}\subset\mathsf{hd}\,R_{\lambda}^{*}\left<n(\lambda)\right>.

From this, it is enough to assume μλ\mu\neq\lambda to conclude that Lλ(j)dL_{\lambda_{(j)}}\left<d\right> does not yield a non-zero module of 𝗁𝖽Rλn(λ)/Lλ\mathsf{hd}\,R_{\lambda}^{*}\left<n(\lambda)\right>/L_{\lambda}. By Theorem 2.8 2), we can assume μ>λ\mu>\lambda. Hence, μ\mu is obtained from λ\lambda by moving one box in the Young diagram to some strictly larger entries.

In case μ\mu is not the shape (mr)(m^{r}), there exists 1k(μ)1\leq k\leq\ell(\mu) such that μ(k)λ(j)\mu_{(k)}\neq\lambda_{(j)} for every 1j(λ)1\leq j\leq\ell(\lambda). It follows that Lλ(j)dLμdRλn(λ)L_{\lambda_{(j)}}\left<d\right>\subset L_{\mu}\left<d\right>\subset R_{\lambda}^{*}\left<n(\lambda)\right> contains a 𝔖n1\mathfrak{S}_{n-1}-module that is not in the head of Rλn(λ)R_{\lambda}^{*}\left<n(\lambda)\right> as A1,n1A_{1,n-1}-modules. Thus, this case does not occur.

In case μ\mu is of the shape (mr)(m^{r}), then we have λ=(mr1,(m1),1)\lambda=(m^{r-1},(m-1),1) and λ(j)=(mr1,(m1))\lambda_{(j)}=(m^{r-1},(m-1)). In this case, we have j=r+1j=r+1. In particular, grading shifts of Rλ(j)R_{\lambda_{(j)}}^{*} appears in the filtration of RλR_{\lambda}^{*} afforded by Theorem 2.7 only once, and its head is a part of LλL_{\lambda} by counting the degree. Therefore, Lλ(j)dL_{\lambda_{(j)}}\left<d\right> contributes zero in 𝗁𝖽Rλn(λ)/Lλ\mathsf{hd}\,R_{\lambda}^{*}\left<n(\lambda)\right>/L_{\lambda}.

From these, we conclude that 𝗁𝖽Rλn(λ)=Lλ\mathsf{hd}\,R_{\lambda}^{*}\left<n(\lambda)\right>=L_{\lambda} by induction hypothesis. This forces Kλ=Rλn(λ)K_{\lambda}=R_{\lambda}^{*}\left<n(\lambda)\right>, and the induction proceeds. ∎

2.2 Identification of the forms

Consider the twisted (graded) Frobenius characteristic map

Ψ:[𝒜][M]μ[M:Lμ]qSμΛq.\Psi:[{\mathcal{A}}]\ni[M]\mapsto\sum_{\mu}[M:L_{\mu}]_{q}\cdot S_{\mu}\in\Lambda_{q}. (2.4)

By Theorem 2.9, we have

Ψ([Kλ])=Qλ(λ𝒫).\Psi([K_{\lambda}])=Q_{\lambda}\hskip 14.22636pt(\lambda\in{\mathcal{P}}). (2.5)
Lemma 2.14.

For a,b𝒜a,b\in{\mathcal{A}}, we have

Ψ(ind(ab))=Ψ(a)Ψ(b),and (Ψ×Ψ)(resa)=Δ(Ψ(a)).\Psi(\mathrm{ind}\,(a\boxtimes b))=\Psi(a)\cdot\Psi(b),\hskip 14.22636pt\text{and}\hskip 14.22636pt(\Psi\times\Psi)(\mathrm{res}\,a)=\Delta(\Psi(a)).
Proof.

This is a straight-forward consequence of Lemma 1.4. The detail is left to the reader. ∎

Proposition 2.15.

We have

[Kλ],[Kμ]EP=Qλ,Qμ=δλ,μbλ.\left<[K_{\lambda}],[K_{\mu}]\right>_{EP}=\left<Q_{\lambda},Q_{\mu}\right>=\delta_{\lambda,\mu}b_{\lambda}.

In particular, we have

a,bEP=Ψ(a),Ψ(b)a,b[𝒜].\left<a,b\right>_{EP}=\left<\Psi(a),\Psi(b)\right>\hskip 14.22636pta,b\in[{\mathcal{A}}]. (2.6)
Remark 2.16.

If we prove the identities in Corollary 2.18 directly, then one can prove (2.6) without appealing to [23, 10] by Proposition 2.17 and its proof.

Proof of Proposition 2.15.

The equations in Theorem 2.10, that are equivalent to the Cauchy identity [19, (4.4)], are special cases of [23, Corollary 4.6]. It is further transformed into the main matrix equality of the so-called Lusztig-Shoji algorithm in [23, Theorem 5.4]. The latter is interpreted as the orthogonality relation with respect to ,EP\left<\bullet,\bullet\right>_{EP} in [10, Theorem 2.10]. In particular, Kostka polynomials defined in [19] and [23] are the same (for symmetric groups and the order \leq on 𝒫{\mathcal{P}}). This implies the first equality in view of (2.5). The second equality is read-off from the relation between QλQ_{\lambda} and QλQ^{\vee}_{\lambda}. The last assertion follows as {Qλ}λ𝒫\{Q_{\lambda}\}_{\lambda\in{\mathcal{P}}} forms a ((q))\mathbb{Q}(\!(q)\!)-basis of Λq\Lambda_{q}, and the Hall inner product is non-degenerate. ∎

Proposition 2.17.

For each λ𝒫\lambda\in{\mathcal{P}}, we have Ψ([Pλ])=sλ\Psi([P_{\lambda}])=s_{\lambda}.

Proof.

For each λ,μ𝒫\lambda,\mu\in{\mathcal{P}}, we have

δλ,μ=sλ,Sμ=sλ,Ψ([Lμ])\delta_{\lambda,\mu}=\left<s_{\lambda},S_{\mu}\right>=\left<s_{\lambda},\Psi([L_{\mu}])\right>

by Theorem 2.10. On the other hand, we have

δλ,μ=gdimhomAn(Pλ,Lμ)=k0(1)kgdimextAnk(Pλ,Lμ)=[Pλ],[Lμ]EP.\delta_{\lambda,\mu}=\mathrm{gdim}\,\mathrm{hom}_{A_{n}}(P_{\lambda},L_{\mu})=\sum_{k\geq 0}(-1)^{k}\mathrm{gdim}\,\mathrm{ext}^{k}_{A_{n}}(P_{\lambda},L_{\mu})=\left<[P_{\lambda}],[L_{\mu}]\right>_{EP}.

As the Hall inner product is non-degenerate (Theorem 2.10) and is the same as the Euler-Poincaré pairing (Proposition 2.15), this forces Ψ([Pλ])=sλ\Psi([P_{\lambda}])=s_{\lambda}. ∎

Corollary 2.18.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, we have

sλ\displaystyle s_{\lambda} =μ𝒫nSμgdimhom𝔖n(Lμ,Pλ)\displaystyle=\sum_{\mu\in{\mathcal{P}}_{n}}S_{\mu}\cdot\mathrm{gdim}\,\mathrm{hom}_{\mathfrak{S}_{n}}(L_{\mu},P_{\lambda})
=μ𝒫nSμgdimhom𝔖n(Lμ,Lλ[X1,,Xn])\displaystyle=\sum_{\mu\in{\mathcal{P}}_{n}}S_{\mu}\cdot\mathrm{gdim}\,\mathrm{hom}_{\mathfrak{S}_{n}}(L_{\mu},L_{\lambda}\otimes{\mathbb{C}}[X_{1},\ldots,X_{n}])
=1(1q)(1q2)(1qn)μ𝒫nSμgdimhom𝔖n(Lμ,LλR(1n)).\displaystyle=\frac{1}{(1-q)(1-q^{2})\cdots(1-q^{n})}\sum_{\mu\in{\mathcal{P}}_{n}}S_{\mu}\cdot\mathrm{gdim}\,\mathrm{hom}_{\mathfrak{S}_{n}}(L_{\mu},L_{\lambda}\otimes R_{(1^{n})}).
Proof.

In view of Proposition 2.17, the first equality is obtained by just expanding [Pλ][P_{\lambda}] using the definition of the twisted Frobenius characteristic. The second and the third equalities follow from

PλLλ[X1,,Xn]LλR(1n)[X1,,Xn]𝔖nP_{\lambda}\cong L_{\lambda}\otimes{\mathbb{C}}[X_{1},\ldots,X_{n}]\cong L_{\lambda}\otimes R_{(1^{n})}\otimes{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}}

as 𝔖n\mathfrak{S}_{n}-modules, where the latter isomorphism is standard ([3]). ∎

Corollary 2.19.

For each MAn𝗀𝗆𝗈𝖽M\in A_{n}\mathchar 45\relax\mathsf{gmod}, we have

Ψ([M])=λ[M],[Kλ]EPQλ.\Psi([M])=\sum_{\lambda}\left<[M],[K_{\lambda}]\right>_{EP}Q_{\lambda}^{\vee}.
Proof.

This follows by Ψ([Kλ])=Qλ\Psi([K_{\lambda}])=Q_{\lambda}, Theorem 2.10, and Proposition 2.15. ∎

2.3 An end\mathrm{end}-estimate

Lemma 2.20.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, the 𝔖n\mathfrak{S}_{n}-module LλL_{\lambda} contains a unique non-zero 𝔖λ\mathfrak{S}_{\lambda}-fixed vector ((up to scalar)).

Proof.

This follows from Theorem 2.8 2) and the Frobenius reciprocity. ∎

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, we set

Aλ\displaystyle A_{\lambda} :=j=1(λ)AλjAn,and\displaystyle:=\bigotimes_{j=1}^{\ell(\lambda)}A_{\lambda_{j}}\subset A_{n},\hskip 14.22636pt\text{and}\hskip 14.22636pt
K~λ+\displaystyle\widetilde{K}^{+}_{\lambda} :=AnAλ(K~(λ1)K~(λ2)K~(λ(λ))).\displaystyle:=A_{n}\otimes_{A_{\lambda}}(\widetilde{K}_{(\lambda_{1})}\boxtimes\widetilde{K}_{(\lambda_{2})}\boxtimes\cdots\boxtimes\widetilde{K}_{(\lambda_{\ell(\lambda)})}). (2.7)
Lemma 2.21.

We have K~(n)L(n)[Y]\widetilde{K}_{(n)}\cong L_{(n)}\otimes{\mathbb{C}}[Y], where [Y]{\mathbb{C}}[Y] is the quotient of the polynomial ring [X1,,Xn]{\mathbb{C}}[X_{1},\ldots,X_{n}] by the submodule generated by degree one part that is complementary to (X1++Xn){\mathbb{C}}(X_{1}+\cdots+X_{n}) as 𝔖n\mathfrak{S}_{n}-modules.

Proof.

We have P(n)[X1,,Xn]P_{(n)}\cong{\mathbb{C}}[X_{1},\ldots,X_{n}]. Its degree one part is L(n)L(n1,1)L_{(n)}\oplus L_{(n-1,1)} as 𝔖n\mathfrak{S}_{n}-modules, and quotient out by L(n1,1)L_{(n-1,1)} yields a polynomial ring [Y]{\mathbb{C}}[Y] generated by the image of (X1++Xn)L(n){\mathbb{C}}(X_{1}+\cdots+X_{n})\cong L_{(n)}. ∎

Lemma 2.22.

Let λ𝒫n\lambda\in{\mathcal{P}}_{n}. We have a unique graded AnA_{n}-module map K~λK~λ+\widetilde{K}_{\lambda}\rightarrow\widetilde{K}^{+}_{\lambda} of degree 0 up to scalar.

Proof.

We have (K~λ+)0=Ind𝔖λ𝔖n𝗍𝗋𝗂𝗏(\widetilde{K}^{+}_{\lambda})_{0}=\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}\,\mathsf{triv}, in which LλL_{\lambda} appears without multiplicity. All the 𝔖λ\mathfrak{S}_{\lambda}-modules appearing in (K~(λ1)K~(λ2))(\widetilde{K}_{(\lambda_{1})}\boxtimes\widetilde{K}_{(\lambda_{2})}\boxtimes\cdots) are trivial. It follows that [K~λ+:Lμ]q0[\widetilde{K}^{+}_{\lambda}:L_{\mu}]_{q}\neq 0 if and only if [Ind𝔖λ𝔖n𝗍𝗋𝗂𝗏:Lμ]0[\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}\,\mathsf{triv}:L_{\mu}]\neq 0. The latter implies λμ\lambda\leq\mu (Theorem 2.8). Therefore, a 𝔖n\mathfrak{S}_{n}-module map Lλ(K~λ+)0L_{\lambda}\to(\widetilde{K}^{+}_{\lambda})_{0} extends uniquely to a graded AnA_{n}-module map K~λK~λ+\widetilde{K}_{\lambda}\rightarrow\widetilde{K}^{+}_{\lambda} by the definition of K~λ\widetilde{K}_{\lambda}. ∎

In the setting of Lemma 2.22, we set

K~λ:=Im(K~λK~λ+).\widetilde{K}_{\lambda}^{\prime}:=\mathrm{Im}(\widetilde{K}_{\lambda}\rightarrow\widetilde{K}^{+}_{\lambda}).

For each 1j(λ)1\leq j\leq\ell(\lambda), we have an endomorphism ψjλ\psi_{j}^{\lambda} on K~λ+\widetilde{K}^{+}_{\lambda} extending

ψjλ(K~(λ1)K~(λ))=K~(λ1)δj,1K~(λ)δj,K~λ+,\psi_{j}^{\lambda}(\widetilde{K}_{(\lambda_{1})}\boxtimes\cdots\boxtimes\widetilde{K}_{(\lambda_{\ell})})=\widetilde{K}_{(\lambda_{1})}\left<\delta_{j,1}\right>\boxtimes\cdots\boxtimes\widetilde{K}_{(\lambda_{\ell})}\left<\delta_{j,\ell}\right>\subset\widetilde{K}_{\lambda}^{+},

obtained by the multiplication of [X1,,Xn]{\mathbb{C}}[X_{1},\ldots,X_{n}]. Consider the group

𝔖(λ):=j1𝔖mj(λ).\mathfrak{S}(\lambda):=\prod_{j\geq 1}\mathfrak{S}_{m_{j}(\lambda)}.
Lemma 2.23.

The group 𝔖(λ)\mathfrak{S}(\lambda) yields automorphisms of K~λ+\widetilde{K}^{+}_{\lambda} as AnA_{n}-modules.

Proof.

The group 𝔖(λ)\mathfrak{S}(\lambda) permutes K~(λj)\widetilde{K}_{(\lambda_{j})}s in (2.7) in such a way the size of the factors (i.e. the values of λj\lambda_{j}) are invariant. These are AλA_{\lambda}-module endomorphisms, and hence K~λ+\widetilde{K}_{\lambda}^{+} inherits these endomorphisms as required. ∎

Let B(λ)B(\lambda) denote the subring of endAn(K~λ+)\mathrm{end}_{A_{n}}(\widetilde{K}^{+}_{\lambda}) generated by {ψjλ}j=1(λ)\{\psi_{j}^{\lambda}\}_{j=1}^{\ell(\lambda)}. The action of 𝔖(λ)\mathfrak{S}(\lambda) permutes ψiλ\psi_{i}^{\lambda} and ψjλ\psi_{j}^{\lambda} such that λi=λj\lambda_{i}=\lambda_{j}. Thus, 𝔖(λ)\mathfrak{S}(\lambda) acts on B(λ)B(\lambda) as automorphisms. The invariant part B(λ)𝔖(λ)B(\lambda)^{\mathfrak{S}(\lambda)} is a polynomial ring.

Lemma 2.24.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, we have

hom𝔖n(Lλ,B(λ)L0)hom𝔖n(Lλ,K~λ+),\mathrm{hom}_{\mathfrak{S}_{n}}(L_{\lambda},B(\lambda)L_{0})\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\mathrm{hom}_{\mathfrak{S}_{n}}(L_{\lambda},\widetilde{K}^{+}_{\lambda}),

where LλL0(K~λ+)0L_{\lambda}\cong L_{0}\subset(\widetilde{K}^{+}_{\lambda})_{0} is the multiplicity one copy as 𝔖n\mathfrak{S}_{n}-modules.

Proof.

By construction, K~λ+\widetilde{K}^{+}_{\lambda} is a direct sum of (grading shifts of) copies of Ind𝔖λ𝔖n𝗍𝗋𝗂𝗏\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}\,\mathsf{triv} as a 𝔖n\mathfrak{S}_{n}-module. We have [Ind𝔖λ𝔖n𝗍𝗋𝗂𝗏:Lλ]=1[\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}\,\mathsf{triv}:L_{\lambda}]=1. The action of B(λ)B(\lambda) preserves the 𝔖n\mathfrak{S}_{n}-isotypic part. As the action of B(λ)B(\lambda) sends (K~λ+)0(\widetilde{K}^{+}_{\lambda})_{0} to all the contributions of Ind𝔖λ𝔖n𝗍𝗋𝗂𝗏\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}\,\mathsf{triv} in K~λ+\widetilde{K}^{+}_{\lambda}, we conclude the assertion. ∎

Proposition 2.25.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, we have

gdimendAn(K~λ)=bλ1and endAn(K~λ)B(λ)𝔖(λ).\mathrm{gdim}\,\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{\prime})=b_{\lambda}^{-1}\hskip 14.22636pt\text{and}\hskip 14.22636pt\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{\prime})\cong B(\lambda)^{\mathfrak{S}(\lambda)}.
Proof.

Since K~λ\widetilde{K}_{\lambda}^{\prime} has (K~λ)0Lλ(\widetilde{K}_{\lambda}^{\prime})_{0}\cong L_{\lambda} as its unique simple graded quotient, endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{\prime}) is determined by the image of (K~λ)0(\widetilde{K}_{\lambda}^{\prime})_{0}. In addition, K~λ\widetilde{K}_{\lambda}^{\prime} is stable under the action of 𝔖(λ)\mathfrak{S}(\lambda) as Lλ(K~λ+)0L_{\lambda}\subset(\widetilde{K}_{\lambda}^{+})_{0} is. Therefore, Lemma 2.24 implies endAn(K~λ)B(λ)𝔖(λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{\prime})\subset B(\lambda)^{\mathfrak{S}(\lambda)}. Thus, we have the inequality \leq in the assertion by

bλ1\displaystyle b_{\lambda}^{-1} =j11(1q)(1qmj(λ))\displaystyle=\prod_{j\geq 1}\frac{1}{(1-q)\cdots(1-q^{m_{j}(\lambda)})}
=j1gdim[x1,,xmj(λ)]𝔖mj(λ)=gdimB(λ)𝔖(λ)\displaystyle=\prod_{j\geq 1}\mathrm{gdim}\,{\mathbb{C}}[x_{1},\ldots,x_{m_{j}(\lambda)}]^{\mathfrak{S}_{m_{j}(\lambda)}}=\mathrm{gdim}\,B(\lambda)^{\mathfrak{S}(\lambda)}

(see Corollary 2.18 for the second equality). We have an identification

(K~(λ1)K~(λ(λ)))e(𝔖(λ)(K~(λ1)K~(λ(λ))))K~λ,(\widetilde{K}_{(\lambda_{1})}\boxtimes\cdots\boxtimes\widetilde{K}_{(\lambda_{\ell(\lambda)})})\cong e\left({\mathbb{C}}\mathfrak{S}(\lambda)\otimes(\widetilde{K}_{(\lambda_{1})}\boxtimes\cdots\boxtimes\widetilde{K}_{(\lambda_{\ell(\lambda)})})\right)\subset\widetilde{K}_{\lambda}, (2.8)

where e=1|𝔖(λ)|w𝔖(λ)we=\frac{1}{|\mathfrak{S}(\lambda)|}\sum_{w\in\mathfrak{S}(\lambda)}w. The actions of ψ1λ,,ψ(λ)λ\psi_{1}^{\lambda},\ldots,\psi_{\ell(\lambda)}^{\lambda} on the first term of (2.8) are induced by the multiplication of [X1,,Xn]{\mathbb{C}}[X_{1},\ldots,X_{n}]. Hence, the action of B(λ)𝔖(λ)𝔖λ=B(λ)𝔖(λ)B(\lambda)^{\mathfrak{S}(\lambda)\ltimes\mathfrak{S}_{\lambda}}=B(\lambda)^{\mathfrak{S}(\lambda)} on the first two terms of (2.8) are realized by the multiplication of [X1,,Xn]{\mathbb{C}}[X_{1},\ldots,X_{n}]. Thus, the inequality must be in fact an equality and endAn(K~λ)=B(λ)𝔖(λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{\prime})=B(\lambda)^{\mathfrak{S}(\lambda)}. ∎

Let us consider the image of the center [X1,,Xn]𝔖n{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}} in endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}) and endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{\prime}) by Z(λ)Z(\lambda) and Z(λ)Z^{\prime}(\lambda), respectively.

Lemma 2.26.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, we have a quotient map

endAn(K~λ)endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda})\longrightarrow\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{\prime})

as an algebra that induces a surjection Z(λ)Z(λ)Z(\lambda)\to Z^{\prime}(\lambda). In addition, Z(λ)Z^{\prime}(\lambda) is precisely the image of [X1,,Xn]𝔖n{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}} in endAn(K~λ+)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{+}).

Proof.

By the construction of K~λ\widetilde{K}_{\lambda}, we have

endAn(K~λ)homAn(K~λ,K~λ)hom𝔖n(Lλ,K~λ).\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda})\longrightarrow\!\!\!\!\!\rightarrow\mathrm{hom}_{A_{n}}(\widetilde{K}_{\lambda},\widetilde{K}_{\lambda}^{\prime})\cong\mathrm{hom}_{\mathfrak{S}_{n}}(L_{\lambda},\widetilde{K}_{\lambda}^{\prime}).

In view of Proposition 2.25 and Lemma 2.24, we have

homAn(K~λ,K~λ)hom𝔖n(Lλ,K~λ)endAn(K~λ).\mathrm{hom}_{A_{n}}(\widetilde{K}_{\lambda},\widetilde{K}_{\lambda}^{\prime})\cong\mathrm{hom}_{\mathfrak{S}_{n}}(L_{\lambda},\widetilde{K}_{\lambda}^{\prime})\cong\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{\prime}).

This proves the first assertion, as the action of [X1,,Xn]𝔖n{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}} on K~λ\widetilde{K}_{\lambda}^{\prime} factors through K~λ\widetilde{K}_{\lambda}. The second assertion follows as the action of [X1,,Xn]𝔖n{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}} on K~λ+\widetilde{K}_{\lambda}^{+} factors through B(λ)𝔖(λ)=endAn(K~λ)B(\lambda)^{\mathfrak{S}(\lambda)}=\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}^{\prime}). ∎

Lemma 2.27.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, the algebra endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}) is a finitely generated module over [X1,,Xn]𝔖n{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}}.

Proof.

Since we have a surjection endAn(Pλ)endAn(K~λ)\mathrm{end}_{A_{n}}(P_{\lambda})\to\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}), it suffices to see that endAn(Pλ)\mathrm{end}_{A_{n}}(P_{\lambda}) is a finitely generated module over [X1,,Xn]𝔖n{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}}. We have

endAn(Pλ)Hom𝔖n(L(n),End(Lλ)[X1,,Xn]).\mathrm{end}_{A_{n}}(P_{\lambda})\cong\mathrm{Hom}_{\mathfrak{S}_{n}}(L_{(n)},\mathrm{End}_{\mathbb{C}}(L_{\lambda})\otimes{\mathbb{C}}[X_{1},\ldots,X_{n}]).

The RHS is a finitely generated module over [X1,,Xn]𝔖n{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}} as required. ∎

For two power series with integer coefficients

f(q)=mfmqm,g(q)=mgmqm((q)),f(q)=\sum_{m}f_{m}q^{m},g(q)=\sum_{m}g_{m}q^{m}\in\mathbb{Z}(\!(q)\!),

we say f(q)g(q)f(q)\leq g(q) if we have fmgmf_{m}\leq g_{m} for every mm\in\mathbb{Z}. We say f(q)g(q)f(q)\ll g(q) if

limmsup{fkkm}sup{gkkm}=0.\lim_{m\to\infty}\frac{\sup\{f_{k}\mid k\leq m\}}{\sup\{g_{k}\mid k\leq m\}}=0. (2.9)
Theorem 2.28 ([20]).

Let RR be a finitely generated graded integral algebra with =R0{\mathbb{C}}=R_{0}, and let SS be its proper graded quotient algebra. For a finitely generated graded SS-module MM, we have

gdimMgdimR.\mathrm{gdim}\,M\ll\mathrm{gdim}\,R.
Proof.

This follows form [20, Theorem 13.4] if we take into account the Krull dimension inequality dimR>dimS\dim\,R>\dim\,S, and the completion with respect to the grading makes RR and SS into local rings. ∎

Lemma 2.29.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n} and an algebra quotient Z(λ)Z^{\prime}(\lambda)\to{\mathbb{C}}, the actions of X1,X2,,XnX_{1},X_{2},\ldots,X_{n} on Z(λ)K~λ{\mathbb{C}}\otimes_{Z^{\prime}(\lambda)}\widetilde{K}_{\lambda}^{\prime} and Z(λ)K~λ+{\mathbb{C}}\otimes_{Z^{\prime}(\lambda)}\widetilde{K}_{\lambda}^{+} have joint eigenvalues of shape

α1==αλ1,αλ1+1==αλ1+λ2,,αnλ(λ)+1==αn\alpha_{1}=\cdots=\alpha_{\lambda_{1}},\alpha_{\lambda_{1}+1}=\cdots=\alpha_{\lambda_{1}+\lambda_{2}},\ldots,\alpha_{n-\lambda_{\ell(\lambda)}+1}=\cdots=\alpha_{n} (2.10)

up to 𝔖n\mathfrak{S}_{n}-permutation.

Proof.

By Lemma 2.27, the modules Z(λ)K~λ{\mathbb{C}}\otimes_{Z^{\prime}(\lambda)}\widetilde{K}_{\lambda}^{\prime} and Z(λ)K~λ+{\mathbb{C}}\otimes_{Z^{\prime}(\lambda)}\widetilde{K}_{\lambda}^{+} must be finite-dimensional. Hence, the actions of X1,,XnX_{1},\ldots,X_{n} have joint eigenvalues. Their values can be read-off from (2.7). ∎

Theorem 2.30.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, we have

gdimker(endAn(K~λ)endAn(K~λ))gdimendAn(K~λ).\mathrm{gdim}\,\ker\left(\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda})\to\mathrm{end}_{A_{n}}(\widetilde{K}^{\prime}_{\lambda})\right)\ll\mathrm{gdim}\,\mathrm{end}_{A_{n}}(\widetilde{K}^{\prime}_{\lambda}).
Proof.

We set Z:=Z(λ)Z:=Z(\lambda) and Z:=Z(λ)Z^{\prime}:=Z^{\prime}(\lambda) during this proof. The specialization ZK~λ{\mathbb{C}}\otimes_{Z}\widetilde{K}_{\lambda} with respect to a maximal ideal 𝔫Z\mathfrak{n}\subset Z decomposes into the generalized eigenspaces with respect to X1,,XnX_{1},\ldots,X_{n}, whose set of joint eigenvalues in {\mathbb{C}} have multiplicities μ1,μ2,,μ\mu_{1},\mu_{2},\ldots,\mu_{\ell} that constitute a partition μ\mu of nn. We have

[ZK~λ:Lγ]𝔖n=0λγ[{\mathbb{C}}\otimes_{Z}\widetilde{K}_{\lambda}:L_{\gamma}]_{\mathfrak{S}_{n}}=0\hskip 14.22636pt\lambda\not\leq\gamma (2.11)

by the definition of K~λ\widetilde{K}_{\lambda} and the fact that 𝔖n\mathfrak{S}_{n} has semi-simple representation theory. Being the cyclic AnA_{n}-module generator, we have [ZK~λ:Lλ]𝔖n0[{\mathbb{C}}\otimes_{Z}\widetilde{K}_{\lambda}:L_{\lambda}]_{\mathfrak{S}_{n}}\neq 0.

We can choose a non-zero generalized eigenspace

MZK~λM\subset{\mathbb{C}}\otimes_{Z}\widetilde{K}_{\lambda}

of X1,,XnX_{1},\ldots,X_{n} that can be regarded as an (ungraded) AλA_{\lambda}-module. We choose

Lμ[1]Lμ[2]Lμ[]M=(μ)L_{\mu^{[1]}}\boxtimes L_{\mu^{[2]}}\boxtimes\cdots\boxtimes L_{\mu^{[\ell]}}\subset M\hskip 14.22636pt\ell=\ell(\mu) (2.12)

as 𝔖μ\mathfrak{S}_{\mu}-modules with partitions μ[1],,μ[]\mu^{[1]},\ldots,\mu^{[\ell]} of μ1,,μ\mu_{1},\ldots,\mu_{\ell}, respectively. Since each piece of the external tensor products of (2.12) have distinct XX-eigenvalues, we deduce

Ind𝔖μ𝔖n(Lμ[1]Lμ[2]Lμ[])M.\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\mu}}(L_{\mu^{[1]}}\boxtimes L_{\mu^{[2]}}\boxtimes\cdots\boxtimes L_{\mu^{[\ell]}})\hookrightarrow M. (2.13)

By the Littlewood-Richardson rule, the smallest label (with respect to \leq) of 𝔖n\mathfrak{S}_{n}-module that appears in the LHS of (2.13) is attained by κ𝒫n\kappa\in{\mathcal{P}}_{n} such that

mi(κ)=j=1mi(μ[j])i1.m_{i}(\kappa)=\sum_{j=1}^{\ell}m_{i}(\mu^{[j]})\hskip 14.22636pti\geq 1.

For an appropriate choice in (2.12), we attain κ=λ\kappa=\lambda by Lemma 1.3. It follows that μ\mu defines a division of entries of λ\lambda into small groups. In view of (2.10), the maximal ideal 𝔫Z\mathfrak{n}\subset Z is the pullback of a maximal ideal of ZZ^{\prime}. In other words, we find that ZZ shares with the same support as ZZ^{\prime} in Spec[X1,,Xn]𝔖n\mbox{\rm Spec}\,{\mathbb{C}}[X_{1},\ldots,X_{n}]^{\mathfrak{S}_{n}}.

We define graded AnA_{n}-modules NrN_{r} (r1r\geq 1) as:

Nr:=ker(endAn(K~λ)endAn(K~λ))r/(ker(endAn(K~λ)endAn(K~λ)))r+1.N_{r}:=\ker\left(\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda})\to\mathrm{end}_{A_{n}}(\widetilde{K}^{\prime}_{\lambda})\right)^{r}/\left(\ker\left(\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda})\to\mathrm{end}_{A_{n}}(\widetilde{K}^{\prime}_{\lambda})\right)\right)^{r+1}.

We show that each NrN_{r} is supported in a proper subset of SpecZ\mbox{\rm Spec}\,Z^{\prime}. Equivalently, we show that general specializations of endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}) and endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}^{\prime}_{\lambda}) with respect to (2.10) are the same. In view of the above construction of the partitions μ\mu and κ\kappa, we have necessarily λ=μ\lambda=\mu and μ(i)=(μi)\mu^{(i)}=(\mu_{i}) for each i1i\geq 1 as otherwise smaller partitions arise. By Lemma 2.21, a thickening of (2.13) as (ungraded) AλA_{\lambda}-modules must be achieved by the actions of

X1++Xλ1,Xλ1+1++Xλ1+λ2,,Xnλ(λ)+1++Xn.X_{1}+\cdots+X_{\lambda_{1}},X_{\lambda_{1}+1}+\cdots+X_{\lambda_{1}+\lambda_{2}},\ldots,X_{n-\lambda_{\ell(\lambda)}+1}+\cdots+X_{n}. (2.14)

As these are contained in the action of B(λ)B(\lambda), we conclude that general specializations of endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}) and endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}^{\prime}_{\lambda}) are the same.

Therefore, Theorem 2.28 implies

gdimNrgdimendAn(K~λ)r>0.\mathrm{gdim}\,N_{r}\ll\mathrm{gdim}\,\mathrm{end}_{A_{n}}(\widetilde{K}^{\prime}_{\lambda})\hskip 14.22636ptr>0.

By Lemma 2.27 (and the support containment), we have only finitely many rr with Nr{0}N_{r}\neq\{0\}. Again using Theorem 2.28, we conclude

gdimendAn(K~λ)gdimendAn(K~λ)=r1gdimNrgdimendAn(K~λ)\mathrm{gdim}\,\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda})-\mathrm{gdim}\,\mathrm{end}_{A_{n}}(\widetilde{K}^{\prime}_{\lambda})=\sum_{r\geq 1}\mathrm{gdim}\,N_{r}\ll\mathrm{gdim}\,\mathrm{end}_{A_{n}}(\widetilde{K}^{\prime}_{\lambda})

as required. ∎

Proposition 2.31.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n}, the module K~λ\widetilde{K}_{\lambda}^{\prime} admits a decreasing separable filtration whose associated graded is the direct sum of grading shifts of KλK_{\lambda}.

Proof.

Consider the submodule N~K~λ+\widetilde{N}\subset\widetilde{K}_{\lambda}^{+} generated by the unique copy L(n)Ind𝔖λ𝔖n𝗍𝗋𝗂𝗏=(K~λ+)0L_{(n)}\subset\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}\mathsf{triv}=(\widetilde{K}_{\lambda}^{+})_{0}. In view of Lemma 2.24, we find

hom𝔖n(L(n),N~)B(λ)𝔖(λ)hom𝔖n(Lλ,K~λ).\hom_{\mathfrak{S}_{n}}(L_{(n)},\widetilde{N})\cong B(\lambda)^{\mathfrak{S}(\lambda)}\cong\hom_{\mathfrak{S}_{n}}(L_{\lambda},\widetilde{K}_{\lambda}^{\prime}). (2.15)

Consequently, we have endAn(N~)=B(λ)𝔖(λ)\mathrm{end}_{A_{n}}(\widetilde{N})=B(\lambda)^{\mathfrak{S}(\lambda)}.

Let NN and KK be the specializations of N~\widetilde{N} and K~λ\widetilde{K}_{\lambda} with respect to a maximal ideal of B(λ)𝔖(λ)B(\lambda)^{\mathfrak{S}(\lambda)} such that the joint eigenvalues αλ1,αλ1+λ2,,α(λ)\alpha_{\lambda_{1}},\alpha_{\lambda_{1}+\lambda_{2}},\ldots,\alpha_{\ell(\lambda)} in Lemma 2.29 are distinct. Let MM be a joint {Xi}i\{X_{i}\}_{i}-eigenspace of KK or NN, that is a 𝔖λ\mathfrak{S}_{\lambda}-module. The distinct eigenvalue condition implies

Ind𝔖λ𝔖nMNor Ind𝔖λ𝔖nMK.\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}M\subset N\hskip 8.53581pt\text{or}\hskip 8.53581pt\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}M\subset K. (2.16)

The 𝔖n\mathfrak{S}_{n}-module LμL_{\mu} appears in NN or KK only if LμInd𝔖λ𝔖n𝗍𝗋𝗂𝗏L_{\mu}\subset\mathrm{Ind}^{\mathfrak{S}_{n}}_{\mathfrak{S}_{\lambda}}\mathsf{triv}. Applying the Littlewood-Richardson rule to the middle term of (2.16), we deduce M𝗍𝗋𝗂𝗏M\cong\mathsf{triv}. In particular, we have [N:Lλ]𝔖n>0<[K:L(n)]𝔖n[N:L_{\lambda}]_{\mathfrak{S}_{n}}>0<[K:L_{(n)}]_{\mathfrak{S}_{n}}. By the semi-continuity of the specializations, we deduce

[0B(λ)𝔖(λ)N~:Lλ]>0,and [0B(λ)𝔖(λ)K~λ:L(n)]>0.[{\mathbb{C}}_{0}\otimes_{B(\lambda)^{\mathfrak{S}(\lambda)}}\widetilde{N}:L_{\lambda}]>0,\hskip 8.53581pt\text{and}\hskip 8.53581pt[{\mathbb{C}}_{0}\otimes_{B(\lambda)^{\mathfrak{S}(\lambda)}}\widetilde{K}_{\lambda}^{\prime}:L_{(n)}]>0. (2.17)

From this, we conclude 0B(λ)𝔖(λ)K~λKλ{\mathbb{C}}_{0}\otimes_{B(\lambda)^{\mathfrak{S}(\lambda)}}\widetilde{K}_{\lambda}^{\prime}\cong K_{\lambda}. Thus, the torsion free B(λ)𝔖(λ)B(\lambda)^{\mathfrak{S}(\lambda)}-action on K~λK~λ+\widetilde{K}_{\lambda}^{\prime}\subset\widetilde{K}_{\lambda}^{+} yields the assertion. ∎

Corollary 2.32.

Keep the setting of Proposition 2.31. We have Ψ([K~λ])=Qλ\Psi([\widetilde{K}^{\prime}_{\lambda}])=Q^{\vee}_{\lambda}.

Proof.

Compare Propositions 2.25 and 2.31 with (2.2) and (2.5). ∎

Corollary 2.33.

For each λ𝒫n\lambda\in{\mathcal{P}}_{n} and an AnA_{n}-module proper quotient MλM_{\lambda} of K~λ\widetilde{K}_{\lambda}^{\prime}, we have [Lλ:Mλ]qbλ1[L_{\lambda}:M_{\lambda}]_{q}\ll b_{\lambda}^{-1}.

Proof.

We borrow the setting of the proof of Proposition 2.31. Since 𝗌𝗈𝖼Kλ=L(n)\mathsf{soc}\,K_{\lambda}=L_{(n)}, we find L(n)mker(K~λMλ)L_{(n)}\left<m\right>\subset\ker\,(\widetilde{K}^{\prime}_{\lambda}\to M_{\lambda}) for some m>0m\in\mathbb{Z}_{>0}. As all the copies of L(n)L_{(n)} and LλL_{\lambda} in K~λ+\widetilde{K}^{+}_{\lambda} are obtained by the B(λ)B(\lambda)-action from unique copies at degree zero, we find m>0m^{\prime}\in\mathbb{Z}_{>0} such that

N~mK~λand K~λmN~inside K~λ+ as An-modules.\widetilde{N}\left<m\right>\subset\widetilde{K}_{\lambda}^{\prime}\hskip 8.53581pt\text{and}\hskip 8.53581pt\widetilde{K}_{\lambda}^{\prime}\left<m^{\prime}\right>\subset\widetilde{N}\hskip 8.53581pt\text{inside $\widetilde{K}^{+}_{\lambda}$ as $A_{n}$-modules.}

This forces K~λm+mN~K~λ\widetilde{K}_{\lambda}^{\prime}\left<m^{\prime}+m\right>\subset\widetilde{N}\subset\widetilde{K}_{\lambda}^{\prime} to be zero in MλM_{\lambda}. Therefore, we have

gdimhom𝔖n(Lλ,Mλ)(1qm+m)gdimhom𝔖n(Lλ,K~λ)=(1qm+m)bλ1.\mathrm{gdim}\,\mathrm{hom}_{\mathfrak{S}_{n}}(L_{\lambda},M_{\lambda})\leq(1-q^{m+m^{\prime}})\mathrm{gdim}\,\mathrm{hom}_{\mathfrak{S}_{n}}(L_{\lambda},\widetilde{K}_{\lambda})=(1-q^{m+m^{\prime}})b_{\lambda}^{-1}.

This implies the assertion. ∎

Corollary 2.34.

Let λ𝒫n\lambda\in{\mathcal{P}}_{n}. Assume that MM is a graded AnA_{n}-module generated by the subspace

Mtopj=1mLλdjMsuch that [M:Lμ]q={bλ1j=1mqdj(μ=λ)0(μλ).M^{\mathrm{top}}\cong\bigoplus_{j=1}^{m}L_{\lambda}\left<d_{j}\right>\subset M\hskip 14.22636pt\text{such that}\hskip 14.22636pt[M:L_{\mu}]_{q}=\begin{cases}b_{\lambda}^{-1}\sum_{j=1}^{m}q^{d_{j}}&(\mu=\lambda)\\ 0&(\mu\not\geq\lambda)\end{cases}.

Then, we have Mj=1mK~λdjM\cong\bigoplus_{j=1}^{m}\widetilde{K}^{\prime}_{\lambda}\left<d_{j}\right>.

Proof.

We have a surjection

f:j=1mK~λdjM.f:\bigoplus_{j=1}^{m}\widetilde{K}_{\lambda}\left<d_{j}\right>\longrightarrow\!\!\!\!\!\rightarrow M.

Consider the quotient MM^{\prime} of MM by j=1mf(ker(K~λK~λ)dj)\sum_{j=1}^{m}f(\ker(\widetilde{K}_{\lambda}\to\widetilde{K}_{\lambda}^{\prime})\left<d_{j}\right>). Let f:j=1mK~λdjMf^{\prime}:\bigoplus_{j=1}^{m}\widetilde{K}_{\lambda}^{\prime}\left<d_{j}\right>\to M^{\prime} be the map induced from ff. Let us choose a maximal subset S{1,,m}S\subset\{1,\ldots,m\} such that jSK~λdj\bigoplus_{j\in S}\widetilde{K}_{\lambda}^{\prime}\left<d_{j}\right> injects into MM^{\prime} by ff^{\prime}. We take the quotient M′′M^{\prime\prime} of MM^{\prime} by this image. Then, the image KjK_{j} of K~λdj\widetilde{K}_{\lambda}^{\prime}\left<d_{j}\right> (jSj\not\in S) in M′′M^{\prime\prime} under the induced map must be a proper quotient of K~λdj\widetilde{K}_{\lambda}^{\prime}\left<d_{j}\right>.

Suppose that S{1,,m}S\neq\{1,\ldots,m\}. Corollary 2.33 and Theorem 2.28 forces

[M′′:Lλ]qjS[Kj:Lλ]q[K~λ:Lλ]q.[M^{\prime\prime}:L_{\lambda}]_{q}\leq\sum_{j\not\in S}[K_{j}:L_{\lambda}]_{q}\ll[\widetilde{K}^{\prime}_{\lambda}:L_{\lambda}]_{q}.

By Theorem 2.30, we have [M:Lλ]q[M:Lλ]qbλ1[M:L_{\lambda}]_{q}-[M^{\prime}:L_{\lambda}]_{q}\ll b_{\lambda}^{-1}. Thus, we have

[M:Lλ]qjSqj[K~λ:Lλ]qjSmqdj[K~λ:Lλ]q,[M:L_{\lambda}]_{q}-\sum_{j\in S}q^{j}[\widetilde{K}^{\prime}_{\lambda}:L_{\lambda}]_{q}\ll\sum_{j\not\in S}^{m}q^{d_{j}}[\widetilde{K}^{\prime}_{\lambda}:L_{\lambda}]_{q},

that is a contradiction by [K~λ:Lλ]q=bλ1[\widetilde{K}^{\prime}_{\lambda}:L_{\lambda}]_{q}=b_{\lambda}^{-1}. Therefore, we have S={1,2,,m}S=\{1,2,\ldots,m\}. This implies that ff^{\prime} is an isomorphism. By Propositions 2.31 and 2.12, we conclude M=MM=M^{\prime} by the comparison of graded multiplicities. ∎

2.4 Proof of Theorem 2.3

We prove Theorem 2.3 and K~λ=K~λ\widetilde{K}_{\lambda}=\widetilde{K}_{\lambda}^{\prime} (λ𝒫n\lambda\in{\mathcal{P}}_{n}) by induction on nn. Theorem 2.3 holds for n=1n=1 as 𝒫1={(1)}{\mathcal{P}}_{1}=\{(1)\}, P(1)=K~(1)=K~(1)=[X]P_{(1)}=\widetilde{K}_{(1)}=\widetilde{K}_{(1)}^{\prime}={\mathbb{C}}[X], K(1)=K_{(1)}={\mathbb{C}}, and

ext[X]k([X],)δk,0.\mathrm{ext}^{k}_{{\mathbb{C}}[X]}({\mathbb{C}}[X],{\mathbb{C}})\cong{\mathbb{C}}^{\delta_{k,0}}.

We assume the assertion for all 1n<n01\leq n<n_{0} and prove the assertion for n=n0n=n_{0}. We fix λ𝒫n01\lambda\in{\mathcal{P}}_{n_{0}-1} and set

ind(λ):=ind1,n01([X]K~λ).\mathrm{ind}(\lambda):=\mathrm{ind}_{1,n_{0}-1}({\mathbb{C}}[X]\boxtimes\widetilde{K}_{\lambda}).

For each μ𝒫n0\mu\in{\mathcal{P}}_{n_{0}} and kk\in\mathbb{Z}, Theorem 1.5 implies

extAn0k(ind(λ),Kμ)extA1,n01k([X]K~λ,Kμ).\mathrm{ext}^{k}_{A_{n_{0}}}(\mathrm{ind}(\lambda),K_{\mu}^{*})\cong\mathrm{ext}^{k}_{A_{1,n_{0}-1}}({\mathbb{C}}[X]\boxtimes\widetilde{K}_{\lambda},K_{\mu}^{*}). (2.18)

Since [X]{\mathbb{C}}[X] is projective as [X]{\mathbb{C}}[X]-modules, Theorem 2.7 implies that

gdimextA1,n01k([X]K~λ,Kμ){1j(μ),λ=μ(j)qn(μ)n(μ(j))+j(k=0)0(k0)\mathrm{gdim}\,\mathrm{ext}^{k}_{A_{1,n_{0}-1}}({\mathbb{C}}[X]\boxtimes\widetilde{K}_{\lambda},K_{\mu}^{*})\cong\begin{cases}\sum_{1\leq j\leq\ell(\mu),\lambda=\mu_{(j)}}q^{n(\mu)-n(\mu_{(j)})+j}&(k=0)\\ 0&(k\neq 0)\end{cases} (2.19)

by the short exact sequences associated to (2.1). In other word, we have

gdimhomA1,n01([X]K~λ,Kμ)=q[mj(μ)]q.\mathrm{gdim}\,\mathrm{hom}_{A_{1,n_{0}-1}}({\mathbb{C}}[X]\boxtimes\widetilde{K}_{\lambda},K_{\mu}^{*})=q^{\star}[m_{j}(\mu)]_{q}.

and it is nonzero if and only if μ(j)=λ\mu_{(j)}=\lambda for some 1j(μ)1\leq j\leq\ell(\mu). This is equivalent to λ(j)=μ\lambda^{(j)}=\mu for some 1j(λ)+11\leq j\leq\ell(\lambda)+1. We set S:={λ(j)}j=1(λ)+1𝒫n0S:=\{\lambda^{(j)}\}_{j=1}^{\ell(\lambda)+1}\subset\mathcal{P}_{n_{0}}.

Note that Lμ=𝗌𝗈𝖼KμL_{\mu}=\mathsf{soc}\,K_{\mu}^{*}, and hence every 0fhomAn0(ind(λ),Kμ)0\neq f\in\mathrm{hom}_{A_{n_{0}}}(\mathrm{ind}(\lambda),K_{\mu}^{*}) satisfies [Imf:Lμ]q0[\mathrm{Im}\,f:L_{\mu}]_{q}\neq 0. In view of Lemma 1.3, we further deduce [Imf:Lμ]=1[\mathrm{Im}\,f:L_{\mu}]=1. Therefore, the image of the map

f+:ind(λ)(Kμ)f^{+}:\mathrm{ind}(\lambda)\longrightarrow\left(K_{\mu}^{*}\right)^{\oplus\star}

obtained by taking the sum of all the maps of homAn0(ind(λ),Kμ)\mathrm{hom}_{A_{n_{0}}}(\mathrm{ind}(\lambda),K_{\mu}^{*}) satisfies

  • 𝗌𝗈𝖼Imf+\mathsf{soc}\,\mathrm{Im}\,f^{+} is the direct sum of LμmL_{\mu}\left<m\right> (mm\in\mathbb{Z});

  • dim(𝗌𝗈𝖼Imf+)=(dimLμ)(dimhomAn0(ind(λ),Kμ))\dim\,(\mathsf{soc}\,\mathrm{Im}\,f^{+})=(\dim\,L_{\mu})\cdot(\dim\,\mathrm{hom}_{A_{n_{0}}}(\mathrm{ind}(\lambda),K_{\mu}^{*})).

We consider an An0A_{n_{0}}-submodule generated by the preimage of (𝗌𝗈𝖼Imf+)(\mathsf{soc}\,\mathrm{Im}\,f^{+}) (considered as the direct sum of grading shifts of LμL_{\mu}), that we denote by NμN_{\mu}. Although the module NμN_{\mu} might depend on the choice of a lift, the number of its An0A_{n_{0}}-module generators is unambiguously determined.

We have λ(j)λ(j+1)\lambda^{(j)}\geq\lambda^{(j+1)} for 1j(λ)1\leq j\leq\ell(\lambda) by inspection. In particular, SS is a totally ordered set with respect to \leq. Moreover, ind(λ)\mathrm{ind}(\lambda) is generated by Ind1,n01Lλ\mathrm{Ind}_{1,n_{0}-1}L_{\lambda} as an An0A_{n_{0}}-module, and an irreducible constituent of Ind1,n01Lλ\mathrm{Ind}_{1,n_{0}-1}L_{\lambda} is of the form Lλ(j)L_{\lambda^{(j)}} for 1j((λ)+1)1\leq j\leq(\ell(\lambda)+1) by the Littlewood-Richardson rule. As a consequence, we find that γSNγ=ind(λ)\sum_{\gamma\in S}N_{\gamma}=\mathrm{ind}(\lambda). For each 1j(λ)+11\leq j\leq\ell(\lambda)+1, we set N(j):=ijNλ(i)N(j):=\sum_{i\geq j}N_{\lambda^{(i)}}. We have N(j+1)N(j)N(j+1)\subset N(j) for 1j(λ)1\leq j\leq\ell(\lambda) and N(1)=ind(λ)N(1)=\mathrm{ind}(\lambda).

By the Littlewood-Richardson rule and Lemma 1.3, we find that

[ind(λ):Lγ]q0only if γλ((λ)+1).[\mathrm{ind}(\lambda):L_{\gamma}]_{q}\neq 0\hskip 14.22636pt\text{only if}\hskip 14.22636pt\gamma\geq\lambda^{(\ell(\lambda)+1)}. (2.20)
Claim A.

We have [N(j)/N(j+1):Lγ]q=0[N(j)/N(j+1):L_{\gamma}]_{q}=0 for γ<λ(j)\gamma<\lambda^{(j)}.

Proof.

Assume to the contrary to deduce contradiction. We have some 1j(λ)1\leq j\leq\ell(\lambda) and γ<λ(j)\gamma<\lambda^{(j)} such that [N(j)/N(j+1):Lγ]q0[N(j)/N(j+1):L_{\gamma}]_{q}\neq 0. Here we have λ((λ)+1)γ<λ(j)\lambda^{(\ell(\lambda)+1)}\leq\gamma<\lambda^{(j)} by (2.20). By rearranging jj, we assume that jj is the minimal number with this property. In particular, we have

[N(l)/N(l+1):Lγ]q=0γ<λ(l)for l<j.[N(l)/N(l+1):L_{\gamma}]_{q}=0\hskip 14.22636pt\gamma<\lambda^{(l)}\hskip 8.53581pt\text{for}\hskip 8.53581ptl<j. (2.21)

This in turn implies that [N(l)/N(j):Lγ]q=0[N(l)/N(j):L_{\gamma}]_{q}=0 for γ<λ(j)\gamma<\lambda^{(j)} for every ljl\leq j. By rearranging γ\gamma if necessary, we can assume that the An0A_{n_{0}}-submodule N(j)N(j)/N(j+1)N^{-}(j)\subset N(j)/N(j+1) generated by 𝔖n0\mathfrak{S}_{n_{0}}-isotypic components LκL_{\kappa} such that κ<λ(j)\kappa<\lambda^{(j)} satisfies Lγm𝗁𝖽N(j)L_{\gamma}\left<m\right>\subset\mathsf{hd}\,N^{-}(j) and the value mm is minimum among all γ<λ(j)\gamma<\lambda^{(j)}. Then, the lift of Lγm𝗁𝖽N(j)L_{\gamma}\left<m\right>\subset\mathsf{hd}\,N^{-}(j) to N(j)N^{-}(j) is uniquely determined as graded 𝔖n0\mathfrak{S}_{n_{0}}-module. It follows that the maximal quotient Lγ+L_{\gamma}^{+} of N(j)/N(j+1)N(j)/N(j+1) (and hence also a quotient of N(j)N(j)) such that 𝗌𝗈𝖼Lγ+=Lγm\mathsf{soc}\,L_{\gamma}^{+}=L_{\gamma}\left<m\right> is finite-dimensional (as the grading must be bounded) and [Lγ+:Lκ]q=0[L_{\gamma}^{+}:L_{\kappa}]_{q}=0 if κ<γ(<λ(j))\kappa<\gamma(<\lambda^{(j)}). By Proposition 2.12 and Theorem 1.6, we find

extAn01(coker(LγLγ+),Kγ)=0\mathrm{ext}^{1}_{A_{n_{0}}}(\mathrm{coker}\,(L_{\gamma}\to L_{\gamma}^{+}),K_{\gamma}^{*})=0

by a repeated applications of the short exact sequences. In particular, the non-zero map LγmKγmL_{\gamma}\left<m\right>\to K_{\gamma}^{*}\left<m\right> prolongs to Lγ+L_{\gamma}^{+}, and hence it gives rise to a map N(j)KγmN(j)\rightarrow K_{\gamma}^{*}\left<m\right>. By (2.21), we additionally have

extAn01(ind(λ)/N(j),Kγ)=0.\mathrm{ext}^{1}_{A_{n_{0}}}(\mathrm{ind}(\lambda)/N(j),K_{\gamma}^{*})=0.

Therefore, we deduce a non-zero map ind(λ)Kγm\mathrm{ind}(\lambda)\rightarrow K_{\gamma}^{*}\left<m\right> from our assumption that does not come from the generator set of Nλ(l)N_{\lambda^{(l)}} for every ll. This is a contradiction, and hence we conclude the result. ∎

We return to the proof of Theorem 2.3. Note that Claim A guarantees that N(j)N(j) (1j(λ+1)1\leq j\leq\ell(\lambda+1)) is defined unambiguously as all the possible generating 𝔖n0\mathfrak{S}_{n_{0}}-isotypical components of N(j)ind(λ)N(j)\subset\mathrm{ind}(\lambda) (i.e. Lλ(k)L_{\lambda^{(k)}} for jk(λ)+1j\leq k\leq\ell(\lambda)+1) must belong to N(j)N(j). In view of the above and Corollary 2.19, we deduce

Ψ([ind(λ)])\displaystyle\Psi([\mathrm{ind}(\lambda)]) =γ𝒫Qγ[ind(λ)],[Kγ]EP\displaystyle=\sum_{\gamma\in{\mathcal{P}}}Q^{\vee}_{\gamma}\cdot\left<[\mathrm{ind}(\lambda)],[K_{\gamma}]\right>_{EP}
=γ𝒫,k(1)kQγgdimextAn0k(ind(λ),Kγ)\displaystyle=\sum_{\gamma\in{\mathcal{P}},k\in\mathbb{Z}}(-1)^{k}Q^{\vee}_{\gamma}\cdot\mathrm{gdim}\,\mathrm{ext}_{A_{n_{0}}}^{k}(\mathrm{ind}(\lambda),K_{\gamma}^{*})^{*}
=γSQγgdimhomAn0(ind(λ),Kγ)\displaystyle=\sum_{\gamma\in S}Q^{\vee}_{\gamma}\cdot\mathrm{gdim}\,\mathrm{hom}_{A_{n_{0}}}(\mathrm{ind}(\lambda),K_{\gamma}^{*})^{*}
=γSbγ1QγgdimhomAn0(ind(λ),Kγ)Λq.\displaystyle=\sum_{\gamma\in S}b_{\gamma}^{-1}\cdot Q_{\gamma}\cdot\mathrm{gdim}\,\mathrm{hom}_{A_{n_{0}}}(\mathrm{ind}(\lambda),K_{\gamma}^{*})^{*}\in\Lambda_{q}. (2.22)

This expansion exhibits positivity (as a formal power series in ((q))\mathbb{Q}(\!(q)\!)).

Claim B.

For each 1j(λ)1\leq j\leq\ell(\lambda), the module N(j)/N(j+1)N(j)/N(j+1) is the direct sum of grading shifts of K~λ(j)\widetilde{K}^{\prime}_{\lambda^{(j)}}.

Proof.

We assume that the assertion holds for all the larger jj (or j=(λ)+1j=\ell(\lambda)+1), and λ(j)λ(j+1)\lambda^{(j)}\neq\lambda^{(j+1)} (and hence λ(j)>λ(j+1)\lambda^{(j)}>\lambda^{(j+1)}). We apply Claim A, and compare Lemma 1.3 and Theorem 2.9 with (2.22) to find

[ind(λ)N(j+1):Lλ(j)]q=[N(j)N(j+1):Lλ(j)]q=bλ(j)1gdimhomAn0(ind(λ),Kλ(j)).\left[\frac{\mathrm{ind}(\lambda)}{N(j+1)}:L_{\lambda^{(j)}}\right]_{q}=\left[\frac{N(j)}{N(j+1)}:L_{\lambda^{(j)}}\right]_{q}=b_{\lambda^{(j)}}^{-1}\cdot\mathrm{gdim}\,\mathrm{hom}_{A_{n_{0}}}(\mathrm{ind}(\lambda),K_{\lambda^{(j)}}^{*})^{*}.

Since Ψ([ind(λ)/N(j+1)])\Psi([\mathrm{ind}(\lambda)/N(j+1)]) must be the sum of QγQ^{\vee}_{\gamma} for γ=λ(k)\gamma=\lambda^{(k)} (kj+1k\leq j+1) by the induction hypothesis and the above formulae, Theorem 2.9 implies

[N(j)/N(j+1):Lμ]q=0if μλ(j).[N(j)/N(j+1):L_{\mu}]_{q}=0\hskip 14.22636pt\text{if}\hskip 14.22636pt\mu\not\geq\lambda^{(j)}.

It follows that N(j)/N(j+1)N(j)/N(j+1) admits a surjection from direct sum of K~λ(j)\widetilde{K}_{\lambda^{(j)}} with its multiplicity gdimhomAn0(ind(λ),Kλ(j))\mathrm{gdim}\,\mathrm{hom}_{A_{n_{0}}}(\mathrm{ind}(\lambda),K_{\lambda^{(j)}}^{*})^{*} (as this latter number counts the number of generators of N(j)/N(j+1)N(j)/N(j+1)). Applying Corollary 2.34, we conclude that N(j)/N(j+1)N(j)/N(j+1) is the direct sum of grading shifts of K~λ(j)\widetilde{K}^{\prime}_{\lambda^{(j)}}. These proceed the induction, and we conclude the result. ∎

Claim C.

Let us enumerate as S={γ1<γ2<<γs}S=\{\gamma_{1}<\gamma_{2}<\cdots<\gamma_{s}\}. We have a finite increasing filtration

{0}=G0G1G2Gs=ind(λ)\{0\}=G_{0}\subset G_{1}\subset G_{2}\subset\cdots\subset G_{s}=\mathrm{ind}(\lambda)

as An0A_{n_{0}}-modules such that each Gi/Gi1G_{i}/G_{i-1} is isomorphic to the direct sum of grading shifts of K~γi\widetilde{K}^{\prime}_{\gamma_{i}}. In addition, each Gs/Gi1G_{s}/G_{i-1} contains a copy of K~γi\widetilde{K}^{\prime}_{\gamma_{i}} as its An0A_{n_{0}}-module direct summand.

Proof.

The first part is a rearrangement of Claim B.

We have LγiInd𝔖n01𝔖n0LλL_{\gamma_{i}}\subset\mathrm{Ind}^{\mathfrak{S}_{n_{0}}}_{\mathfrak{S}_{n_{0}-1}}L_{\lambda} as 𝔖n0\mathfrak{S}_{n_{0}}-modules. If we have [Gs/Gi1:Lμ]q0[G_{s}/G_{i-1}:L_{\mu}]_{q}\neq 0, then Claim B implies [K~γj:Lμ]q0[\widetilde{K}^{\prime}_{\gamma_{j}}:L_{\mu}]_{q}\neq 0 for some ijsi\leq j\leq s. By Lemma 1.3, we conclude that μγi\mu\geq\gamma_{i}. Since Ind𝔖n01𝔖n0Lλ=ind(λ)0\mathrm{Ind}^{\mathfrak{S}_{n_{0}}}_{\mathfrak{S}_{n_{0}-1}}L_{\lambda}=\mathrm{ind}(\lambda)_{0}, we find a degree zero copy of LγiL_{\gamma_{i}} in 𝗁𝖽ind(λ)\mathsf{hd}\,\mathrm{ind}(\lambda). By Propositions 2.31 and 2.12, it must lift to a direct summand K~γiGs/Gi1\widetilde{K}^{\prime}_{\gamma_{i}}\subset G_{s}/G_{i-1}. This implies the second assertion. ∎

Claim D.

For each γS\gamma\in S, we have

extAnk(K~γ,Kμ)={(k=0,γ=μ){0}(else).\mathrm{ext}^{k}_{A_{n}}(\widetilde{K}^{\prime}_{\gamma},K_{\mu}^{*})=\begin{cases}{\mathbb{C}}&(k=0,\gamma=\mu)\\ \{0\}&(\text{else})\end{cases}. (2.23)
Proof.

We prove (2.23) and

extAn>0(Gs/Gj,Kμ)=0\mathrm{ext}^{>0}_{A_{n}}(G_{s}/G_{j},K_{\mu}^{*})=0 (2.24)

for 0js0\leq j\leq s by induction. The j=0j=0 case of (2.24) follows by (2.18). The j=i1j=i-1 case of (2.24) implies (2.23) for γ=γi\gamma=\gamma_{i} and k>0k>0 as Gs/Gi1G_{s}/G_{i-1} contains K~γi\widetilde{K}^{\prime}_{\gamma_{i}} as its direct summand by Claim C. We have

homAn0(K~γ,Kμ)={(γ=μ)0(γμ)\mathrm{hom}_{A_{n_{0}}}(\widetilde{K}^{\prime}_{\gamma},K_{\mu}^{*})=\begin{cases}{\mathbb{C}}&(\gamma=\mu)\\ 0&(\gamma\neq\mu)\end{cases} (2.25)

by Lemma 1.3, 𝗁𝖽K~γ=Lγ\mathsf{hd}\,\widetilde{K}_{\gamma}=L_{\gamma}, and 𝗌𝗈𝖼Kμ=Lμ\mathsf{soc}\,K_{\mu}^{*}=L_{\mu}. By counting the multiplicities of LγiL_{\gamma_{i}}, we deduce

homAn0(Gs/Gj1,Kγj)homAn0((K~γj),Kγj)\mathrm{hom}_{A_{n_{0}}}(G_{s}/G_{j-1},K_{\gamma_{j}}^{*})\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\mathrm{hom}_{A_{n_{0}}}((\widetilde{K}^{\prime}_{\gamma_{j}})^{\oplus\star},K_{\gamma_{j}}^{*}) (2.26)

for 0js0\leq j\leq s from Claim C.

Now a part of the long exact sequence

0\displaystyle 0\rightarrow homAn0(Gs/Gi,Kμ)homAn0(Gs/Gi1,Kμ)homAn0((K~γi),Kμ)\displaystyle\,\mathrm{hom}_{A_{n_{0}}}(G_{s}/G_{i},K_{\mu}^{*})\rightarrow\mathrm{hom}_{A_{n_{0}}}(G_{s}/G_{i-1},K_{\mu}^{*})\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\mathrm{hom}_{A_{n_{0}}}((\widetilde{K}^{\prime}_{\gamma_{i}})^{\oplus\star},K_{\mu}^{*})
\displaystyle\rightarrow extAn01(Gs/Gi,Kμ)extAn01(Gs/Gi1,Kμ)=0\displaystyle\,\mathrm{ext}^{1}_{A_{n_{0}}}(G_{s}/G_{i},K_{\mu}^{*})\rightarrow\mathrm{ext}^{1}_{A_{n_{0}}}(G_{s}/G_{i-1},K_{\mu}^{*})=0

associated to the short exact sequence

0(K~γi)Gs/Gi1Gs/Gi0,0\rightarrow(\widetilde{K}^{\prime}_{\gamma_{i}})^{\oplus\star}\rightarrow G_{s}/G_{i-1}\rightarrow G_{s}/G_{i}\rightarrow 0,

as well as (2.25) and (2.26), yields (2.23) for γ=γi\gamma=\gamma_{i} and (2.24) for j=ij=i from (2.24) for j=i1j=i-1 inductively on ii. ∎

We return to the proof of Theorem 2.3. All elements of 𝒫n0{\mathcal{P}}_{n_{0}} appear as λ(j)\lambda^{(j)} for suitable λ𝒫n01\lambda\in{\mathcal{P}}_{n_{0}-1} and 1j((λ)+1)1\leq j\leq(\ell(\lambda)+1). By rearranging λ\lambda if necessary, we conclude (2.23) for every γ𝒫n0\gamma\in{\mathcal{P}}_{n_{0}}. A repeated use of short exact sequences decomposes {Kμ}γμ\{K_{\mu}\}_{\gamma\leq\mu} into {Lμ}γμ\{L_{\mu}\}_{\gamma\leq\mu} by starting from K(n)=L(n)K_{(n)}=L_{(n)} (see Lemma 1.3). Substituting these to the second factor of (2.23), we deduce

extAn>0(K~γ,Lμ)0implies μ<γ.\mathrm{ext}^{>0}_{A_{n}}(\widetilde{K}^{\prime}_{\gamma},L_{\mu})\neq 0\hskip 8.53581pt\text{implies}\hskip 8.53581pt\mu<\gamma.

This implies K~γ=K~γ\widetilde{K}^{\prime}_{\gamma}=\widetilde{K}_{\gamma} for all γ𝒫n0\gamma\in{\mathcal{P}}_{n_{0}}. Therefore, Proposition 2.25 and Proposition 2.31 imply Theorem 2.3 1) and 2) for n=n0n=n_{0}, and (2.23) is Theorem 2.3 3) for n=n0n=n_{0}.

In view of the above arguments, we find that each ind(λ)\mathrm{ind}(\lambda) (λ𝒫n01)(\lambda\in{\mathcal{P}}_{n_{0}-1}) admits a Δ\Delta-filtration. Since ind1,\mathrm{ind}_{1,\star} preserves projectivity, we deduce that An0A_{n_{0}} admits a filtration by ind(λ)\mathrm{ind}(\lambda) (λ𝒫n01)(\lambda\in{\mathcal{P}}_{n_{0}-1}) by the induction hypothesis. Therefore, An0A_{n_{0}} admits a Δ\Delta-filtration. Since each K~λ\widetilde{K}_{\lambda} is generated by its simple head, applying an idempotent does not separate them out non-trivially. Therefore, we conclude that each projective module of An0A_{n_{0}} also admits a Δ\Delta-filtration. Given this and Theorem 2.3 2) and 3), the latter assertion of Theorem 2.3 4) is standard (see e.g. [11, Corollary 3.12]). This is Theorem 2.3 4) for n=n0n=n_{0}.

This completes the proof of Theorem 2.3.

2.5 Applications of Theorem 2.3

Note that AnA_{n} is a Noetherian ring as a finitely generated AnA_{n}-module is also finitely generated by [X1,,Xn]{\mathbb{C}}[X_{1},\ldots,X_{n}]. The global dimension of AnA_{n} is finite (Theorem 1.7). We have gdimAn[[q]]\mathrm{gdim}\,A_{n}\in\mathbb{Z}[\![q]\!] by inspection.

We introduce a total order \prec on 𝒫n{\mathcal{P}}_{n} that refines \leq and set 𝐞λ:=λμ𝒫neμ\mathbf{e}_{\lambda}:=\sum_{\lambda\succ\mu\in{\mathcal{P}}_{n}}e_{\mu} for each λ𝒫n\lambda\in{\mathcal{P}}_{n}. The two sided ideals An𝐞λAnAnA_{n}\mathbf{e}_{\lambda}A_{n}\subset A_{n} satisfies An𝐞λAnAn𝐞μAnA_{n}\mathbf{e}_{\lambda}A_{n}\subset A_{n}\mathbf{e}_{\mu}A_{n} if μλ\mu\succ\lambda. By Lemma 1.3, we deduce that

(An𝐞λAn)AnPλK~λ(A_{n}\mathbf{e}_{\lambda}A_{n})\otimes_{A_{n}}P_{\lambda}\longrightarrow\widetilde{K}_{\lambda}

is a surjection. By Proposition 2.12 and Theorem 2.3 2), we further deduce

(An𝐞λAn)AnPλK~λ.(A_{n}\mathbf{e}_{\lambda}A_{n})\otimes_{A_{n}}P_{\lambda}\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\widetilde{K}_{\lambda}.

Theorem 2.3 1) implies that endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}) is a graded polynomial ring for each λ𝒫n\lambda\in{\mathcal{P}}_{n}. In conjunction with Theorem 2.3 2), we find that

endAn(Pμ,K~λ)\mathrm{end}_{A_{n}}(P_{\mu},\widetilde{K}_{\lambda})

is a free module over endAn(K~λ)\mathrm{end}_{A_{n}}(\widetilde{K}_{\lambda}) for each λ,μ𝒫n\lambda,\mu\in{\mathcal{P}}_{n}. In particular, that the graded algebra AnA_{n} is an affine quasi-hereditary in the sense of [14, Introduction] with Δλ=K~λ\Delta_{\lambda}=\widetilde{K}_{\lambda} and ¯λ=Kλ\overline{\nabla}_{\lambda}=K_{\lambda}^{*} (λ𝒫n)(\lambda\in{\mathcal{P}}_{n}).

Theorem 2.35 ([14] Theorem 7.21 and Lemma 7.22).

A module MA𝗀𝗆𝗈𝖽M\in A\mathchar 45\relax\mathsf{gmod} admits a Δ\Delta-filtration if and only if

extAn1(M,Kλ)=0λ𝒫n.\mathrm{ext}^{1}_{A_{n}}(M,K_{\lambda}^{*})=0\hskip 14.22636pt\lambda\in{\mathcal{P}}_{n}.

A module MA𝖿𝗆𝗈𝖽M\in A\mathchar 45\relax\mathsf{fmod} admits a Δ¯\overline{\Delta}-filtration if and only if

extAn1(K~λ,M)=0λ𝒫n.\mathrm{ext}^{1}_{A_{n}}(\widetilde{K}_{\lambda},M^{*})=0\hskip 14.22636pt\lambda\in{\mathcal{P}}_{n}.
Corollary 2.36 ([14] §7, particularly Lemma 7.5).

Let MA𝗀𝗆𝗈𝖽M\in A\mathchar 45\relax\mathsf{gmod}. If MM admits a Δ\Delta-filtration, then the multiplicity space of K~λ\widetilde{K}_{\lambda} in MM is given by

homAn(M,Kλ).\mathrm{hom}_{A_{n}}(M,K_{\lambda})^{*}.

If the module MM admits a Δ¯\overline{\Delta}-filtration, then the multiplicity space of of KλK_{\lambda} in MM is given by

homAn(K~λ,M).\mathrm{hom}_{A_{n}}(\widetilde{K}_{\lambda},M^{*})^{*}.
Theorem 2.37.

Fix n0n\geq 0, and 0rn0\leq r\leq n. Let λ𝒫n,μ𝒫r,ν𝒫nr\lambda\in{\mathcal{P}}_{n},\mu\in{\mathcal{P}}_{r},\nu\in{\mathcal{P}}_{n-r}. We have the following:

  1. 1.

    (Garsia-Procesi [6]) The module resr,nrKλ\mathrm{res}_{r,n-r}\,K_{\lambda} admits a Δ¯\overline{\Delta}-filtration;

  2. 2.

    The module indr,nr(PμK~ν)\mathrm{ind}_{r,n-r}\,(P_{\mu}\boxtimes\widetilde{K}_{\nu}) admits a Δ\Delta-filtration.

Remark 2.38.

One cannot swap the roles of {K~λ}λ\{\widetilde{K}_{\lambda}\}_{\lambda} and {Kλ}λ\{K_{\lambda}\}_{\lambda} in Theorem 2.37. In fact, the polynomiality claim in Corollary 2.39 2) is already nontrivial (without a prior knowledge of characters).

Proof of Theorem 2.37.

We prove the first assertion. By the second part of Theorem 2.35, it suffices to check the ext1\mathrm{ext}^{1}-vanishing with respect to LμK~νL_{\mu}\boxtimes\widetilde{K}_{\nu} (μ𝒫r,ν𝒫nr\mu\in{\mathcal{P}}_{r},\nu\in{\mathcal{P}}_{n-r}) as a module over 𝔖rAnr{\mathbb{C}}\mathfrak{S}_{r}\boxtimes A_{n-r} (equivalently, we can check the ext1\mathrm{ext}^{1}-vanishing with respect to PμK~νP_{\mu}\boxtimes\widetilde{K}_{\nu} as a module of Ar,nrA_{r,n-r}; see below). In particular, we do not need to mind the first factor as the 𝔖r\mathfrak{S}_{r}-action is granted by construction. Therefore, the first assertion is just a rr-times repeated application of Theorem 2.7.

We prove the second assertion. For each λ𝒫r,μ𝒫nr\lambda\in{\mathcal{P}}_{r},\mu\in{\mathcal{P}}_{n-r} and ν𝒫n\nu\in{\mathcal{P}}_{n}, we have

extAn(indr,nr(PλK~μ),Kν)extAr,nr(PλK~μ,Kν)\mathrm{ext}^{\bullet}_{A_{n}}(\mathrm{ind}_{r,n-r}(P_{\lambda}\boxtimes\widetilde{K}_{\mu}),K_{\nu}^{*})\cong\mathrm{ext}^{\bullet}_{A_{r,n-r}}(P_{\lambda}\boxtimes\widetilde{K}_{\mu},K_{\nu}^{*}) (2.27)

by Theorem 1.5. Applying Theorem 2.7 to KνK_{\nu}^{*} as many as rr-times, we find that the restriction of KνK_{\nu} to AnrA_{n-r} admits a filtration whose associated graded is the direct sum of grading shifts of {Kγ}γ𝒫nr\{K_{\gamma}\}_{\gamma\in{\mathcal{P}}_{n-r}}. Since PλP_{\lambda} is free over a polynomial ring of rr-variables, we have

extAr,nr(PλK~μ,Kν)ext𝔖rAnr(LλK~μ,Kν).\mathrm{ext}^{\bullet}_{A_{r,n-r}}(P_{\lambda}\boxtimes\widetilde{K}_{\mu},K_{\nu}^{*})\cong\mathrm{ext}^{\bullet}_{{\mathbb{C}}\mathfrak{S}_{r}\boxtimes A_{n-r}}(L_{\lambda}\boxtimes\widetilde{K}_{\mu},K_{\nu}^{*}).

Thus, we derive a natural isomorphism

ext𝔖rAnr1(LλK~μ,Kν)hom𝔖r(Lλ,extAnr1(K~μ,Kν)).\mathrm{ext}^{1}_{{\mathbb{C}}\mathfrak{S}_{r}\boxtimes A_{n-r}}(L_{\lambda}\boxtimes\widetilde{K}_{\mu},K_{\nu}^{*})\stackrel{{\scriptstyle\cong}}{{\longrightarrow}}\mathrm{hom}_{\mathfrak{S}_{r}}(L_{\lambda},\mathrm{ext}_{A_{n-r}}^{1}(\widetilde{K}_{\mu},K_{\nu}^{*})). (2.28)

By Theorem 2.3 3) and Theorem 2.7, the RHS of (2.28) is zero. By the first part of Theorem 2.35, we conclude the second assertion. ∎

Corollary 2.39.

Let λ,μ𝒫\lambda,\mu\in{\mathcal{P}}. We have the following:

  1. 1.

    We have Δ(Qλ)γ,κ0[q](SγQκ)\Delta\,(Q_{\lambda})\in\sum_{\gamma,\kappa}\mathbb{Z}_{\geq 0}[q]\,(S_{\gamma}\otimes Q_{\kappa});

  2. 2.

    We have sλQμγ0[q]Qνs_{\lambda}\cdot Q^{\vee}_{\mu}\in\sum_{\gamma}\mathbb{Z}_{\geq 0}[q]\,Q^{\vee}_{\nu}. In case λ=(1n)\lambda=(1^{n}), it is the Pieri rule.

Proof.

Apply the twisted Frobenius characteristic to Theorem 2.37 using Lemma 2.14. Here the equality s(1n)=Q(1n)s_{(1^{n})}=Q^{\vee}_{(1^{n})} is in [19, I​I​I (2.8)] and the Pieri rule is in [19, I​I​I (3.2)]. ∎

Corollary 2.40.

The skew Hall-Littlewood QQ-function Qλ/νQ_{\lambda/\nu} expands positively with respect to the big Schur function. In addition, we have a graded A|λ||ν|A_{|\lambda|-|\nu|}-module defined as

homA|ν|(K~ν,Kλ),\mathrm{hom}_{A_{|\nu|}}(\widetilde{K}_{\nu},K_{\lambda}^{*})^{*},

such that its image under Ψ\Psi is Qλ/νQ_{\lambda/\nu}.

Proof.

Let λ𝒫n\lambda\in{\mathcal{P}}_{n}. The Hall-Littlewood QQ-polynomial corresponds to the module KλK_{\lambda} by Theorem 2.9. Therefore, its restriction admits a Δ¯\overline{\Delta}-filtration. In particular, we have

[resr,nrKλ]=μ,νcλμ,ν[LμKν]cλμ,ν0[q].[\mathrm{res}_{r,n-r}\,K_{\lambda}]=\sum_{\mu,\nu}c_{\lambda}^{\mu,\nu}[L_{\mu}\boxtimes K_{\nu}]\hskip 14.22636ptc_{\lambda}^{\mu,\nu}\in\mathbb{Z}_{\geq 0}[q].

Here resr,nr\mathrm{res}_{r,n-r} corresponds to Δ\Delta in Λq\Lambda_{q} by Lemma 2.14. Taking [Mac98, I​I​I (5.2)], (2.3), and (2.5) into account, we conclude that

Qλ/ν=μcλμ,νΨ([Lμ]).Q_{\lambda/\nu}=\sum_{\mu}c_{\lambda}^{\mu,\nu}\Psi([L_{\mu}]).

This is the first assertion by Ψ([Lμ])=Sμ\Psi([L_{\mu}])=S_{\mu}, read off from (2.4). In view of Theorem 2.37 1) and Corollary 2.36, we conclude the second assertion. ∎

Acknowledgement: The author presented an intensive lecture course based on the contents of this paper at Nagoya University on Fall 2020. The author thanks the hospitality of Shintarou Yanagida and Nagoya University. He also thanks Kota Murakami for pointing out some typos. This research was supported in part by JSPS KAKENHI Grant Number JP19H01782.

References

  • [1] Sabin Cautis and Anthony M. Licata “Heisenberg categorification and Hilbert schemes” In Duke Mathematical Journal 161.13, 2012, pp. 2469–2547 DOI: 10.1215/00127094-1812726
  • [2] Vyjayanthi Chari and Bogdan Ion “BGG reciprocity for current algebras” In Compos. Math. 151.7, 2015, pp. 1265–1287 DOI: 10.1112/S0010437X14007908
  • [3] Claude Chevalley “Invariants of finite groups generated by reflections” In Amer. J. Math. 77, 1955, pp. 778–782
  • [4] Corrado De Concini and Claudio Procesi “Symmetric functions, conjugacy classes and the flag variety” In Invent. Math. 64, 1981, pp. 203–219
  • [5] Evgeny Feigin, Anton Khoroshkin and Ievgen Makedonskyi “Peter-Weyl, Howe and Schur-Weyl theorems for current groups”, arXiv:1906.03290
  • [6] A. M. Garsia and C. Procesi “On certain graded SnS_{n}-modules and the qq-Kostka polynomials” In Adv. Math. 94.1, 1992, pp. 82–138
  • [7] Alexander Grothendieck “Sur quelques points d’algèbre homologique” In Tohoku Math. J. (2) 9, 1957, pp. 119–221
  • [8] Mark Haiman “Combinatorics, symmetric functions, and Hilbert schemes” In Current developments in mathematics, 2002 Int. Press, Somerville, MA, 2003, pp. 39–111
  • [9] Mark Haiman “Hilbert schemes, polygraphs and the Macdonald positivity conjecture” In J. Amer. Math. Soc. 14.941–1006, 2001
  • [10] Syu Kato “A homological study of Green polynomials” In Ann. Sci. Éc. Norm. Supér. (4) 48.5, 2015, pp. 1035–1074
  • [11] Syu Kato “An algebraic study of extension algebras” arXiv:1207.4640 In Amer. J. Math. 139.3, 2017, pp. 567–615
  • [12] Syu Kato “An exotic Deligne-Langlands correspondence for symplectic groups” In Duke Math. J. 148.2, 2009, pp. 305–371
  • [13] David Kazhdan and George Lusztig “Proof of the Deligne-Langlands conjecture for Hecke algebras” In Invent. Math. 87.1, 1987, pp. 153–215
  • [14] Alexander S. Kleshchev “Affine highest weight categories and affine quasihereditary algebras” In Proc. Lond. Math. Soc. (3) 110.4, 2015, pp. 841–882 DOI: 10.1112/plms/pdv004
  • [15] Steffen König and Changchang Xi “Affine cellular algebras” In Adv. Math. 229.1, 2012, pp. 139–182
  • [16] G. Lusztig “Intersection cohomology complexes on a reductive group” In Invent. Math. 75.2, 1984, pp. 205–272
  • [17] George Lusztig “Cuspidal local systems and graded Hecke algebras. I” In Inst. Hautes Études Sci. Publ. Math. 67, 1988, pp. 145–202
  • [18] George Lusztig “Green functions and character sheaves” In Ann. of Math. (2) 131, 1990, pp. 355–408
  • [19] I. G. Macdonald “Symmetric functions and Hall polynomials” With contributions by A. Zelevinsky, Oxford Science Publications, Oxford Mathematical Monographs The Clarendon Press, Oxford University Press, New York, 1995, pp. x+475
  • [20] Hideyuki Matsumura “Commutative ring theory” Translated from the Japanese by M. Reid 8, Cambridge Studies in Advanced Mathematics Cambridge University Press, Cambridge, 1986, pp. xiv+320
  • [21] John C. McConnell and James C. Robson “Noncommutative Noetherian rings.” 30, Graduate Studies in Math. AMS, 2001
  • [22] Ryosuke Shimoji and Shintarou Yanagida “A study of symmetric functions via derived Hall algebra” In Communications in Algebras to appear, 2020
  • [23] Toshiaki Shoji “Green functions associated to complex reflection groups” In J. Algebra 245.2, 2001, pp. 650–694
  • [24] T. A. Springer “Trigonometric sums, Green functions of finite groups and representations of Weyl groups” In Invent. Math. 36, 1976, pp. 173–207
  • [25] T. A. Springer and A. V. Zelevinsky “Characters of GL(n,𝐅q){\rm GL}(n,\,{\bf F}_{q}) and Hopf algebras” In J. London Math. Soc. (2) 30.1, 1984, pp. 27–43
  • [26] Toshiyuki Tanisaki “Defining ideals of the closures of the conjugaty classes and representations of the Weyl groups” In Tohoku Mathematical Journal 34, 1982, pp. 575–585
  • [27] Nanhua Xi “The based ring of two-sided cells of affine Weyl groups of type A~n1\widetilde{A}_{n-1} In Mem. Amer. Math. Soc. 157.749, 2002, pp. xiv+95
  • [28] Andrey V. Zelevinsky “Representations of finite classical groups” 869, Lecture Notes in Mathematics Springer-Verlag, Berlin-New York, 1981