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

Quasi-hereditary covers of Temperley-Lieb algebras and relative dominant dimension

Tiago Cruz Institut für Algebra und Zahlentheorie, Universität Stuttgart, Germany tiago.cruz@mathematik.uni-stuttgart.de Max-Planck-Institut für Mathematik, Vivatsgasse 7, 53111 Bonn, Germany  and  Karin Erdmann Mathematical Institute, Oxford University, ROQ, Oxford OX2 6GG, United Kingdom erdmann@maths.ox.ac.uk
Abstract.

Many connections and dualities in representation theory can be explained using quasi-hereditary covers in the sense of Rouquier. The concepts of relative dominant and codominant dimension with respect to a module, introduced recently by the first-named author, are important tools to evaluate and classify quasi-hereditary covers.

In this paper, we prove that the relative dominant dimension of the regular module of a quasi-hereditary algebra with a simple preserving duality with respect to a summand QQ of a characteristic tilting module equals twice the relative dominant dimension of a characteristic tilting module with respect to QQ.

To resolve the Temperley-Lieb algebras of infinite global dimension, we apply this result to the class of Schur algebras S(2,d)S(2,d) and Q=VdQ=V^{\otimes d} the dd-tensor power of the 2-dimensional module and we completely determine the relative dominant dimension of the Schur algebra S(2,d)S(2,d) with respect to VdV^{\otimes d}. The qq-analogues of these results are also obtained. As a byproduct, we obtain a Hemmer-Nakano type result connecting the Ringel duals of qq-Schur algebras and Temperley-Lieb algebras. From the point of view of Temperley-Lieb algebras, we obtain the first complete classification of their connection to their quasi-hereditary covers formed by Ringel duals of qq-Schur algebras.

These results are compatible with the integral setup, and we use them to deduce that the Ringel dual of a qq-Schur algebra over the ring of Laurent polynomials over the integers together with some projective module is the best quasi-hereditary cover of the integral Temperley-Lieb algebra.

Key words and phrases:
Quasi-hereditary cover, relative dominant dimension, qq-Schur algebra, Temperley-Lieb algebra, Frobenius twist
2020 Mathematics Subject Classification:
Primary: 16E10, Secondary: 20G43, 16G10, 16G30, 82B20

1. Introduction

The theory of quasi-hereditary covers, introduced in [43], gives a framework to study finite-dimensional algebras of infinite global dimension through algebras having nicer homological properties, for instance, quasi-hereditary algebras via an exact functor known as Schur functor. Quasi-hereditary covers appear naturally and are useful in algebraic Lie theory, representation theory and homological algebra. In particular, they are in the background of Auslander’s correspondence [1] and in Iyama’s proof of finiteness of representation dimension [33]. Further, quasi-hereditary algebras arise quite naturally in the representation theory of algebraic groups ([4, 41]) and algebras of global dimension at most two are quasi-hereditary.

Schur algebras S(n,d)S(n,d) form an important class of quasi-hereditary algebras, they provide a link between polynomial representations of general linear groups and representations of symmetric groups. Classically, when ndn\geq d, the Schur algebra, via the Schur functor, is a quasi-hereditary cover of the group algebra of the symmetric group 𝒮d\mathcal{S}_{d}. This connection is seen as one of the versions of Schur–Weyl duality. Indeed, this formulation clarifies the connection between the representation theory of symmetric groups and the representation theory of Schur algebras, by detecting how their subcategories are related and how the Yoneda extension groups in these subcategories are related by the Schur functor (see also [32]). Further, this connection becomes stronger as the characteristic of the ground field increases. It was first observed in [26] that this behaviour is captured by the classical dominant dimension. However, not all quasi-hereditary covers can be evaluated using classical dominant dimension.

To fix this, the first-named author introduced in [8] the concepts of relative dominant dimension and relative codominant dimension with respect to a module. Further, in [8] these homological invariants were exploited to create new quasi-hereditary covers. With this, the link between Schur algebras and symmetric groups can be regarded as a special case of quasi-hereditary covers of quotients of Iwahori-Hecke algebras.

Temperley-Lieb algebras are among the algebras that can be regarded as quotients of Iwahori-Hecke algebras and they can have infinite global dimension. They were introduced in [44] in the context of statistical mechanics and they were popularised by Jones, in particular, they are used to define the Jones polynomial (see [37]). However, contrary to Iwahori-Hecke algebras no Hemmer-Nakano type result was known for Temperley-Lieb algebras up until now. Both classes of algebras are cellular (see for example [29]) and so an important property that they have in common is the existence of a simple preserving duality.

Quasi-hereditary algebras with a simple preserving duality always have even global dimension. Mazorchuk and Ovsienko have shown this fact in [40] by proving that the global dimension of a quasi-hereditary algebra with a simple preserving duality is exactly twice the projective dimension of the characteristic tilting module. Later, under much stronger conditions, the analog result for dominant dimension was obtained in [26] by Fang and Koenig exploiting that a faithful projective-injective module is a summand of the characteristic tilting module.

The present paper has two aims. First, we will establish that the relative dominant dimension of a quasi-hereditary algebra with respect to any summand of its characteristic tilting module is always twice as large as that of the characteristic tilting module, in the case when the algebra has a simple preserving duality. In particular, this homological invariant is always even for such quasi-hereditary algebras. Further, Fang and Koenig’s result can then be recovered from ours by just fixing the summand to be a projective-injective module. Therefore, we obtain an alternative approach to the classical case of dominant dimension without any further assumptions.

The second aim is to study classes of quasi-hereditary covers of Temperley-Lieb algebras and their link with the representation theory of Temperley-Lieb algebras. In particular, we aim to completely understand such a connection using the representation theory of qq-Schur algebras and how good are the resolutions of Temperley-Lieb algebras by the Ringel duals of qq-Schur algebras.

Questions to be addressed and setup

To make our results precise, we need further notation. In general, assume that BB is a finite-dimensional algebra over an algebraically closed field. A pair (A,P)(A,P) is a quasi-hereditary cover of BB if AA is a quasi-hereditary algebra, PP is a finitely generated projective AA-module such that B=EndA(P)opB=\operatorname{End}_{A}(P)^{op}, and in addition the restriction of the associated Schur functor F:=HomA(P,):AmodBmodF:=\operatorname{Hom}_{A}(P,-)\colon A\!\operatorname{-mod}\rightarrow B\!\operatorname{-mod} to the subcategory of finitely generated projective AA-modules is full and faithful.

Let (Δ)\mathcal{F}(\Delta) be the category of AA-modules which have a filtration by standard modules. We would like the functor FF to be faithful on (Δ)\mathcal{F}(\Delta) and to induce isomorphisms

ExtAj(X,Y)ExtBj(FX,FY){\rm Ext}_{A}^{j}(X,Y)\to{\rm Ext}^{j}_{B}(FX,FY)

for X,YX,Y modules in (Δ)\mathcal{F}(\Delta). If this is the case for 0ji0\leq j\leq i then (A,P)(A,P) is called an i(Δ)i-\mathcal{F}(\Delta) cover of BB. The largest nn such that (A,P)(A,P) is an n(Δ)n-\mathcal{F}(\Delta) cover of BB, is called the Hemmer-Nakano dimension of (Δ)\mathcal{F}(\Delta) in [26]. When BB is self-injective, Fang and Koenig showed that this dimension is controlled by the dominant dimension of a characteristic tilting module. In addition, they proved that if BB is a symmetric algebra and the quasi-hereditary cover admits a certain simple preserving duality, then the dominant dimension of a characteristic tilting module is exactly half of the dominant dimension of AA.

Recently, in [8], the situation was generalised to include cases where BB is not necessarily self-injective. Moreover, it was proved in [8] that the Hemmer-Nakano dimension of (Δ)\mathcal{F}(\Delta) associated with a 0(Δ)0-\mathcal{F}(\Delta) cover can be determined using the relative codominant dimension of a characteristic tilting module with respect to a certain summand of the characteristic tilting module.

The concepts of relative dominant and relative codominant dimension (see the definition below in Subsection 2.3) and the concept of quasi-hereditary cover can be considered in an integral setup, that is, both of these concepts can be studied for Noetherian algebras which are finitely generated and projective as modules over a regular commutative Noetherian ring. In [7], methods were developed to reduce the computations of Hemmer-Nakano dimensions in the integral setup to computations of Hemmer-Nakano dimensions in the setup where the ground ring is an algebraically closed field. So, it will be enough for our purposes to concentrate our attention on the case when the coefficient ring is an algebraically closed field.

The new approach to construct quasi-hereditary covers is [8, Theorem 5.3.1.] and [8, Theorem 8.1.5] when applied to Schur algebras (and qq-Schur algebras). The novelty is that it uses the Ringel dual of a Schur algebra, rather than a Schur algebra, and works for arbitrary parameters n,dn,d.

This can, in particular, be applied to the study of Temperley-Lieb algebras. Indeed, the Temperley-Lieb algebras can be viewed as centraliser algebras of S(2,d)S(2,d) in the endomorphism algebra of the tensor power (K2)d(K^{2})^{\otimes d} (over a field KK) and their qq-analogues. Here S(n,d)S(n,d) can be regarded as the centraliser algebra of 𝒮d\mathcal{S}_{d} in the endomorphism algebra of the tensor power (Kn)d(K^{n})^{\otimes d} over, where (Kn)d(K^{n})^{\otimes d} affords a module structure over 𝒮d\mathcal{S}_{d} by place permutation. Furthermore, Vd:=(Kn)dV^{\otimes d}:=(K^{n})^{\otimes d} belongs in the additive closure of a characteristic tilting module over S(n,d)S(n,d). Our cases of interest have a simple preserving duality, and in such a case, for this situation, we can without ambiguity interchange the concepts: relative dominant dimension and relative codominant dimension.

Denote by QdomdimAXQ\!\operatorname{-domdim}_{A}X the relative dominant dimension of an AA-module XX with respect to QQ. In this context, the following questions arise:

  1. (1)

    What is the value of VddomdimS(2,d)TV^{\otimes d}\!\operatorname{-domdim}_{S(2,d)}T, where TT is a characteristic tilting module of the quasi-hereditary algebra S(2,d)S(2,d)? What happens to this value when we replace a Schur algebra by a qq-Schur algebra?

  2. (2)

    The Ringel duals of Schur algebras as well as Schur algebras have a simple preserving duality. Can we expect, like in the classical case (see [26, Theorem 4.3.]), the equality

    VddomdimS(n,d)=2VddomdimS(n,d)TV^{\otimes d}\!\operatorname{-domdim}S(n,d)=2\cdot V^{\otimes d}\!\operatorname{-domdim}_{S(n,d)}T

    to hold in general?

  3. (3)

    Can we expect the quasi-hereditary cover of the Temperley-Lieb algebra constructed in [8, Theorem 8.1.5] to be unique, in some meaningful way?

Our goal in this paper is to give answers to these three questions.

Main results

Surprisingly, the answer to (2) is positive without using extra structure on S(n,d)S(n,d) besides the quasi-hereditary structure and the existence of a simple preserving duality.

Theorem A.

(see Theorem 3.2.2) Let AA be a quasi-hereditary algebra over a field KK. Suppose that there exists a simple preserving duality ():AmodAmod{}^{\diamond}\!(-)\colon A\!\operatorname{-mod}\to A\!\operatorname{-mod}. Let TT be the characteristic tilting module of AA. Assume that Qadd(T)Q\in{\rm add}(T). Then

QdomdimAA=2QdomdimAT.Q\!\operatorname{-domdim}_{A}A=2\cdot Q\!\operatorname{-domdim}_{A}T.

This result generalises [26, Theorem 4.3.] and our methods give a new proof to their case without using any information on AA being gendo-symmetric, that is, an endomorphism algebra of a faithful module over a symmetric algebra. In particular, our result also works for dominant dimension exactly zero. Our approach exploits basic properties of relative injective dimensions, Δ\Delta-filtration dimensions, some tools that were used to prove the main result of [40] and general properties connecting relative dominant dimensions with relative codominant dimensions with respect to a fixed module. Observe that the left hand side of the equation in Theorem A is exactly the faithful dimension of QQ in sense of [2]. This means that, under these conditions, if the faithful dimension of QQ is greater or equal to 4, then the faithful dimension controls the Hemmer-Nakano dimension of (Δ)\mathcal{F}(\Delta) associated with a quasi-hereditary cover of the endomorphism algebra of QQ. Theorem A is applied to prove a more general case of Conjecture 6.2.4 of [6], that is, that the faithful dimension of a summand of a characteristic tilting module is an upper bound for the dominant dimension of the algebra provided that the former is greater or equal than two.

Combining techniques of Frobenius twisted tensor products with Theorem A we obtain a complete answer to (1):

Theorem B.

(see also Theorem 5.2.2 for the qq-version) Let KK be a field and let AA be the Schur algebra SK(2,d)S_{K}(2,d) and TT be the characteristic tilting module of AA. Then,

VddomdimAA=2VddomdimAT={d, if charK=2 and d is even,+, otherwise .\displaystyle V^{\otimes d}\!\operatorname{-domdim}_{A}A=2\cdot V^{\otimes d}\!\operatorname{-domdim}_{A}T=\begin{cases}d,&\text{ if }\operatorname{char}K=2\text{ and }d\text{ is even},\\ +\infty,&\text{ otherwise }\end{cases}.

The same approach can be used for the qq-analogue. In this case, the algebra AA is the qq-Schur algebra SK,q(2,d)S_{K,q}(2,d), which can be defined as the centraliser of the Hecke algebra Hq(d)H_{q}(d) acting on VdV^{\otimes d}, again for dimV=2\dim V=2 and in the theorem the characteristic is replaced by the quantum characteristic. When we have vRv\in R such that v2=qv^{2}=q and δ=vv1\delta=-v-v^{-1}, the Temperley-Lieb algebra TLK,d(δ)TL_{K,d}(\delta) is a quotient of this action. In both cases, the real difficulty lies in the case in which the characteristic (resp. quantum characteristic) is two.

From Theorem B and its qq-analogue, it follows that the Temperley-Lieb algebra TLK,q(δ)TL_{K,q}(\delta) is quasi-hereditary, and, in fact, it is the Ringel dual of a qq-Schur algebra if δ0\delta\neq 0 or dd is odd. Otherwise, from Theorem B follows the value of Hemmer-Nakano dimension of (Δ)\mathcal{F}(\Delta) associated with the quasi-hereditary cover of TLK,q(0)TL_{K,q}(0) formed by the Ringel dual of a qq-Schur algebra (see [8, Theorem 8.1.5]). So, we have obtained a Hemmer-Nakano type result (see Corollary 6.2.4) now between the Ringel dual of a qq-Schur algebra and the Temperley-Lieb algebra. In particular, this generalises [10, Theorem C (3), (4)] for Temperley-Lieb algebras. In addition, the full subcategory of costandard modules over a qq-Schur algebra is equivalent to the full subcategory of cell modules of the Temperley-Lieb algebra whenever dd is greater or equal to 6.

If d=2d=2, the Temperley-Lieb algebra is exactly an Iwahori-Hecke algebra, so nothing is new for this case. We obtain a positive answer to question (3) when we consider the Laurent polynomial ring over the integers as coefficient ring and d>2d>2 (see Section 7 and Corollary 7.2.1). In such a case, the (integral) Schur functor FF induces an exact equivalence (Δ~)(FΔ~),\mathcal{F}(\tilde{\Delta})\rightarrow\mathcal{F}(F\tilde{\Delta}), where the first category denotes the subcategory of modules admitting a filtration by direct summands of direct sums of standard modules over the Ringel dual of an integral qq-Schur algebra. The quasi-hereditary cover of the integral Temperley-Lieb algebra formed by the Ringel dual of a qq-Schur algebra is the unique quasi-hereditary cover which induces this exact equivalence.

We emphasize that the specialisation of Theorem A to projective-injective modules played a keyrole to determine the dominant dimension of Schur algebras of the form S(n,d)S(n,d) with ndn\geq d in [26] (also their qq-analogues [27]) and it also gives an easier method to determine the dominant dimension of the blocks of the BGG category 𝒪\mathcal{O}. It is our expectation that its use will be crucial to determine, in particular, VddomdimS(n,d)V^{\otimes d}\!\operatorname{-domdim}S(n,d) and domdimS(n,d)\operatorname{domdim}S(n,d) also in the cases 2<n<d2<n<d while the latter is also an open problem for n=2n=2.

The article is organised as follows: In Section 2, we introduce the notation and the main properties of relative dominant dimension with respect to a module, split quasi-hereditary algebras with a simple preserving duality and cover theory to be used throughout the paper. In Section 3, we discuss elementary results on relative injective dimensions and we give the proof of Theorem A. We then deduce that the dominant dimension is a lower bound for the faithful dimension of a summand of a characteristic tilting module fixed by a simple preserving duality provided the latter is at least two (see Proposition 3.2.3). In Section 4, we collect results on the quasi-hereditary structure of Schur algebras S(2,d)S(2,d), in particular, reduction techniques and how to construct partial tilting and standard modules inductively using the Frobenius twist functor. In Section 5, we compute the relative dominant dimension of S(2,d)S(2,d) with respect to VdV^{\otimes d} in terms of VddomdimS(2,d)TV^{\otimes d}\!\operatorname{-domdim}_{S(2,d)}T, where TT is a characteristic tilting module of S(2,d)S(2,d). In particular, we give the proof of Theorem B and its qq-analogue (see Theorem 5.2.2). In Section 6, we recall that all Temperley-Lieb algebras can be realised as the centraliser algebras of qq-Schur algebras in the endomorphism algebra of the tensor power VdV^{\otimes d}. As a consequence, we determine the value of Hemmer-Nakano dimension of (Δ)\mathcal{F}(\Delta) in all cases associated to the cover of the Temperley-Lieb algebra formed by the Ringel dual of a qq-Schur algebra. This computation is contained in Corollary 6.2.4. In Section 7, we determine the Hemmer-Nakano dimension of the above mentioned quasi-hereditary cover in the integral setup, dividing the study into two cases: the coefficient ring having or not a property of being 22-partially qq-divisible (see Subsection 7.1). When the coefficient ring does not have such property, we show that a quasi-hereditary cover with such coefficient ring has better properties. We conclude by addressing the problem of the uniqueness of this cover (see Subsection 7.2).

2. Preliminaries

2.1. The setting

This follows [8]. Throughout we fix a Noetherian commutative ring RR with identity, and AA is an RR-algebra which is finitely generated and projective as an RR-module. We refer to AA as a projective Noetherian RR-algebra. The set of invertible elements of RR is denoted by R×R^{\times}.

We denote by AmodA\!\operatorname{-mod} the category of finitely generated AA-modules. Given MAmodM\in A\!\operatorname{-mod}, we denote by addAM\!\operatorname{add}_{A}M (or just addM\!\operatorname{add}M) the full subcategory of AmodA\!\operatorname{-mod} whose modules are direct summands of a finite direct sum of copies of MM. We also denote addA\!\operatorname{add}A by AprojA\!\operatorname{-proj}.

The endomorphism algebra of a module MAmodM\in A\!\operatorname{-mod} is denoted by EndA(M){\rm End}_{A}(M). We denote by DRD_{R} or just DD the standard duality functor HomR(,R):AmodAopmod{\rm Hom}_{R}(-,R):A\!\operatorname{-mod}\to A^{op}\!\operatorname{-mod} where AopA^{op} is the opposite algebra of AA.

A module MAmodRprojM\in A\!\operatorname{-mod}\cap R\!\operatorname{-proj} is said to be (A,R)(A,R)-injective if it belongs to addDA{\rm add}DA, and we write (A,R)injRproj(A,R)\!\operatorname{-inj}\cap R\!\operatorname{-proj} for the full subcategory of AmodRprojA\!\operatorname{-mod}\cap R\!\operatorname{-proj} whose modules are (A,R)(A,R)-injective.

Furthermore, an exact sequence of AA-modules which is split as an exact sequence of RR-modules is said to be (A,R)(A,R)-exact. In particular, an (A,R)(A,R)-monomorphism is a homomorphism f:MNf:M\to N that fits into an (A,R)(A,R)-exact sequence 0MfN0\to M\stackrel{{\scriptstyle f}}{{\to}}N.

Given a left exact covariant additive functor GG, we say that XX is a GG-acyclic object if Ri>0G(X)=0\operatorname{R}^{i>0}G(X)=0. An exact sequence 0LX0X10\rightarrow L\rightarrow X_{0}\rightarrow X_{1}\rightarrow\cdots is called a GG-acyclic coresolution of LL if all objects X0,X1,X_{0},X_{1},\cdots are GG-acyclic. Given XAmodRprojX\in A\!\operatorname{-mod}\cap R\!\operatorname{-proj}, we denote by XX^{\perp} the full subcategory

{MAmodRproj:ExtAi>0(Z,M)=0,ZaddX},{\{M\in A\!\operatorname{-mod}\cap R\!\operatorname{-proj}\colon\operatorname{Ext}_{A}^{i>0}(Z,M)=0,\forall Z\in\!\operatorname{add}X\}},

and by X{}^{\perp}X the full subcategory {MAmodRproj:ExtAi>0(M,Z)=0,ZaddX}\{M\in A\!\operatorname{-mod}\cap R\!\operatorname{-proj}\colon\operatorname{Ext}_{A}^{i>0}(M,Z)=0,\forall Z\in\!\operatorname{add}X\}.

2.2. Basics on approximations

We recall definitions and some general properties relevant to approximations. Assume that AA is an RR-algebra as above, and QQ is a fixed module in AmodRprojA\!\operatorname{-mod}\cap R\!\operatorname{-proj}.

An AA-homomorphism f:MNf:M\to N is a left addAQ\!\operatorname{add}_{A}Q-approximation of MM provided that NN belongs to addAQ\!\operatorname{add}_{A}Q, and moreover the induced map

HomA(N,X)HomA(M,X){\rm Hom}_{A}(N,X)\to{\rm Hom}_{A}(M,X)

is surjective for every XaddAQX\in\!\operatorname{add}_{A}Q. Dually one defines right addAQ\!\operatorname{add}_{A}Q-approximations. Note that every module MAmodM\in A\!\operatorname{-mod} has a left and a right addAQ\!\operatorname{add}_{A}Q-approximation.

2.3. Relative (co)dominant dimension with respect to a module

We recall from [8] the definition of relative (co)dominant dimensions.

Let Q,XAmodRprojQ,X\in A\!\operatorname{-mod}\cap R\!\operatorname{-proj}. If XX does not admit a left addQ\!\operatorname{add}Q-approximation which is an (A,R)(A,R)-monomorphism then the relative dominant dimension of XX with respect to QQ is zero. Otherwise, the relative dominant dimension of XX with respect to QQ, denoted by Qdomdim(A,R)XQ\!\operatorname{-domdim}_{(A,R)}X, or QdomdimAXQ\!\operatorname{-domdim}_{A}X when RR is a field, is the supremum of all nn\in\mathbb{N} such that there is an (A,R)(A,R)-exact sequence

0XQ1Q2Qn0\to X\to Q_{1}\to Q_{2}\to\ldots\to Q_{n}

with all QiaddQQ_{i}\in\!\operatorname{add}Q, which remains exact under HomA(,Q){\rm Hom}_{A}(-,Q).

Dually one defines the relative codominant dimension, denoted by Qcodomdim(A,R)(X)Q-{\rm codomdim}_{(A,R)}(X) with Q,XQ,X as above: if XX does not admit a surjective right addQ\!\operatorname{add}Q-approximation, then Qcodomdim(A,R)(X)=0{Q\!\operatorname{-codomdim}_{(A,R)}(X)}=0. Otherwise it is the supremum of all nn\in\mathbb{N} such that there is an (A,R)(A,R)-exact sequence

QnQn1Q1X0Q_{n}\to Q_{n-1}\to\ldots\to Q_{1}\to X\to 0

with all QiaddQQ_{i}\in\!\operatorname{add}Q, which remains exact under HomA(Q,){\rm Hom}_{A}(Q,-).

Hence, Qcodomdim(A,R)X=DQdomdim(Aop,R)DXQ\!\operatorname{-codomdim}_{(A,R)}X=DQ\!\operatorname{-domdim}_{(A^{op},R)}DX. By Qdomdim(A,R)Q\!\operatorname{-domdim}{(A,R)} we mean the value Qdomdim(A,R)AQ\!\operatorname{-domdim}_{(A,R)}A. We will write QcodomdimAXQ\!\operatorname{-codomdim}_{A}X to denote Qcodomdim(A,R)XQ\!\operatorname{-codomdim}_{(A,R)}X when RR is a field.

The following gives a criterion towards finding Qdomdim(A,R)MQ\!\operatorname{-domdim}_{(A,R)}M for a given module MM in AmodRprojA\!\operatorname{-mod}\cap R\!\operatorname{-proj}.

Lemma 2.3.1.

Assume MAmodRprojM\in A\!\operatorname{-mod}\cap R\!\operatorname{-proj}, and let QiaddQQ_{i}\in\!\operatorname{add}Q. An exact sequence

0Mα0Q0α1Q1Qt0\to M\xrightarrow{\alpha_{0}}Q_{0}\stackrel{{\scriptstyle\alpha_{1}}}{{\to}}Q_{1}\to\ldots\to Q_{t}

remains exact under HomA(,Q){\rm Hom}_{A}(-,Q) if and only if for every factorisation Qiimαi+1Qi+1Q_{i}\to{\rm im}\alpha_{i+1}\to Q_{i+1} of αi+1\alpha_{i+1}, the (A,R)(A,R)-monomorphism imαi+1Qi+1{\rm im}\alpha_{i+1}\to Q_{i+1} and α0\alpha_{0} are left addQ\!\operatorname{add}Q-approximations.

Proof.

See [8, Lemma 2.1.4.]. ∎

In addition to the assumptions on R,AR,A and QQ, in the following, we also assume that DQAQRprojDQ\otimes_{A}Q\in R\!\operatorname{-proj}.

It is crucial to compare relative dominant dimensions for end terms of a short exact sequence which remains exact under HomA(,Q){\rm Hom}_{A}(-,Q). This is completely described in [8, Lemma 3.1.7], for convenience, we recall part of this.

Lemma 2.3.2.

Let MAmodRprojM\in A\!\operatorname{-mod}\cap R\!\operatorname{-proj} and consider an (A,R)(A,R)-exact sequence

0M1MM200\to M_{1}\to M\to M_{2}\to 0

which remains exact under HomA(,Q){\rm Hom}_{A}(-,Q). Let n=QdomdimAMn=Q\!\operatorname{-domdim}_{A}M and ni=QdomdimAMin_{i}=Q\!\operatorname{-domdim}_{A}M_{i} for i=1,2i=1,2, then:

  1. (a)

    nmin{n1,n2}n\geq{\rm min}\{n_{1},n_{2}\}.

  2. (b)

    If n=n=\infty and n1<n_{1}<\infty then n2=n11n_{2}=n_{1}-1.

Corollary 2.3.3.

Let MiM_{i} for iIi\in I be a finite set of modules in AmodRprojA\!\operatorname{-mod}\cap R\!\operatorname{-proj}. Then

QdomdimA(iIMi)=inf{QdomdimAMiiI}.Q\!\operatorname{-domdim}_{A}\left(\bigoplus_{i\in I}M_{i}\right)={\rm inf}\{Q\!\operatorname{-domdim}_{A}M_{i}\mid i\in I\}.
Proof.

See [8, Lemma 3.1.8]. ∎

Recall Q={MAmodRprojExtAi>0(M,Q)=0}{}^{\perp}Q=\{M\in A\!\operatorname{-mod}\cap R\!\operatorname{-proj}\mid{\rm Ext}_{A}^{i>0}(M,Q)=0\}. The following is proved in [8, Proposition 3.1.11.].

Proposition 2.3.4.

Assume ExtAi>0(Q,Q)=0{\rm Ext}^{i>0}_{A}(Q,Q)=0, and MQM\in{}^{\perp}Q. An exact sequence

0MQ1Qn0\to M\to Q_{1}\to\ldots\to Q_{n}

yields Qdomdim(A,R)(M)nQ\!\operatorname{-domdim}_{(A,R)}(M)\geq n if and only if QiaddQQ_{i}\in\!\operatorname{add}Q and the cokernel of Qn1QnQ_{n-1}\to Q_{n} belongs to Q{}^{\perp}Q.

The following application of Lemma 2.3.2 will be useful later.

Corollary 2.3.5.

Assume QQQ\in{}^{\perp}Q. Let MAmodRprojM\in A\!\operatorname{-mod}\cap R\!\operatorname{-proj}, and consider an (A,R)(A,R)-exact sequence

0MQ1QtX00\to M\to Q_{1}\to\ldots\to Q_{t}\to X\to 0

with QiaddQQ_{i}\in\!\operatorname{add}Q. If ExtAi(X,Q)=0{\rm Ext}_{A}^{i}(X,Q)=0 for 1it1\leq i\leq t, then

Qdomdim(A,R)M=t+Qdomdim(A,R)X.\displaystyle Q\!\operatorname{-domdim}_{(A,R)}M=t+Q\!\operatorname{-domdim}_{(A,R)}X.
Proof.

See [8, Corollary 3.1.12.]. ∎

2.4. Split quasi-hereditary algebras with duality

For the definition and general properties of split quasi-hereditary algebras we refer to [5, 43, 9, 7, 8]. In particular, we follow the notation of [9, 7, 8]. One of the advantages to use such setup stems from the fact that split quasi-hereditary RR-algebras (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) are exactly the algebras so that (SRA,{SRΔ(λ)λΛ})(S\otimes_{R}A,\{S\otimes_{R}\Delta(\lambda)_{\lambda\in\Lambda}\}) are quasi-hereditary algebras for every commutative Noetherian ring SS which is an RR-algebra. Concerning the terminology, we remark the word split arises from the endomorphism algebra EndA(Δ(λ))\operatorname{End}_{A}(\Delta(\lambda)) being isomorphic to the ground ring RR. As it was observed in [43], when (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) is a split quasi-hereditary RR-algebra, the objects T(λ)T(\lambda) satisfying addλΛT(λ)=(Δ~)(~)\!\operatorname{add}\bigoplus_{\lambda\in\Lambda}T(\lambda)=\mathcal{F}(\tilde{\Delta})\cap\mathcal{F}(\tilde{\nabla}) are no longer unique, in contrast to quasi-hereditary algebras over a field. For this reason, we will say that TT is a characteristic tilting module of AA if addT=(Δ~)(~)\!\operatorname{add}T=\mathcal{F}(\tilde{\Delta})\cap\mathcal{F}(\tilde{\nabla}) and TT is the (basic) characteristic tilting module of AA if AA is a quasi-hereditary algebra over a field and T=λΛT(λ)T=\bigoplus_{\lambda\in\Lambda}T(\lambda).

The following prepares the ground for quasi-hereditary covers, constructed from the Ringel dual R(A)=EndA(T)opR(A)={\rm End}_{A}(T)^{op} of a quasi-hereditary algebra AA with a characteristic tilting module TT. To see that Ringel duality is well defined in the integral setup, we refer to [8, Subsection 2.2.3].

Proposition 2.4.1.

Let (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) be a split quasi-hereditary RR-algebra with a characteristic tilting module TT. Denote by R(A)R(A) the Ringel dual EndA(T)op\operatorname{End}_{A}(T)^{op} of AA. Suppose that QaddATQ\in\!\operatorname{add}_{A}T is a partial tilting module. Then,

  1. (i)

    HomA(T,Q)codomdim(R(A),R)DT=Qdomdim(A,R)T\operatorname{Hom}_{A}(T,Q)\!\operatorname{-codomdim}_{(R(A),R)}DT=Q\!\operatorname{-domdim}_{(A,R)}T.

  2. (ii)

    DQdomdim(A,R)=Qcodomdim(A,R)DA=Qdomdim(A,R).DQ\!\operatorname{-domdim}{(A,R)}=Q\!\operatorname{-codomdim}_{(A,R)}DA=Q\!\operatorname{-domdim}{(A,R)}.

Proof.

For (i), see [8, Proposition 6.1.1]. For (ii), see [8, Corollary 3.1.5]. ∎

Recall that HomA(T,DA)DT\operatorname{Hom}_{A}(T,DA)\simeq DT is a characteristic tilting module over R(A)R(A).

Proposition 2.4.2.

Let (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) be a split quasi-hereditary algebra over a field kk. Assume that there exists a simple preserving duality ():AmodAmod{}^{\diamond}\!(-)\colon A\!\operatorname{-mod}\rightarrow A\!\operatorname{-mod}. Let TT be the characteristic tilting module of AA and assume that QaddATQ\in\!\operatorname{add}_{A}T. Then,

  1. (i)

    Δ(λ)(λ){}^{\diamond}\!\Delta(\lambda)\simeq\nabla(\lambda) for all λΛ\lambda\in\Lambda;

  2. (ii)

    T(λ)T(λ){}^{\diamond}\!T(\lambda)\simeq T(\lambda) for all indecomposable modules of TT;

  3. (iii)

    Qdomdim(A,R)T=Qcodomdim(A,R)TQ\!\operatorname{-domdim}_{(A,R)}T=Q\!\operatorname{-codomdim}_{(A,R)}T.

Proof.

(i) and (ii) follow by applying the simple preserving duality to the canonical exact sequences defining Δ(λ)\Delta(\lambda) and T(λ)T(\lambda), respectively. For (iii), see [8, Proposition 3.1.6]. ∎

Let (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) be a split quasi-hereditary algebra over a field kk. The \nabla-filtration dimension of XX, denoted by dim()X\dim_{\mathcal{F}(\nabla)}X, is the minimal n0n\geq 0 such that there exists an exact sequence

0XM0Mn00\rightarrow X\rightarrow M_{0}\rightarrow\cdots\rightarrow M_{n}\rightarrow 0

with M0,,Mn()M_{0},\ldots,M_{n}\in\mathcal{F}(\nabla). Analogously, the Δ\Delta-filtration dimension is defined. The \nabla-filtration dimensions first appeared in [28] in the study of cohomology of algebraic groups.

\nabla and Δ\Delta-filtration dimensions play a crucial role in [23] and [40] establishing that the global dimension of a quasi-hereditary algebra having a simple preserving duality is always an even number. For us, they are of importance due to the following result.

Proposition 2.4.3.

Let (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) be a split quasi-hereditary algebra over a field kk. Assume that there exists a simple preserving duality ():AmodAmod{}^{\diamond}\!(-)\colon A\!\operatorname{-mod}\rightarrow A\!\operatorname{-mod}. If MAmodM\in A\!\operatorname{-mod} satisfying dim(Δ)M=t<+\dim_{\mathcal{F}(\Delta)}M=t<+\infty, then ExtA2t(M,M)0\operatorname{Ext}_{A}^{2t}(M,{}^{\diamond}\!M)\neq 0.

Proof.

See [40, Corollary 6]. ∎

2.5. Cover theory

The concept of a cover, and in particular, of a split quasi-hereditary cover was introduced in [43] to give an abstract framework to connections in representation theory like Schur–Weyl duality. Given a split quasi-hereditary algebra (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) over a commutative Noetherian ring RR and a finitely generated projective AA-module PP, let B:=EndA(P)opB:={\rm End}_{A}(P)^{op}. We say that (A,P)(A,P) is a split quasi-hereditary cover of BB if the restriction of the functor F:=HomA(P,):AmodBmodF:=\operatorname{Hom}_{A}(P,-)\colon A\!\operatorname{-mod}\rightarrow B\!\operatorname{-mod}, known as Schur functor, to AprojA\!\operatorname{-proj} is fully faithful. Given, in addition, i{1,0,+}i\in\mathbb{N}\cup\{-1,0,+\infty\}, following the notation of [7], we say that (A,P)(A,P) is an i(Δ~)i-\mathcal{F}(\tilde{\Delta}) (quasi-hereditary) cover of BB if the following conditions hold:

  • (A,P)(A,P) is a split quasi-hereditary cover of EndA(P)op\operatorname{End}_{A}(P)^{op};

  • The restriction of FF to (Δ~)\mathcal{F}(\tilde{\Delta}) is faithful;

  • The Schur functor FF induces bijections ExtAj(M,N)ExtBj(FM,FN)\operatorname{Ext}_{A}^{j}(M,N)\simeq\operatorname{Ext}^{j}_{B}(FM,FN), for every M,N(Δ~)M,N\in\mathcal{F}(\tilde{\Delta}) and 0ji0\leq j\leq i;

Here (Δ~)\mathcal{F}(\tilde{\Delta}) denotes the resolving subcategory of AmodRprojA\!\operatorname{-mod}\cap R\!\operatorname{-proj} whose modules admit a finite filtration into direct summands of direct sums of standard modules Δ(λ)\Delta(\lambda), λΛ\lambda\in\Lambda.

The optimal value of the quality of a cover is known as the Hemmer-Nakano dimension. More precisely, if (A,P)(A,P) is a (1)(Δ~)(-1)-\mathcal{F}(\tilde{\Delta}) (quasi-hereditary) cover of BB, the Hemmer-Nakano dimension of (Δ~)\mathcal{F}(\tilde{\Delta}) with respect to FF is i{1,0,+}i\in\mathbb{N}\cup\{-1,0,+\infty\} if (A,P)(A,P) is an i(Δ~)i-\mathcal{F}(\tilde{\Delta}) (quasi-hereditary) cover of EndA(P)op\operatorname{End}_{A}(P)^{op} but (A,P)(A,P) is not an (i+1)(Δ~)(i+1)-\mathcal{F}(\tilde{\Delta}) (quasi-hereditary) cover of BB. The Hemmer-Nakano dimension of (Δ~)\mathcal{F}(\tilde{\Delta}) is denoted by HNdimF((Δ~))\operatorname{HNdim}_{F}(\mathcal{F}(\tilde{\Delta})).

Major tools to compute Hemmer-Nakano dimensions are classical dominant dimension and relative dominant dimensions. This idea can be traced back to [26] which was later amplified in several directions in [7] and in [8]. This principle is briefly summarized in the following result proved in [8, Theorem 5.3.1., Corollary 5.3.4.]. Note that HomA(T,Q){\rm Hom}_{A}(T,Q) is projective as a BB-module.

Theorem 2.5.1.

Let RR be a commutative Noetherian ring. Let (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) be a split quasi-hereditary RR-algebra with a characteristic tilting module TT. Denote by R(A)R(A) the Ringel dual EndA(T)op\operatorname{End}_{A}(T)^{op} of AA. Assume that QaddTQ\in\!\operatorname{add}T is a (partial) tilting module of AA. Then, the following assertions hold.

  1. (a)

    If Qcodomdim(A,R)Tn2Q\!\operatorname{-codomdim}_{(A,R)}T\geq n\geq 2, then (R(A),HomA(T,Q))(R(A),\operatorname{Hom}_{A}(T,Q)) is an (n2)(n-2)-(Δ~R(A))\mathcal{F}(\tilde{\Delta}_{R(A)}) split quasi-hereditary cover of EndA(Q)op\operatorname{End}_{A}(Q)^{op}.

  2. (b)

    Assume, in addition, that RR is a field. Then, Qcodomdim(A,R)Tn2Q\!\operatorname{-codomdim}_{(A,R)}T\geq n\geq 2 if and only if (R(A),HomA(T,Q))(R(A),\operatorname{Hom}_{A}(T,Q)) is an (n2)(n-2)-(Δ~R(A))\mathcal{F}(\tilde{\Delta}_{R(A)}) split quasi-hereditary cover of EndA(Q)op\operatorname{End}_{A}(Q)^{op}.

Let BB be a projective Noetherian RR-algebra, (A,P)(A,P) be an (1)(Δ~)(-1)-\mathcal{F}(\tilde{\Delta}) (quasi-hereditary) cover of BB and (A,P)(A^{\prime},P^{\prime}) be an (1)(Δ~)(-1)-\mathcal{F}(\tilde{\Delta}^{\prime}) (quasi-hereditary) cover of BB. We say that (A,P)(A,P) is equivalent to (A,P)(A^{\prime},P^{\prime}) as quasi-hereditary covers if there exists an equivalence functor H:AmodAmodH\colon A\!\operatorname{-mod}\rightarrow A^{\prime}\!\operatorname{-mod} which restricts to an equivalence of categories between (Δ~)\mathcal{F}(\tilde{\Delta}) and (Δ~)\mathcal{F}(\tilde{\Delta}^{\prime}) making the following diagram commutative

AmodBmodAmodBmodHHomA(P,)LHomA(P,),\leavevmode\hbox to101.28pt{\vbox to53.53pt{\pgfpicture\makeatletter\hbox{\hskip 50.64069pt\lower-23.55057pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-50.64069pt}{-19.89087pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\qquad\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-14.16672pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${A\!\operatorname{-mod}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}}}&\qquad\hfil&\hfil\hskip 42.76563pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-14.46011pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${B\!\operatorname{-mod}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\qquad\hfil\cr\vskip 18.00005pt\cr\hfil\qquad\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-15.5695pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${A^{\prime}\!\operatorname{-mod}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\qquad\hfil&\hfil\hskip 42.76563pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-14.46011pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${B\!\operatorname{-mod}}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}&\qquad\hfil\cr}}}\pgfsys@invoke{ }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{ }\pgfsys@endscope}}} {}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{{ {\pgfsys@beginscope \pgfsys@setdash{}{0.0pt}\pgfsys@roundcap\pgfsys@roundjoin{} {}{}{} {}{}{} \pgfsys@moveto{-2.07988pt}{2.39986pt}\pgfsys@curveto{-1.69989pt}{0.95992pt}{-0.85313pt}{0.27998pt}{0.0pt}{0.0pt}\pgfsys@curveto{-0.85313pt}{-0.27998pt}{-1.69989pt}{-0.95992pt}{-2.07988pt}{-2.39986pt}\pgfsys@stroke\pgfsys@endscope}} }{}{}{{}}{}{}{{}}\pgfsys@moveto{-30.76564pt}{9.08672pt}\pgfsys@lineto{-30.76564pt}{-8.11337pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-30.76564pt}{-8.31335pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-28.41287pt}{-2.10497pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{H}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-12.09338pt}{15.44643pt}\pgfsys@lineto{12.50941pt}{15.44643pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{12.7094pt}{15.44643pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-18.91902pt}{22.57835pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\operatorname{Hom}_{A}(P{,}-)}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{31.87503pt}{9.08672pt}\pgfsys@lineto{31.87503pt}{-8.68677pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{31.87503pt}{-8.88675pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{34.2278pt}{-2.39166pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{L}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-10.6906pt}{-17.39087pt}\pgfsys@lineto{12.50941pt}{-17.39087pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{12.7094pt}{-17.39087pt}\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@invoke{ }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-21.35306pt}{-10.17004pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\operatorname{Hom}_{A^{\prime}}(P^{\prime}{,}-)}$} }}\pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope}}} \pgfsys@invoke{ }\pgfsys@endscope \pgfsys@invoke{ }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{ }\pgfsys@endscope\hss}}\endpgfpicture}},

for some equivalence of categories LL. The first application of uniqueness of covers goes back to [43]. Split quasi-hereditary covers with higher values of Hemmer-Nakano dimension associated to them are essentially unique. In fact, this is due to the following result which can be found in [7, Corollary 4.3.6.].

Corollary 2.5.2.

Let BB be a projective Noetherian RR-algebra, (A,P)(A,P) be a 1(Δ~)1-\mathcal{F}(\tilde{\Delta}) (quasi-hereditary) cover of BB and (A,P)(A^{\prime},P^{\prime}) be a 1(Δ~)1-\mathcal{F}(\tilde{\Delta}^{\prime}) (quasi-hereditary) cover of BB. If there exists an exact equivalence L:BmodBmodL\colon B\!\operatorname{-mod}\rightarrow B\!\operatorname{-mod} which restricts to an exact equivalence between (HomA(P,)Δ~)\mathcal{F}(\operatorname{Hom}_{A}(P,-)\tilde{\Delta}) and (HomA(P,)Δ~)\mathcal{F}(\operatorname{Hom}_{A^{\prime}}(P^{\prime},-)\tilde{\Delta}^{\prime}), then (A,P)(A,P) is equivalent as split quasi-hereditary cover to (A,P)(A^{\prime},P^{\prime}).

For a more detailed exposition on cover theory and Hemmer-Nakano dimensions we refer to [7, 8].

3. The main result

The aim of this section is to prove Theorem A.

3.1. Relative injective dimension

The following concept of relative injective dimension will be useful as a tool in the proof of Theorem A.

Definition 3.1.1.

Let 𝒜\mathcal{A} be a full subcategory of AmodA\!\operatorname{-mod}. We define the 𝒜\mathcal{A}-injective dimension of NAmodN\in A\!\operatorname{-mod} (or the relative injective dimension of NN with respect to 𝒜\mathcal{A}) as the value

inf{n{0}:ExtAi>n(M,N)=0,M𝒜}.\displaystyle\inf\{n\in\mathbb{N}\cup\{0\}\colon\operatorname{Ext}_{A}^{i>n}(M,N)=0,\forall M\in\mathcal{A}\}. (1)

We denote by idim𝒜N\operatorname{idim}_{\mathcal{A}}N the 𝒜\mathcal{A}-injective dimension of NN. Analogously, we define the 𝒜\mathcal{A}-projective dimension of NAmodN\in A\!\operatorname{-mod} as the value

inf{n{0}:ExtAi>n(N,M)=0,M𝒜}.\displaystyle\inf\{n\in\mathbb{N}\cup\{0\}\colon\operatorname{Ext}_{A}^{i>n}(N,M)=0,\forall M\in\mathcal{A}\}. (2)
Lemma 3.1.2.

Let AA be a projective Noetherian RR-algebra and let QAmodQ\in A\!\operatorname{-mod} satisfying ExtAi>0(Q,Q)=0\operatorname{Ext}_{A}^{i>0}(Q,Q)=0. Then, the following assertions hold.

  1. (1)

    If there exists an exact sequence 0XYZ00\rightarrow X\rightarrow Y\rightarrow Z\rightarrow 0 with YaddQY\in\!\operatorname{add}Q, then

    idimQX1+idimQZ.\operatorname{idim}_{{}^{\perp}Q}X\leq 1+\operatorname{idim}_{{}^{\perp}Q}Z.
  2. (2)

    If there exists an exact sequence 0XXrX1Z00\rightarrow X\rightarrow X_{r}\rightarrow\cdots\rightarrow X_{1}\rightarrow Z\rightarrow 0 with X1,,XraddQX_{1},\ldots,X_{r}\in\!\operatorname{add}Q, then idimQXr+idimQZ\operatorname{idim}_{{}^{\perp}Q}X\leq r+\operatorname{idim}_{{}^{\perp}Q}Z.

  3. (3)

    If there exists an exact sequence 0XXrX1Z00\rightarrow X\rightarrow X_{r}\rightarrow\cdots\rightarrow X_{1}\rightarrow Z\rightarrow 0 with X1,,XraddQX_{1},\ldots,X_{r}\in\!\operatorname{add}Q, then for every YQY\in Q^{\perp}, ExtAi(X,Y)ExtAi+r(Z,Y)\operatorname{Ext}_{A}^{i}(X,Y)\simeq\operatorname{Ext}_{A}^{i+r}(Z,Y) for all ii\in\mathbb{N}.

Proof.

For each MQM\in{}^{\perp}Q, applying HomA(M,)\operatorname{Hom}_{A}(M,-) yields that ExtAi(M,Z)ExtAi+1(M,X)\operatorname{Ext}_{A}^{i}(M,Z)\simeq\operatorname{Ext}_{A}^{i+1}(M,X) for all i1i\geq 1. Hence, (i) follows. By induction and using (i), (ii) follows. Denote by CiC_{i} the image of Xi+1XiX_{i+1}\rightarrow X_{i} for all i=1,,r1i=1,\ldots,r-1. By applying HomA(,Y)\operatorname{Hom}_{A}(-,Y) we deduce that ExtAi(X,Y)ExtAi+1(Cr1,Y)ExtAi+2(Cr2,Y)ExtAi+r1(C1,Y)ExtAi+r(Z,Y)\operatorname{Ext}_{A}^{i}(X,Y)\simeq\operatorname{Ext}_{A}^{i+1}(C_{r-1},Y)\simeq\operatorname{Ext}_{A}^{i+2}(C_{r-2},Y)\simeq\operatorname{Ext}_{A}^{i+r-1}(C_{1},Y)\simeq\operatorname{Ext}_{A}^{i+r}(Z,Y). ∎

3.2. Computing relative dominant dimension of the regular module using a characteristic tilting module

In general, the relative codominant dimension of a characteristic tilting module with respect to a partial tilting module gives a lower bound to the relative dominant dimension of the regular module with respect to a partial tilting module (see [8, Theorem 5.3.1(a)]). In the following, we will see that this lower bound can be sharpened using also the relative dominant dimension of a characteristic tilting module with respect to a partial tilting module.

Lemma 3.2.1.

Let (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) be a split quasi-hereditary RR-algebra with a characteristic tilting module TT. Denote by R(A)R(A) the Ringel dual EndA(T)op\operatorname{End}_{A}(T)^{op} of AA. Suppose that QaddTQ\in\!\operatorname{add}T. Then,

Qdomdim(A,R)Qdomdim(A,R)T+Qcodomdim(A,R)T.\displaystyle Q\!\operatorname{-domdim}(A,R)\geq Q\!\operatorname{-domdim}_{(A,R)}T+Q\!\operatorname{-codomdim}_{(A,R)}T. (3)
Proof.

Observe that DQAQRprojDQ\otimes_{A}Q\in R\!\operatorname{-proj} (see for example [7, A.4.3.]). By [8, Theorem 5.3.1(a)] and [8, Corollary 3.1.5], we obtain that

Qdomdim(A,R)A=Qcodomdim(A,R)DAQcodomdim(A,R)T.\displaystyle Q\!\operatorname{-domdim}_{(A,R)}A=Q\!\operatorname{-codomdim}_{(A,R)}DA\geq Q\!\operatorname{-codomdim}_{(A,R)}T. (4)

If QdomdimAT=0Q\!\operatorname{-domdim}_{A}T=0, then there is nothing more to prove. Assume that n:=n:= Qdomdim(A,R)T1{Q\!\operatorname{-domdim}_{(A,R)}T}\geq 1. By Proposition 2.4.1(i), HomA(T,Q)codomdimR(A)DT=n\operatorname{Hom}_{A}(T,Q)\!\operatorname{-codomdim}_{R(A)}DT=n. Then there exists an exact sequence

0CXnX1DT0\displaystyle 0\rightarrow C\rightarrow X_{n}\rightarrow\cdots\rightarrow X_{1}\rightarrow DT\rightarrow 0 (5)

with all XiaddHomA(T,Q)X_{i}\in\!\operatorname{add}\operatorname{Hom}_{A}(T,Q), and so they are projective modules over R(A)R(A). The subcategory (Δ~R(A))\mathcal{F}(\tilde{\Delta}_{R(A)}) is closed under kernels of epimorphisms and since DTDT is a characteristic tilting module over R(A)R(A) we obtain that C(Δ~R(A))C\in\mathcal{F}(\tilde{\Delta}_{R(A)}). Thus, (5) remains exact under TR(A)T\otimes_{R(A)} which is left adjoint to HomA(T,)\operatorname{Hom}_{A}(T,-), and we obtain an exact sequence

0C¯Xn¯X1¯TR(A)DT0\displaystyle 0\rightarrow\overline{C}\rightarrow\overline{X_{n}}\rightarrow\cdots\rightarrow\overline{X_{1}}\rightarrow T\otimes_{R(A)}DT\rightarrow 0 (6)

with all Xi¯addTR(A)HomA(T,Q)=addQ\overline{X_{i}}\in\!\operatorname{add}T\otimes_{R(A)}\operatorname{Hom}_{A}(T,Q)=\!\operatorname{add}Q since QaddTQ\in\!\operatorname{add}T. Moreover, TR(A)DTRprojT\otimes_{R(A)}DT\in R\!\operatorname{-proj} by [7, A.4.3.] and so

TR(A)DTTR(A)HomA(T,DA)DAT\otimes_{R(A)}DT\simeq T\otimes_{R(A)}\operatorname{Hom}_{A}(T,DA)\simeq DA

since DTdomdim(A,R)DA=+{DT\!\operatorname{-domdim}_{(A,R)}DA}=+\infty (see [8, Theorem 3.1.1]). By construction, C¯(~)\overline{C}\in\mathcal{F}(\tilde{\nabla}) and so (6) remains exact under HomA(Q,)\operatorname{Hom}_{A}(Q,-). By the dual version of [8, Corollary 3.1.12], we obtain that

Qcodomdim(A,R)DA=n+Qcodomdim(A,R)C¯.{Q\!\operatorname{-codomdim}_{(A,R)}DA}=n+Q\!\operatorname{-codomdim}_{(A,R)}\overline{C}.

By [8, Theorem 5.3.1(a)], Qcodomdim(A,R)C¯Qcodomdim(A,R)T.{Q\!\operatorname{-codomdim}_{(A,R)}\overline{C}}\geq Q\!\operatorname{-codomdim}_{(A,R)}T.

Surprisingly, the following result generalises [26, Theorem 4.3] without using any techniques on symmetric algebras. In particular, for this proof we do not use the fact that the endomorphism algebra of a faithful projective-injective module over a quasi-hereditary algebra with a simple preserving duality is a symmetric algebra.

Theorem 3.2.2.

Let (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) be a split quasi-hereditary algebra over a field kk. Suppose that there exists a simple preserving duality ():AmodAmod{}^{\diamond}\!(-)\colon A\!\operatorname{-mod}\rightarrow A\!\operatorname{-mod}. Let TT be a characteristic tilting module of AA. Assume that QaddTQ\in\!\operatorname{add}T. Then,

QdomdimA=QcodomdimADA=2QcodomdimAT=2QdomdimAT.\displaystyle Q\!\operatorname{-domdim}A=Q\!\operatorname{-codomdim}_{A}DA=2\cdot Q\!\operatorname{-codomdim}_{A}T=2\cdot Q\!\operatorname{-domdim}_{A}T. (7)
Proof.

By [8, Corollary 3.1.5], [8, Proposition 3.1.6] and Lemma 3.2.1, it remains to show that QcodomdimADA2QcodomdimATQ\!\operatorname{-codomdim}_{A}DA\leq 2\cdot Q\!\operatorname{-codomdim}_{A}T. If QcodomdimAT=+Q\!\operatorname{-codomdim}_{A}T=+\infty then there is nothing to prove. Denote by nn the value QdomdimAT=QcodomdimATQ\!\operatorname{-domdim}_{A}T=Q\!\operatorname{-codomdim}_{A}T. Assume first that n>0n>0. So we can consider again exact sequences of the form (5) and (6). Assume, for a contradiction, that QcodomdimADA>2nQ\!\operatorname{-codomdim}_{A}DA>2n. Hence, also QcodomdimAC¯>nQ\!\operatorname{-codomdim}_{A}\overline{C}>n according [8, dual of Corollary 3.1.12]. So there exists an exact sequence

0LX2n+1¯X2n¯Xn+1¯C¯0\displaystyle 0\rightarrow L\rightarrow\overline{X_{2n+1}}\rightarrow\overline{X_{2n}}\rightarrow\cdots\rightarrow\overline{X_{n+1}}\rightarrow\overline{C}\rightarrow 0 (8)

which remains exact under HomA(Q,)\operatorname{Hom}_{A}(Q,-) and Xi¯addQ\overline{X_{i}}\in\!\operatorname{add}Q, i=n+1,,2n+1i=n+1,\ldots,2n+1. In particular, 0LX2n+1¯X1¯DA00\rightarrow L\rightarrow\overline{X_{2n+1}}\rightarrow\cdots\rightarrow\overline{X_{1}}\rightarrow DA\rightarrow 0 is an HomA(Q,)\operatorname{Hom}_{A}(Q,-)-acyclic coresolution of LL, so it can be used to compute ExtAi(Q,L)\operatorname{Ext}_{A}^{i}(Q,L) for all ii. Since it remains exact under HomA(Q,)\operatorname{Hom}_{A}(Q,-) we obtain that ExtAi>0(Q,L)=0\operatorname{Ext}_{A}^{i>0}(Q,L)=0 and so LQL\in Q^{\perp} and LQ{}^{\diamond}\!L\in{}^{\perp}Q. Let DD be the kernel of the map Xn+1¯C¯\overline{X_{n+1}}\rightarrow\overline{C} and consider the exact sequence

0DXn+1¯Xn¯X1¯DA0.\displaystyle 0\rightarrow D\rightarrow\overline{X_{n+1}}\rightarrow\overline{X_{n}}\rightarrow\cdots\rightarrow\overline{X_{1}}\rightarrow DA\rightarrow 0. (9)

By Lemma 3.1.2(2), idimQDn+1\operatorname{idim}_{{}^{\perp}Q}D\leq n+1 and since (9) remains exact under HomA(Q,)\operatorname{Hom}_{A}(Q,-) we have that DQD\in Q^{\perp}. On the other hand, observe that DD cannot belong to ()\mathcal{F}(\nabla) because otherwise (9) would remain exact under HomA(T,)\operatorname{Hom}_{A}(T,-) yielding that n<HomA(T,Q)codomdimR(A)DTn<\operatorname{Hom}_{A}(T,Q)\!\operatorname{-codomdim}_{R(A)}DT contradicting the definition of nn.

So the exact sequence 0DXn+1¯C¯00\rightarrow D\rightarrow\overline{X_{n+1}}\rightarrow\overline{C}\rightarrow 0 yields that dim()D=1\dim_{\mathcal{F}(\nabla)}D=1. Hence, dim(Δ)D=1\dim_{\mathcal{F}(\Delta)}{}^{\diamond}\!D=1. By Corollary 6 of [40] we obtain that 0ExtA2(D,D)ExtA2(D,D)0\neq\operatorname{Ext}_{A}^{2}({}^{\diamond}\!D,{}^{\diamond}\!{}^{\diamond}\!D)\simeq\operatorname{Ext}_{A}^{2}({}^{\diamond}\!D,D). By Lemma 3.1.2(3) on the exact sequence 0DXn+2¯X2n+1¯L00\rightarrow{}^{\diamond}\!D\rightarrow{}^{\diamond}\!\overline{X_{n+2}}\rightarrow\cdots\rightarrow{}^{\diamond}\!\overline{X_{2n+1}}\rightarrow{}^{\diamond}\!L\rightarrow 0 we obtain 0ExtA2(D,D)ExtAn+2(L,D)0\neq\operatorname{Ext}_{A}^{2}({}^{\diamond}\!D,D)\simeq\operatorname{Ext}_{A}^{n+2}({}^{\diamond}\!L,D). This contradicts idimQD\operatorname{idim}_{{}^{\perp}Q}D being at most n+1n+1. We will now treat the case n=0n=0. Assume, for sake of contradiction, that QcodomdimADA1Q\!\operatorname{-codomdim}_{A}DA\geq 1, then there exists an exact sequence 0LX1DA00\rightarrow L\rightarrow X_{1}\rightarrow DA\rightarrow 0 which remains exact under HomA(Q,)\operatorname{Hom}_{A}(Q,-) and X1addQX_{1}\in\!\operatorname{add}Q. Hence, LQL\in Q^{\perp}, LQ{}^{\diamond}\!L\in{}^{\perp}Q, dim()(L)1\dim_{\mathcal{F}(\nabla)}(L)\leq 1 and the Q{}^{\perp}Q-injective dimension of LL is at most one. In particular, ExtA2(L,L)=0\operatorname{Ext}_{A}^{2}({}^{\diamond}\!L,L)=0. By Proposition 2.4.3, we must have that L()L\in\mathcal{F}(\nabla). But, then applying HomA(T,)\operatorname{Hom}_{A}(T,-) to 0LX1DA00\rightarrow L\rightarrow X_{1}\rightarrow DA\rightarrow 0 yields that HomA(T,Q)codomdimR(A)HomA(T,DA)1\operatorname{Hom}_{A}(T,Q)\!\operatorname{-codomdim}_{R(A)}\operatorname{Hom}_{A}(T,DA)\geq 1 which, in turn, implies that n=QdomdimT1n=Q\!\operatorname{-domdim}T\geq 1 by Proposition 2.4.1(i). ∎

The following will in particular give a positive answer to the Conjecture 6.2.4 of [6].

Proposition 3.2.3.

Let (A,{Δ(λ)λΛ})(A,\{\Delta(\lambda)_{\lambda\in\Lambda}\}) be a split quasi-hereditary algebra over a field kk with a simple preserving duality. Let TT be a characteristic tilting module of AA. Assume that QaddTQ\in\!\operatorname{add}T satisfying QdomdimAA2Q\!\operatorname{-domdim}_{A}A\geq 2. Then,

QdomdimAA=2QdomdimAT2domdimAT=domdimA.\displaystyle Q\!\operatorname{-domdim}_{A}A=2\cdot Q\!\operatorname{-domdim}_{A}T\geq 2\operatorname{domdim}_{A}T=\operatorname{domdim}A. (10)
Proof.

Let PP be a faithful projective-injective module over AA. By assumption, QdomdimAA2Q\!\operatorname{-domdim}_{A}A\geq 2, so by [8, Corollary 3.1.8] it follows that QdomdimAP2Q\!\operatorname{-domdim}_{A}P\geq 2. Since PP is injective we must have that PaddQP\in\!\operatorname{add}Q. Hence, HomA(T,P)addHomA(T,Q)\operatorname{Hom}_{A}(T,P)\in\!\operatorname{add}\operatorname{Hom}_{A}(T,Q).

By Proposition 2.4.1(i),

QdomdimAT\displaystyle Q\!\operatorname{-domdim}_{A}T =HomR(A)(T,Q)codomdimR(A)DTHomA(T,P)codomdimR(A)DT\displaystyle=\operatorname{Hom}_{R(A)}(T,Q)\!\operatorname{-codomdim}_{R(A)}DT\geq\operatorname{Hom}_{A}(T,P)\!\operatorname{-codomdim}_{R(A)}DT
=PdomdimAT=domdimAT.\displaystyle=P\!\operatorname{-domdim}_{A}T=\operatorname{domdim}_{A}T. (11)

Applying Theorem 3.2.2 to the partial tilting modules PP and QQ, the result follows. ∎

4. Input from Schur algebras

The main work to prove the second main result, to determine the Hemmer-Nakano dimension of (Δ)\mathcal{F}(\Delta) over the quasi-hereditary cover for the Temperley-Lieb algebra, in Sections 6 and 7, is done for Schur algebras, and we can work over an algebraically closed field. In this section, we give an outline of the background. To keep the notation simple, we do this for the classical case.

Assume KK is an algebraically closed field. The Schur algebra S=SK(n,d)S=S_{K}(n,d) (or just S(n,d)S(n,d)) of degree dd over KK can be defined in different ways. One can start with the symmetric group 𝒮d\mathcal{S}_{d} which acts (on the right) by place permutations on the tensor power VdV^{\otimes d} where VV is an nn-dimensional vector space. Then the Schur algebra S(n,d)S(n,d) is the endomorphism algebra EndK𝒮d(Vd){\rm End}_{K\mathcal{S}_{d}}(V^{\otimes d}). Analogously, the integral Schur algebra SR(2,d)S_{R}(2,d) is defined as the endomorphism algebra EndR𝒮d((R2)d){\rm End}_{R\mathcal{S}_{d}}((R^{2})^{\otimes d}) where (R2)d(R^{2})^{\otimes d} affords a right R𝒮dR\mathcal{S}_{d}-module structure via place permutations. Alternatively one can construct S(n,d)S(n,d) via the general linear group GL(V)GL(V), for details see for example [30] or [15]. The first route shows that the endomorphism algebra of S(n,d)S(n,d) acting on VdV^{\otimes d} is a quotient of K𝒮dK\mathcal{S}_{d}. The second approach allows one use tensor products and Frobenius twists as tools to study representations.

The Schur algebra S(n,d)S(n,d) is quasi-hereditary, with respect to the dominance order on the set Λ+(n,d)\Lambda^{+}(n,d) of partitions of dd with at most nn parts, which is the standard labelling set for simple modules. It has a simple preserving duality (){}^{\diamond}\!(-) (see for example [18, p.83]). For each partition λ\lambda of dd with at most nn parts, the corresponding simple module will be denoted by L(λ)L(\lambda). We denote the standard module with simple top L(λ)L(\lambda) by Δ(λ)\Delta(\lambda), then the costandard module with simple socle L(λ)L(\lambda) is (λ)=Δ(λ)\nabla(\lambda)={}^{\diamond}\!\Delta(\lambda). For background we refer to [25] or [20], [19].

Of central importance for the quasi-hereditary structure is the characteristic tilting module TT: By [42] the indecomposable modules in (Δ)()\mathcal{F}(\Delta)\cap\mathcal{F}(\nabla) are in bijection with the weights. Write T(λ)T(\lambda) for the indecomposable labelled by λΛ+(n,d)\lambda\in\Lambda^{+}(n,d). Then the direct sum T:=λΛ+(n,d)T(λ)T:=\bigoplus_{\lambda\in\Lambda^{+}(n,d)}T(\lambda) (or a module with the same indecomposable summands) is a distinguished tilting module, known as the characteristic tilting module of SS. Its endomorphism algebra R(S):=EndS(T)opR(S):={\rm End}_{S}(T)^{op} is again quasi-hereditary and R(R(S))R(R(S)) is Morita equivalent (as quasi-hereditary algebra) to SS.

For each λ\lambda, there is an associated exact sequence

0Δ(λ)T(λ)X(λ)0\displaystyle 0\to\Delta(\lambda)\to T(\lambda)\to X(\lambda)\to 0 (12)

where X(λ)X(\lambda) has Δ\Delta-filtration where only Δ(μ)\Delta(\mu) with μ<λ\mu<\lambda occur. We will refer to this as a standard sequence.

We follow the usual practice in algebraic Lie theory to refer to a module in add(T){\rm add}(T) as a tilting module, and to TT as a full tilting module (this will not be ambiguous here).

For the connection between Schur algebras and symmetric groups, the tensor space VdV^{\otimes d} is of central importance. As it happens, the tensor space is a direct sum of tilting modules, and T(λ)T(\lambda) occurs as a summand if and only if λ\lambda is pp-regular (that is does not have pp equal parts). For the quantum case, T(λ)T(\lambda) occurs in the tensor space if and only if λ\lambda is \ell-regular where qq is a primitive \ell-th root of 1. This is proved in [25, 4.2], or combining the reasoning of [25, 4.2] with [18, 2.2(1), 4.3, 4.7] respectively. Hence, the following result has become folklore.

Lemma 4.0.1.

Assume that n=2n=2 and dd is a natural number. If charK2\operatorname{char}K\neq 2 or dd is odd, then VdV^{\otimes d} is a characteristic tilting module over SK(2,d)S_{K}(2,d).

Proof.

If KK has characteristic zero, then the Schur algebra S(2,d)S(2,d) is semi-simple (see for example [30, (2.6)e]) and since VdV^{\otimes d} is faithful over S(2,d)S(2,d) it contains the regular module in its additive closure, and in particular, VdV^{\otimes d} is a characteristic tilting module. If KK has positive characteristic, as discussed before VdV^{\otimes d} is a characteristic tilting module over S(2,d)S(2,d) if and only if all partitions of dd in at most 22 parts are charK\operatorname{char}K-regular partitions of dd. Of course, all partitions of dd in at most 22 parts are pp-regular if p>2p>2. If dd is odd, then there are no partitions of dd in exactly two equal parts. ∎

From now on we assume n=2n=2 and charK=2\operatorname{char}K=2, or in the quantum case that =2\ell=2. We also assume dd is even (unless specified differently).

4.1. On the quasi-hereditary structure of S(2,d)S(2,d)

Let S=S(2,d)S=S(2,d), and let e=ξ(d)e=\xi_{(d)} be the idempotent corresponding to the largest weight (in the notation of [30]). Then SeSSeS is an idempotent heredity ideal and S/SeSS/SeS is isomorphic to S(2,d2)S(2,d-2) (for details see for example [24]). Since SeSSeS is a heredity ideal corresponding to (d)(d), factoring it out is compatible with the quasi-hereditary structure. Furthermore, as it is proved in the appendix of [20] computing Exti{\rm Ext}^{i}’s for S/SeSS/SeS-modules is the same whether in SS or in S/SeSS/SeS. In particular, S(2,d)modS(2,d)\!\operatorname{-mod} is the full subcategory of S(2,d+2)modS(2,d+2)\!\operatorname{-mod} consisting of modules whose composition factors are different from those appearing in the top of SeSe.

We work mostly with the restrictions of simple modules, (co)standard modules and tilting modules to SL(2,K)SL(2,K). Recall that L(λ)L(\lambda) and L(μ)L(\mu) are isomorphic as SL(2,K)SL(2,K)-modules if and only if they can be regarded both as S(2,d)S(2,d)-modules for some large dd and they are isomorphic as S(2,d)S(2,d)-modules. This fact can be seen using the canonical surjective map of KSL(2,K)KSL(2,K) onto S(2,d)S(2,d). Since every partition (λ1,λ2)(\lambda_{1},\lambda_{2}) of dd in at most 22 parts is completely determined by the value λ1λ2\lambda_{1}-\lambda_{2}, it follows that L(λ)L(\lambda) and L(μ)L(\mu) are isomorphic as SL(2,K)SL(2,K)-modules if and only if λ1λ2=μ1μ2\lambda_{1}-\lambda_{2}=\mu_{1}-\mu_{2}. Similarly for standard modules and tilting modules.

We therefore label these modules by m=λ1λ2m=\lambda_{1}-\lambda_{2} if λ=(λ1,λ2)\lambda=(\lambda_{1},\lambda_{2}) (such labellings can also be found for example in [22, Subsection 3.2]). This means that we consider Schur algebras S=S(2,d)S=S(2,d), allowing degrees to vary but keeping the parity. We make the convention that we view tacitly modules for S(2,d)S(2,d^{\prime}) with ddd^{\prime}\leq d of the same parity as modules for S(2,d)S(2,d). We say that such a degree dd^{\prime} is admissible for the module defined in degree dd. With this, the weights labelling the simple modules for S(2,d)S(2,d) are precisely all non-negative integers mdm\leq d of the same parity. The dominance order when p=2p=2 and the degree is even, is the linear order.

The tilting module T(0)T(0) is simple, it is the trivial module for SL(2,K)SL(2,K). As a building block, the tilting module T(1)T(1) appears, which is isomorphic to the natural SL(2,K)SL(2,K)-module VV. Furthermore, T(2)V2T(2)\cong V^{\otimes 2}. For d4d\geq 4 we have that VdV^{\otimes d} is the direct sum of T(k)T(k) where all T(k)T(k) occur for kk of the same parity of dd, except that T(0)T(0) does not occur when dd is even. (See for example [25]).

4.2. The category (Δ)\mathcal{F}(\Delta) and projective modules

Non-split extensions of standard modules satisfy a directedness property, that is

ExtS1(Δ(r),Δ(s))0 impliesr<s.\operatorname{Ext}_{S}^{1}(\Delta(r),\Delta(s))\neq 0\ \ \mbox{ implies}\ \ r<s.

This has the following immediate consequence:

Lemma 4.2.1.

Every module in (Δ)\mathcal{F}(\Delta) has a filtration in which weights of Δ\Delta-quotients increase from top to bottom.

Proof.

This follows for example from [20, Lemma 1.4], see also [7, B.1.6]. See [34], for an earlier reference. ∎

Of main interest for us are the indecomposable projective modules. Let Pd(m)P_{d}(m) denote the indecomposable projective of S(2,d)S(2,d) with simple quotient L(m)L(m). Recall Pd(m)P_{d}(m) has a Δ\Delta-filtration, and that the filtration multiplicities [Pd(m):Δ(w)][P_{d}(m):\Delta(w)] are the same as the decomposition numbers. That is,

[Pd(m):Δ(w)]=((w):L(m))=(Δ(w):L(m)).[P_{d}(m):\Delta(w)]=(\nabla(w):L(m))=(\Delta(w):L(m)).

where we write (M:L(m))(M:L(m)) for the multiplicity of L(m)L(m) as a composition factor of the module MM. Note this also shows that projective modules depend on the degree dd. In this case, decomposition numbers are always 0 or 11, see [31, Prop. 2.2, Theorem 3.2.]. We give an example in Figure 1.

01211411161.1181.11110111.111211..111141...1.11161...1.1111811..111.1120111.11..1111221.111...1.11241.11....1.11126111.....111.112811......11..111301.......1...1.11321.......1...1.1113411......11..111.1136111.....111.11..111381.11....1.111...1.11401.111...1.11....1.11142111.11..111.....111.114411..111.11......11..111461...1.111.......1...1.11\begin{array}[]{r|*{4}c|*{4}c|*{8}c|*{8}cc}\hline\cr 0&1&&&&&&&&&&&&&&&&&&&&&&&&\\ 2&1&1&&&&&&&&&&&&&&&&&&&&&&&\\ 4&1&1&1&&&&&&&&&&&&&&&&&&&&&&\\ 6&1&.&1&1&&&&&&&&&&&&&&&&&&&&&\\ \hline\cr 8&1&.&1&1&1&&&&&&&&&&&&&&&&&&&&\\ 10&1&1&1&.&1&1&&&&&&&&&&&&&&&&&&&\\ 12&1&1&.&.&1&1&1&&&&&&&&&&&&&&&&&&\\ 14&1&.&.&.&1&.&1&1&&&&&&&&&&&&&&&&&\\ \hline\cr 16&1&.&.&.&1&.&1&1&1&&&&&&&&&&&&&&&&\\ 18&1&1&.&.&1&1&1&.&1&1&&&&&&&&&&&&&&&\\ 20&1&1&1&.&1&1.&.&1&1&1&1&&&&&&&&&&&&&&\\ 22&1&.&1&1&1&.&.&.&1&.&1&1&&&&&&&&&&&&&\\ 24&1&.&1&1&.&.&.&.&1&.&1&1&1&&&&&&&&&&&&\\ 26&1&1&1&.&.&.&.&.&1&1&1&.&1&1&&&&&&&&&&&\\ 28&1&1&.&.&.&.&.&.&1&1&.&.&1&1&1&&&&&&&&&&\\ 30&1&.&.&.&.&.&.&.&1&.&.&.&1&.&1&1&&&&&&&&&\\ \hline\cr 32&1&.&.&.&.&.&.&.&1&.&.&.&1&.&1&1&1&&&&&&&&\\ 34&1&1&.&.&.&.&.&.&1&1&.&.&1&1&1&.&1&1&&&&&&&\\ 36&1&1&1&.&.&.&.&.&1&1&1&.&1&1&.&.&1&1&1&&&&&&\\ 38&1&.&1&1&.&.&.&.&1&.&1&1&1&.&.&.&1&.&1&1&&&&&\\ 40&1&.&1&1&1&.&.&.&1&.&1&1&.&.&.&.&1&.&1&1&1&&&&\\ 42&1&1&1&.&1&1&.&.&1&1&1&.&.&.&.&.&1&1&1&.&1&1&&&\\ 44&1&1&.&.&1&1&1&.&1&1&.&.&.&.&.&.&1&1&.&.&1&1&1&&\\ 46&1&.&.&.&1&.&1&1&1&.&.&.&.&.&.&.&1&.&.&.&1&.&1&1&\\ \end{array}

\vdots

Figure 1. Decomposition matrix for S(2,46)S(2,46) for p=2p=2. The (m,n)(m,n)-entry denotes (Δ(m):L(n))(\Delta(m):L(n)), the column label is the same as the row label.

It follows that either Pd(m)Pd2(m)P_{d}(m)\cong P_{d-2}(m) as a module for S(2,d)S(2,d), or else there is a non-split exact sequence

0Δ(d)Pd(m)Pd2(m)0.\displaystyle 0\to\Delta(d)\to P_{d}(m)\to P_{d-2}(m)\to 0. (13)

Namely, the top of Pd2(m)P_{d-2}(m) is L(m)L(m), so there is a surjective homomorphism from Pd(m)P_{d}(m) onto Pd2(m)P_{d-2}(m). Recall that (Δ)\mathcal{F}(\Delta) is closed under kernels of epimorphisms. By the filtration property in Lemma 4.2.1 if this is not an isomorphism, then its kernel is a direct sum of copies of Δ(d)\Delta(d) and there is only one since the decomposition numbers are 1\leq 1.

4.3. Twisted tensor product methods

Let ()F(-)^{F} denote the Frobenius twist (see [18, page 64]), this is an exact functor. In our setting, that is for even characteristic, we have the following tools, due to [15]. Odd degrees when p=2p=2 are less important. Namely, each block of S(2,d)S(2,d) for dd odd is Morita equivalent to some block of some Schur algebra S(2,x)S(2,x) with x=(d1)/2x=(d-1)/2 via the functor Δ(1)()F\Delta(1)\otimes(-)^{F}, see for example [21, Lemma 1] or [16, Section 4, Theorem].

  1. (1)
    1. (a)

      Let m=2tm=2t. There is an exact sequence of SS-modules

      0Δ(t1)FΔ(m)Δ(t)F00\to\Delta(t-1)^{F}\to\Delta(m)\to\Delta(t)^{F}\to 0

      Taking contravariant duals gives the analog for costandard modules.

    2. (b)

      Let m=2t+1m=2t+1, then Δ(m)L(1)Δ(t)F\Delta(m)\cong L(1)\otimes\Delta(t)^{F}.

    (See for example [3, Prop. 3.3]).

We note that this determines recursively the decomposition numbers, as input using that Δ(t)\Delta(t) is simple and isomorphic to L(t)L(t) for t=0,1t=0,1, recall p=2p=2. This can also be used to show that when p=2p=2 and dd is even, the algebra S(2,d)S(2,d) is indecomposable. Further, this also implies, by induction, that the decomposition numbers are always 0 and 11 when p=2p=2 and dd is even.

  1. (2)

    We have a complete description of the indecomposable tilting modules in this case. We have already described T(m)T(m) for m2m\leq 2. The following is due to S. Donkin, see [15, Example 2 p. 47].

Proposition 4.3.1.

Let m=2sm=2s and m2m\geq 2, then

T(m)T(2)T(s1)FT(m)\cong T(2)\otimes T(s-1)^{F}

If m=2s+1m=2s+1, then T(r)T(1)T(s)FT(r)\cong T(1)\otimes T(s)^{F}.

This describes recursively all indecomposable tilting modules. Note that tilting modules are not changed if the degree increases.

The following shows that filtration multiplicities [T(m):Δ(w)][T(m):\Delta(w)] are 1\leq 1.

Proposition 4.3.2.

The Δ\Delta-filtration multiplicities of indecomposable tilting modules in even degree can be computed recursively from

0Δ(2t+2)T(2)Δ(t)FΔ(2t)0.0\to\Delta(2t+2)\to T(2)\otimes\Delta(t)^{F}\to\Delta(2t)\to 0.

To prove this, one may specialize [3, Prop. 3.4].

We will see below that modules T(2)XFT(2)\otimes X^{F} for XX in (Δ)\mathcal{F}(\Delta) have infinite relative dominant dimension with respect to VdV^{\otimes d}. This means that we can use Lemma 2.3.2 (from (3.1.7) of [8]) to relate the relative VdV^{\otimes d}-dominant dimension of the end terms, and this suggests a route towards the proof of our second main result.

We define a twisted filtration of a module M(Δ)M\in\mathcal{F}(\Delta) to be a filtration where each quotient is isomorphic to T(2)Δ(t)FT(2)\otimes\Delta(t)^{F} for some tt.

Lemma 4.3.3.

Let m=2s1m=2s\geq 1. Then the tilting module T(m)T(m) has a twisted filtration

0=MkMk1M1M0=T(m)0=M_{k}\subset M_{k-1}\subset\ldots\subset M_{1}\subset M_{0}=T(m)

with Mi1/MiT(2)Δ(si)FM_{i-1}/M_{i}\cong T(2)\otimes\Delta(s_{i})^{F}, with quotients Δ(2si)\Delta(2s_{i}) and Δ(2si+2)\Delta(2s_{i}+2), for s1<s2<<sks_{1}<s_{2}<\ldots<s_{k}.

Proof.

We have T(m)T(2)T(s1)FT(m)\cong T(2)\otimes T(s-1)^{F}. The module T(s1)T(s-1) has a Δ\Delta-filtration

Nk=0Nk1N0=T(s1)N_{k}=0\subset N_{k-1}\subset\ldots\subset N_{0}=T(s-1)

with Ni1/NiΔ(si)N_{i-1}/N_{i}\cong\Delta(s_{i}) and such that s1<s2<<sks_{1}<s_{2}<\ldots<s_{k}, by Lemma 4.2.1. Applying the exact functor T(2)()FT(2)\otimes(-)^{F} gives the claim. ∎

Remark 4.3.4.
  1. (1)

    The algebra S(2,d)S(2,d) is Ringel self-dual for p=2p=2 and dd even if and only if d=2n+12d=2^{n+1}-2 for some nn, see [21] and for a functorial proof see [22]. In [21] and [22] it was identified precisely which tilting modules are projective (and injective) for these degrees.

  2. (2)

    Each T(m)T(m) has a simple top, this is also part of [21], [22].

5. The relative dominant dimension of the regular module with respect to VdV^{\otimes d}

Let S=S(2,d)S=S(2,d) and assume that KK has characteristic pp. Recall that the indecomposable summands of VdV^{\otimes d} are precisely the T(λ)T(\lambda) where λ\lambda is a partition of dd with at most 22 parts, such that λ\lambda does not have pp equal parts. Recall that we identify λ\lambda with m=λ1λ2m=\lambda_{1}-\lambda_{2}. Hence unless p=2p=2 and dd is even, all indecomposable summands of TT occur in VdV^{\otimes d}, and then VddomdimSS=V^{\otimes d}\!\operatorname{-domdim}_{S}S=\infty, by the following:

Lemma 5.0.1.

If VdV^{\otimes d} has all T(λ)T(\lambda) as direct summands, then

inf{VddomdimSM:M(Δ)}=+.\inf\{V^{\otimes d}\!\operatorname{-domdim}_{S}M\colon M\in\mathcal{F}(\Delta)\}=+\infty.
Proof.

Every module in (Δ)\mathcal{F}(\Delta) admits a finite addT\!\operatorname{add}T-coresolution (see for example [42, Lemma 6] or [14]) which, in particular, remains exact under HomS(,Vd)\operatorname{Hom}_{S}(-,V^{\otimes d}). By Corollary 2.3.5, the result follows. ∎

5.1. The characteristic two case

Lemma 5.0.1 leaves us to consider p=2p=2 and dd even (0(\neq 0). In this case, as mentioned above, the components of VdV^{\otimes d} are the T(m)T(m) with m0m\neq 0. The standard sequence (12) is an addT\!\operatorname{add}T-approximation, this follows from a special case of Proposition 2.3.4. In particular,

VddomdimSΔ(d)=1+VddomdimSX(d).\displaystyle V^{\otimes d}\!\operatorname{-domdim}_{S}\Delta(d)=1+V^{\otimes d}\!\operatorname{-domdim}_{S}X(d). (14)
Theorem 5.1.1.

Let d=2s>0d=2s>0. We have VddomdimS(Δ(d))=s+VddomdimS(T(0))V^{\otimes d}\!\operatorname{-domdim}_{S}(\Delta({d}))=s+V^{\otimes d}\!\operatorname{-domdim}_{S}(T(0)).

To prove this, we will use Lemma 2.3.2 on extensions of Δ(t)\Delta(t) by Δ(2+t)\Delta(2+t).

The cases d=2d=2 and d=4d=4 are easy.

  1. (1)

    For d=2d=2 we have the exact sequence 0Δ(2)T(2)Δ(0)=T(0)00\to\Delta(2)\to T(2)\to\Delta(0)=T(0)\to 0, which proves the statement of the Theorem by Corollary 2.3.5.

  2. (2)

    Let d=4d=4, we have the exact sequence 0Δ(4)T(4)Δ(2)00\to\Delta(4)\to T(4)\to\Delta(2)\to 0. Splicing this with the sequence for Δ(2)\Delta(2) gives the claim.

Degrees d6d\geq 6 need more work. The main ingredient is the observation that subquotients of the form T(2)NFT(2)\otimes N^{F} with N(Δ)N\in\mathcal{F}(\Delta) are not relevant for a minimal VdV^{\otimes d}-approximation.

Lemma 5.1.2.

Let XS(2,s)modX\in S(2,s)\!\operatorname{-mod}. Assume that X(Δ)X\in\mathcal{F}(\Delta). Then the module T(2)XFT(2)\otimes X^{F} has infinite relative dominant dimension with respect to VdV^{\otimes d} for any even degree dd greater or equal to 2s+22s+2.

Proof.

By Lemma 2.3.2 it suffices to prove this when X=Δ(s)X=\Delta(s). We proceed by induction on ss. When s=0s=0 or s=1s=1 we see that T(2)Δ(s)FT(2)\otimes\Delta(s)^{F} is a summand of VdV^{\otimes d}. For the inductive step consider the exact sequence

0T(2)Δ(s)FT(2)T(s)FT(2)X(s)F0.0\to T(2)\otimes\Delta(s)^{F}\to T(2)\otimes T(s)^{F}\to T(2)\otimes X(s)^{F}\to 0.

The middle term is isomorphic to T(2+2s)T(2+2s). Since X(s)X(s) has a filtration with quotients Δ(t)\Delta(t) for t<st<s it follows by induction (and Lemma 2.3.2) that the T(2)X(s)FT(2)\otimes X(s)^{F} has infinite VdV^{\otimes d}-dominant dimension for any even degree d2s+2d\geq 2s+2. We deduce that T(2)Δ(s)FT(2)\otimes\Delta(s)^{F} has infinite VdV^{\otimes d}-dominant dimension as well. ∎

Proof of Theorem 5.1.1.

Assume d=2s4d=2s\geq 4. By Proposition 4.3.2, there exists S(2,d)S(2,d)-exact sequences

0Δ(2t+2)T(2)Δ(t)FΔ(2t)0,\displaystyle 0\rightarrow\Delta(2t+2)\rightarrow T(2)\otimes\Delta(t)^{F}\rightarrow\Delta(2t)\rightarrow 0, (15)

for every 0ts10\leq t\leq s-1. Moreover, they remain exact over HomS(2,d)(,Vd)\operatorname{Hom}_{S(2,d)}(-,V^{\otimes d}) since VdV^{\otimes d} is a partial tilting module and so ExtS(2,d)1(Δ(2t+2),Vd)=0\operatorname{Ext}_{S(2,d)}^{1}(\Delta(2t+2),V^{\otimes d})=0. By Lemma 5.1.2, T(2)Δ(t)FT(2)\otimes\Delta(t)^{F} has infinite relative dominant dimension with respect to VdV^{\otimes d} for 0ts10\leq t\leq s-1. By Lemma 2.3.2,

VddomdimS(2,d)Δ(2t+2)=1+VddomdimS(2,d)Δ(2t),0ts1.\displaystyle V^{\otimes d}\!\operatorname{-domdim}_{S(2,d)}\Delta(2t+2)=1+V^{\otimes d}\!\operatorname{-domdim}_{S(2,d)}\Delta(2t),\quad 0\leq t\leq s-1. (16)

Hence, VddomdimS(2,d)Δ(0)=s+VddomdimS(2,d)Δ(d)V^{\otimes d}\!\operatorname{-domdim}_{S(2,d)}\Delta(0)=s+V^{\otimes d}\!\operatorname{-domdim}_{S(2,d)}\Delta(d). ∎

We will now determine the relative dominant dimension of S(2,d)S(2,d) with respect to VdV^{\otimes d}. Let Pd(m)P_{d}(m) be the indecomposable projective S(2,d)S(2,d)-module with homomorphic image L(m)L(m). Throughout dd and mm are even.

Lemma 5.1.3.

Consider a projective module Pd(m)P_{d}(m) where dd and mm are even with m<dm<d. Then one of the following holds.

  1. (a)

    The number of quotients in a Δ\Delta-filtration of Pd(m)P_{d}(m) is even and Pd(m)P_{d}(m) has a twisted filtration.

  2. (b)

    There is an exact sequence

    0Δ(d)Pd(m)Pd2(m)00\to\Delta(d)\to P_{d}(m)\to P_{d-2}(m)\to 0

    and Pd2(m)P_{d-2}(m) has a twisted filtration.

Proof.

Our strategy consists of proving that the projective module Pd(m)P_{d}(m) is a quotient of a tilting module and then to combine this fact with Lemma 4.3.3. Let rn:=2n+12r_{n}:=2^{n+1}-2. There is a unique nn such that rn1<drnr_{n-1}<d\leq r_{n}. By our convention, we can view Pd(m)P_{d}(m) as a module in degree rnr_{n}. Since it has a simple top isomorphic to L(m)L(m), it is isomorphic to a quotient of Prn(m)P_{r_{n}}(m).

By results in [21], [22], we have the following.

  1. (i)

    If 0m<(rn)/20\leq m<(r_{n})/2, then Prn(m)P_{r_{n}}(m) is a tilting module (in fact, it is isomorphic to T(rnm)T(r_{n}-m)).

  2. (ii)

    For (rn)/2<mrn(r_{n})/2<m\leq r_{n}, the projective module Prn(m)P_{r_{n}}(m) is a factor module of the tilting module T(rn+1m)T(r_{n+1}-m).

We exploit this now. Let m^\hat{m} be the weight as above such that Pd(m)P_{d}(m) is a quotient of T(m^)T(\hat{m}). With the notation as in Lemma 4.3.3, since (Δ)\mathcal{F}(\Delta) is closed under kernels of epimorphisms there is a submodule UT(m^)U\subseteq T(\hat{m}) which has a Δ\Delta-filtration, with MiUMi1M_{i}\subseteq U\subset M_{i-1} where 0<ik0<i\leq k, and Pd(m)T(m^)/UP_{d}(m)\cong T(\hat{m})/U.

If U=MiU=M_{i} then we have part (a). Otherwise, Pd(m)P_{d}(m) has the submodule Mi1/UM_{i-1}/U which is isomorphic to Δ(2si)\Delta(2s_{i}) and U/MiΔ(2si+2)U/M_{i}\cong\Delta(2s_{i}+2). Moreover, Pd(m)/Δ(2si)T(m^)/Mi1P_{d}(m)/\Delta(2s_{i})\simeq T(\hat{m})/M_{i-1} which has a twisted filtration. Since Δ(2si)Pd(m)\Delta(2s_{i})\subset P_{d}(m) we deduce 2sid2s_{i}\leq d. Suppose we have 2si<d2s_{i}<d, then 2si+2d2s_{i}+2\leq d. Hence, T(m^)/MiS(2,d)modT(\hat{m})/M_{i}\in S(2,d)\!\operatorname{-mod}. Since Pd(m)P_{d}(m) is a quotient of the indecomposable T(m^)T(\hat{m}), T(m^)T(\hat{m}) has a simple top isomorphic to L(m)L(m). So, the module T(m^)/MiT(\hat{m})/M_{i} has a simple top isomorphic to L(m)L(m) and is in degree dd, and therefore must be a quotient of Pd(m)P_{d}(m). In particular, we would obtain Mi=UM_{i}=U. This is not so in the case considered. Therefore 2si=d2s_{i}=d and the result follows from (13). ∎

Example 5.1.4.

Consider Figure 1, with d=28d=28. Then rn=30r_{n}=30 and the projective modules Pd(m)P_{d}(m) for 15<m<2815<m<28 are as follows. We have (a) when m=18,20,22,26m=18,20,22,26 and we have (b) when m=16,24m=16,24. Note that cases Pd(m)Pd2(m)P_{d}(m)\cong P_{d-2}(m) occur in (a).

Corollary 5.1.5.

With the setting as in Lemma 5.1.3,
if (a) occurs, then VddomdimSPd(m)=+V^{\otimes d}\!\operatorname{-domdim}_{S}P_{d}(m)=+\infty.
If (b) occurs, then for d=2sd=2s we have VddomdimSPd(m)=VddomdimSΔ(d).V^{\otimes d}\!\operatorname{-domdim}_{S}P_{d}(m)=V^{\otimes d}\!\operatorname{-domdim}_{S}\Delta(d). In particular,

VddomdimSS(2,d)=VddomdimSΔ(d)=(d/2)+VddomdimST(0).V^{\otimes d}\!\operatorname{-domdim}_{S}S(2,d)=V^{\otimes d}\!\operatorname{-domdim}_{S}\Delta(d)=(d/2)+V^{\otimes d}\!\operatorname{-domdim}_{S}T(0).
Proof.

This follows directly from Lemma 5.1.3 and Lemma 2.3.2. ∎

This completes the proof of Theorem B for algebraically closed fields. By [8, Lemma 3.2.3], the result also holds over arbitrary fields.

5.2. The quantum case

Remark 5.2.1.

If qq is not a root of unity then SK,q(2,d)S_{K,q}(2,d) is semi-simple ([18, 4.3(7)]) and VdV^{\otimes d} being faithful is a characteristic tilting module. Otherwise, the summands of VdV^{\otimes d} over SK,q(2,d)S_{K,q}(2,d) are the tilting modules labelled by the \ell-regular partitions of dd in at most 22 parts, where qq is an \ell-root of unity. Hence, replacing charK\operatorname{char}K by \ell in Lemma 4.0.1, we obtain that VdV^{\otimes d} is a characteristic tilting module over SK,q(2,d)S_{K,q}(2,d) if q+10q+1\neq 0 or dd is odd.

For the quantum case, it is enough to take S=SK,q(2,d)S=S_{K,q}(2,d) where q+1=0q+1=0. In this case, everything is exactly the same as over S(2,d)S(2,d) when charK=2\operatorname{char}K=2. Namely, we may take SK,q(2,d)S_{K,q}(2,d) as Aq(2,d)A_{q}(2,d)^{*} as it is done in [18] and [3], and also in [11] and [17]. The definition of Aq(2)A_{q}(2) may be found in [11, p. 16]. This means that one takes the quantum group G(2)G(2) as defined in [11] instead of SL(2,K)SL(2,K).

As it is explained in [22, Sections 3.1 and 3.2], we can use the same labelling for weights, in [22]; in that paper the parameter qq is a primitive \ell-th root of 11 and we only need =2\ell=2. We can regard SK,q(2,d)S_{K,q}(2,d) as a factor algebra of SK,q(2,d+2)S_{K,q}(2,d+2), using [18, Section 4.2], and therefore regard modules in degree dd again as modules in degree dd^{\prime} for d>dd^{\prime}>d of the same parity.

There is a Frobenius morphism from the quantum group G(2)G(2) to the classical setting, hence if Δ(m)\Delta(m) (resp. T(m)T(m)) is a standard module (resp. a tilting module) for the classical setting, then Δ(m)F\Delta(m)^{F} (resp. T(m)FT(m)^{F}) is a module for the quantum group, and so are the tensor products T(2)Δ(m)FT(2)\otimes\Delta(m)^{F} and T(2)T(m)FT(2)\otimes T(m)^{F} modules for the quantum group. The qq-analogues of the exact sequences in Subsection 4.3 and Proposition 4.3.2 exist by [3, Prop. 3.3 and 3.4.]. See also [22, Proposition 3.1] (our situation of interest is recovered by fixing l=2l=2 in their setup). The qq-analogue of Proposition 4.3.1 can be found in [18, Section 3.4, page 73, (8)].

We note that 1 of Subsection 4.3 and the qq-analogue imply, by induction, that all decomposition numbers are 0 or 11.

In [11] it is shown that this version of the qq-Schur algebra is the same as our definition, as the endomorphism algebra of the action of the Iwahori-Hecke algebra on the tensor space VdV^{\otimes d}, see Section 6 below. The definition of the Iwahori-Hecke algebra, as we take it is given in 6.2 below. In particular, we denote by HH the Iwahori-Hecke algebra. Their strategy in [11, Section 3] is to show that the action of the Iwahori-Hecke algebra on the tensor space is a comodule homomorphism (see 3.1.6 of [11]). The HH-action in [11, 3.1.6] is not the same as ours, but it is explained in detail (see 4.4.3 of [11]) that the action we use also can be taken.

Hence, the arguments of Section 5 remain valid in the quantum case and therefore, we obtain the following:

Theorem 5.2.2.

Let KK be a field and fix q=u2q=u^{-2} for some uKu\in K. Let SS be the qq-Schur algebra SK,q(2,d)S_{K,q}(2,d) and TT be the characteristic tilting module of SS. Then,

VddomdimSS=2VddomdimST={d, if 1+q=0 and d is even,+, otherwise .\displaystyle V^{\otimes d}\!\operatorname{-domdim}_{S}S=2\cdot V^{\otimes d}\!\operatorname{-domdim}_{S}T=\begin{cases}d,&\text{ if }1+q=0\text{ and }d\text{ is even},\\ +\infty,&\text{ otherwise }\end{cases}.
Remark 5.2.3.

One might want to know for which mm it is true that VddomdimSPd(m)V^{\otimes d}\!\operatorname{-domdim}_{S}P_{d}(m) is finite. In principle, one can answer this, using the formula in [31] for decomposition numbers. Namely this VdV^{\otimes d}- dominant dimension is finite if and only the number of Δ\Delta-quotients of Pd(m)P_{d}(m) is odd, ie the number of 11s in the column of L(m)L(m).

6. Temperley-Lieb algebras

These algebras were introduced as a model for statistical mechanics ([44]), and then became popular through the work of Jones. In particular, he discovered that they occur as quotients of Iwahori-Hecke algebras ([36, 37]). See also [45] for further details. We give the definition and discuss the connections with Schur algebras.

Definition 6.0.1.

Let RR be a commutative ring and δ\delta an element of RR. The Temperley-Lieb algebra TLR,d(δ)TL_{R,d}(\delta) over RR is the RR-algebra generated by elements U1,U2,,Ud1U_{1},U_{2},\ldots,U_{d-1} with defining relations, here 1i,jd11\leq i,j\leq d-1 such that each term is defined:

  1. (a)

    UiUj=UjUiU_{i}U_{j}=U_{j}U_{i} ( |ij|>1|i-j|>1),

  2. (b)

    Ui2=δUiU_{i}^{2}=\delta U_{i},

  3. (c)

    UiUi+1Ui=UiU_{i}U_{i+1}U_{i}=U_{i}, 1in2,1\leq i\leq n-2,

  4. (d)

    UiUi1Ui=UiU_{i}U_{i-1}U_{i}=U_{i}, 2in1.2\leq i\leq n-1.

It can be viewed as a diagram algebra, with a very extensive literature, but we will not give details since we do use diagram calculations.

6.1. The classical case

We will start by considering the class of Temperley-Lieb algebras which can be viewed as quotients of group algebras of the symmetric group.

Lemma 6.1.1.

There is a surjective algebra homomorphism Φ:R𝒮dTLR,d(2)\Phi:R\mathcal{S}_{d}\to TL_{R,d}(-2) taking the generator Ti=(ii+1)T_{i}=(i\ i+1) of 𝒮d\mathcal{S}_{d} to Ui+1U_{i}+1 for 1id11\leq i\leq d-1.

Proof.

Recall that the group algebra R𝒮dR\mathcal{S}_{d} is generated by the TiT_{i} subject to the relations

  1. (a)

    TiTi+1Ti=Ti+1TiTi+1T_{i}T_{i+1}T_{i}=T_{i+1}T_{i}T_{i+1},

  2. (b)

    TiTj=TjTiT_{i}T_{j}=T_{j}T_{i}  (|ij|>1|i-j|>1),

  3. (c)

    Ti2=1T_{i}^{2}=1,

for 1i,jd11\leq i,j\leq d-1 such that each factor is defined. To show that the map is well-defined one has to check that it preserves these relations; this is straightforward. It is clear that Φ\Phi is surjective, noting that Ui=Φ(TiTi2)U_{i}=\Phi(T_{i}-T_{i}^{2}). ∎

The following description of the kernel of Φ\Phi goes back to [Jon87, p. 364].

Theorem 6.1.2.

For each i=1,2,,d2i=1,2,\ldots,d-2 define

xi:=TiTi+1TiTiTi+1Ti+1Ti+Ti+Ti+11R𝒮d.x_{i}:=T_{i}T_{i+1}T_{i}-T_{i}T_{i+1}-T_{i+1}T_{i}+T_{i}+T_{i+1}-1\in R\mathcal{S}_{d}.

Let II be the ideal of R𝒮dR\mathcal{S}_{d} generated by the xix_{i} for 1id21\leq i\leq d-2. Then there is an exact sequence

0IR𝒮dΦTLR,d(2)00\to I\to R\mathcal{S}_{d}\stackrel{{\scriptstyle\Phi}}{{\to}}TL_{R,d}(-2)\to 0
Proof.

One checks that Φ(xi)=0\Phi(x_{i})=0 for each i=1,,d2i=1,\ldots,d-2. So we have a commutative diagram

0kerΦR𝒮dΦTLR,d(2)0ιidR𝒮dπ0IR𝒮dR𝒮d/I0,\begin{CD}0@>{}>{}>{\rm ker}\Phi @>{}>{}>R\mathcal{S}_{d}@>{\Phi}>{}>TL_{R,d}(-2)@>{}>{}>0\cr&&@A{\iota}A{}A@A{id_{R\mathcal{S}_{d}}}A{}A@A{\pi}A{}A\cr 0@>{}>{}>I@>{}>{}>R\mathcal{S}_{d}@>{}>{}>R\mathcal{S}_{d}/I@>{}>{}>0\end{CD}\ ,

where π\pi maps the image of TiT_{i} in RSd/IRS_{d}/I to Φ(Ti)=Ui+1\Phi(T_{i})=U_{i}+1, and ι\iota is the inclusion map. Consider π:TLR,d(2)R𝒮d/I\pi^{\prime}:TL_{R,d}(-2)\to R\mathcal{S}_{d}/I defined by taking UiU_{i} to the image of Ti1T_{i}-1 in R𝒮d/IR\mathcal{S}_{d}/I. One checks that π\pi^{\prime} preserves the defining relations for TLR,d(2)TL_{R,d}(-2), so that it is a well-defined map. Finally,

π(π(Ti+I))=Ti+I,π(π(Ui))=Ui.\pi^{\prime}(\pi(T_{i}+I))=T_{i}+I,\ \ \pi(\pi^{\prime}(U_{i}))=U_{i}.

Therefore kerΦ=I{\rm ker}\Phi=I. ∎

It is nowadays widely known that Temperley-Lieb algebras can be viewed as the centraliser algebras of quantum groups 𝔰𝔩2\mathfrak{sl}_{2} in the endomorphism algebra of a tensor power and it goes back to the work of Martin [39] and Jimbo [35]. Recall that over RR, the Schur algebra SR(2,d)S_{R}(2,d) is defined as the endomorphism algebra EndR𝒮d((R2)d)\operatorname{End}_{R\mathcal{S}_{d}}((R^{2})^{\otimes d}), where (R2)d(R^{2})^{\otimes d} affords a right R𝒮dR\mathcal{S}_{d}-module structure via place permutation. In order to relate the Temperley-Lieb algebra to the Schur algebra, we need a suitable action of the Temperley-Lieb algebra on the tensor space. It is as follows.

Theorem 6.1.3.

Let VV be a free RR-module of rank 22. Then VdV^{\otimes d} is a module over Λ=TLR,d(2)\Lambda=TL_{R,d}(-2) where UiU_{i} acts as idV(i1)τidV(di1){\rm id}_{V}^{\otimes(i-1)}\otimes\tau\otimes{\rm id}_{V}^{\otimes(d-i-1)}. Here τ\tau is the endomorphism of V2V^{\otimes 2} defined by

τ(v1v2)=v2v1v1v2\tau(v_{1}\otimes v_{2})=v_{2}\otimes v_{1}-v_{1}\otimes v_{2}

(for v1,v2Vv_{1},v_{2}\in V). Moreover there is an algebra isomorphism

ΛEndSR(2,d)(Vd)op.\Lambda\to{\rm End}_{S_{R}(2,d)}(V^{\otimes d})^{op}.
Proof.

We know that R𝒮dR\mathcal{S}_{d} acts by place permutations on VdV^{\otimes d} and we can view this as a right action. With this, Ti1T_{i}-1 acts exactly as the action of UiU_{i} as in the statement. This shows that it factors through Λ\Lambda. In particular, to show that ΛEndS(2,d)(Vd)op\Lambda\rightarrow{\rm End}_{S(2,d)}(V^{\otimes d})^{op} is surjective it is enough to check that the canonical map RSdEndS(2,d)(Vd)opRS_{d}\rightarrow{\rm End}_{S(2,d)}(V^{\otimes d})^{op} is surjective. But this follows from classical Schur–Weyl duality (see for example [38]). In fact, this can be seen in the following way: let KK be a field, then the canonical map K𝒮dEndS(2,d)(Vd)opK\mathcal{S}_{d}\rightarrow\operatorname{End}_{S(2,d)}(V^{\otimes d})^{op} fits in the following commutative diagram

K𝒮d{K\mathcal{S}_{d}}EndS(2,d)(Vd)op{\operatorname{End}_{S(2,d)}(V^{\otimes d})^{op}}EndS(d,d)((Kd)d)op{\operatorname{End}_{S(d,d)}((K^{d})^{\otimes d})^{op}}ψ\scriptstyle{\psi}ϕ\scriptstyle{\phi}

Here, ψ\psi is surjective because domdimSK(d,d)2\operatorname{domdim}S_{K}(d,d)\geq 2 and ϕ\phi is surjective by [25, 1.7] and [18, 4.7] because (Kd)d(K^{d})^{\otimes d} is a projective-injective module over S(d,d)S(d,d). Observe that EndSR(2,d)(Vd)opRproj\operatorname{End}_{S_{R}(2,d)}(V^{\otimes d})^{op}\in R\!\operatorname{-proj} (see for example [7, Proposition A.4.3., Corollary A.4.4.]) and it has a base change property (see for example [7, Corollary A.4.6]). In particular, R(𝔪)REndSR(2,d)(Vd)opEndSR(𝔪)(2,d)(Vd)opR(\mathfrak{m})\otimes_{R}\operatorname{End}_{S_{R}(2,d)}(V^{\otimes d})^{op}\simeq\operatorname{End}_{S_{R(\mathfrak{m})}(2,d)}(V^{\otimes d})^{op} for every maximal ideal 𝔪\mathfrak{m} of RR. By the above discussion, the maps R(𝔪)SdEndSR(𝔪)(2,d)(Vd)opR(\mathfrak{m})S_{d}\rightarrow\operatorname{End}_{S_{R(\mathfrak{m})}(2,d)}(V^{\otimes d})^{op} are surjective for every maximal ideal 𝔪\mathfrak{m} of RR. Now, by Nakayama’s Lemma the map RSdEndSR(2,d)(Vd)opRS_{d}\rightarrow\operatorname{End}_{S_{R}(2,d)}(V^{\otimes d})^{op} is surjective.

It remains to show that the action of Λ\Lambda is injective. Let iaiUiΛ\sum_{i}a_{i}U_{i}\in\Lambda acting as zero on VdV^{\otimes d}. The action of iaiUi\sum_{i}a_{i}U_{i} in yky_{k}, defined as the basis element e1e1e2e1e1e_{1}\otimes\ldots\otimes e_{1}\otimes e_{2}\otimes e_{1}\otimes\ldots\otimes e_{1} where e2e_{2} appears in position k+1k+1, yields that ak=ak+1=0a_{k}=a_{k+1}=0. This concludes the proof. ∎

6.2. The qq-analogue

We shall now discuss the general case of Theorem 6.1.3 and its importance for all Temperley-Lieb algebras.

Let RR be a commutative Noetherian ring with an invertible element uRu\in R. We fix a natural number dd, and we set q:=u2q:=u^{-2}. We take the Iwahori-Hecke algebra H=HR,q(d)H=H_{R,q}(d) to be the RR-algebra with basis {T~ww𝒮d}\{\widetilde{T}_{w}\mid w\in\mathcal{S}_{d}\} with relations

T~wT~s={T~wsif l(ws)=l(w)+1(uu1)T~w+T~ws otherwise.\widetilde{T}_{w}\widetilde{T}_{s}=\left\{\begin{array}[]{ll}\widetilde{T}_{ws}&\mbox{if }l(ws)=l(w)+1\cr(u-u^{-1})\widetilde{T}_{w}+\widetilde{T}_{ws}&\mbox{ otherwise.}\end{array}\right.

here ss runs through the set of transpositions S={(ii+1)1i<d}S=\{(i\ i+1)\mid 1\leq i<d\} in 𝒮d\mathcal{S}_{d}, and where l(w)l(w) is the usual length for w𝒮dw\in\mathcal{S}_{d}, that is the minimal number of transpositions needed in a factorisation of ww.

This presentation corresponds to the presentation used in [12, 13], [11], [18] by

T~w=(u)l(w)Tw.\widetilde{T}_{w}=(-u)^{l(w)}T_{w}.

The algebra HH can also be defined by the braid relations, together with T~s2=(uu1)T~s+1\widetilde{T}_{s}^{2}=(u-u^{-1})\widetilde{T}_{s}+1, that is

(T~su)(T~s+u1)=0(sS).(\widetilde{T}_{s}-u)(\widetilde{T}_{s}+u^{-1})=0\ \ (s\in S).
Lemma 6.2.1.

Let δ=uu1\delta=-u-u^{-1}. Then there is a surjective algebra homomorphism

Φ:HR,q(d)TLR,d(δ)\Phi:H_{R,q}(d)\to TL_{R,d}(\delta)

taking the generator T~i:=T~(ii+1)\widetilde{T}_{i}:=\widetilde{T}_{(i\ i+1)} to Ui+uU_{i}+u for 1id11\leq i\leq d-1.

Proof.

We must show that Φ\Phi is well-defined, that is it respects the relations of the Hecke algebra. It is clearly surjective. We work with the presentation via the braid relations, together with (T~iu)(T~i+u1)=0(\widetilde{T}_{i}-u)(\widetilde{T}_{i}+u^{-1})=0 (for 1i<d11\leq i<d-1). With the definition given, Φ(T~i+u1)=Ui+u+u1=Uiδ\Phi(\widetilde{T}_{i}+u^{-1})=U_{i}+u+u^{-1}=U_{i}-\delta and Φ(T~iu)=Ui\Phi(\widetilde{T}_{i}-u)=U_{i}. Hence we have Φ(T~i+u1)Φ(T~iu)=(Uiδ)Ui=0\Phi(\widetilde{T}_{i}+u^{-1})\Phi(\widetilde{T}_{i}-u)=(U_{i}-\delta)U_{i}=0.

To check the braid relations, we compute

(Ui+u)(Ui+1+u)(Ui+u)=Ui+(uδ)Ui+u(UiUi+1+Ui+1Ui)+2u2Ui+u2Ui+1+u3(U_{i}+u)(U_{i+1}+u)(U_{i}+u)=U_{i}+(u\delta)U_{i}+u(U_{i}U_{i+1}+U_{i+1}U_{i})+2u^{2}U_{i}+u^{2}U_{i+1}+u^{3}

The coefficient of UiU_{i} is equal to u2u^{2}. With this, the expression is symmetric in i,i+1i,i+1 and is therefore equal to (Ui+1+u)(Ui+u)(Ui+1+u)(U_{i+1}+u)(U_{i}+u)(U_{i+1}+u). ∎

We want to determine the kernel of Φ\Phi.

Theorem 6.2.2.

For each i=1,2,,d2i=1,2,\ldots,d-2 define

xi:=T~iT~i+1T~iuT~iT~i+1uT~i+1T~i+u2T~i+u2T~i+1u3H.x_{i}:=\widetilde{T}_{i}\widetilde{T}_{i+1}\widetilde{T}_{i}-u\widetilde{T}_{i}\widetilde{T}_{i+1}-u\widetilde{T}_{i+1}\widetilde{T}_{i}+u^{2}\widetilde{T}_{i}+u^{2}\widetilde{T}_{i+1}-u^{3}\in H.

Let II be the ideal of HR,q(d)H_{R,q}(d) generated by the xix_{i} for 1id21\leq i\leq d-2. Fix δ=uu1\delta=-u-u^{-1}, then there is an exact sequence

0IHR,q(d)ΦTLR,d(δ)0.0\to I\to H_{R,q}(d)\stackrel{{\scriptstyle\Phi}}{{\to}}TL_{R,d}(\delta)\to 0.
Proof.

From the proof of Lemma 6.2.1 we see that

Φ(T~i)Φ(T~i+1)Φ(T~i)=u2(Ui+Ui+1)+u(Ui+1Ui+UiUi+1)+u3\Phi(\widetilde{T}_{i})\Phi(\widetilde{T}_{i+1})\Phi(\widetilde{T}_{i})=u^{2}(U_{i}+U_{i+1})+u(U_{i+1}U_{i}+U_{i}U_{i+1})+u^{3}

and with this one gets that Φ(xi)=0\Phi(x_{i})=0 for all ii. Hence Iker(Φ)I\subseteq{\rm ker}(\Phi). Analogously to the proof of Theorem 6.1.2 replacing the map π\pi^{\prime} with the map TLR,d(δ)HR,q(d)TL_{R,d}(\delta)\rightarrow H_{R,q}(d) defined by taking UiU_{i} to the image of T~iu\widetilde{T}_{i}-u in HR,q(d)/IH_{R,q}(d)/I one proves equality. ∎

Let VdV^{\otimes d} be the free RR-module of rank nn over RR (later we will take n=2n=2). Then VdV^{\otimes d} is a right HH-module, which can be thought of as a deformation of the place permutation action of 𝒮d\mathcal{S}_{d}. Denote by I(n,d)I(n,d) the set of maps {1,,d}{1,,n}\{1,\ldots,d\}\rightarrow\{1,\ldots,n\} and by iji_{j} the image i(j)i(j). If 𝐢I(n,d){\bf i}\in I(n,d) labels the basis element e𝐢=ei1ei2eide_{\bf{i}}=e_{i_{1}}\otimes e_{i_{2}}\otimes\ldots\otimes e_{i_{d}} of VdV^{\otimes d} and s=(tt+1)Ss=(t\ t+1)\in S we write e𝐢se_{\bf i}\cdot s for the basis element obtained by interchanging eite_{i_{t}} and eit+1e_{i_{t+1}}. Then

e𝐢T~s:={e𝐢sit<it+1ue𝐢it=it+1(uu1)e𝐢+e𝐢sit>it+1e_{\bf{i}}\cdot\widetilde{T}_{s}:=\left\{\begin{array}[]{ll}e_{\bf{i}}\cdot s&\ \ i_{t}<i_{t+1}\cr ue_{\bf{i}}&\ \ i_{t}=i_{t+1}\cr(u-u^{-1})e_{\bf{i}}+e_{\bf{i}}\cdot s&\ \ i_{t}>i_{t+1}\end{array}\right.

Focussing on the TL algebra, we take n=2n=2. Recall that the qq-Schur algebra Sq(2,d)S_{q}(2,d) is the endomorphism algebra EndH(Vd){\rm End}_{H}(V^{\otimes d}) via the action as above.

Theorem 6.2.3.

The HH-module structure on VdV^{\otimes d} factors through Φ:HΛ=TLR,d(δ),\Phi:H\to\Lambda=TL_{R,d}(\delta), where δ=uu1\delta=-u-u^{-1}. Hence UsU_{s} acts as

e𝐢Us:={e𝐢sue𝐢it<it+1ue𝐢ue𝐢it=it+1(u)1e𝐢+e𝐢sit>it+1e_{\bf i}U_{s}:=\left\{\begin{array}[]{ll}e_{{\bf i}s}-ue_{\bf i}&i_{t}<i_{t+1}\cr ue_{\bf i}-ue_{\bf i}&i_{t}=i_{t+1}\cr(-u)^{-1}e_{\bf i}+e_{{\bf i}s}&i_{t}>i_{t+1}\end{array}\right.

where s=(tt+1)Ss=(t\ t+1)\in S. Moreover there is an algebra isomorphism

ΛEndSR,q(2,d)(Vd)op.\Lambda\to{\rm End}_{S_{R,q}(2,d)}(V^{\otimes d})^{op}.
Proof.

The first statement follows by checking that the elements xix_{i} act as zero on VdV^{\otimes d}.

The element T~iu\widetilde{T}_{i}-u acts exactly as the action of UiU_{i} in VdV^{\otimes d}, so the canonical map HR,q(d)EndSR,q(2,d)(Vd)op{H_{R,q}(d)\rightarrow\operatorname{End}_{S_{R,q}(2,d)}(V^{\otimes d})^{op}} factors through Λ\Lambda, that is, there is an algebra homomorphism ΛEndSR,q(2,d)(Vd)op.\Lambda\to{\rm End}_{S_{R,q}(2,d)}(V^{\otimes d})^{op}. The same argument as the one given in Theorem 6.1.3 works in this case replacing the Schur algebra by the qq-Schur algebra and the group algebra of the symmetric group by the Iwahori-Hecke algebra. The injectivity follows again by considering the action of the elements in Λ\Lambda, acting as zero on VdV^{\otimes d}, on the elements yky_{k} defined in the exactly same way as in Theorem 6.1.3. ∎

Theorem 6.2.3 places VdV^{\otimes d} in a central position in the representation theory of Temperley-Lieb algebras where it plays a role similar to that played by (Rn)d(R^{n})^{\otimes d} in the study the representation theory of symmetric groups via Schur algebras. In fact, Theorem 8.1.5 of [8] specializes to the following.

Corollary 6.2.4.

Let KK be a field and fix q=u2q=u^{-2} for some element uK×u\in K^{\times}. Let TT be a characteristic tilting module of SS and let R(S)R(S) be the Ringel dual of S:=SK,q(2,d)S:=S_{K,q}(2,d) over a field KK. Then, (R(S),HomS(T,Vd)(R(S),\operatorname{Hom}_{S}(T,V^{\otimes d}) is a (VddomdimST2)(V^{\otimes d}\!\operatorname{-domdim}_{S}T-2)-(ΔR(S))\mathcal{F}(\Delta_{R(S)}) quasi-hereditary cover of TLK,d(uu1)TL_{K,d}(-u-u^{-1}), where ΔR(S)\Delta_{R(S)} denotes the set of standard modules over R(S)R(S). Moreover, the following assertions hold:

  1. (i)

    If q+10q+1\neq 0 or dd is odd, then TLK,d(uu1)TL_{K,d}(-u-u^{-1}) is the Ringel dual of SK,q(2,d)S_{K,q}(2,d), and in particular, it is a split quasi-hereditary algebra over KK;

  2. (ii)

    If q+1=0q+1=0 and dd is even, then (R(S),HomS(T,Vd)(R(S),\operatorname{Hom}_{S}(T,V^{\otimes d}) is a (d22)(\frac{d}{2}-2)-(ΔR(S))\mathcal{F}(\Delta_{R(S)}) quasi-hereditary cover of TLK,d(0)TL_{K,d}(0) and HNdimF(ΔR(S))=d22\operatorname{HNdim}_{F}\mathcal{F}(\Delta_{R(S)})=\frac{d}{2}-2. In particular, the Schur functor F:=HomR(S)(HomS(T,Vd),):R(SK,q(2,d))modTLK,d(0)modF:=\operatorname{Hom}_{R(S)}(\operatorname{Hom}_{S}(T,V^{\otimes d}),-)\colon R(S_{K,q}(2,d))\!\operatorname{-mod}\rightarrow TL_{K,d}(0)\!\operatorname{-mod} induces bijections

    ExtR(S)i(M,N)ExtTLK,d(0)i(FM,FN),M,N(ΔR(S)),0id22.\operatorname{Ext}_{R(S)}^{i}(M,N)\simeq\operatorname{Ext}_{TL_{K,d}(0)}^{i}(FM,FN),\quad\forall M,N\in\mathcal{F}(\Delta_{R(S)}),\quad 0\leq i\leq\frac{d}{2}-2.
Proof.

The result follows from Theorem 5.2.2 and [8, Theorem 8.1.5.] and [8, Theorem 6.0.1]. ∎

7. Uniqueness of the quasi-hereditary cover of TLR,d(δ)TL_{R,d}(\delta)

In Corollary 6.2.4, we construct a quasi-hereditary cover of TLK,q(δ)TL_{K,q}(\delta) using the Ringel dual of a qq-Schur algebra. We will argue now that it is the best quasi-hereditary cover of TLK,d(δ)TL_{K,d}(\delta) if d>2d>2. For that, going to the integral case is helpful. Assume that RR is a commutative Noetherian ring. Let uu be an invertible element of RR and fix q=u2q=u^{-2}. If d=1,2d=1,2 then the Temperley-Lieb algebra TLR,q(uu1)TL_{R,q}(-u-u^{-1}) coincides with the Iwahori-Hecke algebra HR,q(d)H_{R,q}(d) and so this case was dealt in [7, Subsection 7.2].

Assume from now on that d>2d>2. Combining Theorem 6.2.3 with Theorem 8.1.5 of [8] we obtain the following:

Corollary 7.0.1.

Let RR be a commutative Noetherian ring. Fix an element uR×u\in R^{\times} and q=u2q=u^{-2}. Let TT be a characteristic tilting module of (SR,q(2,d),{Δ(λ)λΛ+(2,d)})(S_{R,q}(2,d),\{\Delta(\lambda)_{\lambda\in\Lambda^{+}(2,d)}\}). Denote by R(S)R(S) the Ringel dual of (SR,q(2,d),{Δ(λ)λΛ+(2,d)})(S_{R,q}(2,d),\{\Delta(\lambda)_{\lambda\in\Lambda^{+}(2,d)}\}), that is, R(S)=EndSR,q(2,d)(T)opR(S)=\operatorname{End}_{S_{R,q}(2,d)}(T)^{op}.

Then, (R(S),HomSR,q(2,d)(T,Vd))(R(S),\operatorname{Hom}_{S_{R,q}(2,d)}(T,V^{\otimes d})) is a (VddomdimSR,q(2,d),RT2)(Δ~R(S))(V^{\otimes d}\!\operatorname{-domdim}_{S_{R,q}(2,d),R}T-2)-\mathcal{F}(\tilde{\Delta}_{R(S)}) split quasi-hereditary cover of TLR,d(uu1)TL_{R,d}(-u-u^{-1}).

In the following, we will write R(S)R(S) to denote the Ringel dual EndSR,q(2,d)(T)op\operatorname{End}_{S_{R,q}(2,d)}(T)^{op}. Denote by FR,qF_{R,q} the Schur functor associated with the quasi-hereditary cover constructed in Corollary 7.0.1. The aim now is to compute HNdimFR,q(Δ~R(S))\operatorname{HNdim}_{F_{R,q}}\mathcal{F}(\tilde{\Delta}_{R(S)}) and in particular to determine VddomdimSR,q(2,d),RTV^{\otimes d}\!\operatorname{-domdim}_{S_{R,q}(2,d),R}T in terms of the ground ring RR.

Theorem 7.0.2.

Let RR be a commutative Noetherian ring. Fix an element uR×u\in R^{\times} and q=u2q=u^{-2}. Let TT be a characteristic tilting module of (SR,q(2,d),{Δ(λ)λΛ+(2,d)})(S_{R,q}(2,d),\{\Delta(\lambda)_{\lambda\in\Lambda^{+}(2,d)}\}). Then,

Vddomdim(SR,q(2,d),R)T={d2, if 1+qR× and d is even+, otherwise.\displaystyle V^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T=\begin{cases}\dfrac{d}{2},&\text{ if }1+q\notin R^{\times}\text{ and }d\text{ is even}\\ +\infty,&\text{ otherwise}\end{cases}.
Proof.

Since SR,q(2,d)S_{R,q}(2,d) has the base change property: SRSR,q(2,d)SS,1Sq(2,d)S\otimes_{R}S_{R,q}(2,d)\simeq S_{S,1_{S}\otimes q}(2,d) as SS-algebras for every commutative ring SS which is an RR-algebra and the standard modules of SS,1Sq(2,d)S_{S,1_{S}\otimes q}(2,d) are of the form SRΔ(λ)S\otimes_{R}\Delta(\lambda), λΛ+(n,d)\lambda\in\Lambda^{+}(n,d) (see for example [9, Subsection 3.3, Section 5]), the result follows from Theorem 5.2.2, [7, Propositions A.4.7, A.4.3.] and [8, Theorem 3.2.5.]. ∎

7.1. Hemmer-Nakano dimension of (Δ~R(S))\mathcal{F}(\tilde{\Delta}_{R(S)})

Similarly to the classical case (see also [7]), there are two cases to be considered.

Following [7], the commutative Noetherian ring RR is called 22-partially qq-divisible if 1+qR×1+q\in R^{\times} or 1+q=01+q=0.

Theorem 7.1.1.

Let RR be a local regular 22-partially qq-divisible (commutative Noetherian) ring, where q=u2q=u^{-2}, uR×u\in R^{\times}. Let TT is a characteristic tilting module of SR,q(n,d)S_{R,q}(n,d). Then,

HNdimFR,q(Δ~R(S))=Vddomdim(SR,q(2,d),R)T2.\displaystyle\operatorname{HNdim}_{F_{R,q}}\mathcal{F}(\tilde{\Delta}_{R(S)})=V^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T-2. (17)
Proof.

By Corollary 7.0.1, HNdimFR,q(Δ~R(S))Vddomdim(SR,q(2,d),R)T2\operatorname{HNdim}_{F_{R,q}}\mathcal{F}(\tilde{\Delta}_{R(S)})\geq V^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T-2.

If Vddomdim(SR,q(2,d),R)T=+{V^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T}=+\infty, then dd is odd, and then there is nothing to prove. Assume that it is finite. By Theorem 7.0.2, Vddomdim(SR,q(2,d),R)T=d2V^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T=\frac{d}{2}. In particular, dd is even and 1+qR×1+q\notin R^{\times}. Hence, 1+q1+q must be zero. Therefore,

Q(R)RVddomdim(SQ(R),q(2,d),Q(R))Q(R)RT=d2,{Q(R)\otimes_{R}V^{\otimes d}\!\operatorname{-domdim}_{(S_{Q(R),q}(2,d),Q(R))}Q(R)\otimes_{R}T}=\frac{d}{2},

where Q(R)Q(R) is a quotient field of RR.

By [8, Corollary 5.3.6.], HNdimQ(R)RFR,q(Q(R)RΔR(S))\operatorname{HNdim}_{Q(R)\otimes_{R}F_{R,q}}\mathcal{F}(Q(R)\otimes_{R}\Delta_{R(S)}) cannot be higher than Vddomdim(SR,q(2,d),R)T2{V^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T-2}. It follows that

Vddomdim(SR,q(2,d),R)T2=HNdimQ(R)RFR,q(Q(R)RΔR(S))HNdimFR,q(Δ~R(S)).\displaystyle V^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T-2=\operatorname{HNdim}_{Q(R)\otimes_{R}F_{R,q}}\mathcal{F}(Q(R)\otimes_{R}\Delta_{R(S)})\geq\operatorname{HNdim}_{F_{R,q}}\mathcal{F}(\tilde{\Delta}_{R(S)}).

Theorem 7.1.2.

Let RR be a local regular commutative Noetherian ring which is not a 22-partially qq-divisible commutative ring, where q=u2q=u^{-2}, uR×u\in R^{\times}. Let TT is a characteristic tilting module of SR,q(n,d)S_{R,q}(n,d). Then,

HNdimFR,q(Δ~R(S))=Vddomdim(SR,q(2,d),R)T1.\displaystyle\operatorname{HNdim}_{F_{R,q}}\mathcal{F}(\tilde{\Delta}_{R(S)})=V^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T-1. (18)
Proof.

By Corollary 7.0.1, if Vddomdim(SR,q(2,d),R)TV^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T is infinite, then there is nothing to show. So, assume that Vddomdim(SR,q(2,d),R)TV^{\otimes d}\!\operatorname{-domdim}_{(S_{R,q}(2,d),R)}T is finite. By Theorem 7.0.2, dd is even and 1+qR×1+q\notin R^{\times}. By assumption, 1+q01+q\neq 0, otherwise RR would be a 22-partially qq-divisible ring. It follows that Q(R)RVddomdim(Q(R)RSR,q(2,d),R)Q(R)RTQ(R)\otimes_{R}V^{\otimes d}\!\operatorname{-domdim}_{(Q(R)_{\otimes}RS_{R,q}(2,d),R)}Q(R)\otimes_{R}T is infinite by Theorem 7.0.2. The result for d=2d=2 follows from [7, Theorem 7.2.7]. Assume that d4d\geq 4. By Corollary 7.0.1 and [8, Theorem 3.2.5.], HNdimFR(𝔪),q𝔪(R(𝔪)RΔR(S))0\operatorname{HNdim}_{F_{R(\mathfrak{m}),q_{\mathfrak{m}}}}\mathcal{F}(R(\mathfrak{m})\otimes_{R}\Delta_{R(S)})\geq 0 and HNdimQ(R)RFR,q(Q(R)RΔR(S))=+{\operatorname{HNdim}_{Q(R)\otimes_{R}F_{R,q}}\mathcal{F}(Q(R)\otimes_{R}\Delta_{R(S)})}=+\infty, where 𝔪\mathfrak{m} is the unique maximal ideal of RR, and q𝔪q_{\mathfrak{m}} is the image of qq in R/𝔪R/\mathfrak{m}. By [7, Theorem 5.0.9], HNdimFR,q(Δ~R(S))d21\operatorname{HNdim}_{F_{R,q}}\mathcal{F}(\tilde{\Delta}_{R(S)})\geq\frac{d}{2}-1. The Hemmer-Nakano dimension cannot be higher because similarly to the proof of Theorem 7.2.7 of [7] there exists a prime ideal of height one 𝔭\mathfrak{p} such that 1+q𝔭1+q\in\mathfrak{p}. Hence, Q(R/𝔭)RVddomdimSQ(R/𝔭),q𝔭(2,d),Q(R/𝔭)Q(R/\mathfrak{p})\otimes_{R}V^{\otimes d}\!\operatorname{-domdim}_{S_{Q(R/\mathfrak{p}),q_{\mathfrak{p}}}(2,d),Q(R/\mathfrak{p})} is exactly d2\frac{d}{2}, where q𝔭q_{\mathfrak{p}} denotes the image of qq in R/𝔭Q(R/𝔭)R/\mathfrak{p}\subset Q(R/\mathfrak{p}). The result follows from [7, Theorem 5.1.1]. ∎

7.2. Uniqueness

In this part, assume that R=[x,x1]R=\mathbb{Z}[x,x^{-1}] and fix q=x2q=x^{-2}. Assume that d>2d>2. By [7, Proposition 5.0.3] and Theorem 7.1.2, HNdimFR,q(Δ~R(S))d21\operatorname{HNdim}_{F_{R,q}}\mathcal{F}(\tilde{\Delta}_{R(S)})\geq\frac{d}{2}-1. In particular, the Schur functor FR,qF_{R,q} induces an exact equivalence

(Δ~R(S))(FR,qΔ~R(S)).\displaystyle\mathcal{F}(\tilde{\Delta}_{R(S)})\rightarrow\mathcal{F}(F_{R,q}\tilde{\Delta}_{R(S)}). (19)
Corollary 7.2.1.

(R(S),HomSR,q(2,d)(T,Vd))(R(S),\operatorname{Hom}_{S_{R,q}(2,d)}(T,V^{\otimes d})) is the unique split quasi-hereditary cover of TLR,d(xx1)TL_{R,d}(-x-x^{-1}) satisfying the property (19), where TT is a characteristic tilting module of SR,q(2,d)S_{R,q}(2,d) and R(S)R(S) denotes the Ringel dual of SR,q(2,d)S_{R,q}(2,d). In particular, TLR,d(xx1)TL_{R,d}(-x-x^{-1}) is a split quasi-hereditary algebra over RR if and only if dd is odd.

Proof.

The first statement follows from Corollary 2.5.2 together with [7, Proposition 5.0.3] and Theorem 7.1.2. For the second statement see for example [7, Proposition A.4.7.] or [8, Theorem 6.0.1] together with Theorem 7.0.2). ∎

As a consequence, when dd is odd, the Temperley-Lieb algebra TLR,d(xx1)TL_{R,d}(-x-x^{-1}) is exactly a Ringel dual of S[x,x1],q(2,d)S_{\mathbb{Z}[x,x^{-1}],q}(2,d).

Acknowledgements

Part of the collaboration was done during a research visit of the second-named author to the University of Stuttgart, and the authors are grateful for the support given. The authors would like to thank Steffen Koenig for his comments on an earlier version of this manuscript.

References

  • Aus [71] M. Auslander. Representation dimension of Artin algebras. With the assistance of Bernice Auslander. Queen Mary College Mathematics Notes. London: Queen Mary College. 179 p., 1971.
  • BS [98] A. B. Buan and Ø. Solberg. Relative cotilting theory and almost complete cotilting modules. In Algebras and modules II. Eighth international conference on representations of algebras, Geiranger, Norway, August 4–10, 1996, pages 77–92. Providence, RI: American Mathematical Society, 1998.
  • Cox [98] A. Cox. Ext1{\rm Ext}^{1} for Weyl modules for qq-GL(2,k){\rm GL}(2,k). Math. Proc. Cambridge Philos. Soc., 124(2):231–251, 1998. doi:10.1017/S0305004198002679.
  • CPS [88] E. Cline, B. Parshall, and L. Scott. Finite-dimensional algebras and highest weight categories. J. Reine Angew. Math., 391:85–99, 1988.
  • CPS [90] E. Cline, B. Parshall, and L. Scott. Integral and graded quasi-hereditary algebras, I. Journal of Algebra, 131(1):126–160, 1990. doi:10.1016/0021-8693(90)90169-O.
  • Cru [21] T. Cruz. Algebraic analogues of resolution of singularities, quasi-hereditary covers and Schur algebras. PhD thesis, University of Stuttgart, 2021. URL: http://dx.doi.org/10.18419/opus-11835.
  • [7] T. Cruz. On Noetherian algebras, Schur functors and Hemmer-Nakano dimensions, 2022. arXiv:2208.00291.
  • [8] T. Cruz. On split quasi-hereditary covers and Ringel duality, 2022. arXiv:2210.09344.
  • Cru [23] T. Cruz. Cellular Noetherian algebras with finite global dimension are split quasi-hereditary. Journal of Algebra and its Applications, 2023. doi:10.1142/S0219498824501627.
  • CZ [19] K. Coulembier and R. Zhang. Borelic pairs for stratified algebras. Adv. Math., 345:53–115, 2019. doi:10.1016/j.aim.2019.01.002.
  • DD [91] R. Dipper and S. Donkin. Quantum GLnGL_{n}. Proc. Lond. Math. Soc. (3), 63(1):165–211, 1991. doi:10.1112/plms/s3-63.1.165.
  • DJ [89] R. Dipper and G. James. The q-Schur algebra. Proc. Lond. Math. Soc. (3), 59(1):23–50, 1989. doi:10.1112/plms/s3-59.1.23.
  • DJ [91] R. Dipper and G. James. qq-tensor space and qq-Weyl modules. Trans. Am. Math. Soc., 327(1):251–282, 1991. doi:10.2307/2001842.
  • Don [86] S. Donkin. Finite resolutions of modules for reductive algebraic groups. J. Algebra, 101:473–488, 1986. doi:10.1016/0021-8693(86)90206-1.
  • Don [93] S. Donkin. On tilting modules for algebraic groups. Math. Z., 212(1):39–60, 1993. doi:10.1007/BF02571640.
  • Don [94] S. Donkin. On Schur algebras and related algebras. IV: The blocks of the Schur algebras. J. Algebra, 168(2):400–429, 1994. doi:10.1006/jabr.1994.1236.
  • Don [96] S. Donkin. Standard homological properties for quantum GLn{\rm GL}_{n}. J. Algebra, 181(1):235–266, 1996. doi:10.1006/jabr.1996.0118.
  • Don [98] S. Donkin. The qq-Schur algebra, volume 253 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1998. doi:10.1017/CBO9780511600708.
  • DR [89] V. Dlab and C. M. Ringel. Quasi-hereditary algebras. Illinois Journal of Mathematics, 33(2):280–291, 1989. doi:10.1215/ijm/1255988725.
  • DR [92] V. Dlab and C. M. Ringel. The module theoretical approach to quasi-hereditary algebras. In Representations of algebras and related topics. Proceedings of the Tsukuba international conference, held in Kyoto, Japan, 1990, page 200–224. Cambridge: Cambridge University Press, 1992. doi:10.1017/cbo9780511661853.007.
  • EH [02] K. Erdmann and A. Henke. On Ringel duality for Schur algebras. Math. Proc. Camb. Philos. Soc., 132(1):97–116, 2002. doi:10.1017/S0305004101005485.
  • EL [23] K. Erdmann and S. Law. Torsion pairs and Ringel duality for Schur algebras. Algebr. Represent. Theory, 26(2):411–432, 2023. doi:10.1007/s10468-021-10098-y.
  • EP [04] K. Erdmann and A. E. Parker. On the global and \nabla-filtration dimensions of quasi-hereditary algebras. J. Pure Appl. Algebra, 194(1-2):95–111, 2004. doi:10.1016/j.jpaa.2004.04.005.
  • Erd [93] K. Erdmann. Schur algebras of finite type. Quart. J. Math. Oxford Ser. (2), 44(173):17–41, 1993. doi:10.1093/qmath/44.1.17.
  • Erd [94] K. Erdmann. Symmetric groups and quasi-hereditary algebras. In Finite dimensional algebras and related topics. Proceedings of the NATO Advanced Research Workshop on Representations of algebras and related topics. Ottawa, Canada, August 10-18, 1992, page 123–161. Dordrecht: Kluwer Academic Publishers, 1994. doi:10.1007/978-94-017-1556-0_7.
  • FK [11] M. Fang and S. Koenig. Schur functors and dominant dimension. Trans. Am. Math. Soc., 363(3):1555–1576, 2011. doi:10.1090/s0002-9947-2010-05177-3.
  • FM [19] M. Fang and H. Miyachi. Hochschild cohomology and dominant dimension. Trans. Am. Math. Soc., 371(8):5267–5292, 2019. doi:10.1090/tran/7704.
  • FP [86] E. M. Friedlander and B. J. Parshall. Cohomology of Lie algebras and algebraic groups. Am. J. Math., 108:235–253, 1986. doi:10.2307/2374473.
  • GL [96] J. J. Graham and G. I. Lehrer. Cellular algebras. Invent. Math., 123(1):1–34, 1996. doi:10.1007/BF01232365.
  • Gre [80] J. A. Green. Polynomial representations of GLn{\rm GL}_{n}, volume 830 of Lecture Notes in Mathematics. Springer-Verlag, Berlin-New York, 1980.
  • Hen [01] A. E. Henke. The Cartan matrix of the Schur algebra S(2,r)S(2,r). Arch. Math. (Basel), 76(6):416–425, 2001. doi:10.1007/PL00000452.
  • HN [04] D. J. Hemmer and D. K. Nakano. Specht filtrations for Hecke algebras of type AA. J. Lond. Math. Soc., II. Ser., 69(3):623–638, 2004. doi:10.1112/s0024610704005186.
  • Iya [03] O. Iyama. Finiteness of representation dimension. Proc. Am. Math. Soc., 131(4):1011–1014, 2003. doi:10.1090/S0002-9939-02-06616-9.
  • Jan [80] J. C. Jantzen. Darstellungen halbeinfacher Gruppen und ihrer Frobenius-Kerne. J. Reine Angew. Math., 317:157–199, 1980. doi:10.1515/crll.1980.317.157.
  • Jim [86] M. Jimbo. A qq-analogue of U(𝔤𝔩(N+1))U(\mathfrak{gl}(N+1)), Hecke algebra, and the Yang-Baxter equation. Lett. Math. Phys., 11:247–252, 1986. doi:10.1007/BF00400222.
  • Jon [83] V. F. R. Jones. Index for subfactors. Invent. Math., 72:1–25, 1983. doi:10.1007/BF01389127.
  • Jon [85] V. F. R. Jones. A polynomial invariant for knots via von Neumann algebras. Bull. Am. Math. Soc., New Ser., 12:103–111, 1985. doi:10.1090/S0273-0979-1985-15304-2.
  • KSX [01] S. König, I. H. Slungård, and C. Xi. Double centralizer properties, dominant dimension, and tilting modules. Journal of Algebra, 240(1):393–412, 2001. doi:10.1006/jabr.2000.8726.
  • Mar [92] P. P. Martin. On Schur-Weyl duality, AnA_{n} Hecke algebras and quantum sl(N)sl(N) on n+1N\otimes^{n+1}\mathbb{C}^{N}. Int. J. Mod. Phys. A, 7:645–673, 1992. doi:10.1142/S0217751X92003975.
  • MO [04] V. Mazorchuk and S. Ovsienko. Finitistic dimension of properly stratified algebras. Adv. Math., 186(1):251–265, 2004. doi:10.1016/j.aim.2003.08.001.
  • PS [88] B. Parshall and L. Scott. Derived categories, quasi-hereditary algebras, and algebraic groups. Carlton Univ. Lecture Notes in Math., 3(3):1–104, 1988. URL: http://people.virginia.edu/~lls2l/Ottawa.pdf.
  • Rin [91] C. M. Ringel. The category of modules with good filtrations over a quasi-hereditary algebra has almost split sequences. Math. Z., 208(2):209–223, 1991. doi:10.1007/BF02571521.
  • Rou [08] R. Rouquier. qq-Schur algebras and complex reflection groups. Moscow Mathematical Journal, 8(1):119–158, 184, 2008. doi:10.17323/1609-4514-2008-8-1-119-158.
  • TL [71] H. V. N. 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. Proc. R. Soc. Lond., Ser. A, 322:251–280, 1971. doi:10.1098/rspa.1971.0067.
  • Wes [95] B. W. Westbury. The representation theory of the Temperley-Lieb algebras. Math. Z., 219(4):539–565, 1995. doi:10.1007/BF02572380.