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

Generalised Temperley-Lieb algebras of type G(r,1,n)G(r,1,n)

Gus Lehrer and Mengfan Lyu School of Mathematics and Statistics, University of Sydney, N.S.W. 2006, Australia [email protected],[email protected]
Abstract.

In this paper, we define a quotient of the cyclotomic Hecke algebra of type G(r,1,n)G(r,1,n) as a generalisation of the Temperley-Lieb algebras of type AA and BB. We establish a graded cellular structure for the generalised Temperley-Lieb algebra and, using the technology of KLRKLR algebras, determine the corresponding decomposition matrix.

Key words and phrases:
Temperley-Lieb algebra, Hecke algebras, KLR algebras, cellular basis, decomposition numbers
2020 Mathematics Subject Classification:
Primary 16G20, 20C08; Secondary 20G42

1. Introduction

1.1. Temperley-Lieb algebras of types An1A_{n-1} and BnB_{n}

The Temperley-Lieb algebra was originally introduced in [29] in connection with transition matrices in statistical mechanics. Since then, it has been shown to have links with many diverse areas of mathematics, such as operator algebras, quantum groups, categorification and representation theory. It can be defined in several different ways,including as a quotient of the Hecke algebra and as an associative diagram algebra. Steinberg pointed out in 1982 during a lecture at UCLA that the Temperley-Lieb algebra TLn(q)TL_{n}(q) is a quotient of the Hecke algebra Hn(q)H_{n}(q) of type An1A_{n-1} by the ideal generated by some idempotents corresponding to two non-commuting generators. It plays a pivotal role in the polynomial invariant of knots and links discovered by Jones in [18]. In 1990, Kauffman found a diagrammatic presentation for the Temperley-Lieb algebra in [19]. It is defined as an associative diagram algebra modulo planar isotopy for classical unoriented knots and links. In [6], Bernstein, Frenkel and Khovanov ”categorify” the Temperley-Lieb algebra, which is the commutant of the quantum enveloping algebra Uq(𝔰𝔩2)U_{q}(\mathfrak{sl}_{2}) in its action on the nthn^{th} tensor power V1nV_{1}^{\bigotimes n} of the natural two dimensional representation V1V_{1}. More precisely, they realise the Temperley-Lieb algebra via the projective and Zuckerman functors in the usual category 𝒪\mathcal{O} for 𝔤𝔩n\mathfrak{gl}_{n}. Moreover, the Temperley-Lieb algebra also plays an important role in statistical physics models, and readers can find more references in [5],[2] and [1].

The “blob algebra” Bn(q,Q)B_{n}(q,Q) was introduced in [22] by Martin and Saleur in connection with statistical mechanics. It has a similar diagrammatic representation to the original Temperley-Lieb algebra. Algebraically, it can also be identified with a quotient of HBn(q,Q)HB_{n}(q,Q) the two-parameter Hecke algebra of type BnB_{n}. For these reasons, the blob algebra is considered as a type-BnB_{n} analogue of TLn(q)TL_{n}(q) and is also denoted by TLBn(q,Q)TLB_{n}(q,Q). In analogy to the original Temperley-Lieb algebra, TLBn(q,Q)TLB_{n}(q,Q) also fits into a category. In [17], Iohara, Lehrer and Zhang construct a family of equivalences between the Temperley-Lieb category of type B and a certain category of infinite dimensional representations of Uq(𝔰𝔩2)U_{q}(\mathfrak{sl}_{2}) as a generalization of the Temperley-Lieb category TL(q)TL(q).

Cellularity is a property of associative algebras which permits their deformation and enables the computation of their decomposition numbers. The above algebras are shown to be cellular in [13]. Using this result, the standard modules and decomposition numbers are well-known, see [23],[12] and [14].

1.2. Objectives of this paper

In this paper, we define a generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} in Definition 3.1 corresponding to the imprimitive unitary reflection group G(r,1,n)G(r,1,n), which can be regarded as a generalisation of the Coxeter groups of type An1A_{n-1} and BnB_{n}. This generalisation strengthens the connection between Temperley-Lieb algebras and cyclotomic Hecke algebras and brings potential innovations in knot theory, quantum enveloping algebras and statistical physics models. We also construct a cellular structure of our generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} in Theorem 4.9, which provides clues for a potential diagrammatic realisation. Using this cellular basis, we give a condition equivalent to the semisimplicity of TLr,1,nTL_{r,1,n} in Theorem 5.2. Moreover, we determine the associative decomposition matrix in Theorem 6.15.

Our generalisation is inspired by the construction of generalised blob algebras by Martin and Woodcock. They show in [24] that TLBn(q,Q)TLB_{n}(q,Q) can be obtained as the quotient of HBn(q,Q)HB_{n}(q,Q) by the idempotents corresponding to the irreducible modules associated with the bipartitions ((2),)((2),\emptyset) and (,(2))(\emptyset,(2)) and generalised this construction to all cyclotomic Hecke algebras H(r,1,n)H(r,1,n).

Following this method, we define in Definition 3.1 the generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} as a quotient of the corresponding cyclotomic Hecke algebra H(r,1,n)H(r,1,n) by the two-sided ideal generated by certain idempotents. The Temperley-Lieb algebras of both type An1A_{n-1} and BnB_{n} are special cases of TLr,1,nTL_{r,1,n}. Therefore, we refer to this quotient of the cyclotomic Hecke algebra H(r,1,n)H(r,1,n) as the generalised Temperley-Lieb algebra of type G(r,1,n)G(r,1,n).

1.3. The content of this work

As a direct consequence of the definition, our generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} is a quotient of the generalised blob algebra Br,nB_{r,n}. We use this property to construct a graded cellular structure on it.

Lobos and Ryom-Hansen [25] give a graded cellular basis associated with one-column multipartitions of nn for the generalised blob algebras Br,nB_{r,n}. In [7], Bowman gives the decomposition numbers for the standard modules of the generalised blob algebras Br,nB_{r,n}. In Lemma 4.8, we show that the ideal of Br,nB_{r,n} corresponding to its quotient TLr,1,nTL_{r,1,n} is exactly spanned by the elements corresponding to a downward closed subset of the poset in the cellular basis given by Lobos and Ryom-Hansen. Therefore, a graded cellular structure of TLr,1,nTL_{r,1,n} can be obtained by a truncation of that of Br,nB_{r,n}, which is described in Theorem 4.9. The graded cellular basis of TLr,1,nTL_{r,1,n} is associated with one-column rr-partitions of nn which consist of at most two non-empty components.

Using this graded cellular basis, we investigate the representations of the generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n}. Following the calculation of the dimensions of the cell modules, we concentrate on the bilinear form ϕλ(,)\phi_{\lambda}(,) on each cell module. For a one-column multipartition λ\lambda, let W(λ)W(\lambda) be the corresponding cell module of TLr,1,nTL_{r,1,n}. Define a bilinear form ϕλ(,)\phi_{\lambda}(,) on W(λ)W(\lambda) in the usual way.According to Theorem 3.4 in [13], W(λ)W(\lambda) is simple if and only if rad(ϕλ)=0rad(\phi_{\lambda})=0 . Using this method, we obtain a criterion for the semisimplicity of TLr,1,nTL_{r,1,n} in Theorem 5.2 and interpret it as a restriction on the parameters in Corollary 5.3.

As the main result in this paper, we determine the decomposition numbers for the cell modules using this bilinear form. We first give a method to find out the radical of the bilinear form ϕλ(,)\phi_{\lambda}(,) in section 6.2. By showing that this method is not sensitive to the parameter rr, we demonstrate that the dimension of a simple TLr,1,nTL_{r,1,n}-module equals that of a corresponding TLBn(q,Q)TLB_{n}(q,Q)-module with the parameters chosen properly. By showing that the cell and simple modules of TLr,1,nTL_{r,1,n} can be realised as those of a corresponding Temperley-Lieb algebra of type BnB_{n}, we transform the decomposition matrix of TLr,1,nTL_{r,1,n} into a combination of those of TLBn(q,Q)TLB_{n}(q,Q) in Lemma 6.12.

The non-graded decomposition numbers of the Temperley-Lieb algebras of type BnB_{n} are given by Martin and Woodcock in [24] and Graham and Lehrer in [14] and the graded ones are given by Plaza in [27]. Following their results, we give the decomposition numbers of TLr,1,nTL_{r,1,n} in Theorem 6.15.

2. Multipartitions, Cyclotomic Hecke algebras and KLR algebras

In this section, we recall some combinatorial concepts such as multipartitions and tableaux in the context of the representation theory of the cyclotomic Hecke algebra as developed by Arike and Koike in [4]. We will use this language to define our generalized Temperley-Lieb algebra TLr,1,nTL_{r,1,n} as a quotient of the cyclotomic Hecke algebra H(r,1,n)H(r,1,n). We also recall the cyclotomic KLR algebra nΛ\mathcal{R}_{n}^{\Lambda} of type A, which is isomorphic to H(r,1,n)H(r,1,n) according to Theorem 1.1 in [10]. We will construct a graded cellular basis of TLr,1,nTL_{r,1,n} using the KLR generators in section 4.

2.1. Multipartitions and their Young tableaux

A partition of nn is a sequence σ=(σ1σ2)\sigma=(\sigma_{1}\geq\sigma_{2}\geq\dots) of non-negative integers σi\sigma_{i} such that |σ|=i1σi=n|\sigma|=\sum_{i\geq 1}\sigma_{i}=n. Denote σ=(σ1,σ2,,σk)\sigma=(\sigma_{1},\sigma_{2},\dots,\sigma_{k}) if σi=0\sigma_{i}=0 for i>ki>k and σk>0\sigma_{k}>0. A multipartition (more specifically, an rr-partition) of nn is an ordered rr-tuple λ=(λ(1),,λ(r))\lambda=(\lambda^{(1)},\dots,\lambda^{(r)}) of partitions with |λ(1)|++|λ(r)|=n|\lambda^{(1)}|+\dots+|\lambda^{(r)}|=n; if the ithi^{th} component of λ\lambda is empty, |λ(i)|=0|\lambda^{(i)}|=0. Denote by 𝔓n(r)\mathfrak{P}_{n}^{(r)} the set of rr-partitions of nn.

The Young diagram of a multipartition λ𝔓n(r)\lambda\in\mathfrak{P}_{n}^{(r)} is the set of 3-tuples:

(2.1) [λ]:={(a,b,l)|a>0,1bλa(l),1lr}3.[\lambda]:=\{(a,b,l)\;|\;a>0,1\leq b\leq\lambda_{a}^{(l)},1\leq l\leq r\}\subset\mathbb{Z}^{3}.

For fixed ll, the subset [λ(l)]:={(a,b,l)|a>0,1bλa(l)}[\lambda^{(l)}]:=\{(a,b,l)|a>0,1\leq b\leq\lambda_{a}^{(l)}\} is the lthl^{th} component of this Young diagram of the multipartition λ\lambda. It can be regarded as a Young diagram of the partition λ(l)\lambda^{(l)}. If |λ(l)|=0|\lambda^{(l)}|=0, we call it an empty component of the multipartition λ\lambda. As there is a unique Young diagram [λ][\lambda] for each multipartition λ𝔓n(r)\lambda\in\mathfrak{P}_{n}^{(r)} and vice versa, we do not distinguish [λ][\lambda] and λ\lambda in the following sections of this thesis. For each 3-tuple (a,b,l)[λ](a,b,l)\in[\lambda], we call it the node in the atha^{th} row and bthb^{th} column of the lthl^{th} component. A Young tableau is obtained by filling in the nodes with numbers 1,2,,n1,2,\dots,n. More precisely, a Young tableau of shape λ\lambda is a bijective map:

(2.2) t:{1,2,3,,n}[λ].t:\{1,2,3,\dots,n\}\to[\lambda].

For example, let n=9n=9, r=3r=3 and λ=((4,3),0,(1,1))\lambda=((4,3),0,(1,1)). Figure-1 shows a Young tableau of shape λ\lambda.

125834976\emptysetll=123
Figure 1. A standard Young tableau of shape λ\lambda

A standard Young tableau is a Young tableau in which the numbers increase along each row and column in every component. Denote by Std(λ)Std(\lambda) the set of standard tableaux of shape λ\lambda. The tableau in Figure-1 is a standard Young tableau of shape ((4,3),,(1,1))((4,3),\emptyset,(1,1)).

2.2. One-column multipartitions and the partial order

Next we introduce the definition of one-column multipartitions which are the main tool we use to describe the representations of our generalised Temperley-Lieb algebras. A multipartition λ𝔓n(r)\lambda\in\mathfrak{P}_{n}^{(r)} is called a one-column multipartition if each non-empty component consists of only one column. In other words, λ\lambda is called a one-column multipartition if λa(l)1\lambda_{a}^{(l)}\leq 1 for all a>0,1lra>0,1\leq l\leq r. Figure 2 gives an example of the one-column multipartition and one of its standard tableaux. Denote by 𝔅n(r)\mathfrak{B}_{n}^{(r)} the set of one-column rr-partitions of nn. If γ=(i,1,k)\gamma=(i,1,k) and γ=(i,1,k)\gamma^{\prime}=(i^{\prime},1,k^{\prime}) are two nodes of a one-column multipartition λ\lambda, we say γ<γ\gamma<\gamma^{\prime} if either i<ii<i^{\prime} or i=ii=i^{\prime} and k<kk<k^{\prime}. We write γγ\gamma\leq\gamma^{\prime} if γ<γ\gamma<\gamma^{\prime} or γ=γ\gamma=\gamma^{\prime}. For example, let tt be the tableau in Figure 2, then we have t(1)<t(2)<t(4)<t(3)<t(5)<t(6)<t(7)<t(8)<t(9)t(1)<t(2)<t(4)<t(3)<t(5)<t(6)<t(7)<t(8)<t(9). In fact, \leq is a total order on the set of nodes N={(a,1,l)|a>0,1lr}N=\{(a,1,l)|a>0,1\leq l\leq r\}.

145789236\emptyset
Figure 2. A standard tableau of shape λ=((16),(13),0)\lambda=((1^{6}),(1^{3}),0)

For λ,μ𝔅n(r)\lambda,\mu\in\mathfrak{B}_{n}^{(r)}, we say λμ\lambda\unlhd\mu if for each γ0×{1}×{1,2,r}\gamma_{0}\in\mathbb{N}\times\{1\}\times\{1,2\dots,r\} we have

(2.3) |{γ[λ]:γγ0}||{γ[μ]:γγ0}|.|\{\gamma\in[\lambda]:\gamma\leq\gamma_{0}\}|\geq|\{\gamma\in[\mu]:\gamma\leq\gamma_{0}\}|.

Thus, roughly speaking, a multipartition is smaller if its diagram has more ‘small’ nodes with respect to the total order \leq. The relation \unlhd defines a partial order on 𝔅n(r)\mathfrak{B}_{n}^{(r)}. The cellular structure of our generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} will involve this poset. As an example, the diagram in Figure-3 shows the partial order on 𝔅3(3)\mathfrak{B}_{3}^{(3)}.

((1),(1),(1))((1),(1),(1))((12),(1),)((1^{2}),(1),\emptyset)((1),(12),)((1),(1^{2}),\emptyset)((12),,(1))((1^{2}),\emptyset,(1))((1),,(12))((1),\emptyset,(1^{2}))(,(12),(1))(\emptyset,(1^{2}),(1))((13),,)((1^{3}),\emptyset,\emptyset)(,(1),(12))(\emptyset,(1),(1^{2}))(,(13),)(\emptyset,(1^{3}),\emptyset)(,,(13))(\emptyset,\emptyset,(1^{3}))
Figure 3. Partial order on 𝔅3(3)\mathfrak{B}_{3}^{(3)}

For 1kr1\leq k\leq r, denote by 𝔅n(r)(k)\mathfrak{B}_{n}^{(r)}(k) the subset of 𝔅n(r)\mathfrak{B}_{n}^{(r)} consisting of the multipartitions containing exactly kk non-empty components. We next discuss some consequences of the definition of the partial order.

Lemma 2.1.

Let λ𝔅n(r)(k)\lambda\in\mathfrak{B}_{n}^{(r)}(k) and μ𝔅n(r)(l)\mu\in\mathfrak{B}_{n}^{(r)}(l) be two rr-partitions of nn. If λμ\lambda\unlhd\mu, then klk\geq l.

Proof.

Let γ0=(1,1,r)\gamma_{0}=(1,1,r), then we have |{γ[λ]:γγ0}|=k|\{\gamma\in[\lambda]:\gamma\leq\gamma_{0}\}|=k and |{γ[μ]:γγ0}|=l|\{\gamma\in[\mu]:\gamma\leq\gamma_{0}\}|=l. The statement is now an immediate consequence of the definition of the partial order in (2.3)(\ref{podf}). ∎

The converse of this lemma is not always true. Counterexamples can be found in Figure-3.

To study the cellular structure of our generalized Temperley-Lieb algebras, we are particularly interested in those multipartitions with at most two non-empty components. The following lemmas give descriptions of the partial order on the elements of 𝔅n(r)(1)\mathfrak{B}_{n}^{(r)}(1) and 𝔅n(r)(2)\mathfrak{B}_{n}^{(r)}(2).

Lemma 2.2.

Let λ,μ𝔅n(r)(1)\lambda,\mu\in\mathfrak{B}_{n}^{(r)}(1). Suppose the kthk^{th} component of λ\lambda and lthl^{th} component of μ\mu are non-empty; then λμ\lambda\unlhd\mu if and only if klk\leq l.

Lemma 2.3.

Let λ𝔅n(r)(2)\lambda\in\mathfrak{B}_{n}^{(r)}(2) with λ(k1)\lambda^{(k_{1})} and λ(k2)\lambda^{(k_{2})} (k1<k2k_{1}<k_{2}) non-empty and let μ𝔅n(r)(1)\mu\in\mathfrak{B}_{n}^{(r)}(1) with μ(l)\mu^{(l)} as the only non-empty component. Then λμ\lambda\unlhd\mu if and only if k1lk_{1}\leq l.

Denote by λk1,k2[a]𝔅n(r)(2)\lambda_{k_{1},k_{2}}^{[a]}\in\mathfrak{B}_{n}^{(r)}(2) (k1<k2k_{1}<k_{2}, |a|<n|a|<n and ana\equiv n modmod 2 ) the multipartition λ\lambda of which λ(k1)\lambda^{(k_{1})} and λ(k2)\lambda^{(k_{2})} are the two non-empty components and |λ(k1)||λ(k2)|=a|\lambda^{(k_{1})}|-|\lambda^{(k_{2})}|=a. The following lemma gives a description of the partial order on 𝔅n(r)(2)\mathfrak{B}_{n}^{(r)}(2).

Lemma 2.4.

Let λk1,k2[a],μl1,l2[b]𝔅n(r)(2)\lambda_{k_{1},k_{2}}^{[a]},\mu_{l_{1},l_{2}}^{[b]}\in\mathfrak{B}_{n}^{(r)}(2). Then λk1,k2[a]μl1,l2[b]\lambda_{k_{1},k_{2}}^{[a]}\unlhd\mu_{l_{1},l_{2}}^{[b]} if and only if k1l1k_{1}\leq l_{1}, k2l2k_{2}\leq l_{2} and one of the following holds:

1. |a|<|b||a|<|b|;

2.|a|=|b||a|=|b| and aba\geq b;

3. |a|=|b||a|=|b|, a<ba<b and k2l1k_{2}\leq l_{1}.

Proof.

For i=1,2i=1,2, denote by aia_{i} and bib_{i} the number of nodes in λ(ki)\lambda^{(k_{i})} and μ(li)\mu^{(l_{i})} respectively.

We first check the sufficiency. Let c=min(a1,a2,b1,b2)c=min(a_{1},a_{2},b_{1},b_{2}). As |a||b||a|\leq|b|, c=bic=b_{i} for some i{1,2}i\in\{1,2\}. The condition k1l1k_{1}\leq l_{1}, k2l2k_{2}\leq l_{2} implies that

(2.4) |{γ[λ]:γγ0}||{γ[μ]:γγ0}|,|\{\gamma\in[\lambda]:\gamma\leq\gamma_{0}\}|\geq|\{\gamma\in[\mu]:\gamma\leq\gamma_{0}\}|,

for γ0{1,2,,c}×{1}×{1,2,r}\gamma_{0}\in\{1,2,\dots,c\}\times\{1\}\times\{1,2\dots,r\} and the equality holds for γ0=(c,1,r)\gamma_{0}=(c,1,r). Denote by [λ][\lambda^{\prime}] and [μ][\mu^{\prime}] the sub-diagrams of [λk1,k2[a]][\lambda_{k_{1},k_{2}}^{[a]}] and [μl1,l2[b]][\mu_{l_{1},l_{2}}^{[b]}] consisting of the nodes below the cthc^{th} row. If they are not empty, |a||b||a|\leq|b| implies that [μ][\mu^{\prime}] has exactly one non-empty component and [λ][\lambda^{\prime}] has at most two. Lemma 2.2 and 2.3 guarantees the sufficiency of the listed three cases.

On the other hand, if λk1,k2[a]μl1,l2[b]\lambda_{k_{1},k_{2}}^{[a]}\unlhd\mu_{l_{1},l_{2}}^{[b]}, let γ0=(1,1,li)\gamma_{0}=(1,1,l_{i}) with i=1,2i=1,2 in (2.3)(\ref{podf}), then we have

k1l1 and k2l2.k_{1}\leq l_{1}\text{ and }k_{2}\leq l_{2}.

Similarly, let c=min(a1,a2,b1,b2)c=min(a_{1},a_{2},b_{1},b_{2}) and γ0=(c,1,r)\gamma_{0}=(c,1,r). We have

(2.5) |{γ[λ]:γγ0}|=|{γ[μ]:γγ0}|.|\{\gamma\in[\lambda]:\gamma\leq\gamma_{0}\}|=|\{\gamma\in[\mu]:\gamma\leq\gamma_{0}\}|.

Denote by [λ][\lambda^{\prime}] and [μ][\mu^{\prime}] the sub-diagrams of [λk1,k2[a]][\lambda_{k_{1},k_{2}}^{[a]}] and [μl1,l2[b]][\mu_{l_{1},l_{2}}^{[b]}] consisting of the nodes below the cthc^{th} row. Then λk1,k2[a]μl1,l2[b]\lambda_{k_{1},k_{2}}^{[a]}\unlhd\mu_{l_{1},l_{2}}^{[b]} implies λμ\lambda^{\prime}\unlhd\mu^{\prime}. If |a|>|b||a|>|b|, we have λ𝔅n2c(r)(1)\lambda^{\prime}\in\mathfrak{B}_{n-2c}^{(r)}(1) and μ𝔅n2c(r)(2)\mu^{\prime}\in\mathfrak{B}_{n-2c}^{(r)}(2), which is contradictory to λμ\lambda^{\prime}\unlhd\mu^{\prime} by Lemma 2.1. If |a|=|b||a|=|b| and a<0<ba<0<b, then λ𝔅n2c(r)(1)(l1)\lambda^{\prime}\in\mathfrak{B}_{n-2c}^{(r)}(1)(l_{1}) and μ𝔅n2c(r)(1)(k2)\mu^{\prime}\in\mathfrak{B}_{n-2c}^{(r)}(1)(k_{2}). By Lemma 2.2, we have k2l1k_{2}\leq l_{1}. Otherwise, we have either |a|<|b||a|<|b| or |a|=|b||a|=|b| and aba\geq b. This completes the proof. ∎

We next introduce a special standard tableau tλt^{\lambda} for each multipartition λ\lambda which will be used in Section 4 to define the cellular basis of our generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n}. Let tλt^{\lambda} be the tableau of shape λ\lambda such that tλ(i)<tλ(j)t^{\lambda}(i)<t^{\lambda}(j) if 1i<jn1\leq i<j\leq n. As \leq is a total order on the nodes, tλt^{\lambda} is unique for λ𝔅n(r)\lambda\in\mathfrak{B}_{n}^{(r)}. Moreover, for each tStd(λ)t\in Std(\lambda), denote by d(t)𝔖nd(t)\in\mathfrak{S}_{n} the unique permutation such that t=tλd(t)t=t^{\lambda}\circ d(t) where \circ is the natural 𝔖n\mathfrak{S}_{n}-action on the tableaux.

2.3. Cyclotomic Hecke algebras and their representations

We next recall the definition of cyclotomic Hecke algebras and description of their representations following [4] by Ariki and Koike. The irreducible representations are described in the language of multipartions. Our generalised Temperley-Lieb algebras will be defined as quotients of cyclotomic Hecke algebras by certain ideals generated by a set of idempotents corresponding to a set of irreducible representations.

We first recall the presentation of the imprimitive reflection group G(r,1,n)G(r,1,n).

Lemma 2.5.

(cf. [8],Appendix 2) Let rr and nn be positive integers. The unitary reflection group G(r,1,n)G(r,1,n) is the group generated by s0,s1,s2,s_{0},s_{1},s_{2}, s3,,sn1s_{3},\dots,s_{n-1} subject to the following relations:

s0r=s12=s22\displaystyle s_{0}^{r}=s_{1}^{2}=s_{2}^{2} ==sn12=1;\displaystyle=\dots=s_{n-1}^{2}=1;
s0s1s0s1=\displaystyle s_{0}s_{1}s_{0}s_{1}= s1s0s1s0;\displaystyle s_{1}s_{0}s_{1}s_{0};
sisi+1si=si+1sisi+1\displaystyle s_{i}s_{i+1}s_{i}=s_{i+1}s_{i}s_{i+1} for 1in2;\displaystyle\text{ for }1\leq i\leq n-2;
sisj=sjsi\displaystyle s_{i}s_{j}=s_{j}s_{i} for |ij|2.\displaystyle\text{ for }|i-j|\geq 2.

The generators s0,s1,s2,,sn1s_{0},s_{1},s_{2},\dots,s_{n-1} are called simple reflections and the relations above can be represented in the Dynkin diagram in Figure-4.

s0s_{0}s1s_{1}s2s_{2}s3s_{3}\dotssn2s_{n-2}sn1s_{n-1}G(r,1,n):G(r,1,n):rr2222222222
Figure 4. Dynkin diagram of type G(r,1,n)G(r,1,n)

The reflection group G(r,1,n)G(r,1,n) can be viewed as the group consisting of all n×nn\times n monomial matrices such that each non-zero entry is an rthr^{th} root of unity. Moreover, if ζ\zeta\in\mathbb{C}^{*} is a primitive rthr^{th} root of unity, the complex reflection group G(r,1,n)G(r,1,n) is generated by the following elements:

s0\displaystyle s_{0} =ζE1,1+k=2nEk,k,\displaystyle=\zeta E_{1,1}+\sum_{k=2}^{n}E_{k,k},
si\displaystyle s_{i} =1ki1Ek,k+Ei+1,i+Ei,i+1+i+2knEk,k for all 1in1\displaystyle=\sum_{1\leq k\leq i-1}E_{k,k}+E_{i+1,i}+E_{i,i+1}+\sum_{i+2\leq k\leq n}E_{k,k}\text{ for all }1\leq i\leq n-1

where Ei,jE_{i,j} is the elementary n×nn\times n matrix with the (i,j)(i,j)-entry nonzero. Viewed thus, the generators are realised as (pseudo)-reflections in the nn dimensional space over \mathbb{C}, exhibiting G(r,1,n)G(r,1,n) as a complex reflection group.

Let R=[q,q1,v1,v2,,vr]R=\mathbb{Z}[q,q^{-1},v_{1},v_{2},\dots,v_{r}] where q,v1,v2,,vrq,v_{1},v_{2},\dots,v_{r} are indeterminates over \mathbb{C}. The cyclotomic Hecke algebra corresponding to G(r,1,n)G(r,1,n) over RR is defined as follows:

Definition 2.6.

([8],Definition 4.21) Let G(r,1,n)G(r,1,n) and RR be as defined above. The cyclotomic Hecke algebra H(r,1,n)H(r,1,n) corresponding to G(r,1,n)G(r,1,n) over RR is the associative algebra generated by T0,T1,,Tn1T_{0},T_{1},\dots,T_{n-1} subject to the following relations:

(2.6) (T0v1)(T0v2)(T0vr)\displaystyle(T_{0}-v_{1})(T_{0}-v_{2})\dots(T_{0}-v_{r}) =0;\displaystyle=0;
(2.7) (Tiq)(Ti+1)\displaystyle(T_{i}-q)(T_{i}+1) =0 for 1in1;\displaystyle=0\textit{ for }1\leq i\leq n-1;
(2.8) T0T1T0T1\displaystyle T_{0}T_{1}T_{0}T_{1} =T1T0T1T0;\displaystyle=T_{1}T_{0}T_{1}T_{0};
(2.9) TiTi+1Ti\displaystyle T_{i}T_{i+1}T_{i} =Ti+1TiTi+1 for 1in2;\displaystyle=T_{i+1}T_{i}T_{i+1}\textit{ for }1\leq i\leq n-2;
(2.10) TiTj\displaystyle T_{i}T_{j} =TjTi for |ij|2.\displaystyle=T_{j}T_{i}\textit{ for }|i-j|\geq 2.

The irreducible representations of H(r,1,n)H(r,1,n) are described in the language of rr-partitions of nn. Let λ\lambda be an rr-partition of nn and KK be the quotient field of RR. Denote by fλf_{\lambda} the number of standard tableaux of shape λ\lambda and by 𝔜λ={𝕐1,𝕐2,,𝕐fλ}\mathfrak{Y}^{\lambda}=\{\mathbb{Y}_{1},\mathbb{Y}_{2},\dots,\mathbb{Y}_{f_{\lambda}}\} the set of all the standard tableaux of shape λ\lambda. Let 𝕍λ\mathbb{V}_{\lambda} be the vector space over KK with the basis 𝕐1,𝕐2,,𝕐fλ\mathbb{Y}_{1},\mathbb{Y}_{2},\dots,\mathbb{Y}_{f_{\lambda}}; namely,

(2.11) 𝕍λ=iK𝕐i,\mathbb{V}_{\lambda}=\bigoplus_{i}K\mathbb{Y}_{i},

where the sum runs over all the standard tableaux 𝕐i\mathbb{Y}_{i} of shape λ\lambda.

Let γ=(a,b,i)\gamma=(a,b,i) be a node in λ(i)\lambda^{(i)} for some ii. The content c(λ:γ)c(\lambda:\gamma) of γ\gamma in λ\lambda is defined to be the content of γ\gamma in the Young diagram λ(i)\lambda^{(i)}, namely, c(λ:γ)=c(λ(i):γ)=bac(\lambda:\gamma)=c(\lambda^{(i)}:\gamma)=b-a.

For any indeterminate xx and an integer kk, let Δ(k,x)\Delta(k,x) be the polynomial Δ(k,x)=1qkx\Delta(k,x)=1-q^{k}x in qq and xx with coefficients in KK and define M(k,x)M(k,x) as follows:

(2.12) M(k,x)=1Δ(k,x)(q1Δ(k+1,x)qΔ(k1,x)qk(1q)x).M(k,x)=\dfrac{1}{\Delta(k,x)}\left(\begin{matrix}&q-1&\Delta(k+1,x)\\ &q\Delta(k-1,x)&q^{k}(1-q)x\end{matrix}\right).

With the notations above, Ariki and Koike describe a complete set of the irreducible representations of H(r,1,n)H(r,1,n):

Theorem 2.7.

([4],Theorem 3.10)

Let KK be the quotient field of the base ring RR which is defined before Definition 2.6. H(K)=H(r,1,n)KH(K)=H(r,1,n)\bigotimes K is semisimple over KK. The set {𝕍λ|λn}\{\mathbb{V}_{\lambda}|\lambda\vdash n\} is a complete set of non-isomorphic irreducible H(K)H(K)-modules. For each multipartition λ\lambda and standard tableau 𝕐\mathbb{Y} of shape λ\lambda, H(r,1,n)H(r,1,n) acts on 𝕐𝕍λ\mathbb{Y}\in\mathbb{V}_{\lambda} as follows:

(1) The action of T0T_{0}.

(2.13) T0𝕐=vτ(1)𝕐T_{0}\mathbb{Y}=v_{\tau(1)}\mathbb{Y}

where the number 1 occurs in the τ(1)\tau(1)-th component, λ(τ(1))\lambda^{(\tau(1))}, of the tableau 𝕐\mathbb{Y}.

(2)The action of Ti(i=1,2,,n1)T_{i}(i=1,2,\dots,n-1).

Case (2.1). If ii and i+1i+1 appear in the same row of the same component of 𝕐\mathbb{Y}, then

(2.14) Ti𝕐=q𝕐.T_{i}\mathbb{Y}=q\mathbb{Y}.

Case (2.2). If ii and i+1i+1 appear in the same column of the same component of 𝕐\mathbb{Y}, then

(2.15) Ti𝕐=𝕐.T_{i}\mathbb{Y}=-\mathbb{Y}.

Case (2.3). ii and i+1i+1 appear neither in the same row nor in the same column of 𝕐\mathbb{Y}. Let 𝕐\mathbb{Y}^{\prime} be the standard tableau obtained by transposing the letters ii and i+1i+1 in 𝕐\mathbb{Y}. The action of TiT_{i} on the two-dimensional subspace spaned by 𝕐\mathbb{Y} and 𝕐\mathbb{Y}^{\prime} is given by

(2.16) [Ti𝕐,Ti𝕐]=[𝕐,𝕐]M(c(λ:i)c(λ:i+1),vτ(i)vτ(i+1)),[T_{i}\mathbb{Y},T_{i}\mathbb{Y}^{\prime}]=[\mathbb{Y},\mathbb{Y}^{\prime}]M(c(\lambda:i)-c(\lambda:i+1),\dfrac{v_{\tau(i)}}{v_{\tau(i+1)}}),

where M(c(λ:i)c(λ:i+1),vτ(i)vτ(i+1))M(c(\lambda:i)-c(\lambda:i+1),\dfrac{v_{\tau(i)}}{v_{\tau(i+1)}}) is the 2×22\times 2 matrix M(k,x)M(k,x) defined in (2.12) with k=c(λ:i)c(λ:i+1)k=c(\lambda:i)-c(\lambda:i+1) and x=vτ(i)vτ(i+1)x=\dfrac{v_{\tau(i)}}{v_{\tau(i+1)}} and the numbers ii and i+1i+1 occur in the τ(i)\tau(i)-th and τ(i+1)\tau(i+1)-th components of the tableau 𝕐\mathbb{Y}, respectively.

Further, if RR is specialised to a field FF, the following theorem states exactly when H(r,1,n)H(r,1,n) is semisimple:

Theorem 2.8.

([3], main theorem) Suppose RR is a field FF. Then the cyclotomic Hecke algebra H(r,1,n)H(r,1,n) is semisimple if and only if the parameters q,v1,v2,,vrq,v_{1},v_{2},\dots,v_{r} satisfy

qdvivjq^{d}v_{i}\neq v_{j}

for iji\neq j and dd\in\mathbb{Z} with |d|<n|d|<n and

1+q+q2++qj01+q+q^{2}+\dots+q^{j}\neq 0

for 1jn11\leq j\leq n-1.

In the case when the cyclotomic Hecke algebra H(r,1,n)H(r,1,n) is semisimple, each one-dimensional module corresponds to a central idempotent in H(r,1,n)H(r,1,n). We will use these idempotents to define our generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} in the next section.

2.4. The cyclotomic KLR algebras

To show the graded cellularity of our generalised Temperley-Lieb algebras, we shall make use of the cyclotomic KLR algebras, which are isomorphic to the cyclotomic Hecke algebras H(r,1,n)H(r,1,n). In the next section, we will use this isomorphism to give a second interpretation of our generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n}.

The KLR algebra was first introduced independently by Khovanov and Lauda in [20] and by Rouquier in [28] to categorify the quantum deformation of the universal enveloping algebra of the lower-triangular subalgebra of a Kac-Moody algebra 𝔤\mathfrak{g}. Brundan and Kleshchev have shown in [10] that the KLR algebras of type An1A_{n-1} are isomorphic to the cyclotomic Hecke algebras corresponding to the complex reflection groups G(r,1,n)G(r,1,n) or their rational degenerations. In this section, we recall the special cyclotomic KLR algebra which is isomorphic to H(r,1,n)H(r,1,n). We refer readers to [10] for a broader discussion.

Let FF be a fixed ground field and qFq\in F^{*}. Let ee be the smallest positive integer such that 1+q++qe1=01+q+\dots+q^{e-1}=0, setting e:=0e:=0 if no such integer exists. Let Γ\Gamma be the quiver with vertex set I:=/eI:=\mathbb{Z}/e\mathbb{Z}, and a directed edge from ii to jj if j=i+1j=i+1. Thus Γ\Gamma is the quiver of type AA_{\infty} if e=0e=0 or Ae1(1)A_{e-1}^{(1)}, with the following orientation:

A:{A_{\infty}:}{\dots}2{-2}1{-1}0{0}1{1}2{2}{\dots}Ae1(1):{A_{e-1}^{(1)}:}01{0\leftrightarrow 1}0{0}0{0}1{1}{\dots}1{1}2{2}3{3}2{2}

Let (cij)(c_{ij}) be the Cartan matrix associated with Γ\Gamma, so that

(2.17) cij={2 if i=j;0 if ij±1;1 if e2 and i=j±1;2 if e=2 and i=j±1.\displaystyle c_{ij}=\begin{cases}2&\textit{ if }i=j;\\ 0&\textit{ if }i\neq j\pm 1;\\ -1&\textit{ if }e\neq 2\textit{ and }i=j\pm 1;\\ -2&\textit{ if }e=2\textit{ and }i=j\pm 1.\end{cases}

Let {αi|iI}\{\alpha_{i}|i\in I\} be the corresponding set of simple roots and {Λi|iI}\{\Lambda_{i}|i\in I\} the fundamental weights. Let (,)(\;,\;) be the bilinear form determined by

(αi,αj)=cij and (Λi,αj)=δij.\displaystyle(\alpha_{i},\alpha_{j})=c_{ij}\textit{ and }(\Lambda_{i},\alpha_{j})=\delta_{ij}.

Let P+=iIΛiP_{+}=\oplus_{i\in I}\mathbb{N}\Lambda_{i} be the dominant weights in the weight lattice and Q+=iIαiQ_{+}=\oplus_{i\in I}\mathbb{N}\alpha_{i} be the positive roots. For ΛP+\Lambda\in P_{+}, define the length of Λ\Lambda, l(Λ):=iI(Λ,αi)l(\Lambda):=\sum_{i\in I}(\Lambda,\alpha_{i}) and for αQ+\alpha\in Q_{+}, let the height of α\alpha be ht(α):=iI(α,Λi)ht(\alpha):=\sum_{i\in I}(\alpha,\Lambda_{i}). Fix an αQ+\alpha\in Q_{+}, define IαI^{\alpha} as follows:

(2.18) Iα:={iIht(α)|α=αi1+αi2++αiht(α)}.I^{\alpha}:=\{i\in I^{ht(\alpha)}|\alpha=\alpha_{i_{1}}+\alpha_{i_{2}}+\dots+\alpha_{i_{ht(\alpha)}}\}.

Fixing a dominant weight Λ\Lambda, we next recall the definition of the cyclotomic KLR algebra of type An1A_{n-1} associated with the weight Λ\Lambda. The following definition is originally from the work of Khovanov and Lauda in [20] and Rouquier in [28].

Definition 2.9.

The cyclotomic KLR algebra of type An1A_{n-1} associated with the weight Λ\Lambda is defined as nΛ:=ht(α)=nαΛ\mathcal{R}_{n}^{\Lambda}:=\oplus_{ht(\alpha)=n}\mathcal{R}_{\alpha}^{\Lambda}, where αΛ\mathcal{R}_{\alpha}^{\Lambda} is the unital associative FF-algebra with generators

(2.19) {ψ1,,ψn1}{y1,y2,,yn}{e(i)|iIα}\{\psi_{1},\dots,\psi_{n-1}\}\cup\{y_{1},y_{2},\dots,y_{n}\}\cup\{e(i)|i\in I^{\alpha}\}

subject to the following relations

(2.20) y1(Λ,αi1)e(i)=0,y_{1}^{(\Lambda,\alpha_{i_{1}})}e(i)=0,
(2.21) e(i)e(j)=δije(i),e(i)e(j)=\delta_{ij}e(i),
(2.22) iIαe(i)=1,\sum_{i\in I^{\alpha}}e(i)=1,
(2.23) e(i)yr=yre(i),e(i)y_{r}=y_{r}e(i),
(2.24) ψre(i)=e(sri)ψr,\psi_{r}e(i)=e(s_{r}\circ i)\psi_{r},
(2.25) yrys=ysyr,y_{r}y_{s}=y_{s}y_{r},
(2.26) ψrys=ysψr if sr,r+1;\psi_{r}y_{s}=y_{s}\psi_{r}\textit{ if }s\neq r,r+1;
(2.27) ψrψs=ψsψr if |rs|1;\psi_{r}\psi_{s}=\psi_{s}\psi_{r}\textit{ if }|r-s|\neq 1;
(2.28) ψryr+1e(i)={yrψre(i)+e(i) if ir=ir+1;yrψre(i) if irir+1;\displaystyle\psi_{r}y_{r+1}e(i)=\begin{cases}y_{r}\psi_{r}e(i)+e(i)&\textit{ if }i_{r}=i_{r+1};\\ y_{r}\psi_{r}e(i)&\textit{ if }i_{r}\neq i_{r+1};\end{cases}
(2.29) yr+1ψre(i)={ψryre(i)+e(i) if ir=ir+1;ψryre(i) if irir+1;\displaystyle y_{r+1}\psi_{r}e(i)=\begin{cases}\psi_{r}y_{r}e(i)+e(i)&\textit{ if }i_{r}=i_{r+1};\\ \psi_{r}y_{r}e(i)&\textit{ if }i_{r}\neq i_{r+1};\end{cases}
(2.30) ψr2e(i)={0 if ir=ir+1;e(i) if irir+1±1;(yr+1yr)e(i) if irir+1;(yryr+1)e(i) if irir+1;(yr+1yr)(yryr+1)e(i) if irir+1.\displaystyle\psi_{r}^{2}e(i)=\begin{cases}0&\textit{ if }i_{r}=i_{r+1};\\ e(i)&\textit{ if }i_{r}\neq i_{r+1}\pm 1;\\ (y_{r+1}-y_{r})e(i)&\textit{ if }i_{r}\rightarrow i_{r+1};\\ (y_{r}-y_{r+1})e(i)&\textit{ if }i_{r}\leftarrow i_{r+1};\\ (y_{r+1}-y_{r})(y_{r}-y_{r+1})e(i)&\textit{ if }i_{r}\leftrightarrow i_{r+1}.\end{cases}
(2.31) (ψrψr+1ψrψr+1ψrψr+1)e(i)={e(i) if ir+2=irir+1;e(i) if ir+2=irir+1;(yr2yr+1+yr+2)e(i) if ir+2=irir+1;0 otherwise .\displaystyle(\psi_{r}\psi_{r+1}\psi_{r}-\psi_{r+1}\psi_{r}\psi_{r+1})e(i)=\begin{cases}e(i)&\textit{ if }i_{r+2}=i_{r}\rightarrow i_{r+1};\\ -e(i)&\textit{ if }i_{r+2}=i_{r}\leftarrow i_{r+1};\\ (y_{r}-2y_{r+1}+y_{r+2})e(i)&\textit{ if }i_{r+2}=i_{r}\leftrightarrow i_{r+1};\\ 0&\textit{ otherwise }.\end{cases}

where sris_{r}\circ i in (2.24) is the natural action of 𝔖n\mathfrak{S}_{n} on InI^{n} and the arrows in (2.30) and (2.31) are those in the quiver Γ\Gamma.

Note that all the relations above are homogeneous with respect to the following degree function on the generators:

(2.32) deg e(i)=0,deg yr=2 and deg ψse(i)=cis,is+1\text{deg }e(i)=0,\text{deg }y_{r}=2\text{ and deg }\psi_{s}e(i)=-c_{i_{s},i_{s+1}}

where iIni\in I^{n},1rn1\leq r\leq n and 1sn11\leq s\leq n-1. Therefore, the cyclotomic KLR algebra defined above is \mathbb{Z}-graded with respect to the degree function in (2.32).

Let n\mathcal{H}_{n} be the affine Hecke algebra over FF corresponding to 𝔖n\mathfrak{S}_{n}, that is the FF-algebra generated by T1,T2,,Tn1,X1±1,,Xn±1T_{1},T_{2},\dots,T_{n-1},X_{1}^{\pm 1},\dots,X_{n}^{\pm 1} subject to the following relations:

(2.33) XrXs=XsXr,qXr+1=TrXrTr,X_{r}X_{s}=X_{s}X_{r},qX_{r+1}=T_{r}X_{r}T_{r},
(2.34) TrXs=XsTr if sr,r+1,T_{r}X_{s}=X_{s}T_{r}\text{ if }s\neq r,r+1,
(2.35) (Trq)(Tr+1)=0,TrTr+1Tr=Tr+1TrTr+1,(T_{r}-q)(T_{r}+1)=0,T_{r}T_{r+1}T_{r}=T_{r+1}T_{r}T_{r+1},
(2.36) TrTs=TsTr if |rs|>1.T_{r}T_{s}=T_{s}T_{r}\text{ if }|r-s|>1.

Let nΛ\mathcal{H}_{n}^{\Lambda} be the cyclotomic quotient of n\mathcal{H}_{n} corresponding to the weight Λ\Lambda, that is

(2.37) nΛ:=n/iI(X1qi)(Λ,αi).\mathcal{H}_{n}^{\Lambda}:=\mathcal{H}_{n}/\langle\prod_{i\in I}(X_{1}-q^{i})^{(\Lambda,\alpha_{i})}\rangle.

Suppose that MM is a finite dimensional nΛ\mathcal{H}_{n}^{\Lambda}-module. Then MM may be regarded as a \mathcal{H} module, and by [15, Lemma 4.7], when q1q\neq 1, the eigenvalues of XrX_{r} (for all rr) are of form qiq^{i} where iIi\in I. That implies MM decomposes as a direct sum of the joint generalised eigenspaces for X1,X2,,XnX_{1},X_{2},\dots,X_{n}, that is M=iInMiM=\oplus_{i\in I^{n}}M_{i} with

(2.38) Mi:={vM|(Xrqir)dim(M)v=0, for 1rn}.M_{i}:=\{v\in M|(X_{r}-q^{i_{r}})^{dim(M)}v=0,\text{ for }1\leq r\leq n\}.

In particular, let MM be the regular nΛ\mathcal{H}_{n}^{\Lambda} module. Then there is a set of pairwise orthogonal idempotents {e(i)|iIn}\{e(i)|i\in I^{n}\} in nΛ\mathcal{H}_{n}^{\Lambda} such that Mi=e(i)MM_{i}=e(i)M. All but finitely many of the e(i)e(i) are zero and their sum is the identity in nΛ\mathcal{H}_{n}^{\Lambda}. Fix an αQ+\alpha\in Q_{+}, let

(2.39) eα:=iIαe(i)nΛ.e_{\alpha}:=\sum_{i\in I^{\alpha}}e(i)\in\mathcal{H}_{n}^{\Lambda}.

As a consequence of [21] or [9, Theorem 1], eαe_{\alpha} is either zero or a primitive central idempotent in nΛ\mathcal{H}_{n}^{\Lambda}. Let

(2.40) αΛ:=eαnΛ.\mathcal{H}_{\alpha}^{\Lambda}:=e_{\alpha}\mathcal{H}_{n}^{\Lambda}.

Then αΛ\mathcal{H}_{\alpha}^{\Lambda} is either zero or an indecomposable two-sided ideal of nΛ\mathcal{H}_{n}^{\Lambda}.

We will next introduce the isomorphism between αΛ\mathcal{R}_{\alpha}^{\Lambda} and αΛ\mathcal{H}_{\alpha}^{\Lambda} constructed by Brundan and Kleshchev in [10] by defining elements in αΛ\mathcal{H}_{\alpha}^{\Lambda} which satisfy the relations in αΛ\mathcal{R}_{\alpha}^{\Lambda}. Let e(i)e(i) be the idempotent defined above. For 1rn1\leq r\leq n, define

(2.41) yr:=iIα(1qirXr)e(i).y_{r}:=\sum_{i\in I^{\alpha}}(1-q^{-i_{r}}X_{r})e(i).

And for 1rn1\leq r\leq n and iIαi\in I^{\alpha} set

(2.42) yr(i):=qir(1yr)F[y1,y2,,yn].y_{r}(i):=q^{i_{r}}(1-y_{r})\in F[y_{1},y_{2},\dots,y_{n}].

Define power series Pr(i),Qr(i)F[[yr,yr1]]P_{r}(i),Q_{r}(i)\in F[[y_{r},y_{r-1}]] as follows:

(2.43) Pr(i):={1 if ir=ir+1;(1q)(1yr(i)yr+1(i)1)1 if irir+1.\displaystyle P_{r}(i):=\begin{cases}1&\textit{ if }i_{r}=i_{r+1};\\ (1-q)(1-y_{r}(i)y_{r+1}(i)^{-1})^{-1}&\textit{ if }i_{r}\neq i_{r+1}.\end{cases}
(2.44) Qr(i):={1q+qyr+1yr if ir=ir+1;(yr(i)qyr+1(i))/(yr(i)yr+1(i)) if irir+1±1;(yr(i)qyr+1(i))/(yr(i)yr+1(i))2 if irir+1;qir if irir+1;qir/(yr(i)yr+1(i)) if irir+1.\displaystyle Q_{r}(i):=\begin{cases}1-q+qy_{r+1}-y_{r}&\text{ if }i_{r}=i_{r+1};\\ (y_{r}(i)-qy_{r+1}(i))/(y_{r}(i)-y_{r+1}(i))&\text{ if }i_{r}\neq i_{r+1}\pm 1;\\ (y_{r}(i)-qy_{r+1}(i))/(y_{r}(i)-y_{r+1}(i))^{2}&\text{ if }i_{r}\rightarrow i_{r+1};\\ q^{i_{r}}&\text{ if }i_{r}\leftarrow i_{r+1};\\ q^{i_{r}}/(y_{r}(i)-y_{r+1}(i))&\text{ if }i_{r}\leftrightarrow i_{r+1}.\end{cases}

Then for 1rn11\leq r\leq n-1, set

(2.45) ψr:=iIα(Tr+Pr(i))Qr(i)1e(i).\psi_{r}:=\sum_{i\in I^{\alpha}}(T_{r}+P_{r}(i))Q_{r}(i)^{-1}e(i).

We do not distinguish here between the generators of the cyclotomic KLR algebra and the corresponding elements in the cyclotomic Hecke algebra because of the following isomorphism theorem by Brundan and Kleshchev:

Theorem 2.10.

([10, Theorem 1.1]) Let α\alpha be an element in the positive root lattice Q+Q_{+}, such that the two-sided ideal αΛ0\mathcal{H}_{\alpha}^{\Lambda}\neq 0. The map f:αΛαΛf:\mathcal{R}_{\alpha}^{\Lambda}\to\mathcal{H}_{\alpha}^{\Lambda} such that

f(e(i))=e(i),f(yr)=yr and f(ψs)=ψsf(e(i))=e(i),f(y_{r})=y_{r}\text{ and }f(\psi_{s})=\psi_{s}

extends uniquely as an algebra isomorphism.

We identify αΛ\mathcal{R}_{\alpha}^{\Lambda} and αΛ\mathcal{H}_{\alpha}^{\Lambda} via ff in the rest of this paper. Moreover, the inverse of ff is given by

(2.46) f1(Xr)=iIαyr(i)e(i)f^{-1}(X_{r})=\sum_{i\in I^{\alpha}}y_{r}(i)e(i)

and

(2.47) f1(Tr)=iIα(ψrQr(i)Pr(i))e(i).f^{-1}(T_{r})=\sum_{i\in I^{\alpha}}(\psi_{r}Q_{r}(i)-P_{r}(i))e(i).

The cyclotomic Hecke algebra nΛ\mathcal{H}_{n}^{\Lambda} is naturally a subalgebra of n+1Λ\mathcal{H}_{n+1}^{\Lambda} and n+1Λ\mathcal{H}_{n+1}^{\Lambda} can also be regarded as a free nΛ\mathcal{H}_{n}^{\Lambda}-module. This implies that we can define the usual restriction and induction functors between Rep(nΛ)Rep(\mathcal{R}_{n}^{\Lambda}) and Rep(n+1Λ)Rep(\mathcal{R}_{n+1}^{\Lambda}). These functors can be decomposed into ii-restrictions ResiΛRes_{i}^{\Lambda} and ii-inductions IndiΛInd_{i}^{\Lambda}, for iIi\in I, by projecting onto the blocks αΛ\mathcal{R}_{\alpha}^{\Lambda} and α+αiΛ\mathcal{R}_{\alpha+\alpha_{i}}^{\Lambda} where ht(α)=nht(\alpha)=n. For each iIi\in I, define

(2.48) en,i:=jIne(j,i).e_{n,i}:=\sum_{j\in I^{n}}e(j,i).

Let fi:nΛn+1Λf_{i}:\mathcal{R}_{n}^{\Lambda}\to\mathcal{R}_{n+1}^{\Lambda} be the embedding map given by:

(2.49) fi(e(j))=e(j,i),fi(yr)=en,iyr and fi(ψs)=en,iψs,f_{i}(e(j))=e(j,i),f_{i}(y_{r})=e_{n,i}y_{r}\text{ and }f_{i}(\psi_{s})=e_{n,i}\psi_{s},

where jInj\in I^{n}, 1rn1\leq r\leq n and 1sn11\leq s\leq n-1. It should be observed that this embedding is not unital. The restriction and induction functors induced by fif_{i} are:

(2.50) ResiΛ:Rep(n+1Λ)\displaystyle Res_{i}^{\Lambda}:Rep(\mathcal{R}_{n+1}^{\Lambda}) Rep(nΛ)\displaystyle\mapsto Rep(\mathcal{R}_{n}^{\Lambda})
N\displaystyle N Nen,i\displaystyle\mapsto Ne_{n,i}
IndiΛ:Rep(nΛ)\displaystyle Ind_{i}^{\Lambda}:Rep(\mathcal{R}_{n}^{\Lambda}) Rep(n+1Λ)\displaystyle\mapsto Rep(\mathcal{R}_{n+1}^{\Lambda})
M\displaystyle M MnΛen,in+1Λ.\displaystyle\mapsto M\otimes_{\mathcal{R}_{n}^{\Lambda}}e_{n,i}\mathcal{R}_{n+1}^{\Lambda}.

The following lemma indicates how to extend a certain ideal n\mathcal{I}_{n} of nΛ\mathcal{R}_{n}^{\Lambda} to an ideal of n+1Λ\mathcal{R}_{n+1}^{\Lambda}.

Lemma 2.11.

Let αQ+\alpha\in Q^{+} be an element of height nn and n=αΛ\mathcal{I}_{n}=\mathcal{R}_{\alpha}^{\Lambda} be an indecomposable two-sided ideal of nΛ\mathcal{R}_{n}^{\Lambda}. Denote by n+1\mathcal{I}_{n+1} the two-sided ideal of n+1Λ\mathcal{R}_{n+1}^{\Lambda} generated by n\mathcal{I}_{n}. Then we have

n+1=𝐢Iα,iIe(𝐢,i)n+1Λ.\mathcal{I}_{n+1}=\langle\sum_{\mathbf{i}\in I^{\alpha},i\in I}e(\mathbf{i},i)\rangle_{\mathcal{H}_{n+1}^{\Lambda}}.

Proof.

By Theorem 2.10, αΛαΛ\mathcal{R}_{\alpha}^{\Lambda}\cong\mathcal{H}_{\alpha}^{\Lambda}. We can identify αΛ\mathcal{R}_{\alpha}^{\Lambda} and αΛ\mathcal{H}_{\alpha}^{\Lambda}. For 𝐢In\mathbf{i}\in I^{n} (𝐣In+1\mathbf{j}\in I^{n+1}), let e(𝐢)e(\mathbf{i}) (e(𝐣)e(\mathbf{j})) be the idempotent in nΛ\mathcal{H}_{n}^{\Lambda} (n+1Λ\mathcal{H}_{n+1}^{\Lambda}) which corresponds to the generalised eigenspaces M𝐢M_{\mathbf{i}} (N𝐣N_{\mathbf{j}}) in (2.38) of the regular left module. We first prove the following equation:

(2.51) e(𝐢)e(𝐣)={e(𝐣) if 𝐣=(𝐢,i) for some iI;0otherwise.e(\mathbf{i})e(\mathbf{j})=\begin{cases}e(\mathbf{j})&\textit{ if }\mathbf{j}=(\mathbf{i},i)\textit{ for some }i\in I;\\ 0&\textit{otherwise}.\end{cases}

Let 𝐣=(j1,j2,,jn+1)\mathbf{j}=(j_{1},j_{2},\dots,j_{n+1}) and 𝐣(n)=(j1,j2,,jn)\mathbf{j}^{(n)}=(j_{1},j_{2},\dots,j_{n}). By the definition of e(𝐣)e(\mathbf{j}), (Xrqjr)dim(n+1Λ)e(𝐣)=0, for 1rn+1(X_{r}-q^{j_{r}})^{dim(\mathcal{H}_{n+1}^{\Lambda})}e(\mathbf{j})=0,\text{ for }1\leq r\leq n+1. Regard n+1Λ\mathcal{H}_{n+1}^{\Lambda} as a left nΛ\mathcal{H}_{n}^{\Lambda}-module, M(n+1)M^{(n+1)}. Then e(𝐣)M𝐣(n)(n+1)e(\mathbf{j})\in M_{\mathbf{j}^{(n)}}^{(n+1)}. By definition, M𝐣(n)(n+1)=e(𝐣(n))M(n+1)M_{\mathbf{j}^{(n)}}^{(n+1)}=e({\mathbf{j}^{(n)}})M^{(n+1)}. So e(𝐣)=e(𝐣(n))he(\mathbf{j})=e({\mathbf{j}^{(n)}})h for some hn+1Λh\in\mathcal{H}_{n+1}^{\Lambda}. As e(𝐢)e(\mathbf{i})’s are pairwise orthogonal idempotents, e(𝐢)e(𝐣(n))e(\mathbf{i})e({\mathbf{j}^{(n)}}) is e(𝐣(n))e(\mathbf{j}^{(n)}) (when 𝐢=𝐣(n)\mathbf{i}=\mathbf{j}^{(n)}) or zero, which implies (2.51).

Let eαnΛe_{\alpha}\in\mathcal{H}_{n}^{\Lambda} be the primitive central idempotent such that αΛ=eαnΛ\mathcal{H}_{\alpha}^{\Lambda}=e_{\alpha}\mathcal{H}_{n}^{\Lambda}. Then we have

n+1\displaystyle\mathcal{I}_{n+1} =nn+1Λ\displaystyle=\langle\mathcal{I}_{n}\rangle_{\mathcal{R}_{n+1}^{\Lambda}}
=eαnΛn+1Λ\displaystyle=\langle e_{\alpha}\mathcal{H}_{n}^{\Lambda}\rangle_{\mathcal{H}_{n+1}^{\Lambda}}
=eαn+1Λ\displaystyle=\langle e_{\alpha}\rangle_{\mathcal{H}_{n+1}^{\Lambda}}
=𝐢Iαe(𝐢)𝐣In+1e(𝐣)n+1Λ\displaystyle=\langle\sum_{\mathbf{i}\in I^{\alpha}}e(\mathbf{i})\sum_{\mathbf{j}\in I^{n+1}}e(\mathbf{j})\rangle_{\mathcal{H}_{n+1}^{\Lambda}}
=𝐢Iα,iIe(𝐢,i)n+1Λ.\displaystyle=\langle\sum_{\mathbf{i}\in I^{\alpha},i\in I}e(\mathbf{i},i)\rangle_{\mathcal{H}_{n+1}^{\Lambda}}.

3. Generalised Temperley-Lieb algebras TLr,1,nTL_{r,1,n}

In this section, we define our generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} as a quotient of the corresponding cyclotomic Hecke algebra H(r,1,n)H(r,1,n). We will show that the Temperley-Lieb algebras of types An1A_{n-1} and BnB_{n} are both special cases of TLr,1,nTL_{r,1,n}. We will also give an interpretation of our generalised Temperley-Lieb algebra with respect to the KLR generators. This interpretation will be used to construct the graded cellular structure of TLr,1,nTL_{r,1,n} in the next section. In what follows, we require RR to be a field of characteristic 0 and q,v1,v2,,vrRq,v_{1},v_{2},\dots,v_{r}\in R^{*} such that

(3.1) vivj1,q or q2\frac{v_{i}}{v_{j}}\neq 1,q\textbf{ or }q^{2}

for all iji\neq j and

(3.2) (1+q)(1+q+q2)0.(1+q)(1+q+q^{2})\neq 0.

These restrictions on q,v1,v2,,vrRq,v_{1},v_{2},\dots,v_{r}\in R^{*} are equivalent to the semi-simplicity of the parabolic subalgebras Hi,i+1H_{i,i+1} mentioned in the introduction by Theorem 2.8.

3.1. The definition of TLr,1,nTL_{r,1,n} as a quotient of the cyclotomic Hecke algebra

Let H(r,1,n)H(r,1,n) be the cyclotomic Hecke algebra defined in section 2.3 and Hi,i+1H_{i,i+1} be the subalgebra of H(r,1,n)H(r,1,n) which is generated by two non-commuting generators, TiT_{i} and Ti+1T_{i+1} where 0in20\leq i\leq n-2. We call Hi,i+1H_{i,i+1} a parabolic subalgebra of H(r,1,n)H(r,1,n). There are two types of parabolic subalgebras:

Case 1. i=0i=0. H0,1H_{0,1} is a cyclotomic Hecke algebra corresponding to G(r,1,2)G(r,1,2). By Theorem 2.8, it is semisimple given the restriction on parameters in (3.1)(\ref{pmres}) and (3.2)(\ref{pmres1}). It has 2r2r 1-dimensional representations corresponding to the multipartitions (0,0,,(2),,0)(0,0,\dots,(2),\dots,0) and (0,0,,(1,1),,0)(0,0,\dots,(1,1),\dots,0) where (2)(2) and (1,1)(1,1) are in the jthj^{th} component of the rr-partition with 1jr1\leq j\leq r. Denote by E0(j)E_{0}^{(j)} and F0(j)F_{0}^{(j)} the corresponding primitive central idempotents in H0,1H_{0,1}.

Case 2. 1in21\leq i\leq n-2. Hi,i+1H_{i,i+1} is a Hecke algebra of type A2A_{2}. It has 2 1-dimensional representations and the corresponding primitive central idempotents are

Ei\displaystyle E_{i} =a(TiTi+1Ti+TiTi+1+Ti+1Ti+Ti+Ti+1+1);\displaystyle=a(T_{i}T_{i+1}T_{i}+T_{i}T_{i+1}+T_{i+1}T_{i}+T_{i}+T_{i+1}+1);
Fi\displaystyle F_{i} =b(TiTi+1TiqTiTi+1qTi+1Ti+q2Ti+q2Ti+1q3),\displaystyle=b(T_{i}T_{i+1}T_{i}-qT_{i}T_{i+1}-qT_{i+1}T_{i}+q^{2}T_{i}+q^{2}T_{i+1}-q^{3}),

where a=(q3+2q2+2q+1)1a=(q^{3}+2q^{2}+2q+1)^{-1} and b=ab=-a. As a generalisation of the original Temperley-Lieb algebra, TLr,1,nTL_{r,1,n} is defined as the quotient of H(r,1,n)H(r,1,n) by the two-sided ideal generated by half of the central idempotents listed above, that is:

Definition 3.1.

Let H(r,1,n)H(r,1,n) be the cyclotomic Hecke algebra defined in 2.6 where RR is a field of characteristic 0 and q,v1,v2,,vrRq,v_{1},v_{2},\dots,v_{r}\in R^{*} satisfy (3.1) and (3.2). Let Hi,i+1H_{i,i+1} be the parabolic subalgebra generated by TiT_{i} and Ti+1T_{i+1} and EiE_{i}( Ei(j)E_{i}^{(j)}, if i=0i=0) be the primitive central idempotents of Hi,i+1H_{i,i+1} listed above. The generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} is

TLr,1,n:=H(r,1,n)/E0(1),,E0(r),E1,,En2.TL_{r,1,n}:=H(r,1,n)/\langle E_{0}^{(1)},\dots,E_{0}^{(r)},E_{1},\dots,E_{n-2}\rangle.

It should be remarked that the generalised Temperley-Lieb algebra depends heavily on the parameters q,v1,v2,,vrq,v_{1},v_{2},\dots,v_{r}. To make the notations simpler, we omit the parameters without causing confusion. The next two examples show that the usual Temperley-Lieb algebra TLn(q)TL_{n}(q) and the blob algebra TLBn(q,Q)TLB_{n}(q,Q) are special cases of TLr,1,nTL_{r,1,n} where r=1,2r=1,2.

Example 3.2.

r=1r=1. In this case, T0T_{0} is an element in RR, thus T0T_{0} and T1T_{1} commutative. So H0,1H_{0,1} is no longer a parabolic subalgebra in this case. Therefore, TL1,n=Hn1/E1,,En2=Hn1/E1=TLn1(q)TL_{1,n}=H_{n-1}/\langle E_{1},\dots,E_{n-2}\rangle=H_{n-1}/\langle E_{1}\rangle=TL_{n-1}(q).

Example 3.3.

r=2r=2. Let v1=Qv_{1}=Q and v2=Q1v_{2}=-Q^{-1}. We have H(2,1,n)=HBn(Q,q)H(2,1,n)=HB_{n}(Q,q). We next show TL2,n=TLB(Q,q)TL_{2,n}=TLB(Q,q) when 1+Q2R1+Q^{2}\in R^{*}. In Lemma 5.1 of [14], Graham and Lehrer point out that C0X=XC0=u2Q2E0(1)C_{0}X=XC_{0}=u^{-2}Q^{-2}E_{0}^{(1)} which implies E0(1)XE_{0}^{(1)}\in\langle X\rangle as u,Qu,Q are invertible. We observe that

E0(1)Q4E0(2)\displaystyle E_{0}^{(1)}-Q^{4}E_{0}^{(2)}
=\displaystyle= (u2Q2u2Q2)T2T0T2T0+(u2Q+u2Q3)T2T0T2+(uQ2uQ2)T0T2T0\displaystyle(u^{2}Q^{2}-u^{2}Q^{2})T_{2}T_{0}T_{2}T_{0}+(u^{2}Q+u^{2}Q^{3})T_{2}T_{0}T_{2}+(uQ^{2}-uQ^{2})T_{0}T_{2}T_{0}
+(uQ+uQ3)(T0T2+T2T0)+(Q+Q3)T0+(uuQ4)T2+1Q4\displaystyle+(uQ+uQ^{3})(T_{0}T_{2}+T_{2}T_{0})+(Q+Q^{3})T_{0}+(u-uQ^{4})T_{2}+1-Q^{4}
=\displaystyle= (u2Q+u2Q3)T2T0T2+(uQ+uQ3)(T0T2+T2T0)+(Q+Q3)T0\displaystyle(u^{2}Q+u^{2}Q^{3})T_{2}T_{0}T_{2}+(uQ+uQ^{3})(T_{0}T_{2}+T_{2}T_{0})+(Q+Q^{3})T_{0}
+(uuQ4)T2+1Q4\displaystyle+(u-uQ^{4})T_{2}+1-Q^{4}
=\displaystyle= (1+Q2)(u2QT2T0T2+uQT0T2+uQT2T0+QT0+u(1Q2)T2+1Q2)\displaystyle(1+Q^{2})(u^{2}QT_{2}T_{0}T_{2}+uQT_{0}T_{2}+uQT_{2}T_{0}+QT_{0}+u(1-Q^{2})T_{2}+1-Q^{2})
=\displaystyle= (1+Q2)(uT2+1)(QT0+1)(uT2+1)(1+Q2)(u2+u3T2+uQ2T2+Q2)\displaystyle(1+Q^{2})(uT_{2}+1)(QT_{0}+1)(uT_{2}+1)-(1+Q^{2})(u^{2}+u^{3}T_{2}+uQ^{2}T_{2}+Q^{2})
=\displaystyle= u2Q(1+Q2)C2C0C2+u2Q(1+Q2)κC1\displaystyle-u^{2}Q(1+Q^{2})C_{2}C_{0}C_{2}+u^{2}Q(1+Q^{2})\kappa C_{1}
=\displaystyle= u2Q2(1+Q2)X.\displaystyle-u^{2}Q^{2}(1+Q^{2})X.

As Q,uQ,u are invertible in RR, we have:

(3.3) (1+Q2)XE0(1),E0(2).\langle(1+Q^{2})X\rangle\subseteq\langle E_{0}^{(1)},E_{0}^{(2)}\rangle.

On the other hand, the equation above implies

Q4E0(2)=E0(1)+u2Q2(1+Q2)X.Q^{4}E_{0}^{(2)}=E_{0}^{(1)}+u^{2}Q^{2}(1+Q^{2})X.

So we have

E0(1),E0(2)E0(1),(1+Q2)XX.\langle E_{0}^{(1)},E_{0}^{(2)}\rangle\subseteq\langle E_{0}^{(1)},(1+Q^{2})X\rangle\subseteq\langle X\rangle.

Since 1+Q2R1+Q^{2}\in R^{*}, we have X=E0(1),E0(2)\langle X\rangle=\langle E_{0}^{(1)},E_{0}^{(2)}\rangle. So we have

TL2,n\displaystyle TL_{2,n} =H(2,1,n)/E0(1),E0(2),E1,,En2\displaystyle=H(2,1,n)/\langle E_{0}^{(1)},E_{0}^{(2)},E_{1},\dots,E_{n-2}\rangle
=HBn(Q,q)/E0(1),E0(2),E1\displaystyle=HB_{n}(Q,q)/\langle E_{0}^{(1)},E_{0}^{(2)},E_{1}\rangle
=HBn(Q,q)/X,E1\displaystyle=HB_{n}(Q,q)/\langle X,E_{1}\rangle
=TLBn(Q,q).\displaystyle=TLB_{n}(Q,q).

We next show that the generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} is a quotient of the corresponding generalised blob algebra B(r,n)B(r,n). Let 1,n\mathcal{I}_{1,n} be the two-sided ideal of H(r,1,n)H(r,1,n) generated by E0(i)E_{0}^{(i)} for all i=1,2,,ri=1,2,\dots,r. The generalised blob algebra was originally defined by Martin and Woodcock in [24] as follows:

Definition 3.4.

([24], Section 3) Let 1,n\mathcal{I}_{1,n} be the two-sided ideal of H(r,1,n)H(r,1,n) generated by E0(i)E_{0}^{(i)} for all i=1,2,,ri=1,2,\dots,r. The generalised blob algebra Br,nB_{r,n} is defined as the quotient of H(r,1,n)H(r,1,n) by 1,n\mathcal{I}_{1,n}, that is

(3.4) Br,n=H(r,1,n)/1,n.B_{r,n}=H(r,1,n)/\mathcal{I}_{1,n}.

Denote by 2,n\mathcal{I}_{2,n} the ideal of Br,nB_{r,n} generated by the representatives of E1,,En2E_{1},\dots,E_{n-2}. The following lemma is a direct consequence of definitions:

Lemma 3.5.

TLr,1,nBr,n/2,nTL_{r,1,n}\cong B_{r,n}/\mathcal{I}_{2,n}.

3.2. Realisation of TLr,1,nTL_{r,1,n} as a quotient of the cyclotomic KLR algebra

We next give an interpretation of the generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} as a quotient of the cyclotomic KLR algebra nΛ\mathcal{R}_{n}^{\Lambda}. This interpretation will be used to construct our graded cellular basis of TLr,1,nTL_{r,1,n} in the next section.

In similar fashion to the generalised blob algebra Br,nB_{r,n}, TLr,1,nTL_{r,1,n} is defined with respect to a dominant weight Λ\Lambda. We will use a specific notation TLr,1,nΛTL_{r,1,n}^{\Lambda} to indicate which dominant weight we are using if necessary. By definition, the condition (3.1) can be transformed into the following restriction:

(3.5) Λ=Λi1+Λi2++ΛirP+ where |iaib|>2 for ab.\Lambda=\Lambda_{i_{1}}+\Lambda_{i_{2}}+\dots+\Lambda_{i_{r}}\in P^{+}\textit{ where }|i_{a}-i_{b}|>2\textit{ for }a\neq b.

When n=3n=3, H(r,1,3)H(r,1,3) is semisimple by Theorem 2.8. The following lemma gives a realisation of TLr,1,3TL_{r,1,3} as a quotient of the cyclotomic KLR algebra 3Λ\mathcal{R}_{3}^{\Lambda}, which is the cornerstone of the interpretation of TLr,1,nTL_{r,1,n} into the language of KLR generators.

Lemma 3.6.

Let TLr,1,3TL_{r,1,3} be the generalised Temperley-Lieb algebra defined in Definition 3.1 and 3Λ\mathcal{R}_{3}^{\Lambda} be the cyclotomic KLR algebra isomorphic to H(r,1,3)H(r,1,3) indicated in Theorem 2.10. Then

(3.6) TLr,1,33Λ/𝒥3TL_{r,1,3}\cong\mathcal{R}_{3}^{\Lambda}/\mathcal{J}_{3}

where

(3.7) 𝒥3=\displaystyle\mathcal{J}_{3}= jI,(αi,Λ)>0e(i,i+1,j)3Λ\displaystyle\sum_{j\in I,(\alpha_{i},\Lambda)>0}\langle e(i,i+1,j)\rangle_{\mathcal{R}_{3}^{\Lambda}}
+(αij,Λ)>0 for j=1,2,3e(i1,i2,i3)3Λ.\displaystyle+\sum_{(\alpha_{i_{j}},\Lambda)>0\text{ for }j=1,2,3}\langle e(i_{1},i_{2},i_{3})\rangle_{\mathcal{R}_{3}^{\Lambda}}.
Proof.

This can be proved by direct calculation via (2.47) in the isomorphism by Brundan and Kleshchev. We give an alternative proof using the cellular structure of the generalised blob algebra Br,3B_{r,3}. By Theorem 3.8, we have

(3.8) Br,33Λ/jI,(αi,Λ)>0e(i,i+1,j)3Λ.B_{r,3}\cong\mathcal{R}_{3}^{\Lambda}/\sum_{j\in I,(\alpha_{i},\Lambda)>0}\langle e(i,i+1,j)\rangle_{\mathcal{R}_{3}^{\Lambda}}.

Comparing with (3.6), it is enough to prove

(3.9) TLr,1,3Br,3/(αij,Λ)>0 for j=1,2,3e(i1,i2,i3)Br,3.TL_{r,1,3}\cong B_{r,3}/\sum_{(\alpha_{i_{j}},\Lambda)>0\text{ for }j=1,2,3}\langle e(i_{1},i_{2},i_{3})\rangle_{B_{r,3}}.

Let 𝔅3(r)\mathfrak{B}_{3}^{(r)} be the poset of one-column rr-partitions of 33 and 𝔅3(r)(i),i=1,2,3\mathfrak{B}_{3}^{(r)}(i),i=1,2,3 be the subset consisting of the multipartitions with exactly ii non-empty partitions. For λμ𝔅3(r)\lambda\neq\mu\in\mathfrak{B}_{3}^{(r)} and tStd(λ),sStd(μ)t\in Std(\lambda),s\in Std(\mu), the restriction that Λ=Λi1+Λi2++ΛirP+\Lambda=\Lambda_{i_{1}}+\Lambda_{i_{2}}+\dots+\Lambda_{i_{r}}\in P^{+} where |iaib|>2|i_{a}-i_{b}|>2 for aba\neq b implies e(t)e(s)e(t)\neq e(s). As e(i)e(i)’s are orthogonal idempotents, for Ct,tλC_{t^{\prime},t}^{\lambda} and Cs,sμC_{s,s^{\prime}}^{\mu} in the cellular basis in Definition 4.5, we have

(3.10) Ct,tλCs,sμ=0,C_{t^{\prime},t}^{\lambda}C_{s,s^{\prime}}^{\mu}=0,

which implies

(3.11) Br,3=λ𝔅3(r)eλBr,3B_{r,3}=\bigoplus_{\lambda\in\mathfrak{B}_{3}^{(r)}}\langle e_{\lambda}\rangle_{B_{r,3}}

and each eλBr,3\langle e_{\lambda}\rangle_{B_{r,3}} is an indecomposable two-sided ideal of Br,3B_{r,3}. On the other hand, we have

(3.12) TLr,1,3Br,3/E1Br,3TL_{r,1,3}\cong B_{r,3}/\langle E_{1}\rangle_{B_{r,3}}

where E1E_{1} is the idempotent above. The decomposition in (3.11) implies

(3.13) E1Br,3=λ𝔇eλBr,3,\langle E_{1}\rangle_{B_{r,3}}=\bigoplus_{\lambda\in\mathfrak{D}}\langle e_{\lambda}\rangle_{B_{r,3}},

where 𝔇\mathfrak{D} is a subset of 𝔅3(r)\mathfrak{B}_{3}^{(r)}. By Theorem 2.7 and a simple calculation, observe that

(3.14) E1𝕍λ=0;E0(i)𝕍λ=0 for all i=1,2 and λ𝔅3(r)(j),j=1,2.E_{1}\mathbb{V}_{\lambda}=0;E_{0}^{(i)}\mathbb{V}_{\lambda}=0\textit{ for all }i=1,2\textit{ and }\lambda\in\mathfrak{B}_{3}^{(r)}(j),j=1,2.

Therefore, for any λ𝔅3(r)(j),j=1,2\lambda\in\mathfrak{B}_{3}^{(r)}(j),j=1,2, 𝕍λ\mathbb{V}_{\lambda} is a simple module of TLr,1,3TL_{r,1,3}. So we have

dim(TLr,1,3)\displaystyle dim(TL_{r,1,3}) λ𝔅3(r)(1)(dim(𝕍λ))2+λ𝔅3(r)(2)(dim(𝕍λ))2\displaystyle\geq\sum_{\lambda\in\mathfrak{B}_{3}^{(r)}(1)}(dim(\mathbb{V}_{\lambda}))^{2}+\sum_{\lambda\in\mathfrak{B}_{3}^{(r)}(2)}(dim(\mathbb{V}_{\lambda}))^{2}
=3×1+6×32\displaystyle=3\times 1+6\times 3^{2}
=57.\displaystyle=57.

On the other hand, for any λ𝔅3(r)(3)\lambda\in\mathfrak{B}_{3}^{(r)}(3), denote by k1k_{1},k2k_{2} and k3k_{3} the indices of the three non-empty components of λ\lambda. Let ta1a2a3t_{a_{1}a_{2}a_{3}} be the standard tableau of shape λ\lambda with the number aia_{i} in the kithk_{i}^{th} position. Then by Theorem 2.7, {ta1a2a3}\{t_{a_{1}a_{2}a_{3}}\} forms a basis of the irreducible module 𝕍λ\mathbb{V}_{\lambda}, and we have

(3.15) E1t123=\displaystyle E_{1}t_{123}= (qva2va1va2va1+(q1)va3(qva2va1)(va2va1)(va3va2)+(q1)va3(va2qva1)(va2va1)(va3va1))(t123+t213)\displaystyle(\frac{qv_{a_{2}}-v_{a_{1}}}{v_{a_{2}}-v_{a_{1}}}+\frac{(q-1)v_{a_{3}}(qv_{a_{2}}-v_{a_{1}})}{(v_{a_{2}}-v_{a_{1}})(v_{a_{3}}-v_{a_{2}})}+\frac{(q-1)v_{a_{3}}(v_{a_{2}}-qv_{a_{1}})}{(v_{a_{2}}-v_{a_{1}})(v_{a_{3}}-v_{a_{1}})})(t_{123}+t_{213})
+(qva3va1)(qva3va2)(va3va1)(va3va2)(t132+t231+t312+t321).\displaystyle+\frac{(qv_{a_{3}}-v_{a_{1}})(qv_{a_{3}}-v_{a_{2}})}{(v_{a_{3}}-v_{a_{1}})(v_{a_{3}}-v_{a_{2}})}(t_{132}+t_{231}+t_{312}+t_{321}).

Since vivjq\frac{v_{i}}{v_{j}}\neq q for any iji\neq j, (qva3va1)(qva3va2)0(qv_{a_{3}}-v_{a_{1}})(qv_{a_{3}}-v_{a_{2}})\neq 0. Therefore, E1E_{1} acts non-trivially on 𝕍λ\mathbb{V}_{\lambda} if λ𝔅3(r)(3)\lambda\in\mathfrak{B}_{3}^{(r)}(3). So TLr,1,3TL_{r,1,3} has no simple modules of rank 66. Therefore, 𝔅3(r)(3)𝔇\mathfrak{B}_{3}^{(r)}(3)\subseteq\mathfrak{D}. Comparing dimensions, we have 𝔅3(r)(3)=𝔇\mathfrak{B}_{3}^{(r)}(3)=\mathfrak{D} which implies

(3.16) TLr,1,3=λ𝔅3(r)(i),i=1,2eλBr,3.TL_{r,1,3}=\bigoplus_{\lambda\in\mathfrak{B}_{3}^{(r)}(i),i=1,2}\langle e_{\lambda}\rangle_{B_{r,3}}.

Therefore,

TLr,1,3\displaystyle TL_{r,1,3} =Br,3/λ𝔅3(r)(3)eλBr,3\displaystyle=B_{r,3}/\bigoplus_{\lambda\in\mathfrak{B}_{3}^{(r)}(3)}\langle e_{\lambda}\rangle_{B_{r,3}}
=Br,3/(αij,Λ)>0 for j=1,2,3,ijik for jke(i1,i2,i3)Br,3\displaystyle=B_{r,3}/\sum_{(\alpha_{i_{j}},\Lambda)>0\text{ for }j=1,2,3,i_{j}\neq i_{k}\textit{ for }j\neq k}\langle e(i_{1},i_{2},i_{3})\rangle_{B_{r,3}}
=3Λ/𝒥3.\displaystyle=\mathcal{R}_{3}^{\Lambda}/\mathcal{J}_{3}.

For any n3n\geq 3, 3Λ\mathcal{R}_{3}^{\Lambda} can be regarded as a subalgebra of nΛ\mathcal{R}_{n}^{\Lambda}. Denote by 𝒥n\mathcal{J}_{n} the two-sided ideal of nΛ\mathcal{R}_{n}^{\Lambda} which is generated by 𝒥3\mathcal{J}_{3}. Lemma (2.11)(\ref{keylm}) implies

(3.17) 𝒥n(Λ)=\displaystyle\mathcal{J}_{n}(\Lambda)= 𝐢In2,(αi,Λ)>0e(i,i+1,𝐢)nΛ\displaystyle\sum_{\mathbf{i}\in I^{n-2},(\alpha_{i},\Lambda)>0}\langle e(i,i+1,\mathbf{i})\rangle_{\mathcal{R}_{n}^{\Lambda}}
+𝐢In3,(αij,Λ)>0 for j=1,2,3,ijik for jke(i1,i2,i3,𝐢)nΛ.\displaystyle+\sum_{\mathbf{i}\in I^{n-3},(\alpha_{i_{j}},\Lambda)>0\text{ for }j=1,2,3,i_{j}\neq i_{k}\textit{ for }j\neq k}\langle e(i_{1},i_{2},i_{3},\mathbf{i})\rangle_{\mathcal{R}_{n}^{\Lambda}}.

It now follows that the generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} can be realised as a quotient of the KLR algebra, that is

Theorem 3.7.

Let TLr,1,nTL_{r,1,n} be the generalised Temperley-Lieb algebra in Definition 3.1 and nΛ\mathcal{R}_{n}^{\Lambda} be the cyclotomic KLR algebra isomorphic to the corresponding Hecke algebra H(r,1,n)H(r,1,n) with Λ\Lambda satisfying (3.5)(\ref{reslmd}). Then

(3.18) TLr,1,nnΛ/𝒥n,TL_{r,1,n}\cong\mathcal{R}_{n}^{\Lambda}/\mathcal{J}_{n},

where 𝒥n\mathcal{J}_{n} is the two-sided ideal of nΛ\mathcal{R}_{n}^{\Lambda} in (3.17).

In [25], Lobos and Ryom-Hansen introduce a KLR interpretation of the generalised blob algebra Br,nB_{r,n} as a quotient of nΛ\mathcal{R}_{n}^{\Lambda}:

Theorem 3.8.

([25], Theorem 42) Let Br,nB_{r,n} and nΛ\mathcal{R}_{n}^{\Lambda} be as defined above, then Br,n=nΛ/𝒥1,nB_{r,n}=\mathcal{R}_{n}^{\Lambda}/\mathcal{J}_{1,n} where

(3.19) 𝒥1,n=𝐢In2,(αi,Λ)>0e(i,i+1,𝐢)nΛ.\mathcal{J}_{1,n}=\sum_{\mathbf{i}\in I^{n-2},(\alpha_{i},\Lambda)>0}\langle e(i,i+1,\mathbf{i})\rangle_{\mathcal{H}_{n}^{\Lambda}}.

Since 𝒥1,n\mathcal{J}_{1,n} is contained in the ideal 𝒥n\mathcal{J}_{n}, the generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} is realised as a quotient of the generalised blob algebra Br,nB_{r,n}:

Corollary 3.9.

The generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} is a quotient of the generalised blob algebra Br,nB_{r,n} by the two-sided ideal

𝒥2,n=𝐢In3,(αij,Λ)>0 for j=1,2,3,ijik for jke(i1,i2,i3,𝐢)Br,n.\mathcal{J}_{2,n}=\sum_{\mathbf{i}\in I^{n-3},(\alpha_{i_{j}},\Lambda)>0\text{ for }j=1,2,3,i_{j}\neq i_{k}\textit{ for }j\neq k}\langle e(i_{1},i_{2},i_{3},\mathbf{i})\rangle_{B_{r,n}}.

4. The graded cellularity of TLr,1,nTL_{r,1,n}

In this section, we show that TLr,1,nTL_{r,1,n} is a graded cellular algebra when r2r\geq 2. The cellularity of TL1,1,nTL_{1,1,n} is explained in Example 1.4 in [13]. We construct a cellular structure of TLr,1,nTL_{r,1,n} using a truncation of that of the generalised blob algebra Br,nB_{r,n} given by Lobos and Ryom-Hansen in [25].

We begin by recalling the definition of cellular algebras.

Definition 4.1.

([13], Definition 1.1) Let RR be a commutative ring with unit and AA be an RR-algebra. Then AA is called cellular if it has a cell datum (Λ,i,M,C)(\Lambda,i,M,C) consisting of

a) A finite poset Λ\Lambda;

b) An RR-linear anti-involution ii;

c) A finite, non-empty set M(λ)M(\lambda) of indices for every λΛ\lambda\in\Lambda;

d) An injection

C:λΛM(λ)×M(λ)\displaystyle C:\cup_{\lambda\in\Lambda}M(\lambda)\times M(\lambda) A\displaystyle\to A
(S,T)\displaystyle(S,T) CS,Tλ\displaystyle\mapsto C_{S,T}^{\lambda}

satisfying the following conditions:

1) The image of CC forms an RR-basis of AA.

2) i(CS,Tλ)=CT,Sλi(C_{S,T}^{\lambda})=C_{T,S}^{\lambda} for all elements in the basis.

3) Let A<λA_{<\lambda} be the RR-span of all the elements of form CX,YμC_{X,Y}^{\mu} with μ<λ\mu<\lambda in the poset. Then for all aAa\in A,

(4.1) aCS,Tλ=SM(λ)ra(S,S)CS,Tλ mod A<λaC_{S,T}^{\lambda}=\sum_{S^{\prime}\in M(\lambda)}r_{a}(S^{\prime},S)C_{S^{\prime},T}^{\lambda}\text{ }mod\text{ }A_{<\lambda}

with the coefficients ra(S,S)r_{a}(S^{\prime},S) independent of TT.

Let Λ\Lambda^{\prime} be a downward closed subset of Λ\Lambda. Then A(Λ):=CS,Tλ|λΛRA(\Lambda^{\prime}):=\langle C_{S,T}^{\lambda}|\lambda\in\Lambda^{\prime}\rangle_{R} forms a two-sided ideal of AA. Further, if Λ1Λ2\Lambda_{1}\subseteq\Lambda_{2} are two downward closed subsets of Λ\Lambda, then A(Λ1)A(Λ2A(\Lambda_{1})\subseteq A(\Lambda_{2} and A(Λ2)/A(Λ1)A(\Lambda_{2})/A(\Lambda_{1}) can be regarded as an (A,A)(A,A)-bimodule. In particular, for λΛ\lambda\in\Lambda, denote by A({λ})A(\{\lambda\}) the bimodule A(Λ2)/A(Λ1)A(\Lambda_{2})/A(\Lambda_{1}) where Λ2={μΛ|μλ}\Lambda_{2}=\{\mu\in\Lambda|\mu\leq\lambda\} and Λ1={μΛ|μ<λ}\Lambda_{1}=\{\mu\in\Lambda|\mu<\lambda\}.

Let us recall the definition of a cell module which is introduced by Graham and Lehrer in [13].

Definition 4.2.

([13], Definition 2.1) For each λΛ\lambda\in\Lambda define the (left) AA-module W(λ)W(\lambda) as follows:

W(λ)W(\lambda) is a free module with basis {CS|SM(λ)}\{C_{S}|S\in M(\lambda)\} and the AA-action is defined by

(4.2) aCS=SStd(λ)ra(S,S)CSaC_{S}=\sum_{S^{\prime}\in Std(\lambda)}r_{a}(S^{\prime},S)C_{S^{\prime}}

for all aAa\in A and SM(λ)S\in M(\lambda) and ra(S,S)r_{a}(S^{\prime},S) is the coefficient in (4.1)(\ref{55}). The module W(λ)W(\lambda) is called the (left) cell module of AA corresponding to λ\lambda.

Similarly, we can define a right AA-module W(λ)W(\lambda)^{*}. There is a natural isomorphism of RR-modules Cλ:W(λ)W(λ)A({λ})C^{\lambda}:W(\lambda)\otimes W(\lambda)^{*}\mapsto A(\{\lambda\}) defined by (CS,CT)CS,Tλ(C_{S},C_{T})\mapsto C_{S,T}^{\lambda}.

There is a bilinear form ϕλ(,)\phi_{\lambda}(\;,\;) on W(λ)W(\lambda) which is defined as follows and extended bilinearly:

(4.3) CS,TλCS,Tλ=ϕλ(CT,CS)CS,Tλ mod A<λ.C_{S^{\prime},T}^{\lambda}C_{S,T^{\prime}}^{\lambda}=\phi_{\lambda}(C_{T},C_{S})C_{S^{\prime},T^{\prime}}^{\lambda}\text{ }mod\text{ }A_{<\lambda}.

For λΛ\lambda\in\Lambda, define rad(ϕλ):={xW(λ)|ϕλ(x,y)=0 for all yW(λ)}rad(\phi_{\lambda}):=\{x\in W(\lambda)|\phi_{\lambda}(x,y)=0\text{ for all }y\in W(\lambda)\}. Let Λ0={λΛ|ϕλ0}\Lambda_{0}=\{\lambda\in\Lambda|\phi_{\lambda}\neq 0\}. Graham and Lehrer list the simple modules of AA in [13]:

Theorem 4.3.

([13], Proposition 3.2,Theorem 3.4) Let AA be a cellular algebra with the cell datum (Λ,i,M,C)(\Lambda,i,M,C) over a base ring RR which is a field. For λΛ\lambda\in\Lambda, let W(λ)W(\lambda) be the cell module of AA and rad(ϕλ)rad(\phi_{\lambda}) be as defined above. Then we have

(i) rad(ϕλ)rad(\phi_{\lambda}) is an AA-submodule of W(λ)W(\lambda);

(ii) If ϕλ0\phi_{\lambda}\neq 0, the quotient W(λ)/rad(ϕλ)W(\lambda)/rad(\phi_{\lambda}) is absolutely irreducible;

(iii) The set {L(λ)=W(λ)/rad(ϕλ)|λΛ0}\{L(\lambda)=W(\lambda)/rad(\phi_{\lambda})|\lambda\in\Lambda_{0}\} is a complete set of absolutely irreducible AA-modules.

In [16], Hu and Mathas introduce a graded cell datum as a development of the original one, which is equipped with a degree function

deg:λΛM(λ),deg:\cup_{\lambda\in\Lambda}M(\lambda)\mapsto\mathbb{Z},

which satisfies CS,TλC_{S,T}^{\lambda} is homogeneous and of degree deg(CS,Tλ)=deg(S)+deg(T)deg(C_{S,T}^{\lambda})=deg(S)+deg(T) for all λλ\lambda\in\lambda and S,TM(λ)S,T\in M(\lambda).

We next recall the graded cellular basis of Br,nB_{r,n} constructed by Lobos and Ryom-Hansen in [25]. Let 𝔅n(r)\mathfrak{B}_{n}^{(r)} be the poset of one-column multipartitions of nn defined in 2.2. For λ𝔓n(r)\lambda\in\mathfrak{P}_{n}^{(r)}, let tλt^{\lambda} be the unique tableau of shape λ\lambda such that tλ(i)<tλ(j)t^{\lambda}(i)<t^{\lambda}(j) if 1i<jn1\leq i<j\leq n. For each tStd(λ)t\in Std(\lambda), denote by d(t)𝔖nd(t)\in\mathfrak{S}_{n} the permutation such that t=tλd(t)t=t^{\lambda}\circ d(t) where \circ is the natural 𝔖n\mathfrak{S}_{n}-action on the tableaux. For a multipartition λ𝔓n(r)\lambda\in\mathfrak{P}_{n}^{(r)} and a node γ=(a,1,l)[λ]\gamma=(a,1,l)\in[\lambda], define the residue of γ\gamma as

(4.4) ResΛ(γ):=1a+ilRes^{\Lambda}(\gamma):=1-a+i_{l}

where ili_{l} is the subscript of the dominate weight Λ=Λi1+Λi2++Λir\Lambda=\Lambda_{i_{1}}+\Lambda_{i_{2}}+\dots+\Lambda_{i_{r}}. For tTab(λ)t\in Tab(\lambda) and 1kn1\leq k\leq n, set RestΛ(k)=ResΛ(γ)Res_{t}^{\Lambda}(k)=Res^{\Lambda}(\gamma), where γ\gamma is the unique node such that t(γ)=kt(\gamma)=k. Define the residue sequence of tt as follows:

(4.5) resΛ(t):=(RestΛ(1),RestΛ(2),,RestΛ(n))In.res^{\Lambda}(t):=(Res_{t}^{\Lambda}(1),Res_{t}^{\Lambda}(2),\dots,Res_{t}^{\Lambda}(n))\in I^{n}.

For example, let Λ=Λ0+Λ3+Λ7\Lambda=\Lambda_{0}+\Lambda_{3}+\Lambda_{7} and tt be the following tableau, then resΛ(t)=(0,3,2,1,2,1,3,4,5)res^{\Lambda}(t)=(0,3,2,-1,-2,1,-3,-4,-5).

145789236\emptyset
Definition 4.4.

Suppose that λ𝔅n(r)\lambda\in\mathfrak{B}_{n}^{(r)} and tTab(λ)t\in Tab(\lambda) is a tableau of shape λ\lambda. Let t(λ)t^{(\lambda)} be the standard tableau of shape λ\lambda defined above. Set

e(t)\displaystyle e(t) =e(resΛ(t));\displaystyle=e(res^{\Lambda}(t));
eλ\displaystyle e_{\lambda} :=e(resΛ(t(λ))),\displaystyle:=e(res^{\Lambda}(t^{(\lambda)})),

where e(i)e(i) is the idempotent corresponding to the eigenspace MiM_{i} in (2.38).

Let * be the unique RR-linear anti-automorphism of the KLR algebra nΛ\mathcal{R}_{n}^{\Lambda} introduced by Brundan and Kleshchev in section 4.5 of [11] which fixes each of the generators in Definition 2.9.

Definition 4.5.

Suppose that λ𝔅n(r)\lambda\in\mathfrak{B}_{n}^{(r)} and s,tStd(λ)s,t\in Std(\lambda) and fix reduced expressions d(s)=si1si2sikd(s)=s_{i_{1}}s_{i_{2}}\dots s_{i_{k}} and d(t)=sj1sj2sjmd(t)=s_{j_{1}}s_{j_{2}}\dots s_{j_{m}}. Define

(4.6) Cs,tλ=ψd(s)eλψd(t),C_{s,t}^{\lambda}=\psi_{d(s)}^{*}e_{\lambda}\psi_{d(t)},

where ψd(s)=ψi1ψi2ψik\psi_{d(s)}=\psi_{i_{1}}\psi_{i_{2}}\dots\psi_{i_{k}} and ψd(t)=ψj1ψj2ψjm\psi_{d(t)}=\psi_{j_{1}}\psi_{j_{2}}\dots\psi_{j_{m}}.

Define the degree function deg:λ𝔅n(r)Std(λ)deg:\bigsqcup_{\lambda\in\mathfrak{B}_{n}^{(r)}}Std(\lambda)\mapsto\mathbb{Z} as

(4.7) deg(t):=deg(ψd(t)eλ),deg(t):=deg(\psi_{d(t)}^{*}e_{\lambda}),

where the degree function on the right hand side is the one defined in (2.32)(\ref{degf}). We next recall the graded cell datum of 𝔅n(r)\mathfrak{B}_{n}^{(r)} introduced by Lobos and Ryom-Hansen:

Theorem 4.6.

([25], Theorem 38) Let 𝔅n(r)\mathfrak{B}_{n}^{(r)} be the poset of one-column rr-partitions of nn and C,degC,deg be as defined above. (𝔅n(r),,Std,C,deg)(\mathfrak{B}_{n}^{(r)},*,Std,C,deg) is a graded cell datum of Br,nΛB_{r,n}^{\Lambda}. In other words, {Cs,tλ|λ𝔅n(r),s,tStd(λ)}\{C_{s,t}^{\lambda}|\lambda\in\mathfrak{B}_{n}^{(r)},s,t\in Std(\lambda)\} is a graded cellular basis of Br,nΛB_{r,n}^{\Lambda} with respect to the partial order \unlhd.

Readers may notice that the KLR generators yky_{k} are not used in the cellular basis. The following lemma shows that ykeλy_{k}e_{\lambda} can be realised as a combination of lower terms.

Lemma 4.7.

([25], Lemma 17, Lemma 37)

For λ𝔅n(r)\lambda\in\mathfrak{B}_{n}^{(r)} and 1kn1\leq k\leq n, we have

(4.8) ykeλ=eλyk=μλDμy_{k}e_{\lambda}=e_{\lambda}y_{k}=\sum_{\mu\triangleleft\lambda}D_{\mu}

where the sum runs over μ𝔅n(r)\mu\in\mathfrak{B}_{n}^{(r)} and DμD_{\mu} is in the two-sided ideal generated by eμe_{\mu}.

We now concentrate on a truncation of this cellular structure. Let 𝔇n(r):=k3𝔅n(r)(k)\mathfrak{D}_{n}^{(r)}:=\coprod_{k\geq 3}\mathfrak{B}_{n}^{(r)}(k) be the subset of 𝔅n(r)\mathfrak{B}_{n}^{(r)} consisting of multipartitions with at least 3 non-empty components. By Lemma 2.1, 𝔇n(r)\mathfrak{D}_{n}^{(r)} is downward closed with respect to the partial order \unlhd defined by (2.3).

Lemma 4.8.

Denote by 𝔇\mathcal{I}_{\mathfrak{D}} the two-sided ideal of Br,nB_{r,n} generated by {Cs,tλ}\{C_{s,t}^{\lambda}\} where λ\lambda runs over 𝔇n(r)\mathfrak{D}_{n}^{(r)}. Let 𝒥2,n\mathcal{J}_{2,n} be the two-sided ideal in Corollary 3.9. Then we have 𝔇=𝒥2,n\mathcal{I}_{\mathfrak{D}}=\mathcal{J}_{2,n}.

Proof.

By the definition of tλt^{\lambda}, for any λ𝔇n(r)\lambda\in\mathfrak{D}_{n}^{(r)}, e(tλ)e(t^{\lambda}) is of the form e(i1,i2,i3,𝐢)e(i_{1},i_{2},i_{3},\mathbf{i}) where (αij,Λ)>0(\alpha_{i_{j}},\Lambda)>0 for j=1,2,3j=1,2,3 and 𝐢In3\mathbf{i}\in I^{n-3}. Therefore, Cs,tλ𝒥2,nC_{s,t}^{\lambda}\in\mathcal{J}_{2,n}. So 𝔇\mathcal{I}_{\mathfrak{D}} is contained in 𝒥2,n\mathcal{J}_{2,n}.

We next show that e(i1,i2,i3,𝐢)𝔇e(i_{1},i_{2},i_{3},\mathbf{i})\in\mathcal{I}_{\mathfrak{D}} where (αij,Λ)>0(\alpha_{i_{j}},\Lambda)>0 for j=1,2,3j=1,2,3 and 𝐢In3\mathbf{i}\in I^{n-3}. We prove this by showing that the image of e(i1,i2,i3,𝐢)e(i_{1},i_{2},i_{3},\mathbf{i}) in the truncation Br,n/𝔇B_{r,n}/\mathcal{I}_{\mathfrak{D}} is trivial.

Let λk=1,2𝔅n(r)(k)\lambda\in\coprod_{k=1,2}\mathfrak{B}_{n}^{(r)}(k), s,tStd(λ)s,t\in Std(\lambda) and Cs,tλC_{s,t}^{\lambda} is the element defined in the cellular basis in Definition 4.5.

If λ𝔅n(r)(1)\lambda\in\mathfrak{B}_{n}^{(r)}(1), we have s=t=tλs=t=t^{\lambda} and Cs,tλ=eλ=e(j,j1,j2,,j+1n)C_{s,t}^{\lambda}=e_{\lambda}=e(j,j-1,j-2,\dots,j+1-n) for some jIj\in I. The condition (3.1)(\ref{pmres}) implies that |i1i2|1|i_{1}-i_{2}|\neq 1. Therefore, e(i1,i2,i3,𝐢)eλe(i_{1},i_{2},i_{3},\mathbf{i})\neq e_{\lambda} which implies

(4.9) e(i1,i2,i3,𝐢)eλ=e(i1,i2,i3,𝐢)Cs,tλ=0.e(i_{1},i_{2},i_{3},\mathbf{i})e_{\lambda}=e(i_{1},i_{2},i_{3},\mathbf{i})C_{s,t}^{\lambda}=0.

If λ𝔅n(r)(2)\lambda\in\mathfrak{B}_{n}^{(r)}(2), we have

Cs,tλ=ψd(s)eλψd(t)=e(s)ψd(s)ψd(t).C_{s,t}^{\lambda}=\psi_{d(s)}^{*}e_{\lambda}\psi_{d(t)}=e(s)\psi_{d(s)}^{*}\psi_{d(t)}.

Let e(j1,j2,j3,,jn)=e(s)e(j_{1},j_{2},j_{3},\dots,j_{n})=e(s). As sStd(λ)s\in Std(\lambda) and λ𝔅n(r)(2)\lambda\in\mathfrak{B}_{n}^{(r)}(2), two of 1,21,2 and 33 must be in the same column of ss which implies j1=j2+1j_{1}=j_{2}+1, j2=j3+1j_{2}=j_{3}+1 or j1=j3+1j_{1}=j_{3}+1. Comparing with the restriction on i1i_{1}, i2i_{2} and i3i_{3}, we have e(i1,i2,i3,𝐢)e(s)e(i_{1},i_{2},i_{3},\mathbf{i})\neq e(s). Therefore,

(4.10) e(i1,i2,i3,𝐢)Cs,tλ=0.e(i_{1},i_{2},i_{3},\mathbf{i})C_{s,t}^{\lambda}=0.

As Br,n/𝔇B_{r,n}/\mathcal{I}_{\mathfrak{D}} is a truncation of Br,nB_{r,n} with respect to the cellular structure in Theorem 4.6 and the ideal 𝔇n(r):=k3𝔅n(r)(k)\mathfrak{D}_{n}^{(r)}:=\coprod_{k\geq 3}\mathfrak{B}_{n}^{(r)}(k), {Cs,tλ|λk=1,2𝔅n(r)(k)}\{C_{s,t}^{\lambda}|\lambda\in\coprod_{k=1,2}\mathfrak{B}_{n}^{(r)}(k)\} is a cellular basis of it. As Br,nB_{r,n} is an algebra with 11, (4.9)(\ref{orth1}) and (4.10)(\ref{orth2}) imply that the image of e(i1,i2,i3,𝐢)e(i_{1},i_{2},i_{3},\mathbf{i}) in this truncation is 0, so e(i1,i2,i3,𝐢)𝔇e(i_{1},i_{2},i_{3},\mathbf{i})\in\mathcal{I}_{\mathfrak{D}}. Since the ideal 𝒥2,n\mathcal{J}_{2,n} is generated by e(i1,i2,i3,𝐢)e(i_{1},i_{2},i_{3},\mathbf{i}), we have 𝒥2,n𝔇\mathcal{J}_{2,n}\subseteq\mathcal{I}_{\mathfrak{D}}.

Therefore, 𝒥2,n=𝔇\mathcal{J}_{2,n}=\mathcal{I}_{\mathfrak{D}}. ∎

As an immediate consequence of this lemma and Theorem 4.6, a cellular structure of TLr,1,nTL_{r,1,n} can be obtained by a truncation of that of the generalised blob algebra Br,nB_{r,n}. In particular,

Theorem 4.9.

Let 𝔅n(r)\mathfrak{B}_{n}^{(r)} be the set of one-column rr-partitions of nn and 𝔇n(r)\mathfrak{D}_{n}^{(r)} be the subset consisting multipartitions with more than two non-empty components. For λ𝔅n(r)𝔇n(r)\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)} and s,tStd(λ)s,t\in Std(\lambda), let Cs,tλC_{s,t}^{\lambda} be the element defined in (4.6)(\ref{76}) and degdeg be the degree function on the tableaux defined in (4.7)(\ref{degt}). Let TLr,1,nTL_{r,1,n} be the generalised Temperley-Lieb algebra over a field RR of characteristic 0 defined in Definition 3.1 (cf. Theorem 3.7) and * be the anti-automorphism of TLr,1,nTL_{r,1,n} fixing the KLR generators. Then TLr,1,nTL_{r,1,n} is a graded cellular algebra with the cell datum (𝔅n(r)𝔇n(r),,Std,C,deg)(\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)},*,Std,C,deg) with respect to the partial order \unlhd defined by (2.3). Specifically, for any aTLr,1,na\in TL_{r,1,n}, λ𝔅n(r)𝔇n(r)\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)} and s,tStd(λ)s,t\in Std(\lambda), we have

(4.11) aCs,tλ=sStd(λ)ra(s,s)Cs,tλ+μλ,u,vStd(μ)ca(s,t,u,v)Cu,vμaC_{s,t}^{\lambda}=\sum_{s^{\prime}\in Std(\lambda)}r_{a}(s^{\prime},s)C_{s^{\prime},t}^{\lambda}+\sum_{\mu\triangleleft\lambda,u,v\in Std(\mu)}c_{a}(s,t,u,v)C_{u,v}^{\mu}

where ra(s,s)Rr_{a}(s^{\prime},s)\in R does not depend on tt and ca(s,t,u,v)Rc_{a}(s,t,u,v)\in R.

Notice that the cellular bases of Br,nΛB_{r,n}^{\Lambda} and TLr,1,nΛTL_{r,1,n}^{\Lambda} are heavily dependent on the fixed reduced expression for each d(t)d(t). In order to facilitate calculations, we fix a special set of reduced expressions, called “official”, for the d(t)d(t), which satisfy the following property.

Definition 4.10.

Let λ𝔅n(r)\lambda\in\mathfrak{B}_{n}^{(r)} be a one-column multipartition and for tStd(λ)t\in Std(\lambda), denote by d(t)d(t) the permutation such that t=tλd(t)t=t^{\lambda}\circ d(t). A set of fixed reduced expressions ORE(λ):={d(t)=sj1sj2sjm|tStd(λ)}ORE(\lambda):=\{d(t)=s_{j_{1}}s_{j_{2}}\dots s_{j_{m}}|t\in Std(\lambda)\} is called official if for each tStd(λ)t\in Std(\lambda), there is a unique reduced expression d(t)=sj1sj2sjmORE(λ)d(t)=s_{j_{1}}s_{j_{2}}\dots s_{j_{m}}\in ORE(\lambda) and sj1sj2sjm1ORE(λ)s_{j_{1}}s_{j_{2}}\dots s_{j_{m-1}}\in ORE(\lambda) is a reduced expression of d(s)d(s) for some sStd(λ)s\in Std(\lambda).

In order to prove the existence of an official set of reduced expressions, we introduce a partial order on Tab(λ)Tab(\lambda), the set of tableaux of shape λ\lambda. For tTab(λ)t\in Tab(\lambda), denote by t|kt|_{k} the tableau obtained by deleting k+1,k+2,,nk+1,k+2,\dots,n from tt. For example, if tt is the tableau in Figure-2, then t|5t|_{5} is as in Figure-5.

12534\emptyset
Figure 5. A restriction to the tableau tt

For s,tTab(λ)s,t\in Tab(\lambda), we say sts\unlhd t if

(4.12) shape(s|k)shape(t|k)shape(s|_{k})\unlhd shape(t|_{k})

for all 1kn1\leq k\leq n, where \unlhd is the partial order defined by (2.3). Note that shape(s|k)shape(s|_{k}) is not necessarily a multipartition since ss can be non-standard but we can still define a partial order on the set consisting shape(s|k)shape(s|_{k}) for all sTab(λ)s\in Tab(\lambda) using (2.3). Let \leq be the Bruhat order on 𝔖n\mathfrak{S}_{n}. The following Theorem builds a connection between the posets of tableaux and the elements in the symmetric group.

Theorem 4.11.

([25], Theorem 4) Let λ𝔅n(r)\lambda\in\mathfrak{B}_{n}^{(r)} be a one-column multipartition and s,tTab(λ)s,t\in Tab(\lambda). Then d(s)d(t)d(s)\leq d(t) if and only if sts\unlhd t.

We next show how to construct an official set of reduced expressions ORE(λ)ORE(\lambda) for each λ𝔅n(r)\lambda\in\mathfrak{B}_{n}^{(r)}. We do this progressively in the set Std(λ)Std(\lambda), starting with the tableau with a trivial permutation. Denote by SS the set of tableaux corresponding to which we have chosen reduced expressions. We start with S={tλ}S=\{t^{\lambda}\} and the corresponding set of reduced expressions is ore={d(tλ)=1}ore=\{d(t^{\lambda})=1\}.

If S=Std(λ)S=Std(\lambda) then let ORE(λ)=oreORE(\lambda)=ore and ORE(λ)ORE(\lambda) is an official set of reduced expressions.

Otherwise, choose tStd(λ)St\in Std(\lambda)-S such that l(d(t))=ml(d(t))=m is the smallest among permutations corresponding to the left standard tableaux. As tλSt^{\lambda}\in S, ttλt\neq t^{\lambda}. Let 1in1\leq i\leq n be the smallest number such that t(i)>t(i+1)t(i)>t(i+1). Denote s=tsis=t\circ s_{i}. Then by definition, sts\triangleleft t. So d(s)<d(t)d(s)<d(t). On the other hand, t(i)>t(i+1)t(i)>t(i+1) implies that ii is not above i+1i+1, so sStd(λ)s\in Std(\lambda). Since l(d(t))l(d(t)) is the smallest, we have sSs\in S. Thus there exists d(s)=sj1sj2sjm1ored(s)=s_{j_{1}}s_{j_{2}}\dots s_{j_{m-1}}\in ore. Add tt to SS as well as d(t)=sj1sj2sjmd(t)=s_{j_{1}}s_{j_{2}}\dots s_{j_{m}} to oreore where jm=ij_{m}=i defined above. Then check whether S=Std(λ)S=Std(\lambda). If not, repeat this procedure.

As there are finitely many elements in Std(λ)Std(\lambda), the algorithm above confirms the existence of an official set of reduced expressions.

It should be remarked that the official set is not unique. As a counterexample, let λ=((12),(12))\lambda=((1^{2}),(1^{2})). There are 6 standard tableaux of shape λ\lambda. Denote by tijt_{ij} the tableau with ii and jj in the first component. There are two official sets of reduced expressions:

Standard tableaux:\displaystyle\textit{Standard tableaux}: t13\displaystyle t_{13} t12\displaystyle t_{12} t14\displaystyle t_{14} t23\displaystyle t_{23} t24\displaystyle t_{24} t34\displaystyle t_{34}
Official set 1:\displaystyle\textit{ Official set 1}: 1\displaystyle 1 s2\displaystyle s_{2} s3\displaystyle s_{3} s1\displaystyle s_{1} s1s3\displaystyle s_{1}s_{3} s1s3s2\displaystyle s_{1}s_{3}s_{2}
Official set 2:\displaystyle\textit{ Official set 2}: 1\displaystyle 1 s2\displaystyle s_{2} s3\displaystyle s_{3} s1\displaystyle s_{1} s3s1\displaystyle s_{3}s_{1} s3s1s2.\displaystyle s_{3}s_{1}s_{2}.

In the following sections, we fix one official set of reduced expressions ORE(λ)ORE(\lambda) for all λ𝔅n(r)\lambda\in\mathfrak{B}_{n}^{(r)}. Observe that the ‘official’ property is one of the whole set {d(t)=sj1sj2sjm|tStd(λ)}\{d(t)=s_{j_{1}}s_{j_{2}}\dots s_{j_{m}}|t\in Std(\lambda)\} rather than of a single permutation corresponding to some tableau. In the following sections, we always assume that the chosen set of reduced expressions for the cellular basis is official. To make the statement briefer, we also call these reduced expressions the official ones.

5. The cell modules and semisimplicity of TLr,1,nTL_{r,1,n}

In this section, we first calculate the dimensions of the cell modules of TLr,1,nTL_{r,1,n} when r2r\geq 2 and then give several equivalent conditions for the semisimplicity of TLr,1,nTL_{r,1,n}. Let λ𝔅n(r)𝔇n(r)\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)} be a one-column multipartition of nn with at most two non-empty components (cf. Theorem 4.9) and W(λ)W(\lambda) be the corresponding cell module of TLr,1,nTL_{r,1,n} with respect to the cell datum in Theorem 4.9. Denote by a1a_{1} the number of nodes in the first non-empty component of λ\lambda. Then the number of standard tableaux of shape λ\lambda is (na1)\binom{n}{a_{1}}. Hence we have

(5.1) dim(W(λ))=(na1).dim(W(\lambda))=\binom{n}{a_{1}}.

Denoting by TL({λ})TL(\{\lambda\}) the corresponding TLr,1,nTL_{r,1,n}-bimodule, then

(5.2) dim(TL({λ}))=(na1)2.dim(TL(\{\lambda\}))=\binom{n}{a_{1}}^{2}.

Further,

dim(TLr,1,n)\displaystyle dim(TL_{r,1,n}) =λ𝔅n(r)𝔇n(r)dim(TL({λ}))\displaystyle=\sum_{\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)}}dim(TL(\{\lambda\}))
=λ𝔅n(r)(1)dim(TL({λ}))+λ𝔅n(r)(2)dim(TL({λ}))\displaystyle=\sum_{\lambda\in\mathfrak{B}_{n}^{(r)}(1)}dim(TL(\{\lambda\}))+\sum_{\lambda\in\mathfrak{B}_{n}^{(r)}(2)}dim(TL(\{\lambda\}))
=r+(r2)(a1=1n1(na1)2)\displaystyle=r+\binom{r}{2}(\sum_{a_{1}=1}^{n-1}\binom{n}{a_{1}}^{2})
=r+(r2)((2nn)2)\displaystyle=r+\binom{r}{2}(\binom{2n}{n}-2)
=(r2)(2nn)r2+2r.\displaystyle=\binom{r}{2}\binom{2n}{n}-r^{2}+2r.

In the case where r=2r=2, we have dim(TL2,1,n)=(2nn)dim(TL_{2,1,n})=\binom{2n}{n}, which is the dimension of TLBn(q,Q)TLB_{n}(q,Q), the Temperley-Lieb algebra of type BnB_{n}. We next concentrate on the irreducible representations of TLr,1,nTL_{r,1,n}. A direct consequence of the definition of the cellular basis given in (4.6)(\ref{76}) is as follows:

Lemma 5.1.

Let ϕλ(,)\phi_{\lambda}(,) be the bilinear form defined in (4.3)(\ref{81}). For any λ𝔅n(r)𝔇n(r)\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)}, ϕλ0\phi_{\lambda}\neq 0.

Proof.

Let tλt^{\lambda} be the standard tableau of shape λ\lambda such that tλ(i)<tλ(j)t^{\lambda}(i)<t^{\lambda}(j) for all 1i<jn1\leq i<j\leq n. Then we have

Cs,tλλ\displaystyle C_{s,t^{\lambda}}^{\lambda} =ψd(s)eλ\displaystyle=\psi_{d(s)}^{*}e_{\lambda}
Ctλ,sλ\displaystyle C_{t^{\lambda},s^{\prime}}^{\lambda} =eλψd(s).\displaystyle=e_{\lambda}\psi_{d(s^{\prime})}.

Thus,

Cs,tλλCtλ,sλ\displaystyle C_{s,t^{\lambda}}^{\lambda}C_{t^{\lambda},s^{\prime}}^{\lambda} =ψd(s)eλeλψd(s)\displaystyle=\psi_{d(s)}^{*}e_{\lambda}e_{\lambda}\psi_{d(s^{\prime})}
=ψd(s)eλψd(s)\displaystyle=\psi_{d(s)}^{*}e_{\lambda}\psi_{d(s^{\prime})}
=Cs,sλ\displaystyle=C_{s,s^{\prime}}^{\lambda}

which implies ϕλ(tλ,tλ)=1\phi_{\lambda}(t^{\lambda},t^{\lambda})=1. ∎

This lemma guarantees that TLr,1,nTL_{r,1,n} is a quasi-hereditary algebra.

For two residues, aa and bb, in /e\mathbb{Z}/e\mathbb{Z}, We define the difference between them as

(5.3) |ab|:=min{|cd||c,d and ca,db mod e}.|a-b|:=min\{|c-d|\in\mathbb{N}|c,d\in\mathbb{Z}\textit{ and }c\equiv a,d\equiv b\textit{ mod }e\}.

The next theorem gives an equivalent condition for the semi-simplicity of TLr,1,nΛTL_{r,1,n}^{\Lambda} for some specific dominant weight Λ\Lambda.

Theorem 5.2.

Let Λ=Λj1+Λj2++Λjr\Lambda=\Lambda_{j_{1}}+\Lambda_{j_{2}}+\dots+\Lambda_{j_{r}} be a dominant weight satisfying (3.5)(\ref{reslmd}) and TLr,1,nΛTL_{r,1,n}^{\Lambda} be the generalised Temperley-Lieb algebra defined in Theorem 3.7 over a field RR of characteristic 0. Then TLr,1,nΛTL_{r,1,n}^{\Lambda} is semi-simple if and only if |jkjl|n|j_{k}-j_{l}|\geq n for all 1k<lr1\leq k<l\leq r, where |jkjl||j_{k}-j_{l}| is as defined in (5.3).

Before proving this theorem, we remark that the condition |jkjl|n|j_{k}-j_{l}|\geq n is not only a restriction on the dominant weight Λ\Lambda but also on the cardinality of the index set II, which is the smallest positive integer ee such that 1+q+q2++qe1=01+q+q^{2}+\dots+q^{e-1}=0. Since I=/eI=\mathbb{Z}/e\mathbb{Z} if e>0e>0 and all jkj_{k}’s are in II, this condition implies that erne\geq rn if e0e\neq 0.

Proof.

We first check the sufficiency. Suppose Λ=Λj1+Λj2++Λjr\Lambda=\Lambda_{j_{1}}+\Lambda_{j_{2}}+\dots+\Lambda_{j_{r}} satisfying that |jkjl|n|j_{k}-j_{l}|\geq n for all 1k<lr1\leq k<l\leq r. By Theorem 4.3, it is enough to show that rad(ϕλ)=0rad(\phi_{\lambda})=0 for all λ𝔅n(r)𝔇n(r)\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)}. Without losing generality, assume that the last r2r-2 components of λ\lambda are empty.

Let tλt^{\lambda} be the tableau of shape λ\lambda such that tλ(i)<tλ(j)t^{\lambda}(i)<t^{\lambda}(j) if 1i<jn1\leq i<j\leq n. For w𝔖nw\in\mathfrak{S}_{n}, let twλ=tλwt_{w}^{\lambda}=t^{\lambda}\circ w. We claim that ϕλ(twλ,twλ)=1\phi_{\lambda}(t_{w}^{\lambda},t_{w}^{\lambda})=1, if twλStd(λ)t_{w}^{\lambda}\in Std(\lambda). This will be proved by induction on l(w)l(w).

When l(w)=0l(w)=0, twλ=tλt_{w}^{\lambda}=t^{\lambda} and we have ϕλ(tλ,tλ)=1\phi_{\lambda}(t^{\lambda},t^{\lambda})=1 by the proof of Lemma 5.1. For a positive integer kk, assume ϕλ(twλ,twλ)=1\phi_{\lambda}(t_{w}^{\lambda},t_{w}^{\lambda})=1 for all w𝔖nw\in\mathfrak{S}_{n} such that l(w)<kl(w)<k and twλt_{w}^{\lambda} is a standard tableau. For w𝔖nw\in\mathfrak{S}_{n} with length kk, let w=su1su2sukw=s_{u_{1}}s_{u_{2}}\dots s_{u_{k}} be the official reduced expression. Then w=su1su2suk1w^{\prime}=s_{u_{1}}s_{u_{2}}\dots s_{u_{k-1}} is the official reduced expression of ww^{\prime} which is of length k1k-1. By the inductive assumption, we have:

(5.4) eλψwψweλ=ϕλ(twλ,twλ)eλ=eλ.e_{\lambda}\psi_{w^{\prime}}\psi_{w^{\prime}}^{*}e_{\lambda}=\phi_{\lambda}(t_{w^{\prime}}^{\lambda},t_{w^{\prime}}^{\lambda})e_{\lambda}=e_{\lambda}.

For twλt_{w}^{\lambda}, we have

ϕλ(twλ,twλ)eλ\displaystyle\phi_{\lambda}(t_{w}^{\lambda},t_{w}^{\lambda})e_{\lambda} =eλψwψweλ\displaystyle=e_{\lambda}\psi_{w}\psi_{w}^{*}e_{\lambda}
=eλψwψukψukψweλ\displaystyle=e_{\lambda}\psi_{w^{\prime}}\psi_{u_{k}}\psi_{u_{k}}\psi_{w^{\prime}}^{*}e_{\lambda}
=ψwψuk2e(twλ)ψw.\displaystyle=\psi_{w^{\prime}}\psi_{u_{k}}^{2}e(t_{w^{\prime}}^{\lambda})\psi_{w^{\prime}}^{*}.

We next show that ψuk2e(twλ)=e(twλ)\psi_{u_{k}}^{2}e(t_{w^{\prime}}^{\lambda})=e(t_{w^{\prime}}^{\lambda}). Let e(twλ)=e(i1,i2,,in)e(t_{w^{\prime}}^{\lambda})=e(i_{1},i_{2},\dots,i_{n}). By definition, (i1,i2,,in)(i_{1},i_{2},\dots,i_{n}) is a sequence consisting of i1,i11,,i1a+1,i2,i21,,i2n+a1i_{1},i_{1}-1,\dots,i_{1}-a+1,i_{2},i_{2}-1,\dots,i_{2}-n+a-1 if there are aa nodes in λ(1)\lambda^{(1)} and nan-a nodes in λ(2)\lambda^{(2)}. As |i1i2|n|i_{1}-i_{2}|\geq n, they are all distinct numbers. So the position of ll in the tableau twλt_{w^{\prime}}^{\lambda} is uniquely determined by ili_{l}. So for any value of uku_{k}, iukiuk+1i_{u_{k}}\neq i_{u_{k}+1}. As twλt_{w^{\prime}}^{\lambda} is a standard tableau, uku_{k} is not in the node below uk+1u_{k+1}, so iukiuk+11i_{u_{k}}\neq i_{u_{k}+1}-1. As twλ=twλsukt_{w}^{\lambda}=t_{w^{\prime}}^{\lambda}s_{u_{k}} is standard, uku_{k} is not in the node above uk+1u_{k}+1, so iukiuk+1+1i_{u_{k}}\neq i_{u_{k}+1}+1, either. By (2.30)(\ref{25eq}), ψuk2e(twλ)=e(twλ)\psi_{u_{k}}^{2}e(t_{w^{\prime}}^{\lambda})=e(t_{w^{\prime}}^{\lambda}). We have

ϕλ(twλ,twλ)eλ\displaystyle\phi_{\lambda}(t_{w}^{\lambda},t_{w}^{\lambda})e_{\lambda} =ψwψuk2e(twλ)ψw\displaystyle=\psi_{w^{\prime}}\psi_{u_{k}}^{2}e(t_{w^{\prime}}^{\lambda})\psi_{w^{\prime}}^{*}
=ψwe(twλ)ψw\displaystyle=\psi_{w^{\prime}}e(t_{w^{\prime}}^{\lambda})\psi_{w^{\prime}}^{*}
=ϕλ(twλ,twλ)eλ\displaystyle=\phi_{\lambda}(t_{w^{\prime}}^{\lambda},t_{w^{\prime}}^{\lambda})e_{\lambda}
=eλ.\displaystyle=e_{\lambda}.

So ϕλ(twλ,twλ)=1\phi_{\lambda}(t_{w}^{\lambda},t_{w}^{\lambda})=1.

For two different standard tableaux, t,sStd(λ)t,s\in Std(\lambda), let kk be the smallest number such that t(k)s(k)t(k)\neq s(k). As |jkjl|n|j_{k}-j_{l}|\geq n, ResΛ(t(k))ResΛ(s(k))Res^{\Lambda}(t(k))\neq Res^{\Lambda}(s(k)). Thus e(t)e(s)e(t)\neq e(s). So we have

eλψd(t)ψd(s)eλ=ψd(t)e(t)e(s)ψd(s)=0e_{\lambda}\psi_{d(t)}\psi_{d(s)}^{*}e_{\lambda}=\psi_{d(t)}e(t)e(s)\psi_{d(s)}^{*}=0

which implies that ϕλ(t,s)=0\phi_{\lambda}(t,s)=0.

Therefore, ϕλ(t,s)=δst\phi_{\lambda}(t,s)=\delta_{st} for all s,tStd(λ)s,t\in Std(\lambda). Thus, rad(ϕλ)=0rad(\phi_{\lambda})=0 for all λ𝔅n(r)𝔇n(r)\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)} and TLr,1,nΛTL_{r,1,n}^{\Lambda} is semi-simple.

We now turn to the necessity. Without losing generality, let j1j_{1} and j2j_{2} be the two closest elements in all the jj’s and 0<j2j1=b<n0<j_{2}-j_{1}=b<n. If e>0e>0, let bb be such that 2be2b\leq e.

Let λ=((1nb),(1b),,,)\lambda=((1^{n-b}),(1^{b}),\emptyset,\dots,\emptyset) and t0t_{0} be the standard tableau of shape λ\lambda such that 1,2,,b1,2,\dots,b are in the second component. We claim that Ct0rad(ϕλ)C_{t_{0}}\in rad(\phi_{\lambda}), which implies that TLr,1,nΛTL_{r,1,n}^{\Lambda} is not semi-simple according to Theorem 4.3.

We first prove ϕλ(t0,s)=0\phi_{\lambda}(t_{0},s)=0 if st0s\neq t_{0}. By definition,

ϕλ(t0,s)eλ=eλψd(t0)ψd(s)eλ=ψd(t0)e(t0)e(s)ψd(s).\phi_{\lambda}(t_{0},s)e_{\lambda}=e_{\lambda}\psi_{d(t_{0})}^{*}\psi_{d(s)}e_{\lambda}=\psi_{d(t_{0})}^{*}e(t_{0})e(s)\psi_{d(s)}.

As e(𝐢)e(\mathbf{i})’s are orthogonal idempotents, e(t0)e(s)0e(t_{0})e(s)\neq 0 if and only if e(t0)=e(s)e(t_{0})=e(s). Without losing generality, let j1=0j_{1}=0 and j2=bj_{2}=b. Then e(t0)=e(b,b1,b2,,bn)e(t_{0})=e(b,b-1,b-2,\dots,b-n). We prove that for s=t0s=t_{0} if e(t0)=e(s)e(t_{0})=e(s) and sStd(λ)s\in Std(\lambda). It is enough to show that 1,2,,b1,2,\dots,b are in the second component of ss. This is implied by the fact RessΛ(i)=b+1iRes_{s}^{\Lambda}(i)=b+1-i for 1ib1\leq i\leq b and ss is standard. So if st0s\neq t_{0}, ϕλ(t0,s)=0\phi_{\lambda}(t_{0},s)=0.

We next prove ϕλ(t0,t0)=0\phi_{\lambda}(t_{0},t_{0})=0. Let d(t0)=su1su2sukd(t_{0})=s_{u_{1}}s_{u_{2}}\dots s_{u_{k}} be the official reduced expression and t1t_{1} be the standard tableau such that d(t1)=su1su2suk1d(t_{1})=s_{u_{1}}s_{u_{2}}\dots s_{u_{k-1}}. As all of 1,2b1,2\dots b are in the second component of t0t_{0} and both t0t_{0} and t1t_{1} are standard, suk=sbs_{u_{k}}=s_{b} and t1t_{1} is the standard tableau with 1,2,,b1,b+11,2,\dots,b-1,b+1 in the second component. We have

e(t1)=e(b,b1,,2,0,1,1,,bn).e(t_{1})=e(b,b-1,\dots,2,0,1,-1,\dots,b-n).

And we have

eλψd(t0)ψd(t0)eλ\displaystyle e_{\lambda}\psi_{d(t_{0})}^{*}\psi_{d(t_{0})}e_{\lambda}
=eλψd(t1)ψuk2ψd(t1)eλ\displaystyle=e_{\lambda}\psi_{d(t_{1})}^{*}\psi_{u_{k}}^{2}\psi_{d(t_{1})}e_{\lambda}
=ψd(t1)ψb2e(b,b1,,2,0,1,1,,bn)ψd(t1)\displaystyle=\psi_{d(t_{1})}^{*}\psi_{b}^{2}e(b,b-1,\dots,2,0,1,-1,\dots,b-n)\psi_{d(t_{1})}
=ψd(t1)(yb+1yb)e(b,b1,,2,0,1,1,,bn)ψd(t1)\displaystyle=\psi_{d(t_{1})}^{*}(y_{b+1}-y_{b})e(b,b-1,\dots,2,0,1,-1,\dots,b-n)\psi_{d(t_{1})}
=ψd(t1)ψd(t1)(ycyd)eλ\displaystyle=\psi_{d(t_{1})}^{*}\psi_{d(t_{1})}(y_{c}-y_{d})e_{\lambda}

for some cc and dd between b+1b+1 and 1b1-b. The second last equation comes from the third case in (2.30)(\ref{25eq}). The last equation is from (2.26),(2.28)(\ref{21eq}),(\ref{23eq}) and (2.29)(\ref{24eq}) where the case ir=ir+1i_{r}=i_{r+1} is excluded by the fact that 2be2b\leq e and only sls_{l} where l2b1l\leq 2b-1 appears in the official reduced expression of d(t1)d(t_{1}).

And by Lemma 4.7, we have

yreλ=yseλ=0 mod TLr,1,nλ.y_{r}e_{\lambda}=y_{s}e_{\lambda}=0\text{ mod }TL_{r,1,n}^{\triangleleft\lambda}.

Thus,

eλψd(t0)ψd(t0)eλ=0 mod TLr,1,nλ.e_{\lambda}\psi_{d(t_{0})}^{*}\psi_{d(t_{0})}e_{\lambda}=0\text{ mod }TL_{r,1,n}^{\triangleleft\lambda}.

Therefore, we have ϕλ(t0,t0)=0\phi_{\lambda}(t_{0},t_{0})=0 by definition. In conclusion, Ct0rad(ϕλ)C_{t_{0}}\in rad(\phi_{\lambda}), which implies that TLr,1,nΛTL_{r,1,n}^{\Lambda} is not semi-simple. ∎

The following corollary is an immediate consequence of Theorem 5.2 and (2.37)(\ref{32eq}):

Corollary 5.3.

Let TLr,1,nTL_{r,1,n} be the generalised Temperley-Lieb quotient defined in Definition 3.1 with parameters q,v1,v2,,vrq,v_{1},v_{2},\dots,v_{r}. Then TLr,1,n(q,v1,v2,,vr)TL_{r,1,n}(q,v_{1},v_{2},\dots,v_{r}) is semisimple if and only if

vivjql\dfrac{v_{i}}{v_{j}}\neq q^{l}

for all 1ijr1\leq i\leq j\leq r and n+1ln1-n+1\leq l\leq n-1 and

1+q+q2++qi01+q+q^{2}+\dots+q^{i}\neq 0

for 1in11\leq i\leq n-1.

Comparing with Theorem 2.8, the generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n} is semisimple if and only if the same is true of the corresponding Hecke algebra H(r,1,n)H(r,1,n).

6. Irreducible representations and decomposition numbers

In this section, we study the irreducible representations and the decomposition numbers for cell modules of TLr,1,nTL_{r,1,n}. Let λ\lambda be a one-column multipartition of nn consisting of at most two non-empty components. Let W(λ)W(\lambda) be the cell module of TLr,1,nTL_{r,1,n} corresponding to λ\lambda and L(λ)=W(λ)/rad(ϕλ)L(\lambda)=W(\lambda)/rad(\phi_{\lambda}) be the corresponding simple module. We first calculate the dimension of L(λ)L(\lambda). By Theorem 4.3, it is enough to find the rank of the bilinear form ϕλ\phi_{\lambda}.

6.1. Garnir tableaux

When calculating the value of the bilinear form ϕλ(s,t)\phi_{\lambda}(s,t) for s,tStd(λ)s,t\in Std(\lambda), we can meet some non-standard tableaux. Inspired by Lobos and Ryom-Hansen, we use Garnir tableaux as a tool to deal with these non-standard tableaux. This method is originally due to Murphy in [26]. We start with the definition of a Garnir tableau.

Definition 6.1.

Let λ𝔅n(r)𝔇n(r)\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)} be a one-column multipartition of nn and gg be a λ\lambda-tableau. We call gg a Garnir tableau if there exists kk with 1kn11\leq k\leq n-1 such that

(a) gg is not standard but gskg\circ s_{k} is;

(b) if gsigg\circ s_{i}\triangleleft g, then i=ki=k, where \triangleleft is the partial order defined by (4.12).

Equivalently, gg is a Garnir tableau if and only if there is a unique k,  1kn1k,\;\;1\leq k\leq n-1 such that g(k)>g(k+1)g(k)>g(k+1) with respect to the order on nodes and this number kk is in the same column as k+1k+1. Here are some Garnir tableaux of shape λ=((12),(14),)\lambda=((1^{2}),(1^{4}),\emptyset):

213456321456143256\emptyset\emptyset\emptyset;;.
Figure 6. Garnir tableaux of λ\lambda

Comparing the three tableaux, we find that a Garnir tableau is neither uniquely determined by the positions nor by the numbers that cause the “non-standardness”. The next lemma shows that any non-standard tableaux can be transformed into a Garnir one.

Lemma 6.2.

[25, Corollary 22] Let tt be a non-standard tableau of shape λ\lambda. Then there exists a Garnir tableau gg and w𝔖nw\in\mathfrak{S}_{n} such that t=gwt=g\circ w and l(d(t))=l(d(g))+l(w)l(d(t))=l(d(g))+l(w).

Fix a Garnir tableau gg and a dominant weight Λ\Lambda, let e(g):=e(resΛ(g))e(g):=e(res^{\Lambda}(g)) be the idempotent of TLr,1,nΛTL_{r,1,n}^{\Lambda} where resΛ(g)res^{\Lambda}(g) is as defined in (4.5)(\ref{eqres}). The next lemma shows that e(g)e(g) is a composition of lower terms.

Lemma 6.3.

[25, Lemma 35] If gg is a Garnir tableau of shape λ\lambda, let e(g)=e(ResΛ(g))e(g)=e(Res^{\Lambda}(g)) then

(6.1) e(g)=μλDμe(g)=\sum_{\mu\triangleleft\lambda}D_{\mu}

where DμD_{\mu} is in the two-sided ideal of H(r,1,n)H(r,1,n) generated by eμe_{\mu}, which is the idempotent defined before Definition 4.5.

These two lemmas provide us a method to transform e(t)e(t) into a combination of lower terms when tt is not a standard tableau.

6.2. The bilinear form ϕ\phi and the irreducible representations of TLr,1,nTL_{r,1,n}

This subsection concerns the dimensions of the simple modules L(λ)L(\lambda) of TLr,1,nTL_{r,1,n}. According to Theorem 4.3, dim(L(λ))dim(L(\lambda)) equals to the rank of the bilinear form ϕλ(,)\phi_{\lambda}(,). We next introduce a theoretical method to calculate the value of the bilinear form ϕλ(s,t)\phi_{\lambda}(s,t) for s,tStd(λ)s,t\in Std(\lambda). In fact, we will not calculate any values of ϕλ(s,t)\phi_{\lambda}(s,t). This method is used to show the value of the bilinear form ϕλ(s,t)\phi_{\lambda}(s,t) equals to the one on the corresponding cell module of TLB(q,Q)TLB(q,Q).

We first introduce a notation which will be used when calculating the values of ϕλ(s,t)\phi_{\lambda}(s,t). Denote N={1,2,,n1}N=\{1,2,\dots,n-1\}. For U(k)=(U1(k),U2(k),,Uk(k))NkU^{(k)}=(U^{(k)}_{1},U^{(k)}_{2},\dots,U^{(k)}_{k})\in N^{k} where kk is a non-negative integer, let ψU(k)=ψU1(k)ψU2(k)ψUk(k)\psi_{U^{(k)}}=\psi_{U^{(k)}_{1}}\psi_{U^{(k)}_{2}}\dots\psi_{U^{(k)}_{k}} with ψi(iN)\psi_{i}(i\in N) being the KLR generators in Definition 2.9. We use the following equivalence relation to describe the difference between two elements in the cyclotomic KLR algebra.

Definition 6.4.

For any positive integer kk, define an equivalence relation k\stackrel{{\scriptstyle k}}{{\sim}} on NkN^{k} as follows: We write V(k)kU(k)V^{(k)}\stackrel{{\scriptstyle k}}{{\sim}}U^{(k)} if

ψV(k)eλψU(k)eλ=l<kc(W(l))ψW(l)eλ\psi_{V^{(k)}}^{*}e_{\lambda}-\psi_{U^{(k)}}^{*}e_{\lambda}=\sum_{l<k}c(W^{(l)})\psi_{W^{(l)}}^{*}e_{\lambda}

where c(W(l))Rc(W^{(l)})\in R and W(l)W^{(l)} runs over NlN^{l} for all 0l<k0\leq l<k.

The following lemma shows that two reduced expressions of the same element lead to equivalent sequences.

Lemma 6.5.

For w𝔖nw\in\mathfrak{S}_{n}, let sV(k)s_{V^{(k)}} and sU(k)s_{U^{(k)}} be two reduced expressions of ww where k=l(w)k=l(w), then V(k)kU(k)V^{(k)}\stackrel{{\scriptstyle k}}{{\sim}}U^{(k)}.

Proof.

As both sV(k)s_{V^{(k)}} and sU(k)s_{U^{(k)}} are reduced expressions of the same element ww, they can be transformed to each other by braid relations. These relations correspond to (2.27)(\ref{22eq}) and (2.31)(\ref{26eq}) in TLr,1,nΛTL_{r,1,n}^{\Lambda}. The error terms only occur in (2.31)(\ref{26eq}). They lead to strictly shorter sequences. By Definition 6.4, we have V(k)kU(k)V^{(k)}\stackrel{{\scriptstyle k}}{{\sim}}U^{(k)}. ∎

The following lemma provides us the main tool to eliminate the lower terms when calculating the value of the bilinear form.

Lemma 6.6.

For any non-negative integer kk, let U(k)U^{(k)} be a sequence of length kk and λ𝔅n(r)𝔇n(r)\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)}. Then ψU(k)eλ\psi_{U^{(k)}}^{*}e_{\lambda} can be transformed into the following form using the generating relations (2.23)(2.31)(\ref{18eq})-(\ref{26eq}):

(6.2) ψU(k)eλ=sStd(λ)cU(k)(s)ψd(s)eλ+i=1nYi+gGar(λ)Dg\psi_{U^{(k)}}^{*}e_{\lambda}=\sum_{s\in Std(\lambda)}c_{U^{(k)}}(s)\psi_{d(s)}^{*}e_{\lambda}+\sum_{i=1}^{n}Y_{i}+\sum_{g\in Gar(\lambda)}D_{g}

where YiY_{i} is in the two-sided ideal generated by yieλy_{i}e_{\lambda} and DgD_{g} is in the one generated by e(g)e(g) with gg running over the Garnir tableaux of shape λ\lambda.

Proof.

We prove this by induction on kk. If k=0k=0, we can get the form directly. Assume this is true for any sequence U(l)U^{(l)} with l<kl<k for a positive integer kk. For a sequence U(k)U^{(k)}, we consider three cases:

1. sU(k)s_{U^{(k)}} is a reduced expression of some w𝔖nw\in\mathfrak{S}_{n} and tλwt^{\lambda}\circ w is a standard tableau. Denote by ss the standard tableau tλwt^{\lambda}\circ w. Let V(k)V^{(k)} be the sequence such that ψV(k)\psi_{V^{(k)}}^{*} is the chosen ψd(s)\psi_{d(s)}^{*} in the definition of the cellular basis. Then sV(k)s_{V^{(k)}} and sU(k)s_{U^{(k)}} are two reduced expressions of ww with k=l(w)k=l(w), so we have V(k)kU(k)V^{(k)}\stackrel{{\scriptstyle k}}{{\sim}}U^{(k)} by Lemma 6.5. Thus,

(6.3) ψU(k)eλ=ψd(s)eλ+l<kc(W(l))ψW(l)eλ.\psi_{U^{(k)}}^{*}e_{\lambda}=\psi_{d(s)}^{*}e_{\lambda}+\sum_{l<k}c(W^{(l)})\psi_{W^{(l)}}^{*}e_{\lambda}.

By the inductive assumption, ψW(l)eλ\psi_{W^{(l)}}^{*}e_{\lambda} is of the form in (6.2)(\ref{89}). Therefore, ψU(k)eλ\psi_{U^{(k)}}^{*}e_{\lambda} can be transformed into the form in (6.2)(\ref{89}).

2. sU(k)s_{U^{(k)}} is a reduced expression of some w𝔖nw\in\mathfrak{S}_{n}, but tλwt^{\lambda}\circ w is not a standard tableau. By Lemma 6.2, there exists a Garnir tableau gg and an element w1𝔖nw_{1}\in\mathfrak{S}_{n} such that tλw=gw1t^{\lambda}\circ w=g\circ w_{1}. Let V(g)V(g) and V(w1)V(w_{1}) be two sequences consisting of numbers in {1,2,,n1}\{1,2,\dots,n-1\} such that sV(g)s_{V(g)} and sV(w1)s_{V}(w_{1}) are reduced expressions of d(g)d(g) and w1w_{1} respectively. Let V(k)=V(g)V(w1)V^{(k)}=V(g)V(w_{1}) be the combination of these two sequences. Then sV(k)s_{V^{(k)}} is a reduced expression of ww as well. By Lemma 6.5, V(k)kU(k)V^{(k)}\stackrel{{\scriptstyle k}}{{\sim}}U^{(k)}. So we have

ψU(k)eλ\displaystyle\psi_{U^{(k)}}^{*}e_{\lambda} =ψV(k)eλ+l<kc(W(l))ψW(l)eλ\displaystyle=\psi_{V^{(k)}}^{*}e_{\lambda}+\sum_{l<k}c(W^{(l)})\psi_{W^{(l)}}^{*}e_{\lambda}
=ψV(w1)ψV(g)eλ+l<kc(W(l))ψW(l)eλ\displaystyle=\psi_{V(w_{1})}^{*}\psi_{V(g)}^{*}e_{\lambda}+\sum_{l<k}c(W^{(l)})\psi_{W^{(l)}}^{*}e_{\lambda}
=ψV(w1)e(ig)ψV(g)+l<kc(W(l))ψW(l)eλ.\displaystyle=\psi_{V(w_{1})}^{*}e(i^{g})\psi_{V(g)}^{*}+\sum_{l<k}c(W^{(l)})\psi_{W^{(l)}}^{*}e_{\lambda}.

The first term is in the two-sided ideal generated by e(ig)e(i^{g}) and the other terms satisfy that the length of the sequence decreases strictly, so the inductive assumption can be used.

3. If sU(k)s_{U^{(k)}} is not a reduced expression of any w𝔖nw\in\mathfrak{S}_{n}, let ll be the largest number such that sU(k)|ls_{U^{(k)}|_{l}} is a reduced expression of some w𝔖nw\in\mathfrak{S}_{n} where U(k)|lU^{(k)}|_{l} is the sub-sequence of U(k)U^{(k)} consisting of the first ll terms. By the exchange condition, there exists a sequence V(l)V^{(l)} ending with Ul+1(k)U^{(k)}_{l+1} such that sV(l)s_{V^{(l)}} is a reduced expression for sU(k)|ls_{U^{(k)}|_{l}}. By Lemma 6.5, V(l)kU(k)|lV^{(l)}\stackrel{{\scriptstyle k}}{{\sim}}U^{(k)}|_{l}. Therefore, we have

ψU(k)eλ\displaystyle\psi_{U^{(k)}}^{*}e_{\lambda} =(ψU(k)|lψUl+1(k)ψU(kl1))eλ\displaystyle=(\psi_{U^{(k)}|_{l}}\psi_{U^{(k)}_{l+1}}\psi_{U^{(k-l-1)}})^{*}e_{\lambda}
=(ψV(l)ψUl+1(k)ψU(kl1))eλ+l<kc(W(l))ψW(l)eλ\displaystyle=(\psi_{V^{(l)}}\psi_{U^{(k)}_{l+1}}\psi_{U^{(k-l-1)}})^{*}e_{\lambda}+\sum_{l<k}c(W^{(l)})\psi_{W^{(l)}}^{*}e_{\lambda}
=(ψV(l)|l1ψUl+1(k)2ψU(kl1))eλ+l<kc(W(l))ψW(l)eλ\displaystyle=(\psi_{V^{(l)}|_{l-1}}\psi_{U^{(k)}_{l+1}}^{2}\psi_{U^{(k-l-1)}})^{*}e_{\lambda}+\sum_{l<k}c(W^{(l)})\psi_{W^{(l)}}^{*}e_{\lambda}
=ψU(kl1)e(i)ψUl+1(k)2ψV(l)|l1+l<kc(W(l))ψW(l)eλ.\displaystyle=\psi_{U^{(k-l-1)}}^{*}e(i)\psi_{U^{(k)}_{l+1}}^{2}\psi_{V^{(l)}|_{l-1}}^{*}+\sum_{l<k}c(W^{(l)})\psi_{W^{(l)}}^{*}e_{\lambda}.

For the same reason as above, we only need to deal with the first term. By (2.30), e(i)ψUl+1(k)2e(i)\psi_{U^{(k)}_{l+1}}^{2} is 0, e(i)e(i), (ys+1ys)e(i)(y_{s+1}-y_{s})e(i) or (ysys+1)e(i)(y_{s}-y_{s+1})e(i). The first case is trivial. The second case leads to a strictly shorter sequence. If we get (ys+1ys)e(i)(y_{s+1}-y_{s})e(i) or (ysys+1)e(i)(y_{s}-y_{s+1})e(i), (2.28) can help to move ysy_{s} to the right to get some elements YiY_{i}. The error term will also lead to a strictly shorter sequence which is covered by the inductive assumption.

Therefore, for any sequence U(k)U^{(k)} of length kk, ψU(k)\psi_{U^{(k)}} can be transformed into the form in (6.2)(\ref{89}). Thus the lemma has been proved. ∎

It should be remarked that the form we get in the lemma above does not depend on the dominant weight Λ\Lambda, given the fact that only the generating relations (2.23)(2.31)(\ref{18eq})-(\ref{26eq}) are used in the procedure. Therefore, this lemma builds a bridge between a general TLr,1,nΛTL_{r,1,n}^{\Lambda} and a Temperley-Lieb algebra of type BnB_{n}.

To be more precise, let Λ=Λi1+Λi2++Λir\Lambda=\Lambda_{i_{1}}+\Lambda_{i_{2}}+\dots+\Lambda_{i_{r}} be a dominant weight and TLr,1,nΛTL_{r,1,n}^{\Lambda} be the corresponding generalised Temperley-Lieb algebra. Let λ𝔅n(r)𝔇n(r)\lambda\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)} be a multipartition in which the uthu^{th} and vthv^{th} components are non-empty. Denote by W(λ)W(\lambda) the cell module of TLr,1,nΛTL_{r,1,n}^{\Lambda} corresponding to λ\lambda and by ϕλ\phi_{\lambda} the bilinear form on W(λ)W(\lambda).

Let Λ=Λiu+Λiv\Lambda^{\prime}=\Lambda_{i_{u}}+\Lambda_{i_{v}} and TLBnΛTLB_{n}^{\Lambda^{\prime}} be the corresponding Temperley-Lieb algebra of Type BnB_{n}. Let λ\lambda^{\prime} be the bipartition with the same shape as the non-empty part of λ\lambda. For example, if λ=((1a),,(1na),,,)\lambda=((1^{a}),\emptyset,(1^{n-a}),\emptyset,\dots,\emptyset) is the multipartition of nn with aa nodes in the first partition and nan-a in the third, the corresponding bipartition λ=((1a),(1na))\lambda^{\prime}=((1^{a}),(1^{n-a})) and the dominant weight Λ=Λi1+Λi3\Lambda^{\prime}=\Lambda_{i_{1}}+\Lambda_{i_{3}}.

Denote by W(λ)W(\lambda^{\prime}) and ϕλ\phi_{\lambda^{\prime}} the corresponding cell module of TLBnΛTLB_{n}^{\Lambda^{\prime}} and the bilinear form. Let ff be the natural map from Std(λ)Std(\lambda) to Std(λ)Std(\lambda^{\prime}). Then we have

Corollary 6.7.

ϕλ(s,t)=ϕλ(f(s),f(t))\phi_{\lambda}(s,t)=\phi_{\lambda^{\prime}}(f(s),f(t)), for all s,tStd(λ)s,t\in Std(\lambda).

Proof.

By definition, we have

ϕλ(s,t)eλ=eλψd(s)ψd(t)eλ mod TLr,1,nλ.\phi_{\lambda}(s,t)e_{\lambda}=e_{\lambda}\psi_{d(s)}\psi_{d(t)}^{*}e_{\lambda}\text{ mod }TL_{r,1,n}^{\triangleleft\lambda}.

and

ϕλ(f(s),f(t))eλ=eλψd(f(s))ψd(f(t))eλ mod TLBnλ.\phi_{\lambda^{\prime}}(f(s),f(t))e_{\lambda^{\prime}}=e_{\lambda^{\prime}}\psi_{d(f(s))}\psi_{d(f(t))}^{*}e_{\lambda^{\prime}}\text{ mod }TLB_{n}^{\triangleleft\lambda^{\prime}}.

The right hand sides of the two equations share the same expression. But this is not enough to show ϕλ(s,t)=ϕλ(f(s),f(t))\phi_{\lambda}(s,t)=\phi_{\lambda^{\prime}}(f(s),f(t)) because TLr,1,nλTLBnλTL_{r,1,n}^{\triangleleft\lambda}\neq TLB_{n}^{\triangleleft\lambda^{\prime}}. By the fact that e(i)e(i)’s are orthogonal idempotents, we have eλψd(s)ψd(t)eλ=ψd(s)ψd(t)eλe_{\lambda}\psi_{d(s)}\psi_{d(t)}^{*}e_{\lambda}=\psi_{d(s)}\psi_{d(t)}^{*}e_{\lambda} in both TLr,1,nΛTL_{r,1,n}^{\Lambda} and TLBnΛTLB_{n}^{\Lambda^{\prime}} if eλψd(s)ψd(t)eλ0e_{\lambda}\psi_{d(s)}\psi_{d(t)}^{*}e_{\lambda}\neq 0 . Lemma (6.6)(\ref{algo}) implies that the right hand sides of the two equations can be transformed into the same form:

(6.4) ψd(s)ψd(t)eλ=cs,teλ+i=1nYi+gGar(λ)Dg.\psi_{d(s)}\psi_{d(t)}^{*}e_{\lambda}=c_{s,t}e_{\lambda}+\sum_{i=1}^{n}Y_{i}+\sum_{g\in Gar(\lambda)}D_{g}.

As the transformation only depends on the generating relations (2.23)(2.31)(\ref{18eq})-(\ref{26eq}) which are shared by TLr,1,nΛTL_{r,1,n}^{\Lambda} and TLBnΛTLB_{n}^{\Lambda^{\prime}}, we have cs,t=cf(s),f(t)c_{s,t}=c_{f(s),f(t)}. Lemma 4.7 and Lemma 6.3 show that YiY_{i} and DgD_{g} are in TLr,1,nλTL_{r,1,n}^{\triangleleft\lambda} and TLBnλTLB_{n}^{\triangleleft\lambda^{\prime}}. So we have

ϕλ(s,t)=cs,t\phi_{\lambda}(s,t)=c_{s,t}

and

ϕλ(f(s),f(t))=cf(s),f(t).\phi_{\lambda^{\prime}}(f(s),f(t))=c_{f(s),f(t)}.

Therefore, we have

ϕλ(s,t)=ϕλ(f(s),f(t)).\phi_{\lambda}(s,t)=\phi_{\lambda^{\prime}}(f(s),f(t)).

Further, let L(λ)L(\lambda) and L(λ)L(\lambda^{\prime}) be the simple modules of TLr,1,nΛTL_{r,1,n}^{\Lambda} and TLBnΛTLB_{n}^{\Lambda^{\prime}} respectively in Theorem 4.3. As a direct consequence, we have

Corollary 6.8.

dim(L(λ))=dim(L(λ))dim(L(\lambda))=dim(L(\lambda^{\prime})).

6.3. Decomposition numbers

We next concentrate on the decomposition numbers of TLr,1,nΛTL_{r,1,n}^{\Lambda}. In a similar way to the observation above, we claim that the decomposition numbers can be obtained from those of TLBnΛTLB_{n}^{\Lambda^{\prime}} where Λ\Lambda^{\prime} is the dominant weight defined in terms of the cell module CλC_{\lambda} in the last subsection. To prove this, we show that the cell module W(λ)W(\lambda) and the simple module L(λ)L(\lambda) of TLr,1,nΛTL_{r,1,n}^{\Lambda} are isomorphic to those of a Temperley-Lieb algebra of type BnB_{n} as TLBnΛTLB_{n}^{\Lambda^{\prime}}-modules.

Lemma 6.9.

TLBnΛTLB_{n}^{\Lambda^{\prime}} is a quotient of TLr,1,nΛTL_{r,1,n}^{\Lambda} by the two-sided ideal generated by all e(𝐢)=e(𝐢1,𝐢2,,𝐢n)e(\mathbf{i})=e(\mathbf{i}_{1},\mathbf{i}_{2},\dots,\mathbf{i}_{n}) such that (Λ,α𝐢1)=0(\Lambda^{\prime},\alpha_{\mathbf{i}_{1}})=0.

Proof.

By comparing the generators and relations in Definition 2.9, we see that nΛ\mathcal{R}_{n}^{\Lambda^{\prime}} is a quotient of nΛ\mathcal{R}_{n}^{\Lambda} by the two-sided ideal generated by all e(𝐢)e(\mathbf{i}) such that (Λ,α𝐢1)=0(\Lambda^{\prime},\alpha_{\mathbf{i}_{1}})=0.

Fix an idempotent e(i1,i2,i3,𝐣)nΛe(i_{1},i_{2},i_{3},\mathbf{j}^{\prime})\in\mathcal{R}_{n}^{\Lambda} where (αia,Λ)>0(\alpha_{i_{a}},\Lambda)>0 for a=1,2,3a=1,2,3. We next show e(i1,i2,i3,𝐣)e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime}) is in the two-sided ideal of nΛ\mathcal{R}_{n}^{\Lambda},

e(𝐢),e(i,i+1,𝐣),e(i1,i2,i3,𝐣)|(αi,Λ)>0,(αij,Λ)>0,\langle e(\mathbf{i}),e(i,i+1,\mathbf{j}),e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime})|(\alpha_{i},\Lambda^{\prime})>0,(\alpha_{i_{j}},\Lambda^{\prime})>0\rangle,

where 𝐢\mathbf{i} runs over InI^{n} such that (Λ,α𝐢1)=0(\Lambda^{\prime},\alpha_{\mathbf{i}_{1}})=0.

By Theorem 3.7 we have

TLBnΛ\displaystyle TLB_{n}^{\Lambda^{\prime}} =nΛ/e(i,i+1,𝐣),e(i1,i2,i3,𝐣)|(αi,Λ)>0,(αij,Λ)>0 for j=1,2,3\displaystyle=\mathcal{R}_{n}^{\Lambda^{\prime}}/\langle e(i,i+1,\mathbf{j}),e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime})|(\alpha_{i},\Lambda^{\prime})>0,(\alpha_{i_{j}},\Lambda^{\prime})>0\text{ for }j=1,2,3\rangle
=nΛ/e(𝐢),e(i,i+1,𝐣),e(i1,i2,i3,𝐣)|(αi,Λ)>0,(αij,Λ)>0\displaystyle=\mathcal{R}_{n}^{\Lambda}/\langle e(\mathbf{i}),e(i,i+1,\mathbf{j}),e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime})|(\alpha_{i},\Lambda^{\prime})>0,(\alpha_{i_{j}},\Lambda^{\prime})>0\rangle

where 𝐢\mathbf{i} runs over InI^{n} such that (Λ,α𝐢1)=0(\Lambda^{\prime},\alpha_{\mathbf{i}_{1}})=0. Let ff be the quotient map from nΛ\mathcal{R}_{n}^{\Lambda} to TLBnΛTLB_{n}^{\Lambda^{\prime}}. It is enough to show that f(e(i1,i2,i3,𝐣))=0f(e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime}))=0. By the same method as that in the proof of Lemma 4.8, we can get f(e(i1,i2,i3,𝐣))Cs,tλ=0f(e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime}))C_{s,t}^{\lambda}=0 where Cs,tλC_{s,t}^{\lambda} runs over the cellular basis of TLBnΛTLB_{n}^{\Lambda^{\prime}} in Theorem 4.9. So e(i1,i2,i3,𝐣)e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime}) is in the corresponding two-sided ideal of nΛ\mathcal{R}_{n}^{\Lambda}, e(𝐢),e(i,i+1,𝐣),e(i1,i2,i3,𝐣)|(αi,Λ)>0,(αij,Λ)>0\langle e(\mathbf{i}),e(i,i+1,\mathbf{j}),e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime})|(\alpha_{i},\Lambda^{\prime})>0,(\alpha_{i_{j}},\Lambda^{\prime})>0\rangle.

Therefore, we have

TLBnΛ\displaystyle TLB_{n}^{\Lambda^{\prime}} =nΛ/e(𝐢),e(i,i+1,𝐣),e(i1,i2,i3,𝐣)|(αi,Λ)>0,(αij,Λ)>0\displaystyle=\mathcal{R}_{n}^{\Lambda}/\langle e(\mathbf{i}),e(i,i+1,\mathbf{j}),e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime})|(\alpha_{i},\Lambda^{\prime})>0,(\alpha_{i_{j}},\Lambda^{\prime})>0\rangle
=nΛ/e(𝐢),e(i,i+1,𝐣),e(i1,i2,i3,𝐣)|(αi,Λ)>0,(αij,Λ)>0\displaystyle=\mathcal{R}_{n}^{\Lambda}/\langle e(\mathbf{i}),e(i,i+1,\mathbf{j}),e(i_{1},i_{2},i_{3},\mathbf{j}^{\prime})|(\alpha_{i},\Lambda)>0,(\alpha_{i_{j}},\Lambda)>0\rangle
=TLr,1,nΛ/e(𝐢)\displaystyle=TL_{r,1,n}^{\Lambda}/\langle e(\mathbf{i})\rangle

where 𝐢\mathbf{i} runs over InI^{n} such that (Λ,α𝐢1)=0(\Lambda^{\prime},\alpha_{\mathbf{i}_{1}})=0. ∎

We next show that further, the cell module W(λ)W(\lambda) can be regarded as a module of the quotient algebra TLBnΛTLB_{n}^{\Lambda^{\prime}}.

Lemma 6.10.

Let W(λ)W(\lambda) be the cell module of TLr,1,nΛTL_{r,1,n}^{\Lambda} defined above. Then W(λ)W(\lambda) is a TLBnΛTLB_{n}^{\Lambda^{\prime}}-module.

Proof.

For tStd(λ)t\in Std(\lambda), let e(t)=e(j1,j2,,jn)e(t)=e(j_{1},j_{2},\dots,j_{n}). Then by definition, j1=iu or ivj_{1}=i_{u}\text{ or }i_{v}. For any e(𝐢)e(\mathbf{i}) such that (Λ,α𝐢1)=0(\Lambda^{\prime},\alpha_{\mathbf{i}_{1}})=0, we have 𝐢1iu or iv\mathbf{i}_{1}\neq i_{u}\text{ or }i_{v}. So

(6.5) e(𝐢)ψd(t)eλ=e(𝐢)e(t)ψd(t)=0e(\mathbf{i})\psi_{d(t)}^{*}e_{\lambda}=e(\mathbf{i})e(t)\psi_{d(t)}^{*}=0

which implies re(𝐢)(t,t)=0r_{e(\mathbf{i})}(t^{\prime},t)=0 for all t,tStd(λ)t,t^{\prime}\in Std(\lambda). Therefore,

(6.6) e(𝐢)W(λ)=0e(\mathbf{i})W(\lambda)=0

for all e(𝐢)e(\mathbf{i}) such that (Λ,α𝐢1)=0(\Lambda^{\prime},\alpha_{\mathbf{i}_{1}})=0. That is, the ideal which defines TLBnΛTLB_{n}^{\Lambda^{\prime}} as a quotient of TLr,1,nΛTL_{r,1,n}^{\Lambda} acts trivially on W(λ)W(\lambda). Hence W(λ)W(\lambda) is a TLBnΛTLB_{n}^{\Lambda^{\prime}}-module. ∎

As a direct consequence, we have that the simple module L(λ)=W(λ)/rad(ϕλ)L(\lambda)=W(\lambda)/rad(\phi_{\lambda}) is a module of TLBnΛTLB_{n}^{\Lambda^{\prime}}:

Lemma 6.11.

Let L(λ)L(\lambda) be the simple module of TLr,1,nΛTL_{r,1,n}^{\Lambda}. Then L(λ)L(\lambda) is a TLBnΛTLB_{n}^{\Lambda^{\prime}}-module.

Further, the next lemma shows that the two TLBnΛTLB_{n}^{\Lambda^{\prime}}-modules obtained in the last two lemmas are exactly the corresponding cell and simple modules corresponding to an appropriate λ\lambda^{\prime}.

Lemma 6.12.

Let λ\lambda^{\prime} be the bipartition with the same shape as the non-empty part of λ\lambda and W(λ)W(\lambda^{\prime}), L(λ)L(\lambda^{\prime}) be the cell module and simple module of TLBnΛTLB_{n}^{\Lambda^{\prime}} corresponding to λ\lambda^{\prime} respectively. Then W(λ)W(λ)W(\lambda^{\prime})\simeq W(\lambda), L(λ)L(λ)L(\lambda^{\prime})\simeq L(\lambda) as TLBnΛTLB_{n}^{\Lambda^{\prime}}-modules.

Proof.

We first prove W(λ)W(λ)W(\lambda^{\prime})\simeq W(\lambda). Let f:Std(λ)Std(λ)f:Std(\lambda)\to Std(\lambda^{\prime}) be the natural bijection and F:W(λ)W(λ)F:W(\lambda)\to W(\lambda^{\prime}) be the induced map. We will prove that FF is an isomorphism between the two TLBnΛTLB_{n}^{\Lambda^{\prime}}-modules. Since FF is an isomorphism of vector spaces, it is enough to show that FF is a module homomorphism, which is to show F(aCt)=aF(Ct)F(aC_{t})=aF(C_{t}) for all aa in the generating set of TLBnΛTLB_{n}^{\Lambda^{\prime}} and tStd(λ)t\in Std(\lambda). Let ra(s,t)r_{a}(s,t) and ra(f(s),f(t))r^{\prime}_{a}(f(s),f(t)) be the coefficients in the equation (4.1) corresponding to TLr,1,nΛTL_{r,1,n}^{\Lambda} and TLBnΛTLB_{n}^{\Lambda^{\prime}} respectively. We next show that ra(s,t)=ra(f(s),f(t))r_{a}(s,t)=r^{\prime}_{a}(f(s),f(t)) for all aa in the generating set of TLBnΛTLB_{n}^{\Lambda^{\prime}} and deduce that FF is a module homomorphism.

If a=e(𝐢)a=e(\mathbf{i}), we have

ra(s,t)=ra(f(s),f(t))={1 if s=t and a=e(t)0 otherwise.r_{a}(s,t)=r^{\prime}_{a}(f(s),f(t))=\begin{cases}1&\text{ if }s=t\text{ and }a=e(t)\\ 0&\text{ otherwise.}\end{cases}

If a=ψia=\psi_{i}, for tStd(λ)t\in Std(\lambda), denote by U(k)U^{(k)} be the sequence such that ψUi,t(k)=ψiψd(t)\psi_{U_{i,t}^{(k)}}^{*}=\psi_{i}\psi_{d(t)}^{*}. Then by Lemma 6.6, ψiψd(t)eλ\psi_{i}\psi_{d(t)}^{*}e_{\lambda} and ψiψd(f(t))eλ\psi_{i}\psi_{d(f(t))}^{*}e_{\lambda^{\prime}} can be written in the form of (6.2)(\ref{89}) with

ci,t(s)=ci,f(t)(f(s)).c_{i,t}(s)=c_{i,f(t)}(f(s)).

So we have

ra(s,t)=ci,t(s)=ci,f(t)(f(s))=ra(f(s),f(t)).r_{a}(s,t)=c_{i,t}(s)=c_{i,f(t)}(f(s))=r^{\prime}_{a}(f(s),f(t)).

If a=yia=y_{i}, for tStd(λ)t\in Std(\lambda), by (2.26),(2.28) and (2.29), we have

(6.7) yiψd(t)eλ=ψd(t)yjeλ+l(V)<l(d(t))c(V)ψVeλ,y_{i}\psi_{d(t)}^{*}e_{\lambda}=\psi_{d(t)}^{*}y_{j}e_{\lambda}+\sum_{l(V)<l({d(t)})}c(V)\psi_{V}^{*}e_{\lambda},

where c(V)c(V) is 1 or -1. Then by Lemma 6.6, yiψd(t)eλy_{i}\psi_{d(t)}^{*}e_{\lambda} and yiψd(f(t))eλy_{i}\psi_{d(f(t))}^{*}e_{\lambda^{\prime}} can be written in the form of (6.2)(\ref{89}) with the same coefficients. So we have

ra(s,t)=cyi,t(s)=cyi,f(t)(f(s))=ra(f(s),f(t)).r_{a}(s,t)=c_{y_{i},t}(s)=c_{y_{i},f(t)}(f(s))=r^{\prime}_{a}(f(s),f(t)).

To summarise, ra(s,t)=ra(f(s),f(t))r_{a}(s,t)=r^{\prime}_{a}(f(s),f(t)) for all aa in the generating set of TLBnΛTLB_{n}^{\Lambda^{\prime}}. Thus we have F(aCt)=aF(Ct)F(aC_{t})=aF(C_{t}) which implies FF is an isomorphism between the two TLBnΛTLB_{n}^{\Lambda^{\prime}}-modules.

By Corollary 6.7, ϕλ(s,t)=ϕλ(f(s),f(t))\phi_{\lambda}(s,t)=\phi_{\lambda^{\prime}}(f(s),f(t)). Thus ff induces an isomorphism between the two TLBnΛTLB_{n}^{\Lambda^{\prime}}-modules L(λ)L(\lambda^{\prime}) and L(λ)L(\lambda) as well. ∎

The following theorem shows the connection between the decomposition numbers for a generalised Temperley-Lieb algebra and those for a blob algebra.

Theorem 6.13.

Let TLr,1,nΛTL_{r,1,n}^{\Lambda} be a generalised Temperley-Lieb algebra (cf. Theorem 3.7) and λ,μ𝔅n(r)𝔇n(r)\lambda,\mu\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)} be two multipartitions of which all the components except for the uthu^{th} and vthv^{th} are empty. Let TLBnΛTLB_{n}^{\Lambda^{\prime}} and λ,μ\lambda^{\prime},\mu^{\prime} be as defined before Corollary 6.7. We have [W(λ),L(μ)]=[W(λ),L(μ)]TLB[W(\lambda),L(\mu)]=[W(\lambda^{\prime}),L(\mu^{\prime})]_{TLB} where the left hand side is the decomposition number for TLr,1,nΛTL_{r,1,n}^{\Lambda} and the right hand side is that for TLBnΛTLB_{n}^{\Lambda^{\prime}}.

Proof.

Lemma 6.12 shows that L(μ)L(\mu) is a simple TLBnΛTLB_{n}^{\Lambda^{\prime}}-module and [W(λ),L(μ)]TLB=[W(λ),L(μ)]TLB[W(\lambda),L(\mu)]_{TLB}=[W(\lambda^{\prime}),L(\mu^{\prime})]_{TLB} as TLBnΛTLB_{n}^{\Lambda^{\prime}}-modules. The proof of Lemma 6.10 shows that the kernel of the quotient map from TLr,1,nΛTL_{r,1,n}^{\Lambda} to TLBnΛTLB_{n}^{\Lambda^{\prime}} acts trivially on both W(λ)W(\lambda^{\prime}) and L(μ)L(\mu^{\prime}). So HomTLr,1,nΛ(W(λ),L(μ))=HomTLBnΛ(W(λ),L(μ))Hom_{TL_{r,1,n}^{\Lambda}}(W(\lambda),L(\mu))=Hom_{TLB_{n}^{\Lambda^{\prime}}}(W(\lambda),L(\mu)). Therefore, we have [W(λ),L(μ)]=[W(λ),L(μ)]TLB[W(\lambda),L(\mu)]=[W(\lambda^{\prime}),L(\mu^{\prime})]_{TLB}. ∎

The non-graded decomposition numbers of the Temperley-Lieb algebras of type BnB_{n} are given by Martin and Woodcock in [24] and Graham and Lehrer in [14] and the graded ones are given by Plaza in [27]. We give an interpretation of their result.

Theorem 6.14.

(cf.cf. [14],Theorem 10.16) Let TLBnΛTLB_{n}^{\Lambda^{\prime}} the Temperley-Lieb algebra of type BnB_{n} corresponding to the dominant weight Λ\Lambda^{\prime} (cf. Theorem 3.7) and λ,μ\lambda^{\prime},\mu^{\prime} be two bipartitions of nn. Then the decomposition numbers for the cell module W(λ)W(\lambda^{\prime}) of TLBnΛTLB_{n}^{\Lambda^{\prime}},

(6.8) [W(λ),L(μ)]TLB={1 if λμ and there exists tStd(λ) such that e(t)=eμ0 otherwise,[W(\lambda^{\prime}),L(\mu^{\prime})]_{TLB}=\begin{cases}1&\text{ if }\lambda^{\prime}\unlhd\mu^{\prime}\text{ and there exists }t\in Std(\lambda^{\prime})\text{ such that }e(t)=e_{\mu^{\prime}}\\ 0&\text{ otherwise,}\end{cases}

where e(t)e(t) and eμe_{\mu^{\prime}} are the KLR generators in Definition 4.4.

We can now determine the decomposition numbers for the generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n}.

Theorem 6.15.

Let TLr,1,nΛTL_{r,1,n}^{\Lambda} be the generalised Temperley-Lieb algebra (cf. Definition 3.1 and Theorem 3.7) over a field RR of characteristic 0, defined by a dominant weight Λ\Lambda which satisfies (3.5)(\ref{reslmd}). According to Theorem 4.9, it is a graded cellular algebra with the cell datum (𝔅n(r)𝔇n(r),,Std,C,deg)(\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)},*,Std,C,deg). For λ,μ𝔅n(r)𝔇n(r)\lambda,\mu\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)}, let W(λ)W(\lambda) be the cell module and L(μ)L(\mu) be the simple of TLr,1,nΛTL_{r,1,n}^{\Lambda} corresponding to λ\lambda and μ\mu, respectively. Then the decomposition number

(6.9) [W(λ),L(μ)]={1 if λμ and there exists tStd(λ) such that e(t)=eμ0 otherwise,[W(\lambda),L(\mu)]=\begin{cases}1&\text{ if }\lambda\unlhd\mu\text{ and there exists }t\in Std(\lambda)\text{ such that }e(t)=e_{\mu}\\ 0&\text{ otherwise,}\end{cases}

where e(t)e(t) and eμe_{\mu^{\prime}} are the KLR generators in Definition 4.4.

Proof.

If the set of indices of non-empty components of μ\mu is a subset of that of λ\lambda, Theorem 6.13 implies that [W(λ),L(μ)]=1[W(\lambda),L(\mu)]=1 if λμ and there exists tStd(λ)\lambda\unlhd\mu\text{ and there exists }t\in Std(\lambda) such that e(t)=eμe(t)=e_{\mu}. By comparing the dimensions, we can see that these 1’s are the only non-zero decomposition numbers for TLr,1,nΛTL_{r,1,n}^{\Lambda}. If the indices of non-empty components of μ\mu are not covered by those of λ\lambda, by comparing the first two indices in e(t)e(t), we notice there is no standard tableau tt of shape λ\lambda such that e(t)=eμe(t)=e_{\mu}. ∎

It should be remarked that the former theorem shows that for any λ,μ𝔅n(r)𝔇n(r)\lambda,\mu\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)}, if the indices of non-empty components of μ\mu are not covered by those of λ\lambda, then [W(λ),L(μ)]=0[W(\lambda),L(\mu)]=0. Further, we have the following result:

Corollary 6.16.

Let λk1,k2,μl1,l2𝔅n(r)𝔇n(r)\lambda_{k_{1},k_{2}},\mu_{l_{1},l_{2}}\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)} be two one-column multipartitions of nn such that the k1k_{1} and k2k_{2} components in λk1,k2\lambda_{k_{1},k_{2}} and l1l_{1} and l2l_{2} components in μl1,l2\mu_{l_{1},l_{2}} are non-empty. If {k1,k2}{l1,l2}=\{k_{1},k_{2}\}\cap\{l_{1},l_{2}\}=\emptyset, then

Hom(W(λk1,k2),W(μl1,l2))=0.Hom(W(\lambda_{k_{1},k_{2}}),W(\mu_{l_{1},l_{2}}))=0.
Proof.

For any ν𝔅n(r)𝔇n(r)\nu\in\mathfrak{B}_{n}^{(r)}-\mathfrak{D}_{n}^{(r)}, as {k1,k2}{l1,l2}=\{k_{1},k_{2}\}\cap\{l_{1},l_{2}\}=\emptyset, the theorem above implies

(6.10) [W(λk1,k2),L(ν)]×[W(μl1,l2),L(ν)]=0.[W(\lambda_{k_{1},k_{2}}),L(\nu)]\times[W(\mu_{l_{1},l_{2}}),L(\nu)]=0.

If Hom(W(λk1,k2),W(μl1,l2))0Hom(W(\lambda_{k_{1},k_{2}}),W(\mu_{l_{1},l_{2}}))\neq 0, then Hom(W(λk1,k2),L(ν))0Hom(W(\lambda_{k_{1},k_{2}}),L(\nu))\neq 0 for some ν\nu such that [W(μl1,l2),L(ν)]0[W(\mu_{l_{1},l_{2}}),L(\nu)]\neq 0. On the other hand, Hom(W(λk1,k2),L(ν))0Hom(W(\lambda_{k_{1},k_{2}}),L(\nu))\neq 0 implies that [W(λk1,k2),L(ν)]0[W(\lambda_{k_{1},k_{2}}),L(\nu)]\neq 0, which is contradictory to (6.10)(\ref{eq96}). ∎

7. Further problems

One of the most important properties of the Temperley-Lieb algebras is that they may be described in terms of planar diagrams. In [19] Kauffman shows that the Temperley-Lieb algebra TLn(q)TL_{n}(q) has a basis consisting of planar (n,n)(n,n)-diagrams. In [22], Martin and Saleur define the Temperley-Lieb algebra of type BnB_{n} as an associative algebra with blob (n,n)(n,n)-diagrams as a basis.

It is still an open question how to find a diagrammatic description of our generalised Temperley-Lieb algebra TLr,1,nTL_{r,1,n}. Graham and Lehrer reveal the connection between cellular bases and diagrammatic bases of Temperley-Lieb algebras of type BnB_{n} and affine Temperley-Lieb algebras in [14]. Theorem 4.9 gives us a cellular basis of our TLr,1,nTL_{r,1,n} with respect to the KLR generators. Although there is natural grading with this basis, a direct connection to blob diagrams is still not clear.

Another question is whether there is a further generalisation of the Temperley-Lieb algebras corresponding to all unitary reflection groups. We have made a significant progress in this direction. For any imprimitive complex reflection group G(r,p,n)G(r,p,n), the corresponding Temperley-Lieb algebra TLr,p,nTL_{r,p,n} can be defined in a similar way. This work will be described in an upcoming work.

References

  • [1] S. Abramsky. Temperley–Lieb algebra: From knot theory to logic and computation via quantum mechanics. In Mathematics of quantum computation and quantum technology, pages 533–576. Chapman and Hall/CRC, 2007.
  • [2] G. E. Andrews, R. J. Baxter, and P. J. Forrester. Eight-vertex SOS model and generalized Rogers-Ramanujan-type identities. Journal of Statistical Physics, 35:193–266, 1984.
  • [3] S. Ariki. On the semi-simplicity of the Hecke algebra of /r𝔖n\mathbb{Z}/r\mathbb{Z}\wr\mathfrak{S}_{n}. Journal of Algebra, 169:216–225, 1994.
  • [4] S. Ariki and K. Koike. A Hecke algebra of /r𝔖n\mathbb{Z}/r\mathbb{Z}\wr\mathfrak{S}_{n} and construction of its irreducible representations. Advances in Mathematics, 106:216–243, 1994.
  • [5] M. T. Batchelor and A. Kuniba. Temperley-Lieb lattice models arising from quantum groups. Journal of Physics A: Mathematical and General, 24:2599–2614, 1991.
  • [6] J. Bernstein, I. Frenkel, and M. Khovanov. A categorification of the Temperley-Lieb algebra and Schur quotients of U(sl2)U(sl_{2}) via projective and Zuckerman functors. Selecta Mathematica, 5:199–241, 1999.
  • [7] C. Bowman. The many graded cellular bases of Hecke algebras. American Journal of Mathematics, 144:437–504, 2022.
  • [8] M. Broué, G. Malle, and R. Rouquier. Complex reflection groups, braid groups, Hecke algebras . J. Reine Angew. Math., 1998:127–190, 1998.
  • [9] J. Brundan. Centers of degenerate cyclotomic Hecke algebras and parabolic category 𝒪\mathcal{O}. Representation Theory of The American Mathematical Society, 12:236–259, 2008.
  • [10] J. Brundan and A. Kleshchev. Blocks of cyclotomic Hecke algebras and Khovanov-Lauda algebras. Inventiones mathematicae, 178:451–484, 2009.
  • [11] J. Brundan and A. Kleshchev. Graded decomposition numbers for cyclotomic Hecke algebras. Advances in Mathematics, 222:1883–1942, 2009.
  • [12] A. Cox, J. J. Graham, and P.P. Martin. The blob algebra in positive characteristic. Journal of Algebra, 266:584–635, 2003.
  • [13] J.J. Graham and G.I. Lehrer. Cellular algebras. Inventiones mathematicae, 123:1–34, 1996.
  • [14] J.J. Graham and G.I. Lehrer. Diagram algebras, Hecke algebras and decomposition numbers at roots of unity. Annales Scientifiques de l’École Normale Supérieure, 36:479–524, 2003.
  • [15] I. Grojnowski. Affine slpsl_{p} controls the representation theory of the symmetric group and related Hecke algebras. arXiv: math/9907129, 1999.
  • [16] J. Hu and A. Mathas. Graded cellular bases for the cyclotomic Khovanov-Lauda-Rouquier algebras of type AA. Advances in Mathematics, 225:598–642, 2010.
  • [17] K. Iohara, G.I. Lehrer, and R.B. Zhang. Equivalence of a tangle category and a category of infinite dimensional Uq(𝔰𝔩2)\mathrm{U}_{q}(\mathfrak{sl}_{2})-modules. Representation Theory of the American Mathematical Society, 25:265–299, 2021.
  • [18] V. F. R. Jones. A polynomial invariant for knots via von Neumann algebras. Bulletin (New Series) of the American Mathematical Society, 12:103–111, 1985.
  • [19] L. H. Kauffman. An invariant of regular isotopy. Transactions of the American Mathematical Society, 318:417–471, 1990.
  • [20] M. Khovanov and A. D. Lauda. A diagrammatic approach to categorification of quantum groups I. Representation Theory, 13:309–347, 2009.
  • [21] S. Lyle and A. Mathas. Blocks of cyclotomic Hecke algebras. Advances in Mathematics, 216:854–878, 2007.
  • [22] P. P. Martin and H.Saleur. The blob algebra and the periodic Temperley-Lieb algebra. Letters in Mathematical Physics, 30:189–206, 1994.
  • [23] P. P. Martin and D. Woodcock. On the structure of the blob algebra. Journal of Algebra, 225:957–988, 2000.
  • [24] P. P. Martin and D. Woodcock. Generalized blob algebras and alcove geometry. LMS Journal of Computation and Mathematics, 6:249–296, 2003.
  • [25] D. Lobos Maturana and S. Ryom-Hansen. Graded cellular basis and Jucys-Murphy elements for generalized blob algebras. Journal of Pure and Applied Algebra, 224, 2020.
  • [26] G.E. Murphy. The representations of Hecke algebras of type AnA_{n}. Journal of Algebra, 173:97–121, 1995.
  • [27] D. Plaza. Graded decomposition numbers for the blob algebra. Journal of Algebra, 394:182–206, 2013.
  • [28] R. Rouquier. 2-Kac-Moody algebras. arXiv:0812.5023, 2008.
  • [29] H. N. V. Temperley and E. H. Lieb. Relations between the ’Percolation’ and ’Colouring’ Problem and other Graph-Theoretical Problems Associated with Regular Planar Lattices: Some Exact Results for the ’Percolation’ Problem. Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, 322:251–280, 1971.