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

Some remarks on the unrolled quantum group of 𝔰​𝔩​(2){\mathfrak{sl}(2)}

Francesco Costantino Institut de MathΓ©matiques Toulouse
118 route de Narbonne
31062 Toulouse Cedex 9, France
Francesco.Costantino@math.univ-toulouse.fr
,Β  Nathan Geer Mathematics & Statistics
Utah State University
Logan, Utah 84322, USA
nathan.geer@gmail.com
Β andΒ  Bertrand Patureau-Mirand UMR 6205, LMBA, universitΓ© de Bretagne-Sud, universitΓ© europΓ©enne de Bretagne, BP 573, 56017 Vannes, France bertrand.patureau@univ-ubs.fr
Abstract.

In this paper we consider the representation theory of a non-standard quantization of 𝔰​𝔩​(2){\mathfrak{sl}(2)}. This paper contains several results which have applications in quantum topology, including the classification of projective indecomposable modules and a description of morphisms between them. In the process of proving these results the paper acts as a survey of the known representation theory associated to this non-standard quantization of 𝔰​𝔩​(2){\mathfrak{sl}(2)}. The results of this paper are used extensively in [4] to study Topological Quantum Field Theory (TQFT) and have connections with Conformal Field Theory (CFT).

The first author’s research was supported by French ANR project ANR-08-JCJC-0114-01. Research of the second author was partially supported by NSF grants DMS-1007197 and DMS-1308196. All the authors would like to thank the Erwin SchrΓΆdinger Institute for Mathematical Physics in Vienna for support during a stay in the Spring of 2014, where part of this work was done.

1. Introduction

There are many different flavors of quantum 𝔰​𝔩​(2){\mathfrak{sl}(2)} based on a common algebraic presentations. In particular, these presentations depend on two features: (1) if the quantum parameter qq is generic or a root of unity and (2) what part of the center is killed. The associated representation theory varies widely when these features are changed. Two examples, when qq is a root of unity, are the finite-dimensional Hopf algebra commonly known as the small quantum group and the non-restricted quantum group obtained by specializing the De Concini-Kac form (for definitions of these algebras see [7]). The representation theory of the small quantum group leads to a modular category (in particular a finite, semi-simple, ribbon category) which can be used to construct 3-manifold invariants. On the other hand, the representation theory of the non-restricted quantum group contains an infinite class of modules called the cyclic modules.

In this paper we consider an intermediate quotient UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}, which we call the unrolled quantum group, leading to a category UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-mod which is ribbon but not semi-simple or finite. This category has been used to construct quantum link and 3-manifold invariants in several papers [2, 24, 25, 22, 10, 8, 4]. These 3-manifold invariants have powerful new properties, including asymptotic behavior related to the Volume Conjecture and novel quantum representation of mapping class groups (see [8, 4]). The existence of these properties is directly related to the unique representation theory discussed in this paper.

The Hopf algebra UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}} of this paper has an additional generator HH which is not in the usual quantum algebra associated to 𝔰​𝔩​(2){\mathfrak{sl}(2)}. The element HH should be thought of as a logarithm of the usual generator KK. The generator HH is used to define a braiding and a twist on a category of UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-modules. In this category, HH is also responsible for the apparition of an infinite cyclic group of one dimensional invertible objects which play a key role in the topological applications.

The purpose of this paper is to give a survey of the known results about UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-mod while proving some new useful results which have topological applications. In particular, we classify all indecomposable projective modules and define a modified trace on these objects (see Section 6). We give a β€œgraded” quiver which describes the maps between the indecomposable projective modules (see Section 7). We also study the decomposition of the tensor product of certain indecomposable modules (see Section 8). These results are used in [4] in an essential way to build a TQFT for 3-manifolds equipped with a cohomology class. The category UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-mod contains some indecomposable non-projective modules that are not studied in this paper (see for example their use in [9]). Instead here, we focus on semi-simple and projective modules that form together a sub tensor category (see Proposition 8.4).

The category of UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-modules has a grading in the abelian group β„‚/2​℀\mathbb{C}/2\mathbb{Z} and its non semi-simple part is concentrated in degree 0Β―,1Β―\overline{0},\overline{1} blocks. These two blocks form a category similar to the category U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}}-mod of representations of the standard small quantum group U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}}. The category U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}}-mod, equivalent to that of modules over the triplet vertex operator algebra 𝒲​(p)\mathcal{W}(p) (see [26, 28]), has been intensively studied in logarithmic conformal field theories (CFT) associated to the (1,p)(1,p) triplet algebras (see [18, 17, 6, 5, 12]). In particular, some results of Section 6 are similar to the analysis of projective modules in [18].

The category UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-mod has additional modules which do not appear in the representation theory of the small quantum group (in particular, the one dimensional invertible objects mentioned above). Moreover, conjecturally UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-mod is equivalent to the category of representation of the vertex operator algebra called singlet vertex algebra 𝒲​(2,2​pβˆ’1)\mathcal{W}(2,2p-1) (see [1, 11]). Understanding a deeper connection between the representation theory of this paper and CFT deserves some attention. For example, it would be interesting to compare the CFT representations of S​L​(2,β„€)SL(2,\mathbb{Z}) (see [17]) and more generally mapping class group representations (see [19]) with those obtained from UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}} in the TQFT of [4].

1.1. Acknowledgements

We would like to thank Simon Wood and Antun Milas for their useful comments on the relations with the theory of logarithmic CFTs and the organizers of the conference β€œModern trends in topological quantum field theory” at the Erwin SchrΓΆdinger Institut (Vienna) for their kind invitation to the conference.

2. A quantization of 𝔰​𝔩​(2){\mathfrak{sl}(2)} and its associated ribbon category

In this section we recall the algebra UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}} and the category of modules over this algebra. Fix a positive integer rr. Let rβ€²=rr^{\prime}=r if rr is odd and rβ€²=r2r^{\prime}=\frac{r}{2} else. Let β„‚\mathbb{C} be the complex numbers and ¨​ℂ=(β„‚βˆ–β„€)βˆͺr​℀.{\ddot{}\mathbb{C}}=(\mathbb{C}\setminus\mathbb{Z})\cup r\mathbb{Z}. Let q=eΟ€β€‹βˆ’1rq=e^{\frac{\pi\sqrt{-1}}{r}} be a 2​rt​h2r^{th}-root of unity. We use the notation qx=eΟ€β€‹βˆ’1​xrq^{x}=e^{\frac{\pi\sqrt{-1}x}{r}}. For nβˆˆβ„•n\in\mathbb{N}, we also set

{x}=qxβˆ’qβˆ’x,[x]={x}{1},{n}!={n}​{nβˆ’1}​⋯​{1}and[n]!=[n]​[nβˆ’1]​⋯​[1]{\left\{x\right\}}=q^{x}-q^{-x},\quad{\left[x\right]}=\frac{{\left\{x\right\}}}{{\left\{1\right\}}},\quad{\left\{n\right\}}!={\left\{n\right\}}{\left\{n-1\right\}}\cdots{\left\{1\right\}}{\quad\text{and}\quad}{\left[n\right]}!={\left[n\right]}{\left[n-1\right]}\cdots{\left[1\right]}

2.1. The Drinfel’d-Jimbo quantum group

Let Uq​𝔰​𝔩​(2){U_{q}{\mathfrak{sl}(2)}} be the β„‚\mathbb{C}-algebra given by generators E,F,K,Kβˆ’1E,F,K,K^{-1} and relations:

(1) K​Kβˆ’1\displaystyle KK^{-1} =Kβˆ’1​K=1,\displaystyle=K^{-1}K=1, K​E​Kβˆ’1\displaystyle KEK^{-1} =q2​E,\displaystyle=q^{2}E, K​F​Kβˆ’1\displaystyle KFK^{-1} =qβˆ’2​F,\displaystyle=q^{-2}F, [E,F]\displaystyle[E,F] =Kβˆ’Kβˆ’1qβˆ’qβˆ’1.\displaystyle=\frac{K-K^{-1}}{q-q^{-1}}.

The algebra Uq​𝔰​𝔩​(2){U_{q}{\mathfrak{sl}(2)}} is a Hopf algebra where the coproduct, counit and antipode are defined by

(2) Δ​(E)\displaystyle\Delta(E) =1βŠ—E+EβŠ—K,\displaystyle=1\otimes E+E\otimes K, Ρ​(E)\displaystyle\varepsilon(E) =0,\displaystyle=0, S​(E)\displaystyle S(E) =βˆ’E​Kβˆ’1,\displaystyle=-EK^{-1},
(3) Δ​(F)\displaystyle\Delta(F) =Kβˆ’1βŠ—F+FβŠ—1,\displaystyle=K^{-1}\otimes F+F\otimes 1, Ρ​(F)\displaystyle\varepsilon(F) =0,\displaystyle=0, S​(F)\displaystyle S(F) =βˆ’K​F,\displaystyle=-KF,
(4) Δ​(K)\displaystyle\Delta(K) =KβŠ—K\displaystyle=K\otimes K Ρ​(K)\displaystyle\varepsilon(K) =1,\displaystyle=1, S​(K)\displaystyle S(K) =Kβˆ’1.\displaystyle=K^{-1}.

Let UΒ―q​𝔰​𝔩​(2){\overline{U}_{q}{\mathfrak{sl}(2)}} be the algebra Uq​𝔰​𝔩​(2){U_{q}{\mathfrak{sl}(2)}} modulo the relations Er=Fr=0E^{r}=F^{r}=0. Also, let U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}} be the algebra UΒ―q​𝔰​𝔩​(2){\overline{U}_{q}{\mathfrak{sl}(2)}} modulo the relations K2​r=1K^{2r}=1. These relations generate Hopf ideals so UΒ―q​𝔰​𝔩​(2){\overline{U}_{q}{\mathfrak{sl}(2)}} and U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}} inherit a Hopf algebra structure.

As we will now explain, the categories of modules over Uq​𝔰​𝔩​(2),UΒ―q​𝔰​𝔩​(2){U_{q}{\mathfrak{sl}(2)}},{\overline{U}_{q}{\mathfrak{sl}(2)}} and U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}} have very different properties. Let XX-mod be the tensor category of finite dimensional XX-modules for XX equal to Uq​𝔰​𝔩​(2),UΒ―q​𝔰​𝔩​(2){U_{q}{\mathfrak{sl}(2)}},{\overline{U}_{q}{\mathfrak{sl}(2)}} or U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}}. The algebra U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}} is known as the small quantum group and has been well studied, see [7] and the references within. The algebra Uq​𝔰​𝔩​(2){U_{q}{\mathfrak{sl}(2)}} is known as the De Concini-Kac quantum group. It and the category Uq​𝔰​𝔩​(2){U_{q}{\mathfrak{sl}(2)}}-mod have rich structures and have been studied in [13, 14, 15, 16]. This category is not braided nor semi-simple and has an infinite number of simple modules called cyclic modules which are not highest weight modules. Finally, the category UΒ―q​𝔰​𝔩​(2){\overline{U}_{q}{\mathfrak{sl}(2)}}-mod is not semi-simple nor braided and has an infinite number of non-isomorphic simple modules. However, one can easily modify UΒ―q​𝔰​𝔩​(2){\overline{U}_{q}{\mathfrak{sl}(2)}} and obtain a braided category of highest weight modules which has been used to construct invariants of links ([24]), of 3-manifolds ([8]) and TQFTs ([4]). The aim of this paper is to give an overview of the algebraic results related to this modified quantization and prove a few straightforward results.

2.2. A modified version of Uq​𝔰​𝔩​(2){U_{q}{\mathfrak{sl}(2)}}

Let UqH​𝔰​𝔩​(2){U_{q}^{H}{\mathfrak{sl}(2)}} be the β„‚\mathbb{C}-algebra given by generators E,F,K,Kβˆ’1,HE,F,K,K^{-1},H and relations in Equation (1) plus the relations:

H​K\displaystyle HK =K​H,\displaystyle=KH, [H,E]\displaystyle[H,E] =2​E,\displaystyle=2E, [H,F]\displaystyle[H,F] =βˆ’2​F.\displaystyle=-2F.

The algebra UqH​𝔰​𝔩​(2){U_{q}^{H}{\mathfrak{sl}(2)}} is a Hopf algebra where the coproduct, counit and antipode are defined by Equations (2)–(4) and by

Δ​(H)\displaystyle\Delta(H) =HβŠ—1+1βŠ—H,\displaystyle=H\otimes 1+1\otimes H, Ρ​(H)\displaystyle\varepsilon(H) =0,\displaystyle=0, S​(H)\displaystyle S(H) =βˆ’H.\displaystyle=-H.

Define UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}} to be the Hopf algebra UqH​𝔰​𝔩​(2){U_{q}^{H}{\mathfrak{sl}(2)}} modulo the relations Er=Fr=0E^{r}=F^{r}=0.

Let VV be a finite dimensional UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-module. An eigenvalue Ξ»βˆˆβ„‚\lambda\in\mathbb{C} of the operator H:Vβ†’VH:V\to V is called a weight of VV and the associated eigenspace is called a weight space. A vector vv in the Ξ»\lambda-eigenspace of HH is a weight vector of weight Ξ»\lambda, i.e. H​v=λ​vHv=\lambda v. We call VV a weight module if VV splits as a direct sum of weight spaces and qH=K{q}^{H}=K as operators on VV, i.e. K​v=qλ​vKv=q^{\lambda}v for any vector vv of weight Ξ»\lambda. Let π’ž\mathscr{C} be the category of finite dimensional weight UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-modules.

Remark 2.1.

The algebra UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}} does not have a requirement on KrK^{r}, allowing modules in π’ž\mathscr{C} to have non-integral weights. The requirement Er=Fr=0E^{r}=F^{r}=0 forces modules to be highest weight modules. As we will see the generator HH is used to define a braiding on π’ž\mathscr{C}. Here the main point is that one must know the action of HH and not just the action of KK which acts as a kind of exponential of HH.

Since UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}} is a Hopf algebra then π’ž\mathscr{C} is tensor category where the unit 𝕀\mathbb{I} is the 1-dimensional trivial module β„‚\mathbb{C}. Moreover, π’ž\mathscr{C} is β„‚\mathbb{C}-linear: hom-sets are β„‚\mathbb{C}-modules, the composition and tensor product of morphisms are β„‚\mathbb{C}-bilinear, and Endπ’žβ‘(𝕀)=ℂ​Id𝕀\operatorname{End}_{\mathscr{C}}(\mathbb{I})=\mathbb{C}\operatorname{Id}_{\mathbb{I}}. When it is clear we denote the unit 𝕀\mathbb{I} by β„‚\mathbb{C}. We say a module VV is simple if has no proper submodules. If VV is simple then Schur’s lemma implies that Endπ’žβ‘(V)=ℂ​IdV\operatorname{End}_{\mathscr{C}}(V)=\mathbb{C}\operatorname{Id}_{V}. If Endπ’žβ‘(V)=ℂ​IdV\operatorname{End}_{\mathscr{C}}(V)=\mathbb{C}\operatorname{Id}_{V} then for f∈Endπ’žβ‘(V)f\in\operatorname{End}_{\mathscr{C}}(V) we denote ⟨f⟩{\left\langle{f}\right\rangle} as the scalar determined by f=⟨fβŸ©β€‹IdVf={\left\langle{f}\right\rangle}\operatorname{Id}_{V}.

We will now recall that the category π’ž\mathscr{C} is a ribbon category. Let VV and WW be objects of π’ž\mathscr{C}. Let {vi}\{v_{i}\} be a basis of VV and {viβˆ—}\{v_{i}^{*}\} be a dual basis of Vβˆ—=Homℂ⁑(V,β„‚)V^{*}=\operatorname{Hom}_{\mathbb{C}}(V,\mathbb{C}). Then

coevV⟢:\displaystyle\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{coev}}}_{V}: β„‚β†’VβŠ—Vβˆ—,Β given by ​1β†¦βˆ‘viβŠ—viβˆ—,\displaystyle\mathbb{C}\rightarrow V\otimes V^{*},\text{ given by }1\mapsto\sum v_{i}\otimes v_{i}^{*}, evV⟢:\displaystyle\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{ev}}}_{V}: Vβˆ—βŠ—Vβ†’β„‚,Β given by ​fβŠ—w↦f​(w)\displaystyle V^{*}\otimes V\rightarrow\mathbb{C},\text{ given by }f\otimes w\mapsto f(w)

are duality morphisms of π’ž\mathscr{C}. In [29] Ohtsuki truncates the usual formula of the hh-adic quantum 𝔰​𝔩​(2){\mathfrak{sl}(2)} RR-matrix to define an operator on VβŠ—WV\otimes W by

(5) R=qHβŠ—H/2β€‹βˆ‘n=0rβˆ’1{1}2​n{n}!​qn​(nβˆ’1)/2​EnβŠ—Fn.R={q}^{H\otimes H/2}\sum_{n=0}^{r-1}\frac{\{1\}^{2n}}{\{n\}!}{q}^{n(n-1)/2}E^{n}\otimes F^{n}.

where qHβŠ—H/2q^{H\otimes H/2} is the operator given by

qHβŠ—H/2​(vβŠ—vβ€²)=qλ​λ′/2​vβŠ—vβ€²q^{H\otimes H/2}(v\otimes v^{\prime})=q^{\lambda\lambda^{\prime}/2}v\otimes v^{\prime}

for weight vectors vv and vβ€²v^{\prime} of weights of Ξ»\lambda and Ξ»β€²\lambda^{\prime}. The RR-matrix is not an element in UΒ―qH​𝔰​𝔩​(2)βŠ—UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}\otimes{\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}, however the action of RR on the tensor product of two objects of π’ž\mathscr{C} is a well defined linear map on such a tensor product. Moreover, RR gives rise to a braiding cV,W:VβŠ—Wβ†’WβŠ—Vc_{V,W}:V\otimes W\rightarrow W\otimes V on π’ž\mathscr{C} defined by vβŠ—w↦τ​(R​(vβŠ—w))v\otimes w\mapsto\tau(R(v\otimes w)) where Ο„\tau is the permutation xβŠ—y↦yβŠ—xx\otimes y\mapsto y\otimes x. Also, let ΞΈ\theta be the operator given by

(6) ΞΈ=Krβˆ’1β€‹βˆ‘n=0rβˆ’1{1}2​n{n}!​qn​(nβˆ’1)/2​S​(Fn)​qβˆ’H2/2​En\theta=K^{r-1}\sum_{n=0}^{r-1}\frac{\{1\}^{2n}}{\{n\}!}{q}^{n(n-1)/2}S(F^{n}){q}^{-H^{2}/2}E^{n}

where qβˆ’H/2q^{-H/2} is an operator defined by on a weight vector vΞ»v_{\lambda} by qβˆ’H2/2.vΞ»=qβˆ’Ξ»2/2​vΞ».q^{-H^{2}/2}.v_{\lambda}=q^{-\lambda^{2}/2}v_{\lambda}. Ohtsuki shows that the family of maps ΞΈV:Vβ†’V\theta_{V}:V\rightarrow V in π’ž\mathscr{C} defined by vβ†¦ΞΈβˆ’1​vv\mapsto\theta^{-1}v is a twist (see [27, 29]).

Now the ribbon structure on π’ž\mathscr{C} yields right duality morphisms

(7) evV⟡=evV⟢cV,Vβˆ—(ΞΈVβŠ—IdVβˆ—)Β andΒ coevV⟡=(IdVβˆ—βŠ—ΞΈV)cV,Vβˆ—coevV⟢\stackrel{{\scriptstyle\longleftarrow}}{{\operatorname{ev}}}_{V}=\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{ev}}}_{V}c_{V,V^{*}}(\theta_{V}\otimes\operatorname{Id}_{V^{*}})\text{ and }\stackrel{{\scriptstyle\longleftarrow}}{{\operatorname{coev}}}_{V}=(\operatorname{Id}_{V^{*}}\otimes\theta_{V})c_{V,V^{*}}\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{coev}}}_{V}

which are compatible with the left duality morphisms {coevV⟢}V\{\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{coev}}}_{V}\}_{V} and {evV⟢}V\{\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{ev}}}_{V}\}_{V}. These duality morphisms are given by

coev⟡V:\displaystyle\stackrel{{\scriptstyle\longleftarrow}}{{\operatorname{coev}}}{V}: β„‚β†’Vβˆ—βŠ—V,Β where ​1β†¦βˆ‘Krβˆ’1​viβŠ—viβˆ—,\displaystyle\mathbb{C}\rightarrow V^{*}\otimes V,\text{ where }1\mapsto\sum K^{r-1}v_{i}\otimes v_{i}^{*},
evV⟡:\displaystyle\stackrel{{\scriptstyle\longleftarrow}}{{\operatorname{ev}}}_{V}: VβŠ—Vβˆ—β†’β„‚,Β where ​vβŠ—f↦f​(K1βˆ’r​v).\displaystyle V\otimes V^{*}\rightarrow\mathbb{C},\text{ where }v\otimes f\mapsto f(K^{1-r}v).

The quantum dimension qdim⁑(V)\operatorname{qdim}(V) of an object VV in π’ž\mathscr{C} is the qdim(V)=⟨evV⟡∘coevV⟢⟩=βˆ‘viβˆ—(K1βˆ’rvi)\operatorname{qdim}(V)={\left\langle{\stackrel{{\scriptstyle\longleftarrow}}{{\operatorname{ev}}}_{V}\circ\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{coev}}}_{V}}\right\rangle}=\sum v_{i}^{*}(K^{1-r}v_{i}).

For gβˆˆβ„‚/2​℀g\in\mathbb{C}/2\mathbb{Z}, define π’žg\mathscr{C}_{g} as the full sub-category of weight modules whose weights are all in the class gg (mod 2​℀2\mathbb{Z}). Then π’ž={π’žg}gβˆˆβ„‚/2​℀\mathscr{C}=\{\mathscr{C}_{g}\}_{g\in\mathbb{C}/2\mathbb{Z}} is a β„‚/2​℀\mathbb{C}/2\mathbb{Z}-grading (where β„‚/2​℀\mathbb{C}/2\mathbb{Z} is an additive group): Let Vβˆˆπ’žgV\in\mathscr{C}_{g} and Vβ€²βˆˆπ’žgβ€²V^{\prime}\in\mathscr{C}_{g^{\prime}}. Then the weights of VβŠ—Vβ€²V\otimes V^{\prime} are congruent to g+gβ€²mod2​℀g+g^{\prime}\mod 2\mathbb{Z}, and so the tensor product is in π’žg+gβ€²\mathscr{C}_{g+g^{\prime}}. Also if gβ‰ gβ€²g\neq g^{\prime} then Homπ’žβ‘(V,Vβ€²)=0\operatorname{Hom}_{\mathscr{C}}(V,V^{\prime})=0 since morphisms in π’ž\mathscr{C} preserve weights. Finally, for f∈Vβˆ—=Homℂ⁑(V,β„‚)f\in V^{*}=\operatorname{Hom}_{\mathbb{C}}(V,\mathbb{C}) then by definition the action of HH on ff is given by (H​f)​(v)=f​(S​(H)​v)=βˆ’f​(H​v)(Hf)(v)=f(S(H)v)=-f(Hv) and so Vβˆ—βˆˆπ’žβˆ’gV^{*}\in\mathscr{C}_{-g}.

3. Modified traces on the projective modules.

Let π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} be the full subcategory of π’ž\mathscr{C} consisting of projective UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-modules. The subcategory π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} is an ideal (see also [20]): it is closed under retracts (i.e.Β if Wβˆˆπ–―π—‹π—ˆπ—ƒW\in{\mathsf{Proj}} and Ξ±:Xβ†’W\alpha:X\to W and Ξ²:Wβ†’X\beta:W\to X satisfy β∘α=IdX\beta\circ\alpha=\operatorname{Id}_{X}, then Xβˆˆπ–―π—‹π—ˆπ—ƒX\in{\mathsf{Proj}}) and if XX is in π’ž\mathscr{C} and YY is in π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} then XβŠ—YX\otimes Y is in π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}}.

For any objects V,WV,W of π’ž\mathscr{C} and any endomorphism ff of VβŠ—WV\otimes W, set

(8) ptrL(f)=(evVβŸΆβŠ—IdW)∘(IdVβˆ—βŠ—f)∘(coevVβŸ΅βŠ—IdW)∈Endπ’ž(W),\operatorname{ptr}_{L}(f)=(\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{ev}}}_{V}\otimes\operatorname{Id}_{W})\circ(\operatorname{Id}_{V^{*}}\otimes f)\circ(\stackrel{{\scriptstyle\longleftarrow}}{{\operatorname{coev}}}_{V}\otimes\operatorname{Id}_{W})\in\operatorname{End}_{\mathscr{C}}(W),

and

(9) ptrR(f)=(IdVβŠ—evW⟡)∘(fβŠ—IdWβˆ—)∘(IdVβŠ—coevW⟢)∈Endπ’ž(V).\operatorname{ptr}_{R}(f)=(\operatorname{Id}_{V}\otimes\stackrel{{\scriptstyle\longleftarrow}}{{\operatorname{ev}}}_{W})\circ(f\otimes\operatorname{Id}_{W^{*}})\circ(\operatorname{Id}_{V}\otimes\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{coev}}}_{W})\in\operatorname{End}_{\mathscr{C}}(V).
Definition 3.1.

A trace on π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} is a family of linear functions

{𝗍V:Endπ’žβ‘(V)β†’K}\{\operatorname{\mathsf{t}}_{V}:\operatorname{End}_{\mathscr{C}}(V)\rightarrow K\}

where VV runs over all objects of π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} and such that the following two conditions hold.

  1. (1)

    If Uβˆˆπ–―π—‹π—ˆπ—ƒU\in{\mathsf{Proj}} and W∈Ob⁑(π’ž)W\in\operatorname{Ob}(\mathscr{C}) then for any f∈Endπ’žβ‘(UβŠ—W)f\in\operatorname{End}_{\mathscr{C}}(U\otimes W) we have

    (10) 𝗍UβŠ—W⁑(f)=𝗍U⁑(ptrR⁑(f)).\operatorname{\mathsf{t}}_{U\otimes W}\left(f\right)=\operatorname{\mathsf{t}}_{U}\left(\operatorname{ptr}_{R}(f)\right).
  2. (2)

    If U,Vβˆˆπ–―π—‹π—ˆπ—ƒU,V\in{\mathsf{Proj}} then for any morphisms f:Vβ†’Uf:V\rightarrow U and g:Uβ†’Vg:U\rightarrow V in π’ž\mathscr{C} we have

    (11) 𝗍V⁑(g∘f)=𝗍U⁑(f∘g).\operatorname{\mathsf{t}}_{V}(g\circ f)=\operatorname{\mathsf{t}}_{U}(f\circ g).

4. The center of UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}

The center of the small quantum group is known (see [17]) and its dimension is 3​rβˆ’13r-1. The following proposition is a description of a subalgebra of the center of UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}. Let CC be quantum Casimir element defined by

(12) C=F​E+K​q+Kβˆ’1​qβˆ’1{1}2=E​F+K​qβˆ’1+Kβˆ’1​q{1}2.C=FE+\dfrac{Kq+K^{-1}q^{-1}}{{\left\{1\right\}}^{2}}=EF+\dfrac{Kq^{-1}+K^{-1}q}{{\left\{1\right\}}^{2}}.

Also, let 𝒯r\mathcal{T}_{r} be the rrth Chebyshev polynomial determined by 𝒯r​(X+Xβˆ’12)=Xr+Xβˆ’r2\mathcal{T}_{r}(\frac{X+X^{-1}}{2})=\frac{X^{r}+X^{-r}}{2}.

Proposition 4.1.

The center of UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}} contains the β„‚\mathbb{C}-algebra generated by CC and KΒ±rK^{\pm r} with the relation 𝒯r​({1}22​C)=βˆ’Kr+Kβˆ’r2\mathcal{T}_{r}\left(\frac{{\left\{1\right\}}^{2}}{2}C\right)=-\frac{K^{r}+K^{-r}}{2}.

Proof.

First, it is easy to see that the elements CC and KΒ±rK^{\pm r} are central. Next, we will show the relation stated in the proposition holds. Using induction on kβˆˆβ„•k\in\mathbb{N} one can show that

(13) ∏i=0kβˆ’1(Cβˆ’qβˆ’2​iβˆ’1​K+q2​i+1​Kβˆ’1{1}2)=Ek​Fk.\prod_{i=0}^{k-1}\left(C-\dfrac{q^{-2i-1}K+q^{2i+1}K^{-1}}{{\left\{1\right\}}^{2}}\right)=E^{k}F^{k}.

On the other hand, we have

2​(𝒯r​(X+Xβˆ’12)βˆ’π’―r​(Y+Yβˆ’12))=(Xr+Xβˆ’r)βˆ’(Yr+Yβˆ’r)2{\left(\mathcal{T}_{r}{\left(\frac{X+X^{-1}}{2}\right)}-\mathcal{T}_{r}{\left(\frac{Y+Y^{-1}}{2}\right)}\right)}={\left(X^{r}+X^{-r}\right)}-{\left(Y^{r}+Y^{-r}\right)}
=Xβˆ’r​(Xrβˆ’Yr)​(Xrβˆ’Yβˆ’r)=∏i=0rβˆ’1Xβˆ’1​(Xβˆ’q2​i​Y)​(Xβˆ’qβˆ’2​i​Yβˆ’1)=X^{-r}(X^{r}-Y^{r})(X^{r}-Y^{-r})=\prod_{i=0}^{r-1}X^{-1}(X-q^{2i}Y)(X-q^{-2i}Y^{-1})
=∏i=0rβˆ’1(X+Xβˆ’1βˆ’Y​q2​iβˆ’Yβˆ’1​qβˆ’2​i).=\displaystyle{\prod_{i=0}^{r-1}{\left(X+X^{-1}-Yq^{2i}-Y^{-1}q^{-2i}\right)}}.

Combine the last expression with the fact that the product of Equation (13) vanishes for k=rk=r we obtain the following polynomial relation of degree rr for CC:

2​𝒯r​({1}22​C)βˆ’2​𝒯r​(q​K+qβˆ’1​Kβˆ’12)=∏i=0rβˆ’1({1}2​Cβˆ’(q2​i+1​K+qβˆ’2​iβˆ’1​Kβˆ’1))=0.2\mathcal{T}_{r}\left(\frac{{\left\{1\right\}}^{2}}{2}C\right)-2\mathcal{T}_{r}{\left(\frac{qK+q^{-1}K^{-1}}{2}\right)}=\prod_{i=0}^{r-1}\left({{\left\{1\right\}}^{2}}C-{\left(q^{2i+1}K+q^{-2i-1}K^{-1}\right)}\right)=0.

Thus, 𝒯r​({1}22​C)=βˆ’Kr+Kβˆ’r2\mathcal{T}_{r}\left(\frac{{\left\{1\right\}}^{2}}{2}C\right)=-\frac{K^{r}+K^{-r}}{2}. ∎

It can be show that the center center of UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}} contains more complicated elements involving the element HH. We don’t need these elements in the rest of the paper.

5. Simple UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-modules

For each n∈{0,…,rβˆ’1}n\in\{0,\ldots,r-1\} let SnS_{n} be the usual (n+1)(n+1)-dimensional simple highest weight UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-module with highest weight nn. The module SnS_{n} is a highest weight module with a highest weight vector s0s_{0} such that E​s0=0Es_{0}=0 and H​s0=n​s0Hs_{0}=ns_{0}. Then {s0,s1,…,sn}\{s_{0},s_{1},\ldots,s_{n}\} is a basis of SnS_{n} where F​si=si+1Fs_{i}=s_{i+1}, H.si=(nβˆ’2​i)​siH.s_{i}=(n-2i)s_{i}, E.s0=0=Fn+1.s0E.s_{0}=0=F^{n+1}.s_{0} and E.si={i}​{n+1βˆ’i}{1}2​siβˆ’1E.s_{i}=\frac{{\left\{i\right\}}{\left\{n+1-i\right\}}}{{\left\{1\right\}}^{2}}s_{i-1}. The quantum dimension SiS_{i} is qdim⁑(Sn)=(βˆ’1)n​{n+1}{1}\operatorname{qdim}(S_{n})=(-1)^{n}\frac{{\left\{n+1\right\}}}{{\left\{1\right\}}}.

In U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}}-mod the modules SnS_{n} are the only simple modules up to isomorphism. However in π’ž\mathscr{C}, there is a (n+1)(n+1)-dimensional simple UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-module with highest weight n+rn+r which not isomorphic to SiS_{i}, as follows. For kβˆˆβ„€k\in\mathbb{Z}, let β„‚k​rH\mathbb{C}^{H}_{kr} be the one dimensional modules where both EE and FF act by zero and HH acts by k​rkr. The degree of β„‚k​rH\mathbb{C}^{H}_{kr} is k​r​mod​ 2kr\ {\rm mod}\ 2. Then SnβŠ—β„‚k​rHS_{n}\otimes\mathbb{C}^{H}_{kr} is the simple highest weight module with highest weight n+k​rn+kr. As a U~q​𝔰​𝔩​(2){\widetilde{U}_{q}{\mathfrak{sl}(2)}}-module β„‚k​rH\mathbb{C}^{H}_{kr} is isomorphic to the trivial module. The modules β„‚k​rH\mathbb{C}^{H}_{kr} are important tools in the work of [8, 4]. We also use another notation to distinguish among these modules, those that are in the degree 0Β―\overline{0} part of π’ž\mathscr{C}: we define for any kβˆˆβ„€k\in\mathbb{Z},

(14) Οƒk=β„‚2​k​rβ€²Hβˆˆπ’ž0Β―Β whereΒ rβ€²=r/2​ if ​r∈2​℀ andΒ rβ€²=r​ else.\sigma^{k}=\mathbb{C}^{H}_{2kr^{\prime}}\in\mathscr{C}_{\overline{0}}\quad\text{ where }\quad r^{\prime}=r/2\text{ if }r\in 2\mathbb{Z}\quad\text{ and }\quad r^{\prime}=r\text{ else.}

Next we consider a larger class of finite dimensional highest weight modules: for each Ξ±βˆˆβ„‚\alpha\in\mathbb{C} we let VΞ±V_{\alpha} be the rr-dimensional highest weight UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-module of highest weight Ξ±+rβˆ’1\alpha+r-1. The modules VΞ±V_{\alpha} has a basis {v0,…,vrβˆ’1}\{v_{0},\ldots,v_{r-1}\} whose action is given by

(15) H.vi=(Ξ±+rβˆ’1βˆ’2​i)​vi,E.vi={i}​{iβˆ’Ξ±}{1}2​viβˆ’1,F.vi=vi+1.H.v_{i}=(\alpha+r-1-2i)v_{i},\quad E.v_{i}=\frac{{\left\{i\right\}}{\left\{i-\alpha\right\}}}{{\left\{1\right\}}^{2}}v_{i-1},\quad F.v_{i}=v_{i+1}.

For all Ξ±βˆˆβ„‚\alpha\in\mathbb{C}, the quantum dimension of VΞ±V_{\alpha} is zero:

qdim⁑(VΞ±)=βˆ‘i=0rβˆ’1viβˆ—β€‹(K1βˆ’r​vi)=βˆ‘i=0rβˆ’1q(rβˆ’1)​(Ξ±+rβˆ’1βˆ’2​i)=q(rβˆ’1)​(Ξ±+rβˆ’1)​1βˆ’q2​r1βˆ’q2=0.\operatorname{qdim}(V_{\alpha})=\sum_{i=0}^{r-1}v_{i}^{*}(K^{1-r}v_{i})=\sum_{i=0}^{r-1}q^{(r-1)(\alpha+r-1-2i)}=q^{(r-1)(\alpha+r-1)}\frac{1-q^{2r}}{1-q^{2}}=0.

We say VΞ±V_{\alpha} is typical if α∈(β„‚βˆ–β„€)βˆͺr​℀\alpha\in(\mathbb{C}\setminus\mathbb{Z})\cup r\mathbb{Z}, otherwise it is atypical. If VΞ±V_{\alpha} is typical then it is simple, since it is generated by any of the basis vectors viv_{i} (see Equation (15)).

Definition 5.1.

The character of a weight module Vβˆˆπ’žV\in\mathscr{C} is χ​(V)=βˆ‘Ξ±dim(V​(Ξ±))​XΞ±βˆˆβ„€β€‹[β„‚]\chi(V)=\sum_{\alpha}\dim(V(\alpha))X^{\alpha}\in\mathbb{Z}[\mathbb{C}] where V​(Ξ±)V(\alpha) is the Ξ±\alpha-eigenspace of the action of HH on VV and XΞ±X^{\alpha} is a notation for the element Ξ±βˆˆβ„‚\alpha\in\mathbb{C} seen in the group ring ℀​[β„‚]\mathbb{Z}[\mathbb{C}].

Let [k]X=Xkβˆ’1+Xkβˆ’3+β‹―+Xβˆ’(kβˆ’1){\left[k\right]}_{X}=X^{k-1}+X^{k-3}+\cdots+X^{-(k-1)}. Then for Ξ±βˆˆΒ¨β€‹β„‚\alpha\in{\ddot{}\mathbb{C}} and i∈{0,…,rβˆ’1}i\in\{0,\ldots,r-1\}, one has

(16) χ​(VΞ±)=Xα​[r]Xandχ​(Si)=[i+1]X.\chi(V_{\alpha})=X^{\alpha}{\left[r\right]}_{X}{\quad\text{and}\quad}\chi(S_{i})={\left[i+1\right]}_{X}.

Let V,Wβˆˆπ’žV,W\in\mathscr{C} and define

(17) Ξ¦V,W=(IdWβŠ—evV⟡)∘(cV,WβŠ—IdVβˆ—)∘(cW,VβŠ—IdVβˆ—)∘(IdWβŠ—coevV⟢)∈End(W).\Phi_{V,W}=(\operatorname{Id}_{W}\otimes\stackrel{{\scriptstyle\longleftarrow}}{{\operatorname{ev}}}_{V})\circ(c_{V,W}\otimes\operatorname{Id}_{V^{*}})\circ(c_{W,V}\otimes\operatorname{Id}_{V^{*}})\circ(\operatorname{Id}_{W}\otimes\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{coev}}}_{V})\in\operatorname{End}(W).
Theorem 5.2.

(1) If Ξ±Β―βˆˆβ„‚/2β€‹β„€βˆ–β„€/2​℀\overline{\alpha}\in\mathbb{C}/2\mathbb{Z}\setminus\mathbb{Z}/2\mathbb{Z} then π’žΞ±Β―\mathscr{C}_{\overline{\alpha}} is semi-simple.
(2) If Ξ±,Ξ²βˆˆΒ¨β€‹β„‚=(β„‚βˆ–β„€)βˆͺr​℀\alpha,\beta\in{\ddot{}\mathbb{C}}=(\mathbb{C}\setminus\mathbb{Z})\cup r\mathbb{Z} and Ξ±+Ξ²βˆ‰β„€\alpha+\beta\notin\mathbb{Z} then VΞ±βŠ—VΞ²β‰ƒβŠ•k∈HrVΞ±+Ξ²+kV_{\alpha}\otimes V_{\beta}\simeq\oplus_{k\in H_{r}}V_{\alpha+\beta+k} where Hr={βˆ’(rβˆ’1),βˆ’(rβˆ’3),…,rβˆ’1}H_{r}=\{-(r-1),-(r-3),\ldots,r-1\}.
(3) All the typical modules are projective.

Proof.

Let Ξ±βˆˆβ„‚βˆ–β„€\alpha\in\mathbb{C}\setminus\mathbb{Z} and define cΞ±=qΞ±+r+qβˆ’Ξ±βˆ’r{1}2c_{\alpha}=\frac{q^{\alpha+r}+q^{-\alpha-r}}{{\left\{1\right\}}^{2}}. Proposition 4.1 implies that CC satisfies the relation ∏i=0rβˆ’1(Cβˆ’cΞ±+2​i)=0\prod_{i=0}^{r-1}\left(C-c_{\alpha+2i}\right)=0 on π’žΞ±Β―\mathscr{C}_{\overline{\alpha}}. Since cΞ±+2​iβˆ’cΞ±+2​j={iβˆ’j}​{Ξ±+r+i+j}{1}2c_{\alpha+2i}-c_{\alpha+2j}=\frac{{\left\{i-j\right\}}{\left\{\alpha+r+i+j\right\}}}{{\left\{1\right\}}^{2}}, this polynomial has only simple roots. Hence any Wβˆˆπ’žΞ±Β―W\in\mathscr{C}_{\overline{\alpha}} splits as the direct sum of the eigenspaces of CC. It is enough to show that WW is semi-simple when CC acts by a scalar (say cΞ±c_{\alpha}) on WW. Let VV be a maximal semi-simple submodule of WW and suppose Vβ‰ WV\neq W. The weights of WW differ by elements of 2​℀2\mathbb{Z}. In particular, they are totally ordered and there is a weight vector ww of Wβˆ–VW\setminus V of maximal weight Ξ»\lambda. Hence E.w∈VE.w\in V and F​E.w=(Cβˆ’K​q+Kβˆ’1​qβˆ’1{1}2).w=0FE.w={\left(C-\dfrac{Kq+K^{-1}q^{-1}}{{\left\{1\right\}}^{2}}\right)}.w=0 because it is proportional to ww and also in VV. It follows that Ξ»=Ξ±+rβˆ’1\lambda=\alpha+r-1 modulo 2. Then by Equation (13), Erβˆ’1​Frβˆ’1.w=ν​wE^{r-1}F^{r-1}.w=\nu w where Ξ½=∏i=1rβˆ’1(cΞ±βˆ’cΞ±βˆ’2​i)β‰ 0.\nu=\prod_{i=1}^{r-1}(c_{\alpha}-c_{\alpha-2i})\neq 0. Thus, E.w=1ν​Er​Frβˆ’1.w=0E.w=\frac{1}{\nu}E^{r}F^{r-1}.w=0 and ww is an highest weight vector. It follows that ww generates a module Vβ€²V^{\prime} isomorphic to the simple module VΞ»βˆ’r+1V_{\lambda-r+1} where V∩Vβ€²={0}V\cap V^{\prime}=\{0\}. This contradicts the maximality of VV and so V=WV=W.

The direct sum decomposition of VΞ±βŠ—VΞ²V_{\alpha}\otimes V_{\beta} follows from a straightforward calculation using the character formula for a typical module. Finally, we prove the last statement of the theorem in two cases: 1) if VΞ±V_{\alpha} is a typical module with Ξ±βˆˆβ„‚βˆ–β„€\alpha\in\mathbb{C}\setminus\mathbb{Z} then the previous parts of the theorem imply that VΞ±V_{\alpha} is projective. 2) If VΞ±V_{\alpha} is a typical module with Ξ±=r​n\alpha=rn then it can be shown (see Lemma 6.6) that the morphism Ξ¦VΞ²,Vr​n\Phi_{V_{\beta},V_{rn}} defined in (17) is non-zero for any Ξ²βˆˆβ„‚βˆ–β„€\beta\in\mathbb{C}\setminus\mathbb{Z}. This morphism can be decomposed into the composition g∘fg\circ f where f:Vr​nβ†’VΞ²βŠ—Vr​nβŠ—VΞ²βˆ—f:V_{rn}\to V_{\beta}\otimes V_{rn}\otimes V_{\beta}^{*} and g:VΞ²βŠ—Vr​nβŠ—VΞ²βˆ—β†’Vr​ng:V_{\beta}\otimes V_{rn}\otimes V_{\beta}^{*}\to V_{rn} are the obvious morphisms. But VΞ²βŠ—Vr​nβŠ—VΞ²βˆ—V_{\beta}\otimes V_{rn}\otimes V_{\beta}^{*} is projective because it is of the form VΞ²βŠ—WV_{\beta}\otimes W with VΞ²V_{\beta} projective. Furthermore, 1⟨g∘fβŸ©β€‹(g∘f)=IdVr​n\frac{1}{{\left\langle{g\circ f}\right\rangle}}(g\circ f)=\operatorname{Id}_{V_{rn}}. Since the class of projective modules is closed under retracts then Vr​nV_{rn} is projective. ∎

Lemma 5.3.

Every simple module of π’ž\mathscr{C} is isomorphic to exactly one of the modules in the list:

  • β€’

    SnβŠ—β„‚k​rHS_{n}\otimes\mathbb{C}^{H}_{kr}, for n=0,β‹―,rβˆ’2n=0,\cdots,r-2 and kβˆˆβ„€k\in\mathbb{Z},

  • β€’

    VΞ±V_{\alpha} for α∈(β„‚βˆ–β„€)βˆͺr​℀\alpha\in(\mathbb{C}\setminus\mathbb{Z})\cup r\mathbb{Z}.

Proof.

Let WW be a simple UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}}-module in π’ž\mathscr{C}. Then WW is uniquely determined, up to isomorphism, by its highest weight Ξ»βˆˆβ„‚\lambda\in\mathbb{C}. The lemma follows from the fact that the highest weight of modules in the above list is in bijection with elements of β„‚\mathbb{C}. ∎

Note in the above lemma the modules β„‚k​rH\mathbb{C}^{H}_{kr} and Srβˆ’1βŠ—β„‚k​rHS_{r-1}\otimes\mathbb{C}^{H}_{kr} are obtained by the isomorphisms β„‚k​rHβ‰…S0βŠ—β„‚k​rH\mathbb{C}^{H}_{kr}\cong S_{0}\otimes\mathbb{C}^{H}_{kr} and Srβˆ’1βŠ—β„‚k​rHβ‰…Vk​rS_{r-1}\otimes\mathbb{C}^{H}_{kr}\cong V_{kr}, respectively.

Theorem 5.4.

There exists a unique trace on π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} up to multiplication by an element of β„‚\mathbb{C}. In particular, there is a unique trace 𝗍={𝗍V}Vβˆˆπ–―π—‹π—ˆπ—ƒ\operatorname{\mathsf{t}}=\left\{\operatorname{\mathsf{t}}_{V}\right\}_{V\in{\mathsf{Proj}}} on π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} such that for any f∈Endπ’žβ‘(V0)f\in\operatorname{End}_{\mathscr{C}}(V_{0}) we have 𝗍V0⁑(f)=(βˆ’1)rβˆ’1β€‹βŸ¨f⟩\operatorname{\mathsf{t}}_{V_{0}}(f)=(-1)^{r-1}{\left\langle{f}\right\rangle}.

Proof.

The proof follows from results of [20, 24]. Here we explain this proof without recalling the definitions given in these papers: In [24] we show that if Ξ±βˆˆβ„‚βˆ–12​℀\alpha\in\mathbb{C}\setminus\frac{1}{2}\mathbb{Z} then VΞ±V_{\alpha} is an ambidextrous object in π’ž\mathscr{C}. In [20] we show that an ambidextrous object JJ leads to the existence of a unique (up to a constant) trace 𝗍\operatorname{\mathsf{t}} on the ideal ℐJ\mathcal{I}_{J} generated by JJ. When JJ is simple then the trace is uniquely determined by the assignment 𝗍J⁑(f)=cβ€‹βŸ¨f⟩\operatorname{\mathsf{t}}_{J}(f)=c{\left\langle{f}\right\rangle}, where cc is a constant. Since π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} generated by any VΞ±V_{\alpha} with Ξ±βˆˆβ„‚βˆ–12​℀\alpha\in\mathbb{C}\setminus\frac{1}{2}\mathbb{Z} then there exists a trace with the above property. Finally, since π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} is generated by V0V_{0} the theorem follows. ∎

We define the modified quantum dimension function as

𝖽:Ob⁑(π–―π—‹π—ˆπ—ƒ)β†’K​ by ​𝖽⁑(V)=𝗍V⁑(IdV).\operatorname{\mathsf{d}}:\operatorname{Ob}({\mathsf{Proj}})\to K\;\;\text{ by }\operatorname{\mathsf{d}}(V)=\operatorname{\mathsf{t}}_{V}\left(\operatorname{Id}_{V}\right).

We will prove in Lemma 6.8 that the modified quantum dimension function is given by

(18) 𝖽​(VΞ±)=(βˆ’1)rβˆ’1β€‹βˆj=1rβˆ’1{j}{Ξ±+rβˆ’j}=(βˆ’1)rβˆ’1​r​{Ξ±}{r​α}=(βˆ’1)rβˆ’1​rq(1βˆ’r)​α+β‹―+q(rβˆ’3)​α+q(rβˆ’1)​α{\mathsf{d}}(V_{\alpha})=(-1)^{r-1}\prod_{j=1}^{r-1}\frac{{\left\{j\right\}}}{{\left\{\alpha+r-j\right\}}}=(-1)^{r-1}\frac{r\,{\left\{\alpha\right\}}}{{\left\{r\alpha\right\}}}={\color[rgb]{0,1,0}\frac{(-1)^{r-1}r}{q^{(1-r)\alpha}+\cdots+q^{(r-3)\alpha}+q^{(r-1)\alpha}}}

for Ξ±βˆˆΒ¨β€‹β„‚\alpha\in{\ddot{}\mathbb{C}}.

6. Projective modules

Recall that an highest weight vector v∈Vv\in V is a weight vector such that E​v=0Ev=0. We call a weight vector vv dominant if (F​E)2​v=0(FE)^{2}v=0 (in particular, a highest weight vector is dominant). It is well known that a highest weight vector vv of a module VV generates a submodule with basis {Fi​v}\{F^{i}v\}. The following proposition describes the submodule generated by a dominant weight vector.

Proposition 6.1.

Let v∈Vv\in V be a dominant vector of weight i∈{0,1,…,rβˆ’2}i\in\{0,1,\ldots,r-2\} and let j=rβˆ’2βˆ’ij=r-2-i. Consider the following 2​r2r vectors of VV defined by

(19) 𝗐iH=v,𝗐rβˆ’jR=E​𝗐iH,𝗐iS=F​𝗐rβˆ’jR,𝗐jβˆ’rL=Fi+1​𝗐iH,\mathsf{w}^{H}_{i}=v,\qquad\mathsf{w}^{R}_{r-j}=E\mathsf{w}^{H}_{i},\qquad\mathsf{w}^{S}_{i}=F\mathsf{w}^{R}_{r-j},\qquad\mathsf{w}^{L}_{j-r}=F^{i+1}\mathsf{w}^{H}_{i},
(20) 𝗐iβˆ’2​kH\displaystyle\mathsf{w}^{H}_{i-2k} =Fk​𝗐iH\displaystyle=F^{k}\mathsf{w}^{H}_{i} and 𝗐iβˆ’2​kS\displaystyle\mathsf{w}^{S}_{i-2k} =Fk​𝗐iS\displaystyle=F^{k}\mathsf{w}^{S}_{i} for ​k={0​⋯​i},\displaystyle\text{ for }k=\{0\cdots i\},
(21) 𝗐rβˆ’j+2​kR\displaystyle\mathsf{w}^{R}_{r-j+2k} =Ek​𝗐rβˆ’jR\displaystyle=E^{k}\mathsf{w}^{R}_{r-j} and 𝗐jβˆ’2​kβˆ’rL\displaystyle\mathsf{w}^{L}_{j-2k-r} =Fk​𝗐jβˆ’rL\displaystyle=F^{k}\mathsf{w}^{L}_{j-r} for ​k={0​⋯​j}.\displaystyle\text{ for }k=\{0\cdots j\}.

Then the vector space they generate is a submodule of VV and the following relations holds in VV (whenever the involved vectors are defined):

(22) H​𝗐kX\displaystyle H\mathsf{w}^{X}_{k} =k​𝗐kX,\displaystyle=k\mathsf{w}^{X}_{k}, K​𝗐kX\displaystyle K\mathsf{w}^{X}_{k} =qk​𝗐kX​ for ​X∈{L,R,H,S},\displaystyle=q^{k}\mathsf{w}^{X}_{k}\text{ for }X\in\{L,R,H,S\},
(23) E​𝗐kR\displaystyle E\mathsf{w}^{R}_{k} =𝗐k+2R,\displaystyle=\mathsf{w}^{R}_{k+2}, F​𝗐kX\displaystyle F\mathsf{w}^{X}_{k} =𝗐kβˆ’2X​ for ​X∈{H,S,L},\displaystyle=\mathsf{w}^{X}_{k-2}\text{ for }X\in\{H,S,L\},
(24) Fβ€‹π—βˆ’iH\displaystyle F\mathsf{w}^{H}_{-i} =𝗐jβˆ’rL,\displaystyle=\mathsf{w}^{L}_{j-r}, E​𝗐jβˆ’rL\displaystyle E\mathsf{w}^{L}_{j-r} =π—βˆ’iS,E​𝗐j+rR=E​𝗐iS=Fβ€‹π—βˆ’iS=Fβ€‹π—βˆ’jβˆ’rL=0\displaystyle=\mathsf{w}^{S}_{-i},\qquad E\mathsf{w}^{R}_{j+r}=E\mathsf{w}^{S}_{i}=F\mathsf{w}^{S}_{-i}=F\mathsf{w}^{L}_{-j-r}=0
(25) E​𝗐iβˆ’2​kH\displaystyle E\mathsf{w}^{H}_{i-2k} =Ξ³i,k​𝗐iβˆ’2​k+2H+𝗐iβˆ’2​k+2S,\displaystyle=\gamma_{i,k}\mathsf{w}^{H}_{i-2k+2}+\mathsf{w}^{S}_{i-2k+2}, E​𝗐iβˆ’2​kS\displaystyle E\mathsf{w}^{S}_{i-2k} =Ξ³i,k​𝗐iβˆ’2​k+2S\displaystyle=\gamma_{i,k}\mathsf{w}^{S}_{i-2k+2}
(26) F​𝗐rβˆ’j+2​kR\displaystyle F\mathsf{w}^{R}_{r-j+2k} =βˆ’Ξ³j,k​𝗐rβˆ’j+2​kβˆ’2R\displaystyle=-\gamma_{j,k}\mathsf{w}^{R}_{r-j+2k-2} and E​𝗐jβˆ’2​kβˆ’rL\displaystyle E\mathsf{w}^{L}_{j-2k-r} =βˆ’Ξ³j,k​𝗐jβˆ’2​kβˆ’r+2L\displaystyle=-\gamma_{j,k}\mathsf{w}^{L}_{j-2k-r+2}

where Ξ³n,k=[k]​[nβˆ’k+1]=Ξ³n,nβˆ’k+1\gamma_{n,k}={\left[k\right]}{\left[n-k+1\right]}=\gamma_{n,n-k+1}.

π—βˆ’iH⋯𝗐iH⟢Eπ—βˆ’jβˆ’rL⋯𝗐jβˆ’rLπ—βˆ’j+rR⋯𝗐j+rRπ—βˆ’iS⋯𝗐iS⟡Fβ€‹β†™β†˜β†˜β†™\begin{array}[]{|ccccccccc|cc}\cline{1-9}\cr&&&\mathsf{w}^{H}_{-i}&\cdots&\mathsf{w}^{H}_{i}&&&&&\stackrel{{\scriptstyle E}}{{\longrightarrow}}\\ \mathsf{w}^{L}_{-j-r}&\cdots&\mathsf{w}^{L}_{j-r}&&&&\mathsf{w}^{R}_{-j+r}&\cdots&\mathsf{w}^{R}_{j+r}&&\\ &&&\mathsf{w}^{S}_{-i}&\cdots&\mathsf{w}^{S}_{i}&&&&&\stackrel{{\scriptstyle F}}{{\longleftarrow}}\\ \cline{1-9}\cr\end{array}\put(-220.0,8.0){$\swarrow$}\put(-219.0,-9.0){$\searrow$}\put(-143.0,8.0){$\searrow$}\put(-143.0,-9.0){$\swarrow$}
Figure 1. The weight spaces structure of the module PiP_{i} (here j=rβˆ’2βˆ’ij=r-2-i).
Proof.

First, we show that the Relations (22)–(26) hold. The actions of HH and KK are easily deduced from their commutation relations with EE and FF. The formulas in (23) are restatements of (20)–(21).

Let ci=qi+1+qβˆ’iβˆ’1{1}2=βˆ’qj+1+qβˆ’jβˆ’1{1}2c_{i}=\frac{q^{i+1}+q^{-i-1}}{{\left\{1\right\}}^{2}}=-\frac{q^{j+1}+q^{-j-1}}{{\left\{1\right\}}^{2}}. We have C​𝗐iH=ci​𝗐iH+𝗐iSC\mathsf{w}^{H}_{i}=c_{i}\mathsf{w}^{H}_{i}+\mathsf{w}^{S}_{i}. Since F​E​𝗐iS=0FE\mathsf{w}^{S}_{i}=0 then CC acts by the scalar cic_{i} on 𝗐iS\mathsf{w}^{S}_{i}. As CC is central, we have

E​𝗐jβˆ’rL=E​Fβ€‹π—βˆ’iH=(Cβˆ’K​qβˆ’1+Kβˆ’1​q{1}2)​Fi​𝗐iH=Fi​(Cβˆ’ci)​𝗐iH=Fi​𝗐iS=π—βˆ’iS.E\mathsf{w}^{L}_{j-r}=EF\mathsf{w}^{H}_{-i}={\left(C-\frac{Kq^{-1}+K^{-1}q}{{\left\{1\right\}}^{2}}\right)}F^{i}\mathsf{w}^{H}_{i}=F^{i}{\left(C-c_{i}\right)}\mathsf{w}^{H}_{i}=F^{i}\mathsf{w}^{S}_{i}=\mathsf{w}^{S}_{-i}.

Next, Fβ€‹π—βˆ’jβˆ’rL=F​(Fj​Fi+1​𝗐iH)=Fr​𝗐iH=0F\mathsf{w}^{L}_{-j-r}=F(F^{j}F^{i+1}\mathsf{w}^{H}_{i})=F^{r}\mathsf{w}^{H}_{i}=0 so CC acts by cic_{i} on π—βˆ’jβˆ’rL\mathsf{w}^{L}_{-j-r}. Then by induction on k=0​⋯​jβˆ’1k=0\cdots j-1,

E​𝗐2​kβˆ’jβˆ’rL\displaystyle E\mathsf{w}^{L}_{2k-j-r} =E​F​𝗐2​kβˆ’jβˆ’r+2L=(Cβˆ’K​qβˆ’1+Kβˆ’1​q{1}2)​𝗐2​kβˆ’jβˆ’r+2L\displaystyle=EF\mathsf{w}^{L}_{2k-j-r+2}={\left(C-\frac{Kq^{-1}+K^{-1}q}{{\left\{1\right\}}^{2}}\right)}\mathsf{w}^{L}_{2k-j-r+2}
=(ciβˆ’q2​kβˆ’jβˆ’r+1+qβˆ’2​k+j+rβˆ’1{1}2)​𝗐2​kβˆ’jβˆ’r+2L=βˆ’[jβˆ’k]​[k+1]​𝗐2​kβˆ’jβˆ’r+2L\displaystyle={\left(c_{i}-\frac{q^{2k-j-r+1}+q^{-2k+j+r-1}}{{\left\{1\right\}}^{2}}\right)}\mathsf{w}^{L}_{2k-j-r+2}=-{\left[j-k\right]}{\left[k+1\right]}\mathsf{w}^{L}_{2k-j-r+2}
=βˆ’Ξ³j,jβˆ’k​𝗐2​kβˆ’jβˆ’r+2L.\displaystyle=-\gamma_{j,j-k}\mathsf{w}^{L}_{2k-j-r+2}.

By writing 2​kβˆ’j2k-j as jβˆ’2​(jβˆ’k)j-2(j-k) we obtain the second formula in (26). Moreover, CC acts by cic_{i} on 𝗐2​kβˆ’jβˆ’r+2L=βˆ’Ξ³j,jβˆ’kβˆ’1​E​𝗐2​kβˆ’jβˆ’rL\mathsf{w}^{L}_{2k-j-r+2}=-\gamma_{j,j-k}^{-1}E\mathsf{w}^{L}_{2k-j-r}. Now CC also acts by cic_{i} on 𝗐iβˆ’2​kS=Fk​𝗐iS\mathsf{w}^{S}_{i-2k}=F^{k}\mathsf{w}^{S}_{i}. This implies that for k=1​⋯​ik=1\cdots i,

E​𝗐iβˆ’2​kS\displaystyle E\mathsf{w}^{S}_{i-2k} =E​F​𝗐iβˆ’2​k+2S=(Cβˆ’K​qβˆ’1+Kβˆ’1​q{1}2)​𝗐iβˆ’2​k+2S\displaystyle=EF\mathsf{w}^{S}_{i-2k+2}={\left(C-\frac{Kq^{-1}+K^{-1}q}{{\left\{1\right\}}^{2}}\right)}\mathsf{w}^{S}_{i-2k+2}
=(ciβˆ’qiβˆ’2​k+1+qβˆ’i+2​kβˆ’1{1}2)​𝗐iβˆ’2​k+2S=[k]​[iβˆ’k+1]​𝗐iβˆ’2​k+2S.\displaystyle={\left(c_{i}-\frac{q^{i-2k+1}+q^{-i+2k-1}}{{\left\{1\right\}}^{2}}\right)}\mathsf{w}^{S}_{i-2k+2}={\left[k\right]}{\left[i-k+1\right]}\mathsf{w}^{S}_{i-2k+2}.

Since Er=0E^{r}=0 we have E​𝗐iS=E​(βˆ’1)j[i]!2​[j]!2​Erβˆ’1β€‹π—βˆ’rβˆ’jL=0E\mathsf{w}^{S}_{i}=E\frac{(-1)^{j}}{{\left[i\right]}!^{2}{\left[j\right]}!^{2}}E^{r-1}\mathsf{w}^{L}_{-r-j}=0. This implies that E​F​𝗐rβˆ’jR=E​𝗐iS=0EF\mathsf{w}^{R}_{r-j}=E\mathsf{w}^{S}_{i}=0, so CC acts by the scalar cic_{i} on 𝗐rβˆ’jR\mathsf{w}^{R}_{r-j} and on 𝗐rβˆ’j+2​kR=Ek​𝗐rβˆ’jR\mathsf{w}^{R}_{r-j+2k}=E^{k}\mathsf{w}^{R}_{r-j}. Using this, for k=1​⋯​jk=1\cdots j we have

F​𝗐rβˆ’j+2​kR\displaystyle F\mathsf{w}^{R}_{r-j+2k} =F​E​𝗐rβˆ’j+2​kβˆ’2R=(Cβˆ’K​q+Kβˆ’1​qβˆ’1{1}2)​𝗐rβˆ’j+2​kβˆ’2R\displaystyle=FE\mathsf{w}^{R}_{r-j+2k-2}={\left(C-\frac{Kq+K^{-1}q^{-1}}{{\left\{1\right\}}^{2}}\right)}\mathsf{w}^{R}_{r-j+2k-2}
=(ciβˆ’qrβˆ’j+2​kβˆ’1+qβˆ’r+jβˆ’2​k+1{1}2)​𝗐rβˆ’j+2​kβˆ’2R=βˆ’[k]​[jβˆ’k+1]​𝗐rβˆ’j+2​kβˆ’2R.\displaystyle={\left(c_{i}-\frac{q^{r-j+2k-1}+q^{-r+j-2k+1}}{{\left\{1\right\}}^{2}}\right)}\mathsf{w}^{R}_{r-j+2k-2}=-{\left[k\right]}{\left[j-k+1\right]}\mathsf{w}^{R}_{r-j+2k-2}.

Since Fr=0F^{r}=0 we have Fβ€‹π—βˆ’iS=F​(βˆ’1)j([j]!)2​Frβˆ’1​𝗐r+jR=0F\mathsf{w}^{S}_{-i}=F\frac{(-1)^{j}}{({\left[j\right]}!)^{2}}F^{r-1}\mathsf{w}^{R}_{r+j}=0. Then Cβˆ’ciC-c_{i} sends 𝗐iH↦𝗐iS\mathsf{w}^{H}_{i}\mapsto\mathsf{w}^{S}_{i} so it sends 𝗐iβˆ’2​kH=Fk​𝗐iH↦𝗐iβˆ’2​kS=Fk​𝗐iS\mathsf{w}^{H}_{i-2k}=F^{k}\mathsf{w}^{H}_{i}\mapsto\mathsf{w}^{S}_{i-2k}=F^{k}\mathsf{w}^{S}_{i}. This implies that for k=1​⋯​ik=1\cdots i,

E​𝗐iβˆ’2​kH\displaystyle E\mathsf{w}^{H}_{i-2k} =E​F​𝗐iβˆ’2​k+2H=(Cβˆ’K​qβˆ’1+Kβˆ’1​q{1}2)​𝗐iβˆ’2​k+2H\displaystyle=EF\mathsf{w}^{H}_{i-2k+2}={\left(C-\frac{Kq^{-1}+K^{-1}q}{{\left\{1\right\}}^{2}}\right)}\mathsf{w}^{H}_{i-2k+2}
=(ciβˆ’qiβˆ’2​k+1+qβˆ’i+2​kβˆ’1{1}2)​𝗐iβˆ’2​k+2H+𝗐iβˆ’2​k+2S=[k]​[iβˆ’k+1]​𝗐iβˆ’2​k+2H+𝗐iβˆ’2​k+2S.\displaystyle={\left(c_{i}-\frac{q^{i-2k+1}+q^{-i+2k-1}}{{\left\{1\right\}}^{2}}\right)}\mathsf{w}^{H}_{i-2k+2}+\mathsf{w}^{S}_{i-2k+2}={\left[k\right]}{\left[i-k+1\right]}\mathsf{w}^{H}_{i-2k+2}+\mathsf{w}^{S}_{i-2k+2}.

Now Fβ€‹π—βˆ’iH=F​Fi​𝗐iH=𝗐jβˆ’rLF\mathsf{w}^{H}_{-i}=FF^{i}\mathsf{w}^{H}_{i}=\mathsf{w}^{L}_{j-r}.

The final relation to check is E​𝗐j+rR=0E\mathsf{w}^{R}_{j+r}=0. Relation (25) implies

Eiβ€‹π—βˆ’iH=c​𝗐iH+c′​𝗐iSE^{i}\mathsf{w}^{H}_{-i}=c\mathsf{w}^{H}_{i}+c^{\prime}\mathsf{w}^{S}_{i}

where cc and c′c^{\prime} are constants and c≠0c\neq 0. We have

Erβ€‹π—βˆ’iH=Erβˆ’i​(c​𝗐iH+c′​𝗐iS)=c​Erβˆ’iβˆ’1​𝗐rβˆ’jR=c​E​𝗐r+jR.E^{r}\mathsf{w}^{H}_{-i}=E^{r-i}(c\mathsf{w}^{H}_{i}+c^{\prime}\mathsf{w}^{S}_{i})=cE^{r-i-1}\mathsf{w}^{R}_{r-j}=cE\mathsf{w}^{R}_{r+j}.

Thus, we have shown the Relations (19)–(26) hold.

Finally the vector space generated by the vectors of this proposition is clearly stable by the generators of UΒ―qH​𝔰​𝔩​(2){\overline{U}_{q}^{H}{\mathfrak{sl}(2)}} so it is a submodule of VV. ∎

Projective indecomposable weight modules in π’ž0Β―βˆͺπ’ž1Β―\mathscr{C}_{\overline{0}}\cup\mathscr{C}_{\overline{1}} have a highest weight vector. The following proposition classifies the isomorphism classes of these modules.

Proposition 6.2.

Let i∈{0,1,…,rβˆ’2}i\in\{0,1,\ldots,r-2\} and let j=rβˆ’2βˆ’ij=r-2-i. Denote the vectors of the canonical basis of β„‚2​r\mathbb{C}^{2r} by

(𝗐iH,𝗐iβˆ’2H,…,π—βˆ’iH,𝗐r+jR,𝗐r+jβˆ’2R,…,𝗐rβˆ’jR,𝗐jβˆ’rL,𝗐jβˆ’2βˆ’rL,…,π—βˆ’jβˆ’rL,𝗐iS,𝗐iβˆ’2S,…,π—βˆ’iS).(\mathsf{w}^{H}_{i},\mathsf{w}^{H}_{i-2},\ldots,\mathsf{w}^{H}_{-i},\,\mathsf{w}^{R}_{r+j},\mathsf{w}^{R}_{r+j-2},\ldots,\mathsf{w}^{R}_{r-j},\,\mathsf{w}^{L}_{j-r},\mathsf{w}^{L}_{j-2-r},\ldots,\mathsf{w}^{L}_{-j-r},\,\mathsf{w}^{S}_{i},\mathsf{w}^{S}_{i-2},\ldots,\mathsf{w}^{S}_{-i}).

Then Formulas (19)-(26) define a structure of weight module on β„‚2​r\mathbb{C}^{2r} which we denote by PiP_{i}. Here Prβˆ’1=Srβˆ’1=V0P_{r-1}=S_{r-1}=V_{0}. The module PiP_{i} is projective and indecomposable. Any projective indecomposable weight module Pβˆˆπ’ž0Β―βˆͺπ’ž1Β―P\in\mathscr{C}_{\overline{0}}\cup\mathscr{C}_{\overline{1}} with highest weight (k+1)​rβˆ’iβˆ’2(k+1)r-i-2 is isomorphic to PiβŠ—β„‚k​rHP_{i}\otimes\mathbb{C}^{H}_{kr}.

Proof.

A direct computation shows that the commutation relation E​Fβˆ’F​E=Kβˆ’Kβˆ’1{1}EF-FE=\frac{K-K^{-1}}{{\left\{1\right\}}} is satisfied on PiP_{i}, the other relations are consequences of the fact that EE and FF translate the weight spaces (see Figure 1). Hence Formulas (19)-(26) define a structure of weight module on PiP_{i}.

Proposition 6.1 implies PiP_{i} is a module which is generated by its dominant vector 𝗐iH\mathsf{w}^{H}_{i}. Furthermore, if VV is any weight module then

(27) Hom⁑(Pi,V)≃{v∈V:v​ is dominant of weight ​i}.\operatorname{Hom}(P_{i},V)\simeq\{v\in V:v\text{ is dominant of weight }i\}.

In particular, 𝗐iH\mathsf{w}^{H}_{i} and 𝗐iS\mathsf{w}^{S}_{i} are both dominant vector of PiP_{i} of weight ii, thus End⁑(Pi)\operatorname{End}(P_{i}) is a two dimensional vector space generated by Id:𝗐iH↦𝗐iH\operatorname{Id}:\mathsf{w}^{H}_{i}\mapsto\mathsf{w}^{H}_{i} and the nilpotent map xi:𝗐iH↦𝗐iSx_{i}:\mathsf{w}^{H}_{i}\mapsto\mathsf{w}^{S}_{i} given by the action of Cβˆ’ciC-c_{i}. This implies that End⁑(Pi)\operatorname{End}(P_{i}) is a local algebra and that PiP_{i} is indecomposable (see [30, section 5.2]).

Let now Ο•:Vβ†’Pi\phi:V\to P_{i} be a surjective map in π’ž\mathscr{C}. As CC is central, VV splits as the direct sum of the characteristic spaces of CC and only the summand ViV_{i} of VV associated to the eigenvalue cic_{i} is not included in ker⁑ϕ\ker\phi. We claim that ViβŠ‚ker(Cβˆ’ci)|V2V_{i}\subset\ker(C-c_{i})^{2}_{|V}. Indeed Proposition 4.1 implies that on a module of π’ž0Β―\mathscr{C}_{\overline{0}}, 𝒯r​({1}22​C)=βˆ’1=𝒯r​(q+qβˆ’12)\mathcal{T}_{r}\left(\frac{{\left\{1\right\}}^{2}}{2}C\right)=-1=\mathcal{T}_{r}(\frac{q+q^{-1}}{2}) so ∏i=0rβˆ’1(Cβˆ’c2​i)=0\prod_{i=0}^{r-1}{\left(C-c_{2i}\right)}=0 and on a module of π’ž1Β―\mathscr{C}_{\overline{1}}, we have 𝒯r​({1}22​C)=1=𝒯r​(1+12)\mathcal{T}_{r}\left(\frac{{\left\{1\right\}}^{2}}{2}C\right)=1=\mathcal{T}_{r}(\frac{1+1}{2}) so ∏i=0rβˆ’1(Cβˆ’c2​iβˆ’1)=0.\prod_{i=0}^{r-1}{\left(C-c_{2i-1}\right)}=0. In both cases, all roots of the minimal polynomial of CC have multiplicity at most 22 and this proves that ViβŠ‚ker(Cβˆ’ci)|V2V_{i}\subset\ker(C-c_{i})^{2}_{|V}. Any vector v∈Viv\in V_{i} of weight ii satisfy (Cβˆ’ci)2.v=(F​E)2.v=0(C-c_{i})^{2}.v=(FE)^{2}.v=0 thus is dominant. Let vβˆˆΟ•βˆ’1​({𝗐iH})v\in\phi^{-1}(\{\mathsf{w}^{H}_{i}\}). Then Proposition 6.1 implies that there is an unique ψ∈Hom⁑(Pi,Vi)\psi\in\operatorname{Hom}(P_{i},V_{i}) sending 𝗐iH↦v\mathsf{w}^{H}_{i}\mapsto v. Furthermore, Ο•βˆ˜Οˆβ€‹(𝗐iH)=𝗐iH\phi\circ\psi(\mathsf{w}^{H}_{i})=\mathsf{w}^{H}_{i} so Ο•βˆ˜Οˆ=IdPi\phi\circ\psi=\operatorname{Id}_{P_{i}}. Thus Ο•\phi has a section and PiP_{i} is projective.

For all 0≀i≀rβˆ’10\leq i\leq r-1, Hom⁑(Pi,Si)≃{v∈Si:v​ is dominant of weight ​i}\operatorname{Hom}(P_{i},S_{i})\simeq\{v\in S_{i}:v\text{ is dominant of weight }i\} is one dimensional generated by the surjective morphism Ο€i\pi_{i} sending 𝗐iH\mathsf{w}^{H}_{i} to a highest weight vector viv_{i} of SiS_{i}. Let Pβ€²P^{\prime} be an indecomposable projective module and Ο•β€²:Pβ€²β†’S\phi^{\prime}:P^{\prime}\to S a surjective map to a simple module (obtained by taking the quotient of Pβ€²P^{\prime} by a maximal submodule). Then for some kβˆˆβ„€k\in\mathbb{Z} and some 0≀i≀rβˆ’10\leq i\leq r-1, there exist an isomorphism SβŠ—β„‚r​kH≃SiS\otimes\mathbb{C}^{H}_{rk}\simeq S_{i}. If i=rβˆ’1i=r-1 then SS is simple and projective. In this case Ο•β€²\phi^{\prime} has a section and is an isomorphism since Pβ€²P^{\prime} is indecomposable. Assume now i<rβˆ’1i<r-1 and let P=Pβ€²βŠ—β„‚r​kHP=P^{\prime}\otimes\mathbb{C}^{H}_{rk} which is also projective and indecomposable. Let Ο•=Ο•β€²βŠ—Idβ„‚r​kH:Pβ†’Si\phi=\phi^{\prime}\otimes\operatorname{Id}_{\mathbb{C}^{H}_{rk}}:P\to S_{i}. Since PP is projective, there exists ψ:Pβ†’Pi\psi:P\to P_{i} such that Ο•=Ο€i∘ψ\phi=\pi_{i}\circ\psi. Let v∈Pv\in P such that ϕ​(v)=vi∈Si\phi(v)=v_{i}\in S_{i}. Then Οˆβ€‹(v)βˆˆΟ€iβˆ’1​(vi)∩Pi​(i)=𝗐iH+ℂ​𝗐iS\psi(v)\in\pi_{i}^{-1}(v_{i})\cap P_{i}(i)=\mathsf{w}^{H}_{i}+\mathbb{C}\mathsf{w}^{S}_{i} (here Pi​(i)P_{i}(i) is the weight space of weight ii of PiP_{i}). Hence Οˆβ€‹(v)\psi(v) generates PiP_{i} and ψ\psi is surjective. Finally, as PiP_{i} is projective, ψ\psi has a section and is an isomorphism since PP is indecomposable. Thus we have Pβ€²=PβŠ—β„‚βˆ’r​kH≃PiβŠ—β„‚βˆ’r​kHP^{\prime}=P\otimes\mathbb{C}^{H}_{-rk}\simeq P_{i}\otimes\mathbb{C}^{H}_{-rk}. ∎

The module PiP_{i} is also an injective module in π’ž\mathscr{C} which is self dual. Let i∈{0,…,rβˆ’2}i\in\{0,\dots,r-2\}. After identifying both SiS_{i} and PiP_{i} with their duals, the map Ο€iβˆ—:Siβ†’Pi\pi_{i}^{*}:S_{i}\to P_{i} is an injective morphism with image Spanℂ⁑(𝗐iS,…,π—βˆ’iS)\operatorname{Span}_{\mathbb{C}}(\mathsf{w}^{S}_{i},\ldots,\mathsf{w}^{S}_{-i}). The quotient (ker⁑πi)/Ο€iβˆ—β€‹(Si)(\ker\pi_{i})/\pi_{i}^{*}(S_{i}) is isomorphic to (β„‚rHβŠ•β„‚βˆ’rH)βŠ—Sj(\mathbb{C}^{H}_{r}\oplus\mathbb{C}^{H}_{-r})\otimes S_{j} where j=rβˆ’iβˆ’2j=r-i-2. As in the proof of Proposition 6.2, we let xix_{i} be the nilpotent endomorphism of PiP_{i}, that sends 𝗐iH↦𝗐iS\mathsf{w}^{H}_{i}\mapsto\mathsf{w}^{S}_{i}. We have

End⁑(Pi)=ℂ​IdβŠ•β„‚β€‹xi=ℂ​[xi]/(xi2).\operatorname{End}(P_{i})=\mathbb{C}\operatorname{Id}\oplus\mathbb{C}x_{i}=\mathbb{C}[x_{i}]/(x_{i}^{2}).

Finally, the character of PiP_{i} is given by

χ​(Pi)=2​[i+1]X+(Xr+Xβˆ’r)​[rβˆ’iβˆ’1]X=[r]X​(Xrβˆ’iβˆ’1+Xβˆ’r+i+1).\chi(P_{i})=2[i+1]_{X}+(X^{r}+X^{-r})[r-i-1]_{X}=[r]_{X}(X^{r-i-1}+X^{-r+i+1}).
Corollary 6.3.

For all Ξ±βˆˆΒ¨β€‹β„‚\alpha\in{\ddot{}\mathbb{C}}, VΞ±βŠ—Vβˆ’Ξ±V_{\alpha}\otimes V_{-\alpha} is isomorphic to V0βŠ—V0V_{0}\otimes V_{0}.

Proof.

It follows from the previous proposition that the projective modules of π’ž\mathscr{C} are determined up to isomorphism by their characters, so VΞ±βŠ—Vβˆ’Ξ±V_{\alpha}\otimes V_{-\alpha} and V0βŠ—V0V_{0}\otimes V_{0} are isomorphic because they have the same character. ∎

Recall the definitions of Οƒk\sigma^{k} and rβ€²r^{\prime} in Equation (14).

Corollary 6.4.

Let Ξ±βˆˆβ„‚βˆ–β„€\alpha\in\mathbb{C}\setminus\mathbb{Z}. Let Pβˆˆπ’ž0Β―P\in\mathscr{C}_{\overline{0}} be a projective module, then there exist maps fi:Pβ†’V0βŠ—ΟƒniβŠ—V0βˆ—f_{i}:P\to V_{0}\otimes\sigma^{n_{i}}\otimes V_{0}^{*}, gi:V0βŠ—ΟƒniβŠ—V0βˆ—β†’Pg_{i}:V_{0}\otimes\sigma^{n_{i}}\otimes V_{0}^{*}\to P, fjβ€²:Pβ†’VΞ±+2​kjβŠ—Vβˆ’Ξ±f^{\prime}_{j}:P\to V_{\alpha+2k_{j}}\otimes V_{-\alpha} and gjβ€²:VΞ±+2​kjβŠ—Vβˆ’Ξ±β†’Pg^{\prime}_{j}:V_{\alpha+2k_{j}}\otimes V_{-\alpha}\to P such that

IdP=βˆ‘igi​fi+βˆ‘jgj′​fjβ€²\operatorname{Id}_{P}=\sum_{i}g_{i}f_{i}+\sum_{j}g^{\prime}_{j}f^{\prime}_{j}

where niβˆˆβ„€n_{i}\in\mathbb{Z} and kjβˆˆβ„€βˆ–r′​℀k_{j}\in\mathbb{Z}\setminus r^{\prime}\mathbb{Z}.

Proof.

Consider the epimorphism f=IdPβŠ—evVβˆ’Ξ±βŸΆ:PβŠ—Vβˆ’Ξ±βˆ—βŠ—Vβˆ’Ξ±β†’Pf=\operatorname{Id}_{P}\otimes\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{ev}}}_{V_{-\alpha}}:P\otimes V_{-\alpha}^{*}\otimes V_{-\alpha}\to P. Since PP is projective this morphism has a left inverse gg, i.e. f∘g=IdPf\circ g=\operatorname{Id}_{P}. Now PβŠ—Vβˆ’Ξ±βˆ—βˆˆπ’žΞ±+rβˆ’1Β―P\otimes V_{-\alpha}^{*}\in\mathscr{C}_{\overline{\alpha+r-1}} splits as a direct sum of modules isomorphic to VΞ±+2​kV_{\alpha+2k} (kβˆˆβ„€)(k\in\mathbb{Z}). This produces a factorization of IdP\operatorname{Id}_{P} through the modules VΞ±+2​kβŠ—Vβˆ’Ξ±V_{\alpha+2k}\otimes V_{-\alpha}. Finally, if k=r′​n∈r′​℀k=r^{\prime}n\in r^{\prime}\mathbb{Z}, then VΞ±+2​kβŠ—Vβˆ’Ξ±β‰ƒΟƒnβŠ—VΞ±βŠ—Vβˆ’Ξ±β‰ƒΟƒnβŠ—V0βŠ—V0≃V0βŠ—ΟƒnβŠ—V0βˆ—V_{\alpha+2k}\otimes V_{-\alpha}\simeq\sigma^{n}\otimes V_{\alpha}\otimes V_{-\alpha}\simeq\sigma^{n}\otimes V_{0}\otimes V_{0}\simeq V_{0}\otimes\sigma^{n}\otimes V_{0}^{*}. ∎

The modified trace is non degenerate:

Proposition 6.5.

Let Pβˆˆπ’žP\in\mathscr{C} be a projective module and Vβˆˆπ’žV\in\mathscr{C}. Then the pairing

Homπ’žβ‘(V,P)Γ—Homπ’žβ‘(P,V)β†’β„‚(h1,h2)↦𝗍P​(h1​h2)\begin{array}[]{ccl}\operatorname{Hom}_{\mathscr{C}}(V,P)\times\operatorname{Hom}_{\mathscr{C}}(P,V)&\to&\mathbb{C}\\ (h_{1},h_{2})&\mapsto&\mathsf{t}_{P}(h_{1}h_{2})\end{array}

is non degenerate.

Proof.

Let Ξ±βˆˆβ„‚βˆ–β„€\alpha\in\mathbb{C}\setminus\mathbb{Z}. Let f:Pβ†’VΞ±βŠ—Wf:P\to V_{\alpha}\otimes W and g:VΞ±βŠ—Wβ†’Pg:V_{\alpha}\otimes W\to P be morphisms of π’ž\mathscr{C} such that g​f=IdPgf=\operatorname{Id}_{P}. We show that for any non zero h:Pβ†’Vh:P\to V, there exists hβ€²:Vβ†’Ph^{\prime}:V\to P such that 𝗍P​(h′​h)β‰ 0\mathsf{t}_{P}(h^{\prime}h)\neq 0. Indeed, we have h=h​g​fβ‰ 0h=hgf\neq 0 thus we have a non trivial morphism (hgβŠ—IdWβˆ—)∘(IdVΞ±βŠ—coevW⟢):VΞ±β†’VβŠ—Wβˆ—(hg\otimes\operatorname{Id}_{W^{*}})\circ(\operatorname{Id}_{V_{\alpha}}\otimes\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{coev}}}_{W}):V_{\alpha}\to V\otimes W^{*}. But VΞ±V_{\alpha} is in the category π’žΞ±Β―\mathscr{C}_{\overline{\alpha}} which is semi-simple so the previous map has a left inverse k:VβŠ—Wβˆ—β†’VΞ±k:V\otimes W^{*}\to V_{\alpha}. Then we have that

𝗍VΞ±(k∘(hgβŠ—IdWβˆ—)∘(IdVΞ±βŠ—coevW⟢))=𝖽(VΞ±)β‰ 0.\mathsf{t}_{V_{\alpha}}(k\circ(hg\otimes\operatorname{Id}_{W^{*}})\circ(\operatorname{Id}_{V_{\alpha}}\otimes\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{coev}}}_{W}))={\mathsf{d}}(V_{\alpha})\neq 0.

Let kβ€²=(kβŠ—IdW)∘(IdVβŠ—coevW⟡):Vβ†’VΞ±βŠ—Wk^{\prime}=(k\otimes\operatorname{Id}_{W})\circ(\operatorname{Id}_{V}\otimes\stackrel{{\scriptstyle\longleftarrow}}{{\operatorname{coev}}}_{W}):V\to V_{\alpha}\otimes W and hβ€²=g​kβ€²h^{\prime}=gk^{\prime}. Finally,

𝗍P(hβ€²h)=𝗍P(g(kβ€²h))=𝗍VΞ±βŠ—W((kβ€²h)g)=𝗍VΞ±(k∘(hgβŠ—IdWβˆ—)∘(IdVΞ±βŠ—coevW⟢))=𝖽(VΞ±)β‰ 0.\mathsf{t}_{P}(h^{\prime}h)=\mathsf{t}_{P}(g(k^{\prime}h))=\mathsf{t}_{V_{\alpha}\otimes W}((k^{\prime}h)g)=\mathsf{t}_{V_{\alpha}}(k\circ(hg\otimes\operatorname{Id}_{W^{*}})\circ(\operatorname{Id}_{V_{\alpha}}\otimes\stackrel{{\scriptstyle\longrightarrow}}{{\operatorname{coev}}}_{W}))={\mathsf{d}}(V_{\alpha})\neq 0.

∎

Lemma 6.6 (General Hopf links).

Recall the map Ξ¦\Phi given in Equation (17). For all i,j∈{0,1,…,rβˆ’2}i,j\in\{0,1,\ldots,r-2\} and Ξ±,Ξ²βˆˆΒ¨β€‹β„‚=(β„‚βˆ–β„€)βˆͺr​℀\alpha,\beta\in{\ddot{}\mathbb{C}}=(\mathbb{C}\setminus\mathbb{Z})\cup r\mathbb{Z}, one has

(28) Ξ¦VΞ²,VΞ±\displaystyle\Phi_{V_{\beta},V_{\alpha}} =(βˆ’1)rβˆ’1​r𝖽​(VΞ±)​qα​β​IdVΞ±\displaystyle=\frac{(-1)^{r-1}r}{{\mathsf{d}}(V_{\alpha})}q^{\alpha\beta}\operatorname{Id}_{V_{\alpha}} Ξ¦Si,Sj\displaystyle\Phi_{S_{i},S_{j}} =(βˆ’1)i​{(i+1)​(j+1)}{j+1}​IdVj\displaystyle=(-1)^{i}\frac{{\left\{(i+1)(j+1)\right\}}}{{\left\{j+1\right\}}}\operatorname{Id}_{V_{j}}
(29) Ξ¦Si,VΞ±\displaystyle\Phi_{S_{i},V_{\alpha}} ={(i+1)​α}{Ξ±}​IdVΞ±\displaystyle=\frac{{\left\{(i+1)\alpha\right\}}}{{\left\{\alpha\right\}}}\operatorname{Id}_{V_{\alpha}} Ξ¦Pi,VΞ±\displaystyle\Phi_{P_{i},V_{\alpha}} =(βˆ’1)rβˆ’1​r​q(rβˆ’1βˆ’i)​α+qβˆ’(rβˆ’1βˆ’i)​α𝖽​(VΞ±)​IdVΞ±\displaystyle=(-1)^{r-1}r\frac{q^{(r-1-i)\alpha}+q^{-(r-1-i)\alpha}}{{\mathsf{d}}(V_{\alpha})}\operatorname{Id}_{V_{\alpha}}

Moreover, recall the nilpotent xj∈End⁑(Pj)x_{j}\in\operatorname{End}(P_{j}) given by the action of Cβˆ’cjC-c_{j}, then

(30) Ξ¦Si,Pj\displaystyle\Phi_{S_{i},P_{j}} =(βˆ’1)i​{(i+1)​(j+1)}{j+1}​IdPj+(βˆ’1)i​{1}2​i​{(i+2)​(j+1)}βˆ’(i+2)​{i​(j+1)}{j+1}3​xj,\displaystyle=(-1)^{i}\frac{{\left\{(i+1)(j+1)\right\}}}{{\left\{j+1\right\}}}\operatorname{Id}_{P_{j}}+(-1)^{i}{\left\{1\right\}}^{2}\frac{i{\left\{(i+2)(j+1)\right\}}-(i+2){\left\{i(j+1)\right\}}}{{\left\{j+1\right\}}^{3}}\,x_{j},
(31) Ξ¦V0,Pj\displaystyle\Phi_{V_{0},P_{j}} =(βˆ’1)r+j​2​r​{1}2{j+1}2​xj,\displaystyle=(-1)^{r+j}\frac{2r{\left\{1\right\}}^{2}}{{\left\{j+1\right\}}^{2}}x_{j}, Ξ¦Pi,Pj\displaystyle\Phi_{P_{i},P_{j}} =(βˆ’1)i​2​r​{1}2{j+1}2​(q(i+1)​(j+1)+qβˆ’(i+1)​(j+1))​xj.\displaystyle=\frac{(-1)^{i}2r{\left\{1\right\}}^{2}}{{\left\{j+1\right\}}^{2}}{\left(q^{(i+1)(j+1)}+q^{-(i+1)(j+1)}\right)}x_{j}.
Proof.

For Ξ±βˆˆβ„‚\alpha\in\mathbb{C}, let Ψα:℀​[β„‚]β†’β„‚\Psi_{\alpha}:\mathbb{Z}[\mathbb{C}]\to\mathbb{C} be the map sending Xz↦qα​zX^{z}\mapsto q^{\alpha z}. We start by observing the following fact: Let ww be a highest weight vector of WW of weight Ξ±\alpha, then

(32) Ξ¦V,W​(w)=Ψα+1βˆ’r​(χ​(V))​w.\Phi_{V,W}(w)=\Psi_{\alpha+1-r}(\chi(V))w.

Indeed, the map Ξ¦V,W\Phi_{V,W} is given by the partial quantum trace of cV,W∘cW,Vc_{V,W}\circ c_{W,V}. A standard argument shows that on a highest weight vector, this partial trace only depends of the Cartan part qHβŠ—H/2q^{H\otimes H/2} of the RR-matrix. The identity then follows from a direct computation. A detailed presentation of an analogous computation is given in [21, Proposition 2.2]. Equation (32) implies that if WW is simple then Ξ¦V,W=Ψα+1βˆ’r​(χ​(V))​IdW\Phi_{V,W}=\Psi_{\alpha+1-r}(\chi(V))\operatorname{Id}_{W}.

Equations (28) and (29) follow from Equation (32). For example, Ξ¦Pi,VΞ±=λ​IdVΞ±\Phi_{P_{i},V_{\alpha}}=\lambda\operatorname{Id}_{V_{\alpha}} where

Ξ»=Ψα​(χ​(Pi))=Ψα​([r]X​(Xrβˆ’iβˆ’1+Xβˆ’r+i+1))=(βˆ’1)rβˆ’1​r𝖽​(VΞ±)​(q(rβˆ’1βˆ’i)​α+qβˆ’(rβˆ’1βˆ’i)​α).\lambda=\Psi_{\alpha}(\chi(P_{i}))=\Psi_{\alpha}{\left([r]_{X}(X^{r-i-1}+X^{-r+i+1})\right)}=\frac{(-1)^{r-1}r}{{\mathsf{d}}(V_{\alpha})}(q^{(r-1-i)\alpha}+q^{-(r-1-i)\alpha}).

Similarly, to compute Ξ¦S1,Pj\Phi_{S_{1},P_{j}} observe that any endomorphism of PjP_{j} is of the form a​Id+b​xj∈End⁑(Pj)=ℂ​[xj]/(xj2)a\operatorname{Id}+bx_{j}\in\operatorname{End}(P_{j})=\mathbb{C}[x_{j}]/(x_{j}^{2}). Computing as above,

Ξ¦S1,Pj​(𝗐jS)=a​𝗐jS\Phi_{S_{1},P_{j}}(\mathsf{w}^{S}_{j})=a\mathsf{w}^{S}_{j}

where 𝗐jS\mathsf{w}^{S}_{j} is the highest weight vector of PjP_{j} and a=Ξ¨j+1βˆ’r​(X+Xβˆ’1)=βˆ’(qj+1+qβˆ’jβˆ’1)a=\Psi_{j+1-r}(X+X^{-1})=-(q^{j+1}+q^{-j-1}). We now compute bb. Recall that S1S_{1} is generated by two weight vectors π—Œ0,π—Œ1{\mathsf{s}}_{0},{\mathsf{s}}_{1} and E.π—Œ1=π—Œ0E.{\mathsf{s}}_{1}={\mathsf{s}}_{0}, F.π—Œ0=π—Œ1F.{\mathsf{s}}_{0}={\mathsf{s}}_{1}, H.π—Œi=(βˆ’1)iβ€‹π—ŒiH.{\mathsf{s}}_{i}=(-1)^{i}{\mathsf{s}}_{i}. In general,

cW,V=Ο„βˆ˜R=Ο„βˆ˜qHβŠ—H/2​(IdβŠ—Id+(qβˆ’qβˆ’1)​EβŠ—F+β‹―).c_{W,V}=\tau\circ R=\tau\circ q^{H\otimes H/2}(\operatorname{Id}\otimes\operatorname{Id}+(q-q^{-1})E\otimes F+\cdots).

So

cS1,Pj∘cPj,S1​(𝗐jHβŠ—π—Œ0)=cS1,Pj​(qj2β€‹π—Œ0βŠ—π—jH+qβˆ’j+22​(qβˆ’qβˆ’1)β€‹π—Œ1βŠ—π—j+2R)c_{S_{1},P_{j}}\circ c_{P_{j},S_{1}}(\mathsf{w}^{H}_{j}\otimes{\mathsf{s}}_{0})=c_{S_{1},P_{j}}(q^{\frac{j}{2}}{\mathsf{s}}_{0}\otimes\mathsf{w}^{H}_{j}+q^{-\frac{j+2}{2}}(q-q^{-1}){\mathsf{s}}_{1}\otimes\mathsf{w}^{R}_{j+2})
=(qj​𝗐jH+(qβˆ’qβˆ’1)2​qβˆ’1​𝗐jS)βŠ—π—Œ0+β‹―βŠ—π—Œ1={\left(q^{j}\mathsf{w}^{H}_{j}+(q-q^{-1})^{2}q^{-1}\mathsf{w}^{S}_{j}\right)}\otimes{\mathsf{s}}_{0}+\cdots\otimes{\mathsf{s}}_{1}
cS1,Pj∘cPj,S1​(𝗐jHβŠ—π—Œ1)=cS1,Pj​(qβˆ’j2β€‹π—Œ1βŠ—π—jH)=qβˆ’j​𝗐jHβŠ—π—Œ1+β‹―βŠ—π—Œ0.c_{S_{1},P_{j}}\circ c_{P_{j},S_{1}}(\mathsf{w}^{H}_{j}\otimes{\mathsf{s}}_{1})=c_{S_{1},P_{j}}(q^{-\frac{j}{2}}{\mathsf{s}}_{1}\otimes\mathsf{w}^{H}_{j})=q^{-j}\mathsf{w}^{H}_{j}\otimes{\mathsf{s}}_{1}+\cdots\otimes{\mathsf{s}}_{0}.

When taking the quantum trace with respect to S1S_{1} (i.e. the trace on S1S_{1} of the endomorphism composed with IdβŠ—K1βˆ’r\operatorname{Id}\otimes K^{1-r}, we then get that

Ξ¦S1,Pj​(𝗐jH)=βˆ’q​(qj​𝗐jH+(qβˆ’qβˆ’1)2​qβˆ’1​𝗐jS)βˆ’qβˆ’1.qβˆ’j​𝗐jH=βˆ’(qj+1+qβˆ’jβˆ’1)​𝗐jHβˆ’(qβˆ’qβˆ’1)2​𝗐jS\Phi_{S_{1},P_{j}}(\mathsf{w}^{H}_{j})=-q{\left(q^{j}\mathsf{w}^{H}_{j}+(q-q^{-1})^{2}q^{-1}\mathsf{w}^{S}_{j}\right)}-q^{-1}.q^{-j}\mathsf{w}^{H}_{j}=-(q^{j+1}+q^{-j-1})\mathsf{w}^{H}_{j}-(q-q^{-1})^{2}\mathsf{w}^{S}_{j}

and we get a=βˆ’(qj+1+qβˆ’jβˆ’1)a=-(q^{j+1}+q^{-j-1}), b=βˆ’(qβˆ’qβˆ’1)2b=-(q-q^{-1})^{2}. We have

Ξ¦S1,Pj∘ΦSi,Pj=Ξ¦S1βŠ—Si,Pj=Ξ¦Si+1,Pj+Ξ¦Siβˆ’1,Pj\Phi_{S_{1},P_{j}}\circ\Phi_{S_{i},P_{j}}=\Phi_{S_{1}\otimes S_{i},P_{j}}=\Phi_{S_{i+1},P_{j}}+\Phi_{S_{i-1},P_{j}}

so Ξ¦Si,Pj\Phi_{S_{i},P_{j}} is determined by the recurrence relations

Ξ¦Si+1,Pj=(a+b​xj)​ΦSi,Pjβˆ’Ξ¦Siβˆ’1,Pj,Ξ¦S1,Pj=a+b​xjandΞ¦S0,Pj=1.\Phi_{S_{i+1},P_{j}}=(a+bx_{j})\Phi_{S_{i},P_{j}}-\Phi_{S_{i-1},P_{j}},\,\quad\Phi_{S_{1},P_{j}}=a+bx_{j}\quad\text{and}\quad\Phi_{S_{0},P_{j}}=1.

Solving for Ξ¦Si,Pj\Phi_{S_{i},P_{j}} we have the unique solution

Ξ¦Si,Pj=(βˆ’1)i{j+1}​({(i+1)​(j+1)}​IdPj+{1}2​xj{j+1}2​(i​{(i+2)​(j+1)}βˆ’(i+2)​{i​(j+1)})).\Phi_{S_{i},P_{j}}=\frac{(-1)^{i}}{{\left\{j+1\right\}}}{\left({\left\{(i+1)(j+1)\right\}}\operatorname{Id}_{P_{j}}+\frac{{\left\{1\right\}}^{2}x_{j}}{{\left\{j+1\right\}}^{2}}\,\big{(}{i{\left\{(i+2)(j+1)\right\}}-(i+2){\left\{i(j+1)\right\}}}\big{)}\right)}.

In particular, for i=rβˆ’1i=r-1, Si=V0S_{i}=V_{0} and we get

Ξ¦V0,Pj=(βˆ’1)r+j​2​r​{1}2{j+1}2​xj.\Phi_{V_{0},P_{j}}=(-1)^{r+j}\frac{2r{\left\{1\right\}}^{2}}{{\left\{j+1\right\}}^{2}}x_{j}.

Finally the character formulas give the isomorphism of projective modules:

V0βŠ—Srβˆ’iβˆ’1=V0βŠ—Srβˆ’iβˆ’3βŠ•Pi.V_{0}\otimes S_{r-i-1}=V_{0}\otimes S_{r-i-3}\oplus P_{i}.

Thus,

Ξ¦Pi,Pj\displaystyle\Phi_{P_{i},P_{j}} =(βˆ’1)i​2​r​{1}2{j+1}2​(q(i+1)​(j+1)+qβˆ’(i+1)​(j+1))​xj.\displaystyle=\frac{(-1)^{i}2r{\left\{1\right\}}^{2}}{{\left\{j+1\right\}}^{2}}{\left(q^{(i+1)(j+1)}+q^{-(i+1)(j+1)}\right)}x_{j}.

∎

Lemma 6.7.

If PP is a projective module then 𝗍P​(Ξ¦V0,P)=𝗍V0​(Ξ¦P,V0)=(βˆ’1)rβˆ’1β€‹βŸ¨Ξ¦P,V0⟩\mathsf{t}_{P}(\Phi_{V_{0},P})=\mathsf{t}_{V_{0}}(\Phi_{P,V_{0}})=(-1)^{r-1}{\left\langle{\Phi_{P,V_{0}}}\right\rangle}.

Proof.

From the properties of a trace in Definition 3.1 we have

𝗍P⁑(Ξ¦V0,P)=𝗍P⁑(ptrR⁑(cP,V0​cV0,P))=𝗍PβŠ—V0⁑(cV0,P​cP,V0)=𝗍V0βŠ—P⁑(cP,V0​cV0,P)\operatorname{\mathsf{t}}_{P}(\Phi_{V_{0},P})=\operatorname{\mathsf{t}}_{P}(\operatorname{ptr}_{R}(c_{P,V_{0}}c_{V_{0},P}))=\operatorname{\mathsf{t}}_{P\otimes V_{0}}(c_{V_{0},P}c_{P,V_{0}})=\operatorname{\mathsf{t}}_{V_{0}\otimes P}(c_{P,V_{0}}c_{V_{0},P})
=𝗍V0⁑(ptrR⁑(cP,V0​cV0,P))=𝗍V0⁑(Ξ¦P,V0)=(βˆ’1)rβˆ’1β€‹βŸ¨Ξ¦P,V0⟩=\operatorname{\mathsf{t}}_{V_{0}}(\operatorname{ptr}_{R}(c_{P,V_{0}}c_{V_{0},P}))=\operatorname{\mathsf{t}}_{V_{0}}(\Phi_{P,V_{0}})=(-1)^{r-1}{\left\langle{\Phi_{P,V_{0}}}\right\rangle}

where the last equality follows from Theorem 5.4. ∎

Lemma 6.8 (The modified trace on typical modules).

Let VΞ±V_{\alpha} be a typical module. Then for any f∈Endπ’žβ‘(VΞ±)f\in\operatorname{End}_{\mathscr{C}}(V_{\alpha}), trVα⁑(f)=𝖽⁑(VΞ±)β€‹βŸ¨f⟩\operatorname{tr}_{V_{\alpha}}(f)=\operatorname{\mathsf{d}}(V_{\alpha}){\left\langle{f}\right\rangle} where 𝖽⁑(VΞ±)\operatorname{\mathsf{d}}({V_{\alpha}}) is given in Equation (18).

Proof.

First, since VΞ±V_{\alpha} is simple we have 𝗍Vα⁑(f)=⟨fβŸ©β€‹π—Vα⁑(IdVΞ±)=𝖽​(VΞ±)β€‹βŸ¨f⟩\operatorname{\mathsf{t}}_{V_{\alpha}}(f)={\left\langle{f}\right\rangle}\operatorname{\mathsf{t}}_{V_{\alpha}}(\operatorname{Id}_{V_{\alpha}})={\mathsf{d}}(V_{\alpha}){\left\langle{f}\right\rangle} where 𝖽​(VΞ±)=𝗍Vα⁑(IdVΞ±){\mathsf{d}}(V_{\alpha})=\operatorname{\mathsf{t}}_{V_{\alpha}}(\operatorname{Id}_{V_{\alpha}}). From Lemma 6.7 we have 𝗍Vα​(Ξ¦V0,VΞ±)=𝗍V0​(Ξ¦VΞ±,V0)\mathsf{t}_{V_{\alpha}}(\Phi_{V_{0},V_{\alpha}})=\mathsf{t}_{V_{0}}(\Phi_{V_{\alpha},V_{0}}).
So 𝖽⁑(VΞ±)β€‹βŸ¨Ξ¦V0,Vα⟩=𝖽⁑(V0)β€‹βŸ¨Ξ¦VΞ±,V0⟩\operatorname{\mathsf{d}}(V_{\alpha}){\left\langle{\Phi_{V_{0},V_{\alpha}}}\right\rangle}=\operatorname{\mathsf{d}}(V_{0}){\left\langle{\Phi_{V_{\alpha},V_{0}}}\right\rangle} where 𝖽⁑(V0)=(βˆ’1)rβˆ’1\operatorname{\mathsf{d}}(V_{0})=(-1)^{r-1}, and

𝖽⁑(VΞ±)=𝖽⁑(V0)β€‹βŸ¨Ξ¦VΞ±,V0⟩⟨ΦV0,Vα⟩.\operatorname{\mathsf{d}}(V_{\alpha})=\operatorname{\mathsf{d}}(V_{0})\frac{{\left\langle{\Phi_{V_{\alpha},V_{0}}}\right\rangle}}{{\left\langle{\Phi_{V_{0},V_{\alpha}}}\right\rangle}}.

Finally, the formula for 𝖽⁑(VΞ±)\operatorname{\mathsf{d}}(V_{\alpha}) follows from Lemma 6.6. ∎

Lemma 6.9 (The modified trace on PjP_{j}).

We have

𝖽​(Pj)=𝗍Pj​(IdPj)=(βˆ’1)j+1​(qj+1+qβˆ’jβˆ’1)and𝗍Pj​(xj)=(βˆ’1)j+1​{j+1}2{1}2.{\mathsf{d}}(P_{j})=\mathsf{t}_{P_{j}}(\operatorname{Id}_{P_{j}})=(-1)^{j+1}(q^{j+1}+q^{-j-1}){\quad\text{and}\quad}{\color[rgb]{0,0,0}\mathsf{t}_{P_{j}}(x_{j})=(-1)^{j+1}\frac{{\left\{j+1\right\}}^{2}}{{\left\{1\right\}}^{2}}}.
Proof.

As V0V_{0} is projective, so are V0βŠ—Srβˆ’jβˆ’1V_{0}\otimes S_{r-j-1} and V0βŠ—Srβˆ’jβˆ’3V_{0}\otimes S_{r-j-3}. Now by Proposition 6.2, indecomposable projective modules are determined by their highest weight so the isomorphism class of a projective module is determined by its character. Hence the character formulas imply that there exists an isomorphism of projective modules V0βŠ—Srβˆ’jβˆ’1≃V0βŠ—Srβˆ’jβˆ’3βŠ•PjV_{0}\otimes S_{r-j-1}\simeq V_{0}\otimes S_{r-j-3}\oplus P_{j}. Taking the modified traces of the identities of these modules gives 𝖽​(V0)​qdim⁑(Srβˆ’jβˆ’1)=𝖽​(V0)​qdim⁑(Srβˆ’jβˆ’3)+𝖽​(Pj){\mathsf{d}}(V_{0})\operatorname{qdim}(S_{r-j-1})={\mathsf{d}}(V_{0})\operatorname{qdim}(S_{r-j-3})+{\mathsf{d}}(P_{j}). Since qdim⁑(Si)=(βˆ’1)i​{i+1}{1}\operatorname{qdim}(S_{i})=(-1)^{i}\frac{\{i+1\}}{\{1\}} we have

𝖽​(Pj)\displaystyle{\mathsf{d}}(P_{j}) =𝖽​(V0)​(qdim⁑(Srβˆ’jβˆ’1)βˆ’qdim⁑(Srβˆ’jβˆ’3))=(βˆ’1)j​([rβˆ’j]βˆ’[rβˆ’jβˆ’2])\displaystyle={\mathsf{d}}(V_{0}){\left(\operatorname{qdim}(S_{r-j-1})-\operatorname{qdim}(S_{r-j-3})\right)}=(-1)^{j}{\left({\left[r-j\right]}-{\left[r-j-2\right]}\right)}
=(βˆ’1)j+1​(qj+1+qβˆ’jβˆ’1).\displaystyle=(-1)^{j+1}{\left(q^{j+1}+q^{-j-1}\right)}.

Lemma 6.7 implies 𝗍Pj​(Ξ¦V0,Pj)=𝗍V0​(Ξ¦Pj,V0).\mathsf{t}_{P_{j}}(\Phi_{V_{0},P_{j}})=\mathsf{t}_{V_{0}}(\Phi_{P_{j},V_{0}}). Then Lemma 6.6, implies

(βˆ’1)r+j​2​r​𝗍Pj​(xj)​{1}2{j+1}2=(βˆ’1)rβˆ’1​r​2𝖽​(V0)​𝗍V0​(IdV0)(-1)^{r+j}2r\mathsf{t}_{P_{j}}(x_{j}){\color[rgb]{0,0,0}\frac{{\left\{1\right\}}^{2}}{{\left\{j+1\right\}}^{2}}}=(-1)^{r-1}r\frac{2}{{\mathsf{d}}(V_{0})}\mathsf{t}_{V_{0}}(\operatorname{Id}_{V_{0}})

which implies the second relation of the lemma. ∎

Lemma 6.10 (Twist on PjP_{j}).

The action of the twist on PjP_{j} is given by

ΞΈPj=(βˆ’1)j​qj2+2​j2​(1βˆ’(rβˆ’jβˆ’1)​{1}2{j+1}​xj).\theta_{P_{j}}=(-1)^{j}q^{\frac{j^{2}+2j}{2}}(1-(r-j-1)\frac{{\left\{1\right\}}^{2}}{{\left\{j+1\right\}}}x_{j}).

In particular, ΞΈPj\theta_{P_{j}} has infinite order.

Proof.

The twist commutes with the map Ο€j:Pjβ†’Sj\pi_{j}:P_{j}\to S_{j} and Ο€j​xj=0\pi_{j}x_{j}=0. Thus the twist on PjP_{j} is given by ΞΈPj=ΞΈSj​(1+λ​xj)\theta_{P_{j}}=\theta_{S_{j}}(1+\lambda x_{j}) where ΞΈSj=(βˆ’1)j​qj2+2​j2\theta_{S_{j}}=(-1)^{j}q^{\frac{j^{2}+2j}{2}} is the scalar action of the twist on SjS_{j}. Hence

𝗍​(ΞΈPj)=βˆ’qj2+2​j2​(qj+1+qβˆ’jβˆ’1+λ​{j+1}2{1}2).\mathsf{t}(\theta_{P_{j}})=-q^{\frac{j^{2}+2j}{2}}(q^{j+1}+q^{-j-1}+\lambda\frac{{\left\{j+1\right\}}^{2}}{{\left\{1\right\}}^{2}}).

Finally we use again the module V0βŠ—Srβˆ’jβˆ’1≃V0βŠ—Srβˆ’jβˆ’3βŠ•PjV_{0}\otimes S_{r-j-1}\simeq V_{0}\otimes S_{r-j-3}\oplus P_{j} to color the unknot with framing +1+1. Its double is the Hopf link with both components having framing +1+1 and this gives

ΞΈV0​θSrβˆ’jβˆ’1.𝗍​(Φ​(Srβˆ’jβˆ’1,V0))=ΞΈV0​θSrβˆ’jβˆ’3.𝗍​(Φ​(Srβˆ’jβˆ’3,V0))+𝗍​(ΞΈPj)\theta_{V_{0}}\theta_{S_{r-j-1}}.\mathsf{t}(\Phi(S_{r-j-1},V_{0}))=\theta_{V_{0}}\theta_{S_{r-j-3}}.\mathsf{t}(\Phi(S_{r-j-3},V_{0}))+\mathsf{t}(\theta_{P_{j}})

Hence

𝗍​(ΞΈPj)\displaystyle\mathsf{t}(\theta_{P_{j}}) =ΞΈV0​𝖽​(V0)​((rβˆ’j)​θSrβˆ’jβˆ’1βˆ’(rβˆ’jβˆ’2)​θSrβˆ’jβˆ’3)\displaystyle=\theta_{V_{0}}{\mathsf{d}}(V_{0}){\left((r-j)\theta_{S_{r-j-1}}-(r-j-2)\theta_{S_{r-j-3}}\right)}
=βˆ’qj2/2βˆ’1​((rβˆ’j)βˆ’(rβˆ’jβˆ’2)​qβˆ’2​jβˆ’2)\displaystyle=-q^{j^{2}/2-1}((r-j)-(r-j-2)q^{-2j-2})
=βˆ’qj2+2​j2​(βˆ’(rβˆ’jβˆ’2)​qj+1+(rβˆ’j)​qβˆ’jβˆ’1)\displaystyle=-q^{\frac{j^{2}+2j}{2}}(-(r-j-2)q^{j+1}+(r-j)q^{-j-1})

and this gives the announced formula for λ\lambda. ∎

7. The algebra of projective modules

In this section, we define and study two algebras encoding the maps between projective modules of π’ž0Β―\mathscr{C}_{\overline{0}} and π’ž1Β―\mathscr{C}_{\overline{1}} respectively. These are the algebras one would associate to curves in a 1+1+11+1+1-TQFT which would be an extension of the 2+12+1 TQFT given in [4].

7.1. Maps between indecomposable projective modules

We first describe the maps between indecomposable projective modules in the categories π’ž0Β―\mathscr{C}_{\overline{0}} and π’ž1Β―\mathscr{C}_{\overline{1}}:

Proposition 7.1.

Let i,β„“βˆˆ{0,…,rβˆ’2}i,\ell\in\{0,\dots,r-2\} and kβˆˆβ„€k\in\mathbb{Z}. Let P=β„‚k​rHβŠ—Pβ„“P=\mathbb{C}^{H}_{kr}\otimes P_{\ell} be an indecomposable module, then any non zero map Piβ†’PP_{i}\to P is equal to λ​Ii+μ​xi\lambda I_{i}+\mu x_{i}, λ​αi+\lambda\alpha_{i}^{+} or λ​αiβˆ’\lambda\alpha_{i}^{-} where Ξ»,ΞΌβˆˆβ„‚\lambda,\mu\in\mathbb{C} and the maps IiI_{i}, xix_{i}, Ξ±i+\alpha_{i}^{+} and Ξ±iβˆ’\alpha_{i}^{-} are uniquely determined by

Ii:Piβ†’Pi𝗐iH↦𝗐iHxi:Piβ†’Pi𝗐iH↦𝗐iS\begin{array}[t]{rcl}I_{i}:P_{i}&\to&P_{i}\\ \mathsf{w}^{H}_{i}&\mapsto&\mathsf{w}^{H}_{i}\end{array}\quad\begin{array}[t]{rcl}x_{i}:P_{i}&\to&P_{i}\\ \mathsf{w}^{H}_{i}&\mapsto&\mathsf{w}^{S}_{i}\end{array}
Ξ±i+:Piβ†’β„‚rHβŠ—Prβˆ’2βˆ’i𝗐iH↦1βŠ—π—iβˆ’rLΞ±iβˆ’:Piβ†’β„‚βˆ’rHβŠ—Prβˆ’2βˆ’i.𝗐iH↦[i]!βˆ’2​1βŠ—π—i+rR\begin{array}[t]{rcl}\alpha^{+}_{i}:P_{i}&\to&\mathbb{C}^{H}_{r}\otimes P_{{r-2-i}}\\ \mathsf{w}^{H}_{i}&\mapsto&1\otimes\mathsf{w}^{L}_{i-r}\end{array}\quad\begin{array}[t]{rcl}\alpha^{-}_{i}:P_{i}&\to&\mathbb{C}^{H}_{-r}\otimes P_{{r-2-i}}.\\ \mathsf{w}^{H}_{i}&\mapsto&[i]!^{-2}1\otimes\mathsf{w}^{R}_{i+r}\end{array}
Proof.

By Equality (27), the space Homπ’žβ‘(Pi,P)\operatorname{Hom}_{\mathscr{C}}(P_{i},P) is isomorphic to the space of dominant weight vectors of weight ii of PP. Now the space of dominant vectors of β„‚k​rHβŠ—Pβ„“\mathbb{C}^{H}_{kr}\otimes P_{\ell} has dimension 44 and is generated by 1βŠ—π—β„“H1\otimes\mathsf{w}^{H}_{\ell} and 1βŠ—π—β„“S1\otimes\mathsf{w}^{S}_{\ell} of weight β„“+k​r\ell+kr, 1βŠ—π—βˆ’β„“βˆ’2L1\otimes\mathsf{w}^{L}_{-\ell-2} of weight βˆ’β„“βˆ’2+k​r-\ell-2+kr and 1βŠ—π—2​rβˆ’β„“βˆ’2R1\otimes\mathsf{w}^{R}_{2r-\ell-2} of weight (k+2)​rβˆ’β„“βˆ’2{(k+2)r-\ell-2}. The result then follows by analyzing for which k,β„“k,\ell the module β„‚k​rHβŠ—Pβ„“\mathbb{C}^{H}_{kr}\otimes P_{\ell} has dominant weight vectors of weight ii. ∎

Tensoring by β„‚n​rH\mathbb{C}^{H}_{nr} gives canonical isomorphisms Homπ’žβ‘(Pi,β„‚k​rHβŠ—Pβ„“)β‰…Homπ’žβ‘(β„‚n​rHβŠ—Pi,β„‚(n+k)​rHβŠ—Pβ„“)\operatorname{Hom}_{\mathscr{C}}(P_{i},\mathbb{C}^{H}_{kr}\otimes P_{\ell})\cong\operatorname{Hom}_{\mathscr{C}}(\mathbb{C}^{H}_{nr}\otimes P_{i},\mathbb{C}^{H}_{(n+k)r}\otimes P_{\ell}). Then for i∈{0,…,rβˆ’2}i\in\{0,\dots,r-2\} and j=rβˆ’2βˆ’ij=r-2-i, maps between indecomposable projective modules Pik=β„‚k​rHβŠ—PiP_{i}^{k}=\mathbb{C}^{H}_{kr}\otimes P_{i}, Pjk=β„‚k​rHβŠ—PjP_{j}^{k}=\mathbb{C}^{H}_{kr}\otimes P_{j} can be represented by the following periodic quiver:

⋯​[Uncaptioned image]​Pi2PiPi-2Pj1Pj3Pj-1Pj-3​⋯\cdots\,\,\begin{array}[]{c}\hskip-3.69885pt\raisebox{-4.0pt}{\psfig{height=64.58313pt}}\hskip-5.406pt\end{array}\put(-66.0,17.0){\mbox{\tiny$P_{i}^{2}$}}\put(-120.0,17.0){\mbox{\tiny$P_{i}$}}\put(-181.0,17.0){\mbox{\tiny$P_{i}^{-2}$}}\put(-95.0,-15.0){\mbox{\tiny$P_{j}^{1}$}}\put(-40.0,-15.0){\mbox{\tiny$P_{j}^{3}$}}\put(-152.0,-15.0){\mbox{\tiny$P_{j}^{-1}$}}\put(-206.0,-15.0){\mbox{\tiny$P_{j}^{-3}$}}\,\,\cdots

7.2. The algebras of curves

As above, let rβ€²=rr^{\prime}=r if rr is odd and rβ€²=r2r^{\prime}=\frac{r}{2} else. Let Οƒ=β„‚2​rβ€²H\sigma=\mathbb{C}^{H}_{2r^{\prime}} be the one dimensional module where EE and FF act as 0 and HH acts as 2​rβ€²2r^{\prime}. The object Οƒβˆˆπ’ž0Β―\sigma\in\mathscr{C}_{\overline{0}} generates the group of invertible objects of π’ž0Β―\mathscr{C}_{\overline{0}} which is isomorphic to β„€\mathbb{Z}. For kβˆˆβ„€k\in\mathbb{Z}, we just denote by Οƒk\sigma^{k} the module β„‚2​k​rβ€²H\mathbb{C}^{H}_{2kr^{\prime}} so that ΟƒkβŠ—Οƒβ„“=Οƒk+β„“\sigma^{k}\otimes\sigma^{\ell}=\sigma^{k+\ell} and Οƒ0=𝕀\sigma^{0}=\mathbb{I}.

A Οƒ\sigma-invariant module is an infinite dimensional weight module VV with finite dimensional weight spaces and with the property that ΟƒβŠ—V=V\sigma\otimes V=V. Then tensor product by 1βˆˆΟƒ1\in\sigma gives an action of ℀≃{Οƒk:kβˆˆβ„€}\mathbb{Z}\simeq\{\sigma^{k}:k\in\mathbb{Z}\} on VV denoted by v↦σ​vv\mapsto\sigma v. Remark that since VV is infinite dimensional it is not an object of π’ž\mathscr{C}.

Let π’žΟƒ\mathscr{C}^{\sigma} be the category whose objects are Οƒ\sigma-invariant modules and maps from VV to WW are given by the set Homσ⁑(V,W)\operatorname{Hom}_{\sigma}(V,W) of morphism of UqH​𝔰​𝔩​(2){U_{q}^{H}{\mathfrak{sl}(2)}}-modules that commute with the action of Οƒ\sigma.

We study endomorphisms of the Οƒ\sigma-invariant module

β„™=⨁kβˆˆβ„€β¨i=0rβˆ’1β„‚k​rHβŠ—Pi.{\mathbb{P}}=\bigoplus_{k\in\mathbb{Z}}\bigoplus_{i=0}^{r-1}\mathbb{C}^{H}_{kr}\otimes P_{i}.

According to the parity of weights, this module splits into two Οƒ\sigma-invariant modules β„™=β„™0Β―βŠ•β„™1Β―{\mathbb{P}}={\mathbb{P}}_{\overline{0}}\oplus{\mathbb{P}}_{\overline{1}}. Also for ν∈{0Β―,1Β―}\nu\in\{\overline{0},\overline{1}\}, β„™Ξ½{\mathbb{P}}_{\nu} has a β„€\mathbb{Z}-grading for which β„‚k​rHβŠ—Pi\mathbb{C}^{H}_{kr}\otimes P_{i} is of degree kk. In particular, if rr is odd, the action of Οƒ=β„‚2​rH\sigma=\mathbb{C}^{H}_{2r} shifts the degree by 22 whereas for rr even, Οƒ=β„‚rH\sigma=\mathbb{C}^{H}_{r} shifts the degree by 11. In the standard way, the β„€\mathbb{Z}-grading of β„™Ξ½{\mathbb{P}}_{\nu} turns Endσ⁑(β„™Ξ½)\operatorname{End}_{\sigma}({\mathbb{P}}_{\nu}) into a β„€\mathbb{Z}-graded algebra. We call

𝔸0Β―=Endσ⁑(β„™0Β―)and𝔸1Β―=Endσ⁑(β„™1Β―){\mathbb{A}}_{\overline{0}}=\operatorname{End}_{\sigma}({\mathbb{P}}_{\overline{0}}){\quad\text{and}\quad}{\mathbb{A}}_{\overline{1}}=\operatorname{End}_{\sigma}({\mathbb{P}}_{\overline{1}})

these β„€\mathbb{Z}-graded algebras.

We now introduce two algebras AA and BB used to describe 𝔸ν=Endσ⁑(β„™Ξ½){\mathbb{A}}_{\nu}=\operatorname{End}_{\sigma}({\mathbb{P}}_{\nu}). Let AA be the algebra of graded dimension 2​sβˆ’1+4+2​s2s^{-1}+4+2s which is the quotient of the β„‚\mathbb{C}-path algebra associated to the quiver

Ξ“=[Uncaptioned image]pqa+a-b+b-Β by the relations{a+​b+=b+​a+=aβˆ’β€‹bβˆ’=bβˆ’β€‹aβˆ’=0,a+​bβˆ’+aβˆ’β€‹b+=0,b+​aβˆ’+bβˆ’β€‹a+=0.\Gamma=\quad\begin{array}[]{c}\hskip-3.69885pt\raisebox{-4.0pt}{\psfig{height=55.97205pt}}\hskip-5.406pt\end{array}\put(-58.0,2.0){\mbox{\tiny$p$}}\put(-7.0,2.0){\mbox{\tiny$q$}}\put(-33.0,33.0){\mbox{\tiny$a_{+}$}}\put(-33.0,-30.0){\mbox{\tiny$a_{-}$}}\put(-33.0,17.0){\mbox{\tiny$b_{+}$}}\put(-33.0,-16.0){\mbox{\tiny$b_{-}$}}\quad\text{ by the relations}\left\{\begin{array}[]{l}a_{+}b_{+}=b_{+}a_{+}=a_{-}b_{-}=b_{-}a_{-}=0\,,\\ a_{+}b_{-}+a_{-}b_{+}=0\,,\\ b_{+}a_{-}+b_{-}a_{+}=0\,.\end{array}\right.

As a β„€\mathbb{Z}-graded β„‚\mathbb{C}-vector space, AA is spanned in degree 11 by {a+,b+}\{a_{+},b_{+}\}, in degree βˆ’1-1 by {aβˆ’,bβˆ’}\{a_{-},b_{-}\} and in degree 0 by {p,q=1βˆ’p,x=b+​aβˆ’,y=a+​bβˆ’}\{p,q=1-p,x=b_{+}a_{-},y=a_{+}b_{-}\}.

The algebra BB is the quotient of AA obtained by identifying p=qp=q, a+=b+a_{+}=b_{+} and aβˆ’=bβˆ’a_{-}=b_{-} (BB is the exterior algebra of β„‚2\mathbb{C}^{2}). It is also the quotient of the β„‚\mathbb{C}-path algebra associated to the quiver

Ξ“β€²=[Uncaptioned image]pa+a-Β by the relations{a+​a+=aβˆ’β€‹aβˆ’=0,a+​aβˆ’+aβˆ’β€‹a+=0.\Gamma^{\prime}=\quad\begin{array}[]{c}\hskip-3.69885pt\raisebox{-4.0pt}{\psfig{height=55.97205pt}}\hskip-5.406pt\end{array}\put(-10.0,1.0){\mbox{\tiny$p$}}\put(-13.0,22.0){\mbox{\tiny$a_{+}$}}\put(-13.0,-19.0){\mbox{\tiny$a_{-}$}}\quad\text{ by the relations}\left\{\begin{array}[]{l}a_{+}a_{+}=a_{-}a_{-}=0\,,\\ a_{+}a_{-}+a_{-}a_{+}=0\,.\end{array}\right.

The basis of BB is given by {aβˆ’,p=1,x=a+​aβˆ’,a+}\{a_{-},p=1,x=a_{+}a_{-},a_{+}\}.

Theorem 7.2.

There exist isomorphisms of algebras:

(33) If r∈3+2​ℕr\in 3+2\mathbb{N}, 𝔸0Β―=Endσ⁑(β„™0Β―)≃Arβˆ’12×ℂ≃Endσ⁑(β„™1Β―)=𝔸1Β―.\displaystyle{\mathbb{A}}_{\overline{0}}=\operatorname{End}_{\sigma}({\mathbb{P}}_{\overline{0}})\simeq A^{\frac{r-1}{2}}\times\mathbb{C}\simeq\operatorname{End}_{\sigma}({\mathbb{P}}_{\overline{1}})={\mathbb{A}}_{\overline{1}}.
(34) If r∈2+4​ℕr\in 2+4\mathbb{N}, 𝔸0Β―=Endσ⁑(β„™0Β―)≃Arβˆ’24Γ—Band𝔸1Β―=Endσ⁑(β„™1Β―)≃Arβˆ’24Γ—β„‚.\displaystyle{\mathbb{A}}_{\overline{0}}=\operatorname{End}_{\sigma}({\mathbb{P}}_{\overline{0}})\simeq A^{\frac{r-2}{4}}\times B{\quad\text{and}\quad}{\mathbb{A}}_{\overline{1}}=\operatorname{End}_{\sigma}({\mathbb{P}}_{\overline{1}})\simeq A^{\frac{r-2}{4}}\times\mathbb{C}.
(35) If r∈4+4​ℕr\in 4+4\mathbb{N}, 𝔸0Β―=Endσ⁑(β„™0Β―)≃Ar4and𝔸1Β―=Endσ⁑(β„™1Β―)≃Arβˆ’44Γ—BΓ—β„‚.\displaystyle{\mathbb{A}}_{\overline{0}}=\operatorname{End}_{\sigma}({\mathbb{P}}_{\overline{0}})\simeq A^{\frac{r}{4}}{\quad\text{and}\quad}{\mathbb{A}}_{\overline{1}}=\operatorname{End}_{\sigma}({\mathbb{P}}_{\overline{1}})\simeq A^{\frac{r-4}{4}}\times B\times\mathbb{C}.
Proof.

To prove this theorem we build the explicit isomorphisms. If a Οƒ\sigma-invariant module WW splits as W=⨁kβˆˆβ„€ΟƒkβŠ—VW=\bigoplus_{k\in\mathbb{Z}}\sigma^{k}\otimes V for some finite dimensional weight module VV then the action of Οƒ\sigma on WW is free and the restriction map Homσ⁑(W,Wβ€²)β†’HomUqH​𝔰​𝔩​(2)⁑(V,Wβ€²)\operatorname{Hom}_{\sigma}(W,W^{\prime})\to\operatorname{Hom}_{{U_{q}^{H}{\mathfrak{sl}(2)}}}(V,W^{\prime}) is easily seen to be an isomorphism (here HomUqH​𝔰​𝔩​(2)\operatorname{Hom}_{{U_{q}^{H}{\mathfrak{sl}(2)}}} denotes morphisms of UqH​𝔰​𝔩​(2){{U_{q}^{H}{\mathfrak{sl}(2)}}}-modules). Using this fact, we restrict our study to the maps from PiP_{i} to β„™{\mathbb{P}}. For i≀rβˆ’2i\leq r-2, let j=rβˆ’2βˆ’ij=r-2-i. By Proposition 7.1, the space HomUqH​𝔰​𝔩​(2)⁑(Pi,β„™)\operatorname{Hom}_{{U_{q}^{H}{\mathfrak{sl}(2)}}}(P_{i},{\mathbb{P}}) is of dimension 4 generated by the morphisms determined uniquely by

Ii:Piβ†’PiβŠ‚β„™xi:Piβ†’PiβŠ‚β„™\begin{array}[t]{rcl}I_{i}:P_{i}&\to&P_{i}\subset{\mathbb{P}}\end{array}\quad\begin{array}[t]{rcl}x_{i}:P_{i}&\to&P_{i}\subset{\mathbb{P}}\end{array}
Ξ±i+:Piβ†’β„‚rHβŠ—PjβŠ‚β„™Ξ±iβˆ’:Piβ†’β„‚βˆ’rHβŠ—PjβŠ‚β„™\begin{array}[t]{rcl}\alpha^{+}_{i}:P_{i}&\to&\mathbb{C}^{H}_{r}\otimes P_{j}\subset{\mathbb{P}}\end{array}\quad\begin{array}[t]{rcl}\alpha^{-}_{i}:P_{i}&\to&\mathbb{C}^{H}_{-r}\otimes P_{j}\subset{\mathbb{P}}\end{array}

These maps extend to maps of Endσ⁑(β„™)\operatorname{End}_{\sigma}({\mathbb{P}}) on factors β„‚k​rHβŠ—Pi\mathbb{C}^{H}_{kr}\otimes P_{i} by tensoring them by the identity of β„‚k​rH\mathbb{C}^{H}_{kr} and we extend them by 0 on the other factors. We use the same name for these extended maps of Endσ⁑(β„™)\operatorname{End}_{\sigma}({\mathbb{P}}). The composition of these maps is computed by looking at the image of the dominant vector 𝗐iH∈Pi\mathsf{w}^{H}_{i}\in P_{i}. One easily gets

Ξ±j+∘αi+=0=Ξ±jβˆ’βˆ˜Ξ±iβˆ’.\alpha^{+}_{j}\circ\alpha^{+}_{i}=0=\alpha^{-}_{j}\circ\alpha^{-}_{i}.

Now we use that if v∈Vv\in V is a weight vector then in β„‚Β±rHβŠ—V\mathbb{C}^{H}_{\pm r}\otimes V one has E.(1βŠ—v)=1βŠ—(E.v)E.(1\otimes v)=1\otimes(E.v) and F.(1βŠ—v)=1βŠ—(βˆ’F.v)F.(1\otimes v)=1\otimes(-F.v) to compute:

Pi⟢αi+β„‚rHβŠ—Pj⟢αjβˆ’Pi𝗐iH⟼1βŠ—π—iβˆ’rL=(βˆ’F)j+1​(1βŠ—π—jH)⟼[j]!βˆ’2​(βˆ’F)j+1​𝗐j+rR\begin{array}[]{rcccl}P_{i}&\stackrel{{\scriptstyle\alpha_{i}^{+}}}{{\longrightarrow}}&\mathbb{C}^{H}_{r}\otimes P_{j}&\stackrel{{\scriptstyle\alpha_{j}^{-}}}{{\longrightarrow}}&P_{i}\\ \mathsf{w}^{H}_{i}&\longmapsto&1\otimes\mathsf{w}^{L}_{i-r}=(-F)^{j+1}(1\otimes\mathsf{w}^{H}_{j})&\longmapsto&[j]!^{-2}(-F)^{j+1}\mathsf{w}^{R}_{j+r}\end{array}

As Fj+1​𝗐j+rR=∏k=1j(βˆ’Ξ³j,jβˆ’k+1)​𝗐iS=(βˆ’1)j​[j]!2​𝗐iSF^{j+1}\mathsf{w}^{R}_{j+r}=\prod_{k=1}^{j}(-\gamma_{j,j-k+1})\mathsf{w}^{S}_{i}=(-1)^{j}[j]!^{2}\mathsf{w}^{S}_{i}, we get

Ξ±jβˆ’βˆ˜Ξ±i+=βˆ’xi.\alpha_{j}^{-}\circ\alpha_{i}^{+}=-x_{i}.

Similarly,

Pi⟢αiβˆ’β„‚βˆ’rHβŠ—Pj⟢αj+Pi𝗐iH⟼[i]!βˆ’2​1βŠ—π—i+rR=[i]!βˆ’2​Ei+1​(1βŠ—π—jH)⟼[i]!βˆ’2​Ei+1​𝗐jβˆ’rL\begin{array}[]{rcccl}P_{i}&\stackrel{{\scriptstyle\alpha_{i}^{-}}}{{\longrightarrow}}&\mathbb{C}^{H}_{-r}\otimes P_{j}&\stackrel{{\scriptstyle\alpha_{j}^{+}}}{{\longrightarrow}}&P_{i}\\ \mathsf{w}^{H}_{i}&\longmapsto&[i]!^{-2}1\otimes\mathsf{w}^{R}_{i+r}=[i]!^{-2}E^{i+1}(1\otimes\mathsf{w}^{H}_{j})&\longmapsto&[i]!^{-2}E^{i+1}\mathsf{w}^{L}_{j-r}\end{array}

And as Ei+1​𝗐jβˆ’rL=∏k=1i(Ξ³i,k)​𝗐iS=[i]!2​𝗐iSE^{i+1}\mathsf{w}^{L}_{j-r}=\prod_{k=1}^{i}(\gamma_{i,k})\mathsf{w}^{S}_{i}=[i]!^{2}\mathsf{w}^{S}_{i}, we get

Ξ±j+∘αiβˆ’=xi.\alpha_{j}^{+}\circ\alpha_{i}^{-}=x_{i}.

We now explicit the isomorphism of Theorem 7.2. First remark that the maps of Endσ⁑(β„™)\operatorname{End}_{\sigma}({\mathbb{P}}) commute with KrK^{r} thus they restrict to maps of Endσ⁑(β„™Ξ½)\operatorname{End}_{\sigma}({\mathbb{P}}_{\nu}) for ν∈{0Β―,1Β―}\nu\in\{\overline{0},\overline{1}\}. Next the decomposition of endomorphism algebras in Theorem 7.2 follows from the fact that these endomorphisms respect the characteristic spaces of the Casimir element CC whose minimal polynomial is given in Proposition 4.1. For i∈{0​⋯​rβˆ’1}i\in\{0\cdots r-1\}, let ci=qi+1+qβˆ’iβˆ’1{1}2=βˆ’qj+1+qβˆ’jβˆ’1{1}2c_{i}=\frac{q^{i+1}+q^{-i-1}}{{\left\{1\right\}}^{2}}=-\frac{q^{j+1}+q^{-j-1}}{{\left\{1\right\}}^{2}} be the scalar by which CC acts on the simple module SiS_{i}. The action of CC on ΟƒβŠ—Si\sigma\otimes S_{i} and on SiS_{i} are the same if rr is odd, but they are opposite if rr is even.

Let ν∈{0,1}\nu\in\{0,1\}. For i∈2​ℕ+Ξ½i\in 2\mathbb{N}+\nu, i≀rβ€²βˆ’2i\leq r^{\prime}-2, the kernel of (C2βˆ’ci2)2(C^{2}-c_{i}^{2})^{2} on β„™Ξ½{\mathbb{P}}_{\nu} is V=⨁kβˆˆβ„€ΟƒkβŠ—(PiβŠ•Qj)V=\bigoplus_{k\in\mathbb{Z}}\sigma^{k}\otimes(P_{i}\oplus Q_{j}) where QjQ_{j} is PjP_{j} if rr even and Qj=β„‚rHβŠ—PjQ_{j}=\mathbb{C}^{H}_{r}\otimes P_{j} if rr is odd. Then an isomorphism Aβ†’βˆΌEndσ⁑(V)A\stackrel{{\scriptstyle\sim}}{{\to}}\operatorname{End}_{\sigma}(V) is given by

p\displaystyle p ↦Ii\displaystyle\mapsto I_{i} x\displaystyle x ↦xi\displaystyle\mapsto x_{i} a+\displaystyle a^{+} ↦αi+\displaystyle\mapsto\alpha^{+}_{i} aβˆ’\displaystyle a^{-} ↦αiβˆ’\displaystyle\mapsto\alpha^{-}_{i}
q\displaystyle q ↦Ij\displaystyle\mapsto I_{j} y\displaystyle y ↦xj\displaystyle\mapsto x_{j} b+\displaystyle b^{+} ↦αj+\displaystyle\mapsto\alpha^{+}_{j} bβˆ’\displaystyle b^{-} ↦αjβˆ’\displaystyle\mapsto\alpha^{-}_{j}

Now if rr is even, let i=rβˆ’22=rβˆ’2βˆ’ii=\frac{r-2}{2}=r-2-i and Ξ½=i\nu=i mod 22. Then ci=0c_{i}=0 and the kernel of C2C^{2} on β„™Ξ½{\mathbb{P}}_{\nu} is then V=⨁kβˆˆβ„€ΟƒkβŠ—PiV=\bigoplus_{k\in\mathbb{Z}}\sigma^{k}\otimes P_{i}. Then an isomorphism Bβ†’βˆΌEndσ⁑(V)B\stackrel{{\scriptstyle\sim}}{{\to}}\operatorname{End}_{\sigma}(V) is given by

p\displaystyle p ↦Ii\displaystyle\mapsto I_{i} x\displaystyle x ↦xi\displaystyle\mapsto x_{i} a+\displaystyle a^{+} ↦αi+\displaystyle\mapsto\alpha^{+}_{i} aβˆ’\displaystyle a^{-} ↦αiβˆ’\displaystyle\mapsto\alpha^{-}_{i}

Finally the remaining β„‚\mathbb{C} factors in Theorem 7.2 correspond to the eigenspace of CC associated to the simple eigenvalue crβˆ’1c_{r-1}. ∎

In the paper [3], the concepts of Coend, trace and the Hochschild-Mitchell homology in a linear category are related. In [4], a graded TQFT is defined for manifolds equipped with a 1-cohomology class with value in β„‚/2​℀\mathbb{C}/2\mathbb{Z}. The algebras 𝔸ν{\mathbb{A}}_{\nu} would naturally be associated to a curve Ξ³\gamma with cohomology class Ο‰\omega such that ω​([Ξ³])=Ξ½\omega([\gamma])=\nu. Then the graded vector space Tr⁑(𝔸ν)=(𝔸ν)/f​g⁣=g​f\operatorname{Tr}({\mathbb{A}}_{\nu})=({\mathbb{A}}_{\nu})_{/fg=gf} maps surjectively onto the TQFT space of the torus Ξ³Γ—S1\gamma\times S^{1} with cohomology class Ο‰\omega such that Ο‰([Ξ³Γ—βˆ—])=Ξ½\omega([\gamma\times*])=\nu and Ο‰([βˆ—Γ—S1])=0\omega([*\times S^{1}])=0. Here we define a graded version of the trace of 𝔸ν{\mathbb{A}}_{\nu} that surjects on the TQFT space of the torus Ξ³Γ—S1\gamma\times S^{1} with cohomology class Ο‰\omega such that Ο‰([Ξ³Γ—βˆ—])=Ξ½\omega([\gamma\times*])=\nu and Ο‰([βˆ—Γ—S1])=Ξ²\omega([*\times S^{1}])=\beta for any Ξ²βˆˆβ„‚/2​℀\beta\in\mathbb{C}/2\mathbb{Z} (instead of Ξ²\beta, we use z=q2​r′​βz=q^{2r^{\prime}\beta}).

Let zβˆˆβ„‚βˆ—z\in\mathbb{C}^{*}, π’œ\mathcal{A} be a β„€\mathbb{Z}-graded β„‚\mathbb{C}-algebra. If f,gf,g are homogenous elements of degree |f|,|g|βˆˆβ„€|f|,|g|\in\mathbb{Z}, let [f,g]z=f​gβˆ’z|f|​g​f\left[f,g\right]_{z}=fg-z^{|f|}gf. Define the β„€\mathbb{Z}-graded module

Trz⁑(π’œ)=π’œ/[π’œ,π’œ]z.\operatorname{Tr}^{z}(\mathcal{A})=\mathcal{A}_{/[\mathcal{A},\mathcal{A}]_{z}}.

Similarly, if π’œ\mathcal{A} is considered as a super algebra, the bracket is replaced by [f,g]zβˆ’=f​gβˆ’(βˆ’1)|f|​|g|​z|f|​g​f\left[f,g\right]_{z}^{-}=fg-(-1)^{|f||g|}z^{|f|}gf and we define the β„€\mathbb{Z}-graded super module

STrz⁑(π’œ)=π’œ/[π’œ,π’œ]zβˆ’.\operatorname{STr}^{z}(\mathcal{A})=\mathcal{A}_{/[\mathcal{A},\mathcal{A}]_{z}^{-}}.
Proposition 7.3.

Recall the algebras AA and BB above. Then

  1. (1)

    If zβˆˆβ„‚βˆ—βˆ–{Β±1}z\in\mathbb{C}^{*}\setminus\{\pm 1\}, Trz⁑(A)≃ℂ2≃STrz⁑(A)\operatorname{Tr}^{z}(A)\simeq\mathbb{C}^{2}\simeq\operatorname{STr}^{z}(A).

  2. (2)

    TrΒ±1⁑(A)≃ℂ3≃STrΒ±1⁑(A)\operatorname{Tr}^{\pm 1}(A)\simeq\mathbb{C}^{3}\simeq\operatorname{STr}^{\pm 1}(A).

  3. (3)

    If zβ‰ 1z\neq 1 then STrz⁑(B)≃ℂ\operatorname{STr}^{z}(B)\simeq\mathbb{C} and STr1⁑(B)≃B\operatorname{STr}^{1}(B)\simeq B.

Here the spaces β„‚2\mathbb{C}^{2} and β„‚3\mathbb{C}^{3} are concentrated in degree 0. As a consequence, we have the following graded dimensions:

  1. (1)

    If r∈2​℀+1r\in 2\mathbb{Z}+1 and zβ‰ Β±1z\neq\pm 1, then

    dims(Trz⁑(𝔸0Β―))=dims(Trz⁑(𝔸1Β―))=randdims(TrΒ±1⁑(𝔸0Β―))=dims(TrΒ±1⁑(𝔸1Β―))=3​rβˆ’12.\dim_{s}(\operatorname{Tr}^{z}({\mathbb{A}}_{\overline{0}}))=\dim_{s}(\operatorname{Tr}^{z}({\mathbb{A}}_{\overline{1}}))=r{\quad\text{and}\quad}\dim_{s}(\operatorname{Tr}^{\pm 1}({\mathbb{A}}_{\overline{0}}))=\dim_{s}(\operatorname{Tr}^{\pm 1}({\mathbb{A}}_{\overline{1}}))=\frac{3r-1}{2}.
  2. (2)

    If r∈4​℀+2r\in 4\mathbb{Z}+2 and zβ‰ Β±1z\neq\pm 1, then

    dims(STrz⁑(𝔸0Β―))=dims(STrz⁑(𝔸1Β―))=r2,\dim_{s}(\operatorname{STr}^{z}({\mathbb{A}}_{\overline{0}}))=\dim_{s}(\operatorname{STr}^{z}({\mathbb{A}}_{\overline{1}}))=\frac{r}{2},
    dims(STrΒ±1⁑(𝔸1Β―))=dims(STrβˆ’1⁑(𝔸0Β―))=3​rβˆ’24and\dim_{s}(\operatorname{STr}^{\pm 1}({\mathbb{A}}_{\overline{1}}))=\dim_{s}(\operatorname{STr}^{-1}({\mathbb{A}}_{\overline{0}}))=\frac{3r-2}{4}{\quad\text{and}\quad}
    dims(STr1⁑(𝔸0Β―))=sβˆ’1+3​r+24+s\dim_{s}(\operatorname{STr}^{1}({\mathbb{A}}_{\overline{0}}))=s^{-1}+\frac{3r+2}{4}+s

where dims\dim_{s} is the sum for kβˆˆβ„€k\in\mathbb{Z} of sks^{k} times the dimension of the degree kk subspace.

Proof.

Let Ξ΅=Β±1\varepsilon=\pm 1. First remark that for any elements f,gf,g of the algebra, [f,g]zΞ΅+Ξ΅|f|.|g|​z|f|​[g,f]zΞ΅=(1βˆ’z|f|+|g|)​f​g\left[f,g\right]_{z}^{\varepsilon}+\varepsilon^{|f|.|g|}z^{|f|}\left[g,f\right]_{z}^{\varepsilon}=(1-z^{|f|+|g|})fg. Hence if z|f|+|g|β‰ 1z^{|f|+|g|}\neq 1 then f​g=0fg=0 in the quotient, else [g,f]zΞ΅\left[g,f\right]_{z}^{\varepsilon} and [f,g]zΞ΅\left[f,g\right]_{z}^{\varepsilon} are proportional. Finally for g=1g=1, one gets that a map ff vanishes in the quotient unless z|f|=1z^{|f|}=1. Then the relations in AA implies that [A,A]zΞ΅\left[A,A\right]^{\varepsilon}_{z} is generated by the following elements

  • β€’

    [aΒ±,p]zΞ΅=aΒ±\left[a_{\pm},p\right]_{z}^{\varepsilon}=a_{\pm},

  • β€’

    [bΒ±,q]zΞ΅=bΒ±\left[b_{\pm},q\right]_{z}^{\varepsilon}=b_{\pm},

  • β€’

    [b+,aβˆ’]zΞ΅=b+​aβˆ’βˆ’Ξ΅β€‹z​aβˆ’β€‹b+=x+Ρ​z​y\left[b_{+},a_{-}\right]_{z}^{\varepsilon}=b_{+}a_{-}-\varepsilon za_{-}b_{+}=x+\varepsilon zy

  • β€’

    [a+,bβˆ’]zΞ΅=a+​bβˆ’βˆ’Ξ΅β€‹z​bβˆ’β€‹a+=y+Ρ​z​x\left[a_{+},b_{-}\right]_{z}^{\varepsilon}=a_{+}b_{-}-\varepsilon zb_{-}a_{+}=y+\varepsilon zx

If z2β‰ 1z^{2}\neq 1 then x=y=0x=y=0 in the quotient, and if z=Β±1z=\pm 1, then [A,A]zΞ΅\left[A,A\right]^{\varepsilon}_{z} is generated in degree 0 by x+Ρ​z​yx+\varepsilon zy.

Similarly for zβ‰ 1z\neq 1, [B,B]zβˆ’\left[B,B\right]_{z}^{-} is generated by a+a_{+}, aβˆ’a_{-} and the element

[a+,aβˆ’]zβˆ’=a+​aβˆ’+z​aβˆ’β€‹a+=(1βˆ’z)​x.\left[a_{+},a_{-}\right]_{z}^{-}=a_{+}a_{-}+za_{-}a_{+}=(1-z)x.

On the other hand, for z=1z=1 we have [B,B]1βˆ’=0\left[B,B\right]^{-}_{1}=0.

For the last statements, we use Trz⁑(π’œΓ—π’œβ€²)=Trz⁑(π’œ)βŠ•Trz⁑(π’œβ€²)\operatorname{Tr}^{z}(\mathcal{A}\times\mathcal{A}^{\prime})=\operatorname{Tr}^{z}(\mathcal{A})\oplus\operatorname{Tr}^{z}(\mathcal{A}^{\prime}) and STrz⁑(π’œΓ—π’œβ€²)=STrz⁑(π’œ)βŠ•STrz⁑(π’œβ€²)\operatorname{STr}^{z}(\mathcal{A}\times\mathcal{A}^{\prime})=\operatorname{STr}^{z}(\mathcal{A})\oplus\operatorname{STr}^{z}(\mathcal{A}^{\prime}). ∎

8. Decomposition of tensor products

We recall the different notations for the simple self-dual projective module:

Prβˆ’1=V0=Srβˆ’1.P_{r-1}=V_{0}=S_{r-1}.

From Proposition 6.2 any projective indecomposable module of π’ž0Β―βˆͺπ’ž1Β―\mathscr{C}_{\overline{0}}\cup\mathscr{C}_{\overline{1}} is an element of the set {PiβŠ—β„‚k​rH,i∈{0,1,…​rβˆ’1},kβˆˆβ„€}\{P_{i}\otimes\mathbb{C}^{H}_{kr},i\in\{0,1,\ldots r-1\},k\in\mathbb{Z}\}. Let us recall their characters

χ​(PiβŠ—β„‚k​rH)=Xk​r​[r]X​(Xrβˆ’iβˆ’1+Xβˆ’r+i+1)andχ​(V0βŠ—β„‚k​rH)=Xr​k​[r]X\chi(P_{i}\otimes\mathbb{C}^{H}_{kr})=X^{kr}[r]_{X}(X^{r-i-1}+X^{-r+i+1}){\quad\text{and}\quad}\chi(V_{0}\otimes\mathbb{C}^{H}_{kr})=X^{rk}[r]_{X}

where i∈{0,1,…​rβˆ’2}i\in\{0,1,\ldots r-2\}. Observe now that these characters are linearly independent in ℀​[XΒ±1]\mathbb{Z}[X^{\pm 1}] and form a basis of an ideal of polynomials which are divisible by [r]X[r]_{X} (but not of the whole ideal generated by [r]X[r]_{X}).

As a consequence to decompose a projective module PP in direct sum of projective indecomposable ones, it is sufficient to decompose χ​(P)\chi(P) as

χ​(P)=βˆ‘i=0rβˆ’1βˆ‘kiβˆˆβ„€ni,ki​χ​(β„‚ki​rHβŠ—Pi).\chi(P)=\sum_{i=0}^{r-1}\sum_{k_{i}\in\mathbb{Z}}n_{i,k_{i}}\chi{\left(\mathbb{C}^{H}_{k_{i}r}\otimes P_{i}\right)}.

In the following, we write

βˆ‘k=mby ​2nand⨁k=mby ​2n{\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=m\\ \text{by }2\end{array}}}^{n}\!\!\!}}{\quad\text{and}\quad}{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=m\\ \text{by }2\end{array}}}^{n}\!\!\!}}

for the sums where kk is k≀nk\leq n and varies in the set m+2​ℕm+2\mathbb{N}. Similarly, we write

βˆ‘k=nbyΒ βˆ’2mand⨁k=nbyΒ βˆ’2m{\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=n\\ \text{by }-2\end{array}}}^{m}\!\!\!}}{\quad\text{and}\quad}{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=n\\ \text{by }-2\end{array}}}^{m}\!\!\!}}

for the sums where kβ‰₯mk\geq m and varies in the set nβˆ’2​ℕn-2\mathbb{N}.

Lemma 8.1 (Decomposition of tensor products V0βŠ—SiV_{0}\otimes S_{i}).

Let 0≀i≀rβˆ’10\leq i\leq r-1. Then

V0βŠ—Si=⨁k=rβˆ’1βˆ’iby ​2rβˆ’1Pk.V_{0}\otimes S_{i}={\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=r-1-i\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}P_{k}.
Proof.

If ii is odd it holds :

χ​(V0βŠ—Si)=[r]X​[i+1]X=[r]Xβ€‹βˆ‘j=ibyΒ βˆ’21(Xj+Xβˆ’j)=βˆ‘j=ibyΒ βˆ’20χ​(Prβˆ’1βˆ’j).\chi(V_{0}\otimes S_{i})=[r]_{X}[i+1]_{X}=[r]_{X}{\displaystyle{\sum_{\tiny{\begin{array}[]{c}j=i\\ \text{by }-2\end{array}}}^{1}\!\!\!}}(X^{j}+X^{-j})={\displaystyle{\sum_{\tiny{\begin{array}[]{c}j=i\\ \text{by }-2\end{array}}}^{0}\!\!\!}}\chi(P_{r-1-j}).

If ii is even it holds :

χ​(V0βŠ—Si)=[r]X​[i+1]X=[r]X​(1+βˆ‘j=ibyΒ βˆ’22(Xj+Xβˆ’j))=βˆ‘j=ibyΒ βˆ’20χ​(Prβˆ’1βˆ’j).\chi(V_{0}\otimes S_{i})=[r]_{X}[i+1]_{X}=[r]_{X}{\left(1+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}j=i\\ \text{by }-2\end{array}}}^{2}\!\!\!}}(X^{j}+X^{-j})\right)}={\displaystyle{\sum_{\tiny{\begin{array}[]{c}j=i\\ \text{by }-2\end{array}}}^{0}\!\!\!}}\chi(P_{r-1-j}).

∎

Proposition 8.2 (The decomposition of the tensor products PiβŠ—SjP_{i}\otimes S_{j}).

Let 0≀i≀rβˆ’20\leq i\leq r-2 and 0≀j≀rβˆ’10\leq j\leq r-1. It holds:

PiβŠ—Sj=(⨁k=|iβˆ’j|by ​2min⁑(i+j,rβˆ’1)Pk)βŠ•(⨁k=2​rβˆ’2βˆ’iβˆ’jby ​2rβˆ’1Pk)βŠ•(⨁k=r+iβˆ’jby ​2rβˆ’1PkβŠ—(β„‚rHβŠ•β„‚βˆ’rH))P_{i}\otimes S_{j}=\left({\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=|i-j|\\ \text{by }2\end{array}}}^{\min(i+j,r-1)}\!\!\!}}P_{k}\right)\oplus\left({\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=2r-2-i-j\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}P_{k}\right)\oplus\left({\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=r+i-j\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}P_{k}\otimes(\mathbb{C}^{H}_{r}\oplus\mathbb{C}^{H}_{-r})\right)

where the sums are meant to be empty if the lower bound is bigger than the upper bound.

Proof.

It holds :

χ​(PiβŠ—Sj)=[r]X​(Xrβˆ’iβˆ’1+Xβˆ’r+i+1)​[j+1]X=[r]X​([rβˆ’i+j]Xβˆ’[rβˆ’iβˆ’jβˆ’2]X).\chi(P_{i}\otimes S_{j})=[r]_{X}(X^{r-i-1}+X^{-r+i+1})[j+1]_{X}=[r]_{X}([r-i+j]_{X}-[r-i-j-2]_{X}).

Recall that [βˆ’n]=βˆ’[n][-n]=-[n]. We denote the parity of rβˆ’1βˆ’i+jr-1-i+j and rβˆ’3βˆ’iβˆ’jr-3-i-j by p∈{0,1}p\in\{0,1\} (note they coincide). If i>ji>j and i+j≀rβˆ’2i+j\leq r-2 we have :

χ​(PiβŠ—Sj)=[r]X​(βˆ‘l=rβˆ’iβˆ’jβˆ’1by ​2rβˆ’i+jβˆ’1(Xl+Xβˆ’l))=βˆ‘l=rβˆ’iβˆ’jβˆ’1by ​2rβˆ’i+jβˆ’1χ​(Prβˆ’1βˆ’l)=βˆ‘k=iβˆ’jby ​2i+jχ​(Pk).\chi(P_{i}\otimes S_{j})=[r]_{X}\left({\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=r-i-j-1\\ \text{by }2\end{array}}}^{r-i+j-1}\!\!\!}}(X^{l}+X^{-l})\right)={\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=r-i-j-1\\ \text{by }2\end{array}}}^{r-i+j-1}\!\!\!}}\chi(P_{r-1-l})={\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=i-j\\ \text{by }2\end{array}}}^{i+j}\!\!\!}}\chi(P_{k}).

If iβ‰₯ji\geq j and i+jβ‰₯rβˆ’1i+j\geq r-1 and p=1p=1 we have :

χ​(PiβŠ—Sj)=[r]X​(βˆ‘l=pby ​2rβˆ’i+jβˆ’1(Xl+Xβˆ’l)+βˆ‘l=pby ​2i+jβˆ’r+1(Xl+Xβˆ’l))=\displaystyle\chi(P_{i}\otimes S_{j})=[r]_{X}\left({\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=p\\ \text{by }2\end{array}}}^{r-i+j-1}\!\!\!}}(X^{l}+X^{-l})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=p\\ \text{by }2\end{array}}}^{i+j-r+1}\!\!\!}}(X^{l}+X^{-l})\right)=
=βˆ‘l=pby ​2rβˆ’i+jβˆ’1χ​(Prβˆ’1βˆ’l)+βˆ‘l=pby ​2i+jβˆ’r+1χ​(Prβˆ’1βˆ’l)=βˆ‘k=iβˆ’jby ​2rβˆ’1βˆ’pχ​(Pk)+βˆ‘k=2​rβˆ’2βˆ’iβˆ’jby ​2rβˆ’1βˆ’pχ​(Pk).\displaystyle={\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=p\\ \text{by }2\end{array}}}^{r-i+j-1}\!\!\!}}\chi(P_{r-1-l})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=p\\ \text{by }2\end{array}}}^{i+j-r+1}\!\!\!}}\chi(P_{r-1-l})={\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=i-j\\ \text{by }2\end{array}}}^{r-1-p}\!\!\!}}\chi(P_{k})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=2r-2-i-j\\ \text{by }2\end{array}}}^{r-1-p}\!\!\!}}\chi(P_{k}).

Note a similar calculation gives the result above in the case p=0p=0 (just pay attention to the fact that if p=0p=0 the terms X0+Xβˆ’0X^{0}+X^{-0} should be replaced by X0X^{0}).

Let us now suppose that j>ij>i and i+j≀rβˆ’2i+j\leq r-2 and let q∈{0,1}q\in\{0,1\} be the parity of jβˆ’iβˆ’1j-i-1. Then, if q=1q=1 (as above, a similar calculation proves the same final formula if q=0q=0) it holds :

χ​(PiβŠ—Sj)=[r]X​(βˆ‘l=rβˆ’j+i+1by ​2rβˆ’i+jβˆ’1(Xl+Xβˆ’l)+βˆ‘l=rβˆ’iβˆ’jβˆ’1by ​2rβˆ’j+iβˆ’1(Xl+Xβˆ’l))=\displaystyle\chi(P_{i}\otimes S_{j})=[r]_{X}\left({\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=r-j+i+1\\ \text{by }2\end{array}}}^{r-i+j-1}\!\!\!}}(X^{l}+X^{-l})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=r-i-j-1\\ \text{by }2\end{array}}}^{r-j+i-1}\!\!\!}}(X^{l}+X^{-l})\right)=
[r]X​((Xr+Xβˆ’r)β€‹βˆ‘h=qby ​2jβˆ’iβˆ’1(Xh+Xβˆ’h)+βˆ‘l=rβˆ’iβˆ’jβˆ’1by ​2rβˆ’j+iβˆ’1(Xl+Xβˆ’l))=\displaystyle[r]_{X}\left((X^{r}+X^{-r}){\displaystyle{\sum_{\tiny{\begin{array}[]{c}h=q\\ \text{by }2\end{array}}}^{j-i-1}\!\!\!}}(X^{h}+X^{-h})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=r-i-j-1\\ \text{by }2\end{array}}}^{r-j+i-1}\!\!\!}}(X^{l}+X^{-l})\right)=
=(Xr+Xβˆ’r)β€‹βˆ‘k=rβˆ’j+iby ​2rβˆ’1βˆ’qχ​(Pk)+βˆ‘k=jβˆ’iby ​2i+jχ​(Pk).\displaystyle=(X^{r}+X^{-r}){\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=r-j+i\\ \text{by }2\end{array}}}^{r-1-q}\!\!\!}}\chi(P_{k})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=j-i\\ \text{by }2\end{array}}}^{i+j}\!\!\!}}\chi(P_{k}).

Finally suppose that j>ij>i and i+jβ‰₯rβˆ’1i+j\geq r-1 and as before let p∈{0,1}p\in\{0,1\} be the parity of rβˆ’i+jβˆ’1r-i+j-1. If p=1p=1 (and as above if p=0p=0 or q=0q=0 modify the calculation by replacing the terms X0+Xβˆ’0X^{0}+X^{-0} by X0X^{0}, still getting the same final result):

χ​(PiβŠ—Sj)=[r]X​(βˆ‘l=pby ​2rβˆ’i+jβˆ’1(Xl+Xβˆ’l)+βˆ‘l=pby ​2i+j+1βˆ’r(Xl+Xβˆ’l))=\displaystyle\chi(P_{i}\otimes S_{j})=[r]_{X}\left({\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=p\\ \text{by }2\end{array}}}^{r-i+j-1}\!\!\!}}(X^{l}+X^{-l})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=p\\ \text{by }2\end{array}}}^{i+j+1-r}\!\!\!}}(X^{l}+X^{-l})\right)=
[r]X​(βˆ‘h=rβˆ’j+i+1by ​2r+jβˆ’iβˆ’1Xh+βˆ‘h=rβˆ’j+i+1by ​2r+jβˆ’iβˆ’1Xβˆ’h+βˆ‘h=pby ​2rβˆ’j+iβˆ’1(Xh+Xβˆ’h)+βˆ‘l=pby ​2i+j+1βˆ’r(Xl+Xβˆ’l))=\displaystyle[r]_{X}\left({\displaystyle{\sum_{\tiny{\begin{array}[]{c}h=r-j+i+1\\ \text{by }2\end{array}}}^{r+j-i-1}\!\!\!}}X^{h}+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}h=r-j+i+1\\ \text{by }2\end{array}}}^{r+j-i-1}\!\!\!}}X^{-h}+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}h=p\\ \text{by }2\end{array}}}^{r-j+i-1}\!\!\!}}(X^{h}+X^{-h})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=p\\ \text{by }2\end{array}}}^{i+j+1-r}\!\!\!}}(X^{l}+X^{-l})\right)=
=[r]X​((Xr+Xβˆ’r)β€‹βˆ‘s=qby ​2jβˆ’iβˆ’1(Xs+Xβˆ’s)+βˆ‘h=pby ​2rβˆ’j+iβˆ’1(Xh+Xβˆ’h)+βˆ‘l=pby ​2i+j+1βˆ’r(Xl+Xβˆ’l))=\displaystyle=[r]_{X}\left((X^{r}+X^{-r}){\displaystyle{\sum_{\tiny{\begin{array}[]{c}s=q\\ \text{by }2\end{array}}}^{j-i-1}\!\!\!}}(X^{s}+X^{-s})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}h=p\\ \text{by }2\end{array}}}^{r-j+i-1}\!\!\!}}(X^{h}+X^{-h})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}l=p\\ \text{by }2\end{array}}}^{i+j+1-r}\!\!\!}}(X^{l}+X^{-l})\right)=
=(Xr+Xβˆ’r)β€‹βˆ‘k=rβˆ’j+iby ​2rβˆ’1βˆ’qχ​(Pk)+βˆ‘k=jβˆ’iby ​2rβˆ’1βˆ’pχ​(Pk)+βˆ‘k=2​rβˆ’2βˆ’iβˆ’jby ​2rβˆ’1βˆ’pχ​(Pk).\displaystyle=(X^{r}+X^{-r}){\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=r-j+i\\ \text{by }2\end{array}}}^{r-1-q}\!\!\!}}\chi(P_{k})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=j-i\\ \text{by }2\end{array}}}^{r-1-p}\!\!\!}}\chi(P_{k})+{\displaystyle{\sum_{\tiny{\begin{array}[]{c}k=2r-2-i-j\\ \text{by }2\end{array}}}^{r-1-p}\!\!\!}}\chi(P_{k}).

To summarize the above computations, let p,q∈{0,1}p,q\in\{0,1\} be the parities of r+jβˆ’iβˆ’1r+j-i-1 and of jβˆ’iβˆ’1j-i-1, respectively. It holds:

(36) PiβŠ—Sj={⨁k=iβˆ’jby ​2i+jPkif​{iβ‰₯ji+j≀rβˆ’2⨁k=iβˆ’jby ​2rβˆ’1βˆ’pPk​⨁k=2​rβˆ’2βˆ’iβˆ’jby ​2rβˆ’1βˆ’pPkif​{iβ‰₯ji+jβ‰₯rβˆ’1⨁k=jβˆ’iby ​2i+jPk​⨁k=r+iβˆ’jby ​2rβˆ’1βˆ’q(β„‚rHβŠ•β„‚βˆ’rH)βŠ—Pkif​{i<ji+j≀rβˆ’2⨁k=jβˆ’iby ​2rβˆ’1βˆ’pPk​⨁k=2​rβˆ’2βˆ’iβˆ’jby ​2rβˆ’1βˆ’pPk​⨁k=rβˆ’j+iby ​2rβˆ’1βˆ’q(β„‚rHβŠ•β„‚βˆ’rH)βŠ—Pkif​{i<ji+jβ‰₯rβˆ’1.P_{i}\otimes S_{j}=\begin{cases}{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=i-j\\ \text{by }2\end{array}}}^{i+j}\!\!\!}}P_{k}&{\rm if}\ \left\{\begin{array}[]{l}i\geq j\\ i+j\leq r-2\end{array}\right.\\ {\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=i-j\\ \text{by }2\end{array}}}^{r-1-p}\!\!\!}}P_{k}{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=2r-2-i-j\\ \text{by }2\end{array}}}^{r-1-p}\!\!\!}}P_{k}&{\rm if}\ \left\{\begin{array}[]{l}i\geq j\\ i+j\geq r-1\end{array}\right.\\ {\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=j-i\\ \text{by }2\end{array}}}^{i+j}\!\!\!}}P_{k}{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=r+i-j\\ \text{by }2\end{array}}}^{r-1-q}\!\!\!}}(\mathbb{C}_{r}^{H}\oplus\mathbb{C}_{-r}^{H})\otimes P_{k}&{\rm if}\ \left\{\begin{array}[]{l}i<j\\ i+j\leq r-2\end{array}\right.\\ {\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=j-i\\ \text{by }2\end{array}}}^{r-1-p}\!\!\!}}P_{k}{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=2r-2-i-j\\ \text{by }2\end{array}}}^{r-1-p}\!\!\!}}\hskip-8.61108ptP_{k}{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=r-j+i\\ \text{by }2\end{array}}}^{r-1-q}\!\!\!}}(\mathbb{C}_{r}^{H}\oplus\mathbb{C}_{-r}^{H})\otimes P_{k}&{\rm if}\ \left\{\begin{array}[]{l}i<j\\ i+j\geq r-1\end{array}\right..\end{cases}

This is equivalent to the statement of the proposition. ∎

Let us now remark that for each i∈{0,1,…​rβˆ’2}i\in\{0,1,\ldots r-2\},

χ​(Pi)=2​χ​(Si)+(χ​(β„‚rH)+χ​(β„‚βˆ’rH))​χ​(Srβˆ’2βˆ’i).\chi(P_{i})=2\chi(S_{i})+(\chi(\mathbb{C}^{H}_{r})+\chi(\mathbb{C}^{H}_{-r}))\chi(S_{r-2-i}).

This, together with Proposition 8.2 and the fact that the modules PiP_{i} are projective allow us to compute the full tensor decomposition of PiβŠ—PjP_{i}\otimes P_{j} :

Corollary 8.3 (The tensor decomposition of PiβŠ—PjP_{i}\otimes P_{j}).

For each i,j∈{0,1,…​rβˆ’2}i,j\in\{0,1,\ldots r-2\} we have

PiβŠ—Pj=((β„‚rHβŠ•β„‚βˆ’rH)βŠ—(PiβŠ—Srβˆ’2βˆ’j))​⨁2​(PiβŠ—Sj)P_{i}\otimes P_{j}=\left((\mathbb{C}^{H}_{r}\oplus\mathbb{C}^{H}_{-r})\otimes\left(P_{i}\otimes S_{r-2-j}\right)\right)\bigoplus 2\left(P_{i}\otimes S_{j}\right)

and so

PiβŠ—Pj=(2​⨁k=|iβˆ’j|by ​2min⁑(i+j,rβˆ’1)Pk)βŠ•(2​⨁k=2​rβˆ’2βˆ’iβˆ’jby ​2rβˆ’1Pk)βŠ•(2​⨁k=r+iβˆ’jby ​2rβˆ’1PkβŠ—(β„‚rHβŠ•β„‚βˆ’rH))βŠ•P_{i}\otimes P_{j}=\left(2{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=|i-j|\\ \text{by }2\end{array}}}^{\min(i+j,r-1)}\!\!\!}}P_{k}\right)\oplus\left(2{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=2r-2-i-j\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}P_{k}\right)\oplus\left(2{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=r+i-j\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}P_{k}\otimes(\mathbb{C}^{H}_{r}\oplus\mathbb{C}^{H}_{-r})\right)\oplus
βŠ•(⨁k=|i+jβˆ’r+2|by ​2min⁑(i+rβˆ’jβˆ’2,rβˆ’1)PkβŠ—(β„‚rHβŠ•β„‚βˆ’rH))βŠ•(⨁k=rβˆ’i+jby ​2rβˆ’1PkβŠ—(β„‚rHβŠ•β„‚βˆ’rH))βŠ•\oplus\left({\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=|i+j-r+2|\\ \text{by }2\end{array}}}^{\min(i+r-j-2,r-1)}\!\!\!}}P_{k}\otimes(\mathbb{C}^{H}_{r}\oplus\mathbb{C}^{H}_{-r})\right)\oplus\left({\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=r-i+j\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}P_{k}\otimes(\mathbb{C}^{H}_{r}\oplus\mathbb{C}^{H}_{-r})\right)\oplus
βŠ•(⨁k=i+j+2by ​2rβˆ’1PkβŠ—(β„‚2​rHβŠ•2βŠ•β„‚βˆ’2​rH)).\oplus\left({\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=i+j+2\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}P_{k}\otimes(\mathbb{C}^{H}_{2r}\oplus 2\oplus\mathbb{C}^{H}_{-2r})\right).

Similarly V0βŠ—Pj=((β„‚rHβŠ•β„‚βˆ’rH)βŠ—(V0βŠ—Srβˆ’2βˆ’j))​⨁2​(V0βŠ—Sj),and​soV_{0}\otimes P_{j}=\left((\mathbb{C}^{H}_{r}\oplus\mathbb{C}^{H}_{-r})\otimes\left(V_{0}\otimes S_{r-2-j}\right)\right)\bigoplus 2\left(V_{0}\otimes S_{j}\right),\ {\rm and\ so}

V0βŠ—Pj=(⨁k=j+1by ​2rβˆ’1(β„‚rHβŠ•β„‚βˆ’rH)βŠ—Pk)​⨁⨁k=rβˆ’1βˆ’jby ​2rβˆ’12​Pk.V_{0}\otimes P_{j}=\left({\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=j+1\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}(\mathbb{C}^{H}_{r}\oplus\mathbb{C}^{H}_{-r})\otimes P_{k}\right)\bigoplus{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=r-1-j\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}2P_{k}.
Proposition 8.4.

Let i,j∈{0..rβˆ’1}i,j\in\{0..r-1\}. If i+j≀rβˆ’1i+j\leq r-1, then

SiβŠ—Sj=⨁k=|iβˆ’j|by ​2i+jSk.S_{i}\otimes S_{j}={\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=|i-j|\\ \text{by }2\end{array}}}^{i+j}\!\!\!}}S_{k}.

If i+jβ‰₯ri+j\geq r then

SiβŠ—Sj=⨁k=|iβˆ’j|by ​22​rβˆ’4βˆ’iβˆ’jSkβŠ•β¨k=2​rβˆ’2βˆ’iβˆ’jby ​2rβˆ’1Pk.S_{i}\otimes S_{j}={\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=|i-j|\\ \text{by }2\end{array}}}^{2r-4-i-j}\!\!\!}}S_{k}\oplus{\displaystyle{\bigoplus_{\tiny{\begin{array}[]{c}k=2r-2-i-j\\ \text{by }2\end{array}}}^{r-1}\!\!\!}}P_{k}.

In particular, semi-simple and projective modules of π’ž\mathscr{C} form a full sub-tensor category.

Proof.

The proof is by induction on ii using that for j∈{1​⋯​rβˆ’2}j\in\{1\cdots r-2\}, S1βŠ—Sj=Sjβˆ’1βŠ•Sj+1S_{1}\otimes S_{j}=S_{j-1}\oplus S_{j+1} and that Srβˆ’1S_{r-1} is projective. The induction is given by using

S1βŠ—SiβŠ—Sj=(Si+1βŠ—Sj)βŠ•(Siβˆ’1βŠ—Sj).S_{1}\otimes S_{i}\otimes S_{j}=(S_{i+1}\otimes S_{j})\oplus(S_{i-1}\otimes S_{j}).

To see the last point, remark that the tensor product of two simple modules is a direct sum of a semi-simple module direct sum a projective module. Thus, the full subcategory formed by semi-simple and projective modules is stable by tensor product. ∎

9. Multiplicity modules

Here we summarize some known facts about multiplicity modules. The one dimensional Hom\operatorname{Hom} spaces Homπ’žβ‘(β„‚,VΞ±βŠ—Vβˆ’Ξ±)\operatorname{Hom}_{\mathscr{C}}(\mathbb{C},V_{\alpha}\otimes V_{-\alpha}) and Homπ’žβ‘(β„‚,VΞ±βŠ—VΞ²βŠ—VΞ³)\operatorname{Hom}_{\mathscr{C}}(\mathbb{C},V_{\alpha}\otimes V_{\beta}\otimes V_{\gamma}) for Ξ±+Ξ²+γ∈{βˆ’(rβˆ’1),βˆ’(rβˆ’3),…,rβˆ’1}\alpha+\beta+\gamma\in\{-(r-1),-(r-3),\ldots,r-1\} can be equipped with nice basis. By a nice basis of Homπ’žβ‘(β„‚,VΞ±1βŠ—β‹―βŠ—VΞ±n)\operatorname{Hom}_{\mathscr{C}}(\mathbb{C},V_{\alpha_{1}}\otimes\cdots\otimes V_{\alpha_{n}}) we mean a set of basis of these spaces spaces such that

  1. (1)

    it depends analytically of the parameters Ξ±iβˆˆΒ¨β€‹β„‚\alpha_{i}\in{\ddot{}\mathbb{C}} (here we identify VΞ±V_{\alpha} with β„‚r=⨁iβ„‚.vi\mathbb{C}^{r}=\bigoplus_{i}\mathbb{C}.v_{i} as in Equation (15)) and

  2. (2)

    the set of basis is globally permuted by the pivotal isomorphism

    Homπ’žβ‘(β„‚,VΞ±1βŠ—β‹―βŠ—VΞ±n)⟢∼Homπ’žβ‘(β„‚,VΞ±2βŠ—β‹―βŠ—VΞ±nβŠ—VΞ±1).\operatorname{Hom}_{\mathscr{C}}(\mathbb{C},V_{\alpha_{1}}\otimes\cdots\otimes V_{\alpha_{n}})\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\operatorname{Hom}_{\mathscr{C}}(\mathbb{C},V_{\alpha_{2}}\otimes\cdots\otimes V_{\alpha_{n}}\otimes V_{\alpha_{1}}).

The existence of a nice basis has been checked in [22] for rr odd and in [10] for any rr but using a different normalizations. These basis are used in [25, 22, 10, 8, 4] to produce numerical invariant of ¨​ℂ{\ddot{}\mathbb{C}}-colored framed trivalent graphs embedded in S3S^{3}, and numerical 6j-symbols.

The basis of Homπ’žβ‘(β„‚,VΞ±βŠ—Vβˆ’Ξ±)\operatorname{Hom}_{\mathscr{C}}(\mathbb{C},V_{\alpha}\otimes V_{-\alpha}) induce isomorphisms wΞ±:VΞ±β†’Vβˆ’Ξ±βˆ—w_{\alpha}:V_{\alpha}\to V_{-\alpha}^{*} forming what is called a basic data (see [25]). Using these isomorphisms and the modified trace one gets a duality

Homπ’žβ‘(β„‚,VΞ±βŠ—VΞ²βŠ—VΞ³)βŠ—Homπ’žβ‘(β„‚,Vβˆ’Ξ³βŠ—Vβˆ’Ξ²βŠ—Vβˆ’Ξ±)β†’β„‚\operatorname{Hom}_{\mathscr{C}}(\mathbb{C},V_{\alpha}\otimes V_{\beta}\otimes V_{\gamma})\otimes\operatorname{Hom}_{\mathscr{C}}(\mathbb{C},V_{-\gamma}\otimes V_{-\beta}\otimes V_{-\alpha})\to\mathbb{C}

for which the basis are dual to each other.

The version UU of quantum 𝔰​𝔩​(2){\mathfrak{sl}(2)} used in [10] is slightly different from Uq​𝔰​𝔩​(2){U_{q}{\mathfrak{sl}(2)}}. To differentiate these algebras, let us call KU,EU,FU∈UK_{U},E_{U},F_{U}\in U the generators, then there is a morphism of Hopf algebras Uq​𝔰​𝔩​(2)β†’U{U_{q}{\mathfrak{sl}(2)}}\to U given by sending

K,E,F​ to respectively ​KU2,KU​EU,FU​KUβˆ’1.K,E,F\text{ to respectively }K_{U}^{2},K_{U}E_{U},F_{U}K_{U}^{-1}.

through this morphism, the module VaV^{a} of [10] can be identified with the module VΞ±V_{\alpha} where Ξ±=2​aβˆ’r+1\alpha=2a-r+1. Then the nice basis are given in [10] by computing some Clebsch-Gordan coefficients.
Different nice basis were computed in [22]. They were computed recursively using the morphisms X:VΞ±βŠ—VΞ²β†’VΞ±+1βŠ—VΞ²+1X:V_{\alpha}\otimes V_{\beta}\to V_{\alpha+1}\otimes V_{\beta+1} given by

X:viβŠ—vj↦qΞ²+iβˆ’jβˆ’1​{Ξ±βˆ’i}​viβŠ—vj+1+qβˆ’1​{Ξ²βˆ’j}​vi+1βŠ—vj.X:v_{i}\otimes v_{j}\mapsto q^{\beta+i-j-1}{\left\{\alpha-i\right\}}v_{i}\otimes v_{j+1}+q^{-1}{\left\{\beta-j\right\}}v_{i+1}\otimes v_{j}.

More than analytic in the parameters Ξ±i\alpha_{i}, they are given by Laurent polynomials in qΞ±iq^{\alpha_{i}}. But the work of [22] only consider odd values of rr.

10. Odd roots of unity

In this section we briefly discuss the quantum group of Subsection 2.2 when r∈2​ℕ+3r\in 2\mathbb{N}+3 is odd and q=e2β€‹Ο€β€‹βˆ’1rq={\operatorname{e}}^{\frac{2\pi\sqrt{-1}}{r}} is a rt​hr^{th}-root of unity. The reason why this case is not treated with the other are historic, technical, and due to the belief than topological applications won’t differ from the case q=eΟ€β€‹βˆ’1rq={\operatorname{e}}^{\frac{\pi\sqrt{-1}}{r}}.

Here the simple modules are

  1. (1)

    the dimension rr typical modules {VΞ±:Ξ±βˆˆΒ¨β€‹β„‚}\{V_{\alpha}:\alpha\in{\ddot{}\mathbb{C}}\} where now ¨​ℂ=(β„‚βˆ–12​℀)βˆͺr2​℀{\ddot{}\mathbb{C}}=(\mathbb{C}\setminus\frac{1}{2}\mathbb{Z})\cup\frac{r}{2}\mathbb{Z},

  2. (2)

    the dimension 11 invertible modules {β„‚k​r2H:kβˆˆβ„€}\{\mathbb{C}_{k\frac{r}{2}}^{H}:k\in\mathbb{Z}\}, and

  3. (3)

    the simple modules of dimension less than rr: {SiβŠ—β„‚k​r2H:0<i<r,kβˆˆβ„€}\{S_{i}\otimes\mathbb{C}_{k\frac{r}{2}}^{H}:0<i<r,k\in\mathbb{Z}\}, where the highest weight of SiS_{i} is ii.

One difference between the odd/even case is that Ohtsuki in [29] does not treat the case discussed in this subsection. In any case, when r∈2​ℕ+3r\in 2\mathbb{N}+3 the category is still pivotal with the same pivot given by Krβˆ’1K^{r-1}. The fact that the formula (5) still defines a braiding on the category π’ž\mathscr{C} is proven in [23, section 5.8]. The computation of Ohtsuki for the associated twist has never been completed in this case. Still in [23] we show that a full subcategory of π’ž\mathscr{C} that contain typical modules and self-dual modules is ribbon.

If gβˆˆβ„‚/2β€‹β„€βˆ–(12​℀)/2​℀g\in\mathbb{C}/2\mathbb{Z}\setminus(\frac{1}{2}\mathbb{Z})/2\mathbb{Z} then π’žg\mathscr{C}_{g} is semi-simple and π’žgβŠ‚π–―π—‹π—ˆπ—ƒ\mathscr{C}_{g}\subset{\mathsf{Proj}}. Typical modules are projective and there exists a unique trace on π–―π—‹π—ˆπ—ƒ{\mathsf{Proj}} up to a scalar. Its associated modified dimension is given by formula (18).

A nice basis for the multiplicity modules Hom⁑(β„‚,VΞ±βŠ—VΞ²βŠ—VΞ³)\operatorname{Hom}(\mathbb{C},V_{\alpha}\otimes V_{\beta}\otimes V_{\gamma}) is missing in the literature, and the 6​j6j-symbols have not been computed in this case (they have been computed when qq is a 2 times odd root of unity in [22] and for any even root of unity in [10] with a different normalization).

In [25, 23, 8] the authors construct topological invariants of dimension 33 using algebraic data. The case treated with most attention is that of quantum 𝔰​𝔩​(2){\mathfrak{sl}(2)} when qq is a root of unity of order 2​r2r but the case we have discussed in this section is also considered as an example all together with the quantum groups associated to the other simple Lie algebras (also at odd root of unity).

References

  • [1] D. AdamoviΔ‡ and A. Milas - Logarithmic intertwining operators and 𝒲​(2,2​pβˆ’1)\mathscr{W}(2,2p-1) algebras. J. Math. Phys. 48 (2007), no. 7, 073503, 20 p.
  • [2] Y. Akutsu, T. Deguchi, and T. Ohtsuki - Invariants of colored links. J. Knot Theory Ramifications 1 (1992), no. 2, 161–184.
  • [3] A. Beliakova, K. Habiro, A. Lauda, M. Ε½ivkoviΔ‡ - Trace decategorification of categorified quantum s​l​(2)sl(2), arXiv:1404.1806.
  • [4] C. Blanchet, F. Costantino, N. Geer, and B. Patureau-Mirand - Non semi-simple TQFTs, Reidemeister torsion and Kashaev’s invariants. Preprint, arXiv:1404.7289.
  • [5] P. Bushlanov, A. Gainutdinov and I. Tipunin - Kazhdan-Lusztig equivalence and fusion of Kac modules in Virasoro logarithmic models, Nuclear Phys. B 862 (2012), no. 1, 232–269.
  • [6] P. Bushlanov, B. Feigin, A. Gainutdinov, and I. Tipunin - Lusztig limit of quantum sl​(2){\rm sl}(2) at root of unity and fusion of (1,p)(1,p) Virasoro logarithmic minimal models. Nuclear Phys. B 818 (2009), no. 3, 179–195.
  • [7] V. Chari, A. Pressley - A guide to quantum groups, Cambridge University Press, Cambridge, 1994.
  • [8] F. Costantino, N. Geer and B. Patureau-Mirand - Quantum invariants of 3-manifolds via link surgery presentations and non-semi-simple categories. J Topology (2014) doi:10.1112/jtopol/jtu006, arXiv:1202.3553.
  • [9] F. Costantino, N. Geer, B. Patureau-Mirand - Relations between Witten-Reshetikhin-Turaev invariants and non-semi-simple 𝔰​l​(2){\mathfrak{s}l}(2) 33-manifold invariants, preprint arXiv:1310.2735 (2013).
  • [10] F. Costantino, J. Murakami - On S​L​(2,β„‚)SL(2,\mathbb{C}) quantum 6​j6j-symbols and its relation to the hyperbolic volume. Quantum Topology 4 (2013), no. 3, 303–351.
  • [11] T. Creutzig and A. Milas - False Theta Functions and the Verlinde formula, Adv. Math. 262 (2014), 520–545.
  • [12] T. Creutzig, D. Ridout and S. Wood - Coset Constructions of Logarithmic (1,p) Models. Lett. Math. Phys. 104 (2014), no. 5, 553–583.
  • [13] C. De Concini, V.G. Kac - Representations of quantum groups at roots of 11. In Operator algebras, unitary representations, enveloping algebras, and invariant theory. (Paris, 1989), 471–506, Progr. Math., 92, Birkhauser Boston, 1990.
  • [14] C. De Concini, V.G. Kac, C. Procesi - Quantum coadjoint action. J. Amer. Math. Soc. 5 (1992), no. 1, 151–189.
  • [15] C. De Concini, V.G. Kac, C. Procesi - Some remarkable degenerations of quantum groups. Comm. Math. Phys. 157 (1993), no. 2, 405–427.
  • [16] C. De Concini, C. Procesi, N. Reshetikhin, M. Rosso - Hopf algebras with trace and representations. Invent. Math. 161 (2005), no. 1, 1–44.
  • [17] B. Feigin, A. Gainutdinov, A. Semikhatov and I. Tipunin - Modular group representations and fusion in logarithmic conformal field theories and in the quantum group center. Comm. Math. Phys. 265 (2006), no. 1, 47–93.
  • [18] A. GaΔ­nutdinov, A. Semikhatov, I. Tipunin and B. FeΔ­gin - The Kazhdan-Lusztig correspondence for the representation category of the triplet WW-algebra in logorithmic conformal field theories. (Russian) Teoret. Mat. Fiz. 148 (2006), no. 3, 398–427; translation in Theoret. and Math. Phys. 148 (2006), no. 3, 1210–1235
  • [19] J. Fuchs, C. Schweigert and C. Stigner - Higher genus mapping class group invariants from factorizable Hopf algebras, Adv. Math. 250 (2014), 285–319.
  • [20] N. Geer, J. Kujawa, B. Patureau-Mirand - Generalized trace and modified dimension functions on ribbon categories. Selecta Math. (N.S.) 17 (2011), no. 2, 453–504.
  • [21] N. Geer, B. Patureau-Mirand - Multivariable link invariants arising from Lie superalgebras of type I. J. Knot Theory Ramifications 19, Issue 1 (2010) 93–115.
  • [22] N. Geer, B. Patureau-Mirand - Polynomial 6j-Symbols and States Sums. Algebraic & Geometric Topology 11 (2011) 1821–1860.
  • [23] N. Geer, B. Patureau-Mirand - Topological invariants from non-restricted quantum groups. Algebraic & Geometric Topology, 13 (2013), no. 6, 3305–3363.
  • [24] N. Geer, B. Patureau-Mirand, V. Turaev - Modified quantum dimensions and re-normalized link invariants. Compos. Math. 145 (2009), no. 1, 196–212.
  • [25] N. Geer, B. Patureau-Mirand, V. Turaev - Modified 6j-Symbols and 3–manifold invariants. Adv. Math. 228 (2011), no. 2, 1163–1202.
  • [26] H. Kausch - Extended conformal algebras generated by a multiplet of primary fields. Phys. Lett. B, 259 (1991), no. 4, 448–455.
  • [27] J. Murakami - Colored Alexander Invariants and Cone-Manifolds. Osaka J. Math. Volume 45, Number 2 (2008), 265–564.
  • [28] K. Nagatomo and A. Tsuchiya - The triplet vertex operator algebra W​(p)W(p) and the restricted quantum group UΒ―q​(s​l2)\overline{U}_{q}(sl_{2}) at q=eπ​ipq=e^{\frac{\pi i}{p}}. Exploring new structures and natural constructions in mathematical physics, Adv. Stud. Pure Math., 61, Math. Soc. Japan, Tokyo, (2011), 1–49.
  • [29] T. Ohtsuki - Quantum invariants. A study of knots, 3–manifolds, and their sets. Series on Knots and Everything, 29. World Scientific Publishing Co., Inc., River Edge, NJ, (2002).
  • [30] R.S. Pierce - Associative algebras. Graduate Texts in Mathematics, 88. Studies in the History of Modern Science, 9. Springer-Verlag, New York-Berlin, (1982).
  • [31] N. Reshetikhin, V.G. Turaev - Ribbon graphs and their invariants derived from quantum groups. Comm. Math. Phys. 127 (1990), no. 1, 1–26.
  • [32] N. Reshetikhin, V.G. Turaev - Invariants of 3–manifolds via link polynomials and quantum groups. Invent. Math. 103 (1991), no. 3, 547–597.
  • [33] V.G. Turaev - Quantum invariants of knots and 3–manifolds. de Gruyter Studies in Mathematics, 18. Walter de Gruyter & Co., Berlin, (1994).
  • [34] E. Witten - Quantum field theory and Jones polynomial. Comm. Math. Phys. 121 (1989), 351–399.