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

\setstackgap

L.75

Triangulations of prisms and
preprojective algebras of type AA

Osamu Iyama [email protected] https://www.ms.u-tokyo.ac.jp/ iyama/index.html Graduate School of Mathematical Sciences, The University of Tokyo, 3-8-1 Komaba, Meguro-ku, Tokyo 153-8914, Japan  and  Nicholas J. Williams [email protected] https://nchlswllms.github.io/ Department of Mathematics and Statistics, Fylde College, Lancaster University, Lancaster, LA1 4YF, United Kingdom
Abstract.

We show that indecomposable two-term presilting complexes over Πn\Pi_{n}, the preprojective algebra of AnA_{n}, are in bijection with internal nn-simplices in the prism Δn×Δ1\Delta_{n}\times\Delta_{1}, the product of an nn-simplex with a 1-simplex. We show further that this induces a bijection between triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1} and two-term silting complexes over Πn\Pi_{n} such that bistellar flips of triangulations correspond to mutations of two-term silting complexes. These bijections are shown to compatible with the known bijections involving the symmetric group.

Key words and phrases:
Product of simplices, prism, triangulations, preprojective algebras, τ\tau-tilting, silting
2020 Mathematics Subject Classification:
05E10, 16G20, 52B12
OI was supported by JSPS Grant-in-Aid for Scientific Research (B) 16H03923, (C) 18K3209 and (S) 15H05738. NJW was supported by a JSPS International Short-Term Postdoctoral Research Fellowship at the University of Tokyo.

1. Introduction

Cluster algebras are intimately connected with the combinatorics of triangulations. The simplest example of this is the bijection between the clusters in the type AA cluster algebra and triangulations of a convex polygon [FZ02]. Further connections were found in the work of Fomin, Shapiro, and Thurston, who defined cluster algebras using tagged triangulations of surfaces [FST08].

Cluster algebras were first connected with the representation theory of finite-dimensional algebras in [MRZ03], which led to the definition of cluster categories of hereditary algebras in [BMR+06]. Another approach to categorifying cluster algebras uses the representation theory of preprojective algebras [GLS06, GLS07a, GLS07b, GLS08]. Preprojective algebras have connections with Kleinian singularities [CBH98], Nakajima quiver varieties [Nak94, ST11], and crystal bases [KS97].

Cluster categories were subsequently extended to cluster algebras arising from triangulations of surfaces in [Ami09]. Since then, the relation between triangulated surfaces and representation theory in the form of so-called geometric models has been an active subject of research [BZ11, OPS18, BCSo21, CS20].

The triangulations considered thus far are all two-dimensional and it is natural to wonder whether similar phenomena exist in higher dimensions. Indeed, a beautiful connection between representation theory and higher-dimensional triangulations was found in [OT12], where triangulations of even-dimensional cyclic polytopes were shown to be in bijection with cluster-tilting objects in higher cluster categories. In [Wil22] it was shown how odd-dimensional triangulations enter the picture too, through a bijection with equivalence classes of maximal green sequences.

High-dimensional polytopes are in general very complicated, but, besides cyclic polytopes, another example that is well-studied is the Cartesian product Δm×Δn\Delta_{m}\times\Delta_{n} of two simplices. The triangulations of this polytope are interesting for many reasons. The regular triangulations classify the types of tropical polytopes [DS04]. Furthermore, the secondary polytope of Δm×Δn\Delta_{m}\times\Delta_{n} is the Newton polytope of the product of all minors of an (m+1)×(n+1)(m+1)\times(n+1) matrix [BB98], as well as the state polytope of the Segre embedding of m×n\mathbb{P}^{m}\times\mathbb{P}^{n} [Stu91]. Products of simplices are also related to transportation polytopes from operations research [DLKOS09], and arise as strategy spaces for two-player games [vdLT82].

In this paper, we show that combinatorics of the prism Δn×Δ1\Delta_{n}\times\Delta_{1} is closely related to the representation theory of Πn\Pi_{n}, the preprojective algebra of AnA_{n}.

Theorem 1.1 (Proposition 3.2, Corollary 3.17, Proposition 3.19).

There is a bijection between codimension one internal simplices in Δn×Δ1\Delta_{n}\times\Delta_{1} and indecomposable two-term presilting complexes over Πn\Pi_{n} which induces a bijection between triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1} and two-term silting complexes over Πn\Pi_{n}. Mutations of silting complexes correspond to bistellar flips of triangulations.

It is already known from [GKZ94, Chapter 7, Section 3C] that triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1} are in bijection with permutations in the symmetric group 𝔖n+1\mathfrak{S}_{n+1}, and from [Miz14] that permutations in 𝔖n+1\mathfrak{S}_{n+1} were in bijection with two-term silting complexes over Πn\Pi_{n}. We show that our results are compatible with these bijections. Indeed, to summarise the results of the paper and how they fit in with the literature, there are bijections

𝗂𝗇𝗍-𝗌𝗂𝗆n(Δn×Δ1)𝗂𝗇𝖽-𝟤-𝗉𝗌𝗂𝗅𝗍Πn𝗂𝗇𝖽-τ-𝗋𝗂𝗀𝗂𝖽-𝗉𝖺𝗂𝗋Πn\mathsf{int\mbox{-}sim}_{n}(\Delta_{n}\times\Delta_{1})\longleftrightarrow\mathsf{ind\mbox{-}2\mbox{-}psilt}\,\Pi_{n}\longleftrightarrow\mathsf{ind\mbox{-}\tau\mbox{-}rigid\mbox{-}pair}\,\Pi_{n}

which induce the commutative diagram

𝔖n+1\mathfrak{S}_{n+1}𝟤-𝗌𝗂𝗅𝗍Πn\mathsf{2\mbox{-}silt}\,\Pi_{n}𝗌τ-𝗍𝗂𝗅𝗍Πn\mathsf{s\tau\mbox{-}tilt}\,\Pi_{n}𝗍𝗋𝗂(Δn×Δ1)\mathsf{tri}(\Delta_{n}\times\Delta_{1})Pn(𝗂𝗇𝖽-τ-𝗋𝗂𝗀𝗂𝖽-𝗉𝖺𝗂𝗋Πn)P_{n}(\mathsf{ind\mbox{-}\tau\mbox{-}rigid\mbox{-}pair}\,\Pi_{n})Pn(𝗂𝗇𝖽-𝟤-𝗉𝗌𝗂𝗅𝗍Πn)P_{n}(\mathsf{ind\mbox{-}2\mbox{-}psilt}\,\Pi_{n})Pn(𝗂𝗇𝗍-𝗌𝗂𝗆n(Δn×Δ1))P_{n}(\mathsf{int\mbox{-}sim}_{n}(\Delta_{n}\times\Delta_{1}))[Miz14][AIR14]Corollary 3.17[AIR14][GKZ94]Proposition 3.2Proposition 3.2

Here

  • 𝗌τ-𝗍𝗂𝗅𝗍Πn\mathsf{s\tau\mbox{-}tilt}\,\Pi_{n} is the set of basic support τ\tau-tilting pairs over Πn\Pi_{n};

  • Pn(𝗂𝗇𝖽-τ-𝗋𝗂𝗀𝗂𝖽-𝗉𝖺𝗂𝗋Πn)P_{n}(\mathsf{ind\mbox{-}\tau\mbox{-}rigid\mbox{-}pair}\,\Pi_{n}) is the set of size-nn sets of indecomposable τ\tau-rigid pairs over Πn\Pi_{n};

  • 𝟤-𝗌𝗂𝗅𝗍Πn\mathsf{2\mbox{-}silt}\,\Pi_{n} is the set of basic two-term silting complexes over Πn\Pi_{n};

  • Pn(𝗂𝗇𝖽-𝟤-𝗉𝗌𝗂𝗅𝗍Πn)P_{n}(\mathsf{ind\mbox{-}2\mbox{-}psilt}\,\Pi_{n}) is the set of size-nn sets of indecomposable two-term presilting complexes over Πn\Pi_{n};

  • 𝗍𝗋𝗂(Δn×Δ1)\mathsf{tri}(\Delta_{n}\times\Delta_{1}) is the set of triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1};

  • Pn(𝗂𝗇𝗍-𝗌𝗂𝗆n(Δn×Δ1))P_{n}(\mathsf{int\mbox{-}sim}_{n}(\Delta_{n}\times\Delta_{1})) is the set of size-nn sets of internal nn-simplices in Δn×Δ1\Delta_{n}\times\Delta_{1};

  • \longleftrightarrow arrows denote bijections;

  • \dashrightarrow arrows denote direct sum decompositions, or decompositions into internal nn-simplices.

This paper is structured as follows. We give background to the paper in Section 2, first covering representation theory and, in particular, τ\tau-tilting theory, preprojective algebras of type AA, and their relation with 𝔖n+1\mathfrak{S}_{n+1}. We then cover background on triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1} and their relation with 𝔖n+1\mathfrak{S}_{n+1}. We prove our results in Section 3, first showing the bijection between internal simplices in Δn×Δ1\Delta_{n}\times\Delta_{1} and indecomposable two-term presilting complexes over Πn\Pi_{n}. We show how this induces a bijection between triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1} and two-term silting complexes over Πn\Pi_{n}. In order to do this, we make use of a certain factor algebra Π¯n\overline{\Pi}_{n} of Πn\Pi_{n} and prove some intermediate lemmas concerning its representation theory. Finally, we show that our bijections are compatible with the bijections with 𝔖n+1\mathfrak{S}_{n+1}.

2. Background

2.1. Representation theory

In this section, by Λ\Lambda we mean a finite-dimensional algebra over a field KK, and we write 𝗆𝗈𝖽Λ\mathsf{mod}\,\Lambda for the category of right Λ\Lambda-modules. We use τ\tau to denote the Auslander–Reiten translate in 𝗆𝗈𝖽Λ\mathsf{mod}\,\Lambda. We denote by 𝒦Λ:=𝖪𝖻(𝗉𝗋𝗈𝗃Λ)\mathcal{K}_{\Lambda}:=\mathsf{K}^{\sf b}(\mathsf{proj}\,\Lambda) the homotopy category of bounded complexes in 𝗉𝗋𝗈𝗃Λ\mathsf{proj}\,\Lambda.

2.1.1. Support τ\tau-tilting pairs

τ\tau-tilting theory was introduced in [AIR14] as a generalisation of cluster-tilting theory. A Λ\Lambda-module MM is called τ\tau-rigid if HomΛ(M,\operatorname{Hom}_{\Lambda}(M, τM)=0\tau M)=0. A pair (M,P)(M,P) of Λ\Lambda-modules where PP is projective is called τ\tau-rigid if MM is τ\tau-rigid and Hom(P,M)=0\operatorname{Hom}(P,M)=0. A τ\tau-rigid pair (M,P)(M,P) is called support τ\tau-tilting if |M|+|P|=|Λ||M|+|P|=|\Lambda|, where |X||X| denotes the number of non-isomorphic indecomposable direct summands of XX. Here MM is called a support τ\tau-tilting module. We call a τ\tau-rigid pair indecomposable if it is of the form (M,0)(M,0) with MM indecomposable, or of the form (0,P)(0,P) with PP indecomposable. We refer to τ\tau-rigid pairs of the form (0,P)(0,P) as shifted projectives. We write 𝗌τ-𝗍𝗂𝗅𝗍Λ\mathsf{s\tau\mbox{-}tilt}\,\Lambda for the set of (isomorphism classes of) basic support τ\tau-tilting pairs over Λ\Lambda and 𝗂𝗇𝖽-τ-𝗋𝗂𝗀𝗂𝖽-𝗉𝖺𝗂𝗋Λ\mathsf{ind\mbox{-}\tau\mbox{-}rigid\mbox{-}pair}\,\Lambda for the set of (isomorphism classes of) basic indecomposable τ\tau-rigid pairs over Λ\Lambda.

2.1.2. Two-term silting

Two-term silting complexes over Λ\Lambda are in bijection with support τ\tau-tilting pairs over Λ\Lambda and are in many ways nicer to work with [AIR14]. An object TT of 𝒦Λ\mathcal{K}_{\Lambda} is presilting if

Hom𝒦Λ(T,T[i])=0\operatorname{Hom}_{\mathcal{K}_{\Lambda}}(T,T[i])=0

for all i>0i>0. A presilting complex TT is silting if, additionally, thickT=𝒦Λ\operatorname{thick}T=\mathcal{K}_{\Lambda}. Here thickT\operatorname{thick}T denotes the smallest full subcategory of 𝒦Λ\mathcal{K}_{\Lambda} which contains PP and is closed under cones, [±1][\pm 1], direct summands, and isomorphisms. For a two-term complex TT to be presilting, it suffices that Hom𝒦Λ(T,T[1])=0\operatorname{Hom}_{\mathcal{K}_{\Lambda}}(T,T[1])=0. Moreover, for a presilting two-term complex TT to be silting, it suffices that |T|=|Λ||T|=|\Lambda| by [AIR14, Proposition 3.3(b)]. We write 𝟤-𝗌𝗂𝗅𝗍Λ\mathsf{2\mbox{-}silt}\,\Lambda for the set of (isomorphism classes of) two-term silting complexes over Λ\Lambda and 𝗂𝗇𝖽-𝟤-𝗉𝗌𝗂𝗅𝗍Λ\mathsf{ind\mbox{-}2\mbox{-}psilt}\,\Lambda for the set of (isomorphism classes of) indecomposable two-term presilting complexes over Λ\Lambda.

Given two-term silting complexes T,T𝖪[1,0](𝗉𝗋𝗈𝗃Λ)T,T^{\prime}\in\mathsf{K}^{[-1,0]}(\mathsf{proj}\,\Lambda), we say that TT^{\prime} is a mutation of TT if and only if T=EX,T=EYT=E\oplus X,T^{\prime}=E\oplus Y, where XX and YY are indecomposable with X≇YX\not\cong Y.

Adachi, Iyama, and Reiten showed that there was a bijection between 𝗂𝗇𝖽-𝟤-𝗉𝗌𝗂𝗅𝗍Λ\mathsf{ind\mbox{-}2\mbox{-}psilt}\,\Lambda and 𝗂𝗇𝖽-τ-𝗋𝗂𝗀𝗂𝖽-𝗉𝖺𝗂𝗋Λ\mathsf{ind\mbox{-}\tau\mbox{-}rigid\mbox{-}pair}\,\Lambda which induced a bijection between 𝟤-𝗌𝗂𝗅𝗍Λ\mathsf{2\mbox{-}silt}\,\Lambda and 𝗌τ-𝗍𝗂𝗅𝗍Λ\mathsf{s\tau\mbox{-}tilt}\,\Lambda [AIR14, Theorem 3.2]. Here, if (M,P)(M,P) is an indecomposable support τ\tau-rigid pair, then P1PP0P^{-1}\oplus P\to P^{0} is a two-term presilting complex, where P1P0P^{-1}\to P^{0} is a minimal projective presentation of MM.

2.1.3. Preprojective algebras of type AA

The algebras we are interested in are the preprojective algebras of type AA, which are defined as follows using quivers and relations. Preprojective algebras were originally defined by Gel\cprimefand and Ponomarev [GfP79]. They were subsequently generalised to non-simply-laced types in [GLS17]. The preprojective algebra of AnA_{n}, denoted Πn\Pi_{n}, has quiver QnQ_{n}

1{1}2{2}3{3}{\cdots}n1{n-1}n{n}α1\scriptstyle{\alpha_{1}}β1\scriptstyle{\beta_{1}}α2\scriptstyle{\alpha_{2}}β2\scriptstyle{\beta_{2}}α3\scriptstyle{\alpha_{3}}β3\scriptstyle{\beta_{3}}αn2\scriptstyle{\alpha_{n-2}}αn1\scriptstyle{\alpha_{n-1}}βn2\scriptstyle{\beta_{n-2}}βn1\scriptstyle{\beta_{n-1}}

with relations βiαi=αi+1βi+1\beta_{i}\alpha_{i}=\alpha_{i+1}\beta_{i+1} for i{1,2,,n1}i\in\{1,2,\dots,n-1\}, and α1β1=βn1αn1=0\alpha_{1}\beta_{1}=\beta_{n-1}\alpha_{n-1}=0. We compose arrows using the convention 𝛾𝛿=γδ\xrightarrow{\gamma}\xrightarrow{\delta}=\gamma\delta. We denote the idempotent at vertex ii by eie_{i} and the indecomposable projective Πn\Pi_{n}-module at vertex ii by PiP_{i}. The module PiP_{i} is also the indecomposable injective at vertex ni+1n-i+1.

In order to study the τ\tau-tilting theory of Πn\Pi_{n}, it will be useful to introduce a particular quotient of it. We let

z=i=1n1αiβiz=\sum_{i=1}^{n-1}\alpha_{i}\beta_{i}

in Πn\Pi_{n}. This is a central element of Πn\Pi_{n}. We define

Π¯n=Πn/z.\overline{\Pi}_{n}=\Pi_{n}/\langle z\rangle.

Then Π¯n\overline{\Pi}_{n} is the quotient of Πn\Pi_{n} by the ideal generated by all two-cycles. That is, Π¯n=Πn/I\overline{\Pi}_{n}=\Pi_{n}/I, where I=αiβi,βiαii{1,2,,n1}I=\langle\alpha_{i}\beta_{i},\beta_{i}\alpha_{i}\mid i\in\{1,2,\dots,n-1\}\rangle. This algebra was considered in [EJR18, BCZ19, DIR+17]. Denote by P¯i\overline{P}_{i} the projective Π¯n\overline{\Pi}_{n}-module at vertex ii. By [EJR18, Theorem 11], the functor ΠnΠ¯n:𝒦Πn𝒦Π¯n-\otimes_{\Pi_{n}}\overline{\Pi}_{n}\colon\mathcal{K}_{\Pi_{n}}\to\mathcal{K}_{\overline{\Pi}_{n}} induces a bijection 𝟤-𝗌𝗂𝗅𝗍Πn𝟤-𝗌𝗂𝗅𝗍Π¯n\mathsf{2\mbox{-}silt}\,\Pi_{n}\to\mathsf{2\mbox{-}silt}\,\overline{\Pi}_{n}. Given a two-term silting complex PP^{\bullet} of Πn\Pi_{n}, we write P¯\overline{P}^{\bullet} for the corresponding two-term silting complex over Π¯n\overline{\Pi}_{n}.

2.1.4. gg-vectors

Let P=P1P0P^{\bullet}=P^{-1}\to P^{0} be a two-term complex over projectives over Πn\Pi_{n}, where P0=i=1nPiriP^{0}=\bigoplus_{i=1}^{n}P_{i}^{r_{i}} and P1=i=1nPisiP^{-1}=\bigoplus_{i=1}^{n}P_{i}^{s_{i}}. Then the gg-vector of PP^{\bullet} is

gP:=(r1s1,r2s2,,rnsn).g^{P^{\bullet}}:=(r_{1}-s_{1},r_{2}-s_{2},\dots,r_{n}-s_{n}).

Note that this is the class of PP^{\bullet} in the Grothendieck group K0(𝒦Πn)=K0(𝗉𝗋𝗈𝗃Πn)K_{0}(\mathcal{K}_{\Pi_{n}})=K_{0}(\mathsf{proj}\,\Pi_{n}), identified with n\mathbb{Z}^{n}. Given a τ\tau-rigid pair (M,P)(M,P), we define g(M,P):=gPg^{(M,P)}:=g^{P^{\bullet}} where PP^{\bullet} is the corresponding two-term presilting complex. It is known that two-term presilting complexes and support τ\tau-rigid pairs are uniquely determined by their gg-vectors [AIR14, Theorem 5.5].

2.1.5. Relation with permutations

It was shown in [Miz14] (see also [BIRS09]) that support τ\tau-tilting modules over Πn\Pi_{n} were in bijection with permutations in 𝔖n+1\mathfrak{S}_{n+1}. This was generalised to non-simply-laced types in [FG19, Mur22]. The bijection is constructed as follows. Define Ii:=Πn(1ei)ΠnI_{i}:=\Pi_{n}(1-e_{i})\Pi_{n}, the principal two-sided ideal of Πn\Pi_{n} generated by (1ei)(1-e_{i}). Then, for w𝔖n+1w\in\mathfrak{S}_{n+1} with reduced expression w=si1si2sikw=s_{i_{1}}s_{i_{2}}\dots s_{i_{k}}, define

Iw=Ii1Ii2Iik.I_{w}=I_{i_{1}}I_{i_{2}}\dots I_{i_{k}}.

The map wIww\mapsto I_{w} then defines a bijection from 𝔖n+1\mathfrak{S}_{n+1} to the support τ\tau-tilting modules over Πn\Pi_{n}. In the case w=ew=e, the identity permutation, the corresponding support τ\tau-tilting module is the regular module Πn\Pi_{n}. We write PwP_{w} for the projective module PwP_{w} such that (Iw,Pw)(I_{w},P_{w}) is a support τ\tau-tilting pair.

Convention 2.1.

We shall use the convention that the base set for the symmetric group 𝔖n+1\mathfrak{S}_{n+1} is {0,1,,n}\{0,1,\dots,n\} and that the simple reflection sis_{i} is given by the transposition (i1i)(i-1\ i).

2.2. Convex polytopes

We now give background on the side of the paper concerning triangulations of the prism Δn×Δ1\Delta_{n}\times\Delta_{1}.

2.2.1. Prisms of simplices

We now explain the relevant background on prisms of simplices, following [DLRS10, Section 6.2.1]. The prism of the nn-simplex Δn\Delta_{n} is the polytope Δn×Δ1\Delta_{n}\times\Delta_{1}. The particular geometric realisation of this polytope is not important for the results of this paper, but for the sake of clarity, we take Δn×Δ1\Delta_{n}\times\Delta_{1} to be the convex hull of the points given by the column vectors of the (n+3)×(2n+2)(n+3)\times(2n+2) matrix

(100100010010001001111000000111).\begin{pmatrix}1&0&\dots&0&1&0&\dots&0\\ 0&1&\dots&0&0&1&\dots&0\\ \vdots&\vdots&\ddots&\vdots&\vdots&\vdots&\ddots&\vdots\\ 0&0&\dots&1&0&0&\dots&1\\ 1&1&\dots&1&0&0&\dots&0\\ 0&0&\dots&0&1&1&\dots&1\end{pmatrix}.

We label the points given by the first n+1n+1 columns of this matrix by a0,a1,,ana_{0},a_{1},\dots,a_{n} and the points given by the last n+1n+1 columns by b0,b1,,bnb_{0},b_{1},\dots,b_{n}. We write V={a0,a1,,an,b0,b1,,bn}V=\{a_{0},a_{1},\dots,a_{n},b_{0},b_{1},\dots,b_{n}\} for the set of vertices of the prism Δn×Δ1\Delta_{n}\times\Delta_{1}. One can specify kk-simplices in the prism Δn×Δ1\Delta_{n}\times\Delta_{1} by giving subsets of VV of size k+1k+1, provided the chosen vertices are affinely independent. An internal kk-simplex is one which does not lie in the boundary of Δn×Δ1\Delta_{n}\times\Delta_{1}. We write 𝗂𝗇𝗍-𝗌𝗂𝗆n(Δn×Δ1)\mathsf{int\mbox{-}sim}_{n}(\Delta_{n}\times\Delta_{1}) for the set of internal nn-simplices of Δn×Δ1\Delta_{n}\times\Delta_{1}.

2.2.2. Triangulations

We again follow [DLRS10]. A polyhedral subdivision 𝒮\mathcal{S} of Δn×Δ1\Delta_{n}\times\Delta_{1} is a set of (n+1)(n+1)-dimensional convex polytopes {L1,,Lr}\{L_{1},\dots,L_{r}\} such that

  • i=1rLi=Δn×Δ1\bigcup_{i=1}^{r}L_{i}=\Delta_{n}\times\Delta_{1} and

  • LiLjL_{i}\cap L_{j} is a face of both LiL_{i} and LjL_{j} for all i,ji,j.

A polyhedral subdivision 𝒯\mathcal{T} is a triangulation if LiL_{i} is a simplex for every ii. We write 𝗍𝗋𝗂(Δn×Δ1)\mathsf{tri}(\Delta_{n}\times\Delta_{1}) for the set of triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1}.

One polyhedral subdivision 𝒮={L1,L2,,Lr}\mathcal{S}=\{L_{1},L_{2},\dots,L_{r}\} refines another polyhedral subdivision 𝒮={L1,L2,,Lr}\mathcal{S}^{\prime}=\{L^{\prime}_{1},L^{\prime}_{2},\dots,L^{\prime}_{r^{\prime}}\} if for every LiL_{i} we have LiLjL_{i}\subseteq L^{\prime}_{j} for some jj, and 𝒮𝒮\mathcal{S}\neq\mathcal{S}^{\prime}. A polyhedral subdivision is an almost triangulation if all of its refinements are triangulations. Two triangulations 𝒯\mathcal{T} and 𝒯\mathcal{T}^{\prime} are related by a bistellar flip if there is an almost triangulation 𝒮\mathcal{S} whose only two refinements are 𝒯\mathcal{T} and 𝒯\mathcal{T}^{\prime}.

Remark 2.2.

Bistellar flips are the generalisation of the operation of flipping a diagonal inside a quadrilateral in a triangulated convex polygon. Here, the almost triangulation is given by a subdivision consisting of triangles and one quadrilateral. The two triangulations of the quadrilateral then give the two refinements of this subdivision which are the triangulations related by a bistellar flip.

A circuit of Δn×Δ1\Delta_{n}\times\Delta_{1} is a pair (Z,Z)(Z,Z^{\prime}) of two disjoint vertex subsets ZZ and ZZ^{\prime} such that conv(Z)conv(Z)\operatorname{conv}(Z)\cap\operatorname{conv}(Z^{\prime})\neq\varnothing and such that ZZ and ZZ^{\prime} are minimal with respect to this property. Here ‘conv\operatorname{conv}’ denotes the convex hull. The circuits of Δn×Δ1\Delta_{n}\times\Delta_{1} are given by ({ai,bj},{aj,bi})(\{a_{i},b_{j}\},\{a_{j},b_{i}\}) for iji\neq j.

2.2.3. Relation with permutations

Triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1} are in bijection with permutations in the symmetric group 𝔖n+1\mathfrak{S}_{n+1}, as shown in [GKZ94, Chapter 7, Section 3C]. Indeed, for a permutation w=i0i1in𝔖n+1w=i_{0}i_{1}\dots i_{n}\in\mathfrak{S}_{n+1}, we have that

{{ai0,,aij,bij,,bin}0jn}\{\{a_{i_{0}},\dots,a_{i_{j}},b_{i_{j}},\dots,b_{i_{n}}\}\mid 0\leqslant j\leqslant n\}

is the set of (n+1)(n+1)-simplices of the corresponding triangulation, and all triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1} arise in this way. For a permutation w𝔖n+1w\in\mathfrak{S}_{n+1}, we write 𝒯w\mathcal{T}_{w} for the corresponding triangulation of Δn×Δ1\Delta_{n}\times\Delta_{1}.

3. Results

In this section, we show our main result that there is a bijection between the sets 𝗂𝗇𝖽-𝟤-𝗉𝗌𝗂𝗅𝗍Πn\mathsf{ind\mbox{-}2\mbox{-}psilt}\,\Pi_{n} of indecomposable two-term presilting complexes over Πn\Pi_{n} and 𝗂𝗇𝗍-𝗌𝗂𝗆n(Δn×Δ1)\mathsf{int\mbox{-}sim}_{n}(\Delta_{n}\times\Delta_{1}) of internal nn-simplices in Δn×Δ1\Delta_{n}\times\Delta_{1} which induces a bijection between the sets 𝟤-𝗌𝗂𝗅𝗍Πn\mathsf{2\mbox{-}silt}\,\Pi_{n} of two-term silting complexes over Πn\Pi_{n} and 𝗍𝗋𝗂(Δn×Δ1)\mathsf{tri}(\Delta_{n}\times\Delta_{1}) of triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1}.

3.1. Bijection for simplices

We begin by showing the first bijection. We write

𝗌𝖾𝗊n+1(a,b):={a,b}n+1{an+1,bn+1}\mathsf{seq}_{n+1}(a,b):=\{a,b\}^{n+1}\setminus\{a^{n+1},b^{n+1}\}

for the set of words of length n+1n+1 in the alphabet {a,b}\{a,b\} which use both letters. Recall that a facet of a polytope is a face of codimension one.

Lemma 3.1.

There is a bijection between 𝗂𝗇𝗍-𝗌𝗂𝗆n(Δn×Δ1)\mathsf{int\mbox{-}sim}_{n}(\Delta_{n}\times\Delta_{1}) and 𝗌𝖾𝗊n+1(a,b)\mathsf{seq}_{n+1}(a,b).

Proof.

Clearly {a0,a1,,an}\{a_{0},a_{1},\dots,a_{n}\} and {b0,b1,,bn}\{b_{0},b_{1},\dots,b_{n}\} both correspond to facets of Δn×Δ1\Delta_{n}\times\Delta_{1}. Furthermore, V{ai,bi}V\setminus\{a_{i},b_{i}\} corresponds to a facet of Δn×Δ1\Delta_{n}\times\Delta_{1} for all ii, since it is the Cartesian product of a facet of Δn\Delta_{n} with Δ1\Delta_{1}. Thus, an internal nn-simplex in Δn×Δ1\Delta_{n}\times\Delta_{1} is given by choosing aia_{i} or bib_{i} for each ii, excluding the cases {a0,a1,,an}\{a_{0},a_{1},\dots,a_{n}\} and {b0,b1,,bn}\{b_{0},b_{1},\dots,b_{n}\}. Such a set of vertices is, moreover, affinely independent, and so indeed has an nn-simplex as its convex hull. Hence, internal nn-simplices in Δn×Δ1\Delta_{n}\times\Delta_{1} correspond bijectively to words of length n+1n+1 in the alphabet {a,b}\{a,b\} which use both letters. ∎

Given X𝗌𝖾𝗊n+1(a,b)X\in\mathsf{seq}_{n+1}(a,b), we write ΔX\Delta_{X} for the corresponding internal nn-simplex in 𝗂𝗇𝗍-𝗌𝗂𝗆n(Δn×Δ1)\mathsf{int\mbox{-}sim}_{n}(\Delta_{n}\times\Delta_{1}).

Proposition 3.2.

The sets 𝗂𝗇𝗍-𝗌𝗂𝗆n(Δn×Δ1)\mathsf{int\mbox{-}sim}_{n}(\Delta_{n}\times\Delta_{1}) and 𝗂𝗇𝖽-τ-𝗋𝗂𝗀𝗂𝖽-𝗉𝖺𝗂𝗋Πn\mathsf{ind\mbox{-}\tau\mbox{-}rigid\mbox{-}pair}\,\Pi_{n} are in bijection with each other. Moreover, the sets 𝗂𝗇𝗍-𝗌𝗂𝗆n(Δn×Δ1)\mathsf{int\mbox{-}sim}_{n}(\Delta_{n}\times\Delta_{1}) and 𝗂𝗇𝖽-𝟤-𝗉𝗌𝗂𝗅𝗍Πn\mathsf{ind\mbox{-}2\mbox{-}psilt}\,\Pi_{n} are also in bijection with each other.

Proof.

We argue in terms of indecomposable τ\tau-rigid pairs. It follows from [Miz14] that the indecomposable τ\tau-rigid modules over Πn\Pi_{n} correspond to submodules of indecomposable injectives—see also [IRRT18]. Thus, the indecomposable τ\tau-rigid modules with socle ii correspond to Young diagrams lying in an (n+1i)×i(n+1-i)\times i grid, rotated so that the north-west corner of the Young diagram is lying at the bottom. Such Young diagrams are determined by the path given by their upper contour. Let us label upwards steps in the path by aa and downwards steps in the path by bb and orient the paths from left to right. Such Young diagrams correspond to words in the alphabet {a,b}\{a,b\} with iibb’s and n+1in+1-iaa’s, such that at least one ‘aa’ precedes a ‘bb’. This can be seen in Example 3.4 and Figure 1.

We associate the shifted projective with top ii to the word given by iibb’s followed by n+1in+1-iaa’s. Such a word gives an empty Young diagram, and so is not included in the words corresponding to the τ\tau-rigid modules.

We obtain that the indecomposable τ\tau-rigid pairs over Πn\Pi_{n} are in bijection with 𝗌𝖾𝗊n+1(a,b)\mathsf{seq}_{n+1}(a,b), and so are in bijection with the internal nn-simplices in Δn×Δ1\Delta_{n}\times\Delta_{1} by Proposition 3.1.

By [AIR14], we obtain the second statement. ∎

Given a indecomposable τ\tau-rigid pair (M,0)(M,0), we write word(M,0)\mathrm{word}(M,0) for the corresponding word in the alphabet {a,b}\{a,b\}, usually abbreviating this to word(M)\mathrm{word}(M). We do likewise for word(0,P)\mathrm{word}(0,P).

Remark 3.3.

In [IRRT18], it was shown that join-irreducible permutations in 𝔖n+1\mathfrak{S}_{n+1} were in bijection with τ\tau-rigid modules over Πn\Pi_{n}. The bijection between τ\tau-rigid modules and 𝗌𝖾𝗊n+1(a,b){bkan+1k1kn}\mathsf{seq}_{n+1}(a,b)\setminus\{b^{k}a^{n+1-k}\mid 1\leqslant k\leqslant n\} gives a neat way of seeing this. Indeed, the join-irreducible permutations are the ones with precisely one descent. That is, if the permutation is i0i1in𝔖n+1i_{0}i_{1}\dots i_{n}\in\mathfrak{S}_{n+1}, there is some l{1,2,,n}l\in\{1,2,\dots,n\} such that

i0<<il1>il<<in.i_{0}<\dots<i_{l-1}>i_{l}<\dots<i_{n}.

Such a permutation corresponds to the internal nn-simplex ΔX\Delta_{X} in Δn×Δ1\Delta_{n}\times\Delta_{1} with word X=x0x1xnX=x_{0}x_{1}\dots x_{n} where xij=ax_{i_{j}}=a for j{0,1,,l1}j\in\{0,1,\dots,l-1\} and xij=bx_{i_{j}}=b otherwise. This then gives a τ\tau-rigid module via Proposition 3.2. Note that the words bkan+1kb^{k}a^{n+1-k} corresponding to shifted projectives do not arise via this construction, since il1>ili_{l-1}>i_{l} guarantees that at least one ‘aa’ entry lies to the right of a ‘bb’ entry.

Example 3.4.

We give the example of the indecomposable τ\tau-rigid pairs of Π2\Pi_{2}, as shown in Figure 1. We use [1][1] to denote the indecomposable τ\tau-rigid pairs of the form (0,P)(0,P). The words baabaa and bbabba correspond to the indecomposable τ\tau-rigid pairs

(0,\tabbedCenterstack21) and (0,\tabbedCenterstack12).\left(0,{\tabbedCenterstack{2\\ 1}}\right)\text{ and }\left(0,{\tabbedCenterstack{1\\ 2}}\right).

The internal 2-simplices in Δ2×Δ1\Delta_{2}\times\Delta_{1} are shown in Figure 2. These internal 2-simplices are in the same position in the figure as their corresponding indecomposable τ\tau-rigid pairs in Figure 1.

Figure 1. Indecomposable τ\tau-rigid pairs for Π2\Pi_{2}
12aaaabb12aabbaa12bbaaaa[1][1]12aabbbb12bbaabb12bbbbaa[1][1]
Figure 2. Internal 22-simplices in Δ2×Δ1\Delta_{2}\times\Delta_{1}
a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}
Proposition 3.5.

Given a τ\tau-rigid pair (M,P)(M,P) of Πn\Pi_{n} corresponding to an internal nn-simplex x0x1xnx_{0}x_{1}\dots x_{n} in Δn×Δ1\Delta_{n}\times\Delta_{1}, the gg-vector g(M,P)=(g1,g2,,gn)g^{(M,P)}=(g_{1},g_{2},\dots,g_{n}) has entries

gi={1 if xi1xi=ab,1 if xi1xi=ba,0 otherwise.g_{i}=\left\{\begin{array}[]{cl}1&\text{ if }x_{i-1}x_{i}=ab,\\ -1&\text{ if }x_{i-1}x_{i}=ba,\\ 0&\text{ otherwise.}\end{array}\right.
Proof.

Consider an indecomposable τ\tau-rigid module MM over Πn\Pi_{n} with its dimension vector displayed as in Figure 1. Let P1P0P^{-1}\to P^{0} be the projective presentation of MM. The indecomposable projective with top ii is a direct summand of P0P^{0} if and only if ii is a peak in the upper contour of MM. Similarly, the indecomposable projective with top ii is a direct summand of P1P^{-1} if and only if ii is a trough of the upper contour of MM. Since peaks in the upper contour correspond to baba and troughs correspond to abab, we obtain the result. ∎

Remark 3.6.

Proposition 3.5 gives the gg-vector of the indecomposable two-term silting complex P1P0P^{-1}\to P^{0} corresponding to an internal nn-simplex in Δn×Δ1\Delta_{n}\times\Delta_{1}, and so gives the indecomposable summands of P1P^{-1} and P0P^{0}. The map P1P0P^{-1}\to P^{0} can then be described as follows. If the indecomposable summands of P1P^{-1} are Pi1,Pi2,,PilP_{i_{1}},P_{i_{2}},\dots,P_{i_{l}}, then P0P_{0} must either have l1l-1, ll, or l+1l+1 summands. The three cases behave quite similarly. If P0P^{0} has l+1l+1 summands Pj1,,Pjl+1P_{j_{1}},\dots,P_{j_{l+1}}, then

j1<i1<j2<i2<<jl<il<jl+1.j_{1}<i_{1}<j_{2}<i_{2}<\dots<j_{l}<i_{l}<j_{l+1}.

Moreover, the component of the map from P1P0P^{-1}\to P^{0} from PirP_{i_{r}} to PjsP_{j_{s}} is zero unless s{r,r+1}s\in\{r,r+1\}, in which case, the component is the map PirPjsP_{i_{r}}\to P_{j_{s}} which has as large an image as possible. Such a map is unique up to scalar.

One can therefore use Proposition 3.5 to construct the bijection between the sets 𝗂𝗇𝖽-𝟤-𝗉𝗌𝗂𝗅𝗍Πn\mathsf{ind\mbox{-}2\mbox{-}psilt}\,\Pi_{n} and 𝗌𝖾𝗊n+1(a,b)\mathsf{seq}_{n+1}(a,b) directly, without using Proposition 3.2. We also write word(P)\mathrm{word}(P^{\bullet}) for the word corresponding to an indecomposable two-term presilting complex PP^{\bullet}.

3.2. Representation theory of Π¯n\overline{\Pi}_{n}

In this section, we give some results describing the representation theory of Π¯n\overline{\Pi}_{n} combinatorially, which will then be used to prove the relation between Πn\Pi_{n} and triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1}. We first note the following.

Lemma 3.7.

Every indecomposable Π¯n\overline{\Pi}_{n}-module is τ\tau-rigid.

Proof.

We have from [BCZ19, Proposition 4.1.2] that every indecomposable Π¯n\overline{\Pi}_{n}-module is a brick, meaning that every non-zero endomorphism is an isomorphism. It follows from [DIJ19, Theorem 4.1] that every indecomposable Π¯n\overline{\Pi}_{n}-module must also be τ\tau-rigid, since Π¯n\overline{\Pi}_{n} is representation-finite by [BCZ19, Proposition 4.1.]. ∎

The indecomposable τ\tau-rigid pairs over Π¯n\overline{\Pi}_{n} can be described in a way analogous to those of Πn\Pi_{n} [DIR+17].

Lemma 3.8.

There is a bijection between 𝗂𝗇𝖽-𝟤-𝗉𝗌𝗂𝗅𝗍Π¯n\mathsf{ind\mbox{-}2\mbox{-}psilt}\,\overline{\Pi}_{n}, 𝗂𝗇𝖽-τ-𝗋𝗂𝗀𝗂𝖽-𝗉𝖺𝗂𝗋Π¯n\mathsf{ind\mbox{-}\tau\mbox{-}rigid\mbox{-}pair}\,\overline{\Pi}_{n}, and 𝗌𝖾𝗊n+1(a,b)\mathsf{seq}_{n+1}(a,b), given in the same way as in Proposition 3.2 and Proposition 3.5.

Proof.

This bijection follows from Proposition 3.5 and [EJR18, Theorem 11], which says that the gg-vectors of indecomposable presilting complexes for Π¯n\overline{\Pi}_{n} must be the same as those for Πn\Pi_{n}. The bijection between 𝗌𝖾𝗊n+1(a,b)\mathsf{seq}_{n+1}(a,b) and 𝗂𝗇𝖽-τ-𝗋𝗂𝗀𝗂𝖽-𝗉𝖺𝗂𝗋Π¯n\mathsf{ind\mbox{-}\tau\mbox{-}rigid\mbox{-}pair}\,\overline{\Pi}_{n} then follows from [AIR14], and it is clear that it can be constructed in a similar way to Proposition 3.2. ∎

For indecomposable τ\tau-rigid Π¯n\overline{\Pi}_{n}-modules MM and indecomposable presilting complexes P¯\overline{P}^{\bullet} over Π¯n\overline{\Pi}_{n}, we write word(M)\mathrm{word}(M) and word(P¯)\mathrm{word}(\overline{P}^{\bullet}) for the corresponding words in 𝗌𝖾𝗊n+1(a,b)\mathsf{seq}_{n+1}(a,b), as we do with Πn\Pi_{n}.

Example 3.9.

Consider the Π¯6\overline{\Pi}_{6}-module MM given by

\tabbedCenterstack213465.{\tabbedCenterstack{\phantom{1}2\phantom{3}\phantom{4}\phantom{5}\phantom{6}\\ 1\phantom{2}3\phantom{4}\phantom{5}\phantom{6}\\ \phantom{1}\phantom{2}\phantom{3}4\phantom{5}6\\ \phantom{1}\phantom{2}\phantom{3}\phantom{4}5\phantom{6}}}\,.

Then word(M)=aabbbab\mathrm{word}(M)=aabbbab, using Lemma 3.8 and Proposition 3.2. Furthermore, using Lemma 3.8 and Proposition 3.5, the projective presentation of MM is given by P¯5P¯2P¯6\overline{P}_{5}\to\overline{P}_{2}\oplus\overline{P}_{6}.

Similarly, the Π¯6\overline{\Pi}_{6}-module NN given by

\tabbedCenterstack243{\tabbedCenterstack{2\phantom{3}4\\ \phantom{2}3\phantom{4}}}

has word word(N)=bababaa\mathrm{word}(N)=bababaa and projective presentation P¯1P¯3P¯5P¯2P¯4\overline{P}_{1}\oplus\overline{P}_{3}\oplus\overline{P}_{5}\to\overline{P}_{2}\oplus\overline{P}_{4}.

The following criterion for identifying the support of an indecomposable Π¯n\overline{\Pi}_{n}-module from its word is very useful.

Lemma 3.10.

An indecomposable Π¯n\overline{\Pi}_{n}-module MM with word(M)=x0x1xn\mathrm{word}(M)=x_{0}x_{1}\dots x_{n} is supported at vertex jj if and only if there exist ii and kk such that ij<ki\leqslant j<k such that xi=ax_{i}=a and xk=bx_{k}=b.

Proof.

It suffices to note that the composition factors of MM always form an interval [i+1,k][i+1,k] in [n][n] and that the first of these composition factors i+1i+1 will occur when xix_{i} is the first occurrence of ‘aa’ and the last composition factor kk will occur when xkx_{k} is the last occurrence of ‘bb’. Before the first occurrence of ‘aa’ and after the first occurrence of ‘bb’, the contour of the module MM will go along the boundary of the Young diagram, and so MM will not be supported at the corresponding simple module. Between these occurrence, the contour of MM never reaches the boundary of the Young diagram again, so MM is supported at all of the simples. ∎

We can describe the Auslander–Reiten translate over Π¯n\overline{\Pi}_{n}. This can also be seen from the interpretation of the Auslander–Reiten translate for string algebras in terms of adding and removing hooks and cohooks from [BR87].

Lemma 3.11.

Let MM be an indecomposable non-projective Π¯n\overline{\Pi}_{n}-module with

word(M)=arbwabs,\mathrm{word}(M)=a^{r}\,b\,w\,a\,b^{s},

where ww is some arbitrary, possibly empty, subword, and possibly r=0r=0 and possibly s=0s=0. Then

word(τM)=brawbas.\mathrm{word}(\tau M)=b^{r}\,a\,w\,b\,a^{s}.
Proof.

Let P1P0M0P^{-1}\to P^{0}\to M\to 0 be the projective presentation of MM. Then there is an exact sequence 0τMνP1νP00\to\tau M\to\nu P^{-1}\to\nu P^{0}. Note that ν\nu sends a projective at a given vertex to the injective at the same vertex. Hence, τM\tau M has an injective presentation by the injectives at the same vertices as the projectives in the projective presentation of MM. Using the dual of Lemma 3.8, we conclude that the lower contour of τM\tau M is given by word(M)\mathrm{word}(M). Hence, word(τM)\mathrm{word}(\tau M) is obtained from word(τM)\mathrm{word}(\tau M) as shown in Figure 3, which establishes the result. (The cases where only one of rr and ss is non-zero are easily extrapolated from those shown in the figure.) ∎

Example 3.12.

We give an example where we compute the Auslander–Reiten translate of an indecomposable Π¯n\overline{\Pi}_{n}-module. We consider the Π¯5\overline{\Pi}_{5}-module MM given by

\tabbedCenterstack2413.{\tabbedCenterstack{\phantom{1}2\phantom{3}4\\ 1\phantom{2}3\phantom{4}}}.

Then word(M)=aababa\mathrm{word}(M)=aababa, so word(τM)=bbaabb\mathrm{word}(\tau M)=bbaabb. Hence τM\tau M is the module

\tabbedCenterstack435.{\tabbedCenterstack{\phantom{3}4\phantom{5}\\ 3\phantom{4}5}}.

This can be visualised in Figure 4.

Figure 3. Illustration of Lemma 3.11
aaaabbbbwwwwara^{r}brb^{r}bsb^{s}asa^{s}r0r\neq 0 and s0s\neq 0 caseaaaabbbbwwwwr=0r=0 and s=0s=0 case
Figure 4. Illustration of Example 3.12
1234345

The following description of submodules and factor modules of indecomposable Π¯n\overline{\Pi}_{n}-modules is key. Here,

X=ub𝑖vX=u\overset{i}{b}v

means that b=xib=x_{i}, where X=x0xnX=x_{0}\dots x_{n}. We use Π¯[i+1,j]\overline{\Pi}_{[i+1,j]} to denote the algebra Π¯ji\overline{\Pi}_{j-i} with the vertex set given by [i+1,j]:={i+1,i+2,,j}[i+1,j]:=\{i+1,i+2,\dots,j\}, in the natural way.

Lemma 3.13.

Let MM be an indecomposable Π¯n\overline{\Pi}_{n}-module. Then MM has a submodule supported on the vertices {i+1,,j}\{i+1,\dots,j\} if and only if one of the following four cases hold.

  1. (1)

    word(M)=bsaub𝑖wa𝑗vbar\mathrm{word}(M)=b^{s}\,a\,u\,\overset{i}{b}\,w\,\overset{j}{a}\,v\,b\,a^{r}, where possibly r=0r=0 or s=0s=0.

  2. (2)

    word(M)=bia𝑖wa𝑗vbar\mathrm{word}(M)=b^{i}\,\overset{i}{a}\,w\,\overset{j}{a}\,v\,b\,a^{r}, where possibly r=0r=0 or i=0i=0.

  3. (3)

    word(M)=bsaub𝑖wb𝑗anj\mathrm{word}(M)=b^{s}\,a\,u\,\overset{i}{b}\,w\,\overset{j}{b}\,a^{n-j}, where possibly s=0s=0 or j=nj=n.

  4. (4)

    word(M)=bia𝑖wb𝑗anj\mathrm{word}(M)=b^{i}\,\overset{i}{a}\,w\,\overset{j}{b}\,a^{n-j}, where possibly i=0i=0 or j=nj=n.

In each case, if LL is the Π¯[i+1,j]\overline{\Pi}_{[i+1,j]}-module corresponding to the Π¯n\overline{\Pi}_{n}-submodule of MM, then word(L)=awb\mathrm{word}(L)=awb. Note that ww is possibly empty, as are uu and vv.

Dually, MM has a factor module supported on the vertices {i+1,,j}\{i+1,\dots,j\} if and only if one of the following four cases hold.

  1. (1)

    word(M)=bsaua𝑖wb𝑗vbar\mathrm{word}(M)=b^{s}\,a\,u\,\overset{i}{a}\,w\,\overset{j}{b}\,v\,b\,a^{r}, where possibly r=0r=0 or s=0s=0.

  2. (2)

    word(M)=bia𝑖wb𝑗vbar\mathrm{word}(M)=b^{i}\,\overset{i}{a}\,w\,\overset{j}{b}\,v\,b\,a^{r}, where possibly r=0r=0 or i=0i=0.

  3. (3)

    word(M)=bsaua𝑖wb𝑗anj\mathrm{word}(M)=b^{s}\,a\,u\,\overset{i}{a}\,w\,\overset{j}{b}\,a^{n-j}, where possibly s=0s=0 or j=nj=n.

  4. (4)

    word(M)=bia𝑖wb𝑗anj\mathrm{word}(M)=b^{i}\,\overset{i}{a}\,w\,\overset{j}{b}\,a^{n-j}, where possibly i=0i=0 or j=nj=n.

In each case, if NN is the Π¯[i+1,j]\overline{\Pi}_{[i+1,j]}-module corresponding to the Π¯n\overline{\Pi}_{n}-submodule of MM, then word(N)=awb\mathrm{word}(N)=awb. Again, ww is possibly empty, as are uu and vv; and possibly r=0r=0 or s=0s=0.

Proof.

We focus on the submodule case. The four cases correspond as follows.

  1. (1)

    i0i\neq 0 and jnj\neq n and MM is supported on ii and j+1j+1.

  2. (2)

    i=0i=0 or MM is not supported on ii, but jnj\neq n and MM is supported on j+1j+1.

  3. (3)

    i0i\neq 0 and MM is supported at ii, but j=nj=n or MM is not supported on j+1j+1.

  4. (4)

    i=0i=0 or MM is not supported at ii, and j=nj=n or MM is not supported at j+1j+1.

These cases are clearly exhaustive. We deal with them one by one. We let word(M)=x0x1xn\mathrm{word}(M)=x_{0}x_{1}\dots x_{n}.

Case (1): It follows from Lemma 3.10 that if MM is supported on ii and j+1j+1, then ‘aa’ must occur before ii and ‘bb’ must occur after jj. Since the vector subspace LL of MM supported on {i+1,,j}\{i+1,\dots,j\} is a submodule, then we must have xi=bx_{i}=b and xj=ax_{j}=a if LL is to be closed under the action of Π¯n\overline{\Pi}_{n}.

Case (2): As in case (1), we conclude that ‘bb’ occurs after jj, and that xj=ax_{j}=a. If i=0i=0, then s=0s=0; moreover, x0=xi=ax_{0}=x_{i}=a by Lemma 3.10, since MM is supported on i+1i+1 by assumption. If i0i\neq 0, then, since MM is not supported on ii, by Lemma 3.10, we must have that x0xi1=bix_{0}\dots x_{i-1}=b^{i}.

Cases (3) and (4) can be shown in a similar way to case (2). In all cases, we have that word(L)=awb\mathrm{word}(L)=awb by Lemma 3.10, since LL is sincere by assumption, and must have the shape prescribed by the word ww.

The dual cases concerning a factor module of MM follow similarly. Case (1) is straightforwardly dual. In case (2), one can reason that x0xi=biax_{0}\dots x_{i}=b^{i}a in the same way to in the submodule case. Showing that xjxn=vbasx_{j}\dots x_{n}=vba^{s} is then done in the same way as in the factor module case (1). The remaining cases can be done in an analogous way. ∎

We can now prove a criterion for non-vanishing of HomΠ¯n(N,τM)0\operatorname{Hom}_{\overline{\Pi}_{n}}(N,\tau M)\neq 0 for indecomposable Π¯n\overline{\Pi}_{n}-modules MM and NN.

Lemma 3.14.

Let MM and NN be indecomposable Π¯n\overline{\Pi}_{n}-modules with word(M)=x0x1xn\mathrm{word}(M)=x_{0}x_{1}\dots x_{n} and word(N)=y0y1yn\mathrm{word}(N)=y_{0}y_{1}\dots y_{n}. Then HomΠ¯n(N,τM)0\operatorname{Hom}_{\overline{\Pi}_{n}}(N,\tau M)\neq 0 if and only if there exist ii and jj such that xi=bx_{i}=b, xj=ax_{j}=a, yi=ay_{i}=a, and yj=by_{j}=b.

Proof.

First note that if MM is projective, then τM=0\tau M=0, so HomΠ¯n(N,τM)=0\operatorname{Hom}_{\overline{\Pi}_{n}}(N,\tau M)=0. In this case, word(M)=albnl+1\mathrm{word}(M)=a^{l}b^{n-l+1} for some l{1,2,,n}l\in\{1,2,\dots,n\}, so it is clear that there can be no i<ji<j with xi=bx_{i}=b and xj=ax_{j}=a.

Excluding the case where MM is projective, we have that HomΠ¯n(N,τM)0\operatorname{Hom}_{\overline{\Pi}_{n}}(N,\tau M)\neq 0 if and only if there is a non-zero factor module of NN which is a submodule of τM\tau M, namely the image of the non-zero map. Hence, this is the case if and only if word(τM)\mathrm{word}(\tau M) falls into one of the submodule cases of Lemma 3.13 and word(N)\mathrm{word}(N) falls into one of the factor module cases. Hence, there exist ii and jj such that yi=ay_{i}=a and yj=by_{j}=b, whilst word(τM)\mathrm{word}(\tau M) falls into one of the following cases.

  1. (1)

    word(τM)=bsaub𝑖wa𝑗vbar\mathrm{word}(\tau M)=b^{s}\,a\,u\,\overset{i}{b}\,w\,\overset{j}{a}\,v\,b\,a^{r}, where possibly s=0s=0 or r=0r=0.

  2. (2)

    word(τM)=bia𝑖wa𝑗vbar\mathrm{word}(\tau M)=b^{i}\,\overset{i}{a}\,w\,\overset{j}{a}\,v\,b\,a^{r}, where possibly i=0i=0 or r=0r=0.

  3. (3)

    word(τM)=bsaub𝑖wb𝑗anj\mathrm{word}(\tau M)=b^{s}\,a\,u\,\overset{i}{b}\,w\,\overset{j}{b}\,a^{n-j}, where possibly s=0s=0 or j=nj=n.

  4. (4)

    word(τM)=bia𝑖wb𝑗anj\mathrm{word}(\tau M)=b^{i}\,\overset{i}{a}\,w\,\overset{j}{b}\,a^{n-j}, where possibly i=0i=0 or j=nj=n.

Applying Lemma 3.11, we have these four cases holds if and only if the corresponding one of the following four options hold for word(M)\mathrm{word}(M).

  1. (1)

    word(M)=asbub𝑖wa𝑗vabr\mathrm{word}(M)=a^{s}\,b\,u\,\overset{i}{b}\,w\,\overset{j}{a}\,v\,a\,b^{r}, where possibly s=0s=0 or r=0r=0.

  2. (2)

    word(M)=aib𝑖wa𝑗vabr\mathrm{word}(M)=a^{i}\,\overset{i}{b}\,w\,\overset{j}{a}\,v\,a\,b^{r}, where possibly i=0i=0 or r=0r=0.

  3. (3)

    word(M)=asbub𝑖wa𝑗bnj\mathrm{word}(M)=a^{s}\,b\,u\,\overset{i}{b}\,w\,\overset{j}{a}\,b^{n-j}, where possibly s=0s=0 or j=nj=n.

  4. (4)

    word(M)=aib𝑖wa𝑗bnj\mathrm{word}(M)=a^{i}\,\overset{i}{b}\,w\,\overset{j}{a}\,b^{n-j}, where possibly i=0i=0 or j=nj=n.

Thus, if any of these four cases hold, then we have xi=bx_{i}=b and xj=ax_{j}=a.

Conversely, if we have xi=bx_{i}=b, xj=ax_{j}=a, yi=ay_{i}=a, and yj=by_{j}=b, then we can choose ii and jj as close together as possible. It follows that xixj=bwax_{i}\dots x_{j}=bwa and yiyj=awby_{i}\dots y_{j}=awb for some common, possibly empty, subword ww, otherwise we can find some ii and jj closer together. Then word(M)\mathrm{word}(M) must lie in one of the cases above and word(N)\mathrm{word}(N) must lie in one of the factor module cases in Lemma 3.13. One can then apply Lemma 3.11 and the logic runs backwards. ∎

The proof is related to the description of maps between string modules in [CB89].

3.3. Bijection for triangulations

The next result shows how extensions between indecomposable presilting complexes are encoded in the words of the corresponding internal nn-simplices.

Proposition 3.15.

Suppose that PP^{\bullet} and QQ^{\bullet} are indecomposable two-term presilting complexes over Πn\Pi_{n}, where word(P)=X=x0x1xn\mathrm{word}(P^{\bullet})=X=x_{0}x_{1}\dots x_{n} and word(Q)=Y=y0y1yn\mathrm{word}(Q^{\bullet})=Y=y_{0}y_{1}\dots y_{n}. Then Hom𝒦Πn(P,Q[1])0\operatorname{Hom}_{\mathcal{K}_{\Pi_{n}}}(P^{\bullet},Q^{\bullet}[1])\neq 0 if and only if there exist ii and jj with i<ji<j such that xi=bx_{i}=b, xj=ax_{j}=a, yi=ay_{i}=a, and yj=by_{j}=b.

Proof.

By [EJR18, Theorem 11], we have that Hom𝒦Πn(P,Q)0\operatorname{Hom}_{\mathcal{K}_{\Pi_{n}}}(P^{\bullet},Q^{\bullet})\neq 0 if and only if Hom𝒦Π¯n(P¯,Q¯)0\operatorname{Hom}_{\mathcal{K}_{\overline{\Pi}_{n}}}(\overline{P}^{\bullet},\overline{Q}^{\bullet})\neq 0, so we reason in terms of Π¯n\overline{\Pi}_{n} instead. Let (M,M)(M,M^{\prime}) and (N,N)(N,N^{\prime}) be the indecomposable τ\tau-rigid pairs over Π¯n\overline{\Pi}_{n} corresponding to P¯\overline{P}^{\bullet} and Q¯\overline{Q}^{\bullet}.

We first deal with the cases where either M=0M=0 or N=0N=0. If N=0N=0, then Q¯=(Q¯10)\overline{Q}^{\bullet}=(\overline{Q}^{-1}\to 0) and Y=bkan+1kY=b^{k}a^{n+1-k}. Then Hom𝒦Π¯n(P¯,Q¯[1])=0\operatorname{Hom}_{\mathcal{K}_{\overline{\Pi}_{n}}}(\overline{P}^{\bullet},\overline{Q}^{\bullet}[1])=0 and there can be no i<ji<j with yi=ay_{i}=a and yj=by_{j}=b, which solves this case.

If M=0M=0 and N0N\neq 0, then, by [AIR14], we have that Hom𝒦Π¯n(P¯,Q¯)0\operatorname{Hom}_{\mathcal{K}_{\overline{\Pi}_{n}}}(\overline{P}^{\bullet},\overline{Q}^{\bullet})\neq 0 if and only if HomΠ¯n(M,N)0\operatorname{Hom}_{\overline{\Pi}_{n}}(M^{\prime},N)\neq 0. We have that M=P¯kM^{\prime}=\overline{P}_{k} for some kk. Hence HomΠ¯n(M,N)0\operatorname{Hom}_{\overline{\Pi}_{n}}(M^{\prime},N)\neq 0 if and only if NN is supported at the vertex kk, which is the case if and only if there exist ik<ji\leqslant k<j such that yi=ay_{i}=a and yj=by_{j}=b. Since X=word(0,P¯k)=bkan+1kX=\mathrm{word}(0,\overline{P}_{k})=b^{k}a^{n+1-k}, we have that xi=bx_{i}=b and xj=ax_{j}=a, as desired.

We can now assume that M=0M^{\prime}=0 and N=0N^{\prime}=0, so that Hom𝒦Π¯n(P¯,Q¯)0\operatorname{Hom}_{\mathcal{K}_{\overline{\Pi}_{n}}}(\overline{P}^{\bullet},\overline{Q}^{\bullet})\neq 0 if and only if Hom𝒦Π¯n(N,τM)0\operatorname{Hom}_{\mathcal{K}_{\overline{\Pi}_{n}}}(N,\tau M)\neq 0. By Lemma 3.14, this is the case if and only if there exist ii and jj with i<ji<j such that xi=bx_{i}=b, xj=ax_{j}=a, yi=ay_{i}=a, and yj=by_{j}=b. ∎

This proposition therefore gives a simple criterion for the existence of an extension between indecomposable presilting complexes over Πn\Pi_{n} and Π¯n\overline{\Pi}_{n}.

Corollary 3.16.

Suppose that PP^{\bullet} and QQ^{\bullet} are indecomposable two-term presilting complexes over Πn\Pi_{n} corresponding to internal nn-simplices ΔX\Delta_{X} and ΔY\Delta_{Y} in Δn×Δ1\Delta_{n}\times\Delta_{1}. Then PQP^{\bullet}\oplus Q^{\bullet} is presilting if and only if ΔX\Delta_{X} and ΔY\Delta_{Y} do not intersect in their interiors.

Proof.

The circuits of Δn×Δ1\Delta_{n}\times\Delta_{1} correspond to ({ai,bj},{aj,bi})(\{a_{i},b_{j}\},\{a_{j},b_{i}\}) where iji\neq j. These are crossing diagonals in the face {ai,aj,bi,bj}\{a_{i},a_{j},b_{i},b_{j}\} of Δn×Δ1\Delta_{n}\times\Delta_{1}. Since internal nn-simplices Δx0x1xn\Delta_{x_{0}x_{1}\dots x_{n}} and Δy0y1yn\Delta_{y_{0}y_{1}\dots y_{n}} intersect in their interiors precisely if each contains one half of a circuit, we obtain that these simplices intersect if and only if there exist ii and jj such that xi=yj=ax_{i}=y_{j}=a and xj=yi=bx_{j}=y_{i}=b. The result then follows by Proposition 3.15. ∎

Corollary 3.17.

There are bijections between 𝗍𝗋𝗂(Δn×Δ1)\mathsf{tri}(\Delta_{n}\times\Delta_{1}), 𝟤-𝗌𝗂𝗅𝗍Πn\mathsf{2\mbox{-}silt}\,\Pi_{n}, and 𝗌τ-𝗍𝗂𝗅𝗍Πn\mathsf{s\tau\mbox{-}tilt}\,\Pi_{n}.

Proof.

It follows from [AIR14, Proposition 3.3] and [Aih13] that two-term silting complexes over Πn\Pi_{n} are precisely two-term presilting complexes over Πn\Pi_{n} with nn non-isomorphic indecomposable summands. Hence, by Proposition 3.2 and Proposition 3.15, we have that two-term silting complexes over Πn\Pi_{n} correspond to collections of nn internal nn-simplices in Δn×Δ1\Delta_{n}\times\Delta_{1} which do not intersect each other’s interiors. It follows from [GKZ94, Chapter 7, Proposition 3.10(a)] that triangulations of Δn×Δ1\Delta_{n}\times\Delta_{1} correspond to sets of nn internal nn-simplices which do not intersect each other’s interiors. ∎

Given a support τ\tau-tilting pair (M,P)(M,P) over Πn\Pi_{n}, we write 𝒯(M,P)\mathcal{T}(M,P) for the corresponding triangulation of Δn×Δ1\Delta_{n}\times\Delta_{1}. Likewise, we write 𝒯(P)\mathcal{T}(P^{\bullet}) for the triangulation of Δn×Δ1\Delta_{n}\times\Delta_{1} corresponding to a two-term silting complex PP^{\bullet} over Πn\Pi_{n}.

Example 3.18.

In each row of Figure 5, we show a support τ\tau-tilting pair over Π2\Pi_{2}, a two-term silting complex over Π2\Pi_{2}, a permutation in 𝔖3\mathfrak{S}_{3}, a pair of three-letter words in the alphabet {a,b}\{a,b\}, and a triangulation of Δ2×Δ1\Delta_{2}\times\Delta_{1}, all of which correspond to each other under the bijections.

Figure 5. Support τ\tau-tilting pairs, two-term silting complexes, permutations, words, and triangulations of Δ2×Δ1\Delta_{2}\times\Delta_{1}
a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}aab,abbaab,abb123123(0\tabbedCenterstack21)(0\tabbedCenterstack12)\left(0\to{\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\right)\oplus\left(0\to{\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\right)(\tabbedCenterstack21\tabbedCenterstack12, 0)\left({\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\oplus{\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}},\,0\right)a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}aab,babaab,bab213213(0\tabbedCenterstack21)(\tabbedCenterstack12\tabbedCenterstack21)\left(0\to{\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\right)\oplus\left({\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\to{\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\right)(\tabbedCenterstack212, 0)\left({\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\oplus 2,\,0\right)a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}aba,abbaba,abb132132(\tabbedCenterstack21\tabbedCenterstack12)(0\tabbedCenterstack12)\left({\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\to{\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\right)\oplus\left(0\to{\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\right)(1\tabbedCenterstack12, 0)\left(1\oplus{\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}},\,0\right)a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}baa,babbaa,bab231231(\tabbedCenterstack120)(\tabbedCenterstack12\tabbedCenterstack21)\left({\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\to 0\right)\oplus\left({\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\to{\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\right)(2,\tabbedCenterstack12)\left(2,\,{\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\right)a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}aba,bbaaba,bba312312(\tabbedCenterstack21\tabbedCenterstack12)(\tabbedCenterstack210)\left({\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\to{\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\right)\oplus\left({\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\to 0\right)(1,\tabbedCenterstack21)\left(1,\,{\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\right)a0a_{0}a1a_{1}a2a_{2}b0b_{0}b1b_{1}b2b_{2}baa,bbabaa,bba321321(\tabbedCenterstack120)(\tabbedCenterstack210)\left({\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\to 0\right)\oplus\left({\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\to 0\right)(0,\tabbedCenterstack12\tabbedCenterstack21)\left(0,\,{\tabbedCenterstack{1\phantom{2}\\ \phantom{1}2}}\oplus{\tabbedCenterstack{\phantom{1}2\\ 1\phantom{2}}}\right)
Proposition 3.19.

Under the bijection between 𝟤-𝗌𝗂𝗅𝗍Πn\mathsf{2\mbox{-}silt}\,\Pi_{n} and 𝗍𝗋𝗂(Δn×Δ1)\mathsf{tri}(\Delta_{n}\times\Delta_{1}), mutations of two-term silting complexes correspond to bistellar flips of triangulations.

Proof.

Since two-term silting complexes are related by mutation if and only if they differ by only one indecomposable summand, it follows from Corollary 3.17 that two silting complexes are related by a mutation if and only if the corresponding triangulations differ by only one codimension one internal simplex. In turn, it is then true that two triangulations 𝒯\mathcal{T} and 𝒯\mathcal{T}^{\prime} are bistellar flips of each other if and only if they differ by an internal nn-simplex. Indeed, if 𝒯\mathcal{T} and 𝒯\mathcal{T}^{\prime} are two triangulations, with 𝒮\mathcal{S} the subdivision given by their common internal nn-simplices, then 𝒮\mathcal{S} is an almost triangulation if and only if it contains n1n-1 internal nn-simplices. ∎

3.4. Compatibility with permutations

In this section, we show that our bijection between the support τ\tau-tilting pairs 𝗌τ-𝗍𝗂𝗅𝗍Πn\mathsf{s\tau\mbox{-}tilt}\,\Pi_{n} and triangulations 𝗍𝗋𝗂(Δn×Δ1)\mathsf{tri}(\Delta_{n}\times\Delta_{1}) is compatible with the existing bijections between 𝔖n+1\mathfrak{S}_{n+1} and 𝗌τ-𝗍𝗂𝗅𝗍Πn\mathsf{s\tau\mbox{-}tilt}\,\Pi_{n} and between 𝔖n+1\mathfrak{S}_{n+1} and 𝗍𝗋𝗂(Δn×Δ1)\mathsf{tri}(\Delta_{n}\times\Delta_{1}).

Proposition 3.20.

For all w𝔖n+1w\in\mathfrak{S}_{n+1}, we have that 𝒯(Iw,Pw)=𝒯w\mathcal{T}(I_{w},P_{w})=\mathcal{T}_{w}.

Proof.

We show this by induction on the length of ww. The base case is the identity permutation ee. We have that 𝒯e\mathcal{T}_{e} has (n+1)(n+1)-simplices

{{a0,,aj,bj,,bn}0jn}.\{\{a_{0},\dots,a_{j},b_{j},\dots,b_{n}\}\mid 0\leqslant j\leqslant n\}.

The internal nn-simplices of this triangulation are

{{a0,,aj,bj+1,,bn}0j<n}.\{\{a_{0},\dots,a_{j},b_{j+1},\dots,b_{n}\}\mid 0\leqslant j<n\}.

We have that (Ie,Pe)=(Πn,0)(I_{e},P_{e})=(\Pi_{n},0), and by Proposition 3.2, the word corresponding to the indecomposable projective PjP_{j} is

aajbbnj.\underbrace{a\dots a}_{j}\underbrace{b\dots b}_{n-j}.

These are the words of the internal nn-simplices of 𝒯e\mathcal{T}_{e}, so we have 𝒯e=𝒯(Ie,Pe)\mathcal{T}_{e}=\mathcal{T}(I_{e},P_{e}).

We now suppose that we have w𝔖n+1w\in\mathfrak{S}_{n+1} such that w=wsiw=w^{\prime}s_{i} for some ii, so that Iw=IwIiI_{w}=I_{w^{\prime}}I_{i}. Hence, IwI_{w} is obtained from IwI_{w^{\prime}} by removing composition factors in the top given by SiS_{i}, the simple Πn\Pi_{n}-module at the vertex ii. Given an indecomposable summand MM^{\prime} of IwI_{w^{\prime}}, we have, by Proposition 3.5 that SiS_{i} occurs in the top of MM^{\prime} if and only if xi1xi=abx_{i-1}x_{i}=ab for word(M)=x0x1xn\mathrm{word}(M^{\prime})=x_{0}x_{1}\dots x_{n}. Then, if MM is the corresponding indecomposable summand of IwI_{w}, we have that word(M)=x0x1xi2baxi+1xn\mathrm{word}(M)=x_{0}x_{1}\dots x_{i-2}bax_{i+1}\dots x_{n}.

On the other hand, considering the permutations, if we let w=k0kj1(i1)kj+1kl1ikl+1knw^{\prime}=k_{0}\dots k_{j-1}(i-1)k_{j+1}\dots k_{l-1}ik_{l+1}\dots k_{n}, then w=k0kj1ikj+1kl1(i1)kl+1knw=k_{0}\dots k_{j-1}ik_{j+1}\dots k_{l-1}(i-1)k_{l+1}\dots k_{n}. Note that, by assumption, we have j<lj<l, since the length of ww is greater than the length of ww^{\prime}. Using the description of 𝒯w\mathcal{T}_{w} and 𝒯w\mathcal{T}_{w^{\prime}} from Section 2.2.3, we have that if X=x0x1xnX=x_{0}x_{1}\dots x_{n} is the word of a simplex in 𝒯w\mathcal{T}_{w^{\prime}}, then ΔX\Delta_{X} is a simplex of 𝒯w\mathcal{T}_{w} too if and only if xi1xi=aax_{i-1}x_{i}=aa or xi1xi=bbx_{i-1}x_{i}=bb. Furthermore, if xi1xi=abx_{i-1}x_{i}=ab, then x0xi2baxi+1xnx_{0}\dots x_{i-2}bax_{i+1}\dots x_{n} is the word of a simplex in 𝒯w\mathcal{T}_{w}. The case xi1xi=bax_{i-1}x_{i}=ba is not possible since i1i-1 precedes ii in ww. Comparing with the previous paragraph, we see that the summands of (Iw,Pw)(I_{w^{\prime}},P_{w^{\prime}}) which change to give (Iw,Pw)(I_{w},P_{w}) correspond to the internal nn-simplices of 𝒯w\mathcal{T}_{w^{\prime}} which change to give 𝒯w\mathcal{T}_{w}. Moreover, the change in the words of the simplices corresponds precisely to the change in the upper contours of the summands of the τ\tau-tilting pair. Hence, we obtain that 𝒯(Iw,Pw)=𝒯w\mathcal{T}(I_{w},P_{w})=\mathcal{T}_{w} and the result follows by induction. ∎

Remark 3.21.

Note that our results therefore also give a different way of obtaining the support τ\tau-tilting pair over Πn\Pi_{n} corresponding to a permutation in 𝔖n+1\mathfrak{S}_{n+1} to the description from [Miz14]. Namely, given a permutation, one uses the description of the corresponding triangulation of Δn×Δ1\Delta_{n}\times\Delta_{1} from Section 2.2.3 to obtain a set of words in 𝗌𝖾𝗊n+1(a,b)\mathsf{seq}_{n+1}(a,b). One then uses Proposition 3.2 to translate this into a support τ\tau-tilting pair over Πn\Pi_{n} by using these words to give the upper contours of the indecomposable τ\tau-rigid summands. This does not require a reduced expression for the permutation, whereas the description from [Miz14] does.

References

  • [Aih13] Takuma Aihara. Tilting-connected symmetric algebras. Algebr. Represent. Theory, 16(3):873–894, 2013.
  • [AIR14] Takahide Adachi, Osamu Iyama, and Idun Reiten. τ\tau-tilting theory. Compos. Math., 150(3):415–452, 2014.
  • [Ami09] Claire Amiot. Cluster categories for algebras of global dimension 2 and quivers with potential. Ann. Inst. Fourier (Grenoble), 59(6):2525–2590, 2009.
  • [BB98] Eric K. Babson and Louis J. Billera. The geometry of products of minors. Discrete Comput. Geom., 20(2):231–249, 1998.
  • [BCSo21] Karin Baur and Raquel Coelho Simões. A geometric model for the module category of a gentle algebra. Int. Math. Res. Not. IMRN, (15):11357–11392, 2021.
  • [BCZ19] Emily Barnard, Andrew Carroll, and Shijie Zhu. Minimal inclusions of torsion classes. Algebr. Comb., 2(5):879–901, 2019.
  • [BIRS09] A. B. Buan, O. Iyama, I. Reiten, and J. Scott. Cluster structures for 2-Calabi-Yau categories and unipotent groups. Compos. Math., 145(4):1035–1079, 2009.
  • [BMR+06] Aslak Bakke Buan, Bethany Marsh, Markus Reineke, Idun Reiten, and Gordana Todorov. Tilting theory and cluster combinatorics. Adv. Math., 204(2):572–618, 2006.
  • [BR87] Michael C. R. Butler and Claus Michael Ringel. Auslander–Reiten sequences with few middle terms and applications to string algebras. Comm. Algebra, 15(1-2):145–179, 1987.
  • [BZ11] Thomas Brüstle and Jie Zhang. On the cluster category of a marked surface without punctures. Algebra Number Theory, 5(4):529–566, 2011.
  • [CB89] W. W. Crawley-Boevey. Maps between representations of zero-relation algebras. J. Algebra, 126(2):259–263, 1989.
  • [CBH98] William Crawley-Boevey and Martin P. Holland. Noncommutative deformations of Kleinian singularities. Duke Math. J., 92(3):605–635, 1998.
  • [CS20] Wen Chang and Sibylle Schroll. A geometric realization of silting theory for gentle algebras. arXiv preprint arXiv:2012.12663, 2020.
  • [DIJ19] Laurent Demonet, Osamu Iyama, and Gustavo Jasso. τ\tau-tilting finite algebras, bricks, and gg-vectors. Int. Math. Res. Not. IMRN, (3):852–892, 2019.
  • [DIR+17] Laurent Demonet, Osamu Iyama, Nathan Reading, Idun Reiten, and Hugh Thomas. Lattice theory of torsion classes. arXiv preprint arXiv:1711.01785, 2017.
  • [DLKOS09] Jesús A. De Loera, Edward D. Kim, Shmuel Onn, and Francisco Santos. Graphs of transportation polytopes. J. Combin. Theory Ser. A, 116(8):1306–1325, 2009.
  • [DLRS10] Jesús A. De Loera, Jörg Rambau, and Francisco Santos. Triangulations, volume 25 of Algorithms and Computation in Mathematics. Springer-Verlag, Berlin, 2010. Structures for algorithms and applications.
  • [DS04] Mike Develin and Bernd Sturmfels. Tropical convexity. Doc. Math., 9:1–27, 2004.
  • [EJR18] Florian Eisele, Geoffrey Janssens, and Theo Raedschelders. A reduction theorem for τ\tau-rigid modules. Math. Z., 290(3-4):1377–1413, 2018.
  • [FG19] Changjian Fu and Shengfei Geng. Tilting modules and support τ\tau-tilting modules over preprojective algebras associated with symmetrizable Cartan matrices. Algebr. Represent. Theory, 22(5):1239–1260, 2019.
  • [FST08] Sergey Fomin, Michael Shapiro, and Dylan Thurston. Cluster algebras and triangulated surfaces. I. Cluster complexes. Acta Math., 201(1):83–146, 2008.
  • [FZ02] Sergey Fomin and Andrei Zelevinsky. Cluster algebras. I. Foundations. J. Amer. Math. Soc., 15(2):497–529 (electronic), 2002.
  • [GfP79] I. M. Gel\cprime fand and V. A. Ponomarev. Model algebras and representations of graphs. Funktsional. Anal. i Prilozhen., 13(3):1–12, 1979.
  • [GKZ94] Israel M. Gel\cprimefand, Mikhail M. Kapranov, and Andrei V. Zelevinsky. Discriminants, resultants, and multidimensional determinants. Mathematics: Theory & Applications. Birkhäuser Boston, Inc., Boston, MA, 1994.
  • [GLS06] Christof Geiß, Bernard Leclerc, and Jan Schröer. Rigid modules over preprojective algebras. Invent. Math., 165(3):589–632, 2006.
  • [GLS07a] Christof Geiss, Bernard Leclerc, and Jan Schröer. Auslander algebras and initial seeds for cluster algebras. J. Lond. Math. Soc. (2), 75(3):718–740, 2007.
  • [GLS07b] Christof Geiss, Bernard Leclerc, and Jan Schröer. Semicanonical bases and preprojective algebras. II. A multiplication formula. Compos. Math., 143(5):1313–1334, 2007.
  • [GLS08] Christof Geiss, Bernard Leclerc, and Jan Schröer. Partial flag varieties and preprojective algebras. Ann. Inst. Fourier (Grenoble), 58(3):825–876, 2008.
  • [GLS17] Christof Geiss, Bernard Leclerc, and Jan Schröer. Quivers with relations for symmetrizable Cartan matrices I: Foundations. Invent. Math., 209(1):61–158, 2017.
  • [IRRT18] Osamu Iyama, Nathan Reading, Idun Reiten, and Hugh Thomas. Lattice structure of Weyl groups via representation theory of preprojective algebras. Compos. Math., 154(6):1269–1305, 2018.
  • [KS97] Masaki Kashiwara and Yoshihisa Saito. Geometric construction of crystal bases. Duke Math. J., 89(1):9–36, 1997.
  • [Miz14] Yuya Mizuno. Classifying τ\tau-tilting modules over preprojective algebras of Dynkin type. Math. Z., 277(3-4):665–690, 2014.
  • [MRZ03] Bethany Marsh, Markus Reineke, and Andrei Zelevinsky. Generalized associahedra via quiver representations. Trans. Amer. Math. Soc., 355(10):4171–4186, 2003.
  • [Mur22] Kota Murakami. On the module categories of generalized preprojective algebras of Dynkin type. Osaka J. Math., 59(2):387–402, 2022.
  • [Nak94] Hiraku Nakajima. Instantons on ALE spaces, quiver varieties, and Kac-Moody algebras. Duke Math. J., 76(2):365–416, 1994.
  • [OPS18] Sebastian Opper, Pierre-Guy Plamondon, and Sibylle Schroll. A geometric model for the derived category of gentle algebras. arXiv preprint arXiv:1801.09659, 2018.
  • [OT12] Steffen Oppermann and Hugh Thomas. Higher-dimensional cluster combinatorics and representation theory. J. Eur. Math. Soc. (JEMS), 14(6):1679–1737, 2012.
  • [ST11] Alistair Savage and Peter Tingley. Quiver Grassmannians, quiver varieties and the preprojective algebra. Pacific J. Math., 251(2):393–429, 2011.
  • [Stu91] Bernd Sturmfels. Gröbner bases of toric varieties. Tohoku Math. J. (2), 43(2):249–261, 1991.
  • [vdLT82] Gerard van der Laan and Adolphus J. J. Talman. On the computation of fixed points in the product space of unit simplices and an application to noncooperative NN person games. Math. Oper. Res., 7(1):1–13, 1982.
  • [Wil22] Nicholas J. Williams. New interpretations of the higher Stasheff–Tamari orders. Adv. Math., 407:108552, 2022.