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

\DefineSimpleKey

bibarxiv \DefineSimpleKeybibhow

Polynomial families of quantum semisimple coajoint orbits via deformed quantum enveloping algebras

Mao Hoshino Department of Mathematical Sciences, The University of Tokyo
Komaba 3-8-1, Tokyo 153-8914, Japan
mhoshino@ms.u-tokyo.ac.jp
Abstract.

We construct a polynomial family of semisimple left module categories over the representation category of the Drinfeld-Jimbo deformation, with the fusion rule of the representation category of each Levi subalgebra. In this construction we perform a kind of generalized parabolic induction using a deformed quantum enveloping algebra, whose definition depends on an arbitrary choice of a positive system and corresponds to De Commer’s definition for the standard positive system. These algebras define a sheaf of algebras on the toric variety associated to the root system, which contains the moduli of equivariant Poisson brackets. This fact finally produces the family of 22-cocycle. We also obtain a comparison theorem between our module categories and module categories induced from our construction for intermediate Levi subalgebras. The construction of deformed quantum enveloping algebras and the comparison theorem are discussed in the integral setting of Lusztig’s sense.

Key words and phrases:
quantum group, deformation quantization, representation theory, toric variety
2020 Mathematics Subject Classification:
Primary 17B37, Secondary 17B10
This work was supported by JSPS KAKENHI Grant Number JP23KJ0695 and WINGS-FoPM Program at the University of Tokyo.

1. Introduction

Deformation quantization is one of rigorous formulations of quantization, which naturally arose in quantum physics. This physical background is a reason why we typically consider an algebra of suitable functions on a space with a Poisson structure as an algebra to be quantized, though it is possible to introduce the notion of deformation quantization for general Poisson algebras. Actually finding and classifying deformation quantizations of such function algebras is a natural problem to consider and has been attracting many physicists and mathematicians. One of most successful results was accomplished by Kontsevich ([MR2062626]), who proved the existence of deformation quantizations of each Poisson manifold and their classification theorem. See [MR1321655] for developments before Kontsevich’s work, including Fedosov’s construction for deformation quantizations of symplectic manifolds ([MR1293654]).

Theory of quantum groups is also a relevant and important branch of mathematical physics around quantization. Not surprisingly, deformation quantization deeply relates to it. For instance a Drinfeld-Jimbo deformation Uq(𝔤)U_{q}(\mathfrak{g}) of a complex semisimple Lie algebra 𝔤\mathfrak{g} is a kind of dual of a quantization 𝒪q(G)\mathcal{O}_{q}(G) of the function algebra on the complex semisimple Lie group GG with respect to the so-called standard Poisson bracket. This fact invokes a further topic: equivariant quantization of actions of complex semisimple Lie groups as actons of Uq(𝔤)U_{q}(\mathfrak{g}), or equivalently coactions of 𝒪q(G)\mathcal{O}_{q}(G), on quantized spaces. One of important researches on this direction is Podleś’s study [MR0919322] on equivariant quantizations of 22-sphere. Nowadays his work can be understood as an example of quantized compact symmetric spaces, initiated by Letzter ([MR1913438]).

Nontheless, in this paper, we would like to explore an alternative approach to generalizing the Podleś quantum spheres, focusing on quantum semisimple coadjoint orbits, or equivalently quantum flag manifolds. The latter terminology is used in papers considering the algebra of smooth or continuous functions on partial flag manifolds with the action of the compact real form KK of GG, for instance ([MR3376147, MR1697598]). In this case a \ast-structure also seems to be taken in consideration. On the other hand, the former is used by researchers considering L\GL\backslash G with a Levi subgroup LL and the algebra 𝒪(L\G)\mathcal{O}(L\backslash G) of algebraic functions on L\GL\backslash G, usually which is not equipped with any \ast-structure.

Quantization of such actions has already been studied by several researchers. In [MR1817512] Donin gives an explicit description of the moduli of all Poisson structures compatible with the action and also shows that such deformation quantizaions can be classified by the set of formal paths on the moduli space. We explain this fact in more detail. Fix a Cartan subalgebra 𝔥\mathfrak{h} of the Lie algebra 𝔤\mathfrak{g} and consider the associated root system RR with a positive system R+R^{+}. The set of simple roots is denoted by Δ\Delta. We also choose a subset SS of Δ\Delta, from which we can generate a closed subsystem RSR_{S} and a Levi subgroup LSL_{S}. In this setting, Donin shows that the set of Poisson brackets on LS\GL_{S}\backslash G compatible with the action of the Poisson-Lie group GG can be identified with the set XLS\GX_{L_{S}\backslash G} of φ=(φα)αRRSRRS\varphi=(\varphi_{\alpha})_{\alpha\in R\setminus R_{S}}\in\mathbb{C}^{R\setminus R_{S}} satisfying the following relations:

  1. (i)

    φα=φα\varphi_{-\alpha}=-\varphi_{\alpha} for αRRS\alpha\in R\setminus R_{S},

  2. (ii)

    φαφβ+1=φα+β(φα+φβ)\varphi_{\alpha}\varphi_{\beta}+1=\varphi_{\alpha+\beta}(\varphi_{\alpha}+\varphi_{\beta}) when α,β,α+βRRS\alpha,\beta,\alpha+\beta\in R\setminus R_{S},

  3. (iii)

    φα=φβ\varphi_{\alpha}=\varphi_{\beta} when α,βRRS\alpha,\beta\in R\setminus R_{S} and αβRS\alpha-\beta\in R_{S}.

He also shows the existence of a holomorphic family {𝒪h,φ(LS\G)}φXLS\G\{\mathcal{O}_{h,\varphi}(L_{S}\backslash G)\}_{\varphi\in X_{L_{S}\backslash G}} of equivariant deformation quantizations and that any equivariant deformation quantization is equivalent to 𝒪h,φ(h)(LS\G)\mathcal{O}_{h,\varphi(h)}(L_{S}\backslash G), where φ(h)hRRS\varphi(h)\in\mathbb{C}\llbracket h\rrbracket^{R\setminus R_{S}} satisfies the same relations (i), (ii), (iii). We would like to emphasize here that his result was establised as a consequence of some vanishing theorems of relevant cohomological obstructions. In particular any explicit construction of those deformations was not presented, at least in the paper.

If φXLS\G\varphi\in X_{L_{S}\backslash G} is generic in the sense that φα1\varphi_{\alpha}\neq-1 for all αRRS\alpha\in R\setminus R_{S}, Enriquez-Etingof-Marshall ([MR2126485, MR2349621]) and Mudrov ([MR2304470]) give an explicit construction of corresponding equivariant deformation quantizations. Namely they construct left Rephf𝔤\operatorname{\mathrm{Rep}}^{\mathrm{f}}_{h}\mathfrak{g}-module categories using the parabolic induction twisted by hλ/h\mathbb{C}\llbracket h\rrbracket_{\lambda/h}, where λ𝔥\lambda\in\mathfrak{h}^{*} is an appropriate weight. It should be mentioned that this approach also appears in the investigation of equivariant quantization with respect to the trivial Poisson-Lie structure ([MR1952112, MR2182701, MR2141466]).

De Commer ([MR3208147]) also followed these pieces of research, with a slight and significant modification, which enables us to discuss quantizations with respect to φ\varphi with φα1\varphi_{\alpha}\neq-1 for αR+RS+\alpha\in R^{+}\setminus R_{S}^{+}. There he uses a deformed quantum enveloping algebra instead of the parabolic induction twisted by hλ/h\mathbb{C}\llbracket h\rrbracket_{\lambda/h}, which cannot be formulated in this case as the generic case. He also discusses some operator algebraic aspects, motivated by construction of real semisimple quantum groups as locally compact quantum groups.

Contents of this paper

The main purpose is to give an explicit way to construct an algebraic family of 22-cocyle twists of left Rephf𝔤\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{g}-module categories Rephf𝔩S\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{l}_{S}, parametrized by XLS\GX_{L_{S}\backslash G}. To achieve this we define a deformed quantum enveloping algebra (Definition 3.6), which is inspired by De Commer’s work. Namely such a deformed quantum enveloping algebra is defined whenever we fix a positive system R0+R_{0}^{+} and a character on 2Q02Q_{0}^{-}, the cone generated by 2R0+-2R_{0}^{+}. If R0+=R+R_{0}^{+}=R^{+}, this definition corresponds to De Commer’s definition. It is worth to remark here that the PBW theorem still holds for such algebras, which involves some observations on quantum PBW vectors in the usual quantum enveloping algebras. This enables us to investigate the twisted parabolic induction defined in Subsection 3.3. Any construction in Section 3 is conducted in the integral setting of Lusztig’s sense.

In Section 4, we discuss a kind of semisimplicity of the twisted parabolic induction in the formal setting. Using this fact we construct left Rephf𝔤\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{g}-module categories which are 22-cocycle twists of Rephf𝔩S\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{l}_{S}. We also construct quantum semisimple coadjoint orbits using these 22-cocycles and determine their semi-classical limits.

In Section 5, we discuss the compatibility of the 22-cocycles constructed from different choices of R0+R_{0}^{+}. At the beginning we discuss an immersion of the moduli space XLS\GX_{L_{S}\backslash G} to the toric scheme XRX_{R} associated to the root system R+R^{+}. Then we construct a sheaf 𝒰q,XR(𝔤)\mathscr{U}_{q,X_{R}}(\mathfrak{g}{}^{\sim}) of deformed quantum enveloping algebras on XRX_{R}, which also induces a sheaf 𝒰q,XLS\G(𝔤)\mathscr{U}_{q,X_{L_{S}\backslash G}}(\mathfrak{g}{}^{\sim}). This fact immediately implies that the 22-cocycles constructed in Section 44 can be glued up to a 22-cocycle with coefficients in 𝒪(XLS\G)\mathcal{O}(X_{L_{S}\backslash G}).

Finally, in Section 6, we compare quantum semisimple coadjoint orbits constructed in Section 4 and those induced from quantum semisimple coadjoint orbits of intermediate Levi subgroups. Namely we show that an induction functor becomes a left Rephf𝔤\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{g}-module functor in the specific situation.

2. Preliminaries

In this paper 𝔤\mathfrak{g} and 𝔥\mathfrak{h} denote a complex semisimple Lie algebra and its Cartan subalgebra respectively. The associated set of roots is denoted by RR, which forms a root system with a bilinear form (,)(\textendash,\textendash) on 𝔥\mathfrak{h}^{*} induced by a normalized Killing form so that (α,α)=2(\alpha,\alpha)=2 for a short root α\alpha. The reflection with respect to αR\alpha\in R is denoted by sαs_{\alpha}. The associated Weyl group is denoted by WW.

We fix a positive system R+R^{+}, which induces a triangular decomposition 𝔤=𝔫𝔥𝔫+\mathfrak{g}=\mathfrak{n}^{-}\oplus\mathfrak{h}\oplus\mathfrak{n}^{+} and defines a set Δ={ε1,ε2,,εr}\Delta=\{\varepsilon_{1},\varepsilon_{2},\dots,\varepsilon_{r}\} of simple roots. Here rr, the number of simple roots, is the rank of 𝔤\mathfrak{g}. We also set N=|R+|N=\lvert R^{+}\rvert. The set of reflections with respect to simple roots generates WW, and defines the length function :W0\ell\colon W\longrightarrow\mathbb{Z}_{\geq 0}. The unique longest element is denoted by w0w_{0}, whose length is NN.

We set dα,α,aijd_{\alpha},\alpha^{\vee},a_{ij} as follows:

dα=(α,α)2,α=dα1α,aij=(δi,δj).d_{\alpha}=\frac{(\alpha,\alpha)}{2},\quad\alpha^{\vee}=d_{\alpha}^{-1}\alpha,\quad a_{ij}=(\delta_{i}^{\vee},\delta_{j}).

The fundamental weights, which are dual to (εi)i(\varepsilon_{i}^{\vee})_{i} with respect to (,)(\textendash,\textendash), are denoted by ϖi\varpi_{i}. The root lattice QQ (resp. PP) is \mathbb{Z}-linear span of Δ\Delta (resp. (ϖi)i(\varpi_{i})_{i}). We also use the positive cone Q+Q^{+} and P+P^{+}:

Q+\displaystyle Q^{+} =0ε1+0ε2++0εr,\displaystyle=\mathbb{Z}_{\geq 0}\varepsilon_{1}+\mathbb{Z}_{\geq 0}\varepsilon_{2}+\cdots+\mathbb{Z}_{\geq 0}\varepsilon_{r},
P+\displaystyle P^{+} =0ϖ1+0ϖ2++0ϖr.\displaystyle=\mathbb{Z}_{\geq 0}\varpi_{1}+\mathbb{Z}_{\geq 0}\varpi_{2}+\cdots+\mathbb{Z}_{\geq 0}\varpi_{r}.

We usually replace εi\varepsilon_{i} by the symbol ii when εi\varepsilon_{i} appears as a subscript. For instance we use si,di,Kis_{i},d_{i},K_{i} instead of using sεi,dεi,Kεis_{\varepsilon_{i}},d_{\varepsilon_{i}},K_{\varepsilon_{i}}.

For SΔS\subset\Delta, we define RSR_{S} as RspanSR\cap\operatorname{\mathrm{span}}_{\mathbb{Z}}S and RS+R_{S}^{+} as R+RSR^{+}\cap R_{S}. The number of positive roots in RSR_{S} is denote by NSN_{S}. Then we have Lie subalgebras 𝔫S±=αRS+𝔤±α\mathfrak{n}_{S}^{\pm}=\bigoplus_{\alpha\in R_{S}^{+}}\mathfrak{g}_{\pm\alpha}, 𝔟S±=𝔥𝔫S±\mathfrak{b}_{S}^{\pm}=\mathfrak{h}\oplus\mathfrak{n}_{S}^{\pm}, 𝔩S=𝔫S𝔥𝔫S+\mathfrak{l}_{S}=\mathfrak{n}_{S}^{-}\oplus\mathfrak{h}\oplus\mathfrak{n}_{S}^{+}, 𝔲S±=αR+RS+𝔤±α\mathfrak{u}_{S}^{\pm}=\bigoplus_{\alpha\in R^{+}\setminus R_{S}^{+}}\mathfrak{g}_{\pm\alpha}, 𝔭S=𝔩S𝔲S+\mathfrak{p}_{S}=\mathfrak{l}_{S}\oplus\mathfrak{u}_{S}^{+}. We also define QS,PS,QS+,PS+Q_{S},P_{S},Q_{S}^{+},P_{S}^{+} for each subset SΔS\subset\Delta.

For qq-integers, we use the following symbols:

qα=qdα,[n]q=qnqnqq1,[n]q!=[1]q[2]q[n]q,[nk]q=[n]q![k]q![nk]q!q_{\alpha}=q^{d_{\alpha}},\quad[n]_{q}=\frac{q^{n}-q^{-n}}{q-q^{-1}},\quad[n]_{q}!=[1]_{q}[2]_{q}\cdots[n]_{q},\quad\begin{bmatrix}n\\ k\end{bmatrix}_{q}=\frac{[n]_{q}!}{[k]_{q}![n-k]_{q}!}

A quantum commutator is defined for vectors in suitable algebras which admits a weight space decomposition as stated in the next section:

[x,y]q=xyq(wtx,wty)yx.\displaystyle[x,y]_{q}=xy-q^{-(\operatorname{\mathrm{wt}}{x},\operatorname{\mathrm{wt}}{y})}yx.

Finally we prepare some notations on a multi-index Λ=(λi)i0n\Lambda=(\lambda_{i})_{i}\in\mathbb{Z}_{\geq 0}^{n}.

  • |Λ|=iλi\lvert\Lambda\rvert=\sum_{i}\lambda_{i}.

  • suppΛ={iλi0}\operatorname{\mathrm{supp}}\Lambda=\{i\mid\lambda_{i}\neq 0\}.

  • Λ(k,l)defsuppΛ{k+1,k+2,,l1}\Lambda\subset(k,l)\overset{\text{def}}{\Longleftrightarrow}\operatorname{\mathrm{supp}}\Lambda\subset\{k+1,k+2,\cdots,l-1\}. For an interval II, like [k,l][k,l], ΛI\Lambda\subset I is defined in a similar way.

  • Λ<kdefΛ(0,k)\Lambda<k\overset{\text{def}}{\Longleftrightarrow}\Lambda\subset(0,k). Similarly Λk,Λ>k,Λk\Lambda\leq k,\Lambda>k,\Lambda\geq k are defined.

  • Λα=iλiαi\Lambda\cdot\alpha=\sum_{i}\lambda_{i}\alpha_{i} for a sequence (αi)i(\alpha_{i})_{i} of vectors.

  • xΛ=x1λ1x2λ2xnλnx^{\Lambda}=x_{1}^{\lambda_{1}}x_{2}^{\lambda_{2}}\cdots x_{n}^{\lambda_{n}} for a sequence (xi)i(x_{i})_{i} in a (possibly non-commutative) ring.

2.1. Quantum enveloping algebras of semisimple Lie algebras

Basically we refer the convension in [MR4162277] and [MR1492989]. Though our main purpose is to construct deformation quantizations, we will present a definition of the deformed quantum enveloping algebra in the integral form for generality. Therefore begin with the definition of the Drinfeld-Jimbo deformation over (s)\mathbb{Q}(s), the function field of one variables over \mathbb{Q} with a deformation parameter q=s2Lq=s^{2L} where LL is the smallest positive integer such that (λ,μ)L1(\lambda,\mu)\in L^{-1}\mathbb{Z} for any λ,μP\lambda,\mu\in P. We set qr:=s2Lrq^{r}:=s^{2Lr} for r(2L)1r\in(2L)^{-1}\mathbb{Z} and qi=qdiq_{i}=q^{d_{i}}. The Drinfeld-Jimbo deformation of 𝔤\mathfrak{g} is a Hopf algebra Uq(𝔤)U_{q}(\mathfrak{g}) generated by Ei,Fi,KλE_{i},F_{i},K_{\lambda} for 1ir1\leq i\leq r and λP\lambda\in P, with relations

K0=1,\displaystyle K_{0}=1, KλEiKλ1=q(λ,εi)Ei,\displaystyle K_{\lambda}E_{i}K_{\lambda}^{-1}=q^{(\lambda,\varepsilon_{i})}E_{i}, [Ei,Fj]=δijKiKi1qiqi1,\displaystyle[E_{i},F_{j}]=\delta_{ij}\frac{K_{i}-K_{i}^{-1}}{q_{i}-q_{i}^{-1}},
KλKμ=Kλ+μ,\displaystyle K_{\lambda}K_{\mu}=K_{\lambda+\mu}, KλFiKλ1=q(λ,εi)Fi,\displaystyle K_{\lambda}F_{i}K_{\lambda}^{-1}=q^{-(\lambda,\varepsilon_{i})}F_{i},

and the quantum Serre relations:

k=01aij(1)k[1aijk]qiEi1aijkEjEik=0,\displaystyle\sum_{k=0}^{1-a_{ij}}(-1)^{k}\begin{bmatrix}1-a_{ij}\\ k\end{bmatrix}_{q_{i}}E_{i}^{1-a_{ij}-k}E_{j}E_{i}^{k}=0,
k=01aij(1)k[1aijk]qiFi1aijkFjFik=0.\displaystyle\sum_{k=0}^{1-a_{ij}}(-1)^{k}\begin{bmatrix}1-a_{ij}\\ k\end{bmatrix}_{q_{i}}F_{i}^{1-a_{ij}-k}F_{j}F_{i}^{k}=0.

The coproduct Δ\Delta, the antipode SS and the counit ε\varepsilon are given as follows on the generators:

Δ(Kλ)=KλKλ,\displaystyle\Delta(K_{\lambda})=K_{\lambda}\otimes K_{\lambda}, S(Kλ)=Kλ1,\displaystyle S(K_{\lambda})=K_{\lambda}^{-1}, ε(Kλ)=1,\displaystyle\varepsilon(K_{\lambda})=1,
Δ(Ei)=EiKi+1Ei,\displaystyle\Delta(E_{i})=E_{i}\otimes K_{i}+1\otimes E_{i}, S(Ei)=EiKi1,\displaystyle S(E_{i})=-E_{i}K_{i}^{-1}, ε(Ei)=0,\displaystyle\varepsilon(E_{i})=0,
Δ(Fi)=Fi1+Ki1Fi,\displaystyle\Delta(F_{i})=F_{i}\otimes 1+K_{i}^{-1}\otimes F_{i}, S(Fi)=KiFi,\displaystyle S(F_{i})=-K_{i}F_{i}, ε(Fi)=0.\displaystyle\varepsilon(F_{i})=0.

Next we introduce some subalgebras of Uq(𝔤)U_{q}(\mathfrak{g}). The most fundamental ones are Uq(𝔫+),Uq(𝔫)U_{q}(\mathfrak{n}^{+}),U_{q}(\mathfrak{n}^{-}) and Uq(𝔥)U_{q}(\mathfrak{h}), which are genereted by Ei,Fi,KλE_{i},F_{i},K_{\lambda} respectively. These allow us to decompose Uq(𝔤)U_{q}(\mathfrak{g}) into the tensor products Uq(𝔫±)Uq(𝔥)Uq(𝔫)U_{q}(\mathfrak{n}^{\pm})\otimes U_{q}(\mathfrak{h})\otimes U_{q}(\mathfrak{n}^{\mp}) via the multiplication maps. We also use Uq(𝔟±)U_{q}(\mathfrak{b}^{\pm}) for the subalgebras generated by Uq(𝔥)U_{q}(\mathfrak{h}) and Uq(𝔫±)U_{q}(\mathfrak{n}^{\pm}) respectively. Note that Uq(𝔟±)U_{q}(\mathfrak{b}^{\pm}) are Hopf subalgebras of Uq(𝔤)U_{q}(\mathfrak{g}).

For a Uq(𝔥)U_{q}(\mathfrak{h})-module MM and vMv\in M, we say that vv is a weight vector of weight λP\lambda\in P when Kμv=q(μ,λ)vK_{\mu}v=q^{(\mu,\lambda)}v for all μP\mu\in P. In this case λ\lambda is denoted by wtv\operatorname{\mathrm{wt}}v. The submodule of elements of weight λ\lambda is denoted by MλM_{\lambda}. To consider the weight of an element of Uq(𝔤)U_{q}(\mathfrak{g}), we regard Uq(𝔤)U_{q}(\mathfrak{g}) as a Uq(𝔥)U_{q}(\mathfrak{h})-module by the restriction of the left adjoint action xy=x(1)yS(x(2))x\triangleright y=x_{(1)}yS(x_{(2)}).

Next we describe the braid group action on Uq(𝔤)U_{q}(\mathfrak{g}) and the quantum PBW bases. At first we have an algebra automorphism 𝒯i\mathcal{T}_{i} on Uq(𝔤)U_{q}(\mathfrak{g}) for each δiΔ\delta_{i}\in\Delta, which satisfies

𝒯i(Kλ)=Ksi(λ),𝒯i(Ei)=KiFi,𝒯i(Fi)=EiKi1\displaystyle\mathcal{T}_{i}(K_{\lambda})=K_{s_{i}(\lambda)},\quad\mathcal{T}_{i}(E_{i})=-K_{i}F_{i},\quad\mathcal{T}_{i}(F_{i})=-E_{i}K_{i}^{-1}

and other formulae in [MR4162277, Theorem 3.58] which determine 𝒯i\mathcal{T}_{i} uniquely.

Then the family (𝒯i)i(\mathcal{T}_{i})_{i} satisfies the Coxeter relations and defines a braid group action on Uq(𝔤)U_{q}(\mathfrak{g}). Especially we have 𝒯w\mathcal{T}_{w} for each wWw\in W, which is given by 𝒯w=𝒯i1𝒯i2𝒯i(w)\mathcal{T}_{w}=\mathcal{T}_{i_{1}}\mathcal{T}_{i_{2}}\cdots\mathcal{T}_{i_{\ell(w)}} where w=si1si2si(w)w=s_{i_{1}}s_{i_{2}}\cdots s_{i_{\ell(w)}} is a reduced expression.

This action produces PBW bases of Uq(𝔤)U_{q}(\mathfrak{g}). Let w0w_{0} be the longest element in WW and fix its reduced expression w0=s𝒊=si1si2siNw_{0}=s_{\boldsymbol{i}}=s_{i_{1}}s_{i_{2}}\cdots s_{i_{N}}, where 𝒊=(i1,i2,,iN)\boldsymbol{i}=(i_{1},i_{2},\dots,i_{N}). Then each αR+\alpha\in R^{+} has a unique positive integer kNk\leq N with α=αk𝒊:=si1si2sik1(εik)\alpha=\alpha^{\boldsymbol{i}}_{k}:=s_{i_{1}}s_{i_{2}}\cdots s_{i_{k-1}}(\varepsilon_{i_{k}}). Finally we set E𝒊,αE_{\boldsymbol{i},\alpha} and F𝒊,αF_{\boldsymbol{i},\alpha}, the quantum root vectors, as follows:

E𝒊,α\displaystyle E_{\boldsymbol{i},\alpha} =E𝒊,k:=𝒯si1si2sik1(Eik)=𝒯si1𝒯si2𝒯sik1(Eik),\displaystyle=E_{\boldsymbol{i},k}:=\mathcal{T}_{s_{i_{1}}s_{i_{2}}\cdots s_{i_{k-1}}}(E_{i_{k}})=\mathcal{T}_{s_{i_{1}}}\mathcal{T}_{s_{i_{2}}}\cdots\mathcal{T}_{s_{i_{k-1}}}(E_{i_{k}}),
F𝒊,α\displaystyle F_{\boldsymbol{i},\alpha} =F𝒊,k:=𝒯si1si2sik1(Fik)=𝒯si1𝒯si2𝒯sik1(Fik).\displaystyle=F_{\boldsymbol{i},k}:=\mathcal{T}_{s_{i_{1}}s_{i_{2}}\cdots s_{i_{k-1}}}(F_{i_{k}})=\mathcal{T}_{s_{i_{1}}}\mathcal{T}_{s_{i_{2}}}\cdots\mathcal{T}_{s_{i_{k-1}}}(F_{i_{k}}).

Though these elements depend on 𝒊\boldsymbol{i}, we still have an analogue of the Poincaré-Birkhoff-Witt theorem in Uq(𝔤)U_{q}(\mathfrak{g}) i.e. {F𝒊ΛKμE𝒊Λ+}Λ±,μ\{F_{\boldsymbol{i}}^{\Lambda^{-}}K_{\mu}E_{\boldsymbol{i}}^{\Lambda^{+}}\}_{\Lambda^{\pm},\mu} forms a basis of Uq(𝔤)U_{q}(\mathfrak{g}). Each element of this family is called a quantum PBW vector.

2.2. The Drinfeld pairing and the quantum Killing form

Next we recall the Drinfeld pairing and the quantum Killing form. The Drinfeld pairing is a unique bilinear form τ:Uq(𝔟+)×Uq(𝔟)k\tau\colon U_{q}(\mathfrak{b}^{+})\times U_{q}(\mathfrak{b}^{-})\longrightarrow k with

τ(Kλ,Kμ)=q(λ,μ),τ(Ei,Kλ)=0=τ(Fi,Kμ),τ(Ei,Fj)=δijqiqi1\displaystyle\tau(K_{\lambda},K_{\mu})=q^{-(\lambda,\mu)},\quad\tau(E_{i},K_{\lambda})=0=\tau(F_{i},K_{\mu}),\quad\tau(E_{i},F_{j})=-\frac{\delta_{ij}}{q_{i}-q_{i}^{-1}}

and the following conditions:

τ(xx,y)=τ(x,y(1))τ(x,y(2)),\displaystyle\tau(xx^{\prime},y)=\tau(x,y_{(1)})\tau(x^{\prime},y_{(2)}),\quad τ(x,1)=ε(x),\displaystyle\tau(x,1)=\varepsilon(x),\quad
τ(x,yy)=τ(x(1),y)τ(x(2),y),\displaystyle\tau(x,yy^{\prime})=\tau(x_{(1)},y^{\prime})\tau(x_{(2)},y),\quad τ(1,y)=ε(y).\displaystyle\tau(1,y)=\varepsilon(y).

Here Δ(x)=x(1)x(2)\Delta(x)=x_{(1)}\otimes x_{(2)} is Sweedler’s notation.

The PBW bases are orthogonal with respect to this form:

τ(E𝒊Λ+,F𝒊Λ)=δΛ+Λk=1N(1)λk+qαk𝒊λk+(λk+1)/2[λk+]qαk𝒊!(qαk𝒊qαk𝒊1)λk+.\displaystyle\tau(E_{\boldsymbol{i}}^{\Lambda^{+}},F_{\boldsymbol{i}}^{\Lambda^{-}})=\delta_{\Lambda^{+}\Lambda^{-}}\prod_{k=1}^{N}(-1)^{\lambda_{k}^{+}}q_{\alpha^{\boldsymbol{i}}_{k}}^{\lambda^{+}_{k}(\lambda^{+}_{k}-1)/2}\frac{[\lambda^{+}_{k}]_{q_{\alpha^{\boldsymbol{i}}_{k}}!}}{(q_{\alpha^{\boldsymbol{i}}_{k}}-q^{-1}_{\alpha^{\boldsymbol{i}}_{k}})^{\lambda^{+}_{k}}}.

The following formulae will be used frequently:

(1) τ(xKμ,yKν)=q(μ,ν)τ(x,y),\displaystyle\tau(xK_{\mu},yK_{\nu})=q^{-(\mu,\nu)}\tau(x,y), xUq(𝔫+),yUq(𝔫),\displaystyle x\in U_{q}(\mathfrak{n}^{+}),y\in U_{q}(\mathfrak{n}^{-}),
(2) xy=τ(S(x(1)),y(1))y(2)x(2)τ(x(3),y(3)),\displaystyle xy=\tau(S(x_{(1)}),y_{(1)})y_{(2)}x_{(2)}\tau(x_{(3)},y_{(3)}), xUq(𝔟+),yUq(𝔟).\displaystyle x\in U_{q}(\mathfrak{b}^{+}),y\in U_{q}(\mathfrak{b}^{-}).

The quantum Killing form κ:Uq(𝔤)×Uq(𝔤)k\kappa\colon U_{q}(\mathfrak{g})\times U_{q}(\mathfrak{g})\longrightarrow k is given by the following formula:

κ(XKμS1(X+),YKνS(Y+))=q(μ,ν)/2τ(X+,Y)τ(Y+,X).\displaystyle\kappa(X^{-}K_{\mu}S^{-1}(X^{+}),Y^{-}K_{\nu}S(Y^{+}))=q^{(\mu,\nu)/2}\tau(X^{+},Y^{-})\tau(Y^{+},X^{-}).

Ad-invariance, an important property of the classical Killing form, still holds in this case:

κ(ZX,Y)=κ(X,S(Z)Y)\kappa(Z\triangleright X,Y)=\kappa(X,S(Z)\triangleright Y)

2.3. The restricted integral form

At the end of this section we recall the restricted integral form Uq𝒜(𝔤)U_{q}^{\mathcal{A}}(\mathfrak{g}). Set 𝒜:=[s2,s2](s)\mathcal{A}:=\mathbb{Z}[s^{2},s^{-2}]\subset\mathbb{Q}(s). Then Uq(𝔤)U_{q}(\mathfrak{g}) has a Hopf subalgebra Uq𝒜(𝔤)U_{q}^{\mathcal{A}}(\mathfrak{g}) over 𝒜\mathcal{A}, called the restricted integral form, which is generated by the following elements:

Ei(r)=1[r]qi!Eir,Fi(r)=1[r]qi!Fir,Kλ,[Ki;0]qi=KiKi1qiqi1.\displaystyle E_{i}^{(r)}=\frac{1}{[r]_{q_{i}}!}E_{i}^{r},\quad F_{i}^{(r)}=\frac{1}{[r]_{q_{i}}!}F_{i}^{r},\quad K_{\lambda},\quad[K_{i};0]_{q_{i}}=\frac{K_{i}-K_{i}^{-1}}{q_{i}-q_{i}^{-1}}.

The upper (resp. lower) triangular subalgebra Uq𝒜(𝔫±)U_{q}^{\mathcal{A}}(\mathfrak{n}^{\pm}) are defined similarly to the non-integral case. They are free over 𝒜\mathcal{A} since the following elements form a basis:

E𝒊(Λ)=E𝒊,1(λ1)E𝒊,2(λ2)E𝒊,N(λN),F𝒊(Λ)=F𝒊,1(λ1)F𝒊,2(λ2)F𝒊,N(λN)\displaystyle E_{\boldsymbol{i}}^{(\Lambda)}=E_{\boldsymbol{i},1}^{(\lambda_{1})}E_{\boldsymbol{i},2}^{(\lambda_{2})}\cdots E_{\boldsymbol{i},N}^{(\lambda_{N})},\quad F_{\boldsymbol{i}}^{(\Lambda)}=F_{\boldsymbol{i},1}^{(\lambda_{1})}F_{\boldsymbol{i},2}^{(\lambda_{2})}\cdots F_{\boldsymbol{i},N}^{(\lambda_{N})}

We also have the Cartan subalgebra and the Borel subalgebras, defined as Uq𝒜(𝔥)=Uq(𝔥)Uq𝒜(𝔤)U_{q}^{\mathcal{A}}(\mathfrak{h})=U_{q}(\mathfrak{h})\cap U_{q}^{\mathcal{A}}(\mathfrak{g}) and Uq𝒜(𝔟±)=Uq𝒜(𝔫±)Uq𝒜(𝔤)U_{q}^{\mathcal{A}}(\mathfrak{b}^{\pm})=U_{q}^{\mathcal{A}}(\mathfrak{n}^{\pm})U_{q}^{\mathcal{A}}(\mathfrak{g}) respectively.

Finally, for a commutative 𝒜\mathcal{A}-algebra kk, we define Uqk(𝔤)U_{q}^{k}(\mathfrak{g}) as k𝒜Uq𝒜(𝔤)k\otimes_{\mathcal{A}}U_{q}^{\mathcal{A}}(\mathfrak{g}). we also define Uqk(𝔫±),Uqk(𝔥)U_{q}^{k}(\mathfrak{n}^{\pm}),U_{q}^{k}(\mathfrak{h}) and Uqk(𝔟±)U_{q}^{k}(\mathfrak{b}^{\pm}) in a similar way.

3. Deformed quantum enveloping algebras

3.1. More on quantum PBW bases

Before defining deformed quantum enveloping algebras, we collect some technical facts on quantum PBW vectors in the usual quantum enveloping algebra. For a reduced expression s𝒊s_{\boldsymbol{i}} of w0w_{0}, we set E´𝒊,α=E𝒊,αKα1,K´μ=Kμ\acute{E}_{\boldsymbol{i},\alpha}=E_{\boldsymbol{i},\alpha}K_{\alpha}^{-1},\acute{K}_{\mu}=K_{\mu} and F´𝒊,α=F𝒊,α\acute{F}_{\boldsymbol{i},\alpha}=F_{\boldsymbol{i},\alpha}.

Remark 3.1.

In the following, various expansion formulae with respect to the modified quantum PBW vectors F´𝒊(Λ)KμE´𝒊(Λ+)\acute{F}_{\boldsymbol{i}}^{(\Lambda^{-})}K_{\mu}\acute{E}_{\boldsymbol{i}}^{(\Lambda^{+})} are stated. We sometimes apply them to other generators, like E^𝒊,α,K^μ,F^𝒊,α\hat{E}_{\boldsymbol{i},\alpha},\hat{K}_{\mu},\hat{F}_{\boldsymbol{i},\alpha}. Moreover we use their variations: for instance, using E^𝒊Λ+\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}} instead of E^𝒊(Λ+)\hat{E}_{\boldsymbol{i}}^{(\Lambda^{+})}. These application can be validated by considering the “easiest” case, in which these generators are scalar multiple of the original generators. Then the other cases are reduced to this case by changing the base ring.

The following lemma is fundamental in this subsection.

Lemma 3.2.

Let s𝐢s_{\boldsymbol{i}} be a reduced expression of w0w_{0} and set w=si1si2sikw=s_{i_{1}}s_{i_{2}}\cdots s_{i_{k}} for a fixed kk with 1kN1\leq k\leq N. Then there is another reduced expression s𝐣s_{\boldsymbol{j}} which satisfies the following:

𝒯w(E´𝒋,l)\displaystyle\mathcal{T}_{w}(\acute{E}_{\boldsymbol{j},l}) ={E´𝒊,k+l(1lNk)qαl(Nk)𝒊2F´𝒊,l(Nk)K´2αl(Nk)𝒊(Nk<lN),\displaystyle=\begin{cases}\acute{E}_{\boldsymbol{i},k+l}&(1\leq l\leq N-k)\\ -q_{\alpha_{l-(N-k)}^{\boldsymbol{i}}}^{-2}\acute{F}_{\boldsymbol{i},l-(N-k)}\acute{K}_{2\alpha^{\boldsymbol{i}}_{l-(N-k)}}&(N-k<l\leq N),\end{cases}
𝒯w(F´𝒋,l)\displaystyle\mathcal{T}_{w}(\acute{F}_{\boldsymbol{j},l}) ={F´𝒊,k+l(1lNk)E´𝒊,l(Nk)(Nk<lN),\displaystyle=\begin{cases}\acute{F}_{\boldsymbol{i},k+l}&(1\leq l\leq N-k)\\ -\acute{E}_{\boldsymbol{i},l-(N-k)}&(N-k<l\leq N),\end{cases}
Proof.

At first we assume k=1k=1. Set ε=siNsiN1si2(εi1)\varepsilon=s_{i_{N}}s_{i_{N-1}}\cdots s_{i_{2}}(\varepsilon_{i_{1}}). Then this is a simple root and s𝒋=si2si3siNsεs_{\boldsymbol{j}}=s_{i_{2}}s_{i_{3}}\cdots s_{i_{N}}s_{\varepsilon} is a reduced expression of w0w_{0}, which has the required property. Actually we can see from the construction that 𝒯i1(E´𝒋,l)=E´𝒊,l+1\mathcal{T}_{i_{1}}(\acute{E}_{\boldsymbol{j},l})=\acute{E}_{\boldsymbol{i},l+1} and 𝒯i1(F´𝒋,l)=F´𝒊,l+1\mathcal{T}_{i_{1}}(\acute{F}_{\boldsymbol{j},l})=\acute{F}_{\boldsymbol{i},l+1} for 1lN11\leq l\leq N-1.

For the case l=Nl=N, E´𝒋,N=E´i1\acute{E}_{\boldsymbol{j},N}=\acute{E}_{i_{1}} and F´𝒋,N=F´i1\acute{F}_{\boldsymbol{j},N}=\acute{F}_{i_{1}} hold since αN𝒋=εi1\alpha_{N}^{\boldsymbol{j}}=\varepsilon_{i_{1}} is a simple root. Hence we have

𝒯i1(E´𝒋,N)\displaystyle\mathcal{T}_{i_{1}}(\acute{E}_{\boldsymbol{j},N}) =𝒯i1(E´i1)=Ki1F´i1Ki1=qα1𝒊2F´𝒊,1K´2α1𝒊,\displaystyle=\mathcal{T}_{i_{1}}(\acute{E}_{i_{1}})=-K_{i_{1}}\acute{F}_{i_{1}}K_{i_{1}}=-q_{\alpha^{\boldsymbol{i}}_{1}}^{-2}\acute{F}_{\boldsymbol{i},1}\acute{K}_{2\alpha^{\boldsymbol{i}}_{1}},
𝒯i1(F´𝒋,N)\displaystyle\mathcal{T}_{i_{1}}(\acute{F}_{\boldsymbol{j},N}) =𝒯i1(F´i1)=E´i1=E´𝒊,1.\displaystyle=\mathcal{T}_{i_{1}}(\acute{F}_{i_{1}})=-\acute{E}_{i_{1}}=-\acute{E}_{\boldsymbol{i},1}.

Now we show the statement by induction on kk. Assume the statement with k=k0k=k_{0} holds for all reduced expressions. Then we can take 𝒊\boldsymbol{i^{\prime}} for 𝒊\boldsymbol{i} and k=k0k=k_{0}, which automatically satisfies i1=ik+1i^{\prime}_{1}=i_{k+1}. Then we can take 𝒋\boldsymbol{j} for 𝒊\boldsymbol{i^{\prime}} and k=1k=1, which is the required one for 𝒊\boldsymbol{i} and k=k0+1k=k_{0}+1. ∎

Next we investigate quantum commutators, whose definition in this paper is [x,y]q:=xyq(wtx,wty)yx[x,y]_{q}:=xy-q^{-(\operatorname{\mathrm{wt}}x,\operatorname{\mathrm{wt}}y)}yx for weight vectors x,yx,y.

The following is an easy consequence of a well-known result, called the Levendörskii-Soibelman relation in literature.

Proposition 3.3 (c.f. [MR1116413, Proposition 5.5.2]).

For 1k<lN1\leq k<l\leq N, we have the following expansion with CΛ±𝒜C_{\Lambda}^{\pm}\in\mathcal{A}:

[E´𝒊,l,E´𝒊,k]q=Λ(k,l)Λα𝒊=αk𝒊+αl𝒊CΛ+E´𝒊(Λ),[F´𝒊,l,F´𝒊,k]q=Λ(k,l)Λα𝒊=αk𝒊+αl𝒊CΛF´𝒊(Λ),\displaystyle[\acute{E}_{\boldsymbol{i},l},\acute{E}_{\boldsymbol{i},k}]_{q}=\sum_{\begin{subarray}{c}\Lambda\subset(k,l)\\ \Lambda\cdot\alpha^{\boldsymbol{i}}=\alpha^{\boldsymbol{i}}_{k}+\alpha^{\boldsymbol{i}}_{l}\end{subarray}}C_{\Lambda}^{+}\acute{E}_{\boldsymbol{i}}^{(\Lambda)},\quad[\acute{F}_{\boldsymbol{i},l},\acute{F}_{\boldsymbol{i},k}]_{q}=\sum_{\begin{subarray}{c}\Lambda\subset(k,l)\\ \Lambda\cdot\alpha^{\boldsymbol{i}}=\alpha^{\boldsymbol{i}}_{k}+\alpha^{\boldsymbol{i}}_{l}\end{subarray}}C_{\Lambda}^{-}\acute{F}_{\boldsymbol{i}}^{(\Lambda)},

Combining this with Lemma 3.2, we can obtain a “mixed” version of these formulae.

Proposition 3.4.

Fix a reduced expression s𝐢s_{\boldsymbol{i}} of the longest element w0w_{0}. and positive integers 1k<lN1\leq k<l\leq N.

  1. (i)

    The following 𝒜\mathcal{A}-submodules are closed under multiplication:

    span𝒜{F´𝒊ΛE´𝒊Λ+Λ+k<lΛ},\displaystyle\operatorname{\mathrm{span}}_{\mathcal{A}}\{\acute{F}_{\boldsymbol{i}}^{\Lambda^{-}}\acute{E}_{\boldsymbol{i}}^{\Lambda^{+}}\mid\Lambda^{+}\leq k<l\leq\Lambda^{-}\},
    span𝒜{E´𝒊Λ+K´μF´𝒊ΛμP,Λl<kΛ+}.\displaystyle\operatorname{\mathrm{span}}_{\mathcal{A}}\{\acute{E}_{\boldsymbol{i}}^{\Lambda^{+}}\acute{K}_{\mu}\acute{F}_{\boldsymbol{i}}^{\Lambda^{-}}\mid\mu\in P,\,\Lambda^{-}\leq l<k\leq\Lambda^{+}\}.
  2. (ii)

    The latter subalgebra coincides with the following:

    span𝒜{F´𝒊(Λ)K´μE´𝒊(Λ+)μP,Λl<kΛ+}.\displaystyle\operatorname{\mathrm{span}}_{\mathcal{A}}\{\acute{F}_{\boldsymbol{i}}^{(\Lambda^{-})}\acute{K}_{\mu}\acute{E}_{\boldsymbol{i}}^{(\Lambda^{+})}\mid\mu\in P,\Lambda^{-}\leq l<k\leq\Lambda^{+}\}.
  3. (iii)

    We have the following expansion with Ck,l(Λ,Λ+),Cl,k(Λ,Λ+)𝒜C_{k,l}(\Lambda^{-},\Lambda^{+}),C_{l,k}(\Lambda^{-},\Lambda^{+})\in\mathcal{A}:

    (3) [E´𝒊,k,F´𝒊,l]q\displaystyle[\acute{E}_{\boldsymbol{i},k},\acute{F}_{\boldsymbol{i},l}]_{q} =Λ+<k<l<Λ,μ:=αk𝒊Λ+α𝒊=αl𝒊Λα𝒊Q+Ck,l(Λ,Λ+)F´𝒊(Λ)E´𝒊(Λ+),\displaystyle=\sum_{\begin{subarray}{c}\Lambda^{+}<k<l<\Lambda^{-},\\ \mu:=\alpha_{k}^{\boldsymbol{i}}-\Lambda^{+}\cdot\alpha^{\boldsymbol{i}}=\alpha_{l}^{\boldsymbol{i}}-\Lambda^{-}\cdot\alpha^{\boldsymbol{i}}\in Q^{+}\end{subarray}}C_{k,l}(\Lambda^{-},\Lambda^{+})\acute{F}_{\boldsymbol{i}}^{(\Lambda^{-})}\acute{E}_{\boldsymbol{i}}^{(\Lambda^{+})},
    (4) [E´𝒊,l,F´𝒊,k]q\displaystyle[\acute{E}_{\boldsymbol{i},l},\acute{F}_{\boldsymbol{i},k}]_{q} =Λ<k<l<Λ+μ:=αk𝒊Λ+α𝒊=αl𝒊Λα𝒊Q+Cl,k(Λ,Λ+)F´𝒊(Λ)K´2μ1E´𝒊(Λ+).\displaystyle=\sum_{\begin{subarray}{c}\Lambda^{-}<k<l<\Lambda^{+}\\ \mu:=\alpha_{k}^{\boldsymbol{i}}-\Lambda^{+}\cdot\alpha^{\boldsymbol{i}}=\alpha_{l}^{\boldsymbol{i}}-\Lambda^{-}\cdot\alpha^{\boldsymbol{i}}\in Q^{+}\end{subarray}}C_{l,k}(\Lambda^{-},\Lambda^{+})\acute{F}_{\boldsymbol{i}}^{(\Lambda^{-})}\acute{K}_{2\mu}^{-1}\acute{E}_{\boldsymbol{i}}^{(\Lambda^{+})}.
Proof.

(i) This follows from Lemma 3.2 immediately.

(ii) Note that the latter subalgebra in (i) is generated by (E´𝒊,n)nl(\acute{E}_{\boldsymbol{i},n})_{n\leq l}, (F´𝒊,m)mk(\acute{F}_{\boldsymbol{i},m})_{m\geq k} and (K´μ)μP(\acute{K}_{\mu})_{\mu\in P}. Hence it suffices to show that the following 𝒜\mathcal{A}-submodules are closed under multiplication for all 0ml0\leq m\leq l:

m:=span𝒜{F´𝒊(Λ1)E´𝒊(Λ+)K´μF´𝒊(Λ2)μP,Λ2m<Λ1l<kΛ+}.\displaystyle\mathcal{I}_{m}:=\operatorname{\mathrm{span}}_{\mathcal{A}}\{\acute{F}_{\boldsymbol{i}}^{(\Lambda^{-}_{1})}\acute{E}_{\boldsymbol{i}}^{(\Lambda^{+})}\acute{K}_{\mu}\acute{F}_{\boldsymbol{i}}^{(\Lambda^{-}_{2})}\mid\mu\in P,\Lambda^{-}_{2}\leq m<\Lambda^{-}_{1}\leq l<k\leq\Lambda^{+}\}.

If m=lm=l, there is nothing to prove. For the downward induction step, assume that the statement holds for a fixed mm. It suffices to show m1\mathcal{I}_{m-1} is invariant under the right multiplication by E´𝒊,m(r)\acute{E}_{\boldsymbol{i},m}^{(r)} for all r1r\geq 1. There is nothing to prove when r=0r=0. For a general rr, we can see m1E´𝒊,m(r)m1E´𝒊,m(r1)\mathcal{I}_{m-1}\acute{E}_{\boldsymbol{i},m}^{(r)}\subset\mathcal{I}_{m-1}\acute{E}_{\boldsymbol{i},m}^{(r-1)} using Lemma 3.2 and Proposition 3.3. Hence an induction argument works.

(iii) Take 𝒋\boldsymbol{j} using Lemma 3.2 for 𝒊\boldsymbol{i} and kk. Then we have

E´𝒊,kF´𝒊,lq(αk𝒊,αl𝒊)F´𝒊,lE´𝒊,k=q(αk𝒊,αl𝒊)𝒯w(F´𝒋,lkF´𝒋,nq(αlk𝒋,αn𝒋)F´𝒋,nF´𝒋,lk).\displaystyle\acute{E}_{\boldsymbol{i},k}\acute{F}_{\boldsymbol{i},l}-q^{-(\alpha^{\boldsymbol{i}}_{k},\alpha^{\boldsymbol{i}}_{l})}\acute{F}_{\boldsymbol{i},l}\acute{E}_{\boldsymbol{i},k}=q^{(\alpha_{k}^{\boldsymbol{i}},\alpha_{l}^{\boldsymbol{i}})}\mathcal{T}_{w}(\acute{F}_{\boldsymbol{j},l-k}\acute{F}_{\boldsymbol{j},n}-q^{(\alpha_{l-k}^{\boldsymbol{j}},\alpha_{n}^{\boldsymbol{j}})}\acute{F}_{\boldsymbol{j},n}\acute{F}_{\boldsymbol{j},l-k}).

Now Proposition 3.3 leads us to the desired expansion without the constraint on μ\mu. A similar argument works for the other case.

Finally we show the positivity of μ\mu. Since Δ(Fi)=Fi1+Ki1Fi\Delta(F_{i})=F_{i}\otimes 1+K_{i}^{-1}\otimes F_{i} holds, Δ(F𝒊,α)\Delta(F_{\boldsymbol{i},\alpha}) is an 𝒜\mathcal{A}-linear combination of F𝒊(Λ1)KΛ2α𝒊1F𝒊(Λ2)F_{\boldsymbol{i}}^{(\Lambda_{1})}K_{\Lambda_{2}\cdot\alpha^{\boldsymbol{i}}}^{-1}\otimes F_{\boldsymbol{i}}^{(\Lambda_{2})} with Λ1α𝒊+Λ2α𝒊=α\Lambda_{1}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{2}\cdot\alpha^{\boldsymbol{i}}=\alpha. Using this fact twice we can see that (Δid)Δ(F´𝒊,α)(\Delta\otimes\operatorname{\mathrm{id}})\Delta(\acute{F}_{\boldsymbol{i},\alpha}) is an 𝒜\mathcal{A}-linear combination of F𝒊(Λ1)KΛ2α𝒊+Λ3α𝒊1F´𝒊(Λ2)K´Λ3α𝒊1F𝒊(Λ3)F_{\boldsymbol{i}}^{(\Lambda_{1})}K_{\Lambda_{2}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{3}\cdot\alpha^{\boldsymbol{i}}}^{-1}\otimes\acute{F}_{\boldsymbol{i}}^{(\Lambda_{2})}\acute{K}_{\Lambda_{3}\cdot\alpha^{\boldsymbol{i}}}^{-1}\otimes F_{\boldsymbol{i}}^{(\Lambda_{3})} with Λ1α𝒊+Λ2α𝒊+Λ3α𝒊=α\Lambda_{1}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{2}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{3}\cdot\alpha^{\boldsymbol{i}}=\alpha. A similar fact holds for (Δid)Δ(E´𝒊,β)(\Delta\otimes\operatorname{\mathrm{id}})\Delta(\acute{E}_{\boldsymbol{i},\beta}). Then the formula (2) implies that E´𝒊,α+F´𝒊,α\acute{E}_{\boldsymbol{i},\alpha^{+}}\acute{F}_{\boldsymbol{i},\alpha^{-}} is an 𝒜\mathcal{A}-linear combination of elements of the following form:

τ(S(E𝒊(Λ1+)Kα+1),F𝒊(Λ1)KΛ2α𝒊+Λ3α𝒊1)τ(E𝒊(Λ3+)KΛ3+α𝒊1,F𝒊(Λ3))F´𝒊(Λ2)K´Λ3α𝒊+Λ3+α𝒊1E´𝒊(Λ2+)\displaystyle\tau(S(E_{\boldsymbol{i}}^{(\Lambda_{1}^{+})}K_{\alpha^{+}}^{-1}),F_{\boldsymbol{i}}^{(\Lambda_{1}^{-})}K_{\Lambda_{2}^{-}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{3}^{-}\cdot\alpha^{\boldsymbol{i}}}^{-1})\tau(E_{\boldsymbol{i}}^{(\Lambda_{3}^{+})}K_{\Lambda_{3}^{+}\cdot\alpha^{\boldsymbol{i}}}^{-1},F_{\boldsymbol{i}}^{(\Lambda_{3}^{-})})\acute{F}_{\boldsymbol{i}}^{(\Lambda_{2}^{-})}\acute{K}_{\Lambda_{3}^{-}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{3}^{+}\cdot\alpha^{\boldsymbol{i}}}^{-1}\acute{E}_{\boldsymbol{i}}^{(\Lambda_{2}^{+})}

with Λ1±α𝒊+Λ2±α𝒊+Λ3±α𝒊=α±\Lambda_{1}^{\pm}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{2}^{\pm}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{3}^{\pm}\cdot\alpha^{\boldsymbol{i}}=\alpha^{\pm}. Since this term does not vanish only when Λ3+=Λ3\Lambda_{3}^{+}=\Lambda_{3}^{-}, we have μ=Λ3±α𝒊Q+\mu=\Lambda_{3}^{\pm}\cdot\alpha^{\boldsymbol{i}}\in Q^{+}. ∎

As an application we can obtain formulae on the coproduct.

Proposition 3.5.

Fix a reduced expression s𝐢s_{\boldsymbol{i}} of w0w_{0}. Then we have the following expansion with Ck±(Λl,Λr)𝒜C_{k}^{\pm}(\Lambda^{l},\Lambda^{r})\in\mathcal{A}:

(5) Δ(E´𝒊,k)=E𝒊,kKαk𝒊11+Kαk𝒊1E´𝒊,k+Λr<k<Λl,Λlα𝒊+Λrα𝒊=αk𝒊Ck+(Λl,Λr)Kαk𝒊1E𝒊(Λl)E´𝒊(Λr),\displaystyle\begin{aligned} \Delta(\acute{E}_{\boldsymbol{i},k})=E_{\boldsymbol{i},k}&K_{\alpha^{\boldsymbol{i}}_{k}}^{-1}\otimes 1+K_{\alpha^{\boldsymbol{i}}_{k}}^{-1}\otimes\acute{E}_{\boldsymbol{i},k}\\ &+\sum_{\Lambda_{r}<k<\Lambda_{l},\Lambda_{l}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{r}\cdot\alpha^{\boldsymbol{i}}=\alpha^{\boldsymbol{i}}_{k}}C_{k}^{+}(\Lambda_{l},\Lambda_{r})K_{\alpha^{\boldsymbol{i}}_{k}}^{-1}E_{\boldsymbol{i}}^{(\Lambda_{l})}\otimes\acute{E}_{\boldsymbol{i}}^{(\Lambda_{r})},\end{aligned}
(6) Δ(F´𝒊,k)=F𝒊,k1+Kαk𝒊1F´𝒊,k+Λl<k<Λr,Λlα𝒊+Λrα𝒊=αk𝒊Ck(Λl,Λr)F𝒊(Λl)KΛrα𝒊1F´𝒊(Λr).\displaystyle\begin{aligned} \Delta(\acute{F}_{\boldsymbol{i},k})=F_{\boldsymbol{i},k}&\otimes 1+K_{\alpha^{\boldsymbol{i}}_{k}}^{-1}\otimes\acute{F}_{\boldsymbol{i},k}\\ &+\sum_{\Lambda_{l}<k<\Lambda_{r},\Lambda_{l}\cdot\alpha^{\boldsymbol{i}}+\Lambda_{r}\cdot\alpha^{\boldsymbol{i}}=\alpha^{\boldsymbol{i}}_{k}}C_{k}^{-}(\Lambda_{l},\Lambda^{r})F_{\boldsymbol{i}}^{(\Lambda^{l})}K_{\Lambda^{r}\cdot\alpha^{\boldsymbol{i}}}^{-1}\otimes\acute{F}_{\boldsymbol{i}}^{(\Lambda^{r})}.\end{aligned}

We need some preparations to prove this proposition. Recall the quantum exponential function:

expqx=k=0qk(k1)/2[k]q!xk\displaystyle\exp_{q}x=\sum_{k=0}^{\infty}\frac{q^{k(k-1)/2}}{[k]_{q}!}x^{k}

Let s𝒊s_{\boldsymbol{i}} be a reduced expression of an element of WW.

Δ(𝒯i(z))=Ai(𝒯i𝒯i)(Δ(z))Ai1,\displaystyle\Delta(\mathcal{T}_{i}(z))=A_{i}(\mathcal{T}_{i}\otimes\mathcal{T}_{i})(\Delta(z))A_{i}^{-1},

where Ai:=expqi((qiqi1)KiFiE´i)A_{i}:=\exp_{q_{i}}((q_{i}-q_{i}^{-1})K_{i}F_{i}\otimes\acute{E}_{i}), whose inverse is given by Ai1=expqi1((qiqi1)KiFiE´i)A_{i}^{-1}=\exp_{q_{i}^{-1}}(-(q_{i}-q_{i}^{-1})K_{i}F_{i}\otimes\acute{E}_{i}). Applying this formula iteratively to a reduced expression s𝒊s_{\boldsymbol{i}} of an element wWw\in W, we also have

Δ(𝒯w(z))=A𝒊,1A𝒊,2A𝒊,(w)(𝒯w𝒯w)(Δ(z))A𝒊,(w)1A𝒊,(w)11A𝒊,11\displaystyle\Delta(\mathcal{T}_{w}(z))=A_{\boldsymbol{i},1}A_{\boldsymbol{i},2}\cdots A_{\boldsymbol{i},\ell(w)}(\mathcal{T}_{w}\otimes\mathcal{T}_{w})(\Delta(z))A_{\boldsymbol{i},\ell(w)}^{-1}A_{\boldsymbol{i},\ell(w)-1}^{-1}\cdots A_{\boldsymbol{i},1}^{-1}

where

A𝒊,l\displaystyle A_{\boldsymbol{i},l} =expqαl𝒊((qαl𝒊qαl𝒊1)Kαl𝒊F𝒊,lE´𝒊,l),\displaystyle=\exp_{q_{\alpha^{\boldsymbol{i}}_{l}}}((q_{\alpha^{\boldsymbol{i}}_{l}}-q_{\alpha^{\boldsymbol{i}}_{l}}^{-1})K_{\alpha^{\boldsymbol{i}}_{l}}F_{\boldsymbol{i},l}\otimes\acute{E}_{\boldsymbol{i},l}),
A𝒊,l1\displaystyle A_{\boldsymbol{i},l}^{-1} =expqαl𝒊1((qαl𝒊qαl𝒊1)Kαl𝒊F𝒊,lE´𝒊,l).\displaystyle=\exp_{q_{\alpha^{\boldsymbol{i}}_{l}}^{-1}}(-(q_{\alpha^{\boldsymbol{i}}_{l}}-q_{\alpha^{\boldsymbol{i}}_{l}}^{-1})K_{\alpha^{\boldsymbol{i}}_{l}}F_{\boldsymbol{i},l}\otimes\acute{E}_{\boldsymbol{i},l}).
Proof.

Let s𝒊s_{\boldsymbol{i}} be a reduced expression of the longest element compatible with R0+R_{0}^{+} and wk=si1si2sikw_{k}=s_{i_{1}}s_{i_{2}}\cdots s_{i_{k}}. Then, for any 1kn1\leq k\leq n, we have

Δ(E´𝒊,k)=A𝒊,1A𝒊,2A𝒊,k1(𝒯wk1𝒯wk1)(Δ(E´ik))A𝒊,k11A𝒊,k21A𝒊,11.\displaystyle\Delta(\acute{E}_{\boldsymbol{i},k})=A_{\boldsymbol{i},1}A_{\boldsymbol{i},2}\cdots A_{\boldsymbol{i},k-1}(\mathcal{T}_{w_{k-1}}\otimes\mathcal{T}_{w_{k-1}})(\Delta(\acute{E}_{i_{k}}))A_{\boldsymbol{i},k-1}^{-1}A_{\boldsymbol{i},k-2}^{-1}\cdots A_{\boldsymbol{i},1}^{-1}.

On the other hand we have

(𝒯wk1𝒯wk1)(Δ(E´ik))=E𝒊,kKαk𝒊11+Kαk𝒊1E´𝒊,k\displaystyle(\mathcal{T}_{w_{k-1}}\otimes\mathcal{T}_{w_{k-1}})(\Delta(\acute{E}_{i_{k}}))=E_{\boldsymbol{i},k}K_{\alpha^{\boldsymbol{i}}_{k}}^{-1}\otimes 1+K_{\alpha^{\boldsymbol{i}}_{k}}^{-1}\otimes\acute{E}_{\boldsymbol{i},k}

Combining with Proposition 3.3, Proposition 3.4 (i) and Δ(E´𝒊,k)Uq𝒜(𝔟+)Uq𝒜(𝔟+)\Delta(\acute{E}_{\boldsymbol{i},k})\in U_{q}^{\mathcal{A}}(\mathfrak{b}^{+})\otimes U_{q}^{\mathcal{A}}(\mathfrak{b}^{+}), we can see (5).

The same argument works for (6) since the following formula holds:

(𝒯wk1𝒯wk1)(Δ(F´ik(r)))=j=0rqαk𝒊j(rj)F𝒊,k(rj)Kαk𝒊jF´𝒊,k(j).\displaystyle(\mathcal{T}_{w_{k-1}}\otimes\mathcal{T}_{w_{k-1}})(\Delta(\acute{F}_{i_{k}}^{(r)}))=\sum_{j=0}^{r}q_{\alpha^{\boldsymbol{i}}_{k}}^{j(r-j)}F_{\boldsymbol{i},k}^{(r-j)}K_{\alpha^{\boldsymbol{i}}_{k}}^{-j}\otimes\acute{F}_{\boldsymbol{i},k}^{(j)}.

For a reference, see [MR4162277, Proof of Lemma 3.18]. ∎

3.2. Deformed quantum enveloping algebra

We fix a positive system R0+R_{0}^{+}, possibly different from R+R^{+}. Then there is a unique element wWw\in W such that w(R+)=R0+w(R^{+})=R_{0}^{+}. Note R+R0+={αR+w1(α)R+}R^{+}\setminus R_{0}^{+}=\{\alpha\in R^{+}\mid w^{-1}(\alpha)\in-R^{+}\} for such ww.

We say that a reduced expression s𝒊s_{\boldsymbol{i}} of the longest element is compatible with R0+R_{0}^{+} if it begins with a reduced expression of ww i.e. w=si1si2si(w)w=s_{i_{1}}s_{i_{2}}\cdots s_{i_{\ell(w)}}. Such an expression always exists since (w0)=(w)+(w1w0)\ell(w_{0})=\ell(w)+\ell(w^{-1}w_{0}). Note that s𝒊s_{\boldsymbol{i}} is compatible with R0+R_{0}^{+} if and only if R+R0+={αn𝒊}n=1(w)R^{+}\setminus R_{0}^{+}=\{\alpha^{\boldsymbol{i}}_{n}\}_{n=1}^{\ell(w)}.

For a monoid MM, the monoid algebra with coefficients in a commutative ring kk is denoted by k[M]k[M]. It has a canonical kk-basis {em}mM\{e_{m}\}_{m\in M}, for which we have emem=emme_{m}e_{m^{\prime}}=e_{mm^{\prime}}.

Let Q0Q_{0}^{-} be the additive submonoid of QQ generated by R0+-R_{0}^{+}. By the universal property of monoid algebras, we can identify an 𝒜[2Q0]\mathcal{A}[2Q_{0}^{-}]-algebra with a pair of an 𝒜\mathcal{A}-algebra kk and a monoid homomorphism χ:2Q0k;λχλ\chi\colon 2Q_{0}^{-}\longrightarrow k;\lambda\longmapsto\chi_{\lambda} with respect to the multiplication on kk. In light of this fact, we refer such a pair as a commutative 𝒜[2Q0]\mathcal{A}[2Q_{0}^{-}]-algebra.

Definition 3.6.

Let (k,χ)(k,\chi) be a commutative 𝒜[2Q0]\mathcal{A}[2Q_{0}^{-}]-algebra and fix a reduced expression s𝒊s_{\boldsymbol{i}} compatible with R0+R_{0}^{+}. We define Uq,χk(𝔤)U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim}) as k𝒜[2Q0]Uq,e𝒜[2Q0](𝔤)k\otimes_{\mathcal{A}[2Q_{0}^{-}]}U_{q,e}^{\mathcal{A}[2Q_{0}^{-}]}(\mathfrak{g}{}^{\sim}), where Uq,e𝒜[2Q0](𝔤)U_{q,e}^{\mathcal{A}[2Q_{0}^{-}]}(\mathfrak{g}{}^{\sim}) is an 𝒜[2Q0]\mathcal{A}[2Q_{0}^{-}]-subalgebra of Uq𝒜[P](𝔤)U_{q}^{\mathcal{A}[P]}(\mathfrak{g}) generated by the following elements:

  • F^𝒊,α(r):={e2αrF´𝒊,α(r)for αR+R0+,r0F´𝒊,α(r)for αR+R0+,r0,{\displaystyle\hat{F}_{\boldsymbol{i},\alpha}^{(r)}:=\begin{cases}e_{2\alpha}^{r}\acute{F}_{\boldsymbol{i},\alpha}^{(r)}&\text{for }\alpha\in R^{+}\setminus R_{0}^{+},r\geq 0\\ \acute{F}_{\boldsymbol{i},\alpha}^{(r)}&\text{for }\alpha\in R^{+}\cap R_{0}^{+},r\geq 0,\end{cases}}

  • K^λ:=eλK´λ\hat{K}_{\lambda}:=e_{-\lambda}\acute{K}_{\lambda} for λP\lambda\in P.

  • E^𝒊,α:=(qαqα1)E´𝒊,α\hat{E}_{\boldsymbol{i},\alpha}:=(q_{\alpha}-q_{\alpha}^{-1})\acute{E}_{\boldsymbol{i},\alpha} for αR+\alpha\in R^{+}.

We also define Uq,χk(𝔤)U_{q,\chi}^{k}({}^{\sim}\mathfrak{g}) by using Fˇ𝒊,α=(qαqα1)F^𝒊,α,Kˇλ=K^λ,Eˇ𝒊,α(r)=E´𝒊,α(r)\check{F}_{\boldsymbol{i},\alpha}=(q_{\alpha}-q_{\alpha}^{-1})\hat{F}_{\boldsymbol{i},\alpha},\check{K}_{\lambda}=\hat{K}_{\lambda},\check{E}_{\boldsymbol{i},\alpha}^{(r)}=\acute{E}_{\boldsymbol{i},\alpha}^{(r)}.

Remark 3.7.

These definitions actually do not depend on the choice of s𝒊s_{\boldsymbol{i}} since span𝒜{F𝒊(Λ)}Λ|R+R0+|=Uq𝒜(𝔫)𝒯w1(Uq𝒜(𝔟+))\mathrm{span}_{\mathcal{A}}\{F_{\boldsymbol{i}}^{(\Lambda)}\}_{\Lambda\leq\lvert R^{+}\setminus R_{0}^{+}\rvert}=U_{q}^{\mathcal{A}}(\mathfrak{n}^{-})\cap\mathcal{T}_{w}^{-1}(U_{q}^{\mathcal{A}}(\mathfrak{b}^{+})) and span𝒜{F𝒊(Λ)}Λ>|R+R0+|=Uq𝒜(𝔫)𝒯w(Uq𝒜(𝔫))\mathrm{span}_{\mathcal{A}}\{F_{\boldsymbol{i}}^{(\Lambda)}\}_{\Lambda>\lvert R^{+}\setminus R_{0}^{+}\rvert}=U_{q}^{\mathcal{A}}(\mathfrak{n}^{-})\cap\mathcal{T}_{w}(U_{q}^{\mathcal{A}}(\mathfrak{n}^{-})) are independent of s𝒊s_{\boldsymbol{i}}, where ww is the unique element with w(R+)=R0+w(R^{+})=R_{0}^{+}.

On the other hand, though we can consider a subalgebra of Uq𝒜[P](𝔤)U_{q}^{\mathcal{A}[P]}(\mathfrak{g}) generated by the elements above constructed from a non-compatible reduced expression s𝒊s_{\boldsymbol{i}}, it can be different from Uq,e𝒜[2Q0](𝔤)U_{q,e}^{\mathcal{A}[2Q_{0}^{-}]}(\mathfrak{g}{}^{\sim}). As an example we consider 𝔤=𝔰𝔩3\mathfrak{g}=\mathfrak{sl}_{3}. The simple roots is denoted by α\alpha and β\beta. Then the reduced expressions of the longest element are s𝒊=sαsβsαs_{\boldsymbol{i}}=s_{\alpha}s_{\beta}s_{\alpha} and s𝒋:=sβsαsβs_{\boldsymbol{j}}:=s_{\beta}s_{\alpha}s_{\beta}. Consider R0+:={β,α,αβ}R_{0}^{+}:=\{\beta,-\alpha,-\alpha-\beta\}. Then s𝒊s_{\boldsymbol{i}} is compatible with R0+R_{0}^{+} and s𝒋s_{\boldsymbol{j}} is not. In this case we have

e2(α+β)F𝒋,α+β=e2(α+β)(FβFαq1FαFβ)=q1F^𝒊,α+βe2β(qq1)F^βF^α.\displaystyle e_{2(\alpha+\beta)}F_{\boldsymbol{j},\alpha+\beta}=e_{2(\alpha+\beta)}(F_{\beta}F_{\alpha}-q^{-1}F_{\alpha}F_{\beta})=-q^{-1}\hat{F}_{\boldsymbol{i},\alpha+\beta}-e_{2\beta}(q-q^{-1})\hat{F}_{\beta}\hat{F}_{\alpha}.

Since e2βe_{2\beta} is not invertible in 𝒜[2Q0]\mathcal{A}[2Q_{0}^{-}] and the PBW theorem holds for Uq,e𝒜[2Q0](𝔤)U_{q,e}^{\mathcal{A}[2Q_{0}^{-}]}(\mathfrak{g}{}^{\sim}), as shown in Proposition 3.11, e2(α+β)F𝒊,α+βe_{2(\alpha+\beta)}F_{\boldsymbol{i},\alpha+\beta} is not in Uq,e𝒜[2Q0](𝔤)U_{q,e}^{\mathcal{A}[2Q_{0}^{-}]}(\mathfrak{g}{}^{\sim}).

Remark 3.8.

We cannot remove F^𝒊,α\hat{F}_{\boldsymbol{i},\alpha} with non-simple root α\alpha from the set of generators. As an example, we consider 𝔰𝔩3\mathfrak{sl}_{3} and a positive system R0+:={α+β,β,α}R_{0}^{+}:=\{\alpha+\beta,\beta,-\alpha\}. Then s𝒊s_{\boldsymbol{i}} is compatible with R0+R_{0}^{+}. Then we can calculate some commutation relations of the generators as follows:

[F^α,F^β]q=e2αF^α+β,[E^α+β,F^α]q=K^2α1E^β,[E^α+β,F^β]q=E^α,\displaystyle[\hat{F}_{\alpha},\hat{F}_{\beta}]_{q}=e_{2\alpha}\hat{F}_{\alpha+\beta},\,[\hat{E}_{\alpha+\beta},\hat{F}_{\alpha}]_{q}=\hat{K}_{2\alpha}^{-1}\hat{E}_{\beta},\,[\hat{E}_{\alpha+\beta},\hat{F}_{\beta}]_{q}=-\hat{E}_{\alpha},

here we omit the subscript 𝒊\boldsymbol{i}. Since e2αe_{2\alpha} is not invertible in 𝒜[2Q0]\mathcal{A}[2Q_{0}^{-}], the subalgebra generated by F^α(r),F^β(r),K^λ,E^α\hat{F}_{\alpha}^{(r)},\hat{F}_{\beta}^{(r)},\hat{K}_{\lambda},\hat{E}_{\alpha}, E^α+β,E^β\hat{E}_{\alpha+\beta},\hat{E}_{\beta} does not contain F^α+β\hat{F}_{\alpha+\beta}.

Proposition 3.9.

The canonical left coaction of Uq𝒜[2Q0](𝔤)U_{q}^{\mathcal{A}[2Q_{0}^{-}]}(\mathfrak{g}) on itself restricts to a left coaction on Uq,e𝒜[2Q0](𝔤)U_{q,e}^{\mathcal{A}[2Q_{0}^{-}]}(\mathfrak{g}{}^{\sim}). In particular Uq,χk(𝔤)U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim}) has a canonical left Uqk(𝔤)U_{q}^{k}(\mathfrak{g})-coaction. Similarly Uq,χk(𝔤)U_{q,\chi}^{k}({}^{\sim}\mathfrak{g}) has a canonical left Uqk(𝔤)U_{q}^{k}(\mathfrak{g})-coaction.

Proof.

Proof of Proposition 3.5 shows the statement since Δ(F^𝒊,α(r))\Delta(\hat{F}_{\boldsymbol{i},\alpha}^{(r)}), Δ(K^μ),Δ(E^𝒊,α)\Delta(\hat{K}_{\mu}),\Delta(\hat{E}_{\boldsymbol{i},\alpha}) and ±KαF𝒊,αE^𝒊,α\pm K_{\alpha}F_{\boldsymbol{i},\alpha}\otimes\hat{E}_{\boldsymbol{i},\alpha}, whose quantum exponential is A𝒊,n±1A_{\boldsymbol{i},n}^{\pm 1}, are contained in Uq𝒜(𝔤)Uq,e𝒜[2Q0](𝔤)U_{q}^{\mathcal{A}}(\mathfrak{g})\otimes U_{q,e}^{\mathcal{A}[2Q_{0}^{-}]}(\mathfrak{g}{}^{\sim}). A similar argument works for Uq,χk(𝔤)U_{q,\chi}^{k}({}^{\sim}\mathfrak{g}). ∎

Next we prove a PBW-type result for Uq,χk(𝔤)U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim}). Let us define E~𝒊,α,K~μ,F~𝒊,αUq𝒜[P](𝔤)\widetilde{E}_{\boldsymbol{i},\alpha},\widetilde{K}_{\mu},\widetilde{F}_{\boldsymbol{i},\alpha}\in U_{q}^{\mathcal{A}[P]}(\mathfrak{g}) as (qαqα1)E´𝒊,α,eμKμ,(qαqα1)F´𝒊,α(q_{\alpha}-q_{\alpha}^{-1})\acute{E}_{\boldsymbol{i},\alpha},e_{-\mu}K_{\mu},(q_{\alpha}-q_{\alpha}^{-1})\acute{F}_{\boldsymbol{i},\alpha} respectively.

Lemma 3.10.

Set 𝒰:=span𝒜[2Q]{F~𝐢ΛK~μE~𝐢Λ+}(Λ±,μ)Uq𝒜[P](𝔤)\mathcal{U}:=\mathrm{span}_{\mathcal{A}[2Q^{-}]}\{\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}\widetilde{K}_{\mu}\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{+}}\}_{(\Lambda^{\pm},\mu)}\subset U_{q}^{\mathcal{A}[P]}(\mathfrak{g}). Then this is closed under multiplication. Moreover, for all 1lN1\leq l\leq N, it coincides with the 𝒜[2Q]\mathcal{A}[2Q^{-}]-linear span of elements of the form E~𝐢Λ2+F~𝐢ΛK~μE~𝐢Λ1+\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{2}^{+}}\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}\widetilde{K}_{\mu}\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{1}^{+}} where Λ1+l<Λ2+\Lambda_{1}^{+}\leq l<\Lambda_{2}^{+}.

Proof.

It is not difficult to see that 𝒰:=span𝒜[2Q]{F~𝒊ΛK~μ}\mathcal{U}^{-}:=\mathrm{span}_{\mathcal{A}[2Q^{-}]}\{\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}\widetilde{K}_{\mu}\} and 𝒰+:=span𝒜[2Q]{K~μE~𝒊Λ+}\mathcal{U}^{+}:=\mathrm{span}_{\mathcal{A}[2Q^{-}]}\{\widetilde{K}_{\mu}\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{+}}\} are closed under multiplication. To see xy𝒰xy\in\mathcal{U} for x𝒰+x\in\mathcal{U}^{+} and y𝒰y\in\mathcal{U}^{-}, one should note that the argument in the proof of Proposition 3.4 (iii) works for computing the expansion of E~𝒊Λ+F~𝒊Λ\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{+}}\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}. Hence this product can be expressed as an 𝒜\mathcal{A}-linear combination of elements of the form F~𝒊ΓK2μ1E~𝒊Γ+\widetilde{F}_{\boldsymbol{i}}^{\Gamma^{-}}K_{2\mu}^{-1}\widetilde{E}_{\boldsymbol{i}}^{\Gamma^{+}} with μQ+\mu\in Q^{+}, which shows the required property since K2μ1=e2μK~2μ1K_{2\mu}^{-1}=e_{-2\mu}\widetilde{K}_{2\mu}^{-1}.

Next we show the latter half of the statement by downward induction on ll. If l=Nl=N, there is nothing to prove. For the induction step, assume the statement holds for a fixed ll. It suffices to show that 𝒰:=span𝒜[2Q]{E~𝒊Λ2+F~𝒊ΛK~μE~𝒊Λ1+Λ1+<lΛ2+}\mathcal{U}^{\prime}:=\mathrm{span}_{\mathcal{A}[2Q^{-}]}\{\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{2}^{+}}\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}\widetilde{K}_{\mu}\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{1}^{+}}\mid\Lambda_{1}^{+}<l\leq\Lambda_{2}^{+}\} is closed under multiplication, which, combining with the induction hypothesis, implies that the linear span is the 𝒜[2Q]\mathcal{A}[2Q^{-}]-subalgebra generated by (F~𝒊,α)αR+,(K~μ)μP(\widetilde{F}_{\boldsymbol{i},\alpha})_{\alpha\in R^{+}},(\widetilde{K}_{\mu})_{\mu\in P} and (E~𝒊,α)αR+(\widetilde{E}_{\boldsymbol{i},\alpha})_{\alpha\in R^{+}}, which is nothing but 𝒰\mathcal{U}.

We only prove that 𝒰\mathcal{U}^{\prime} is invariant under the right multiplication by E~𝒊,l\widetilde{E}_{\boldsymbol{i},l} since this fact and the induction hypothesis imply the invariance for other elements.

Set 𝒰l±\mathcal{U}_{l}^{\pm} as follows:

𝒰l\displaystyle\mathcal{U}_{l}^{-} =span𝒜[2Q]{F~𝒊ΛK~μE~Λ+μP,Λ+<l<Λ}\displaystyle=\operatorname{\mathrm{span}}_{\mathcal{A}[2Q^{-}]}\{\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}\widetilde{K}_{\mu}\widetilde{E}^{\Lambda^{+}}\mid\mu\in P,\Lambda^{+}<l<\Lambda^{-}\}
𝒰l+\displaystyle\mathcal{U}_{l}^{+} =span𝒜[2Q]{E~𝒊Λ+K~μF~𝒊ΛμP,Λ<l<Λ+}.\displaystyle=\operatorname{\mathrm{span}}_{\mathcal{A}[2Q^{-}]}\{\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{+}}\widetilde{K}_{\mu}\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}\mid\mu\in P,\Lambda^{-}<l<\Lambda^{+}\}.

These are closed under multiplication by Proposition 3.4 (i). Also note that 𝒰\mathcal{U}^{\prime} is spanned by elements of the form E~𝒊,lpx+F~𝒊,lqx\widetilde{E}_{\boldsymbol{i},l}^{p}x^{+}\widetilde{F}_{\boldsymbol{i},l}^{q}x^{-} with x±𝒰l±x^{\pm}\in\mathcal{U}_{l}^{\pm} and we have

F~𝒊,lE~𝒊,l=qαl𝒊2E~𝒊,lF~𝒊,l(qαl𝒊qαl𝒊1)(1e2αl𝒊K~2αl𝒊).\displaystyle\widetilde{F}_{\boldsymbol{i},l}\widetilde{E}_{\boldsymbol{i},l}=q_{\alpha^{\boldsymbol{i}}_{l}}^{-2}\widetilde{E}_{\boldsymbol{i},l}\widetilde{F}_{\boldsymbol{i},l}-(q_{\alpha^{\boldsymbol{i}}_{l}}-q_{\alpha^{\boldsymbol{i}}_{l}}^{-1})(1-e_{-2\alpha^{\boldsymbol{i}}_{l}}\widetilde{K}_{-2\alpha^{\boldsymbol{i}}_{l}}).

Therefore it suffices to show that 𝒰l±E~𝒊,lE~𝒊,l𝒰l±+𝒰l±\mathcal{U}_{l}^{\pm}\widetilde{E}_{\boldsymbol{i},l}\subset\widetilde{E}_{\boldsymbol{i},l}\mathcal{U}_{l}^{\pm}+\mathcal{U}_{l}^{\pm}. At first take x𝒰lx^{-}\in\mathcal{U}_{l}^{-}. By using the triangular decomposition and Proposition 3.3, we may assume x=F~𝒊Λx^{-}=\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}} with Λ>l\Lambda^{-}>l. If Λ=δm\Lambda^{-}=\delta_{m} for some m>lm>l, the mixed Levendörskii-Soibelman relation (3) proves F~𝒊,mE~𝒊,lq(αl𝒊,αm𝒊)E~𝒊,lF~𝒊,m𝒰l\widetilde{F}_{\boldsymbol{i},m}\widetilde{E}_{\boldsymbol{i},l}-q^{-(\alpha^{\boldsymbol{i}}_{l},\alpha^{\boldsymbol{i}}_{m})}\widetilde{E}_{\boldsymbol{i},l}\widetilde{F}_{\boldsymbol{i},m}\in\mathcal{U}_{l}^{-}. Then we have

F~𝒊,m1F~𝒊,m2F~𝒊,mnE~𝒊,lq(αl𝒊,αm1𝒊+αm2𝒊++αml𝒊)E~𝒊,lF~𝒊,m1F~𝒊,m2F~𝒊,mn𝒰l\displaystyle\widetilde{F}_{\boldsymbol{i},m_{1}}\widetilde{F}_{\boldsymbol{i},m_{2}}\cdots\widetilde{F}_{\boldsymbol{i},m_{n}}\widetilde{E}_{\boldsymbol{i},l}-q^{-(\alpha^{\boldsymbol{i}}_{l},\alpha^{\boldsymbol{i}}_{m_{1}}+\alpha^{\boldsymbol{i}}_{m_{2}}+\cdots+\alpha^{\boldsymbol{i}}_{m_{l}})}\widetilde{E}_{\boldsymbol{i},l}\widetilde{F}_{\boldsymbol{i},m_{1}}\widetilde{F}_{\boldsymbol{i},m_{2}}\cdots\widetilde{F}_{\boldsymbol{i},m_{n}}\in\mathcal{U}_{l}^{-}

for any sequence (m1,m2,,mn)(m_{1},m_{2},\cdots,m_{n}) with mj>lm_{j}>l, which is nothing but the required property. A similar argument works for x+𝒰l+x^{+}\in\mathcal{U}_{l}^{+}. ∎

Proposition 3.11.

Fix a reduced expression s𝐢s_{\boldsymbol{i}} of the longest element compatible with R0+R_{0}^{+}. Then {F^𝐢(Λ)K^μE^𝐢Λ+}(Λ±,μ)\{\hat{F}_{\boldsymbol{i}}^{(\Lambda^{-})}\hat{K}_{\mu}\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}}\}_{(\Lambda^{\pm},\mu)} is a basis of Uq,χk(𝔤)U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim}).

Proof.

We may assume k=𝒜[2Q0]k=\mathcal{A}[2Q_{0}^{-}]. Set l:=|R+R0+|l:=\lvert R^{+}\cap R_{0}^{+}\rvert. Then there is wWw\in W and a reduced expression s𝒋s_{\boldsymbol{j}} of w0w_{0} as Lemma 3.2 for s𝒊s_{\boldsymbol{i}} and NlN-l. Let twt_{w} be an automorphism of 𝒜[P]\mathcal{A}[P] which sends tw(eλ)=ew(λ)t_{w}(e_{\lambda})=e_{w(\lambda)} and φ\varphi be an automorphism on Uq𝒜[P]U_{q}^{\mathcal{A}[P]} defined as tw𝒯wt_{w}\otimes\mathcal{T}_{w}. Then φ(𝒰)\varphi(\mathcal{U}) is closed under multiplication and spanned by elements of the form 𝒯w(E~𝒋)Λ2+𝒯w(F~𝒋)ΛK~μ𝒯w(E~𝒋)Λ1+\mathcal{T}_{w}(\widetilde{E}_{\boldsymbol{j}})^{\Lambda_{2}^{+}}\mathcal{T}_{w}(\widetilde{F}_{\boldsymbol{j}})^{\Lambda^{-}}\widetilde{K}_{\mu}\mathcal{T}_{w}(\widetilde{E}_{\boldsymbol{j}})^{\Lambda_{1}^{+}} with Λ1+l<Λ2+\Lambda_{1}^{+}\leq l<\Lambda_{2}^{+}. Hence Lemma 3.10 implies that spank{Fˇ𝒊ΛK^μE^𝒊Λ+}(Λ±,μ)\mathrm{span}_{k}\{\check{F}_{\boldsymbol{i}}^{\Lambda^{-}}\hat{K}_{\mu}\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}}\}_{(\Lambda^{\pm},\mu)} is closed under multiplication.

Now fix (Λ1±,μ1)(\Lambda_{1}^{\pm},\mu_{1}) and (Λ2±,μ2)(\Lambda_{2}^{\pm},\mu_{2}) and consider the following two expansions with CΛ±,μ𝒜C_{\Lambda^{\pm},\mu}\in\mathcal{A}:

(F´𝒊(Λ1)Kμ1E~𝒊Λ1+)(F´𝒊(Λ2)Kμ2E~𝒊Λ2+)\displaystyle(\acute{F}_{\boldsymbol{i}}^{(\Lambda_{1}^{-})}K_{\mu_{1}}\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{1}^{+}})(\acute{F}_{\boldsymbol{i}}^{(\Lambda_{2}^{-})}K_{\mu_{2}}\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{2}^{+}}) =Λ±,μCΛ±,μF´𝒊(Λ)KμE~Λ+.\displaystyle=\sum_{\Lambda^{\pm},\mu}C_{\Lambda^{\pm},\mu}\acute{F}_{\boldsymbol{i}}^{(\Lambda^{-})}K_{\mu}\widetilde{E}^{\Lambda^{+}}.

Then there is ν(Λ±,μ)P\nu(\Lambda^{\pm},\mu)\in P for each (Λ±,μ)(\Lambda^{\pm},\mu) with which we have

(F^𝒊(Λ1)K^μ1E^𝒊Λ1+)(F^𝒊(Λ2)K^μ2E^𝒊Λ2+)=Λ±,μeν(Λ±,μ)CΛ±,μF^𝒊(Λ)K^μE^𝒊Λ+.\displaystyle(\hat{F}_{\boldsymbol{i}}^{(\Lambda_{1}^{-})}\hat{K}_{\mu_{1}}\hat{E}_{\boldsymbol{i}}^{\Lambda_{1}^{+}})(\hat{F}_{\boldsymbol{i}}^{(\Lambda_{2}^{-})}\hat{K}_{\mu_{2}}\hat{E}_{\boldsymbol{i}}^{\Lambda_{2}^{+}})=\sum_{\Lambda^{\pm},\mu}e_{\nu(\Lambda^{\pm},\mu)}C_{\Lambda^{\pm},\mu}\hat{F}_{\boldsymbol{i}}^{(\Lambda^{-})}\hat{K}_{\mu}\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}}.

On the other hand, this weight appears also in the following comparison:

(F~𝒊Λ1Kμ1E~𝒊Λ1+)(F~𝒊Λ2Kμ2E~𝒊Λ2+)\displaystyle(\widetilde{F}_{\boldsymbol{i}}^{\Lambda_{1}^{-}}K_{\mu_{1}}\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{1}^{+}})(\widetilde{F}_{\boldsymbol{i}}^{\Lambda_{2}^{-}}K_{\mu_{2}}\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{2}^{+}}) =Λ±,μcΛ±,μF~𝒊ΛKμE~Λ+,\displaystyle=\sum_{\Lambda^{\pm},\mu}c_{\Lambda^{\pm},\mu}\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}K_{\mu}\widetilde{E}^{\Lambda^{+}},
(Fˇ𝒊Λ1K^μ1E^𝒊Λ1+)(Fˇ𝒊Λ2K^μ2E^𝒊Λ2+)\displaystyle(\check{F}_{\boldsymbol{i}}^{\Lambda_{1}^{-}}\hat{K}_{\mu_{1}}\hat{E}_{\boldsymbol{i}}^{\Lambda_{1}^{+}})(\check{F}_{\boldsymbol{i}}^{\Lambda_{2}^{-}}\hat{K}_{\mu_{2}}\hat{E}_{\boldsymbol{i}}^{\Lambda_{2}^{+}}) =Λ±,μe(Λ±,μ)cΛ±,μFˇ𝒊ΛK^μE^𝒊Λ+.\displaystyle=\sum_{\Lambda^{\pm},\mu}e_{(\Lambda^{\pm},\mu)}c_{\Lambda^{\pm},\mu}\check{F}_{\boldsymbol{i}}^{\Lambda^{-}}\hat{K}_{\mu}\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}}.

Then the discussion above shows that each ν(Λ±,μ)\nu(\Lambda^{\pm},\mu) is contained in the image of tw(2Q)=2Q0t_{w}(2Q^{-})=2Q_{0}^{-}, which completes the proof. ∎

A similar argument works to prove the following.

Proposition 3.12.

Fix a reduced expression s𝐢s_{\boldsymbol{i}} of the longest element compatible with R0+R_{0}^{+}. Then {Fˇ𝐢ΛKˇμEˇ𝐢(Λ+)}(Λ±,μ)\{\check{F}_{\boldsymbol{i}}^{\Lambda^{-}}\check{K}_{\mu}\check{E}_{\boldsymbol{i}}^{(\Lambda^{+})}\}_{(\Lambda^{\pm},\mu)} is a basis of Uqk(𝔤)U_{q}^{k}({}^{\sim}\mathfrak{g}).

3.3. Twisted parabolic induction via deformed QEA

At the end of this section, we define twisted parabolic induction, which play an important role to construct quantum semisimple coadjoint orbits.

Let SS be a set of simple roots and R0+R_{0}^{+} be a positive system of RR containing RS+R_{S}^{+}. Fix a reduced expression s𝒊s_{\boldsymbol{i}} compatible with R0+R_{0}^{+}. We also assume that RS+={αk𝒊}k=NNS+1NR_{S}^{+}=\{\alpha^{\boldsymbol{i}}_{k}\}_{k=N-N_{S}+1}^{N} holds:

α1𝒊,α2𝒊,,αl𝒊R+R0+,αl+1𝒊,αl+2𝒊,,αNNS𝒊R0+RS+,αNNS+1𝒊,αNNS+2𝒊,αN𝒊RS+.\displaystyle\underbrace{\alpha^{\boldsymbol{i}}_{1},\,\alpha^{\boldsymbol{i}}_{2},\,\dots,\,\alpha^{\boldsymbol{i}}_{l}}_{R^{+}\setminus R_{0}^{+}},\,\underbrace{\alpha^{\boldsymbol{i}}_{l+1},\,\alpha^{\boldsymbol{i}}_{l+2},\,\dots,\alpha^{\boldsymbol{i}}_{N-N_{S}}}_{R_{0}^{+}\setminus R_{S}^{+}},\,\underbrace{\alpha^{\boldsymbol{i}}_{N-N_{S}+1},\,\alpha^{\boldsymbol{i}}_{N-N_{S}+2}\dots,\,\alpha^{\boldsymbol{i}}_{N}}_{R_{S}^{+}}.

where l=|R+R0+|l=\lvert R^{+}\setminus R_{0}^{+}\rvert. For an 𝒜[2Q0]\mathcal{A}[2Q_{0}^{-}]-algebra (k,χ)(k,\chi) with χ2α=1\chi_{-2\alpha}=1 for αRS+\alpha\in R_{S}^{+}, we introduce the following kk-subalgebras:

  • Uq,χk(𝔭S)U_{q,\chi}^{k}(\mathfrak{p}_{S}{}^{\sim}), generated by (K^μ)μP,(E^𝒊,α)αR+(\hat{K}_{\mu})_{\mu\in P},\,(\hat{E}_{\boldsymbol{i},\alpha})_{\alpha\in R^{+}} and (F^𝒊,α(r))αRS+,r0(\hat{F}_{\boldsymbol{i},\alpha}^{(r)})_{\alpha\in R_{S}^{+},r\geq 0}.

  • Uq,χk(𝔩S)U_{q,\chi}^{k}(\mathfrak{l}_{S}{}^{\sim}), generated by (K^μ)μP,(E^𝒊,α)αRS+(\hat{K}_{\mu})_{\mu\in P},\,(\hat{E}_{\boldsymbol{i},\alpha})_{\alpha\in R_{S}^{+}} and (F^𝒊,α(r))αRS+,r0(\hat{F}_{\boldsymbol{i},\alpha}^{(r)})_{\alpha\in R_{S}^{+},r\geq 0}.

  • Uq,χk(𝔲S+)U_{q,\chi}^{k}(\mathfrak{u}_{S}^{+}{}^{\sim}), generated by (E^𝒊,α)αR+RS+(\hat{E}_{\boldsymbol{i},\alpha})_{\alpha\in R^{+}\setminus R_{S}^{+}}.

  • Uq,χk(𝔲S)U_{q,\chi}^{k}(\mathfrak{u}_{S}^{-}{}^{\sim}), generated by (F^𝒊,α(r))αR+RS+,r0(\hat{F}_{\boldsymbol{i},\alpha}^{(r)})_{\alpha\in R^{+}\setminus R_{S}^{+},r\geq 0}.

Due to the additional requirement on s𝒊s_{\boldsymbol{i}}, these do not depend on the choice of such a reduced expression. Moreover these are spanned by the PBW basis introduced in Proposition 3.11. Hence the multiplication map induces the following tensor product decompositions:

(7) Uq,χk(𝔭S)\displaystyle U_{q,\chi}^{k}(\mathfrak{p}_{S}{}^{\sim}) Uq,χk(𝔩S)Uq,χk(𝔲S+),\displaystyle\cong U_{q,\chi}^{k}(\mathfrak{l}_{S}{}^{\sim})\otimes U_{q,\chi}^{k}(\mathfrak{u}_{S}^{+}{}^{\sim}),
(8) Uq,χk(𝔤)\displaystyle U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim}) Uq,χk(𝔲S)Uq,χk(𝔭S).\displaystyle\cong U_{q,\chi}^{k}(\mathfrak{u}_{S}^{-}{}^{\sim})\otimes U_{q,\chi}^{k}(\mathfrak{p}_{S}{}^{\sim}).

For χ1\chi\equiv 1, these algebras are simply denoted by Uqk(𝔤),Uqk(𝔭S),Uqk(𝔩S)U_{q}^{k}(\mathfrak{g}{}^{\sim}),U_{q}^{k}(\mathfrak{p}_{S}{}^{\sim}),U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim}) and Uqk(𝔲S±)U_{q}^{k}(\mathfrak{u}_{S}^{\pm}{}^{\sim}) respectively. These can be thought as subalgebras of Uqk(𝔤)U_{q}^{k}(\mathfrak{g}) by the identifications F^𝒊,α(r)=F𝒊,α(r),K^μ=Kμ,E^𝒊,α=E~𝒊,α\hat{F}_{\boldsymbol{i},\alpha}^{(r)}=F_{\boldsymbol{i},\alpha}^{(r)},\hat{K}_{\mu}=K_{\mu},\hat{E}_{\boldsymbol{i},\alpha}=\widetilde{E}_{\boldsymbol{i},\alpha}.

Lemma 3.13.

There is a canonical isomorphism from Uq,χk(𝔭S)U_{q,\chi}^{k}(\mathfrak{p}_{S}{}^{\sim}) to Uqk(𝔭S)U_{q}^{k}(\mathfrak{p}_{S}{}^{\sim}), which is defined as follows on the generators:

F^𝒊,α(r)F𝒊,α(r),K^μKμ,E^𝒊,αE~𝒊,α.\displaystyle\hat{F}_{\boldsymbol{i},\alpha}^{(r)}\longmapsto F_{\boldsymbol{i},\alpha}^{(r)},\quad\hat{K}_{\mu}\longmapsto K_{\mu},\quad\hat{E}_{\boldsymbol{i},\alpha}\longmapsto\widetilde{E}_{\boldsymbol{i},\alpha}.
Remark 3.14.

In light of this fact, we identify Uq,χk(𝔭S)U_{q,\chi}^{k}(\mathfrak{p}_{S}{}^{\sim}) with Uqk(𝔭S)U_{q}^{k}(\mathfrak{p}_{S}{}^{\sim}).

Proof.

Using Proposition 3.11 we can define a kk-linear map from Uq,χk(𝔭S)U_{q,\chi}^{k}(\mathfrak{p}_{S}{}^{\sim}) to Uqk(𝔭S)U_{q}^{k}(\mathfrak{p}_{S}{}^{\sim}) sending F^𝒊(Λ)K^μE^𝒊Λ+\hat{F}_{\boldsymbol{i}}^{(\Lambda^{-})}\hat{K}_{\mu}\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}} to F𝒊(Λ)KμE~𝒊Λ+F_{\boldsymbol{i}}^{(\Lambda^{-})}K_{\mu}\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{+}}. What we have to show is multiplicativity of this map. By Proposition 3.4, E~𝒊Λ0+F𝒊(Λ0)\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{0}^{+}}F_{\boldsymbol{i}}^{(\Lambda_{0}^{-})} with Λ+NNS<Λ\Lambda^{+}\leq N-N_{S}<\Lambda^{-} is a linear combination of F𝒊(Λ)E~𝒊Λ+F_{\boldsymbol{i}}^{(\Lambda^{-})}\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{+}} with Λ0+NNS<Λ0\Lambda_{0}^{+}\leq N-N_{S}<\Lambda_{0}^{-}. Hence we have

F𝒊(Λ1)Kμ1E~𝒊Λ1+F𝒊(Λ1)Kμ1E~𝒊Λ1+=Λ±,μQSΛ>NNSCΛ±,μF𝒊(Λ)Kμ1+μ2+μE~𝒊Λ+\displaystyle F_{\boldsymbol{i}}^{(\Lambda_{1}^{-})}K_{\mu_{1}}\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{1}^{+}}F_{\boldsymbol{i}}^{(\Lambda_{1}^{-})}K_{\mu_{1}}\widetilde{E}_{\boldsymbol{i}}^{\Lambda_{1}^{+}}=\sum_{\begin{subarray}{c}\Lambda^{\pm},\mu\in Q_{S}\\ \Lambda^{-}>N-N_{S}\end{subarray}}C_{\Lambda^{\pm},\mu}F_{\boldsymbol{i}}^{(\Lambda^{-})}K_{\mu_{1}+\mu_{2}+\mu}\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{+}}

when Λi>NNS\Lambda_{i}^{-}>N-N_{S}. Now the multiplicativity follows from this expansion and the definition of the generators F^𝒊,α(r),K^μ,E^𝒊,α\hat{F}_{\boldsymbol{i},\alpha}^{(r)},\hat{K}_{\mu},\hat{E}_{\boldsymbol{i},\alpha}. ∎

Note that there is a canonical homomorphism πS:Uqk(𝔭S)Uqk(𝔩S)\pi_{S}\colon U_{q}^{k}(\mathfrak{p}_{S}{}^{\sim})\longrightarrow U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim}) which is identical on Uqk(𝔩S)U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim}) and equal to ε\varepsilon on Uqk(𝔲S+)U_{q}^{k}(\mathfrak{u}_{S}^{+}{}^{\sim}). All Uqk(𝔩S)U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim})-modules are regarded as Uqk(𝔭S)U_{q}^{k}(\mathfrak{p}_{S}{}^{\sim})-modules via this homomorphism.

Definition 3.15.

We define a functor ind𝔩S,q𝔤,χ:Uqk(𝔩S)-ModUq,χk(𝔤)-Mod\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}\colon U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim})\text{-}\mathrm{Mod}\longrightarrow U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})\text{-}\mathrm{Mod} as Uq,χk(𝔤)Uqk(𝔭S)U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})\otimes_{U_{q}^{k}(\mathfrak{p}_{S}{}^{\sim})}\textendash. For a Uq,χk(𝔤)U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})-module VV, ind𝔩S,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V is called the (R0+,χ)(R_{0}^{+},\chi)-twisted parabolic induction module of VV.

Remark 3.16.

By the tensor product decomposition (8), ind𝔩S,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V can be naturally identified with Uq,χk(𝔲S)VU_{q,\chi}^{k}(\mathfrak{u}_{S}^{-}{}^{\sim})\otimes V as Uq,χk(𝔲S)U_{q,\chi}^{k}(\mathfrak{u}_{S}^{-}{}^{\sim})-modules.

Let MM be a Uq,χk(𝔤)U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})-module. We say that mMm\in M is SS-maximal when xm=πS(x)mxm=\pi_{S}(x)m for xUqk(𝔭S)x\in U_{q}^{k}(\mathfrak{p}_{S}{}^{\sim}), which is the same thing as saying that E^𝒊,βm=0\hat{E}_{\boldsymbol{i},\beta}m=0 for βR+RS+\beta\in R^{+}\setminus R_{S}^{+}. The following universal property of (R0+,χ)(R_{0}^{+},\chi)-twisted parabolic induction is proven in an analogous way to the usual setting.

Lemma 3.17.

Let MM be a Uqk(𝔤)U_{q}^{k}(\mathfrak{g}{}^{\sim})-module. Then MS-maxM_{S\text{-}\max}, the set of SS-maximal vectors in MM, forms a Uqk(𝔩S)U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim})-module. Moreover we have a canonical isomorphism:

HomUqk(𝔩S)(V,MS-max)HomUq,χk(𝔤)(ind𝔩S,q𝔤,χV,M).\displaystyle\operatorname{\mathrm{Hom}}_{U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim})}(V,M_{S\text{-}\max})\cong\operatorname{\mathrm{Hom}}_{U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})}(\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V,M).

Let ind𝔭S,q𝔤:Uqk(𝔩S)-ModUqk(𝔤)-Mod\mathrm{ind}_{\mathfrak{p}_{S},q}^{\mathfrak{g}}\colon U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim})\text{-}\mathrm{Mod}\longrightarrow U_{q}^{k}(\mathfrak{g}{}^{\sim})\text{-}\mathrm{Mod} be the usual parabolic induction Uqk(𝔤)Uqk(𝔭S)U_{q}^{k}(\mathfrak{g}{}^{\sim})\otimes_{U_{q}^{k}(\mathfrak{p}_{S}{}^{\sim})}\textendash. In order to compare the (R0+,χ)(R_{0}^{+},\chi)-twisted parabolic induction with the usual parabolic induction, we consider subalgebras Uq,χk(𝔤)2QU_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})_{2Q} defined by restricting the Cartan part spank{K^μ}μP\operatorname{\mathrm{span}}_{k}\{\hat{K}_{\mu}\}_{\mu\in P} to spank{K^2μ}μQ\operatorname{\mathrm{span}}_{k}\{\hat{K}_{2\mu}\}_{\mu\in Q} and similar variants for the other subalgebras. Note that ind𝔩S,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V is naturally isomorphic to Uq,χk(𝔤)2QUq,χk(𝔭S)2QVU_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})_{2Q}\otimes_{U_{q,\chi}^{k}(\mathfrak{p}_{S}{}^{\sim})_{2Q}}V as Uq,χk(𝔤)2QU_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})_{2Q}-module and the same thing can be said for ind𝔭S,q𝔤V\mathrm{ind}_{\mathfrak{p}_{S},q}^{\mathfrak{g}}V.

Lemma 3.18.

Assume that χ:2Qk\chi\colon 2Q^{-}\longrightarrow k extends to a group homomorphism χ:2Qk×\chi\colon 2Q\longrightarrow k^{\times} with χ|2QS1\chi|_{2Q_{S}}\equiv 1. Let kχk_{\chi} be a Uqk(𝔩S)U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim})-module which is kk as a kk-module and F^𝐢,α(r),K^2μ,E^𝐢,β\hat{F}_{\boldsymbol{i},\alpha}^{(r)},\hat{K}_{2\mu},\hat{E}_{\boldsymbol{i},\beta} act on it as 0,χ2μ,00,\chi_{2\mu},0 respectively.

  1. (i)

    There is an isomorphism Uq,χk(𝔤)2QUqk(𝔤)2QU_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})_{2Q}\cong U_{q}^{k}(\mathfrak{g}{}^{\sim})_{2Q} with:

    F^𝒊,α(r){χ2αrF𝒊,α(r)(αR+R0+)F𝒊,α(r)(αR+R0+),K^2μχ2μK2μ,E^𝒊,αE~𝒊,α.\displaystyle\hat{F}_{\boldsymbol{i},\alpha}^{(r)}\longmapsto\begin{cases}\chi_{2\alpha}^{r}F_{\boldsymbol{i},\alpha}^{(r)}&(\alpha\in R^{+}\setminus R_{0}^{+})\\ F_{\boldsymbol{i},\alpha}^{(r)}&(\alpha\in R^{+}\cap R_{0}^{+})\end{cases},\quad\hat{K}_{2\mu}\longmapsto\chi_{-2\mu}K_{2\mu},\quad\hat{E}_{\boldsymbol{i},\alpha}\longmapsto\widetilde{E}_{\boldsymbol{i},\alpha}.
  2. (ii)

    Let VV be a Uqk(𝔩S)U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim})-module. Then ind𝔩S,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V is canonically isomorphic to ind𝔭S,q𝔤(Vkχ)\mathrm{ind}_{\mathfrak{p}_{S},q}^{\mathfrak{g}}(V\otimes k_{\chi}) as a Uqk(𝔤)2QU_{q}^{k}(\mathfrak{g}{}^{\sim})_{2Q}-module.

Proof.

(i) This can be immediately verified by the definition of Uq,χk(𝔤)U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim}).

(ii) This can be seen from the universal property of the (R0+,χ)(R_{0}^{+},\chi)-twisted parabolic induction and that of the usual parabolic induction. ∎

4. SS-maximal vectors in the formal setting

4.1. Notation

In Section 4 and 5, we work on (kh,s=eh/2L)(k\llbracket h\rrbracket,s=e^{h/2L}) where kk is a \mathbb{Q}-algebra. In this case it is convenient to consider hh-adically completed tensor products, which is denoted by ^\textendash\hat{\otimes}\textendash in this paper.

In the formal setting we consider the hh-adic Drinfeld-Jimbo deformation Uhk(𝔤):=kh^Uh(𝔤)U_{h}^{k}(\mathfrak{g}):=k\llbracket h\rrbracket\hat{\otimes}U_{h}(\mathfrak{g}), where the definition of Uh(𝔤)U_{h}(\mathfrak{g}) is described in [MR1492989, Section 6.1.3, Definition 2] for instance. This is isomorphic to (kU(𝔤))h(k\otimes U(\mathfrak{g}))\llbracket h\rrbracket as topological khk\llbracket h\rrbracket-algebra. Moreover we can see that a topological Uhk(𝔤)U_{h}^{k}(\mathfrak{g})-module admitting a finite khk\llbracket h\rrbracket-basis is isomorphic to (kV)h(k\otimes V)\llbracket h\rrbracket where VV is a finite dimensional representation of 𝔤\mathfrak{g} ([MR1492989, Section 7.1.3, Proposition 10]). Such a module is called a finite integrable module in this paper and the category of finite integrable modules is denoted by Rephf𝔤\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{g}.

Fix S,R0+,s𝒊S,R_{0}^{+},s_{\boldsymbol{i}} and χ\chi as those in Subsection 3.3. In the formal setting it is rather appropriate to add H^α\hat{H}_{\alpha} to Uq,χkh(𝔤)U_{q,\chi}^{k\llbracket h\rrbracket}(\mathfrak{g}{}^{\sim}) so that K^α=exphdαH^α\hat{K}_{\alpha}=\exp{hd_{\alpha}\hat{H}_{\alpha}}. The precise definition is the following.

Definition 4.1.

Set Uk(𝔥):=kU(𝔥)U^{k}(\mathfrak{h}):=k\otimes U(\mathfrak{h}) and Uhk(𝔥)U_{h}^{k}(\mathfrak{h}) be the hh-adic completion of Uk(𝔥)khU^{k}(\mathfrak{h})\otimes k\llbracket h\rrbracket. This is equipped with a canonical coproduct such that Δ(H)=H1+1H\Delta(H)=H\otimes 1+1\otimes H for H𝔥H\in\mathfrak{h}. Consider Uq,χkh(𝔤)khUhk(𝔥)U_{q,\chi}^{k\llbracket h\rrbracket}(\mathfrak{g}{}^{\sim})\otimes_{k\llbracket h\rrbracket}U_{h}^{k}(\mathfrak{h}) with the following multiplication:

(xf)(yg)=(wty)(f(1))(xyf(2)g),(x\otimes f)(y\otimes g)=(\operatorname{\mathrm{wt}}y)(f_{(1)})(xy\otimes f_{(2)}g),

where wty\operatorname{\mathrm{wt}}y, which is an element of 𝔥\mathfrak{h}^{*}, is considered as a character on Uhk(𝔥)U_{h}^{k}(\mathfrak{h}). Then We define Uh,χk(𝔤)U_{h,\chi}^{k}(\mathfrak{g}{}^{\sim}) as a quotient of Uq,χkh(𝔤)khUhk(𝔥)U_{q,\chi}^{k\llbracket h\rrbracket}(\mathfrak{g}{}^{\sim})\otimes_{k\llbracket h\rrbracket}U_{h}^{k}(\mathfrak{h}) by the ideal generated by K^λ11exp(hλ)\hat{K}_{\lambda}\otimes 1-1\otimes\exp(h\lambda), where λ\lambda is considered as an element of 𝔥\mathfrak{h} uniquely determined by α(λ)=(α,λ)\alpha(\lambda)=(\alpha,\lambda) for α𝔥\alpha\in\mathfrak{h}^{*}.

We refer definitions and properties developed in Section 3 to apply them to Uh,χk(𝔤)U_{h,\chi}^{k}(\mathfrak{g}{}^{\sim}) in a suitably modified form. For instance Uh,χk(𝔤)U_{h,\chi}^{k}(\mathfrak{g}{}^{\sim}) has a canonical left coaction of Uhk(𝔤)U_{h}^{k}(\mathfrak{g}), the hh-adic Drinfeld-Jimbo deformation of 𝔤\mathfrak{g}.

4.2. Generalized maximal vector

Though what we would like to show in this section is a kind of semisimplicity of the tensor product of finite integrable Uhk(𝔤)U_{h}^{k}(\mathfrak{g})-modules and (R0+,χ)(R_{0}^{+},\chi)-twisted parabolically induced modules, we show a stronger result: there is an operator in Uhk(𝔤×𝔩S)U_{h}^{k}(\mathfrak{g}\times\mathfrak{l}_{S}) which produces SS-maximal vectors in the tensor product modules. To discuss such an operator, we introduce the following generalization of SS-maximal vectors. Like the Sweedler notation, x(𝔲S)x(𝔭S)x_{(\mathfrak{u}_{S}^{-})}\otimes x_{(\mathfrak{p}_{S})} denotes the image of xUh,χk(𝔤)x\in U_{h,\chi}^{k}(\mathfrak{g}{}^{\sim}) under the isomorphism Uh,χk(𝔤)Uh,χk(𝔲S)Uh,χk(𝔭S)U_{h,\chi}^{k}(\mathfrak{g}{}^{\sim})\cong U_{h,\chi}^{k}(\mathfrak{u}_{S}^{-}{}^{\sim})\otimes U_{h,\chi}^{k}(\mathfrak{p}_{S}{}^{\sim}). Similarly xW^Vx\in W\hat{\otimes}V is denoted by x(W)x(V)x_{(W)}\otimes x_{(V)}. For instance, the image of xx under a linear mapping wvwavw\otimes v\longmapsto w\otimes a\otimes v is denoted by x(W)ax(V)x_{(W)}\otimes a\otimes x_{(V)}.

Definition 4.2.

Let VV be a complete Uhk(𝔩S)U_{h}^{k}(\mathfrak{l}_{S})-module and WW be a complete Uhk(𝔤)U_{h}^{k}(\mathfrak{g})-module. We say that (mΛ)ΛNNSW^V(m_{\Lambda})_{\Lambda\leq N-N_{S}}\subset W\hat{\otimes}V is a generalized SS-maximal vector if it satisfies the following equation for all βR+RS+\beta\in R^{+}\setminus R_{S}^{+}:

(9) ΛE^𝒊,β,(1)mΛ,(W)(E^𝒊,β,(2)F^𝒊(Λ))(𝔲S)πS((E^𝒊,β,(2)F^𝒊(Λ))(𝔭S))mΛ,(V)=0.\displaystyle\sum_{\Lambda}\hat{E}_{\boldsymbol{i},\beta,(1)}m_{\Lambda,(W)}\otimes(\hat{E}_{\boldsymbol{i},\beta,(2)}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{u}_{S}^{-})}\otimes\pi_{S}((\hat{E}_{\boldsymbol{i},\beta,(2)}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{p}_{S})})m_{\Lambda,(V)}=0.
Remark 4.3.

Strictly speaking, these equalities should be interpreted after expanding (E^𝒊,β,(2)F^𝒊(Λ))(𝔲S)(\hat{E}_{\boldsymbol{i},\beta,(2)}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{u}_{S}^{-})} along the PBW basis. Then we obtain well-defined equations since, for any fixed Γ\Gamma, only finitely many terms produce F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma)}.

In the rest of this subsection, we assume χ2α=1\chi_{-2\alpha}=1 for αRS+\alpha\in R_{S}^{+} and χ2α1kh×\chi_{-2\alpha}-1\in k\llbracket h\rrbracket^{\times} for αR0+RS+\alpha\in R_{0}^{+}\setminus R_{S}^{+}.

Proposition 4.4.

Let VV (resp. W) be a complete Uhk(𝔩S)U_{h}^{k}(\mathfrak{l}_{S})-module (resp. Uhk(𝔤)U_{h}^{k}(\mathfrak{g})-module). Then, for any mW^Vm\in W\hat{\otimes}V, a generalized SS-maximal vector (mΛ)Λ(m_{\Lambda})_{\Lambda} with m0=mm_{0}=m exists and is unique.

To prove this proposition, we consider a generalized version. Let (pΓ)ΓNNS(p_{\Gamma})_{\Gamma\leq N-N_{S}} be the dual basis of the PBW basis (F^𝒊(Γ))ΓNNS(\hat{F}_{\boldsymbol{i}}^{(\Gamma)})_{\Gamma\leq N-N_{S}} of Uh,χk(𝔲S)U_{h,\chi}^{k}(\mathfrak{u}_{S}^{-}{}^{\sim}). Then, for xW^Uh,χk(𝔲S)^Vx\in W\hat{\otimes}U_{h,\chi}^{k}(\mathfrak{u}_{S}^{-}{}^{\sim})\hat{\otimes}V, we define the coefficient of F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma)} in xx as (idpΓid)(x)W^V(\operatorname{\mathrm{id}}\otimes p_{\Gamma}\otimes\operatorname{\mathrm{id}})(x)\in W\hat{\otimes}V.

We say that (mΛ)ΛNNS(m_{\Lambda})_{\Lambda\leq N-N_{S}} is a weak solution of (9) for β=αl𝒊\beta=\alpha^{\boldsymbol{i}}_{l} if the coefficient of F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma)} in the RHS is 0 for all Γ\Gamma with Γl.\Gamma\geq l.

Lemma 4.5.

Fix a positive integer 0lNNS0\leq l\leq N-N_{S}.

  1. (i)

    If (mΛ)l<ΛNNS(m_{\Lambda})_{l<\Lambda\leq N-N_{S}} is given, this extends a unique family (mΛ)ΛNNS(m_{\Lambda})_{\Lambda\leq N-N_{S}} which is a weak solution of (9) for all αm𝒊\alpha^{\boldsymbol{i}}_{m} with 1ml1\leq m\leq l.

  2. (ii)

    If (mΛ)ΛNNS(m_{\Lambda})_{\Lambda\leq N-N_{S}} is a weak solution of (9) for all αm𝒊\alpha^{\boldsymbol{i}}_{m} with 1ml1\leq m\leq l, it actually satisfies (9) for αm𝒊\alpha^{\boldsymbol{i}}_{m} with 1ml1\leq m\leq l.

Let Π:QQΔS\Pi\colon Q\longrightarrow Q_{\Delta\setminus S} be the projection with respect to a basis Δ\Delta.

Proof.

(i) Fix a multi-index Γ\Gamma with lΓ|R+RS+|l\not<\Gamma\leq\lvert R^{+}\setminus R_{S}^{+}\rvert and consider a decomposition Γ=Γ+δi\Gamma=\Gamma^{\prime}+\delta_{i} such that γ1=γ2==γi1=0\gamma_{1}=\gamma_{2}=\cdots=\gamma_{i-1}=0 and γi=γi+1\gamma_{i}=\gamma^{\prime}_{i}+1. In this case ili\leq l holds.

We look at the coefficient of F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma^{\prime})} in the RHS of (9) with β=αi𝒊\beta=\alpha^{\boldsymbol{i}}_{i}. Note that Δ(E^𝒊,i)\Delta(\hat{E}_{\boldsymbol{i},i}) is a linear combination of KΛrα𝒊1E~𝒊ΛlE^𝒊ΛrK_{\Lambda^{r}\cdot\alpha^{\boldsymbol{i}}}^{-1}\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{l}}\otimes\hat{E}_{\boldsymbol{i}}^{\Lambda^{r}} with ΛriΛl\Lambda^{r}\leq i\leq\Lambda^{l} and Λlα𝒊+Λrα𝒊=αi𝒊\Lambda^{l}\cdot\alpha^{\boldsymbol{i}}+\Lambda^{r}\cdot\alpha^{\boldsymbol{i}}=\alpha^{\boldsymbol{i}}_{i} as we can see in (5).

If Π(Λlα𝒊)0\Pi(\Lambda^{l}\cdot\alpha^{\boldsymbol{i}})\neq 0 holds, pΓ((E^𝒊ΛrF^𝒊(Λ))(𝔲S))πS((E^𝒊ΛrF^𝒊(Λ))(𝔭S))0p_{\Gamma^{\prime}}((\hat{E}_{\boldsymbol{i}}^{\Lambda^{r}}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{u}_{S}^{-})})\pi_{S}((\hat{E}_{\boldsymbol{i}}^{\Lambda^{r}}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{p}_{S})})\neq 0 implies Π(Λα𝒊)=Π(Γα𝒊)+Π(Λrα𝒊)<Π(Γα𝒊)\Pi(\Lambda\cdot\alpha^{\boldsymbol{i}})=\Pi(\Gamma^{\prime}\cdot\alpha^{\boldsymbol{i}})+\Pi(\Lambda^{r}\cdot\alpha^{\boldsymbol{i}})<\Pi(\Gamma\cdot\alpha^{\boldsymbol{i}}).

If Π(Λlα𝒊)=0\Pi(\Lambda^{l}\cdot\alpha^{\boldsymbol{i}})=0 and λ1=λ2==λi1=0\lambda_{1}=\lambda_{2}=\cdots=\lambda_{i-1}=0 holds, Π(Λα𝒊)=Π(Γα𝒊)\Pi(\Lambda\cdot\alpha^{\boldsymbol{i}})=\Pi(\Gamma\cdot\alpha^{\boldsymbol{i}}) is necessary for pΓ((E^𝒊ΛrF^𝒊(Λ))(𝔲S))πS((E^𝒊ΛrF^𝒊(Λ))(𝔭S))0p_{\Gamma^{\prime}}((\hat{E}_{\boldsymbol{i}}^{\Lambda^{r}}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{u}_{S}^{-})})\pi_{S}((\hat{E}_{\boldsymbol{i}}^{\Lambda^{r}}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{p}_{S})})\neq 0 by the discussion above. Moreover Λr<i\Lambda^{r}<i holds when Λl0\Lambda^{l}\neq 0, since Λlα𝒊+Λrα𝒊=αi𝒊\Lambda^{l}\cdot\alpha^{\boldsymbol{i}}+\Lambda^{r}\cdot\alpha^{\boldsymbol{i}}=\alpha^{\boldsymbol{i}}_{i}. Then Proposition 3.4 (ii) implies pΓ((E^𝒊ΛrF^𝒊(Λ))(𝔲S))πS((E^𝒊ΛrF^𝒊(Λ))(𝔭S))h|Λl|Uhk(𝔩S)p_{\Gamma^{\prime}}((\hat{E}_{\boldsymbol{i}}^{\Lambda^{r}}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{u}_{S}^{-})})\pi_{S}((\hat{E}_{\boldsymbol{i}}^{\Lambda^{r}}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{p}_{S})})\in h^{\lvert\Lambda^{l}\rvert}U_{h}^{k}(\mathfrak{l}_{S}{}^{\sim}). On the other hand Λr=δi\Lambda^{r}=\delta_{i} holds when Λl=0\Lambda^{l}=0, since iΛri\leq\Lambda^{r} and Λα𝒊αi𝒊\Lambda^{\prime}\cdot\alpha^{\boldsymbol{i}}\neq\alpha^{\boldsymbol{i}}_{i} when i<Λi<\Lambda^{\prime} by [MR1169886, Theorem, p.662]. Hence we have to look at the coefficient of F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma^{\prime})}in the following summation:

ΛKβ1mΛ,(W)[E^𝒊,β,F^𝒊(Λ)]q,(𝔲S)πS([E^𝒊,β,F^𝒊(Λ)]q,(𝔭S))mΛ,(V).\displaystyle\sum_{\Lambda}K_{\beta}^{-1}m_{\Lambda,(W)}\otimes[\hat{E}_{\boldsymbol{i},\beta},\hat{F}_{\boldsymbol{i}}^{(\Lambda)}]_{q,(\mathfrak{u}_{S}^{-})}\otimes\pi_{S}([\hat{E}_{\boldsymbol{i},\beta},\hat{F}_{\boldsymbol{i}}^{(\Lambda)}]_{q,(\mathfrak{p}_{S})})m_{\Lambda,(V)}.

Consider a decomposition Λ=λiδi+Λ>i\Lambda=\lambda_{i}\delta_{i}+\Lambda_{>i} with Λ>i>i\Lambda_{>i}>i. Then we have

[E^𝒊,β,F^𝒊(Λ)]q=[E^𝒊,β,F^𝒊,β(λi)]qF^𝒊(Λ>i)+qβ2λiF^𝒊,i(λi)[E^𝒊,β,F^𝒊(Λ>i)]q.\displaystyle[\hat{E}_{\boldsymbol{i},\beta},\hat{F}_{\boldsymbol{i}}^{(\Lambda)}]_{q}=[\hat{E}_{\boldsymbol{i},\beta},\hat{F}_{\boldsymbol{i},\beta}^{(\lambda_{i})}]_{q}\hat{F}_{\boldsymbol{i}}^{(\Lambda_{>i})}+q_{\beta}^{2\lambda_{i}}\hat{F}_{\boldsymbol{i},i}^{(\lambda_{i})}[\hat{E}_{\boldsymbol{i},\beta},\hat{F}_{\boldsymbol{i}}^{(\Lambda_{>i})}]_{q}.

When Λ=Γ\Lambda=\Gamma, the first term produces F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma^{\prime})} and the second term does not. Moreover we can determine its coefficient completely since we have

[E^𝒊,β,F^𝒊,β(r)]q={qβ2rF^𝒊,β(r1)(qβ1rqβr1e2βK^2β)(βR0+),qβ2rF^𝒊,β(r1)(qβ1re2βqβr1K^2β)(βR0+).\displaystyle[\hat{E}_{\boldsymbol{i},\beta},\hat{F}_{\boldsymbol{i},\beta}^{(r)}]_{q}=\begin{cases}q_{\beta}^{2r}\hat{F}_{\boldsymbol{i},\beta}^{(r-1)}(q_{\beta}^{1-r}-q_{\beta}^{r-1}e_{-2\beta}\hat{K}_{-2\beta})&(\beta\in R_{0}^{+}),\\ q_{\beta}^{2r}\hat{F}_{\boldsymbol{i},\beta}^{(r-1)}(q_{\beta}^{1-r}e_{2\beta}-q_{\beta}^{r-1}\hat{K}_{-2\beta})&(\beta\not\in R_{0}^{+}).\end{cases}

Also note that the coefficient of Kβ1E^𝒊,βK_{\beta}^{-1}\otimes\hat{E}_{\boldsymbol{i},\beta} in Δ(E^𝒊,β)\Delta(\hat{E}_{\boldsymbol{i},\beta}) is 11 by (5).

When ΛΓ\Lambda\neq\Gamma, the coefficient of F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma^{\prime})} in [E^𝒊,β,F^𝒊,β(λi)]qF^𝒊(Λ>i)[\hat{E}_{\boldsymbol{i},\beta},\hat{F}_{\boldsymbol{i},\beta}^{(\lambda_{i})}]_{q}\hat{F}_{\boldsymbol{i}}^{(\Lambda_{>i})} is 0 and that in F^𝒊,i(λi)[E^𝒊,β,F^𝒊(Λ>i)]q\hat{F}_{\boldsymbol{i},i}^{(\lambda_{i})}[\hat{E}_{\boldsymbol{i},\beta},\hat{F}_{\boldsymbol{i}}^{(\Lambda_{>i})}]_{q} is in hUhk(𝔩S)hU_{h}^{k}(\mathfrak{l}_{S}) by Proposition 3.4 (i).

Take νQΔS+\nu\in Q_{\Delta\setminus S}^{+} and set Bl,ν:={ΓΠ(Γα𝒊)=ν,lΓ|R+RS+|}B_{l,\nu}:=\{\Gamma\mid\Pi(\Gamma\cdot\alpha^{\boldsymbol{i}})=\nu,\,l\not<\Gamma\leq\lvert R^{+}\setminus R_{S}^{+}\rvert\}. Then nl,ν:=|Bl,ν|<n_{l,\nu}:=\lvert B_{l,\nu}\rvert<\infty. We also consider an enumeration {Γi}i\{\Gamma_{i}\}_{i} on Bl,νB_{l,\nu} along the lexicographic order. Then the consideration above implies that we can find a nl,ν×nl,νn_{l,\nu}\times n_{l,\nu}-matrix Al,νA_{l,\nu} and (ui)i(W^V)nl,ν(u_{i})_{i}\in(W\hat{\otimes}V)^{n_{l,\nu}} such that

  • the ii-th entry of Al,ν(mΓi)i+(ui)iA_{l,\nu}(m_{\Gamma_{i}})_{i}+(u_{i})_{i} is the coefficient of F^𝒊(Γi)\hat{F}_{\boldsymbol{i}}^{(\Gamma^{\prime}_{i})} in the LHS of (9),

  • each entry of Al,νA_{l,\nu} is in Uhk(𝔤×𝔩S)U_{h}^{k}(\mathfrak{g}\times\mathfrak{l}_{S}),

  • each uiu_{i} is in Uhk(𝔤×𝔩S)U_{h}^{k}(\mathfrak{g}\times\mathfrak{l}_{S})-submodule generated by mΛm_{\Lambda} with Π(Λα𝒊)<ν\Pi(\Lambda\cdot\alpha^{\boldsymbol{i}})<\nu,

  • the matrix Al,νA_{l,\nu} is of the following form modulo hh:

    (σ1(1e2σ1β1)0σ2(1e2σ2β2)00σnl,ν(1e2σnl,νβnl,ν)),\displaystyle\begin{pmatrix}\sigma_{1}(1-e_{-2\sigma_{1}\beta_{1}})&\ast&\cdots&\ast\\ 0&\sigma_{2}(1-e_{-2\sigma_{2}\beta_{2}})&\cdots&\ast\\ \vdots&\vdots&\ddots&\vdots\\ 0&0&\cdots&\sigma_{n_{l,\nu}}(1-e_{-2\sigma_{n_{l,\nu}}\beta_{n_{l,\nu}}})\end{pmatrix},

    where σi=1\sigma_{i}=1 when βiR+R0+\beta_{i}\in R^{+}\cap R_{0}^{+} and σi=1\sigma_{i}=-1 when βiR+R0+\beta_{i}\in R^{+}\setminus R_{0}^{+}.

Since 1e2σiβi1-e_{2\sigma_{i}\beta_{i}} is also invertible for all ii, Al,νA_{l,\nu} is invertible. Hence we can solve the equations Al,ν(mΓi)i+(ui)i=0A_{l,\nu}(m_{\Gamma_{i}})_{i}+(u_{i})_{i}=0 inductively on ν\nu and obtain a unique weak solution.

(ii) We show the statement by induction on ll. If l=1l=1, there is nothing to prove. For induction step, we assume that the statement holds for l1l-1 with l2l\geq 2. Then any weak solution of (9)(\ref{eq:recursion formula}) for all αm𝒊\alpha^{\boldsymbol{i}}_{m} with 1ml1\leq m\leq l is actually a solution for all 1m<l1\leq m<l. Now we define mΛm^{\prime}_{\Lambda} as the coefficient of F^𝒊(Λ)\hat{F}_{\boldsymbol{i}}^{(\Lambda)} in the LHS of (9) for αl𝒊\alpha^{\boldsymbol{i}}_{l}. What we would like to show is mΛ=0m_{\Lambda}=0 for all Λ\Lambda. For this, it suffices to show that (mΛ)Λ(m^{\prime}_{\Lambda})_{\Lambda} is asolution of (9) for αm𝒊\alpha^{\boldsymbol{i}}_{m} with 1ml11\leq m\leq l-1 since the uniqueness part of (i) and mΛ=0m^{\prime}_{\Lambda}=0 for l<ΛNNSl<\Lambda\leq N-N_{S} implies mΛ=0m^{\prime}_{\Lambda}=0 for all Λ\Lambda. This follows from Proposition 3.3, which states that E^𝒊,mE^𝒊,l\hat{E}_{\boldsymbol{i},m}\hat{E}_{\boldsymbol{i},l} can be expressed as a linear combination of E^𝒊,lE^𝒊,m\hat{E}_{\boldsymbol{i},l}\hat{E}_{\boldsymbol{i},m} and other PBW vectors of the form E^𝒊Λ+\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}} with m<Λ+<lm<\Lambda^{+}<l. ∎

Proof of Proposition 4.4.

This is a special case of Lemma 4.5, namely the case of l=NNSl=N-N_{S}. ∎

Definition 4.6.

Set V=Uhk(𝔩S)V=U_{h}^{k}(\mathfrak{l}_{S}) and W=Uhk(𝔤)W=U_{h}^{k}(\mathfrak{g}). A generalized SS-maximal vector (UΛ)Λ(U_{\Lambda})_{\Lambda} with U0=1U_{0}=1 is called the SS-maximizer.

The SS-maximizer is a universal solution of (9) in the following sense.

Proposition 4.7.

Let (UΛ)Λ(U_{\Lambda})_{\Lambda} be the SS-maximizer and (mΛ)Λ(m_{\Lambda})_{\Lambda} is a generalized SS-maximal vector. Then mΛ=UΛm0m_{\Lambda}=U_{\Lambda}m_{0} holds for any Λ\Lambda.

Proof.

The statements immediately follows from the definition of generalized SS-maximal vector and the uniqueness part of Proposition 4.4. ∎

4.3. Construction of the 2-cocycles

Next we discuss a further property of the SS-maximizer. We begin with the following preparation.

Lemma 4.8.

Let 𝒪\mathcal{O} be the subalgebra of Uq𝒜(𝔤)U_{q}^{\mathcal{A}}(\mathfrak{g}) generated by {F~𝐢,α}αR+\{\widetilde{F}_{\boldsymbol{i},\alpha}\}_{\alpha\in R^{+}}, {E~𝐢,α}αR\{\widetilde{E}_{\boldsymbol{i},\alpha}\}_{\alpha\in R} and {Kμ}μP\{K_{\mu}\}_{\mu\in P}.

  1. (i)

    A family {F~𝒊ΛKμE~𝒊Λ+}Λ±,μ\{\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}K_{\mu}\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{+}}\}_{\Lambda^{\pm},\mu} is an 𝒜\mathcal{A}-basis of 𝒪\mathcal{O}.

  2. (ii)

    For any x,y𝒪x,y\in\mathcal{O}, [x,y]q±1(q1)𝒪[x,y]_{q^{\pm 1}}\in(q-1)\mathcal{O}.

Proof.

(i) This follows from Lemma 3.10 by specializing e2αe_{-2\alpha} to 11 for all αR+\alpha\in R^{+}.

(ii) We may assume both of x,yx,y are generators in the statement. Moreover we may assume x,yx,y are not KμK_{\mu} for any μ\mu since it is easy to see the statement in such a case.

We may assume that x=E~𝒊,αx=\widetilde{E}_{\boldsymbol{i},\alpha} with αΔ\alpha\in\Delta and y=F~𝒊,βy=\widetilde{F}_{\boldsymbol{i},\beta} with βR+\beta\in R^{+}, using Lemma 3.2. In particular yUq𝒜(𝔫)y\in U_{q}^{\mathcal{A}}(\mathfrak{n}^{-}). Hence [E´𝒊,α,y]q[\acute{E}_{\boldsymbol{i},\alpha},y]_{q} is an element of Uq𝒜(𝔟)U_{q}^{\mathcal{A}}(\mathfrak{b}^{-}). On the other hand span𝒜{F~𝒊ΛKμE´𝒊(Λ+)}\mathrm{span}_{\mathcal{A}}\{\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}K_{\mu}\acute{E}_{\boldsymbol{i}}^{(\Lambda^{+})}\} is closed under the multiplication. Hence [E´𝒊,α,y]q[\acute{E}_{\boldsymbol{i},\alpha},y]_{q} is in the intersection of these subalgebras, which is spanned by {F~𝒊ΛKμ}Λ,μ\{\widetilde{F}_{\boldsymbol{i}}^{\Lambda^{-}}K_{\mu}\}_{\Lambda^{-},\mu}. This means [x,y](qαqα1)𝒪[x,y]\in(q_{\alpha}-q_{\alpha}^{-1})\mathcal{O} since x=(qαqα1)E´𝒊,αx=(q_{\alpha}-q_{\alpha}^{-1})\acute{E}_{\boldsymbol{i},\alpha}. ∎

Proposition 4.9.

Let (UΛ)Λ(U_{\Lambda})_{\Lambda} be the SS-maximizer. Then UΛh|Λ|Uhk(𝔤×𝔩S)U_{\Lambda}\in h^{\lvert\Lambda\rvert}U_{h}^{k}(\mathfrak{g}\times\mathfrak{l}_{S}) for any Λ\Lambda.

Proof.

We construct a weak solution of (9) for all βR+RS+\beta\in R^{+}\setminus R_{S}^{+} by another twisted parabolic induction using Uh,χk(𝔤)U_{h,\chi}^{k}({}^{\sim}\mathfrak{g}) i.e. Uh,χk(𝔤)Uh,χk(𝔭S)VU_{h,\chi}^{k}({}^{\sim}\mathfrak{g})\otimes_{U_{h,\chi}^{k}({}^{\sim}\mathfrak{p}_{S})}V. Note that this can be naturally embedded in ind𝔩S,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V.

In this case we consider the following equation for each βR+R0+\beta\in R^{+}\setminus R_{0}^{+}:

(10) ΛEˇ𝒊,β,(1)m~Λ,(W)(Eˇ𝒊,β,(2)Fˇ𝒊Λ)(𝔲S)πS((Eˇ𝒊,β,(2)Fˇ𝒊Λ)(𝔭S))m~Λ,(V)=0.\displaystyle\sum_{\Lambda}\check{E}_{\boldsymbol{i},\beta,(1)}\widetilde{m}_{\Lambda,(W)}\otimes(\check{E}_{\boldsymbol{i},\beta,(2)}\check{F}_{\boldsymbol{i}}^{\Lambda})_{(\mathfrak{u}_{S}^{-})}\otimes\pi_{S}((\check{E}_{\boldsymbol{i},\beta,(2)}\check{F}_{\boldsymbol{i}}^{\Lambda})_{(\mathfrak{p}_{S})})\widetilde{m}_{\Lambda,(V)}=0.

Then we can define the notion of a weak solution of (10) in a similar manner to that of (9). If this equation has a weak solution for any m=m~0m=\widetilde{m}_{0}, we can see that the statement is valid since we can obtain a weak solution (mΛ)Λ(m_{\Lambda})_{\Lambda} of (9) for all βR+RS+\beta\in R^{+}\setminus R_{S}^{+} by setting mΛ:=(qα𝒊qα𝒊1)Λ[Λ]qα𝒊!m~Λm_{\Lambda}:=(q_{\alpha^{\boldsymbol{i}}}-q_{\alpha^{\boldsymbol{i}}}^{-1})^{\Lambda}[\Lambda]_{q_{\alpha^{\boldsymbol{i}}}}!\widetilde{m}_{\Lambda}. Then Lemma 4.5 (ii) implies that this is actually a generalized SS-maximal vector. In particular we can see mΛh|Λ|W^Vm_{\Lambda}\in h^{\lvert\Lambda\rvert}W\hat{\otimes}V.

The outline of the proof is almost the same as that of the proof of Lemma 4.5 with l=NNSl=N-N_{S}. The only different point is that AνA_{\nu}, the matrix of the coefficient of Fˇ𝒊Γ\check{F}_{\boldsymbol{i}}^{\Gamma^{\prime}} in the term of Λ\Lambda with Π(Λα𝒊)=Π(Γα𝒊)=ν\Pi(\Lambda\cdot\alpha^{\boldsymbol{i}})=\Pi(\Gamma\cdot\alpha^{\boldsymbol{i}})=\nu, is a lower triangular matrix with invertible diagonal entries modulo hh. In the following, we prove this fact.

Fix Γ\Gamma and take Γ\Gamma^{\prime} and β=αi𝒊\beta=\alpha^{\boldsymbol{i}}_{i} in the same way as Proof of Lemma 4.5. Noting that Uh(𝔤)U_{h}(\mathfrak{g}) is cocommutative modulo hh, we can see that Δ(Eˇ𝒊,β)=Eˇ𝒊,β1+Kβ1Eˇ𝒊,β\Delta(\check{E}_{\boldsymbol{i},\beta})=\check{E}_{\boldsymbol{i},\beta}\otimes 1+K_{\beta}^{-1}\otimes\check{E}_{\boldsymbol{i},\beta} modulo hh and the first term does not affect to the entries of AνA_{\nu}. Hence it suffices to look at the coefficient of Fˇ𝒊Γ\check{F}_{\boldsymbol{i}}^{\Gamma^{\prime}} in the following:

ΛKβ1m~Λ,(W)[Eˇ𝒊,β,Fˇ𝒊Λ]q,(𝔲S)πS([Eˇ𝒊,β,Fˇ𝒊Λ]q,(𝔭S))m~Λ,(V).\displaystyle\sum_{\Lambda}K_{\beta}^{-1}\widetilde{m}_{\Lambda,(W)}\otimes[\check{E}_{\boldsymbol{i},\beta},\check{F}_{\boldsymbol{i}}^{\Lambda}]_{q,(\mathfrak{u}_{S}^{-})}\otimes\pi_{S}([\check{E}_{\boldsymbol{i},\beta},\check{F}_{\boldsymbol{i}}^{\Lambda}]_{q,(\mathfrak{p}_{S})})\widetilde{m}_{\Lambda,(V)}.

We consider the following decomposition of a quantum commutator:

(11) =q[Eˇ𝒊,β,Fˇ𝒊Λ<i]qFˇ𝒊λiFˇ𝒊Λ>i\displaystyle{}_{q}=[\check{E}_{\boldsymbol{i},\beta},\check{F}_{\boldsymbol{i}}^{\Lambda_{<i}}]_{q}\check{F}_{\boldsymbol{i}}^{\lambda_{i}}\check{F}_{\boldsymbol{i}}^{\Lambda_{>i}} +q(β,Λ<iα𝒊)Fˇ𝒊Λ<i[Eˇ𝒊,β,Fˇ𝒊,iλi]qFˇ𝒊Λ>i\displaystyle+q^{(\beta,\Lambda_{<i}\cdot\alpha^{\boldsymbol{i}})}\check{F}_{\boldsymbol{i}}^{\Lambda_{<i}}[\check{E}_{\boldsymbol{i},\beta},\check{F}_{\boldsymbol{i},i}^{\lambda_{i}}]_{q}\check{F}_{\boldsymbol{i}}^{\Lambda_{>i}}
+q(β,Λ<iα𝒊+λiαi𝒊)Fˇ𝒊Λ<iFˇ𝒊,iλi[Eˇ𝒊,β,Fˇ𝒊Λ>i]q\displaystyle+q^{(\beta,\Lambda_{<i}\cdot\alpha^{\boldsymbol{i}}+\lambda_{i}\alpha^{\boldsymbol{i}}_{i})}\check{F}_{\boldsymbol{i}}^{\Lambda_{<i}}\check{F}_{\boldsymbol{i},i}^{\lambda_{i}}[\check{E}_{\boldsymbol{i},\beta},\check{F}_{\boldsymbol{i}}^{\Lambda_{>i}}]_{q}

Assume Λ<i0\Lambda_{<i}\neq 0. It is easy to see that the second and third terms do not affect the coefficient of Fˇ𝒊Γ\check{F}_{\boldsymbol{i}}^{\Gamma^{\prime}}. We show that even the first term does not modulo hh. By Proposition 3.4 and Lemma 4.8 we have the following expansion:

[Eˇ𝒊,β,Fˇ𝒊Λ<i]q=Λ<iΛ+,μβΛ+α𝒊=Λiα𝒊Λα𝒊CΛ±,μFˇ𝒊ΛKμE^𝒊Λ+\displaystyle[\check{E}_{\boldsymbol{i},\beta},\check{F}_{\boldsymbol{i}}^{\Lambda_{<i}}]_{q}=\sum_{\begin{subarray}{c}\Lambda^{-}<i\leq\Lambda^{+},\mu\\ \beta-\Lambda^{+}\cdot\alpha^{\boldsymbol{i}}=\Lambda_{i}\cdot\alpha^{\boldsymbol{i}}-\Lambda^{-}\cdot\alpha^{\boldsymbol{i}}\end{subarray}}C_{\Lambda^{\pm},\mu}\check{F}_{\boldsymbol{i}}^{\Lambda^{-}}K_{\mu}\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}}

Then Lemma 4.8 allows us to expand the first term as follows modulo hh:

Λ<iΛ+,μβΛ+α𝒊=Λiα𝒊Λα𝒊CΛ±,μFˇ𝒊ΛFˇ𝒊λiFˇ𝒊Λ>iKμE^𝒊Λ+.\displaystyle\sum_{\begin{subarray}{c}\Lambda^{-}<i\leq\Lambda^{+},\mu\\ \beta-\Lambda^{+}\cdot\alpha^{\boldsymbol{i}}=\Lambda_{i}\cdot\alpha^{\boldsymbol{i}}-\Lambda^{-}\cdot\alpha^{\boldsymbol{i}}\end{subarray}}C_{\Lambda^{\pm},\mu}\check{F}_{\boldsymbol{i}}^{\Lambda^{-}}\check{F}_{\boldsymbol{i}}^{\lambda_{i}}\check{F}_{\boldsymbol{i}}^{\Lambda_{>i}}K_{\mu}\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}}.

In this summation Fˇ𝒊Γ\check{F}_{\boldsymbol{i}}^{\Gamma^{\prime}} appears as the Uh,χk(𝔲S)U_{h,\chi}^{k}({}^{\sim}\mathfrak{u}_{S}^{-})-part only when Λ=0\Lambda^{-}=0, in which case Λ+0\Lambda^{+}\neq 0 since Λ+α𝒊Λα𝒊=βΛ<iα𝒊0\Lambda^{+}\cdot\alpha^{\boldsymbol{i}}-\Lambda^{-}\cdot\alpha^{\boldsymbol{i}}=\beta-\Lambda_{<i}\cdot\alpha^{\boldsymbol{i}}\neq 0 by [MR1169886, Theorem, p. 662]. If Λ+>NNS\Lambda^{+}>N-N_{S} holds additionally, Uhk(𝔩S)U_{h}^{k}(\mathfrak{l}_{S})-part of such a term is 0modh0\operatorname{\mathrm{mod}}h since E^𝒊Λ+hUhk(𝔩S)\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}}\in hU_{h}^{k}(\mathfrak{l}_{S}). If ΛNNS\Lambda^{-}\not>N-N_{S} holds, such a term is removed by πS\pi_{S}. Hence we can ignore Λ\Lambda with λ1=λ2==λi1=0\lambda_{1}=\lambda_{2}=\cdots=\lambda_{i-1}=0.

Next assume Λ<i=0\Lambda_{<i}=0. The case of Λ=Γ\Lambda=\Gamma can be treated in the same way as the proof of Lemma 4.5. When ΛΓ\Lambda\neq\Gamma, the second term does not affect the coefficient of Fˇ𝒊Γ\check{F}_{\boldsymbol{i}}^{\Gamma^{\prime}} and the third term does only when λi=γi1<γi\lambda_{i}=\gamma_{i}-1<\gamma_{i}.

Therefore AνA_{\nu} is of the following form modulo hh:

(12) ([γ1,i1]qβ1σ1(1e2σ1β1)0[γl,inν]qβnνσnν(1e2σnνβnν)),\displaystyle\begin{pmatrix}[\gamma_{1,i_{1}}]_{q_{\beta_{1}}}\sigma_{1}(1-e_{-2\sigma_{1}\beta_{1}})&\cdots&0\\ \vdots&\ddots&\vdots\\ \ast&\cdots&[\gamma_{l,i_{n_{\nu}}}]_{q_{\beta_{n_{\nu}}}}\sigma_{n_{\nu}}(1-e_{-2\sigma_{n_{\nu}}\beta_{n_{\nu}}})\end{pmatrix},

where Γk=Γk+δik\Gamma_{k}=\Gamma^{\prime}_{k}+\delta_{i_{k}} and βk=αik𝒊\beta_{k}=\alpha^{\boldsymbol{i}}_{i_{k}} are taken as before. Since [γk,ik]qβk[\gamma_{k,i_{k}}]_{q_{\beta_{k}}} is invertible in our setting, we can see the existence of a weak solution of (10) for all βR+RS+\beta\in R^{+}\setminus R_{S}^{+}. ∎

Corollary 4.10.

Let VV be a Uhk(𝔩S)U_{h}^{k}(\mathfrak{l}_{S})-module and WW be a Uhk(𝔤)U_{h}^{k}(\mathfrak{g})-module. Then there is a unique SS-maximal vector of the form m(W)^(1m(V))+m_{(W)}\hat{\otimes}(1\otimes m_{(V)})+\cdots for any mW^Vm\in W\hat{\otimes}V.

Proof.

By the definition of the completion, we can consider Λ(UΛm)(W)F^𝒊(Λ)(UΛm)(V)\sum_{\Lambda}(U_{\Lambda}m)_{(W)}\otimes\hat{F}_{\boldsymbol{i}}^{(\Lambda)}\otimes(U_{\Lambda}m)_{(V)}, which is the SS-maximizer. ∎

Remark 4.11.

If VV and WW are finite integrable, there is no need to take the completion.

Remark 4.12.

Since 1v1\otimes v is SS-maximal in ind𝔩S,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V, the uniqueness implies that an SS-maximal vector in ind𝔩S,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V must be of the form 1v1\otimes v.

Definition 4.13.

Let Reph,χf𝔩S\operatorname{\mathrm{Rep}}_{h,\chi}^{\mathrm{f}}\mathfrak{l}_{S} be the full subcategory of Uh,χk(𝔤)-ModU_{h,\chi}^{k}(\mathfrak{g}{}^{\sim})\text{-}\mathrm{Mod} consisting of objects which are isomorphic to ind𝔩S,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V for some VRephf𝔩SV\in\operatorname{\mathrm{Rep}}^{\mathrm{f}}_{h}\mathfrak{l}_{S}.

Proposition 4.14.

The following facts hold:

  1. (i)

    The functor ind𝔩S,q𝔤,χ\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi} restricts to an equivalence of category between Rephf𝔩S\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{l}_{S} and Reph,χf𝔩S\operatorname{\mathrm{Rep}}_{h,\chi}^{\mathrm{f}}\mathfrak{l}_{S}.

  2. (ii)

    For VRephf𝔩SV\in\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{l}_{S} and WRephf𝔤W\in\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{g}, there is a natural isomorphism from ind𝔩S,q𝔤,χ(WV)\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}(W\otimes V) to Wind𝔩S,q𝔤,χVW\otimes\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V such that 1(wv)w(1v)+1\otimes(w\otimes v)\longmapsto w\otimes(1\otimes v)+\cdots.

Therefore Reph,χf𝔩S\operatorname{\mathrm{Rep}}_{h,\chi}^{\mathrm{f}}\mathfrak{l}_{S} naturally becomes a semisimple left Rephf𝔤\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{g}-module category.

Proof.

Using the universal property of (R0+,χ)(R_{0}^{+},\chi)-parabolic induction (Lemma 3.17), we can deduce (i) from Remark 4.12.

Next we prove (ii). By the universal property and Corollary 4.10 we can see that a homomorphism stated in (ii) actually exists and is unique. Hence it suffices to show that this is an isomorphism. Consider the following composition of homomorphisms:

Wind𝔩S,q𝔤,χV\displaystyle W\otimes\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V Wind𝔩S,q𝔤,χ(WWV)\displaystyle\longrightarrow W\otimes\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}(W^{*}\otimes W\otimes V)
WWind𝔩S,q𝔤,χ(WV)ind𝔩S,q𝔤,χ(WV),\displaystyle\longrightarrow W\otimes W^{*}\otimes\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}(W\otimes V)\longrightarrow\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}(W\otimes V),

where khWWk\llbracket h\rrbracket\longrightarrow W^{*}\otimes W is given by 1ieiei1\longmapsto\sum_{i}e^{i}\otimes e_{i} with (ei)i(e_{i})_{i} is a basis of WW and (ei)i(e^{i})_{i} is its dual basis of WW^{*}, and WWkhW\otimes W^{*}\longrightarrow k\llbracket h\rrbracket is given by wff(K2ρw)w\otimes f\longmapsto f(K_{2\rho}w) where ρ\rho is the half sum of positive roots. Then the composition of this homomorphism and ind𝔩S,q𝔤,χ(WV)Wind𝔩S,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}(W\otimes V)\longrightarrow W\otimes\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V maps 1(wv)1\otimes(w\otimes v) to 1(wv)modh1\otimes(w\otimes v)\operatorname{\mathrm{mod}}h and is isomorphism. This fact implies the injectivity of the homomorphism in the statement. We can see surjectivity in a similar way. ∎

We rewrite Reph,χf𝔩S\operatorname{\mathrm{Rep}}_{h,\chi}^{\mathrm{f}}\mathfrak{l}_{S} as a 22-cocycle twist of Rephf𝔩S\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{l}_{S}. For VRephf𝔩SV\in\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{l}_{S} and W,WRephf𝔤W,W^{\prime}\in\operatorname{\mathrm{Rep}}_{h}^{\mathrm{f}}\mathfrak{g}, we consider the following isomorphism:

(13) ind𝔩S,q𝔤,χ(WWV)\displaystyle\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}(W\otimes W^{\prime}\otimes V) Wind𝔩S,q𝔤,χ(WV)\displaystyle\cong W\otimes\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}(W^{\prime}\otimes V)
(WW)ind𝔩S,q𝔤,χVind𝔩S,q𝔤,χ(WWV).\displaystyle\cong(W\otimes W^{\prime})\otimes\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V\cong\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}(W\otimes W^{\prime}\otimes V).

Proposition 4.14 (i) implies that this automorphism is induced from an automorphism on WWVW\otimes W^{\prime}\otimes V. Actually we have a stronger statement. The centralizer of Uhk(𝔩S)U_{h}^{k}(\mathfrak{l}_{S}) embedded in Uhk(𝔤×𝔤×𝔩S)U_{h}^{k}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S}) via (Δid)Δ(\Delta\otimes\operatorname{\mathrm{id}})\Delta is denoted by Uhk(𝔤×𝔤×𝔩S)𝔩SU_{h}^{k}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S})^{\mathfrak{l}_{S}}.

Proposition 4.15.

There is a unique Ψh,χkUh(𝔤×𝔤×𝔩S)𝔩S\Psi_{h,\chi}^{k}\in U_{h}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S})^{\mathfrak{l}_{S}} such that the isomorphism (13) is induced from the mulitplication by Ψh,χ\Psi_{h,\chi}. Moreover Ψh,χ\Psi_{h,\chi} satisfies the cocyle identities

(14) (idΔid)(Ψh,χ)(1Ψh,χ)=(Δidid)(Ψh,χ)(ididΔ)(Ψh,χ),\displaystyle(\operatorname{\mathrm{id}}\otimes\Delta\otimes\operatorname{\mathrm{id}})(\Psi_{h,\chi})(1\otimes\Psi_{h,\chi})=(\Delta\otimes\operatorname{\mathrm{id}}\otimes\operatorname{\mathrm{id}})(\Psi_{h,\chi})(\operatorname{\mathrm{id}}\otimes\operatorname{\mathrm{id}}\otimes\Delta)(\Psi_{h,\chi}),
(15) (εidid)(Ψh,χ)=(idεid)(Ψh,χ)=1\displaystyle(\varepsilon\otimes\operatorname{\mathrm{id}}\otimes\operatorname{\mathrm{id}})(\Psi_{h,\chi})=(\operatorname{\mathrm{id}}\otimes\varepsilon\otimes\operatorname{\mathrm{id}})(\Psi_{h,\chi})=1

and has the following expansion:

(16) Ψh,χ=1h\displaystyle\Psi_{h,\chi}=1-h {βR+R0+RS+2dβ1χ2βE𝒊,βF𝒊,β1\displaystyle\left\{\sum_{\beta\in R^{+}\cap R_{0}^{+}\setminus R_{S}^{+}}\frac{2d_{\beta}}{1-\chi_{-2\beta}}E_{\boldsymbol{i},\beta}\otimes F_{\boldsymbol{i},\beta}\otimes 1\right.
(17) +βR+R0+2dβχ2βχ2β1E𝒊,βF𝒊,β1}+.\displaystyle\hskip 50.00008pt\left.+\sum_{\beta\in R^{+}\setminus R_{0}^{+}}\frac{2d_{\beta}\chi_{2\beta}}{\chi_{2\beta}-1}E_{\boldsymbol{i},\beta}\otimes F_{\boldsymbol{i},\beta}\otimes 1\right\}+\cdots.
Proof.

Fix uWWVu\in W\otimes W^{\prime}\otimes V. Let uu^{\prime} be the image of 1u1\otimes u at WWind𝔩S,q𝔤,χVW\otimes W^{\prime}\otimes\mathrm{ind}_{\mathfrak{l}_{S},q}^{\mathfrak{g},\chi}V in (13) and ΨW,W,VEndUhk(𝔩S)(WWV)\Psi_{W,W^{\prime},V}\in\mathrm{End}_{U_{h}^{k}(\mathfrak{l}_{S})}(W\otimes W^{\prime}\otimes V) be the isomorphism corresponding to (13). Then we have

u=ΨW,W,V(u)(WW)(1ΨW,W,V(u)(V))+.\displaystyle u^{\prime}=\Psi_{W,W^{\prime},V}(u)_{(W\otimes W^{\prime})}\otimes(1\otimes\Psi_{W,W^{\prime},V}(u)_{(V)})+\cdots.

On the other hand, by straightforward calculation, we also have

u=Λ\displaystyle u^{\prime}=\sum_{\Lambda} ((idΔ)(UΛ)u)(W)\displaystyle((\operatorname{\mathrm{id}}\otimes\Delta)(U_{\Lambda})u)_{(W)}
F^𝒊(Λ)(((idΔ)(UΛ)u)(W)(1((idΔ)(UΛ)u)(V))+).\displaystyle\otimes\hat{F}_{\boldsymbol{i}}^{(\Lambda)}(((\operatorname{\mathrm{id}}\otimes\Delta)(U_{\Lambda})u)_{(W^{\prime})}\otimes(1\otimes((\operatorname{\mathrm{id}}\otimes\Delta)(U_{\Lambda})u)_{(V)})+\cdots).

Hence we can obtain Ψh,χ\Psi_{h,\chi} as follows:

(18) Ψh,χ=Λ(1F^𝒊,(1)ΛπS(F^𝒊,(2)Λ))(idΔ)(UΛ).\displaystyle\Psi_{h,\chi}=\sum_{\Lambda}(1\otimes\hat{F}_{\boldsymbol{i},(1)}^{\Lambda}\otimes\pi_{S}(\hat{F}_{\boldsymbol{i},(2)}^{\Lambda}))(\operatorname{\mathrm{id}}\otimes\Delta)(U_{\Lambda}).

Note that this summation is well-defined since UΛh|Λ|Uhk(𝔤×𝔤×𝔩S)U_{\Lambda}\in h^{\lvert\Lambda\rvert}U_{h}^{k}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S}).

The cocycle identities (14), (15) can be obtained as consequences of direct calculations. To prove the expansion, we use the formula (18). Since UΛh|Λ|Uhk(𝔤×𝔤×𝔩S)U_{\Lambda}\in h^{\lvert\Lambda\rvert}U_{h}^{k}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S}) and

Δ(F^𝒊,β)={F𝒊,β1+Kβ1F𝒊,βmodh(βR+R0+),χ2βF𝒊,β1+Kβ1F𝒊,βmodh(βR+R0+),\displaystyle\Delta(\hat{F}_{\boldsymbol{i},\beta})=\begin{cases}F_{\boldsymbol{i},\beta}\otimes 1+K_{\beta}^{-1}\otimes F_{\boldsymbol{i},\beta}&\operatorname{\mathrm{mod}}h\quad(\beta\in R^{+}\cap R_{0}^{+}),\\ \chi_{2\beta}F_{\boldsymbol{i},\beta}\otimes 1+K_{\beta}^{-1}\otimes F_{\boldsymbol{i},\beta}&\operatorname{\mathrm{mod}}h\quad(\beta\in R^{+}\setminus R_{0}^{+}),\end{cases}

it suffices to calculate the following summation:

βR+R0+(1F𝒊,β1)(idΔ)(Uβ)+βR+R0+χ2β(1F𝒊,β1)(idΔ)(Uβ),\displaystyle\sum_{\beta\in R^{+}\cap R_{0}^{+}}(1\otimes F_{\boldsymbol{i},\beta}\otimes 1)(\operatorname{\mathrm{id}}\otimes\Delta)(U_{\beta})+\sum_{\beta\in R^{+}\setminus R_{0}^{+}}\chi_{2\beta}(1\otimes F_{\boldsymbol{i},\beta}\otimes 1)(\operatorname{\mathrm{id}}\otimes\Delta)(U_{\beta}),

where Uβ:=UδkU_{\beta}:=U_{\delta_{k}} with β=αk𝒊\beta=\alpha^{\boldsymbol{i}}_{k}. To determine UβU_{\beta} modulo hh for each β=αk𝒊\beta=\alpha^{\boldsymbol{i}}_{k}, we go back to (12) and show that entries in the ii-th row of AΠ(β)A_{\Pi(\beta)}, where Γi=δk\Gamma_{i}=\delta_{k}, are 0 modulo hh except the diagonal entry. Let Λ\Lambda be a multi-index with Π(Λα𝒊)=Π(β)\Pi(\Lambda\cdot\alpha^{\boldsymbol{i}})=\Pi(\beta) and Λk\Lambda\geq k and Λδik\Lambda\neq\delta_{i_{k}}. Then we have Λ>k\Lambda>k, in which case we can use Proposition 3.4 to obtain the following expansion:

[Eˇ𝒊,k,Fˇ𝒊Λ]=Λ+<kΛCΛ±Fˇ𝒊ΛEˇ𝒊(Λ+).\displaystyle[\check{E}_{\boldsymbol{i},k},\check{F}_{\boldsymbol{i}}^{\Lambda}]=\sum_{\Lambda^{+}<k\leq\Lambda^{-}}C_{\Lambda^{\pm}}\check{F}_{\boldsymbol{i}}^{\Lambda^{-}}\check{E}_{\boldsymbol{i}}^{(\Lambda^{+})}.

Since πS(Eˇ𝒊Λ+)=0\pi_{S}(\check{E}_{\boldsymbol{i}}^{\Lambda^{+}})=0 whenever Λ+0\Lambda^{+}\neq 0, it suffices to look at the terms with Λ+\Lambda^{+}. Moreover Γi=0+δk\Gamma_{i}=0+\delta_{k} implies that we may additionally assume Π(Λα𝒊)=0\Pi(\Lambda^{-}\cdot\alpha^{\boldsymbol{i}})=0. Such a term is 0 modulo hh in Uhk(𝔭S)U_{h}^{k}(\mathfrak{p}_{S}) if Λ0\Lambda^{-}\neq 0, which is automatically satisfied since Λα𝒊αk𝒊\Lambda\cdot\alpha^{\boldsymbol{i}}\neq\alpha^{\boldsymbol{i}}_{k} by [MR1169886, Theorem, p.662] and [MR1890629, Chapter VI, Proposition 19].

Therefore we have

Uβ=σqβqβ11χ2σβE𝒊,βKβ11=σ2dβ1χ2σβE𝒊,β1modh\displaystyle U_{\beta}=-\sigma\frac{q_{\beta}-q_{\beta}^{-1}}{1-\chi_{-2\sigma\beta}}E_{\boldsymbol{i},\beta}K_{\beta}^{-1}\otimes 1=-\sigma\frac{2d_{\beta}}{1-\chi_{-2\sigma\beta}}E_{\boldsymbol{i},\beta}\otimes 1\quad\operatorname{\mathrm{mod}}h

where σ=1\sigma=1 when βR0+\beta\in R_{0}^{+} and σ=1\sigma=-1 when βR0+\beta\not\in R_{0}^{+} and obtain the expansion (16). ∎

4.4. Construction of quantum semisimple coadjoint orbits

Let GG be a 11-connected complex Lie group with the Lie algebra 𝔤\mathfrak{g} and LSL_{S} be its connected complex Lie subgroup with the Lie algebra 𝔩S\mathfrak{l}_{S}. In this subsection we use Ψh,χ\Psi_{h,\chi} to construct a quantum semisimple coadjoint orbit, which is a deformation quantization of 𝒪(LS\G)\mathcal{O}(L_{S}\backslash G) with an action of Uh(𝔤)U_{h}(\mathfrak{g}). Let Irrh𝔤\operatorname{\mathrm{Irr}}_{h}\mathfrak{g} be the set of equivalence classes of irreducible finite integrable Uhk(𝔤)U_{h}^{k}(\mathfrak{g}) modules. We also fix a representative π\pi of each class and sometimes identify Irrh𝔤\operatorname{\mathrm{Irr}}_{h}\mathfrak{g} as the set of all representatives.

For a Uhk(𝔤)U_{h}^{k}(\mathfrak{g})-module VV, its Uhk(𝔩S)U_{h}^{k}(\mathfrak{l}_{S})-invariant part is denoted by V𝔩SV^{\mathfrak{l}_{S}}.

Set 𝒪h,χk(LS\G)\mathcal{O}_{h,\chi}^{k}(L_{S}\backslash G) as follows:

(19) 𝒪h,χk(LS\G)=πIrrh𝔤HomUhk(𝔩S)(Vπ,kh)Vπ.\displaystyle\mathcal{O}_{h,\chi}^{k}(L_{S}\backslash G)=\bigoplus_{\pi\in\operatorname{\mathrm{Irr}}_{h}\mathfrak{g}}\operatorname{\mathrm{Hom}}_{U_{h}^{k}(\mathfrak{l}_{S})}(V_{\pi},k\llbracket h\rrbracket)\otimes V_{\pi}.

Note that there is a unique linear map pV:HomUhk(𝔩S)(V,kh)V𝒪h,χk(LS\G)p_{V}\colon\operatorname{\mathrm{Hom}}_{U_{h}^{k}(\mathfrak{l}_{S})}(V,k\llbracket h\rrbracket)\otimes V\longrightarrow\mathcal{O}_{h,\chi}^{k}(L_{S}\backslash G) for each finite integrable module VV which is natural in the sense that pV(fTv)=pW(fT(v))p_{V}(f\circ T\otimes v)=p_{W}(f\otimes T(v)) for any fHomUhk(𝔩S,kh),vVf\in\operatorname{\mathrm{Hom}}_{U_{h}^{k}}(\mathfrak{l}_{S},k\llbracket h\rrbracket),v\in V and THom𝔤h(V,W)T\in\operatorname{\mathrm{Hom}}_{\mathfrak{g}_{h}}(V,W). For ease to read, we use fvf\otimes v to represent pV(fv)p_{V}(f\otimes v).

Using this we construct a product h,χ\ast_{h,\chi} on 𝒪h,χk(LS\G)\mathcal{O}_{h,\chi}^{k}(L_{S}\backslash G) defined as follows by using Ψh,χ,0=(ididε)(Ψh,χ)\Psi_{h,\chi,0}=(\operatorname{\mathrm{id}}\otimes\operatorname{\mathrm{id}}\otimes\varepsilon)(\Psi_{h,\chi}), which is thought as a Uhk(𝔩S)U_{h}^{k}(\mathfrak{l}_{S})-homomorphism:

(fv)h,χ(gw)=(fg)Ψh,χ,01(vw)\displaystyle(f\otimes v)\ast_{h,\chi}(g\otimes w)=(f\otimes g)\Psi_{h,\chi,0}^{-1}\otimes(v\otimes w)

This product is actually associative. To see this consider the product of three elements in different order:

((fu)h,χ(gv))\displaystyle((f\otimes u)\ast_{h,\chi}(g\otimes v)) h,χ(hw)\displaystyle\ast_{h,\chi}(h\otimes w)
=(fgh)(Ψh,χ,011)(Δid)(Ψh,χ,01)(uvw),\displaystyle=(f\otimes g\otimes h)(\Psi_{h,\chi,0}^{-1}\otimes 1)(\Delta\otimes\operatorname{\mathrm{id}})(\Psi_{h,\chi,0}^{-1})\otimes(u\otimes v\otimes w),
(fu)h,χ((gv)\displaystyle(f\otimes u)\ast_{h,\chi}((g\otimes v) h,χ(hw))\displaystyle\ast_{h,\chi}(h\otimes w))
=(fgh)(1Ψh,χ,01)(idΔ)(Ψh,χ,01)(uvw).\displaystyle=(f\otimes g\otimes h)(1\otimes\Psi_{h,\chi,0}^{-1})(\operatorname{\mathrm{id}}\otimes\Delta)(\Psi_{h,\chi,0}^{-1})\otimes(u\otimes v\otimes w).

Since hx=ε(x)hhx=\varepsilon(x)h for all xUhk(𝔩S)x\in U_{h}^{k}(\mathfrak{l}_{S}), we can see that (fgh)(Ψh,χ,011)=(fgh)Ψh,χ1(f\otimes g\otimes h)(\Psi_{h,\chi,0}^{-1}\otimes 1)=(f\otimes g\otimes h)\Psi_{h,\chi}^{-1}. Then the associativity follows from the cocycle identity (14).

When k=k=\mathbb{C}, 𝒪h,χ(LS\G)\mathcal{O}_{h,\chi}(L_{S}\backslash G) has a natural structure of deformation quantizations of 𝒪(LS\G)\mathcal{O}(L_{S}\backslash G). To determine the semi-classical limit of 𝒪h,χ(LS\G)\mathcal{O}_{h,\chi}(L_{S}\backslash G), we recall the identification of Uh(𝔤)U_{h}(\mathfrak{g}) and the twist of U(𝔤)hU(\mathfrak{g})\llbracket h\rrbracket. For any subset SΔS\subset\Delta we can take an isomorphism T:U(𝔤)hUh(𝔤)T\colon U(\mathfrak{g})\llbracket h\rrbracket\longrightarrow U_{h}(\mathfrak{g}) as topological h\mathbb{C}\llbracket h\rrbracket-algebra such that

  1. (i)

    T(U(𝔩S)h)=Uh(𝔩S)T(U(\mathfrak{l}_{S})\llbracket h\rrbracket)=U_{h}(\mathfrak{l}_{S}).

  2. (ii)

    There is FU(𝔤×𝔤)hF\in U(\mathfrak{g}\times\mathfrak{g})\llbracket h\rrbracket such that F=1modhF=1\operatorname{\mathrm{mod}}h and (TT)1ΔT(x)=FΔ(x)F1(T\otimes T)^{-1}\circ\Delta\circ T(x)=F\Delta(x)F^{-1}

The standard rr-matrix is given by the following formula:

r=αR+dαEαFα,\displaystyle r=\sum_{\alpha\in R^{+}}d_{\alpha}E_{\alpha}\wedge F_{\alpha},

here 𝔤𝔤\mathfrak{g}\wedge\mathfrak{g} is embedded into 𝔤𝔤\mathfrak{g}\otimes\mathfrak{g} by xy12(xyyx)x\wedge y\longmapsto\frac{1}{2}(x\otimes y-y\otimes x). Then the anti-symmetric part of the first coefficient of FF is given by r-r.

Under this identification of Uh(𝔤)U_{h}(\mathfrak{g}) with U(𝔤)hU(\mathfrak{g})\llbracket h\rrbracket, VV is finite integrable Uh(𝔤)U_{h}(\mathfrak{g})-module if and only if it is isomorphic to LhL\otimes\mathbb{C}\llbracket h\rrbracket where LL is finite dimensional representation of 𝔤\mathfrak{g}. In particular we can identify Irrh𝔤\operatorname{\mathrm{Irr}}_{h}\mathfrak{g} with Irr𝔤\operatorname{\mathrm{Irr}}\mathfrak{g}.

Using this picture of Uh(𝔤)U_{h}(\mathfrak{g}) we can compare 𝒪h,χ(LS\G)\mathcal{O}_{h,\chi}(L_{S}\backslash G) with 𝒪(LS\G)h\mathcal{O}(L_{S}\backslash G)\llbracket h\rrbracket. At first one should note that we can decompose 𝒪(LS\G)\mathcal{O}(L_{S}\backslash G) as follows:

𝒪(LS\G)=[π]Irr𝔤Hom𝔩S(Lπ,)Lπ,\displaystyle\mathcal{O}(L_{S}\backslash G)=\bigoplus_{[\pi]\in\operatorname{\mathrm{Irr}}\mathfrak{g}}\operatorname{\mathrm{Hom}}_{\mathfrak{l}_{S}}(L_{\pi},\mathbb{C})\otimes L_{\pi},
(fv)(gw)=(fg)(vw).\displaystyle(f\otimes v)(g\otimes w)=(f\otimes g)\otimes(v\otimes w).

Here the isomorphism is given by fvf(π()v)f\otimes v\longmapsto f(\pi(\textendash)v).

Hence we have a canonical identification of 𝒪h,χ(LS\G)\mathcal{O}_{h,\chi}(L_{S}\backslash G) with 𝒪(LS\G)h\mathcal{O}(L_{S}\backslash G)\llbracket h\rrbracket. To interpret h,χ\ast_{h,\chi} as a product on 𝒪(LS\G)h\mathcal{O}(L_{S}\backslash G)\llbracket h\rrbracket, one should note that Hom𝔤(W,WW′′)HomUh(𝔤)(W,WW′′)\operatorname{\mathrm{Hom}}_{\mathfrak{g}}(W,W^{\prime}\otimes W^{\prime\prime})\cong\operatorname{\mathrm{Hom}}_{U_{h}(\mathfrak{g})}(W,W^{\prime}\otimes W^{\prime\prime}) via TFTT\longmapsto F\circ T. Therefore we have

(fv)h,χ(gw)=(fg)Ψh,χ,01FF1(vw).\displaystyle(f\otimes v)\ast_{h,\chi}(g\otimes w)=(f\otimes g)\Psi_{h,\chi,0}^{-1}F\otimes F^{-1}(v\otimes w).

Now we can obtain the semi-classical limit of 𝒪h,χ(LS\G)\mathcal{O}_{h,\chi}(L_{S}\backslash G) by a direct calculation. For θ𝔤𝔤\theta\in\mathfrak{g}\wedge\mathfrak{g}, the right (resp. left) invariant bi-vector field corresponding to θ\theta is denoted by L(θ)L_{*}(\theta) (resp. R(θ)R_{*}(\theta)).

Proposition 4.16.

The algebra 𝒪h,χ(LS\G)\mathcal{O}_{h,\chi}(L_{S}\backslash G) is an equivariant deformation quantization of 𝒪(LS\G)\mathcal{O}(L_{S}\backslash G). The associated Poisson structure is given by

{F,G}=(L(π)+R(r))(F,G)\displaystyle\{F,G\}=(L_{*}(\pi)+R_{*}(r))(F,G)

where

π=βR+R0+dβ1+χ2β1χ2βEβFβ+βR+R0+dβχ2β+1χ2β1EβFβ.\displaystyle\pi=\sum_{\beta\in R^{+}\cap R_{0}^{+}}d_{\beta}\frac{1+\chi_{-2\beta}}{1-\chi_{-2\beta}}E_{\beta}\wedge F_{\beta}+\sum_{\beta\in R^{+}\setminus R_{0}^{+}}d_{\beta}\frac{\chi_{2\beta}+1}{\chi_{2\beta}-1}E_{\beta}\wedge F_{\beta}.

5. Universal families of 22-cocycles

5.1. The toric variety associated to a root system

For general treatment of toric varieties, see [MR1234037, telen2022]. Let us recall the definition of the fan associated to a root system. For a root system RR, QQ^{\vee} denotes the dual lattice Hom(Q,)\operatorname{\mathrm{Hom}}_{\mathbb{Z}}(Q,\mathbb{Z}). We also set 𝔥\mathfrak{h}_{\mathbb{R}}^{*} (resp. 𝔥\mathfrak{h}_{\mathbb{R}}^{**}) as QQ\otimes_{\mathbb{Z}}\mathbb{R} (resp. QQ^{\vee}\otimes_{\mathbb{Z}}\mathbb{R}), which are thought as an \mathbb{R}-linear subspace of 𝔥\mathfrak{h}^{*} (resp. 𝔥\mathfrak{h}^{**}).

Then the fan associated to RR is the collection ΣR\Sigma_{R} of the following cones:

σR0+,S:={λ𝔥λ|R0+0,λ|S0=0},\displaystyle\sigma_{R_{0}^{+},S}:=\{\lambda\in\mathfrak{h}^{*}_{\mathbb{R}}\mid\lambda|_{R_{0}^{+}}\geq 0,\lambda|_{S_{0}}=0\},

where R0+R_{0}^{+} is a positive system and S0S_{0} is a set of its simple roots. It is not difficult to see that this collection is actually a fan i.e. satisfies the following conditions:

  1. (i)

    For any σΣR\sigma\in\Sigma_{R}, σσ={0}\sigma\cap-\sigma=\{0\} holds.

  2. (ii)

    For any σΣR\sigma\in\Sigma_{R} and mσ={x𝔥λ(x)0 for λσ}m\in\sigma^{\vee}=\{x\in\mathfrak{h}_{\mathbb{R}}\mid\lambda(x)\geq 0\text{ for }\lambda\in\sigma\}, the face σm:={λσλ(m)=0}\sigma_{m}:=\{\lambda\in\sigma\mid\lambda(m)=0\} is also contained in ΣR\Sigma_{R}.

  3. (iii)

    For any σ,σΣR\sigma,\sigma^{\prime}\in\Sigma_{R}, σσ\sigma\cap\sigma^{\prime} is of the form σm\sigma_{m} for some mσm\in\sigma^{\vee}.

Let kk be a commutative ring. The affine toric scheme Uσ(k)U_{\sigma}(k) associated to σΣR\sigma\in\Sigma_{R} is defined as Speck[σ2Q]\operatorname{\mathrm{Spec}}k[\sigma^{\vee}\cap 2Q], here we use 2Q2Q instead of QQ by a notational reason. Since any face of σ\sigma is of the form σm\sigma_{m} with mσ2Qm\in\sigma^{\vee}\cap 2Q, a face τ\tau of σ\sigma gives a canonical open subscheme, namely Speck[σ2Q]em=Speck[τ2Q]\operatorname{\mathrm{Spec}}k[\sigma^{\vee}\cap 2Q]_{e_{m}}=\operatorname{\mathrm{Spec}}k[\tau^{\vee}\cap 2Q] for τ=σm\tau=\sigma_{m} with mσ2Qm\in\sigma^{\vee}\cap 2Q. Then the toric scheme associated to ΣR\Sigma_{R} over kk, denoted by XR(k)X_{R}(k), can be obtained by gluing {Uσ}σΣR\{U_{\sigma}\}_{\sigma\in\Sigma_{R}} along with the gluing data {UστUσ}σ,τΣR\{U_{\sigma\cap\tau}\rightarrow U_{\sigma}\}_{\sigma,\tau\in\Sigma_{R}}.

5.2. The moduli space of equivariant Poisson brackets on LS\GL_{S}\backslash G

Let 𝒪(XLS\G)\mathcal{O}(X_{L_{S}\backslash G}) be the quotient of [φα]αRRS\mathbb{C}[\varphi_{\alpha}]_{\alpha\in R\setminus R_{S}} divided by the following relations:

  1. (i)

    φα=φα\varphi_{-\alpha}=-\varphi_{\alpha} for αRRS\alpha\in R\setminus R_{S},

  2. (ii)

    φαφβ+1=φα+β(φα+φβ)\varphi_{\alpha}\varphi_{\beta}+1=\varphi_{\alpha+\beta}(\varphi_{\alpha}+\varphi_{\beta}) when α,β,α+βRRS\alpha,\beta,\alpha+\beta\in R\setminus R_{S},

  3. (iii)

    φα=φβ\varphi_{\alpha}=\varphi_{\beta} when α,βRRS\alpha,\beta\in R\setminus R_{S} and αβRS\alpha-\beta\in R_{S}.

As pointed out in [MR1817512, MR2141466, MR2349621], this is the moduli of Poisson structures on LS\GL_{S}\backslash G equivariant with respect to the Poisson-Lie group structure on GG. Actually we can associate an equivariant Poisson bracket {,}φ\{\textendash,\textendash\}_{\varphi} to each φXLS\G\varphi\in X_{L_{S}\backslash G} as follows:

{F,G}φ=(L(πφ)+R(r))(F,G), where πφ=βR+RS+dβφβEβFβ.\displaystyle\{F,G\}_{\varphi}=(L_{*}(\pi_{\varphi})+R_{*}(r))(F,G),\text{ where }\pi_{\varphi}=\sum_{\beta\in R^{+}\setminus R_{S}^{+}}d_{\beta}\varphi_{\beta}E_{\beta}\wedge F_{\beta}.

In the following we describe a canonical immersion from XLS\G(k)X_{L_{S}\backslash G}(k) into the toric scheme XR(k)X_{R}(k) for an arbitrary commutative ring kk.

Let 𝒪(XLS\G(k))\mathcal{O}(X_{L_{S}\backslash G}(k)) be the quotient of k[ψα]αRRSk[\psi_{\alpha}]_{\alpha\in R\setminus R_{S}} divided by the following relations:

  1. (iψ)

    ψα+ψα=1\psi_{\alpha}+\psi_{-\alpha}=1 for αRRS\alpha\in R\setminus R_{S},

  2. (iiψ)

    ψαψβ=ψα+β(ψα+ψβ1)\psi_{\alpha}\psi_{\beta}=\psi_{\alpha+\beta}(\psi_{\alpha}+\psi_{\beta}-1) when α,β,α+βRRS\alpha,\beta,\alpha+\beta\in R\setminus R_{S},

  3. (iiiψ)

    ψα=ψβ\psi_{\alpha}=\psi_{\beta} when α,βRRS\alpha,\beta\in R\setminus R_{S} satisfy αβspanS\alpha-\beta\in\operatorname{\mathrm{span}}_{\mathbb{Z}}S.

Then we set XLS\G(k)=Spec𝒪(XLS\G(k))X_{L_{S}\backslash G}(k)=\operatorname{\mathrm{Spec}}\mathcal{O}(X_{L_{S}\backslash G}(k)). Note that these relations coincide with (i), (ii), (iii) under the transformation φα=2ψα1\varphi_{\alpha}=2\psi_{\alpha}-1 if 2k×2\in k^{\times}. In particular we can canonically identify XLS\G()X_{L_{S}\backslash G}(\mathbb{C}) with XLS\GX_{L_{S}\backslash G}.

Lemma 5.1.

For a positive system R0+R_{0}^{+}, we set fR0+=αR0+RSψαf_{R_{0}^{+}}=\prod_{\alpha\in R_{0}^{+}\setminus R_{S}}\psi_{\alpha} and U(R0+):={𝔭XLS\G(k)fR0+𝔭}U(R_{0}^{+}):=\{\mathfrak{p}\in X_{L_{S}\backslash G}(k)\mid f_{R_{0}^{+}}\not\in\mathfrak{p}\}. Then {U(R0+)}R0+RS+\{U(R_{0}^{+})\}_{R_{0}^{+}\supset R_{S}^{+}} is an open covering of XLS\G(k)X_{L_{S}\backslash G}(k).

Proof.

Take a prime ideal 𝔭\mathfrak{p}. What we have to show is the existence of a positive system R0+R_{0}^{+} containing RS+R_{S}^{+} such that ψα𝔭\psi_{\alpha}\not\in\mathfrak{p} for all αR0+RS+\alpha\in R_{0}^{+}\setminus R_{S}^{+}, which is equivalent to ψα0\psi_{\alpha}\neq 0 in 𝒪(XLS\G(k))/𝔭\mathcal{O}(X_{L_{S}\backslash G}(k))/\mathfrak{p}. Since this algebra is a domain, the following lemma can be applied. ∎

Lemma 5.2 (c.f. [MR2141466, Lemma 11]).

Let kk be a domain and consider ψ={ψα}αkRRS\psi=\{\psi_{\alpha}\}_{\alpha}\in k^{R\setminus R_{S}} satisfying (iψ), (iiψ), (iiiψ). Then there is a positive system R0+R_{0}^{+} contained in RS+{αRRSψα0}R_{S}^{+}\cup\{\alpha\in R\setminus R_{S}\mid\psi_{\alpha}\neq 0\}.

Proof.

At first we show that P:=RS{αRRSψα0}P:=R_{S}\cup\{\alpha\in R\setminus R_{S}\mid\psi_{\alpha}\neq 0\} contains a positive system. By [MR1890629, Chapter VI, Proposition 20], it suffices to show that the following properties hold:

  1. (a)

    For any αR\alpha\in R, either of αP\alpha\in P or αP-\alpha\in P holds.

  2. (b)

    For any α,βP\alpha,\beta\in P, α+βP\alpha+\beta\in P whenever α+βR\alpha+\beta\in R.

The condition (a) follows from (iψ). To see (b), we assume either of α\alpha or β\beta is in RSR_{S} at first. Then (iiiψ) shows (b) in this case. Next we assume that both of them are not in RSR_{S} and α+βR\alpha+\beta\in R. Moreover we also assume α+βRS\alpha+\beta\not\in R_{S} since there is nothing to prove when α+βRS\alpha+\beta\in R_{S}. In this case, using (iiψ), we have

ψα+β(ψα+ψβ1)=ψαψβ0\psi_{\alpha+\beta}(\psi_{\alpha}+\psi_{\beta}-1)=\psi_{\alpha}\psi_{\beta}\neq 0

since kk is a domain. Hence we have ψα+β0\psi_{\alpha+\beta}\neq 0, which implies α+βP\alpha+\beta\in P.

Take a positive system R1+R_{1}^{+} of RR contained in PP. Then R1+RSR_{1}^{+}\cap R_{S} is also a positive system of RSR_{S}. Hence there is ww, a product of sαs_{\alpha} with αRS\alpha\in R_{S}, such that w(R1+RS)=R+RSw(R_{1}^{+}\cap R_{S})=R^{+}\cap R_{S}. Now R0+:=w(R1+)R_{0}^{+}:=w(R_{1}^{+}) meets the required conditions since ww preserves a decompostion R=RS(RRS)R=R_{S}\sqcup(R\setminus R_{S}) and ψw(α)=ψα\psi_{w(\alpha)}=\psi_{\alpha} holds for αRRS\alpha\in R\setminus R_{S}. ∎

Lemma 5.3.

Let 𝒪(U(R0+))\mathcal{O}(U(R_{0}^{+})) be the localization of 𝒪(XLS\G(k))\mathcal{O}(X_{L_{S}\backslash G}(k)) by fR0+f_{R_{0}^{+}} and R0+\mathcal{B}_{R_{0}^{+}} be the localization of k[2Q0]/(e2α1)αRS+k[2Q_{0}^{-}]/(e_{-2\alpha}-1)_{\alpha\in R_{S}^{+}} by (1e2α)αR0+RS+(1-e_{-2\alpha})_{\alpha\in R_{0}^{+}\setminus R_{S}^{+}}. Then there is a canonical isomorphism between these algebras, as constructed in the proof below.

Proof.

We give the constructions of u:𝒪(U(R0+))R0+u\colon\mathcal{O}(U(R_{0}^{+}))\longrightarrow\mathcal{B}_{R_{0}^{+}} and v:R0+𝒪XLS\G(U(R0+))v\colon\mathcal{B}_{R_{0}^{+}}\longrightarrow\mathcal{O}_{X_{L_{S}\backslash G}}(U(R_{0}^{+})) which are mutually inverse. It suffices to give the images of the generators {ψα}αR0+RS+\{\psi_{\alpha}\}_{\alpha\in R_{0}^{+}\setminus R_{S}^{+}} and {e2α}αR0+RS+\{e_{-2\alpha}\}_{\alpha\in R_{0}^{+}\setminus R_{S}^{+}}:

(20) u(ψα)=11e2α,v(e2α)=11ψα.\displaystyle u(\psi_{\alpha})=\frac{1}{1-e_{-2\alpha}},\quad v(e_{-2\alpha})=1-\frac{1}{\psi_{\alpha}}.

It can be checked by direct calculations that these elements satisfy the required relations. ∎

Now we have a locally closed immersion XLS\G(k)XR(k)X_{L_{S}\backslash G}(k)\subset X_{R}(k).

Proposition 5.4.

There is a canonical immersion from XLS\G(k)X_{L_{S}\backslash G}(k) to XR(k)X_{R}(k), as constructed in the proof below.

Proof.

We can obtain a closed subscheme of XR(k)X_{R}(k) by gluing Speck[σ2Q]/(e2α1)ασQS\operatorname{\mathrm{Spec}}k[\sigma^{\vee}\cap 2Q]/(e_{-2\alpha}-1)_{\alpha\in\sigma^{\vee}\cap Q_{S}} and also obtain its open subscheme by gluing SpecR0+\operatorname{\mathrm{Spec}}\mathcal{B}_{R_{0}^{+}} over all R0RS+R_{0}\supset R_{S}^{+}. Hence it suffices to check the compatibility of the isomorphisms constructed in Lemma 5.3. Since SpecR0+SpecR1+=SpecR0+,R1+\operatorname{\mathrm{Spec}}\mathcal{B}_{R_{0}^{+}}\cap\operatorname{\mathrm{Spec}}\mathcal{B}_{R_{1}^{+}}=\operatorname{\mathrm{Spec}}\mathcal{B}_{R_{0}^{+},R_{1}^{+}}, where R0+,R1+\mathcal{B}_{R_{0}^{+},R_{1}^{+}} is the localization of k[2Q0+2Q1]/(1e2α)αRS+k[2Q_{0}^{-}+2Q_{1}^{-}]/(1-e_{-2\alpha})_{\alpha\in R_{S}^{+}} by (1e2α)αR0+R1+RS+(1-e_{-2\alpha})_{\alpha\in R_{0}^{+}\cup R_{1}^{+}\setminus R_{S}^{+}}, it suffices to construct an isomorphism between R0+,R1+\mathcal{B}_{R_{0}^{+},R_{1}^{+}} and 𝒪(XLS\G(k))fR0+fR1+\mathcal{O}(X_{L_{S}\backslash G}(k))_{f_{R_{0}^{+}}f_{R_{1}^{+}}} which makes the following diagram commutative:

This can be proven in the completely same way as the proof of Lemma 5.3. ∎

5.3. The sheaf of deformed quantum enveloping algebra on XR(k)X_{R}(k)

In this subsection we construct a sheaf of algebras on the toric scheme using deformed quantum enveloping algebras. For this purpose we have to compare deformed quantum envelpoing algebras based on different positive systems.

We begin with some preparations on root systems. We say that LR+L\subset R^{+} is admissible if it is of the form R+R0+R^{+}\setminus R_{0}^{+} for some positive system R0+R_{0}^{+}, or equivalently of the form R+R1R^{+}\cap R_{1}^{-} for some positive system R1+R_{1}^{+}. The admissibility can be characterized by the following conditions ([MR1169886, Theorem, p. 663]):

  1. (i)

    α+βL\alpha+\beta\in L for α,βL\alpha,\beta\in L.

  2. (ii)

    If α,βR+\alpha,\beta\in R^{+} and α+βL\alpha+\beta\in L, either of αL\alpha\in L or βL\beta\in L holds.

Note that R+LR^{+}\setminus L is also admissible when LL is admissible and that LL contains a simple root if LL\neq\emptyset.

Lemma 5.5.

Let LR+L\subset R^{+} be an admissible subset. Then there is a unique subset SΔS\subset\Delta such that spanLR+=RS+\operatorname{\mathrm{span}}_{\mathbb{Z}}L\cap R^{+}=R_{S}^{+}.

In the following proof, suppα:={εΔcε0}\operatorname{\mathrm{supp}}\alpha:=\{\varepsilon\in\Delta\mid c_{\varepsilon}\neq 0\} where α=εΔcεεQ\alpha=\sum_{\varepsilon\in\Delta}c_{\varepsilon}\varepsilon\in Q.

Proof.

By induction on |L|\lvert L\rvert. If |L|=0,1\lvert L\rvert=0,1, there is nothing to prove since L=,{ε}L=,\{\varepsilon\} for some εΔ\varepsilon\in\Delta respectively. Next we proceed to the induction step. At first we show the following claim: if AR+A\subset R^{+} and SΔS\subset\Delta satisfy spanAR+=RS+\operatorname{\mathrm{span}}_{\mathbb{Z}}A\cap R^{+}=R_{S}^{+}, spanA{ε}=RS{ε}+\operatorname{\mathrm{span}}_{\mathbb{Z}}A\cup\{\varepsilon\}=R_{S\cup\{\varepsilon\}}^{+} also holds for any εΔ\varepsilon\in\Delta. This can be checked easily if εS\varepsilon\in S. Hence we may assume εΔS\varepsilon\in\Delta\setminus S. Since SS is a union of {suppα}αA\{\operatorname{\mathrm{supp}}\alpha\}_{\alpha\in A}, it is easy to see that span(A{ε})R+RS{ε}+\operatorname{\mathrm{span}}_{\mathbb{Z}}(A\cup\{\varepsilon\})\cap R^{+}\subset R_{S\cup\{\varepsilon\}}^{+}. The other inclusion is trivial.

Now we take a simple root εL\varepsilon\in L. Then L{ε}L\setminus\{\varepsilon\} is admissible in sε(R+)s_{\varepsilon}(R^{+}). Hence there is a subset S0ΔS_{0}\subset\Delta with span(L{ε})sε(R+)=sε(R+)sε(S0)\operatorname{\mathrm{span}}_{\mathbb{Z}}(L\setminus\{\varepsilon\})\cap s_{\varepsilon}(R^{+})=s_{\varepsilon}(R^{+})_{s_{\varepsilon}(S_{0})}. Since ε-\varepsilon is simple in sε(R+)s_{\varepsilon}(R^{+}), the claim above implies that spanLsε(R+)=sε(R+)sε(S0){ε}\operatorname{\mathrm{span}}_{\mathbb{Z}}L\cap s_{\varepsilon}(R^{+})=s_{\varepsilon}(R^{+})_{s_{\varepsilon}(S_{0})\cup\{-\varepsilon\}}. Since spanL\operatorname{\mathrm{span}}_{\mathbb{Z}}L is invariant under sεs_{\varepsilon}, we can obtain the statement for LL with S=S0{ε}S=S_{0}\cup\{\varepsilon\}. ∎

Corollary 5.6.

Let R0+R_{0}^{+} and R1+R_{1}^{+} be positive systems of RR. Then there is m𝔥m\in\mathfrak{h}_{\mathbb{R}}^{**} with the following conditions:

  1. (i)

    {αRm(α)>0}R0+R1+\{\alpha\in R\mid m(\alpha)>0\}\subset R_{0}^{+}\cap R_{1}^{+},

  2. (ii)

    {αRm(α)=0}=spanR0+R1+\{\alpha\in R\mid m(\alpha)=0\}=\operatorname{\mathrm{span}}_{\mathbb{Z}}R_{0}^{+}\triangle R_{1}^{+},

  3. (iii)

    {αRm(α)<0}R0R1\{\alpha\in R\mid m(\alpha)<0\}\subset R_{0}^{-}\cap R_{1}^{-}.

Proof.

We may assume R1+=R+R_{1}^{+}=R^{+}. Take SΔS\subset\Delta by using Lemma 5.5 for L=R+R0+L=R^{+}\setminus R_{0}^{+}. Then m=εΔSεm=\sum_{\varepsilon\in\Delta\setminus S}\varepsilon^{\vee} satisfies the condition. ∎

Lemma 5.7.

Let R0+R_{0}^{+} and R1+R_{1}^{+} be positive systems of RR. There are reduced expressioins s𝐢s_{\boldsymbol{i}} and s𝐣s_{\boldsymbol{j}} of w0w_{0}, compatible with R0+R_{0}^{+} and R1+R_{1}^{+} respectively, and positive integers 1k<lN1\leq k<l\leq N with the following property:

  1. (i)

    For 1nk1\leq n\leq k, αn𝒊=αn𝒋R+(R0+R1+)\alpha^{\boldsymbol{i}}_{n}=\alpha^{\boldsymbol{j}}_{n}\in R^{+}\setminus(R_{0}^{+}\cup R_{1}^{+}).

  2. (ii)

    For k<nlk<n\leq l, αn𝒊,αn𝒋spanR0+R1+\alpha^{\boldsymbol{i}}_{n},\alpha^{\boldsymbol{j}}_{n}\in\operatorname{\mathrm{span}}_{\mathbb{Z}}R_{0}^{+}\triangle R_{1}^{+}.

  3. (iii)

    For l<nNl<n\leq N, αn𝒊=αn𝒋R+R0+R1+\alpha^{\boldsymbol{i}}_{n}=\alpha^{\boldsymbol{j}}_{n}\in R^{+}\cap R_{0}^{+}\cap R_{1}^{+}.

Proof.

By using m𝔥m\in\mathfrak{h}_{\mathbb{R}}^{**} in Corollary 5.6, we can see the admissiblilty of the following sets.

L±:=R+R0±R1±V={αR+±m(α)>0}.\displaystyle L^{\pm}:=R^{+}\cap R_{0}^{\pm}\cap R_{1}^{\pm}\setminus V=\{\alpha\in R^{+}\mid\pm m(\alpha)>0\}.

Take w±Ww^{\pm}\in W such that L±=R+w±(R+)L^{\pm}=R^{+}\setminus w^{\pm}(R^{+}) and also take w,wWw,w^{\prime}\in W such that R0+=w(R+)R_{0}^{+}=w(R^{+}) and R1+=w(R+)R_{1}^{+}=w^{\prime}(R^{+}). Consider a reduced expression of ww^{-}. This expression can be extended to reduced expressions ww and ww^{\prime}. Similarly a reduced expression of w+w^{+} can be extended to reduced expressions of ww0ww_{0} and ww0w^{\prime}w_{0}. Then we obtain enumerations {αni}\{\alpha^{i}_{n}\} on R+Ri+R^{+}\setminus R_{i}^{+} and {βmi}\{\beta^{i}_{m}\} on R+Ri+R^{+}\cap R_{i}^{+} from these reduced expressions. Then, as a consequence of [MR1169886, Theorem, p.662], we can find reduced expressions s𝒊s_{\boldsymbol{i}} and s𝒋s_{\boldsymbol{j}} which satisfy the following conditions respectively:

αn𝒊={αn0(1n(w)),βNn+10((w)<nN),αn𝒋={αn1(1n(w)),βNn+11((w)<nN).\displaystyle\alpha^{\boldsymbol{i}}_{n}=\begin{cases}\alpha^{0}_{n}&(1\leq n\leq\ell(w)),\\ \beta^{0}_{N-n+1}&(\ell(w)<n\leq N),\end{cases}\quad\alpha^{\boldsymbol{j}}_{n}=\begin{cases}\alpha^{1}_{n}&(1\leq n\leq\ell(w^{\prime})),\\ \beta^{1}_{N-n+1}&(\ell(w^{\prime})<n\leq N).\end{cases}

By construction these reduced expressions satisfy the required conditions. ∎

Remark 5.8.

Let ww^{-} be as defined in the proof above. Then (w)1(R+spanR0+R1+)(w^{-})^{-1}(R^{+}\cap\operatorname{\mathrm{span}}_{\mathbb{Z}}R_{0}^{+}\triangle R_{1}^{+}) is of the form RS+R_{S}^{+} for some SΔS\subset\Delta since the image is {αR+m(w(α))=0}\{\alpha\in R^{+}\mid m(w^{-}(\alpha))=0\} and its complement in R+R^{+} is {αR+m(w(α))>0}\{\alpha\in R^{+}\mid m(w^{-}(\alpha))>0\}. This fact implies that the linear span of {F𝒊(Λ)}k<Λl\{F_{\boldsymbol{i}}^{(\Lambda)}\}_{k<\Lambda\leq l} coincides with the linear span of {F𝒋(Λ)}k<Λl\{F_{\boldsymbol{j}}^{(\Lambda)}\}_{k<\Lambda\leq l}.

Proposition 5.9.

Let kk be an 𝒜\mathcal{A}-algebra and χ:2Q0+2Q1k\chi\colon 2Q_{0}^{-}+2Q_{1}^{-}\longrightarrow k be a character. Then there is a canonical isomorphism between Uq,χ0k(𝔤)U_{q,\chi_{0}}^{k}(\mathfrak{g}{}^{\sim}) and Uq,χ1k(𝔤)U_{q,\chi_{1}}^{k}(\mathfrak{g}{}^{\sim}) as constructed in the proof, where χi\chi_{i} is the restriction of χ\chi on 2Qi2Q_{i}^{-}.

Proof.

Set Ui=Uq,ei𝒜[2Q0+2Q1+](𝔤)U_{i}=U_{q,e_{i}}^{\mathcal{A}[2Q_{0}^{-}+2Q_{1}^{+}]}(\mathfrak{g}{}^{\sim}). It suffices to show U0=U1U_{0}=U_{1} in Uq𝒜[P](𝔤)U_{q}^{\mathcal{A}[P]}(\mathfrak{g}), where e:2Q0+2Q1𝒜[2Q0+2Q1]e\colon 2Q_{0}-+2Q_{1}\longrightarrow\mathcal{A}[2Q_{0}^{-}+2Q_{1}^{-}] is the canonical character.

Take a reduced expressions s𝒊s_{\boldsymbol{i}} and s𝒋s_{\boldsymbol{j}} of w0w_{0} as Lemma 5.7. Since e2αe_{2\alpha} with αR+spanR0+R1+\alpha\in R^{+}\cap\operatorname{\mathrm{span}}_{\mathbb{Z}}R_{0}^{+}\triangle R_{1}^{+} is invertible in 𝒜[2Q0+2Q1]\mathcal{A}[2Q_{0}^{-}+2Q_{1}^{-}], U0U_{0} is generated by the following elements:

{(qαqα1)E´𝒊,α}αR+,{eλK´λ}λP,{e2αn𝒊F´𝒊,n(r)(1nk),F´𝒊,n(r)(k<nl),F´𝒊,n(r)(l<nN).\displaystyle\{(q_{\alpha}-q_{\alpha}^{-1})\acute{E}_{\boldsymbol{i},\alpha}\}_{\alpha\in R^{+}},\quad\{e_{-\lambda}\acute{K}_{\lambda}\}_{\lambda\in P},\quad\begin{cases}e_{2\alpha^{\boldsymbol{i}}_{n}}\acute{F}_{\boldsymbol{i},n}^{(r)}&(1\leq n\leq k),\\ \acute{F}_{\boldsymbol{i},n}^{(r)}&(k<n\leq l),\\ \acute{F}_{\boldsymbol{i},n}^{(r)}&(l<n\leq N).\end{cases}

The same thing can be said for U1U_{1}. By the property of s𝒊s_{\boldsymbol{i}} and s𝒋s_{\boldsymbol{j}}, only different generators are F´𝒊,n(r)\acute{F}_{\boldsymbol{i},n}^{(r)} and F´𝒋,n(r)\acute{F}_{\boldsymbol{j},n}^{(r)} with k<nlk<n\leq l. But the algebras generated by these elements are independent of 𝒊\boldsymbol{i} and 𝒋\boldsymbol{j} as stated in the remark above. Hence we have U0=U1U_{0}=U_{1}. ∎

Remark 5.10.

Take another positive system R2+R_{2}^{+} and consider a character χ\chi from 2Q0+2Q1+2Q22Q_{0}^{-}+2Q_{1}^{-}+2Q_{2}^{-}. In this case we have three isomorphisms Uq,χ0k(𝔤)Uq,χ1k(𝔤)U_{q,\chi_{0}}^{k}(\mathfrak{g}{}^{\sim})\cong U_{q,\chi_{1}}^{k}(\mathfrak{g}{}^{\sim}), Uq,χ1k(𝔤)Uq,χ2k(𝔤)U_{q,\chi_{1}}^{k}(\mathfrak{g}{}^{\sim})\cong U_{q,\chi_{2}}^{k}(\mathfrak{g}{}^{\sim}) and Uq,χ0k(𝔤)Uq,χ2k(𝔤)U_{q,\chi_{0}}^{k}(\mathfrak{g}{}^{\sim})\cong U_{q,\chi_{2}}^{k}(\mathfrak{g}{}^{\sim}). By construction of the isomorphisms, we can see the commutativity of the following diagram:

Now we can define a sheaf of deformed quantum enveloping algebras on XR(k)X_{R}(k) and XLS\G(k)X_{L_{S}\backslash G}(k).

Definition 5.11.

Let kk be an 𝒜\mathcal{A}-algebra. We define 𝒰q,XRk(𝔤)\mathscr{U}_{q,X_{R}}^{k}(\mathfrak{g}{}^{\sim}), a sheaf of algebras on XR(k)X_{R}(k), by gluing {Uq,ek[2Q0](𝔤)}R0+\{U_{q,e}^{k[2Q_{0}^{-}]}(\mathfrak{g}{}^{\sim})\}_{R_{0}^{+}} by the isomorpshims in Proposition 5.9. We also define 𝒰q,XLS\Gk(𝔤)\mathscr{U}_{q,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim}) as the inverse image of 𝒰q,XRk(𝔤)\mathscr{U}_{q,X_{R}}^{k}(\mathfrak{g}{}^{\sim}) by the immersion XLS\G(k)XR(k)X_{L_{S}\backslash G}(k)\longrightarrow X_{R}(k) defined in Proposition 5.4.

Remark 5.12.

By construction of the isomorphisms in Proposition 5.9, we can obtain an injection from fUqk(𝔩S)~f^{*}\widetilde{U_{q}^{k}(\mathfrak{l}_{S}{}^{\sim})} to 𝒰q,XLS\Gk(𝔤)\mathscr{U}_{q,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim}) where f:XLS\G(k)Speckf\colon X_{L_{S}\backslash G}(k)\longrightarrow\operatorname{\mathrm{Spec}}k is the canonical morphism.

Remark 5.13.

Since XLS\G(k)X_{L_{S}\backslash G}(k) is an affine kk-scheme and 𝒰q,XLS\Gk(𝔤)\mathscr{U}_{q,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim}) is quasi-coherent, we can reconstruct the sheaf from its global sections Uq,XLS\Gk(𝔤):=𝒰q,XLS\Gk(𝔤)(XLS\G(k))U_{q,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim}):=\mathscr{U}_{q,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim})(X_{L_{S}\backslash G}(k)).

By construction of 𝒰q,XRk(𝔤)\mathscr{U}_{q,X_{R}}^{k}(\mathfrak{g}{}^{\sim}), we have the local freeness of Uq,XLS\Gk(𝔤)U_{q,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim}). For SΔS\subset\Delta with |ΔS|=1\lvert\Delta\setminus S\rvert=1, we can say a stronger statement: there is a “PBW basis” for Uq,XLS\Gk(𝔤)U_{q,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim}), constructed as follows.

At first one should note that XLS\GX_{L_{S}\backslash G} is covered by U(R+)U(R^{+}) and U(R0+)U(R_{0}^{+}) where R0+=RS+(RRS)R_{0}^{+}=R_{S}^{+}\cup(R^{-}\setminus R_{S}^{-}). Take a reduced expression s𝒊s_{\boldsymbol{i}} of w0w_{0} compatible with R0+R_{0}^{+}. By definition of deformed quantum enveloping algebra, we have E^𝒊\hat{E}_{\boldsymbol{i}} and K^λ\hat{K}_{\lambda} as elements of Uq,XLS\Gk(𝔤)U_{q,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim}). For F^𝒊,α\hat{F}_{\boldsymbol{i},\alpha}, we define a global section by gluing ψαF^𝒊,α\psi_{\alpha}\hat{F}_{\boldsymbol{i},\alpha} on U(R+)U(R^{+}) and ψαF^𝒊,α-\psi_{\alpha}\hat{F}_{\boldsymbol{i},\alpha} on U(R0+)U(R_{0}^{+}). Then these global sections behave as “quantum root vectors” in Uq,XLS\Gk(𝔤)U_{q,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim}).

For general 𝔩S𝔤\mathfrak{l}_{S}\subset\mathfrak{g}, it seems to be still possible to show the freeness. For example we can define Fα+βUq,XH\SL3k(𝔰𝔩3)F_{\alpha+\beta}\in U_{q,X_{H\backslash SL_{3}}}^{k}(\mathfrak{sl}_{3}^{\sim}) as follows:

Fα+β=ψα+βFβFαψα+β(qψα+q1ψα)FαFβ.\displaystyle F_{\alpha+\beta}=\psi_{\alpha+\beta}F_{\beta}F_{\alpha}-\psi_{\alpha+\beta}(q\psi_{\alpha}+q^{-1}\psi_{-\alpha})F_{\alpha}F_{\beta}.

But we do not have any idea to obtain good quantum root vectors in general case. Even in the case of 𝔰𝔩3\mathfrak{sl}_{3}, we do not know which element should be taken as Fα+βF_{\alpha+\beta}.

5.4. Explicit construction of all quantizations

Let kk be a \mathbb{Q}-algebra. In this case we can consider 𝒰h,XRk(𝔤)\mathscr{U}_{h,X_{R}}^{k}(\mathfrak{g}{}^{\sim}) and 𝒰h,XLS\Gk(𝔤)\mathscr{U}_{h,X_{L_{S}\backslash G}}^{k}(\mathfrak{g}{}^{\sim}) on XR(k)X_{R}(k) and XLS\G(k)X_{L_{S}\backslash G}(k) respectively. Also note that we can use φα:=2ψα1\varphi_{\alpha}:=2\psi_{\alpha}-1 as generators of 𝒪(XLS\G(k))\mathcal{O}(X_{L_{S}\backslash G}(k)) since 2k×2\in k^{\times}.

Let us recall the construction of the 22-cocycle Ψh,χk\Psi_{h,\chi}^{k}. This can be characterized as a unique element inducing the isomorphism (13), which only depends on the QQ^{-}-grading of the (R0+,χ)(R_{0}^{+},\chi)-twisted parabolically induced modules. Since the identification in Proposition 5.9 preserves the triangular decomposition and the QQ-grading, we can see that Ψh,χk\Psi_{h,\chi}^{k} does not depend on the choice of R0+R_{0}^{+}. This leads us to the following conclusion. Note that there is a canonical homomorphism from 𝒪(XLS\G(k))\mathcal{O}(X_{L_{S}\backslash G}(k)) to kk when we take a character χ:2Q0k\chi\colon 2Q_{0}^{-}\longrightarrow k with RS+R0+,χ2α=1R_{S}^{+}\subset R_{0}^{+},\chi_{-2\alpha}=1 for αRS+\alpha\in R_{S}^{+} and χ2α1k×\chi_{-2\alpha}-1\in k^{\times} for αR0+RS+\alpha\in R_{0}^{+}\setminus R_{S}^{+}.

Theorem 5.14.

There is a unique 22-cocycle Ψh,φkUh𝒪(XLS\G(k))(𝔤×𝔤×𝔩S)𝔩S\Psi_{h,\varphi}^{k}\in U_{h}^{\mathcal{O}(X_{L_{S}\backslash G}(k))}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S})^{\mathfrak{l}_{S}} which coincides with Ψh,χk\Psi_{h,\chi}^{k} under the homomorphism above for all (k,χ)(k,\chi). Moreover it has the following expansion:

(21) Ψh,φ=1hαR+RS+dα(φα+1)EαFα1+.\displaystyle\Psi_{h,\varphi}=1-h\sum_{\alpha\in R^{+}\setminus R_{S}^{+}}d_{\alpha}(\varphi_{\alpha}+1)E_{\alpha}\otimes F_{\alpha}\otimes 1+\cdots.

As an immediate corollary, we can prove [MR1817512, Proposition 5.3] in a constructible way.

Corollary 5.15.

There is an algebraic family {𝒪h,φ(LS\G)}φXLS\G\{\mathcal{O}_{h,\varphi}(L_{S}\backslash G)\}_{\varphi\in X_{L_{S}\backslash G}} of equivariant deformation quantization of 𝒪(LS\G)\mathcal{O}(L_{S}\backslash G) with the associated Poisson bracket {,}φ\{\textendash,\textendash\}_{\varphi}.

By using [MR1817512, Proposition 5.6] we also obtain the classification theorem. A formal path on XLS\GX_{L_{S}\backslash G} is (φα(h))αRRShRRS(\varphi_{\alpha}(h))_{\alpha\in R\setminus R_{S}}\in\mathbb{C}\llbracket h\rrbracket^{R\setminus R_{S}} satisfying the relations (i), (ii), (iii). Then we can consider a homomorphism 𝒪(XLS\G)h\mathcal{O}(X_{L_{S}\backslash G})\longrightarrow\mathbb{C}\llbracket h\rrbracket by setting φαφα(h)\varphi_{\alpha}\longmapsto\varphi_{\alpha}(h) and obtain a deformation quantization 𝒪h,φ(h)(LS\G)\mathcal{O}_{h,\varphi(h)}(L_{S}\backslash G)

Corollary 5.16.

Let 𝒪h(LS\G)\mathcal{O}_{h}(L_{S}\backslash G) be an equivariant deformation quantization of 𝒪(LS\G)\mathcal{O}(L_{S}\backslash G). Then there is a unique formal path φ(h)\varphi(h) on XLS\GX_{L_{S}\backslash G} such that 𝒪h(LS\G)\mathcal{O}_{h}(L_{S}\backslash G) is equivalent to 𝒪h,φ(h)(LS\G)\mathcal{O}_{h,\varphi(h)}(L_{S}\backslash G) as an equivariant deformation quantization.

6. Comparison theorem

Let SS0ΔS\subset S_{0}\subset\Delta be subsets. The Levi subalgebra of 𝔤\mathfrak{g} (resp. Levi subgroup of GG) associated to S0S_{0} is denoted by 𝔩0\mathfrak{l}_{0} (resp. L0L_{0}). Then, by the results in the previous section, we have the 22-cocycle Ψ0Uh𝒪(XLS\L0)(𝔩0×𝔩0×𝔩S)𝔩S\Psi_{0}\in U_{h}^{\mathcal{O}(X_{L_{S}\backslash L_{0}})}(\mathfrak{l}_{0}\times\mathfrak{l}_{0}\times\mathfrak{l}_{S})^{\mathfrak{l}_{S}}. Then we can thought Ψ0\Psi_{0} as a 22-cocycle in Uh𝒪(XLS\L0)(𝔤×𝔤×𝔩S)𝔩SU_{h}^{\mathcal{O}(X_{L_{S}\backslash L_{0}})}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S})^{\mathfrak{l}_{S}} and obtain a deformation quantization by the evaluation at an element of XLS\L0X_{L_{S}\backslash L_{0}}. In light of Corollary 5.16, it is natural to compare this quantization and quantizations coming from ΨUh𝒪(XLS\G)(𝔤×𝔤×𝔩S)𝔩S\Psi\in U_{h}^{\mathcal{O}(X_{L_{S}\backslash G})}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S})^{\mathfrak{l}_{S}}. For such a purpose we consider a homomorphism π:𝒪(XLS\G(k))𝒪(XLS\L0(k))\pi\colon\mathcal{O}(X_{L_{S}\backslash G}(k))\longrightarrow\mathcal{O}(X_{L_{S}\backslash L_{0}}(k)) defined by

π(φα)={φα(αRS0RS),1(αR+RS0+),1(αRRS0).\displaystyle\pi(\varphi_{\alpha})=\begin{cases}\varphi_{\alpha}&(\alpha\in R_{S_{0}}\setminus R_{S}),\\ -1&(\alpha\in R^{+}\setminus R_{S_{0}}^{+}),\\ 1&(\alpha\in R^{-}\setminus R_{S_{0}}^{-}).\end{cases}
Proposition 6.1.

Let ΨUh𝒪(XLS\G(k))(𝔤×𝔤×𝔩S)𝔩\Psi\in U_{h}^{\mathcal{O}(X_{L_{S}\backslash G}(k))}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S})^{\mathfrak{l}} be the 22-cocycle constructed in the previous section. Then πΨ=Ψ0\pi_{*}\Psi=\Psi_{0}.

We prove this fact as a corollary of the fully faithfulness of a functor introduced in the following discussion, which holds in the integral setting.

Let R0+R_{0}^{+} be a positive system of RR containing RS+(RRS0)R_{S}^{+}\cup(R^{-}\setminus R_{S_{0}}^{-}). Take a reduced expression s𝒊s_{\boldsymbol{i}} compatible with R0+R_{0}^{+}. Then we can introduce the notion of generalized S0S_{0}-maximal vector in the similar manner to Definition 4.2.

We say that a Uqk(𝔤)U_{q}^{k}(\mathfrak{g})-module VV is locally strongly Uqk(𝔫+)U_{q}^{k}(\mathfrak{n}^{+})-finite if, for any vVv\in V, Uqk(𝔫+)μv=0U_{q}^{k}(\mathfrak{n}^{+})_{\mu}v=0 for all μQ+\mu\in Q^{+} with finitely many exceptions.

Theorem 6.2.

Let (k,χ)(k,\chi) be an 𝒜[2Q0]\mathcal{A}[2Q_{0}^{-}]-algebra with χ2α=0\chi_{-2\alpha}=0 for αRRS0\alpha\in R^{-}\setminus R_{S_{0}}^{-}and χ2α=1\chi_{-2\alpha}=1 for αRS+\alpha\in R_{S}^{+}.

  1. (i)

    For any VUq,χk-Mod(𝔩0)V\in U_{q,\chi}^{k}\text{-}\mathrm{Mod}(\mathfrak{l}_{0}{}^{\sim}) and WUqk(𝔤)-ModW\in U_{q}^{k}(\mathfrak{g})\text{-}\mathrm{Mod}, each mWVm\in W\otimes V has a unique generalized S0S_{0}-maximal vector (mΛ)Λ(m_{\Lambda})_{\Lambda} with m0=mm_{0}=m.

  2. (ii)

    The functor ind𝔩0,q𝔤,χ:Uq,χk(𝔩0)-ModUq,χk(𝔤)-Mod,VUq,χk(𝔤)Uq,χk(𝔭0)V\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}\colon U_{q,\chi}^{k}(\mathfrak{l}_{0}{}^{\sim})\text{-}\mathrm{Mod}\longrightarrow U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})\text{-}\mathrm{Mod},V\longmapsto U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})\otimes_{U_{q,\chi}^{k}(\mathfrak{p}_{0}{}^{\sim})}V is fully faithful.

  3. (iii)

    If a Uq,χk(𝔤)U_{q,\chi}^{k}(\mathfrak{g}{}^{\sim})-module WW is locally strongly Uqk(𝔫+)U_{q}^{k}(\mathfrak{n}^{+})-finite, there is a unique homomorphism ind𝔩0,q𝔤,χ(WV)Wind𝔩0,q𝔤,χV\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}(W\otimes V)\longrightarrow W\otimes\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}V which sends mm to a S0S_{0}-maximal vector of the form m(W)(1m(V))+m_{(W)}\otimes(1\otimes m_{(V)})+\cdots. Moreover this is an isomorphism.

  4. (iv)

    Let W,WW,W^{\prime} be locally strongly Uqk(𝔟+)U_{q}^{k}(\mathfrak{b}^{+})-finite Uqk(𝔤)U_{q}^{k}(\mathfrak{g})-module and VV be a Uq,χk(𝔩0)U_{q,\chi}^{k}(\mathfrak{l}_{0}{}^{\sim})-module. Then the following diagram commutes.

Proof.

(i) The proof is very similar to that of Proposition 4.4, hence that of Lemma 4.5. The only different point is to show the upper diagonality and the invertibility of diagonal entries in the usual sense.

Fix a multi-index Γ\Gamma and consider its decomposition Γ=Γ+δi\Gamma=\Gamma^{\prime}+\delta_{i} such that γ1=γ2==γi1=0\gamma_{1}=\gamma_{2}=\cdot=\gamma_{i-1}=0. Then we look at the coefficient of F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma^{\prime})} in (9) with β=αi𝒊\beta=\alpha^{\boldsymbol{i}}_{i}:

ΛE^𝒊,β,(1)mΛ,(W)(E^𝒊,β,(2)F^𝒊(Λ))(𝔲S)πS((E^𝒊,β,(2)F^𝒊(Λ))(𝔭S))mΛ,(V).\displaystyle\sum_{\Lambda}\hat{E}_{\boldsymbol{i},\beta,(1)}m_{\Lambda,(W)}\otimes(\hat{E}_{\boldsymbol{i},\beta,(2)}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{u}_{S}^{-})}\otimes\pi_{S}((\hat{E}_{\boldsymbol{i},\beta,(2)}\hat{F}_{\boldsymbol{i}}^{(\Lambda)})_{(\mathfrak{p}_{S})})m_{\Lambda,(V)}.

By (5), we can replace Δ(E^𝒊,β)\Delta(\hat{E}_{\boldsymbol{i},\beta}) by KΛrα𝒊1E^𝒊ΛrK_{\Lambda^{r}\cdot\alpha^{\boldsymbol{i}}}^{-1}\otimes\hat{E}_{\boldsymbol{i}}^{\Lambda^{r}} with ΛriΛl\Lambda^{r}\leq i\leq\Lambda^{l}. Moreover it suffices to consider the terms with Π(Λlα𝒊)=0\Pi(\Lambda^{l}\cdot\alpha^{\boldsymbol{i}})=0 and Λrδi\Lambda^{r}\neq\delta_{i} to show the upper diagonality. Note that Λr<i\Lambda^{r}<i holds in this case.

Consider Λ\Lambda such that ΛBl,Π(Γα𝒊)\Lambda\in B_{l,\Pi(\Gamma\cdot\alpha^{\boldsymbol{i}})} and λ1=λ2==λi1=0\lambda_{1}=\lambda_{2}=\cdots=\lambda_{i-1}=0. Then we have the following expansion by Proposition 3.4 (i):

E~𝒊ΛrF´𝒊(Λ)=Λ+<iΛΛrα𝒊Λ+α𝒊=Λα𝒊Λα𝒊Q+CΛ±F´𝒊(Λ)E~𝒊Λ+.\displaystyle\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{r}}\acute{F}_{\boldsymbol{i}}^{(\Lambda)}=\sum_{\begin{subarray}{c}\Lambda^{+}<i\leq\Lambda^{-}\\ \Lambda^{r}\cdot\alpha^{\boldsymbol{i}}-\Lambda^{+}\cdot\alpha^{\boldsymbol{i}}=\Lambda\cdot\alpha^{\boldsymbol{i}}-\Lambda^{-}\cdot\alpha^{\boldsymbol{i}}\in Q^{+}\end{subarray}}C_{\Lambda^{\pm}}\acute{F}_{\boldsymbol{i}}^{(\Lambda^{-})}\widetilde{E}_{\boldsymbol{i}}^{\Lambda^{+}}.

Hence we also have the following:

E^𝒊ΛrF^𝒊(Λ)=Λ+<iΛΛrα𝒊Λ+α𝒊=Λα𝒊Λα𝒊Q+cΛ±F^𝒊(Λ)E^𝒊Λ+.\displaystyle\hat{E}_{\boldsymbol{i}}^{\Lambda^{r}}\hat{F}_{\boldsymbol{i}}^{(\Lambda)}=\sum_{\begin{subarray}{c}\Lambda^{+}<i\leq\Lambda^{-}\\ \Lambda^{r}\cdot\alpha^{\boldsymbol{i}}-\Lambda^{+}\cdot\alpha^{\boldsymbol{i}}=\Lambda\cdot\alpha^{\boldsymbol{i}}-\Lambda^{-}\cdot\alpha^{\boldsymbol{i}}\in Q^{+}\end{subarray}}c_{\Lambda^{\pm}}\hat{F}_{\boldsymbol{i}}^{(\Lambda^{-})}\hat{E}_{\boldsymbol{i}}^{\Lambda^{+}}.

If Λ+0\Lambda^{+}\neq 0, such a term is removed by πS0\pi_{S_{0}} since Λ+<i\Lambda^{+}<i. It means that we only have to look at the term with Λ+=0\Lambda^{+}=0 and Λ=Γ\Lambda^{-}=\Gamma^{\prime} to determine the coefficient of F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma^{\prime})}. On the other hand, the coefficient of such a term can be calculated as follows:

c0,Γ=C0,Γχ2Λα𝒊2Λα𝒊=C0,Γχ2Λrα𝒊=0.\displaystyle c_{0,\Gamma^{\prime}}=C_{0,\Gamma^{\prime}}\chi_{2\Lambda\cdot\alpha^{\boldsymbol{i}}-2\Lambda^{-}\cdot\alpha^{\boldsymbol{i}}}=C_{0,\Gamma^{\prime}}\chi_{2\Lambda^{r}\cdot\alpha^{\boldsymbol{i}}}=0.

Hence our Λ\Lambda does not affect the coefficient of F^𝒊(Γ)\hat{F}_{\boldsymbol{i}}^{(\Gamma^{\prime})}. This shows the upper diagonality of Al,νA_{l,\nu}.

To show the invertibility, consider the case of Λ=Γ\Lambda=\Gamma. By the consideration above we may assume Λr=δi\Lambda^{r}=\delta_{i}. Then a completely same argument in the proof of Lemma 4.5 works to prove the invertibility of the diagonal entry. Then the upper triangularity implies the invertibility of Al,νA_{l,\nu}.

(ii) This is a corollary of (i).

(iii) To see the former half of the statement, it suffices to show that mΛ=0m_{\Lambda}=0 except finitely many Λ\Lambda for any generalized S0S_{0}-maximal vector (mΛ)Λ(m_{\Lambda})_{\Lambda} since it implies the existence and uniqueness of required S0S_{0}-maximal vectors and also the existence and uniqueness of the required homomorphism.

By setting W=Uqk(𝔤)W=U_{q}^{k}(\mathfrak{g}) and V=Uq,χk(𝔩0)V=U_{q,\chi}^{k}(\mathfrak{l}_{0}{}^{\sim}), we can obtain the S0S_{0}-maximizer (UΛ)Λ(U_{\Lambda})_{\Lambda}. Moreover the consideration above implies that UΛΠ(ν)=Π(Λα𝒊)Uqk(𝔟+)νUq,χk(𝔩0)U_{\Lambda}\in\sum_{\Pi(\nu)=\Pi(\Lambda\cdot\alpha^{\boldsymbol{i}})}U_{q}^{k}(\mathfrak{b}^{+})_{\nu}\otimes U_{q,\chi}^{k}(\mathfrak{l}_{0}{}^{\sim}). Hence our assumption on WW implies that mΛ=UΛm0=0m_{\Lambda}=U_{\Lambda}m_{0}=0 except for finitely many Λ\Lambda.

Next we show the bijectivity. Note that we have Δ(F^𝒊Λ)=KΛα𝒊1F^𝒊Λ\Delta(\hat{F}_{\boldsymbol{i}}^{\Lambda})=K_{\Lambda\cdot\alpha^{\boldsymbol{i}}}^{-1}\otimes\hat{F}_{\boldsymbol{i}}^{\Lambda}. Hence the image of w(F^𝒊(Γ)v)w\otimes(\hat{F}_{\boldsymbol{i}}^{(\Gamma)}\otimes v) is

ΛKΓα𝒊1UΛ,(Uqk(𝔟+))wF^𝒊ΓF^𝒊ΛUΛ,(Uq,χk(𝔩0))v.\displaystyle\sum_{\Lambda}K_{\Gamma\cdot\alpha^{\boldsymbol{i}}}^{-1}U_{\Lambda,(U_{q}^{k}(\mathfrak{b}^{+}))}w\otimes\hat{F}_{\boldsymbol{i}}^{\Gamma}\hat{F}_{\boldsymbol{i}}^{\Lambda}\otimes U_{\Lambda,(U_{q,\chi}^{k}(\mathfrak{l}_{0}{}^{\sim}))}v.

Now we regard the homomorphism as an endomorphism TT on Uq,χk(𝔲S0)WVU_{q,\chi}^{k}(\mathfrak{u}_{S_{0}}^{-}{}^{\sim})\otimes W\otimes V. Then Uq,χk(𝔲S0)Uqk(𝔟+)wVU_{q,\chi}^{k}(\mathfrak{u}_{S_{0}}^{-}{}^{\sim})\otimes U_{q}^{k}(\mathfrak{b}^{+})w\otimes V is stable under this map. Moreover our assumption implies the existence of a positive integer nn such that (Tid)n=0(T-\operatorname{\mathrm{id}})^{n}=0 on this module. This fact shows the bijectivity.

(iv) By (i) we have an automorphism TT on WWVW\otimes W^{\prime}\otimes V such that ind𝔩0,q𝔤,χT\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}T coincides with the following composition:

ind𝔩0,q𝔤,χ(WWV)\displaystyle\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}(W\otimes W^{\prime}\otimes V) Wind𝔩0,q𝔤,χ(WV)\displaystyle\longrightarrow W\otimes\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}(W^{\prime}\otimes V)
WWind𝔩0,q𝔤,χVind𝔩0,q𝔤,χ(WWV).\displaystyle\longrightarrow W\otimes W^{\prime}\otimes\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}V\longrightarrow\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}(W\otimes W^{\prime}\otimes V).

Then the image of 1(wwv)1\otimes(w\otimes w^{\prime}\otimes v) at WWind𝔩0,q𝔤,χVW\otimes W^{\prime}\otimes\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}V is of the form T(wwv)(WW)(1T(wwv)(V))+T(w\otimes w^{\prime}\otimes v)_{(W\otimes W^{\prime})}\otimes(1\otimes T(w\otimes w^{\prime}\otimes v)_{(V)})+\cdots. On the other hand, by using the S0S_{0}-maximizer, we can calculate the image and see that it is of the form ww(1v)+w\otimes w^{\prime}\otimes(1\otimes v)+\cdots. Hence we have T=idT=\operatorname{\mathrm{id}} and the diagram commutes. ∎

Proof of Proposition 6.1.

By the construction of the cocycles, it suffices to show the commutativity of the following diagram:

Since the both images of 1(wv)1\otimes(w\otimes v) at the upper right corner is of the form w(1v)+w\otimes(1\otimes v)+\cdots, it suffices to show that these are S0S_{0}-maximal vectors. The image under the upper isomorphism is trivially SS-maximal. To see the maximality of another image, we show more general statement: Let WW be a locally strongly Uqk(𝔫+)U_{q}^{k}(\mathfrak{n}^{+})-finite module and VV be a Uq,χk(𝔩0)U_{q,\chi}^{k}(\mathfrak{l}_{0}{}^{\sim})-module. If mWVm\in W\otimes V is SS-maximal, the S0S_{0}-maximal vector m~=m(W)(1m(V))+Wind𝔩0,q𝔤,χV\widetilde{m}=m_{(W)}\otimes(1\otimes m_{(V)})+\cdots\in W\otimes\mathrm{ind}_{\mathfrak{l}_{0},q}^{\mathfrak{g},\chi}V is also SS-maximal.

Take E^𝒊,β\hat{E}_{\boldsymbol{i},\beta} with βRS0+RS+\beta\in R_{S_{0}}^{+}\setminus R_{S}^{+}. Then E^𝒊,βm~\hat{E}_{\boldsymbol{i},\beta}\widetilde{m} is still of the form (E^𝒊,βm)(W)(1(E^𝒊,βm)(V))+(\hat{E}_{\boldsymbol{i},\beta}m)_{(W)}\otimes(1\otimes(\hat{E}_{\boldsymbol{i},\beta}m)_{(V)})+\cdots. By a discussion similar to Lemma 3.17, E^𝒊,βm~\hat{E}_{\boldsymbol{i},\beta}\widetilde{m} is again S0S_{0}-maximal. This implies E^𝒊,βm~=0\hat{E}_{\boldsymbol{i},\beta}\widetilde{m}=0 as a consequence of the uniqueness for a maximal vector and E^𝒊,βm=0\hat{E}_{\boldsymbol{i},\beta}m=0. ∎

At last we interpret this comparison theorem as a statement on deformation quantizations. Let us recall the Uhk(𝔤)U_{h}^{k}(\mathfrak{g})-bimodule structure on 𝒪hk(G)\mathcal{O}_{h}^{k}(G), induced from the embedding 𝒪hk(G)Uhk(𝔤)\mathcal{O}_{h}^{k}(G)\subset U_{h}^{k}(\mathfrak{g})^{*}. Then, for a Uhk(𝔩0)U_{h}^{k}(\mathfrak{l}_{0})-module algebra BB, we define the induced Uhk(𝔤)U_{h}^{k}(\mathfrak{g})-module algebra indL0,hGhB\mathrm{ind}_{L_{0,h}}^{G_{h}}B as follows:

indL0,hGhB={a𝒪hk(G)B(r(a)id)(x)=(idl(x))(a) for all xUhk(𝔩0)},\displaystyle\mathrm{ind}_{L_{0,h}}^{G_{h}}B=\{a\in\mathcal{O}_{h}^{k}(G)\otimes B\mid(r(a)\otimes\operatorname{\mathrm{id}})(x)=(\operatorname{\mathrm{id}}\otimes l(x))(a)\text{ for all }x\in U_{h}^{k}(\mathfrak{l}_{0})\},

where r(x)r(x) (resp. l(x)l(x)) is the right (resp. left) multiplication by xUhk(𝔩0)x\in U_{h}^{k}(\mathfrak{l}_{0}).

Corollary 6.3.

Let (kh,χ)(k\llbracket h\rrbracket,\chi) be an 𝒜[2Q0]\mathcal{A}[2Q_{0}^{-}]-algebra same with χ2α=0\chi_{-2\alpha}=0 for αRRS0\alpha\in R^{-}\setminus R_{S_{0}}^{-} and χ2α=1\chi_{-2\alpha}=1 for αRS+\alpha\in R_{S}^{+}. Then 𝒪h,χk(LS\G)indL0,hGh𝒪h,χ(LS\L0)\mathcal{O}_{h,\chi}^{k}(L_{S}\backslash G)\cong\mathrm{ind}_{L_{0,h}}^{G_{h}}\mathcal{O}_{h,\chi}(L_{S}\backslash L_{0}).

Proof.

The 22-cocycle arising from Uhk(𝔩S)Uh,χk(𝔩0)U_{h}^{k}(\mathfrak{l}_{S}{}^{\sim})\subset U_{h,\chi}^{k}(\mathfrak{l}_{0}{}^{\sim}) (resp. Uh,χ(𝔤)U_{h,\chi}(\mathfrak{g}{}^{\sim})) is denoted by Ψ0\Psi_{0} (resp. Ψ\Psi). By Proposition 6.1, we have Ψ0=Ψ\Psi_{0}=\Psi throught the embedding Uhk(𝔩0×𝔩0×𝔩S)Uhk(𝔤×𝔤×𝔩S)U_{h}^{k}(\mathfrak{l}_{0}\times\mathfrak{l}_{0}\times\mathfrak{l}_{S})\subset U_{h}^{k}(\mathfrak{g}\times\mathfrak{g}\times\mathfrak{l}_{S}).

In light of the spectral decomposition (19) and the Peter-Weyl theorem for 𝒪hk(G)\mathcal{O}_{h}^{k}(G), indL0,hGh𝒪h,χ(LS\L0)\mathrm{ind}_{L_{0,h}}^{G_{h}}\mathcal{O}_{h,\chi}(L_{S}\backslash L_{0}) has the following expression:

indL0,hGh𝒪h,χ(LS\L0)πIrrh𝔤,ρIrrh𝔩0HomUhk(𝔩S)(Vρ,kh)HomUhk(𝔩0)(Vπ,Vρ)Vπ,\displaystyle\mathrm{ind}_{L_{0,h}}^{G_{h}}\mathcal{O}_{h,\chi}(L_{S}\backslash L_{0})\cong\bigoplus_{\pi\in\operatorname{\mathrm{Irr}}_{h}\mathfrak{g},\rho\in\operatorname{\mathrm{Irr}}_{h}\mathfrak{l}_{0}}\operatorname{\mathrm{Hom}}_{U_{h}^{k}(\mathfrak{l}_{S})}(V_{\rho},k\llbracket h\rrbracket)\otimes\operatorname{\mathrm{Hom}}_{U_{h}^{k}(\mathfrak{l}_{0})}(V_{\pi},V_{\rho})\otimes V_{\pi},
(fTv)(gSw)=(fg)Ψ01(TS)(vw).\displaystyle(f\otimes T\otimes v)\ast(g\otimes S\otimes w)=(f\otimes g)\Psi_{0}^{-1}\otimes(T\otimes S)\otimes(v\otimes w).

Note that we have

(fg)Ψ01(TS)=(fg)(TS)Ψ1=(fTgS)Ψ1.\displaystyle(f\otimes g)\Psi_{0}^{-1}(T\otimes S)=(f\otimes g)(T\otimes S)\Psi^{-1}=(f\circ T\otimes g\circ S)\Psi^{-1}.

On the other hand, we have

HomUhk(𝔩S)(Vπ,kh)ρIrrh𝔩0HomUhk(𝔩S)(Vρ,kh)HomUhk(𝔩0)(Vπ,Vρ),\displaystyle\operatorname{\mathrm{Hom}}_{U_{h}^{k}(\mathfrak{l}_{S})}(V_{\pi},k\llbracket h\rrbracket)\cong\bigoplus_{\rho\in\operatorname{\mathrm{Irr}}_{h}\mathfrak{l}_{0}}\operatorname{\mathrm{Hom}}_{U_{h}^{k}(\mathfrak{l}_{S})}(V_{\rho},k\llbracket h\rrbracket)\otimes\operatorname{\mathrm{Hom}}_{U_{h}^{k}(\mathfrak{l}_{0})}(V_{\pi},V_{\rho}),

in which fTf\otimes T corresponds to fTf\circ T. Combining these facts we obtain the desired isomorphism. ∎

Acknowlegements. The author is grateful to Yasuyuki Kawahigashi for comments on this paper and his invaluable supports. He also thanks Yuki Arano for encouraging him to write this paper.

References