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

Quantum affine algebras, graded limits and flags

Matheus Brito Departamento de Matematica, UFPR, Curitiba - PR - Brazil, 81530-015 mbrito@ufpr.br Vyjayanthi Chari Department of Mathematics, University of California, Riverside, 900 University Ave., Riverside, CA 92521, USA chari@math.ucr.edu Deniz Kus University of Bochum, Faculty of Mathematics, Universitätsstr. 150, 44801 Bochum, Germany deniz.kus@rub.de  and  R. Venkatesh Department of Mathematics, Indian Institute of Science, Bangalore 560012, India rvenkat@iisc.ac.in
Abstract.

In this survey, we review some of the recent connections between the representation theory of (untwisted) quantum affine algebras and the representation theory of current algebras. We mainly focus on the finite-dimensional representations of these algebras. This connection arises via the notion of the graded and classical limit of finite-dimensional representations of quantum affine algebras. We explain how this study has led to interesting connections with Macdonald polynomials and discuss a BGG-type reciprocity result. We also discuss the role of Demazure modules in this theory and several recent results on the presentation, structure and combinatorics of Demazure modules.

V.C. was partially supported by DMS-1719357, the Simons Fellows Program and by the Infosys Visiting Chair position at the Indian Institute of Science.
D.K. was partially funded by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation)– grant 446246717.
R.V. was partially supported by the Infosys Young Investigator Award grant.

1. Introduction

Quantized enveloping algebras were introduced independently by Drinfeld (1985) and Jimbo (1986) in the context of integrable systems and solvable lattice models and give a systematic way to construct solutions to the quantum Yang-Baxter equation. The quantized algebra associated to an affine Lie algebra is called a quantum affine algebra. The representation theory of these has been intensively studied for nearly thirty-five years since its introduction. It has connections with many research areas of mathematics and physics, for example, statistical mechanics, cluster algebras, dynamical systems, the geometry of quiver varieties, Macdonald polynomials to name a few. In this survey, we mainly focus on the category of finite-dimensional representations q\mathcal{F}_{q} of quantum affine algebras and their connections to graded representations of current algebras. The fact that this category is not semi-simple gives a very rich structure and has many interesting consequences. The category is studied via the Drinfeld realization of quantum affine algebras and irreducible objects are parametrized in terms of Drinfeld polynomials. The classical version of q\mathcal{F}_{q} was studied previously in [19], [33], and the irreducible finite-dimensional representations of the affine algebra and the loop algebra were classified in those papers.

However, we still have limited information on the structure of finite-dimensional representations of quantum affine algebras except for a few special cases. For example, we do not even know the dimension formulas in general. One way to study these representations is to go from quantum level to classical level by forming the classical limit, see for instance [39] for a necessary and sufficient condition for the existence of the classical limit. The classical limit (when it exists) is a finite-dimensional module for the corresponding affine Lie algebra. By restricting and suitably twisting this classical limit, we obtain the graded limit which is a graded representation of the corresponding current algebra, see Section 4 for more details. Most of the time we get a reducible indecomposable representation of affine Lie algebra (or current algbera) on passing to the classical (graded) limit. A similar phenomenon is observed in modular representation theory: an irreducible finite-dimensional representation in characteristic zero becomes reducible on passing to characteristic pp. Many interesting families of representations from q\mathcal{F}_{q} admit this graded limit, for instance, the local Weyl modules, Kirillov-Reshetikhin modules, minimal affinizations, and some of the prime representations coming from the work of Hernandez and Leclerc on monoidal categorification.

In [39] the authors introduced the notion of local Weyl modules for a quantum affine algebra. They are given by generators and relations, are highest weight modules in a suitable sense and satisfy a canonical universal property. In particular, any irreducible module in q\mathcal{F}_{q} is a quotient of some Weyl module. It was conjectured in [39] (and proved there for 𝔰𝔩2\mathfrak{sl}_{2}) that any local Weyl module has a tensor product decomposition into fundamental local Weyl modules, see Section 3.2.6 for more details. This conjecture stimulated a lot of research on this topic and the general case was established through the work of [30, 58, 88]. The work of Kirillov and Reshetikhin [76] has a connection with the irreducible representations of quantum affine algebras corresponding to a multiple of a fundamental weight. These modules are referred as Kirillov-Reshetikhin modules in the literature, because they conjectured the classical decomposition of these modules in their paper. The study of Kirillov-Reshetikhin modules has been of immense interest in recent years due to their rich combinatorial structures and several applications to mathematical physics [64, 65]. Many important conjectures on the character formulas of these modules and their fusion products were made from physical considerations and they stimulated lots of research, see [59, 91, 92] and the references therein.

One of the very natural questions that arises from the work of [61] is: what is the smallest representation from q\mathcal{F}_{q} that corresponds to a given finite-dimensional irreducible representation of the underlying simple Lie algebra 𝔤\mathfrak{g}? The second author introduced the notion of a minimal affinization in [20] with this motivation and it was further studied in [21, 34, 35]. One introduces a poset for each dominant integral weight, such that each element of the poset determines a family of irreducible representations in q\mathcal{F}_{q}. The irreducible modules that correspond to the minimal elements of this poset are minimal affinizations. Kirillov-Reshetikhin modules are the minimal affinizations of multiples of the fundamental weights. Our final example of graded limits comes from the work of Hernandez and Leclerc [67] on the monoidal categorification of cluster algebras. The authors defined an interesting monoidal subcategory of q\mathcal{F}_{q} in simply-laced type and proved that for 𝔤\mathfrak{g} of type AnA_{n} and D4D_{4}, it categorifies a cluster algebra of the same type, i.e., its Grothendieck ring admits a cluster algebra structure of the same type as 𝔤\mathfrak{g}. The prime real representations of this subcategory are the cluster variables and these are called the HL-modules. A more detailed discussion of HL-modules can be found in Section 4.1.8.

Even though the study of graded representations of current algebras is mainly motivated by their connection with the representations of quantum affine algebras (via graded limits), they are now of independent interest as they have found many applications in number theory, combinatorics, and mathematical physics. They have connections with mock theta functions, cone theta functions [12, 13, 15], symmetric Macdonald polynomials [14, 27, 71], the X=MX=M conjecture [2, 59, 91], and Schur positivity [55, 101] etc. One of the very important families of graded representations of current algebras comes from 𝔤\mathfrak{g}-stable Demazure modules. A Demazure module by definition is a module of the Borel subalgebra of the affine Lie algebra. If it is 𝔤\mathfrak{g}-stable, then it naturally becomes a module of the maximal parabolic subalgebra which contains the current algebra. By restriction, we get a graded module of the current algebra. These modules are parametrized by pairs consisting of a positive integer and a dominant weight, and given such a pair (,λ)(\ell,\lambda), the corresponding 𝔤\mathfrak{g}-stable Demazure module of the current algebra is denoted by D(,λ)D(\ell,\lambda). These modules include all well-known families of graded representations of current algebras. For example, any local Weyl module of the current algebra is isomorphic to a level one Demazure module D(1,λ)D(1,\lambda) when 𝔤\mathfrak{g} is simply-laced.

The limit of a tensor product of quantum affine algebra modules is not necessarily isomorphic to the tensor product of their classical limits. So, we need to replace the tensor product with something else in order to study the limit of a tensor product of quantum affine algebra modules. Examples suggest that the fusion product introduced by Feigin and Loktev [50] is the correct notion that should replace the tensor product. It is a very important and seemingly very hard problem to understand the fusion products of 𝔤\mathfrak{g}-stable Demazure modules of various levels. One would like to find the generators and relations and the graded character of these modules, but very limited cases are known [3, 42, 54, 91].

The survey is organized as follows. We begin by stating the foundational results, including the definition of local Weyl modules and the classification of irreducible modules in Section 2. In Section 3, we discuss various well-studied families of finite-dimensional representations of quantum affine algebras and review the literature on the presentation of these modules, their classical limit, and the closely related graded limits. Later we move on to the study of graded finite-dimensional representations of current algebras. We relate the local Weyl modules to the 𝔤\mathfrak{g}-stable Demazure modules and discuss the connection between the characters of the local Weyl modules and Macdonald polynomials. We also discuss the BGG-type reciprocity results and briefly mention some recent developments on tilting modules, generalized Weyl modules, and global Demazure modules. In the end, we collect together some results on Demazure modules.

Acknowledgements. Part of this paper was written while the second author was visiting the Max Planck Institute, Bonn, in Fall 2021. She gratefully acknowledges the financial support and the excellent working conditions provided by the institute.

2. The simple and untwisted affine Lie algebras

In this section we collect the notation and some well-known results on the structure and representation theory of affine Lie algebras.

2.1. Conventions

We let \mathbb{C} (resp. \mathbb{Q}, \mathbb{Z}, +\mathbb{Z}_{+}, \mathbb{N}) be the set of complex numbers (resp. rational numbers, integers, non-negative integers, positive integers). We adopt the convention that given two complex vector spaces V,WV,W the corresponding tensor product VWV\otimes_{\mathbb{C}}W will be just denoted as VWV\otimes W.

Given an indeterminate tt we let [t]\mathbb{C}[t] (resp. [t,t1]\mathbb{C}[t,t^{-1}], (t)\mathbb{C}(t)) be the ring of polynomials (resp. Laurent polynomials, rational functions) in the variable t.t. For s,m,r+s\in\mathbb{Z},m,r\in\mathbb{Z}_{+} with mrm\geq r, set

[s]t=tststt1,[m]t!=[m]t[m1]t[1]t,[mr]t=[m]t![r]t![mr]t!.[s]_{t}=\frac{t^{s}-t^{-s}}{t-t^{-1}},\ \ [m]_{t}!=[m]_{t}[m-1]_{t}\cdots[1]_{t},\ \ {\genfrac{[}{]}{0.0pt}{0}{m}{r}}_{t}=\frac{[m]_{t}!}{[r]_{t}![m-r]_{t}!}.

For any complex Lie algebra 𝔞\mathfrak{a} we let 𝕌(𝔞)\mathbb{U}(\mathfrak{a}) be the corresponding universal enveloping algebra. Given any commutative associative algebra AA over \mathbb{C} define a Lie algebra structure on 𝔞A\mathfrak{a}\otimes A by

[xa,yb]=[x,y]ab,x,y𝔞,a,bA.[x\otimes a,y\otimes b]=[x,y]\otimes ab,\ \ x,y\in\mathfrak{a},\ a,b\in A.

In the special case when AA is [t]\mathbb{C}[t] or [t,t1]\mathbb{C}[t,t^{-1}] we set

𝔞[t]=𝔞[t],L(𝔞)=𝔞[t±1].\mathfrak{a}[t]=\mathfrak{a}\otimes\mathbb{C}[t],\ \ L(\mathfrak{a})=\mathfrak{a}\otimes\mathbb{C}[t^{\pm 1}].

2.2. Simple and affine Lie algebras

2.2.1. The simple Lie algebra 𝔤\mathfrak{g}

Let 𝔤\mathfrak{g} denote a simple finite-dimensional Lie algebra over \mathbb{C} and let 𝔥\mathfrak{h} be a fixed Cartan subalgebra of 𝔤\mathfrak{g} and RR the corresponding set of roots. Let I={1,,n}I=\{1,\dots,n\} be an index set for the set of simple roots {α1,,αn}\{\alpha_{1},\dots,\alpha_{n}\} of RR and {ω1,,ωn}\{\omega_{1},\dots,\omega_{n}\} a set of fundamental weights. Given λ,μ𝔥\lambda,\mu\in\mathfrak{h}^{*} we say that

λμλμiI+αi.\lambda\geq\mu\iff\lambda-\mu\in\sum_{i\in I}\mathbb{Z}_{+}\alpha_{i}.

Let PP, QQ (resp. P+,Q+P^{+},Q^{+}) be the \mathbb{Z}-span (resp. +\mathbb{Z}_{+}-span) of the fundamental weights and simple roots respectively and let R+=RQ+R^{+}=R\cap Q^{+}. We denote by θR+\theta\in R^{+} the highest root in R+R^{+} and let (,)(\ ,\ ) be the form on 𝔥\mathfrak{h}^{*} induced by the restriction of the Killing form κ:𝔤𝔤\kappa:\mathfrak{g}\otimes\mathfrak{g}\to\mathbb{C} of 𝔤\mathfrak{g}. We assume that it is normalized so that (θ,θ)=2(\theta,\theta)=2 and set

dα=2/(α,α),di=dαi,αi=diαi,ωi=diωi,ai,j=(αj,αi), 1i,jn.d_{\alpha}=2/(\alpha,\alpha),\ \ d_{i}=d_{\alpha_{i}},\ \ \alpha_{i}^{\vee}=d_{i}\alpha_{i},\ \ \omega_{i}^{\vee}=d_{i}\omega_{i},\ \ a_{i,j}=(\alpha_{j},\alpha_{i}^{\vee}),\ \ 1\leq i,j\leq n.

Let WW be the Weyl group of 𝔤\mathfrak{g}; recall that it is the subgroup of Aut(𝔥){\rm{Aut}}(\mathfrak{h}^{*}) generated by the reflections si,iIs_{i},i\in I, defined by:

si(λ)=λ(λ,αi)αi,iI.s_{i}(\lambda)=\lambda-(\lambda,\alpha_{i}^{\vee})\alpha_{i},\ \ i\in I.

Fix a Chevalley basis {xα±,hi:αR+,iI}\{x_{\alpha}^{\pm},\ h_{i}:\alpha\in R^{+},i\in I\} of 𝔤\mathfrak{g} and set for simplicity xi±=xαi±x_{i}^{\pm}=x_{\alpha_{i}}^{\pm}. The elements xi±,hix_{i}^{\pm},h_{i}, iIi\in I generate 𝔤\mathfrak{g} as a Lie algebra. Given α=i=1nriαiR+\alpha=\sum_{i=1}^{n}r_{i}\alpha_{i}\in R^{+} let hα𝔥h_{\alpha}\in\mathfrak{h} be given by hα=dαi=1nridihih_{\alpha}=d_{\alpha}\sum_{i=1}^{n}\frac{r_{i}}{d_{i}}h_{i} and note that the elements sα,αR+s_{\alpha},\alpha\in R^{+}, defined by sα(λ)=λλ(hα)αs_{\alpha}(\lambda)=\lambda-\lambda(h_{\alpha})\alpha are elements of WW and we have si=sαis_{i}=s_{\alpha_{i}}.

Let 𝔫±\mathfrak{n}^{\pm} be the subalgebra generated by the elements {xi±:iI}\{x_{i}^{\pm}:i\in I\}. Then,

𝔫±=αR+xα±,𝔟±=𝔥𝔫±,𝔤=𝔟±𝔫.\mathfrak{n}^{\pm}=\bigoplus_{\alpha\in R^{+}}\mathbb{C}x_{\alpha}^{\pm},\ \ \ \ \mathfrak{b}^{\pm}=\mathfrak{h}\oplus\mathfrak{n}^{\pm},\ \ \mathfrak{g}=\mathfrak{b}^{\pm}\oplus\mathfrak{n}^{\mp}.

We have a corresponding decomposition of 𝕌(𝔤)\mathbb{U}(\mathfrak{g}) as vector spaces

𝕌(𝔤)𝕌(𝔫)𝕌(𝔥)𝕌(𝔫+)𝕌(𝔫)𝕌(𝔟+).\mathbb{U}(\mathfrak{g})\cong\mathbb{U}(\mathfrak{n}^{-})\otimes\mathbb{U}(\mathfrak{h})\otimes\mathbb{U}(\mathfrak{n}^{+})\cong\mathbb{U}(\mathfrak{n}^{-})\otimes\mathbb{U}(\mathfrak{b}^{+}).

2.2.2. The affine Lie algebra

The (untwisted) affine Lie algebra 𝔤^\widehat{\mathfrak{g}} and its Cartan subalgebra 𝔥^\widehat{\mathfrak{h}} are defined as follows:

𝔤^=L(𝔤)cd,𝔥^=𝔥cd\widehat{\mathfrak{g}}=L(\mathfrak{g})\oplus\mathbb{C}c\oplus\mathbb{C}d,\ \ \ \widehat{\mathfrak{h}}=\mathfrak{h}\oplus\mathbb{C}c\oplus\mathbb{C}d

with commutator given by requiring cc to be central and

[xtr,yts]=[x,y]tr+s+rδr+s,0κ(x,y)c,[d,xtr]=rxtr.[x\otimes t^{r},y\otimes t^{s}]=[x,y]\otimes t^{r+s}+r\delta_{r+s,0}\kappa(x,y)c,\ \ [d,x\otimes t^{r}]=rx\otimes t^{r}.

Here x,y𝔤x,y\in\mathfrak{g}, r,sr,s\in\mathbb{Z} and δn,m\delta_{n,m} is the Kronecker delta symbol. Setting h0=hθ+ch_{0}=-h_{\theta}+c we see that the set {hi,d:0in}\{h_{i},d:0\leq i\leq n\} is a basis for 𝔥^\widehat{\mathfrak{h}}.

Regard an element λ𝔥\lambda\in\mathfrak{h}^{*} as an element of 𝔥^\widehat{\mathfrak{h}}^{*} by setting λ(c)=0=λ(d)\lambda(c)=0=\lambda(d). Define elements δ\delta, α0\alpha_{0} and the affine fundamental weights Λi\Lambda_{i}, 0in0\leq i\leq n, of 𝔥^\widehat{\mathfrak{h}}^{*} by:

δ(d)=1,δ(𝔥c)=0,α0=θ+δ,Λ0(c)=1,Λ0(𝔥d)=0,\delta(d)=1,\ \ \delta(\mathfrak{h}\oplus\mathbb{C}c)=0,\ \ \alpha_{0}=-\theta+\delta,\ \ \Lambda_{0}(c)=1,\ \ \Lambda_{0}(\mathfrak{h}\oplus\mathbb{C}d)=0,
Λi(hj)=δi,j,Λi(d)=0,iI,j{0,,n}.\Lambda_{i}(h_{j})=\delta_{i,j},\ \ \Lambda_{i}(d)=0,\ \ i\in I,\ \ j\in\{0,\dots,n\}.

The subset

R^={α+rδ:αR{0},r}\{0}𝔥^,\widehat{R}=\{\alpha+r\delta:\alpha\in R\cup\{0\},r\in\mathbb{Z}\}\backslash\{0\}\subseteq\widehat{\mathfrak{h}}^{*},

is called the set of affine roots. The set of affine simple roots is {αi:iI^}\{\alpha_{i}:i\in\widehat{I}\} where I^={0,1,,n}\widehat{I}=\{0,1,\dots,n\}. The corresponding set of positive roots is given by:

R^+={±α+(r+1)δ:αR+{0},r+}R+.\widehat{R}^{+}=\{\pm\alpha+(r+1)\delta:\alpha\in R^{+}\cup\{0\},r\in\mathbb{Z}_{+}\}\ \cup R^{+}.

Set

P^=i=0nΛi+δ,P^+=i=0n+Λi+δ\widehat{P}=\sum_{i=0}^{n}\mathbb{Z}\Lambda_{i}+\mathbb{Z}\delta,\ \ \ \widehat{P}^{+}=\sum_{i=0}^{n}\mathbb{Z}_{+}\Lambda_{i}+\mathbb{Z}\delta

Let Q^\widehat{Q} (resp. Q^+)\widehat{Q}^{+}) be the \mathbb{Z}-span (resp. +\mathbb{Z}_{+}-span) of the affine simple roots. The affine Weyl group W^\widehat{W} is the subgroup of Aut(𝔥^){\rm{Aut}}(\widehat{\mathfrak{h}}^{*}) generated by the set {si:iI^}\{s_{i}:i\in\widehat{I}\} where

si(λ)=λλ(hi)αi,iI^,λ𝔥^.s_{i}(\lambda)=\lambda-\lambda(h_{i})\alpha_{i},\ \ i\in\widehat{I},\ \ \lambda\in\widehat{\mathfrak{h}}^{*}.

Clearly WW is a subgroup of W^\widehat{W} and we have an isomorphism

W^WiIαi.\widehat{W}\cong W\ltimes\sum_{i\in I}\mathbb{Z}\alpha^{\vee}_{i}.

It will also be convenient to introduce the extended affine Weyl group W~=WiIωi.\widetilde{W}=W\ltimes\sum_{i\in I}\mathbb{Z}\omega^{\vee}_{i}.

Setting x0±=xθt±1x_{0}^{\pm}=x_{\theta}^{\mp}\otimes t^{\pm 1} we observe that 𝔤^\widehat{\mathfrak{g}} is generated by the elements {xi±,hi:iI^}{d}\{x_{i}^{\pm},h_{i}:i\in\widehat{I}\}\cup\{d\}. The root space corresponding to an element ±α+sδR^\pm\alpha+s\delta\in\widehat{R} with αR+\alpha\in R^{+} is (xα±ts)\mathbb{C}(x^{\pm}_{\alpha}\otimes t^{s}) and to an element rδr\delta is (𝔥tr)(\mathfrak{h}\otimes t^{r}), s,rs,r\in\mathbb{Z} and r0r\neq 0. We shall just denote a non-zero element of the one-dimensional root space corresponding to ±α\pm\alpha, αR^+δ\alpha\in\widehat{R}^{+}\setminus\mathbb{N}\delta by xα±x_{\alpha}^{\pm} and let hα=[xα+,xα]h_{\alpha}=[x_{\alpha}^{+},x_{\alpha}^{-}]. The subalgebras 𝔫^±\widehat{\mathfrak{n}}^{\pm} and 𝔟^\widehat{\mathfrak{b}} are defined in the obvious way and we have

𝔫^±=𝔤t±1[t]𝔫±,𝔟^+=𝔥^𝔫^+.\widehat{\mathfrak{n}}^{\pm}=\mathfrak{g}\otimes t^{\pm 1}\mathbb{C}[t]\oplus\mathfrak{n}^{\pm},\ \ \widehat{\mathfrak{b}}^{+}=\widehat{\mathfrak{h}}\oplus\widehat{\mathfrak{n}}^{+}.

This gives rise to an analogous triangular decomposition

𝕌(𝔤^)𝕌(𝔫^)𝕌(𝔥^)𝕌(𝔫^+)𝕌(𝔫^)𝕌(𝔟^+).\mathbb{U}(\widehat{\mathfrak{g}})\cong\mathbb{U}(\widehat{\mathfrak{n}}^{-})\otimes\mathbb{U}(\widehat{\mathfrak{h}})\otimes\mathbb{U}(\widehat{\mathfrak{n}}^{+})\cong\mathbb{U}(\widehat{\mathfrak{n}}^{-})\otimes\mathbb{U}(\widehat{\mathfrak{b}}^{+}).

2.2.3. The loop algebra L(𝔤)L(\mathfrak{g}) and the current algebra 𝔤[t]\mathfrak{g}[t]

It is trivial to see that L(𝔤)dL(\mathfrak{g})\oplus\mathbb{C}d is the quotient of 𝔤^\widehat{\mathfrak{g}} by the center c\mathbb{C}c. The action of dd obviously induces a \mathbb{Z}-grading on L(𝔤)L(\mathfrak{g}) and the current algebra 𝔤[t]\mathfrak{g}[t] is a graded subalgebra of L(𝔤)L(\mathfrak{g}). Moreover 𝔤[t]cd\mathfrak{g}[t]\oplus\mathbb{C}c\oplus\mathbb{C}d can also be regarded as a maximal parabolic subalgebra of 𝔤^\widehat{\mathfrak{g}}, namely

𝔤[t]cd𝔟+^𝔫.\mathfrak{g}[t]\oplus\mathbb{C}c\oplus\mathbb{C}d\cong\widehat{\mathfrak{b}^{+}}\oplus\mathfrak{n}^{-}.

We make the grading on L(𝔤)L(\mathfrak{g}) and 𝔤[t]\mathfrak{g}[t] explicit for the reader’s convenience. For rr\in\mathbb{Z} we declare 𝔤tr\mathfrak{g}\otimes t^{r} to be the rr-th graded piece and note that for r,sr,s\in\mathbb{Z}, we have [𝔤tr,𝔤ts]=𝔤tr+s[\mathfrak{g}\otimes t^{r},\mathfrak{g}\otimes t^{s}]=\mathfrak{g}\otimes t^{r+s}. This grading induces a grading on the corresponding enveloping algebras as well once we declare a monomial of the form (a1tr1)(aptrp)(a_{1}\otimes t^{r_{1}})\cdots(a_{p}\otimes t^{r_{p}}), as𝔤a_{s}\in\mathfrak{g}, rsr_{s}\in\mathbb{Z}, 1sp1\leq s\leq p to have grade (r1++rp)(r_{1}+\cdots+r_{p}).

2.2.4. Ideals in L(𝔤)L(\mathfrak{g})

The affine Lie algebra is clearly not simple; the center spans a one-dimensional ideal. One can prove using the explicit realization that this and the derived algebra L(𝔤)cL(\mathfrak{g})\oplus\mathbb{C}c are the only non trivial proper ideals in 𝔤^\widehat{\mathfrak{g}}. The following result is well-known, a proof can be found for instance in [25, Lemma 1].

Lemma.

For all f[t,t1]f\in\mathbb{C}[t,t^{-1}] the subspace 𝔤f[t,t1]\mathfrak{g}\otimes f\mathbb{C}[t,t^{-1}] is an ideal in L(𝔤)L(\mathfrak{g}). Moreover any ideal in L(𝔤)L(\mathfrak{g}) must be of this form. In particular all ideals are of finite codimension. Writing f=(ta1)r1(tak)rkf=(t-a_{1})^{r_{1}}\cdots(t-a_{k})^{r_{k}} with arasa_{r}\neq a_{s} for 1rsk1\leq r\neq s\leq k we see that we have an isomorphism of Lie algebras

L(𝔤)𝔤f[t,t1]𝔤[t,t1](f)s=1k(𝔤[t,t1](tas)rs).\frac{L(\mathfrak{g})}{\mathfrak{g}\otimes f\mathbb{C}[t,t^{-1}]}\cong\mathfrak{g}\otimes\frac{\mathbb{C}[t,t^{-1}]}{(f)}\cong\bigoplus_{s=1}^{k}\left(\mathfrak{g}\otimes\frac{\mathbb{C}[t,t^{-1}]}{(t-a_{s})^{r_{s}}}\right).

We shall sometimes refer to the finite-dimensional quotient of L(𝔤)L(\mathfrak{g}) defined by f[t,t1]f\in\mathbb{C}[t,t^{-1}] as the truncation of L(𝔤)L(\mathfrak{g}) at f.f.

2.3. Representations of simple and affine Lie algebras

2.3.1. Finite-dimensional representations of 𝔤\mathfrak{g}

We say that a 𝔤\mathfrak{g}-module VV is a weight module if,

V=μ𝔥Vμ,Vμ={vV:hv=μ(h)v,h𝔥}V=\bigoplus_{\mu\in\mathfrak{h}^{*}}V_{\mu},\ \ V_{\mu}=\{v\in V:hv=\mu(h)v,\ \forall h\in\mathfrak{h}\}

and we let wt(V)={μ𝔥:Vμ0}.{\rm{wt}}(V)=\{\mu\in\mathfrak{h}^{*}:V_{\mu}\neq 0\}. If wt(V)P{\rm wt}(V)\subset P and dimVμ<\dim V_{\mu}<\infty for all μP\mu\in P, we let ch(V){\rm{ch}}(V) be the formal sum μP(dimVμ)eμ\sum_{\mu\in P}(\dim V_{\mu})\ e_{\mu} where eμe_{\mu} varies over a basis of the group ring [P]\mathbb{Z}[P].

Given λP+\lambda\in P^{+} let V(λ)V(\lambda) be the 𝔤\mathfrak{g}-module generated by an element vλv_{\lambda} with defining relations:

hivλ=λ(hi)vλ,xi+vλ=0,(xi)λ(hi)+1vλ=0, 1in.h_{i}v_{\lambda}=\lambda(h_{i})v_{\lambda},\ \ x_{i}^{+}v_{\lambda}=0,\ \ (x_{i}^{-})^{\lambda(h_{i})+1}v_{\lambda}=0,\ \ 1\leq i\leq n.

It is well-known [69] that V(λ)V(\lambda) is a weight module with wt(V(λ))λQ+{\rm{wt}}(V(\lambda))\subseteq\lambda-Q^{+} and that it is also an irreducible and finite-dimensional 𝔤\mathfrak{g}-module. The set wt(V(λ)){\rm{wt}}(V(\lambda)) is WW-invariant and dimV(λ)μ=dimV(λ)wμ\dim V(\lambda)_{\mu}=\dim V(\lambda)_{w\mu} for all wWw\in W.

Any finite-dimensional 𝔤\mathfrak{g}-module is isomorphic to a direct sum of copies of V(λ)V(\lambda), λP+\lambda\in P^{+}.

2.3.2. Integrable and positive level representations of 𝔤^\widehat{\mathfrak{g}}

The notion of a weight module and its character for 𝔤^\widehat{\mathfrak{g}} are defined as for simple Lie algebras with 𝔥\mathfrak{h} replaced by 𝔥^\widehat{\mathfrak{h}}. We say that a 𝔤^\widehat{\mathfrak{g}}-module VV is integrable if it is a weight module and the elements xi±x_{i}^{\pm}, iI^i\in\widehat{I}, act locally nilpotently. We say that VV is of level rr\in\mathbb{Z} if cv=rvcv=rv for all vVv\in V; if r>0r>0 (resp. r<0r<0) then we say that VV is of positive level (resp. negative level).

Given λP^+\lambda\in\widehat{P}^{+} let V(λ)V(\lambda) be the 𝔤^\widehat{\mathfrak{g}}-module generated by an element vλv_{\lambda} with defining relations:

hivλ=λ(hi)vλ,xi+vλ=0,(xi)λ(hi)+1vλ=0, 0in.h_{i}v_{\lambda}=\lambda(h_{i})v_{\lambda},\ \ x_{i}^{+}v_{\lambda}=0,\ \ (x_{i}^{-})^{\lambda(h_{i})+1}v_{\lambda}=0,\ \ 0\leq i\leq n.

Again it is well-known [73] that V(λ)V(\lambda) is an integrable irreducible module with positive level λ(c)\lambda(c) and

dimV(λ)μ0μλQ^+,dimV(λ)μ<,μP^,\dim V(\lambda)_{\mu}\neq 0\implies\mu\in\lambda-\widehat{Q}^{+},\ \ \dim V(\lambda)_{\mu}<\infty,\ \ \mu\in\widehat{P},
dimV(λ)μ=dimV(λ)wμ,wW^,μP^.\dim V(\lambda)_{\mu}=\dim V(\lambda)_{w\mu},\ \ w\in\widehat{W},\ \ \mu\in\widehat{P}.

Notice that the preceding properties show immediately that V(λ)V(\lambda) is infinite-dimensional as long as λδ\lambda\notin\mathbb{Z}\delta. In the case when VV is irreducible the following was proved in [19]. The complete reducibility statement was proved in [48].

Theorem.

Any positive level integrable 𝔤^\widehat{\mathfrak{g}}-module VV with the property that dimVμ<\dim V_{\mu}<\infty for all μP^\mu\in\widehat{P} is isomorphic to a direct sum of modules of the form V(λ)V(\lambda), λP^+\lambda\in\widehat{P}^{+}. ∎

There is a completely similar theory for negative level modules.

2.3.3. Demazure modules

Given λP+\lambda\in P^{+} and wWw\in W (resp. λP^+,wW^\lambda\in\widehat{P}^{+},w\in\widehat{W}) the Demazure module Vw(λ)V_{w}(\lambda) is the 𝔟\mathfrak{b}-submodule (resp. 𝔟^\widehat{\mathfrak{b}}-submodule) of V(λ)V(\lambda) generated by the one dimensional subspace V(λ)wλV(\lambda)_{w\lambda}. The Demazure modules are always finite-dimensional; in the case of 𝔤\mathfrak{g} this statement is trivial while for 𝔤^\widehat{\mathfrak{g}} the statement follows from the fact that wt(V(λ))λQ^+{\rm wt}(V(\lambda))\subseteq\lambda-\widehat{Q}^{+}. For λP+\lambda\in P^{+} and wWw\in W, (resp. λP^+\lambda\in\widehat{P}^{+}, wW^w\in\widehat{W}) it is a result from [72, 85] that as a 𝕌(𝔟)\mathbb{U}(\mathfrak{b})-module (resp. 𝕌(𝔟^)\mathbb{U}(\widehat{\mathfrak{b}})-module) Vw(λ)V_{w}(\lambda) is generated by vwλv_{w\lambda} with relations: hvwλ=(wλ)(h)vwλ,hv_{w\lambda}=(w\lambda)(h)v_{w\lambda}, for all h𝔥h\in\mathfrak{h} (resp. for all h𝔥^h\in\widehat{\mathfrak{h}}) and

(xα+)p+1vwλ=0,pmax{0,wλ(hα)},αR+(x_{\alpha}^{+})^{p+1}v_{w\lambda}=0,\ \ p\geq\max\{0,-w\lambda(h_{\alpha})\},\ \ \alpha\in R^{+}
(resp.(𝔥tr+1)vwλ=0,r+,(xα+)p+1vwλ=0,pmax{0,wλ(hα)},αR^+δ).({\rm{resp.}}(\mathfrak{h}\otimes t^{r+1})v_{w\lambda}=0,\ \ r\in\mathbb{Z}_{+},\ (x^{+}_{\alpha})^{p+1}v_{w\lambda}=0,\ \ p\geq\max\{0,-w\lambda(h_{\alpha})\},\ \ \alpha\in\widehat{R}^{+}\setminus\mathbb{N}\delta).

These relations were simplified in [42] and [80] and we refer to Theorem Theorem for a precise statement.

Lemma.

Suppose that λP+\lambda\in P^{+}, wWw\in W, and iIi\in I are such that (wλ)(hi)0(w\lambda)(h_{i})\leq 0. Then xiV(λ)wλ=0.x_{i}^{-}V(\lambda)_{w\lambda}=0. An analogous statement holds for V(λ)V(\lambda) with λP^+\lambda\in\widehat{P}^{+}.

Proof.

If wλαiwt(V(λ))w\lambda-\alpha_{i}\in{\rm wt}(V(\lambda)), then λw1αiwt(V(λ))\lambda-w^{-1}\alpha_{i}\in{\rm wt}(V(\lambda)). On the other hand our assumptions force (wλ)(hi)=0(w\lambda)(h_{i})=0 or w1αiRw^{-1}\alpha_{i}\in R^{-} where the latter condition ends in a contradiction to wt(V(λ))λQ+{\rm wt}(V(\lambda))\subseteq\lambda-Q^{+}. Hence (wλ)(hi)=0(w\lambda)(h_{i})=0 and xiV(λ)wλ=0x_{i}^{-}V(\lambda)_{w\lambda}=0 follows. ∎

As a consequence of this lemma we see immediately that

λP^+,wW^,(wλ)(hi)+,iI𝔤[t]Vw(λ)Vw(λ).\lambda\in\widehat{P}^{+},\ \ w\in\widehat{W},\ \ (w\lambda)(h_{i})\in-\mathbb{Z}_{+},\ \forall\ i\in I\implies\ \ \mathfrak{g}[t]V_{w}(\lambda)\subseteq V_{w}(\lambda).

We call these 𝔟^\widehat{\mathfrak{b}}-submodules the 𝔤\mathfrak{g}-stable Demazure modules.

2.3.4. Level zero modules for 𝔤^\widehat{\mathfrak{g}} and finite-dimensional modules for L(𝔤)L(\mathfrak{g}) and 𝔤[t]\mathfrak{g}[t]

A level zero module for 𝔤^\widehat{\mathfrak{g}} is one on which the center acts trivially, in particular it can be regarded as a module for L(𝔤)dL(\mathfrak{g})\oplus\mathbb{C}d. More generally given any representation VV of 𝔤\mathfrak{g} one can define a L(𝔤)dL(\mathfrak{g})\oplus\mathbb{C}d-module structure on L(V)=V[t,t1]L(V)=V\otimes\mathbb{C}[t,t^{-1}] by:

(xtr)(vts)=xvtr+s,d(vtr)=rvtr,r,s,x𝔤.(x\otimes t^{r})(v\otimes t^{s})=xv\otimes t^{r+s},\ \ d(v\otimes t^{r})=rv\otimes t^{r},\ \ r,s\in\mathbb{Z},\ x\in\mathfrak{g}.

The only finite-dimensional representations of 𝔤^\widehat{\mathfrak{g}} are those on which L(𝔤)cL(\mathfrak{g})\oplus\mathbb{C}c acts trivially. We give a proof of this fact for the reader’s convenience. Note that by working with the Jordan–Holder series it suffices to prove this for irreducible finite-dimensional representations. Thus, let VV be a finite-dimensional irreducible representation of 𝔤^\widehat{\mathfrak{g}}. Then it is easily seen that there exists a vector 0vV0\neq v\in V such that the following hold:

(xα+tr)v=0,(hits)v=ai,sv,αR+,iI,r,s+,cv=v,dv=av,(x_{\alpha}^{+}\otimes t^{r})v=0,\ \ (h_{i}\otimes t^{s})v=a_{i,s}v,\ \ \alpha\in R^{+},\ \ i\in I,r\in\mathbb{Z},\ \ s\in\mathbb{Z}_{+},\ \ cv=\ell v,\ \ dv=av,

where a,ai,sa,a_{i,s}\in\mathbb{C} for s>0s>0 and ai,0+a_{i,0}\in\mathbb{Z}_{+} and \ell\in\mathbb{Z}. Then

[d,hits]v=s(hits)v0=sai,svai,s=0,s>0.[d,h_{i}\otimes t^{s}]v=s(h_{i}\otimes t^{s})v\implies 0=sa_{i,s}v\implies a_{i,s}=0,\ \ s>0.

Working with the 𝔰𝔩2\mathfrak{sl}_{2}-triple {xi+ts,xits,hisκ(xi+,xi)c}\{x_{i}^{+}\otimes t^{-s},x_{i}^{-}\otimes t^{s},h_{i}-s\kappa(x_{i}^{+},x_{i}^{-})c\} we see that we must have ai,0sκ(xi+,xi)0a_{i,0}-s\ell\kappa(x_{i}^{+},x_{i}^{-})\geq 0 for all s0s\geq 0 and this forces =0\ell=0. Hence VV must be a finite-dimensional representation of L(𝔤)dL(\mathfrak{g})\oplus\mathbb{C}d. Suppose that ai,0>0a_{i,0}>0 for some iIi\in I. Then working again with the triple {xi+ts,xits,hisκ(xi+,xi)}\{x_{i}^{+}\otimes t^{-s},x_{i}^{-}\otimes t^{s},h_{i}-s\kappa(x_{i}^{+},x_{i}^{-})\} we see that (xits)v0(x_{i}^{-}\otimes t^{s})v\neq 0 for all s>0s>0. Since these elements have dd-eigenvalues a+sa+s they must be linearly independent which is a contradiction. It follows that ai,0=0a_{i,0}=0 for all iIi\in I and then it is easily seen that VV must be the trivial L(𝔤)cL(\mathfrak{g})\oplus\mathbb{C}c-module.

Consider however, the commutator subalgebra L(𝔤)cL(\mathfrak{g})\oplus\mathbb{C}c of 𝔤^\widehat{\mathfrak{g}}. The preceding arguments show that the center must act trivially on any finite-dimensional representation. Hence it suffices to study finite-dimensional representations of L(𝔤)L(\mathfrak{g}). To construct examples we introduce for each a×a\in\mathbb{C}^{\times} the evaluation homomorphism eva:L(𝔤)𝔤\operatorname{ev}_{a}:L(\mathfrak{g})\to\mathfrak{g} which sends xtrarxx\otimes t^{r}\to a^{r}x for x𝔤x\in\mathfrak{g} and rr\in\mathbb{Z}. Given a representation VV of 𝔤\mathfrak{g} let evaV\operatorname{ev}_{a}V be the pull-back L(𝔤)L(\mathfrak{g})-module. The following was proved in [32, 19].

Proposition.
  1. (i)

    Any irreducible finite-dimensional representation of L(𝔤)L(\mathfrak{g}) is isomorphic to a tensor product of the form eva1V(λ1)evakV(λk)\operatorname{ev}_{a_{1}}V(\lambda_{1})\otimes\cdots\otimes\operatorname{ev}_{a_{k}}V(\lambda_{k}) for some k1k\geq 1, pairwise distinct elements a1,,ak×a_{1},\dots,a_{k}\in\mathbb{C}^{\times}, and elements λ1,,λkP+\lambda_{1},\dots,\lambda_{k}\in P^{+}. Moreover any tensor product of irreducible finite-dimensional representations is either irreducible or completely reducible.

  2. (ii)

    Suppose that VV is an irreducible finite-dimensional module of L(𝔤)L(\mathfrak{g}). Then L(V)L(V) is a direct sum of irreducible modules for L(𝔤)dL(\mathfrak{g})\oplus\mathbb{C}d. Any level zero integrable irreducible module for 𝔤^\widehat{\mathfrak{g}} with finite-dimensional weight spaces is obtained as a direct summand of L(V)L(V).

  3. (iii)

    A similar result holds for finite-dimensional modules of 𝔤[t]\mathfrak{g}[t] once we also allow the module ev0V(λ)\operatorname{ev}_{0}V(\lambda). ∎

Level zero modules for 𝔤^\widehat{\mathfrak{g}} are not completely reducible. The simplest example is the adjoint representation where the center is a proper submodule which does not have a complement. Finite-dimensional modules for L(𝔤)L(\mathfrak{g}) and 𝔤[t]\mathfrak{g}[t] are also not completely reducible. For instance the 𝔤\mathfrak{g}-stable Demazure modules are usually reducible and indecomposable. A simple exercise, left to the reader, is to verify this in the case when 𝔤^=𝔰𝔩2^\widehat{\mathfrak{g}}=\widehat{\mathfrak{sl}_{2}}, λ=Λ0\lambda=\Lambda_{0} and w=s1s0w=s_{1}s_{0}.

We shall frequently be interested in \mathbb{Z}-graded modules for 𝔤[t]\mathfrak{g}[t]. These are \mathbb{Z}-graded vector spaces V=mV[m]V=\bigoplus_{m\in\mathbb{Z}}V[m] which admit an action of 𝔤[t]\mathfrak{g}[t] satisfying (𝔤tr)V[m]V[m+r](\mathfrak{g}\otimes t^{r})V[m]\subseteq V[m+r] for all mm\in\mathbb{Z} and r+r\in\mathbb{Z}_{+}. In particular each graded piece V[m]V[m] is a module for 𝔤\mathfrak{g}. If dimV[m]<\dim V[m]<\infty for all mm\in\mathbb{Z} we define the graded character as the formal sum

chgrV=mch𝔤V[m]qm.\operatorname{ch}_{\operatorname{gr}}V=\sum_{m\in\mathbb{Z}}\operatorname{ch}_{\mathfrak{g}}V[m]q^{m}.

The module evaV(λ)\operatorname{ev}_{a}V(\lambda) for 𝔤[t]\mathfrak{g}[t] is graded if and only if a=0a=0. The 𝔤\mathfrak{g}-stable Demazure modules are also graded where the grading is given by the action of dd; namely Vw(λ)[r]={vVw(λ):dv=rv}.V_{w}(\lambda)[r]=\{v\in V_{w}(\lambda):dv=rv\}.

Given a \mathbb{Z}-graded vector space VV and an integer ss\in\mathbb{Z} let τsV\tau_{s}V be the grade shifted vector space obtained by declaring (τsV)[r]=V[rs].(\tau_{s}V)[r]=V[r-s].

2.3.5. The monoid 𝒫+\mathcal{P}^{+} and an alternate parametrization of finite-dimensional modules

Let 𝒫+\mathcal{P}^{+} be the free abelian multiplicative monoid generated by elements 𝝎i,a{\mbox{\boldmath$\omega$}}_{i,a}, iIi\in I, a×a\in\mathbb{C}^{\times} and let 𝒫\mathcal{P} be the corresponding group generated by these elements.

For λP+\lambda\in P^{+} let 𝝎λ,a=i=1n𝝎i,aλ(hi).{\mbox{\boldmath$\omega$}}_{\lambda,a}=\prod_{i=1}^{n}{\mbox{\boldmath$\omega$}}_{i,a}^{\lambda(h_{i})}. Clearly any element of 𝒫+\cal P^{+} can be written uniquely as a product 𝝎λ1,a1𝝎λk,ak{\mbox{\boldmath$\omega$}}_{\lambda_{1},a_{1}}\cdots{\mbox{\boldmath$\omega$}}_{\lambda_{k},a_{k}} for some multisubset {λ1,,λk}P+\{\lambda_{1},\dots,\lambda_{k}\}\subseteq P^{+} and distinct elements as×a_{s}\in\mathbb{C}^{\times}, 1sk1\leq s\leq k. Then part (i) of Proposition 2.3.4 can be reformulated as follows.

Proposition.

There exists a bijective correspondence between 𝒫+\mathcal{P}^{+} and isomorphism classes of finite-dimensional irreducible representations of L(𝔤)L(\mathfrak{g}) given by

𝝎=𝝎λ1,a1𝝎λk,ak[V(𝝎)]=[eva1V(λ1)evakV(λk)].{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}_{\lambda_{1},a_{1}}\cdots{\mbox{\boldmath$\omega$}}_{\lambda_{k},a_{k}}\longrightarrow[V({\mbox{\boldmath$\omega$}})]=[\operatorname{ev}_{a_{1}}V(\lambda_{1})\otimes\cdots\otimes~{}\operatorname{ev}_{a_{k}}V(\lambda_{k})].

Moreover if 𝝎,𝝎𝒫+{\mbox{\boldmath$\omega$}},{\mbox{\boldmath$\omega$}}^{\prime}\in\cal P^{+} then V(𝝎)V(𝝎)V({\mbox{\boldmath$\omega$}})\otimes V({\mbox{\boldmath$\omega$}}^{\prime}) is completely reducible and has V(𝝎𝝎)V({\mbox{\boldmath$\omega$}}{\mbox{\boldmath$\omega$}}^{\prime}) as a summand. ∎

2.3.6. Annihilating ideals

Suppose that VV is a finite-dimensional representation of L(𝔤)L(\mathfrak{g}). Then the discussion in Section 2.2.4 shows that there exists f[t,t1]f\in\mathbb{C}[t,t^{-1}] such that

{aL(𝔤):av=0forallvV}=𝔤f[t,t1].\{a\in L(\mathfrak{g}):av=0\ \ {\rm{for\ all}}\ \ v\in V\}=\mathfrak{g}\otimes f\mathbb{C}[t,t^{-1}].

In particular VV becomes a module for the truncated Lie algebra 𝔤[t,t1]/(f)\mathfrak{g}\otimes\mathbb{C}[t,t^{-1}]/(f). In the case when VV is irreducible the discussion so far proves that f=(ta1)(tak)f=(t-a_{1})\cdots(t-a_{k}) for some distinct element a1,,ak×a_{1},\dots,a_{k}\in\mathbb{C}^{\times}.

Suppose that V1V_{1} and V2V_{2} are modules for the truncation of L(𝔤)L(\mathfrak{g}) at f1f_{1} and f2f_{2} respectively and let ff be the least common multiple of the pair. Then V1V2V_{1}\otimes V_{2} is a module for the truncation at ff.

3. The quantized simple and affine enveloping algebras

3.1. Definitions and the Hopf algebra structure

3.1.1. The Drinfeld-Jimbo presentation

Let qq be an indeterminate and set qi=qdiq_{i}=q^{d_{i}}, iIi\in I, and q0=qq_{0}=q. The quantized enveloping algebra 𝕌q(𝔤^)\mathbb{U}_{q}(\widehat{\mathfrak{g}}) (also called the quantum affine algebra) is the associative algebra over (q)\mathbb{C}(q) generated by elements Xi±,Ki±1X_{i}^{\pm},K_{i}^{\pm 1}, D±1D^{\pm 1}, iI^i\in\widehat{I} and relations:

KiKi1=1,DD1=1,KiKj=KjKi,DKi=KiD,i,jI^\displaystyle K_{i}K_{i}^{-1}=1,\ \ \ DD^{-1}=1,\ \ \ K_{i}K_{j}=K_{j}K_{i},\ \ \ DK_{i}=K_{i}D,\ \ \ i,j\in\widehat{I}
KiXj±Ki1=qi±aijXj±,DXj±D1=q±δ0,jXj±,[Xi+,Xj]=δi,jKiKi1qiqi1,i,jI^,\displaystyle K_{i}X_{j}^{\pm}K_{i}^{-1}=q_{i}^{\pm a_{ij}}X_{j}^{\pm},\ \ DX_{j}^{\pm}D^{-1}=q^{\pm\delta_{0,j}}X_{j}^{\pm},\ \ \ \ [X_{i}^{+},X_{j}^{-}]=\delta_{i,j}\frac{K_{i}-K_{i}^{-1}}{q_{i}-q_{i}^{-1}},\ \ \ i,j\in\widehat{I},
m=01ai,j(1)m[1ai,jm]qi(Xi±)1ai,jmXj±(Xi±)m=0,i,jI^,ij.\displaystyle\sum_{m=0}^{1-a_{i,j}}(-1)^{m}{\genfrac{[}{]}{0.0pt}{0}{1-a_{i,j}}{m}}_{q_{i}}(X_{i}^{\pm})^{1-a_{i,j}-m}X_{j}^{\pm}(X_{i}^{\pm})^{m}=0,\ \ \ i,j\in\widehat{I},\ i\neq j.

The Hopf structure on this algebra is given by

Δ(D)=DD,Δ(Ki)=KiKi,\displaystyle\Delta(D)=D\otimes D,\ \ \ \Delta(K_{i})=K_{i}\otimes K_{i},
Δ(Xi+)=Xi+Ki+1Xi+,Δ(Xi)=Xi1+Ki1Xi,\displaystyle\Delta(X_{i}^{+})=X_{i}^{+}\otimes K_{i}+1\otimes X_{i}^{+},\ \ \ \Delta(X_{i}^{-})=X_{i}^{-}\otimes 1+K_{i}^{-1}\otimes X_{i}^{-},
S(Ki)=Ki1,S(D)=D1,S(Xi+)=Xi+Ki1,S(Xi)=KiXi,\displaystyle S(K_{i})=K_{i}^{-1},\ \ \ S(D)=D^{-1},\ \ \ S(X_{i}^{+})=-X_{i}^{+}K_{i}^{-1},\ \ \ S(X_{i}^{-})=-K_{i}X_{i}^{-},
ϵ(Ki)=1=ϵ(D),ϵ(Xi±)=0.\displaystyle\epsilon(K_{i})=1=\epsilon(D),\ \ \ \epsilon(X_{i}^{\pm})=0.

Set

C=K0Kθ,Kθ=iIKici,C=K_{0}K_{\theta},\ \ K_{\theta}=\prod_{i\in I}K_{i}^{c_{i}},

where cic_{i} are such that hθ=iIcihih_{\theta}=\sum_{i\in I}c_{i}h_{i}. The quantized enveloping algebra 𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g}) is the Hopf subalgebra generated by the elements Xi±,Ki±1X_{i}^{\pm},K_{i}^{\pm 1}, iIi\in I.

3.1.2. The Drinfeld presentation of 𝕌q(𝔤^)\mathbb{U}_{q}(\widehat{\mathfrak{g}})

An alternate presentation of the quantum affine algebra was given by Drinfeld.

The algebra 𝕌q(𝔤^)\mathbb{U}_{q}(\widehat{\mathfrak{g}}) is isomorphic to the (q)\mathbb{C}(q)-associative algebra with unit given by generators c±1/2c^{\pm 1/2}, xi,r±x_{i,r}^{\pm}, ki±1k_{i}^{\pm 1}, d±1d^{\pm 1} hi,sh_{i,s}, for iIi\in I, r,sr,s\in\mathbb{Z} with s0s\neq 0 subject to the following relations:

c1/2c1/2=1=dd1=kiki1=ki1ki,c±1/2arecentral,\displaystyle c^{1/2}c^{-1/2}=1=dd^{-1}=k_{i}k_{i}^{-1}=k_{i}^{-1}k_{i},\ \ \ c^{\pm 1/2}\ {\rm are\ central},
kikj=kjki,kihj,r=hj,rki,dki=kid,dhi,rd1=qrhi,r\displaystyle k_{i}k_{j}=k_{j}k_{i},\ \ \ k_{i}h_{j,r}=\ h_{j,r}k_{i},\ \ \ dk_{i}=k_{i}d,\ \ dh_{i,r}d^{-1}=q^{r}h_{i,r}
kixj,r±ki1=qi±ai,jxj,r±,dxi,r±d1=qrxi,r±,\displaystyle k_{i}x_{j,r}^{\pm}k_{i}^{-1}=q_{i}^{{}\pm a_{i,j}}x_{j,r}^{{}\pm{}},\ \ \ \ dx_{i,r}^{\pm}d^{-1}=q^{r}x_{i,r}^{\pm},
[hi,r,hj,s]=δr,s1r[rai,j]qicrcrqjqj1,[hi,r,xj,s±]=±1r[rai,j]qic|r|/2xj,r+s±,\displaystyle[h_{i,r},h_{j,s}]=\delta_{r,-s}\frac{1}{r}[ra_{i,j}]_{q_{i}}\frac{c^{r}-c^{-r}}{q_{j}-q_{j}^{-1}},\ \ \ [h_{i,r},x_{j,s}^{{}\pm{}}]=\pm\frac{1}{r}[ra_{i,j}]_{q_{i}}c^{\mp|r|/2}x_{j,r+s}^{{}\pm{}},
xi,r+1±xj,s±qi±ai,jxj,s±xi,r+1±=qi±ai,jxi,r±xj,s+1±xj,s+1±xi,r±,\displaystyle x_{i,r+1}^{{}\pm{}}x_{j,s}^{{}\pm{}}-q_{i}^{{}\pm a_{i,j}}x_{j,s}^{{}\pm{}}x_{i,r+1}^{{}\pm{}}=\ q_{i}^{{}\pm a_{i,j}}x_{i,r}^{{}\pm{}}x_{j,s+1}^{{}\pm{}}-x_{j,s+1}^{{}\pm{}}x_{i,r}^{{}\pm{}},
[xi,r+,xj,s]=δi,jc(rs)/2ϕi,r+s+c(rs)/2ϕi,r+sqiqi1,\displaystyle[x_{i,r}^{+},x_{j,s}^{-}]=\delta_{i,j}\ \frac{c^{(r-s)/2}\phi_{i,r+s}^{+}-c^{-(r-s)/2}\phi_{i,r+s}^{-}}{q_{i}-q_{i}^{-1}},
σSmk=0m(1)k[mk]qixi,nσ(1)±xi,nσ(k)±xj,s±xi,nσ(k+1)±xi,nσ(m)±=0,if ij,\displaystyle\sum_{\sigma\in S_{m}}\sum_{k=0}^{m}(-1)^{k}\genfrac{[}{]}{0.0pt}{0}{m}{k}_{q_{i}}x_{i,n_{\sigma(1)}}^{\pm}\ldots x_{i,n_{\sigma(k)}}^{{}\pm{}}x_{j,s}^{{}\pm{}}x_{i,n_{\sigma(k+1)}}^{{}\pm{}}\ldots x_{i,n_{\sigma(m)}}^{{}\pm{}}=0,\ \ \text{if $i\neq j$},

for all sequences of integers n1,,nmn_{1},\ldots,n_{m}, where m=1ai,jm=1-a_{i,j}, i,jIi,j\in I, SmS_{m} is the symmetric group on mm letters, and the ϕi,r±\phi_{i,r}^{{}\pm{}} are determined by equating powers of uu in the formal power series

ϕi±(u)=r=0ϕi,±r±u±r=ki±1exp(±(qiqi1)r=1hi,±ru±r).\phi_{i}^{\pm}(u)=\sum_{r=0}^{\infty}\phi_{i,\pm r}^{\pm}u^{\pm r}=k_{i}^{\pm 1}\exp\left(\pm(q_{i}-q_{i}^{-1})\sum_{r=1}^{\infty}h_{i,\pm r}u^{\pm r}\right).

Note that ϕi,r±=0\phi_{i,\mp r}^{\pm}=0 for r>0r>0. The quantized loop algebra 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g})) is the algebra with generators xi,r±x_{i,r}^{\pm}, ki±1k_{i}^{\pm 1}, hi,sh_{i,s}, r,sr,s\in\mathbb{Z} with s0s\neq 0 and iIi\in I and the same relations as above where we replace c1/2c^{1/2} by 1. Moreover we have a canonical inclusion 𝕌q(𝔤)𝕌q(L(𝔤))\mathbb{U}_{q}(\mathfrak{g})\hookrightarrow\mathbb{U}_{q}(L(\mathfrak{g})) given by mapping Xi±xi,0±X_{i}^{\pm}\to x_{i,0}^{\pm}, ki±Ki±k^{\pm}_{i}\to K^{\pm}_{i} iIi\in I.

Explicit formulae for the Hopf algebra structure in terms of these generators are not known. However the following partial information on the coproduct is often enough for our purposes [37, 44].

Proposition.

Let

X±=iI,r(q)xi,r±,X±(i)=jI{i},r(q)xj,r±,iI.X^{\pm}=\sum_{i\in I,\ r\in\mathbb{Z}}\mathbb{C}(q)x_{i,r}^{\pm},\ \ \ X^{\pm}(i)=\sum_{j\in I\setminus\{i\},\ r\in\mathbb{Z}}\mathbb{C}(q)x_{j,r}^{\pm},\ \ \ i\in I.

Then

  • (i)

    Modulo 𝕌q(𝔤^)X𝕌q(𝔤^)(X+)2+𝕌q(𝔤^)X𝕌q(𝔤^)X+(i)\mathbb{U}_{q}(\widehat{\mathfrak{g}})X^{-}\otimes\mathbb{U}_{q}(\widehat{\mathfrak{g}})(X^{+})^{2}+\mathbb{U}_{q}(\widehat{\mathfrak{g}})X^{-}\otimes\mathbb{U}_{q}(\widehat{\mathfrak{g}})X^{+}(i), we have

    Δ(xi,k+)=xi,k+1+kixi,k++j=1kϕi,j+xi,kj+,k0,\Delta(x_{i,k}^{+})=x_{i,k}^{+}\otimes 1+k_{i}\otimes x_{i,k}^{+}+\sum_{j=1}^{k}\phi_{i,j}^{+}\otimes x_{i,k-j}^{+},\ \ \ k\geq 0,
    Δ(xi,k+)=xi,k+1+ki1xi,k++j=1k1ϕi,jxi,k+j+,k>0.\Delta(x_{i,-k}^{+})=x_{i,-k}^{+}\otimes 1+k_{i}^{-1}\otimes x_{i,-k}^{+}+\sum_{j=1}^{k-1}\phi_{i,-j}^{-}\otimes x_{i,-k+j}^{+},\ \ \ k>0.
  • (ii)

    Modulo 𝕌q(𝔤^)(X)2𝕌q(𝔤^)X++𝕌q(𝔤^)X𝕌qX+(i)\mathbb{U}_{q}(\widehat{\mathfrak{g}})(X^{-})^{2}\otimes\mathbb{U}_{q}(\widehat{\mathfrak{g}})X^{+}+\mathbb{U}_{q}(\widehat{\mathfrak{g}})X^{-}\otimes\mathbb{U}_{q}X^{+}(i), we have

    Δ(xi,k)=xi,kki+1xi,k+j=1k1xi,kjϕi,j+,k>0,\Delta(x_{i,k}^{-})=x_{i,k}^{-}\otimes k_{i}+1\otimes x_{i,k}^{-}+\sum_{j=1}^{k-1}x_{i,k-j}^{-}\otimes\phi_{i,j}^{+},\ \ \ k>0,
    Δ(xi,k)=xi,kki1+1xi,k+j=1kxi,k+jϕi,j,k0.\Delta(x_{i,-k}^{-})=x_{i,-k}^{-}\otimes k_{i}^{-1}+1\otimes x_{i,-k}^{-}+\sum_{j=1}^{k}x_{i,-k+j}^{-}\otimes\phi_{i,-j}^{-},\ \ \ k\geq 0.
  • (iii)

    Modulo 𝕌q(𝔤^)X𝕌q(𝔤^)X+\mathbb{U}_{q}(\widehat{\mathfrak{g}})X^{-}\otimes\mathbb{U}_{q}(\widehat{\mathfrak{g}})X^{+}, we have

    Δ(hi,k)=hi,k1+1hi,k,k{0}.\Delta(h_{i,k})=h_{i,k}\otimes 1+1\otimes h_{i,k},\ \ k\in\mathbb{Z}\setminus\{0\}.

3.2. Representations of quantum algebras

The classification of finite-dimensional and integrable representations of the quantum algebras is essentially the same as that of the corresponding Lie algebras, once we impose certain restrictions. Thus we shall only be interested in type 1 modules for these algebras; namely we require that the elements Ki±1K_{i}^{\pm 1} act semi-simply on the module with eigenvalues in qq^{\mathbb{Z}}. The character of such a representation VV is given by ch(V)=μdim(Vμ)eμ\operatorname{ch}(V)=\sum_{\mu}\dim(V_{\mu})e_{\mu} where Vμ={vV:Ki±1v=qi±μ(hi)v,iI}.V_{\mu}=\{v\in V:K_{i}^{\pm 1}v=q_{i}^{\pm\mu(h_{i})}v,\ i\in I\}.

3.2.1. The irreducible modules

Let 𝒫q+\mathcal{P}_{q}^{+} (resp. 𝒫q\mathcal{P}_{q}) be the free abelian monoid (resp. group) generated by elements 𝝎i,a{\mbox{\boldmath$\omega$}}_{i,a} with a(q)×a\in\mathbb{C}(q)^{\times}. Clearly we can regard 𝒫+\mathcal{P}^{+} as a submonoid of 𝒫q+\mathcal{P}_{q}^{+}. Define wt:𝒫q+P+{\rm{wt}}{:}\ \mathcal{P}_{q}^{+}\to P^{+} by extending the assignment wt𝝎i,a=ωi{\rm{wt}}{\ }{\mbox{\boldmath$\omega$}}_{i,a}=\omega_{i} to a morphism of monoids.

Part (i) of the next result was proved in [33] while part (ii) was proved in [83], [98].

Theorem.
  1. (i)

    There is a bijection 𝝎[Vq(𝝎)]{\mbox{\boldmath$\omega$}}\to[V_{q}({\mbox{\boldmath$\omega$}})] between elements of 𝒫q+\mathcal{P}_{q}^{+} and isomorphism classes of finite-dimensional irreducible representations of 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g})). Moreover if 𝝎,𝝎𝒫q+{\mbox{\boldmath$\omega$}},{\mbox{\boldmath$\omega$}}^{\prime}\in\mathcal{P}_{q}^{+} then Vq(𝝎𝝎)V_{q}({\mbox{\boldmath$\omega$}}{\mbox{\boldmath$\omega$}}^{\prime}) is a subquotient of Vq(𝝎)Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}})\otimes V_{q}({\mbox{\boldmath$\omega$}}^{\prime}).

  2. (ii)

    Given λP+\lambda\in P^{+} there exists a unique (up to isomorphism) finite-dimensional irreducible 𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g})-module Vq(λ)V_{q}(\lambda). Moreover chVq(λ)=chV(λ){\rm ch}V_{q}(\lambda)={\rm ch}V(\lambda) and for λ,μP+\lambda,\mu\in P^{+} we have that Vq(λ+μ)V_{q}(\lambda+\mu) is a summand of Vq(λ)Vq(μ)V_{q}(\lambda)\otimes V_{q}(\mu). An analogous statement holds for positive level integrable representations of 𝕌q(𝔤^)\mathbb{U}_{q}(\widehat{\mathfrak{g}}).

Remark.
  1. (1)

    It is worth emphasizing that in general chVq(𝝎)chV(𝝎){\rm{ch}}V_{q}({\mbox{\boldmath$\omega$}})\neq{\rm{ch}}V({\mbox{\boldmath$\omega$}}) for 𝝎𝒫+{\mbox{\boldmath$\omega$}}\in\mathcal{P}^{+}.

  2. (2)

    The elements 𝝎i,a𝒫𝓆+{\mbox{\boldmath$\omega$}}_{i,a}\in\cal P_{q}^{+} are called the fundamental weights and the associated representations are called fundamental representations.

3.2.2. The category 𝓆\cal F_{q}

We shall be interested primarily in the category 𝓆\cal F_{q} of finite-dimensional representations of 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g})). The Hopf algebra structure of 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g})) ensures that 𝓆\cal F_{q} is a monoidal tensor category. Roughly speaking this means that 𝓆\cal F_{q} is an abelian category which is closed under taking tensor products and duals. However, since the Hopf algebra is not co-commutative it is not true in general that the modules VWV\otimes W and WVW\otimes V are isomorphic. One has also to be careful to distinguish between left and right duals say VV^{*} and V{}^{*}V; in one case we have an inclusion of VV\mathbb{C}\hookrightarrow V\otimes V^{*} and in the other case a projection VV0{}^{*}V\otimes V\to\mathbb{C}\to 0. The tensor product defines a ring structure on the Grothendieck group of this category. A very interesting fact proved in [60] is that the Grothendieck ring is always commutative.

We shall say that an irreducible representation in 𝓆\cal F_{q} is prime if it cannot be written in a nontrivial way as a tensor product of two objects of 𝓆\cal F_{q}. It is trivially true that any irreducible representation is isomorphic to a tensor product of prime representations. It follows that to understand irreducible representations it is enough to understand the prime ones. However, outside 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2} the classification of prime objects seems to be a very hard and perhaps wild problem. We shall nevertheless, give various examples of families of prime representations in this section and the next, including some coming from the connection with cluster algebras.

Our next definition is entirely motivated by the connection with cluster algebras. Namely we shall say that an object VV of 𝓆\cal F_{q} is real if VrV^{\otimes r} is irreducible for all r1r\geq 1. As a consequence of the main result of [66] it is enough to require V2V^{\otimes 2} to be irreducible. Again, it is hard to characterize real representations. A well-known example of Leclerc shows in [81] that there are prime representations which are not real.

The notion of prime and real objects can obviously be defined for the finite-dimensional module category of any Hopf algebra, and in particular for irreducible finite-dimensional representations of 𝔤\mathfrak{g} and L(𝔤)L(\mathfrak{g}). For 𝔤\mathfrak{g} it is an exercise to prove that the representations V(λ)V(\lambda), λP+\lambda\in P^{+} are prime and not real if λ0\lambda\neq 0. In the case of L(𝔤)L(\mathfrak{g}) it then follows from Proposition 2.3.4 that the prime irreducible representations are precisely evaV(λ)\operatorname{ev}_{a}V(\lambda), λP+\lambda\in P^{+}, aa\in\mathbb{C}. Moreover, it also follows that these representations are never real if λ0\lambda\neq 0. So these notions are uninteresting in these examples.

As in the case of L(𝔤)L(\mathfrak{g}) the objects of 𝓆\cal F_{q} are not completely reducible. In fact these categories behave more like the category 𝒪\cal O for simple Lie algebras and some of these similarities are explored in this article.

3.2.3. Representations of quantum loop 𝔰𝔩2\mathfrak{sl}_{2}

In this case the study of the irreducible objects in 𝓆\cal F_{q} is well-understood and we briefly review the main results from [33]. Given r+r\in\mathbb{Z}_{+} and a(q)×a\in\mathbb{C}(q)^{\times} let

𝝎1,a,r=𝝎1,aqr1𝝎1,aqr3𝝎1,aqr+1.{\mbox{\boldmath$\omega$}}_{1,a,r}={\mbox{\boldmath$\omega$}}_{1,aq^{r-1}}{\mbox{\boldmath$\omega$}}_{1,aq^{r-3}}\cdots{\mbox{\boldmath$\omega$}}_{1,aq^{-r+1}}.

Note that 𝝎1,a,0{\mbox{\boldmath$\omega$}}_{1,a,0} is the unit element of the monoid 𝒫q+\mathcal{P}_{q}^{+} for all a(q)×a\in\mathbb{C}(q)^{\times}. Then

Vq(𝝎1,a,r)𝕌q(𝔰𝔩2)Vq(rω1).V_{q}({\mbox{\boldmath$\omega$}}_{1,a,r})\cong_{\mathbb{U}_{q}(\mathfrak{sl}_{2})}V_{q}(r\omega_{1}).

Moreover

Vq(𝝎1,a,r)Vq(𝝎1,b,s)Vq(𝝎1,a,r𝝎1,b,s)ab1{q±(r+s2p):0p<min{r,s}}.V_{q}({\mbox{\boldmath$\omega$}}_{1,a,r})\otimes V_{q}({\mbox{\boldmath$\omega$}}_{1,b,s})\cong V_{q}({\mbox{\boldmath$\omega$}}_{1,a,r}{\mbox{\boldmath$\omega$}}_{1,b,s})\iff ab^{-1}\notin\{q^{\pm(r+s-2p)}:0\leq p<\min\{r,s\}\}.

In particular, the modules Vq(𝝎1,a,r)V_{q}({\mbox{\boldmath$\omega$}}_{1,a,r}) are prime and real. If ab1=q±(r+s2p)ab^{-1}=q^{\pm(r+s-2p)} for some 0p<min{r,s}0\leq p<\min\{r,s\} then we have a non-split short exact sequence,

0V1Vq(𝝎1,a,r)Vq(𝝎1,b,s)V20,0\to V_{1}\to V_{q}({\mbox{\boldmath$\omega$}}_{1,a,r})\otimes V_{q}({\mbox{\boldmath$\omega$}}_{1,b,s})\to V_{2}\to 0,

where

V1Vq(𝝎1,aq(rp),p)Vq(𝝎1,bq(pr),r+sp),V2Vq(𝝎1,aqp1,rp1)Vq(𝝎1,aqp+1,sp1)V_{1}\cong V_{q}({\mbox{\boldmath$\omega$}}_{1,aq^{(r-p)},p})\otimes V_{q}({\mbox{\boldmath$\omega$}}_{1,bq^{(p-r)},r+s-p}),\ \ \ V_{2}\cong V_{q}({\mbox{\boldmath$\omega$}}_{1,aq^{-p-1},r-p-1})\otimes V_{q}({\mbox{\boldmath$\omega$}}_{1,aq^{p+1},s-p-1})

if ab1=q(r+s2p)ab^{-1}=q^{-(r+s-2p)} while if ab1=q(r+s2p)ab^{-1}=q^{(r+s-2p)} then

V1Vq(𝝎1,aqp+1,rp1)Vq(𝝎1,aqp1,sp1),V2Vq(𝝎1,aq(pr),p)Vq(𝝎1,bq(rp),r+sp).V_{1}\cong V_{q}({\mbox{\boldmath$\omega$}}_{1,aq^{p+1},r-p-1})\otimes V_{q}({\mbox{\boldmath$\omega$}}_{1,aq^{-p-1},s-p-1}),\ \ V_{2}\cong V_{q}({\mbox{\boldmath$\omega$}}_{1,aq^{(p-r)},p})\otimes V_{q}({\mbox{\boldmath$\omega$}}_{1,bq^{(r-p)},r+s-p}).

Any irreducible module in 𝓆\cal F_{q} is isomorphic to a tensor product of representations of the form Vq(𝝎1,a,r)V_{q}({\mbox{\boldmath$\omega$}}_{1,a,r}), r+r\in\mathbb{Z}_{+}, a(q)×a\in\mathbb{C}(q)^{\times}. More precisely, if 𝝎𝒫𝓆+{\mbox{\boldmath$\omega$}}\in\cal P_{q}^{+} then

V(𝝎)Vq(𝝎1,a1,r1)Vq(𝝎1,ak,rk),V({\mbox{\boldmath$\omega$}})\cong V_{q}({\mbox{\boldmath$\omega$}}_{1,a_{1},r_{1}})\otimes\cdots\otimes V_{q}({\mbox{\boldmath$\omega$}}_{1,a_{k},r_{k}}),

for a unique choice of k1k\geq 1 and pairs (as,rs)(a_{s},r_{s}), 1sk1\leq s\leq k satisfying asam1q±(rs+rm2p)a_{s}a_{m}^{-1}\neq q^{\pm(r_{s}+r_{m}-2p)} for any 0p<min{rs,rm}0\leq p<\min\{r_{s},r_{m}\}, 1smk1\leq s\neq m\leq k.

3.2.4. Local Weyl modules for quantum loop algebras

We identify the monoid 𝒫q+\mathcal{P}_{q}^{+} with the monoid consisting of II-tuples of polynomials (πi(u))iI(\pi_{i}(u))_{i\in I}, πi(u)(q)[u]\pi_{i}(u)\in\mathbb{C}(q)[u] πi(0)=1\pi_{i}(0)=1, via

𝝎i,a(1δi,jau)jI.{\mbox{\boldmath$\omega$}}_{i,a}\mapsto(1-\delta_{i,j}au)_{j\in I}.

For 𝝎𝒫q+{\mbox{\boldmath$\omega$}}\in\mathcal{P}_{q}^{+} the local Weyl module Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) is the 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g}))-module generated by an element v𝝎v_{\mbox{{\boldmath{\small$\omega$}}}} with relations

xi,r+v𝝎=0=(xi,0)degπi(u)+1v𝝎,ϕi,r±v𝝎=γi,r±v𝝎,rx_{i,r}^{+}v_{{\mbox{{\boldmath{\small$\omega$}}}}}=0=(x_{i,0}^{-})^{\deg\pi_{i}(u)+1}v_{\mbox{{\boldmath{\small$\omega$}}}},\ \ \ \phi_{i,r}^{\pm}v_{{\mbox{{\boldmath{\small$\omega$}}}}}=\gamma_{i,r}^{\pm}v_{\mbox{{\boldmath{\small$\omega$}}}},\ \ r\in\mathbb{Z}

where γi,r±(q)\gamma_{i,r}^{\pm}\in\mathbb{C}(q) are defined by

r=0γi,±r±u±r=qidegπiπi(qi1u)πi(qiu),𝝎=(πi(u))iI.\sum_{r=0}^{\infty}\gamma_{i,\pm r}^{\pm}u^{\pm r}=q_{i}^{\deg\pi_{i}}\frac{\pi_{i}(q_{i}^{-1}u)}{\pi_{i}(q_{i}u)},\ \ \ \ {\mbox{\boldmath$\omega$}}=(\pi_{i}(u))_{i\in I}.

The following was proved in [39].

Proposition.

Let 𝝎𝒫q+{\mbox{\boldmath$\omega$}}\in\mathcal{P}_{q}^{+}. Then dimWq(𝝎)<\dim W_{q}({\mbox{\boldmath$\omega$}})<\infty and Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) has a unique irreducible quotient Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}). If 𝝎𝒫q+{\mbox{\boldmath$\omega$}}^{\prime}\in\mathcal{P}_{q}^{+} then the modules Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) and Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}^{\prime}) are isomorphic if and only if 𝝎=𝝎{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}^{\prime}. The module Vq(𝝎𝝎)V_{q}({\mbox{\boldmath$\omega$}}{\mbox{\boldmath$\omega$}}^{\prime}) is a subquotient of Wq(𝝎)Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}})\otimes W_{q}({\mbox{\boldmath$\omega$}}^{\prime}). ∎

3.2.5. The fundamental local Weyl modules

It was proved in [22] that Wq(𝝎i,a)Vq(𝝎i,a)W_{q}({\mbox{\boldmath$\omega$}}_{i,a})\cong V_{q}({\mbox{\boldmath$\omega$}}_{i,a}). In general it is not true that local Weyl modules are irreducible but we will discuss conditions for these later in the paper.

3.2.6. Local Weyl modules and tensor products

Suppose that 𝝎,𝝎𝒫q+{\mbox{\boldmath$\omega$}},{\mbox{\boldmath$\omega$}}^{\prime}\in\mathcal{P}_{q}^{+} and let MM, MM^{\prime} be any quotient of Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) and Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}^{\prime}) respectively. Let v𝝎v_{\mbox{{\boldmath{\small$\omega$}}}} and v𝝎v_{{\mbox{{\boldmath{\small$\omega$}}}}^{\prime}} also denote the images of these elements in MM and MM^{\prime}. Then using the formulae for comultiplication one can prove that we have the following sequence of surjective maps:

Wq(𝝎𝝎)𝕌q(L(𝔤))(v𝝎v𝝎)Vq(𝝎𝝎).W_{q}({\mbox{\boldmath$\omega$}}{\mbox{\boldmath$\omega$}}^{\prime})\twoheadrightarrow\mathbb{U}_{q}(L(\mathfrak{g}))(v_{{\mbox{{\boldmath{\small$\omega$}}}}}\otimes v_{{\mbox{{\boldmath{\small$\omega$}}}}^{\prime}})\twoheadrightarrow V_{q}({\mbox{\boldmath$\omega$}}{\mbox{\boldmath$\omega$}}^{\prime}).

In particular Vq(𝝎𝝎)V_{q}({\mbox{\boldmath$\omega$}}{\mbox{\boldmath$\omega$}}^{\prime}) is a subquotient of MMM\otimes M^{\prime}.

Suppose that 𝝎=𝝎i1,a1𝝎ik,ak𝒫𝓆+{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}_{i_{1},a_{1}}\cdots{\mbox{\boldmath$\omega$}}_{i_{k},a_{k}}\in\cal P_{q}^{+}; the discussion so far establishes the existence of a map of 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g}))-modules

ϕ𝝎:Wq(𝝎)Vq(𝝎i1,a1)Vq(𝝎ik,ak).\phi_{{\mbox{\boldmath$\omega$}}}:W_{q}({\mbox{\boldmath$\omega$}})\to V_{q}({\mbox{\boldmath$\omega$}}_{i_{1},a_{1}})\otimes\cdots\otimes V_{q}({\mbox{\boldmath$\omega$}}_{i_{k},a_{k}}).

If we assume that a1,,aka_{1},\dots,a_{k} are such that aj/asqa_{j}/a_{s}\notin q^{\mathbb{N}} for all 1j<sk1\leq j<s\leq k, then the results of [1, 23, 100] show that this map is surjective. It was conjectured in [39] that ϕ𝝎\phi_{\mbox{\boldmath$\omega$}} is an isomorphism. Clearly to prove the conjecture it suffices to establish an equality of dimensions. This equality was proved in that paper for 𝔰𝔩2\mathfrak{sl}_{2}. In the general case the conjecture was established through the work of [30, 58, 88]. We will discuss this further in the next section.

Assuming from now on that ϕ𝝎\phi_{\mbox{\boldmath$\omega$}} is an isomorphism we study the question of the irreducibility of Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}). A sufficient condition for Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) to be irreducible is to require that aj/asqa_{j}/a_{s}\notin q^{\mathbb{Z}} for all 1jsk1\leq j\neq s\leq k; this was known through the work of [1, 23]. A precise statement was given in [23, Corollary 5.1] when 𝔤\mathfrak{g} is of classical type and for some of the exceptional nodes. The following result summarizes the results of [23] for classical cases.

Theorem.
  1. (i)

    Suppose that 𝔤\mathfrak{g} is of classical type and i,jIi,j\in I and a,b(q)×a,b\in\mathbb{C}(q)^{\times}. Then

    ab1q±S(i,j)Vq(𝝎i,a𝝎j,b)Vq(𝝎i,a)Vq(𝝎j,b)Wq(𝝎i,a𝝎j,b),ab^{-1}\notin q^{\pm S(i,j)}\implies V_{q}({\mbox{\boldmath$\omega$}}_{i,a}{\mbox{\boldmath$\omega$}}_{j,b})\cong V_{q}({\mbox{\boldmath$\omega$}}_{i,a})\otimes V_{q}({\mbox{\boldmath$\omega$}}_{j,b})\cong W_{q}({\mbox{\boldmath$\omega$}}_{i,a}{\mbox{\boldmath$\omega$}}_{j,b}),

    where S(i,j)S(i,j) is given as follows:

    If 𝔤\mathfrak{g} is of type AnA_{n}:

    S(i,j)={2+2kij:max{i,j}kmin{i+j1,n}}.S(i,j)=\{2+2k-i-j:\max\{i,j\}\leq k\leq\min\{i+j-1,n\}\}.

    If 𝔤\mathfrak{g} is of type BnB_{n}, and α1\alpha_{1} is short:

    • S(1,1)={4k2:1kn}S(1,1)=\{4k-2:1\leq k\leq n\},

    • S(i,1)=S(i,1)={4k2i+1:ikn}S(i,1)=S(i,1)=\{4k-2i+1:\ i\leq k\leq n\},

    • S(i,j)={4+4k2i2j:max{i,j}kn}{4k22|ji|:max{i,j}kn}S(i,j)=\{4+4k-2i-2j:\ \max\{i,j\}\leq k\leq n\}\cup\{4k-2-2|j-i|:\ \max\{i,j\}\leq k\leq n\}, i,j>1i,j>1.

    If 𝔤\mathfrak{g} is of type CnC_{n}, and α1\alpha_{1} is long:

    • S(1,1)={2k+2: 1kn}S(1,1)=\{2k+2:\ 1\leq k\leq n\}

    • S(i,1)=S(1,i)={2ki+3:1kn}S(i,1)=S(1,i)=\{2k-i+3:1\leq k\leq n\}, i>1i>1,

    • S(i,j)={2+2kij:max{i,j}kn}{2+2k|ij|:max{i,j}kn}S(i,j)=\{2+2k-i-j:\ \max\{i,j\}\leq k\leq n\}\cup\{2+2k-|i-j|:\ \max\{i,j\}\leq k\leq n\}, i,j>1i,j>1.

    If 𝔤\mathfrak{g} is of type DnD_{n}, and 11 and 22 denote the spin nodes:

    • S(1,1)=S(2,2)={2k2:2kn,k0mod2}S(1,1)=S(2,2)=\{2k-2:2\leq k\leq n,\ \ k\equiv 0\mod 2\}

    • S(1,2)=S(2,1)={2k2:3kn,k1mod2}S(1,2)=S(2,1)=\{2k-2:3\leq k\leq n,\ \ k\equiv 1\mod 2\}

    • S(1,j)=S(2,j)={2kj:jkn}=S(j,2)=S(j,1)S(1,j)=S(2,j)=\{2k-j:j\leq k\leq n\}=S(j,2)=S(j,1),   j3j\geq 3.

    • S(i,j)={2+2kij:max{i,j}kn}{2+2k|ij|:max{i,j}kn}S(i,j)=\{2+2k-i-j:\ \max\{i,j\}\leq k\leq n\}\cup\{-2+2k-|i-j|:\ \max\{i,j\}\leq k\leq n\}, i,j3i,j\geq 3.

  2. (ii)

    Given 𝝎=𝝎i1,a1𝝎ik,ak{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}_{i_{1},a_{1}}\cdots{\mbox{\boldmath$\omega$}}_{i_{k},a_{k}} the module Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) is irreducible if

    aras1q±S(ir,is), 1r<sk.a_{r}a_{s}^{-1}\notin q^{\pm S(i_{r},i_{s})},\ \ 1\leq r<s\leq k.

Remark.

More recently an alternative approach to describing the set S(i,j)S(i,j) was given in [62, Theorem 2.10] and [63, Section 6] by relating it to the poles of the universal RR-matrix. It is nontrivial to see that those conditions are equivalent to the explicit description given in the preceding theorem.

3.2.7. 𝔸\mathbb{A}-forms and classical limits

Let 𝔸=[q,q1]\mathbb{A}=\mathbb{Z}[q,q^{-1}] and define 𝕌𝔸(𝔤^)\mathbb{U}_{\mathbb{A}}(\widehat{\mathfrak{g}}) to be the 𝔸\mathbb{A}-subalgebra of 𝕌q(𝔤^)\mathbb{U}_{q}(\widehat{\mathfrak{g}}) generated by the elements (Xi±)r/[r]q!(X_{i}^{\pm})^{r}/[r]_{q}!, iI^i\in\widehat{I}. Then

𝕌q(𝔤^)𝕌𝔸(𝔤^)𝔸(q).\mathbb{U}_{q}(\widehat{\mathfrak{g}})\cong\mathbb{U}_{\mathbb{A}}(\widehat{\mathfrak{g}})\otimes_{\mathbb{A}}\mathbb{C}(q).

For ϵ×\epsilon\in\mathbb{C}^{\times} we let ϵ\mathbb{C}_{\epsilon} be the 𝔸\mathbb{A}-module obtained by letting qq act as ϵ\epsilon and set

𝕌ϵ(𝔤^)=𝕌𝔸(𝔤^)𝔸ϵ.\mathbb{U}_{\epsilon}(\widehat{\mathfrak{g}})=\mathbb{U}_{\mathbb{A}}(\widehat{\mathfrak{g}})\otimes_{\mathbb{A}}\mathbb{C}_{\epsilon}.

The algebra 𝕌(𝔤^)\mathbb{U}(\widehat{\mathfrak{g}}) is isomorphic to the quotient of 𝕌1(𝔤^)\mathbb{U}_{1}(\widehat{\mathfrak{g}}) by the ideal generated by Ki1,D1,iI^K_{i}-1,D-1,\ i\in\widehat{I}. Similar assertions hold for 𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g}) and 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g})) as well. Part (i) of the following was proved in [83], [98] while part (ii) was proved in [39].

Theorem.
  1. (i)

    Suppose that λP^+\lambda\in\widehat{P}^{+} and ϵ×\epsilon\in\mathbb{C}^{\times}. There exists a 𝕌𝔸(𝔤^)\mathbb{U}_{\mathbb{A}}(\widehat{\mathfrak{g}})-submodule V𝔸(λ)V_{\mathbb{A}}(\lambda) of Vq(λ)V_{q}(\lambda) such that

    Vq(λ)V𝔸(λ)𝔸(q).V_{q}(\lambda)\cong V_{\mathbb{A}}(\lambda)\otimes_{\mathbb{A}}\mathbb{C}(q).

    In particular Vϵ(λ)=V𝔸(λ)𝔸ϵV_{\epsilon}(\lambda)=V_{\mathbb{A}}(\lambda)\otimes_{\mathbb{A}}\mathbb{C}_{\epsilon} is a module for 𝕌ϵ(𝔤^)\mathbb{U}_{\epsilon}(\widehat{\mathfrak{g}}), and if ϵ=1\epsilon=1 or if ϵ\epsilon is not a root of unity then we have chVϵ(λ)=chVq(λ).{\rm{ch}}V_{\epsilon}(\lambda)={\rm{ch}}V_{q}(\lambda). Analogous statements hold for the 𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g}) representations Vq(λ)V_{q}(\lambda), λP+\lambda\in P^{+}.

  2. (ii)

    Let 𝒫𝔸+\mathcal{P}_{\mathbb{A}}^{+} be the submonoid of 𝒫+\mathcal{P}^{+} generated by elements 𝝎i,a{\mbox{\boldmath$\omega$}}_{i,a}, a𝔸a\in\mathbb{A} and let 𝝎𝒫𝔸+{\mbox{\boldmath$\omega$}}\in\mathcal{P}_{\mathbb{A}}^{+}. Then Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) admits a 𝕌𝔸(L(𝔤))\mathbb{U}_{\mathbb{A}}(L(\mathfrak{g}))-submodule W𝔸(𝝎)W_{\mathbb{A}}({\mbox{\boldmath$\omega$}}) and

    Wq(𝝎)W𝔸(𝝎)𝔸(q).W_{q}({\mbox{\boldmath$\omega$}})\cong W_{\mathbb{A}}({\mbox{\boldmath$\omega$}})\otimes_{\mathbb{A}}\mathbb{C}(q).

    In particular Wϵ(𝝎)=W𝔸(𝝎)𝔸ϵW_{\epsilon}({\mbox{\boldmath$\omega$}})=W_{\mathbb{A}}({\mbox{\boldmath$\omega$}})\otimes_{\mathbb{A}}\mathbb{C}_{\epsilon} is a 𝕌ϵ(L(𝔤))\mathbb{U}_{\epsilon}(L(\mathfrak{g}))-module, and chWϵ(𝝎)=Wq(𝝎){\rm{ch}}W_{\epsilon}({\mbox{\boldmath$\omega$}})=W_{q}({\mbox{\boldmath$\omega$}}) if ϵ=1\epsilon=1 or if ϵ\epsilon is not a root of unity. If MqM_{q} is any 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g}))-module quotient of Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) let M𝔸M_{\mathbb{A}} be the image of W𝔸(𝝎)W_{\mathbb{A}}({\mbox{\boldmath$\omega$}}). Then,

    MqM𝔸𝔸(q),MϵM𝔸𝔸ϵ,M_{q}\cong M_{\mathbb{A}}\otimes_{\mathbb{A}}\mathbb{C}(q),\ \ M_{\epsilon}\cong M_{\mathbb{A}}\otimes_{\mathbb{A}}\mathbb{C}_{\epsilon},

    and MϵM_{\epsilon} is a canonical quotient of Wϵ(𝝎)W_{\epsilon}({\mbox{\boldmath$\omega$}}).

Remark.

  • The modules V1(λ)V_{1}(\lambda), λP^+\lambda\in\widehat{P}^{+} and V1(𝝎)V_{1}({\mbox{\boldmath$\omega$}}), 𝝎𝒫𝔸+{\mbox{\boldmath$\omega$}}\in\mathcal{P}^{+}_{\mathbb{A}} are modules for the universal enveloping algebra of 𝕌(L(𝔤))\mathbb{U}(L(\mathfrak{g})) and are called the classical limit of Vq(λ)V_{q}(\lambda) and Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}) respectively.

  • It is worth noting again, that in part (i) of the theorem the module Vϵ(λ)V_{\epsilon}(\lambda) is irreducible for 𝕌ϵ(𝔤^)\mathbb{U}_{\epsilon}(\widehat{\mathfrak{g}}) if ϵ\epsilon is not a root of unity. This is false in part (ii).

4. Classical and Graded limits

In this section we discuss various well-studied families of finite-dimensional representations of quantum affine algebras. We review the literature on the presentation of these modules, their classical limits and the closely related graded limits.

4.1. Classical and Graded Limits of the Quantum Local Weyl modules

4.1.1. Relations in the classical limit

Suppose that 𝝎𝒫𝔸+{\mbox{\boldmath$\omega$}}\in\mathcal{P}^{+}_{\mathbb{A}} and let Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}) be the unique irreducible quotient of Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) (see Section 3.2.4). Since Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) is finite-dimensional the corresponding classical limits (see Theorem 3.2.7) W1(𝝎)W_{1}({\mbox{\boldmath$\omega$}}) and V1(𝝎)V_{1}({\mbox{\boldmath$\omega$}}) are finite-dimensional modules for L(𝔤)L(\mathfrak{g}). Let v¯𝝎=v𝝎1W1(𝝎){\bar{v}_{{\mbox{\boldmath$\omega$}}}}=v_{\mbox{\boldmath$\omega$}}\otimes 1\in W_{1}({\mbox{\boldmath$\omega$}}). The following was proved in [39].

Lemma.

Suppose that 𝝎=𝝎i1,a1𝝎ik,ak𝒫𝔸+{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}_{i_{1},a_{1}}\cdots{\mbox{\boldmath$\omega$}}_{i_{k},a_{k}}\in\cal P^{+}_{\mathbb{A}}. The following relations hold in W1(𝝎)W_{1}({\mbox{\boldmath$\omega$}}):

hv¯𝝎=wt𝝎(h)v¯𝝎,(h(ta1(1))(tak(1))[t,t1])v¯𝝎=0,h\bar{v}_{\mbox{\boldmath$\omega$}}={\rm{wt}}{\ }{\mbox{\boldmath$\omega$}}(h)\bar{v}_{\mbox{\boldmath$\omega$}},\ \ \left(h\otimes(t-a_{1}(1))\cdots(t-a_{k}(1))\mathbb{C}[t,t^{-1}]\right){\bar{v}_{{\mbox{\boldmath$\omega$}}}}=0,
(xα+[t,t1])v¯𝝎=0,h𝔥,αR+.(x_{\alpha}^{+}\otimes\mathbb{C}[t,t^{-1}]){\bar{v}_{{\mbox{\boldmath$\omega$}}}}=0,\ \ h\in\mathfrak{h},\ \ \alpha\in R^{+}.

Remark.

With a little more work one can actually prove that there exists an integer NN\in\mathbb{N} such that W1(𝝎)W_{1}({\mbox{\boldmath$\omega$}}) is a module for the truncation of L(𝔤)L(\mathfrak{g}) by the polynomial (ta1(1))N(tak(1))N(t-a_{1}(1))^{N}\cdots(t-a_{k}(1))^{N}.

4.1.2. Graded Limits: the modules VlocV_{\operatorname{loc}}

In the rest of this section we shall restrict our attention to the submonoid 𝒫+\cal P^{+}_{\mathbb{Z}} which is generated by the elements 𝛚i,qr{\mbox{\boldmath$\omega$}}_{i,q^{r}} for iIi\in I and rr\in\mathbb{Z}.

In this case the results of Section 4.1.1 imply that W1(𝝎)W_{1}({\mbox{\boldmath$\omega$}}) is a module for the truncation of L(𝔤)L(\mathfrak{g}) at (t1)N(t-1)^{N} for some NN sufficiently large. Using the isomorphism of Lie algebras

𝔤[t,t1](t1)N𝔤[t](t1)N𝔤[t](tN)\mathfrak{g}\otimes\frac{\mathbb{C}[t,t^{-1}]}{(t-1)^{N}}\cong\mathfrak{g}\otimes\frac{\mathbb{C}[t]}{(t-1)^{N}}\cong\mathfrak{g}\otimes\frac{\mathbb{C}[t]}{(t^{N})}

we see that we can regard W1(𝝎)W_{1}({\mbox{\boldmath$\omega$}}) as a module for the truncation of 𝔤[t]\mathfrak{g}[t] at tNt^{N}. We shall denote this module by Wloc(𝝎)W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}). We call this the graded limit of Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}). In fact if MM is any quotient of Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) we can define a corresponding module MlocM_{\operatorname{loc}} for 𝔤[t]\mathfrak{g}[t] using the isomorphisms of Lie algebras and call this the graded limit of MM.

This terminology of course requires justification. Recall that the action of dd on 𝔤[t]\mathfrak{g}[t] defines a +\mathbb{Z}_{+}-grading on it: the rr-th graded piece is 𝔤tr\mathfrak{g}\otimes t^{r}. The adjoint action of dd on 𝕌(𝔤[t])\mathbb{U}(\mathfrak{g}[t]) also gives a +\mathbb{Z}_{+}-grading. Hence one can define the notion of a graded 𝔤[t]\mathfrak{g}[t]-module VV to be one which admits a compatible \mathbb{Z}-grading namely:

V=sV[s],(𝔤tr)V[s]V[r+s].V=\bigoplus_{s\in\mathbb{Z}}V[s],\ \ (\mathfrak{g}\otimes t^{r})V[s]\subseteq V[r+s].

The general belief is that when MM is a quotient of Wq(𝝎)W_{q}({\mbox{\boldmath$\omega$}}) with 𝝎𝒫+{\mbox{\boldmath$\omega$}}\in\cal P^{+}_{\mathbb{Z}} then MlocM_{\operatorname{loc}} is a graded 𝔤[t]\mathfrak{g}[t]-module. This is far from clear in general and is hard to prove even in specific cases. In the rest of the section we will discuss certain families of modules where the corresponding graded limit is in fact a graded 𝔤[t]\mathfrak{g}[t]-module.

Remark.

In the discussion that follows we shall see that the classical or graded limit depends only on wt𝝎{\rm{wt}}{~{}}{\mbox{\boldmath$\omega$}}. So there is a substantial loss of information when we go to the limits. However, the character and the underlying 𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g})-module is the same and this is one reason for our interest in this study.

4.1.3. Kirillov–Reshetikhin modules

We begin by discussing this particular family of modules since this was essentially the motivation for the interest in graded limits.

Given iIi\in I, rr\in\mathbb{N} and ss\in\mathbb{Z} set

𝝎i,s,r=𝝎i,qis+r1𝝎i,qis+r3𝝎i,qisr+1.{\mbox{\boldmath$\omega$}}_{i,s,r}={\mbox{\boldmath$\omega$}}_{i,q_{i}^{s+r-1}}{\mbox{\boldmath$\omega$}}_{i,q_{i}^{s+r-3}}\cdots{\mbox{\boldmath$\omega$}}_{i,q_{i}^{s-r+1}}.

Notice that in the case i=1i=1 these elements were introduced in Section 3.2.3 in the case of 𝔰𝔩2\mathfrak{sl}_{2} where they were denoted as 𝝎1,qs,r{\mbox{\boldmath$\omega$}}_{1,q^{s},r} since we were working in a more general situation. The corresponding irreducible 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g}))-module is called a Kirillov–Reshetikhin module. This is because of an important conjecture that they had made; they predicted the existence of certain modules for the quantum loop algebra with a specific decomposition as 𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g})-modules (see also [64]). In [22] it was proved that the conjectured modules were of the form Vq(𝝎i,s,r)V_{q}({\mbox{\boldmath$\omega$}}_{i,s,r}) for all classical Lie algebras and for some iIi\in I in the exceptional cases. Moreover the following presentation was given (see Corollary 2.1 of [22]) for the module V1(𝝎i,s,r)V_{1}({\mbox{\boldmath$\omega$}}_{i,s,r}).

Theorem.

The L(𝔤)L(\mathfrak{g})-module V1(𝝎i,s,r)V_{1}({\mbox{\boldmath$\omega$}}_{i,s,r}) is generated by an element vi,s,rv_{i,s,r} with relations:

xα+vi,s,r=0,(htk)vi,s,r=rωi(h)vi,s,r,((xitk)xi1)vi,s,r=0,(xα1)rωi(hα)+1vi,s,r=0,x_{\alpha}^{+}v_{i,s,r}=0,\ \ (h\otimes t^{k})v_{i,s,r}=r\omega_{i}(h)v_{i,s,r},\ \ ((x_{i}^{-}\otimes t^{k})-x_{i}^{-}\otimes 1)v_{i,s,r}=0,\ \ (x_{\alpha}^{-}\otimes 1)^{r\omega_{i}(h_{\alpha})+1}v_{i,s,r}=0,

where αR+\alpha\in R^{+}, h𝔥h\in\mathfrak{h} and kk\in\mathbb{Z}. ∎

Here we have used the fact that wt𝝎i,s,r=rωi{\rm{wt}}{~{}}{\mbox{\boldmath$\omega$}}_{i,s,r}=r\omega_{i}. Notice that these relations are independent of ss. Moreover,

((htk)h)vi,s,r=0(h(t1)k)vi,s,r=0, 0k,((h\otimes t^{k})-h)v_{i,s,r}=0\implies(h\otimes(t-1)^{k}){v}_{i,s,r}=0,\ \ \ 0\neq k\in\mathbb{Z},

and similarly for the third relation in the presentation above. It follows that Vloc(𝝎i,s,r)V_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}_{i,s,r}) is the 𝔤[t]\mathfrak{g}[t]-quotient of Wloc(𝝎i,s,r)W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}_{i,s,r}) by imposing the additional relation: (xit)v𝝎i,s,r=0.(x_{i}^{-}\otimes t)v_{{\mbox{{\boldmath{\small$\omega$}}}}_{i,s,r}}=0.

Later, in [31], a more systematic self contained study of these modules was developed and the graded 𝔤\mathfrak{g}-module decomposition of these modules was calculated. One can think of this as a graded version of the Kirillov–Reshetikhin character formula. The results of [31] led to the definition of graded limits and more generally resulted in the development of the subject of graded (not necessarily finite-dimensional) representations of 𝔤[t]\mathfrak{g}[t]. We say more about this study in later sections of the paper.

4.1.4. Minimal Affinizations

The notion of minimal affinizations was introduced and further studied in [21, 34, 35]. Perhaps the simplest place to explain what this notion means is in the case of 𝔰𝔩2\mathfrak{sl}_{2}. Since we have only one simple root we denote the generators of 𝒫𝓆+\cal P_{q}^{+} by 𝝎1,a{\mbox{\boldmath$\omega$}}_{1,a}. If 𝝎=𝝎1,a1𝝎1,ak𝒫𝓆+{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}_{1,a_{1}}\cdots{\mbox{\boldmath$\omega$}}_{1,a_{k}}\in\cal P_{q}^{+} then it is not hard to see that there exists a 𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g})-module MM such that

Vq(𝝎)𝕌q(𝔤)Vq(kω)M.V_{q}({\mbox{\boldmath$\omega$}})\cong_{\mathbb{U}_{q}(\mathfrak{g})}V_{q}(k\omega)\oplus M.

Moreover, it was shown in [33] that M0M\neq 0 unless Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}) is a Kirillov–Reshetikhin module:

M=0𝝎=𝝎1,aqk1𝝎1,aqk+1,a(q)×.M=0\iff{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}_{1,aq^{k-1}}\cdots{\mbox{\boldmath$\omega$}}_{1,aq^{-k+1}},\ \ a\in\mathbb{C}(q)^{\times}.

Using the results of [33] we can also give precise conditions under which Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}) and Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}^{\prime}) are isomorphic as 𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g})-modules.

It is natural to ask what analogs of these results hold in the higher rank case. It was known essentially from the beginning (see [45]) that if 𝔤\mathfrak{g} is not of type AA, there does not exist a corresponding 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g}))-module structure on Vq(λ)V_{q}(\lambda). On the other hand it is also clear that there were many pairs 𝝎,𝝎𝒫𝓆+{\mbox{\boldmath$\omega$}},{\mbox{\boldmath$\omega$}}^{\prime}\in\cal P_{q}^{+} with Vq(𝝎)𝕌q(𝔤)Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}})\cong_{\mathbb{U}_{q}(\mathfrak{g})}V_{q}({\mbox{\boldmath$\omega$}}^{\prime}). So this motivated the question: given λP+\lambda\in P^{+}, is there a “smallest”  𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g})-module containing a copy of Vq(λ)V_{q}(\lambda) which admits an action of the quantum loop algebra. This question can be more formally stated as follows.

Given 𝝎,𝝎𝒫𝓆+{\mbox{\boldmath$\omega$}},{\mbox{\boldmath$\omega$}}^{\prime}\in\cal P_{q}^{+} we say that Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}) is equivalent to Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}^{\prime}) if they are isomorphic as 𝕌q(𝔤)\mathbb{U}_{q}(\mathfrak{g})-modules. Denote the equivalence class corresponding to 𝝎\omega by [Vq(𝝎)]𝔤[V_{q}({\mbox{\boldmath$\omega$}})]_{\mathfrak{g}}. In particular,

[Vq(𝝎)]𝔤=[Vq(𝝎)]𝔤wt𝝎=wt𝝎.[V_{q}({\mbox{\boldmath$\omega$}})]_{\mathfrak{g}}=[V_{q}({\mbox{\boldmath$\omega$}}^{\prime})]_{\mathfrak{g}}\implies{\rm{wt}}{~{}}{\mbox{\boldmath$\omega$}}={\rm{wt}}{~{}}{\mbox{\boldmath$\omega$}}^{\prime}.

The converse statement is definitely false, this is already the case in 𝔰𝔩2\mathfrak{sl}_{2}.

Define a partial order on the set of equivalence classes by: [Vq(𝝎)]𝔤[Vq(𝝎)]𝔤[V_{q}({\mbox{\boldmath$\omega$}})]_{\mathfrak{g}}\leq[V_{q}({\mbox{\boldmath$\omega$}}^{\prime})]_{\mathfrak{g}} if for all μP+\mu\in P^{+} either

dimHom𝕌q(𝔤)(Vq(μ),Vq(𝝎))dimHom𝕌q(𝔤)(Vq(μ),Vq(𝝎))\dim\operatorname{Hom}_{\mathbb{U}_{q}(\mathfrak{g})}(V_{q}(\mu),V_{q}({\mbox{\boldmath$\omega$}}))\leq\dim\operatorname{Hom}_{\mathbb{U}_{q}(\mathfrak{g})}(V_{q}(\mu),V_{q}({\mbox{\boldmath$\omega$}}^{\prime}))

or there exists ν>μ\nu>\mu (i.e. νμQ+\{0}\nu-\mu\in Q^{+}\backslash\{0\}) such that

dimHom𝕌q(𝔤)(Vq(ν),Vq(𝝎))<dimHom𝕌q(𝔤)(Vq(ν),Vq(𝝎)).\dim\operatorname{Hom}_{\mathbb{U}_{q}(\mathfrak{g})}(V_{q}(\nu),V_{q}({\mbox{\boldmath$\omega$}}))<\dim\operatorname{Hom}_{\mathbb{U}_{q}(\mathfrak{g})}(V_{q}(\nu),V_{q}({\mbox{\boldmath$\omega$}}^{\prime})).

It was proved in [21] that minimal elements exist in this order and an irreducible representation corresponding to a minimal element was called a minimal affinization. When 𝔤\mathfrak{g} is not of type DD or EE the explicit expression for the elements 𝝎𝒫𝓆+{\mbox{\boldmath$\omega$}}\in\cal P_{q}^{+} which give a minimal affinization are given by

𝝎=𝝎i1,s1,r1𝝎ik,sk,rk,i1<i2<<ik,\displaystyle{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}_{i_{1},s_{1},r_{1}}\cdots{\mbox{\boldmath$\omega$}}_{i_{k},s_{k},r_{k}},\ \ i_{1}<i_{2}<\cdots<i_{k},
sp+1sp=ϵ(diprp+dip+1rp+1+j=ipip+11(dj1aj,j+1)), 1pk1,\displaystyle s_{p+1}-s_{p}=\epsilon\left(d_{i_{p}}r_{p}+d_{i_{p+1}}r_{p+1}+\sum_{j=i_{p}}^{i_{p+1}-1}(d_{j}-1-a_{j,j+1})\right),\ \ 1\leq p\leq k-1,

where either ϵ=1\epsilon=1 for all pp or ϵ=1\epsilon=-1 for all pp and 𝝎i,s,r{\mbox{\boldmath$\omega$}}_{i,s,r} is the element of 𝒫𝓆+\cal P_{q}^{+} which was introduced in Section 4.1.3. In types DD and EE the preceding formulae still correspond to minimal affinizations under suitable restrictions. Unfortunately these are far from being all of them; the difficulty lies in the existence of the trivalent node (see [36, 37]). The problem of classifying all the minimal elements was studied in [95] but the full details are still to appear.

The equivalence classes of Kirillov–Reshetikhin modules are clearly minimal affinizations. The following result was conjectured in [86] and proved in [82, 89, 90]. It again justifies the use of the term graded limit. We do not state the result in full generality in type DD but restrict our attention to the minimal affinizations discussed here.

Theorem.

Assume that 𝝎𝒫𝓆+{\mbox{\boldmath$\omega$}}\in\cal P_{q}^{+} is as in the discussion just preceding the theorem. Then Vloc(𝝎)V_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}) is the 𝔤[t]\mathfrak{g}[t]-module generated by an element v𝝎v_{\mbox{\boldmath$\omega$}} with relations:

xi+v𝝎=0,(hitk)v𝝎=δk,0wt𝝎(hi)v𝝎(xβt)v𝝎=0(xi1)wt𝝎(hi)+1v𝝎=0,x_{i}^{+}v_{\mbox{\boldmath$\omega$}}=0,\ \ (h_{i}\otimes t^{k})v_{\mbox{\boldmath$\omega$}}=\delta_{k,0}{\rm{wt}}{~{}}{\mbox{\boldmath$\omega$}}(h_{i})v_{\mbox{\boldmath$\omega$}}\ \ (x_{\beta}^{-}\otimes t)v_{\mbox{\boldmath$\omega$}}=0\ \ (x_{i}^{-}\otimes 1)^{{\rm{wt}}{~{}}{\mbox{\boldmath$\omega$}}(h_{i})+1}v_{\mbox{\boldmath$\omega$}}=0,

for all iIi\in I and for all β=i=1nsiαiR+\beta=\sum_{i=1}^{n}s_{i}\alpha_{i}\in R^{+} with si1s_{i}\leq 1. ∎

The 𝔤\mathfrak{g}-module decomposition of the minimal affinizations was computed in rank two in [21]. In the case of AnA_{n} the graded limit is irreducible and so its character is just the character of Vq(wt𝝎)V_{q}({\rm{wt}}{~{}}{\mbox{\boldmath$\omega$}}). The 𝔤\mathfrak{g}-module decomposition in type BB and D4D_{4} was partially given in [86] and the result in complete generality is in [89, 90] for types B,CB,C and for certain minimal affinizations in type DD. Moreover, in [99] Sam proved a conjecture made in [26] that the character of minimal affinizations in types BCDBCD are given by a Jacobi–Trudi determinant.

4.1.5. Tensor products and Fusion products

Before continuing with our justification for the term graded limit, we discuss the following natural question. Suppose that MM and MM^{\prime} are 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g}))-modules which have a classical (resp. graded) limit. Is it true that MMM\otimes M^{\prime} has a classical limit and how does this relate to the tensor product of the classical (resp. graded) limits? The answer to this question is far from straightforward; even if MMM\otimes M^{\prime} does have a classical limit it is easy to generate examples where (MM)1(M\otimes M^{\prime})_{1} is not isomorphic to M1M1M_{1}\otimes M^{\prime}_{1} as L(𝔤)L(\mathfrak{g})-modules. For instance if we take 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2} and M=Vq(𝝎1,q2)M=V_{q}({\mbox{\boldmath$\omega$}}_{1,q^{2}}), M=Vq(𝝎1,1)M^{\prime}=V_{q}({\mbox{\boldmath$\omega$}}_{1,1}) then the module MMM\otimes M^{\prime} is a cyclic 𝕌q(L(𝔰𝔩2))\mathbb{U}_{q}(L(\mathfrak{sl}_{2}))-module and so the classical limit is a cyclic indecomposable module for L(𝔰𝔩2)L(\mathfrak{sl}_{2}). However the tensor product of the classical limits is V(𝝎1)V(𝝎1)V({\mbox{\boldmath$\omega$}}_{1})\otimes V({\mbox{\boldmath$\omega$}}_{1}) (equivalently ev1V(ω)ev1V(ω)\operatorname{ev}_{1}V(\omega)\otimes\operatorname{ev}_{1}V(\omega)) which is completely reducible.

The 𝔤\mathfrak{g}-module structure however is unchanged in the process. This is because it is known that the process of taking classical limits preserves tensor products for the simple Lie algebras. If we work with the graded limit then again it is false that the graded character of (MM)loc(M\otimes M^{\prime})_{\operatorname{loc}} is the same as MlocMlocM_{\operatorname{loc}}\otimes M^{\prime}_{\operatorname{loc}}. However in many examples the graded limit of the tensor product coincides with an operation called the fusion product defined on graded 𝔤[t]\mathfrak{g}[t]-modules. This notion was introduced in [50] and we now recall this construction.

Let VV be a finite-dimensional cyclic 𝔤[t]\mathfrak{g}[t]-module generated by an element vv and for r+r\in\mathbb{Z}_{+} set

FrV=(0sr𝕌(𝔤[t])[s])vF^{r}V=\left(\bigoplus_{0\leq s\leq r}\mathbb{U}(\mathfrak{g}[t])[s]\right)\cdot v

Clearly FrVF^{r}V is a 𝔤\mathfrak{g}-submodule of VV and we have a finite 𝔤\mathfrak{g}-module filtration

0F0VF1VFkV=V,0\subseteq F^{0}V\subseteq F^{1}V\subseteq\cdots\subseteq F^{k}V=V,

for some k+k\in\mathbb{Z}_{+}. The associated graded vector space grV\operatorname{gr}V acquires a graded 𝔤[t]\mathfrak{g}[t]-module structure in a natural way and is generated by the image of vv in grV\operatorname{gr}V. Given a 𝔤[t]\mathfrak{g}[t]-module VV and zz\in\mathbb{C}, let VzV^{z} be the 𝔤[t]\mathfrak{g}[t]-module with action

(xtr)w=(x(t+z)r)w,x𝔤,r+,wV.(x\otimes t^{r})w=(x\otimes(t+z)^{r})w,\ \ x\in\mathfrak{g},\ \ r\in\mathbb{Z}_{+},\ w\in V.

Let VsV_{s}, 1sp1\leq s\leq p, be cyclic finite-dimensional 𝔤[t]\mathfrak{g}[t]-modules with cyclic vectors vsv_{s}, 1sp1\leq s\leq p and let z1,,zpz_{1},\dots,z_{p} be distinct complex numbers. Then the module V1z1VpzpV_{1}^{z_{1}}\otimes\cdots\otimes V_{p}^{z_{p}} is cyclic with cyclic generator v1vp.v_{1}\otimes\cdots\otimes v_{p}. The fusion product V1z1VpzpV_{1}^{z_{1}}*\cdots*V_{p}^{z_{p}} is defined to be grV1z1Vpzp.\operatorname{gr}V_{1}^{z_{1}}\otimes\cdots\otimes V_{p}^{z_{p}}. For ease of notation we shall use V1VkV_{1}*\cdots*V_{k} for V1z1VkzkV_{1}^{z_{1}}*\cdots*V_{k}^{z_{k}}.

It is conjectured in [50] that under some suitable conditions on VsV_{s} and vsv_{s}, the fusion product is independent of the choice of the complex numbers zsz_{s}, 1sk1\leq s\leq k, and this conjecture is verified in many special cases by various people (see for instance [30], [41], [49], [51] [58], [75], [91]). In all these cases the conjecture is proved by exhibiting a graded presentation of the fusion product which is independent of all parameters. This is much like what we have been doing to justify the use of graded limit and the coincidence is not accidental. In almost all of these papers the proof of the Feigin–Loktev conjecture involves giving a presentation of the graded limit of certain 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g}))-modules.

4.1.6. A presentation of Wloc(𝝎)W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}})

We return to our discussion in Section 3.2.6. Recall that we had discussed that given 𝝎𝒫𝓆+{\mbox{\boldmath$\omega$}}\in\cal P_{q}^{+} we can write 𝝎=𝝎i1,a1𝝎ik,ak{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}_{i_{1},a_{1}}\cdots{\mbox{\boldmath$\omega$}}_{i_{k},a_{k}} so that there is a surjective map of 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g}))-modules

Wq(𝝎)Vq(𝝎i1,a1)Vq(𝝎ik,ak)0.W_{q}({\mbox{\boldmath$\omega$}})\to V_{q}({\mbox{\boldmath$\omega$}}_{i_{1},a_{1}})\otimes\cdots\otimes V_{q}({\mbox{\boldmath$\omega$}}_{i_{k},a_{k}})\to 0.

It has been conjectured in [38] that

dimWq(𝝎)=s=1kdimVq(𝝎is,as)=s=1kdimWq(𝝎is,as),\dim W_{q}({\mbox{\boldmath$\omega$}})=\prod_{s=1}^{k}\dim V_{q}({\mbox{\boldmath$\omega$}}_{i_{s},a_{s}})=\prod_{s=1}^{k}\dim W_{q}({\mbox{\boldmath$\omega$}}_{i_{s},a_{s}}), (4.1)

where the second equality is a consequence of Section 3.2.5. Since the dimension is unchanged when passing to the graded limit it suffices to prove that

dimWloc(𝝎)=s=1kdimVloc(𝝎is,as).\dim W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}})=\prod_{s=1}^{k}\dim V_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}_{i_{s},a_{s}}). (4.2)

Using Lemma 4.1.1 and the discussion in Section 4.1.2 we see that Wloc(𝝎)W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}) is the quotient of the module W~loc(wt𝝎)\tilde{W}_{\operatorname{loc}}({\rm{wt}}{{\mbox{\boldmath$\omega$}}}) which is generated as a 𝔤[t]\mathfrak{g}[t]-module by an element w𝝎w_{{\mbox{\boldmath$\omega$}}} with defining relations:

(ht[t])w𝝎=0,hv𝝎=wt𝝎(h)w𝝎,xi+w𝝎=0,(xi)wt𝝎(hi)+1v𝝎=0.(h\otimes t\mathbb{C}[t])w_{\mbox{\boldmath$\omega$}}=0,\ \ hv_{\mbox{{\boldmath{\small$\omega$}}}}={\rm{wt}}{{\mbox{\boldmath$\omega$}}}(h)w_{\mbox{{\boldmath{\small$\omega$}}}},\ \ x_{i}^{+}w_{{\mbox{{\boldmath{\small$\omega$}}}}}=0,\ (x_{i}^{-})^{{\rm{wt}}{{\mbox{{\boldmath{\small$\omega$}}}}}(h_{i})+1}v_{{\mbox{{\boldmath{\small$\omega$}}}}}=0.

Notice that by Theorem 4.1.3 we know that

W~loc(ωi)Wloc(𝝎i,a).\tilde{W}_{\operatorname{loc}}(\omega_{i})\cong W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}_{i,a}).

Choosing distinct scalars z1,,zkz_{1},\dots,z_{k} consider the fusion product W~loc(ωi1)z1W~loc(ωik)zk\tilde{W}_{\operatorname{loc}}(\omega_{i_{1}})^{z_{1}}*\cdots*\tilde{W}_{\operatorname{loc}}(\omega_{i_{k}})^{z_{k}}. It is not too hard to prove that this module is a quotient of W~loc(wt𝝎)\tilde{W}_{\operatorname{loc}}({\rm{wt}}{{\mbox{\boldmath$\omega$}}}). We get

dimW~loc(wt𝝎)dimWloc(𝝎)s=1kdimW~loc(ωis).\dim\tilde{W}_{\operatorname{loc}}({\rm{wt}}{{\mbox{\boldmath$\omega$}}})\geq\dim W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}})\geq\prod_{s=1}^{k}\dim\tilde{W}_{\operatorname{loc}}(\omega_{i_{s}}).

The following result was established in [39] for 𝔰𝔩2\mathfrak{sl}_{2}. Using this the result was established in [30] in the case of 𝔰𝔩n+1\mathfrak{sl}_{n+1} where a Gelfand–Tsetlin type basis was also given for Wloc(𝝎)W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}). These bases were further studied in [96, 97]. In [58] the theorem was proved for simply-laced Lie algebras. Finally in [88] the result was established for non-simply laced types.

Theorem.

We have an isomorphism of 𝔤[t]\mathfrak{g}[t]-modules:

W~loc(wt𝝎)Wloc(𝝎)W~loc(ωi1)z1W~loc(ωik)zk,wt𝝎=ωi1++ωik.\tilde{W}_{\operatorname{loc}}({\rm{wt}}{{\mbox{\boldmath$\omega$}}})\cong W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}})\cong\tilde{W}_{\operatorname{loc}}(\omega_{i_{1}})^{z_{1}}*\cdots*\tilde{W}_{\operatorname{loc}}(\omega_{i_{k}})^{z_{k}},\ \ {\rm{wt}}{{\mbox{\boldmath$\omega$}}}=\omega_{i_{1}}+\cdots+\omega_{i_{k}}.

Clearly this theorem establishes the conjecture in [22] and also the conjecture of Feigin–Loktev for this particular family of modules.

Remark.

Although the preceding theorem is uniformly stated the methods of proof are very different. In [30, 39] the proof goes by writing down a basis and then doing a dimension count. In [58] the proof proceeds by showing that W~loc(wt𝝎)\tilde{W}_{\operatorname{loc}}({\rm{wt}}{{\mbox{\boldmath$\omega$}}}) is isomorphic to a stable Demazure module in a level one representation of the affine Lie algebra (see Section 2.3.2 for the relevant definitions). This isomorphism fails in the non-simply laced case. Instead it is proved in [88] that the module has a flag by stable level one Demazure modules and this plays a key role in the proof. We return to these ideas in the later sections of this paper.

4.1.7. Tensor products of Kirillov–Reshetikhin modules

It was proved in [23] that the tensor products of Kirillov–Reshetikhin modules Vq(𝝎i1,s1,r1)Vq(𝝎ik,sk,rk)V_{q}({\mbox{\boldmath$\omega$}}_{i_{1},s_{1},r_{1}})\otimes\cdots\otimes V_{q}({\mbox{\boldmath$\omega$}}_{i_{k},s_{k},r_{k}}) is irreducible as long as sisps_{i}-s_{p}, 1ipk1\leq i\neq p\leq k lie outside a finite set. A precise description of this set was also given in that paper when 𝔤\mathfrak{g} is classical. Set

𝕍=Vq(𝝎i1,s1,r1)Vq(𝝎ik,sk,rk),λ=s=1krsωis.\mathbb{V}=V_{q}({\mbox{\boldmath$\omega$}}_{i_{1},s_{1},r_{1}})\otimes\cdots\otimes V_{q}({\mbox{\boldmath$\omega$}}_{i_{k},s_{k},r_{k}}),\ \ \lambda=\sum_{s=1}^{k}r_{s}\omega_{i_{s}}.

We now discuss the results of [91] on the structure of 𝕍loc\mathbb{V}_{\operatorname{loc}}. Thus, let 𝕍~loc\tilde{\mathbb{V}}_{\operatorname{loc}} be the 𝔤[t]\mathfrak{g}[t]-module generated by a vector vv satisfying the relations:

𝔫+[t]v=0=(𝔥t[t])v,hv=λ(h)v,h𝔥,\displaystyle\mathfrak{n}^{+}[t]v=0=\left(\mathfrak{h}\otimes t\mathbb{C}[t]\right)v,\ \ \ hv=\lambda(h)v,\ \ h\in\mathfrak{h},
(Fi(z)r)sv=0,iI,r>0,s<p:ip=imin{r,rp},\displaystyle\left(F_{i}(z)^{r}\right)_{s}v=0,\ \ i\in I,\ r>0,\ s<-\sum_{p:\ i_{p}=i}\min\{r,r_{p}\},

where (Fi(z)r)s(F_{i}(z)^{r})_{s} denotes the coefficient of zsz^{s} in the rr-th power of

Fi(z)=m=0(xitm)zm1𝕌(𝔤[t])[[z1]].F_{i}(z)=\sum_{m=0}^{\infty}(x_{i}^{-}\otimes t^{m})z^{-m-1}\in\mathbb{U}(\mathfrak{g}[t])[[z^{-1}]].

The following is the main result of [91].

Theorem.

We have an isomorphism of graded 𝔤[t]\mathfrak{g}[t]-modules

𝕍locVloc(𝝎i1,s1,r1)z1Vloc(𝝎ik,sk,rk)zk𝕍~loc.\mathbb{V}_{\operatorname{loc}}\cong V_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}_{i_{1},s_{1},r_{1}})^{z_{1}}*\cdots*V_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}_{i_{k},s_{k},r_{k}})^{z_{k}}\cong\tilde{\mathbb{V}}_{\operatorname{loc}}.

Again, the conjecture of Feigin–Loktev for this family of modules is a consequence of this presentation. The proof of the Feigin–Loktev conjecture when r1=r2==rkr_{1}=r_{2}=\cdots=r_{k} and 𝔤\mathfrak{g} simply-laced was proved earlier in [58] by identifying the fusion product with a 𝔤\mathfrak{g}-stable Demazure module. In general the connection with Demazure modules or the existence of a Demazure flag (as in the case of local Weyl modules) is not known.

4.1.8. Monoidal categorification and HL-modules

Our final example of graded limits comes from the work of David Hernandez and Bernard Leclerc on monoidal categorification of cluster algebras. We refer the reader to [103] for a quick introduction to cluster algebras. For the purposes of this article it is enough for us to recall that a cluster algebra is a commutative ring with certain distinguished generators called cluster variables and certain algebraically independent subsets of cluster variables called clusters. Monomials in the cluster variables belonging to a cluster are called cluster monomials. There is also an operation called mutation; this is a way to produce a new cluster by replacing exactly one element of the original cluster by another cluster variable.

The remarkable insight of Hernandez–Leclerc was to relate these ideas to the representation theory of quantum affine algebras associated to simply-laced Lie algebras. Broadly speaking they prove that the Grothendieck ring of a suitable tensor subcategory admits the structure of a cluster algebra. A cluster variable is a prime real representation in this category (see Section 3.2.2 for the definitions) and we call these the HL-modules. Suppose that V,VV,V^{\prime} are irreducible modules in this subcategory. Assume that their isomorphism classes correspond to cluster variables which belong to the same cluster. Then VVV\otimes V^{\prime} is an irreducible module. The operation of mutation in this language corresponds to the Jordan–Holder decomposition of the corresponding tensor product.

We now give one specific example of their work and relate it to our study of graded limits. We assume that 𝔤\mathfrak{g} is of type AnA_{n}. Let κ:{1,,n}\kappa:\{1,\dots,n\}\rightarrow\mathbb{Z} be a height function; namely it satisfies |κ(i+1)κ(i)|=1|\kappa(i+1)-\kappa(i)|=1 for 1in1\leq i\leq n. Let 𝒫κ+\cal P^{+}_{\kappa} be the submonoid of 𝒫𝓆+\cal P^{+}_{q} generated by elements 𝝎i,qκ(i)±1{\mbox{\boldmath$\omega$}}_{i,q^{\kappa(i)\pm 1}}, iIi\in I. Let κ\cal F_{\kappa} be the full subcategory of 𝓆\cal F_{q} consisting of finite-dimensional 𝕌q(L(𝔤))\mathbb{U}_{q}(L(\mathfrak{g}))-modules whose Jordan–Holder constituents are isomorphic to Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}) for some 𝝎𝒫κ+{\mbox{\boldmath$\omega$}}\in\cal P_{\kappa}^{+}. It was shown in [67] that κ\cal F_{\kappa} is closed under taking tensor products and that its Grothendieck ring has the structure of a cluster algebra of type AnA_{n}. The following result was proved in [67] when κ(i)=imod2\kappa(i)=i\mod 2, in [68] when κ(i)=i\kappa(i)=i and in complete generality in [16].

Theorem.

Suppose that Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}) is a prime real object of κ\cal F_{\kappa}. Then 𝝎\omega must be one of the following:

𝝎i,qκ(i)±1,𝝎i,qκ(i)+1𝝎i,qκ(i)1,iI,\displaystyle{\mbox{\boldmath$\omega$}}_{i,q^{\kappa(i)\pm 1}},\ \ {\mbox{\boldmath$\omega$}}_{i,q^{\kappa(i)+1}}{\mbox{\boldmath$\omega$}}_{i,q^{\kappa(i)-1}},\ \ i\in I,
𝝎i,a1𝝎i2,a2𝝎ik1,ak1𝝎j,ak, 1i<jn,\displaystyle{\mbox{\boldmath$\omega$}}_{i,a_{1}}{\mbox{\boldmath$\omega$}}_{i_{2},a_{2}}\cdots{\mbox{\boldmath$\omega$}}_{i_{k-1},a_{k-1}}{\mbox{\boldmath$\omega$}}_{j,a_{k}},\ 1\leq i<j\leq n,

where i2<<ik1i_{2}<\cdots<i_{k-1} is an ordered enumeration of {p:i<p<j,κ(p1)=κ(p+1)}\{p:i<p<j,\ \kappa(p-1)=\kappa(p+1)\} and a1=qκ(i)±1a_{1}=q^{\kappa(i)\pm 1} if κ(i+1)=κ(i)1\kappa(i+1)=\kappa(i)\mp 1 and as=qκ(is)±1a_{s}=q^{\kappa(i_{s})\pm 1} if κ(is)=κ(is1)±1\kappa(i_{s})=\kappa(i_{s}-1)\pm 1 for s2s\geq 2. Conversely the irreducible representation associated to any 𝝎\omega as above is a real prime object of κ\cal F_{\kappa}. ∎

4.1.9. Graded Limits of HL-modules in κ\cal F_{\kappa}

Continue to assume that 𝔤\mathfrak{g} is of type AnA_{n} and for 1ijn1\leq i\leq j\leq n set αi,j=αi++αjR+\alpha_{i,j}=\alpha_{i}+\cdots+\alpha_{j}\in R^{+}. It follows from the discussion in Section 3.2.5 that Vloc(𝝎i,qκ(i)±1)Wloc(𝝎i,qκ(i)±1)V_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}_{i,q^{\kappa(i)\pm 1}})\cong W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}_{i,q^{\kappa(i)\pm 1}}). The discussion in Section 4.1.3 gives a presentation for Vloc(𝝎i,qκ(i)+1𝝎i,qκ(i)1)V_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}_{i,q^{\kappa(i)+1}}{\mbox{\boldmath$\omega$}}_{i,q^{\kappa(i)-1}}) since this is a special example of a Kirillov–Reshetikhin module. The following was proved in [17] and shows that the graded limits of HL-modules are indeed graded.

Theorem.

Suppose that 𝔤\mathfrak{g} is of type AnA_{n} and 𝝎=𝝎i,a1𝝎j,ak𝒫+{\mbox{\boldmath$\omega$}}={\mbox{\boldmath$\omega$}}_{i,a_{1}}\cdots{\mbox{\boldmath$\omega$}}_{j,a_{k}}\in\cal P^{+} is as in Theorem 4.1.8. Then Vloc(𝝎)V_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}) is the quotient of Wloc(𝝎)W_{\operatorname{loc}}({\mbox{\boldmath$\omega$}}) by the submodule generated by the additional relations:

(xαt)w𝝎=0,α{αi,i2,αi2,i3,,αik1,j}.(x_{\alpha}^{-}\otimes t)w_{{\mbox{\boldmath$\omega$}}}=0,\ \ \alpha\in\{\alpha_{i,i_{2}},\alpha_{i_{2},i_{3}},\cdots,\alpha_{i_{k-1},j}\}.

We remark that the result in [17] is more general in the sense that it gives a presentation of the graded limit of the tensor product of an HL-module with the Kirillov–Reshetikhin modules in this category. Here again the result shows that tensor products specialize to fusion products. A problem that has not been studied so far is to understand the graded limit of a tensor product of Vq(𝝎)Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}})\otimes V_{q}({\mbox{\boldmath$\omega$}}^{\prime}) for an arbitrary pair 𝝎,𝝎𝒫κ+{\mbox{\boldmath$\omega$}},{\mbox{\boldmath$\omega$}}^{\prime}\in\cal P^{+}_{\kappa} and the connection with the fusion product of the graded limits of Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}) and Vq(𝝎)V_{q}({\mbox{\boldmath$\omega$}}^{\prime}).

The graded characters of the limits of HL-modules have been studied in [14] and [4] in different ways. In the first paper a character formula was given as an explicit linear combination of Macdonald polynomials. In [4] the authors studied the 𝔤\mathfrak{g}-module decomposition of the graded limit. The multiplicity of a particular 𝔤\mathfrak{g}-type is given by the number of certain lattice points in a convex polytope. Moreover, considering a particular face of that polytope encodes in fact the graded multiplicity.

A comparable study of HL-modules in other types is only partially explored. A first step was taken in [24] in type DnD_{n} but it does not capture all the prime objects in the category κ\cal F_{\kappa}. There are important differences from the AnA_{n} case and some new ideas seem to be necessary.

4.1.10. Further Remarks

As we said, there are other subcategories of representations of 𝓆\cal F_{q} which were shown by Hernandez-Leclerc to be monoidal categorifications of (infinite rank) cluster algebras. However, it is far from clear what subset of 𝒫𝓆+\cal P^{+}_{q} is an index set for the prime representations corresponding to the cluster variables. Hence little is known about the characters or the graded limits of these representations.

Another example of prime representations comes from the theory of snake modules studied in types AnA_{n} and BnB_{n} in [87, 18]. Again the problem of studying the graded limits of these modules is wide open.

5. Demazure modules, Projective modules and Global Weyl modules

Our focus in this section will be on the study of graded representations of 𝔤[t]\mathfrak{g}[t]. We begin by establishing the correct category 𝒢\cal G of representations of the current algebra and introduce the projective objects and the global Weyl modules. We then relate the study of local Weyl modules in Section 4 to the 𝔤\mathfrak{g}-stable Demazure modules introduced in Section Lemma. Next we discuss the characters of the local Weyl modules and relate them to Macdonald polynomials. Finally, we discuss BGG-type reciprocity results. We conclude the section with some comments on the more recent work of [47], [52], [46] [74].

5.1. The category 𝒢\cal G

The study of this category was initiated in [26] and we recall several ideas from that paper. Recall from Section 4.1.2 that we have a +\mathbb{Z}_{+}-grading on 𝔤[t]\mathfrak{g}[t] and its universal enveloping algebra. Define 𝒢\cal G to be the category whose objects are \mathbb{Z}-graded representations V=mV[m]V=\oplus_{m\in\mathbb{Z}}V[m] of 𝔤[t]\mathfrak{g}[t] with dimV[m]<\dim V[m]<\infty for all mm\in\mathbb{Z}. The morphisms in the category are grade preserving maps of 𝔤[t]\mathfrak{g}[t]-modules.

Define the restricted dual of an object VV in 𝒢\mathcal{G} by

V=mV[m],V[m]=V[m].V^{*}=\bigoplus_{m\in\mathbb{Z}}V^{*}[m],\ \ \ V^{*}[m]=V[-m]^{*}.

Clearly VV^{*} is again an object of 𝒢\cal G.

For any object VV of 𝒢\cal G, each graded subspace V[m]V[m] is a finite-dimensional 𝔤\mathfrak{g}-module and we define the graded 𝔤\mathfrak{g}-character of VV to be the element of [P][[q±1]]\mathbb{Z}[P][[q^{\pm 1}]]:

chgrV=λP+mdimHom𝔤(V(λ),V[m])qmchV(λ)=μPmdimV[m]μqmeμ=μPpμ(q)eμ,pμ(q)+[[q±1]].\begin{split}\operatorname{ch}_{\operatorname{gr}}V&=\sum_{\lambda\in P^{+}}\sum_{m\in\mathbb{Z}}\dim\operatorname{Hom}_{\mathfrak{g}}(V(\lambda),V[m])q^{m}\operatorname{ch}V(\lambda)\\ &=\sum_{\mu\in P}\sum_{m\in\mathbb{Z}}\dim V[m]_{\mu}q^{m}e_{\mu}=\sum_{\mu\in P}p_{\mu}(q)e_{\mu},\ \ p_{\mu}(q)\in\mathbb{Z}_{+}[[q^{\pm 1}]].\end{split}

It is clear that for all rr\in\mathbb{Z} we have chgr(τrV)=qrchgrV\operatorname{ch}_{\operatorname{gr}}(\tau_{r}V)=q^{r}\operatorname{ch}_{\operatorname{gr}}V, where τr\tau_{r} is as defined in Section 2.3.4).

Finally, note that 𝒢\cal G is an abelian category and is closed under taking restricted duals. If VV and VV^{\prime} are objects of 𝒢\cal G then VVV\otimes V^{\prime} is again an object in 𝒢\cal G if dimV<\dim V<\infty.

5.1.1. Finite-dimensional objects of 𝒢\cal G

It is straightforward that if VV is a simple object of 𝒢\cal G, then VV is concentrated in a single grade. In particular VV must be a finite-dimensional irreducible 𝔤\mathfrak{g}-module. In other words Vτmev0V(λ)V\cong\tau_{m}\operatorname{ev}_{0}V(\lambda) for some λP+\lambda\in P^{+} where ev0\operatorname{ev}_{0} is the evaluation 𝔤[t]𝔤\mathfrak{g}[t]\to\mathfrak{g}, xtrδ0,rxx\otimes t^{r}\mapsto\delta_{0,r}x. From now we set

V(λ,m)=τmev0V(λ).V(\lambda,m)=\tau_{m}\operatorname{ev}_{0}V(\lambda).

Another example of finite-dimensional modules in 𝒢\cal G are the 𝔤\mathfrak{g}-stable Demazure modules Vw(λ)V_{w}(\lambda), λP^+\lambda\in\widehat{P}^{+} (see Section Lemma) and the local Weyl modules studied in Section 4.1.6. We give a direct definition of those objects as 𝔤[t]\mathfrak{g}[t]-modules here for the reader’s convenience, and we also drop the ~\tilde{} for ease of notation.

Given λP+\lambda\in P^{+} the local Weyl module Wloc(λ)W_{\operatorname{loc}}(\lambda) is the 𝔤[t]\mathfrak{g}[t]-module generated by an element vλv_{\lambda} and relations:

xi+vλ=0,(htr)vλ=δr,0λ(h)vλ,(xi)λ(hi)+1vλ=0,iI,h𝔥.x_{i}^{+}v_{\lambda}=0,\ \ (h\otimes t^{r})v_{\lambda}=\delta_{r,0}\lambda(h)v_{\lambda},\ \ (x_{i}^{-})^{\lambda(h_{i})+1}v_{\lambda}=0,\ \ \ i\in I,\ \ h\in\mathfrak{h}.

Setting grvλ=r\operatorname{gr}v_{\lambda}=r we see that Wloc(λ)W_{\operatorname{loc}}(\lambda) can be regarded as an object of 𝒢\cal G and we denote this as Wloc(λ,r)W_{\operatorname{loc}}(\lambda,r). Clearly Wloc(λ,r)=τrWloc(λ,0)W_{\operatorname{loc}}(\lambda,r)=\tau_{r}W_{\operatorname{loc}}(\lambda,0). It was proved in [39] that the local Weyl modules are finite-dimensional with unique irreducible quotients.

Given μP+\mu\in P^{+} and \ell\in\mathbb{N}, let D(,μ)D(\ell,\mu) be the quotient of Wloc(μ)W_{\operatorname{loc}}(\mu) by the submodule generated by elements

(xαtsα1)mα+1vμ, if mα<dα,(xαtsα)vμ,αR+(x_{\alpha}^{-}\otimes t^{s_{\alpha}-1})^{m_{\alpha}+1}v_{\mu},\ \text{ if $m_{\alpha}<d_{\alpha}\ell$},\ \ (x_{\alpha}^{-}\otimes t^{s_{\alpha}})v_{\mu},\ \ \alpha\in R^{+}

where sαs_{\alpha} and mαm_{\alpha} are determined by

μ(hα)=(sα1)dα+mα, 0<mαdα.\mu(h_{\alpha})=(s_{\alpha}-1)d_{\alpha}\ell+m_{\alpha},\ \ 0<m_{\alpha}\leq d_{\alpha}\ell.

The following was proved in [42].

Proposition.

Suppose that λP^+\lambda\in\widehat{P}^{+} and wW^w\in\widehat{W} is such that wλ(hi)0w\lambda(h_{i})\leq 0 for all iIi\in I and assume that λ(c)=\lambda(c)=\ell. The module Vw(λ)V_{w}(\lambda) is isomorphic to τrD(,μ)\tau_{r}D(\ell,\mu) where μP+\mu\in P^{+} is given by μ(hi)=wwλ(hi)\mu(h_{i})=-w_{\circ}w\lambda(h_{i}) and r=wλ(d)r=w\lambda(d). ∎

Remark.

An analogous presentation of non-stable Demazure modules is given in [80] and we discuss this in the next section. These modules however are not objects of 𝒢\cal G. ∎

5.1.2. Relation between local Weyl and Demazure modules

The following corollary of Proposition 5.1.1 is easily established.

Corollary.

If 𝔤\mathfrak{g} is simply-laced then

D(1,μ)Wloc(μ),μP+.D(1,\mu)\cong W_{\operatorname{loc}}(\mu),\ \ \mu\in P^{+}.
Proof.

It suffices to prove that the following relation holds in Wloc(μ)W_{\operatorname{loc}}(\mu):

(xαtμ(hα))vμ=0,αR+.(x_{\alpha}^{-}\otimes t^{\mu(h_{\alpha})})v_{\mu}=0,\ \ \alpha\in R^{+}.

But this follows by using

(xαtμ(hα))vμ=(xα+t)μ(hα)(xα)μ(hα)+1vμ=0.(x_{\alpha}^{-}\otimes t^{\mu(h_{\alpha})})v_{\mu}=(x_{\alpha}^{+}\otimes t)^{\mu(h_{\alpha})}(x_{\alpha}^{-})^{\mu(h_{\alpha})+1}v_{\mu}=0.

Here the first equality is established by a simple calculation and using the relations in Wloc(μ)W_{\operatorname{loc}}(\mu). The second equality holds since Wloc(μ)W_{\operatorname{loc}}(\mu) is finite-dimensional and

xα+vμ=0(xα)μ(hα)+1vμ=0.x_{\alpha}^{+}v_{\mu}=0\implies(x_{\alpha}^{-})^{\mu(h_{\alpha})+1}v_{\mu}=0.

In the non-simply laced case it is not true in general that Wloc(μ)D(1,μ)W_{\operatorname{loc}}(\mu)\cong D(1,\mu). However it was proved in [88] that Wloc(μ)W_{\operatorname{loc}}(\mu) admits a decreasing filtration where the successive quotients are isomorphic to τrD(1,μr)\tau_{r}D(1,\mu_{r}) for some rr\in\mathbb{Z} and μrP+\mu_{r}\in P^{+}. In fact one can make a more precise statement which can be found in Section 6.2.

5.1.3. Projective modules and Global Weyl modules

Given (λ,r)P+×(\lambda,r)\in P^{+}\times\mathbb{Z} set

P(λ,r)=𝕌(𝔤[t])𝐔(𝔤)V(λ,r).P(\lambda,r)=\mathbb{U}(\mathfrak{g}[t])\otimes_{\mathbf{U}(\mathfrak{g})}V(\lambda,r).

It is not hard to check that P(λ,r)P(\lambda,r) is an indecomposable projective object of 𝒢\cal G and that there exists a surjective map P(λ,r)V(λ,r)0P(\lambda,r)\to V(\lambda,r)\to 0 of 𝔤[t]\mathfrak{g}[t]-modules. Equivalently P(λ,r)P(\lambda,r) is the 𝔤[t]\mathfrak{g}[t]-module generated by an element vλv_{\lambda} of grade rr subject to the relations:

xi+vλ=0,hvλ=λ(h)vλ,(xi)λ(hi)+1vλ=0,iI,h𝔥.x_{i}^{+}v_{\lambda}=0,\ \ hv_{\lambda}=\lambda(h)v_{\lambda},\ \ (x_{i}^{-})^{\lambda(h_{i})+1}v_{\lambda}=0,\ \ i\in I,h\in\mathfrak{h}.

The global Weyl module W(λ,r)W(\lambda,r) is the maximal quotient of P(λ,r)P(\lambda,r) such that wtW(λ,r)λQ+\mathrm{wt}W(\lambda,r)\subseteq\lambda-Q^{+}. Equivalently it is the quotient of P(λ,r)P(\lambda,r) obtained by imposing the additional relations (xi+tk)vλ=0(x_{i}^{+}\otimes t^{k})v_{\lambda}=0 for all iIi\in I and k0k\geq 0. Clearly we have the following sequence of surjective maps

P(λ,r)W(λ,r)Wloc(λ,r)V(λ,r).P(\lambda,r)\twoheadrightarrow W(\lambda,r)\twoheadrightarrow W_{\operatorname{loc}}(\lambda,r)\twoheadrightarrow V(\lambda,r).

5.1.4. The algebra 𝔸λ\mathbb{A}_{\lambda} and the bimodule structure on W(λ,r)W(\lambda,r)

Let 𝔥[t]+=𝔥t[t]\mathfrak{h}[t]_{+}=\mathfrak{h}\otimes t\mathbb{C}[t] and for λP+\lambda\in P^{+} and vλW(λ,r)λv_{\lambda}\in W(\lambda,r)_{\lambda} non-zero of grade rr let

𝕀λ={u𝕌(𝔥[t]+):uvλ=0},𝔸λ=𝕌(𝔥[t]+)/𝕀λ.\mathbb{I}_{\lambda}=\{u\in\mathbb{U}(\mathfrak{h}[t]_{+}):uv_{\lambda}=0\},\ \ \mathbb{A}_{\lambda}=\mathbb{U}(\mathfrak{h}[t]_{+})/\mathbb{I}_{\lambda}.

Clearly 𝔸λ\mathbb{A}_{\lambda} is commutative and graded. Moreover W(λ,r)W(\lambda,r) is a (𝔤[t],𝔸λ)(\mathfrak{g}[t],\mathbb{A}_{\lambda})-bimodule where the right action of 𝔸λ\mathbb{A}_{\lambda} is given by:

(gvλ)a=gavλ,g𝕌(𝔤[t]),a𝔸λ.(gv_{\lambda})a=gav_{\lambda},\ \ g\in\mathbb{U}(\mathfrak{g}[t]),\ \ a\in\mathbb{A}_{\lambda}.

To see that the action is well-defined, one must prove that

(𝔫+[t])(hf)vλ=0,(hλ(h))(hf)vλ=0,(xi)λ(hi)+1(hf)vλ=0(\mathfrak{n}^{+}\otimes\mathbb{C}[t])(h\otimes f)v_{\lambda}=0,\ \ (h^{\prime}-\lambda(h^{\prime}))(h\otimes f)v_{\lambda}=0,\ \ (x_{i}^{-})^{\lambda(h_{i})+1}(h\otimes f)v_{\lambda}=0

for all iIi\in I, h,h𝔥h,h^{\prime}\in\mathfrak{h} and f[t]f\in\mathbb{C}[t]. However, all relations are immediate to check. It was proved in [39] (for the loop algebra; the proof is essentially the same for the current algebra) that 𝔸λ\mathbb{A}_{\lambda} can be realized as a ring of invariants as follows. Consider the polynomial ring [xi,r:iI,1rλ(hi)]\mathbb{C}[x_{i,r}:i\in I,1\leq r\leq\lambda(h_{i})]. The direct product of symmetric groups

𝒮λ=Sλ(h1)××Sλ(hn)\mathcal{S}_{\lambda}=S_{\lambda(h_{1})}\times\cdots\times S_{\lambda(h_{n})}

acts on this ring in an obvious way and we have

𝔸λ[xi,r:iI,1rλ(hi)]𝒮λ.\mathbb{A}_{\lambda}\cong\mathbb{C}[x_{i,r}:i\in I,1\leq r\leq\lambda(h_{i})]^{\cal S_{\lambda}}.

The grading on 𝔸λ\mathbb{A}_{\lambda} is given by requiring the grade of xi,rx_{i,r} being rr. Let 𝕀λ\mathbb{I}_{\lambda} be the maximal graded ideal in 𝔸λ\mathbb{A}_{\lambda}. The local Weyl module can then be realized as follows:

Wloc(λ,r)=W(λ,r)𝔸λ𝔸λ/𝕀λ.W_{\operatorname{loc}}(\lambda,r)=W(\lambda,r)\otimes_{\mathbb{A}_{\lambda}}{\mathbb{A}_{\lambda}}/{\mathbb{I}_{\lambda}}.

A nontrivial consequence of the dimension conjecture discussed in Section 4.1.6 (see [39], [25] for more details) is the following result.

Proposition.

The global Weyl module W(λ)W(\lambda) is a free 𝔸λ\mathbb{A}_{\lambda}-module of rank equal to the dimension of Wloc(λ)W_{\operatorname{loc}}(\lambda). ∎

The algebra 𝔸λ\mathbb{A}_{\lambda} plays an important role in the rest of the section.

5.2. The category 𝒪\cal O for 𝔤\mathfrak{g}

Before continuing our study of the category 𝒢\cal G, we discuss briefly, the resemblance of the theory with that of the well-known category 𝒪\cal O for semi-simple Lie algebras.

The objects of 𝒪\cal O are finitely generated weight modules (with finite-dimensional weight spaces) for 𝔤\mathfrak{g} which are locally nilpotent for the action of 𝔫+\mathfrak{n}^{+}. The morphisms are just 𝔤\mathfrak{g}-module maps. Given λ𝔥\lambda\in\mathfrak{h}^{*} one can associate to it a Verma module M(λ)M(\lambda) which is defined as

M(λ)=𝕌(𝔤)𝕌(𝔟)vλ,M(\lambda)=\mathbb{U}(\mathfrak{g})\otimes_{\mathbb{U}(\mathfrak{b})}\mathbb{C}v_{\lambda},

where vλ\mathbb{C}v_{\lambda} is the one-dimensional 𝔟\mathfrak{b}-module given by hvλ=λ(h)vλhv_{\lambda}=\lambda(h)v_{\lambda} and 𝔫+vλ=0\mathfrak{n}^{+}v_{\lambda}=0. It is not hard to prove that M(λ)M(\lambda) is infinite-dimensional and has a unique irreducible quotient denoted by V(λ)V(\lambda) and any irreducible object in 𝒪\cal O is isomorphic to some V(λ)V(\lambda). Moreover V(λ)V(\lambda) is finite-dimensional if and only if λP+\lambda\in P^{+}; in particular M(λ)M(\lambda) is reducible if λP+\lambda\in P^{+}.

The modules M(λ)M(\lambda) have finite length and the multiplicity of V(μ)V(\mu) in the Jordan–Hölder series of M(λ)M(\lambda) is denoted by [M(λ):V(μ)][M(\lambda):V(\mu)]. The study of these multiplicities has been of great interest and there is extensive literature on the subject. Perhaps the starting point for this study is the famous result of Bernstein–Gelfand–Gelfand (BGG) which we now recall.

The category 𝒪\cal O has enough projectives, which means that for λ𝔥\lambda\in\mathfrak{h}^{*} there exists an indecomposable module P(λ)P(\lambda) which is projective in 𝒪\cal O and we have surjective maps

P(λ)M(λ)V(λ).P(\lambda)\twoheadrightarrow M(\lambda)\twoheadrightarrow V(\lambda).

The following theorem (known as BGG-reciprocity) was proved in [10].

Theorem.

Given λ0𝔥\lambda_{0}\in\mathfrak{h}^{*} there exist λ1,,λr𝔥\lambda_{1},\dots,\lambda_{r}\in\mathfrak{h}^{*} such that the module P(λ0)P(\lambda_{0}) has a decreasing filtration P0=P(λ0)P1P2PrPr+1={0},P_{0}=P(\lambda_{0})\supseteq P_{1}\supseteq P_{2}\supseteq\cdots\supseteq P_{r}\supseteq P_{r+1}=\{0\}, and

Pi/Pi+1M(λi), 0ir.P_{i}/P_{i+1}\cong M(\lambda_{i}),\ \ 0\leq i\leq r.

Moreover if we let [P(λ):M(μ)][P(\lambda):M(\mu)] be the multiplicity of M(μ)M(\mu) in this filtration then we have [P(λ):M(μ)]=[M(μ):V(λ)].[P(\lambda):M(\mu)]=[M(\mu):V(\lambda)].

Remark.

Although the filtration is not unique in general, a comparison of formal characters shows that the filtration length and the multiplicity [P(λ),M(μ)][P(\lambda),M(\mu)] (see [70, Section 3.7]) is independent of the choice of the filtration.

More generally a module in 𝒪\cal O which admits a decreasing sequence of submodules where the successive quotients are Verma modules is said to admit a standard filtration. In the rest of this section we shall discuss an analog of this result for current algebras.

We will also explore other ideas stemming from the formal similarity between 𝒪\cal O and 𝒢\cal G. For instance it is known that dimHom𝔤(M(λ),M(μ))1\dim\operatorname{Hom}_{\mathfrak{g}}(M(\lambda),M(\mu))\leq 1 and that any non-zero map between Verma modules is injective and we shall discuss its analog for current algebras. We shall also discuss an analog of tilting modules; in the category 𝒪\cal O these are defined to be modules which admit a filtration where the successive quotients are Verma modules and also a filtration where the successive quotients are the restricted duals of Verma modules. It is known that for each λ𝔥\lambda\in\mathfrak{h}^{*} there exists a unique indecomposable tilting module which contains a copy of M(λ)M(\lambda).

5.3. BGG reciprocity in 𝒢\cal G

In the category 𝒢\cal G the role of the Verma module is played by the global Weyl module. However, in general the global Weyl module W(λ,r)W(\lambda,r), λP+\lambda\in P^{+} does not have a unique finite-dimensional quotient in 𝒢\cal G; for instance the modules Wloc(λ,r)W_{\operatorname{loc}}(\lambda,r) and V(λ,r)V(\lambda,r) are usually not isomorphic and we have

W(λ,r)Wloc(λ,r)V(λ,r).W(\lambda,r)\twoheadrightarrow W_{\operatorname{loc}}(\lambda,r)\twoheadrightarrow V(\lambda,r).

However both quotients have a uniqueness property; Wloc(λ,r)W_{\operatorname{loc}}(\lambda,r) is unique in the sense that any finite-dimensional quotient of W(λ,r)W(\lambda,r) is actually a quotient of Wloc(λ,r)W_{\operatorname{loc}}(\lambda,r) and V(λ,r)V(\lambda,r) is the unique irreducible quotient of W(λ,r)W(\lambda,r). The further difference from the category 𝒪\cal O situation is that the global Weyl module is not of finite length. In spite of these differences, one is still able to formulate the appropriate version of BGG-reciprocity. Such a formulation was first conjectured in [9] and proved there for 𝔰𝔩2[t]\mathfrak{sl}_{2}[t]. The result was proved in complete generality in [27] for twisted and untwisted current algebras; as usual the case of A2n(2)A_{2n}^{(2)} is much more difficult and one has to work with the hyperspecial current algebra. A key ingredient in the proof is to relate the character of the local Weyl module to specializations of (non)symmetric Macdonald polynomials (see Section 5.4.1 for a brief review).

The following is the main result of [27].

Theorem.

Let (λ,r)P+×+(\lambda,r)\in P^{+}\times\mathbb{Z}_{+}. The module P(λ,r)P(\lambda,r) admits a decreasing series of submodules: P0=P(λ,r)P1P2P_{0}=P(\lambda,r)\supseteq P_{1}\supseteq P_{2}\supseteq\cdots such that

Pi/Pi+1W(μi,si),forsome(μi,si)P+×+,P_{i}/P_{i+1}\cong W(\mu_{i},s_{i}),\ \ {\rm{for\ some}}\ \ (\mu_{i},s_{i})\in P^{+}\times\mathbb{Z}_{+},

and

[P(λ,r):W(μi,si)]=[Wloc(μi,si):V(λ,r)].[P(\lambda,r):W(\mu_{i},s_{i})]=[W_{\operatorname{loc}}(\mu_{i},s_{i}):V(\lambda,r)].

5.3.1. Tilting modules

We discuss the construction of tilting modules and some of their properties. These ideas were developed in [5, 6, 7] and one works in a suitable subcategory of 𝒢\cal G. Thus, let 𝒢bdd\mathcal{G}_{\mathrm{bdd}} be the full subcategory of objects MM of 𝒢\mathcal{G} such that M[j]=0M[j]=0 for all j0j\gg 0 and

wt(M)i=1sconvWμi,μ1,,μsP+\mathrm{wt}(M)\subseteq\bigcup_{i=1}^{s}\mathrm{conv}\ W\mu_{i},\ \ \mu_{1},\dots,\mu_{s}\in P^{+}

where convWμ\mathrm{conv}\ W\mu denotes the convex hull of the Weyl group orbit WμW\mu. An object MM in the category 𝒢bdd\mathcal{G}_{\mathrm{bdd}} is called tilting if it admits two increasing filtrations:

M0M1Mr,M0M1Mr\displaystyle M_{0}\subseteq M_{1}\subseteq\cdots\subseteq M_{r}\subseteq\cdots,\ \ \ \ \ M^{0}\subseteq M^{1}\subseteq\cdots\subseteq M^{r}\subseteq\cdots
M=r0Mr=r0Mr,\displaystyle M=\bigcup_{r\geq 0}M_{r}=\bigcup_{r\geq 0}M^{r},

such that Mi+1/MiM_{i+1}/M_{i} (resp. Mi+1/Mi)M^{i+1}/M^{i}) is isomorphic to a finite direct sum of modules of the form Wloc(λ,r)W_{\operatorname{loc}}(\lambda,r) (resp. to a sum of dual global Weyl modules W(λ,r)W(\lambda,r)^{*}) where (λ,r)P+×(\lambda,r)\in P^{+}\times\mathbb{Z}. One can also work with a dual definition of tilting modules, where one requires that the module has decreasing filtrations and the successive quotients are isomorphic to the dual local Weyl modules and the global Weyl modules respectively.

The following was proved in [6, Section 2].

Theorem.

For (λ,r)P+×(\lambda,r)\in P^{+}\times\mathbb{Z} there exists an indecomposable tilting module T(λ,r)T(\lambda,r) in 𝒢bdd\mathcal{G}_{\mathrm{bdd}} which maps onto the local Weyl module Wloc(λ,r)W_{\operatorname{loc}}(\lambda,r) and such that

τrT(λ,0)=T(λ,r),T(λ,r)T(μ,s)(λ,r)=(μ,s).\tau_{r}T(\lambda,0)=T(\lambda,r),\ \ T(\lambda,r)\cong T(\mu,s)\Leftrightarrow(\lambda,r)=(\mu,s).

Any indecomposable tilting module in 𝒢bdd\mathcal{G}_{\mathrm{bdd}} is isomorphic to T(μ,s)T(\mu,s) for some (μ,s)P+×(\mu,s)\in P^{+}\times\mathbb{Z} and any tilting module in 𝒢bdd\mathcal{G}_{\mathrm{bdd}} is isomorphic to a direct sum of indecomposable tilting modules. ∎

The proof of the theorem relies on the following necessary and sufficient condition for an object of 𝒢bdd\mathcal{G}_{\mathrm{bdd}} to admit a filtration by dual global Weyl modules. Namely:

MM admits a filtration by costandard modules if and only if Ext𝒢1(Wloc(μ,s),M)=0,(μ,s)P+×.\mathrm{Ext}^{1}_{\mathcal{G}}(W_{\mathrm{loc}}(\mu,s),M)=0,\ \forall(\mu,s)\in P^{+}\times\mathbb{Z}.

Remark.

This equivalence was established in [6] in the case when 𝔤\mathfrak{g} is of type AA. This is because the proof depended on knowing Theorem Theorem which at that time had only been proved when 𝔤\mathfrak{g} is of type AA. However the proof given there goes through verbatim for any 𝔤\mathfrak{g}.

5.3.2. Tilting modules for 𝔰𝔩2\mathfrak{sl}_{2} and in Serre subcategories

The existence of tilting modules is proved in a very abstract way. In [7] an explicit realization of the dual modules was given in the case of 𝔰𝔩2\mathfrak{sl}_{2}. In this case we identify P+P^{+} with +\mathbb{Z}_{+}. Recall also the algebra 𝔸λ\mathbb{A}_{\lambda} defined in Section 5.1.4; in this special case it is just the ring of symmetric polynomials in λ\lambda-variables.

Theorem.

Suppose that 𝔤=𝔰𝔩2\mathfrak{g}=\mathfrak{sl}_{2} and let (λ,r)+×+.(\lambda,r)\in\mathbb{Z}_{+}\times\mathbb{Z}_{+}. The module T(λ,r)T(\lambda,r)^{*} is a free right 𝔸λ\mathbb{A}_{\lambda}-module and T(λ,r)τrT(λ,0)T(\lambda,r)\cong\tau_{r}\ T(\lambda,0). Moreover,

T(1,0)W(1,0),T(λ,0)τrλλW(1,0),rλ=(λ2),λ2.T(1,0)^{*}\cong W(1,0),\ \ T(\lambda,0)^{*}\cong\tau_{-r_{\lambda}}\ {\bigwedge}^{\lambda}W(1,0),\ \ r_{\lambda}=\binom{\lambda}{2},\ \ \lambda\geq 2.
chgrT(λ,0)=s=0λ2ts(sλ)(1:t)schgrW(λ2s,0)=trλ(1:t)λchgrWloc(λ,rλ)\mathrm{ch}_{\mathrm{gr}}T(\lambda,0)^{*}=\sum_{s=0}^{\lfloor\frac{\lambda}{2}\rfloor}t^{s(s-\lambda)}(1:t)_{s}\operatorname{ch}_{\mathrm{gr}}W(\lambda-2s,0)=t^{r_{\lambda}}(1:t)_{\lambda}\operatorname{ch}_{\mathrm{gr}}W_{\mathrm{loc}}(\lambda,r_{\lambda})^{*}

where

(1:t)n=1(1t)(1t2)(1tn)(1:t)_{n}=\frac{1}{(1-t)(1-t^{2})\cdots(1-t^{n})}

Very little is known about the structure or the character of the tilting modules in general.

A theory of tilting modules was also developed for Serre subcategories of 𝒢\mathcal{G} which are defined as follows. Given a subset ΓP+×\Gamma\subseteq P^{+}\times\mathbb{Z}, we define a full subcategory 𝒢(Γ)\mathcal{G}(\Gamma) whose objects MM satisfy additionally

[M:V(λ,r)]0(λ,r)Γ.[M:V(\lambda,r)]\neq 0\Rightarrow(\lambda,r)\in\Gamma.

The category 𝒢(Γ)bdd\mathcal{G}(\Gamma)_{\mathrm{bdd}} is now defined in an obvious way. If Γ=P+×J\Gamma=P^{+}\times J, where JJ is an (possibly infinite) interval in \mathbb{Z}, then the existence of tilting modules holds with 𝒢bdd\mathcal{G}_{\mathrm{bdd}} and P+×P^{+}\times\mathbb{Z} replaced by 𝒢(Γ)bdd\mathcal{G}(\Gamma)_{\mathrm{bdd}} and Γ\Gamma respectively (see [5, Proposition 4.2 and Theorem 4.3]). The local and dual global Weyl modules in this setting are obtained by applying a certain natural functor to the standard and costandard modules in 𝒢\mathcal{G}.

5.3.3. Socle and Radical Filtration for local Weyl modules

The local Weyl module Wloc(λ)W_{\mathrm{loc}}(\lambda) has a natural increasing grading filtration induced from its graded module structure. This filtration coincides with the radical filtration (see [77, Proposition 3.5]) which is defined as follows. For a module MM of 𝐔(𝔤[t])\mathbf{U}(\mathfrak{g}[t]) the radical filtration is given by

radk(M)rad1(M)rad0(M)=M\cdots\subseteq\mathrm{rad}^{k}(M)\subseteq\cdots\subseteq\mathrm{rad}^{1}(M)\subseteq\mathrm{rad}^{0}(M)=M

where rad(M)\mathrm{rad}(M) is the smallest submodule of MM such that the quotient M/rad(M)M/\mathrm{rad}(M) is semi-simple and radk(M)\mathrm{rad}^{k}(M) is defined inductively by

radk(M)=rad(radk1(M)).\mathrm{rad}^{k}(M)=\mathrm{rad}(\mathrm{rad}^{k-1}(M)).

In particular,

rad1(Wloc(λ))=s>0𝐔(𝔤[t])[s]vλ.\mathrm{rad}^{1}(W_{\mathrm{loc}}(\lambda))=\bigoplus_{s>0}\mathbf{U}(\mathfrak{g}[t])[s]v_{\lambda}.

There is another natural filtration on a module MM, called the socle filtration. It is given as follows

0=soc0(M)soc1(M)sock(M),0=\mathrm{soc}^{0}(M)\subseteq\mathrm{soc}^{1}(M)\subseteq\cdots\subseteq\mathrm{soc}^{k}(M)\subseteq\cdots,

where soc(M)=soc1(M)\mathrm{soc}(M)=\mathrm{soc}^{1}(M) is the largest semi-simple submodule of MM and sock(M)\mathrm{soc}^{k}(M) is defined inductively by

sock(M)/sock1(M)=soc(M/sock1(M)).\mathrm{soc}^{k}(M)/\mathrm{soc}^{k-1}(M)=\mathrm{soc}(M/\mathrm{soc}^{k-1}(M)).

A module MM of 𝐔(𝔤[t])\mathbf{U}(\mathfrak{g}[t]) is called rigid if the socle filtration coincides with the radical filtration. This is in particular the case, if MM is a finite-dimensional graded module such that M/rad(M)M/\mathrm{rad}(M) and soc(M)\mathrm{soc}(M) are both simple. We remind the reader that when 𝔤\mathfrak{g} is of type ADEADE the local Weyl module is isomorphic to a level one Demazure module and hence embeds in a highest weight module for the affine Lie algebra. Given λP+\lambda\in P^{+}, let wW^w\in\widehat{W} be such that Λ=w1(w0λ+Λ0)P^+\Lambda=w^{-1}(w_{0}\lambda+\Lambda_{0})\in\widehat{P}^{+}. Since V(Λ)V(\Lambda) is an irreducible integrable module for 𝔤^\widehat{\mathfrak{g}}, it follows that any 𝔟^\widehat{\mathfrak{b}}-submodule of M=Vw(Λ)M=V_{w}(\Lambda) must contain the highest weight vector vΛv_{\Lambda}. Hence any 𝔤[t]\mathfrak{g}[t]-submodule of MM contains the 𝔤[t]\mathfrak{g}[t]-module U(𝔤[t])vΛU(\mathfrak{g}[t])v_{\Lambda}. In other words soc(M){\rm{soc}}(M) must be simple and we must have soc(M)=U(𝔤[t])vΛV(Λ|𝔥,0){\rm{soc}}(M)=U(\mathfrak{g}[t])v_{\Lambda}\cong V(\Lambda|_{\mathfrak{h}},0). So, we get

Lemma.

Let 𝔤\mathfrak{g} be of type ADEADE. Then soc(Wloc(λ))V(Λ|𝔥,0){\rm{soc}}(W_{\mathrm{loc}}(\lambda))\cong V(\Lambda|_{\mathfrak{h}},0). ∎

It was proved in [77] that when 𝔤\mathfrak{g} is of type ADEADE the local Weyl module is rigid. However, in general the socle of the local Weyl module is not simple and we give the counterexample given in [77, Example 3.12]. In type C2C_{2}, the socle of the local Weyl module Wloc(2ω1+ω2)W_{\mathrm{loc}}(2\omega_{1}+\omega_{2}) (the short root is α1\alpha_{1}) is isomorphic to V(0,3)V(ω2,2)V(0,3)\oplus V(\omega_{2},2).

5.3.4. Maps between local Weyl modules

We apply the discussion on the socle of Wloc(λ,r)W_{\operatorname{loc}}(\lambda,r) to study morphisms between local Weyl modules when 𝔤\mathfrak{g} is of type ADEADE. For the purposes of this section it will be convenient to think of Wloc(λ,r)W_{\operatorname{loc}}(\lambda,r) as modules for the subalgebra 𝔤[t]cd\mathfrak{g}[t]\oplus\mathbb{C}c\oplus\mathbb{C}d of 𝔤^\widehat{\mathfrak{g}} where we let cc act as 11 and the action of dd is given by the grading. Recall the Bruhat order on W^\widehat{W} given by uwu\leq w if some substring of some reduced word for ww is a reduced word for uu.

Proposition.

Assume that 𝔤\mathfrak{g} is of type ADEADE and let (λ,r),(μ,s)P+×(\lambda,r),(\mu,s)\in P^{+}\times\mathbb{Z}. Then,

dimHom𝒢(Wloc(λ,r),Wloc(μ,s))1\dim\mathrm{Hom}_{\mathcal{G}}(W_{\mathrm{loc}}(\lambda,r),W_{\mathrm{loc}}(\mu,s))\leq 1

with equality holding if and only if there exist w1,w2W^w_{1},w_{2}\in\widehat{W} and ΛP^+\Lambda\in\widehat{P}^{+} such that the following hold:

w2w1,w1(λ+Λ0+rδ)=Λ=w2(μ+Λ0+sδ).w_{2}\leq w_{1},\ \ w_{1}(\lambda+\Lambda_{0}+r\delta)=\Lambda=w_{2}(\mu+\Lambda_{0}+s\delta).

Moreover, any non-zero map between local Weyl modules is injective.

Proof.

Let φ:Wloc(λ,r)Wloc(μ,s)\varphi:W_{\mathrm{loc}}(\lambda,r)\rightarrow W_{\mathrm{loc}}(\mu,s) be a non-zero homomorphism. We first prove that this implies that there exist w1,w2W^w_{1},w_{2}\in\widehat{W} and ΛP^+\Lambda\in\widehat{P}^{+} with w1(λ+Λ0+rδ)=Λ=w2(μ+Λ0+sδ)w_{1}(\lambda+\Lambda_{0}+r\delta)=\Lambda=w_{2}(\mu+\Lambda_{0}+s\delta). To see this assume that w1(λ+Λ0+rδ)=Λw_{1}(\lambda+\Lambda_{0}+r\delta)=\Lambda and w2(μ+Λ0+sδ)=Λw_{2}(\mu+\Lambda_{0}+s\delta)=\Lambda^{\prime} with Λ,ΛP^+\Lambda,\Lambda^{\prime}\in\widehat{P}^{+} and let Wloc(μ,s)V(Λ)W_{\operatorname{loc}}(\mu,s)\hookrightarrow V(\Lambda^{\prime}) be the inclusion which exists since Wloc(μ,s)W_{\operatorname{loc}}(\mu,s) is isomorphic to a stable Demazure module in V(Λ)V(\Lambda^{\prime}). Since the image of φ\varphi is non-zero it must include the simple socle of Wloc(μ,s)W_{\operatorname{loc}}(\mu,s). This in turn implies that Λ\Lambda^{\prime} is a weight of Wloc(λ,r)V(Λ)W_{\operatorname{loc}}(\lambda,r)\hookrightarrow V(\Lambda). It follows that ΛΛ\Lambda-\Lambda^{\prime} must be a sum of affine positive roots. On the other hand since φ(wλ)0\varphi(w_{\lambda})\neq 0 it follows that λ+Λ0+rδ\lambda+\Lambda_{0}+r\delta must be a weight of V(Λ)V(\Lambda^{\prime}) and hence Λ\Lambda is also a weight of V(Λ)V(\Lambda^{\prime}). This forces ΛΛ\Lambda^{\prime}-\Lambda to be a sum of positive affine roots and also shows that Λ=Λ\Lambda=\Lambda^{\prime}. To see that φ\varphi is injective, we note that otherwise both the kernel and the cokernel of φ\varphi would have to contain vΛv_{\Lambda} which is absurd.

Finally to see that the dimension of the homomorphism space is at most one, it suffices to note that dimV(Λ)wΛ=1\dim V(\Lambda)_{w\Lambda}=1 for all wW^w\in\widehat{W}.

5.3.5. Morphisms between global Weyl modules

The study of the homomorphism space between global Weyl modules has also been studied in [8] and confirms further the phenomenon that the global Weyl module plays a role similar to that of the Verma modules in category 𝒪\mathcal{O}. The following result can be found in [8, Theorem 3].

Theorem.

Let λ,μP+\lambda,\mu\in P^{+} and assume that μ(hi)=0\mu(h_{i})=0 for all iIi\in I with ωi(hθ)1\omega_{i}(h_{\theta})\neq 1. Then

Hom𝒢(W(λ),W(μ))=0,if λμ,Hom𝒢(W(μ),W(μ))𝔸μ.\mathrm{Hom}_{\mathcal{G}}(W(\lambda),W(\mu))=0,\ \ \text{if }\lambda\neq\mu,\ \ \mathrm{Hom}_{\mathcal{G}}(W(\mu),W(\mu))\cong\mathbb{A}_{\mu}.

Moreover any non-zero map φ:W(μ)W(μ)\varphi:W(\mu)\to W(\mu) is injective. ∎

The restriction on μ\mu is necessary (see [8, Remark 6.1]). For instance, in types BnB_{n} and DnD_{n} (n6n\geq 6) we have Hom𝒢(W(ω2),W(ω4))0\mathrm{Hom}_{\mathcal{G}}(W(\omega_{2}),W(\omega_{4}))\neq 0. However the second statement, namely the injectivity of any non-zero map, is still expected to hold in general.

Remark.

This theorem is quite unlike the analogous theorem for local Weyl modules which was discussed in the preceding section.

5.4. Generalized Weyl modules, Global Demazure modules and other directions.

We now discuss generalizations of some of the ideas presented earlier in this section. This is a brief and far from complete discussion of the papers of [52], [46], [47] and we refer the interested readers to those papers for greater detail. We begin by elucidating the connection between local Weyl modules and specializations of Macdonald polynomials which was briefly mentioned in Section 5.3. These polynomials are those associated with (anti) dominant weights. We then discuss the work of [52] who introduced the notion of generalized Weyl modules for 𝔫^+\widehat{\mathfrak{n}}^{+} and showed that their characters are again related to specializations of Macdonald polynomials associated with any integral weight.

We then move on to discuss the notion of global Demazure modules introduced by Dumanski and Feigin and state a few open problems regarding the homomorphism spaces between these objects. The aim is to generalize the global-local picture of Weyl modules for wider families of modules and develop some modifications of results in this broader setting.

5.4.1. Local Weyl modules, Generalized Weyl modules and Macdonald polynomials

Let

q,t=(q,t)[eλ:λP]and q=(q)[eλ:λP]\mathcal{R}_{q,t}=\mathbb{Q}(q,t)[e_{\lambda}:\lambda\in P]\ \ \text{and }\ \mathcal{R}_{q}=\mathbb{Q}(q)[e_{\lambda}:\lambda\in P]

respectively be the group algebra of the weight lattice with coefficients in (q,t)\mathbb{Q}(q,t) and (q)\mathbb{Q}(q) respectively. Consider Rq,tWR_{q,t}^{W} the subring of WW-invariants where the action is induced from the action of WW on PP and define RqWR_{q}^{W} similarly. For fRq,tWf\in R_{q,t}^{W}, we denote by [f][f] its constant term (i.e. the coefficient in front of e0e_{0}) and set

(q,t)=α(R++δ)1eα1t1eα,eδ=q1,Δ(q,t)=(q,t)[(q,t)].\displaystyle\nabla(q,t)=\prod_{\alpha\in(R\ +\ \mathbb{Z}_{+}\delta)}\frac{1-e_{\alpha}}{1-t^{-1}e_{\alpha}},\ \ e_{\delta}=q^{-1},\ \ \Delta(q,t)=\frac{\nabla(q,t)}{[\nabla(q,t)]}.

This ring Rq,tWR_{q,t}^{W} and RqWR_{q}^{W} both admit a scalar product

f,gq,t=[fg¯Δ(q,t)]f,gRq,tW,f,gq=[fι(g)Δ(q,)]f,gRqW\langle f,g\rangle_{q,t}=[f\overline{g}\Delta(q,t)]\ \ f,g\in R_{q,t}^{W},\ \ \ \langle f,g\rangle_{q}=[f\iota(g)\Delta(q,\infty)]\ \ f,g\in R_{q}^{W}

where ¯\overline{\cdot} is the involution on Rq,tWR_{q,t}^{W} given by tt1,qq1,eλeλt\mapsto t^{-1},\ q\mapsto q^{-1},\ e_{\lambda}\mapsto e_{-\lambda} and ι\iota is the involution of q\mathcal{R}_{q} fixing qq and mapping eλe_{\lambda} to ewλe_{-w_{\circ}\lambda}. Moreover, we have a natural basis {mλ}λP+\{m_{\lambda}\}_{\lambda\in P^{+}} given by

mλ(q,t)=μWλeμm_{\lambda}(q,t)=\sum_{\mu\in W\cdot\lambda}e_{\mu}

The symmetric Macdonald polynomials {Pλ(q,t)}λP+\{P_{\lambda}(q,t)\}_{\lambda\in P^{+}} are uniquely defined by the following two properties

  1. (1)

    Pλ(q,t)=mλ(q,t)+μ<λμP+cλ,μmμ(q,t),cλ,μ(q,t),\displaystyle P_{\lambda}(q,t)=m_{\lambda}(q,t)+\sum_{\begin{subarray}{c}\mu<\lambda\\ \mu\in P^{+}\end{subarray}}c_{\lambda,\mu}m_{\mu}(q,t),\ \ c_{\lambda,\mu}\in\mathbb{Q}(q,t),

  2. (2)

    Pλ(q,t),Pμ(q,t)q,t=0,λμ\langle P_{\lambda}(q,t),P_{\mu}(q,t)\rangle_{q,t}=0,\ \ \lambda\neq\mu

These polynomials have the property that the limit tt\rightarrow\infty exists which we denote by Pλ(q,)=limtPλ(q,t)P_{\lambda}(q,\infty)=\lim_{t\rightarrow\infty}P_{\lambda}(q,t). The following result can be found in [27, Theorem 4.2].

Theorem.

The family {Pλ(q,)}λP+\{P_{\lambda}(q,\infty)\}_{\lambda\in P^{+}} forms an orthogonal basis of RqWR^{W}_{q} with respect to the form ,q\langle\cdot,\cdot\rangle_{q}. Moreover

Pλ(q,)=chgrWloc(λ),λP+P_{\lambda}(q,\infty)=\operatorname{ch}_{\mathrm{gr}}W_{\operatorname{loc}}(\lambda),\ \ \lambda\in P^{+}

In the case when 𝔤\mathfrak{g} is simply-laced it was already proved in [71] that the graded character of the stable Demazure module was given by the specialization of the Macdonald polynomial as in the above theorem; recall the connection between local Weyl modules and stable Demazure modules first made in [30] in the case of 𝔰𝔩n+1\mathfrak{sl}_{n+1} and then in [58] for 𝔤\mathfrak{g} simply-laced.

At that time it was also known that this formula could not hold when 𝔤\mathfrak{g} was not simply-laced. In the non-simply laced case the result of [88] showed that the local Weyl module had a flag where the successive quotients were stable Demazure modules. The corresponding results for the twisted current algebras were studied in [57]. However in the case of the twisted A2n(2)A_{2n}^{(2)} one has to work with a different current algebra [28], called the hyperspecial current algebra.

5.4.2. Nonsymmetric Macdonald polynomials

There is another family of polynomials {Eλ(q,t)}λP\{E_{\lambda}(q,t)\}_{\lambda\in P} indexed by the weight lattice called the nonsymmetric Macdonald polynomials. They were introduced by Opdam [93] and Cherednik [43]. First we define a new order on the set of weights P.P. Consider the level one action of W^\widehat{W} on 𝔥\mathfrak{h}^{*} defined as follows (the action differs only for s0s_{0})

s0μ:=sθ(μ)+θ,μ𝔥.s_{0}\circ\mu:=s_{\theta}(\mu)+\theta,\ \mu\in\mathfrak{h}^{*}.

Given λP\lambda\in P, we denote by wλw_{\lambda} the unique minimal length element of W^\widehat{W} such that wλλw_{\lambda}\circ\lambda is either miniscule or zero. For λ,μP\lambda,\mu\in P, we say μ<bλ\mu<_{b}\lambda if and only if wμ<wλw_{\mu}<w_{\lambda} with respect to the Bruhat order.

Again we can define a scalar product (,)q,t(\cdot,\cdot)_{q,t} on q,t\mathcal{R}_{q,t} and the family {Eλ(q,t)}λP\{E_{\lambda}(q,t)\}_{\lambda\in P} is uniquely determined by the following two properties

  1. (1)

    Eλ(q,t)=eλ+μ<bλcλ,μeμ,cλ,μ(q,t)E_{\lambda}(q,t)=e_{\lambda}+\sum\limits_{\mu<_{b}\lambda}c_{\lambda,\mu}e_{\mu},\,\,\,c_{\lambda,\mu}\in\mathbb{Q}(q,t),

  2. (2)

    (Eλ(q,t),eμ)q,t=0ifμ<bλ.(E_{\lambda}(q,t),e_{\mu})_{q,t}=0\ \ \text{if}\ \mu<_{b}\lambda.

For a dominant weight λ\lambda we have Pλ(q,)=limtEwλ(q,t)=Ewλ(q,)P_{\lambda}(q,\infty)=\lim_{t\rightarrow\infty}E_{w_{\circ}\lambda}(q,t)=E_{w_{\circ}\lambda}(q,\infty) and hence by the above theorem the characters of local Weyl modules appear also as specializations of nonsymmetric Macdonald polynomials for anti-dominant weights

Ewλ(q,)=chgrWloc(λ),λP+E_{w_{\circ}\lambda}(q,\infty)=\operatorname{ch}_{\mathrm{gr}}W_{\operatorname{loc}}(\lambda),\ \ \lambda\in P^{+}

The natural question is whether other specializations are also meaningful in the sense that they have a representation theoretic interpretation. This leads to the definition of generalized local Weyl modules which can be found in [52].

The reader should be warned that there are several notions of generalized Weyl modules, e.g. in [25, 56, 78] when the polynomial algebra is replaced by an arbitrary commutative algebra. But these are not the modules under consideration in this discussion.

Definition.

Given μP\mu\in P let WμW_{\mu} be the 𝔫^+\widehat{\mathfrak{n}}^{+}-module generated by vμv_{\mu} with relations,

(htr+1)vμ=0,r0,(xα+1)max{μ(hα),0}+1vμ=0,(xαt)max{μ(hα),0}+1vμ=0,αR+.(h\otimes t^{r+1})v_{\mu}=0,\ r\geq 0,\ \ \ (x^{+}_{\alpha}\otimes 1)^{\max\{-\mu(h_{\alpha}),0\}+1}v_{\mu}=0,\ \ (x^{-}_{\alpha}\otimes t)^{\max\{\mu(h_{\alpha}),0\}+1}v_{\mu}=0,\ \ \alpha\in R^{+}.

These are called the generalized local Weyl modules.

Note that for anti-dominant weights we obviously have WμWloc(wμ)W_{\mu}\cong W_{\operatorname{loc}}(w_{\circ}\mu) as 𝔫^+\widehat{\mathfrak{n}}^{+}-modules and hence the character is again a specialized nonsymmetric Macdonald polynomial. The characters of WλW_{\lambda} for λP+\lambda\in P^{+} are related to the Orr–Shimozono specialization of Ew0λ(q,t)E_{w_{0}\lambda}(q,t). The first part of the next proposition is proved in [94] and the second part in [52].

Proposition.

Let λP+\lambda\in P^{+}.

  1. (1)

    The limit Ewλ(q1,0):=limt0Ewλ(q1,t)E_{w_{\circ}\lambda}(q^{-1},0):=\lim_{t\to 0}E_{w_{\circ}\lambda}(q^{-1},t) exists and admits an explicit combinatorial formula in terms of quantum alcove paths.

  2. (2)

    The character of WλW_{\lambda} is given by wEwλ(q1,0)w_{\circ}E_{w_{\circ}\lambda}(q^{-1},0).

5.4.3. Recovering the global Weyl module from the local Weyl module

We recall a general construction which was first introduced in [8]. Namely, if VV is any 𝔤[t]\mathfrak{g}[t]-module, one can define an action of 𝔤[t]\mathfrak{g}[t] on V[t]:=V[t]V[t]:=V\otimes\mathbb{C}[t], by

(xtr)(vts)=j=0r(rj)((xtrj)v)ts+j.(x\otimes t^{r})(v\otimes t^{s})=\sum_{j=0}^{r}\binom{r}{j}((x\otimes t^{r-j})v)\otimes t^{s+j}.

In fact it was introduced in a more general context in Section 4 of [8] by replacing [t]\mathbb{C}[t] by any commutative associative Hopf algebra AA. Notice that if VV is generated by an element vv then V[t]V[t] is generated by v1v\otimes 1. It was shown also that the fundamental global Weyl modules could be realized in this way by taking VV to be Wloc(ωi)W_{\operatorname{loc}}(\omega_{i}) for some iIi\in I. Moreover, it was shown in Proposition 6.2 of that paper that if μ=i=1nsiωi\mu=\sum_{i=1}^{n}s_{i}\omega_{i} and μ(hi)=0\mu(h_{i})=0 if ωi(hθ)1\omega_{i}(h_{\theta})\neq 1 then there exists an injective map

W(μ)W(ω1)s1W(ωn)sn,W(\mu)\to W(\omega_{1})^{\otimes s_{1}}\otimes\cdots\otimes W(\omega_{n})^{\otimes s_{n}},

extending the assignment wμwω1s1wωnsnw_{\mu}\to w_{\omega_{1}}^{s_{1}}\otimes\cdots\otimes w_{\omega_{n}}^{s_{n}}. This result was then established without the restriction on μ\mu in [74] by different methods.

5.4.4. A generalization

Suppose now that W1,,WrW_{1},\dots,W_{r} are graded 𝔤[t]\mathfrak{g}[t]-modules with generators w1,,wrw_{1},\dots,w_{r}. Then the associated global module is defined to be the submodule of W1[t]Wr[t]W_{1}[t]\otimes\cdots\otimes W_{r}[t] generated by the tensor product of (w11)(wr1)(w_{1}\otimes 1)\otimes\cdots\otimes(w_{r}\otimes 1). This notion was introduced in [53, Section 1.3] and the resulting module is denoted as R(W1,,Wr)R(W_{1},\dots,W_{r}). In the case when the additional relations

(𝔥t[t])wi=0,hwi=λi(h)wi,(\mathfrak{h}\otimes t\mathbb{C}[t])w_{i}=0,\ \ hw_{i}=\lambda_{i}(h)w_{i}, (5.1)

hold we can define (as in the case of global Weyl modules) a right action of 𝐔(𝔥[t])\mathbf{U}(\mathfrak{h}[t]) on R(W1,,Wr)R(W_{1},\dots,W_{r}). The algebra 𝒜(λ1,,λr)\mathcal{A}(\lambda_{1},\dots,\lambda_{r}) is as usual the quotient of 𝐔(𝔥[t]+)\mathbf{U}(\mathfrak{h}[t]_{+}) by the annihilating ideal of the cyclic vector (w11)(wr1)(w_{1}\otimes 1)\otimes\cdots\otimes(w_{r}\otimes 1). This algebra is harder to understand although one does have an embedding

𝒜(λ1,,λr)i=1r𝒜(λi),𝒜(λ𝒾)[𝓏𝒾]\mathcal{A}(\lambda_{1},\dots,\lambda_{r})\hookrightarrow\bigotimes_{i=1}^{r}\mathcal{A}(\lambda_{i}),\ \ \cal A(\lambda_{i})\cong\mathbb{C}[z_{i}]

given by

(htm)i=1r(1htm1)λ1(h)z1m++λr(h)zrm.(h\otimes t^{m})\mapsto\sum_{i=1}^{r}(1\otimes\cdots\otimes ht^{m}\otimes\cdots\otimes 1)\mapsto\lambda_{1}(h)z_{1}^{m}+\cdots+\lambda_{r}(h)z_{r}^{m}.

Given uu\in\mathbb{C}, set

Wi(u)=Wi[t]𝒜(λ𝒾),W_{i}(u)=W_{i}[t]\otimes_{\cal A(\lambda_{i})}\mathbb{C},

where we regard \mathbb{C} as an 𝒜(λ𝒾)\cal A(\lambda_{i})-module by letting ziz_{i} act as uu.

The following was proved in [46], [47].

Theorem.

Let W1,,WrW_{1},\dots,W_{r} be as above.

  • (i)

    The algebra 𝒜(λ1,,λr)\mathcal{A}(\lambda_{1},\dots,\lambda_{r}) is isomorphic to the subalgebra in [z1,,zr]\mathbb{C}[z_{1},\dots,z_{r}] generated by the polynomials λ1(h)z1m++λr(h)zrm,h𝔥,m0\lambda_{1}(h)z_{1}^{m}+\cdots+\lambda_{r}(h)z_{r}^{m},\ h\in\mathfrak{h},\ m\geq 0.

  • (ii)

    There exists a nonempty Zariski open subset UU of r\mathbb{C}^{r} (in particular 0U0\notin U) such that for all 𝕦=(u1,,ur)U\mathbb{u}=(u_{1},\dots,u_{r})\in U

    R(W1,,Wr)𝒜(λ1,,λr)𝐮𝔤[t]i=1rWi(ui)R(W_{1},\cdots,W_{r})\otimes_{\mathcal{A}(\lambda_{1},\dots,\lambda_{r})}\mathbb{C}_{\mathbf{u}}\ \ \cong_{\mathfrak{g}[t]}\ \ \bigotimes_{i=1}^{r}W_{i}(u_{i})

    where 𝐮\mathbb{C}_{\mathbf{u}} denotes the quotient of 𝒜(λ1,,λr)\mathcal{A}(\lambda_{1},\dots,\lambda_{r}) by the maximal ideal corresponding to 𝐮\mathbf{u}.

  • (iii)

    The module R(W1,,Wr)R(W_{1},\dots,W_{r}) is a finitely-generated 𝒜(λ1,,λr)\mathcal{A}(\lambda_{1},\dots,\lambda_{r})-module.

5.4.5.

The following conjecture of [46] generalizes the known results for local Weyl modules.

Conjecture.

The 𝔤[t]\mathfrak{g}[t]-module

R(W1,,Wr)𝒜(λ1,,λr)0R(W_{1},\ldots,W_{r})\otimes_{\mathcal{A}(\lambda_{1},\dots,\lambda_{r})}\mathbb{C}_{0}

is isomorphic to the fusion product of the modules Wi(ci)W_{i}(c_{i}) for (c1,,cr)(c_{1},\dots,c_{r}) in some Zariski open subset of r\mathbb{C}^{r}.

In [46] the authors prove this conjecture for a certain families of Demazure modules when 𝔤\mathfrak{g} is of type ADEADE, and in [47] they drop the assumption on the type of 𝔤\mathfrak{g}. Given a collection of dominant integral weights λ¯=(λ1,,λr)\underline{\lambda}=(\lambda_{1},\dots,\lambda_{r}) we set

𝔻,λ¯=R(D(,λ1),,D(,λr))\mathbb{D}_{\ell,\ell\underline{\lambda}}=R(D(\ell,\ell\lambda_{1}),\dots,D(\ell,\ell\lambda_{r}))

and let vv be the generating vector of 𝔻,λ¯\mathbb{D}_{\ell,\ell\underline{\lambda}}. The following theorem has been proved for the tuple λ¯=(ω1,,ω1,,ωn,,ωn)\underline{\lambda}=(\omega_{1},\dots,\omega_{1},\dots,\omega_{n},\dots,\omega_{n}) by Dumanski–Feigin and extended later by Dumanski–Feigin–Finkelberg to arbitrary tuples.

Theorem.

Let λP+\lambda\in P^{+} and λ¯=(λ1,,λr)\underline{\lambda}=(\lambda_{1},\cdots,\lambda_{r}) be such that λ=i=1rλi\lambda=\sum_{i=1}^{r}\lambda_{i}. Then, we have an isomorphism

D(,λ)𝔻,λ¯𝒜(λ1,,λr)0.D(\ell,\ell\lambda)\rightarrow\mathbb{D}_{\ell,\ell\underline{\lambda}}\otimes_{\mathcal{A}(\ell\lambda_{1},\dots,\ell\lambda_{r})}\mathbb{C}_{0}.

An interesting question would be to determine the generators and relations for global Demazure modules.

Remark.

Dumanski–Feigin–Finkelberg also prove that 𝔻,λ¯\mathbb{D}_{\ell,\ell\underline{\lambda}} is free over 𝒜(λ1,,λr)\mathcal{A}(\ell\lambda_{1},\dots,\ell\lambda_{r}) and that there exists a tensor product decomposition

𝔻,(λ¯μ¯)(c,d)(𝔻,λ¯c)(𝔻,μ¯d)\mathbb{D}_{\ell,\ell(\underline{\lambda}\cup\underline{\mu})}\otimes\mathbb{C}_{(c,d)}\cong(\mathbb{D}_{\ell,\ell\underline{\lambda}}\otimes\mathbb{C}_{c})\otimes(\mathbb{D}_{\ell,\ell\underline{\mu}}\otimes\mathbb{C}_{d})

provided that cc and dd have no common entries. This is analogous to the well-known factorization of local Weyl modules which was proved in [38].

An interesting direction of research would be to study the homomorphisms between global Demazure modules and observe the analogues to homomorphisms between global Weyl modules discussed earlier in this section.

5.5.

As we mentioned earlier the current algebra is the derived algebra of the standard maximal parabolic subalgebra of the affine Lie algebra. A natural problem is to develop an analogous theory for other parabolic subalgebras. This has been attacked for the first time in [29] and the two important families of local and global Weyl modules have been intensively studied, but many problems are still open. The global Weyl modules cotinue to be parametrized by dominant integral weights of a semi-simple subalgebra of 𝔤\mathfrak{g} depending on the choice of the maximal parabolic algebra. However, the following interesting differences appear.

  • The algebra 𝔸λ\mathbb{A}_{\lambda} (modulo its Jacobson radical) is a Stanley–Reisner ring; in particular it has relations and is not a polynomial algebra (see [29, Theorem 1]).

  • The algebra 𝔸λ\mathbb{A}_{\lambda} and the global Weyl module can be finite-dimensional and this happens if and only if 𝔸λ\mathbb{A}_{\lambda} is a local ring.

  • The global Weyl module is not a free 𝔸λ\mathbb{A}_{\lambda} module in general. However we expect the global Weyl module to be free over a suitable quotient algebra of 𝔸λ\mathbb{A}_{\lambda} corresponding to the coordinate ring of one of the irreducible subvarieties of 𝔸λ\mathbb{A}_{\lambda}.

The dimension of the local Weyl module depends on the choice of the maximal ideal of 𝔸λ\mathbb{A}_{\lambda}. This was in the current algebra case one of the key observations together with the Quillen–Suslin theorem to obtain the freeness of global Weyl modules. It is still an open and interesting question to find the maximal ideals of 𝔸λ\mathbb{A}_{\lambda} producing the local Weyl modules of maximal dimension. An example has been discussed in [29, Section 7.1].

6. Fusion product Decompositions, Demazure flags and connections to combinatorics and Hypergeometric series

In this section we collect together several results on Demazure modules which are of independent interest.

6.1. Demazure modules revisited

6.1.1. A simplified presentation of Demazure modules.

Recall that following [72, 85], we gave in Section 2.3.3 a presentation of Demazure modules involving infinitely many relations. On the other hand we also discussed in Section Corollary that when 𝔤\mathfrak{g} is simply-laced the local Weyl module Wloc(μ,r)W_{\operatorname{loc}}(\mu,r) is isomorphic to a Demazure module occuring in a level one highest weight representation. The local Weyl module by definition has only finitely many relations. It turns out that this remains true for arbitrary Demazure modules. The following result was first proved in [42] for 𝔤\mathfrak{g}-stable Demazure modules (see Proposition Proposition) and was recently proved for arbitrary Demazure modules in [80].

Theorem.

Suppose that (λ,w)P^+×W^(\lambda,w)\in\widehat{P}^{+}\times\widehat{W} and assume that λ(c)=\lambda(c)=\ell, λ(d)=r\lambda(d)=r and wλ|𝔥=μw\lambda|_{\mathfrak{h}}=\mu. The module Vw(λ)V_{w}(\lambda) is isomorphic to a cyclic 𝐔(𝔟^)\mathbf{U}(\widehat{\mathfrak{b}})-module generated by a non-zero vector vv with the following relations:

(hts)v=δs,0μ(h)v, for all h𝔥,dv=rv,cv=v,\displaystyle(h\otimes t^{s})v=\delta_{s,0}\cdot\mu(h)v,\text{ for all }h\in\mathfrak{h},\,\,dv=rv,\,\,cv=\ell v,

and for αR(μ)={αR+:μ(hα)+}\alpha\in R^{\mp}(\mu)=\{\alpha\in R^{+}:\mu(h_{\alpha})\in\mp\mathbb{Z}_{+}\} we have

(xα±tsα±1)mα±+1v=0,if mα±<dα;(xα±tsα±)v=0,\big{(}x_{\alpha}^{\pm}\otimes t^{s^{\pm}_{\alpha}-1}\big{)}^{m^{\pm}_{\alpha}+1}v=0,\ \text{if $m^{\pm}_{\alpha}<d_{\alpha}\ell$};\ \ \big{(}x_{\alpha}^{\pm}\otimes t^{s^{\pm}_{\alpha}}\big{)}v=0,
(xα+[t])v=0,(xαt)max{0,μ(hα)dα}+1v=0, if αR+(μ)\left(x_{\alpha}^{+}\otimes\mathbb{C}[t]\right)v=0,\ \left(x_{\alpha}^{-}\otimes t\right)^{\mathrm{max}\{0,\ \mu(h_{\alpha})-d_{\alpha}\ell\}+1}v=0,\ \text{ if $\alpha\in R^{+}(\mu)$}
(xαt[t])v=0,(xα+1)μ(hα)+1v=0, if αR(μ)\left(x_{\alpha}^{-}\otimes t\mathbb{C}[t]\right)v=0,\ \left(x_{\alpha}^{+}\otimes 1\right)^{-\mu(h_{\alpha})+1}v=0,\ \text{ if $\alpha\in R^{-}(\mu)$}

where sα±,mα±+s^{\pm}_{\alpha},m^{\pm}_{\alpha}\in\mathbb{Z}_{+} are the unique integers such that

μ(hα)=(sα±1)dα+mα±, 0<mα±dα.\mp\mu(h_{\alpha})=(s^{\pm}_{\alpha}-1)d_{\alpha}\ell+m^{\pm}_{\alpha},\ \ \ 0<m^{\pm}_{\alpha}\leq d_{\alpha}\ell.

6.1.2. A tensor product theorem for 𝔤\mathfrak{g}-stable Demazure modules

Recall that we discussed in Section 4.1.6 the realization of local Weyl modules as a fusion product of fundamental local Weyl modules. We also discussed in Section 4.1.7 the results of [91] which gave the generators of the fusion products of Kirillov–Reshetikhin modules. These modules in the simply-laced case are known to be just 𝔤\mathfrak{g}-stable Demazure modules associated to weights of the form λ\ell\lambda with \ell\in\mathbb{N} and λP+\lambda\in P^{+}. We remark here that in the simply-laced case, the results of [91] are a vast generalization of the results of [58] where a presentation was given of the fusion product of Demazure modules of a fixed level. This was achieved by showing that the fusion product was isomorphic to a Demazure module. The following theorem which may be viewed as a Steinberg type decomposition theorem for 𝔤\mathfrak{g}-stable Demazure modules was proved in [41] (see also [102]) and completes the picture studied in [58].

Theorem.

Let 𝔤\mathfrak{g} be a finite-dimensional simple Lie algebra. Given kk\in\mathbb{N}, let λP+\lambda\in P^{+}, \ell\in\mathbb{N} and suppose that λ=(i=1kλi)+λ0\lambda=\ell\ (\sum_{i=1}^{k}\lambda_{i})+\lambda_{0} with λ0P+\lambda_{0}\in P^{+} and λi\lambda_{i} in the +\mathbb{Z}_{+}-span of the ωj\omega_{j}^{\vee} for 1jk1\leq j\leq k. Then there is an isomorphism of 𝔤[t]\mathfrak{g}[t]-modules

D(,λ)D(,λ0)z0D(,λ1)z1D(,λk)zkD(\ell,\lambda)\cong D(\ell,\lambda_{0})^{z_{0}}*D(\ell,\ell\lambda_{1})^{z_{1}}*\cdots*D(\ell,\ell\lambda_{k})^{z_{k}}

where z0,,zkz_{0},\dots,z_{k} are distinct complex parameters. In particular the fusion product is independent of the choice of parameters. ∎

The proof of the theorem relies on the simplified presentation of D(,λ)D(\ell,\lambda) given in Theorem Theorem and a character computation, using the Demazure character formula. This allows one to show that both modules in the theorem have the same 𝔥\mathfrak{h}-characters which is the crucial step to establish the theorem. The analogous theorem for twisted current algebras was proved in [79].

6.1.3.

We discuss an interesting consequence of Theorem Theorem. Say that a 𝔤[t]\mathfrak{g}[t]-module is prime if it is not isomorphic to a fusion product of non-trivial 𝔤[t]\mathfrak{g}[t]-modules. The interested reader should compare this definition with that given in Section 3.2.2 where an analogous definition was made in the context of quantum affine algebras. The following factorization result is a consequence of Theorem Theorem:

Corollary.

Given 1\ell\geq 1 and λP+\lambda\in P^{+} write

λ=(i=1ndimiωi)+λ0,λ0=i=1nriωi, 0ri<di,mi+.\lambda=\ell\,\left(\sum_{i=1}^{n}d_{i}m_{i}\omega_{i}\right)+\lambda_{0},\ \ \lambda_{0}=\sum_{i=1}^{n}r_{i}\omega_{i},\ \ 0\leq r_{i}<d_{i}\ell,\ \ m_{i}\in\mathbb{Z}_{+}.

Then D(,λ)D(\ell,\lambda) has the following fusion product factorization:

D(,λ)𝔤[t]D(,d1ω1)m1D(,d2ω2)m2D(,dnωn)mnD(,λ0).D(\ell,\lambda)\cong_{\mathfrak{g}[t]}D(\ell,\ell d_{1}\omega_{1})^{*m_{1}}*D(\ell,\ell d_{2}\omega_{2})^{*m_{2}}*\cdots*D(\ell,\ell d_{n}\omega_{n})^{*m_{n}}*D(\ell,\lambda_{0}). (6.1)

In addition, if we assume that 𝔤\mathfrak{g} is simply-laced then 6.1 gives prime factorization of D(,λ)D(\ell,\lambda) (i.e., each module on the ritht hand side is prime).

6.2. Demazure flags

In this section we explain a connection between modules admitting Demazure flags and combinatorics and hypergeometric series.

Say that a finite-dimensional 𝔤[t]\mathfrak{g}[t]-module MM has a level mm-Demazure flag if it admits a decreasing family of submodules, M=M0M1Mr0,M=M_{0}\supseteq M_{1}\supseteq\cdots\supseteq M_{r}\supseteq 0, such that

Mj/Mj+1τrjD(m,μj),rj,μjP+.M_{j}/M_{j+1}\cong\tau_{r_{j}}D(m,\mu_{j}),\ \ r_{j}\in\mathbb{Z},\ \ \mu_{j}\in P^{+}.

It is not hard to see by working with graded characters, that if M=M0M1Ms0M=M_{0}^{\prime}\supseteq M_{1}^{\prime}\supseteq\cdots\supset M_{s}^{\prime}\supseteq 0 is another level mm-Demazure flag then r=sr=s and the multiplicity of τrD(m,μ)\tau_{r}D(m,\mu) in both flags is the same. Hence we define [M:τrD(m,μ)][M:\tau_{r}D(m,\mu)] to be the number of times τrD(m,μ)\tau_{r}D(m,\mu) occurs in a level mm-Demazure flag of MM.

The study of Demazure flags goes back to the work of Naoi [88] on local Weyl modules in the non-simply laced case. It was proved in that paper that these modules admit a level one Demazure flag. This was done by first showing that in the simply-laced case every 𝔤[t]\mathfrak{g}[t]-stable Demazure module of level \ell admits a Demazure flag of level mm if m1m\geq\ell\geq 1. In the case when =1\ell=1 and m=2m=2 the multiplicities occurring in this flag can be explicitly related to the multiplicity of the level one flag of the local Weyl module for non-simply laced Lie algebras. However, the methods do not lead to precise formulae for the multiplicity. In this section we discuss how one might approach this problem using different kinds of generating series.

6.2.1. The case of 𝔰𝔩2\mathfrak{sl}_{2} and level two flags

A first step to calculate the multiplicity was taken in [40] for the Lie algebra 𝔰𝔩2\mathfrak{sl}_{2} when m=2m=2 and =1\ell=1. Then the graded multiplicities can be expressed by qq-binomial coefficients [40, Theorem 3.3]:

[D(1,μω):D(2,νω)]q=p0[D(1,μω):τpD(2,νω)]qp={q(μν)/2μ/2[μ/2(μν)/2]q,μν2+,0otherwise.\begin{split}[D(1,\mu\omega):D(2,\nu\omega)]_{q}&=\sum_{p\geq 0}[D(1,\mu\omega):\tau_{p}D(2,\nu\omega)]\ q^{p}\\ &=\begin{cases}q^{(\mu-\nu)/2\lceil{\mu/2\rceil}}\genfrac{[}{]}{0.0pt}{0}{\lfloor\mu/2\rfloor}{(\mu-\nu)/2}_{q},\ \ \mu-\nu\in 2\mathbb{Z}_{+},\\ 0\ \ {\rm{otherwise}}.\end{cases}\end{split}

6.2.2. The case of 𝔰𝔩2\mathfrak{sl}_{2} and arbitrary level

A more general approach in the 𝔰𝔩2\mathfrak{sl}_{2} case was taken in the articles [12, 15] and in the A2(2)A_{2}^{(2)} case in [11]. In those papers, the authors found a connection to algebraic combinatorics and number theory. We first need to introduce more notation. Define a family of generating series by

Anm(x,q)=k0[D(,(n+2k)ω):D(m,nω)]qxk,n0.A^{\ell\rightarrow m}_{n}(x,q)=\sum_{k\geq 0}\ [D(\ell,(n+2k)\omega):D(m,n\omega)]_{q}\cdot x^{k},\,\,n\geq 0.

Introduce the partial theta function θ(q,z)=k=0qk2zk\theta(q,z)=\sum_{k=0}^{\infty}q^{k^{2}}z^{k} and let

Pn(x)=k=0n2(1)k(nkk)xkP_{n}(x)=\sum_{k=0}^{\left\lfloor\frac{n}{2}\right\rfloor}(-1)^{k}\,{n-k\choose k}\,x^{k}

and note that the polynomials Pn(x)P_{n}(x) are related to the Chebyshev polynomials Un(x)U_{n}(x) of the second kind as follows Pn(x2)=xnUn((2x)1).P_{n}(x^{2})=x^{n}\,U_{n}(\,(2x)^{-1}). The following theorem can be found in [12, Theorem 1.6] and [12, Corollary 1.3] respectively.

Theorem.
  • (i)

    Let n,m+n,m\in\mathbb{Z}_{+} and write n=ms+rn=ms+r with s+s\in\mathbb{Z}_{+} and 0r<m0\leq r<m. Then

    An1m(x,1)=Pmr1(x)Pm(x)s+1.A^{1\rightarrow m}_{n}(x,1)=\frac{P_{m-r-1}(x)}{P_{m}(x)^{s+1}}.
  • (ii)

    The specializations

    A113(1,q),qA113(q,q),A013(1,q2)+qA213(1,q2),q4A213(q2,q2)+qA013(q2,q2)A^{1\rightarrow 3}_{1}(1,q),\ \ \ q\cdot A^{1\rightarrow 3}_{1}(q,q),\ \ \ A^{1\rightarrow 3}_{0}(1,q^{2})+qA^{1\rightarrow 3}_{2}(1,q^{2}),\ \ \ q^{4}A^{1\rightarrow 3}_{2}(q^{2},q^{2})+qA^{1\rightarrow 3}_{0}(q^{2},q^{2})

    coincide with fifth order mock theta functions of Ramanujan.

  • (iii)

    The series An12(x,q)A^{1\rightarrow 2}_{n}(x,q) and An23(x,q)A^{2\rightarrow 3}_{n}(x,q) can be expressed as a linear combination of specializations of the partial theta function θ\theta whose coefficients are given by products of qq-binomial coefficients.

6.2.3. Combinatorics of Dyck paths and the functions An1mA_{n}^{1\to m}

We further discuss the 𝔰𝔩2\mathfrak{sl}_{2} case and its connection to the combinatorics of Dyck paths. In [15] a combinatorial formula has been obtained whose ingredients we will now explain. A Dyck path is a diagonal lattice path from the origin (0,0)(0,0) to (s,n)(s,n) for some non-negative integrs s,n+s,n\in\mathbb{Z}_{+}, such that the path never goes below the x-axis. We encode such a path by a 01-word, where 11 encodes the up-steps and 0 the down-steps. We denote by 𝒟nN\mathcal{D}^{N}_{n} the set of Dyck paths that end at height nn and which do not cross the line y=Ny=N. The following picture is an example of an element in 𝒟14\mathcal{D}^{4}_{1}.

For n,m+n,m\in\mathbb{Z}_{+} let n0,n1+n_{0},n_{1}\in\mathbb{Z}_{+} be such that n0<mn_{0}<m and n=mn1+n0n=mn_{1}+n_{0}. If n<mn<m we set A(m,n)=A(m,n)=\emptyset and otherwise define

A(m,n):={(i1,m),(i2,m+1),,(inm+1,n)}+2A(m,n):=\{(i_{1},m),(i_{2},m+1),\dots,(i_{n-m+1},n)\}\subseteq\mathbb{Z}_{+}^{2}

where i1<<inm+1i_{1}<\cdots<i_{n-m+1} is the natural ordering of the set

{0,,n}\{pn1+n0+min{0,(p1)n0}, 1pm}.\{0,\dots,n\}\backslash\{pn_{1}+n_{0}+\min\{0,(p-1)-n_{0}\},\ 1\leq p\leq m\}.

Given a pair of non-negative integers (a,b)+2(a,b)\in\mathbb{Z}^{2}_{+}, we say that P𝒟nmax{m1,n}P\in\mathcal{D}_{n}^{\max\{m-1,n\}} is (a,b)(a,b)-admissible if and only if PP satisfies the following property. If PP has a peak at height bb, the subsequent path is strictly above the line y=ay=a. For example, the path above is not (0,3)(0,3)-admissible.

Let 𝒟m,n\mathcal{D}_{m,n} be the set Dyck paths in 𝒟nmax{m1,n}\mathcal{D}_{n}^{\max\{m-1,n\}}, which are (a,b)(a,b)-admissible for all (a,b)A(m,n)(a,b)\in A(m,n).

The major statistics of a Dyck path was studied first by MacMahon [84] in his interpretation of the qq-Catalan numbers. Let P=a1asP=a_{1}\cdots a_{s}, ai{0,1}a_{i}\in\{0,1\} be a Dyck path of length ss. The major and comajor index are defined by

maj(P)=1i<s,ai>ai+1i,comaj(P)=1i<s,ai>ai+1(si).\text{maj}(P)=\sum_{\begin{subarray}{c}1\leq i<s,\\ a_{i}>a_{i+1}\end{subarray}}i,\ \ \ \text{comaj}(P)=\sum_{\begin{subarray}{c}1\leq i<s,\\ a_{i}>a_{i+1}\end{subarray}}(s-i).

The following was proved in [15, Theorem 4].

Theorem.

Let mm\in\mathbb{N}, n+n\in\mathbb{Z}_{+}. We have,

An1m(x,q)=P𝒟m,nqcomaj(P)xd(P)A^{1\rightarrow m}_{n}(x,q)=\sum_{P\in\mathcal{D}_{m,n}}q^{\text{comaj}(P)}\ x^{d(P)}

where d(P)d(P) denotes the number of down-steps of PP. ∎

In the twisted case graded and weighted generating functions encode again the multiplicity of a given Demazure module. For small ranks these generating functions are completely determined in [11] and they define hypergeometric series and are related to the qq–Fibonacci polynomials defined by Carlitz. For more details we refer the reader to [11, Section 2].

6.2.4. The general case

It is still an open problem to come up with closed or even recursive formulas for the generating series for other finite-dimensional simply-laced Lie algebras; the multiplicities and generating functions are defined in the obvious way. However, some progress has been made in [14] for the Lie algebra 𝔰𝔩n+1\mathfrak{sl}_{n+1} and the connection to Macdonald polynomials was established. The following result can be derived from [14].

Theorem.

Let 𝔤=𝔰𝔩n+1\mathfrak{g}=\mathfrak{sl}_{n+1} and λ,μP+\lambda,\mu\in P^{+} such that λμ=i=1nkiαi,ki+\lambda-\mu=\sum_{i=1}^{n}k_{i}\alpha_{i},\ k_{i}\in\mathbb{Z}_{+}. Then,

[D(1,λ):D(2,μ)]q=i=1n[D(1,(μ(hi)+2ki)ω):D(2,μ(hi)ω)]q[D(1,\lambda):D(2,\mu)]_{q}=\prod_{i=1}^{n}\big{[}D(1,(\mu(h_{i})+2k_{i})\omega):D(2,\mu(h_{i})\omega)\big{]}_{q}

where ω\omega is the corresponding fundamental weight for 𝔰𝔩2\mathfrak{sl}_{2}. ∎

So combining the above theorem with the combinatorial formula in Theorem Theorem gives a combinatorial formula for graded multiplicities of level 2 Demazure modules in level one flags.

References

  • [1] Tatsuya Akasaka and Masaki Kashiwara. Finite-dimensional representations of quantum affine algebras. Publ. Res. Inst. Math. Sci., 33(5):839–867, 1997.
  • [2] Eddy Ardonne and Rinat Kedem. Fusion products of Kirillov-Reshetikhin modules and fermionic multiplicity formulas. J. Algebra, 308(1):270–294, 2007.
  • [3] Leon Barth and Deniz Kus. Graded decompositions of fusion products in rank two. Kyoto J. Math (to appear), 2020.
  • [4] Leon Barth and Deniz Kus. Prime representations in the Hernandez-Leclerc category: classical decompositions. arXiv:2012.15334, 2020.
  • [5] Matthew Bennett and Angelo Bianchi. Tilting modules in truncated categories. SIGMA Symmetry Integrability Geom. Methods Appl., 10:Paper 030, 23pp, 2014.
  • [6] Matthew Bennett and Vyjayanthi Chari. Tilting modules for the current algebra of a simple Lie algebra. In Recent developments in Lie algebras, groups and representation theory, volume 86 of Proc. Sympos. Pure Math., pages 75–97. Amer. Math. Soc., Providence, RI, 2012.
  • [7] Matthew Bennett and Vyjayanthi Chari. Character formulae and a realization of tilting modules for 𝔰𝔩2[t]\mathfrak{sl}_{2}[t]. J. Algebra, 441:216–242, 2015.
  • [8] Matthew Bennett, Vyjayanthi Chari, Jacob Greenstein, and Nathan Manning. On homomorphisms between global Weyl modules. Represent. Theory, 15:733–752, 2011.
  • [9] Matthew Bennett, Vyjayanthi Chari, and Nathan Manning. BGG reciprocity for current algebras. Adv. Math., 231(1):276–305, 2012.
  • [10] I. N. Bernšteĭn, I. M. Gel’fand, and S. I. Gel’fand. A certain category of 𝔤{\mathfrak{g}}-modules. Funkcional. Anal. i Priložen., 10(2):1–8, 1976.
  • [11] Rekha Biswal, Vyjayanthi Chari, and Deniz Kus. Demazure flags, qq-Fibonacci polynomials and hypergeometric series. Res. Math. Sci., 5(1):Paper No. 12, 34, 2018.
  • [12] Rekha Biswal, Vyjayanthi Chari, Lisa Schneider, and Sankaran Viswanath. Demazure flags, Chebyshev polynomials, partial and mock theta functions. J. Combin. Theory Ser. A, 140:38–75, 2016.
  • [13] Rekha Biswal, Vyjayanthi Chari, Peri Shereen, and Jeffrey Wand. Cone theta functions and Demazure flags in higher rank. in preparation.
  • [14] Rekha Biswal, Vyjayanthi Chari, Peri Shereen, and Jeffrey Wand. Macdonald polynomials and level two Demazure modules for affine 𝔰𝔩n+1\mathfrak{sl}_{n+1}. J. Algebra, 575:159–191, 2021.
  • [15] Rekha Biswal and Deniz Kus. A combinatorial formula for graded multiplicities in excellent filtrations. Transform. Groups, 26(1):81–114, 2021.
  • [16] Matheus Brito and Vyjayanthi Chari. Tensor products and qq-characters of HL-modules and monoidal categorifications. J. Éc. polytech. Math., 6:581–619, 2019.
  • [17] Matheus Brito, Vyjayanthi Chari, and Adriano Moura. Demazure modules of level two and prime representations of quantum affine 𝔰𝔩n+1\mathfrak{sl}_{n+1}. J. Inst. Math. Jussieu, 17(1):75–105, 2018.
  • [18] Matheus Brito and Evgeny Mukhin. Representations of quantum affine algebras of type BN{B}_{N}. Transactions of the American Mathematical Society, 369:2775–2806, 2014.
  • [19] Vyjayanthi Chari. Integrable representations of affine Lie-algebras. Invent. Math., 85(2):317–335, 1986.
  • [20] Vyjayanthi Chari. Minimal affinizations of representations of quantum groups: the rank 22 case. Publ. Res. Inst. Math. Sci., 31(5):873–911, 1995.
  • [21] Vyjayanthi Chari. Minimal affinizations of representations of quantum groups: the rank 22 case. Publ. Res. Inst. Math. Sci., 31(5):873–911, 1995.
  • [22] Vyjayanthi Chari. On the fermionic formula and the Kirillov-Reshetikhin conjecture. Internat. Math. Res. Notices, 12:629–654, 2001.
  • [23] Vyjayanthi Chari. Braid group actions and tensor products. Int. Math. Res. Not., 7:357–382, 2002.
  • [24] Vyjayanthi Chari, Justin Davis, and Ryan Moruzzi Jr. Generalized Demazure modules and prime representations in type Dn{D}_{n}. arXiv:1911.07155, 2019.
  • [25] Vyjayanthi Chari, Ghislain Fourier, and Tanusree Khandai. A categorical approach to Weyl modules. Transform. Groups, 15(3):517–549, 2010.
  • [26] Vyjayanthi Chari and Jacob Greenstein. Minimal affinizations as projective objects. Journal of Geometry and Physics, 61, 03 2011.
  • [27] Vyjayanthi Chari and Bogdan Ion. BGG reciprocity for current algebras. Compos. Math., 151(7):1265–1287, 2015.
  • [28] Vyjayanthi Chari, Bogdan Ion, and Deniz Kus. Weyl modules for the hyperspecial current algebra. Int. Math. Res. Not. IMRN, 15:6470–6515, 2015.
  • [29] Vyjayanthi Chari, Deniz Kus, and Matt Odell. Borel–de Siebenthal pairs, global Weyl modules and Stanley-Reisner rings. Math. Z., 290(1-2):649–681, 2018.
  • [30] Vyjayanthi Chari and Sergei Loktev. Weyl, Demazure and fusion modules for the current algebra of 𝔰𝔩r+1\mathfrak{sl}_{r+1}. Adv. Math., 207(2):928–960, 2006.
  • [31] Vyjayanthi Chari and Adriano Moura. The restricted Kirillov-Reshetikhin modules for the current and twisted current algebras. Comm. Math. Phys., 266(2):431–454, 2006.
  • [32] Vyjayanthi Chari and Andrew Pressley. New unitary representations of loop groups. Mathematische Annalen, 275:87–104, 1986.
  • [33] Vyjayanthi Chari and Andrew Pressley. Quantum affine algebras. Comm. Math. Phys., 142(2):261–283, 1991.
  • [34] Vyjayanthi Chari and Andrew Pressley. Minimal affinizations of representations of quantum groups: the nonsimply-laced case. Lett. Math. Phys., 35(2):99–114, 1995.
  • [35] Vyjayanthi Chari and Andrew Pressley. Quantum affine algebras and their representations. In Representations of groups (Banff, AB, 1994), volume 16 of CMS Conf. Proc., pages 59–78. Amer. Math. Soc., Providence, RI, 1995.
  • [36] Vyjayanthi Chari and Andrew Pressley. Minimal affinizations of representations of quantum groups: the irregular case. Lett. Math. Phys., 36(3):247–266, 1996.
  • [37] Vyjayanthi Chari and Andrew Pressley. Minimal affinizations of representations of quantum groups: the simply laced case. J. Algebra, 184(1):1–30, 1996.
  • [38] Vyjayanthi Chari and Andrew Pressley. Integrable and Weyl modules for quantum affine sl2{\rm sl}_{2}. In Quantum groups and Lie theory (Durham, 1999), volume 290 of London Math. Soc. Lecture Note Ser., pages 48–62. Cambridge Univ. Press, Cambridge, 2001.
  • [39] Vyjayanthi Chari and Andrew Pressley. Weyl modules for classical and quantum affine algebras. Represent. Theory, 5:191–223 (electronic), 2001.
  • [40] Vyjayanthi Chari, Lisa Schneider, Peri Shereen, and Jeffrey Wand. Modules with Demazure Flags and Character Formulae. SIGMA Symmetry Integrability Geom. Methods Appl., 10, 2014.
  • [41] Vyjayanthi Chari, Peri Shereen, R. Venkatesh, and Jeffrey Wand. A Steinberg type decomposition theorem for higher level Demazure modules. J. Algebra, 455:314–346, 2016.
  • [42] Vyjayanthi Chari and R. Venkatesh. Demazure modules, fusion products and QQ-systems. Comm. Math. Phys., 333(2):799–830, 2015.
  • [43] Ivan Cherednik. Double affine Hecke algebras and Macdonald’s conjectures. Ann. of Math. (2), 141(1):191–216, 1995.
  • [44] Ilaria Damiani. La RR-matrice pour les algèbres quantiques de type affine non tordu. Ann. Sci. École Norm. Sup. (4), 31(4):493–523, 1998.
  • [45] V. G. Drinfeld. A new realization of Yangians and quantized affine algebras. Sov. Math. Dokl., 36:212–216, 1988.
  • [46] Ilya Dumanski and Evgeny Feigin. Reduced arc schemes for Veronese embeddings and global Demazure modules. arXiv:1912.07988, 2021.
  • [47] Ilya Dumanski, Evgeny Feigin, and Michael Finkelberg. Beilinson-Drinfeld Schubert varieties and global Demazure modules. Forum Math. Sigma, 9:Paper No. e42, 25pp, 2021.
  • [48] S. Eswara Rao. Complete reducibility of integrable modules for the affine Lie (super)algebras. J. Algebra, 264(1):269–278, 2003.
  • [49] B. Feigin and E. Feigin. QQ-characters of the tensor products in 𝔰𝔩2\mathfrak{sl}_{2}-case. Mosc. Math. J., 2(3):567–588, 2002. Dedicated to Yuri I. Manin on the occasion of his 65th birthday.
  • [50] Boris Feigin and Sergei Loktev. On generalized Kostka polynomials and the quantum Verlinde rule. In Differential topology, infinite-dimensional Lie algebras, and applications, volume 194 of Amer. Math. Soc. Transl. Ser. 2, pages 61–79. Amer. Math. Soc., Providence, RI, 1999.
  • [51] Boris Feigin and Sergei Loktev. Multi-dimensional Weyl modules and symmetric functions. Comm. Math. Phys., 251(3):427–445, 2004.
  • [52] Evgeny Feigin and Ievgen Makedonskyi. Generalized Weyl modules, alcove paths and Macdonald polynomials. Selecta Math. (N.S.), 23(4):2863–2897, 2017.
  • [53] Evgeny Feigin and Ievgen Makedonskyi. Vertex algebras and coordinate rings of semi-infinite flags. Comm. Math. Phys., 369(1):221–244, 2019.
  • [54] Ghislain Fourier. New homogeneous ideals for current algebras: filtrations, fusion products and Pieri rules. Mosc. Math. J., 15(1):49–72, 181, 2015.
  • [55] Ghislain Fourier and David Hernandez. Schur positivity and Kirillov-Reshetikhin modules. SIGMA Symmetry Integrability Geom. Methods Appl., 10:Paper 058, 9, 2014.
  • [56] Ghislain Fourier, Tanusree Khandai, Deniz Kus, and Alistair Savage. Local Weyl modules for equivariant map algebras with free abelian group actions. J. Algebra, 350:386–404, 2012.
  • [57] Ghislain Fourier and Deniz Kus. Demazure modules and Weyl modules: The twisted current case. Trans. Amer. Math. Soc., 365(11):6037–6064, 2013.
  • [58] Ghislain Fourier and Peter Littelmann. Weyl modules, Demazure modules, KR-modules, crystals, fusion products and limit constructions. Adv. Math., 211(2):566–593, 2007.
  • [59] Philippe Di Francesco and Rinat Kedem. Proof of the combinatorial Kirillov-Reshetikhin conjecture. Int. Math. Res. Not. IMRN, 7:Art. ID rnn006, 57, 2008.
  • [60] Edward Frenkel and Nicolai Reshetikhin. The qq-characters of representations of quantum affine algebras and deformations of 𝒲\mathcal{W}-algebras. In Recent developments in quantum affine algebras and related topics (Raleigh, NC, 1998), volume 248 of Contemp. Math., pages 163–205. Amer. Math. Soc., Providence, RI, 1999.
  • [61] I. B. Frenkel and N. Yu. Reshetikhin. Quantum affine algebras and holonomic difference equations. Comm. Math. Phys., 146(1):1–60, 1992.
  • [62] Ryo Fujita. Graded quiver varieties and singularities of normalized R-matrices for fundamental modules. Selecta Math. (N.S.), 28(1):Paper No. 2, 45, 2022.
  • [63] Ryo Fujita and Se-jin Oh. Q-data and representation theory of untwisted quantum affine algebras. Comm. Math. Phys., 384(2):1351–1407, 2021.
  • [64] G. Hatayama, A. Kuniba, M. Okado, T. Takagi, and Y. Yamada. Remarks on fermionic formula. In Recent developments in quantum affine algebras and related topics (Raleigh, NC, 1998), volume 248 of Contemp. Math., pages 243–291. Amer. Math. Soc., Providence, RI, 1999.
  • [65] David Hernandez. The Kirillov-Reshetikhin conjecture and solutions of TT-systems. J. Reine Angew. Math., 596:63–87, 2006.
  • [66] David Hernandez. Simple tensor products. Invent. Math., 181(3):649–675, 2010.
  • [67] David Hernandez and Bernard Leclerc. Cluster algebras and quantum affine algebras. Duke Math. J., 154(2):265–341, 2010.
  • [68] David Hernandez and Bernard Leclerc. Monoidal categorifications of cluster algebras of type AA and DD. In Symmetries, integrable systems and representations, volume 40 of Springer Proc. Math. Stat., pages 175–193. Springer, Heidelberg, 2013.
  • [69] James E. Humphreys. Introduction to Lie Algebras and Representation Theory, volume 9 of Graduate Texts in Mathematics. Springer-Verlag, Berlin, 1980.
  • [70] James E. Humphreys. Representations of semisimple Lie algebras in the BGG category O, volume 94 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2008.
  • [71] Bogdan Ion. Nonsymmetric Macdonald polynomials and Demazure characters. Duke Math. J., 116(2):299–318, 2003.
  • [72] Anthony Joseph. On the Demazure character formula. Annales scientifiques de l’École Normale Supérieure, Ser. 4, 18(3):389–419, 1985.
  • [73] Victor G. Kac. Infinite-dimensional Lie algebras. Cambridge University Press, Cambridge, third edition, 1990.
  • [74] Syu Kato. Demazure character formula for semi-infinite flag varieties. Math. Ann., 371(3-4):1769–1801, 2018.
  • [75] Rinat Kedem. A pentagon of identities, graded tensor products, and the Kirillov-Reshetikhin conjecture. In New trends in quantum integrable systems, pages 173–193. World Sci. Publ., Hackensack, NJ, 2011.
  • [76] A. N. Kirillov and N. Yu. Reshetikhin. Representations of Yangians and multiplicities of the inclusion of the irreducible components of the tensor product of representations of simple Lie algebras. Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), 160(Anal. Teor. Chisel i Teor. Funktsiĭ. 8):211–221, 301, 1987.
  • [77] Ryosuke Kodera and Katsuyuki Naoi. Loewy series of Weyl modules and the Poincaré polynomials of quiver varieties. Publ. Res. Inst. Math. Sci., 48(3):477–500, 2012.
  • [78] Deniz Kus and Peter Littelmann. Fusion products and toroidal algebras. Pacific J. Math., 278(2):427–445, 2015.
  • [79] Deniz Kus and R. Venkatesh. Twisted Demazure modules, fusion product decomposition and twisted QQ-systems. Represent. Theory, 20:94–127, 2016.
  • [80] Deniz Kus and R. Venkatesh. Simplified presentations and embeddings of Demazure modules. arXiv:2112.14830, 2021.
  • [81] Bernard Leclerc. Imaginary vectors in the dual canonical basis of uq(n)u_{q}(n). Transformation Groups, 8:95–104, 2002.
  • [82] Jian-Rong Li and Katsuyuki Naoi. Graded limits of minimal affinizations over the quantum loop algebra of type G2G_{2}. Algebr. Represent. Theory, 19(4):957–973, 2016.
  • [83] George Lusztig. Introduction to quantum groups. Modern Birkhäuser Classics. Birkhäuser/Springer, New York, 2010. Reprint of the 1994 edition.
  • [84] Percy A. MacMahon. Combinatory analysis. Chelsea Publishing Co., New York, 1960. Two volumes (bound as one).
  • [85] Olivier Mathieu. Formules de caractères pour les algèbres de Kac-Moody générales. Number 159-160 in Astérisque. Société mathématique de France, 1988.
  • [86] Adriano Moura. Restricted limits of minimal affinizations. Pacific J. Math., 244(2):359–397, 2010.
  • [87] Evgeny Mukhin and Charles A. S. Young. Path description of type B{B} qq-characters. Advances in Mathematics, 231:1119–1150, 2012.
  • [88] Katsuyuki Naoi. Weyl modules, Demazure modules and finite crystals for non-simply laced type. Adv. Math., 229(2):875–934, 2012.
  • [89] Katsuyuki Naoi. Demazure modules and graded limits of minimal affinizations. Represent. Theory, 17:524–556, 2013.
  • [90] Katsuyuki Naoi. Graded limits of minimal affinizations in type DD. SIGMA Symmetry Integrability Geom. Methods Appl., 10:Paper 047, 20, 2014.
  • [91] Katsuyuki Naoi. Tensor products of Kirillov-Reshetikhin modules and fusion products. Int. Math. Res. Not. IMRN, 18:5667–5709, 2017.
  • [92] Masato Okado and Anne Schilling. Existence of Kirillov-Reshetikhin crystals for nonexceptional types. Represent. Theory, 12:186–207, 2008.
  • [93] Eric M. Opdam. Harmonic analysis for certain representations of graded Hecke algebras. Acta Math., 175(1):75–121, 1995.
  • [94] Daniel Orr and Mark Shimozono. Specializations of nonsymmetric Macdonald-Koornwinder polynomials. J. Algebraic Combin., 47(1):91–127, 2018.
  • [95] Fernanda Pereira. Classification of the type D irregular minimal affinizations. PhD thesis, UNICAMP, 2014.
  • [96] K. N. Raghavan, B. Ravinder, and Sankaran Viswanath. Stability of the Chari-Pressley-Loktev bases for local Weyl modules of 𝔰𝔩2[t]\mathfrak{sl}_{2}[t]. Algebr. Represent. Theory, 18(3):613–632, 2015.
  • [97] K. N. Raghavan, B. Ravinder, and Sankaran Viswanath. On Chari-Loktev bases for local Weyl modules in type AA. J. Combin. Theory Ser. A, 154:77–113, 2018.
  • [98] Marc Rosso. Finite-dimensional representations of the quantum analog of the enveloping algebra of a complex simple Lie algebra. Comm. Math. Phys., 117(4):581–593, 1988.
  • [99] Steven Sam. Jacobi-Trudi determinants and characters of minimal affinizations. Pacific Journal of Mathematics, 272, 07 2013.
  • [100] M. Varagnolo and E. Vasserot. Standard modules of quantum affine algebras. Duke Math. J., 111(3):509–533, 2002.
  • [101] R. Venkatesh. Fusion product structure of Demazure modules. Algebr. Represent. Theory, 18(2):307–321, 2015.
  • [102] R. Venkatesh and S. Viswanath. A note on the fusion product decomposition of Demazure modules. Journal of Lie Theory, 32, No. 1:261–266, 2022.
  • [103] Lauren K. Williams. Cluster algebras: an introduction. Bull. Amer. Math. Soc. (N.S.), 51(1):1–26, 2014.