Hostname: page-component-7bb8b95d7b-lvwk9 Total loading time: 0 Render date: 2024-09-28T19:08:34.928Z Has data issue: false hasContentIssue false

Mutating signed $\tau$-exceptional sequences

Published online by Cambridge University Press:  12 October 2023

Aslak Bakke Buan*
Affiliation:
Department of Mathematical Sciences, Norwegian University of Science and Technology, 7491 Trondheim, Norway.
Bethany Rose Marsh
Affiliation:
School of Mathematics, University of Leeds, Leeds, LS2 9JT, UK.
*
Corresponding author: Aslak Bakke Buan; Email: [email protected]
Rights & Permissions [Opens in a new window]

Abstract

We establish some properties of $\tau$-exceptional sequences for finite-dimensional algebras. In an earlier paper, we established a bijection between the set of ordered support $\tau$-tilting modules and the set of complete signed $\tau$-exceptional sequences. We describe the action of the symmetric group on the latter induced by its natural action on the former. Similarly, we describe the effect on a $\tau$-exceptional sequence obtained by mutating the corresponding ordered support $\tau$-tilting module via a construction of Adachi-Iyama-Reiten.

Type
Research Article
Copyright
© The Author(s), 2023. Published by Cambridge University Press on behalf of Glasgow Mathematical Journal Trust

1. Introduction

The usual notion of exceptional sequences in a module category over a finite-dimensional algebra [Reference Crawley-Boevey8, Reference Ringel23] has some drawbacks. In particular, for some non-hereditary algebras, complete exceptional sequences do not exist (see e.g. [Reference Buan and Marsh6, Introduction]). In [Reference Buan and Marsh6], we introduced the notion of $\tau$ -exceptional sequences, motivated by $\tau$ -tilting theory [Reference Adachi, Iyama and Reiten2]. Such sequences can be regarded as an alternative generalisation of exceptional sequences to the non-hereditary case with the property that complete $\tau$ -exceptional sequences always exist. We also introduced signed $\tau$ -exceptional sequences, motivated by the concept of signed exceptional sequences for hereditary algebras [Reference Igusa and Todorov17], and the link to picture groups [Reference Igusa and Todorov17, Reference Igusa, Todorov and Weyman18].

The aim of this paper is to establish further properties of (signed) $\tau$ -exceptional sequences, which we now proceed to discuss in more detail. Recall that a subcategory of a module category is said to be a wide subcategory if it is closed under kernels, cokernels and extensions (and therefore inherits an abelian structure). Let $\Lambda$ be the path algebra of an acyclic quiver with $n$ vertices and ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ the category of finite-dimensional left $\Lambda$ -modules. A $\Lambda$ -module $X$ is said to be exceptional if $\mathop{\textrm{Ext}}\nolimits ^1\!(X,X)=0$ . If $X$ is exceptional, then the subcategory $X^{\perp _{0,1}}$ consisting of modules $Y$ such that $\mathop{\textrm{Hom}}\nolimits\! (X,Y)=0$ and $\mathop{\textrm{Ext}}\nolimits ^1\!(X,Y)=0$ is known as the perpendicular category of $X$ [Reference Geigle and Lenzing14, §1]. By [Reference Geigle and Lenzing14, Prop. 1.1], [Reference Schofield24, Thm. 2.3], $X^{\perp _{0,1}}$ is a wide subcategory of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ equivalent to the module category of the path algebra of a quiver with $n-1$ vertices. An exceptional sequence is a sequence $(X_1,\ldots,X_r)$ where $X_r$ is an indecomposable exceptional $\Lambda$ -module and $(X_1,\ldots,X_{r-1})$ is an exceptional sequence in $X_r^{\perp _{0,1}}$ .

If $r=n$ , then an exceptional sequence $(X_1,X_2,\ldots,X_r)$ is said to be complete. Note that a complete exceptional sequence gives rise to a flag of wide subcategories

\begin{equation*}0=\mathcal {C}_0\subseteq \mathcal {C}_1\subseteq \cdots \subseteq \mathcal {C}_n={\mathop {\textrm {mod}}\nolimits }\, \Lambda \end{equation*}

where, for $1 \leq i \lt n$ , we define $\mathcal{C}_i=(X_{i+1}\amalg X_{i+2}\amalg \cdots \amalg X_n)^{\perp _{0,1}}$ .

The article [Reference Igusa and Todorov17] introduced the notion of a signed exceptional sequence. Let $D^b(\Lambda )$ denote the bounded derived category of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ . Since $\Lambda$ is hereditary, every indecomposable object in $D^b(\Lambda )$ is of the form $X[i]$ , where $[i]$ denotes the $i$ th power of the shift and $X$ is an indecomposable $\Lambda$ -module. We write $|X[i]|=X$ .

A signed exceptional sequence in ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ is a sequence $(X_1,X_2,\ldots,X_r)$ of indecomposable objects in $D^b(\Lambda )$ which are each of the form $Y[j]$ for $j=0$ or $j=1$ for some $\Lambda$ -module $Y$ , where $X_i=Y[1]$ is allowed only if $|X_i|$ is relatively projective in $(|X_{i+1}|\amalg |X_{i+2}|\amalg \cdots \amalg |X_r|)^{\perp _{0,1}}$ , and where $(|X_1|,|X_2|,\ldots, |X_r|)$ is an exceptional sequence.

In [Reference Igusa and Todorov17], signed exceptional sequences were introduced in order to define the cluster morphism category of $\Lambda$ , whose objects are the wide subcategories of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ . The morphisms are described by the signed exceptional sequences. It is shown that the classifying space of the cluster morphism category is a $K(\pi,1)$ , where $\pi$ is the picture group [Reference Igusa, Todorov and Weyman18] of $\Lambda$ .

Now let $\Lambda$ be an arbitrary finite-dimensional algebra over a field. Suppose that $\Lambda$ has $n$ simple modules. The linchpin of the definition of $\tau$ -exceptional sequence is the notion of a $\tau$ -perpendicular category [Reference Jasso19, §1], which plays the role of the Geigle-Lenzing perpendicular category in the general case. A $\Lambda$ -module $X$ is said to be $\tau$ -rigid if $\mathop{\textrm{Hom}}\nolimits\! (X,\tau X)=0$ . Then, the $\tau$ -perpendicular category of $X$ is the subcategory $J(X)$ consisting of modules $Y$ such that $\mathop{\textrm{Hom}}\nolimits\! (X,Y)=0$ and $\mathop{\textrm{Hom}}\nolimits\! (Y,\tau X)=0$ . By [Reference Brüstle, Smith and Treffinger4, Cor. 3.22], [Reference Demonet, Iyama, Reading, Reiten and Thomas10, Thm. 4.12] $J(X)$ is a wide subcategory of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ . By [Reference Jasso19, Thm. 3.8], $J(X)$ is equivalent to the module category of an algebra, which has $n-1$ non-isomorphic simple modules if $X$ is indecomposable. A $\tau$ -exceptional sequence in ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ is a sequence $(X_1,X_2,\ldots,X_r)$ of $\Lambda$ -modules where $X_r$ is $\tau$ -rigid and $(X_1,\ldots,X_{r-1})$ is a $\tau$ -exceptional sequence in $J(X_r)$ (regarded as a module category). Signed $\tau$ -exceptional sequences are then defined in a similar way to signed exceptional sequences (see above, or Section 2). A $\tau$ -exceptional sequence, or signed $\tau$ -exceptional sequence, is said to be complete if it has $n$ terms. Clearly, complete $\tau$ -exceptional sequences and signed $\tau$ -exceptional sequences exist for any finite-dimensional algebra. In [Reference Mendoza and Treffinger22], an interesting interpretation in terms of standardly stratifying systems was given.

Also, $\tau$ -exceptional sequences were used in [Reference Buan and Marsh7] to define the morphisms in the $\tau$ -cluster morphism category of the module category of a $\tau$ -tilting-finite algebra, whose objects are the wide subcategories of the module category. This was extended to an arbitrary finite-dimensional algebra in [Reference Buan and Hanson5]. In [Reference Hanson and Igusa15, Thm. 4.16], it was shown that, if $\Lambda$ is a Nakayama algebra, the classifying space of the $\tau$ -cluster morphism category is a $K(\pi,1)$ for the picture group [Reference Igusa, Todorov and Weyman18] of $\Lambda$ .

In this paper, we study some properties of $\tau$ -exceptional sequences. In Section 3, we prove our first main result, Theorem 3.1, which is restricted to the case of $\tau$ -tilting finite algebras, i.e algebras with a finite number of basic $\tau$ -tilting modules. Under this assumption, we show that if $(X_1,X_2,\ldots,X_i,\ldots X_n)$ and $(X_1,X_2,\ldots,X^{\prime}_i,\ldots,X_n)$ are complete $\tau$ -exceptional sequences, then $X^{\prime}_i \cong X_i$ . We conjecture that this result holds without this assumption.

Suppose now that $\Lambda$ is an arbitrary finite-dimensional algebra. In [Reference Buan and Marsh6, Thm. 5.4], it was shown that there is a bijection between complete signed $\tau$ -exceptional sequences in ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ and ordered support $\tau$ -tilting objects in ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ . Here a support $\tau$ -tilting object is a pair $(P,M)$ where $P$ is projective, $M$ is $\tau$ -rigid and $\mathop{\textrm{Hom}}\nolimits\! (P,M)=0$ , and an ordered support $\tau$ -tilting object is an ordering of the indecomposable summands of $P$ and $M$ (retaining the information as to whether each object is a summand of $P$ or $M$ ). Thus, the symmetric group acts naturally on the set of ordered support $\tau$ -tilting objects and hence, via the bijection, on the set of complete signed $\tau$ -exceptional sequences. In Section 4, we give an explicit description of the action of a simple transposition.

Support $\tau$ -tilting objects can be mutated (see [Reference Adachi, Iyama and Reiten2, §2.3]) and thus so can be ordered $\tau$ -tilting objects. In Section 5, we describe the effect on the corresponding complete $\tau$ -exceptional sequences, translated via the bijection above. We also combine the action of the symmetric group and mutations to give an action of the larger mutation group considered in [Reference King and Pressland20].

Since the braid group on $n$ strands acts transitively on the set of exceptional sequences over a hereditary algebra [Reference Crawley-Boevey8, Reference Ringel23] with $n$ simple modules up to isomorphism, a natural question is whether this braid group acts transitively on the set of all signed $\tau$ -exceptional sequences. In Section 6, we show that, for the Kronecker algebra, there is no transitive action of the braid group on $2$ strands (i.e. the infinite cyclic group) on the set of signed $\tau$ -exceptional sequences which factors through the action of the mutation group referred to above on the set of such sequences (although the mutation group itself does act transitively). We also give an example showing that the obvious generalisation of the definition of the braid action on exceptional sequences to the (signed) $\tau$ -exceptional case does not work, at least without substantial modification.

We remark that after this paper appeared on arxiv.org, a paper by Hanson and Thomas has appeared [Reference Hanson and Thomas16], where the authors used the theory of stability conditions to prove Conjecture 3.7, that is they showed that the uniqueness property in Theorem 3.1 holds for arbitrary finite-dimensional algebras. We thank the referee for their helpful comments on an earlier version of this paper.

2. Background

Let $\Lambda$ be a finite-dimensional basic algebra and denote by ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ the category of finite-dimensional left $\Lambda$ -modules. We let $\tau$ denote the Auslander-Reiten translate on ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ . We assume any subcategories $\mathcal{X}$ to be full and closed under isomorphism; we define $\mathcal{X}^{\perp } = \{Y \in{\mathop{\textrm{mod}}\nolimits }\, \Lambda \mid \mathop{\textrm{Hom}}\nolimits\! (\mathcal{X}, Y) = 0\}$ and define $^{\perp }\mathcal{X}$ dually.

Consider $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda ) ={\mathop{\textrm{mod}}\nolimits }\, \Lambda \amalg{\mathop{\textrm{mod}}\nolimits }\, \Lambda [1]$ as a full subcategory of the bounded derived category $D^b({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ . For an indecomposable object $U$ in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ , we set $\left \lvert U\right \rvert = U$ if $U$ is in ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ and $\left \lvert U\right \rvert = U[\!-\!1]$ if $U$ is in ${\mathop{\textrm{mod}}\nolimits }\, \Lambda [1]$ . If $U$ in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\,\Lambda )$ is basic, we denote by $\mathop{\textrm{rk}}\nolimits\! (U)$ the number of indecomposable summands of $U$ .

We recall some notions from [Reference Adachi, Iyama and Reiten2, §0] (in some cases stated slightly differently, but equivalently). A $\Lambda$ -module $M$ is called $\tau$ -rigid if $\mathop{\textrm{Hom}}\nolimits\! (M,\tau M) = 0$ . A (usually assumed basic) object $M \amalg P[1]$ in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ is said to be a support $\tau$ -rigid object if $M$ is a $\tau$ -rigid $\Lambda$ -module and $P$ is a projective $\Lambda$ -module with $\mathop{\textrm{Hom}}\nolimits _\Lambda\! (P, M)= 0$ . An object $U = M \amalg P[1]$ is said to be a support $\tau$ -tilting object if $\mathop{\textrm{rk}}\nolimits\! (U) =\mathop{\textrm{rk}}\nolimits\!(\Lambda )$ . If, in addition, $P=0$ , $U$ is said to be a $\tau$ -tilting module.

Recall that a subcategory $\mathsf{W}$ of ${\mathop{\textrm{mod}}\nolimits }\,\Lambda$ is said to be wide if it is closed under kernels, cokernels and extensions. If a wide subcategory $\mathsf{W}$ of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ is equivalent to a module category ${\mathop{\textrm{mod}}\nolimits }\, \Lambda ^{\prime}$ , we set $\mathop{\textrm{rk}}\nolimits \mathsf{W} \,:\!= \mathop{\textrm{rk}}\nolimits \Lambda ^{\prime}$ .

Objects which are $\tau$ -rigid give rise to a particular class of wide subcategories.

Definition 2.1. ([Reference Jasso19, Defn. 3.3]). For a support $\tau$ -rigid object $U = M \amalg P[1]$ in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ the category

\begin{equation*}J(U) = (M \amalg P)^{\perp } \cap {^\perp {(\tau M)}}\end{equation*}

is called a $\tau$ -perpendicular subcategory.

In the following Theorem, (a) is from [Reference Demonet, Iyama, Reading, Reiten and Thomas10, Thm. 4.12], [Reference Brüstle, Smith and Treffinger4, Cor. 3.22] and (c) is from [Reference Jasso19, Thm. 3.8]. For (b), see [Reference Enomoto11, Prop. 4.12] and [Reference Enomoto and Sakai12, Lemma 4.7].

Theorem 2.2. A $\tau$ -perpendicular subcategory $J(U)$ of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ is:

  1. (a) wide;

  2. (b) functorially finite;

  3. (c) equivalent to ${\mathop{\textrm{mod}}\nolimits }\, \Lambda _U$ for some finite-dimensional algebra $\Lambda _U$ with $\mathop{\textrm{rk}}\nolimits\! (\Lambda ) = \mathop{\textrm{rk}}\nolimits\! (U) + \mathop{\textrm{rk}}\nolimits\! (\Lambda _U)$ .

Let $\mathsf{W}$ be a $\tau$ -perpendicular subcategory of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ . Since $\mathsf{W}$ is equivalent to a module category, we can also consider the $\tau$ -tilting theory of $\mathsf{W}$ . Let $\mathcal{C}(\mathsf{W}) = \mathsf{W} \amalg \mathsf{W}[1]$ , as a subcategory of $D^b({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ . Note that since $\mathsf{W}$ is an exact subcategory of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ , there is a canonical isomorphism

\begin{equation*}\mathop {\textrm {Hom}}\nolimits _{D^b(\mathsf {W})}(X,Y[1]) \simeq \mathop {\textrm {Hom}}\nolimits _{D^b({\mathop {\textrm {mod}}\nolimits } \Lambda )}(X,Y[1]),\end{equation*}

for modules $X,Y$ in $\mathsf{W}$ , so we can also consider $\mathcal{C}(\mathsf{W})$ as a subcategory of $D^b(\mathsf{W})$ .

Note that in general $\tau _{\mathsf{W}} X \not \simeq \tau X$ for a module $X$ in $\mathsf{W}$ , and hence in general, there exist modules which are $\tau$ -rigid in $\mathsf{W}$ but not $\tau$ -rigid in ${\mathop{\textrm{mod}}\nolimits }\,\Lambda$ (see e.g. [Reference Buan and Marsh6, §1] or the end of Section 6). But we do have the following. For a support $\tau$ -rigid object $V = N \amalg Q[1]$ in $\mathcal{C}(\mathsf{W}) \subseteq \mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ , set

\begin{equation*}J_{\mathsf{W}}(V) = (N \amalg Q)^{\perp } \cap {^{\perp } {(\tau _{\mathsf{W}} N)}}\cap \mathsf{W}.\end{equation*}

The following useful Lemma follows from [Reference Auslander and Smalø1, Prop. 5.8] (see also [Reference Adachi, Iyama and Reiten2, Prop. 1.2]).

Lemma 2.3. Let $\mathsf{W}$ be a $\tau$ -perpendicular subcategory of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ and assume $X, Y$ lie in $\mathsf{W}$ . Then, the following hold:

  1. (a) $\mathop{\textrm{Hom}}\nolimits\! (Y, \tau _{\mathsf{W}} X) = 0$ if and only if $\mathop{\textrm{Ext}}\nolimits ^1\!(X, \mathsf{W} \cap \mathop{\textrm{Gen}}\nolimits Y) = 0$ .

  2. (b) $X$ is $\tau$ -rigid in $\mathsf{W}$ if and only if $\mathop{\textrm{Ext}}\nolimits ^1\!(X, \mathsf{W} \cap \mathop{\textrm{Gen}}\nolimits X) =0$ .

Lemma 2.4. Let $\mathsf{W}^{\prime} \subseteq \mathsf{W}$ be $\tau$ -perpendicular subcategories of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ , and let $X$ be an object in $\mathsf{W}^{\prime}$ . If $X$ is $\tau$ -rigid in $\mathsf{W}$ , then $X$ is also $\tau$ -rigid in $\mathsf{W}^{\prime}$ .

Proof. By Lemma 2.3, $X$ is $\tau$ -rigid in $\mathsf{W}$ implies $\mathop{\textrm{Ext}}\nolimits ^1\!(X,\mathsf{W} \cap \mathop{\textrm{Gen}}\nolimits X ) = 0$ . Hence, also $\mathop{\textrm{Ext}}\nolimits ^1\!(X,\mathsf{W}^{\prime} \cap \mathop{\textrm{Gen}}\nolimits X ) = 0$ , and applying Lemma 2.3 again we obtain that $X$ is $\tau$ -rigid in $\mathsf{W}^{\prime}$ .

The following bijection is crucial. It was proved in [Reference Buan and Marsh6] and can be seen as a refinement of [Reference Jasso19, Thm. 3.16].

Theorem 2.5 ([Reference Buan and Marsh6, Prop. 5.6]). Let $\mathsf{W}$ be a $\tau$ -perpendicular subcategory of ${\mathop{\textrm{mod}}\nolimits }\,\Lambda$ , and let $U$ be a support $\tau$ -rigid object in $\mathcal{C}(\mathsf{W})$ . Then, there is a bijection $\mathcal{E}^{\mathsf{W}}_{U}$ from

\begin{equation*}\left\{X \in \mathop {\textrm {ind}}\nolimits\! (\mathcal {C}(\mathsf {W})) \mid X \amalg U \textit { support } \tau \textit {-rigid in } \mathcal {C}(\mathsf {W})\right\} \setminus \mathop {\textrm {ind}}\nolimits U \end{equation*}

to

\begin{equation*}\left\{X \in \mathop {\textrm {ind}}\nolimits\! (\mathcal {C}(J_{\mathsf {W}}(U)) \mid \textit {$X$ is support $\tau $-rigid in $\mathcal {C}(J_{\mathsf {W}}(U))$} \right\}.\end{equation*}

We denote the inverse of $\mathcal{E}^{\mathsf{W}}_{U}$ by $\mathcal{F}^{\mathsf{W}}_{U}$ and, when $\mathsf{W} ={\mathop{\textrm{mod}}\nolimits }\, \Lambda$ , we denote the map in Theorem 2.5 and its inverse simply by $\mathcal{E}_{U}$ and $\mathcal{F}_{U}$ .

Using the bijection in Theorem 2.5, the following was proved in [Reference Buan and Marsh7, Thms. 1.4, 1.7] for the $\tau$ -tilting finite case. It was generalised in [Reference Buan and Hanson5, Thms. 6.4, 6.12] to arbitrary finite-dimensional algebras.

Theorem 2.6 ([Reference Buan and Hanson5, Reference Buan and Marsh7]). Let $U \amalg V$ be a $\tau$ -rigid object in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ .

  1. (a) We have $J_{J(U)}\, (\mathcal{E}_{U}(V))= J(U \amalg V)$ .

  2. (b) We have $\mathcal{E}_{U \amalg V} = (\mathcal{E}^{J(U)}_{\mathcal{E}_{U}(V)})\, \mathcal{E}_{U}$ :

We also recall the following.

Lemma 2.7 ([Reference Buan and Marsh7, Lemma 4.5]). Let $\mathsf{V}$ and $\mathsf{W}$ be wide subcategories of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ with $\mathsf{V} \subseteq \mathsf{W}$ . Then, $\mathsf{V}$ is a wide subcategory of $\mathsf{W}$ .

3. Uniqueness

Let $n$ be the number of simple $\Lambda$ -modules. Recall that a complete $\tau$ -exceptional sequence in ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ is a sequence $(X_1,X_2,\ldots,X_n)$ of indecomposable $\Lambda$ -modules where $X_n$ is $\tau$ -rigid and $(X_1,\ldots,X_{n-1})$ is a $\tau$ -exceptional sequence in $J(X_n)$ . Moreover, a sequence $(X_1,X_2,\ldots,X_n)$ of indecomposable objects in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ is a signed $\tau$ -exceptional sequence, if (i) $X_n$ is either a $\tau$ -rigid module or of the form $P[1]$ for some projective $\Lambda$ -module $P$ and (ii) $(X_1,X_2,\ldots,X_{n-1})$ is a signed $\tau$ -exceptional sequence in $J(\left \lvert X_n\right \rvert )$ . Note that this means that $X_{n-1}$ is either $\tau$ -rigid in $J(\left \lvert X_n\right \rvert )$ (i.e. $\tau$ -rigid in the equivalent module category), or $X_{n-1}= P^{\prime}[1]$ , where $P^{\prime}$ is (relative) projective in $J(\left \lvert X_n\right \rvert )$ , and so on.

Recall that $\Lambda$ is said to be $\tau$ -tilting finite if it only has a finite number of indecomposable $\tau$ -rigid modules. In this section, we shall prove the following uniqueness result for $\tau$ -exceptional sequences over such algebras:

Theorem 3.1. Let $\Lambda$ be a $\tau$ -tilting finite algebra. Then, the following hold.

  1. (a) Let $(A_1, A_2, \dots, A_n)$ and $(B_1, B_2, \dots, B_n)$ be complete $\tau$ -exceptional sequences in ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ . If, for some $t \in \{1, \dots, n\}$ , we have $A_i= B_i$ for all $i \neq t$ , then also $A_t = B_t$ .

  2. (b) Let $(A_1, A_2, \dots, A_n)$ and $(B_1, B_2, \dots, B_n)$ be complete signed $\tau$ -exceptional sequences in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ . If, for some $t \in \{1, \dots, n\}$ , we have $\left \lvert A_i\right \rvert = \left \lvert B_i\right \rvert$ for all $i \neq t$ , then also $\left \lvert A_t\right \rvert = \left \lvert B_t\right \rvert$ .

We first recall the following:

Theorem 3.2. Let $\Lambda$ be a $\tau$ -tilting finite algebra. Then, the following hold for any wide subcategory $\mathsf{W}$ of ${{\textrm{mod}}}\, \Lambda$ .

  1. (a) [Reference Demonet, Iyama, Reading, Reiten and Thomas10, Thm. 4.18] We have $\mathsf{W} = J(U)$ for some support $\tau$ -rigid object $U$ in $\mathcal{C}({{\textrm{mod}} }\, \Lambda )$ .

  2. (b) [Reference Jasso19, Thm. 3.8, Thm. 3.16] The wide subcategory $\mathsf{W}$ is $\tau$ -tilting finite.

We next make the following observation, which holds for all finite-dimensional algebras:

Lemma 3.3. If $(A_1, A_2, \dots, A_n)$ is a complete signed $\tau$ -exceptional sequence, then $(\!\left \lvert A_1\right \rvert\!, \left \lvert A_2\right \rvert\!, \dots, \left \lvert A_n\right \rvert\! )$ is a complete (unsigned) $\tau$ -exceptional sequence.

Proof. We first claim that $(A_1, A_2, \dots, A_{n-1}, \left \lvert A_n\right \rvert )$ is a signed $\tau$ -exceptional sequence. If $\left \lvert A_n\right \rvert$ is projective, then $J(\left \lvert A_n\right \rvert [1]) = J(\left \lvert A_n\right \rvert )$ . Hence, the initial claim follows from the definition of signed $\tau$ -exceptional sequences. The same argument gives that also $(A_1, A_2, \dots, A_{n-2}, \left \lvert A_{n-1}\right \rvert\!, \left \lvert A_n\right \rvert )$ is a signed $\tau$ -exceptional sequence, and so on.

It is clear that Lemma 3.3 and Theorem 3.1 (a) imply Theorem 3.1 (b), so it is enough to prove Theorem 3.1 (a).

In the remainder of this section, we will prove Theorem 3.1 (a). So, we assume for the remainder of the section that $\Lambda$ is $\tau$ -tilting finite.

We then have the following:

Lemma 3.4. Let $\mathsf{W}$ , $\mathsf{W}^{\prime}$ be wide subcategories of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ with $\mathsf{W}^{\prime} \subseteq \mathsf{W}$ . Then:

  1. (a) We have $\mathop{\textrm{rk}}\nolimits \mathsf{W}^{\prime}\leq \mathop{\textrm{rk}}\nolimits \mathsf{W}$ ;

  2. (b) If $\mathop{\textrm{rk}}\nolimits \mathsf{W}=\mathop{\textrm{rk}}\nolimits \mathsf{W}^{\prime}$ then $\mathsf{W}=\mathsf{W}^{\prime}$ .

Proof. For (a), we note that, by Lemma 2.7, $\mathsf{W}^{\prime}$ is a wide subcategory of $\mathsf{W}$ . By Theorem 3.2, $\mathsf{W}^{\prime}$ is of the form $J_{\mathsf{W}}(U)$ for some $\tau$ -rigid object $U$ in $\mathsf{W}$ . Hence, by Theorem 2.2, $\mathop{\textrm{rk}}\nolimits \mathsf{W}^{\prime}=\mathop{\textrm{rk}}\nolimits \mathsf{W}-r\leq \mathop{\textrm{rk}}\nolimits \mathsf{W}$ , where $r$ is the number of non-isomorphic indecomposable direct summands of $U$ .

For (b) suppose, in addition, that $\mathop{\textrm{rk}}\nolimits \mathsf{W}=\mathop{\textrm{rk}}\nolimits \mathsf{W}^{\prime}$ . Then, $r=0$ in the above, so $U=0$ , and we have $\mathsf{W}=\mathsf{W}^{\prime}$ as required.

We give an alternative proof of (a) at the end of this section.

Lemma 3.5. If $(A_1, A_2, \dots, A_{n-1}, A_n)$ and $\left(A_1, A_2, \dots, A_{n-1}, A^{\prime}_n\right)$ are complete $\tau$ -exceptional sequences in ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ , then $A_n = A^{\prime}_n$ .

Proof. The result is clear for $n=1$ , so we may assume that $n\geq 2$ . Let $\mathsf{W}_n={\mathop{\textrm{mod}}\nolimits }\, \Lambda$ , $\mathsf{W}_{n-1}= J_{\mathsf{W}_n}(A_n)=J(A_n)$ , $\mathsf{W}_{n-2}= J_{\mathsf{W}_{n-1}}(A_{n-1})$ , $\ldots$ , $\mathsf{W}_1= J_{\mathsf{W}_{2}}(A_2)$ . By Proposition 2.2, we have $\mathop{\textrm{rk}}\nolimits \mathsf{W}_i=i$ for $1\leq i\leq n$ . For $0\leq i\leq n-1$ , let $\mathcal{X}_i=\mathsf{W}_{i+1}\cap J(A^{\prime}_n)$ , so $\mathcal{X}_0\subseteq \mathcal{X}_1\subseteq \cdots \mathcal{X}_{n-1}=J(A^{\prime}_n)$ .

By Proposition 2.2, we have $\mathop{\textrm{rk}}\nolimits J(A_n)=\mathop{\textrm{rk}}\nolimits J(A^{\prime}_n)=n-1$ . We have $\mathcal{X}_{n-2}=J(A_n)\cap J(A^{\prime}_n)$ . Assume, for a contradiction, that $\mathop{\textrm{rk}}\nolimits \mathcal{X}_{n-2}\leq n-2$ . Note that $A_1,\ldots,A_{n-1}\in J(A^{\prime}_n)$ . Fix $1\leq i\leq n-2$ . Since $A_{i+1}\in \mathsf{W}_{i+1}=J_{\mathsf{W}_{i+2}}(A_{i+2})$ , we have $A_{i+1}\in \mathcal{X}_i=\mathsf{W}_{i+1}\cap J(A^{\prime}_n)$ . However, $A_{i+1}\not \in \mathsf{W}_i=J_{\mathsf{W}_{i+1}}(A_{i+1})$ , since $\mathop{\textrm{Hom}}\nolimits\! (A_{i+1},A_{i+1})\not =0$ , so $A_{i+1}\not \in \mathcal{X}_{i-1}=\mathsf{W}_i\cap J(A^{\prime}_n)$ . Since $\mathsf{W}_i\subseteq \mathsf{W}_{i+1}$ we see that $\mathcal{X}_{i-1}\subseteq \mathcal{X}_i$ but $\mathcal{X}_{i-1}\not =\mathcal{X}_i$ . Hence, by Lemma 3.4, we have $\mathop{\textrm{rk}}\nolimits \mathcal{X}_{i-1}\lt \mathop{\textrm{rk}}\nolimits \mathcal{X}_i$ . Since $\mathop{\textrm{rk}}\nolimits \mathcal{X}_{n-2}\leq n-2$ , it follows that $\mathop{\textrm{rk}}\nolimits \mathcal{X}_i\leq i$ for $0\leq i\leq n-2$ .

In particular, this means that $\mathcal{X}_0=\mathsf{W}_1\cap J(A^{\prime}_n)=J_{\mathsf{W}_2}(A_2)\cap J(A^{\prime}_n)$ is zero. But this gives a contradiction, since $0\not =A_1\in \mathcal{X}_0$ . Hence, we must have $\mathop{\textrm{rk}}\nolimits \mathcal{X}_{n-2}\geq n-1$ .

Since $\mathcal{X}_{n-2}\subseteq J(A_n)$ , it follows again from Lemma 3.4 that $\mathop{\textrm{rk}}\nolimits \mathcal{X}_{n-2}\leq \mathop{\textrm{rk}}\nolimits J(A_n)=n-1$ , so $\mathop{\textrm{rk}}\nolimits \mathcal{X}_{n-2}=n-1$ . Since $\mathcal{X}_{n-2}\subseteq J(A_n)$ and $\mathcal{X}_{n-2}\subseteq J(A^{\prime}_n)$ , Lemma 3.4 implies that $J(A_n)=J(A_n)\cap J(A^{\prime}_n)=J(A^{\prime}_n)$ , and hence, $A_n=A^{\prime}_n$ by [Reference Buan and Marsh7, Proposition 10.7].

We note the following:

Corollary 3.6. Let $\mathsf{W}$ and $\mathsf{W}^{\prime}$ be wide subcategories of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ and assume that $(A_1, A_2, \dots, A_m)$ is a complete $\tau$ -exceptional sequence in both $\mathsf{W}$ and $\mathsf{W}^{\prime}$ . Then, we have $\mathsf{W} = \mathsf{W}^{\prime}$ .

Proof. This follows using the same argument as in the proof of Lemma 3.5, replacing $J(A_n)$ with $\mathsf{W}$ , $J(A^{\prime}_n)$ with $\mathsf{W}^{\prime}$ and replacing $A_1,A_2,\ldots A_{n-1}$ with $A_1,A_2,\ldots,A_m$ .

We can now complete the proof of Theorem 3.1 (a) and hence the main theorem of this section.

Proof of Theorem 3.1 (a). If $t= n$ , this follows directly from Lemma 3.5. Assume $t \in \{1, \dots, n-1\}$ . Let $W^A_n = J(A_n)$ , and for $j \in \{t, \dots, n-1\}$ define recursively $\mathsf{W}^A_j = J_{\mathsf{W}^A_{j+1}}(A_j)$ . Define similarly $\mathsf{W}^B_n = J(B_n)$ and $\mathsf{W}^B_j = J_{\mathsf{W}^B_{j+1}}(B_j)$ . Then, $\mathsf{W}^A_{t+1} = \mathsf{W}^B_{t+1} \,:\!= \mathsf{W}^{\prime}$ , and $\mathsf{W}^{\prime} \simeq{\mathop{\textrm{mod}}\nolimits }\, \Lambda ^{\prime}$ for a finite-dimensional algebra $\Lambda ^{\prime}$ , and we have that $(A_1, A_2, \dots, A_{t-1}, A_t)$ and $(B_1, B_2, \dots, B_{t-1}, B_t) = (A_1, A_2, \dots, A_{t-1}, B_t)$ are complete exceptional sequences in $\mathsf{W}^{\prime}$ . Hence, we obtain that $A_t = B_t$ by Lemma 3.5.

We note that the uniqueness property of Theorem 3.1 also holds for arbitrary finite-dimensional hereditary algebras, by [Reference Crawley-Boevey8, Lemma 2], and we conjecture that the assumption on $\tau$ -tilting finiteness should not be necessary.

Conjecture 3.7. The statement of Theorem 3.1 holds for all finite-dimensional algebras.

An alternative proof of Lemma 3.4 (a) can be given using the theory of bricks. Recall that a $\Lambda$ -module $M$ is called a brick if $\mathop{\textrm{End}}\nolimits\! (M)$ is a division algebra. A set of isoclasses of pairwise Hom-orthogonal bricks is called a semibrick. Let $C$ be a full subcategory of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ . We denote by $T(C)$ the smallest torsion class containing $C$ , by $\mathop{\textrm{Gen}}\nolimits\! (C)$ the collection of modules obtained as quotients of finite direct sums of modules in $C$ , and by $\mathop{\textrm{Filt}}\nolimits\! (C)$ the category of modules with filtrations by modules in $C$ . By the argument in [Reference Marks and Šťovíček21, Lemma 3.1], $T(C)=\mathop{\textrm{Filt}}\nolimits\! (\!\mathop{\textrm{Gen}}\nolimits\! (C))$ . A semibrick $S$ is called left finite [Reference Asai3, Definition 1.2] if $T(S)$ is functorially finite. Let $n_{\Lambda }$ denote the number of isomorphism classes of simple $\Lambda$ modules. We recall:

Proposition 3.8 ([Reference Asai3, 1.10]). If $S$ is a left finite semibrick, then $|S|\leq n_{\Lambda }$ .

Proposition 3.9 ([Reference Demonet, Iyama and Jasso9, 1.2]). Let $A$ be a $\tau$ -tilting finite algebra, and let $T$ be a torsion class in ${\mathop{\textrm{mod}}\nolimits }\, A$ . Then, $T$ is functorially finite.

Proof of Lemma 3.4 (a). Note that, by [Reference Jasso19, Thm. 3.8], $\mathsf{W}\simeq{\mathop{\textrm{mod}}\nolimits }\, \Lambda ^{\prime}$ for some finite-dimensional algebra $\Lambda ^{\prime}$ . Let $S$ be the set of isoclasses of simple objects in $\mathsf{W}^{\prime}$ . Then $S$ is a semibrick in $\mathsf{W}$ . By [Reference Buan and Marsh7, Prop. 4.2(b)], $\Lambda ^{\prime}$ is $\tau$ -tilting finite. Hence, by Proposition 3.9, $S$ is left finite in $\mathsf{W}$ . By Proposition 3.8, $\mathop{\textrm{rk}}\nolimits\! (\mathsf{W}^{\prime})=|S|\leq \mathop{\textrm{rk}}\nolimits\! (\mathsf{W})$ .

4. Transposition

We now return to the general case, where $\Lambda$ is an arbitrary finite-dimensional algebra. A sequence $(T_1, T_2, \dots, T_r)$ of indecomposable objects in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ is called an ordered $\tau$ -rigid object if $\amalg _{i=1}^r T_i$ is a $\tau$ -rigid object, and an ordered support $\tau$ -tilting object if $r=n\,:\!= \mathop{\textrm{rk}}\nolimits \Lambda$ . The symmetric group acts on the set of ordered support $\tau$ -tilting objects in a wide subcategory $\mathsf{W}$ of ${\mathop{\textrm{mod}}\nolimits }\, \Lambda$ , by reordering. We recall the following theorem from [Reference Buan and Marsh6].

Theorem 4.1 ([Reference Buan and Marsh6, Thm. 5.4]). For each $\tau$ -perpendicular subcategory $\mathsf{W}$ of ${\mathop{\textrm{mod}}\nolimits }\,\Lambda$ , there are mutually inverse bijections

\begin{equation*}\{\textit {ordered support $\tau $-tilting objects in $\mathsf {W}$ } \}\end{equation*}
\begin{equation*}\psi ^{\mathsf {W}} \downarrow \text { } \uparrow \phi ^{\mathsf {W}}\end{equation*}
\begin{equation*}\{\textit {complete signed $\tau $-exceptional sequences in $\mathsf {W}$} \}\end{equation*}

In the case $\mathsf{W}={\mathop{\textrm{mod}}\nolimits }\,\Lambda$ , we write $\psi$ for $\psi ^{{\mathop{\textrm{mod}}\nolimits }\, \Lambda }$ and $\phi$ for $\phi ^{{\mathop{\textrm{mod}}\nolimits }\, \Lambda }$ . In this section, we will describe the action of the symmetric group on complete signed $\tau$ -exceptional sequences in a $\tau$ -perpendicular subcategory $\mathsf{W}$ of ${\mathop{\textrm{mod}}\nolimits }\,\Lambda$ induced by the bijections above.

Remark 4.2. If $\psi (T_1, T_2, \dots, T_n) = (A_1, \dots, A_n)$ , define (as in the proof of Lemma 3.5 ):

\begin{align*} \mathsf{W}_{n-1} & = J(A_n),\\ \mathsf{W}_{n-2} &= J_{\mathsf{W}_{n-1}}(A_{n-1}), \\ & \, \vdots \\ \mathsf{W}_j & = J_{\mathsf{W}_{j+1}}(A_{j+1}), \\ & \, \vdots \\ \mathsf{W}_1 &= J_{\mathsf{W}_{2}}(A_2). \end{align*}

Then, we have

\begin{align*} A_n &= T_n, \\ A_{n-1} &= \mathcal{E}_{A_n}(T_{n-1}), \\ A_{n-2} &= \mathcal{E}^{\mathsf{W}_{n-1}}_{A_{n-1}} \mathcal{E}_{A_n} (T_{n-2}), \\ & \, \vdots \\ A_{n-j} &= \mathcal{E}_{A_{n-j+1}}^{\mathsf{W}_{n-j+1}} \dots \mathcal{E}^{\mathsf{W}_{n-1}}_{A_{n-1}} \mathcal{E}_{A_n}(T_{n-j}), \\ &\, \vdots \\ A_1 &= \mathcal{E}_{A_2}^{\mathsf{W}_2} \dots \mathcal{E}^{\mathsf{W}_{n-1}}_{A_{n-1}} \mathcal{E}_{A_n}(T_1). \end{align*}

Lemma 4.3. With notation as above, we have

\begin{equation*}\mathsf {W}_{n-i} = J(T_n \amalg T_{n-1} \amalg \dots \amalg T_{n-{i+1}}),\end{equation*}

for $i= 1, \dots, n-1$ .

Proof. This follows from repeated use of Theorem 2.6 (a).

Let $\Lambda$ be an algebra of rank $n$ , and let $\mathcal{T}_o^{\Lambda }$ be the set of ordered basic support $\tau$ -tilting objects in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ . There is a natural action of the symmetric group $S_n$ on $T_o^{\Lambda }$ , given by $\pi _i (T_1, \dots, T_n) = (T_1, \dots, T_{i-1}, T_{i+1}, T_i, T_{i+2}, \dots, T_n)$ , where $\pi _i$ denotes the simple transposition $(i\ i+1)$ .

Theorem 4.4.

  1. (a) If $\mathcal{S} = (A_1, \dots, A_n)$ is a signed $\tau$ -exceptional sequence, then for $i \in \{1, \dots, n-1\}$ we have that

    \begin{equation*}\widetilde {\pi _i}(\mathcal {S})\,:\!= \left(A_1, \dots, A_{i-1}, \mathcal {E}^{(i)} (A_{i+1}), \mathcal {F}^{(i)}(A_i), A_{i+2}, \dots, A_n\right)\end{equation*}
    is a signed $\tau$ -exceptional sequence, where
    \begin{equation*}\mathcal {F}^{(i)}\,:\!= \mathcal {F}_{A_{i+1}}^{\mathsf {W}_{i+1}}\end{equation*}
    and
    \begin{equation*}\mathcal {E}^{(i)}\,:\!= \mathcal {E}^{\mathsf {W}_{i+1}}_{\mathcal {F}^{(i)}(A_i)}\end{equation*}
    for $i \in \{1, \dots, n-2\}$ and where $\mathcal{F}^{(n-1)}\,:\!= \mathcal{F}_{A_n}$ and $\mathcal{E}^{(n-1)}\,:\!= \mathcal{E}_{\mathcal{F}^{(n-1)}(A_{n-1})}$ .
  2. (b) For each $i \in \{1, \dots, n-1\}$ we have $ \widetilde{\pi _i} \psi = \psi \pi _i$ .

Lemma 4.5. Let $(B,C)$ be an ordered $\tau$ -rigid object in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ . Then, we have that $\mathcal{E}^{J(C)}_{\mathcal{E}_C(B)}\mathcal{E}_C = \mathcal{E}^{J(B)}_{\mathcal{E}_B(C)}\mathcal{E}_B$ .

Proof. By applying Theorem 2.6 (b) twice, we obtain

\begin{equation*} \mathcal {E}^{J(C)}_{\mathcal {E}_C(B)}\mathcal {E}_C = \mathcal {E}_{C \amalg B} = \mathcal {E}_{B \amalg C} = \mathcal {E}^{J(B)}_{\mathcal {E}_B(C)}\mathcal {E}_B. \end{equation*}

Proposition 4.6. Let $(T_1, \dots, T_n)$ be an ordered support $\tau$ -tilting object in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ and assume that $\psi (T_1, \dots, T_n) = (A_1, \dots, A_n)$ . Then

\begin{equation*}\psi (\pi _{n-1}(T_1, \dots, T_n)) = \left(A_1, \dots A_{n-2}, \mathcal {E}_{\mathcal {F}_{A_{n}} (A_{n-1})}(A_n), \mathcal {F}_{A_n}(A_{n-1})\right).\end{equation*}

Proof. Let $\ (B_1, B_2, \dots, B_n) = \psi (\pi _{n-1}(T_1, \dots, T_n)) = \psi (T_1, \dots, T_{n-2}, T_n, T_{n-1})$ . We need to show that

  1. (i) $B_n = \mathcal{F}_{A_n}(A_{n-1})$ ,

  2. (ii) $B_{n-1} = \mathcal{E}_{\mathcal{F}_{A_{n}}(A_{n-1})}(A_n)$ , and

  3. (iii) $B_j = A_j$ , for $1 \leq j \leq n-2$ .

Note that, by Remark 4.2, we have that $B_n=T_{n-1}$ , $A_n=T_n$ and that $A_{n-1} = \mathcal{E}_{T_n}(T_{n-1})$ . Hence, we have $\mathcal{F}_{T_n}(A_{n-1}) = \mathcal{F}_{T_n}\mathcal{E}_{T_n}(T_{n-1}) = T_{n-1} = B_n$ , which proves claim (i). Moreover, it also follows from Remark 4.2 that $B_{n-1} = \mathcal{E}_{T_{n-1}}(T_n) = \mathcal{E}_{\mathcal{F}_{T_n}(A_{n-1})}(A_n)$ , which proves claim (ii).

It remains to prove that $B_{j} = A_j$ for $j \leq n-2$ . Apply Lemma 4.5, with $B= T_{n-1}$ and $C = T_n$ to obtain that

\begin{equation*}\mathcal {E}^{J(T_n)}_{\mathcal {E}_{T_n}(T_{n-1})}\mathcal {E}_{T_n} = \mathcal {E}^{J(T_{n-1})}_{\mathcal {E}_{T_{n-1}}(T_n)}\mathcal {E}_{T_{n-1}}.\end{equation*}

It now follows directly, from Remark 4.2 and Lemma 4.3, that $B_{j} = A_j$ for $j \leq n-2$ .

Proof of Theorem 4.4. By Proposition 4.6, it follows that both (a) and (b) hold for $i = n-1$ . Assume $i \lt n-1$ . Then, $(A_1, \dots, A_{i-1})$ is a complete signed $\tau$ -exceptional sequence for the $\tau$ -perpendicular subcategory $\mathsf{W}_{i-1}$ , as defined in Remark 4.2. Finally, Proposition 4.6 implies that both (a) and (b) hold also in this case.

5. Mutation

For a fixed positive integer $n$ , consider the group $G_n = \langle \mu _1, \dots, \mu _n \mid \mu _i^2 = e \rangle$ (as in [Reference King and Pressland20, §1]). Let $\Lambda$ be a fixed algebra of rank $n$ .

Mutation of support $\tau$ -tilting objects (as in [Reference Adachi, Iyama and Reiten2, Thm. 2.18]) induces a mutation on the set $\mathcal{T}^{\Lambda }_o$ of ordered basic support $\tau$ -tilting objects in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ , which can be regarded as an action of $G_n$ on $\mathcal{T}^{\Lambda }_o$ and hence, via the bijections in Theorem 4.1, an action on the set of complete signed $\tau$ -exceptional sequences.

The following result follows from [Reference Adachi, Iyama and Reiten2, Thm. 2.18].

Proposition 5.1 ([Reference Adachi, Iyama and Reiten2, Thm. 2.18]). Let $T = (T_1, \dots, T_n)$ be an ordered support $\tau$ -tilting object in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ . Let $i \in \{1, \dots, n\}$ . Then, there is a unique indecomposable object $T_i^{\ast }$ in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\, \Lambda )$ such that $T(i)=(T_1, \dots, T_{i-1}, T_i^{\ast }, T_{i+1}, \dots, T_n)$ is an ordered support $\tau$ -tilting object with $T_i^{\ast } \not \simeq T_i$ .

With $T$ and $T(i)$ as above, we set $\mu _i(T) = T(i)$ . This defines a $G$ -action on $\mathcal{T}^{\Lambda }_o$ . We now describe the corresponding action on the set of complete signed $\tau$ -exceptional sequences.

For a complete signed $\tau$ -exceptional sequence $\mathcal{S} = (A_1, \dots, A_n)$ , let

\begin{equation*}s_1(\mathcal {S}) = (\widetilde {A_1}, A_2, \dots, A_n),\end{equation*}

where

\begin{equation*}\widetilde {A_1} = \begin {cases} A_1[1] & \text { if } A_1 \in {\mathop {\textrm {mod}}\nolimits } \Lambda ; \\ A_1[\!-\!1] & \text { if } A_1 \in {\mathop {\textrm {mod}}\nolimits } \Lambda [1]. \end {cases} \end{equation*}

Moreover, for $j\gt 1$ , let $s_j(\mathcal{S}) = \widetilde{\pi }_{j-1} \widetilde{\pi }_{j-2} \dots \widetilde{\pi _1} s_1 \widetilde{\pi }_1 \dots \widetilde{\pi }_{j-2} \widetilde{\pi }_{j-1}(\mathcal{S})$ .

We make the following observation:

Lemma 5.2. Let $\Lambda$ be an algebra with a unique simple module. Then, the projective cover of the simple module is the unique indecomposable $\tau$ -rigid $\Lambda$ -module.

Proof. Let $S$ be the unique simple $\Lambda$ -module, and suppose that $X$ is a non-projective indecomposable $\tau$ -rigid $\Lambda$ -module. Then, $\mathop{\textrm{Ext}}\nolimits ^1\!(X,M)\not =0$ for some $\Lambda$ -module $M$ . Since $M$ must be constructed from $S$ by repeated extensions with $S$ , it follows that $\mathop{\textrm{Ext}}\nolimits ^1\!(X,S)\not =0$ . Since $X$ is also constructed from $S$ by repeated extensions with $S$ , we have that $S$ is a factor of $X$ , so $\mathop{\textrm{Ext}}\nolimits ^1\!(X,\mathop{\textrm{Gen}}\nolimits X)\not =0$ . By Lemma 2.3, $X$ is not $\tau$ -rigid.

We now have:

Lemma 5.3. If $\mathcal{S} = (A_1, A_2, \dots, A_n)$ is a signed $\tau$ -exceptional sequence, then so is $s_j(\mathcal{S})$ for each $j= 1, \dots, n$ . Moreover, $\mathcal{S}$ and $s_1(\mathcal{S})$ are the only signed $\tau$ -exceptional sequences of the form $(X, A_2, \dots, A_n)$ for some object $X$ in $\mathcal{C}({\mathop{\textrm{mod}}\nolimits }\,\Lambda )$ .

Proof. Define $W^A_j$ , for $j=1,\ldots,n$ , as in the proof of Theorem 3.1 (a). By repeated application of Theorem 2.2, $W^A_j$ is equivalent to a module category over a finite-dimensional algebra of rank $j-1$ for $j=1,\ldots,n$ . Hence, $\mathsf{W}^A_2$ is equivalent to the module category of a finite-dimensional algebra with a unique simple module. By Lemma 5.2, $\mathsf{W}^A_2$ , regarded as a module category, has a unique indecomposable $\tau$ -rigid module, given by the unique indecomposable projective module. This proves that $s_1(\mathcal{S})$ is a signed $\tau$ -exceptional sequence, and also that $\mathcal{S}$ and $s_1(\mathcal{S})$ are the only signed $\tau$ -exceptional sequences of the form $(X, A_2, \dots, A_n)$ , that is the claim for $j=1$ .

The claim for $j\gt 1$ follows by combining this with Theorem 4.4 (a).

Proposition 5.4. With notation as above $s_i \psi = \psi \mu _i$ holds for all $i =1, \dots, n$ .

Proof. Consider first the case $i=1$ , and assume that $\psi (M_1, M_2, \dots, M_n) = (A_1, A_2, \dots,A_n)$ . We have $\mu _1(M_1, M_2, \dots, M_n) = (M_1^\ast, M_2, \dots, M_n)$ , where $M^\ast _1 \not \simeq M_1$ , and so $\psi \mu _1(M_1, M_2, \dots, M_n) = \psi (M_1^\ast, M_2, \dots, M_n) =(X, A_2, \dots,A_n)$ , for some object $X$ . The claim for $i=1$ now follows from Lemma 5.3.

For $j\gt 1$ , we first note that $\mu _j = \pi _{j-1} \pi _{j-2} \dots \pi _1 \mu _1 \pi _1\dots \pi _{j-2}\pi _{j-1}$ .

Next we note that, by repeated applications of Theorem 4.4, we have that

\begin{equation*}\widetilde {\pi _1} \dots \widetilde {\pi _j} \psi = \psi \pi _1 \dots \pi _j.\end{equation*}

Combining this with the above, we obtain

\begin{align*} s_j \psi &= (\widetilde{\pi }_{j-1} \widetilde{\pi }_{j-2} \dots \widetilde{\pi }_1 s_1 \widetilde{\pi }_1 \dots \widetilde{\pi }_{j-2} \widetilde{\pi }_{j-1}) \psi \\ &= (\widetilde{\pi }_{j-1} \widetilde{\pi }_{j-2} \dots \widetilde{\pi }_1 s_1) (\widetilde{\pi }_1 \dots \widetilde{\pi }_{j-2} \widetilde{\pi }_{j-1} \psi ) \\ &= (\widetilde{\pi }_{j-1} \widetilde{\pi }_{j-2} \dots \widetilde{\pi }_1 s_1) (\psi \pi _1 \dots \pi _{j-2} \pi _{j-1}) \\ &= (\widetilde{\pi }_{j-1} \widetilde{\pi }_{j-2} \dots \widetilde{\pi }_1 )(s_1 \psi ) (\pi _1 \dots \pi _{j-2} \pi _{j-1}) \\ &= (\widetilde{\pi }_{j-1} \widetilde{\pi }_{j-2} \dots \widetilde{\pi }_1 )(\psi \mu _1) (\pi _1 \dots \pi _{j-2} \pi _{j-1}) \\ &= (\widetilde{\pi }_{j-1} \widetilde{\pi }_{j-2} \dots \widetilde{\pi }_1 \psi ) (\mu _1\pi _1 \dots \pi _{j-2} \pi _{j-1}) \\ & = (\psi \pi _{j-1} \pi _{j-2} \dots \pi _1) (\mu _1 \pi _1\dots \pi _{j-2}\pi _{j-1}) \\ & = \psi (\pi _{j-1} \pi _{j-2} \dots \pi _1\mu _1 \pi _1\dots \pi _{j-2}\pi _{j-1}) \\ &= \psi \mu _j. \end{align*}

King and Pressland [Reference King and Pressland20, Defn. 1.2] consider the following group:

Definition 5.5 ([Reference King and Pressland20]). Let $M_n = S_n \ltimes G_n$ be the mutation group of degree $n$ , where $S_n$ acts on $G_n$ via $\sigma (\mu _i)=\mu _{\sigma (i)}$ .

They show that this group acts naturally on labelled (i.e. ordered) seeds in a cluster algebra [Reference Fomin and Zelevinsky13] via permutation and mutation. The mutation of support- $\tau$ -tilting objects in [Reference Adachi, Iyama and Reiten2, §2.3] can be regarded as a generalisation of cluster mutation, so it is natural to consider the action of the mutation group in this context. Note that, for an ordered support $\tau$ -tilting object $T$ , we have $\sigma (\mu _i T) = \mu _{\sigma (i)} (\sigma T)$ for a permutation $\sigma$ and mutation $\mu _i$ , so $M_n$ acts on the set of all ordered support $\tau$ -tilting objects, $\mathcal{T}_o^{\Lambda }$ .

We get an induced action of the mutation group on the set of signed $\tau$ -exceptional sequences.

Theorem 5.6. Let $\mathcal{S} = (A_1, A_2, \dots, A_n)$ be a signed $\tau$ -exceptional sequence. The operations

\begin{equation*}\widetilde {\pi _i}(\mathcal {S})\,:\!= (A_1, \dots, A_{i-1}, \mathcal {E}^{(i)} (A_{i+1}), \mathcal {F}^{(i)}(A_i), A_{i+2}, \dots, A_n),\end{equation*}
\begin{equation*}s_1(\mathcal {S}) = (\widetilde {A_1}, A_2, \dots, A_n),\end{equation*}

and, for $j=2\ldots,n$ ,

\begin{equation*}s_j(\mathcal {S}) = \widetilde {\pi _{j-1}} \widetilde {\pi _{j-2}} \dots \widetilde {\pi _1} s_1 \widetilde {\pi _1} \dots \widetilde {\pi _{j-2}} \widetilde {\pi _{j-1}} (\mathcal {S}),\end{equation*}

define an action of the mutation group $M_n$ on the set of signed $\tau$ -exceptional sequences.

Proof. As already noted, $M_n$ acts on the set of ordered support $\tau$ -tilting objects, and the result hence follows directly from combining Theorem 4.4 and Proposition 5.4 with the fact that $\psi$ is a bijection between the set of ordered support $\tau$ -tilting objects and the set of signed $\tau$ -exceptional sequences (Theorem 4.1).

6. Examples relating to braid actions

Note that the braid group, $B_n$ , on $n$ strands, has the symmetric group $S_n$ as a quotient. Since $S_n$ is a subgroup of the mutation group $M_n=S_n \ltimes G_n$ , it follows that $B_n$ acts naturally on the set of all ordered support $\tau$ -tilting objects $\mathcal{T}_o^{\Lambda }$ and thus on the set of all complete signed $\tau$ -exceptional sequences for $\Lambda$ , by Theorem 4.1. However, this action is highly non-transitive in general, since the braid group is only permuting the possible orderings of each support $\tau$ -tilting object.

It is therefore natural to ask whether there is a transitive action. In the first part of this section, we give an example to show that, at least via the mutation group, this is not possible: we give an algebra for which there is no transitive action of $B_2$ which factors through the action of $M_2$ on $\mathcal{T}_o^{\Lambda }$ .

Let $Q$ be the Kronecker quiver , and let $\Lambda$ be the corresponding path algebra. Let $P_i$ (respectively, $I_i$ ), for $i=1,2$ be the indecomposable projective (respectively, injective) $\Lambda$ -modules corresponding to the vertices of $Q$ . Then, the $\tau$ -tilting (equivalently, tilting) $\Lambda$ -modules are the modules $\tau ^{-r}P_1\amalg \tau ^{-r}P_2$ , $\tau ^{-r}P_1\amalg \tau ^{-(r+1)}P_2$ , $\tau ^r I_1 \amalg \tau ^r I_2$ and $\tau ^{r+1} I_1 \amalg \tau ^rI_2$ , for $r=0,1,2,\ldots$ . The support $\tau$ -tilting (equivalently, support tilting) objects over $\Lambda$ which are not $\tau$ -tilting are $I_1\amalg P_2[1]$ , $P_1[1]\amalg P_2$ and $P_1[1]\amalg P_2[1]$ . In particular, note that $\Lambda$ is not $\tau$ -tilting finite.

The mutation group (see Definition 5.5) is $M_2=S_2\ltimes G_2$ , where $S_2=\{1,\sigma \}$ is the symmetric group of degree $2$ and $G_2=\langle \mu _1,\mu _2 \ :\ \mu _1^2=\mu _2^2=e\rangle$ . We have $\sigma \mu _1=\mu _2\sigma$ and $\sigma \mu _2=\mu _1\sigma$ . The action of $M_2$ on the set $\mathcal{T}_o^{\Lambda }$ of ordered basic support $\tau$ -tilting objects is shown in Figure 1, from which it can be seen that this action is transitive. However, we have the following.

Figure 1. The action of $M_2$ on $T_o^{\Lambda }$ for the Kronecker algebra $\Lambda$ .

Proposition 6.1. There is no transitive action of the braid group on two strands on $\mathcal{T}_o^{\Lambda }$ which factors through the action of $M_2$ on $\mathcal{T}_o^{\Lambda }$ .

Proof. Note that the braid group $B_2$ on two strands is isomorphic to the infinite cyclic group. If there was a transitive action of $B_2$ on $\mathcal{T}_o^{\Lambda }$ factoring through the action of $M_2$ on $\mathcal{T}_o^{\Lambda }$ , then there would be a subgroup of $M_2$ which is a quotient of $B_2$ which acted transitively on $\mathcal{T}_o^{\Lambda }$ . Such a subgroup would have to be cyclic.

The elements of $M_2$ are of the form

\begin{equation*}\sigma ^\varepsilon \mu _i\mu _{\sigma (i)}\mu _i\cdots \mu _i\end{equation*}

and

\begin{equation*}\sigma ^\varepsilon \mu _i\mu _{\sigma (i)}\mu _i\cdots \mu _{\sigma (i)},\end{equation*}

where $i\in \{1,2\}$ and $\varepsilon \in \{0,1\}$ . It is easy to check from the description of the action of $M_2$ (see Figure 1) that, for each of these elements, the (infinite) cyclic group it generates does not act transitively on $\mathcal{T}_o^{\Lambda }$ .

Corollary 6.2. There is no transitive action of the braid group on two strands on the set of all signed $\tau$ -exceptional sequences which factors through the action of $M_2$ on the set of all such sequences.

The transitive braid group action on the set of complete exceptional sequences for a hereditary algebra arises in the following way [Reference Crawley-Boevey8, Lemma 9, Theorem], [Reference Ringel23, §5, §7]. Write the braid group $B_n$ on $n$ strands in the usual way as

\begin{equation*}B_n=\left \langle \sigma _1,\ldots,\sigma _{n-1}\,:\,\begin {array}{cc} \sigma _i\sigma _j=\sigma _j\sigma _i, & |i-j|\gt 1; \\ \sigma _i\sigma _j\sigma _i=\sigma _j\sigma _j\sigma _i, & |i-j|=1 \end {array} \right \rangle. \end{equation*}

Given a complete exceptional sequence $(X_1,\ldots,X_n)$ and $1\leq i\leq n$ , there is a unique complete exceptional sequence of the form $(X_1,\ldots,X_{i-1},X_{i+1},Y,X_{i+2},\ldots,X_n)$ for some exceptional indecomposable module $Y$ . The left version of the action of the braid group is given by:

\begin{equation*}\sigma _i(X_1,\ldots,X_n)= (X_1,\ldots,X_{i-1},X_{i+1},Y,X_{i+2},\ldots,X_n).\end{equation*}

If $n=2$ , this means, in particular, that if $(X_1,X_2)$ is a complete exceptional sequence then there is an exceptional sequence of the form $(X_2,Y)$ .

However, if $(X_1,X_2)$ is a complete $\tau$ -exceptional sequence, it can be the case that there is no $\tau$ -exceptional sequence of the form $(X_2,Y)$ , that is it can happen that $X_2\not \in J(Y)$ for all indecomposable $\tau$ -rigid modules $Y$ . We illustrate this point with the following example from [Reference Buan and Marsh6, §6.2].

Let $\Gamma$ be the algebra given by the path algebra of the quiver , subject to the relation $\beta \alpha =0$ . The $\tau$ -exceptional sequences over $\Gamma$ were given in [Reference Buan and Marsh6, §6.2]. They are as follows:

\begin{equation*}\left (1,\begin {matrix} 2 \\ 1 \\ 2\end {matrix}\right ), \left (2,\begin {matrix} 1 \\ 2 \end {matrix}\right ), \left (\begin {matrix} 2 \\ 1 \end {matrix},2\right ), \left (\begin {matrix} 1 \\ 2 \end {matrix},1\right ).\end{equation*}

We make the following observation.

Lemma 6.3. For the algebra $\Gamma$ , which has complete $\tau$ -exceptional sequences of length $2$ , there is a $\tau$ -rigid indecomposable module $X$ such that there is no complete $\tau$ -exceptional sequence of the form $(X,Y)$ for some $\tau$ -rigid module $Y$ .

Proof. This can be seen from the list of $\tau$ -exceptional sequences in ${\mathop{\textrm{mod}}\nolimits }\,\Gamma$ above, since the module $P_2=\begin{smallmatrix} 2 \\ 1 \\ 2 \end{smallmatrix}$ , despite being $\tau$ -rigid, does not occur as the first term in any of the complete $\tau$ -exceptional sequences in the list.

Note that it follows that $P_2$ also does not occur as the first term in a complete signed $\tau$ -exceptional sequence.

The phenomenon described above also occurs with the right version [Reference Crawley-Boevey8, Reference Ringel23] of the action of the braid group. Given a complete exceptional sequence $(X_1,\ldots,X_n)$ and $1\leq i\leq n$ , there is a unique complete exceptional sequence of the form $(X_1,\ldots,X_{i-1},Y,X_{i+1},X_{i+2},\ldots,X_n)$ for some exceptional indecomposable module $Y$ . The right version of the action of the braid group is given by:

\begin{equation*}\sigma _i(X_1,\ldots,X_n)= (X_1,\ldots,X_{i-1},Y,X_{i+1},X_{i+2},\ldots,X_n).\end{equation*}

If $n=2$ , this means, in particular, that if $(X_1,X_2)$ is a complete exceptional sequence then there is an exceptional sequence of the form $(Y,X_1)$ .

We can see from the list above that, although $(\begin{smallmatrix} 2 \\ 1 \end{smallmatrix},2)$ is a complete $\tau$ -exceptional sequence for $\Gamma$ , there is no such sequence of the form $(Y,\begin{smallmatrix} 2 \\ 1\end{smallmatrix})$ . This is because, although $\begin{smallmatrix} 2 \\ 1 \end{smallmatrix}$ is $\tau$ -rigid in $J(2)$ , it is not a $\tau$ -rigid $\Gamma$ -module. In the hereditary case, where a module $M$ is $\tau$ -rigid if and only if $\mathop{\textrm{Ext}}\nolimits ^1\!(M,M) = 0$ , a module in $J(Y) = Y^{\perp _{0,1}}$ is $\tau$ -rigid in $J(Y)$ if and only if it is $\tau$ -rigid in ${\mathop{\textrm{mod}}\nolimits }\,\Lambda$ , but, as we see here, this is not true in general.

Acknowledgement

The first named author would like to thank Eric Hanson for insightful conversations related to this work.

Footnotes

Dedicated to the memory of Helmut Lenzing

This work was supported by FRINAT grant number 301375 from the Norwegian Research Council. This work was partially supported by a grant from the Simons Foundation. The authors would like to thank the Isaac Newton Institute for Mathematical Sciences, Cambridge, for support and hospitality during the programme Cluster Algebras and Representation Theory where work on this paper was undertaken. Both authors would like to thank the Centre of Advanced Study, Oslo for support and hospitality during the programme Representation Theory: Combinatorial Aspects and Applications.

References

Auslander, M. and Smalø, S. O., Almost split sequences in subcategories, J. Algebra 69 (1981), 426454. Addendum: J. Algebra 71 (1981), 592–594.Google Scholar
Adachi, T., Iyama, O. and Reiten, I., τ-tilting theory, Compos. Math. 150(3) (2015), 415452.Google Scholar
Asai, S., Semibricks, Int Math. Res. Not. 16 (2020), 49935054.Google Scholar
Brüstle, T., Smith, D. and Treffinger, H., Wall and chamber structure for finite-dimensional algebras, Adv. Math. 354 (2019), 31pp.Google Scholar
Buan, A. B. and Hanson, E., τ-perpendicular wide subcategories. Preprint arXiv: 2107.01141 [math.RT] (2021).Google Scholar
Buan, A. B. and Marsh, B. R., τ-exceptional sequences, J. Algebra 585 (2021), 3668.Google Scholar
Buan, A. B. and Marsh, B. R., A category of wide subcategories, Int. Math. Res. Not. 13 (2021), 1027810338.Google Scholar
Crawley-Boevey, W., Exceptional sequences of representations of quivers, in Proceedings of the Sixth International Conference on Representations of Algebras (Ottawa, ON, 1992), Carleton-Ottawa Math. Lecture Note Ser., vol. 14 (Carleton Univ., Ottawa, ON, 1992), 7pp.Google Scholar
Demonet, L., Iyama, O. and Jasso, G., τ-tilting finite algebras, bricks, and g-vectors, Int. Math. Res. Not. 3 (2019), 852892.Google Scholar
Demonet, L., Iyama, O., Reading, N., Reiten, I. and Thomas, H., Lattice theory of torsion classes: Beyond $\tau$ -tilting theory, Trans. Am. Math. Soc. Ser. B 10 (2023), 542612.Google Scholar
Enomoto, H., Rigid modules and ICE-closed subcategories in quiver representations, J. Algebra 594 (2022), 364388.CrossRefGoogle Scholar
Enomoto, H. and Sakai, A., ICE-closed subcategories an wide $\tau$ -tilting modules, Math. Z. 300(1) (2022), 541577.Google Scholar
Fomin, S. and Zelevinsky, A., Cluster algebras. I. Foundations, J. Am. Math. Soc. 15(2) (2002), 497529.Google Scholar
Geigle, W. and Lenzing, H., Perpendicular categories with applications to representations and sheaves, J. Algebra 144(2) (1991), 273343.Google Scholar
Hanson, E. J. and Igusa, K., $\tau$ -cluster morphism categories and picture groups, Commun. Algebra 49(10) (2021), 43764415.Google Scholar
Hanson, E. J. and Thomas, H., A uniqueness property of $\tau$ -exceptional sequences. Preprint arXiv: 2302.07118 [math.RT] (2023).Google Scholar
Igusa, K. and Todorov, G., Signed exceptional sequences and the cluster morphism category. Preprint arXiv: 1706.02041 [math.RT] (2017).Google Scholar
Igusa, K., Todorov, G. and Weyman, J., Picture groups of finite type and cohomology in type $A_n$ . Preprint arXiv: 1609.02636 [math.RT] (2016).Google Scholar
Jasso, G., Reduction of $\tau$ -tilting modules and torsion pairs, Int Math. Res. Not. 16 (2015), 71907237.Google Scholar
King, A. and Pressland, M., Labelled seeds and the mutation group, Math. Proc. Camb. Philos. Soc. 163(2) (2017), 193217.Google Scholar
Marks, F. and Šťovíček, J., Torsion classes, wide subcategories and localisations, Bull. Lond. Math. Soc. 49(3) (2017), 405416.Google Scholar
Mendoza, H. O. and Treffinger, H., Stratifying systems through $\tau$ -tilting theory, Doc. Math. 25 (2020), 701720.Google Scholar
Ringel, C. M., The braid group action on the set of exceptional sequences of a hereditary Artin algebra, in Abelian Group Theory and Related Topics (Oberwolfach, 1993), Contemp. Math., vol. 171 (Amer. Math. Soc., Providence, RI, 1994), 339352.Google Scholar
Schofield, A., Semi-invariants of quivers, J. Lond. Math. Soc. (2) 43(3) (1991), 385395.Google Scholar
Figure 0

Figure 1. The action of $M_2$ on $T_o^{\Lambda }$ for the Kronecker algebra $\Lambda$.