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

Unique Factorization For Tensor Products of Parabolic Verma Modules

K.N. Raghavan The Institute of Mathematical Sciences, A CI of Homi Bhabha National Institute, Chennai 600113, India [email protected] V. Sathish Kumar The Institute of Mathematical Sciences, A CI of Homi Bhabha National Institute, Chennai 600113, India [email protected] R. Venkatesh Department of Mathematics, Indian Institute of Science, Bangalore 560012 [email protected]  and  Sankaran Viswanath The Institute of Mathematical Sciences, A CI of Homi Bhabha National Institute, Chennai 600113, India [email protected]
Abstract.

Let 𝔤\mathfrak{g} be a symmetrizable Kac-Moody Lie algebra with Cartan subalgebra 𝔥\mathfrak{h}. We prove a unique factorization property for tensor products of parabolic Verma modules. More generally, we prove unique factorization for products of characters of parabolic Verma modules when restricted to certain subalgebras of 𝔥\mathfrak{h}. These include fixed point subalgebras of 𝔥\mathfrak{h} under subgroups of diagram automorphisms of 𝔤\mathfrak{g} and twisted graph automorphisms in the affine case.

Key words and phrases:
Unique factorization, Kac-Moody Lie algebras, Parabolic Verma modules, tensor products
2020 Mathematics Subject Classification:
17B67, 17B10
The first, second and fourth authors acknowledge partial funding from a DAE Apex Project grant to the Institute of Mathematical Sciences, Chennai.

1. Introduction

Investigating whether a given family of elements from a ring has the unique factorization property (UFP) is a well-studied problem in basic ring theory. In this paper, we are interested in studying a similar phenomenon for representations of Kac-Moody algebras. Let 𝔤\mathfrak{g} be a Kac-Moody algebra. Suppose {Mi}i=1r\{M_{i}\}_{i=1}^{r} and {Nj}j=1s\{N_{j}\}_{j=1}^{s} are 𝔤\mathfrak{g}–modules from a suitable category satisfying

i=1rMij=1sNj\bigotimes\limits_{i=1}^{r}M_{i}\cong\bigotimes\limits_{j=1}^{s}N_{j} (1)

as 𝔤\mathfrak{g}–modules, then we have the following natural questions:

  1. (1)

    Are the number of factors on both sides of (1) equal, i.e., is r=sr=s ?

  2. (2)

    If they are equal, can we compare the highest weights of MiM_{i}’s and NjN_{j}’s if they are highest-weight modules?

  3. (3)

    Can we prove the unique factorization property for their characters (when defined and viewed as elements in the character ring)?

  4. (4)

    Further, what more can one say about the individual modules MiM_{i} and NjN_{j}? For example, are they isomorphic up to a permutation of factors ?

In the literature, unique factorization theorems study these questions. For example, C. S. Rajan proved a unique factorization property for tensor products of finite-dimensional simple modules of a finite-dimensional simple Lie algebra 𝔤\mathfrak{g} in [7]. Later in [9] and [8], the authors extended Rajan’s result suitably beyond the realm of finite dimensional simple Lie algebras.

All these papers [7, 9, 8] study only the unique factorization property of tensor products of simple modules in some suitable categories. In this paper, we will consider some families of typically reducible modules and study their unique factorization properties. More precisely, we consider the following two families of modules of 𝔤=𝔤(A)\mathfrak{g}=\mathfrak{g}(A) symmetrizable Kac-Moody algebra (where AA is a n×nn\times n symmetrizable generalized Cartan matrix):

  1. (1)

    Parabolic Verma modules of 𝔤\mathfrak{g}: these are highest-weight modules subsuming the class of simple integrable highest-weight modules (but are typically neither simple nor integrable).

  2. (2)

    Restrictions of parabolic Verma modules to suitable subalgebras of 𝔤\mathfrak{g}, for example, to fixed point subalgebras of Dynkin diagram automorphisms.

Let 𝔥\mathfrak{h} be a fixed Cartan subalgebra of 𝔤\mathfrak{g} and {αi:1in}\{\alpha_{i}^{\vee}:1\leq i\leq n\} be co-roots corresponding to the simple roots with respect to 𝔥.\mathfrak{h}. The parabolic Verma modules of 𝔤\mathfrak{g} are indexed by (λ,I)(\lambda,I), where λ𝔥\lambda\in\mathfrak{h}^{*} and I{1,,n}I\subseteq\{1,\ldots,n\} and we denote by M(λ,I)M(\lambda,I) the parabolic Verma module corresponding to the tuple (λ,I).(\lambda,I). We call a subset I{1,,n}I\subseteq\{1,\ldots,n\} connected if when II is thought of as a subset of the nodes of the Dynkin diagram associated with AA, the subgraph induced by II is connected. Here is our main theorem for parabolic Verma modules.

Theorem 1.

Let AA be a n×nn\times n symmetrizable generalized Cartan matrix. Let 𝔤=𝔤(A)\mathfrak{g}=\mathfrak{g}(A) be the Kac-Moody Lie algebra associated with AA and let 𝔥\mathfrak{h} be the Cartan subalgebra. Suppose that

k=1rM(λk,Ik)k=1rM(μk,Jk)\bigotimes\limits_{k=1}^{r}M(\lambda_{k},I_{k})\cong\bigotimes\limits_{k=1}^{r}M(\mu_{k},J_{k}) (2)

where for each 1kn1\leq k\leq n we have

  1. (1)

    λk,μk𝔥\lambda_{k},\mu_{k}\in\mathfrak{h}^{*} and Ik,JkI_{k},J_{k} are connected subsets of {1,,n}\{1,\ldots,n\},

  2. (2)

    λk(αi),μk(αj)\lambda_{k}(\alpha_{i}^{\vee}),\mu_{k}(\alpha_{j}^{\vee}) are positive integers for all iIki\in I_{k} and jJkj\in J_{k}.

Then, k=1rλk=k=1rμk\sum_{k=1}^{r}\lambda_{k}=\sum_{k=1}^{r}\mu_{k} and there exists a permutation σ𝔖r\sigma\in\mathfrak{S}_{r} such that

Ik=JσkI_{k}=J_{\sigma k} and λk(αi)=μσk(αi)\lambda_{k}(\alpha_{i}^{\vee})=\mu_{\sigma k}(\alpha_{i}^{\vee}) for all iIki\in I_{k}.

Since M(λ,I)M(\lambda,I) are simple integrable highest weight modules when λ\lambda is dominant and I={1,,n}I=\{1,\ldots,n\}, this may be viewed as an extension of the results of [7, 9].

Next, let Γ\Gamma be a subgroup of Dynkin diagram automorphisms of 𝔤\mathfrak{g}. Consider the fixed point subalgebra 𝔤Γ\mathfrak{g}^{\Gamma} of 𝔤\mathfrak{g} with respect to Γ\Gamma. Then we can restrict the modules in Equation (2) to 𝔤Γ\mathfrak{g}^{\Gamma} and ask whether the unique factorization property holds for these 𝔤Γ\mathfrak{g}^{\Gamma} modules. Under some natural conditions, we answer this question affirmatively. More precisely, we prove the following:

Theorem 2.

Let AA be a n×nn\times n symmetrizable generalized Cartan matrix of finite, affine, or hyperbolic type. Let 𝔤=𝔤(A)\mathfrak{g}=\mathfrak{g}(A) be the associated Kac-Moody Lie algebra, with Cartan subalgebra 𝔥\mathfrak{h}. Let Γ\Gamma be a group of diagram automorphisms of 𝔤\mathfrak{g} and let 𝔤Γ\mathfrak{g}^{\Gamma} be the fixed point subalgebra. Suppose that

k=1rRes𝔤ΓM(λk,Ik)=k=1rRes𝔤ΓM(μk,Jk)\bigotimes_{k=1}^{r}Res_{\mathfrak{g}^{\Gamma}}M(\lambda_{k},I_{k})=\bigotimes_{k=1}^{r}Res_{\mathfrak{g}^{\Gamma}}M(\mu_{k},J_{k})

where for each 1kr1\leq k\leq r, we have

  1. (1)

    λk,μk(𝔥)Γ={ν𝔥:ν(ω(h))=ν(h)for allωΓ,h𝔥}\lambda_{k},\mu_{k}\in(\mathfrak{h}^{*})^{\Gamma}=\{\nu\in\mathfrak{h}^{*}:\nu(\omega(h))=\nu(h)\;\text{for all}\;\omega\in\Gamma,\,h\in\mathfrak{h}\} and Ik,JkI_{k},J_{k} are connected subsets of {1,,n}\{1,\ldots,n\},

  2. (2)

    for each iIki\in I_{k} and jJkj\in J_{k} we have λk(αi)\lambda_{k}(\alpha_{i}^{\vee}), μk(αj)\mu_{k}(\alpha_{j}^{\vee}) are positive integers, and

  3. (3)

    IkI_{k} and JkJ_{k} are Γ\Gamma-stable, i.e., are unions of Γ\Gamma-orbits.

Then, k=1rλk=k=1rμk\sum_{k=1}^{r}\lambda_{k}=\sum_{k=1}^{r}\mu_{k} and there exists a permutation σ𝔖r\sigma\in\mathfrak{S}_{r} such that

Ik=JσkI_{k}=J_{\sigma k} and λk(αi)=μσk(αi)\lambda_{k}(\alpha_{i}^{\vee})=\mu_{\sigma k}(\alpha_{i}^{\vee}) for all iIki\in I_{k}.

We in fact prove stronger versions of Theorems   1, 2, see §4 and §5\S 4\text{ and }\S 5 for more precise statements. It is to be noted that in [6] a similar theorem is proved in the setting of simple modules for finite-dimensional simple Lie algebras with a completely different set of hypotheses. It is easy to see that the converse of Theorems 1, 2 hold at the level of characters. Further, if we assume that complete reducibility holds for the tensor products, then we can also prove the converse of the Theorems 1, 2. For example, the converse is true (see [5, Page No. 180, Corollary 10.7]) when

  1. (1)

    all λk\lambda_{k}’s and μk\mu_{k}’s are integral dominant weights and Ik=Jk={1,,n}I_{k}=J_{k}=\{1,\ldots,n\} (for all 1kn1\leq k\leq n) in Theorem 1 and

  2. (2)

    the fixed point subalgebra 𝔤Γ\mathfrak{g}^{\Gamma} is again Kac-Moody type and all λk\lambda_{k}’s and μk\mu_{k}’s are integral dominant weights and Ik=Jk={1,,n}I_{k}=J_{k}=\{1,\ldots,n\} (for all 1kn1\leq k\leq n) in Theorem 2.

The paper is organized as follows: In Section 2, we set up the notation and preliminaries. In Section 3, we prove some key technical results that will be needed to prove our main theorems. In Section 4, we prove our main theorem for parabolic Verma modules. In Section 5, we consider and prove unique factorization properties for the characters of restricted parabolic Verma modules. We apply this in Section 6 and prove unique factorization properties for parabolic Verma modules when they are restricted to 𝔤Γ\mathfrak{g}^{\Gamma}, where 𝔤\mathfrak{g} is general and Γ\Gamma is a subgroup of Dynkin diagram automorphisms or 𝔤\mathfrak{g} is of affine type and Γ=τ\Gamma=\langle\tau\rangle for some τ\tau twisted graph automorphism.

2. Preliminaries

All vector spaces are assumed to be defined over complex numbers \mathbb{C} throughout the article. For a Lie algebra 𝔤\mathfrak{g}, we denote by U(𝔤)U(\mathfrak{g}) the universal enveloping algebra of 𝔤.\mathfrak{g}. For a vector space VV over \mathbb{C}, we denote by VV^{*} its dual space.

2.1. Structure Theory of Symmetrizable Kac-Moody algebras

In this subsection, we fix some notation and review the structure theory of Kac-Moody algebras, closely following [5]. Let nn be a positive integer and A=(aij)n×nA=(a_{ij})_{n\times n} a generalized Cartan matrix (GCM). That is,

  1. (1)

    aii=2a_{ii}=2, for all 1in1\leq i\leq n,

  2. (2)

    aija_{ij} is a non-positive integer for all 1ijn1\leq i\neq j\leq n, and

  3. (3)

    aij=0a_{ij}=0 if and only if aji=0a_{ji}=0 for all 1i,jn1\leq i,j\leq n.

We say AA is symmetrizable if there exists a diagonal matrix D=diag(d1,,dn)D=\mathrm{diag}(d_{1},\ldots,d_{n}), with did_{i}’s being positive real numbers, such that DADA is symmetric. Let us denote S={1,,n}S=\{1,\ldots,n\}.

Let AA be a symmetrizable GCM and let 𝔤:=𝔤(A)\mathfrak{g}:=\mathfrak{g}(A) be the Kac-Moody Lie algebra associated with AA and 𝔥\mathfrak{h} be a fixed Cartan subalgebra of 𝔤\mathfrak{g}. The Cartan subalgebra 𝔥\mathfrak{h} acts semisimply on 𝔤\mathfrak{g} via the adjoint action. Denoting by Δ\Delta the set of roots of (𝔤,𝔥)(\mathfrak{g},\mathfrak{h}), the corresponding root space decomposition is

𝔤=𝔥αΔ𝔤α,\mathfrak{g}=\mathfrak{h}\oplus\bigoplus_{\alpha\in\Delta}\mathfrak{g}_{\alpha},

where 𝔤α:={x𝔤:[h,x]=α(h)xh𝔥}\mathfrak{g}_{\alpha}:=\{x\in\mathfrak{g}:[h,x]=\alpha(h)x\;\forall h\in\mathfrak{h}\} for αΔ\alpha\in\Delta.

Let the simple system (i.e., simple roots) of Δ\Delta coming from the realization of AA be Π:={α1,,αn}\Pi:=\{\alpha_{1},\cdots,\alpha_{n}\}. We denote by Δ+\Delta_{+} the set of positive roots of Δ\Delta with respect to Π\Pi. The root lattice QQ is the set of all integer linear combinations of elements of Π\Pi. Any αQ\alpha\in Q can be written uniquely as i=1ntiαi\sum_{i=1}^{n}t_{i}\alpha_{i} where tit_{i} are integers. The set of all αQ\alpha\in Q for which all the tit_{i} are non-negative integers is denoted Q+Q^{+}. The support of αQ\alpha\in Q denoted by supp(α)\operatorname{supp}(\alpha) is the set of all kSk\in S for which tk0t_{k}\neq 0. For αΠ\alpha\in\Pi, let α𝔥\alpha^{\vee}\in\mathfrak{h} denote the coroot corresponding to α.\alpha. Let

𝔥{ei,fi:iS}\mathfrak{h}\cup\{e_{i},f_{i}:i\in S\}

be the Chevalley generators of 𝔤\mathfrak{g}. Note that the derived subalgebra [𝔤,𝔤][\mathfrak{g},\mathfrak{g}] is generated by {ei,fi:iS}\{e_{i},f_{i}:i\in S\} and [𝔤,𝔤]𝔥=span{α1,,αn}[\mathfrak{g},\mathfrak{g}]\cap\mathfrak{h}=\text{span}\{\alpha_{1}^{\vee},\ldots,\alpha_{n}^{\vee}\}. The Weyl group of 𝔤\mathfrak{g} is the subgroup of GL(𝔥)GL(\mathfrak{h}^{*}) generated by the reflections {si:iS}\{s_{i}:i\in S\}, where si:𝔥𝔥s_{i}:\mathfrak{h}^{*}\to\mathfrak{h}^{*} is defined by

si(ν)=νν(αi)αiν𝔥s_{i}(\nu)=\nu-\nu(\alpha_{i}^{\vee})\alpha_{i}\quad\quad\forall\nu\in\mathfrak{h}^{*}

The parabolic subgroup WIW_{I} corresponding to ISI\subseteq S is the subgroup of WW generated by {si:iI}\{s_{i}:i\in I\}. It is a fact that WIW_{I} is a Coxeter group with Coxeter generators {si:iI}\{s_{i}:i\in I\}. Given wWIw\in W_{I}, the length of ww is (w):=min{k:w=si1sik}\ell(w):=\mathrm{min}\{k:w=s_{i_{1}}\cdots s_{i_{k}}\}; an expression si1siks_{i_{1}}\cdots s_{i_{k}} for ww is said to be reduced if k=(w)k=\ell(w). Fix a reduced expression si1siks_{i_{1}}\cdots s_{i_{k}} of wWIw\in W_{I}. Define the support of ww by

I(w):={si1,,sik}.I(w):=\{s_{i_{1}},\ldots,s_{i_{k}}\}.

It is a well-known fact (Tits theorem) that I(w)I(w) is independent of the choice of the chosen reduced expression.

A subset ISI\subseteq S is said to be connected if the submatrix of the GCM indexed by II is indecomposable (or equivalently, the subgraph of the Dynkin graph of 𝔤\mathfrak{g} induced by II is connected). Note that any subset II of SS can be written as a finite disjoint union of connected subsets, called connected components, of II and this decomposition is unique up to a permutation of the connected components.

2.2. Parabolic Verma Modules

For λ𝔥\lambda\in\mathfrak{h}^{*}, we denote by M(λ)M(\lambda) the Verma module associated to λ\lambda. The integrability of λ\lambda is defined by

Jλ:={iS|λ(αi) is a non-negative integer}J_{\lambda}:=\{i\in S\;|\;\lambda(\alpha_{i}^{\vee})\text{ is a non-negative integer}\}

For λ𝔥\lambda\in\mathfrak{h}^{*} and IJλ,I\subseteq J_{\lambda}, the Parabolic Verma Module corresponding to (λ,I)(\lambda,I) is defined by

M(λ,I):=M(λ)iIU(𝔤)fαiλ(αi)+1mλM(\lambda,I):=\frac{M(\lambda)}{\sum_{i\in I}U(\mathfrak{g})f_{\alpha_{i}}^{\lambda(\alpha_{i}^{\vee})+1}m_{\lambda}}

where mλm_{\lambda} is the cyclic generator (or highest weight vector) of M(λ)M(\lambda). The parabolic Verma modules are modules in category 𝒪\mathcal{O} (see [2, 4] for more details). When λ\lambda is dominant and integral (i.e., Jλ=SJ_{\lambda}=S) we see that M(λ,S)=V(λ)M(\lambda,S)=V(\lambda), the unique simple 𝔤\mathfrak{g} module with highest weight λ\lambda. When I=I=\emptyset we have M(λ,I)=M(λ)M(\lambda,I)=M(\lambda) for all λ𝔥\lambda\in\mathfrak{h}^{*}. So, the parabolic Verma modules interpolate between the Verma modules and simple modules in category 𝒪\mathcal{O}.

2.3. Characters and their restrictions

Let 𝔤¯\bar{\mathfrak{g}} be a Lie subalgebra of 𝔤\mathfrak{g} and 𝔥¯:=𝔤¯𝔥\bar{\mathfrak{h}}:=\bar{\mathfrak{g}}\cap\mathfrak{h}. Suppose VV is a 𝔥\mathfrak{h}–weight module and

V=ν𝔥VνV=\bigoplus\limits_{\nu\in\mathfrak{h}^{*}}V_{\nu}

is the 𝔥\mathfrak{h}–weight space decomposition of VV. Note that by definition, we have dimVν<\mathrm{dim}V_{\nu}<\infty for all ν𝔥.\nu\in\mathfrak{h}^{*}. The 𝔥\mathfrak{h}–character of VV is defined by

ch𝔥(V)=ν𝔥dim(Vν)eν.\mathrm{ch}_{\mathfrak{h}}(V)=\sum_{\nu\in\mathfrak{h}^{*}}\mathrm{dim}(V_{\nu})e^{\nu}.

Let p:𝔥𝔥¯p:\mathfrak{h}^{*}\to\bar{\mathfrak{h}}^{*} be the restriction map. The restriction of the character of VV to 𝔥¯\bar{\mathfrak{h}} is defined as

ch𝔥¯(V)=ν(𝔥¯)({ν𝔥:p(ν)=ν}dimVν)eν.\mathrm{ch}_{\bar{\mathfrak{h}}}(V)=\sum_{\nu^{\prime}\in(\bar{\mathfrak{h}})^{*}}\left(\sum_{\{\nu\in\mathfrak{h}^{*}\,:\,p(\nu)=\nu^{\prime}\}}\mathrm{dim}V_{\nu}\right)e^{\nu^{\prime}}.

Note that the inner sum {ν𝔥:p(ν)=ν}dimVν\sum_{\{\nu\in\mathfrak{h}^{*}\,:\,p(\nu)=\nu^{\prime}\}}\mathrm{dim}V_{\nu} need not be finite always. We use this definition whenever it makes sense, i.e., whenever this inner sum is finite. This inner sum will be finite and the restricted character ch𝔥¯(V)\mathrm{ch}_{\bar{\mathfrak{h}}}(V) will be well-defined in all our examples.

2.4. The Weyl-Kac character formula

For λP+\lambda\in P^{+}, the Weyl-Kac character formula gives

ch𝔥V(λ)=wW(1)(w)ew(λ+ρ)ραΔ+(1eα)dim𝔤α\mathrm{ch}_{\mathfrak{h}}\,V(\lambda)=\frac{\sum\limits_{w\in W}(-1)^{\ell(w)}e^{w(\lambda+\rho)-\rho}}{\prod_{\alpha\in\Delta^{+}}(1-e^{-\alpha})^{\mathrm{dim}\,\mathfrak{g}_{\alpha}}}

where ρ𝔥\rho\in\mathfrak{h}^{*} is some fixed functional that satisfies ρ(αi)=1\rho(\alpha_{i}^{\vee})=1. A similar formula is known for parabolic Verma modules. The following proposition can be found in [2, Proposition 7.10]. Let λ𝔥\lambda\in\mathfrak{h}^{*} and IJλI\subseteq J_{\lambda}. Then

ch𝔥M(λ,I)=wWI(1)(w)ew(λ+ρ)ραΔ+(1eα)dim𝔤α.\mathrm{ch}_{\mathfrak{h}}\ M(\lambda,I)=\frac{\sum_{w\in W_{I}}(-1)^{\ell(w)}e^{w(\lambda+\rho)-\rho}}{\prod_{\alpha\in\Delta^{+}}(1-e^{-\alpha})^{\mathrm{dim}\,\mathfrak{g}_{\alpha}}}. (3)

Let λ𝔥\lambda\in\mathfrak{h}^{*} and IJλI\subseteq J_{\lambda}. We define the normalised Weyl numerator corresponding to the tuple (λ,I)(\lambda,I) by

U(λ,I):=eλwWI(1)(w)ew(λ+ρ)ρU(\lambda,I):=\;e^{-\lambda}\sum_{w\in W_{I}}(-1)^{\ell(w)}e^{w(\lambda+\rho)-\rho}

In this notation we can rewrite Equation (3) as

ch𝔥M(λ,I)=eλU(λ,I)U(0,S).\mathrm{ch}_{\mathfrak{h}}\ M(\lambda,I)=e^{\lambda}\frac{U(\lambda,I)}{U(0,S)}.

3. Technical results

In this section, we will prove some technical results that will be used later to prove our main theorems. We freely use the notations that were developed in the previous section. Let us define 𝒫:={(λ,I)|λ𝔥 and IJλ}\mathcal{P}:=\{(\lambda,I)|\lambda\in\mathfrak{h}^{*}\text{ and }I\subseteq J_{\lambda}\}. This is the indexing set for the Parabolic Verma modules of 𝔤\mathfrak{g}.

3.1.

The following is elementary; see, for example, [9, §4.1 (in particular, Lemma 2)].

Proposition 1.

Let (λ,I)𝒫(\lambda,I)\in\mathcal{P} and wWIw\in W_{I}. Then

  1. (1)

    λ+ρw(λ+ρ)Q+,\lambda+\rho-w(\lambda+\rho)\in Q^{+},

  2. (2)

    supp(w(λ+ρ)(λ+ρ))=I(w)\operatorname{supp}(w(\lambda+\rho)-(\lambda+\rho))=I(w)

Note that U(λ,I)U(\lambda,I) can be viewed as a formal power series in the variables {xi:=eαi:iI}\{x_{i}:=e^{-\alpha_{i}}:i\in I\} by Proposition 1. For (λ,I)𝒫(\lambda,I)\in\mathcal{P}, we let

L(λ,I):=log(U(λ,I))=αQcαλ,IeαL(\lambda,I):=-log(U(\lambda,I))=\sum\limits_{\alpha\in Q}c^{\lambda,I}_{\alpha}\;e^{-\alpha}

Here, the logarithm is applied to U(λ,I)U(\lambda,I) treating it as a formal power series whose constant term is 11. Note that

cαλ,I=0ifαQ+.c^{\lambda,I}_{\alpha}=0\,\text{if}\,\alpha\notin Q^{+}.

We need some additional notations which we collect here:

𝒞:={(λ,I)𝒫|Iis connected and nonempty}\displaystyle\mathcal{C}:=\{(\lambda,I)\in\mathcal{P}|\;I\;\text{is connected and nonempty}\}
β(λ,I):=iI(λ+ρ)(αi)αiQ+ for (λ,I)𝒫.\displaystyle\beta(\lambda,I):=\sum\limits_{i\in I}(\lambda+\rho)(\alpha_{i}^{\vee})\,\alpha_{i}\in Q^{+}\text{ for }(\lambda,I)\in\mathcal{P}.

3.2.

We need the following key results.

Proposition 2.

Let (λ,I)𝒫(\lambda,I)\in\mathcal{P} and let I=I1˙˙IrI=I_{1}\dot{\cup}\cdots\dot{\cup}I_{r} be the decomposition of II into connected components. Then

  1. (1)

    (λ,J)𝒫(\lambda,J)\in\mathcal{P} for all JIJ\subseteq I.

  2. (2)

    L(λ,I)=L(μ,J)U(λ,I)=U(μ,J)β(λ,I)=β(μ,J)I=Jandλ(αi)=μ(αi)L(\lambda,I)=L(\mu,J)\iff U(\lambda,I)=U(\mu,J)\iff\beta(\lambda,I)=\beta(\mu,J)\iff I=J\;and\;\lambda(\alpha_{i}^{\vee})=\mu(\alpha_{i}^{\vee}) for all iIi\in I.

  3. (3)

    L(λ,I)=k=1rL(λ,Ik)L(\lambda,I)=\sum\limits_{k=1}^{r}L(\lambda,I_{k}).

  4. (4)

    cαλ,I=cαλ,supp(α)c^{\lambda,I}_{\alpha}=c^{\lambda,\operatorname{supp}(\alpha)}_{\alpha} for all αQ\alpha\in Q and Isupp(α)I\supseteq\operatorname{supp}(\alpha).

  5. (5)

    cαλ,I0c^{\lambda,I}_{\alpha}\neq 0 implies that supp(α)\operatorname{supp}(\alpha) is a connected subset of II.

  6. (6)

    cαλ,I0c^{\lambda,I}_{\alpha}\neq 0 and supp(α)=I\operatorname{supp}(\alpha)=I implies αβ(λ,I)Q+\alpha-\beta(\lambda,I)\in Q^{+}.

Proof.

  1. (1)

    Immediate from the definition.

  2. (2)

    Comparing monomials of the form ekαi,iSe^{-k\alpha_{i}},\;i\in S gives the equivalence.

  3. (3)

    Suppose I=I1I2I=I_{1}\cup I_{2} where αi1(αi2)=0\alpha_{i_{1}}(\alpha_{i_{2}}^{\vee})=0 for all i1I1i_{1}\in I_{1} and i2I2i_{2}\in I_{2}. Then WI=WI1×WI2W_{I}=W_{I_{1}}\times W_{I_{2}} because the simple reflections sαi1s_{\alpha_{i_{1}}} and sαi2s_{\alpha_{i_{2}}} commute for i1I1i_{1}\in I_{1} and i2I2i_{2}\in I_{2}. For wWIw\in W_{I}, there exists unique w1WI1w_{1}\in W_{I_{1}} and w2WI2w_{2}\in W_{I_{2}} such that w=w1w2w=w_{1}w_{2}. Therefore we have

    w(λ+ρ)(λ+ρ)\displaystyle w(\lambda+\rho)-(\lambda+\rho) =w1w2(λ+ρ)(λ+ρ)\displaystyle=w_{1}w_{2}(\lambda+\rho)-(\lambda+\rho)
    =w1w2(λ+ρ)w1(λ+ρ)+w1(λ+ρ)(λ+ρ)\displaystyle=w_{1}w_{2}(\lambda+\rho)-w_{1}(\lambda+\rho)+w_{1}(\lambda+\rho)-(\lambda+\rho)
    =w1[w2(λ+ρ)(λ+ρ)]+[w1(λ+ρ)(λ+ρ)]\displaystyle=w_{1}[w_{2}(\lambda+\rho)-(\lambda+\rho)]+[w_{1}(\lambda+\rho)-(\lambda+\rho)]

    But supp(w2(λ+ρ)(λ+ρ))I2\operatorname{supp}(w_{2}(\lambda+\rho)-(\lambda+\rho))\subseteq I_{2} and hence w1w_{1} fixes it. Therefore we have

    w(λ+ρ)(λ+ρ)=w1(λ+ρ)(λ+ρ)+w2(λ+ρ)(λ+ρ)w(\lambda+\rho)-(\lambda+\rho)=w_{1}(\lambda+\rho)-(\lambda+\rho)+w_{2}(\lambda+\rho)-(\lambda+\rho)

    Hence,

    U(λ,I)\displaystyle U(\lambda,I) =wWI(1)l(w)ew(λ+ρ)(λ+ρ)=(w1,w2)WI1×WI2(1)l(w1w2)ew1w2(λ+ρ)(λ+ρ)\displaystyle=\sum_{w\in W_{I}}(-1)^{l(w)}e^{w(\lambda+\rho)-(\lambda+\rho)}\quad\quad=\sum_{(w_{1},w_{2})\in W_{I_{1}}\times W_{I_{2}}}(-1)^{l(w_{1}w_{2})}e^{w_{1}w_{2}(\lambda+\rho)-(\lambda+\rho)}
    =(w1WI1(1)l(w1)ew1(λ+ρ)(λ+ρ))(w2WI2(1)l(w2)ew2(λ+ρ)(λ+ρ))\displaystyle=\left(\sum_{w_{1}\in W_{I_{1}}}(-1)^{l(w_{1})}e^{w_{1}(\lambda+\rho)-(\lambda+\rho)}\right)\left(\sum_{w_{2}\in W_{I_{2}}}(-1)^{l(w_{2})}e^{w_{2}(\lambda+\rho)-(\lambda+\rho)}\right)
    =U(λ,I1)U(λ,I2)\displaystyle=U(\lambda,I_{1})\cdot U(\lambda,I_{2})

    Now taking log-log on both sides, we get

    L(λ,I)=L(λ,I1)+L(λ,I2)L(\lambda,I)=L(\lambda,I_{1})+L(\lambda,I_{2})
  4. (4)

    Fix αQ.\alpha\in Q. Let ψ:=U(λ,I)1ζ\psi:=U(\lambda,I)-1-\zeta where

    ζ:=wWsupp(α)\{e}(1)l(w)ew(λ+ρ)(λ+ρ)\zeta:=\sum_{w\in W_{\operatorname{supp}(\alpha)}\backslash\{e\}}(-1)^{l(w)}e^{w(\lambda+\rho)-(\lambda+\rho)}

    Then L(λ,I)=k1(1)k(ζ+ψ)k/kL(\lambda,I)=\sum_{k\geq 1}(-1)^{k}(\zeta+\psi)^{k}/k. But since the support of any monomial in ψ\psi is not a subset of supp(α)\operatorname{supp}(\alpha), it follows that eαe^{-\alpha} has contributions only from k1(1)kζk/k=L(λ,supp(α))\sum_{k\geq 1}(-1)^{k}\zeta^{k}/k=L(\lambda,\operatorname{supp}(\alpha)). Therefore cαλ,I=cαλ,supp(α)c^{\lambda,I}_{\alpha}=c^{\lambda,\operatorname{supp}(\alpha)}_{\alpha}.

  5. (5)

    The fact that supp(α)I\operatorname{supp}(\alpha)\subseteq I follows from part (2) of Proposition 1. Now, suppose that supp(α)\operatorname{supp}(\alpha) is disconnected. Let supp(α)=I1˙I2\operatorname{supp}(\alpha)=I_{1}\dot{\cup}I_{2}, where I1I_{1} and I2I_{2} are proper non-empty subsets of supp(α)\operatorname{supp}(\alpha) and αi1(αi2)=0\alpha_{i_{1}}(\alpha_{i_{2}}^{\vee})=0 for all i1I1i_{1}\in I_{1} and i2I2i_{2}\in I_{2}. By part (3) of this proposition, we have L(λ,supp(α))=L(λ,I1)+L(λ,I2)L(\lambda,\operatorname{supp}(\alpha))=L(\lambda,I_{1})+L(\lambda,I_{2}). Therefore, cαλ,I=cαλ,supp(α)=cαλ,I1+cαλ,I2c^{\lambda,I}_{\alpha}=c^{\lambda,\operatorname{supp}(\alpha)}_{\alpha}=c^{\lambda,I_{1}}_{\alpha}+c^{\lambda,I_{2}}_{\alpha}. But since supp(α)\operatorname{supp}(\alpha) is not a subset of I1I_{1} or I2I_{2} we have cαλ,I1=cαλ,I2=0c^{\lambda,I_{1}}_{\alpha}=c^{\lambda,I_{2}}_{\alpha}=0.

  6. (6)

    The coefficient of αi\alpha_{i} in λ+ρw(λ+ρ)\lambda+\rho-w(\lambda+\rho) is either 0 or greater than or equal to (λ+ρ)(αi)(\lambda+\rho)(\alpha_{i}^{\vee}) for all iSi\in S. This fact is elementary to prove. For example, see [9, Lemma 2(b)] where it is proved by induction on the length of ww. Now this means that the variable eαie^{-\alpha_{i}} has degree equal to 0 or greater than (λ+ρ)(αi)(\lambda+\rho)(\alpha_{i}^{\vee}) in any monomial in U(λ,I)U(\lambda,I). But since any monomial in L(λ,I)L(\lambda,I) is a product of monomials from U(λ,I)U(\lambda,I), the proof follows.

The following proposition is very crucial and follows from Propositions 3 and 7 of [9] (also see part (4) of Proposition 2, see also Exercise 1.2 in [5]). We give a sketch of the proof for reader’s convenience.

Proposition 3.

Let (λ,I)𝒫(\lambda,I)\in\mathcal{P} as before.

  1. (1)

    cβ(λ,I)λ,Ic^{\lambda,I}_{\beta(\lambda,I)} is independent of λ\lambda. i.e., cβ(λ,I)λ,I=cβ(μ,I)μ,Ic^{\lambda,I}_{\beta(\lambda,I)}=c^{\mu,I}_{\beta(\mu,I)} if (μ,I)𝒫(\mu,I)\in\mathcal{P} with IJλJμI\subseteq J_{\lambda}\cap J_{\mu}.

  2. (2)

    cβ(λ,J)λ,I>0c^{\lambda,I}_{\beta(\lambda,J)}>0 for any non-empty, connected subset JJ of II.

  3. (3)

    In particular, if (λ,I)𝒞(\lambda,I)\in\mathcal{C} then cβ(λ,I)λ,I>0c^{\lambda,I}_{\beta(\lambda,I)}>0.

Proof.
  1. (1)

    Consider the subgraph 𝒢I\mathcal{G}_{I} of the Dynkin diagram of 𝔤\mathfrak{g} induced by II. We then have

    cβ(λ,I)λ,I=(1)|I|k1(1)k|𝒫k(𝒢I)|kc^{\lambda,I}_{\beta(\lambda,I)}=(-1)^{|I|}\sum_{k\geq 1}\frac{(-1)^{k}\;|\mathcal{P}_{k}(\mathcal{G}_{I})|}{k}

    where 𝒫k(𝒢I)\mathcal{P}_{k}(\mathcal{G}_{I}) is the set of all kk-tuples (J1,,Jk)(J_{1},\ldots,J_{k}) of pairwise disjoint subsets of II such that

    1. (a)

      J1˙˙Jk=IJ_{1}\;\dot{\cup}\;\cdots\;\dot{\cup}\;J_{k}=I

    2. (b)

      JiJ_{i} is totally disconnected (i.e., x,yJi\forall x,y\in J_{i} there is no edge between xx and yy in 𝒢I\mathcal{G}_{I}).

    The RHS is clearly independent of λ\lambda, see [9, §4.3] for more details.

  2. (2)

    By (1), it is enough to consider λ=0\lambda=0. By the Weyl denominator identity (for the parabolic subalgebra 𝔤I\mathfrak{g}_{I}) we have

    U(0,I)=αΔ+(I)(1eα)mult(α)U(0,I)=\prod_{\alpha\in\Delta_{+}(I)}(1-e^{-\alpha})^{\mathrm{mult}(\alpha)}

    where, Δ+(I)=-span{αi:iI}Δ+\Delta_{+}(I)=\mathbb{Z}\text{-span}\{\alpha_{i}:i\in I\}\cap\Delta_{+} or equivalently the set of positive roots for the parabolic subalgebra 𝔤I\mathfrak{g}_{I}. Applying log-log we get

    L(0,I)=αΔ+(I)mult(α)k1ekαkL(0,I)=\sum_{\alpha\in\Delta_{+}(I)}\mathrm{mult}(\alpha)\sum_{k\geq 1}\frac{e^{-k\alpha}}{k}

    The coefficient of β(0,I)=iIαi\beta(0,I)=\sum_{i\in I}\alpha_{i} in L(0,I)L(0,I) is then the multiplicity of iIαi\sum_{i\in I}\alpha_{i}. But since II is connected we have this multiplicity to be a positive integer (see [5, Lemma 1.6] and [9, Proposition 4 & 7]).

  3. (3)

    Follows immediately from (2).

4. Unique factorization for Parabolic Verma Modules

In this section, we will prove the unique factorization of tensor products of parabolic Verma modules of 𝔤.\mathfrak{g}. First, we analyze when a sum of finitely many L(λ,I)L(\lambda,I)’s can be equal to another such sum.

4.1.

The following relation \succeq on 𝒫\mathcal{P} will play an important role in this paper. Define (λ,I)(μ,J)(\lambda,I)\succeq(\mu,J) if:

either IJI\supsetneq J   or   I=JI=J and β(μ,J)β(λ,I)Q+\beta(\mu,J)-\beta(\lambda,I)\in Q^{+}.

Observe that the latter part of this condition may be exchanged with

I=JI=J and λ(αi)μ(αi)for alliI\lambda(\alpha_{i}^{\vee})\leq\mu(\alpha_{i}^{\vee})\;\;\text{for all}\;i\in I.

Note that this relation is reflexive, and transitive but not anti-symmetric. i.e., (λ,I)(μ,J)(\lambda,I)\succeq(\mu,J) and (μ,J)(λ,I)(\mu,J)\succeq(\lambda,I) does not imply that (λ,I)=(μ,J)(\lambda,I)=(\mu,J).

For (λ,I)(\lambda,I) and (μ,J)𝒫(\mu,J)\in\mathcal{P} we write (λ,I)(μ,J)(\lambda,I)\approx(\mu,J) if I=JI=J and λ(αi)=μ(αi)\lambda(\alpha_{i}^{\vee})=\mu(\alpha_{i}^{\vee}) for all iIi\in I. This defines an equivalence relation on 𝒫\mathcal{P}. Observe that (λ,I)(μ,J)(\lambda,I)\approx(\mu,J) means that these pairs satisfy the equivalent conditions of part (2) of Proposition 2. The relation \succeq now defines a partial order on 𝒫/\mathcal{P}/\!\approx.

Caveat: Even though \succeq does not form a partial order on 𝒫\mathcal{P} we will find it convenient nevertheless to talk about maximal elements in a subset of 𝒫\mathcal{P}. What we actually mean by saying (λ1,I1)(\lambda_{1},I_{1}) is maximal among {(λ1,I1),\{(\lambda_{1},I_{1}), (λ2,I2),(\lambda_{2},I_{2}), ,\ldots, (λk,Ik)}(\lambda_{k},I_{k})\} is that:

if (λj,Ij)(λ1,I1)(\lambda_{j},I_{j})\succeq(\lambda_{1},I_{1}) for some jj then (λ1,I1)(λj,Ij)(\lambda_{1},I_{1})\approx(\lambda_{j},I_{j})

or equivalently, when thought of as elements of 𝒫/\mathcal{P}/\!\!\approx, (λ1,I1)(\lambda_{1},I_{1}) is maximal among {(λ1,I1),\{(\lambda_{1},I_{1}), (λ2,I2),(\lambda_{2},I_{2}), ,\ldots, (λk,Ik)}(\lambda_{k},I_{k})\}.

Lemma 1.

Let r0r\geq 0. Let (λ1,I1),,(λr,Ir)(\lambda_{1},I_{1}),\ldots,(\lambda_{r},I_{r}) be (not necessarily distinct) elements of 𝒞\mathcal{C} and let L:=k=1rL(λk,Ik)L:=\sum_{k=1}^{r}L(\lambda_{k},I_{k}).

  1. (1)

    For (μ,J)𝒫(\mu,J)\in\mathcal{P}, if the coefficient of eβ(μ,J)e^{-\beta(\mu,J)} in LL is non-zero, then (λj,Ij)(μ,J)(\lambda_{j},I_{j})\succeq(\mu,J) for some 1jr1\leq j\leq r.

  2. (2)

    If (λj,Ij)(\lambda_{j},I_{j}) is a maximal element with respect to \succeq among (λ1,I1),,(λr,Ir)(\lambda_{1},I_{1}),\ldots,(\lambda_{r},I_{r}) then the coefficient of eβ(λj,Ij)e^{-\beta(\lambda_{j},I_{j})} in LL is positive. In particular, L0L\neq 0 if r0r\neq 0.

Proof.

  1. (1)

    By hypothesis we have cβ(μ,J)λj,Ij0c^{\lambda_{j},I_{j}}_{\beta(\mu,J)}\neq 0 for some 1jr1\leq j\leq r. By Proposition 2 (5), we have JIjJ\subseteq I_{j}. If JIjJ\subsetneq I_{j} then (λj,Ij)(μ,J)(\lambda_{j},I_{j})\succeq(\mu,J). Suppose J=IjJ=I_{j}. By part (4) and (6) of Proposition 2 we have β(μ,J)β(λ,J)Q+\beta(\mu,J)-\beta(\lambda,J)\in Q^{+}. Therefore (λj,Ij)(μ,J)(\lambda_{j},I_{j})\succeq(\mu,J).

  2. (2)

    By part (1) and maximality of (λj,Ij)(\lambda_{j},I_{j}), if the coefficient of eβ(λj,Ij)e^{-\beta(\lambda_{j},I_{j})} in L(λk,Ik)L(\lambda_{k},I_{k}) is non-zero then (λj,Ij)(λk,Ik)(\lambda_{j},I_{j})\approx(\lambda_{k},I_{k}). In such a case, β(λj,Ij)=β(λk,Ik)\beta(\lambda_{j},I_{j})=\beta(\lambda_{k},I_{k}) and hence Proposition 3 implies that the coefficient of eβ(λj,Ij)e^{-\beta(\lambda_{j},I_{j})} in L(λk,Ik)L(\lambda_{k},I_{k}) is positive.

Theorem 3.

Let r,s0r,s\geq 0. Let (λ1,I1)(\lambda_{1},I_{1}), (λ2,I2),,(λr,Ir)(\lambda_{2},I_{2}),\ldots,(\lambda_{r},I_{r}) and (μ1,J1)(\mu_{1},J_{1}), (μ2,J2),(\mu_{2},J_{2}), ,\ldots, (μs,Js)𝒞(\mu_{s},J_{s})\in\mathcal{C}. We have

k=1rL(λk,Ik)=k=1sL(μk,Jk)\sum_{k=1}^{r}L(\lambda_{k},I_{k})=\sum_{k=1}^{s}L(\mu_{k},J_{k}) (4)

if and only if r=sr=s and there exists a permutation σ𝔖r\sigma\in\mathfrak{S}_{r} such that (λk,Ik)(μσ(k),Jσ(k))(\lambda_{k},I_{k})\approx(\mu_{\sigma(k)},J_{\sigma(k)}) for 1kr1\leq k\leq r.

Proof.

The reverse implication easily follows from part(2) of Proposition 2. We prove the forward implication by induction on m:=min{r,s}m:=\min\{r,s\}. If m=0m=0, then r=s=0r=s=0 follows from (2) of Lemma 1. Suppose m1m\geq 1. Without loss of generality assume that (λ1,I1)(\lambda_{1},I_{1}) is maximal among {(λ1,I1)\{(\lambda_{1},I_{1}) ,,(λr,Ir),(μ1,J1),,(μs,Js)},\cdots,(\lambda_{r},I_{r}),(\mu_{1},J_{1}),\cdots,(\mu_{s},J_{s})\} viewed as elements of 𝒫/\mathcal{P}/\approx. By (2) of Lemma 1, we see that the coefficient of eβ(λ1,I1)e^{-\beta(\lambda_{1},I_{1})} is non-zero in the left-hand side of the Eq. (4), and therefore also on the right-hand side. Now by (1) of Lemma 1 there exists kk such that (μk,Jk)(λ1,I1)(\mu_{k},J_{k})\succeq(\lambda_{1},I_{1}). But by maximality of (λ1,I1)(\lambda_{1},I_{1}) we conclude that (μk,Jk)(λ1,I1)(\mu_{k},J_{k})\approx(\lambda_{1},I_{1}). Therefore we may cancel L(λ1,I1)=L(μk,Jk)L(\lambda_{1},I_{1})=L(\mu_{k},J_{k}) from both sides of Eq. (4), thereby reducing the value of mm by 11. ∎

Corollary 1.

Let (λ1,I1)(\lambda_{1},I_{1}), (λ2,I2),,(λr,Ir)(\lambda_{2},I_{2}),\ldots,(\lambda_{r},I_{r}) and (μ1,J1)(\mu_{1},J_{1}), (μ2,J2),,(μr,Jr)𝒫(\mu_{2},J_{2}),\ldots,(\mu_{r},J_{r})\in\mathcal{P} such that all Ik,JkI_{k},J_{k} are connected (possibly empty). Then

k=1rch𝔥(M(λk,Ik))=k=1rch𝔥(M(μk,Jk))\prod_{k=1}^{r}ch_{\mathfrak{h}}(M(\lambda_{k},I_{k}))=\prod_{k=1}^{r}ch_{\mathfrak{h}}(M(\mu_{k},J_{k})) (5)

if and only if k=1rλk=k=1rμk\sum_{k=1}^{r}\lambda_{k}=\sum_{k=1}^{r}\mu_{k} and there exists σ𝔖r\sigma\in\mathfrak{S}_{r} such that (λk,Ik)(μσ(k),Jσ(k))(\lambda_{k},I_{k})\approx(\mu_{\sigma(k)},J_{\sigma(k)}) for all 1kr1\leq k\leq r.

Proof.

We will prove only the forward direction, the converse following easily from Proposition 2. Rewriting (5) using the character formula for parabolic Verma modules, we get

k=1reλkU(λk,Ik)U(0,S)=k=1reμkU(μk,Jk)U(0,S)\prod_{k=1}^{r}e^{\lambda_{k}}\frac{U(\lambda_{k},I_{k})}{U(0,S)}=\prod_{k=1}^{r}e^{\mu_{k}}\frac{U(\mu_{k},J_{k})}{U(0,S)} (6)

Comparing the highest weights on both sides of (5), we get k=1rλk=k=1rμk\sum_{k=1}^{r}\lambda_{k}=\sum_{k=1}^{r}\mu_{k}. Therefore (6) gives:

k=1rU(λk,Ik)=k=1rU(μk,Jk)\prod_{k=1}^{r}U(\lambda_{k},I_{k})=\prod_{k=1}^{r}U(\mu_{k},J_{k})

Note that U(λ,I)=1U(\lambda,I)=1 iff I=I=\emptyset. Ignoring such trivial terms in the above product on both sides we have up to a relabelling

k=1tU(λk,Ik)=k=1sU(μk,Jk)\prod_{k=1}^{t}U(\lambda_{k},I_{k})=\prod_{k=1}^{s}U(\mu_{k},J_{k})

where, now {(λ1,I1),,(λt,It),(μ1,J1),,(μs,Js)}𝒞\{(\lambda_{1},I_{1}),\ldots,(\lambda_{t},I_{t}),(\mu_{1},J_{1}),\ldots,(\mu_{s},J_{s})\}\subseteq\mathcal{C}. Taking log on both sides, and applying Theorem 3 we get t=st=s and there is a permutation σ𝔖t\sigma\in\mathfrak{S}_{t} such that (λk,Ik)(μσk,Jσk)(\lambda_{k},I_{k})\approx(\mu_{\sigma k},J_{\sigma k}). But s=ts=t implies that the number of trivial terms on both sides was also equal to begin with. Extending σ\sigma trivially to a bijection of {1,,r}\{1,\ldots,r\}, we get the required permutation (because (λ,)(μ,)(\lambda,\emptyset)\approx(\mu,\emptyset) for any λ,μ𝔥\lambda,\mu\in\mathfrak{h}^{*}). ∎

In particular, Theorem 1 is now immediate from the above corollary.

5. Unique Factorization for Restricted Parabolic Verma Modules

In this section, we prove the unique factorization of tensor products for certain classes of parabolic Verma modules restricted to compatible subalgebras of 𝔤\mathfrak{g}.

5.1.

We begin with some auxiliary results. Recall that 𝔤\mathfrak{g} is a symmetrizable Kac-Moody algebra whose simple roots are {α1,,αn}\{\alpha_{1},\ldots,\alpha_{n}\}. Fix an equivalence relation \sim on S={1,,n}S=\{1,\ldots,n\}. This gives rise to a set partition of SS.

Definition 1.

Let KSK\subseteq S be such that KK is a union of equivalence classes. We say that K^\widehat{K} is a lift of KK if

  1. (1)

    K^K\widehat{K}\subseteq K

  2. (2)

    K^\widehat{K} is connected

  3. (3)

    K^\widehat{K} meets every equivalence class in KK.

Definition 2.

Let KSK\subseteq S be such that KK is a union of equivalence classes. We say that KK is equiconnected if there exists a lift K^\widehat{K} of KK such that given any lift K¯\bar{K} of KK and any equivalence class EE, |K¯E||K^E||\bar{K}\cap E|\geq|\widehat{K}\cap E|. Any such lift K^\widehat{K} will be referred to as a lean lift.

Remark 1.

Note that if K^\widehat{K} and K^\widehat{K}^{\prime} are two lean lifts of KK, then given any equivalence class EE we have

|K^E|=|K^E||\widehat{K}\cap E|=|\widehat{K}^{\prime}\cap E| (7)
\bullet\bullet\bullet\bullet\bullet1122334455
Figure 1. The equivalence classes are given by {1}\{1\}, {2}\{2\}, {3}\{3\} and {4,5}\{4,5\}.

In figure 1, there are two lean lifts for S={1,2,3,4,5}S=\{1,2,3,4,5\} namely: {1,2,3,4}\{1,2,3,4\} and {1,2,3,5}\{1,2,3,5\}.

\bullet55\bullet44\bullet33\bullet66\bullet22\bullet11\bullet88\bullet99\bullet77
Figure 2. The equivalence classes are given by {5},{3,4,6}\{5\},\{3,4,6\}, {2,7,8}\{2,7,8\} and {1,9}\{1,9\}.

In figure 2, for K={1,,9}K=\{1,\ldots,9\}, the subsets K1={1,2,3,4,5}K_{1}=\{1,2,3,4,5\} and K2={5,6,7,8,9}K_{2}=\{5,6,7,8,9\} are both lifts but |K1{3,4,6}|>|K2{3,4,6}||K_{1}\cap\{3,4,6\}|>|K_{2}\cap\{3,4,6\}| while |K1{2,8,7}|<|K2{2,8,7}||K_{1}\cap\{2,8,7\}|<|K_{2}\cap\{2,8,7\}|. It follows that KK is not equiconnected.

Definition 3.

We call λ𝔥\lambda\in\mathfrak{h}^{*} to be symmetric if i,jS\forall\;i,j\in S such that iji\sim j, we have λ(αi)=λ(αj)\lambda(\alpha_{i}^{\vee})=\lambda(\alpha_{j}^{\vee}).

Remark 2.

Suppose K^\widehat{K} is a lean lift and K¯\bar{K} is some lift of KK. It is immediate from Remark 1 that if λ\lambda is symmetric, then β(λ,K¯)β(λ,K^)Q+\beta(\lambda,\bar{K})-\beta(\lambda,\widehat{K})\in Q^{+}. Further β(λ,K¯)=β(λ,K^)\beta(\lambda,\bar{K})=\beta(\lambda,\widehat{K}) if and only if K¯\bar{K} is a lean lift of KK.

Define 𝒞¯\bar{\mathcal{C}} to be the set of pairs (λ,I)𝒞(\lambda,I)\in\mathcal{C} satisfying

  1. (1)

    λ\lambda is symmetric

  2. (2)

    II is a union of equivalence classes of \sim

  3. (3)

    II is equiconnected.

Define Q¯:=[j]S/γ[j]\overline{Q}:=\oplus_{[j]\in S/\!\sim}\;\mathbb{Z}\gamma_{[j]}. The set of all non-negative integer linear combinations of {γ[j]:[j]S/}\{\gamma_{[j]}\;:\;[j]\in S/\!\!\sim\} is denoted by Q¯+\overline{Q}^{+}. Define the map

π:QQ¯\pi:Q\rightarrow\overline{Q}

where αi\alpha_{i} maps to (the formal symbol) γ[i]\gamma_{[i]}. This induces a map from [[{eαi|iS}]]\mathbb{Z}[[\{e^{-\alpha_{i}}|i\in S\}]] to [[{eγ[i]|[i]S/}]]\mathbb{Z}[[\{e^{-\gamma_{[i]}}|[i]\in S/\!\!\sim\}]] which we again denote by π\pi. For (λ,K)𝒞¯(\lambda,K)\in\bar{\mathcal{C}}, we define

β¯(λ,K):=π(β(λ,K^))\bar{\beta}(\lambda,K):=\pi(\beta(\lambda,\widehat{K}))

for any lean lift K^\widehat{K} of KK as in Definition 2. Observe that this does not depend on the choice of K^\widehat{K} by (7).

Lemma 2.

If (λ,I),(μ,J)𝒞¯(\lambda,I),(\mu,J)\in\bar{\mathcal{C}} are such that β¯(λ,I)=β¯(μ,J)\bar{\beta}(\lambda,I)=\bar{\beta}(\mu,J), then (λ,I)(μ,J)(\lambda,I)\approx(\mu,J).

Proof.

Since β¯(λ,I)=β¯(μ,J)\bar{\beta}(\lambda,I)=\bar{\beta}(\mu,J), we have for any choice of I^\hat{I} and J^\hat{J}:

π(iI^(λ+ρ)(αi)αi)=π(iJ^(μ+ρ)(αi)αi)\pi\left(\sum_{i\in\hat{I}}(\lambda+\rho)(\alpha_{i}^{\vee})\alpha_{i}\right)=\pi\left(\sum_{i\in\hat{J}}(\mu+\rho)(\alpha_{i}^{\vee})\alpha_{i}\right)
[i]I/(j[i]I^(λ+ρ)(αi))γ[i]=[i]J/(j[i]J^(μ+ρ)(αi))γ[i]\sum_{[i]\in I/\sim}\left(\sum_{j\in[i]\cap\hat{I}}(\lambda+\rho)(\alpha_{i}^{\vee})\right)\gamma_{[i]}=\sum_{[i]\in J/\sim}\left(\sum_{j\in[i]\cap\hat{J}}(\mu+\rho)(\alpha_{i}^{\vee})\right)\gamma_{[i]}

Since λ+ρ\lambda+\rho and μ+ρ\mu+\rho are regular dominant, γ[i]\gamma_{[i]}’s are linearly independent and I,JI,J are unions of equivalence classes of \sim, it follows that I=JI=J. Now,

[i]I/|[i]I^|(λ+ρ)(αi)γ[i]=[i]J/|[i]J^|(μ+ρ)(αi)γ[i]\sum_{[i]\in I/\sim}|[i]\cap\hat{I}|\cdot(\lambda+\rho)(\alpha_{i}^{\vee})\gamma_{[i]}=\sum_{[i]\in J/\sim}|[i]\cap\hat{J}|\cdot(\mu+\rho)(\alpha_{i}^{\vee})\gamma_{[i]}

Since, II and JJ were equiconnected we have, for all [i][i],

|[i]I^|=|[i]J^||[i]\cap\hat{I}|=|[i]\cap\hat{J}|

By comparing coefficients of γ[i]\gamma_{[i]} one sees that (λ,I)(μ,J)(\lambda,I)\approx(\mu,J). ∎

Remark 3.

The relation \succeq on 𝒫\mathcal{P} can be restricted to 𝒞¯\bar{\mathcal{C}}. It is elementary to check that for (λ,I),(μ,J)𝒞¯(\lambda,I),(\mu,J)\in\bar{\mathcal{C}} we have (λ,I)(μ,J)(\lambda,I)\succeq(\mu,J) if and only if one of the following holds:

  • IJI\supsetneq J or

  • I=JI=J and β¯(μ,J)β¯(λ,I)Q¯+\bar{\beta}(\mu,J)-\bar{\beta}(\lambda,I)\in\overline{Q}^{+}

For (λ,I)𝒫(\lambda,I)\in\mathcal{P} let us write π(L(λ,I))=γQ¯dγλ,Ieγ\pi(L(\lambda,I))=\sum\limits_{\gamma\in\overline{Q}}d^{\lambda,I}_{\gamma}e^{-\gamma}. Observe that dγλ,I=αcαλ,Id^{\lambda,I}_{\gamma}=\sum\limits_{\alpha}c^{\lambda,I}_{\alpha}, where the sum runs over all αQ\alpha\in Q for which π(α)=γ\pi(\alpha)=\gamma. The following is the analogue of Propositions 2 and 3:

Proposition 4.

Suppose (λ,I)𝒞¯(\lambda,I)\in\bar{\mathcal{C}}.

  1. (1)

    If γQ¯\gamma\in\overline{Q} is such that supp(γ)=I/\operatorname{supp}(\gamma)=I/\!\!\sim and dγλ,I0d^{\lambda,I}_{\gamma}\neq 0 then γβ¯(λ,I)Q¯+\gamma-\bar{\beta}(\lambda,I)\in\overline{Q}^{+}.

  2. (2)

    The coefficient of eβ¯(λ,I)e^{-\bar{\beta}(\lambda,I)} in π(L(λ,I))\pi(L(\lambda,I)) is positive (i.e., dβ¯(λ,I)λ,I>0d^{\lambda,I}_{\bar{\beta}(\lambda,I)}>0) and independent of λ\lambda for fixed II.

Proof.

Suppose αQ\alpha\in Q is such that cαλ,I0c^{\lambda,I}_{\alpha}\neq 0 and π(α)=γ\pi(\alpha)=\gamma. This implies that supp(α)\operatorname{supp}(\alpha) is a lift of II (see part (5) in Proposition 2). If I^\hat{I} is any lean lift of II then equiconnectedness of II implies that π(β(λ,supp(α)))π(β(λ,I^))Q¯+\pi(\beta(\lambda,\operatorname{supp}(\alpha)))-\pi(\beta(\lambda,\hat{I}))\in\overline{Q}^{+}. Also, by (4) and (6) of Proposition 2, we have αβ(λ,supp(α))Q+\alpha-\beta(\lambda,\operatorname{supp}(\alpha))\in Q^{+}. Therefore by applying π\pi and combining with the previous observation we see that γβ¯(λ,I)Q¯+\gamma-\bar{\beta}(\lambda,I)\in\overline{Q}^{+}. This proves (1)(1).

Suppose if α\alpha is such that cαλ,I0c^{\lambda,I}_{\alpha}\neq 0 and π(α)=β¯(λ,I)\pi(\alpha)=\bar{\beta}(\lambda,I). It follows from equiconnectedness of II and part (5) in Proposition 2 that α=β(λ,I^)\alpha=\beta(\lambda,\hat{I}) for some lean lift I^\hat{I} of II. Therefore

dβ(λ,I)¯λ,I=I^cβ(λ,I^)λ,Id^{\lambda,I}_{\bar{\beta(\lambda,I)}}=\sum_{\hat{I}}c^{\lambda,I}_{{\beta}(\lambda,\hat{I})}

where the sum runs over all lean lifts I^\hat{I} of II as in Definition 2. Part (2) of the lemma now follows from part (4) of Proposition 2 and Proposition 3. ∎

Lemma 3.

Let (λ1,I1),,(λr,Ir)(\lambda_{1},I_{1}),\ldots,(\lambda_{r},I_{r}) be (not necessarily distinct) elements of 𝒞¯\bar{\mathcal{C}} and let L:=i=1rL(λk,Ik)L:=\sum_{i=1}^{r}L(\lambda_{k},I_{k}).

  1. (1)

    For (μ,J)𝒞¯(\mu,J)\in\bar{\mathcal{C}}, if the coefficient of eβ¯(μ,J)e^{-\bar{\beta}(\mu,J)} in π(L)\pi(L) is non-zero, then (λk,Ik)(μ,J)(\lambda_{k},I_{k})\succeq(\mu,J) for some 1kr1\leq k\leq r.

  2. (2)

    If (λj,Ij)(\lambda_{j},I_{j}) is maximal among (λ1,I1),,(λr,Ir)(\lambda_{1},I_{1}),\ldots,(\lambda_{r},I_{r}) then the coefficient of eβ¯(λj,Ij)e^{-\bar{\beta}(\lambda_{j},I_{j})} in π(L)\pi(L) is positive. In particular, π(L)0\pi(L)\neq 0 if r>0r>0.

Proof.

Suppose that the coefficient of eβ¯(μ,J)e^{-\bar{\beta}(\mu,J)} is non-zero, then this monomial must come from π(L(λk,Ik))\pi(L(\lambda_{k},I_{k})) for some IkJI_{k}\supseteq J. If this containment is proper then we are done. Suppose now that J=IkJ=I_{k}, then by part (1) of Proposition 4 and Remark 3, it follows that (λk,Ik)(μ,J)(\lambda_{k},I_{k})\succeq(\mu,J). By maximality of (λj,Ij)(\lambda_{j},I_{j}), part (2) of the lemma follows from part (1) together with Proposition 4 and Remark 3. ∎

5.2.

We are now ready to state and prove our main theorem for the restricted normalized Weyl numerators.

Theorem 4.

Let {(λk,Ik)}k=1r\{(\lambda_{k},I_{k})\}_{k=1}^{r} and {(μk,Jk)}k=1s\{(\mu_{k},J_{k})\}_{k=1}^{s} be subsets of 𝒞¯\bar{\mathcal{C}}. Then the following are equivalent:

  1. (1)

    k=1rL(λk,Ik)=k=1sL(μk,Jk)\sum_{k=1}^{r}L(\lambda_{k},I_{k})=\sum_{k=1}^{s}L(\mu_{k},J_{k})

  2. (2)

    k=1rπ(L(λk,Ik))=k=1sπ(L(μk,Jk))\sum_{k=1}^{r}\pi(L(\lambda_{k},I_{k}))=\sum_{k=1}^{s}\pi(L(\mu_{k},J_{k}))

  3. (3)

    r=sr=s and there exists σ𝔖r\sigma\in\mathfrak{S}_{r} such that (λk,Ik)(μσ(k),Jσ(k))(\lambda_{k},I_{k})\approx(\mu_{\sigma(k)},J_{\sigma(k)}) for all kk.

Proof.

The statement (2) follows from (1) by applying π\pi. The statement (1) follows from (3) by Proposition 2. We now prove that statement (2) implies (3). We proceed by induction on m:=min{r,s}m:=\min\{r,s\}. If m=0m=0, then r=s=0r=s=0 follows from part (2) of Lemma 3. Suppose m1m\geq 1. Without loss of generality we assume that (λ1,I1)(\lambda_{1},I_{1}) is maximal among {(λ1,I1),\{(\lambda_{1},I_{1}), ,\cdots, (λr,Ir),(\lambda_{r},I_{r}), (μ1,J1),,(μs,Js)}(\mu_{1},J_{1}),\ldots,(\mu_{s},J_{s})\}. By part (2) of Lemma 3, we see that the coefficient of eβ¯(λ1,I1)e^{-\bar{\beta}(\lambda_{1},I_{1})} is non-zero in the LHS of  (4). Now by part (1) of Lemma 3 there exists kk such that (μk,Jk)(λ1,I1)(\mu_{k},J_{k})\succeq(\lambda_{1},I_{1}). But by maximality of (λ1,I1)(\lambda_{1},I_{1}) we conclude that (μk,Jk)(λ1,I1)(\mu_{k},J_{k})\approx(\lambda_{1},I_{1}). Therefore we may cancel π(L(λ1,I1))=π(L(μk,Jk))\pi(L(\lambda_{1},I_{1}))=\pi(L(\mu_{k},J_{k})) from both sides of  (4), thereby reducing the value of mm by 11. ∎

5.3.

Now start with a Dynkin diagram GG. Let 𝔤\mathfrak{g} be the associated KMA. Recall that S={1,,n}S=\{1,\ldots,n\} is the vertex set of GG (or equivalently the set of simple roots). Given an equivalence relation \sim on SS we define the subspace 𝔨\mathfrak{k} of 𝔥\mathfrak{h} as follows:

𝔨:=ijKer(αiαj)\mathfrak{k}:=\cap_{i\sim j}Ker(\alpha_{i}-\alpha_{j}) (8)

Then on 𝔨\mathfrak{k} we have αi=αj\alpha_{i}=\alpha_{j} if iji\sim j (actually, 𝔨\mathfrak{k} is the largest subspace of 𝔥\mathfrak{h} where we have αi=αj\alpha_{i}=\alpha_{j} whenever iji\sim j). Moreover, for a given iSi\in S the element ω[i]:=kiωk𝔨\omega_{[i]}^{\vee}:=\sum_{k\sim i}\omega_{k}^{\vee}\in\mathfrak{k} satisfies αj(ω[i])=δ[i],[j]\alpha_{j}(\omega_{[i]}^{\vee})=\delta_{[i],[j]}. Here ωk\omega_{k}^{\vee} is a fixed choice of fundamental co-weight associated to the simple root αk\alpha_{k}. i.e., αl(ωk)=δk,ll\alpha_{l}(\omega_{k}^{\vee})=\delta_{k,l}\;\forall l. Therefore we have αi=αj\alpha_{i}=\alpha_{j} on 𝔨\mathfrak{k} iff iji\sim j. Moreover, we conclude the following:

Proposition 5.

Any collection of simple roots corresponding to distinct \sim orbit representatives forms a linearly independent set when restricted to 𝔨\mathfrak{k}. ∎

Let 𝔰\mathfrak{s} be any subspace of 𝔥\mathfrak{h} such that

  1. (1)

    αi|𝔰=αj|𝔰\alpha_{i}|_{\mathfrak{s}}=\alpha_{j}|_{\mathfrak{s}} whenever iji\sim j.

  2. (2)

    {αi|𝔰:iS/}\{\alpha_{i}|_{\mathfrak{s}}\;:\;i\in S/\!\!\sim\} is a linearly independent subset of 𝔰\mathfrak{s}^{*}.

Note that 𝔨\mathfrak{k} is one such subspace of 𝔥\mathfrak{h}. Denote the restriction map by p:𝔥𝔰p:\mathfrak{h}^{*}\rightarrow\mathfrak{s}^{*}. This map extends uniquely to a map (which we again denote by pp) from [[{eαi:iS}]]\mathbb{C}[[\{e^{-\alpha_{i}}:i\in S\}]] to [[{ep(αi):[i]S/}]]\mathbb{C}[[\{e^{-p(\alpha_{i})}:[i]\in S/\!\!\sim\}]]. Observe that the map pp is the same as the map π\pi when one identifies γ[i]\gamma_{[i]} with p(αi)p(\alpha_{i}). Therefore we have,

Corollary 2.

Let (λ1,I1)(\lambda_{1},I_{1}), (λ2,I2),,(λr,Ir)(\lambda_{2},I_{2}),\ldots,(\lambda_{r},I_{r}) and (μ1,J1)(\mu_{1},J_{1}), (μ2,J2),,(μr,Jr)𝒞¯(\mu_{2},J_{2}),\ldots,(\mu_{r},J_{r})\in\bar{\mathcal{C}} except that we now allow the Ik,JkI_{k},J_{k} to be empty. Then

p(ch𝔥(k=1rM(λk,Ik)))=p(ch𝔥(k=1rM(μk,Jk)))p(ch_{\mathfrak{h}}(\bigotimes_{k=1}^{r}M(\lambda_{k},I_{k})))=p(ch_{\mathfrak{h}}(\bigotimes_{k=1}^{r}M(\mu_{k},J_{k}))) (9)

if and only if

  1. (1)

    p(k=1rλk)=p(k=1rμkp(\sum_{k=1}^{r}\lambda_{k})=p(\sum_{k=1}^{r}\mu_{k})

  2. (2)

    σ𝔖r\exists\>\sigma\in\mathfrak{S}_{r} such that (λk,Ik)(μσ(k),Jσ(k))(\lambda_{k},I_{k})\approx(\mu_{\sigma(k)},J_{\sigma(k)}) for all kk.

Proof.

Rewriting (9) using the character formula for parabolic Verma modules, we get

k=1rp(eλk)p(U(λk,Ik))p(U(0,S))=k=1rp(eμk)p(U(μk,Jk))p(U(0,S))\prod_{k=1}^{r}p(e^{\lambda_{k}})\frac{p(U(\lambda_{k},I_{k}))}{p(U(0,S))}=\prod_{k=1}^{r}p(e^{\mu_{k}})\frac{p(U(\mu_{k},J_{k}))}{p(U(0,S))}

Comparing the highest weights on both sides of (9), we get p(k=1sλk)=p(k=1sμk)p(\sum_{k=1}^{s}\lambda_{k})=p(\sum_{k=1}^{s}\mu_{k}). Therefore,

k=1rp(U(λk,Ik))=k=1rp(U(μk,Jk))\prod_{k=1}^{r}p(U(\lambda_{k},I_{k}))=\prod_{k=1}^{r}p(U(\mu_{k},J_{k}))

We now proceed as in the proof of Corollary 1. Note that p(U(λ,I))=1p(U(\lambda,I))=1 iff I=I=\emptyset. Ignoring these trivial terms in the above product and relabelling

k=1sp(U(λk,Ik))=k=1tp(U(μk,Jk))\prod_{k=1}^{s}p(U(\lambda_{k},I_{k}))=\prod_{k=1}^{t}p(U(\mu_{k},J_{k}))

where, now {(λ1,I1),,(λs,Is),(μ1,J1),,(μt,Jt)}𝒞¯\{(\lambda_{1},I_{1}),\ldots,(\lambda_{s},I_{s}),(\mu_{1},J_{1}),\ldots,(\mu_{t},J_{t})\}\subseteq\bar{\mathcal{C}}. Taking logarithm on both sides of the above equation, and applying Theorem 4 we get s=ts=t and there is a permutation σ𝔖t\sigma\in\mathfrak{S}_{t} such that (λk,Ik)(μσk,Jσk)(\lambda_{k},I_{k})\approx(\mu_{\sigma k},J_{\sigma k}). The rest of the argument is exactly as in Corollary 1.

For the converse part, the second condition implies that U(λk,Ik)=U(μσ(k),Jσ(k))U(\lambda_{k},I_{k})=U(\mu_{\sigma(k)},J_{\sigma(k)}). Therefore we have

k=1rU(λk,Ik)U(0,S)=k=1rU(μk,Jk)U(0,S)\prod_{k=1}^{r}\frac{U(\lambda_{k},I_{k})}{U(0,S)}=\prod_{k=1}^{r}\frac{U(\mu_{k},J_{k})}{U(0,S)}

Now by applying pp on the both sides of above equation and multiplying ep(k=1rλk)e^{p(\sum_{k=1}^{r}\lambda_{k})} on the left hand side and multiplying ep(k=1rμk)e^{p(\sum_{k=1}^{r}\mu_{k})} on the right hand side of the equation gives us

k=1rp(eλk)p(U(λk,Ik))p(U(0,S))=k=1rp(eμk)p(U(μk,Jk))p(U(0,S)).\prod_{k=1}^{r}p(e^{\lambda_{k}})\frac{p(U(\lambda_{k},I_{k}))}{p(U(0,S))}=\prod_{k=1}^{r}p(e^{\mu_{k}})\frac{p(U(\mu_{k},J_{k}))}{p(U(0,S))}.

Now using the character formula, we conclude the result. ∎

6. Unique Factorization Of Restricted Parabolic Vermas

In this section, we will apply the results of the previous section to the special case of fixed point subalgebras of Dynkin diagram automorphisms and twisted graph automorphisms when 𝔤\mathfrak{g} is of untwisted affine type.

6.1. Graph automorphisms

Proposition 6.

Let G=(V(G),E(G))G=(V(G),E(G)) be a connected graph and Γ\Gamma be a subgroup of the group of all automorphisms of GG. Then there exists a connected subgraph of GG whose vertex set intersects every Γ\Gamma orbit in GG at exactly one point.

Proof.

Let 𝒜\mathcal{A} denote the set of all subsets of V(G)V(G) that intersect any Γ\Gamma orbit in GG in at most one point and whose induced subgraph is connected. Clearly 𝒜\mathcal{A} is non-empty because it contains all the singleton subsets of V(G)V(G).

Let M𝒜M\in\mathcal{A} be a maximal element with respect to the containment partial order. For any graph automorphism ωΓ\omega\in\Gamma we see that, ω(M)\omega(M) also belongs to 𝒜\mathcal{A}. Suppose that MM does not intersect a Γ\Gamma-orbit in GG. This means that

N:=ωΓω(M)V(G).N:=\bigcup\limits_{\omega\in\Gamma}\omega(M)\neq V(G).

But since GG is connected, there exist elements xNx\in N and yV(G)Ny\in V(G)-N such that (x,y)E(G)(x,y)\in E(G). But there exists some ωΓ\omega\in\Gamma for which ω(x)M\omega(x)\in M. Therefore we would have M{ω(y)}𝒜M\cup\{\omega(y)\}\in\mathcal{A} which is a contradiction to the assumption that MM was maximal. So, MM intersects every Γ\Gamma-orbit in GG. ∎

Any graph automorphism ω\omega of GG induces a Lie algebra automorphism of 𝔤\mathfrak{g} (which will be referred to as diagram automorphisms) described as follows: It maps the generators ei,hie_{i},h_{i} and fif_{i} to eωi,hωie_{\omega i},h_{\omega i} and fωif_{\omega i} respectively for all iSi\in S. This assignment extends uniquely to a Lie algebra automorphism of the derived subalgebra of 𝔤\mathfrak{g}. This map can be extended to an automorphism of 𝔤\mathfrak{g} in a unique way if we impose the condition that it preserves the standard invariant bilinear form and has order same as that of ω\omega. Such an automorphism preserves 𝔥\mathfrak{h} and its induced action on 𝔥\mathfrak{h}^{*} permutes the simple roots. See [3, §3.2].

Let Γ\Gamma be a subgroup of diagram automorphisms of 𝔤\mathfrak{g}. Denote by 𝔤Γ\mathfrak{g}^{\Gamma} (resp. 𝔥Γ\mathfrak{h}^{\Gamma}) the fixed point subalgebra of 𝔤\mathfrak{g} (resp. 𝔥\mathfrak{h}) with respect to Γ\Gamma.

Proposition 7.

Let AA be a GCM whose nullity is at most 11. Then for 𝔤=𝔤(A)\mathfrak{g}=\mathfrak{g}(A) we have,

𝔥Γ=wΓ,iSKer(αiαw(i))\mathfrak{h}^{\Gamma}=\bigcap_{w\in\Gamma,\;i\in S}Ker(\alpha_{i}-\alpha_{w(i)})

i.e., 𝔥Γ=𝔨\mathfrak{h}^{\Gamma}=\mathfrak{k} as in the notation of (8).

Proof.

It is easy to see that 𝔥Γ\mathfrak{h}^{\Gamma} is a subset of 𝔨\mathfrak{k}, since for all h𝔥h\in\mathfrak{h} and iSi\in S we have

αi(h)=αω(i)(ω(h))\alpha_{i}(h)=\alpha_{\omega(i)}(\omega(h))

for any graph automorphism ω\omega. Now one checks that the dimension of 𝔥Γ\mathfrak{h}^{\Gamma} is either the number of orbits of Γ\Gamma’s action on the Dynkin diagram or one more to it depending on the nullity of A being 0 or 11 (see the construction in [3, §3.2]. Basically, in the nullity 11 case one can find a d𝔥\span{αi:iS}d\in\mathfrak{h}\backslash\mathrm{span}\{\alpha_{i}^{\vee}\;:\;i\in S\} such that ω(d)=d\omega(d)=d). In both cases, it matches the dimension of 𝔨\mathfrak{k}. ∎

Example 1.

It can be checked that for the following graph, the conclusion of the above proposition is not true. Note that the nullity of the GCM associated to this graph is 22.

\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet112233445566778899

Define an equivalence relation on the set of nodes SS of the Dynkin diagram of 𝔤\mathfrak{g} as follows:

ijωΓ=Aut(G) such that ω(i)=ji\sim j\iff\;\exists\;\omega\in\Gamma=\text{Aut}(G)\text{ such that }\omega(i)=j

In the view of Proposition 6, for the equivalence relation \sim induced by Γ\Gamma, any connected subset of SS which is a union of equivalence classes is indeed equiconnected. Also, it is elementary to check that λ𝔥\lambda\in\mathfrak{h}^{*} is symmetric if and only if λ(ω(h))=λ(h)\lambda(\omega(h))=\lambda(h) for all ωΓ\omega\in\Gamma and h𝔥h\in\mathfrak{h}.

Corollary 3.

Let AA be a symmetrizable Generalised Cartan Matrix whose nullity is at most 11. Suppose (λ1,I1),,(λr,Ir),(μ1,J1),,(μr,Jr)𝒞¯(\lambda_{1},I_{1}),\ldots,(\lambda_{r},I_{r}),(\mu_{1},J_{1}),\ldots,(\mu_{r},J_{r})\in\bar{\mathcal{C}} with the exception that the IkI_{k} and JkJ_{k} could be empty. Then

k=1rch𝔥ΓM(λk,Ik)=k=1rch𝔥ΓM(μk,Jk)\prod_{k=1}^{r}ch_{\mathfrak{h}^{\Gamma}}M(\lambda_{k},I_{k})=\prod_{k=1}^{r}ch_{\mathfrak{h}^{\Gamma}}M(\mu_{k},J_{k})

if and only if

  1. (1)

    k=1rλk=k=1rμk\sum_{k=1}^{r}\lambda_{k}=\sum_{k=1}^{r}\mu_{k}

  2. (2)

    there exists σ𝔖r\sigma\in\mathfrak{S}_{r} such that (λk,Ik)(μσ(k),Jσ(k))(\lambda_{k},I_{k})\approx(\mu_{\sigma(k)},J_{\sigma(k)}).

Proof.

Part (2) follows from Proposition 7 and Corollary 2. We also have by the same that k=1rλk=k=1rμk\sum_{k=1}^{r}\lambda_{k}=\sum_{k=1}^{r}\mu_{k} when restricted to 𝔥Γ\mathfrak{h}^{\Gamma}. But since the λk\lambda_{k}’s and μk\mu_{k}’s are Γ\Gamma-invariant (in other words symmetric) we have k=1rλk=k=1rμk\sum_{k=1}^{r}\lambda_{k}=\sum_{k=1}^{r}\mu_{k} on the whole of 𝔥\mathfrak{h}. The converse part follows from the character formula  (3). ∎

In particular, Theorem 2 now follows from the above corollary.

6.2. Twisted graph automorphisms

Let 𝔤\mathfrak{g} be an untwisted affine Lie algebra. Then 𝔤\mathfrak{g} can be realized very explicitly as follows:

𝔤=𝔤0[t,t1]cd,\mathfrak{g}=\mathfrak{g}_{0}\otimes\mathbb{C}[t,t^{-1}]\oplus\mathbb{C}c\oplus\mathbb{C}d,

where 𝔤0\mathfrak{g}_{0} is the underlying finite-dimensional simple Lie algebra, cc is the central element and dd is the derivation. Let σ\sigma be a diagram automorphism of the underlying finite-dimensional simple algebra 𝔤0\mathfrak{g}_{0}. This induces an automorphism τ\tau of 𝔤\mathfrak{g} called the twisted diagram automorphism described as follows:

τ(xtk):=e2kπi/nσ(x)tk,x𝔤0,τ(c):=cτ(d):=d\tau(x\otimes t^{k}):=e^{-2k\pi i/n}\sigma(x)\otimes t^{k},x\in\mathfrak{g}_{0},\quad\quad\tau(c):=c\quad\quad\tau(d):=d

where nn is the order of σ\sigma. But σ\sigma also induces a diagram automorphism τ¯\bar{\tau} of 𝔤\mathfrak{g} which is obtained by fixing the affine node. That is,

τ¯(xtk):=σ(x)tkτ¯(c):=cτ¯(d):=d,\bar{\tau}(x\otimes t^{k}):=\sigma(x)\otimes t^{k}\quad\quad\bar{\tau}(c):=c\quad\quad\bar{\tau}(d):=d,

see [1, §9.5, §18.3 and §18.4] for more details.

It is well known that we can obtain the twisted affine Lie algebras from the untwisted ones as fixed point subalgebras of twisted diagram automorphisms (see [1, §18.4]). We thus have the following corollary concerning the tensor products of parabolic Verma modules of an untwisted affine Lie algebra restricted to the corresponding twisted affine algebra (obtained as a fixed point subalgebra):

Corollary 4.

Let 𝔤\mathfrak{g} be an untwisted affine Lie algebra. Let τ\tau be a twisted diagram automorphism of 𝔤\mathfrak{g} (and τ¯\bar{\tau} be the associated diagram automorphism). Suppose (λ1,I1),,(λr,Ir),(\lambda_{1},I_{1}),\ldots,(\lambda_{r},I_{r}), (μ1,J1),,(\mu_{1},J_{1}),\ldots, (μr,Jr)𝒞¯(\mu_{r},J_{r})\in\bar{\mathcal{C}} (with respect to τ¯\bar{\tau}) with the exception that the IkI_{k} and JkJ_{k} could be empty. We have

k=1rch𝔥τM(λk,Ik)=k=1rch𝔥τM(μk,Jk)\prod_{k=1}^{r}ch_{\mathfrak{h}^{\tau}}M(\lambda_{k},I_{k})=\prod_{k=1}^{r}ch_{\mathfrak{h}^{\tau}}M(\mu_{k},J_{k})

if and only if

  1. (1)

    k=1rλk=k=1rμk\sum_{k=1}^{r}\lambda_{k}=\sum_{k=1}^{r}\mu_{k}

  2. (2)

    there exists σ𝔖r\sigma\in\mathfrak{S}_{r} such that (λk,Ik)(μσ(k),Jσ(k))(\lambda_{k},I_{k})\approx(\mu_{\sigma(k)},J_{\sigma(k)}).

Proof.

The proof is immediate from Corollary 3, because τ=τ¯\tau=\bar{\tau} when restricted to 𝔥\mathfrak{h}. ∎

References

  • [1] R. W. Carter. Lie algebras of finite and affine type, volume 96 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2005.
  • [2] Gurbir Dhillon and Apoorva Khare. The weights of simple modules in category 𝒪\mathcal{O} for Kac-Moody algebras. J. Algebra, 603:164–200, 2022.
  • [3] Jürgen Fuchs, Bert Schellekens, and Christoph Schweigert. From Dynkin diagram symmetries to fixed point structures. Comm. Math. Phys., 180(1):39–97, 1996.
  • [4] James E. Humphreys. Representations of semisimple Lie algebras in the BGG category 𝒪\mathcal{O}, volume 94 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2008.
  • [5] Victor G. Kac. Infinite-dimensional Lie algebras. Cambridge University Press, Cambridge, third edition, 1990.
  • [6] Santosh Nadimpalli and Santosha Pattanayak. On uniqueness of branching to fixed point Lie subalgebras. Forum Math., 34(6):1663–1678, 2022.
  • [7] C. S. Rajan. Unique decomposition of tensor products of irreducible representations of simple algebraic groups. Ann. of Math. (2), 160(2):683–704, 2004.
  • [8] Shifra Reif and R. Venkatesh. On tensor products of irreducible integrable representations. J. Algebra, 592:402–423, 2022.
  • [9] R. Venkatesh and Sankaran Viswanath. Unique factorization of tensor products for Kac-Moody algebras. Adv. Math., 231(6):3162–3171, 2012.