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

\hypersetup

colorlinks=true, linkcolor=RoyalBlue, citecolor=RoyalBlue

On tensor products of representations of Lie superalgebras

Abhishek Das Department of mathematics & statistics
indian institute of technology, kanpur, u.p 208016.
[email protected]
 and  Santosha Pattanayak Department of mathematics & statistics, indian institute of technology, kanpur, u.p 208016
FB 577.
[email protected]
Abstract.

We consider typical finite dimensional complex irreducible representations of a basic classical simple Lie superalgebra, and give a sufficient condition on when unique factorization of finite tensor products of such representations hold. We also prove unique factorization of tensor products of singly atypical finite dimensional irreducible modules for 𝔰𝔩(m+1,n+1)\mathfrak{sl}(m+1,n+1), 𝔬𝔰𝔭(2,2n)\mathfrak{osp}(2,2n), G(3)G(3) and F(4)F(4) under an additional assumption. This result is a Lie superalgebra analogue of Rajan’s fundamental result  [MR2123935] on unique factorization of tensor products for finite dimensional complex simple Lie algebras.

Key words and phrases:
Lie superalgebras, typical representations, singly atypical weights, tensor products.
2020 Mathematics Subject Classification:
17B05, 17B10, 17B65

1. Introduction

In  [MR2123935], Rajan proved that for a complex simple Lie algebra 𝔤\mathfrak{g}, the following is true:

Theorem 1.1.

Let V1,,VmV_{1},\dots,V_{m} and W1,,WnW_{1},\dots,W_{n} be non trivial finite dimensional irreducible representations of 𝔤\mathfrak{g} such that

V1VmW1Wn.V_{1}\otimes\dots\otimes V_{m}\cong W_{1}\otimes\dots\otimes W_{n}.

Then m=nm=n, and there is a permutation σ\sigma of {1,,n}\{1,\dots,n\} such that VkWσ(k)V_{k}\cong W_{\sigma(k)} as 𝔤\mathfrak{g}-modules for every k{1,,n}k\in\{1,\dots,n\}.

Proof of the above theorem is based on judicious application of the Weyl character formula and proceeding via inductive arguments upon fixing one of the variables. Authors of  [MR2980495] extended this result for integrable representations of symmetrizable Kac-Moody algebras up to a one dimensional twist as these algebras admit non-trivial one-dimensional representations. Their proof relies on techniques involving formal logarithm, and the main idea being comparison of an appropriate monomial.

In  [MR4504111], for a complex semisimple Lie algebra 𝔤\mathfrak{g}, the authors consider a more general question of determining all the pairs (V1,V2)(V_{1},V_{2}) consisting of two finite dimensional irreducible representations of 𝔤\mathfrak{g} such that Res𝔤0V1Res𝔤0V2\text{Res}_{\mathfrak{g}_{0}}V_{1}\cong\text{Res}_{\mathfrak{g}_{0}}V_{2}, where 𝔤0\mathfrak{g}_{0} is the fixed point subalgebra of 𝔤\mathfrak{g} with respect to a finite order automorphism. In  [MR4343717], the authors gave a sufficient condition for the uniqueness of tensor products when 𝔤\mathfrak{g} is a Borcherds-Kac-Moody algebra.

Our goal of this paper is to obtain an analogous result when 𝔤\mathfrak{g} is a basic classical simple Lie superalgebra whose theory is most like that of simple Lie algebras. Nevertheless, there are certain subtle differences that have resulted in many aspects of their representation theory only partly explored. For example, any finite dimensional representation of 𝔤\mathfrak{g} is completely reducible if and only if 𝔤=𝔬𝔰𝔭(1,2n)\mathfrak{g}=\mathfrak{osp}(1,2n). Kac showed that there are two disjoint classes of finite-dimensional irreducible representations of any basic classical Lie superalgebra which he gave appellations as typical and atypical. Typical representations share many properties in common with finite-dimensional representations of simple Lie algebras. In particular, they can be built up explicitly by an induced module construction that further allows a straightforward determination of their characters and dimensions. On the other hand, the situation with atypical representations is far more complicated and they are still not well understood. A serious difficulty being unlike the typical case, here an irreducible representation is not determined by its central character. The only class of atypical representations that is by so far tractable is the class of singly atypical representations. So in this note we restrict ourselves to the class of finite dimensional typical and singly atypical representations.

We first consider typical irreducible finite dimensional representations of 𝔤\mathfrak{g} as they admit a Weyl-Kac character formula. We give a sufficient condition on when unique factorization of finite tensor products of such representations hold. We follow techniques involving formal logarithms as developed in  [MR2980495]. The main strategy is to interpret the given tensor products in terms of products of normalized Weyl numerators. For any typical irreducible finite dimensional representation, we show that, the normalized Weyl numerator factors in accordance with the number of connected components of the set of even simple roots of 𝔤\mathfrak{g} and the factorization doesn’t depend on the highest weight that defines the representation. We then apply logarithm on both sides and compare an appropriate monomial that carries all crucial informations of the tensor products. Our result (\autorefThm:tnsrpdt) can be described as follows:

Let V(λ1),,V(λr),V(μ1),,V(μs)V(\lambda_{1}),\dots,V(\lambda_{r}),V(\mu_{1}),\dots,V(\mu_{s}) be typical irreducible finite dimensional representations of a basic classical simple Lie superalgebra 𝔤\mathfrak{g} such that the following isomorphism of 𝔤\mathfrak{g}-modules holds,

V(λ1)V(λr)V(μ1)V(μs).V(\lambda_{1})\otimes\dots\otimes V(\lambda_{r})\cong V(\mu_{1})\otimes\dots\otimes V(\mu_{s}).

Then r=sr=s, i.e. the number of tensor constituents on both sides is same. We prove that under some additional hypothesis, each constituent in the left hand side of the above isomorphism is 𝔤\mathfrak{g}-module isomorphic to some constituent in the right hand side, i.e. unique factorization of tensor products holds in this case (see \autorefThm:tnsrpdt for details). As a corollary, we show that unique factorization of tensor products of irreducible finite dimensional typical representations holds for superalgebras of types A(n,0)A(n,0), B(0,n)B(0,n), CnC_{n}, F(4)F(4) and G(3)G(3) (cf. \autorefunique-coro). To demonstrate the requirement of the hypothesis in \autorefThm:tnsrpdt, we refer to \autorefexample.

For 𝔰𝔩(m+1,n+1)\mathfrak{sl}(m+1,n+1), 𝔬𝔰𝔭(2,2n)\mathfrak{osp}(2,2n), G(3)G(3), and F(4)F(4), character formulae for singly atypical finite dimensional irreducible modules are known in the literature (see \autorefsec atyp for details). We also addressed the same question of unique factorization of tensor products of such modules and found its answer in affirmative under an additional assumption on the highest weights. We state our result (\autorefThm:atyp) below:

Let 𝔤\mathfrak{g} be any Lie superalgebra among 𝔰𝔩(m+1,n+1)\mathfrak{sl}(m+1,n+1), 𝔬𝔰𝔭(2,2n)\mathfrak{osp}(2,2n), G(3)G(3) and F(4)F(4). Suppose we are given the following isomorphism of 𝔤\mathfrak{g}-modules

V(ν1)V(νr)V(μ1)V(μs),V(\nu_{1})\otimes\dots\otimes V(\nu_{r})\cong V(\mu_{1})\otimes\dots\otimes V(\mu_{s}),

where each νi\nu_{i}’s and μj\mu_{j}’s are dominant integral singly atypical of type γ\gamma. Then r=sr=s, and there is a permutation σ\sigma of {1,,r}\{1,\dots,r\} such that V(νk)V(μσ(k))V(\nu_{k})\cong V(\mu_{\sigma(k)}) for all kk.

We briefly outline the contents of each section. \autorefsecprelims contains preliminaries on Lie superalgebras. In \autoreflemmas, we prove some preliminary lemmas which are crucial for the proof of the main theorem. \autorefsecfacto contains results involving formal logarithm and analysis of normalized Weyl numerators. In \autorefsecmainthm we provide a proof of our main theorem, and finally in \autorefsec atyp, we establish unique decomposition of tensor products for singly atypical finite dimensional irreducible modules for the Lie superalgebras 𝔰𝔩(m+1,n+1)\mathfrak{sl}(m+1,n+1), 𝔬𝔰𝔭(2,2n)\mathfrak{osp}(2,2n), G(3)G(3) and F(4)F(4) under an additional assumption on the weights.

Notation.

Throughout this paper we work over the field of complex numbers \mathbb{C}. All modules and algebras are defined over \mathbb{C} and in addition all the modules are of finite dimension. We write 2={0,1}\mathbb{Z}_{2}=\{0,1\} and use its standard field structure.

2. Preliminaries

In this section we recall few basic definitions and results pertaining to Lie superalgebras and their representations. We mostly follow the notations of  [MR3012224] and  [MR2906817]. To begin with, a superalgebra AA is a 2\mathbb{Z}_{2}-graded vector space A=A0A1A=A_{0}\oplus A_{1} together with a bilinear multiplication satisfying AiAjAi+jA_{i}A_{j}\subseteq A_{i+j}, for i,j2i,j\in\mathbb{Z}_{2}. Degree of a homogeneous element, say aa, is denoted by |a||a|.

2.1.

A Lie superalgebra 𝔤=𝔤0𝔤1\mathfrak{g}=\mathfrak{g}_{0}\oplus\mathfrak{g}_{1} is a superalgebra equipped with a bilinear product [,][\cdot\,,\cdot] (bracket) satisfying the following two axioms: for homogeneous elements a,b,c𝔤a,b,c\in\mathfrak{g},

Skew-supersymmetry:

[a,b]=(1)|a||b|[b,a][a,b]=-(-1)^{|a||b|}[b,a].

Super Jacobi identity:

[a,[b,c]]=[[a,b],c]+(1)|a||b|[b,[a,c]][a,[b,c]]=[[a,b],c]+(-1)^{|a||b|}[b,[a,c]].

The direct summands 𝔤0\mathfrak{g}_{0} (resp. 𝔤1\mathfrak{g}_{1}) are called even (resp. odd) part of 𝔤\mathfrak{g}, and from the definition it follows that, 𝔤0\mathfrak{g}_{0} is a Lie algebra and 𝔤1\mathfrak{g}_{1} is a 𝔤0\mathfrak{g}_{0}-module.

2.2.

A natural analogue of the ordinary simple Lie algebras in the super world are the basic classical Lie superalgebras. In particular, they can be described (with the exception of A(1,1)=𝔭𝔰𝔩(2,2)A(1,1)=\mathfrak{psl}(2,2)) in terms of a Cartan matrix and generalized root systems. They are defined as follows:

A Lie superalgebra 𝔤=𝔤0𝔤1\mathfrak{g}=\mathfrak{g}_{0}\oplus\mathfrak{g}_{1} is called basic classical if it satisfies the following conditions:

  1. a)

    𝔤\mathfrak{g} is simple,

  2. b)

    the Lie algebra 𝔤0\mathfrak{g}_{0} is a reductive subalgebra of 𝔤\mathfrak{g},

  3. c)

    there exists a nondegenerate invariant even supersymmetric bilinear form on 𝔤\mathfrak{g}.

Kac (see  [MR0519631], Proposition 1.1) proved that the complete list of basic classical Lie superalgebras, which are not Lie algebras, consists of Lie superalgebras of the type A(m,n)A(m,n), B(m,n)B(m,n), C(n),D(m,n),F(4),G(3),D(2,1,α)C(n),D(m,n),F(4),G(3),D(2,1,\alpha). We note that the even part of a basic classical Lie superalgebra 𝔤\mathfrak{g} is a reductive Lie algebra.

2.3.

Following  [MR3012224], a Cartan subalgebra 𝔥\mathfrak{h} of 𝔤\mathfrak{g} is defined to be a Cartan subalgebra of the even part 𝔤0\mathfrak{g}_{0} and the Weyl group 𝒲\mathscr{W} of 𝔤\mathfrak{g} is simply defined to be the Weyl group of the Lie algebra 𝔤0\mathfrak{g}_{0}. We can choose a non-degenerate even invariant supersymmetric bilinear form (.|.)(.|.) on 𝔤\mathfrak{g} such that its restriction to 𝔥×𝔥\mathfrak{h}\times\mathfrak{h} is non-degenerate and 𝒲\mathscr{W}-invariant. We pull back this non-degenerate bilinear form on 𝔥\mathfrak{h} to get a non-degenerate bilinear form (.,.)(.,.) on 𝔥\mathfrak{h}^{*}.

Let 𝔥\mathfrak{h} be a Cartan subalgebra of 𝔤\mathfrak{g}. For α𝔥\alpha\in\mathfrak{h}^{*}, let

𝔤α{x𝔤:[h,x]=α(h)x,h𝔥}.\displaystyle\mathfrak{g}_{\alpha}\coloneqq\{x\in\mathfrak{g}\colon[h,x]=\alpha(h)x,\forall h\in\mathfrak{h}\}.
The root system for 𝔤\mathfrak{g} is defined to be
Φ{α𝔥:𝔤α0,α0}.\displaystyle\Phi\coloneqq\{\alpha\in\mathfrak{h}^{*}\colon\mathfrak{g}_{\alpha}\not=0,\alpha\not=0\}.
Define the sets of even and odd roots, respectively, to be
Φ0{αΦ:𝔤α𝔤00},Φ1{αΦ:𝔤α𝔤10}.\displaystyle\Phi_{0}\coloneqq\{\alpha\in\Phi\colon\mathfrak{g}_{\alpha}\cap\mathfrak{g}_{0}\not=0\},\quad\Phi_{1}\coloneqq\{\alpha\in\Phi\colon\mathfrak{g}_{\alpha}\cap\mathfrak{g}_{1}\not=0\}.

For a root αΦ\alpha\in\Phi we have kαΦk\alpha\in\Phi for an integer k±1k\neq\pm 1 if and only if αΦ1\alpha\in\Phi_{1} and (α,α)0(\alpha,\alpha)\neq 0; in this case k=±2k=\pm 2. A root α\alpha is called isotropic if (α,α)=0(\alpha,\alpha)=0. For any root α\alpha, the nondegenerate form on 𝔥\mathfrak{h} gives rise to a unique element hα𝔥h_{\alpha}\in\mathfrak{h}, called the coroot corresponding to α\alpha, such that α(h)=(hα,h)\alpha(h)=(h_{\alpha},h) for every h𝔥h\in\mathfrak{h}. Let EE be the real vector space spanned by Φ\Phi. Then 𝔥=E\mathfrak{h}^{*}=E\otimes_{\mathbb{R}}\mathbb{C} for all basic classical simple Lie superalgebras. We fix a total ordering on \leqslant on EE compatible with the real vector space structure: v1w1v_{1}\leqslant w_{1} and v2w2v_{2}\leqslant w_{2} imply that v1+v2w1+w2v_{1}+v_{2}\leqslant w_{1}+w_{2}, v2v1-v_{2}\leqslant-v_{1}, and for any positive real number cc, cv1cv2cv_{1}\leqslant cv_{2} for all vi,wjEv_{i},w_{j}\in E. We fix such a total order and denote by Φ+\Phi^{+} (resp. Φ\Phi^{-}) the subsets of roots αΦ\alpha\in\Phi such that 0<α0<\alpha (resp. α<0\alpha<0). Φ+\Phi^{+} is called a positive system. The corresponding set of positive even roots is denoted by Φ0+\Phi_{0}^{+}. For further details we refer to Section 1.3.1 of  [MR3012224].

Lie superalgebras of Type I contains a one dimensional center, say span{z}\text{span}\{z\}, of 𝔤0\mathfrak{g}_{0}. In these cases 𝔥=𝔥1span{z}\mathfrak{h}=\mathfrak{h}_{1}\oplus\text{span}\{z\}, where 𝔥1\mathfrak{h}_{1}^{*} is the span of even roots. We extend an element λ\lambda in the span of even roots to 𝔥\mathfrak{h}^{*} by defining λ(z)\lambda(z) to be 0.

2.4.

Simple roots are defined in the same way as in the Lie algebra case; but here not all simple systems are conjugate under 𝒲\mathscr{W} action due to presence of odd roots. Any element of 𝒲\mathscr{W} is a product of simple reflections corresponding to simple even roots. A simple system containing least number of isotropic roots is called distinguished or standard. Dynkin diagram corresponding to the standard simple system is called standard Dynkin diagram. For each basic classical simple Lie superalgebra, we can choose a distinguished system of simple roots containing only one isotropic root; this is possible for any basic classical Lie superalgebra except B(0,n)B(0,n), which has no isotropic roots (see  [MR0486011]). So, for each basic classical simple Lie superalgebra, we fix such a standard simple system Π\Pi and let Π0\Pi_{0}, Π1\Pi_{1} denote the set of even (resp. odd) simple roots in Π\Pi.

2.5.

We have a triangular decomposition 𝔤=𝔫𝔥𝔫+\mathfrak{g}=\mathfrak{n}^{-}\oplus\mathfrak{h}\oplus\mathfrak{n}^{+}, where 𝔫±=αΦ±𝔤α\mathfrak{n}^{\pm}=\oplus_{\alpha\in\Phi^{\pm}}\mathfrak{g}_{\alpha}. Let’s recall some basic facts about the representation theory of basic classical simple Lie superalgebras. Let 𝔤\mathfrak{g} be any such Lie superalgebra. For any λ𝔥\lambda\in\mathfrak{h}^{*}, there is an irreducible (unique up to isomorphism) highest weight 𝔤\mathfrak{g}-module of highest weight λ\lambda. We denote this module by V(λ)V(\lambda). The weight λ\lambda is called dominant integral if V(λ)V(\lambda) is finite dimensional (see  [MR3751124]*page 141). It is well known that any finite dimensional irreducible representation of 𝔤\mathfrak{g} is of the form V(λ)V(\lambda) for some dominant integral weight λ\lambda. Note that if λ𝔥\lambda\in\mathfrak{h}^{*} is dominant integral, then necessarily we have 2(λ,α)(α,α)0\smash{\frac{2(\lambda,\alpha)}{\phantom{2}(\alpha,\alpha)}}\in\mathbb{Z}_{\geqslant 0} for all αΠ0\alpha\in\Pi_{0}  [MR0519631]*Proposition 2.3; and we denote this integer by λ,α\langle\lambda,\alpha\rangle.

Remark 1.

If 𝔤\mathfrak{g} is of Type I, i.e of type A(m,n)A(m,n) and CnC_{n}, then for a weight λ𝔥\lambda\in\mathfrak{h}^{*}, the condition that λ,α0\langle\lambda,\alpha\rangle\in\mathbb{Z}_{\geqslant 0} is also sufficient for being dominant integral (see  [MR3751124] page 132).

2.6.

The Weyl vector ρ𝔥\rho\in\mathfrak{h}^{*} is defined by:

ρ12αΦ0+α12αΦ1+α,\displaystyle\rho\coloneqq\frac{1}{2}\sum_{\alpha\in\Phi^{+}_{0}}\alpha-\frac{1}{2}\sum_{\alpha\in\Phi^{+}_{1}}\alpha,
and for any positive simple root β\beta, it satisfies (see  [MR3012224], Proposition 1.33)
(ρ,β)=12(β,β).\displaystyle(\rho,\beta)=\tfrac{1}{2}(\beta,\beta). (2.1)

In particular, (ρ,β)=0(\rho,\beta)=0 if β\beta is isotropic.

A weight λ𝔥\lambda\in\mathfrak{h}^{*} is said to be typical if (λ+ρ,α)0(\lambda+\rho,\alpha)\not=0 for all isotropic roots αΦ1+\alpha\in\Phi_{1}^{+}, and it is called atypical otherwise. A representation associated to a typical weight is called a typical representation.

2.7.

Let V(λ)V(\lambda) be the finite dimensional irreducible highest weight 𝔤\mathfrak{g}-module of highest weight λ\lambda. It admits a weight space decomposition: V(λ)=μ𝔥V(λ)μV(\lambda)=\oplus_{\mu\in\mathfrak{h}^{*}}V(\lambda)_{\mu}, where V(λ)μV(\lambda)_{\mu} is the weight space corresponding to the weight μ\mu. The formal character of V(λ)V(\lambda) is defined by:

chV(λ)\displaystyle\operatorname{\textnormal{ch}}V(\lambda) μ𝔥dimV(λ)μeμ.\displaystyle\coloneqq\sum_{\mu\in\mathfrak{h}^{*}}\dim V(\lambda)_{\mu}e^{\mu}.

A weight space V(λ)μV(\lambda)_{\mu} is zero unless μ=λαΦ+nαα,nα0\mu=\lambda-\sum_{\alpha\in\Phi^{+}}n_{\alpha}\alpha,\ n_{\alpha}\in\mathbb{Z}_{\geqslant 0}  [MR3012224]*Section 1.5.3. We note that the finite dimensional irreducible representations of 𝔤\mathfrak{g} are completely determined by their characters (cf.  [MR2776360]*Proposition 4.2).

2.8.

If λ\lambda is a typical dominant weight, then we have the Weyl Kac character formula (see  [MR0519631]) for V(λ)V(\lambda) given by:

chV(λ)=D1D2w𝒲(1)(w)ew(λ+ρ),\displaystyle\operatorname{\textnormal{ch}}V(\lambda)=\frac{D_{1}}{D_{2}}\sum_{w\in\mathscr{W}}(-1)^{\ell(w)}e^{w(\lambda+\rho)}, (2.2)

where D1=αΦ1+(eα/2+eα/2)D_{1}=\prod_{\alpha\in\Phi^{+}_{1}}(e^{\alpha/2}+e^{-\alpha/2}) and D2=αΦ0+(eα/2eα/2)D_{2}=\prod_{\alpha\in\Phi^{+}_{0}}(e^{\alpha/2}-e^{-\alpha/2}). By definition of ρ\rho, the expression D1/D2D_{1}/D_{2} can also be written as:

D1D2=eραΦ1+(1+eα)αΦ0+(1eα).\displaystyle{}\frac{D_{1}}{D_{2}}=e^{-\rho}\frac{\prod_{\alpha\in\Phi^{+}_{1}}(1+e^{-\alpha})}{\prod_{\alpha\in\Phi^{+}_{0}}(1-e^{-\alpha})}. (2.3)

We now define the normalized character χλ\chi_{\lambda} and the normalized Weyl numerator U(λ)U(\lambda) of V(λ)V(\lambda) respectively by:

χλeλchV(λ),\displaystyle\chi_{\lambda}\coloneqq e^{-\lambda}\operatorname{\textnormal{ch}}V(\lambda), (2.4)
U(λ)e(λ+ρ)w𝒲(1)(w)ew(λ+ρ).\displaystyle U(\lambda)\coloneqq e^{-(\lambda+\rho)}\sum_{w\in\mathscr{W}}(-1)^{\ell(w)}e^{w(\lambda+\rho)}. (2.5)
By Weyl Kac character formula,
DU(λ)=eλchV(λ)=χλ,\displaystyle D\cdot U(\lambda)=e^{-\lambda}\operatorname{\textnormal{ch}}V(\lambda)=\chi_{\lambda}, (2.6)

where D=eρD1/D2D=e^{\rho}D_{1}/D_{2}. (cf. Equation (2.3)).

3. Preparatory Lemmas

In this section we prove some preliminary lemmas which are needed for the proof of the main theorem.

Lemma 3.1.

Let λ\lambda be a typical dominant integral weight and w𝒲w\in\mathscr{W} be arbitrary. Then λ+ρw(λ+ρ)\lambda+\rho-w(\lambda+\rho) is a sum of positive even roots with non negative integral coefficients.

Proof.

Put λ+ρ=η\lambda+\rho=\eta. If (w)=1\ell(w)=1, say w=sαw=s_{\alpha}, the simple reflection corresponding to the positive even simple root α\alpha; then ηsαη=η,αα\eta-s_{\alpha}\eta=\langle\eta,\alpha\rangle\alpha. As λ\lambda is dominant, Equation (2.1) shows that η,α\langle\eta,\alpha\rangle is positive. Consequently the claim is true in this case. Take a reduced expression of ww and write w=usβ with (u)=(w)1w=us_{\beta}\text{ with }\ell(u)=\ell(w)-1. This implies u(β)u(\beta) is positive even root as the set of even roots is 𝒲\mathscr{W}-invariant. Therefore by induction on (w)\ell(w), ηwη=ηusβη=ηuη+η,βu(β)\eta-w\eta=\eta-us_{\beta}\eta=\eta-u\eta+\langle\eta,\beta\rangle u(\beta) is also a sum of positive even roots with non negative integral coefficients. ∎

For a typical dominant integral weight λ\lambda of 𝔤\mathfrak{g}, we put aαλ+ρ,αa_{\alpha}\coloneqq\langle\lambda+\rho,\alpha\rangle for each even simple root α\alpha. Equation (2.1) shows that aαa_{\alpha} is a positive integer. By \autorefsopr, for w𝒲w\in\mathscr{W}, we can write

λ+ρw(λ+ρ)=αΠ0cα(w)α,where each cα(w)0.\lambda+\rho-w(\lambda+\rho)=\sum_{\alpha\in\Pi_{0}}c_{\alpha}(w)\alpha,\quad\text{where each }c_{\alpha}(w)\in\mathbb{Z}_{\geqslant 0}.

We set X(w,λ)αΠ0Xαcα(w)=ew(λ+ρ)(λ+ρ)X(w,\lambda)\coloneqq\prod_{\alpha\in\Pi_{0}}X_{\alpha}^{c_{\alpha}(w)}=e^{w(\lambda+\rho)-(\lambda+\rho)}. Then we have

U(λ)=w𝒲(1)(w)X(w,λ)U(\lambda)=\sum_{w\in\mathscr{W}}(-1)^{\ell(w)}X(w,\lambda) (3.1)

3.1.

We now recall few definitions from  [MR2980495]. The underlying graph 𝒢\mathcal{G} of 𝔤\mathfrak{g} is defined to be the graph with vertex set Π\Pi: two vertices α\alpha and β\beta are joined by an edge iff (α,β)0(\alpha,\beta)\not=0. For any subset CC of the vertex set, the subgraph spanned by CC is just the graph having CC as the vertex set. A nonempty subset KΠ0K\subset\Pi_{0} is called totally disconnected if it comprises of simple roots that are all mutually orthogonal: (α,β)=0(\alpha,\beta)=0 for every distinct α,βK\alpha,\beta\in K. For any w𝒲w\in\mathscr{W}, we take a reduced expression, say w~\tilde{w} of ww. Let I(w)I(w) be the set defined by I(w){αΠ0:sα appears in w~}I(w)\coloneqq\{\alpha\in\Pi_{0}\colon s_{\alpha}\text{ appears in }\tilde{w}\}. This is a well defined subset of Π0\Pi_{0}, (see  [MR1066460]). Let {1w𝒲:I(w) is totally disconnected}\mathcal{I}\coloneqq\{1\not=w\in\mathscr{W}\colon I(w)\text{ is totally disconnected}\}. Given a totally disconnected subset KΠK\in\Pi, there is a unique element w(K)w(K)\in\mathcal{I} such that I(w(K))=KI(w(K))=K; w(K)w(K) is precisely the product of the commuting simple reflections {sα:αK}\{s_{\alpha}\colon\alpha\in K\}. This establishes a natural bijection between \mathcal{I} and the set of all totally disconnected subsets of Π0\Pi_{0}. Proof of the following lemma follows the same line of argument as in  [MR2980495]*Lemma 2.

Lemma 3.2.

For w𝒲w\in\mathscr{W} we have

  1. (a)

    I(w)={αΠ0:cα(w)0}I(w)=\{\alpha\in\Pi_{0}\colon c_{\alpha}(w)\not=0\}, i.e X(w,λ)=αI(w)Xαcα(w)X(w,\lambda)=\prod_{\alpha\in I(w)}X_{\alpha}^{c_{\alpha}(w)}.

  2. (b)

    For every αI(w),cα(w)aα\alpha\in I(w),c_{\alpha}(w)\geqslant a_{\alpha}.

  3. (c)

    If ww\in\mathcal{I}, then cα(w)=aαc_{\alpha}(w)=a_{\alpha} for every αI(w)\alpha\in I(w).

  4. (d)

    If ww is not in {1}\mathcal{I}\cup\{1\}, then there exists βI(w)\beta\in I(w) such that cβ>aβc_{\beta}>a_{\beta}.

We recall the following definition from  [MR2980495] which will be used in the next section.

Definition 3.3.

Let kk be a positive integer. A kk-partition 𝒥\mathcal{J} of the graph 𝒢\mathcal{G} is an ordered kk-tuple 𝒥(J1,,Jk)\mathcal{J}\coloneqq(J_{1},\dots,J_{k}) such that each JiJ_{i} is a nonempty totally disconnected subset of the vertex set Π\Pi; JiJj=J_{i}\cap J_{j}=\emptyset for iji\not=j, and lastly; i=1kJi=Π\bigcup_{i=1}^{k}J_{i}=\Pi. For each such partition, we define w(𝒥)w(J1)w(Jk)𝒲w(\mathcal{J})\coloneqq w(J_{1})\cdots w(J_{k})\in\mathscr{W}.

Denote by Pk(𝒢)P_{k}(\mathcal{G}) the set of all kk-partitions of 𝒢\mathcal{G} and put ck(𝒢)|Pk(𝒢)|c_{k}(\mathcal{G})\coloneqq|P_{k}(\mathcal{G})|, the cardinality of Pk(𝒢)P_{k}(\mathcal{G}). We also denote by k(𝒢)k(\mathcal{G}) the number (1)|𝒢|k=1|𝒢|(1)kck(𝒢)k(-1)^{|\mathcal{G}|}\sum_{k=1}^{|\mathcal{G}|}(-1)^{k}\frac{c_{k}(\mathcal{G})}{k}. Let the symbol XαX_{\alpha} denote eαe^{-\alpha} for αΠ0\alpha\in\Pi_{0}, and consider the algebra of formal power series 𝒜[[Xα:αΠ0]]\mathcal{A}\coloneqq\mathbb{C}[[X_{\alpha}\colon\alpha\in\Pi_{0}]]. \autorefsopr shows that U(λ)𝒜U(\lambda)\in\mathcal{A} with constant term 1 corresponding to w=1w=1. The formal logarithm for any element ζ𝒜\zeta\in\mathcal{A} with constant term 1 is defined by:

logζk1(1ζ)k/k.\log\zeta\coloneqq-\sum_{k\geqslant 1}(1-\zeta)^{k}/k.

The following lemma will be used in the proof of the main theorem.

Lemma 3.4.

Let 𝔤\mathfrak{g} be a basic classical simple Lie superalgebra and λ,μ𝔥\lambda,\mu\in\mathfrak{h}^{*} be typical dominant integral weights. Then the following statements are equivalent.
(a) χλ=χμ\chi_{\lambda}=\chi_{\mu}, (b) U(λ)=U(μ)U(\lambda)=U(\mu), (c) λ=μ\lambda=\mu, (d) V(λ)V(μ)V(\lambda)\cong V(\mu), (e) chV(λ)=chV(μ)\operatorname{\textnormal{ch}}V(\lambda)=\operatorname{\textnormal{ch}}V(\mu).

Proof.

The equivalence of (a)(a) and (b)(b) directly follows from Equation (2.6). Assume (b). Put λ+ρ=η\lambda+\rho=\eta. For a simple reflection w=sαw=s_{\alpha} corresponding to the even simple root α\alpha we have

U(λ)\displaystyle U(\lambda) =esαηη\displaystyle=-e^{s_{\alpha}\eta-\eta} +sαw𝒲(1)(w)ewηη\displaystyle+\sum_{s_{\alpha}\not=w\in\mathscr{W}}(-1)^{\ell(w)}e^{w\eta-\eta}\phantom{.}
=Xαη,α\displaystyle=-X_{\alpha}^{\langle\eta,\alpha\rangle} +sαw𝒲(1)(w)ewηη.\displaystyle+\sum_{s_{\alpha}\not=w\in\mathscr{W}}(-1)^{\ell(w)}e^{w\eta-\eta}.

Since no monomial of the form XαmX_{\alpha}^{m} appears in the remaining summation, it follows that for any even simple root β\beta we have,

Xαλ+ρ,β=Xαμ+ρ,β.X_{\alpha}^{\langle\lambda+\rho,\beta\rangle}=X_{\alpha}^{\langle\mu+\rho,\beta\rangle}.

This implies that λ+ρ,β=μ+ρ,β\langle\lambda+\rho,\beta\rangle=\langle\mu+\rho,\beta\rangle and hence we have (λ,β)=(μ,β)(\lambda,\beta)=(\mu,\beta) for all even simple root β\beta. As the restriction of the invariant form of 𝔤\mathfrak{g} on 𝔥\mathfrak{h} is non degenerate (𝔤\mathfrak{g} is basic), this implies that λ=μ\lambda=\mu in the subspace spanned by the even simple roots. Since the action of λ,μ\lambda,\mu is defined to be zero on the center of 𝔤0\mathfrak{g}_{0}, which is one-dimensional for types A(m,n)A(m,n) and CnC_{n}, it follows that λ=μ\lambda=\mu in 𝔥\mathfrak{h}^{*}.

To prove (c) implies (b), we observe that the monomial X(w,λ)X(w,\lambda) appearing in the expression of U(λ)U(\lambda) in (3.1) is a product of Xαcα(w)X_{\alpha}^{c_{\alpha}(w)}, αΠ0\alpha\in\Pi_{0}. By the proof of \autorefsopr, the exponent cα(w)c_{\alpha}(w) of XαX_{\alpha} depends only on the integers λ+ρ,γ\langle\lambda+\rho,\gamma\rangle where γΠ0\gamma\in\Pi_{0} is arbitrary. Therefore, equality of λ\lambda and μ\mu would imply equality of all such exponents. In other words, we have X(w,λ)=X(w,μ)X(w,\lambda)=X(w,\mu) for every w𝒲w\in\mathscr{W}; and consequently, U(λ)=U(μ)U(\lambda)=U(\mu). The equivalence of (c) and (d), and (d) and (e) follow from  [MR0519631]*Proposition 2.2 and  [MR2776360]*Proposition 4.2 respectively. ∎

4. Factorization of \texorpdfstringU(λ)U(\lambda)lg

4.1.

Consider the element λ+ρw(λ+ρ)\lambda+\rho-w(\lambda+\rho) for any w𝒲w\in\mathscr{W}. By \autorefsopr, this element can be written as:

λ+ρw(λ+ρ)=αΠ0cα(w)α.cα(w)0\lambda+\rho-w(\lambda+\rho)=\sum_{\alpha\in\Pi_{0}}c_{\alpha}(w)\alpha.\quad c_{\alpha}(w)\in\mathbb{Z}_{\geqslant 0}

By definition and Equation (3.1), we have

U(λ)=w𝒲(1)(w)X(w,λ)=1(1w𝒲(1)(w)X(w,λ))=1ζ,\displaystyle U(\lambda)=\sum_{w\in\mathscr{W}}(-1)^{\ell(w)}X(w,\lambda)=1-\left(-\sum_{1\not=w\in\mathscr{W}}(-1)^{\ell(w)}X(w,\lambda)\right)\ =1-\zeta, (4.1)

where ζ\zeta stands for the term in parenthesis. Therefore, logU(λ)=k1ζk/k-\log U(\lambda)=\sum_{k\geqslant 1}\zeta^{k}/k and no monomial in this expansion can include an odd root in its support. In other words, we have to work only with the set of even simple roots Π0\Pi_{0}. Note that the subgraph spanned by Π0\Pi_{0} is not always connected; in fact, it is connected only for types A(n,0)A(n,0), B(0,n)B(0,n), CnC_{n}, F(4)F(4), G(3)G(3); and for other types, it is union of two connected components.

4.2.

We now show that U(λ)U(\lambda) factors in accordance with the number of connected components of Π0\Pi_{0}. Recall that the Weyl group 𝒲\mathscr{W} of 𝔤\mathfrak{g} is defined to be the Weyl group of the even part 𝔤0\mathfrak{g}_{0}. If Π0\Pi_{0} is union of two connected components, say Π0=C1C2\Pi_{0}=C_{1}\cup C_{2}, then 𝒲\mathscr{W} will be a direct product 𝒲=𝒲1×𝒲2\mathscr{W}=\mathscr{W}_{1}\times\mathscr{W}_{2} of two subgroups where 𝒲i\mathscr{W}_{i} is the group generated by simple reflections {sαi:αCi}\{s_{\alpha_{i}}\colon\alpha\in C_{i}\} for i=1,2i=1,2.

Proposition 4.1.

Put η=λ+ρ\eta=\lambda+\rho. Then with the above notation we have

U(λ)=(u𝒲1(1)(u)euηη)(v𝒲2(1)(v)evηη)\displaystyle U(\lambda)=\left(\sum_{u\in\mathscr{W}_{1}}(-1)^{\ell(u)}e^{u\eta-\eta}\right)\cdot\left(\sum_{v\in\mathscr{W}_{2}}(-1)^{\ell(v)}e^{v\eta-\eta}\right) (4.2)
Proof.

Let C1,C2C_{1},C_{2} be the two connected components of Π0\Pi_{0} and let 𝒲i\mathscr{W}_{i} be the group generated by the simple reflections in CiC_{i} for i=1,2i=1,2. Since the roots belonging to C1C_{1} are mutually orthogonal to those in C2C_{2}, every w𝒲w\in\mathscr{W} can be written uniquely as w=uvw=uv, for some u𝒲1,v𝒲2u\in\mathscr{W}_{1},v\in\mathscr{W}_{2}. Therefore, to prove the proposition, it suffices to show that ewηη=euηηevηηe^{w\eta-\eta}=e^{u\eta-\eta}\cdot e^{v\eta-\eta}. Let uηη=αC1cα(u)αu\eta-\eta=\sum_{\alpha\in C_{1}}c_{\alpha}(u)\alpha and vηη=βC2dβ(v)βv\eta-\eta=\sum_{\beta\in C_{2}}d_{\beta}(v)\beta. We then have,

wηη\displaystyle w\eta-\eta =uvηη=u(vηη)+uηη\displaystyle=uv\eta-\eta=u(v\eta-\eta)+u\eta-\eta
=u(βC2dβ(v)β)+αC1cα(u)α=βC2dβ(v)β+αC1cα(u)α\displaystyle=u\left(\sum_{\beta\in C_{2}}d_{\beta}(v)\beta\right)+\sum_{\alpha\in C_{1}}c_{\alpha}(u)\alpha=\sum_{\beta\in C_{2}}d_{\beta}(v)\beta+\sum_{\alpha\in C_{1}}c_{\alpha}(u)\alpha
=(vηη)+(uηη).\displaystyle=(v\eta-\eta)+(u\eta-\eta).

The penultimate equality holds because uu can be written as u=sα1sα2sαku=s_{\alpha_{1}}s_{\alpha_{2}}\cdots s_{\alpha_{k}} where each αiC1\alpha_{i}\in C_{1}; and as C1C_{1} is orthogonal to C2C_{2}, each sαis_{\alpha_{i}} fixes β\beta for βC2\beta\in C_{2}. ∎

We now define a monomial that will be of particular importance. Let CΠ0C\subseteq\Pi_{0} be any subset and let λ\lambda be a typical dominant integral weight. We define

Xλ(C):-αCXαλ+ρ,α.X^{\lambda}(C)\coloneq\prod_{\alpha\in C}X_{\alpha}^{\langle\lambda+\rho,\alpha\rangle}.
Proposition 4.2.

Let λ\lambda be a typical dominant integral weight of 𝔤\mathfrak{g}, and C1,C2C_{1},C_{2} be the two connected components of Π0\Pi_{0}. For i{1,2}i\in\{1,2\}, denote by Ui(λ)U_{i}(\lambda), the factor corresponding to CiC_{i} of U(λ)U(\lambda). Then we have:

  1. (a)

    The support of any monomial, say αΠ0Xαcα\prod_{\alpha\in\Pi_{0}}X_{\alpha}^{c_{\alpha}} that appears in logUi(λ)-\log U_{i}(\lambda) with nonzero coefficient is contained in CiC_{i}, i.e:

    supp(αΠ0Xαcα){αΠ0:cα0}Ci.\operatorname{\text{supp}}\left(\prod_{\alpha\in\Pi_{0}}X_{\alpha}^{c_{\alpha}}\right)\coloneqq\{\alpha\in\Pi_{0}\colon c_{\alpha}\not=0\}\subseteq C_{i}.
  2. (b)

    For any CCiC\subseteq C_{i}, the coefficient of the monomial Xλ(C)X^{\lambda}(C) depends only on CC.

  3. (c)

    Xλ(C)X^{\lambda}(C) appears in logUi(λ)-\log U_{i}(\lambda) with nonzero coefficient if and only if CC is a connected subset of CiC_{i}. In particular, Xλ(Ci)X^{\lambda}(C_{i}) appears in logUi(λ)-\log U_{i}(\lambda).

Proof.

As in Equation (4.1), we write Ui(λ)=1τU_{i}(\lambda)=1-\tau, where τ=1w𝒲i(1)(w)X(w,λ)\tau=-\sum_{1\not=w\in\mathscr{W}_{i}}(-1)^{\ell(w)}X(w,\lambda). Since every w𝒲iw\in\mathscr{W}_{i} is a product of simple reflections sαs_{\alpha}, where αCi\alpha\in C_{i}, the proof of \autorefsopr shows that the support of X(w,λ)X(w,\lambda) for any w𝒲iw\in\mathscr{W}_{i} is contained in CiC_{i}.

Now logUi(λ)=k1τk/k-\log U_{i}(\lambda)=\sum_{k\geqslant 1}\nicefrac{{\tau^{k}}}{{k}}, and every power of τ\tau is just a sum of products of X(w,λ)X(w,\lambda)’s such that ww varying over 𝒲i\mathscr{W}_{i}, it follows that the support of any monomial in logUi(λ)-\log U_{i}(\lambda) is a subset of CiC_{i}. This proves (a).

To show (b), we write τ=τ1+τ2\tau=\tau_{1}+\tau_{2} with

τ1w(1)(w)X(w,λ)andτ2w(1)(w)X(w,λ).\displaystyle\tau_{1}\coloneqq-\sum_{w\in\mathcal{I}}(-1)^{\ell(w)}X(w,\lambda)\quad\text{and}\quad\tau_{2}\coloneqq-\sum_{w\notin\mathcal{I}}(-1)^{\ell(w)}X(w,\lambda).
Then we have,
logUi(λ)=k1(τ1+τ2)kk.\displaystyle-\log U_{i}(\lambda)=\sum_{k\geqslant 1}\frac{(\tau_{1}+\tau_{2})^{k}}{k}.

Recall that for a subset CC of Π0\Pi_{0}, Xλ(C)=αCXαλ+ρ,αX^{\lambda}(C)=\prod_{\alpha\in C}X_{\alpha}^{\langle\lambda+\rho,\alpha\rangle}. \autorefkeylem, Part (d) shows that τ2\tau_{2} does not contribute to the appearance of Xλ(C)X^{\lambda}(C) in the expansion of logUi(λ)-\log U_{i}(\lambda). In other words, the coefficient of Xλ(C)X^{\lambda}(C) in k1τk/k\sum_{k\geqslant 1}\nicefrac{{\tau^{k}}}{{k}} is same as that in k1τ1k/k\sum_{k\geqslant 1}\nicefrac{{\tau_{1}^{k}}}{{k}}. Therefore, it is sufficient to compute the coefficient of Xλ(C)X^{\lambda}(C) in τ1k/k\nicefrac{{\tau_{1}^{k}}}{{k}}. We have that

τ1k=(1)kwj(1)(wj)j=1kX(wj,λ).\tau_{1}^{k}=(-1)^{k}\sum_{w_{j}\in\mathcal{I}}(-1)^{\sum\ell(w_{j})}\prod_{j=1}^{k}X(w_{j},\lambda).

By \autorefkeylem, Part (c), the product j=1kX(wj,λ)\prod_{j=1}^{k}X(w_{j},\lambda) equals Xλ(C)X^{\lambda}(C) only when j=1kI(wj)=C\cup_{j=1}^{k}I(w_{j})=C; and each I(wj)I(w_{j}) is totally disconnected with I(wj)I(wl)=I(w_{j})\cap I(w_{l})=\emptyset for every ili\not=l. By \autorefdef k partition, this means that (I(w1),,I(wk))(I(w_{1}),\dots,I(w_{k})) is a kk-partition of CC. In particular, for this kk-partition the coefficient of Xλ(C)X^{\lambda}(C) in τ1k\tau_{1}^{k} is (1)k(1)(wj)(-1)^{k}(-1)^{\sum\ell(w_{j})}. As each I(wj)I(w_{j}) is totally disconnected, j=1k(wj)=(w1wk)\sum_{j=1}^{k}\ell(w_{j})=\ell(w_{1}\cdots w_{k}). If we denote the kk-partition (I(w1),,I(wk))(I(w_{1}),\dots,I(w_{k})) by 𝒥\mathcal{J}, then w1wkw_{1}\cdots w_{k} is just w(𝒥)w(\mathcal{J}); and (1)(w(𝒥))=(1)|C|(-1)^{\ell(w(\mathcal{J}))}=(-1)^{|C|}. We obtain that, the coefficient of Xλ(C)X^{\lambda}(C) in logUi(λ)-\log U_{i}(\lambda) is:

k1𝒥Pk(𝒞)(1)k(1)(w(𝒥))k=(1)|C|k=1|C|(1)kck(𝒞)k=k(C),\sum_{k\geqslant 1}\sum_{\mathcal{J}\in P_{k}(\mathcal{C})}\frac{(-1)^{k}\cdot(-1)^{\ell{(w(\mathcal{J}))}}}{k}=(-1)^{|C|}\sum_{k=1}^{|C|}(-1)^{k}\frac{c_{k}(\mathcal{C})}{k}=k(C),

where 𝒞\mathcal{C} is the graph spanned by CC. By  [MR2980495]*Proposition 2, k(C)k(C) is a positive integer if and only if 𝒞\mathcal{C} is connected, otherwise it is zero. Evidently k(C)k(C) does not depend on λ\lambda, and it is determined completely by CC. This completes the proof of (b) and (c). ∎

Remark 2.

The coefficient k(C)k(C) above is actually 1 when CC is connected  [MR2980495]*Corollary 1.

Following  [MR4343717], we define for a connected subset CC of Π0\Pi_{0}, a linear operator ΘC:𝒜𝒜\Theta_{C}\colon\mathcal{A}\to\mathcal{A} by

f=𝒎𝑿𝒎ΘC(f)𝒎supp(𝒎)=C𝑿𝒎,f=\sum_{\boldsymbol{m}}\boldsymbol{X^{m}}\rightsquigarrow\Theta_{C}(f)\coloneqq\sum_{\begin{subarray}{c}\boldsymbol{m}\\ \operatorname{\text{supp}}(\boldsymbol{m})=C\end{subarray}}\boldsymbol{X^{m}},

where 𝒎=(mα:αΠ0)\boldsymbol{m}=(m_{\alpha}\colon\alpha\in\Pi_{0}) is an nn-tuple, where n=Card(Π0)n=\text{Card}(\Pi_{0}), supp(𝒎){αΠ0:mα0}\operatorname{\text{supp}}(\boldsymbol{m})\coloneqq\{\alpha\in\Pi_{0}\colon m_{\alpha}\neq 0\} and 𝑿𝒎=Xα1mα1Xαtmαt\boldsymbol{X^{m}}=X_{\alpha_{1}}^{m_{\alpha_{1}}}\cdots X_{\alpha_{t}}^{m_{\alpha_{t}}} for 𝒎=(mα1,,mαt)\boldsymbol{m}=(m_{\alpha_{1}},\dots,m_{\alpha_{t}}).

Proof of the following proposition follows directly from \autorefmainprop.

Proposition 4.3.

Let λ\lambda be a typical dominant integral weight and let CC be a connected component of Π0\Pi_{0}. Then

ΘC(logUi(λ))=k(C)Xλ(C)+monomials of degree > degXλ(C).\displaystyle\Theta_{C}(-\log U_{i}(\lambda))=k(C)X^{\lambda}(C)+\text{monomials of degree $>$ }\deg X^{\lambda}(C). (4.3)

where the constant k(C)k(C) depends only on CC and degXλ(C)αCλ+ρ,α\deg X^{\lambda}(C)\coloneqq\sum_{\alpha\in C}\langle\lambda+\rho,\alpha\rangle.

The following lemma gives conditions for which Ui(λ)U_{i}(\lambda) and Uj(μ)U_{j}(\mu) are equal.

Lemma 4.4.

Let λ,μ\lambda,\mu be typical dominant integral weights and let C1,C2C_{1},C_{2} be the two connected components of Π0\Pi_{0}. Then following statements are equivalent:

  1. (1)

    Xλ(Ci)=Xμ(Cj)X^{\lambda}(C_{i})=X^{\mu}(C_{j}),

  2. (2)

    Ci=CjC_{i}=C_{j} and λ,αk=μ,αk\langle\lambda,\alpha_{k}\rangle=\langle\mu,\alpha_{k}\rangle for every αkCi\alpha_{k}\in C_{i},

  3. (3)

    Ui(λ)=Uj(μ)U_{i}(\lambda)=U_{j}(\mu).

Proof.

Assume (1). Then all the variables with their exponents must be same; and as a consequence, their corresponding supports are equal, hence (2) follows from (1).

Now if Ci=CjC_{i}=C_{j} we get that 𝒲i=𝒲j\mathscr{W}_{i}=\mathscr{W}_{j}, where 𝒲i\mathscr{W}_{i} (resp. 𝒲j\mathscr{W}_{j}) is the group generated by sαs_{\alpha}, αCi\alpha\in C_{i} (resp. αCj\alpha\in C_{j}). By \autorefsopr the exponents of all the monomials appearing in Ui(λ)U_{i}(\lambda) (resp. Uj(μ)U_{j}(\mu)) depend only on the integers λ+ρ,αk\langle\lambda+\rho,\alpha_{k}\rangle (resp. μ+ρ,αk\langle\mu+\rho,\alpha_{k}\rangle) for all αkCi=Cj\alpha_{k}\in C_{i}=C_{j}, whence (3) follows from (2). Finally, assume (3). After applying log-\log and then ΘCi\Theta_{C_{i}} to both sides of the equation, we get that

ΘCi(logUi(λ))=ΘCi(logUj(μ)).\displaystyle\Theta_{C_{i}}(-\log U_{i}(\lambda))=\Theta_{C_{i}}(-\log U_{j}(\mu)). (4.4)

By \autoreflowestdeg we know that Xλ(Ci)X^{\lambda}(C_{i}) must occur in ΘCi(logUj(μ))\Theta_{C_{i}}(-\log U_{j}(\mu)) with nonzero coefficient. On the other hand, support of any monomial appearing in logUj(μ)-\log U_{j}(\mu) is a subset of CjC_{j} by \autorefmainprop, Part (a). This is possible only when Ci=CjC_{i}=C_{j}. Then using Proposition 4.3 once again we get that Xλ(Ci)=Xμ(Cj)X^{\lambda}(C_{i})=X^{\mu}(C_{j}). ∎

5. The main theorem

We have seen that U(λ)U(\lambda) factorizes as a product: U(λ)=U1(λ)U2(λ)U(\lambda)=U_{1}(\lambda)\cdot U_{2}(\lambda) of two factors. Let λ1,,λr,μ1,,μs\lambda_{1},\dots,\lambda_{r},\mu_{1},\dots,\mu_{s} be typical dominant integral weights and suppose that the following equality holds:

U(λ1)U(λr)=U(μ1)U(μs).U(\lambda_{1})\cdots U(\lambda_{r})=U(\mu_{1})\cdots U(\mu_{s}).

After factoring them further one obtains,

U1(λ1)U2(λ1)U1(λr)U2(λr)=U1(μ1)U2(μ1)U1(μs)U2(μs).\displaystyle{}U_{1}(\lambda_{1})U_{2}(\lambda_{1})\cdots U_{1}(\lambda_{r})U_{2}(\lambda_{r})=U_{1}(\mu_{1})U_{2}(\mu_{1})\cdots U_{1}(\mu_{s})U_{2}(\mu_{s}). (5.1)
Theorem 5.1.

With the notations as above, we have r=sr=s and the factors in (5.1) are all equal up to a permutation, i.e., there exists a bijection σ\sigma of {1,2}×{1,,r}\{1,2\}\times\{1,\dots,r\} such that Ui(λp)=Uj(λq)U_{i}(\lambda_{p})=U_{j}(\lambda_{q}), where i,j{1,2}i,j\in\{1,2\}, p,q{1,,r}p,q\in\{1,\dots,r\} and σ(i,p)=(j,q)\sigma(i,p)=(j,q).

Proof.

Applying log-\log on both sides of Equation (5.1), we obtain

p=1ri=12logUi(λp)=q=1sj=12logUj(μq).\displaystyle\sum_{p=1}^{r}\sum_{i=1}^{2}-\log U_{i}(\lambda_{p})=\sum_{q=1}^{s}\sum_{j=1}^{2}-\log U_{j}(\mu_{q}). (5.2)
Pick a connected component of Π0\Pi_{0}, say C1C_{1}. We choose λk\lambda_{k} such that the monomial Xλk(C1)X^{\lambda_{k}}(C_{1}) is of minimal degree among all the monomials with support C1C_{1} in the left hand side of Equation (5.2). Applying the operator ΘC1\Theta_{C_{1}} to Equation (5.2) yields:
ΘC1(p=1ri=12logUi(λp))=ΘC1(q=1sj=12logUj(μq)).\displaystyle\Theta_{C_{1}}\left(\sum_{p=1}^{r}\sum_{i=1}^{2}-\log U_{i}(\lambda_{p})\right)=\Theta_{C_{1}}\left(\sum_{q=1}^{s}\sum_{j=1}^{2}-\log U_{j}(\mu_{q})\right). (5.3)

By \autorefmainprop, the support of any monomial that appears in logUi(λp)-\log U_{i}(\lambda_{p}) is contained in CiC_{i}. Therefore, if i1i\neq 1, then ΘC1(logUi(λp))=0\Theta_{C_{1}}(-\log U_{i}(\lambda_{p}))=0. On the other hand if i=1i=1, then Xλp(Ci)X^{\lambda_{p}}(C_{i}) is the minimal degree monomial in ΘC1(logUi(λp))\Theta_{C_{1}}(-\log U_{i}(\lambda_{p})) which is nonzero in this case.

Since the monomial Xλk(C1)X^{\lambda_{k}}(C_{1}) occurs in the left hand side of Equation (5.3), by minimality of degree, it must appear in the right hand side with nonzero coefficient. In other words, there exists 1qs1\leqslant q\leqslant s and j{1,2}j\in\{1,2\} such that Xλk(C1)=Xμq(Cj)X^{\lambda_{k}}(C_{1})=X^{\mu_{q}}(C_{j}). \autorefmainlem now gives credence to the conclusion that U1(λk)=Uj(μq)U_{1}(\lambda_{k})=U_{j}(\mu_{q}). To obtain the desired claim we cancel U1(λk)andUj(μq)U_{1}(\lambda_{k})\ \text{and}\ U_{j}(\mu_{q}) in Equation (5.1) and proceed by induction.

To prove r=sr=s, we observe that the number of factors in the left hand side of Equation (5.1) is 2r2r and that in the right hand side is 2s2s. So if r>sr>s, then on the left hand side we have a product of 2(rs)2(r-s) number of factors whereas in the right hand side we have 11, which is a contradiction. So we conclude that r=sr=s. ∎

We now give a sufficient condition for which two tensor products of irreducible typical representations are isomorphic to each other.

Theorem 5.2.

Let λ1,,λr,μ1,,μs\lambda_{1},\dots,\lambda_{r},\mu_{1},\dots,\mu_{s} be typical dominant integral weights and assume that

V(λ1)V(λr)V(μ1)V(μs).\displaystyle{}V(\lambda_{1})\otimes\dots\otimes V(\lambda_{r})\cong V(\mu_{1})\otimes\dots\otimes V(\mu_{s}). (5.4)

Then r=sr=s. Suppose the bijection σ\sigma in \autorefpermfac has the following additional property: if Ui(λp)=Uj(μq)U_{i}(\lambda_{p})=U_{j}(\mu_{q}), then σ(j,p)=(i,q)\sigma(j,p)=(i,q) for all i,j{1,2}i,j\in\{1,2\}, p,q{1,,r}p,q\in\{1,\dots,r\} and σ(i,p)=(j,q)\sigma(i,p)=(j,q). Then V(λp)V(μq)V(\lambda_{p})\cong V(\mu_{q}) for all p,q{1,,r}p,q\in\{1,\dots,r\}.

Proof.

To prove r=sr=s, we assume on contrary that r>sr>s. The maximal weights that occur on both sides of the isomorphism (5.4) are equal, i.e, i=1rλi=j=1sμjγ,\sum_{i=1}^{r}\lambda_{i}=\sum_{j=1}^{s}\mu_{j}\coloneqq\gamma, say. Taking formal character on both sides of Equation (5.4) yields:

chV(λ1)chV(λr)=chV(μ1)chV(μs).\displaystyle\operatorname{\textnormal{ch}}V(\lambda_{1})\cdots\operatorname{\textnormal{ch}}V(\lambda_{r})=\operatorname{\textnormal{ch}}V(\mu_{1})\cdots\operatorname{\textnormal{ch}}V(\mu_{s}).
Multiplying both sides of the above equation by eγe^{-\gamma} and grouping the corresponding highest weights we obtain:
i=1reλichV(λi)=j=1seμjchV(μj).\displaystyle\prod_{i=1}^{r}e^{-\lambda_{i}}\cdot\operatorname{\textnormal{ch}}V(\lambda_{i})=\prod_{j=1}^{s}e^{-\mu_{j}}\cdot\operatorname{\textnormal{ch}}V(\mu_{j}).
By Equation (2.6), this simplifies to
U(λ1)U(λr)Drs=U(μ1)U(μs),where D=αΦ1+(1+eα)αΦ0+(1eα).\displaystyle U(\lambda_{1})\cdots U(\lambda_{r})D^{r-s}=U(\mu_{1})\cdots U(\mu_{s}),\quad\text{where $D=\tfrac{\prod_{\alpha\in\Phi^{+}_{1}}(1+e^{-\alpha})}{\prod_{\alpha\in\Phi^{+}_{0}}(1-e^{-\alpha})}$}.
As every U(λk)U(\lambda_{k}) (resp. U(μl)U(\mu_{l})) is a product of two factors, we get that:
U1(λ1)U2(λ1)U1(λr)U2(λr)Drs=U1(μ1)U2(μ1)U1(μs)U2(μs).\displaystyle U_{1}(\lambda_{1})U_{2}(\lambda_{1})\cdots U_{1}(\lambda_{r})U_{2}(\lambda_{r})D^{r-s}=U_{1}(\mu_{1})U_{2}(\mu_{1})\cdots U_{1}(\mu_{s})U_{2}(\mu_{s}). (5.5)

By \autorefpermfac, we conclude that Ui(λp)=Uj(μq)U_{i}(\lambda_{p})=U_{j}(\mu_{q}), for some j{1,2}j\in\{1,2\} and q{1,,s}q\in\{1,\dots,s\}. Therefore, arguing as in the proof of \autorefpermfac we get a contradiction. Hence r=sr=s.

Now suppose the bijection σ\sigma satisfies the condition mentioned in the hypotheses of the theorem. Then we get that

i=12Ui(λp)=j=12Uj(μq).\prod_{i=1}^{2}U_{i}(\lambda_{p})=\prod_{j=1}^{2}U_{j}(\mu_{q}).

This implies that U(λp)=U(μq)U(\lambda_{p})=U(\mu_{q}). By \autorefmainlem, we get that V(λp)V(μq)V(\lambda_{p})\cong V(\mu_{q}). ∎

The following corollary says that the unique factorization of tensor products of irreducible finite dimensional typical representations holds for superalgebras of types B(0,n),Cn,F(4)B(0,n),C_{n},F(4) and G(3)G(3).

Corollary 5.3.

Let 𝔤\mathfrak{g} be of type A(n,0),B(0,n),Cn,F(4)A(n,0),B(0,n),C_{n},F(4) or G(3)G(3) and let λ1,,λr,μ1,,μs\lambda_{1},\dots,\lambda_{r},\mu_{1},\dots,\mu_{s} be typical dominant integral weights of 𝔤\mathfrak{g}. Suppose we are given an isomorphism of representations:

V(λ1)V(λr)V(μ1)V(μs).\displaystyle{}V(\lambda_{1})\otimes\dots\otimes V(\lambda_{r})\cong V(\mu_{1})\otimes\dots\otimes V(\mu_{s}). (5.6)

Then r=sr=s, and there is a permutation σ\sigma of {1,,r}\{1,\dots,r\} such that V(λk)V(μσ(k))V(\lambda_{k})\cong V(\mu_{\sigma(k)}) for all kk.

Proof.

By \autorefThm:tnsrpdt, we have r=sr=s. Proceeding as in its proof we further obtain that

U(λ1)U(λr)=U(μ1)U(μr)\displaystyle U(\lambda_{1})\cdots U(\lambda_{r})=U(\mu_{1})\cdots U(\mu_{r}) (5.7)

The standard Dynkin diagrams of these types are given in  [MR0519631]*Table 1, page 606. Evidently Π0\Pi_{0} is connected in all of these cases as the node corresponding to the odd simple root appears at the edge of the diagrams. So removal of this node will not render the diagram disconnected. This means that U2(λk)=1U_{2}(\lambda_{k})=1 and U2(μl)=1U_{2}(\mu_{l})=1 for all k,l{1,,r}k,l\in\{1,\dots,r\}. Then the corollary follows from \autorefpermfac and \autorefmainlem.

For type B(0,n)B(0,n) an alternate proof can be given. It is known that if λ\lambda is the highest weight of an irreducible representation of 𝔬𝔰𝔭(1,2n)\mathfrak{osp}(1,2n), then it is also the highest weight of a non spinorial irreducible representation of 𝔰𝔬(2n+1)\mathfrak{so}(2n+1); and characters of both the representations are same. For details we refer to  [MR0648354]. Taking formal character on both sides of the isomorphism (5.6) yields:

chV(λ1)chV(λr)=chV(μ1)chV(μr).\operatorname{\textnormal{ch}}V(\lambda_{1})\cdots\operatorname{\textnormal{ch}}V(\lambda_{r})=\operatorname{\textnormal{ch}}V(\mu_{1})\cdots\operatorname{\textnormal{ch}}V(\mu_{r}).

The above discussion shows that this equality can be considered an equality for 𝔰𝔬(2n+1)\mathfrak{so}(2n+1)-modules. By  [MR2123935]*Theorem 1, this implies that V(λi)V(μσ(i))V(\lambda_{i})\cong V(\mu_{\sigma(i)}) as 𝔰𝔬(2n+1)\mathfrak{so}(2n+1)-modules for some permutation σ\sigma of {1,,r}\{1,\dots,r\}; so chV(λi)=chV(μσ(i))\operatorname{\textnormal{ch}}V(\lambda_{i})=\operatorname{\textnormal{ch}}V(\mu_{\sigma(i)}). Hence V(λi)V(μσ(i))V(\lambda_{i})\cong V(\mu_{\sigma(i)}) as 𝔬𝔰𝔭(1,2n)\mathfrak{osp}(1,2n)-modules. ∎

The example below illustrates that the additional hypothesis on the bijection σ\sigma in \autorefThm:tnsrpdt is essential for the conclusion of unique factorization of the tensor products.

Example 5.4.

We take 𝔤=𝔰𝔩(3,2)\mathfrak{g}=\mathfrak{sl}(3,2). In this case Π0={α1,α2,α3}\Pi_{0}=\{\alpha_{1},\alpha_{2},\alpha_{3}\}, and C1={α1,α2}C_{1}=\{\alpha_{1},\alpha_{2}\}, C2={α3}C_{2}=\{\alpha_{3}\} are the two connected components of Π0\Pi_{0}. The subgroups of the Weyl group of 𝔤\mathfrak{g} generated by C1C_{1} and C2C_{2} are 𝒲1={1,s1,s2,s1s2,s2s1,s1s2s1}\mathscr{W}_{1}=\{1,s_{1},s_{2},\ s_{1}s_{2},\ s_{2}s_{1},\ s_{1}s_{2}s_{1}\} and 𝒲2={1,s3}\mathscr{W}_{2}=\{1,s_{3}\} respectively. Here sis_{i} is the simple reflection corresponding to αi\alpha_{i}. The set of all positive odd roots Φ1+\Phi_{1}^{+} is given by:

Φ1+={εiδj:1i3; 1j2},\Phi_{1}^{+}=\{\varepsilon_{i}-\delta_{j}\colon 1\leqslant i\leqslant 3;\ 1\leqslant j\leqslant 2\},

where εi(diag(a1,a2,a3))=ai\varepsilon_{i}(\text{diag}(a_{1},a_{2},a_{3}))=a_{i} and δj(diag(b1,b2))=bj\delta_{j}(\text{diag}(b_{1},b_{2}))=b_{j} for all i,ji,j.

The sum of all elements in Φ1+\Phi_{1}^{+} is τ:-2(ε1+ε2+ε3)3(δ1+δ2)𝔥\tau\coloneq 2(\varepsilon_{1}+\varepsilon_{2}+\varepsilon_{3})-3(\delta_{1}+\delta_{2})\in\mathfrak{h}^{*}. Here all odd roots are isotropic. We have that (τ,εiδj)=5(\tau,\varepsilon_{i}-\delta_{j})=5. Consider now the following weights:

λ1=ω1+2ω2+3ω3+τλ2=ω1+4ω2+5ω3+τandμ1=ω1+4ω2+3ω3+τμ2=ω1+2ω2+5ω3+τ\begin{gathered}\lambda_{1}=\omega_{1}+2\omega_{2}+3\omega_{3}+\tau\\ \lambda_{2}=\omega_{1}+4\omega_{2}+5\omega_{3}+\tau\end{gathered}\qquad\text{and}\qquad\begin{gathered}\mu_{1}=\omega_{1}+4\omega_{2}+3\omega_{3}+\tau\\ \mu_{2}=\omega_{1}+2\omega_{2}+5\omega_{3}+\tau\end{gathered}

where ωi\omega_{i} is the fundamental dominant weight corresponding to αi\alpha_{i}. All of these weights are typical dominant integral (cf. \autorefrmkdominant). After computing and factoring the normalized Weyl numerators we find:

U1(λ1)=1Xα12Xα23+Xα15Xα23+Xα12Xα25Xα15Xα25U1(λ2)=1Xα12Xα25+Xα17Xα25+Xα12Xα27Xα17Xα27U1(μ1)=1Xα12Xα25+Xα17Xα25+Xα12Xα27Xα17Xα27U1(μ2)=1Xα12Xα23+Xα15Xα23+Xα12Xα25Xα15Xα25andU2(λ1)=1Xα34U2(λ2)=1Xα36U2(μ1)=1Xα34U2(μ2)=1Xα36\begin{gathered}U_{1}(\lambda_{1})=1-X_{\alpha_{1}}^{2}-X_{\alpha_{2}}^{3}+X_{\alpha_{1}}^{5}X_{\alpha_{2}}^{3}+X_{\alpha_{1}}^{2}X_{\alpha_{2}}^{5}-X_{\alpha_{1}}^{5}X_{\alpha_{2}}^{5}\\ U_{1}(\lambda_{2})=1-X_{\alpha_{1}}^{2}-X_{\alpha_{2}}^{5}+X_{\alpha_{1}}^{7}X_{\alpha_{2}}^{5}+X_{\alpha_{1}}^{2}X_{\alpha_{2}}^{7}-X_{\alpha_{1}}^{7}X_{\alpha_{2}}^{7}\\ U_{1}(\mu_{1})=1-X_{\alpha_{1}}^{2}-X_{\alpha_{2}}^{5}+X_{\alpha_{1}}^{7}X_{\alpha_{2}}^{5}+X_{\alpha_{1}}^{2}X_{\alpha_{2}}^{7}-X_{\alpha_{1}}^{7}X_{\alpha_{2}}^{7}\\ U_{1}(\mu_{2})=1-X_{\alpha_{1}}^{2}-X_{\alpha_{2}}^{3}+X_{\alpha_{1}}^{5}X_{\alpha_{2}}^{3}+X_{\alpha_{1}}^{2}X_{\alpha_{2}}^{5}-X_{\alpha_{1}}^{5}X_{\alpha_{2}}^{5}\end{gathered}\qquad\text{and}\qquad\begin{gathered}U_{2}(\lambda_{1})=1-X_{\alpha_{3}}^{4}\\ U_{2}(\lambda_{2})=1-X_{\alpha_{3}}^{6}\\ U_{2}(\mu_{1})=1-X_{\alpha_{3}}^{4}\\ U_{2}(\mu_{2})=1-X_{\alpha_{3}}^{6}\end{gathered}

Evidently,

U1(λ1)U2(λ1)U1(λ2)U2(λ2)\displaystyle U_{1}(\lambda_{1})U_{2}(\lambda_{1})U_{1}(\lambda_{2})U_{2}(\lambda_{2}) =U1(μ1)U2(μ1)U1(μ2)U2(μ2)\displaystyle=U_{1}(\mu_{1})U_{2}(\mu_{1})U_{1}(\mu_{2})U_{2}(\mu_{2})
U(λ1)U(λ2)\displaystyle\Rightarrow U(\lambda_{1})U(\lambda_{2}) =U(μ1)U(μ2).\displaystyle=U(\mu_{1})U(\mu_{2}).
Using the fact that λ1+λ2=μ2+μ2\lambda_{1}+\lambda_{2}=\mu_{2}+\mu_{2}, and multiplying both sides by D1/D2D_{1}/D_{2} (cf. Equation (2.3)) we obtain
chV(λ1)chV(λ2)\displaystyle\operatorname{\textnormal{ch}}V(\lambda_{1})\operatorname{\textnormal{ch}}V(\lambda_{2}) =chV(μ1)chV(μ2).\displaystyle=\operatorname{\textnormal{ch}}V(\mu_{1})\operatorname{\textnormal{ch}}V(\mu_{2}).
From here we get that
V(λ1)V(λ2)\displaystyle V(\lambda_{1})\otimes V(\lambda_{2}) V(μ1)V(μ2).\displaystyle\cong V(\mu_{1})\otimes V(\mu_{2}).

So, unique decomposition of tensor products does not hold.

6. Atypical representations

6.1.

In this section we focus on some atypical representations of 𝔰𝔩(m+1,n+1),Cn+1=𝔬𝔰𝔭(2,2n),G(3)\mathfrak{sl}(m+1,n+1),\ C_{n+1}=\mathfrak{osp}(2,2n),\ G(3) and F(4)F(4) for which a character formula is known in the literature. A weight λ\lambda is called singly atypical if there is a unique γΦ1+\gamma\in\Phi_{1}^{+} such that (λ+ρ,γ)=0(\lambda+\rho,\gamma)=0. In this case, λ\lambda is said to have atypicality type γ\gamma. For 𝔤=𝔰𝔩(m+1,n+1)\mathfrak{g}=\mathfrak{sl}(m+1,n+1), the class of singly atypical finite dimensional irreducible 𝔤\mathfrak{g}-modules admit a character formula closely resembling that of the typical ones. Details can be found in  [MR1063989]. It is known that any dominant integral weight of type Cn+1C_{n+1}, and the exceptional Lie superalgebras F(4)F(4) and G(3)G(3) is either typical or singly atypical; and a character formula for singly atypical finite dimensional irreducible representations for type CnC_{n} can be found in  [MR1092559]. The same for the exceptional Lie superalgebras is obtained in  [MR3253284].

Our goal of this section is to establish unique decomposition of tensor products of singly atypical finite dimensional irreducible modules for the above mentioned Lie superalgebras. Let 𝔤\mathfrak{g} be any such Lie superalgebra and λ𝔥\lambda\in\mathfrak{h}^{*} be a dominant integral singly atypical weight of type β\beta. For G(3)G(3) and F(4)F(4), we first assume that λλ1,λ2\lambda\neq\lambda_{1},\lambda_{2} where λ1\lambda_{1} and λ2\lambda_{2} are the special weights (see  [MR3253284]*Theorem 2.6 for a description of these two weights). Then from  [MR1063989, MR1092559, MR3253284], the character formula for V(λ)V(\lambda) is given by:

chV(λ)\displaystyle\operatorname{\textnormal{ch}}V(\lambda) =D1D2w𝒲(1)(w)11+ewβew(λ+ρ).\displaystyle=\frac{D_{1}}{D_{2}}\sum_{w\in\mathscr{W}}(-1)^{\ell(w)}\frac{1}{1+e^{-w\beta}}\cdot e^{w(\lambda+\rho)}. (6.1)
When λ\lambda is either λ1\lambda_{1} or λ2\lambda_{2} for type G(3)G(3) and F(4)F(4), the character formula for V(λ)V(\lambda) is given by:
chV(λ)\displaystyle\operatorname{\textnormal{ch}}V(\lambda) =D1D2w𝒲(1)(w)22+ewβ1+ewβew(λ+ρ),\displaystyle=\frac{D_{1}}{D_{2}}\sum_{w\in\mathscr{W}}\frac{(-1)^{\ell(w)}}{2}\frac{2+e^{-w\beta}}{1+e^{-w\beta}}\cdot e^{w(\lambda+\rho)}, (6.2)

where D1=αΦ1+(eα/2+eα/2)D_{1}=\prod_{\alpha\in\Phi^{+}_{1}}(e^{\alpha/2}+e^{-\alpha/2}) and D2=αΦ0+(eα/2eα/2)D_{2}=\prod_{\alpha\in\Phi^{+}_{0}}(e^{\alpha/2}-e^{-\alpha/2}).

As in \autorefsecprelims, we define the normalized Weyl numerator U(λ)U(\lambda) for λλ1,λ2\lambda\neq\lambda_{1},\lambda_{2} by:

U(λ)w𝒲(1)(w)ew(λ+ρ)(λ+ρ)1+ewβ=w𝒲(1)(w)11+ewβX(λ,w).\displaystyle U(\lambda)\coloneqq\sum_{w\in\mathscr{W}}\frac{(-1)^{\ell(w)}e^{w(\lambda+\rho)-(\lambda+\rho)}}{1+e^{-w\beta}}=\sum_{w\in\mathscr{W}}(-1)^{\ell(w)}\frac{1}{1+e^{-w\beta}}\cdot X(\lambda,w). (6.3)
For λ=λ1,λ2\lambda=\lambda_{1},\lambda_{2} in type G(3)G(3) and F(4)F(4), we define:
U(λ)w𝒲(1)(w)22+ewβ1+ewβX(λ,w)\displaystyle U(\lambda)\coloneqq\sum_{w\in\mathscr{W}}\frac{(-1)^{\ell(w)}}{2}\frac{2+e^{-w\beta}}{1+e^{-w\beta}}\cdot X(\lambda,w) (6.4)

where Xα=eαX_{\alpha}=e^{-\alpha} for αΠ0\alpha\in\Pi_{0}, and X(λ,w)=αΠ0Xαcα(w)=ew(λ+ρ)(λ+ρ)X(\lambda,w)=\prod_{\alpha\in\Pi_{0}}X_{\alpha}^{c_{\alpha}(w)}=e^{w(\lambda+\rho)-(\lambda+\rho)}.

The definition of the normalized character is same as before. Equation (2.6) remains valid here as well. Let us recall that the monomial Xλ(Π0)X^{\lambda}(\Pi_{0}) is defined by Xλ(Π0):-αΠ0Xαλ+ρ,αX^{\lambda}(\Pi_{0})\coloneq\prod_{\alpha\in\Pi_{0}}X_{\alpha}^{\langle\lambda+\rho,\alpha\rangle} (see 4.2). As before, our primary task is to show that the coefficient of this monomial in the expansion of logU(λ)-\log U(\lambda) is nonzero. We achieve this by means of case by case consideration. First we record the following lemma that shows Xλ(Π0)X^{\lambda}(\Pi_{0}) does determine the representation V(λ)V(\lambda).

Notation.

In what follows we denote the monomial Xλ(Π0)X^{\lambda}(\Pi_{0}) by just XλX^{\lambda} for every λ𝔥\lambda\in\mathfrak{h}^{*}.

Lemma 6.1.

Let 𝔤\mathfrak{g} be any Lie superalgebra among 𝔰𝔩(m+1,n+1),𝔬𝔰𝔭(2,2n),G(3)\mathfrak{sl}(m+1,n+1),\ \mathfrak{osp}(2,2n),\ G(3), or F(4)F(4). Let λ,μ𝔥\lambda,\mu\in\mathfrak{h}^{*} be dominant integral singly atypical weights of the same type, say γ\gamma. Then the following statements are equivalent:

(a) Xλ=XμX^{\lambda}=X^{\mu}, (b) λ=μ\lambda=\mu, (c) U(λ)=U(μ)U(\lambda)=U(\mu).
Proof.

Assume (a). By definition, this means that λ+ρ,α=μ+ρ,α\langle\lambda+\rho,\alpha\rangle=\langle\mu+\rho,\alpha\rangle, for every αΠ0\alpha\in\Pi_{0}. Now an argument as in the proof of \autorefmainlem gives us λ=μ\lambda=\mu. If we have λ=μ\lambda=\mu to begin with, then Xλ=XμX^{\lambda}=X^{\mu} follows form the definition. This proves the equivalence of (a) and (b).

Suppose (b) is given. Then (c) follows from definition. Now we assume U(λ)=U(μ)U(\lambda)=U(\mu) where λ,μλ1,λ2\lambda,\mu\neq\lambda_{1},\lambda_{2} in case when 𝔤=G(3),F(4)\mathfrak{g}=G(3),F(4). For any αΠ0\alpha\in\Pi_{0}, Xαλ+ρ,αX_{\alpha}^{\langle\lambda+\rho,\alpha\rangle} (resp. Xαμ+ρ,αX_{\alpha}^{\langle\mu+\rho,\alpha\rangle}) is the only monomial of the form XαmX_{\alpha}^{m} in U(λ)U(\lambda) (resp. U(μ)U(\mu)) with coefficient (1+eγ)1(1+e^{-\gamma})^{-1}. This gives that (1+eγ)1Xαλ+ρ,α=(1+eγ)1Xαμ+ρ,α(1+e^{-\gamma})^{-1}X_{\alpha}^{\langle\lambda+\rho,\alpha\rangle}=(1+e^{\gamma})^{-1}X_{\alpha}^{\langle\mu+\rho,\alpha\rangle} for all even simple roots α\alpha. So we have λ+ρ,α=μ+ρ,α\langle\lambda+\rho,\alpha\rangle=\langle\mu+\rho,\alpha\rangle for all αΠ0\alpha\in\Pi_{0}. Now arguing as in the proof of \autorefmainlem, we conclude that λ=μ\lambda=\mu. This shows the equivalence of (b) and (c).

If λ,μ\lambda,\mu are among the special weights λ1,λ2\lambda_{1},\lambda_{2} for G(3)G(3) and F(4)F(4), then the above proof works with (1+eγ)1(1+e^{-\gamma})^{-1} replaced by 12(2+eγ)(1+eγ)1\frac{1}{2}(2+e^{-\gamma})(1+e^{-\gamma})^{-1}. ∎

Remark 3.

Notice that omission of the additional condition on the highest weights having same atypicality type in the above lemma poses a technical difficulty. Indeed, suppose λ\lambda is of type γ1\gamma_{1} and μ\mu is of type γ2\gamma_{2} with γ1γ2\gamma_{1}\neq\gamma_{2}. Then from Xλ=XμX^{\lambda}=X^{\mu}, we only get equality of numerators of the term corresponding to wi𝒲w_{i}\in\mathscr{W} in U(λ)U(\lambda) and U(μ)U(\mu). Denominator of the term corresponding to wiw_{i} in U(λ)U(\lambda) is 1+ewiγ11+e^{-w_{i}\gamma_{1}} whereas the same in U(μ)U(\mu) is 1+ewiγ21+e^{-w_{i}\gamma_{2}}. Therefore, in this case Xλ=XμX^{\lambda}=X^{\mu} does not imply U(λ)=U(μ)U(\lambda)=U(\mu), which precludes us from any conclusion about isomorphism of the corresponding representations.

The proof that the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is nonzero for all 𝔤\mathfrak{g} in \autoref x lam to u lam atyp is given in the following subsections. First we have singled out the 𝔰𝔩(m,1)\mathfrak{sl}(m,1) case inasmuch as the proofs for the other types closely resemble it. At the end we have treated the superalgebra 𝔰𝔩(m+1,n+1)\mathfrak{sl}(m+1,n+1).

Notation.

We denote by \mathcal{F} the formal power series algebra [[Zγ:γΦ1]]\mathbb{C}[[Z_{\gamma}\colon\gamma\in\Phi_{1}]], where Zγ=eγZ_{\gamma}=e^{-\gamma}.

6.2.

In this subsection we assume that 𝔤=𝔰𝔩(m,1)\mathfrak{g}=\mathfrak{sl}(m,1) and we show that the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is nonzero. We start by listing down the positive roots of 𝔤\mathfrak{g}:

Φ0+={εiεj:1i<jm},\displaystyle\Phi_{0}^{+}=\{\varepsilon_{i}-\varepsilon_{j}\colon 1\leqslant i<j\leqslant m\},\quad andΦ1+={εiδ1:1im}.\displaystyle\text{and}\quad\Phi_{1}^{+}=\{\varepsilon_{i}-\delta_{1}\colon 1\leqslant i\leqslant m\}. (6.5)
The invariant form on 𝔥\mathfrak{h}^{*} is determined by
(εi,εj)=δij,\displaystyle(\varepsilon_{i},\varepsilon_{j})=\delta_{ij},\quad and(εi,δ1)=0.\displaystyle\text{and}\quad(\varepsilon_{i},\delta_{1})=0. (6.6)

Denote by αi\alpha_{i} (resp. βi\beta_{i}) the even simple root εiεi+1\varepsilon_{i}-\varepsilon_{i+1}, (resp. the isotropic root εiδ1\varepsilon_{i}-\delta_{1}). In this case all odd roots are isotropic. The simple reflection corresponding to αi\alpha_{i} is denoted by sis_{i}. Let λ𝔥\lambda\in\mathfrak{h}^{*} be a dominant integral weight which is singly atypical of type β=βt\beta=\beta_{t}, for some 1tm1\leqslant t\leqslant m.

The following Proposition describes the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda). From the standard Dynkin diagram of 𝔰𝔩(m,1)\mathfrak{sl}(m,1), we see that Π0\Pi_{0} is connected. Moreover, in this case both logU(λ)-\log U(\lambda) and U(λ)U(\lambda) are elements of [[Xα:αΠ0]]\mathcal{F}[[X_{\alpha}\colon\alpha\in\Pi_{0}]]. The proof closely parallels that of \autorefmainprop, so we shall be little concise here. We put K:-k(Π0)=(1)|𝒢|k=1|Π0|(1)kck(Π0)kK\coloneq k(\Pi_{0})=(-1)^{|\mathcal{G}|}\sum_{k=1}^{|\Pi_{0}|}(-1)^{k}\frac{c_{k}(\Pi_{0})}{k}, where ck(Π0)c_{k}(\Pi_{0}) is the number of kk-partitions of Π0\Pi_{0}.

Proposition 6.2.

With the notations as above, the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is given by

{K1+Zβ11+Zβ2for β=β1,K1+Zβm1+Zβm1for β=βm,K(1+Zβt)2(1+Zβt1)(1+Zβt+1)for β=βtβ1,βm.\begin{dcases}\qquad K\frac{1+Z_{\beta_{1}}}{1+Z_{\beta_{2}}}&\quad\text{for $\beta=\beta_{1}$},\\[3.0pt] \qquad K\frac{1+Z_{\beta_{m}}\hfill}{1+Z_{\beta_{m-1}}}&\quad\text{for $\beta=\beta_{m}$},\\[3.0pt] \frac{K(1+Z_{\beta_{t}})^{2}}{(1+Z_{\beta_{t-1}})(1+Z_{\beta_{t+1}})}&\quad\text{for $\beta=\beta_{t}\neq\beta_{1},\beta_{m}$}.\end{dcases}
Proof.

In order to calculate logU(λ)-\log U(\lambda), first we express U(λ)U(\lambda) in the following form:

U(λ)=w𝒲(1)(w)X(λ,w)1+Zwβ=11+Zβ+1w𝒲(1)(w)X(λ,w)1+Zwβ=11+Zβ(1+1w𝒲(1)(w)1+Zβ1+ZwβX(λ,w)ξ)=11+Zβ(1ξ)\displaystyle\begin{aligned} U(\lambda)=\sum_{w\in\mathscr{W}}\frac{(-1)^{\ell(w)}X(\lambda,w)}{1+Z_{w\beta}}&=\frac{1}{1+Z_{\beta}}+\sum_{1\neq w\in\mathscr{W}}\frac{(-1)^{\ell(w)}X(\lambda,w)}{1+Z_{w\beta}}\\ &=\frac{1}{1+Z_{\beta}}\left(1+\underbrace{\sum_{1\neq w\in\mathscr{W}}(-1)^{\ell(w)}\frac{1+Z_{\beta}\hfill}{1+Z_{w\beta}}X(\lambda,w)}_{-\xi}\right)\\ &=\frac{1}{1+Z_{\beta}}(1-\xi)\end{aligned} (6.7)

We have 11+Zβ=(1+Zβ)1=1Zβ+Zβ2\frac{1}{1+Z_{\beta}}=(1+Z_{\beta})^{-1}=1-Z_{\beta}+Z_{\beta}^{2}-\cdots, and 1+Zβ1+Zwβ=(1+Zβ)(1Zwβ+Zwβ2)\frac{1+Z_{\beta}\hfill}{1+Z_{w\beta}}=(1+Z_{\beta})(1-Z_{w\beta}+Z_{w\beta}^{2}-\cdots). As 𝒲\mathscr{W} leaves Φ1\Phi_{1} invariant, both of these elements are members of \mathcal{F} with constant term 1. Like in the typical case, we write ξ=ξ1+ξ2\xi=\xi_{1}+\xi_{2} with

ξ1w(1)(w)1+Zβ1+ZwβX(λ,w)andξ2w(1)(w)1+Zβ1+ZwβX(λ,w),\displaystyle\xi_{1}\coloneqq-\sum_{w\in\mathcal{I}}(-1)^{\ell(w)}\frac{1+Z_{\beta}\hfill}{1+Z_{w\beta}}X(\lambda,w)\quad\text{and}\quad\xi_{2}\coloneqq-\sum_{w\notin\mathcal{I}}(-1)^{\ell(w)}\frac{1+Z_{\beta}\hfill}{1+Z_{w\beta}}X(\lambda,w), (6.8)

where ={1w𝒲:I(w) is totally disconnected}\mathcal{I}=\{1\not=w\in\mathscr{W}\colon I(w)\text{ is totally disconnected}\}, and I(w)I(w) is the subset {αΠ0:α\alpha\in\Pi_{0}\colon\alpha appears in a reduced expression of ww}. Now, logU(λ)=k1ξk/k\log U(\lambda)=\sum_{k\geqslant 1}\nicefrac{{\xi^{k}}}{{k}}. Proceeding as in the proof of \autorefmainprop, we find that the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is same as that in k1ξ1k/k\sum_{k\geqslant 1}\nicefrac{{\xi_{1}^{k}}}{{k}}; therefore it is enough to compute the coefficient of XλX^{\lambda} in ξ1k\xi_{1}^{k}. For any k1k\geqslant 1, we have that

ξ1k=(1)kwi(1)(wi)(1+Zβ)ki=1kX(λ,wi)(1+Zwiβ).\displaystyle{}\xi_{1}^{k}=(-1)^{k}{\sum_{w_{i}\in\mathcal{I}}}(-1)^{\sum\ell(w_{i})}(1+Z_{\beta})^{k}{\prod_{i=1}^{k}}\frac{X(\lambda,w_{i})}{(1+Z_{w_{i}\beta})}. (6.9)

By \autorefkeylem, the product i=1kX(λ,wi)\prod_{i=1}^{k}X(\lambda,w_{i}) equals XλX^{\lambda} only when 𝒬(I(w1),,I(wk))\mathcal{Q}\coloneqq(I(w_{1}),\dots,I(w_{k})) forms a kk-partition of Π0\Pi_{0}.

From the description of the odd roots of 𝔤\mathfrak{g} given in (6.5), we have that for t1,mt\neq 1,m

siβ1={β2for i=1β1for i1\displaystyle s_{i}\beta_{1}=\begin{cases}\beta_{2}&\text{for $i=1$}\\ \beta_{1}&\text{for $i\neq 1$}\end{cases}\quad siβm={βm1for i=mβmfor im\displaystyle\quad s_{i}\beta_{m}=\begin{cases}\beta_{m-1}&\text{for $i=m$}\\ \beta_{m}&\text{for $i\neq m$}\end{cases}\qquad andsiβt={βt+1for i=tβt1for i=t1βtfor it,t1.\displaystyle\text{and}\quad s_{i}\beta_{t}=\begin{cases}\beta_{t+1}&\text{for $i=t$}\\ \beta_{t-1}&\text{for $i=t-1$}\\ \beta_{t}&\text{for $i\neq t,t-1$}.\end{cases}

First we discuss the case where β{β1,βm}\beta\notin\{\beta_{1},\beta_{m}\}. As j=1kI(wj)=Π0\cup_{j=1}^{k}I(w_{j})=\Pi_{0}, it follows that αt1I(wp)\alpha_{t-1}\in I(w_{p}) and αtI(wq)\alpha_{t}\in I(w_{q}) for some p,q{1,,k}p,q\in\{1,\dots,k\} with pqp\neq q. Also by \autorefdef k partition, each I(wj)I(w_{j}) is totally disconnected, whence wjw_{j} is a product of commuting simple reflections. In particular, wp=st1uw_{p}=s_{t-1}u, for some u𝒲u\in\mathscr{W} such that αt1I(u)\alpha_{t-1}\notin I(u). Then wpβ=βt1w_{p}\beta=\beta_{t-1}. Similarly, wqβ=βt+1w_{q}\beta=\beta_{t+1}, and wjβ=βw_{j}\beta=\beta for all jp,qj\neq p,q. This observation immediately implies that for k2k\geqslant 2 we have: (see \autorefcoef sl2,1)

(1+Zβ)ki=1k(1+Zwiβ)=(1+Zβ)2(1+Zβt1)(1+Zβt+1)M(β), say.\displaystyle{}\frac{(1+Z_{\beta})^{k}}{\prod_{i=1}^{k}(1+Z_{w_{i}\beta})}=\frac{(1+Z_{\beta})^{2}}{(1+Z_{\beta_{t-1}})(1+Z_{\beta_{t+1}})}\coloneqq M(\beta),\text{ say.} (6.10)

Since 𝒬\mathcal{Q} is a kk-partition, we get j=1k(wj)=(w1wk)=(w(𝒬))\sum_{j=1}^{k}\ell(w_{j})=\ell(w_{1}\cdots w_{k})=\ell(w(\mathcal{Q})), and (1)(w(𝒬))=(1)|Π0|(-1)^{\ell(w(\mathcal{Q}))}=(-1)^{|\Pi_{0}|}. Note that M(β)M(\beta) is independent of 𝒬\mathcal{Q}, therefore summing over all kk-partitions of Π0\Pi_{0} we find that the coefficient of XλX^{\lambda} in ξ1k/k\nicefrac{{\xi_{1}^{k}}}{{k}} is given by:

𝒬Pk(Π0)(1)k(1)(w(𝒬))M(β)k=(1)k(1)|Π0|ck(Π0)M(β)k.\displaystyle\sum_{\mathcal{Q}\in P_{k}(\Pi_{0})}\frac{(-1)^{k}\cdot(-1)^{\ell(w(\mathcal{Q}))}M(\beta)}{k}=\frac{(-1)^{k}\cdot(-1)^{|\Pi_{0}|}c_{k}(\Pi_{0})M(\beta)}{k}.
Equation (6.10) remains valid for any positive integer k>1k>1, so finally we obtain that the required coefficient in k1ξ1k/k\sum_{k\geqslant 1}\nicefrac{{\xi_{1}^{k}}}{{k}} equals:
M(β)(1)|Π0|k1|Π0|(1)kck(Π0)k=M(β)k(Π0).\displaystyle M(\beta)(-1)^{|\Pi_{0}|}\sum_{k\geqslant 1}^{|\Pi_{0}|}\frac{(-1)^{k}\cdot c_{k}(\Pi_{0})}{k}=M(\beta)k(\Pi_{0}).

This completes the proof for the case when β{β1,βm}\beta\notin\{\beta_{1},\beta_{m}\}.

Now assume β=β1=ε1δ1\beta=\beta_{1}=\varepsilon_{1}-\delta_{1}. Since si(β1)=β1s_{i}(\beta_{1})=\beta_{1} for i1i\neq 1 and s1(β1)=β2s_{1}(\beta_{1})=\beta_{2}, Equation (6.10) now takes the following form:

(1+Zβ)ki=1k(1+Zwiβ)=1+Zβ1+Zβ2N(β), say.\frac{(1+Z_{\beta})^{k}}{\prod_{i=1}^{k}(1+Z_{w_{i}\beta})}=\frac{1+Z_{\beta}\hfill}{1+Z_{\beta_{2}}}\coloneqq N(\beta),\text{ say.}

Arguing as above we conclude that the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is just N(β)k(Π0)N(\beta)k(\Pi_{0}). The case when β=βm=εmδ1\beta=\beta_{m}=\varepsilon_{m}-\delta_{1} is similar. ∎

Remark 4.

In the above proof we have suppressed the fact that Equation (6.10) is not valid for k=1k=1. Indeed, the number of 1-partitions is nonzero only when 𝔤=𝔰𝔩(2,1)\mathfrak{g}=\mathfrak{sl}(2,1). However, in this case we can directly see that the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is (1+Zβ)/(1+Zsβ)\smash{-\nicefrac{{(1+Z_{\beta}\hfill)}}{{(1+Z_{s\beta})}}}, where β\beta is the atypicality type of λ\lambda, and ss being the unique simple reflection in 𝒲\mathscr{W}.

6.3.

In this subsection we assume that 𝔤\mathfrak{g} is of type Cn+1C_{n+1}, i.e., 𝔤=𝔬𝔰𝔭(2,2n)\mathfrak{g}=\mathfrak{osp}(2,2n) for n>1n>1. The root system of 𝔤\mathfrak{g} is given by:

Φ0={±2δi;±δi±δj}ij,Φ1={±ε1±δj}.\displaystyle\Phi_{0}=\{\pm 2\delta_{i};\ \pm\delta_{i}\pm\delta_{j}\}_{i\neq j},\quad\Phi_{1}=\{\pm\varepsilon_{1}\pm\delta_{j}\}.
The invariant form on 𝔥\mathfrak{h}^{*} is determined by:
(δi,δj)=δijand(ε1,δj)=0.\displaystyle(\delta_{i},\delta_{j})=-\delta_{ij}\quad\text{and}\quad(\varepsilon_{1},\delta_{j})=0.
The standard simple system is given by:
Π={δ1δ2,,δn1δn,2δn;ε1δ1}.\displaystyle\Pi=\{\delta_{1}-\delta_{2},\dots,\delta_{n-1}-\delta_{n},2\delta_{n};\ \varepsilon_{1}-\delta_{1}\}.

Denote by αj\alpha_{j} the simple root δjδj+1\delta_{j}-\delta_{j+1}, for 1jn11\leqslant j\leqslant n-1; and we put αn=2δn,αn+1=ε1δ1\alpha_{n}=2\delta_{n},\ \alpha_{n+1}=\varepsilon_{1}-\delta_{1}. We also have that Π0=Π\{αn+1}\Pi_{0}=\Pi\backslash\{\alpha_{n+1}\}, and Φ1+={ε1±δj}\Phi_{1}^{+}=\{\varepsilon_{1}\pm\delta_{j}\} (all are isotropic). The simple reflection corresponding to αj\alpha_{j} is denoted by sjs_{j}.

Let γpε1+δp\gamma_{p}\coloneqq\varepsilon_{1}+\delta_{p} and γpε1δp\gamma_{p}^{\prime}\coloneqq\varepsilon_{1}-\delta_{p}. Consider a dominant integral singly atypical weight λ\lambda of 𝔤\mathfrak{g} of type γ=γp\gamma=\gamma_{p} (the case for γ=γp\gamma=\gamma_{p}^{\prime} is similar). The character formula for V(λ)V(\lambda) is given by Equation (6.1) with β\beta replaced by γ\gamma, and the symbols being understood in 𝔬𝔰𝔭(2,2n)\mathfrak{osp}(2,2n). Note that sjγp=γps_{j}\gamma_{p}=\gamma_{p} for jp,p1j\neq p,p-1; while sp1γp=γp1,spγp=γp+1s_{p-1}\gamma_{p}=\gamma_{p-1},\ s_{p}\gamma_{p}=\gamma_{p+1} and sn1γn=γn1,snγn=γns_{n-1}\gamma_{n}=\gamma_{n-1},\ s_{n}\gamma_{n}=\gamma_{n}^{\prime}. Now arguing as in the proof of \autorefcoef for sl n 1, we obtain the following Proposition.

Proposition 6.3.

With notations as above, the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is given by

{K1+Zγ11+Zγ2for γ=γ1,K(1+Zγn)2(1+Zγn1)(1+Zγn)for γ=γn,K(1+Zγp)2(1+Zγp1)(1+Zγp+1)for γ=γpγ1,γn.\begin{dcases}\qquad K\frac{1+Z_{\gamma_{1}}}{1+Z_{\gamma_{2}}}&\quad\text{for $\gamma=\gamma_{1}$},\\[3.0pt] \frac{K(1+Z_{\gamma_{n}})^{2}}{(1+Z_{\gamma_{n-1}})(1+Z_{\gamma_{n}^{\prime}})}&\quad\text{for $\gamma=\gamma_{n}$},\\[3.0pt] \frac{K(1+Z_{\gamma_{p}})^{2}}{(1+Z_{\gamma_{p-1}})(1+Z_{\gamma_{p+1}})}&\quad\text{for $\gamma=\gamma_{p}\neq\gamma_{1},\gamma_{n}$}.\end{dcases}

6.4.

In this subsection we assume that 𝔤=G(3)\mathfrak{g}=G(3). The roots of 𝔤\mathfrak{g} are expressed in terms of the elements ε1,ε2,ε3,δ𝔥\varepsilon_{1},\varepsilon_{2},\varepsilon_{3},\delta\in\mathfrak{h}^{*}, where ε1+ε2+ε3=0\varepsilon_{1}+\varepsilon_{2}+\varepsilon_{3}=0, and the invariant form is determined by:

(ε1,ε1)=(ε2,ε2)=2(ε1,ε2)=(δ,δ)=2and(εi,δ)=0.\displaystyle(\varepsilon_{1},\varepsilon_{1})=(\varepsilon_{2},\varepsilon_{2})=-2(\varepsilon_{1},\varepsilon_{2})=-(\delta,\delta)=2\quad\text{and}\quad(\varepsilon_{i},\delta)=0.
The standard simple system of 𝔤\mathfrak{g} is given by
Π={α1=ε1,α2=ε2ε1,α3=ε3+δ}.\displaystyle\Pi=\{\alpha_{1}=\varepsilon_{1},\ \alpha_{2}=\varepsilon_{2}-\varepsilon_{1},\ \alpha_{3}=\varepsilon_{3}+\delta\}.
We also have that
Π0={α1,α2} and Φ1+={δ,±εi+δ}.\displaystyle\Pi_{0}=\{\alpha_{1},\alpha_{2}\}\text{\quad and\quad}\Phi_{1}^{+}=\{\delta,\ \pm\varepsilon_{i}+\delta\}.

All positive odd roots are isotropic with the exception of δ\delta. We denote the simple reflection corresponding to αi\alpha_{i} by sis_{i} for i=1,2i=1,2.

Let λλ1,λ2\lambda\neq\lambda_{1},\lambda_{2} be a dominant integral singly atypical weight of atypicality type β\beta. From  [MR3253284], we see that the character of V(λ)V(\lambda) is again given by Equation (6.1) with the symbols in G(3)G(3). Since in this case Π0\Pi_{0} has only two elements which are not mutually orthogonal, the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is same as that in ξ12/2\nicefrac{{\xi_{1}^{2}}}{{2}}, (cf. Equation 6.8). For λ=λ1,λ2\lambda=\lambda_{1},\lambda_{2}, the analogues of Equations (6.7) and (6.9) are respectively:

U(λ)=2+Zβ2(1+Zβ)(1+1w𝒲(1)(w)1+Zβ2+Zβ2+Zwβ1+ZwβX(λ,w)ξ)=2+Zβ2(1+Zβ)(1ξ).\displaystyle U(\lambda)=\frac{\phantom{2}2+Z_{\beta}}{2(1+Z_{\beta})}\left(1+\underbrace{\sum_{1\neq w\in\mathscr{W}}(-1)^{\ell(w)}\frac{1+Z_{\beta}}{2+Z\beta}\cdot\frac{2+Z_{w\beta}}{1+Z_{w\beta}}X(\lambda,w)}_{-\xi}\right)=\frac{\phantom{2}2+Z_{\beta}}{2(1+Z_{\beta})}(1-\xi). (6.7a)
ξ1k=(1)kwi(1)(wi)(1+Zβ2+Zβ)ki=1k(2+Zwiβ1+Zwiβ)X(λ,wi).\displaystyle\xi_{1}^{k}=(-1)^{k}{\sum_{w_{i}\in\mathcal{I}}}(-1)^{\sum\ell(w_{i})}\left(\frac{1+Z_{\beta}}{2+Z_{\beta}}\right)^{k}{\prod_{i=1}^{k}}\left(\frac{2+Z_{w_{i}\beta}}{1+Z_{w_{i}\beta}}\right)X(\lambda,w_{i}). (6.9a)

Now from Equations (6.9) and (6.9a), we obtain the following:

Proposition 6.4.

With notations as above, the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is given by

{(1+Zβ)2(1+Zs1β)(1+Zs2β)for λλ1,λ2,M2(2+Zs1β)(2+Zs2β)(1+Zs1β)(1+Zs2β)for λ=λ1,λ2.\begin{dcases}\phantom{M^{2}}\frac{(1+Z_{\beta})^{2}}{(1+Z_{s_{1}\beta})(1+Z_{s_{2}\beta})}&\quad\text{for $\lambda\neq\lambda_{1},\lambda_{2}$},\\[5.0pt] M^{2}\frac{(2+Z_{s_{1}\beta})(2+Z_{s_{2}\beta})}{(1+Z_{s_{1}\beta})(1+Z_{s_{2}\beta})}&\quad\text{for $\lambda=\lambda_{1},\lambda_{2}$}.\end{dcases}

Here M=1+Zβ2+ZβM=\frac{1+Z_{\beta}}{2+Z_{\beta}}.

6.5.

We now consider 𝔤=F(4)\mathfrak{g}=F(4). The invariant form on 𝔥\mathfrak{h}^{*} is determined by

(εi,εj)=δij,(δ,δ)=3,and (εi,δ)=0for all i,j{1,2,3}.\displaystyle(\varepsilon_{i},\varepsilon_{j})=\delta_{ij},\quad(\delta,\delta)=-3,\quad\text{and\quad}(\varepsilon_{i},\delta)=0\quad\text{for all }i,j\in\{1,2,3\}.
The standard simple system is given by
Π={α1=ε1ε2,α2=ε2ε3,α3=ε3,α4=12(ε1ε2ε3+δ)}.\displaystyle\Pi=\{\alpha_{1}=\varepsilon_{1}-\varepsilon_{2},\ \alpha_{2}=\varepsilon_{2}-\varepsilon_{3},\ \alpha_{3}=\varepsilon_{3},\ \alpha_{4}=\tfrac{1}{2}(-\varepsilon_{1}-\varepsilon_{2}-\varepsilon_{3}+\delta)\}.
We also have that
Π0={α1,α2,α3}andΦ1+={12(±ε1±ε2±ε3+δ)}.\displaystyle\Pi_{0}=\{\alpha_{1},\alpha_{2},\alpha_{3}\}\quad\text{and}\quad\Phi_{1}^{+}=\{\tfrac{1}{2}(\pm\varepsilon_{1}\pm\varepsilon_{2}\pm\varepsilon_{3}+\delta)\}.

Here all the odd roots are isotropic. As before the simple reflection corresponding to αi\alpha_{i} is denoted by sis_{i} for i=1,2,3i=1,2,3. Consider now a dominant integral singly atypical weight λλ1,λ2\lambda\neq\lambda_{1},\lambda_{2} of 𝔤\mathfrak{g} with atypicality type β\beta. Observe that no simple reflection fixes β\beta. The character formula for the 𝔤\mathfrak{g}-module V(λ)V(\lambda) is same as that for G(3)G(3). Here α1\alpha_{1} and α3\alpha_{3} are mutually orthogonal. In this case, we need to compute the coefficient of XλX^{\lambda} in ξ12/2+ξ13/3\nicefrac{{\xi_{1}^{2}}}{{2}}+\nicefrac{{\xi_{1}^{3}}}{{3}}. The coefficient in ξ13/3\nicefrac{{\xi_{1}^{3}}}{{3}} corresponds to the 3-partitions of Π0\Pi_{0}, and it is equal to 2(1+Zβ)3(1+Zs1β)(1+Zs2β)(1+Zs3β)\frac{2(1+Z_{\beta})^{3}}{(1+Z_{s_{1}\beta})(1+Z_{s_{2}\beta})(1+Z_{s_{3}\beta})}. There is only one 2-partition of Π0\Pi_{0}, namely (I1,I2)(I_{1},I_{2}) with I1={α1,α3}I_{1}=\{\alpha_{1},\alpha_{3}\} and I2={α2}I_{2}=\{\alpha_{2}\}. This contributes to the coefficient of XλX^{\lambda} in ξ12/2\nicefrac{{\xi_{1}^{2}}}{{2}} which is equal to (1+Zβ)2(1+Zs1s3β)(1+Zs2β)\frac{-(1+Z_{\beta})^{2}}{(1+Z_{s_{1}s_{3}\beta})(1+Z_{s_{2}\beta})}. The case when λ=λ1,λ2\lambda=\lambda_{1},\lambda_{2} goes over verbatim with the exception that Equation (6.9a) is used instead of Equation (6.9) to compute the coefficient of ξ1k/k\nicefrac{{\xi_{1}^{k}}}{{k}}. Consequently, We have the following:

Proposition 6.5.

With the notations as above, the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is given by

{2(1+Zβ)3(1+Zs1β)(1+Zs2β)(1+Zs3β)(1+Zβ)2(1+Zs1s3β)(1+Zs2β)for λλ1,λ2,2M3(2+Zs1β)(2+Zs2β)(2+Zs3β)(1+Zs1β)(1+Zs2β)(1+Zs3β)M2(2+Zs1s3β)(2+Zs2β)(1+Zs1s3β)(1+Zs2β)for λ=λ1,λ2.\begin{dcases}\phantom{2M^{3}}\frac{2(1+Z_{\beta})^{3}}{(1+Z_{s_{1}\beta})(1+Z_{s_{2}\beta})(1+Z_{s_{3}\beta})}-\frac{(1+Z_{\beta})^{2}}{(1+Z_{s_{1}s_{3}\beta})(1+Z_{s_{2}\beta})}&\quad\text{for $\lambda\neq\lambda_{1},\lambda_{2}$},\\[5.0pt] 2M^{3}\frac{(2+Z_{s_{1}\beta})(2+Z_{s_{2}\beta})(2+Z_{s_{3}\beta})}{(1+Z_{s_{1}\beta})(1+Z_{s_{2}\beta})(1+Z_{s_{3}\beta})}-M^{2}\frac{(2+Z_{s_{1}s_{3}\beta})(2+Z_{s_{2}\beta})}{(1+Z_{s_{1}s_{3}\beta})(1+Z_{s_{2}\beta})}&\quad\text{for $\lambda=\lambda_{1},\lambda_{2}$}.\end{dcases}

Here M=1+Zβ2+ZβM=\frac{1+Z_{\beta}}{2+Z_{\beta}}.

6.6.

We now treat the Lie superalgebra 𝔤=𝔰𝔩(m+1,n+1)\mathfrak{g}=\mathfrak{sl}(m+1,n+1) for n>0n>0. In this case we have Π0=ST\Pi_{0}=S\sqcup T, with S={α1,,αm}S=\{\alpha_{1},\dots,\alpha_{m}\} and T={β1,,βn}T=\{\beta_{1},\dots,\beta_{n}\}, where αi=εiεi+1\alpha_{i}=\varepsilon_{i}-\varepsilon_{i+1} and βj=δjδj+1\beta_{j}=\delta_{j}-\delta_{j+1}. The simple reflection corresponding to αi\alpha_{i} is denoted by sis_{i} and that corresponding to βj\beta_{j} by tjt_{j}. The sets SS and TT are mutually orthogonal. We have that Φ1+={γijεiδj:1im, 1jn}\Phi_{1}^{+}=\{\gamma_{ij}\coloneqq\varepsilon_{i}-\delta_{j}\colon 1\leqslant i\leqslant m,\ 1\leqslant j\leqslant n\}, and all of these roots are isotropic.

Consider now λ𝔥\lambda\in\mathfrak{h}^{*}, a dominant integral singly atypical weight of 𝔤\mathfrak{g} of type γpq\gamma_{pq}. Note that γpq\gamma_{pq} is fixed by all simple reflections except sp1,sp,tq1s_{p-1},s_{p},t_{q-1} and tqt_{q}. Our aim is to calculate the coefficient of XλX^{\lambda} in the expansion of logU(λ)-\log U(\lambda). Since (αi,βj)=0(\alpha_{i},\beta_{j})=0, the contribution of a kk-partition to the coefficient of XλX^{\lambda} is contingent upon how αp1,αp,βq1\alpha_{p-1},\alpha_{p},\beta_{q-1} and βq\beta_{q} appear in that particular partition. Therefore, all possible combinations of kk-partitions have to be treated separately for any k>1k>1. We put Bp,q{αp1,αp,βq1,βq}B_{p,q}\coloneqq\{\alpha_{p-1},\alpha_{p},\beta_{q-1},\beta_{q}\}. Below we list down all distinct combinations of how αp1,αp,βq1\alpha_{p-1},\alpha_{p},\beta_{q-1} and βq\beta_{q} can occur in a kk-partition of Π0\Pi_{0}.

I1I_{1} I2I_{2} Contribution to the coeff. of XλX^{\lambda} in ξ1k/k\nicefrac{{\xi_{1}^{k}}}{{k}}
αp,βq\alpha_{p},\beta_{q} αp1,βq1\alpha_{p-1},\beta_{q-1} f1=(1+Zγpq)2(1+Zγp+1,q+1)(1+Zγp1,q1)f_{1}=\frac{(1+Z_{\gamma_{pq}})^{2}}{(1+Z_{\gamma_{p+1,q+1}})(1+Z_{\gamma_{p-1,q-1}})}
αp,βq1\alpha_{p},\beta_{q-1} αp1,βq\alpha_{p-1},\beta_{q} f2=(1+Zγpq)2(1+Zγp+1,q1)(1+Zγp1,q+1)f_{2}=\frac{(1+Z_{\gamma_{pq}})^{2}}{(1+Z_{\gamma_{p+1,q-1}})(1+Z_{\gamma_{p-1,q+1}})}
Table 1. 2 parts

In \autoref2parts, we describe all possible kk-partitions (I1,,Ik)(I_{1},\dots,I_{k}) where all the elements of Bp,qB_{p,q} occur as a pair. Note that, the number of occurrences of the two types of kk-partitions in \autoref*2parts are same and we denote this number by rk(2)r_{k}^{(2)}. Then from Equation (6.9), it follows that the aggregate contribution of all these partitions to the coefficient of XλX^{\lambda} in ξ1k/k\nicefrac{{\xi_{1}^{k}}}{{k}} is (1)m+n(1)kkrk(2)(f1+f2)(-1)^{m+n}\frac{(-1)^{k}}{k}r_{k}^{(2)}(f_{1}+f_{2}).

In \autoref3parts, we consider all kk-partitions where two elements of Bp,qB_{p,q} occur as a pair and the rest occur separately.

I1I_{1} I2I_{2} I3I_{3} Contribution to the coeff. of XλX^{\lambda} in ξ1k/k\nicefrac{{\xi_{1}^{k}}}{{k}}
αp1,βq1\alpha_{p-1},\beta_{q-1} αp\alpha_{p} βq\beta_{q} g1=(1+Zγpq)3(1+Zγp1,q1)(1+Zγp+1,q)(1+Zγp,q+1)g_{1}=\frac{(1+Z_{\gamma_{pq}})^{3}}{(1+Z_{\gamma_{p-1,q-1}})(1+Z_{\gamma_{p+1,q}})(1+Z_{\gamma_{p,q+1}})}
αp1,βq\alpha_{p-1},\beta_{q} αp\alpha_{p} βq1\beta_{q-1} g2=(1+Zγpq)3(1+Zγp1,q+1)(1+Zγp+1,q)(1+Zγp,q1)g_{2}=\frac{(1+Z_{\gamma_{pq}})^{3}}{(1+Z_{\gamma_{p-1,q+1}})(1+Z_{\gamma_{p+1,q}})(1+Z_{\gamma_{p,q-1}})}
αp,βq1\alpha_{p},\beta_{q-1} αp1\alpha_{p-1} βq\beta_{q} g3=(1+Zγpq)3(1+Zγp+1,q1)(1+Zγp1,q)(1+Zγp,q+1)g_{3}=\frac{(1+Z_{\gamma_{pq}})^{3}}{(1+Z_{\gamma_{p+1,q-1}})(1+Z_{\gamma_{p-1,q}})(1+Z_{\gamma_{p,q+1}})}
αp,βq\alpha_{p},\beta_{q} αp1\alpha_{p-1} βq1\beta_{q-1} g4=(1+Zγpq)3(1+Zγp+1,q+1)(1+Zγp1,q)(1+Zγp,q1)g_{4}=\frac{(1+Z_{\gamma_{pq}})^{3}}{(1+Z_{\gamma_{p+1,q+1}})(1+Z_{\gamma_{p-1,q}})(1+Z_{\gamma_{p,q-1}})}
Table 2. 3 parts

As before, we denote the common number of all such kk-partitions by rk(3)r_{k}^{(3)}. Then the aggregate contribution of all these partitions to the coefficient of XλX^{\lambda} is (1)m+n(1)kkrk(3)(g1+g2+g3+g4)(-1)^{m+n}\frac{(-1)^{k}}{k}r_{k}^{(3)}(g_{1}+g_{2}+g_{3}+g_{4}).

I1I_{1} I2I_{2} I3I_{3} I4I_{4} Contribution to the coeff. of XλX^{\lambda} in ξ1k/k\nicefrac{{\xi_{1}^{k}}}{{k}}
αp1\alpha_{p-1} βq1\beta_{q-1} αp\alpha_{p} βq\beta_{q} h1=(1+Zγpq)4(1+Zγp1,q)(1+Zγp,q1)(1+Zγp+1,q)(1+Zγp,q+1)h_{1}=\frac{(1+Z_{\gamma_{pq}})^{4}}{(1+Z_{\gamma_{p-1,q}})(1+Z_{\gamma_{p,q-1}})(1+Z_{\gamma_{p+1,q}})(1+Z_{\gamma_{p,q+1}})}
Table 3. 4 parts

Finally, in \autoref4parts, we describe all kk-partitions of Π0\Pi_{0} where each element of Bp,qB_{p,q} occurs separately. Let rk(4)r_{k}^{(4)} denote the total number of all such partitions. Then the aggregate contribution of all the kk-partitions depicted in \autoref*4parts to the coefficient of XλX^{\lambda} is given by (1)m+n(1)kkrk(4)h1(-1)^{m+n}\frac{(-1)^{k}}{k}r_{k}^{(4)}h_{1}. As both (αp1,αp)(\alpha_{p-1},\alpha_{p}) and (βq1,βq)(\beta_{q-1},\beta_{q}) are nonzero, by definition of a kk-partition we conclude that tables 1, 2 and 3 exhaust all possible combinations.

We denote by AkA_{k} the coefficient of XλX^{\lambda} in ξ1k/k\nicefrac{{\xi_{1}^{k}}}{{k}}. It now follows that

Ak=(1)m+n(1)kk[rk(2)(f1+f2)+rk(3)(g1+g2+g3+g4)+rk(4)h1].\displaystyle A_{k}=(-1)^{m+n}\cdot\frac{(-1)^{k}}{k}[r_{k}^{(2)}(f_{1}+f_{2})+r_{k}^{(3)}(g_{1}+g_{2}+g_{3}+g_{4})+r_{k}^{(4)}h_{1}]. (6.11)

With the notations just introduced, we find that the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is given by AA2++Am+nA\coloneqq A_{2}+\dots+A_{m+n}, and it is a \mathbb{C}-linear combination of f1,f2,g1,g2,g3,g4,h1f_{1},f_{2},g_{1},g_{2},g_{3},g_{4},h_{1}. It remains to prove that A0A\neq 0, which is contention of the next proposition.

Proposition 6.6.

With the notations as above, the coefficient of XλX^{\lambda} in logU(λ)-\log U(\lambda) is nonzero, i.e. A0A\neq 0.

Proof.

It is not difficult to see that the set {f1,f2,g1,g2,g3,g4,h1}\{f_{1},f_{2},g_{1},g_{2},g_{3},g_{4},h_{1}\} is linearly independent over \mathbb{C}, therefore it suffices to show that the coefficient of f1f_{1} in AA is nonzero. From Equation (6.11), we find that this coefficient is (1)m+nk2(1)kkrk(2)(-1)^{m+n}\sum_{k\geqslant 2}\frac{(-1)^{k}}{k}r_{k}^{(2)}, where the sum runs up to k=m+nk=m+n. Recall that Pk(Π0)P_{k}(\Pi_{0}) is the set of all kk-partitions of Π0\Pi_{0}. Let QkQ_{k} be the subset of Pk(Π0)P_{k}(\Pi_{0}) defined by

Qk{𝒬=(I1,,Ik)Pk(Π0):αp1,βq1Ii1 and αp,βqIi2 for some 1i1i2k}.Q_{k}\coloneqq\{\mathcal{Q}=(I_{1},\dots,I_{k})\in P_{k}(\Pi_{0})\colon\alpha_{p-1},\beta_{q-1}\in I_{i_{1}}\text{ and }\alpha_{p},\beta_{q}\in I_{i_{2}}\text{ for some }1\leqslant i_{1}\neq i_{2}\leqslant k\}.

By definition, rk(2)r_{k}^{(2)} is the cardinality of QkQ_{k}. Notice that rk(2)=0r_{k}^{(2)}=0 if k>m+n2k>m+n-2. Since both {αp1,βq1}\{\alpha_{p-1},\beta_{q-1}\} and {αp,βq}\{\alpha_{p},\beta_{q}\} occur as pairs in any 𝒬Qk\mathcal{Q}\in Q_{k}, they can be treated as a single symbol. Let νp1,q1\nu_{p-1,q-1} (resp. νp,q\nu_{p,q}) stand for the pairs {αp1,βq1}\{\alpha_{p-1},\beta_{q-1}\} (resp. {αp,βq}\{\alpha_{p},\beta_{q}\}). Now we consider the following set:

Gp,q{α1,,αp2,α^p1,α^p,,αm,β1,,βq2,β^q1,β^q,,βn}{νp1,q1,νp,q},G_{p,q}\coloneqq\{\alpha_{1},\dots,\alpha_{p-2},\hat{\alpha}_{p-1},\hat{\alpha}_{p},\dots,\alpha_{m},\beta_{1},\dots,\beta_{q-2},\hat{\beta}_{q-1},\hat{\beta}_{q},\dots,\beta_{n}\}\sqcup\{\nu_{p-1,q-1},\nu_{p,q}\},

where hat denotes omission. Then |Gp,q|=m+n2|G_{p,q}|=m+n-2. We intend to construct a graph 𝒢p,q\mathcal{G}_{p,q} having Gp,qG_{p,q} as the set of vertices. The vertices in Gp,q\{νp1,q1,νp,q}G_{p,q}\backslash\{\nu_{p-1,q-1},\nu_{p,q}\} being a subset of Π0\Pi_{0}, are joined in accordance with the usual rule (see 3.1); αi\alpha_{i} is joined to νp1,q1\nu_{p-1,q-1} (resp. to νp,q\nu_{p,q}) by an edge if (αi,αp1)0(\alpha_{i},\alpha_{p-1})\neq 0 (resp. (αi,αp)0(\alpha_{i},\alpha_{p})\neq 0). Similarly, βj\beta_{j} is joined to νp1,q1\nu_{p-1,q-1} (resp. to νp,q\nu_{p,q}) by an edge if (βj,βq1)0(\beta_{j},\beta_{q-1})\neq 0 (resp. (βj,βq)0(\beta_{j},\beta_{q})\neq 0). As both (αp1,αp)(\alpha_{p-1},\alpha_{p}) and (βq1,βq)(\beta_{q-1},\beta_{q}) are nonzero, νp1,q1\nu_{p-1,q-1} and νp,q\nu_{p,q} is connected by an edge. These rules define the graph 𝒢p,q\mathcal{G}_{p,q}. We see that νp1,q1\nu_{p-1,q-1} is joined to both αp2\alpha_{p-2} and βq2\beta_{q-2}, while νp,q\nu_{p,q} is connected to νp1,q1\nu_{p-1,q-1}, βq+1\beta_{q+1} and αp+1\alpha_{p+1}. This means that 𝒢p,q\mathcal{G}_{p,q} is a tree.

Evidently, rk(2)=Pk(𝒢p,q)r_{k}^{(2)}=P_{k}(\mathcal{G}_{p,q}) for 2km+n22\leqslant k\leqslant m+n-2, and P1(𝒢p,q)=0P_{1}(\mathcal{G}_{p,q})=0. This implies that

(1)m+n2k=2m+n2(1)kkrk(2)=(1)|Gp,q|k=2|Gp,q|(1)kkPk(𝒢p,q)=k(𝒢p,q).(-1)^{m+n-2}\sum\nolimits_{k=2}^{m+n-2}\frac{(-1)^{k}}{k}r_{k}^{(2)}=(-1)^{|G_{p,q}|}\sum\nolimits_{k=2}^{|G_{p,q}|}\frac{(-1)^{k}}{k}P_{k}(\mathcal{G}_{p,q})=k(\mathcal{G}_{p,q}).

Since 𝒢p,q\mathcal{G}_{p,q} is a tree, by  [MR2980495]*Corollary 1 we conclude that k(𝒢p,q)=1k(\mathcal{G}_{p,q})=1. ∎

Summarizing propositions 6.2, 6.3, 6.4, 6.5 and 6.6, we obtain the following important Lemma.

Lemma 6.7.

Let 𝔤\mathfrak{g} be any one among the Lie superalgebras mentioned in \autoref x lam to u lam atyp. Let λ𝔥\lambda\in\mathfrak{h}^{*} be dominant integral singly atypical of type say γ\gamma. Then the coefficient of XλX^{\lambda} in the expansion of logU(λ)-\log U(\lambda) is nonzero.

The next theorem will be used in the proof of the main result of this section (cf. \autorefpermfac).

Theorem 6.8.

Let 𝔤\mathfrak{g} be as in \autoref x lam to u lam atyp. Let ν1,,νr;μ1,,μs\nu_{1},\dots,\nu_{r};\mu_{1},\dots,\mu_{s} be dominant integral singly atypical weights of 𝔤\mathfrak{g} of atypicality type γ\gamma. Suppose we are given the equality

U(ν1)U(νr)=U(μ1)U(μs).\displaystyle U(\nu_{1})\cdots U(\nu_{r})=U(\mu_{1})\cdots U(\mu_{s}). (6.12)

Then r=sr=s, and there is a permutation σ\sigma of {1,,r}\{1,\dots,r\} such that U(νk)=U(μσ(k))U(\nu_{k})=U(\mu_{\sigma(k)}).

Proof.

First we assume that none of these weights are either λ1\lambda_{1} or λ2\lambda_{2} in case when 𝔤=G(3),F(4)\mathfrak{g}=G(3),F(4). As in Equation (6.7), we express U(νi)U(\nu_{i}) in the form U(νi)=11+Zγ(1ξi)U(\nu_{i})=\frac{1}{1+Z_{\gamma}}(1-\xi_{i}), and similarly U(μj)=11+Zγ(1ζj)U(\mu_{j})=\frac{1}{1+Z_{\gamma}}(1-\zeta_{j}). Taking negative logarithm on both sides of Equation (6.12) yields:

i=1rlog(1ξi)=j=1slog(1ζj)+(sr)log(1+Zγ).\displaystyle\sum_{i=1}^{r}-\log(1-\xi_{i})=\sum_{j=1}^{s}-\log(1-\zeta_{j})+(s-r)\log(1+Z_{\gamma}). (6.13)
After applying ΘΠ0\Theta_{\Pi_{0}} (see 4.2) on both sides we obtain (cf. \autoreflowestdeg)
i=1r(LXνi+ϑi)=j=1s(LXμj+φj),\displaystyle\sum_{i=1}^{r}(L\cdot X^{\nu_{i}}+\vartheta_{i})=\sum_{j=1}^{s}(L\cdot X^{\mu_{j}}+\varphi_{j}),

where ϑi\vartheta_{i} (resp. φj\varphi_{j}) is the sum of monomials of total degree bigger than degXνi\deg X^{\nu_{i}} (resp. XμjX^{\mu_{j}}), and LL is the coefficient of XνiX^{\nu_{i}} and XμjX^{\mu_{j}} for all 1ir, 1js1\leqslant i\leqslant r,\ 1\leqslant j\leqslant s. Since all the weights are of same atypicality type, LL does not depend on the weights, and it is nonzero by \autorefnonzero lem atyp.

Among all the XνiX^{\nu_{i}}’s, we choose one having lowest possible degree. Without any loss of generality, say this element is Xν1X^{\nu_{1}}; hence, this monomial must appear in the right hand side of the above equation. By minimality of degree, we must have Xν1=XμkX^{\nu_{1}}=X^{\mu_{k}}, for some 1ks1\leqslant k\leqslant s. It follows that U(ν1)=U(μk)U(\nu_{1})=U(\mu_{k}) by \autoref x lam to u lam atyp.

To show r=sr=s, we assume on contrary that r>sr>s. Proceeding via induction, we see that all the factors on the right side get canceled with some in the left side, forcing r=sr=s, as a product of U(νi)U(\nu_{i})’s is not identically 1.

Now for 𝔤=G(3) and F(4)\mathfrak{g}=G(3)\text{ and }F(4), we assume that λ1\lambda_{1} and λ2\lambda_{2} appear collectively mm times on the left hand side, and nn times on the right hand side of Equation (6.12). In this case, Equation (6.13) will take the following form (cf. Equation 6.7a):

i=1rlog(1ξi)=j=1slog(1ζj)+(sr)log(2+2Zγ)+(mn)log(2+Zγ).\sum_{i=1}^{r}-\log(1-\xi_{i})=\sum_{j=1}^{s}-\log(1-\zeta_{j})+(s-r)\log(2+2Z_{\gamma})+(m-n)\log(2+Z_{\gamma}).

The rest of the proof now goes over verbatim. ∎

We are now ready to prove the main result of this section.

Theorem 6.9.

Let 𝔤\mathfrak{g} be any Lie superalgebra among 𝔰𝔩(m+1,n+1),𝔬𝔰𝔭(2,2n),G(3)\mathfrak{sl}(m+1,n+1),\ \mathfrak{osp}(2,2n),\ G(3), or F(4)F(4). Suppose we are given the following isomorphism of 𝔤\mathfrak{g}-modules

V(ν1)V(νr)V(μ1)V(μs),V(\nu_{1})\otimes\dots\otimes V(\nu_{r})\cong V(\mu_{1})\otimes\dots\otimes V(\mu_{s}),

where each νi\nu_{i}’s and μj\mu_{j}’s are dominant integral singly atypical of type γ\gamma. Then r=sr=s, and there is a permutation σ\sigma of {1,,r}\{1,\dots,r\} such that V(νk)V(μσ(k))V(\nu_{k})\cong V(\mu_{\sigma(k)}) for all kk.

Proof.

After taking characters on both sides of the above isomorphism and proceeding as in the proof of \autorefThm:tnsrpdt, we get that

U(ν1)U(νr)Drs=U(μ1)U(μs),where D=αΦ1+(1+eα)αΦ0+(1eα).U(\nu_{1})\cdots U(\nu_{r})D^{r-s}=U(\mu_{1})\cdots U(\mu_{s}),\quad\text{where $D=\tfrac{\prod_{\alpha\in\Phi^{+}_{1}}(1+e^{-\alpha})}{\prod_{\alpha\in\Phi^{+}_{0}}(1-e^{-\alpha})}$}.

By \autorefthm: perm of ulmbda atyp, we conclude that r=sr=s and U(νk)=U(μσ(k))U(\nu_{k})=U(\mu_{\sigma(k)}) for some permutation σ\sigma of {1,,r}\{1,\dots,r\} and 1kr1\leqslant k\leqslant r. This implies by \autoref x lam to u lam atyp that νk=μσ(k)\nu_{k}=\mu_{\sigma(k)}, whence it follows that V(νk)V(μσ(k))V(\nu_{k})\cong V(\mu_{\sigma(k)}). ∎

References