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

Pop-Stack Operators for Torsion Classes and Cambrian Lattices

Emily Barnard Department of Mathematical Sciences, DePaul University, Chicago, IL 60604, USA e.barnard@depaul.edu Colin Defant Department of Mathematics, Harvard University, Cambridge, MA 02139, USA colindefant@gmail.com  and  Eric J. Hanson Department of Mathematics, North Carolina State University, Raleigh, NC 27695, USA ejhanso3@ncsu.edu
Abstract.

The pop-stack operator of a finite lattice LL is the map popL:LL\mathrm{pop}^{\downarrow}_{L}\colon L\to L that sends each element xLx\in L to the meet of {x}covL(x)\{x\}\cup\text{cov}_{L}(x), where covL(x)\text{cov}_{L}(x) is the set of elements covered by xx in LL. We study several properties of the pop-stack operator of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda, the lattice of torsion classes of a τ\tau-tilting finite algebra Λ\Lambda over a field KK. We describe the pop-stack operator in terms of certain mutations of 2-term simple-minded collections. This allows us to describe preimages of a given torsion class under the pop-stack operator.

We then specialize our attention to Cambrian lattices of a finite irreducible Coxeter group WW. Using tools from representation theory, we provide simple Coxeter-theoretic and lattice-theoretic descriptions of the image of the pop-stack operator of a Cambrian lattice (which can be stated without representation theory). When specialized to a bipartite Cambrian lattice of type A, this result settles a conjecture of Choi and Sun. We also settle a related enumerative conjecture of Defant and Williams. When LL is an arbitrary lattice quotient of the weak order on WW, we prove that the maximum size of a forward orbit under the pop-stack operator of LL is at most the Coxeter number of WW; when LL is a Cambrian lattice, we provide an explicit construction to show that this maximum forward orbit size is actually equal to the Coxeter number.

1. Introduction

1.1. Pop-stack

Let LL be a finite lattice with meet operation \wedge and join operation \vee. The pop-stack operator popL:LL\mathrm{pop}^{\downarrow}_{L}\colon L\to L and the dual pop-stack operator popL:LL\mathrm{pop}^{\uparrow}_{L}\colon L\to L are defined by

popL(x)=x({yy<x})andpopL(x)=x({yx<y}),\mathrm{pop}^{\downarrow}_{L}(x)=x\wedge\left(\bigwedge\{y\mid y{\,\,<\!\!\!\!\cdot\,\,\,}x\}\right)\quad\text{and}\quad\mathrm{pop}^{\uparrow}_{L}(x)=x\vee\left(\bigvee\{y\mid x{\,\,<\!\!\!\!\cdot\,\,\,}y\}\right),

where we write uvu\lessdot v to mean that uu is covered by vv in LL. These operators have appeared in various contexts; they serve as both useful tools [AP22, BH, Eno23, ES22, Hana, Müh19, Rea11, Sak] and objects of interest in their own right [CS, CG19, CGP21, Def22b, Def22a, DW23, Hon22, PS19, Ung82]. When the lattice LL is understood, we will omit subscripts and simply denote these operators by pop\mathrm{pop}^{\downarrow} and pop\mathrm{pop}^{\uparrow}.

In [CG19, Def22b, Def22a, Hon22, PS19, Ung82], the pop-stack operator has been considered as a dynamical system. Given a map f:LLf\colon L\to L and an element xLx\in L, the forward orbit of xx under ff is the set

Orbf(x)={x,f(x),f2(x),f3(x),},\mathrm{Orb}_{f}(x)=\{x,f(x),f^{2}(x),f^{3}(x),\ldots\},

where ftf^{t} is the map obtained by composing ff with itself tt times. To ease notation, let us write

𝒪L(x)=OrbpopL(x)\mathcal{O}_{L}(x)=\mathrm{Orb}_{\mathrm{pop}^{\downarrow}_{L}}(x)

for the forward orbit of xx under popL\mathrm{pop}^{\downarrow}_{L}. If tt is sufficiently large, then (popL)t(x)(\mathrm{pop}^{\downarrow}_{L})^{t}(x) is equal to the minimal element 0^\hat{0} of LL (which is a fixed point of popL\mathrm{pop}^{\downarrow}_{L}). Thus, |𝒪L(x)|1\lvert\mathcal{O}_{L}(x)\rvert-1 is equal to the number of iterations of popL\mathrm{pop}^{\downarrow}_{L} needed to send xx to 0^\hat{0}. Given an interesting lattice LL, one of the primary problems one can consider about its pop-stack operator is that of computing

maxxL|𝒪L(x)|.\max_{x\in L}|\mathcal{O}_{L}(x)|.

For a fixed tt, one can also consider the tt-pop-stack sortable elements of LL, which are the elements xLx\in L such that (popL)t(x)=0^(\mathrm{pop}^{\downarrow}_{L})^{t}(x)=\hat{0}.

Defant and Williams [DW23] also found that it is fruitful to study the image of the pop-stack operator when LL is a semidistrim lattice; this is because the image of pop\mathrm{pop}^{\downarrow} has numerous interesting properties, some of which relate to a certain bijective rowmotion operator row:LL\mathrm{row}\colon L\to L. For example, |pop(L)|\lvert\mathrm{pop}^{\downarrow}(L)\rvert and |pop(L)|\lvert\mathrm{pop}^{\uparrow}(L)\rvert are both equal to the number of elements xLx\in L such that row(x)x\mathrm{row}(x)\leq x. The images of pop\mathrm{pop}^{\downarrow} and pop\mathrm{pop}^{\uparrow} are also naturally in bijection with the set of facets of a certain simplicial complex called the canonical join complex of LL.

1.2. Lattices of torsion classes

In this article, we take a representation-theoretic perspective and consider a finite-dimensional basic algebra Λ\Lambda over a field KK. The set of torsion classes (see Section 3.2 for the definition) of finitely-generated (right) Λ\Lambda-modules forms a complete lattice [IRTT15] that we denote by 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda. Examples of lattices arising in this way include the weak order and Cambrian lattices associated to any finite crystallographic Coxeter group [AHI+, Miz14, IT09]. This paper continues a rich tradition of using lattices of torsion classes as a tool for proving new results in both representation theory and in algebraic combinatorics; see, e.g., [DIR+23, Eno23, IRRT18, RST21, TW19b, TW19a] and many others.

We focus our attention on the case when Λ\Lambda is τ\tau-tilting finite (in the sense of [DIJ17]); this is equivalent to assuming 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda is finite. Our goal is to study the image and the dynamical properties of the pop-stack operator of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda. While pop\mathrm{pop}^{\downarrow} has already appeared (sometimes under different names) in the theory of lattices of torsion classes [AP22, BH, Eno23, ES22, Hana, Sak], it has primarily been used as a tool rather than a dynamical operator worthy of its own investigation. Let us remark that the article [BTZ21] initiated the study of dynamical combinatorics of torsion classes by considering rowmotion.

In Section 4, we consider certain pairs of sets of modules called 2-term simple-minded collections and semibrick pairs. 2-term simple-minded collections were first introduced in the special case when Λ\Lambda is a symmetric algebra in [Ric02]; they are special cases of the more general simple-minded collections introduced in [AN09] (under the name spherical collections). Semibrick pairs are a generalization of 2-term simple-minded collections that were introduced in [HI21b] as a tool for studying a generalization of the picture group of [ITW]. As we recall in Section 3.3, there is a bijection between 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda (with Λ\Lambda τ\tau-tilting finite) and the set 2-𝗌𝗆𝖼Λ\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda of 2-term simple-minded collections [Asa20]. By the results of [BCZ19], this bijection encodes information about cover relations and canonical join representations in 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda.

The set 2-𝗌𝗆𝖼Λ\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda also comes equipped with a set of mutation operators that categorify the mutation theory of cluster algebras. See [BY13, Section 3.7] and [KY14, Section 7.2] (special cases also appeared earlier in [KS, Section 8.1] and [KQ15, Section 3]). These mutation formulas can also be applied to many of the more general semibrick pairs; see [BH22, HI21a]. Moreover, under the bijection between 2-𝗌𝗆𝖼Λ\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda and 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda, the mutation operators give a representation-theoretic formula for how the corresponding 2-term simple-minded collection changes when one traverses a cover relation in 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda.

For each semibrick pair (𝒳,𝒴)(\mathcal{X},\mathcal{Y}), there is classically one mutation operator for each module in 𝒳𝒴\mathcal{X}\cup\mathcal{Y}. In the present paper, we more generally define a mutation operator for each nonempty subset 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X} or 𝒴𝒴\mathcal{Y}^{\prime}\subseteq\mathcal{Y}. Our first main result (Theorem 5.1) states that, under the bijection between 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda and 2-𝗌𝗆𝖼Λ\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda, applying the pop-stack operator and its dual to a torsion class corresponds to performing certain mutations on the associated 2-term simple-minded collections.

In Section 5.2, we characterize the preimages of a prescribed torsion class under pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda} and pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}. As corollaries, we obtain descriptions of the 11-pop-stack sortable and the 22-pop-stack sortable elements of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda (Corollaries 5.4 and 5.5).

1.3. Cambrian lattices

Let WW be a finite irreducible Coxeter group, and let Weak(W)\mathrm{Weak}(W) denote the (right) weak order on WW. Given a Coxeter element cc of WW, one can construct the cc-Cambrian lattice Cambc\mathrm{Camb}_{c}, which is the sublattice of Weak(W)\mathrm{Weak}(W) consisting of Reading’s cc-sortable elements [Rea06, Rea07]. (The cc-Cambrian lattice is also a lattice quotient of Weak(W)\mathrm{Weak}(W).) When WW is crystallographic, we can realize Cambc\mathrm{Camb}_{c} as the lattice of torsion classes of the tensor algebra KQcKQ_{c} associated to the weighted Dynkin quiver associated to cc [IT09].

Two very special Cambrian lattices are the Tamari lattice and the type-B Tamari lattice; the images of the pop-stack operators on these lattices were characterized and enumerated in [Hon22] and [CS], respectively. On the other hand, analogous results for the Cambrian lattice Cambc×\mathrm{Camb}_{c^{\times}} associated to a type-A bipartite Coxeter element c×c^{\times} have remained elusive: a conjectural enumeration of the image was formulated by Defant and Williams [DW23], while a characterization of the image was conjectured by Choi and Sun [CS].

It turns out that our representation-theoretic perspective is quite useful for understanding the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}. In Theorem 5.10, we provide a representation-theoretic characterization of the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda} whenever Λ\Lambda is hereditary. When the Coxeter group WW is crystallographic, the tensor algebra KQcKQ_{c} is hereditary, and we can reformulate our description of the image of the pop-stack operator in purely lattice-theoretic and Coxeter-theoretic terms. We then check directly (by hand and by computer) that these lattice-theoretic and Coxeter-theoretic descriptions still holds when WW is not crystallographic.

Our Coxeter-theoretic description of the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} is surprisingly simple: we will prove (Theorem 7.8) that a cc-sortable element wCambcw\in\mathrm{Camb}_{c} is in the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} if and only if the right descents of ww all commute and ww has no left inversions in common with c1c^{-1}.

To state our purely lattice-theoretic description of the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}, let us write s1,,sns_{1},\ldots,s_{n} for the simple reflections of WW; these are the atoms of Cambc\mathrm{Camb}_{c}. For 1in1\leq i\leq n, let pip_{i} be the unique maximal element of the set

{xCambcsix and sjx for all sjS{si}}.\{x\in\mathrm{Camb}_{c}\mid s_{i}\leq x\text{ and }s_{j}\not\leq x\text{ for all }s_{j}\in S\setminus\{s_{i}\}\}.

We will prove that an element wCambcw\in\mathrm{Camb}_{c} is in the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} if and only if the interval [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] (in Cambc\mathrm{Camb}_{c}) is Boolean and piwp_{i}\not\leq w for all 1in1\leq i\leq n. When WW is crystallographic, the elements p1,,pnp_{1},\ldots,p_{n} correspond in a natural way to the projective indecomposable modules of the tensor algebra KQcKQ_{c}; this is one reason why our representation-theoretic perspective was so useful for discovering and proving our description of the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}.

Along the way to proving our characterization of the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}, we prove (as part of Theorem 7.6) that for wCambcw\in\mathrm{Camb}_{c}, the right descents of ww all commute with each other if and only if the interval [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] of Cambc\mathrm{Camb}_{c} is equal to the interval [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] of Weak(W)\mathrm{Weak}(W). By combining this surprising result with our characterization of the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}, we obtain an equally surprising dynamical corollary (Corollary 7.10); namely, for wCambcw\in\mathrm{Camb}_{c}, we have

(1.1) (popWeak(W))t(popCambc(w))=(popCambc)t+1(w)(\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)})^{t}(\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w))=(\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}})^{t+1}(w)

for all t0t\geq 0.

When WW is of type AnA_{n} and c=c×c=c^{\times} is a bipartite Coxeter element, our description of the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} resolves the aforementioned conjecture of Choi and Sun [CS]. We will also construct a bijection from the image of popCambc×\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c^{\times}}} to a certain set of Motzkin paths (Theorem 8.10); this allows us to resolve the aforementioned enumerative conjecture of Defant and Williams [DW23]. This result, the bijective proof of which utilizes the combinatorics of arc diagrams, provides an exact enumeration of the facets of the canonical join complex of a bipartite type-A Cambrian lattice.

Finally, we turn our attention to forward orbits under popL\mathrm{pop}^{\downarrow}_{L} when LL is a lattice quotient of the weak order on a finite irreducible Coxeter group WW. In this setting, we first show that maxxL|𝒪L(x)|h\max_{x\in L}\lvert\mathcal{O}_{L}(x)\rvert\leq h, where hh is the Coxeter number of WW. We then prove that this inequality is actually an equality when L=CambcL=\mathrm{Camb}_{c} for some Coxeter element cc of WW. To do so, we utilize the combinatorial properties of the cc-sorting word for the long element of WW to construct an element 𝐳cCambc\mathbf{z}_{c}\in\mathrm{Camb}_{c} such that |𝒪Cambc(𝐳c)|=h\lvert\mathcal{O}_{\mathrm{Camb}_{c}}(\mathbf{z}_{c})\rvert=h.

1.4. Organization

Sections 2 and 3 provide necessary background on posets and representation theory, respectively. In Section 4, we discuss the theory of mutation of semibrick pairs, and we extend the theory to allow mutation at multiple bricks simultaneously; the proof of one of the main results of this section (Theorem 4.6) is postponed until Appendix A. Section 5 provides a representation-theoretic description of the pop-stack operator on 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda, describes preimages of a torsion class under pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}, and characterizes the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}. Beginning in Section 6, we fixate on Cambrian lattices. Section 6 provides background on Coxeter groups, root systems, and Cambrian lattices (including their realizations as lattices of torsion classes). Section 7 is devoted to characterizing the image of the pop-stack operator on an arbitrary Cambrian lattice and deducing Equation 1.1. In Section 8, we provide explicit combinatorial characterizations and enumerations of the images of pop-stack operators on Cambrian lattices of type A. In Section 9, we prove our results concerning the maximum sizes of forward orbits under the pop-stack operators on lattice quotients of the weak order. Section 10 collects several ideas for future work.

2. Posets and Lattices

Let PP be a finite poset. For x,yPx,y\in P with xyx\leq y, the interval between xx and yy is the induced subposet [x,y]={zPxzy}[x,y]=\{z\in P\mid x\leq z\leq y\} of PP. We say yy covers xx and write xyx\lessdot y if x<yx<y and [x,y]={x,y}[x,y]=\{x,y\}. The Hasse diagram of PP is the graph with vertex set PP in which xx and yy are adjacent whenever xyx\lessdot y or yxy\lessdot x. We say PP is connected if its Hasse diagram is connected. A rank function on PP is a function rank:P\mathrm{rank}\colon P\to\mathbb{Z} such that rank(y)=rank(x)+1\mathrm{rank}(y)=\mathrm{rank}(x)+1 whenever xyx\lessdot y. We say PP is ranked if there exists a rank function on PP. An antichain is a poset PP in which any two distinct elements are incomparable (that is, if x,yPx,y\in P are such that xyx\leq y, then x=yx=y). The dual of PP is the poset PP^{*} with the same underlying set as PP defined so that xyx\leq y in PP^{*} if and only if yxy\leq x in PP. A subset XX of PP is called convex if for all x,yXx,y\in X and all zPz\in P with xzyx\leq z\leq y, we have zXz\in X. An order ideal is a subset II of PP such that if xPx\in P and yIy\in I satisfy xyx\leq y, then xIx\in I. An upper set of PP is a subset UPU\subseteq P such that PUP\setminus U is an order ideal. We write J(P)J(P) for the set of order ideals of PP, ordered by inclusion.

A lattice is a poset LL such that any two elements x,yLx,y\in L have a greatest lower bound, which is called their meet and denoted by xyx\wedge y, and a least upper bound, which is called their join and denoted by xyx\vee y. A finite lattice is distributive if it is isomorphic to J(P)J(P) for some finite poset PP. A lattice is Boolean if it is isomorphic to the power set of a set, ordered by inclusion. Hence, if PP is an antichain, then J(P)J(P) is Boolean.

A lattice LL is complete if every (possibly infinite) subset has a meet (i.e., greatest lower bound) and a join (i.e., least upper bound). We write X\bigwedge X and X\bigvee X for the meet and join, respectively, of a subset XX of a complete lattice. Given lattices LL and LL^{\prime}, a lattice homomorphism is a map ϕ:LL\phi\colon L\to L^{\prime} such that ϕ(xy)=ϕ(x)ϕ(y)\phi(x\wedge y)=\phi(x)\wedge\phi(y) and ϕ(xy)=ϕ(x)ϕ(y)\phi(x\vee y)=\phi(x)\vee\phi(y) for all x,yLx,y\in L. We say LL^{\prime} is a lattice quotient if there is a surjective lattice homomorphism from LL to LL^{\prime}.

Assume LL is a finite lattice. Then LL has a unique minimal element 0^=L\hat{0}=\bigwedge L and a unique maximal element 1^=L\hat{1}=\bigvee L. An element of LL that covers 0^\hat{0} is called an atom, while an element covered by 1^\hat{1} is called a coatom. An element jLj\in L is called join-irreducible if there does not exist a set XL{j}X\subseteq L\setminus\{j\} such that j=Xj=\bigvee X. Equivalently, jj is join-irreducible if it covers exactly one element of LL. (Note that 0^\hat{0} is not join-irreducible because it is equal to \bigvee\emptyset.) Dually, an element mLm\in L is meet-irreducible if there does not exist a set XL{m}X\subseteq L\setminus\{m\} such that j=Xj=\bigwedge X. Equivalently, mm is meet-irreducible if it is covered by exactly one element of LL. (Note that 1^=\hat{1}=\bigwedge\emptyset is not meet-irreducible.) Let JIrrL\mathrm{JIrr}_{L} (resp. MIrrL\mathrm{MIrr}_{L}) be the set of join-irreducible (resp. meet-irreducible) elements of LL. For jJIrrLj\in\mathrm{JIrr}_{L} and mMIrrm\in\mathrm{MIrr}, let jj_{*} be the unique element covered by jj, and let mm^{*} be the unique element that covers mm. A set ALA\subseteq L is join-irredundant (resp. meet-irredundant) if A<A\bigvee A^{\prime}<\bigvee A (resp. A>A\bigwedge A^{\prime}>\bigwedge A) for every proper subset AA^{\prime} of AA. The canonical join representation of an element xLx\in L (if it exists) is the unique join-irredundant set AJIrrLA\subseteq\mathrm{JIrr}_{L} satisfying the following:

  • x=Ax=\bigvee A.

  • For every join-irredundant set BJIrrLB\subseteq\mathrm{JIrr}_{L} such that x=Bx=\bigvee B, there exist aAa\in A and bBb\in B such that aba\leq b.

Dually, the canonical meet representation of xx (if it exists) is the unique meet-irredundant set AMIrrLA\subseteq\mathrm{MIrr}_{L} satisfying the following:

  • x=Ax=\bigwedge A.

  • For every meet-irredundant set BMIrrLB\subseteq\mathrm{MIrr}_{L} such that x=Bx=\bigvee B, there exist aAa\in A and bBb\in B such that aba\geq b.

We say a lattice LL is semidistributive if for all x,y,zLx,y,z\in L, we have

xy=xzxy=x(yz)andxy=xzxy=x(yz).x\wedge y=x\wedge z\implies x\wedge y=x\wedge(y\vee z)\quad\text{and}\quad x\vee y=x\vee z\implies x\vee y=x\vee(y\wedge z).

Suppose LL is finite and semidistributive. It is known that every element vv of LL has a canonical join representation 𝒟(v)\mathcal{D}(v) and a canonical meet representation 𝒰(v)\mathcal{U}(v); in fact, the existence of both representations for every vLv\in L is equivalent to semidistributivity (see [FJN95, Theorem 2.24]). Moreover, the collection of canonical join representations (resp. canonical meet representations) of elements of LL forms a simplicial complex called the canonical join complex (resp. canonical meet complex) of LL. There is a unique bijection κ:JIrrLMIrrL\kappa\colon\mathrm{JIrr}_{L}\to\mathrm{MIrr}_{L} such that κ(j)j=j\kappa(j)\wedge j=j_{*} and κ(j)j=(κ(j))\kappa(j)\vee j=(\kappa(j))^{*} for all jJIrrLj\in\mathrm{JIrr}_{L}. The map Aκ(A)A\mapsto\kappa(A) is an isomorphism from canonical join complex of LL to the canonical meet complex of LL [Bar19, Corollary 5]. Moreover, the number of facets in each of these simplicial complexes is equal to both |popL(L)|\lvert\mathrm{pop}^{\downarrow}_{L}(L)\rvert and |popL(L)|\lvert\mathrm{pop}^{\uparrow}_{L}(L)\rvert by [DW23, Theorem 9.13]. Indeed, the facets of the canonical join complex (resp. canonical meet complex) of LL are precisely the canonical meet representations (resp. canonical join representations) of the elements of popL(L)\mathrm{pop}^{\downarrow}_{L}(L) (resp. popL(L)\mathrm{pop}^{\uparrow}_{L}(L)). Let 𝐏L(q)\mathbf{P}_{L}(q) be the generating function that counts the facets of the canonical join complex (equivalently, the canonical meet complex) according to their sizes. Then

(2.1) 𝐏L(q)=vpopL(L)q|𝒰(v)|=vpopL(L)q|𝒟(v)|.\mathbf{P}_{L}(q)=\sum_{v\in\mathrm{pop}^{\downarrow}_{L}(L)}q^{|\mathcal{U}(v)|}=\sum_{v\in\mathrm{pop}^{\uparrow}_{L}(L)}q^{|\mathcal{D}(v)|}.

The canonical join complex of LL is equal to the canonical meet complex of the dual lattice LL^{*}, so

𝐏L(q)=𝐏L(q).\mathbf{P}_{L}(q)=\mathbf{P}_{L^{*}}(q).

The Galois graph of a finite semidistributive lattice LL is the loopless directed graph with vertex set JIrrL\mathrm{JIrr}_{L} in which there is an arrow jjj\to j^{\prime} if and only if jjj\neq j^{\prime} and jκ(j)j\not\leq\kappa(j^{\prime}). For each edge xyx\lessdot y in the Hasse diagram of LL, there is a unique join-irreducible element jx,yJIrrLj_{x,y}\in\mathrm{JIrr}_{L} such that jx,yyj_{x,y}\leq y and κ(jx,y)x\kappa(j_{x,y})\geq x (see [Bar19, Proposition 2.2 & Lemma 3.3]). We call jx,yj_{x,y} the shard label of the edge xyx\lessdot y. The canonical join representation and canonical meet representation of an element vLv\in L are given by 𝒟(v)={jx,v:xv}\mathcal{D}(v)=\{j_{x,v}:x\lessdot v\} and 𝒰(v)={κ(jv,y):vy}\mathcal{U}(v)=\{\kappa(j_{v,y}):v\lessdot y\}. Moreover, the canonical join complex of LL is just the collection of independent sets of the Galois graph of LL.

Note that intervals in semidistributive lattices are semidistributive. The following lemma was stated in [DW23] for semidistrim lattices, which are more general than semidistributive lattices, but we only need to consider semidistributive lattices. (See also [RST21, Section 4].)

Lemma 2.1 ([DW23, Corollary 7.10]).

Let LL be a finite semidistributive lattice, and let [u,v][u,v] be an interval in LL. There is a bijection from {jJIrrLjv and κ(j)u}\{j\in\mathrm{JIrr}_{L}\mid j\leq v\text{ and }\kappa(j)\geq u\} to JIrr[u,v]\mathrm{JIrr}_{[u,v]} given by jujj\mapsto u\vee j. This bijection is an isomorphism from an induced subgraph of the Galois graph of LL to the Galois graph of [u,v][u,v]. If xyx\lessdot y is an edge in [u,v][u,v] whose shard label in LL is jx,yj_{x,y}, then the shard label of xyx\lessdot y in [u,v][u,v] is ujx,yu\vee j_{x,y}.

Example 2.2.

Let PP be a finite antichain, and consider the Boolean lattice J(P)J(P). The join-irreducible elements of J(P)J(P) are the singleton subsets of PP. If ABA\lessdot B is an edge in J(P)J(P), then there exists bBb\in B such that A=B{b}A=B\setminus\{b\}. The shard label of ABA\lessdot B is jA,B={b}j_{A,B}=\{b\}. The Galois graph of J(P)J(P) has no edges.

Example 2.3.

Let LL be the lattice obtained by adding a minimal element 0^\hat{0} and a maximal element 1^\hat{1} to a disjoint union of two chains a1am1a_{1}\lessdot\cdots\lessdot a_{m-1} and ama2m2a_{m}\lessdot\cdots\lessdot a_{2m-2} (so a1a_{1} and ama_{m} are the atoms of LL, while am1a_{m-1} and a2m2a_{2m-2} are the coatoms). This lattice is isomorphic to the weak order on the dihedral group I2(m)I_{2}(m) (see Section 6). We have JIrrL=MIrrL={a1,,a2m2}\mathrm{JIrr}_{L}=\mathrm{MIrr}_{L}=\{a_{1},\ldots,a_{2m-2}\}, and the bijection κ\kappa is given by κ(ai)=ai1\kappa(a_{i})=a_{i-1}, where we let a0=a2m2a_{0}=a_{2m-2}. For 1im11\leq i\leq m-1, there is an arrow from aia_{i} to aia_{i^{\prime}} in the Galois graph of LL if and only if 1ii11\leq i^{\prime}\leq i-1 or m+1i2m2m+1\leq i^{\prime}\leq 2m-2. For mi2m2m\leq i\leq 2m-2, there is an arrow from aia_{i} to aia_{i^{\prime}} in the Galois graph of LL if and only if 2im12\leq i^{\prime}\leq m-1 or mii1m\leq i^{\prime}\leq i-1. See Figure 1.

Refer to caption
Figure 1. The Hasse diagram (left) and Galois graph (right) of the lattice LL from Example 2.3 with m=5m=5. The red arrows on the left indicate the bijection κ\kappa.

3. Finite-Dimensional Algebras

In this section, we recall background information on the representation theory of quivers and finite-dimensional algebras. We refer to the textbooks [ARS95, ASS06, Sch14] as standard background references.

Let Λ\Lambda be a finite-dimensional basic111Recall that a module MM is basic if there does not exist a nonzero module NN such that NNN\oplus N is a direct summand of MM. An algebra Λ\Lambda is basic if it is basic as a free Λ\Lambda-module. algebra over a field KK. In the second half of the paper, we will mainly consider the case where Λ\Lambda is the tensor algebra over a weighted Dynkin quiver; discussion of this special case is deferred to Section 6.3.

We denote by 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda the category of finitely-generated right Λ\Lambda-modules and by 𝒟b(𝗆𝗈𝖽Λ)\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda) the bounded derived category of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda with shift functor [1][1]. We recall that Λ\Lambda is hereditary if ExtΛm(M,N)=0\mathrm{Ext}^{m}_{\Lambda}(M,N)=0 for all M,N𝗆𝗈𝖽ΛM,N\in\operatorname{\mathsf{mod}}\Lambda and m>1m>1. Given a basic object M𝗆𝗈𝖽ΛM\in\operatorname{\mathsf{mod}}\Lambda, we denote by |M||M| the number of indecomposable direct summands of MM. For M,N𝗆𝗈𝖽ΛM,N\in\operatorname{\mathsf{mod}}\Lambda, we will sometimes write MNM\hookrightarrow N (resp. MNM\twoheadrightarrow N) to represent a monomorphism/injection (resp. epimorphism/surjection) in HomΛ(M,N)\mathrm{Hom}_{\Lambda}(M,N). A module X𝗆𝗈𝖽ΛX\in\operatorname{\mathsf{mod}}\Lambda is called a brick if every nonzero endomorphism of XX is invertible. In particular, every brick must be indecomposable.

Let n=|Λ|n=|\Lambda|, and choose an indexing P(1),,P(n)P(1),\ldots,P(n) of the indecomposable direct summands of Λ\Lambda. Then for every indecomposable and projective module Q𝗆𝗈𝖽ΛQ\in\operatorname{\mathsf{mod}}\Lambda, there exists a unique ii such that QP(i)Q\cong P(i). Moreover, up to isomorphism, 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda contains nn simple modules. These can be indexed S(1),,S(n)S(1),\ldots,S(n) so that HomΛ(P(i),S(j))0\mathrm{Hom}_{\Lambda}(P(i),S(j))\neq 0 if and only if i=ji=j. Moreover, each S(i)S(i) is a brick, and we have HomΛ(P(i),S(i))EndΛ(S(i))\mathrm{Hom}_{\Lambda}(P(i),S(i))\cong\mathrm{End}_{\Lambda}(S(i)). Given an arbitrary module M𝗆𝗈𝖽ΛM\in\operatorname{\mathsf{mod}}\Lambda, we define its dimension vector to be

(3.1) dim¯M=(dimEndΛ(S(i))HomΛ(P(i),M))i[n]n.\underline{\mathrm{dim}}M=\left(\mathrm{dim}_{\mathrm{End}_{\Lambda}(S(i))}\mathrm{Hom}_{\Lambda}(P(i),M)\right)_{i\in[n]}\in\mathbb{N}^{n}.
Remark 3.1.

As defined, the dimension vector dim¯M\underline{\mathrm{dim}}M depends on the indexing of the projective modules. To avoid this dependency, one can instead define dim¯M\underline{\mathrm{dim}}M to be the class [M][M] of MM in the Grothendieck group K0(𝗆𝗈𝖽Λ)K_{0}(\operatorname{\mathsf{mod}}\Lambda). Indeed, this group is free abelian and has {[S(i)]i[n]}\{[S(i)]\mid i\in[n]\} as a basis. Thus, identifying each [S(i)]K0(𝗆𝗈𝖽Λ)[S(i)]\in K_{0}(\operatorname{\mathsf{mod}}\Lambda) with the standard basis element eine_{i}\in\mathbb{Z}^{n} identifies the class [M][M] with the vector dim¯M\underline{\mathrm{dim}}M as defined above.

The purpose of the remainder of this section is to recall the definitions of torsion pairs and 2-term simple-minded collections, as well as the relationship between them. In doing so, we follow much of the exposition of [BH, Section 3].

3.1. Subcategories and functorial finiteness

By a subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda, we will always mean a subcategory that is full and closed under isomorphisms. For a subcategory 𝒞𝗆𝗈𝖽Λ\mathcal{C}\subseteq\operatorname{\mathsf{mod}}\Lambda, we define several additional subcategories as follows:

  • 𝖺𝖽𝖽(𝒞)\operatorname{\mathsf{add}}(\mathcal{C}) is the subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda consisting of all direct summands of finite direct sums of the objects in 𝒞\mathcal{C}.

  • 𝖦𝖾𝗇(𝒞)\operatorname{\mathsf{Gen}}(\mathcal{C}) is the subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda consisting of all objects that can be written as quotients of objects in 𝖺𝖽𝖽(𝒞)\operatorname{\mathsf{add}}(\mathcal{C}).

  • 𝖢𝗈𝗀𝖾𝗇(𝒞)\operatorname{\mathsf{Cogen}}(\mathcal{C}) is the subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda consisting of all objects that can be written as subobjects of objects in 𝖺𝖽𝖽(𝒞)\operatorname{\mathsf{add}}(\mathcal{C}).

  • 𝖥𝗂𝗅𝗍(𝒞)\operatorname{\mathsf{Filt}}(\mathcal{C}) is the subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda consisting of all objects that can be written as iterated extensions of objects in 𝒞\mathcal{C}; that is, M𝖥𝗂𝗅𝗍(𝒞)M\in\operatorname{\mathsf{Filt}}(\mathcal{C}) if and only if there exists a chain

    0=M0M1Mk=M0=M_{0}\subsetneq M_{1}\subsetneq\cdots\subsetneq M_{k}=M

    of modules such that each factor Mi/Mi1M_{i}/M_{i-1} is in 𝖺𝖽𝖽(𝒞)\operatorname{\mathsf{add}}(\mathcal{C}).

  • 𝖳(𝒞)=𝖥𝗂𝗅𝗍(𝖦𝖾𝗇(𝒞))\mathsf{T}(\mathcal{C})=\operatorname{\mathsf{Filt}}(\operatorname{\mathsf{Gen}}(\mathcal{C})).

  • 𝒞=𝒞0={Y𝗆𝗈𝖽ΛHomΛ(,Y)|𝒞=0}\mathcal{C}^{\perp}=\mathcal{C}^{\perp_{0}}=\{Y\in\operatorname{\mathsf{mod}}\Lambda\mid\mathrm{Hom}_{\Lambda}(-,Y)|_{\mathcal{C}}=0\}.

  • 𝒞=𝒞0={Y𝗆𝗈𝖽ΛHomΛ(Y,)|𝒞=0}\prescript{\perp}{}{\mathcal{C}}=\prescript{\perp_{0}}{}{\mathcal{C}}=\{Y\in\operatorname{\mathsf{mod}}\Lambda\mid\mathrm{Hom}_{\Lambda}(Y,-)|_{\mathcal{C}}=0\}.

  • 𝒞1={Y𝗆𝗈𝖽ΛExtΛ1(,Y)|𝒞=0}\mathcal{C}^{\perp_{1}}=\{Y\in\operatorname{\mathsf{mod}}\Lambda\mid\mathrm{Ext}^{1}_{\Lambda}(-,Y)|_{\mathcal{C}}=0\}

  • 𝒞1={Y𝗆𝗈𝖽ΛExtΛ1(Y,)|𝒞=0}\prescript{\perp_{1}}{}{\mathcal{C}}=\{Y\in\operatorname{\mathsf{mod}}\Lambda\mid\mathrm{Ext}^{1}_{\Lambda}(Y,-)|_{\mathcal{C}}=0\}.

  • 𝒞0,1=𝒞𝒞1\mathcal{C}^{\perp_{0,1}}=\mathcal{C}^{\perp}\cap\mathcal{C}^{\perp_{1}} and 𝒞0,1=𝒞𝒞1\prescript{\perp_{0,1}}{}{\mathcal{C}}=\prescript{\perp}{}{\mathcal{C}}\cap\prescript{\perp_{1}}{}{\mathcal{C}}.

We also apply the above definitions to the objects of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda by considering (the isomorphism class of) a given object as a subcategory.

Now let 𝒞𝗆𝗈𝖽Λ\mathcal{C}\subseteq\operatorname{\mathsf{mod}}\Lambda be an arbitrary subcategory and M𝗆𝗈𝖽ΛM\in\operatorname{\mathsf{mod}}\Lambda. A minimal left 𝒞\mathcal{C}-approximation of MM is a morphism gM,𝒞:MCg_{M,\mathcal{C}}\colon M\rightarrow C with C𝒞C\in\mathcal{C} such that

  1. (i)

    the induced map (gM,𝒞):HomΛ(C,D)HomΛ(M,D)(g_{M,\mathcal{C}})^{*}\colon\mathrm{Hom}_{\Lambda}(C,D)\rightarrow\mathrm{Hom}_{\Lambda}(M,D) is surjective for all D𝒞D\in\mathcal{C} and

  2. (ii)

    if f:CCf\colon C\rightarrow C is such that fgM,𝒞=gM,𝒞f\circ g_{M,\mathcal{C}}=g_{M,\mathcal{C}}, then ff is an isomorphism.

Dually, a minimal right 𝒞\mathcal{C}-approximation of MM is a morphism g𝒞,M:CMg_{\mathcal{C},M}\colon C\rightarrow M with C𝒞C\in\mathcal{C} such that

  1. (i*)

    the induced map (g𝒞,M):HomΛ(D,C)HomΛ(D,M)(g_{\mathcal{C},M})_{*}\colon\mathrm{Hom}_{\Lambda}(D,C)\rightarrow\mathrm{Hom}_{\Lambda}(D,M) is surjective for all D𝒞D\in\mathcal{C} and

  2. (ii*)

    if f:CCf\colon C\rightarrow C is such that g𝒞,Mf=g𝒞,Mg_{\mathcal{C},M}\circ f=g_{\mathcal{C},M}, then ff is an isomorphism.

It is well known that minimal left and right 𝒞\mathcal{C}-approximations are unique up to isomorphism whenever they exist. If every object of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda admits both a minimal left 𝒞\mathcal{C}-approximation and a minimal right 𝒞\mathcal{C}-approximation, then 𝒞\mathcal{C} is said to be functorially finite. Almost all of the subcategories we consider in this paper will be functorially finite.

3.2. Torsion classes

A pair (𝒯,)(\mathcal{T},\mathcal{F}) of subcategories of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda is called a torsion pair if

  1. (i)

    HomΛ(T,F)=0\mathrm{Hom}_{\Lambda}(T,F)=0 for all T𝒯T\in\mathcal{T} and FF\in\mathcal{F} and

  2. (ii)

    every M𝗆𝗈𝖽ΛM\in\operatorname{\mathsf{mod}}\Lambda is the middle term of an exact sequence t𝒯MMfMt_{\mathcal{T}}M\hookrightarrow M\twoheadrightarrow f_{\mathcal{F}}M with t𝒯M𝒯t_{\mathcal{T}}M\in\mathcal{T} and fMf_{\mathcal{F}}M\in\mathcal{F}.

In this case, we say that 𝒯\mathcal{T} is a torsion class and that \mathcal{F} is a torsion-free class. We denote by 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda and 𝗍𝗈𝗋𝗌𝖿Λ\operatorname{\mathsf{torsf}}\Lambda the sets of torsion classes and torsion-free classes in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda, respectively. Torsion pairs were introduced by Dickson in [Dic66] as a generalization of the torsion and torsion-free abelian groups. We refer to the expository article [Tho21] for background information on torsion pairs, including formal statements for the facts mentioned in this section.

It is well known that a subcategory 𝒯𝗆𝗈𝖽Λ\mathcal{T}\subseteq\operatorname{\mathsf{mod}}\Lambda (resp. 𝗆𝗈𝖽Λ\mathcal{F}\subseteq\operatorname{\mathsf{mod}}\Lambda) is a torsion class (resp. torsion-free class) if and only if it is closed under extensions and quotients (resp. extensions and submodules). Moreover, every torsion class determines a unique torsion pair and vice versa via the associations 𝒯(𝒯,𝒯)\mathcal{T}\mapsto(\mathcal{T},\mathcal{T}^{\perp}) and (𝒯,)𝒯(\mathcal{T},\mathcal{F})\mapsto\mathcal{T}. Similarly, every torsion-free class determines a unique torsion pair and vice versa via the associations (,)\mathcal{F}\mapsto(\prescript{\perp}{}{\mathcal{F}},\mathcal{F}) and (𝒯,)(\mathcal{T},\mathcal{F})\mapsto\mathcal{F}.

For 𝒞𝗆𝗈𝖽Λ\mathcal{C}\subseteq\operatorname{\mathsf{mod}}\Lambda an arbitrary subcategory, we have that 𝖳(𝒞)\mathsf{T}(\mathcal{C}) is the smallest torsion class containing 𝒞\mathcal{C}; the corresponding torsion pair is (𝖳(𝒞),𝒞)(\mathsf{T}(\mathcal{C}),\mathcal{C}^{\perp}). Dually, 𝖥𝗂𝗅𝗍(𝖢𝗈𝗀𝖾𝗇(𝒞))\operatorname{\mathsf{Filt}}(\operatorname{\mathsf{Cogen}}(\mathcal{C})) is the smallest torsion-free class containing 𝒞\mathcal{C}, and the corresponding torsion pair is (𝒞,𝖥𝗂𝗅𝗍(𝖢𝗈𝗀𝖾𝗇(𝒞)))(\prescript{\perp}{}{\mathcal{C}},\operatorname{\mathsf{Filt}}(\operatorname{\mathsf{Cogen}}(\mathcal{C}))).

We consider 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda and 𝗍𝗈𝗋𝗌𝖿Λ\operatorname{\mathsf{torsf}}\Lambda as posets under the containment relation. Each of these is a complete lattice whose meet operation is given by intersection and whose join operation is given by 𝒜=𝖥𝗂𝗅𝗍(𝒯𝒜𝒯)\bigvee\mathcal{A}=\operatorname{\mathsf{Filt}}\left(\bigcup_{\mathcal{T}\in\mathcal{A}}\mathcal{T}\right). The associations 𝒯𝒯\mathcal{T}\mapsto\mathcal{T}^{\perp} and \mathcal{F}\mapsto\prescript{\perp}{}{\mathcal{F}} thus give lattice anti-isomorphisms between 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda and 𝗍𝗈𝗋𝗌𝖿Λ\operatorname{\mathsf{torsf}}\Lambda. In particular, the join operation in 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda can be rewritten as 𝒯𝒰=(𝒯𝒰){\mathcal{T}\vee\mathcal{U}=\prescript{\perp}{}{(}{\mathcal{T}^{\perp}\cap\mathcal{U}^{\perp}})}. The maximum element of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda is 1^=𝗆𝗈𝖽Λ\hat{1}=\operatorname{\mathsf{mod}}\Lambda (the torsion class consisting of all modules), and the minimum element is 0^=0\hat{0}=0 (the torsion class consisting only of the zero module). The atoms of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda are precisely those torsion classes of the form 𝖥𝗂𝗅𝗍(S)\operatorname{\mathsf{Filt}}(S) for S𝗆𝗈𝖽ΛS\in\operatorname{\mathsf{mod}}\Lambda a simple module.

In this paper, we restrict our attention to algebras Λ\Lambda for which the lattice 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda is finite. These are precisely the τ\tau-tilting finite algebras introduced in [DIJ17]. They are also precisely the algebras Λ\Lambda for which every torsion class in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda is functorially finite. A related class is that of representation-finite algebras, which are characterized by the property that 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda contains only finitely many indecomposable modules (up to isomorphism). If Λ\Lambda is hereditary, then it is τ\tau-tilting finite if and only if it is representation-finite. For Λ\Lambda not hereditary, representation-finiteness implies τ\tau-tilting finiteness, but not vice versa. Another special class is that of silting discrete algebras [AM17]; an algebra Λ\Lambda belongs to this class when every algebra that is derived equivalent to Λ\Lambda (including Λ\Lambda itself) is τ\tau-tilting finite. All representation-finite hereditary algebras are silting discrete.

From now on, Λ\Lambda will always denote a τ\tau-tilting finite algebra. While many of the results of the following subsections admit generalizations to arbitrary finite-dimensional algebras, this allows us to streamline the exposition.

3.3. Semibricks and 2-term simple-minded collections

Recall that a module X𝗆𝗈𝖽ΛX\in\operatorname{\mathsf{mod}}\Lambda is called a brick if EndΛ(X)\mathrm{End}_{\Lambda}(X) is a division ring. A set 𝒳\mathcal{X} of bricks is called a semibrick if XYYX\in Y^{\perp}\cap\prescript{\perp}{}{Y} for all distinct X,Y𝒳X,Y\in\mathcal{X}. We adopt the convention of using the term 𝒳\mathcal{X} to represent both a set of bricks and the module X𝒳X\bigoplus_{X\in\mathcal{X}}X. We denote by 𝖻𝗋𝗂𝖼𝗄Λ\operatorname{\mathsf{brick}}\Lambda and 𝗌𝖻𝗋𝗂𝖼𝗄Λ\operatorname{\mathsf{sbrick}}\Lambda the sets of (isomorphism classes of) bricks and semibricks in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda, respectively. This leads to the following definition, which we state in this form in light of [BY13, Remark 4.11]. See Section 1.2 for historical notes and references.

Definition 3.2.

Let 𝒳,𝒴𝗌𝖻𝗋𝗂𝖼𝗄(Λ)\mathcal{X},\mathcal{Y}\in\operatorname{\mathsf{sbrick}}(\Lambda).

  1. (1)

    We say (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) is a semibrick pair if 𝒴𝒳0,1\mathcal{Y}\in\mathcal{X}^{\perp_{0,1}}.

  2. (2)

    We say a semibrick pair (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) is a 2-term simple-minded collection if the only triangulated subcategory of 𝒟b(𝗆𝗈𝖽Λ)\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda) that contains 𝒳𝒴[1]\mathcal{X}\oplus\mathcal{Y}[1] and is closed under direct summands is 𝒟b(𝗆𝗈𝖽Λ)\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda) itself.

  3. (3)

    We say a semibrick pair (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) is completable if there exists a 2-term simple-minded collection (𝒳,𝒴)(\mathcal{X}^{\prime},\mathcal{Y}^{\prime}) with 𝒳𝒳\mathcal{X}\subseteq\mathcal{X}^{\prime} and 𝒴𝒴\mathcal{Y}\subseteq\mathcal{Y}^{\prime}.

We denote by sbpΛ\operatorname{\mathrm{sbp}}\Lambda the set of semibrick pairs and by 2-𝗌𝗆𝖼Λ\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda the set of 2-term simple-minded collections.

Remark 3.3.
  1. (1)

    Not every semibrick pair is completable, even in the τ\tau-tilting finite case; see, e.g., [HI21b, Counterexample 1.9].

  2. (2)

    If (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) is a 2-term simple-minded collection, then |𝒳|+|𝒴|=|Λ||\mathcal{X}|+|\mathcal{Y}|=|\Lambda|. If the algebra Λ\Lambda is silting discrete, then the converse also holds by [HW, Corollary 6.12]; that is, if (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) is a semibrick pair with |𝒳|+|𝒴|=|Λ||\mathcal{X}|+|\mathcal{Y}|=|\Lambda|, then (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) must be a 2-term simple-minded collection.

3.4. Semidistributivity and the brick labeling

The lattice of torsion classes is known to be a (completely) semidistributive lattice [DIR+23, GM19]. Therefore, each edge in the Hasse diagram of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda has a join-irreducible shard label as described at the end of Section 2. In this section, we recall from [Asa20] (τ\tau-tilting finite case) and [BCZ19, DIR+23] (general case) how this labeling can be formulated and interpreted representation-theoretically. We note that many of the results recalled in this section have been specialized to the τ\tau-tilting finite case.

Theorem 3.4 ([BCZ19, BTZ21]).

Let Λ\Lambda be a τ\tau-tilting finite algebra.

  1. (1)

    There is a bijection 𝖻𝗋𝗂𝖼𝗄ΛJIrr𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{brick}}\Lambda\rightarrow\mathrm{JIrr}_{\operatorname{\mathsf{tors}}\Lambda} given by X𝖳(X)X\mapsto\mathsf{T}(X).

  2. (2)

    There is a bijection 𝖻𝗋𝗂𝖼𝗄ΛMIrr𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{brick}}\Lambda\rightarrow\mathrm{MIrr}_{\operatorname{\mathsf{tors}}\Lambda} given by XXX\mapsto\prescript{\perp}{}{X}.

  3. (3)

    The bijection κ:JIrr𝗍𝗈𝗋𝗌ΛMIrr𝗍𝗈𝗋𝗌Λ\kappa\colon\mathrm{JIrr}_{\operatorname{\mathsf{tors}}\Lambda}\to\mathrm{MIrr}_{\operatorname{\mathsf{tors}}\Lambda} is given by κ(𝖳(X))=X\kappa(\mathsf{T}(X))=\prescript{\perp}{}{X} for all X𝖻𝗋𝗂𝖼𝗄ΛX\in\operatorname{\mathsf{brick}}\Lambda.

The brick label of a cover relation 𝒯<𝒯\mathcal{T}{\,\,<\!\!\!\!\cdot\,\,\,}\mathcal{T}^{\prime} in 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda is the unique brick XX for which j𝒯,𝒯=𝖳(X)j_{\mathcal{T},\mathcal{T}^{\prime}}=\mathsf{T}(X). For a fixed 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda, we (slightly abusing notation) denote by 𝒟(𝒯)\mathcal{D}(\mathcal{T}) and 𝒰(𝒯)\mathcal{U}(\mathcal{T}) the sets of bricks that label cover relations of the form 𝒯<𝒯\mathcal{T}^{\prime}{\,\,<\!\!\!\!\cdot\,\,\,}\mathcal{T} and 𝒯<𝒯\mathcal{T}{\,\,<\!\!\!\!\cdot\,\,\,}\mathcal{T}^{\prime}, respectively.

Lemma 3.5 ([BCZ19, Corollary 3.9]).

Let 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda.

  1. (1)

    If Y𝒰(𝒯)Y\in\mathcal{U}(\mathcal{T}) and YYY\twoheadrightarrow Y^{\prime} is a surjection with nonzero kernel, then Y𝒯Y\in\mathcal{T}.

  2. (2)

    If X𝒟(𝒯)X\in\mathcal{D}(\mathcal{T}) and XXX^{\prime}\hookrightarrow X is an injection with nonzero cokernel, then X𝒯X^{\prime}\in\mathcal{T}^{\perp}.

Theorems 3.4 and 3.5 immediately yield the following characterization of the pop-stack operators.

Corollary 3.6.

For 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda, we have

pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝒯𝒟(𝒯)andpop𝗍𝗈𝗋𝗌Λ(𝒯)=𝖥𝗂𝗅𝗍(𝒯𝒰(𝒯)).\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\mathcal{T}\cap\prescript{\perp}{}{\mathcal{D}(\mathcal{T})}\quad\text{and}\quad\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\operatorname{\mathsf{Filt}}(\mathcal{T}\cup\mathcal{U}(\mathcal{T})).

We now recall how Theorem 3.4 relates to 2-term simple-minded collections. Once again, we note that the following theorem is simplified since we consider only the case where every torsion class is functorially finite. This result is essentially contained in [Asa20, Theorems 2.3 and 2.12], but we give a short argument utilizing [BTZ21, Corollary 5.1.8] and other results from [Asa20]. Note that it is implicitly included in Theorem 3.71 that the sets 𝒟(𝒯)\mathcal{D}(\mathcal{T}) and 𝒰(𝒯)\mathcal{U}(\mathcal{T}) are semibricks for any 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda.

Theorem 3.7.
  1. (1)

    There are bijections 𝗌𝖻𝗋𝗂𝖼𝗄Λ𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{sbrick}}\Lambda\rightarrow\operatorname{\mathsf{tors}}\Lambda given by 𝒳𝖳(𝒳)\mathcal{X}\mapsto\mathsf{T}(\mathcal{X}) and 𝒳𝒳\mathcal{X}\mapsto\prescript{\perp}{}{\mathcal{X}}. Their inverses are given by 𝒯𝒟(𝒯)\mathcal{T}\mapsto\mathcal{D}(\mathcal{T}) and 𝒯𝒰(𝒯)\mathcal{T}\mapsto\mathcal{U}(\mathcal{T}), respectively.

  2. (2)

    If (𝒳,𝒴)sbpΛ(\mathcal{X},\mathcal{Y})\in\operatorname{\mathrm{sbp}}\Lambda, then 𝖳(𝒳)𝒴\mathsf{T}(\mathcal{X})\subseteq\prescript{\perp}{}{\mathcal{Y}}, with equality if and only if (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda. In particular, there is a bijection 2-𝗌𝗆𝖼Λ𝗍𝗈𝗋𝗌Λ\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda\rightarrow\operatorname{\mathsf{tors}}\Lambda given by (𝒳,𝒴)𝖳(𝒳)=𝒴(\mathcal{X},\mathcal{Y})\mapsto\mathsf{T}(\mathcal{X})=\prescript{\perp}{}{\mathcal{Y}}. The inverse is given by 𝒯(𝒟(𝒯),𝒰(𝒯))\mathcal{T}\mapsto(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T})).

  3. (3)

    There are bijections 2-𝗌𝗆𝖼Λ𝗌𝖻𝗋𝗂𝖼𝗄Λ\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda\rightarrow\operatorname{\mathsf{sbrick}}\Lambda given by (𝒳,𝒴)𝒳(\mathcal{X},\mathcal{Y})\mapsto\mathcal{X} and (𝒳,𝒴)𝒴(\mathcal{X},\mathcal{Y})\mapsto\mathcal{Y}.

Proof.

1 The fact that these associations are bijections is from [Asa20, Theorem 2.3]. The explicit description of the inverses follows from [Asa20, Proposition 1.26] and [BTZ21, Corollary 5.1.8].

2 That 𝖳(X)𝒴\mathsf{T}(X)\subseteq\prescript{\perp}{}{\mathcal{Y}} follows immediately from the definition. The fact that we have equality if and only if (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda then follows from [Asa20, Theorem 2.3].

3 This follows from [Asa20, Theorem 2.3(i)]. ∎

In particular, Theorem 3.7 implies that any semibrick pair of the form (𝒳,)(\mathcal{X},\emptyset) or (,𝒴)(\emptyset,\mathcal{Y}) is completable. For (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda, we write 𝖳(𝒳,𝒴)=𝖳(𝒳)=𝒴\mathsf{T}(\mathcal{X},\mathcal{Y})=\mathsf{T}(\mathcal{X})=\prescript{\perp}{}{\mathcal{Y}}.

3.5. Wide and Serre subcategories

A subcategory 𝒲𝗆𝗈𝖽Λ\mathcal{W}\subseteq\operatorname{\mathsf{mod}}\Lambda is called wide if it is closed under extensions, kernels, and cokernels. Alternatively, a subcategory is wide if it is an exact-embedded abelian subcategory. We denote by 𝗐𝗂𝖽𝖾Λ\operatorname{\mathsf{wide}}\Lambda the set of wide subcategories of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda.

It is a classical result of [Rin76] that there is a bijection 𝗌𝖻𝗋𝗂𝖼𝗄Λ𝗐𝗂𝖽𝖾Λ\operatorname{\mathsf{sbrick}}\Lambda\rightarrow\operatorname{\mathsf{wide}}\Lambda given by 𝒳𝖥𝗂𝗅𝗍(𝒳)\mathcal{X}\mapsto\operatorname{\mathsf{Filt}}(\mathcal{X}). The inverse sends a wide subcategory 𝒲\mathcal{W} to the set of objects that are simple in 𝒲\mathcal{W} (i.e., those X𝒲X\in\mathcal{W} that admit no proper submodules in 𝒲\mathcal{W}). Combining this with Theorem 3.7, we see that, under the τ\tau-tilting finite hypothesis, there are two bijections 𝗍𝗈𝗋𝗌Λ𝗐𝗂𝖽𝖾Λ\operatorname{\mathsf{tors}}\Lambda\rightarrow\operatorname{\mathsf{wide}}\Lambda given by 𝒯𝖥𝗂𝗅𝗍(𝒟(𝒯))\mathcal{T}\mapsto\operatorname{\mathsf{Filt}}(\mathcal{D}(\mathcal{T})) and 𝒯𝖥𝗂𝗅𝗍(𝒰(𝒯))\mathcal{T}\mapsto\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T})). The inverses of these bijections are given by 𝒲𝖳(𝒲)\mathcal{W}\mapsto\mathsf{T}(\mathcal{W}) and 𝒲𝒲\mathcal{W}\mapsto\prescript{\perp}{}{\mathcal{W}}, respectively. Moreover, these bijections correspond to the so-called Ingalls–Thomas bijections of [IT09, MŠ17]; that is, we have the following. (See also [Eno23, Remark 4.20], which relates this result to the rowmotion operators.)

Proposition 3.8 ([Asa20, Theorem 2.3]).

Let 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda.

  1. (1)

    We have 𝖥𝗂𝗅𝗍(𝒟(𝒯))={M𝒯kerf𝒯 for all N𝒯 and f:NM}\operatorname{\mathsf{Filt}}(\mathcal{D}(\mathcal{T}))=\{M\in\mathcal{T}\mid\mathrm{ker}f\in\mathcal{T}\text{ for all }N\in\mathcal{T}\text{ and }f\colon N\rightarrow M\}.

  2. (2)

    We have 𝖥𝗂𝗅𝗍(𝒰(𝒯))={M𝒯cokerf𝒯 for all N𝒯 and f:MN}\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T}))=\{M\in\mathcal{T}^{\perp}\mid\mathrm{coker}f\in\mathcal{T}^{\perp}\text{ for all }N\in\mathcal{T}^{\perp}\text{ and }f\colon M\rightarrow N\}.

Remark 3.9.

Since Λ\Lambda is τ\tau-tilting finite, every wide subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda is functorially finite by [MŠ17, Corollary 3.11]. Equivalently (see [Eno22, Proposition 4.12]), for every 𝒲𝗐𝗂𝖽𝖾Λ\mathcal{W}\in\operatorname{\mathsf{wide}}\Lambda, there exists a finite-dimensional algebra Λ\Lambda^{\prime} such that 𝒲\mathcal{W} is exact-equivalent to 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda^{\prime}.

We conclude this section with a brief discussion of Serre subcategories. A subcategory 𝒲𝗆𝗈𝖽Λ{\mathcal{W}\subseteq\operatorname{\mathsf{mod}}\Lambda} is called Serre if it is closed under extensions, submodules, and quotient modules. The following is well known.

Proposition 3.10.

Let 𝒲𝗆𝗈𝖽Λ\mathcal{W}\subseteq\operatorname{\mathsf{mod}}\Lambda. Then the following are equivalent.

  1. (1)

    The subcategory 𝒲\mathcal{W} is Serre.

  2. (2)

    The subcategory 𝒲\mathcal{W} is at least two of the following: a wide subcategory, a torsion class, and a torsion-free class.

  3. (3)

    The subcategory 𝒲\mathcal{W} is a wide subcategory, a torsion class, and a torsion-free class.

  4. (4)

    We have 𝒲=𝖥𝗂𝗅𝗍(SS)\mathcal{W}=\operatorname{\mathsf{Filt}}(\SS) for some set SS\SS of modules that are simple in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda.

  5. (5)

    There exists a projective module P𝗆𝗈𝖽ΛP\in\operatorname{\mathsf{mod}}\Lambda such that 𝒲=P\mathcal{W}=P^{\perp}.

  6. (6)

    There exists an injective module I𝗆𝗈𝖽ΛI\in\operatorname{\mathsf{mod}}\Lambda such that 𝒲=I\mathcal{W}=\prescript{\perp}{}{I}.

  7. (7)

    The subcategory 𝒲\mathcal{W} is wide, and every object that is simple in 𝒲\mathcal{W} is also simple in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda.

Given a wide subcategory 𝒲\mathcal{W}, we will also say that a wide subcategory 𝒲𝒲\mathcal{W}^{\prime}\subseteq\mathcal{W} is Serre in 𝒲\mathcal{W} if every object that is simple in 𝒲\mathcal{W}^{\prime} is also simple in 𝒲\mathcal{W}.

4. Mutation of Semibrick Pairs

In this section, we extend the theory of mutation of semibrick pairs established in [HI21a, Section 3] and [BH22, Section 3.4]. This extends the mutation formulas for 2-term simple-minded collections from [KY14, Section 7.2] and [BY13, Section 3.7]. In those papers, one mutates a semibrick pair (resp. 2-term simple-minded collection) at a single brick. In the present paper, we extend this to be able to mutate at multiple bricks simultaneously.

Recall that Λ\Lambda denotes a τ\tau-tilting finite algebra, and let (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) be a semibrick pair. For X𝒳X\in\mathcal{X} and 𝒴𝒴\mathcal{Y}^{\prime}\subseteq\mathcal{Y}, denote by g𝒴,X:YXXg_{\mathcal{Y}^{\prime},X}\colon Y^{\prime}_{X}\rightarrow X a minimal right 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y}^{\prime})-approximation of XX. (Note that YX=0Y^{\prime}_{X}=0 if 𝒴=\mathcal{Y}^{\prime}=\emptyset.) Likewise for Y𝒴Y\in\mathcal{Y} and 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}, denote by gY,𝒳:YXYg_{Y,\mathcal{X}^{\prime}}\colon Y\rightarrow X^{\prime}_{Y} a minimal left 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})-approximation of YY. (Note that XY=0X^{\prime}_{Y}=0 if 𝒳=\mathcal{X}^{\prime}=\emptyset.) We note that both g𝒴,Xg_{\mathcal{Y}^{\prime},X} and gY,𝒳g_{Y,\mathcal{X}^{\prime}} exist and are unique up to isomorphism because 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}) and 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y}^{\prime}) are functorially finite wide subcategories; see Sections 3.1 and 3.5.

Definition 4.1.

We say a semibrick pair (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) is singly mutation (SM) compatible if for all X𝒳X\in\mathcal{X}, Y𝒴Y\in\mathcal{Y}, 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}, and 𝒴𝒴\mathcal{Y}^{\prime}\subseteq\mathcal{Y}, the map g𝒴,Xg_{\mathcal{Y}^{\prime},X} is either injective or surjective and the map gY,𝒳g_{Y,\mathcal{X}^{\prime}} is either injective or surjective.

Remark 4.2.

Note that SM compatibility is stronger than, but similar to, the notions of singly left mutation compatible and singly right mutation compatible used in [BH22, HI21b, HI21a]

Proposition 4.3.

Every completable semibrick pair is SM compatible.

Proof.

It suffices to prove the result only for 2-term simple-minded collections. We use an argument similar to that of [BTZ21, Proposition 5.2.1]. Let (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda.

Let (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda, and denote 𝒯:=𝖳(𝒳)\mathcal{T}:=\mathsf{T}(\mathcal{X}) and :=𝖥𝗂𝗅𝗍(𝖢𝗈𝗀𝖾𝗇𝒴)\mathcal{F}:=\operatorname{\mathsf{Filt}}(\operatorname{\mathsf{Cogen}}\mathcal{Y}). Recall from Theorem 3.7 that (𝒯,)(\mathcal{T},\mathcal{F}) is a torsion pair, that 𝒯=𝒴\mathcal{T}=\prescript{\perp}{}{\mathcal{Y}} and =𝒳\mathcal{F}=\mathcal{X}^{\perp}, and that 𝒳=𝒟(𝒯)\mathcal{X}=\mathcal{D}(\mathcal{T}) and 𝒴=𝒰(𝒯)\mathcal{Y}=\mathcal{U}(\mathcal{T}).

Let 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X} and Y𝒴Y\in\mathcal{Y}, and suppose that gY,𝒳g_{Y,\mathcal{X}^{\prime}} is not injective. Then Z:=im(gY,𝒳)𝒯Z:=\mathrm{im}(g_{Y,\mathcal{X}^{\prime}})\in\mathcal{T} by Lemma 3.5. Denote by ι:ZXY\iota\colon Z\rightarrow X^{\prime}_{Y} the inclusion map. Then for W𝒯W\in\mathcal{T} and f:WZf\colon W\rightarrow Z, we have ker(f)=ker(ιf)\mathrm{ker}(f)=\mathrm{ker}(\iota\circ f). Since XY𝖥𝗂𝗅𝗍(𝒳)𝖥𝗂𝗅𝗍(𝒟(𝒯))X^{\prime}_{Y}\in\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})\subseteq\operatorname{\mathsf{Filt}}(\mathcal{D}(\mathcal{T})), Proposition 3.8 implies that ker(f)𝒯\mathrm{ker}(f)\in\mathcal{T}, and therefore that Z𝖥𝗂𝗅𝗍(𝒳)Z\in\operatorname{\mathsf{Filt}}(\mathcal{X}). The fact that 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}) is Serre in 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) thus implies that Z𝖥𝗂𝗅𝗍(𝒳)Z\in\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}). By the minimality of gY,𝒳g_{Y,\mathcal{X}^{\prime}}, we conclude that Z=XYZ=X^{\prime}_{Y}; i.e., gY,𝒳g_{Y,\mathcal{X}^{\prime}} is surjective. The argument that each map gX,𝒴g_{X,\mathcal{Y}^{\prime}} is either injective or surjective is dual. ∎

We now prepare to define the mutation of an SM compatible semibrick pair (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) at either a subset 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X} or a subset 𝒴𝒴\mathcal{Y}^{\prime}\subseteq\mathcal{Y}. We will first formulate the mutation formulas within the category 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda (Definition 4.4) and then explain how they can be restated using the language of the bounded derived category (Remark 4.5).

Fix an SM compatible semibrick pair (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) and subsets 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X} and 𝒴𝒴\mathcal{Y}^{\prime}\subseteq\mathcal{Y}. Let X𝒳𝒳X\in\mathcal{X}\setminus\mathcal{X}^{\prime}. Since the wide subcategory 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) is functorially finite, it is well known that it contains a projective generator; that is, there exists P𝒳𝖥𝗂𝗅𝗍(𝒳)P_{\mathcal{X}}\in\operatorname{\mathsf{Filt}}(\mathcal{X}) such that 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) is the subcategory obtained by closing 𝖺𝖽𝖽(P𝒳)\operatorname{\mathsf{add}}(P_{\mathcal{X}}) under cokernels. See [Eno22, Proposition 4.12] for an explicit proof. Thus, we can consider a short exact sequence

Ω𝒳XP𝒳,XX,\Omega_{\mathcal{X}}X\hookrightarrow P_{\mathcal{X},X}\twoheadrightarrow X,

where P𝒳,X𝖺𝖽𝖽(P𝒳)P_{\mathcal{X},X}\in\operatorname{\mathsf{add}}(P_{\mathcal{X}}) and Ω𝒳X𝖥𝗂𝗅𝗍(𝒳)\Omega_{\mathcal{X}}X\in\operatorname{\mathsf{Filt}}(\mathcal{X}) are the projective cover and first syzygy of XX in the wide subcategory 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}). Let γX,𝒳:Ω𝒳XXX\gamma_{X,\mathcal{X}^{\prime}}\colon\Omega_{\mathcal{X}}X\rightarrow X^{\prime}_{X} be a minimal left 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})-approximation of Ω𝒳X\Omega_{\mathcal{X}}X. Then γX,𝒳\gamma_{X,\mathcal{X}^{\prime}} is surjective because 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}) is a Serre subcategory of 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}). Hence, we can form the following pushout diagram:

(4.1) Ω𝒳X{\Omega_{\mathcal{X}}X}P𝒳,X{P_{\mathcal{X},X}}X{X}ηX,𝒳:{\eta_{X,\mathcal{X}^{\prime}}:}XX{X^{\prime}_{X}}EX,𝒳{E_{X,\mathcal{X}^{\prime}}}X.{X.}γX,𝒳\scriptstyle{\gamma_{X,\mathcal{X}^{\prime}}}

We can consider the short exact sequence ηX,𝒳ExtΛ1(X,XX)\eta_{X,\mathcal{X}^{\prime}}\in\mathrm{Ext}^{1}_{\Lambda}(X,X^{\prime}_{X}) as a morphism XXX[1]X\rightarrow X^{\prime}_{X}[1] in the bounded derived category 𝒟b(𝗆𝗈𝖽Λ)\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda). Then ηX,𝒳[1]\eta_{X,\mathcal{X}^{\prime}}[1] is a minimal left 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})-approximation of X[1]X[-1] with cone EX,𝒳E_{X,\mathcal{X}^{\prime}}.

Dually, let Y𝒴𝒴Y\in\mathcal{Y}\setminus\mathcal{Y}^{\prime}. Since 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y}) is functorially finite, it contains an injective cogenerator; that is, there exists I𝒴𝖥𝗂𝗅𝗍(𝒴)I_{\mathcal{Y}}\in\operatorname{\mathsf{Filt}}(\mathcal{Y}) such that 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y}) is the subcategory obtained by closing 𝖺𝖽𝖽(I𝒴)\operatorname{\mathsf{add}}(I_{\mathcal{Y}}) under kernels. Thus, we can consider a short exact sequence

YI𝒴,YΣ𝒴Y,Y\hookrightarrow I_{\mathcal{Y},Y}\twoheadrightarrow\Sigma_{\mathcal{Y}}Y,

where Y𝒴,YY_{\mathcal{Y},Y} and Σ𝒴Y\Sigma_{\mathcal{Y}}Y are the injective envelope and first cosyzygy of YY in the wide subcategory 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y}). Let γ𝒴,Y:YYΣ𝒴Y\gamma_{\mathcal{Y}^{\prime},Y}\colon Y^{\prime}_{Y}\rightarrow\Sigma_{\mathcal{Y}}Y be a minimal right 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y}^{\prime})-approximation of Σ𝒴Y\Sigma_{\mathcal{Y}}Y. Then γ𝒴,Y\gamma_{\mathcal{Y}^{\prime},Y} is injective because 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y}^{\prime}) is a Serre subcategory of 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y}). Hence, we can form the following pullback diagram:

(4.2) η𝒴,Y:{\eta_{\mathcal{Y}^{\prime},Y}:}Y{Y}E𝒴,Y{E_{\mathcal{Y}^{\prime},Y}}YY{Y^{\prime}_{Y}}Y{Y}I𝒴,Y{I_{\mathcal{Y},Y}}Σ𝒴Y.{\Sigma_{\mathcal{Y}}Y.}γ𝒴,Y\scriptstyle{\gamma_{\mathcal{Y}^{\prime},Y}}

As above, we can consider the short exact sequence η𝒴,YExtΛ1(YY,Y)\eta_{\mathcal{Y}^{\prime},Y}\in\mathrm{Ext}^{1}_{\Lambda}(Y^{\prime}_{Y},Y) as a morphism YYY[1]Y^{\prime}_{Y}\rightarrow Y[1] in the bounded derived category 𝒟b(𝗆𝗈𝖽Λ)\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda). Then η𝒴,Y\eta_{\mathcal{Y}^{\prime},Y} is a minimal right 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y})-approximation of Y[1]Y[1] with cocone E𝒴,YE_{\mathcal{Y}^{\prime},Y}.

We now present our generalized definition.

Definition 4.4.

Let (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) be an SM compatible semibrick pair.

  1. (1)

    Let 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}. Let

    μ𝒳(𝒳,𝒴)d\displaystyle\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{d} =\displaystyle= {coker(gY,𝒳)gY,𝒳 is injective}{EX,𝒳X𝒳𝒳};\displaystyle\{\mathrm{coker}(g_{Y,\mathcal{X}^{\prime}})\mid g_{Y,\mathcal{X}^{\prime}}\text{ is injective}\}\cup\{E_{X,\mathcal{X}^{\prime}}\mid X\in\mathcal{X}\setminus\mathcal{X}^{\prime}\};
    μ𝒳(𝒳,𝒴)u\displaystyle\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{u} =\displaystyle= {ker(gY,𝒳)gY,𝒳 is surjective}𝒳.\displaystyle\{\mathrm{ker}(g_{Y,\mathcal{X}^{\prime}})\mid g_{Y,\mathcal{X}^{\prime}}\text{ is surjective}\}\cup\mathcal{X}^{\prime}.

    The left mutation of (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) at 𝒳\mathcal{X}^{\prime} is μ𝒳(𝒳,𝒴)=(μ𝒳(𝒳,𝒴)d,μ𝒳(𝒳,𝒴)u)\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})=(\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{d},\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{u}).

  2. (2)

    Let 𝒴𝒴\mathcal{Y}^{\prime}\subseteq\mathcal{Y}. Let

    μ𝒴(𝒳,𝒴)d\displaystyle\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y})_{d} =\displaystyle= {coker(g𝒴,X)g𝒴,X is injective}𝒴;\displaystyle\{\mathrm{coker}(g_{\mathcal{Y}^{\prime},X})\mid g_{\mathcal{Y}^{\prime},X}\text{ is injective}\}\cup\mathcal{Y}^{\prime};
    μ𝒴(𝒳,𝒴)u\displaystyle\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y})_{u} =\displaystyle= {ker(g𝒴,X)g𝒴,X is surjective}{EYY𝒴𝒴}.\displaystyle\{\mathrm{ker}(g_{\mathcal{Y}^{\prime},X})\mid g_{\mathcal{Y}^{\prime},X}\text{ is surjective}\}\cup\{E_{Y}\mid Y\in\mathcal{Y}\setminus\mathcal{Y}^{\prime}\}.

    The right mutation of (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) at 𝒴\mathcal{Y}^{\prime} is μ𝒴(𝒳,𝒴)=(μ𝒴(𝒳,𝒴)d,μ𝒴(𝒳,𝒴)u)\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y})=(\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y})_{d},\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y})_{u}).

We note that μ𝒳(𝒳,𝒴)\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y}) and μ𝒴(𝒳,𝒴)\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y}) are well defined by the existence and uniqueness (up to isomorphism) of projective covers, injective envelopes, minimal left 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})-approximations, minimal right 𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{Y}^{\prime})-approximations, finite limits, and finite colimits. Note also that μ(𝒳,𝒴)=(𝒳,𝒴){\mu_{\emptyset}(\mathcal{X},\mathcal{Y})=(\mathcal{X},\mathcal{Y})} by the definition of either left or right mutation. Since 𝒳𝒴=\mathcal{X}\cap\mathcal{Y}=\emptyset for any semibrick pair, there is therefore no ambiguity in the notation.

Recalling that if f:MNf\colon M\rightarrow N is a monomorphism (resp. epimorphism) in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda, then its cone (resp. cocone) in 𝒟b(𝗆𝗈𝖽Λ)\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda) is cokerf\mathrm{coker}f (resp. kerf[1]\mathrm{ker}f[1]), we have the following reformulation of Definition 4.4.

Remark 4.5.

Let (𝒳,𝒴)sbpΛ(\mathcal{X},\mathcal{Y})\in\operatorname{\mathrm{sbp}}\Lambda be SM compatible, and consider 𝒳𝒴[1]𝒟b(𝗆𝗈𝖽Λ)\mathcal{X}\oplus\mathcal{Y}[1]\in\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda).

  1. (1)

    Let 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}. For Z(𝒳𝒳)𝒴[1]Z\in(\mathcal{X}\setminus\mathcal{X}^{\prime})\cup\mathcal{Y}[1], let gZ,𝒳:Z[1]XZg_{Z,\mathcal{X}^{\prime}}:Z[-1]\rightarrow X^{\prime}_{Z} be a minimal left 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})-approximation in 𝒟b(𝗆𝗈𝖽Λ)\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda). Then

    μ𝒳(𝒳,𝒴)dμ𝒳(𝒳,𝒴)u={cone(gZ,𝒳)Z(𝒳𝒳)𝒴[1]}𝒳[1]}.\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{d}\oplus\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{u}=\{\mathrm{cone}(g_{Z,\mathcal{X}^{\prime}})\mid Z\in(\mathcal{X}\setminus\mathcal{X}^{\prime})\cup\mathcal{Y}[1]\}\oplus\mathcal{X}^{\prime}[1]\}.
  2. (2)

    Let 𝒴𝒴\mathcal{Y}^{\prime}\subseteq\mathcal{Y}. For Z(𝒴[1]𝒴[1])𝒳Z\in(\mathcal{Y}[1]\setminus\mathcal{Y}^{\prime}[1])\cup\mathcal{X}, let g𝒴,Z[1]:𝒴Z[1]Z[1]g_{\mathcal{Y}^{\prime},Z}[1]\colon\mathcal{Y}^{\prime}_{Z}[1]\rightarrow Z[1] be a minimal right 𝖥𝗂𝗅𝗍(𝒴)[1]\operatorname{\mathsf{Filt}}(\mathcal{Y}^{\prime})[1]-approximation in 𝒟b(𝗆𝗈𝖽Λ)\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda). Then

    μ𝒴(𝒳,𝒴)dμ𝒴(𝒳,𝒴)u={cocone(g𝒴,Z[1])Z(𝒴[1]𝒴[1])𝒳}𝒴.\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y})_{d}\oplus\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y})_{u}=\{\mathrm{cocone}(g_{\mathcal{Y}^{\prime},Z}[1])\mid Z\in(\mathcal{Y}[1]\setminus\mathcal{Y}^{\prime}[1])\cup\mathcal{X}\}\oplus\mathcal{Y}^{\prime}.

In particular, in the cases |𝒳|=1|\mathcal{X}^{\prime}|=1 and |𝒴|=1|\mathcal{Y}^{\prime}|=1, Items 1 and 2 of Definition 4.4 correspond to the notions of left and right mutations of semibrick pairs from [HI21a, Section 3].

We conclude this section by stating the following result, which justifies the name mutation and tabulates several useful facts about mutations of semibrick pairs. In Appendix A, we prove Theorem 4.6 in multiple steps using arguments similar to those appearing in [HI21a, Section 3] and [KY14, Section 7.2].

Theorem 4.6.

Let (𝒳,𝒴)sbpΛ(\mathcal{X},\mathcal{Y})\in\operatorname{\mathrm{sbp}}\Lambda be SM compatible. Let 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}, and denote 𝒳1=μ𝒳(𝒳,𝒴)d\mathcal{X}_{1}=\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{d} and 𝒴1=μ𝒳(𝒳,𝒴)u\mathcal{Y}_{1}=\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{u}. Then the following hold.

  1. (1)

    (𝒳1,𝒴1)(\mathcal{X}_{1},\mathcal{Y}_{1}) is a semibrick pair.

  2. (2)

    If (𝒳1,𝒴1)(\mathcal{X}_{1}^{\prime},\mathcal{Y}_{1}^{\prime}) is an SM compatible semibrick pair such that 𝒳1𝒳1\mathcal{X}_{1}\subseteq\mathcal{X}^{\prime}_{1} and 𝒴1𝒴1\mathcal{Y}_{1}\subseteq\mathcal{Y}^{\prime}_{1}, then 𝒳μ𝒳(𝒳1,𝒴1)d\mathcal{X}\subseteq\mu_{\mathcal{X}^{\prime}}(\mathcal{X}^{\prime}_{1},\mathcal{Y}^{\prime}_{1})_{d} and 𝒴μ𝒳(𝒳1,𝒴1)u\mathcal{Y}\subseteq\mu_{\mathcal{X}^{\prime}}(\mathcal{X}^{\prime}_{1},\mathcal{Y}^{\prime}_{1})_{u}.

  3. (3)

    (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda if and only if (𝒳1,𝒴1)2-𝗌𝗆𝖼Λ(\mathcal{X}_{1},\mathcal{Y}_{1})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda.

  4. (4)

    (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) is completable if and only if (𝒳1,𝒴1)(\mathcal{X}_{1},\mathcal{Y}_{1}) is completable.

  5. (5)

    If (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda, then 𝖳(μ𝒳(𝒳,𝒴))=𝖳(𝒳,𝒴)𝒳.\mathsf{T}(\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y}))=\mathsf{T}(\mathcal{X},\mathcal{Y})\cap\prescript{\perp}{}{\mathcal{X}^{\prime}}.

Similarly, let 𝒴𝒴\mathcal{Y}^{\prime}\subseteq\mathcal{Y}, and denote 𝒳2=μ𝒴(𝒳,𝒴)d\mathcal{X}_{2}=\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y})_{d} and 𝒴2=μ𝒴(𝒳,𝒴)u\mathcal{Y}_{2}=\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y})_{u}. Then the following hold.

  1. (1*)

    (𝒳2,𝒴2)(\mathcal{X}_{2},\mathcal{Y}_{2}) is a semibrick pair.

  2. (2*)

    If (𝒳2,𝒴2)(\mathcal{X}_{2}^{\prime},\mathcal{Y}_{2}^{\prime}) is an SM compatible semibrick pair such that 𝒳2𝒳2\mathcal{X}_{2}\subseteq\mathcal{X}^{\prime}_{2} and 𝒴2𝒴2\mathcal{Y}_{2}\subseteq\mathcal{Y}^{\prime}_{2}, then 𝒳μ𝒴(𝒳2,𝒴2)d\mathcal{X}\subseteq\mu_{\mathcal{Y}^{\prime}}(\mathcal{X}^{\prime}_{2},\mathcal{Y}^{\prime}_{2})_{d} and 𝒴μ𝒴(𝒳2,𝒴2)u\mathcal{Y}\subseteq\mu_{\mathcal{Y}^{\prime}}(\mathcal{X}^{\prime}_{2},\mathcal{Y}^{\prime}_{2})_{u}.

  3. (3*)

    (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda if and only if (𝒳2,𝒴2)2-𝗌𝗆𝖼Λ(\mathcal{X}_{2},\mathcal{Y}_{2})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda.

  4. (4*)

    (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) is completable if and only if (𝒳2,𝒴2)(\mathcal{X}_{2},\mathcal{Y}_{2}) is completable.

  5. (5*)

    If (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda, then 𝖳(μ𝒴(𝒳,𝒴))=𝖥𝗂𝗅𝗍(𝒴𝖳(𝒳,𝒴))\mathsf{T}(\mu_{\mathcal{Y}^{\prime}}(\mathcal{X},\mathcal{Y}))=\operatorname{\mathsf{Filt}}(\mathcal{Y}^{\prime}\cup\mathsf{T}(\mathcal{X},\mathcal{Y})).

5. Pop-Stack Operators for Torsion Classes

Recall the definition of the pop-stack and dual pop-stack operators from Section 1. For Λ\Lambda a τ\tau-tilting finite algebra, recall the description of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda} and pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda} from Corollary 3.6. In this section, we study further properties of these operators. In Section 5.1, we explain how the pop-stack operators interact with the mutation of 2-term simple-minded collection. In Section 5.2, we describe preimages under the pop-stack operators. In Section 5.3, we describe the images of the pop-stack operators.

5.1. Pop-stack and mutation

We now prove our first main result, which describes the relationship between the pop-stack operators and the mutation of 2-term simple-minded collections.

Theorem 5.1.

For (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda, let μ(𝒳,𝒴):=μ𝒳(𝒳,𝒴)\mu_{\downarrow}(\mathcal{X},\mathcal{Y}):=\mu_{\mathcal{X}}(\mathcal{X},\mathcal{Y}) and μ(𝒳,𝒴):=μ𝒴(𝒳,𝒴)\mu_{\uparrow}(\mathcal{X},\mathcal{Y}):=\mu_{\mathcal{Y}}(\mathcal{X},\mathcal{Y}). There are commutative diagrams as follows:

2-𝗌𝗆𝖼Λ{\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda}𝗍𝗈𝗋𝗌Λ{\operatorname{\mathsf{tors}}\Lambda}2-𝗌𝗆𝖼Λ{\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda}𝗍𝗈𝗋𝗌Λ{\operatorname{\mathsf{tors}}\Lambda}2-𝗌𝗆𝖼Λ{\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda}𝗍𝗈𝗋𝗌Λ{\operatorname{\mathsf{tors}}\Lambda}2-𝗌𝗆𝖼Λ{\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda}𝗍𝗈𝗋𝗌Λ.{\operatorname{\mathsf{tors}}\Lambda.}𝖳()\scriptstyle{\mathsf{T}(-)}μ()\scriptstyle{\mu_{\downarrow}(-)}pop()\scriptstyle{\mathrm{pop}^{\downarrow}(-)}𝖳()\scriptstyle{\mathsf{T}(-)}μ()\scriptstyle{\mu_{\uparrow}(-)}pop()\scriptstyle{\mathrm{pop}^{\uparrow}(-)}𝖳()\scriptstyle{\mathsf{T}(-)}𝖳()\scriptstyle{\mathsf{T}(-)}
Proof.

Let (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda. Then by Theorem 4.65 and Corollary 3.6, we have that

𝖳(μ(𝒳,𝒴))\displaystyle\mathsf{T}(\mu_{\downarrow}(\mathcal{X},\mathcal{Y})) =\displaystyle= 𝖳(𝒳)𝒳\displaystyle\mathsf{T}(\mathcal{X})\cap\prescript{\perp}{}{\mathcal{X}}
=\displaystyle= 𝖳(𝒳)({𝖳(𝒳)XX𝒳})\displaystyle\mathsf{T}(\mathcal{X})\cap\left(\bigcap\left\{\mathsf{T}(\mathcal{X})\cap\prescript{\perp}{}{X}\mid X\in\mathcal{X}\right\}\right)
=\displaystyle= pop(𝖳(𝒳,𝒴)).\displaystyle\mathrm{pop}^{\downarrow}(\mathsf{T}(\mathcal{X},\mathcal{Y})).

This shows that the left diagram commutes. Similarly, Theorem 4.65 and Corollary 3.6 imply that

𝖳(μ(𝒳,𝒴))\displaystyle\mathsf{T}(\mu_{\uparrow}(\mathcal{X},\mathcal{Y})) =\displaystyle= 𝖥𝗂𝗅𝗍(𝒴𝒴)\displaystyle\operatorname{\mathsf{Filt}}(\prescript{\perp}{}{\mathcal{Y}}\cup\mathcal{Y})
=\displaystyle= 𝖥𝗂𝗅𝗍(𝒴({𝖥𝗂𝗅𝗍(𝒴Y)Y𝒴}))\displaystyle\operatorname{\mathsf{Filt}}\left(\prescript{\perp}{}{\mathcal{Y}}\cup\left(\bigcup\left\{\operatorname{\mathsf{Filt}}(\prescript{\perp}{}{\mathcal{Y}}\cup Y)\mid Y\in\mathcal{Y}\right\}\right)\right)
=\displaystyle= pop(𝖳(𝒳,𝒴)).\displaystyle\mathrm{pop}^{\uparrow}(\mathsf{T}(\mathcal{X},\mathcal{Y})).

This shows that the right diagram commutes. ∎

Remark 5.2.

One can also generalize the proof of Theorem 5.1 to show that the torsion classes of the form 𝖳(μ𝒳(𝒳,𝒴))\mathsf{T}(\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})) (with 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}) are precisely those obtained by intersecting 𝖳(𝒳)\mathsf{T}(\mathcal{X}) with a subset of the torsion classes it covers. (The dual result likewise holds for right mutation.) In the language of [DL, DKW], these are precisely the torsion classes obtained by applying Ungar moves to 𝖳(𝒳)\mathsf{T}(\mathcal{X}). Intervals of the form [𝖳(μ𝒳(𝒳,𝒴)),𝖳(𝒳)][\mathsf{T}(\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})),\mathsf{T}(\mathcal{X})] have also appeared in many of the previously-cited papers on lattices of torsion classes such as [AP22, Hana].

5.2. Preimages under the pop-stack operators

We now characterize the preimages of a given torsion class under the pop-stack operator and its dual. As a consequence, we describe the 1-pop-stack sortable elements and the 2-pop-stack sortable elements.

Theorem 5.3.

Let 𝒯,𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T},\mathcal{T}^{\prime}\in\operatorname{\mathsf{tors}}\Lambda. Then pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝒯\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}^{\prime})=\mathcal{T} if and only if both of the following hold.

  1. (1)

    There is an inclusion 𝒟(𝒯)𝒰(𝒯)\mathcal{D}(\mathcal{T}^{\prime})\subseteq\mathcal{U}(\mathcal{T}).

  2. (2)

    For every X𝒟(𝒯)X\in\mathcal{D}(\mathcal{T}), there exists Z𝖥𝗂𝗅𝗍(𝒟(𝒯))Z\in\operatorname{\mathsf{Filt}}(\mathcal{D}(\mathcal{T}^{\prime})) admitting a surjection ZXZ\twoheadrightarrow X.

Moreover, if pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝒯\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}^{\prime})=\mathcal{T}, then (𝒟(𝒯),𝒰(𝒯))=μ𝒟(𝒯)(𝒟(𝒯),𝒰(𝒯))(\mathcal{D}(\mathcal{T}^{\prime}),\mathcal{U}(\mathcal{T}^{\prime}))=\mu_{\mathcal{D}(\mathcal{T}^{\prime})}(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T})).

Dually, pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝒯\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\mathcal{T}^{\prime} if and only if both of the following hold.

  1. (1*)

    There is an inclusion 𝒰(𝒯)𝒟(𝒯)\mathcal{U}(\mathcal{T})\subseteq\mathcal{D}(\mathcal{T}^{\prime}).

  2. (2*)

    For every Y𝒰(𝒯)Y\in\mathcal{U}(\mathcal{T}^{\prime}), there exists Z𝖥𝗂𝗅𝗍(𝒰(𝒯))Z\in\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T})) admitting an injection XZX\hookrightarrow Z.

Moreover, if pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝒯\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\mathcal{T}^{\prime}, then (𝒟(𝒯),𝒰(𝒯))=μ𝒰(𝒯)(𝒟(𝒯),𝒰(𝒯))(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T}))=\mu_{\mathcal{U}(\mathcal{T})}(\mathcal{D}(\mathcal{T}^{\prime}),\mathcal{U}(\mathcal{T}^{\prime})).

Proof.

We prove only the first half of the theorem; the proof of the second half is dual.

First suppose pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝒯\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}^{\prime})=\mathcal{T}. Then

(𝒟(𝒯),𝒰(𝒯))=μ(𝒟(𝒯),𝒰(𝒯))=μ𝒟(𝒯)(𝒟(𝒯),𝒰(𝒯))(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T}))=\mu_{\downarrow}(\mathcal{D}(\mathcal{T}^{\prime}),\mathcal{U}(\mathcal{T}^{\prime}))=\mu_{\mathcal{D}(\mathcal{T}^{\prime})}(\mathcal{D}(\mathcal{T}^{\prime}),\mathcal{U}(\mathcal{T}^{\prime}))

by Theorem 5.1, so in particular, 1 holds. It follows from Theorem 4.62 that

μ𝒟(𝒯)(𝒟(𝒯),𝒰(𝒯))=(𝒟(𝒯),𝒰(𝒯)).{\mu_{\mathcal{D}(\mathcal{T}^{\prime})}(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T}))=(\mathcal{D}(\mathcal{T}^{\prime}),\mathcal{U}(\mathcal{T}^{\prime}))}.

Finally, because μ𝒟(𝒯)(𝒟(𝒯),𝒰(𝒯))=(𝒟(𝒯),𝒰(𝒯))\mu_{\mathcal{D}(\mathcal{T}^{\prime})}(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T}))=(\mathcal{D}(\mathcal{T}^{\prime}),\mathcal{U}(\mathcal{T}^{\prime})), 2 must hold by Definition 4.4.

Suppose now that 1 and 2 both hold. Then μ𝒟(𝒯)(𝒟(𝒯),𝒰(𝒯))d=𝒟(𝒯)\mu_{\mathcal{D}(\mathcal{T}^{\prime})}(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T}))_{d}=\mathcal{D}(\mathcal{T}^{\prime}) by Definition 4.4, and therefore 𝖳(μ𝒟(𝒯)(𝒟(𝒯),𝒰(𝒯)))=𝒯\mathsf{T}(\mu_{\mathcal{D}(\mathcal{T}^{\prime})}(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T})))=\mathcal{T}^{\prime}. Theorem 4.62 then implies that

μ(𝒟(𝒯),𝒰(𝒯))=μ𝒟(𝒯)(𝒟(𝒯),𝒰(𝒯))=(𝒟(𝒯),𝒰(𝒯)).\mu_{\downarrow}(\mathcal{D}(\mathcal{T}^{\prime}),\mathcal{U}(\mathcal{T}^{\prime}))=\mu_{\mathcal{D}(\mathcal{T}^{\prime})}(\mathcal{D}(\mathcal{T}^{\prime}),\mathcal{U}(\mathcal{T}^{\prime}))=(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T})).

We conclude that pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝒯\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}^{\prime})=\mathcal{T} by Theorem 5.1. ∎

We now characterize the 1-pop-stack sortable elements of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda.

Corollary 5.4.

Let 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda. The following are equivalent.

  1. (1)

    We have pop𝗍𝗈𝗋𝗌Λ(𝒯)=0^\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\hat{0}.

  2. (2)

    Every brick in 𝒟(𝒯)\mathcal{D}(\mathcal{T}) is simple.

  3. (3)

    The torsion class 𝒯\mathcal{T} is a Serre subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda.

  4. (4)

    There exists a nonzero injective module II such that 𝒯=I\mathcal{T}=\prescript{\perp}{}{I}.

Dually, the following are equivalent.

  1. (1*)

    We have pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝗆𝗈𝖽Λ\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\operatorname{\mathsf{mod}}\Lambda.

  2. (2*)

    Every brick in 𝒰(𝒯)\mathcal{U}(\mathcal{T}) is simple.

  3. (3*)

    The torsion-free class 𝒯\mathcal{T}^{\perp} is a Serre subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda.

  4. (4*)

    There exists a nonzero projective module PP such that 𝒯=𝖦𝖾𝗇(P)\mathcal{T}=\operatorname{\mathsf{Gen}}(P).

Proof.

We prove only the first result as the second is dual. The equivalences among 2, 3, and 4 are contained in Proposition 3.10.

To see that 1 implies 2, suppose pop𝗍𝗈𝗋𝗌Λ(𝒯)=0^\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\hat{0}. Then every brick in 𝒰(pop𝗍𝗈𝗋𝗌Λ(𝒯))\mathcal{U}(\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})) is simple. But 𝒟(𝒯)𝒰(pop𝗍𝗈𝗋𝗌Λ(𝒯))\mathcal{D}(\mathcal{T}^{\prime})\subseteq\mathcal{U}(\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})) by Theorem 5.3.

We now show that 3 implies 1. Suppose 𝒯\mathcal{T} is a Serre subcategory. Then in particular, 𝒯\mathcal{T} is closed under submodules. Thus, for Y𝒰(𝒯)Y\in\mathcal{U}(\mathcal{T}), the map gY,𝒳g_{Y,\mathcal{X}} must be surjective. Theorem 5.1 then implies that 𝒟(pop𝗍𝗈𝗋𝗌Λ(𝒯))=\mathcal{D}(\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))=\emptyset, so pop𝗍𝗈𝗋𝗌Λ(𝒯)=0\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=0. ∎

The following characterizes all 2-pop-stack sortable elements of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda; it also describes the image under pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda} of each 2-pop-stack sortable element.

Corollary 5.5.

Let 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda, and let 𝒮\mathcal{S} be a set of simple modules. Then pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝖥𝗂𝗅𝗍(SS)\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\operatorname{\mathsf{Filt}}(\SS) if and only if the following all hold.

  1. (1)

    We have HomΛ(SS,𝒟(𝒯))=0=ExtΛ1(SS,𝒟(𝒯))\mathrm{Hom}_{\Lambda}(\SS,\mathcal{D}(\mathcal{T}))=0=\mathrm{Ext}^{1}_{\Lambda}(\SS,\mathcal{D}(\mathcal{T})).

  2. (2)

    The socle222Recall that the socle socM\mathrm{soc}M of a module MM is the sum of all of its semisimple submodules. soc(𝒟(𝒯))\mathrm{soc}(\mathcal{D}(\mathcal{T})) of 𝒟(𝒯)\mathcal{D}(\mathcal{T}) satisfies 𝒟(𝒯)/soc(𝒟(𝒯))𝖥𝗂𝗅𝗍(SS)\mathcal{D}(\mathcal{T})/\mathrm{soc}(\mathcal{D}(\mathcal{T}))\in\operatorname{\mathsf{Filt}}(\SS).

  3. (3)

    There does not exist SSSS\in\SS with HomΛ(𝒟(𝒯),S)=0\mathrm{Hom}_{\Lambda}(\mathcal{D}(\mathcal{T}),S)=0. (Equivalently, there is a surjection 𝒟(𝒯)SS\mathcal{D}(\mathcal{T})\twoheadrightarrow\SS.)

Proof.

Suppose first that pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝖥𝗂𝗅𝗍(SS)\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\operatorname{\mathsf{Filt}}(\SS). Since 𝒟(𝖥𝗂𝗅𝗍(SS))=SS\mathcal{D}(\operatorname{\mathsf{Filt}}(\SS))=\SS, it follows from Theorem 5.3 that 𝒟(𝒯)𝒰(𝖥𝗂𝗅𝗍(SS))\mathcal{D}(\mathcal{T})\subseteq\mathcal{U}(\operatorname{\mathsf{Filt}}(\SS)). Then the fact that 1 holds follows from the definition of a 2-term simple-minded collection. Theorem 5.3 also implies that for every SSSS\in\SS, there exists Z𝖥𝗂𝗅𝗍(𝒟(𝒯))Z\in\operatorname{\mathsf{Filt}}(\mathcal{D}(\mathcal{T})) that admits a surjection ZSZ\twoheadrightarrow S. This shows that 3 holds. To prove 2, consider X𝒟(𝒯)𝒰(𝖥𝗂𝗅𝗍(SS)){X\in\mathcal{D}(\mathcal{T})\subseteq\mathcal{U}(\operatorname{\mathsf{Filt}}(\SS))}. Since 𝖥𝗂𝗅𝗍(SS)\operatorname{\mathsf{Filt}}(\SS) is closed under submodules, the map gX,SS:XSXg_{X,\SS}\colon X\rightarrow S_{X} must be surjective. Moreover, we have that pop𝗍𝗈𝗋𝗌Λ(𝖥𝗂𝗅𝗍(SS))=0\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\operatorname{\mathsf{Filt}}(\SS))=0 by Corollary 5.4. Theorem 5.1 implies that kergX,SS\mathrm{ker}g_{X,\SS} is in 𝒰(0)\mathcal{U}(0), which is precisely the set of simple modules. If follows that kergX,SSsoc(X)\mathrm{ker}g_{X,\SS}\subseteq\mathrm{soc}(X) and therefore that X/soc(X)𝖥𝗂𝗅𝗍(SS)X/\mathrm{soc}(X)\in\operatorname{\mathsf{Filt}}(\SS).

To prove the converse, suppose 𝒯\mathcal{T} satisfies Items 1, 2 and 3. Then 1 implies that (SS,𝒟(𝒯))(\SS,\mathcal{D}(\mathcal{T})) is a semibrick pair. Moreover, for X𝒟(𝒯)X\in\mathcal{D}(\mathcal{T}), the map gX,SSg_{X,\SS} must be surjective since 𝖥𝗂𝗅𝗍(SS)\operatorname{\mathsf{Filt}}(\SS) is closed under submodules. This map must also have a semisimple kernel by 2. We then have that μSS(SS,𝒟(𝒯))=(,SSSS)\mu_{\SS}(\SS,\mathcal{D}(\mathcal{T}))=(\emptyset,\SS\cup\SS^{\prime}) for SS={kergX,SSX𝒟(𝒯)}\SS^{\prime}=\{\mathrm{ker}g_{X,\SS}\mid X\in\mathcal{D}(\mathcal{T})\}. Proposition A.3 therefore implies that kergX,SS\mathrm{ker}g_{X,\SS} is simple for all XX. Letting SS′′\SS^{\prime\prime} denote the set of all simple modules, we find that (,SS′′)(\emptyset,\SS^{\prime\prime}) is a 2-term simple-minded collection satisfying SSSSSS′′\SS\cup\SS^{\prime}\subseteq\SS^{\prime\prime}. By Proposition A.4, (SS,𝒴):=μSS(,SS′′)(\SS,\mathcal{Y}):=\mu_{\SS}(\emptyset,\SS^{\prime\prime}) is a 2-term simple-minded collection satisfying 𝒟(𝒯)𝒴\mathcal{D}(\mathcal{T})\subseteq\mathcal{Y}. Now note that 𝒴=𝒰(𝖥𝗂𝗅𝗍(SS))\mathcal{Y}=\mathcal{U}(\operatorname{\mathsf{Filt}}(\SS)) by Theorem 3.7. Hence, 𝒟(𝒯)𝒰(𝖥𝗂𝗅𝗍(SS))\mathcal{D}(\mathcal{T})\subseteq\mathcal{U}(\operatorname{\mathsf{Filt}}(\SS)). Together with 3, this allows us to use Theorem 5.3 to conclude that pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝖥𝗂𝗅𝗍(SS)\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\operatorname{\mathsf{Filt}}(\SS). ∎

5.3. The image of pop-stack

We now build toward a complete description of the images of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda} and pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda} (and, hence, the facets of the canonical join complex of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda) for any τ\tau-tilting finite algebra Λ\Lambda. Our first result is proven more generally for arbitrary semidistrim lattices in [DW23, Section 9]. We give a new representation-theoretic proof for lattices of torsion classes.

Proposition 5.6.

Let 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda.

  1. (1)

    We have pop𝗍𝗈𝗋𝗌Λ(pop𝗍𝗈𝗋𝗌Λ(𝒯))𝒯\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))\subseteq\mathcal{T}.

  2. (2)

    We have pop𝗍𝗈𝗋𝗌Λ(pop𝗍𝗈𝗋𝗌Λ(𝒯))𝒯\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))\supseteq\mathcal{T}.

  3. (3)

    We have pop𝗍𝗈𝗋𝗌Λ(pop𝗍𝗈𝗋𝗌Λ(pop𝗍𝗈𝗋𝗌Λ(𝒯)))=pop𝗍𝗈𝗋𝗌Λ(𝒯)\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})))=\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}).

  4. (4)

    We have pop𝗍𝗈𝗋𝗌Λ(pop𝗍𝗈𝗋𝗌Λ(pop𝗍𝗈𝗋𝗌Λ(𝒯)))=pop𝗍𝗈𝗋𝗌Λ(𝒯)\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})))=\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}).

Proof.

We prove only 1 and 3 as the proofs of 2 and 4 are similar.

We have 𝒰(𝒯)𝒟(pop𝗍𝗈𝗋𝗌Λ(𝒯))𝒰(pop𝗍𝗈𝗋𝗌Λ(pop𝗍𝗈𝗋𝗌Λ(𝒯)))\mathcal{U}(\mathcal{T})\subseteq\mathcal{D}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))\subseteq\mathcal{U}(\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))) by Theorem 5.1. By Theorem 3.4, this implies 1.

To prove 3, let (𝒳,𝒴)=(𝒟(𝒯),𝒰(𝒯))(\mathcal{X},\mathcal{Y})=(\mathcal{D}(\mathcal{T}),\mathcal{U}(\mathcal{T})). Let (𝒳1,𝒴1)=μ(𝒳,𝒴)(\mathcal{X}_{1},\mathcal{Y}_{1})=\mu_{\downarrow}(\mathcal{X},\mathcal{Y}) and (𝒳2,𝒴2)=μ(𝒳1,𝒴1)(\mathcal{X}_{2},\mathcal{Y}_{2})=\mu_{\uparrow}(\mathcal{X}_{1},\mathcal{Y}_{1}). By Theorem 5.1, we need to show that μ(𝒳2,𝒴2)=(𝒳1,𝒴1)\mu_{\downarrow}(\mathcal{X}_{2},\mathcal{Y}_{2})=(\mathcal{X}_{1},\mathcal{Y}_{1}). By Definition 4.4, this is equivalent to showing that 𝒳2=𝒴1\mathcal{X}_{2}=\mathcal{Y}_{1}, which is equivalent to showing that g𝒴1,Zg_{\mathcal{Y}_{1},Z} is surjective for every Z𝒳1Z\in\mathcal{X}_{1}.

Now recall that 𝒳1={coker(gY,𝒳)Y𝒴 and gY,𝒳 is injective}\mathcal{X}_{1}=\{\mathrm{coker}(g_{Y,\mathcal{X}})\mid Y\in\mathcal{Y}\text{ and }g_{Y,\mathcal{X}}\text{ is injective}\}. Thus, for Z𝒳1Z\in\mathcal{X}_{1}, we have a quotient map XYZX_{Y}\twoheadrightarrow Z for some XY𝖥𝗂𝗅𝗍(𝒳)X_{Y}\in\operatorname{\mathsf{Filt}}(\mathcal{X}). Since 𝒳𝒴1\mathcal{X}\subseteq\mathcal{Y}_{1}, this map must factor through the minimal right 𝖥𝗂𝗅𝗍(𝒴1)\operatorname{\mathsf{Filt}}(\mathcal{Y}_{1})-approximation g𝒴1,Z:(Y1)ZZg_{\mathcal{Y}_{1},Z}\colon(Y_{1})_{Z}\rightarrow Z. This implies that g𝒴1,Zg_{\mathcal{Y}_{1},Z} must be surjective. ∎

As a consequence, we can characterize the image of the pop-stack operators as follows.

Corollary 5.7.

Let 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda. Then the following are equivalent.

  1. (1)

    There exists 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}^{\prime}\in\operatorname{\mathsf{tors}}\Lambda such that 𝒯=pop𝗍𝗈𝗋𝗌Λ(𝒯)\mathcal{T}=\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}^{\prime}).

  2. (2)

    We have pop𝗍𝗈𝗋𝗌Λ(pop𝗍𝗈𝗋𝗌Λ(𝒯))=𝒯\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))=\mathcal{T}.

  3. (3)

    The semibricks 𝒰(𝒯)\mathcal{U}(\mathcal{T}) and 𝒟(pop𝗍𝗈𝗋𝗌Λ(𝒯))\mathcal{D}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})) coincide.

  4. (4)

    For all X𝒟(𝒯)X\in\mathcal{D}(\mathcal{T}), the map g𝒰(𝒯),Xg_{\mathcal{U}(\mathcal{T}),X} is surjective.

  5. (5)

    For all X𝒟(𝒯)X\in\mathcal{D}(\mathcal{T}), there exists Z𝖥𝗂𝗅𝗍(𝒰(𝒯))Z\in\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T})) such that there is a surjective map ZXZ\twoheadrightarrow X.

Dually, the following are equivalent.

  1. (1*)

    There exists 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}^{\prime}\in\operatorname{\mathsf{tors}}\Lambda such that 𝒯=pop𝗍𝗈𝗋𝗌Λ(𝒯)\mathcal{T}=\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}^{\prime}).

  2. (2*)

    We have pop𝗍𝗈𝗋𝗌Λ(pop𝗍𝗈𝗋𝗌Λ(𝒯))=𝒯\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))=\mathcal{T}.

  3. (3*)

    The semibricks 𝒟(𝒯)\mathcal{D}(\mathcal{T}) and 𝒰(pop𝗍𝗈𝗋𝗌Λ(𝒯))\mathcal{U}(\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})) coincide.

  4. (4*)

    For all Y𝒰(𝒯)Y\in\mathcal{U}(\mathcal{T}), the map gY,𝒟(𝒯)g_{Y,\mathcal{D}(\mathcal{T})} is injective.

  5. (5*)

    For all Y𝒰(𝒯)Y\in\mathcal{U}(\mathcal{T}), there exists Z𝖥𝗂𝗅𝗍(𝒟(𝒯))Z\in\operatorname{\mathsf{Filt}}(\mathcal{D}(\mathcal{T})) such that there is an injective map YZY\hookrightarrow Z.

We now wish to further characterize the image of the pop-stack operators for representation-finite hereditary algebras. We first prove the following result (still in the generality of an arbitrary τ\tau-tilting finite algebra).

Lemma 5.8.

Let 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda.

  1. (1)

    If there exists a nonzero projective module P𝒯P\in\mathcal{T}, then 𝒯\mathcal{T} is not in the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}.

  2. (2)

    If there exists a nonzero injective module I𝒯I\in\mathcal{T}^{\perp}, then 𝒯\mathcal{T} is not in the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}.

Proof.

We prove only 1 since 2 is dual. It suffices to consider the case where PP is indecomposable.

We claim that there exists X𝒟(𝒯)X\in\mathcal{D}(\mathcal{T}) such that there is a surjection PXP\twoheadrightarrow X and there does not exist Z𝖥𝗂𝗅𝗍(𝒰(𝒯))Z\in\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T})) admitting a surjection ZXZ\twoheadrightarrow X. We prove the claim by induction on the length of the shortest path from 𝖦𝖾𝗇(P)\operatorname{\mathsf{Gen}}(P) to 𝒯\mathcal{T} in 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda.

For the base case, we consider 𝒯=𝖦𝖾𝗇(P)\mathcal{T}=\operatorname{\mathsf{Gen}}(P). In this case, there is a short exact sequence

0MP𝑓X00\rightarrow M\rightarrow P\xrightarrow{f}X\rightarrow 0

such that M𝒯M\in\mathcal{T} and 𝒟(𝒯)=X\mathcal{D}(\mathcal{T})=X. (If PP is a brick, then X=PX=P and M=0M=0. Otherwise, MM is the sum of the images of the noninvertable endomorphisms of PP.) Now suppose for a contradiction that there exist Z𝖥𝗂𝗅𝗍(𝒰(𝒯))Z\in\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T})) and a surjection Z𝑔XZ\xrightarrow{g}X. Then by the definition of projective, there is a nonzero map h:PZh\colon P\rightarrow Z such that f=ghf=g\circ h. But Z𝖥𝗂𝗅𝗍(𝒰(𝒯))PZ\in\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T}))\subseteq P^{\perp}, which is a contradiction.

For the induction step, suppose 𝖦𝖾𝗇(P)𝒯\operatorname{\mathsf{Gen}}(P)\subsetneq\mathcal{T}, and choose some 𝒯<𝒯\mathcal{T}^{\prime}{\,\,<\!\!\!\!\cdot\,\,\,}\mathcal{T} such that 𝖦𝖾𝗇(P)𝒯\operatorname{\mathsf{Gen}}(P)\subseteq\mathcal{T}^{\prime}. Let Y𝒰(𝒯)Y\in\mathcal{U}(\mathcal{T}^{\prime}) be the brick label of the cover relation 𝒯<𝒯\mathcal{T}^{\prime}{\,\,<\!\!\!\!\cdot\,\,\,}\mathcal{T}. By the induction hypothesis, there exists X𝒟(𝒯)X\in\mathcal{D}(\mathcal{T}^{\prime}) such that there is a surjection PXP\twoheadrightarrow X and there does not exist Z𝖥𝗂𝗅𝗍(𝒰(𝒯))Z\in\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T}^{\prime})) admitting a surjection ZXZ\twoheadrightarrow X. Let gY,X:YXXg_{Y,X}\colon Y_{X}\rightarrow X be a minimal right 𝖥𝗂𝗅𝗍(Y)\operatorname{\mathsf{Filt}}(Y)-approximation of XX. Then gY,Xg_{Y,X} is injective since there is no surjection from an object in 𝖥𝗂𝗅𝗍(𝒰(𝒯))\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T}^{\prime})) onto XX. Thus, coker(gY,X)𝒟(𝒯)\mathrm{coker}(g_{Y,X})\in\mathcal{D}(\mathcal{T}) by Theorem 4.6. Moreover, coker(gY,X)\mathrm{coker}(g_{Y,X}) is a quotient of PP since XX is, and we have that 𝒯P\mathcal{T}^{\perp}\subseteq P^{\perp}. Consequently, no map from an object in 𝖥𝗂𝗅𝗍(𝒰(𝒯))\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T})) to coker(gY,X)\mathrm{coker}(g_{Y,X}) can be surjective for the same reason as in the base case.

Given the claim, the result follows from Corollary 5.7. ∎

Note that a torsion class 𝒯\mathcal{T} contains a projective module PP if and only if 𝖦𝖾𝗇(P)𝒯\operatorname{\mathsf{Gen}}(P)\subseteq\mathcal{T}. Moreover, recall that the atoms of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda are precisely the torsion classes of the form 𝖥𝗂𝗅𝗍(S)\operatorname{\mathsf{Filt}}(S) for S𝗆𝗈𝖽ΛS\in\operatorname{\mathsf{mod}}\Lambda simple. We can thus give a combinatorial interpretation of what it means for a torsion class to contain a projective module as follows.

Proposition 5.9.

For 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda, the following are equivalent.

  1. (1)

    There exists a projective module PP such that 𝒯=𝖦𝖾𝗇(P)\mathcal{T}=\operatorname{\mathsf{Gen}}(P).

  2. (2)

    There exists a set 𝒜\mathcal{A} of atoms of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda such that 𝒯\mathcal{T} is the maximal element of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda that lies (weakly) above all of the atoms in 𝒜\mathcal{A} and does not lie (weakly) above any of the atoms that are not in 𝒜\mathcal{A}.

  3. (3)

    There exists a set 𝒮\mathcal{S} of simple modules such that 𝒯\mathcal{T} is the largest torsion class that contains all of the modules in 𝒮\mathcal{S} and does not contain any simple module that is not in 𝒮\mathcal{S}.

When these conditions hold, the semibrick 𝒮\mathcal{S} is the top of the module PP (equivalently, PP is the projective cover of 𝒮\mathcal{S}); in particular, P=0P=0 if and only if 𝒜==𝒮\mathcal{A}=\emptyset=\mathcal{S}.

Proof.

The equivalence of 2 and 3 follows from the description of the atoms of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda.

Let us prove that 1 implies 3. Let PP be a projective module, and let SS\SS be the set of simple modules that are direct summands of top(P)\mathrm{top}(P). Then SS\SS is a quotient of PP, so 𝖥𝗂𝗅𝗍(SS)𝖦𝖾𝗇(P)\operatorname{\mathsf{Filt}}(\SS)\subseteq\operatorname{\mathsf{Gen}}(P). Moreover, for a simple SSSS\notin\SS, we have HomΛ(P,S)=0\mathrm{Hom}_{\Lambda}(P,S)=0, so S𝖦𝖾𝗇(P)S\notin\operatorname{\mathsf{Gen}}(P). It remains to show that 𝖦𝖾𝗇(P)\operatorname{\mathsf{Gen}}(P) is maximal with respect to these properties. Let 𝒯\mathcal{T} be another torsion class satisfying SS𝒯\SS\in\mathcal{T} and S𝒯S\notin\mathcal{T} for any simple SSSS\notin\SS. Then top(𝒟(𝒯))𝖺𝖽𝖽(SS)𝖥𝗂𝗅𝗍(SS)\mathrm{top}(\mathcal{D}(\mathcal{T}))\in\operatorname{\mathsf{add}}(\SS)\subseteq\operatorname{\mathsf{Filt}}(\SS). Thus, the projective cover of 𝒟(𝒯)\mathcal{D}(\mathcal{T}) lies in 𝖺𝖽𝖽(P)\operatorname{\mathsf{add}}(P), so 𝒟(𝒯)𝖦𝖾𝗇(P)\mathcal{D}(\mathcal{T})\in\operatorname{\mathsf{Gen}}(P). It follows that 𝒯𝖦𝖾𝗇(P)\mathcal{T}\subseteq\operatorname{\mathsf{Gen}}(P).

We now show that 3 implies 1. Let SS\SS be a set of simple modules, and let 𝒯\mathcal{T} be the largest torsion class of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda that contains SS\SS and does not contain any simple that does not lie in SS\SS. Then, as shown above, we have 𝒯=𝖦𝖾𝗇(P)\mathcal{T}=\operatorname{\mathsf{Gen}}(P), where PP is the projective cover of SS\SS. ∎

We now give a characterization of the image of the pop-stack operator when Λ\Lambda is hereditary (and representation-finite). We give a combinatorial interpretation of this characterization in Theorem 7.8.

Theorem 5.10.

Suppose Λ\Lambda is hereditary, and let 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda. Then 𝒯\mathcal{T} is in the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda} if and only if both of the following hold:

  1. (1)

    ExtΛ1(X,X)=0\mathrm{Ext}^{1}_{\Lambda}(X,X^{\prime})=0 for all X,X𝒟(𝒯)X,X^{\prime}\in\mathcal{D}(\mathcal{T}).

  2. (2)

    There does not exist a nonzero projective module P𝒯P\in\mathcal{T}.

Dually, 𝒯\mathcal{T} is in the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda} if and only if both of the following hold.

  1. (1*)

    ExtΛ1(Y,Y)=0\mathrm{Ext}^{1}_{\Lambda}(Y,Y^{\prime})=0 for all Y,Y𝒰(𝒯)Y,Y^{\prime}\in\mathcal{U}(\mathcal{T}).

  2. (2*)

    There does not exist a nonzero injective module I𝒯I\in\mathcal{T}^{\perp}.

Proof.

We prove only the first statement since the second is dual. We first note that 2 is necessary by Lemma 5.8.

Suppose that 1 holds and that 𝒯\mathcal{T} is not in the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}. By Corollary 5.7, this means there exists X𝒟(𝒯)X\in\mathcal{D}(\mathcal{T}) such that g𝒰(𝒯),Xg_{\mathcal{U}(\mathcal{T}),X} is injective. We will prove that XX is projective by induction on |𝒟(𝒯)||\mathcal{D}(\mathcal{T})|. First suppose |𝒟(𝒯)|=1|\mathcal{D}(\mathcal{T})|=1. Then Theorem 5.1 implies that |𝒟(pop𝗍𝗈𝗋𝗌Λ(𝒯))|=|Λ|{|\mathcal{D}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))|=|\Lambda|}. Now, |𝒟(pop𝗍𝗈𝗋𝗌Λ(𝒯))|+|𝒰(pop𝗍𝗈𝗋𝗌Λ(𝒯))|=|Λ||\mathcal{D}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))|+|\mathcal{U}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))|=|\Lambda| by Theorem 3.7 and [KY14, Corollary 5.5]333This is a classical and well-known fact that was proved much earlier than [KY14], but this logical deduction suits the organization of this paper., so pop𝗍𝗈𝗋𝗌Λ(𝒯)=𝗆𝗈𝖽Λ\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=\operatorname{\mathsf{mod}}\Lambda. In particular, 𝒰(𝒯)𝒟(pop𝗍𝗈𝗋𝗌Λ(𝒯))\mathcal{U}(\mathcal{T})\subseteq\mathcal{D}(\mathrm{pop}^{\uparrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})) consists of only simple modules, so 𝒯\mathcal{T}^{\perp} is a Serre subcategory. By Proposition 3.10, this means 𝒯=X=P\mathcal{T}^{\perp}=X^{\perp}=P^{\perp} for some projective module PP. It is straightforward to show that X=PX=P must be the projective cover of the unique simple module that does not lie in 𝒰(𝒯)\mathcal{U}(\mathcal{T}).

For the induction step, suppose that the result holds for |𝒟(𝒯)|=k|\mathcal{D}(\mathcal{T})|=k and that |𝒟(𝒯)|=k+1|\mathcal{D}(\mathcal{T})|=k+1. Let XZ𝒟(𝒯)X\neq Z\in\mathcal{D}(\mathcal{T}), and let 𝒟:=𝒟(𝒯){Z}\mathcal{D}^{\prime}:=\mathcal{D}(\mathcal{T})\setminus\{Z\}. By the assumption of 2, we have that 𝒟𝒰(𝒯)Z0,1\mathcal{D}^{\prime}\cup\mathcal{U}(\mathcal{T})\subseteq Z^{\perp_{0,1}}. Furthermore, since Λ\Lambda is hereditary, Z0,1Z^{\perp_{0,1}} is a wide subcategory444The perpendicular subcategory Z0,1Z^{\perp_{0,1}} was first considered in [GL91].. Consequently, 𝒯:=Z𝒯\mathcal{T}^{\prime}:=Z^{\perp}\cap\mathcal{T} is a torsion class of Z0,1Z^{\perp_{0,1}} that satisfies 𝒟(𝒯)=𝒟\mathcal{D}(\mathcal{T}^{\prime})=\mathcal{D}^{\prime} and 𝒰(𝒯)=𝒰(𝒯)\mathcal{U}(\mathcal{T}^{\prime})=\mathcal{U}(\mathcal{T}). In particular, 1 holds for the torsion class 𝒯\mathcal{T}^{\prime}, and the injective map gX,𝒰(𝒯)g_{X,\mathcal{U}(\mathcal{T})} is a minimal right 𝖥𝗂𝗅𝗍(𝒰(𝒯))\operatorname{\mathsf{Filt}}(\mathcal{U}(\mathcal{T}^{\prime}))-approximation. By the induction hypothesis, we conclude that XX is projective in Z0,1Z^{\perp_{0,1}}; i.e., ExtΛ1(X,M)=0\mathrm{Ext}^{1}_{\Lambda}(X,M)=0 for all MZ0,1M\in Z^{\perp_{0,1}}.

Recall from Section 3.2 that every M𝗆𝗈𝖽ΛM\in\operatorname{\mathsf{mod}}\Lambda admits a short exact sequence t𝒯MMfM{t_{\mathcal{T}}M\hookrightarrow M\twoheadrightarrow f_{\mathcal{F}}M} with t𝒯M𝒯t_{\mathcal{T}}M\in\mathcal{T} and fM𝒯=𝖥𝗂𝗅𝗍(𝖢𝗈𝗀𝖾𝗇(𝒰(𝒯)))f_{\mathcal{F}}M\in\mathcal{T}^{\perp}=\operatorname{\mathsf{Filt}}(\operatorname{\mathsf{Cogen}}(\mathcal{U}(\mathcal{T}))). The module t𝒯Mt_{\mathcal{T}}M then fits into a short exact sequence t𝖦𝖾𝗇(Z)(t𝒯M)t𝒯MfZ(t𝒯M)t_{\operatorname{\mathsf{Gen}}(Z)}(t_{\mathcal{T}}M)\hookrightarrow t_{\mathcal{T}}M\twoheadrightarrow f_{Z^{\perp}}(t_{\mathcal{T}}M) with t𝖦𝖾𝗇(Z)(t𝒯M)𝖦𝖾𝗇(Z)t_{\operatorname{\mathsf{Gen}}(Z)}(t_{\mathcal{T}}M)\in\operatorname{\mathsf{Gen}}(Z) and fZ(t𝒯M)Zf_{Z^{\perp}}(t_{\mathcal{T}}M)\in Z^{\perp}. Note that fZ(t𝒯M)f_{Z^{\perp}}(t_{\mathcal{T}}M) is in 𝒯\mathcal{T}^{\prime} since it is a quotient of t𝒯Mt_{\mathcal{T}}M and t𝒯M𝒯t_{\mathcal{T}}M\in\mathcal{T}. By the above paragraph, we have ExtΛ1(X,fZ(t𝒯M))=0\mathrm{Ext}^{1}_{\Lambda}(X,f_{Z^{\perp}}(t_{\mathcal{T}}M))=0. Similarly, ExtΛ1(X,t𝖦𝖾𝗇(Z)(t𝒯M))=0\mathrm{Ext}^{1}_{\Lambda}(X,t_{\operatorname{\mathsf{Gen}}(Z)}(t_{\mathcal{T}}M))=0 since ExtΛ1(X,Z)=0\mathrm{Ext}^{1}_{\Lambda}(X,Z)=0 (by assumption) and Λ\Lambda is hereditary. Thus, to show that XX is projective in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda, it suffices to show that ExtΛ1(X,N)=0\mathrm{Ext}^{1}_{\Lambda}(X,N)=0 for all N𝖢𝗈𝗀𝖾𝗇(𝒰(𝒯))N\in\operatorname{\mathsf{Cogen}}(\mathcal{U}(\mathcal{T})). Let Y𝒰(𝒯)Y\in\mathcal{U}(\mathcal{T}), and let NYN\subseteq Y be a submodule. Then NZN\in Z^{\perp}. If ExtΛ1(Z,N)=0\mathrm{Ext}^{1}_{\Lambda}(Z,N)=0, we already know that ExtΛ1(X,N)=0\mathrm{Ext}^{1}_{\Lambda}(X,N)=0. Otherwise, the fact that ExtΛ1(Z,Z)=0\mathrm{Ext}^{1}_{\Lambda}(Z,Z)=0 and HomΛ(Z,N)=0\mathrm{Hom}_{\Lambda}(Z,N)=0 means that there exist Zk𝖺𝖽𝖽(Z)Z^{k}\in\operatorname{\mathsf{add}}(Z) and a short exact sequence

0NEZk00\rightarrow N\rightarrow E\rightarrow Z^{k}\rightarrow 0

such that the induced map HomΛ(Z,Zk)ExtΛ1(Z,N)\mathrm{Hom}_{\Lambda}(Z,Z^{k})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,N) is surjective. If we take kk to be minimal with respect to this property, then there is an induced long exact sequence

0=HomΛ(Z,E)HomΛ(Z,Zk)ExtΛ1(Z,N)ExtΛ1(Z,E)ExtΛ1(Z,Z)=0.0=\mathrm{Hom}_{\Lambda}(Z,E)\rightarrow\mathrm{Hom}_{\Lambda}(Z,Z^{k})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,N)\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,E)\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,Z)=0.

(We know that the first term is 0 because ZZ is a brick and HomΛ(Z,N)=0\mathrm{Hom}_{\Lambda}(Z,N)=0, so any nonzero morphism ZEZ\rightarrow E would have trivial intersection with NN and would consequently be a splitting of a composition EZkZE\twoheadrightarrow Z^{k}\twoheadrightarrow Z.) It follows that EZ0,1E\in Z^{\perp_{0,1}} and therefore that ExtΛ1(X,E)=0\mathrm{Ext}^{1}_{\Lambda}(X,E)=0 (since XX is projective in that wide subcategory). Hence, there is an induced exact sequence

0=HomΛ(X,Zk)ExtΛ1(X,N)ExtΛ1(X,E)=0.0=\mathrm{Hom}_{\Lambda}(X,Z^{k})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(X,N)\rightarrow\mathrm{Ext}^{1}_{\Lambda}(X,E)=0.

We conclude that ExtΛ1(X,N)=0\mathrm{Ext}^{1}_{\Lambda}(X,N)=0. Since 𝖢𝗈𝗀𝖾𝗇(𝒰(𝒯))\operatorname{\mathsf{Cogen}}(\mathcal{U}(\mathcal{T})) is the closure under 𝖥𝗂𝗅𝗍\operatorname{\mathsf{Filt}} of all indecomposable quotients of the bricks in 𝒰(𝒯)\mathcal{U}(\mathcal{T}), this proves that XX is projective in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda.

It remains to show that if 𝒯\mathcal{T} is in the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}, then 1 holds. To see this, suppose that 𝒯\mathcal{T} is in the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}, and let X1,X2𝒟(𝒯)X_{1},X_{2}\in\mathcal{D}(\mathcal{T}). If X1=X2X_{1}=X_{2}, then there is nothing to show since all bricks over Λ\Lambda are also rigid. Otherwise, by Corollary 5.7, the map g𝒰(𝒯),X2:UX2X2g_{\mathcal{U}(\mathcal{T}),X_{2}}:U_{X_{2}}\rightarrow X_{2} must be injective. Therefore, there is an induced exact sequence

ExtΛ1(X1,ker(g𝒰(𝒯),X2))ExtΛ1(X1,UX2)ExtΛ1(X1,X2)0,\mathrm{Ext}^{1}_{\Lambda}(X_{1},\mathrm{ker}(g_{\mathcal{U}(\mathcal{T}),X_{2}}))\rightarrow\mathrm{Ext}^{1}_{\Lambda}(X_{1},U_{X_{2}})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(X_{1},X_{2})\rightarrow 0,

where the last 0 comes from the fact that Λ\Lambda is hereditary. Since ExtΛ1(X1,UX2)=0\mathrm{Ext}^{1}_{\Lambda}(X_{1},U_{X_{2}})=0, it follows that ExtΛ1(X1,X2)=0\mathrm{Ext}^{1}_{\Lambda}(X_{1},X_{2})=0; i.e., 1 holds. ∎

Remark 5.11.

If Λ\Lambda is not hereditary, then the Ext-orthogonality of the bricks in 𝒟(𝒯)\mathcal{D}(\mathcal{T}) is generally not a necessary or sufficient condition for 𝒯\mathcal{T} to be in the image of pop𝗍𝗈𝗋𝗌Λ\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}. See Remark 7.9 for an explicit example.

We conclude this section by combining several known results into a lattice-theoretic interpretation of the Ext-orthogonality condition in Theorem 5.10. As we discuss in Remark 7.7, a special case of this result can also be interpreted (and proved) using Coxeter combinatorics.

Lemma 5.12.

Let Λ\Lambda be an arbitrary τ\tau-tilting finite algebra, and let 𝒯𝗍𝗈𝗋𝗌Λ\mathcal{T}\in\operatorname{\mathsf{tors}}\Lambda. The following are equivalent.

  1. (1)

    ExtΛ1(X,X)=0\mathrm{Ext}^{1}_{\Lambda}(X,X^{\prime})=0 for all distinct X,X𝒟(𝒯)X,X^{\prime}\in\mathcal{D}(\mathcal{T}).

  2. (2)

    The interval [pop𝗍𝗈𝗋𝗌Λ(𝒯),𝒯]𝗍𝗈𝗋𝗌Λ[\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}),\mathcal{T}]\subseteq\operatorname{\mathsf{tors}}\Lambda is a distributive lattice.

  3. (3)

    The interval [pop𝗍𝗈𝗋𝗌Λ(𝒯),𝒯]𝗍𝗈𝗋𝗌Λ[\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}),\mathcal{T}]\subseteq\operatorname{\mathsf{tors}}\Lambda is a Boolean lattice.

  4. (4)

    The map 𝒳𝖥𝗂𝗅𝗍(𝒯𝒳)\mathcal{X}\mapsto\operatorname{\mathsf{Filt}}(\mathcal{T}\cup\mathcal{X}) is an isomorphism from the Boolean lattice of subsets of 𝒟(𝒯)\mathcal{D}(\mathcal{T}) to the interval [pop𝗍𝗈𝗋𝗌Λ(𝒯),𝒯][\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}),\mathcal{T}] of 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda.

  5. (5)

    There exist a set of finite-dimensional local KK-algebras {ΛXX𝒟(𝒯)}\{\Lambda_{X}\mid X\in\mathcal{D}(\mathcal{T})\} and an equivalence of categories (pop𝗍𝗈𝗋𝗌Λ(𝒯))𝒯𝗆𝗈𝖽(X𝒟(𝒯)ΛX)(\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}))^{\perp}\cap\mathcal{T}\cong\operatorname{\mathsf{mod}}(\prod_{X\in\mathcal{D}(\mathcal{T})}\Lambda_{X}).

Proof.

Denote 𝒰=pop𝗍𝗈𝗋𝗌Λ(𝒯)\mathcal{U}=\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T}). By [AP22, Theorem 6.3], Corollary 3.6, and Proposition 3.8, we have that 𝒰𝒯=𝖥𝗂𝗅𝗍(𝒟(𝒯))\mathcal{U}^{\perp}\cap\mathcal{T}=\operatorname{\mathsf{Filt}}(\mathcal{D}(\mathcal{T})). Since this is a functorially finite wide subcategory, there exists a finite-dimensional algebra Λ\Lambda^{\prime} such that 𝖥𝗂𝗅𝗍(𝒟(𝒯))\operatorname{\mathsf{Filt}}(\mathcal{D}(\mathcal{T})) is equivalent to 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda^{\prime}. Moreover, [AP22, Theorem 1.4] says that the interval [𝒰,𝒯][\mathcal{U},\mathcal{T}] is isomorphic to 𝗍𝗈𝗋𝗌Λ\operatorname{\mathsf{tors}}\Lambda^{\prime}. (See also [Jas15, Theorem 3.12] and [DIR+23, Section 4.2].) It therefore suffices to assume 𝒯=𝗆𝗈𝖽Λ\mathcal{T}=\operatorname{\mathsf{mod}}\Lambda, in which case 𝒟(𝒯)\mathcal{D}(\mathcal{T}) is the set of simple modules and pop𝗍𝗈𝗋𝗌Λ(𝒯)=0\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}\Lambda}(\mathcal{T})=0. The result then follows from [LW, Theorem 1.1]. ∎

6. The Weak Order and Cambrian Lattices

6.1. Coxeter groups

A standard reference for much of the material in this subsection is [BB05].

Let (W,S)(W,S) be a finite Coxeter system. This means that SS is a finite set and that WW is a finite group with a presentation of the form

S(ss)m(s,s)=e for all s,sS,\langle S\mid(ss^{\prime})^{m(s,s^{\prime})}=e\text{ for all }s,s^{\prime}\in S\rangle,

where ee is the identity element of WW and we have

m(s,s)=1andm(s,s)=m(s,s){2,3,}m(s,s)=1\quad\text{and}\quad m(s,s^{\prime})=m(s^{\prime},s)\in\{2,3,\ldots\}

for all distinct s,sSs,s^{\prime}\in S. (We often refer only to the Coxeter group, tacitly assuming SS is part of the data of WW.)

The elements of SS are called the simple reflections. A reflection is an element of the form wsw1wsw^{-1} for wWw\in W and sSs\in S. The Coxeter graph of WW is the graph ΓW\Gamma_{W} with vertex set SS in which two simple reflections ss and ss^{\prime} are connected by an edge whenever m(s,s)3m(s,s^{\prime})\geq 3; this edge is labeled with the number m(s,s)m(s,s^{\prime}) if m(s,s)4m(s,s^{\prime})\geq 4. We will assume that WW is irreducible, which means that ΓW\Gamma_{W} is connected. We say WW is simply-laced if m(s,s)3m(s,s^{\prime})\leq 3 for all s,sSs,s^{\prime}\in S. There is a well known characterization of finite irreducible Coxeter groups (see [BB05, Appendix A1]).

We will use sans serif font when we write words over the alphabet SS; this allows us to distinguish a word 𝗌1𝗌k\mathsf{s}_{1}\cdots\mathsf{s}_{k} from the element s1skWs_{1}\cdots s_{k}\in W that it represents.

A reduced word for an element wWw\in W is a word over the alphabet SS that represents ww and is as short as possible. The number of letters in a reduced word for ww is called the length of ww and is denoted (w)\ell(w). A right inversion (resp. left inversion) of ww is a reflection tt such that (wt)<(w)\ell(wt)<\ell(w) (resp. (tw)<(w)\ell(tw)<\ell(w)). The (right) weak order is the partial order \leq on WW defined so that uvu\leq v if and only if there exists a reduced word for vv that has a reduced word for uu as a prefix. Equivalently, uvu\leq v if and only if every left inversion of uu is a left inversion of vv. Let Weak(W)\mathrm{Weak}(W) denote the poset (W,)(W,\leq). It is well known that Weak(W)\mathrm{Weak}(W) is a ranked semidistributive lattice (\ell is a rank function). In fact, if WW is crystallographic, then Weak(W)\mathrm{Weak}(W) is isomorphic to the lattice of torsion classes of the preprojective algebra of type WW (see Remark 7.7).

A cover reflection of an element wWw\in W is a reflection tt of WW such that (tw)=(w)1\ell(tw)=\ell(w)-1 and tw=wstw=ws for some sSs\in S. Thus, the cover reflections of ww corresponds bijectively to the elements covered by ww in Weak(W)\mathrm{Weak}(W). Because Weak(W)\mathrm{Weak}(W) is a semidistributive lattice, each element wWw\in W has a canonical join representation

𝒟(w)={jtw,wt is a cover reflection of w}.\mathcal{D}(w)=\{j_{tw,w}\mid t\text{ is a cover reflection of }w\}.

The following result, which appears as [RS11, Theorem 8.1] (see also [BR18, Proposition 4.13]), describes the elements of 𝒟(w)\mathcal{D}(w) more explicitly.

Proposition 6.1.

Let tt be a cover reflection of an element wWw\in W. The shard label jtw,wj_{tw,w} of the edge twwtw\lessdot w in Weak(w)\mathrm{Weak}(w) is the unique minimal element of

{xWeak(W)xw and t is a left inversion of x}.\{x\in\mathrm{Weak}(W)\mid x\leq w\text{ and }t\text{ is a left inversion of }x\}.

The long element of WW is the unique element ww_{\circ} of WW that has maximum length. It is known that w2=ew_{\circ}^{2}=e. If JSJ\subseteq S, then the subgroup WJW_{J} of WW generated by JJ is called a (standard) parabolic subgroup of WW. The pair (WJ,J)(W_{J},J) is a finite Coxeter system; the long element of WJW_{J} is denoted by w(J)w_{\circ}(J).

A descent of an element wWw\in W is a simple reflection ss such that (ws)<(w)\ell(ws)<\ell(w); in other words, it is a simple reflection that is also a right inversion of ww. Let Des(w)\mathrm{Des}(w) denote the set of descents of ww. A simple reflection is a descent of ww if and only if there is a reduced word for ww that ends in ss. If u1v1u^{-1}\leq v^{-1}, then Des(u)Des(v)\mathrm{Des}(u)\subseteq\mathrm{Des}(v). The pop-stack operator on Weak(W)\mathrm{Weak}(W) has an alternative description [Def22b] given by

(6.1) popWeak(W)(x)=xw(Des(x)).\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(x)=xw_{\circ}(\mathrm{Des}(x)).

Let n=|S|n=|S|. For α=(α1,,αn)\alpha=(\alpha_{1},\ldots,\alpha_{n}) and β=(β1,,βn)\beta=(\beta_{1},\ldots,\beta_{n}) in n\mathbb{R}^{n}, let

(6.2) (α,β)=i,j[n]2cos(π/m(i,j))αiβj.(\alpha,\beta)=\sum_{i,j\in[n]}-2\cos(\pi/m(i,j))\cdot\alpha_{i}\beta_{j}.

Because WW is finite, it is known that (,)(-,-) is a positive definite symmetric bilinear form. For i[n]i\in[n], we let eie_{i} denote the ii-th standard basis vector of n\mathbb{R}^{n}, and we let ei=2ei/(ei,ei)e_{i}^{\vee}=2e_{i}/(e_{i},e_{i}). By abuse of notation, we use siSs_{i}\in S to denote the linear transformation si:nns_{i}\colon\mathbb{R}^{n}\rightarrow\mathbb{R}^{n} given by si(α)=α(ei,α)eis_{i}(\alpha)=\alpha-(e_{i}^{\vee},\alpha)e_{i}. This definition extends to a linear representation of WW on n\mathbb{R}^{n}. A vector αn\alpha\in\mathbb{Z}^{n} is called a root of WW if there exists wWw\in W such that α=w(ei)\alpha=w(e_{i}) for some ii. A root is positive if its coordinates are all nonnegative. We denote by ΦW\Phi_{W} and ΦW+\Phi_{W}^{+} the sets of roots and positive roots of WW, respectively.

The action of WW induces a permutation on the set of roots; that is, ΦW={w(α)αΦW}\Phi_{W}=\{w(\alpha)\mid\alpha\in\Phi_{W}\} for all wWw\in W. Given a reflection tt in WW, we can write t=wsiw1t=ws_{i}w^{-1} for some wWw\in W and siSs_{i}\in S; let βt\beta_{t} be the unique positive root in {±w(ei)}\{\pm w(e_{i})\}. It is known that βt\beta_{t} does not depend on the choices of ww and sis_{i}. Moreover, the map tβtt\mapsto\beta_{t} is a bijection from the set of reflections of WW to the set ΦW+\Phi_{W}^{+} of positive roots.

For wWw\in W, let

inv(w)={αΦW+w1(α)ΦW+}.\mathrm{inv}(w)=\{\alpha\in\Phi^{+}_{W}\mid w^{-1}(\alpha)\in-\Phi_{W}^{+}\}.

It is known that βtinv(w)\beta_{t}\in\mathrm{inv}(w) if and only if tt is a left inversion of ww. Thus, for u,vWu,v\in W, we have uvu\leq v in the weak order if and only if inv(u)inv(v)\mathrm{inv}(u)\subseteq\mathrm{inv}(v).

6.2. Coxeter elements and Cambrian lattices

A (standard) Coxeter element of WW is an element cc obtained by multiplying the simple reflections in some order (with each appearing once in the product). Thus, a reduced word for cc is a word representing cc in which each simple reflection appears exactly once. Let us orient each edge {s,s}\{s,s^{\prime}\} in ΓW\Gamma_{W} from ss to ss^{\prime} if and only if ss appears before ss^{\prime} in some (equivalently, every) reduced word for cc. The result is an acyclic orientation of ΓW\Gamma_{W}. This construction establishes a one-to-one correspondence between Coxeter elements of WW and acyclic orientations of ΓW\Gamma_{W}. We denote the acyclic orientation corresponding to a Coxeter element cc by QcQ_{c}. As is standard in representation theory, we will sometimes call QcQ_{c} a quiver instead of a directed graph. We write S={s1,,sn}S=\{s_{1},\ldots,s_{n}\} and denote by (Qc)0=[n](Q_{c})_{0}=[n] the set of vertices of QcQ_{c}, identifying each index i[n]i\in[n] with sis_{i}. Likewise, we denote by (Qc)1(Q_{c})_{1} the set of arrows of QcQ_{c}. Our convention is to use the notation ai,j(Qc)1a_{i,j}\in(Q_{c})_{1} to represent an arrow pointing from the vertex ii to the vertex jj.

Fix a reduced word 𝖼\mathsf{c} for a Coxeter element cc, and consider the infinite word 𝖼=𝖼(1)𝖼(2)\mathsf{c}^{\infty}=\mathsf{c}^{(1)}\mathsf{c}^{(2)}\cdots, where each 𝖼(k)\mathsf{c}^{(k)} is a copy of 𝖼\mathsf{c}. Following Reading [Rea07], we define the 𝖼\mathsf{c}-sorting word of an element wWw\in W to be the reduced word 𝗌𝗈𝗋𝗍𝖼(w)\mathsf{sort}_{\mathsf{c}}(w) for ww that is lexicographically first as a subword of 𝖼\mathsf{c}^{\infty}. Let 𝐈c(k)(w){\bf I}_{c}^{(k)}(w) be the set of simple reflections that are taken from 𝖼(k)\mathsf{c}^{(k)} when we form 𝗌𝗈𝗋𝗍𝖼(w)\mathsf{sort}_{\mathsf{c}}(w) as the lexicographically first subword of 𝖼\mathsf{c}^{\infty}. Although 𝐈c(k)(w){\bf I}_{c}^{(k)}(w) depends on the Coxeter element cc, it does not depend on the choice of the reduced word 𝖼\mathsf{c}. The element ww is called cc-sortable if 𝐈c(1)(w)𝐈c(2)(w){\bf I}_{c}^{(1)}(w)\supseteq{\bf I}_{c}^{(2)}(w)\supseteq\cdots.

The set of cc-sortable elements of WW forms a sublattice of Weak(W)\mathrm{Weak}(W) called the cc-Cambrian lattice, which we denote by Cambc\mathrm{Camb}_{c}. We will recall in Section 6.3 that Cambc\mathrm{Camb}_{c} can be realized as the lattice of torsion classes of a finite-dimensional (hereditary) algebra when WW is crystallographic.

Example 6.2.

A basic yet important example of a Coxeter group is the dihedral group I2(m)I_{2}(m), whose Coxeter graph is ​​​s2s1m\begin{array}[]{l}\leavevmode\hbox to56.83pt{\vbox to17.56pt{\pgfpicture\makeatletter\hbox{\hskip 28.41632pt\lower-6.388pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ } \par{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{17.59583pt}{-1.25055pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$s_{2}$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-25.08331pt}{-1.25055pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$s_{1}$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.39006pt}{3.533pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{{$m$}} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \par{ {}{}{}}{}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.8pt}\pgfsys@invoke{ }{}\pgfsys@moveto{14.06282pt}{0.0pt}\pgfsys@lineto{-14.06282pt}{0.0pt}\pgfsys@stroke\pgfsys@invoke{ } \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}\end{array}​​​ (if m=2m=2, then there is no edge). Let c=s1s2c=s_{1}s_{2}. The cc-sortable elements of I2(m)I_{2}(m) are s2s_{2} and the m+1m+1 elements e,s1,s1s2,s1s2s1,,we,s_{1},s_{1}s_{2},s_{1}s_{2}s_{1},\ldots,w_{\circ}.

For each wWw\in W, the set Cambc{vWvw}\mathrm{Camb}_{c}\cap\{v\in W\mid v\leq w\} has a unique maximal element in the weak order; we denote this element by πc(w)\pi_{\downarrow}^{c}(w). The map πc\pi_{\downarrow}^{c} is a surjective lattice homomorphism from Weak(W)\mathrm{Weak}(W) to Cambc\mathrm{Camb}_{c}, so Cambc\mathrm{Camb}_{c} is a lattice quotient of Weak(W)\mathrm{Weak}(W) [Rea06, Rea07]. The fibers of πc\pi_{\downarrow}^{c} are the equivalence classes of an equivalence relation on WW known as the cc-Cambrian congruence.

According to [Def22a, Theorem 3.2], we have

(6.3) popCambc=πcpopWeak(W).\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}=\pi_{\downarrow}^{c}\circ\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}.

The following lemma, which is immediate from the definition of a cc-sortable element (see also [Rea07, Lemmas 2.4 & 2.5]), gives a recursive characterization of cc-sortable elements.

Lemma 6.3.

Let WW be a finite Coxeter group, and let 𝖼\mathsf{c} be a reduced word for a Coxeter element cc of WW. Let ss be the first letter in ss. Let wWw\in W. If (sw)>(w)\ell(sw)>\ell(w), then ww is cc-sortable if and only if it is an (sc)(sc)-sortable element of the standard parabolic subgroup WS{s}W_{S\setminus\{s\}}. If (sw)<(w)\ell(sw)<\ell(w), then ww is cc-sortable if and only if swsw is (scs)(scs)-sortable.

We will also make use of the following characterization of cc-sortable elements in terms of canonical join representations.

Lemma 6.4.

Let cc be a Coxeter element of a finite Coxeter group WW. An element wWw\in W is cc-sortable if and only if every element of the canonical join representation of ww in Weak(W)\mathrm{Weak}(W) is cc-sortable. If ww is cc-sortable, then the canonical join representation of ww in Weak(W)\mathrm{Weak}(W) is equal to the canonical join representation of ww in Cambc\mathrm{Camb}_{c}.

Proof.

According to [RS11, Proposition 8.2], if ww is cc-sortable, then each element of its canonical join representation is cc-sortable. On the other hand, if each element of the canonical join representation of ww is cc-sortable, then ww is cc-sortable because Cambc\mathrm{Camb}_{c} is a sublattice of the weak order. ∎

6.3. Cambrian lattices as lattices of torsion classes

Suppose for this subsection that WW is crystallographic (i.e., WW is not of type H3H_{3}, H4H_{4}, or I2(m)I_{2}(m) for m{3,4,6}m\notin\{3,4,6\}). Fix a Coxeter element cWc\in W. Associated to the acyclic orientation QcQ_{c} of cc is a finite-dimensional algebra KQcKQ_{c} known as the tensor algebra of QcQ_{c} (More precisely, the definition of KQcKQ_{c} depends on a choice of symmetrizable Cartan matrix with Weyl group WW.) In this subsection, we recall the properties of KQcKQ_{c} that we will use to interpret Theorem 5.10 in terms of Coxeter combinatorics. Notably, the definition of the algebra KQcKQ_{c} is not needed in order to state these results and is thus omitted from this paper. Readers interested in more details are referred to [DR76].

Recall the notation of the projective modules P(i)P(i) and simple modules S(i)S(i) from Section 3. The indexing of these modules can be chosen so that dim¯(S(i))=ei=βsi\underline{\mathrm{dim}}(S(i))=e_{i}=\beta_{s_{i}} for all i[n]i\in[n]. We fix this indexing for the remainder of this section. For i[n]i\in[n], we then denote

(6.4) ρi=snsn1si+1(ei).\rho_{i}=s_{n}s_{n-1}\cdots s_{i+1}(e_{i}).

Note that {ρii[n]}=inv(c1)\{\rho_{i}\mid i\in[n]\}=\mathrm{inv}(c^{-1}). The roots ρi\rho_{i} are sometimes called projective roots, a name that is justified by Item 5 in Proposition 6.5 below.

We now recall the following well-known properties of the algebra KQcKQ_{c}. Note that Items 2, 3 and 4 together constitute Gabriel’s Theorem. These results can be found in the classical references [DR75, DR76].

Proposition 6.5.
  1. (1)

    The algebra KQcKQ_{c} is hereditary.

  2. (2)

    The algebra KQcKQ_{c} is representation-finite.

  3. (3)

    The association Mdim¯MM\mapsto\underline{\mathrm{dim}}M is a bijection from the set of (isomorphism classes of) indecomposable modules in 𝗆𝗈𝖽KQc\operatorname{\mathsf{mod}}KQ_{c} to the set ΦW+\Phi_{W}^{+} of positive roots.

  4. (4)

    Every indecomposable module in 𝗆𝗈𝖽KQc\operatorname{\mathsf{mod}}KQ_{c} is a brick.

  5. (5)

    For i[n]i\in[n], the dimension vector of the indecomposable projective module P(i)𝗆𝗈𝖽KQc{P(i)\in\operatorname{\mathsf{mod}}KQ_{c}} is dim¯P(i)=ρi\underline{\mathrm{dim}}P(i)=\rho_{i}.

Remark 6.6.

Let 𝒯𝗍𝗈𝗋𝗌KQc\mathcal{T}\in\operatorname{\mathsf{tors}}KQ_{c} be a torsion class that satisfies ExtKQc1(X,X)=0\mathrm{Ext}_{KQ_{c}}^{1}(X,X^{\prime})=0 for all X,X𝒟(𝒯){X,X^{\prime}\in\mathcal{D}(\mathcal{T})}. Then Item 4 in Proposition 6.5 implies that there exists a nonzero projective module P𝒯P\in\mathcal{T} if and only if there exists an indecomposable projective module P(i)𝒟(𝒯)P(i)\in\mathcal{D}(\mathcal{T}).

We now recall the explicit bijection between the Cambrian lattice Cambc\mathrm{Camb}_{c} and the lattice of torsion classes 𝗍𝗈𝗋𝗌(KQc)\operatorname{\mathsf{tors}}(KQ_{c}) established in [IT09]. Note that while the result is only proved explicitly for the simply-laced case in [IT09], it is remarked in [IT09, Section 4.4] that the results can be generalized using folding arguments.

Theorem 6.7 ([IT09, Theorem 4.3]).

There is a lattice isomorphism φc:Cambc𝗍𝗈𝗋𝗌KQc\varphi_{c}\colon\mathrm{Camb}_{c}\rightarrow\operatorname{\mathsf{tors}}KQ_{c} characterized by the condition that

φc(w)𝖻𝗋𝗂𝖼𝗄KQc={M𝖻𝗋𝗂𝖼𝗄KQcdim¯Minv(w)}\varphi_{c}(w)\cap\operatorname{\mathsf{brick}}KQ_{c}=\{M\in\operatorname{\mathsf{brick}}KQ_{c}\mid\underline{\mathrm{dim}}M\in\mathrm{inv}(w)\}

for every cc-sortable element wWw\in W.

7. The Image of Pop-Stack on a Cambrian Lattice

In this section, we recast Theorem 5.10 as a combinatorial description of the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} for any finite irreducible Coxeter group WW and any Coxeter element cc. As we discuss in the proof of Theorem 7.8, the results of Section 5 only yield a proof for WW of crystallographic type, while the non-crystallographic types can be verified by computer. Thus, when not directly specified, we assume that WW is an arbitrary finite Coxeter group (not necessarily crystallographic). When WW is crystallographic, we use the notation KQcKQ_{c} as in Section 6.3. In any case, we denote by cc a fixed Coxeter element of WW.

We first give an interpretation of the Ext-orthogonality condition in Theorem 5.101. For each wCambcw\in\mathrm{Camb}_{c}, Lemmas 5.12 and 6.7 say that the torsion class φc(w)\varphi_{c}(w) satisfies the Ext-orthogonality condition if and only if [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] is Boolean. Theorem 7.6 below provides several equivalent conditions characterizing when this holds. Before we state and prove this result, we require some lemmas.

Lemma 7.1.

Let LL be a finite lattice, and let xLx\in L. If the interval [popL(x),x][\mathrm{pop}^{\downarrow}_{L}(x),x] is distributive, then it is Boolean.

Proof.

Suppose [popL(x),x][\mathrm{pop}^{\downarrow}_{L}(x),x] is distributive, and let PP be a finite poset such that [popL(x),x][\mathrm{pop}^{\downarrow}_{L}(x),x] is isomorphic to J(P)J(P). The bottom and top elements of J(P)J(P) are \emptyset and PP, respectively, so =popJ(P)(P)=ymax(P)(P{y}){\emptyset=\mathrm{pop}^{\downarrow}_{J(P)}(P)=\bigcap_{y\in\max(P)}(P\setminus\{y\})}, where max(P)\max(P) is the set of maximal elements of PP. This implies that P=max(P)P=\max(P), so PP is an antichain. Hence, J(P)J(P) is Boolean. ∎

Lemma 7.2.

If LL is a finite semidistributive lattice with kk coatoms, then |L|2k|L|\geq 2^{k}.

Proof.

The map v𝒟(v)v\mapsto\mathcal{D}(v) is a bijection from LL to the canonical join complex of LL. Since |𝒟(1^)|=k|\mathcal{D}(\hat{1})|=k, the canonical join complex of LL must have at least 2k2^{k} faces. ∎

The next result follows from [Rea04, Lemma 3.8]; we include a short self-contained proof.

Lemma 7.3 ([Rea04, Lemma 3.8]).

If y1z1y_{1}\lessdot z_{1} and y2z2y_{2}\lessdot z_{2} are two edges of Weak(W)\mathrm{Weak}(W) with the same shard label, then z1y11=z2y21z_{1}y_{1}^{-1}=z_{2}y_{2}^{-1}.

Proof.

Let t1=z1y11t_{1}=z_{1}y_{1}^{-1} and t2=z2y21t_{2}=z_{2}y_{2}^{-1}. Proposition 6.1 tells us that the join-irreducible element jy1,z1j_{y_{1},z_{1}} has t1t_{1} as a left inversion while (jy1,z1)(j_{y_{1},z_{1}})_{*} does not have t1t_{1} as a left inversion. If follows that t1t_{1} is the unique cover reflection of jy1,z1j_{y_{1},z_{1}}. Similarly, t2t_{2} is the unique cover reflation of jy2,z2j_{y_{2},z_{2}}. Since jy1,z1=jy2,z2j_{y_{1},z_{1}}=j_{y_{2},z_{2}}, we have t1=t2t_{1}=t_{2}. ∎

The following lemma is a special case of [Rea11, Proposition 5.7].

Lemma 7.4 ([Rea11, Proposition 5.7]).

If ww and zz are elements of Weak(W)\mathrm{Weak}(W) such that the canonical join representation of ww contains that of zz, then the set of shard labels of edges in the interval [popWeak(W)(z),z][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(z),z] is contained in the set of shard labels of the edges in the interval [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w].

Lemma 7.5.

Suppose ww is a cc-sortable element of WW whose descents commute. The interval [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] of Weak(W)\mathrm{Weak}(W) is Boolean, and popCambc(w)\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w) is cc-sortable.

Proof.

The fact that the interval [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] is Boolean is immediate from the observation that it consists of the elements of the form wv1wv^{-1} such that vv is the product of some subset of Des(w)\mathrm{Des}(w). Let us now show that popWeak(W)(w)\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w) is cc-sortable. Our proof will proceed by induction on |S||S| and (w)\ell(w) (the base cases are trivial). To ease notation, let us write w=popWeak(W)(w)w^{\prime}=\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w).

Let ss be the first letter in some reduced word for cc. Suppose first that (sw)>(w)\ell(sw)>\ell(w). According to Lemma 6.3, ww is an (sc)(sc)-sortable element of WS{s}W_{S\setminus\{s\}}. By induction, popWeak(WS{s})(w)\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W_{S\setminus\{s\}})}(w) is an (sc)(sc)-sortable element of WS{s}W_{S\setminus\{s\}}. Invoking Lemma 6.3 again, we find that popWeak(WS{s})(w)\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W_{S\setminus\{s\}})}(w) is cc-sortable. The desired result now follows from the fact that w=popWeak(WS{s})(w)w^{\prime}=\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W_{S\setminus\{s\}})}(w).

Now assume that (sw)<(w)\ell(sw)<\ell(w). Then (sw)1w1(sw)^{-1}\leq w^{-1}, so every left inversion of (sw)1(sw)^{-1} is a left inversion of w1w^{-1}. Hence, every right inversion of swsw is a right inversion of ww. It follows that Des(sw)Des(w)\mathrm{Des}(sw)\subseteq\mathrm{Des}(w). We consider two cases.

Case 1. Suppose sws\leq w^{\prime} (equivalently, (sw)=(w)1\ell(sw^{\prime})=\ell(w^{\prime})-1). Let v=swv=sw^{\prime}. We have w=svw(Des(w))w=svw_{\circ}(\mathrm{Des}(w)), and (w)=(v)+(w(Des(w)))+1\ell(w)=\ell(v)+\ell(w_{\circ}(\mathrm{Des}(w)))+1. Hence, sw=vw(Des(w))sw=vw_{\circ}(\mathrm{Des}(w)), and (sw)=(v)+(w(Des(w)))\ell(sw)=\ell(v)+\ell(w_{\circ}(\mathrm{Des}(w))). This shows that swsw has a reduced word that contains a reduced word for w(Des(w))w_{\circ}(\mathrm{Des}(w)) as a suffix. Hence, Des(w)=Des(w(Des(w)))Des(sw)\mathrm{Des}(w)=\mathrm{Des}(w_{\circ}(\mathrm{Des}(w)))\subseteq\mathrm{Des}(sw).

This shows that Des(w)=Des(sw)\mathrm{Des}(w)=\mathrm{Des}(sw), so popWeak(W)(sw)=sww(Des(w))=v\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(sw)=sww_{\circ}(\mathrm{Des}(w))=v. Because ww is cc-sortable, we know by Lemma 6.3 that swsw is (scs)(scs)-sortable. Furthermore, the descents of swsw commute. Because (sw)<(w)\ell(sw)<\ell(w), we can use induction (replacing cc by scsscs) to find that popWeak(W)(sw)=v\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(sw)=v is (scs)(scs)-sortable. But since v=swv=sw^{\prime} and (sw)<(w)\ell(sw^{\prime})<\ell(w^{\prime}), we can invoke Lemma 6.3 once again to find that ww^{\prime} is cc-sortable.

Case 2. Suppose sws\not\leq w^{\prime} (equivalently, (sw)=(w)+1\ell(sw^{\prime})=\ell(w^{\prime})+1). Since ww^{\prime} is the meet of the elements covered by ww in Weak(W)\mathrm{Weak}(W), there must exist an element xx that is covered by ww such that sxs\not\leq x. The unique left inversion of ww that is not a left inversion of xx is wx1wx^{-1}. Since ss is a left inversion of ww but not a left inversion xx, we must have wx1=swx^{-1}=s. Hence, x=swx=sw.

We have seen that Des(sw)Des(w)\mathrm{Des}(sw)\subseteq\mathrm{Des}(w), and we have just shown that swwsw\lessdot w. There exists sSs^{\prime}\in S such that sw=wssw=ws^{\prime}. Because [w,w][w^{\prime},w] is Boolean, we must have Des(sw)=Des(w){s}\mathrm{Des}(sw)=\mathrm{Des}(w)\setminus\{s^{\prime}\}, and it follows from Lemmas 2.1 and 2.2 that 𝒟(sw)=𝒟(w){s}\mathcal{D}(sw)=\mathcal{D}(w)\setminus\{s\}. Hence,

popWeak(W)(sw)=sws′′Des(sw)s′′=wss′′Des(w){s}s′′=ws′′Des(w)s′′=w.\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(sw)=sw\prod_{s^{\prime\prime}\in\mathrm{Des}(sw)}s^{\prime\prime}=ws^{\prime}\prod_{s^{\prime\prime}\in\mathrm{Des}(w)\setminus\{s^{\prime}\}}s^{\prime\prime}=w\prod_{s^{\prime\prime}\in\mathrm{Des}(w)}s^{\prime\prime}=w^{\prime}.

Since ww is cc-sortable, it follows from Lemma 6.4 that swsw is also cc-sortable. The descents of swsw commute, so we can use induction to find that popWeak(W)(sw)=w\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(sw)=w^{\prime} is cc-sortable. ∎

Theorem 7.6.

Let cc be a Coxeter element of a finite Coxeter group WW. For every cc-sortable element ww, the following are equivalent.

  1. (1)

    The descents of ww commute.

  2. (2)

    The interval [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] of Weak(W)\mathrm{Weak}(W) is a distributive lattice.

  3. (3)

    The interval [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] of Weak(W)\mathrm{Weak}(W) is a Boolean lattice.

  4. (4)

    The interval [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] of Cambc\mathrm{Camb}_{c} is a distributive lattice.

  5. (5)

    The interval [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] of Cambc\mathrm{Camb}_{c} is a Boolean lattice.

  6. (6)

    The interval [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] of Cambc\mathrm{Camb}_{c} equals the interval [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] of Weak(W)\mathrm{Weak}(W).

Proof.

It is immediate from Lemma 7.1 that 2 and 3 are equivalent and that 4 and 5 are equivalent. Lemma 7.5 tells us that 1 implies 3. To see that 3 implies 1, suppose [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] is Boolean, and suppose ss and ss^{\prime} are distinct descents of WW. Let w=wswsw^{\prime}=ws\wedge ws^{\prime}. The interval [w,w][w^{\prime},w] in Weak(W)\mathrm{Weak}(W) is contained in [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w], so [w,w][w^{\prime},w] is Boolean. We also know that [w,w][w^{\prime},w] is isomorphic to the weak order on the dihedral group I2(m(s,s))I_{2}(m(s,s^{\prime})). This implies that m(s,s)=2m(s,s^{\prime})=2, so ss and ss^{\prime} commute.

According to Lemma 6.4, the canonical join representations of ww in Weak(W)\mathrm{Weak}(W) and Cambc\mathrm{Camb}_{c} are equal; thus, we can unambiguously write 𝒟(w)\mathcal{D}(w) for this canonical join representation.

Let us show that 1 implies 6. Suppose the descents of ww commute. Let k=𝒟(w)=|Des(w)|{k=\mathcal{D}(w)=|\mathrm{Des}(w)|}. According to Lemma 7.5, the interval [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] in Weak(W)\mathrm{Weak}(W) is Boolean, and the element popWeak(W)(w)\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w) is cc-sortable. This implies that [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] is contained in [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] and that [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] has size 2k2^{k}. But [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] is a semidistributive lattice with kk coatoms, so Lemma 7.2 tells us that it has size at least 2k2^{k}. This implies that [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] equals [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w].

In [Tho06], Thomas introduced the notion of a trim lattice; he proved that intervals in trim lattices are trim and that a lattice is distributive if and only if it is ranked and trim. Ingalls and Thomas [IT09] proved that Cambrian lattices are trim. These facts allow us to prove that 6 implies 2. Indeed, [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] is an interval in the weak order on WW, so it is ranked. If [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w] is equal to an interval of Cambc\mathrm{Camb}_{c}, then it is trim, so it must be distributive.

We now know that Items 1, 2, 3 and 6 are equivalent and that Items 4 and 5 are equivalent. It is also obvious that 3 and 6 together imply 5. Therefore, to complete the proof, it suffices to show that 5 implies 1.

Assume [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] is Boolean. Suppose by way of contradiction that ss and ss^{\prime} are descents of ww that do not commute. Because {jws,w,jws,w}\{j_{ws,w},j_{ws^{\prime},w}\} is contained in 𝒟(w)\mathcal{D}(w), it follows from Lemma 6.4 that there exists a cc-sortable element zz of Weak(W)\mathrm{Weak}(W) whose canonical join representation is {jws,w,jws,w}\{j_{ws,w},j_{ws^{\prime},w}\}. Let yy and yy^{\prime} be the two elements covered by zz, and assume without loss of generality that jy,z=jws,wj_{y,z}=j_{ws,w} and jy,z=jws,wj_{y^{\prime},z}=j_{ws^{\prime},w}. Deleting the last letter in the cc-sorting word for zz yields the cc-sorting word for a cc-sortable element covered by zz; this element must be yy or yy^{\prime}. Assume without loss of generality that yy is cc-sortable. By Lemma 7.3, we have zy1=wsw1zy^{-1}=wsw^{-1} and z(y)1=wsw1z(y^{\prime})^{-1}=ws^{\prime}w^{-1}. It follows that zy1zy^{-1} and z(y)1z(y^{\prime})^{-1} do not commute. Hence, the interval [popWeak(W)(z),z]=[yy,z][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(z),z]=[y\wedge y^{\prime},z] is isomorphic to the weak order on the dihedral group I2(m)I_{2}(m) for some m3m\geq 3. Let xx be the unique element of [popWeak(W)(z),z][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(z),z] covered by yy. Since yy is cc-sortable, Lemma 6.4 tells us that the shard label jx,yj_{x,y} is cc-sortable. According to Lemma 7.4, jx,yj_{x,y} is the shard label of an edge in [popWeak(W)(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(w),w]. Since jx,yj_{x,y} is cc-sortable, it is also a shard label of an edge in [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w]. Because [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] is Boolean, it follows from Lemma 2.1 that jx,y𝒟(w)j_{x,y}\in\mathcal{D}(w). Because 𝒟(w)\mathcal{D}(w) is an independent set of the Galois graph of Weak(W)\mathrm{Weak}(W) and jy,z=jws,w𝒟(w)j_{y,z}=j_{ws,w}\in\mathcal{D}(w), there cannot be an arrow from jy,zj_{y,z} to jx,yj_{x,y} in the Galois graph of Weak(W)\mathrm{Weak}(W). However, since [yy,z][y\wedge y^{\prime},z] is isomorphic to the weak order on I2(m)I_{2}(m), it follows from Lemmas 2.1 and 2.3 that there is an arrow from jy,zj_{y,z} to jx,yj_{x,y} in the Galois graph of Weak(W)\mathrm{Weak}(W). This is our desired contradiction. ∎

Remark 7.7.

It is also possible to prove many of the equivalences in Theorem 7.6 using arguments from representation theory, or conversely to use Theorem 7.6 to prove special cases of Lemma 5.12. Since these arguments highlight further connections between Coxeter groups and lattices of torsion classes, we give a short overview of them here.

In addition to the hereditary algebras KQcKQ_{c}, one can also associate a (non-hereditary) preprojective algebra Π(W)\Pi(W) to each finite crystallographic Coxeter group WW. (For background outside the simply-laced case, see [Kül17, Section 4] and the references therein.) It is shown in [Miz14] (simply-laced case) and [AHI+, Section 7] (general case) that there is an isomorphism φ:Weak(W)𝗍𝗈𝗋𝗌(Π(W))\varphi:\mathrm{Weak}(W)\rightarrow\operatorname{\mathsf{tors}}(\Pi(W)) induced by the symmetric bilinear form (,)(-,-).

For any Coxeter element cc, the hereditary algebra KQcKQ_{c} is a quotient of the preprojective algebra. Thus the algebra quotient Π(W)KQc\Pi(W)\rightarrow KQ_{c} induces a lattice quotient

Weak(W)𝗍𝗈𝗋𝗌(Π(W))𝗍𝗈𝗋𝗌(KQc)Cambc,\mathrm{Weak}(W)\cong\operatorname{\mathsf{tors}}(\Pi(W))\rightarrow\operatorname{\mathsf{tors}}(KQ_{c})\cong\mathrm{Camb}_{c},

and it is shown (in the simply-laced case) in [MT20] that this lattice quotient coincides with the map πc\pi_{\downarrow}^{c}. The algebra quotient also induces a (fully faithful) inclusion 𝖻𝗋𝗂𝖼𝗄KQc𝖻𝗋𝗂𝖼𝗄Π(W)\operatorname{\mathsf{brick}}KQ_{c}\subseteq\operatorname{\mathsf{brick}}\Pi(W), which is known to preserve Ext-orthogonality. Given this, Theorem 7.6 becomes equivalent to the specialization of Items 1, 2, 3 and 4 in Lemma 5.12 to the cases Λ=KQc\Lambda=KQ_{c} and Λ=Π(W)\Lambda=\Pi(W).

We now turn toward interpreting Item 2 in Theorem 5.10. Recall that SS is precisely the set of atoms of Cambc\mathrm{Camb}_{c}. In the crystallographic case, Theorem 6.7 says that these atoms can be indexed (by [n][n]) so that φc(si)=𝖥𝗂𝗅𝗍(Si)\varphi_{c}(s_{i})=\operatorname{\mathsf{Filt}}(S_{i}) for all ii. In the non-crystallographic case, we fix an arbitrary indexing of SS by [n][n]. In either case, for siSs_{i}\in S, we let

Θi={xCambcsix and sjx for all sjS{si}}\Theta_{i}=\{x\in\mathrm{Camb}_{c}\mid s_{i}\leq x\text{ and }s_{j}\not\leq x\text{ for all }s_{j}\in S\setminus\{s_{i}\}\}

and let pi=Θip_{i}=\bigvee\Theta_{i}. (One can show that pip_{i} is actually the unique maximal element of Θi\Theta_{i}.) If WW is crystallographic, then Propositions 5.9 and 6.7 imply that φc(pi)=𝖦𝖾𝗇(P(i))\varphi_{c}(p_{i})=\operatorname{\mathsf{Gen}}(P(i)) for each ii. In particular, Remark 6.6 implies that, for wCambcw\in\mathrm{Camb}_{c}, the torsion class φc(w)\varphi_{c}(w) contains a projective module if and only if piwp_{i}\leq w for some ii.

We now state the main result of this section, which provides purely lattice-theoretic and Coxeter-theoretic descriptions of the image of the pop-stack operator on any Cambrian lattice.

Theorem 7.8.

Let WW be a finite irreducible Coxeter group, and let cWc\in W be a Coxeter element. For wCambcw\in\mathrm{Camb}_{c}, the following are equivalent.

  1. (1)

    ww is in the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}.

  2. (2)

    The descents of ww all commute, and ww has no left inversions in common with c1c^{-1}.

  3. (3)

    The interval [popCambc(w),w][\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w),w] is Boolean, and piwp_{i}\not\leq w for all i[n]i\in[n].

Proof.

Suppose first that WW is crystallographic. By Theorem 6.7, we have that ww is in the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} if and only if φc(w)\varphi_{c}(w) is in the image of pop𝗍𝗈𝗋𝗌KQc\mathrm{pop}^{\downarrow}_{\operatorname{\mathsf{tors}}KQ_{c}}. Furthermore, it was established above that φc(w)\varphi_{c}(w) contains a nonzero projective module if and only if there exists i[n]i\in[n] such that piwp_{i}\leq w, and we know that piwp_{i}\leq w if and only if the root ρi\rho_{i} (defined in Equation 6.4) is in inv(w)\mathrm{inv}(w). Finally, we have that the descents of ww all commute with one another if and only if ExtKQc1(X,X)=0\mathrm{Ext}^{1}_{KQ_{c}}(X,X^{\prime})=0 for all distinct X,X𝒟(φc(w))X,X^{\prime}\in\mathcal{D}(\varphi_{c}(w)) by Lemmas 5.12 and 7.6. Therefore, the desired result follows from Theorems 5.10 and 7.6 and the fact that {ρii[n]}=inv(c1)\{\rho_{i}\mid i\in[n]\}=\mathrm{inv}(c^{-1}).

It remains to consider the non-crystallographic cases. We verified the theorem when WW is of type H3H_{3} or H4H_{4} using Sage [The23]. When WW is of type I2(m)I_{2}(m), it is straightforward to check the desired result by hand. ∎

Remark 7.9.

The naive analogue of Theorem 7.8 fails for the weak order on AnA_{n}. For example, the pop-stack operator on Weak(A4)\mathrm{Weak}(A_{4}) sends the permutation 5234152341 to the permutation 2531425314, but the descents of 2531425314 (s2s_{2} and s3s_{3}) do not commute with each other. In the language of representation theory, this means that there exist bricks X,X𝒟(φ(25314))X,X^{\prime}\in\mathcal{D}(\varphi(25314)) such that ExtΠ(A4)1(X,X)0{\mathrm{Ext}^{1}_{\Pi(A_{4})}(X,X^{\prime})\neq 0} (the notation φ\varphi is from Remark 7.7). Thus, the naive analogue of Theorem 5.10 likewise fails for the preprojective algebra Π(A4)\Pi(A_{4}).

It is also worth mentioning that the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} is not necessarily contained in the image of popWeak(W)\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}. For example, if W=A9W=A_{9} and c=s1s3s5s7s9s2s4s6s8c=s_{1}s_{3}s_{5}s_{7}s_{9}s_{2}s_{4}s_{6}s_{8}, then one can use [ABB+19, Theorem 1] and Corollary 8.5 below to show that the permutation 1,2,8,10,6,9,4,5,3,71,2,8,10,6,9,4,5,3,7 is in the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} but not the image of popWeak(W)\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}.

The following surprising consequence of Theorem 7.6 tells us that when we compute a forward orbit of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}, all applications of the pop-stack operator after the first can be computed in the weak order. We believe this could have interesting further implications concerning the dynamical properties of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}.

Corollary 7.10.

Let cc be a Coxeter element of a finite irreducible crystallographic Coxeter group WW. If wCambcw\in\mathrm{Camb}_{c}, then

(popWeak(W))t(popCambc(w))=(popCambc)t+1(w)(\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)})^{t}(\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w))=(\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}})^{t+1}(w)

for all t0t\geq 0.

Proof.

The result is trivial if t=0t=0, so we may assume t1t\geq 1 and assume inductively that

(popWeak(W))t1(popCambc(w))=(popCambc)t(w).(\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)})^{t-1}(\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(w))=(\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}})^{t}(w).

Let u=(popCambc)t(w)u=(\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}})^{t}(w). Since uu is in the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}, we know by Theorem 7.8 that all of the descent of uu commute with each other. Therefore, it follows from Theorem 7.6 that popWeak(W)(u)=popCambc(u)\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(u)=\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(u); this is equivalent to the desired result. ∎

8. Cambrian Lattices in Type A

In this section, we restrict to Cambrian lattices of type A. We first recall the permutation model for the Coxeter group AnA_{n} and the definition of bipartite Coxeter elements. In Sections 8.3 and 8.4, we recall two classes of combinatorial objects: arc diagrams and Motzkin paths. We use these in Section 8.5 to resolve a conjecture (stated in Equation 8.1) of Defant and Williams from [DW23]. This yields an explicit formula for the generating function that counts the images of the pop-stack operators on bipartite Cambrian lattices of type A.

8.1. Permutations and Coxeter elements

The Coxeter group AnA_{n} is the same as the symmetric group whose elements are permutations of the set [n+1]={1,,n+1}[n+1]=\{1,\ldots,n+1\}. We will frequently represent a permutation wAnw\in A_{n} in one-line notation as the word w(1)w(n+1)w(1)\cdots w(n+1). The simple reflections of AnA_{n} are s1,,sns_{1},\ldots,s_{n}, where sis_{i} is the transposition that swaps ii and i+1i+1. The Coxeter graph ΓAn\Gamma_{A_{n}} is a path that contains an (unlabeled) edge {si,si+1}\{s_{i},s_{i+1}\} for each i[n]i\in[n]. A simple reflection sis_{i} is a descent of ww if and only if w(i)>w(i+1)w(i)>w(i+1); when this is the case, we will also refer to the index ii as a descent of ww. A left inversion of ww is a transposition (ij)(i\,\,j) such that i<ji<j and w1(i)>w1(i)w^{-1}(i)>w^{-1}(i).

Let

c1=i[n]i oddsiandc2=i[n]i evensi.c_{1}=\prod_{\begin{subarray}{c}i\in[n]\\ i\text{ odd}\end{subarray}}s_{i}\quad\text{and}\quad c_{2}=\prod_{\begin{subarray}{c}i\in[n]\\ i\text{ even}\end{subarray}}s_{i}.

The bipartite Coxeter elements of AnA_{n} are c(n)×=c1c2c_{(n)}^{\times}=c_{1}c_{2} and c(n)××=c2c1c_{(n)}^{\times\times}=c_{2}c_{1}. These will be the focus of much (but not all) of this section. The following result shows that it generally suffices to consider only c(n)×c^{\times}_{(n)}.

Proposition 8.1.

The canonical join complexes of Cambc(n)×\mathrm{Camb}_{c_{(n)}^{\times}} and Cambc(n)××\mathrm{Camb}_{c_{(n)}^{\times\times}} are isomorphic.

Proof.

The map xwxwx\mapsto w_{\circ}xw_{\circ} is an automorphism of Weak(An)\mathrm{Weak}(A_{n}); when nn is even, it restricts to an isomorphism from Cambc(n)×\mathrm{Camb}_{c_{(n)}^{\times}} to Cambc(n)××\mathrm{Camb}_{c_{(n)}^{\times\times}}. Thus, the desired result is immediate when nn is even. The map xwxx\mapsto w_{\circ}x is an antiautomorphism of Weak(An)\mathrm{Weak}(A_{n}); when nn is odd, it restricts to an isomorphism from Cambc(n)×\mathrm{Camb}_{c_{(n)}^{\times}} to Cambc(n)××\mathrm{Camb}_{c_{(n)}^{\times\times}}. The canonical join complex of the dual of Cambc(n)××\mathrm{Camb}_{c_{(n)}^{\times\times}} is equal to the canonical meet complex of Cambc(n)××\mathrm{Camb}_{c_{(n)}^{\times\times}}. Therefore, when nn is odd, the desired result follows from the fact that the canonical join complex and the canonical meet complex of a semidistributive lattice are isomorphic. ∎

Recall that if LL is a finite semidistributive lattice, then the generating function

𝐏L(q)=vpopL(L)q|𝒰(v)|=vpopL(L)q|𝒟(v)|\mathbf{P}_{L}(q)=\sum_{v\in\mathrm{pop}^{\downarrow}_{L}(L)}q^{|\mathcal{U}(v)|}=\sum_{v\in\mathrm{pop}^{\uparrow}_{L}(L)}q^{|\mathcal{D}(v)|}

defined in Equation 2.1 counts the facets of the canonical join complex of LL according to their sizes. In [DW23, Conjecture 11.2], Defant and Williams conjectured555This conjecture was stated in [DW23] in terms of an explicit formula for 𝐏Cambc(n)×(q)\mathbf{P}_{\mathrm{Camb}_{c^{\times}_{(n)}}}(q) for each particular nn, but we prefer to write it here in terms of the generating function that encompasses all n1n\geq 1. that

(8.1) n1𝐏Cambc(n)×(q)zn=1qz(21qz(12z)+1+q2z22qz(1+2z)1)1.\sum_{n\geq 1}\mathbf{P}_{\mathrm{Camb}_{c^{\times}_{(n)}}}(q)z^{n}=\frac{1}{qz}\left(\frac{2}{1-qz(1-2z)+\sqrt{1+q^{2}z^{2}-2qz(1+2z)}}-1\right)-1.

We will prove this conjecture in Section 8.5.

8.2. The image of pop-stack in AnA_{n}

For this subsection, we let cc denote an arbitrary (not necessarily bipartite) Coxeter element of AnA_{n}. Define νc:{2,,n}{𝐀,𝐁}\nu_{c}\colon\{2,\ldots,n\}\to\{\mathbf{A},\mathbf{B}\} by

(8.2) νc(i)={𝐀if si precedes si1 in every reduced word for c;𝐁if si1 precedes si in every reduced word for c.\nu_{c}(i)=\begin{cases}\mathbf{A}&\mbox{if }s_{i}\mbox{ precedes }s_{i-1}\mbox{ in every reduced word for }c;\\ \mathbf{B}&\mbox{if }s_{i-1}\mbox{ precedes }s_{i}\mbox{ in every reduced word for }c.\end{cases}

The map cνcc\mapsto\nu_{c} is a bijection from the set of Coxeter elements of AnA_{n} to the set of functions from {2,,n}\{2,\ldots,n\} to {𝐀,𝐁}\{\mathbf{A},\mathbf{B}\}.

Theorem 8.2 ([Rea15, Example 4.9]).

Let cc be a Coxeter element of AnA_{n}. An element wAnw\in A_{n} is cc-sortable if and only if the following hold for all i,j[n+1]i,j\in[n+1] such that w(j+1)<w(i)<w(j)w(j+1)<w(i)<w(j):

  • If νc(i)=𝐀\nu_{c}(i)=\mathbf{A}, then j<ij<i.

  • If νc(i)=𝐁\nu_{c}(i)=\mathbf{B}, then i<ji<j.

For wAnw\in A_{n}, we recall that a simple reflection sis_{i} is in Des(w)\mathrm{Des}(w) if and only if w(i)>w(i+1)w(i)>w(i+1). We say that ii is a double descent of ww if ini\neq n and w(i)>w(i+1)>w(i+2)w(i)>w(i+1)>w(i+2). Note that the descents of ww all commute with one another if and only if ww does not have any double descents.

Example 8.3.

Let c=s1s2snAnc=s_{1}s_{2}\cdots s_{n}\in A_{n}. Then νc(i)=𝐁\nu_{c}(i)=\mathbf{B} for all i{2,,n}i\in\{2,\ldots,n\}. A permutation wAnw\in A_{n} is cc-sortable if and only if it avoids the pattern 312312 (i.e., there do not exist indices i1<i2<i3i_{1}<i_{2}<i_{3} such that w(i2)<w(i3)<w(i1)w(i_{2})<w(i_{3})<w(i_{1})). The cc-Cambrian lattice is often called the (n+1)(n+1)-st Tamari lattice.

The one-line notation of c1c^{-1} is (n+1)123n(n+1)123\cdots n. Hence, the left inversions of c1c^{-1} are the transpositions of the form (in+1)(i\,\,n+1) for i[n]i\in[n]. It follows from Theorem 7.8 that a cc-sortable permutation ww is in the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} if and only if ww has no double descents and the one-line notation of ww ends with n+1n+1. This recovers one of the main results of [Hon22].

Example 8.4.

Let cc be the bipartite Coxeter element c(7)×c^{\times}_{(7)} of A7A_{7}. We have

νc(i)={𝐀if i1(mod2);𝐁if i0(mod2)\nu_{c}(i)=\begin{cases}\mathbf{A}&\text{if }i\equiv 1\pmod{2};\\ \mathbf{B}&\text{if }i\equiv 0\pmod{2}\end{cases}

for all i{2,,7}i\in\{2,\ldots,7\}.

One can readily compute that the left inversions of c1c^{-1} are

s2=(2  3),s4=(4  5),s6=(6  7),s_{2}=(2\,\,3),\quad s_{4}=(4\,\,5),\quad s_{6}=(6\,\,7),
s2s1s2=(1  3),s2s4s3s4s2=(2  5),s4s6s5s6s4=(4  7),s6s7s6=(6  8).s_{2}s_{1}s_{2}=(1\,\,3),\quad s_{2}s_{4}s_{3}s_{4}s_{2}=(2\,\,5),\quad s_{4}s_{6}s_{5}s_{6}s_{4}=(4\,\,7),\quad s_{6}s_{7}s_{6}=(6\,\,8).

According to Theorem 7.8, a cc-sortable permutation ww is in the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} if and it has no double descents and does not have any of the seven transpositions in this list as left inversion.

A straightforward extension of Example 8.4 yields the following corollary to Theorem 7.8, which settles [CS, Conjecture 4.2].

Corollary 8.5.

An element wCambc(n)×w\in\mathrm{Camb}_{c^{\times}_{(n)}} is in the image of popCambc(n)×\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c^{\times}_{(n)}}} if and only if all of the following hold:

  • ww does not have any double descents;

  • w1(2k)<w1(2k+1)w^{-1}(2k)<w^{-1}(2k+1) for all integers kk with 32k+1n+13\leq 2k+1\leq n+1;

  • w1(2k)<w1(2k+3)w^{-1}(2k)<w^{-1}(2k+3) for all integers kk with 52k+3n+15\leq 2k+3\leq n+1;

  • w1(n1)<w1(n+1)w^{-1}(n-1)<w^{-1}(n+1) if n1(mod2)n\equiv 1\pmod{2};

  • w1(1)<w1(3)w^{-1}(1)<w^{-1}(3) if n1n\neq 1.

8.3. Arc diagrams

Arrange n+1n+1 points along a horizontal line, and identify them with the numbers 1,,n+11,\ldots,n+1 from left to right. An arc is a curve that moves monotonically rightward from a point ii to another point jj (for some i<ji<j), passing above or below each of the points i+1,,j1i+1,\ldots,j-1. Two arcs are considered to be the same if they have the same endpoints and they pass above the same collection of numbered points. A noncrossing arc diagram (of type AnA_{n}) is a collection of arcs that can be drawn so that no two arcs have the same left endpoint or have the same right endpoint or cross in their interiors. We write |δ||\delta| for the number of arcs in a noncrossing arc diagram δ\delta. Let ADn\mathrm{AD}_{n} be the set of noncrossing arc diagrams of type AnA_{n}.

Given a permutation wAnw\in A_{n}, form a noncrossing arc diagram Δ(w)ADn\Delta(w)\in\mathrm{AD}_{n} as follows. For each descent ii of ww, draw an arc from w(i+1)w(i+1) to w(i)w(i) such that for each integer kk satisfying w(i+1)<k<w(i){w(i+1)<k<w(i)}, the arc passes above (resp. below) the point kk if w1(k)>i+1w^{-1}(k)>i+1 (resp. w1(k)<i{w^{-1}(k)<i}). This defines a map Δ:AnADn\Delta\colon A_{n}\to\mathrm{AD}_{n}, and it is straightforward to check that Δ\Delta is a bijection. (See [Rea15].)

Given a Coxeter element cc of AnA_{n}, say an arc 𝔞\mathfrak{a} with left endpoint ii and right endpoint jj is cc-sortable if for every k{i+1,,j1}k\in\{i+1,\ldots,j-1\}, 𝔞\mathfrak{a} passes above (resp. below) kk if νc(k)=𝐀\nu_{c}(k)=\mathbf{A} (resp. νc(k)=𝐁\nu_{c}(k)=\mathbf{B}). Note that for all 1i<jn+11\leq i<j\leq n+1, there is a unique cc-sortable arc from ii to jj. Let AD(c)=Δ(Cambc)\mathrm{AD}(c)=\Delta(\mathrm{Camb}_{c}) be the set of noncrossing arc diagrams of cc-sortable elements of AnA_{n}. It is immediate from Theorem 8.2 that a noncrossing arc diagram is in AD(c)\mathrm{AD}(c) if and only if all of its arcs are cc-sortable. Hence, AD(c)\mathrm{AD}(c) is a simplicial complex whose vertices are the cc-sortable arcs.

Example 8.6.

Suppose n=8n=8, and let c=s1s2s4s8s3s5s7s6c=s_{1}s_{2}s_{4}s_{8}s_{3}s_{5}s_{7}s_{6}. Then νc1(𝐀)={4,7,8}\nu_{c}^{-1}(\mathbf{A})=\{4,7,8\}, and νc1(𝐁)={2,3,5,6}\nu_{c}^{-1}(\mathbf{B})=\{2,3,5,6\}. Let ww be the permutation 325148679325148679. The noncrossing arc diagram Δ(w)\Delta(w) is depicted in Figure 2. In this figure, each of the points i[9]i\in[9] is represented by a circle filled with the number ii; for 2i82\leq i\leq 8, a blue semicircle appears on the top (resp. bottom) of the circle if νc(i)=𝐀\nu_{c}(i)=\mathbf{A} (resp. if νc(i)=𝐁\nu_{c}(i)=\mathbf{B}). All of the arcs in Δ(w)\Delta(w) are cc-sortable, so ww is a cc-sortable permutation.

Refer to caption
Figure 2. The noncrossing arc diagram Δ(325148679)\Delta(325148679). For each 2i82\leq i\leq 8, a blue semicircle appears on the top (resp. bottom) of the circle containing ii if νc(i)=𝐀\nu_{c}(i)=\mathbf{A} (resp. if νc(i)=𝐁\nu_{c}(i)=\mathbf{B}), where c=s1s2s4s8s3s5s7s6c=s_{1}s_{2}s_{4}s_{8}s_{3}s_{5}s_{7}s_{6}.
Remark 8.7.

Noncrossing arc diagrams and other similar objects have been used to model the bricks over hereditary and preprojective algebras of type A. In particular, (cc-sortable) arcs are in bijection with bricks, and (cc-sortable) noncrossing arc diagrams are in bijection with semibricks. One can then use noncrossing arc diagrams to compute information about morphisms and extensions between bricks, and thus also to model 2-term simple-minded collections. See, e.g., [BCZ19, BH22, Eno21, GIMO19, Hanb, HY23, Miz22] and references therein.

Cambrian lattices are semidistributive, so we can consider the canonical join complex and the canonical meet complex of Cambc\mathrm{Camb}_{c} (and we know these simplicial complexes are isomorphic by [Bar19, Corollary 5]). An element vCambcv\in\mathrm{Camb}_{c} is join-irreducible if and only if it has exactly one descent, and this occurs if and only if Δ(v)\Delta(v) contains a single arc. Therefore, Δ\Delta establishes a one-to-one correspondence between the join-irreducible elements of Cambc\mathrm{Camb}_{c} and the cc-sortable arcs. Then for each wCambcw\in\mathrm{Camb}_{c}, the noncrossing arc diagram Δ(w)\Delta(w) corresponds to the canonical join representation of ww. It follows that the simplicial complex AD(c)\mathrm{AD}(c) is isomorphic to the canonical join complex of Cambc\mathrm{Camb}_{c}. Say a noncrossing arc diagram in AD(c)\mathrm{AD}(c) is maximal if it is a facet of AD(c)\mathrm{AD}(c). In other words, a noncrossing arc diagram in AD(c)\mathrm{AD}(c) is maximal if it is not properly contained in another noncrossing arc diagram in AD(c)\mathrm{AD}(c). Let MAD(c)\mathrm{MAD}(c) denote the set of maximal noncrossing arc diagrams in AD(c)\mathrm{AD}(c).

The preceding discussion yields the identity

(8.3) 𝐏Cambc(q)=δMAD(c)q|δ|.\mathbf{P}_{\mathrm{Camb}_{c}}(q)=\sum_{\delta\in\mathrm{MAD}(c)}q^{|\delta|}.

8.4. Motzkin paths

A Motzkin path is a lattice path in the plane that consists of up (i.e., (1,1)(1,1)) steps, down (i.e., (1,1)(1,-1)) steps, and horizontal (i.e., (1,0)(1,0)) steps, starts at the origin, never passes below the horizontal axis, and ends on the horizontal axis. Let 𝚄\mathtt{U}, 𝙳\mathtt{D}, and 𝙷\mathtt{H} denote up, down, and horizontal steps, respectively. Given a word PP over the alphabet {𝚄,𝙳,𝙷}\{\mathtt{U},\mathtt{D},\mathtt{H}\}, let #𝚄(P)\#_{\mathtt{U}}(P), #𝙳(P)\#_{\mathtt{D}}(P), and #𝙷(P)\#_{\mathtt{H}}(P) denote the number of 𝚄s\mathtt{U}^{\prime}s, the number of 𝙳\mathtt{D}’s, and the number of 𝙷\mathtt{H}’s in PP, respectively. We can think of a Moztkin path as a word MM over the alphabet {𝚄,𝙳,𝙷}\{\mathtt{U},\mathtt{D},\mathtt{H}\} such that #𝚄(M)=#𝙳(M)\#_{\mathtt{U}}(M)=\#_{\mathtt{D}}(M) and #𝚄(P)#𝙳(P)\#_{\mathtt{U}}(P)\geq\#_{\mathtt{D}}(P) for every prefix PP of MM. For example, the word 𝚄𝙷𝙷𝙳𝙷𝚄𝙳𝚄𝙷𝚄𝙳𝙳\mathtt{U}\mathtt{H}\mathtt{H}\mathtt{D}\mathtt{H}\mathtt{U}\mathtt{D}\mathtt{U}\mathtt{H}\mathtt{U}\mathtt{D}\mathtt{D} represents the Motzkin path in Figure 3.

Refer to caption
Figure 3. The Motzkin path 𝚄𝙷𝙷𝙳𝙷𝚄𝙳𝚄𝙷𝚄𝙳𝙳\mathtt{U}\mathtt{H}\mathtt{H}\mathtt{D}\mathtt{H}\mathtt{U}\mathtt{D}\mathtt{U}\mathtt{H}\mathtt{U}\mathtt{D}\mathtt{D}.

A peak of a Motzkin path MM is a point (j,k)(j,k) where an up step in MM ends and a down step in MM begins; the height of this peak is the number kk. If we view MM as a word over {𝚄,𝙳,𝙷}\{\mathtt{U},\mathtt{D},\mathtt{H}\}, then a peak corresponds to a consecutive occurrence of 𝚄𝙳\mathtt{U}\mathtt{D}, and the height of the peak is #𝚄(P)#𝙳(P)\#_{\mathtt{U}}(P)-\#_{\mathtt{D}}(P), where PP is the prefix of MM that ends with the up step involved in the peak. The two peaks of the Motzkin path in Figure 3 are the points (6,1)(6,1) and (10,2)(10,2).

Let n\mathcal{M}_{n} denote the set of Motzkin paths of length nn, and let ¯n\overline{\mathcal{M}}_{n} be the subset of \mathcal{M} consisting of Motzkin paths that do not have any peaks of height 11. Let =n0n\mathcal{M}=\bigcup_{n\geq 0}\mathcal{M}_{n} and ¯=n0¯n\overline{\mathcal{M}}=\bigcup_{n\geq 0}\overline{\mathcal{M}}_{n}. Let

𝐌(q,z)=n0Mnq#𝚄(M)znand𝐌¯(q,z)=n0M¯nq#𝚄(M)zn.{\bf M}(q,z)=\sum_{n\geq 0}\sum_{M\in\mathcal{M}_{n}}q^{\#_{\mathtt{U}}(M)}z^{n}\quad\text{and}\quad{\overline{\bf M}}(q,z)=\sum_{n\geq 0}\sum_{M\in\overline{\mathcal{M}}_{n}}q^{\#_{\mathtt{U}}(M)}z^{n}.

If MM is a nonempty Motzkin path, then either M=𝚄M𝙳M′′M=\mathtt{U}M^{\prime}\mathtt{D}M^{\prime\prime} for some M,M′′M^{\prime},M^{\prime\prime}\in\mathcal{M}, or M=𝙷M′′′M=\mathtt{H}M^{\prime\prime\prime} for some M′′′M^{\prime\prime\prime}\in\mathcal{M}. This translates into the functional equation

𝐌(q,z)1=qz2𝐌(q,z)2+z𝐌(q,z),{\bf M}(q,z)-1=qz^{2}{\bf M}(q,z)^{2}+z{\bf M}(q,z),

which we can solve to find that

𝐌(q,z)=1z12z+(14q)z22qz2.{\bf M}(q,z)=\frac{1-z-\sqrt{1-2z+(1-4q)z^{2}}}{2qz^{2}}.

If MM is a nonempty Motzkin path with no peaks of height 11, then either M=𝚄M𝙳M′′M=\mathtt{U}M^{\prime}\mathtt{D}M^{\prime\prime} for some nonempty MM^{\prime}\in\mathcal{M} and some M′′¯M^{\prime\prime}\in\overline{\mathcal{M}}, or M=𝙷M′′′M=\mathtt{H}M^{\prime\prime\prime} for some M′′′¯M^{\prime\prime\prime}\in\overline{\mathcal{M}}. This translates into the functional equation

𝐌¯(q,z)1=qz2(𝐌(q,z)1)𝐌¯(q,z)+z𝐌¯(q,z).\overline{{\bf M}}(q,z)-1=qz^{2}({\bf M}(q,z)-1)\overline{\bf M}(q,z)+z\overline{{\bf M}}(q,z).

We can solve this equation and use the above formula for 𝐌(q,z){\bf M}(q,z) to obtain

𝐌¯(q,z)\displaystyle\overline{{\bf M}}(q,z) =11zqz2(𝐌(z,q)1)\displaystyle=\frac{1}{1-z-qz^{2}({\bf M}(z,q)-1)}
(8.4) =21z+2qz2+12z+(14q)z2.\displaystyle=\frac{2}{1-z+2qz^{2}+\sqrt{1-2z+(1-4q)z^{2}}}.

Using Equation 8.4, one can readily check that the expression on the right-hand side of Equation 8.1 is

1qz(𝐌¯(1/q,qz)1)1=n1M¯n+1qn#𝚄(M)zn.\frac{1}{qz}\left(\overline{\bf M}(1/q,qz)-1\right)-1=\sum_{n\geq 1}\sum_{M\in\overline{\mathcal{M}}_{n+1}}q^{n-\#_{\mathtt{U}}(M)}z^{n}.

Therefore, in order to prove Equation 8.1, it suffices (by Equation 8.3) to exhibit a bijection Ψ:MAD(c(n)×)¯n+1\Psi\colon\mathrm{MAD}(c_{(n)}^{\times})\to\overline{\mathcal{M}}_{n+1} such that |δ|=n#𝚄(Ψ(δ))|\delta|=n-\#_{\mathtt{U}}(\Psi(\delta)) for every δMAD(c(n)×)\delta\in\mathrm{MAD}(c_{(n)}^{\times}).

8.5. The bijection

Throughout the remainder of this section, let us fix a positive integer nn and write c×=c(n)×c^{\times}=c_{(n)}^{\times}. The map νc×:{2,,n}{𝐀,𝐁}\nu_{c^{\times}}\colon\{2,\ldots,n\}\to\{\mathbf{A},\mathbf{B}\} is given by

νc×(i)={𝐀if i is odd;𝐁if i is even.\nu_{c^{\times}}(i)=\begin{cases}\mathbf{A}&\mbox{if }i\mbox{ is odd};\\ \mathbf{B}&\mbox{if }i\mbox{ is even}.\end{cases}

As in Section 8.3, we consider noncrossing arc diagrams drawn on n+1n+1 points that are arranged horizontally and identified with the numbers 1,,n+11,\ldots,n+1 from left to right.

Lemma 8.8.

If δMAD(c×)\delta\in\mathrm{MAD}(c^{\times}) and i{2,,n}i\in\{2,\ldots,n\}, then i1i-1 is the left endpoint of an arc in δ\delta, or i+1i+1 is the right endpoint of an arc in δ\delta (or both).

Proof.

Suppose instead that i1i-1 is not a left endpoint of an arc in δ\delta and that i+1i+1 is not a right endpoint of an arc in δ\delta. Let 𝔞\mathfrak{a} be the unique c×c^{\times}-sortable arc whose endpoints are i1i-1 and i+1i+1. Because νc×(i1)=νc×(i+1)νc×(i)\nu_{c^{\times}}(i-1)=\nu_{c^{\times}}(i+1)\neq\nu_{c^{\times}}(i), it is straightforward to check that δ{𝔞}\delta\sqcup\{\mathfrak{a}\} is a noncrossing arc diagram in AD(c×)\mathrm{AD}(c^{\times}). This contradicts the maximality of δ\delta. ∎

Suppose δMAD(c×)\delta\in\mathrm{MAD}(c^{\times}). Let Ψ(δ)\Psi(\delta) be the word 𝙼1𝙼n+1\mathtt{M}_{1}\cdots\mathtt{M}_{n+1}, where for 1in+11\leq i\leq n+1, we define

(8.5) 𝙼i={𝚄if in and i+1 is not the right endpoint of an arc in δ;𝙳if i2 and i1 is not the left endpoint of an arc in δ;𝙷otherwise.\mathtt{M}_{i}=\begin{cases}\mathtt{U}&\mbox{if }i\leq n\mbox{ and }i+1\mbox{ is not the right endpoint of an arc in }\delta;\\ \mathtt{D}&\mbox{if }i\geq 2\mbox{ and }i-1\mbox{ is not the left endpoint of an arc in }\delta;\\ \mathtt{H}&\mbox{otherwise}.\end{cases}

Lemma 8.8 guarantees that Ψ(δ)\Psi(\delta) is well defined (i.e., that no letter in Ψ(δ)\Psi(\delta) can be both 𝚄\mathtt{U} and 𝙳\mathtt{D}). See Figure 4 for an illustration.

Refer to caption
Figure 4. When n=13n=13, the map Ψ\Psi sends a noncrossing arc diagram of type A13A_{13} to a Motzkin path of length 1414 with no peaks of height 11. For each 2i132\leq i\leq 13, a blue semicircle appears on the top (resp. bottom) of the circle containing ii if νc×(i)=𝐀\nu_{c^{\times}}(i)=\mathbf{A} (resp. if νc×(i)=𝐁\nu_{c^{\times}}(i)=\mathbf{B}). (The letters drawn below the noncrossing arc diagram represent the Moztkin path; they are not part of the noncrossing arc diagram.)
Proposition 8.9.

For each δMAD(c×)\delta\in\mathrm{MAD}(c^{\times}), we have Ψ(δ)¯n+1\Psi(\delta)\in\overline{\mathcal{M}}_{n+1} and |δ|=n#𝚄(Ψ(δ))|\delta|=n-\#_{\mathtt{U}}(\Psi(\delta)).

Proof.

Let Ψ(δ)=𝙼1𝙼n+1\Psi(\delta)=\mathtt{M}_{1}\cdots\mathtt{M}_{n+1}, where the steps 𝙼1,,𝙼n+1\mathtt{M}_{1},\ldots,\mathtt{M}_{n+1} are as defined in Equation 8.5. It is immediate from Equation 8.5 that n#𝚄(Ψ(δ))n-\#_{\mathtt{U}}(\Psi(\delta)) and n#𝙳(Ψ(δ))n-\#_{\mathtt{D}}(\Psi(\delta)) are both equal to |δ||\delta|. In particular, #𝚄(Ψ(δ))=#𝙳(Ψ(δ))\#_{\mathtt{U}}(\Psi(\delta))=\#_{\mathtt{D}}(\Psi(\delta)).

Suppose k{0,,n}k\in\{0,\ldots,n\} is such that the prefix P=𝙼1𝙼kP=\mathtt{M}_{1}\cdots\mathtt{M}_{k} of Ψ(δ)\Psi(\delta) satisfies #U(P)=#𝙳(P)\#_{U}(P)=\#_{\mathtt{D}}(P). If k1k\geq 1, then k#𝚄(P)k-\#_{\mathtt{U}}(P) is the number of elements of {2,,k+1}\{2,\ldots,k+1\} that are right endpoints of arcs in δ\delta. Similarly, if k1k\geq 1, then k1#𝙳(P)k-1-\#_{\mathtt{D}}(P) is the number of elements of {1,,k1}\{1,\ldots,k-1\} that are left endpoints of arcs in δ\delta. It follows that either k=0k=0, or there is an arc from kk to k+1k+1 in δ\delta. In either case, it is immediate from Equation 8.5 that 𝙼k+1𝙳\mathtt{M}_{k+1}\neq\mathtt{D}. Moreover, none of the arcs with left endpoints in {1,,k}\{1,\ldots,k\} have right endpoints in {k+2,,n+1}\{k+2,\ldots,n+1\}. If we had 𝙼k+1=𝚄\mathtt{M}_{k+1}=\mathtt{U} and 𝙼k+2=𝙳\mathtt{M}_{k+2}=\mathtt{D}, then we could add the arc from k+1k+1 to k+2k+2 to δ\delta in order to form a larger noncrossing arc diagram in AD(c×)\mathrm{AD}(c^{\times}), contradicting the maximality of δ\delta. Together, this shows that Ψ(δ)\Psi(\delta) cannot pass below the horizontal axis and cannot have any peaks of height 11. Hence, Ψ(δ)¯n+1\Psi(\delta)\in\overline{\mathcal{M}}_{n+1}. ∎

We can now state and prove the main theorem of this section; as mentioned at the end of Section 8.4, this theorem and Proposition 8.9 imply the identity in Equation 8.1, thereby settling the conjecture of Defant and Williams.

Theorem 8.10.

The map Ψ:MAD(c×)¯n+1\Psi\colon\mathrm{MAD}(c^{\times})\to\overline{\mathcal{M}}_{n+1} is a bijection.

Proof.

Given an arc 𝔞\mathfrak{a} and a vertical line 𝔩\mathfrak{l} that lies strictly between the endpoints of 𝔞\mathfrak{a}, let 𝔞𝔩\mathfrak{a}_{\mathfrak{l}} be the curve obtained from 𝔞\mathfrak{a} by deleting the portion of 𝔞\mathfrak{a} that lies to the right of 𝔩\mathfrak{l}. We call a curve 𝔞𝔩\mathfrak{a}_{\mathfrak{l}} obtained in this manner a partial arc.

Consider M=𝙼1𝙼n+1¯n+1M=\mathtt{M}_{1}\cdots\mathtt{M}_{n+1}\in\overline{\mathcal{M}}_{n+1}. Let us try to construct δMAD(c×){\delta\in\mathrm{MAD}(c^{\times})} such that Ψ(δ)=M\Psi(\delta)=M. We can build δ\delta step by step from left to right. This process is illustrated in Figure 5. For the first step, we create a small partial arc that starts at the point 11. In general, at the kk-th step, we consider the point kk. At this point in time, there may be some partial arcs that were produced in the previous k1k-1 steps. If k2k\geq 2 and 𝙼k1𝚄\mathtt{M}_{k-1}\neq\mathtt{U}, then we extend one of these partial arcs and attach it to the point kk in order to create an arc, and we extend the rest of the partial arcs above (if kk is odd) or below (if kk is even) the point kk. The condition that arcs cannot cross uniquely determines which partial arc we must attach to the point kk. During the same kk-th step, if knk\leq n and 𝙼k+1𝙳\mathtt{M}_{k+1}\neq\mathtt{D}, then we create a new small partial arc that starts at the point kk.

To see that the above procedure succeeds in producing a noncrossing arc diagram in AD(c×)\mathrm{AD}(c^{\times}), we must check two things:

  • If k2k\geq 2 and 𝙼k1𝚄\mathtt{M}_{k-1}\neq\mathtt{U}, then there is actually a partial arc from the first k1k-1 steps that we can attach to the point kk.

  • The total number of steps during which we create a new partial arc is equal to the total number of steps during which we attach a partial arc to a point to create an arc.

For the first item, note that the number of partial arcs remaining after the first k1k-1 steps is (k1#𝙳(𝙼2𝙼k))(k2#𝚄(𝙼1𝙼k2))=#𝚄(𝙼1𝙼k2)#𝙳(𝙼2𝙼k)+1(k-1-\#_{\mathtt{D}}(\mathtt{M}_{2}\cdots\mathtt{M}_{k}))-(k-2-\#_{\mathtt{U}}(\mathtt{M}_{1}\cdots\mathtt{M}_{k-2}))=\#_{\mathtt{U}}(\mathtt{M}_{1}\cdots\mathtt{M}_{k-2})-\#_{\mathtt{D}}(\mathtt{M}_{2}\cdots\mathtt{M}_{k})+1. Because MM is a Motzkin path, this quantity is nonnegative; furthermore, it can only be 0 if the steps 𝙼k1\mathtt{M}_{k-1} and 𝙼k\mathtt{M}_{k} are 𝚄\mathtt{U} and 𝙳\mathtt{D} (respectively) and form a peak of height 11. But MM has no peaks of height 11, so the number of partial arcs remaining after k1k-1 steps must actually be at least 11. The second bulleted item follows from the fact that #𝚄(M)=#𝙳(M)\#_{\mathtt{U}}(M)=\#_{\mathtt{D}}(M).

Let us now show that the noncrossing arc diagram δ\delta constructed through the above procedure is actually maximal in AD(c×)\mathrm{AD}(c^{\times}); it will then follow immediately from the construction that Ψ(δ)=M\Psi(\delta)=M. Note also that the steps outlined above for constructing δ\delta were forced upon us; in other words, Ψ\Psi is injective.

Let 𝔞i,j\mathfrak{a}_{i,j} denote the unique c×c^{\times}-sortable arc from ii to jj. Suppose by way of contradiction that there exist i<ji<j such that 𝔞i,j\mathfrak{a}_{i,j} is not in δ\delta and δ{𝔞i,j}\delta\sqcup\{\mathfrak{a}_{i,j}\} is a noncrossing arc diagram. We will assume ii is even; the case when ii is odd is almost identical. We consider three cases.

Case 1. Assume j=i+1j=i+1. Then we have 𝙼i=𝚄\mathtt{M}_{i}=\mathtt{U} and 𝙼i+1=𝙳\mathtt{M}_{i+1}=\mathtt{D}, so the steps 𝙼i\mathtt{M}_{i} and 𝙼i+1\mathtt{M}_{i+1} form a peak. Since 𝔞i,i+1\mathfrak{a}_{i,i+1} does not cross any of the arcs in δ\delta, none of the arcs in δ\delta can pass below ii and above i+1i+1. Hence, every arc in δ\delta whose left endpoint is in {1,,i1}\{1,\ldots,i-1\} has its right endpoint in {2,,i}\{2,\ldots,i\}. It follows from the construction of δ\delta that #𝚄(𝙼1𝙼i1)=#𝙳(𝙼1𝙼i1)\#_{\mathtt{U}}(\mathtt{M}_{1}\cdots\mathtt{M}_{i-1})=\#_{\mathtt{D}}(\mathtt{M}_{1}\cdots\mathtt{M}_{i-1}). But this implies that the peak formed by 𝙼i\mathtt{M}_{i} and 𝙼i+1\mathtt{M}_{i+1} has height 11, which contradicts the fact that MM has no peaks of height 11.

Case 2. Assume j=i+2j=i+2. Then the point ii is not the left endpoint of an arc in δ\delta, and the point i+2i+2 is not the right endpoint of an arc in δ\delta. However, this implies that 𝙼i+1\mathtt{M}_{i+1} is both 𝚄\mathtt{U} and 𝙳\mathtt{D}, which is impossible.

Case 3. Assume ji+3j\geq i+3. Note that 𝙼i+1=𝙳\mathtt{M}_{i+1}=\mathtt{D}. Since 𝙼i+1𝚄\mathtt{M}_{i+1}\neq\mathtt{U}, there must be an arc 𝔞\mathfrak{a}^{\prime} in δ\delta whose right endpoint is i+2i+2. The arc 𝔞i,j\mathfrak{a}_{i,j} passes above i+1i+1 and below i+2i+2. The left endpoint of 𝔞\mathfrak{a}^{\prime} cannot be i+1i+1 since, if it were, the arcs 𝔞i,j\mathfrak{a}_{i,j} and 𝔞\mathfrak{a}^{\prime} would cross. Therefore, 𝔞\mathfrak{a}^{\prime} passes above i+1i+1. The lower endpoint of 𝔞\mathfrak{a}^{\prime} also cannot be ii because ii is the left endpoint of 𝔞i,j\mathfrak{a}_{i,j}. Therefore, 𝔞\mathfrak{a}^{\prime} passes below ii. However, this forces 𝔞i,j\mathfrak{a}_{i,j} and 𝔞\mathfrak{a}^{\prime} to cross, which is a contradiction. ∎

Refer to caption
Figure 5. The step-by-step procedure for constructing the noncrossing arc diagram δ\delta such that Ψ(δ)=𝚄𝙷𝚄𝚄𝙳𝙷𝙳𝙳\Psi(\delta)=\mathtt{U}\mathtt{H}\mathtt{U}\mathtt{U}\mathtt{D}\mathtt{H}\mathtt{D}\mathtt{D}.

9. Maximum-Size Pop-Stack Orbits

Let WW be a finite irreducible Coxeter group. All Coxeter elements of WW are conjugate to each other, so they have the same group-theoretic order; this order is called the Coxeter number of WW and is denoted by hh. (It is also known that h=2N/nh=2N/n, where NN and nn are the number of reflection in WW and the number of simple reflections in WW, respectively.) Recall that, given a finite lattice LL, we write 𝒪L(x)\mathcal{O}_{L}(x) for the forward orbit of an element xLx\in L under the pop-stack operator. If cc is a Coxeter element of WW, then Cambc\mathrm{Camb}_{c} is a quotient of Weak(W)\mathrm{Weak}(W). Let LL be an arbitrary lattice quotient WW_{\equiv} on WW. We begin this section by showing that hh is an upper bound for maxxW|𝒪W(x)|\max_{x\in W_{\equiv}}|\mathcal{O}_{W_{\equiv}}(x)| when LL is an arbitrary quotient of Weak(W)\mathrm{Weak}(W) (Theorem 9.6). In the process, we also provide a formula for computing popW\mathrm{pop}_{W_{\equiv}} for any lattice quotient WW_{\equiv} of the weak order. We then prove that this upper bound is tight when WW_{\equiv} is a Cambrian lattice (Theorem 9.17). We also discuss how the techniques used in this section relate back to representation theory in Section 9.4.

9.1. Quotients of the weak order

As above, let LL be a finite lattice. An equivalence relation \equiv on LL is a lattice congruence provided that equivalence classes respect the meet and join operations in LL. That is, \equiv is a lattice congruence if for all x,y,zLx,y,z\in L satisfying xyx\equiv y, we have xzyzx\vee z\equiv y\vee z and xzyzx\wedge z\equiv y\wedge z.

Let \equiv be a lattice congruence. Let [x][x]_{\equiv} denote the equivalence class of an element xLx\in L under \equiv. One can define a lattice structure on the set L/L/\equiv of equivalence classes as follows. We set [x][y][x]_{\equiv}\leq[y]_{\equiv} if and only if there exist x[x]x^{\prime}\in[x]_{\equiv} and y[y]y^{\prime}\in[y]_{\equiv} such that xyx^{\prime}\leq y^{\prime}. We then have [x][y]=[xy][x]_{\equiv}\wedge[y]_{\equiv}=[x\wedge y]_{\equiv} and [x][y]=[xy][x]_{\equiv}\vee[y]_{\equiv}=[x\vee y]_{\equiv}. The lattice L/L/\equiv is called a lattice quotient of LL. Observe that there is a surjective lattice homomorphism η:LL/\eta_{\equiv}\colon L\to L/\equiv that sends each element xLx\in L to its equivalence class [x][x]_{\equiv}. (See [Rea16, Proposition 9.5-4].)

Given an equivalence relation \equiv on LL, one can check that \equiv is a lattice congruence order-theoretically using the following result.

Proposition 9.1 ([Rea16, Proposition 9-5.2]).

An equivalence relation \equiv on a finite lattice LL is a lattice congruence if and only if the following three conditions hold:

  • Each equivalence class is an interval of LL.

  • The map π\pi_{\downarrow}^{\equiv} that sends each element xLx\in L to the unique minimal element of its equivalence class is order-preserving.

  • The map π\pi^{\uparrow}_{\equiv} that sends each element xLx\in L to the unique maximal element of its equivalence class is order-preserving.

Finally, we note that if \equiv is a lattice congruence on LL, then the map π:LL\pi_{\downarrow}^{\equiv}\colon L\to L is a lattice homomorphism whose image π(L)\pi_{\downarrow}^{\equiv}(L) (taken as an induced subposet of LL) is isomorphic to the quotient L/L/\equiv. (Dually, the image of π\pi^{\uparrow}_{\equiv} is also isomorphic to the quotient L/L/\equiv.)

Remark 9.2.

Although the subposet π(L)\pi_{\downarrow}^{\equiv}(L) is a lattice in its own right, it meet and join operations may not coincide with those of LL. Said another way, π(L)\pi_{\downarrow}^{\equiv}(L) is not generally a sublattice of LL. In general, the join operation of π(L)\pi_{\downarrow}^{\equiv}(L) does coincide with that of LL; dually, the meet operation of π(L)\pi^{\uparrow}_{\equiv}(L) does coincide with that of LL. However, each cc-Cambrian lattice (of type WW) is both a lattice quotient and sublattice of the weak order on WW.

Below, we realize an arbitrary lattice quotient of the weak order as the subposet induced by the image of π\pi_{\downarrow}^{\equiv}, where \equiv is a lattice congruence. We have two main tasks as we consider the pop-stack operator on π(W)\pi_{\downarrow}^{\equiv}(W). First, we must we identify the lower covers of each xπ(W)x\in\pi_{\downarrow}^{\equiv}(W); this is done in the following lemma.

Lemma 9.3.

Let LL be a finite lattice, and let \equiv be a lattice congruence on LL. If xπ(L)x\in\pi_{\downarrow}^{\equiv}(L), then the map π\pi_{\downarrow}^{\equiv} restricts to a bijection from the set of elements covered by xx in LL to the set of elements covered by xx in π(L)\pi_{\downarrow}^{\equiv}(L).

Proof.

First assume that axa\lessdot x in LL, and let a=π(a)a^{\prime}=\pi_{\downarrow}^{\equiv}(a). Since π:LL\pi_{\downarrow}^{\equiv}\colon L\to L is order-preserving, we have axa^{\prime}\leq x. Suppose that there exists bπ(L)b\in\pi_{\downarrow}^{\equiv}(L) with a<b<xa^{\prime}<b<x. We cannot have a<b<aa^{\prime}<b<a because aa^{\prime} is the largest element of π(L)\pi_{\downarrow}^{\equiv}(L) lying below aa. Moreover, we cannot have a<a<ba^{\prime}<a<b because a<xa{\,\,<\!\!\!\!\cdot\,\,\,}x. Therefore, aa and bb are incomparable. This implies that ab=xa\vee b=x (because a<xa{\,\,<\!\!\!\!\cdot\,\,\,}x and b<xb<x). Since aaa\equiv a^{\prime}, we have x=abab=bx=a\vee b\equiv a^{\prime}\vee b=b. But this contradicts the fact that bb and xx are distinct elements of π(L)\pi_{\downarrow}^{\equiv}(L). From this contradiction, we deduce that a<xa^{\prime}{\,\,<\!\!\!\!\cdot\,\,\,}x in π(L)\pi_{\downarrow}^{\equiv}(L).

We have shown that π\pi_{\downarrow}^{\equiv} restricts to a map from the set of elements covered by xx in LL to the set of elements covered by xx in π(L)\pi_{\downarrow}^{\equiv}(L); we must show that it is bijective. To show it is injective, suppose instead that there exist distinct elements a1,a2a_{1},a_{2} that are covered by xx in LL such that π(a1)=π(a2)\pi_{\downarrow}^{\equiv}(a_{1})=\pi_{\downarrow}^{\equiv}(a_{2}). Then a1a2a_{1}\equiv a_{2}, so a1a1a2=xa_{1}\equiv a_{1}\vee a_{2}=x, contradicting the fact that xx is the unique minimal element of [x][x]_{\equiv}. To prove surjectivity, suppose d<xd{\,\,<\!\!\!\!\cdot\,\,\,}x in π(L)\pi_{\downarrow}^{\equiv}(L). Then there exists dLd^{\prime}\in L such that dd<xd\leq d^{\prime}{\,\,<\!\!\!\!\cdot\,\,\,}x. Since π\pi_{\downarrow}^{\equiv} is order-preserving, we have dπ(d)<xd\leq\pi_{\downarrow}^{\equiv}(d^{\prime})<x. Thus, d=π(d)d=\pi_{\downarrow}^{\equiv}(d^{\prime}). ∎

Now that we have identified the lower covers of the elements of a lattice quotient π(L)\pi_{\downarrow}^{\equiv}(L), we describe the pop-stack operator on a lattice quotient of LL. It will be convenient to write L\wedge_{L} and \wedge_{\equiv} for the meet operation of LL and the meet operation of π(L)\pi_{\downarrow}^{\equiv}(L), respectively.

Proposition 9.4.

Let LL be a finite lattice, and let \equiv be a lattice congruence on LL. For xπ(L)x\in\pi_{\downarrow}^{\equiv}(L), we have

popπ(L)(x)=π(L{π(a)a<x in L}).\mathrm{pop}^{\downarrow}_{\pi_{\downarrow}^{\equiv}(L)}(x)=\pi_{\downarrow}^{\equiv}\left(\bigwedge\nolimits_{L}\{\pi_{\downarrow}^{\equiv}(a)\mid a{\,\,<\!\!\!\!\cdot\,\,\,}x\text{ in $L$}\}\right).
Proof.

By Lemma 9.3, it suffices to show that for x,yπ(L)x,y\in\pi_{\downarrow}^{\equiv}(L), we have xy=π(xLy)x\wedge_{\equiv}y=\pi_{\downarrow}^{\equiv}(x\wedge_{L}y). Observe that xLyxyπ(xLy)x\wedge_{L}y\geq x\wedge_{\equiv}y\geq\pi_{\downarrow}^{\equiv}(x\wedge_{L}y) in LL. Since xyx\wedge_{\equiv}y and π(xLy)\pi_{\downarrow}^{\equiv}(x\wedge_{L}y) are both in the image of π\pi_{\downarrow}^{\equiv}, the desired result follows from the fact that π(xLy)\pi_{\downarrow}^{\equiv}(x\wedge_{L}y) is the largest element of π(L)\pi_{\downarrow}^{\equiv}(L) lying below xLyx\wedge_{L}y. ∎

Remark 9.5.

Let LL be a the weak order on a finite Coxeter group WW. Let \equiv be the cc-Cambrian congruence on WW. The fact that each cc-Cambrian lattice is a sublattice of the weak order implies that π\pi_{\downarrow}^{\equiv} and the meet operation of the weak order commute (see, e.g., [Def22a, Lemma 3.1]). Hence,

popCambc(x)\displaystyle\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(x) =π(L{π(a):a<x in L})=π({π(a):a<x in L})\displaystyle=\pi_{\downarrow}^{\equiv}\left(\bigwedge\nolimits_{L}\{\pi_{\downarrow}^{\equiv}(a):a{\,\,<\!\!\!\!\cdot\,\,\,}x\text{ in $L$}\}\right)=\pi_{\downarrow}^{\equiv}\left(\bigwedge\nolimits_{\equiv}\{\pi_{\downarrow}^{\equiv}(a):a{\,\,<\!\!\!\!\cdot\,\,\,}x\text{ in $L$}\}\right)
={π(a):a<x in L}=L{π(a):a<x in L}\displaystyle=\bigwedge\nolimits_{\equiv}\{\pi_{\downarrow}^{\equiv}(a):a{\,\,<\!\!\!\!\cdot\,\,\,}x\text{ in $L$}\}=\bigwedge\nolimits_{L}\{\pi_{\downarrow}^{\equiv}(a):a{\,\,<\!\!\!\!\cdot\,\,\,}x\text{ in $L$}\}
=π(L{a:a<x in W})=πpopL(x).\displaystyle=\pi_{\downarrow}^{\equiv}\left(\bigwedge\nolimits_{L}\{a:a{\,\,<\!\!\!\!\cdot\,\,\,}x\text{ in $W$}\}\right)=\pi_{\downarrow}^{\equiv}\circ\mathrm{pop}^{\downarrow}_{L}(x).

We obtain equality from the first line to the second line because the meet taken in the cc-Cambrian lattice stays in the cc-Cambrian lattice (so the outermost π\pi_{\downarrow}^{\equiv} acts as the identity). The equality from the second to the third line results from the fact that π\pi_{\downarrow}^{\equiv} and the meet operation commute. Thus, we obtain the simpler expression for popCambc\mathrm{pop}_{\mathrm{Camb}_{c}} from Equation 6.3.

We now specialize to the case where LL is the weak order on WW and prove our main result of this subsection. Let WW_{\equiv} denote a lattice quotient of WW coming from a congruence \equiv.

Theorem 9.6.

If WW_{\equiv} is a lattice quotient of Weak(W)\mathrm{Weak}(W), then

maxxW|𝒪W(x)|h.\max_{x\in W_{\equiv}}\left\lvert\mathcal{O}_{W_{\equiv}}(x)\right\rvert\leq h.
Proof.

We may assume that WW_{\equiv} is the subposet π(W)\pi_{\downarrow}^{\equiv}(W) of Weak(W)\mathrm{Weak}(W), where \equiv is a lattice congruence on Weak(W)\mathrm{Weak}(W). We write \wedge_{\equiv} and \lessdot_{\equiv} to denote meets and cover relations, respectively, in WW_{\equiv} (while \wedge and \lessdot will denote meets and cover relations in Weak(W)\mathrm{Weak}(W)). Define a map g:WWg\colon W\to W by

g(x)={popW(x)if xW;popWeak(W)(x) otherwise.g(x)=\begin{cases}\mathrm{pop}^{\downarrow}_{W_{\equiv}}(x)&\text{if }x\in W_{\equiv};\\ \mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(x)&\text{ otherwise}.\end{cases}

Note that

maxxW|𝒪W(x)|maxw6W|Orbg(w)|\max_{x\in W_{\equiv}}\left\lvert\mathcal{O}_{W_{\equiv}}(x)\right\rvert\leq\max_{w\in 6W}\left\lvert\mathrm{Orb}_{g}(w)\right\rvert

(recall that Orbg(w)\mathrm{Orb}_{g}(w) denotes the forward orbit of ww under gg).

Following [Def22b], we say a map f:WWf\colon W\to W is compulsive if f(x)popWeak(W)(x)f(x)\leq\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(x) for all xWx\in W. [Def22b, Theorem 1.5] states that if ff is compulsive, then

maxwW|Orbf(w)|h.\max_{w\in W}\left\lvert\mathrm{Orb}_{f}(w)\right\rvert\leq h.

Hence, it suffices to show that gg is compulsive.

Consider xW0^x\in W_{\equiv}\setminus\hat{0}. By Lemma 9.3, we have

g(x)=popW(x)={yWyx}={π(a)ax}.g(x)=\mathrm{pop}^{\downarrow}_{W_{\equiv}}(x)=\bigwedge\nolimits_{\equiv}\{y\in W_{\equiv}\mid y\lessdot_{\equiv}x\}=\bigwedge\nolimits_{\equiv}\{\pi_{\downarrow}^{\equiv}(a)\mid a\lessdot x\}.

But then since π(a)a\pi_{\downarrow}^{\equiv}(a)\leq a for all aa, we have

{π(a)ax}{π(a)ax}{aax}=popWeak(W)(x).\bigwedge\nolimits_{\equiv}\{\pi_{\downarrow}^{\equiv}(a)\mid a\lessdot x\}\leq\bigwedge\{\pi_{\downarrow}^{\equiv}(a)\mid a\lessdot x\}\leq\bigwedge\{a\mid a\lessdot x\}=\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(x).

On the other hand, g(z)=popWeak(W)(z)g(z)=\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(z) whenever zWWz\in W\setminus W_{\equiv} or z=0^z=\hat{0}. This shows that gg is compulsive. ∎

Remark 9.7.

One cannot weaken the hypothesis in Theorem 9.6 by assuming that LL is just a meet-semilattice quotient or a join-semilattice quotient. For example, suppose W=A2W=A_{2}. Let \equiv be the meet-semilattice congruence on Weak(W)\mathrm{Weak}(W) with one nontrivial equivalence class consisting of 0^\hat{0}, s1s_{1}, and s1s2s_{1}s_{2}. The lattice L=W/L=W/\equiv is a 44-element chain, so maxxL|𝒪L(x)|=4>3=h\max_{x\in L}|\mathcal{O}_{L}(x)|=4>3=h.

9.2. Heaps and combinatorial AR quivers

Our goal is now to show that the upper bound in Theorem 9.6 is tight when LL is a Cambrian lattice Cambc\mathrm{Camb}_{c}. We will do this by constructing an explicit element 𝐳cCambc\mathbf{z}_{c}\in\mathrm{Camb}_{c} whose forward orbit under popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} has size exactly hh. Our construction makes use of the combinatorial AR quivers of [STW, Section 9]. As we explain in Section 9.4, these are intimately related with the representation theory of the tensor algebra KQcKQ_{c}.

It will be convenient to think of the letters in a word over the alphabet SS as being distinct entities even if they represent the same simple reflection. For example, the two occurrences of 𝗌2\mathsf{s}_{2} in the word 𝗌2𝗌1𝗌2𝗌3\mathsf{s}_{2}\mathsf{s}_{1}\mathsf{s}_{2}\mathsf{s}_{3} are considered to be different. If 𝗌\mathsf{s} and 𝗌\mathsf{s}^{\prime} are letters representing simple reflections ss and ss^{\prime}, then we let m(𝗌,𝗌)=m(s,s)m(\mathsf{s},\mathsf{s}^{\prime})=m(s,s^{\prime}) (recall that m(s,s)m(s,s^{\prime}) is the order of ssss^{\prime}). We can apply a commutation move to a word by swapping two consecutive letters 𝗌\mathsf{s} and 𝗌\mathsf{s}^{\prime} if m(𝗌,𝗌)=2m(\mathsf{s},\mathsf{s}^{\prime})=2 (this move is not allowed if m(𝗌,𝗌)=1m(\mathsf{s},\mathsf{s}^{\prime})=1). The commutation class of a word 𝖰\mathsf{Q} over SS is the set of words that can be obtained from 𝖰\mathsf{Q} via a sequence of commutation moves. We write 𝖰𝖰\mathsf{Q}\equiv\mathsf{Q}^{\prime} if 𝖰\mathsf{Q} and 𝖰\mathsf{Q}^{\prime} are in the same commutation class.

Let 𝖰\mathsf{Q} be a word over SS. The heap of 𝖰\mathsf{Q} is the poset Heap(𝖰)\mathrm{Heap}(\mathsf{Q}) whose elements are the letters in 𝖰\mathsf{Q} (which we view as distinct from one another) and whose order relation is defined so that if 𝗌\mathsf{s} and 𝗌\mathsf{s}^{\prime} are two letters (which could represent the same simple reflection), then 𝗌<𝗌\mathsf{s}<\mathsf{s}^{\prime} if and only if 𝗌\mathsf{s} appears to the left of 𝗌\mathsf{s}^{\prime} in every word in the commutation class of 𝖰\mathsf{Q}. A word 𝖰\mathsf{Q}^{\prime} is in the same commutation class as 𝖰\mathsf{Q} if and only if Heap(𝖰)=Heap(𝖰)\mathrm{Heap}(\mathsf{Q})=\mathrm{Heap}(\mathsf{Q}^{\prime}). More generally, Heap(𝖰)\mathrm{Heap}(\mathsf{Q}^{\prime}) is an order ideal of Heap(𝖰)\mathrm{Heap}(\mathsf{Q}) if and only if there exists a word 𝖰′′\mathsf{Q}^{\prime\prime} such that 𝖰𝖰′′𝖰\mathsf{Q}^{\prime}\mathsf{Q}^{\prime\prime}\equiv\mathsf{Q}.

Suppose H\mathrm{H} is the heap of a word 𝗐\mathsf{w}. We write ζ(H)\zeta(\mathrm{H}) for the element of WW represented by 𝗐\mathsf{w} and write Z(H)Z(\mathrm{H}) for the set of simple reflections that appear in 𝗐\mathsf{w}. If 𝗐𝗐\mathsf{w}\equiv\mathsf{w}^{\prime}, then 𝗐\mathsf{w} and 𝗐\mathsf{w^{\prime}} represent the same element of WW; therefore, ζ(H)\zeta(\mathrm{H}) and Z(H)Z(\mathrm{H}) depend only on H\mathrm{H} (and not the particular word 𝗐\mathsf{w}). Let 𝒥(H)\mathcal{J}(\mathrm{H}) denote the set of order ideals of H\mathrm{H}.

Let 𝖼\mathsf{c} be a reduced word for a Coxeter element cc, and let 𝖼k\mathsf{c}^{k} denote the word obtained by concatenating 𝖼\mathsf{c} with itself kk times. Let S={s1,,sn}S=\{s_{1},\ldots,s_{n}\}, and let 𝗌i(j)\mathsf{s}_{i}^{(j)} denote the jj-th letter in 𝖼k\mathsf{c}^{k} that represents sis_{i}. For example, if 𝖼=𝗌1𝗌3𝗌2\mathsf{c}=\mathsf{s}_{1}\mathsf{s}_{3}\mathsf{s}_{2}, then

𝖼3=𝗌1(1)𝗌3(1)𝗌2(1)𝗌1(2)𝗌3(2)𝗌2(2)𝗌1(3)𝗌3(3)𝗌2(3).\mathsf{c}^{3}=\mathsf{s}_{1}^{(1)}\mathsf{s}_{3}^{(1)}\mathsf{s}_{2}^{(1)}\mathsf{s}_{1}^{(2)}\mathsf{s}_{3}^{(2)}\mathsf{s}_{2}^{(2)}\mathsf{s}_{1}^{(3)}\mathsf{s}_{3}^{(3)}\mathsf{s}_{2}^{(3)}.
Lemma 9.8.

For each k1k\geq 1, the poset Heap(𝖼k)\mathrm{Heap}(\mathsf{c}^{k}) is ranked.

Proof.

Because ΓW\Gamma_{W} is a tree, it is straightforward to see that Heap(𝖼)\mathrm{Heap}(\mathsf{c}) is ranked. Let rank:S\mathrm{rank}\colon S\to\mathbb{Z} be the unique rank function of Heap(𝖼)\mathrm{Heap}(\mathsf{c}) whose minimum value is 11. Let us extend this rank function to a map rank:Heap(𝖼k)\mathrm{rank}\colon\mathrm{Heap}(\mathsf{c}^{k})\to\mathbb{Z} by letting rank(𝗌i(j))=rank(si)+2j2\mathrm{rank}(\mathsf{s}_{i}^{(j)})=\mathrm{rank}(s_{i})+2j-2. We claim that this map is a rank function on Heap(𝖼k)\mathrm{Heap}(\mathsf{c}^{k}). To see this, suppose 𝗌i(j)𝗌i(j)\mathsf{s}_{i}^{(j)}\lessdot\mathsf{s}_{i^{\prime}}^{(j^{\prime})} is a cover relation in Heap(𝖼k)\mathrm{Heap}(\mathsf{c}^{k}). Then {si,si}\{s_{i},s_{i^{\prime}}\} is an edge in ΓW\Gamma_{W}. If 𝗌i\mathsf{s}_{i} appears before 𝗌i\mathsf{s}_{i^{\prime}} in 𝖼\mathsf{c}, then we have rank(si)=rank(si)+1\mathrm{rank}(s_{i^{\prime}})=\mathrm{rank}(s_{i})+1 and j=jj^{\prime}=j. If 𝗌i\mathsf{s}_{i} appears after 𝗌i\mathsf{s}_{i^{\prime}} in 𝖼\mathsf{c}, then we have rank(si)=rank(si)1\mathrm{rank}(s_{i^{\prime}})=\mathrm{rank}(s_{i})-1 and j=j+1j^{\prime}=j+1. In either case, we compute that rank(𝗌i(j))=rank(𝗌i(j))+1\mathrm{rank}(\mathsf{s}_{i^{\prime}}^{(j^{\prime})})=\mathrm{rank}(\mathsf{s}_{i}^{(j)})+1. ∎

Let rank:Heap(𝖼h)\mathrm{rank}\colon\mathrm{Heap}(\mathsf{c}^{h})\to\mathbb{Z} be the rank function on Heap(𝖼h)\mathrm{Heap}(\mathsf{c}^{h}) that was constructed in the proof of Lemma 9.8. When we consider a heap H\mathrm{H} that can be naturally identified with a convex subset of Heap(𝖼h)\mathrm{Heap}(\mathsf{c}^{h}), the rank of an element 𝗌\mathsf{s} of H\mathrm{H} will simply be rank(𝗌)\mathrm{rank}(\mathsf{s}), the rank of 𝗌\mathsf{s} in Heap(𝖼h)\mathrm{Heap}(\mathsf{c}^{h}). We will also find it convenient to write

Hk={𝗌Hrank(𝗌)=k}andHk={𝗌Hrank(𝗌)k}.\mathrm{H}^{k}=\{\mathsf{s}\in\mathrm{H}\mid\mathrm{rank}(\mathsf{s})=k\}\quad\text{and}\quad\mathrm{H}^{\leq k}=\{\mathsf{s}\in\mathrm{H}\mid\mathrm{rank}(\mathsf{s})\leq k\}.

There is an involution ψ:SS\psi\colon S\to S given by ψ(s)=wsw\psi(s)=w_{\circ}sw_{\circ}; this map is an automorphism of ΓW\Gamma_{W}. We can extend ψ\psi to the set of words over SS by letting ψ(𝗌1𝗌M)=ψ(𝗌1)ψ(𝗌M)\psi(\mathsf{s}_{1}\cdots\mathsf{s}_{M})=\psi(\mathsf{s}_{1})\cdots\psi(\mathsf{s}_{M}).

Let Hc=Heap(𝗌𝗈𝗋𝗍𝖼(w))\mathrm{H}_{c}=\mathrm{Heap}(\mathsf{sort}_{\mathsf{c}}(w_{\circ})) and H~c=Heap(ψ(𝗌𝗈𝗋𝗍𝖼(w)))\widetilde{\mathrm{H}}_{c}=\mathrm{Heap}(\psi(\mathsf{sort}_{\mathsf{c}}(w_{\circ}))); note that Hc\mathrm{H}_{c} and H~c\widetilde{\mathrm{H}}_{c} depend only on the Coxeter element cc (and not the reduced word 𝖼\mathsf{c}). It follows from [STW, Lemma 2.6.5] that

(9.1) 𝗌𝗈𝗋𝗍𝖼(w)ψ(𝗌𝗈𝗋𝗍𝖼(w))𝖼h.\mathsf{sort}_{\mathsf{c}}(w_{\circ})\psi(\mathsf{sort}_{\mathsf{c}}(w_{\circ}))\equiv\mathsf{c}^{h}.

Therefore, we can think of Hc\mathrm{H}_{c} as an order ideal of Heap(𝖼h)\mathrm{Heap}(\mathsf{c}^{h}) whose complement (in Heap(𝖼h)\mathrm{Heap}(\mathsf{c}^{h})) is H~c\widetilde{\mathrm{H}}_{c}. Define

𝐳c=ζ(Hch1).\mathbf{z}_{c}=\zeta(\mathrm{H}_{c}^{\leq h-1}).
Remark 9.9.

We have defined 𝐳c\mathbf{z}_{c} using Coxeter–Catalan combinatorics because we deemed this approach to be the most amenable for proving that |𝒪Cambc(𝐳c)|=h|\mathcal{O}_{\mathrm{Camb}_{c}}(\mathbf{z}_{c})|=h. However, there is also a more direct lattice-theoretic way to define 𝐳c\mathbf{z}_{c}. The spine of Cambc\mathrm{Camb}_{c}, denoted spine(Cambc)\mathrm{spine}(\mathrm{Camb}_{c}), is the union of the maximum-length chains of Cambc\mathrm{Camb}_{c}. Thomas and Williams [TW19b] proved that spine(Cambc)\mathrm{spine}(\mathrm{Camb}_{c}) is a distributive sublattice of Cambc\mathrm{Camb}_{c}. It is also known [TW19b] that

spine(Cambc)={ζ(I)I𝒥(Hc)}.\mathrm{spine}(\mathrm{Camb}_{c})=\{\zeta(I)\mid I\in\mathcal{J}(\mathrm{H}_{c})\}.

One can show that 𝐳c=(popspine(Cambc))h1(e)\mathbf{z}_{c}=(\mathrm{pop}^{\uparrow}_{\mathrm{spine}(\mathrm{Camb}_{c})})^{h-1}(e) (where e=0^e=\hat{0} is the identity element).

Example 9.10.

Suppose W=A8W=A_{8}. The edges in ΓW\Gamma_{W} are {si,si+1}\{s_{i},s_{i+1}\} for 1i71\leq i\leq 7. The Coxeter number of WW is h=9h=9. The map ψ\psi is given by ψ(sk)=s9k\psi(s_{k})=s_{9-k}. Let c=s1s3s2s4s6s5s7s8c=s_{1}s_{3}s_{2}s_{4}s_{6}s_{5}s_{7}s_{8} and 𝖼=𝗌1𝗌3𝗌2𝗌4𝗌6𝗌5𝗌7𝗌8\mathsf{c}=\mathsf{s}_{1}\mathsf{s}_{3}\mathsf{s}_{2}\mathsf{s}_{4}\mathsf{s}_{6}\mathsf{s}_{5}\mathsf{s}_{7}\mathsf{s}_{8}. The acyclic orientation of ΓW\Gamma_{W} corresponding to cc is

Qc=([Uncaptioned image]).Q_{c}=\left(\includegraphics[height=7.11317pt]{CambrianPopPIC2}\right).

Figure 6 shows the Hasse diagram of Heap(𝖼9)\mathrm{Heap}(\mathsf{c}^{9}); it is drawn sideways so that each cover relation 𝗌i(j)𝗌i(j)\mathsf{s}_{i}^{(j)}\lessdot\mathsf{s}_{i^{\prime}}^{(j^{\prime})} is drawn with 𝗌i(j)\mathsf{s}_{i}^{(j)} to the left of 𝗌i(j)\mathsf{s}_{i^{\prime}}^{(j^{\prime})}. As explained in Section 9.4, this Hasse diagram is referred to as the combinatorial AR quiver of 𝖼9\mathsf{c}^{9} in [STW, Section 9.2]. Note that Heap(𝖼9)\mathrm{Heap}(\mathsf{c}^{9}) is drawn by lining up 99 copies of Heap(𝖼)\mathrm{Heap}(\mathsf{c}) (depicted in 99 different colors) in a row and adding (black) edges as appropriate. The columns are labeled from 11 to 2020 from left to right; the column labeled ii consists of the elements of rank ii. The region shaded in yellow is Hc\mathrm{H}_{c}, while the unshaded region is H~c\widetilde{\mathrm{H}}_{c}. The region outlined in blue is Hch1=Hc8\mathrm{H}_{c}^{\leq h-1}=\mathrm{H}_{c}^{\leq 8}. Thus,

𝐳c=s1s3s2s4s6s5s7s8s1s3s2s4s6s5s7s8s1s3s2s4s6s5s7s8s1s3s2s4s6.\mathbf{z}_{c}={\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}s_{1}s_{3}s_{2}s_{4}s_{6}s_{5}s_{7}s_{8}}{\color[rgb]{0,0.86328125,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.86328125,0}s_{1}s_{3}s_{2}s_{4}s_{6}s_{5}s_{7}s_{8}}{\color[rgb]{0,0.71484375,1}\definecolor[named]{pgfstrokecolor}{rgb}{0,0.71484375,1}s_{1}s_{3}s_{2}s_{4}s_{6}s_{5}s_{7}s_{8}}{\color[rgb]{0.70703125,0,1}\definecolor[named]{pgfstrokecolor}{rgb}{0.70703125,0,1}s_{1}s_{3}s_{2}s_{4}s_{6}}.
Refer to caption
Figure 6. The Hasse diagram of Heap(𝖼h)\mathrm{Heap}(\mathsf{c}^{h}) from Example 9.10, in which W=A8W=A_{8}. The heap Hc\mathrm{H}_{c} is shaded in yellow, while Hch1\mathrm{H}_{c}^{\leq h-1} is outlined in blue.

9.3. Realization of the upper bound

We need a little more preparation before we can prove that |𝒪Cambc(𝐳c)|=h|\mathcal{O}_{\mathrm{Camb}_{c}}(\mathbf{z}_{c})|=h.

For kk\in\mathbb{Z}, let

ε(k)={1if k is odd;2if k is even.\varepsilon(k)=\begin{cases}1&\mbox{if }k\mbox{ is odd};\\ 2&\mbox{if }k\mbox{ is even}.\end{cases}

Note that for each 𝗌i(j)Heap(𝖼h)\mathsf{s}_{i}^{(j)}\in\mathrm{Heap}(\mathsf{c}^{h}), the parity of rank(𝗌i(j))\mathrm{rank}(\mathsf{s}_{i}^{(j)}) only depends on sis_{i} (and not jj). Let

X1={siSrank(𝗌i(j)) is odd for all j}andX2={siSrank(𝗌i(j)) is even for all j}.X_{1}=\{s_{i}\in S\mid\mathrm{rank}(\mathsf{s}_{i}^{(j)})\text{ is odd for all }j\}\quad\text{and}\quad X_{2}=\{s_{i}\in S\mid\mathrm{rank}(\mathsf{s}_{i}^{(j)})\text{ is even for all }j\}.

Then S=X1X2S=X_{1}\sqcup X_{2} is a bipartition of the tree ΓW\Gamma_{W}.

Lemma 9.11.

We have Z(Hch1)=Xε(h1)Z(\mathrm{H}_{c}^{h-1})=X_{\varepsilon(h-1)}.

Proof.

We already know that Z(Hch1)Xε(h1)Z(\mathrm{H}_{c}^{h-1})\subseteq X_{\varepsilon(h-1)}, so we just need to prove the reverse containment. We will show that each element of H~c\widetilde{\mathrm{H}}_{c} has rank at least h+1h+1; this will imply that the elements of Heap(𝖼h)\mathrm{Heap}(\mathsf{c}^{h}) with rank h1h-1 must belong to Hc\mathrm{H}_{c}.

Fix a simple reflection siSs_{i}\in S, and let si=ψ(si)s_{i^{\prime}}=\psi(s_{i}). Let jj be the smallest integer such that 𝗌i(j)H~c\mathsf{s}_{i}^{(j)}\in\widetilde{\mathrm{H}}_{c}. Let jj^{\prime} be the smallest integer such that 𝗌i(j)H~c\mathsf{s}_{i^{\prime}}^{(j^{\prime})}\in\widetilde{\mathrm{H}}_{c}. Then j1j^{\prime}-1 is the number of letters in 𝗌𝗈𝗋𝗍𝖼(w)\mathsf{sort}_{\mathsf{c}}(w_{\circ}) that represent sis_{i^{\prime}}, so it is also the number of letters in ψ(𝗌𝗈𝗋𝗍𝖼(w))\psi(\mathsf{sort}_{\mathsf{c}}(w_{\circ})) that represent sis_{i}. But the number of letters in ψ(𝗌𝗈𝗋𝗍𝖼(w))\psi(\mathsf{sort}_{\mathsf{c}}(w_{\circ})) that represent sis_{i} is hj+1h-j+1, so j+j2=hj+j^{\prime}-2=h. Hence,

rank(𝗌i(j))=rank(si)+2j2=rank(si)+2h2j+2.\mathrm{rank}(\mathsf{s}_{i}^{(j)})=\mathrm{rank}(s_{i})+2j-2=\mathrm{rank}(s_{i})+2h-2j^{\prime}+2.

Because ψ:SS\psi\colon S\to S is an automorphism of ΓW\Gamma_{W}, the map ψ:HcH~c\psi\colon\mathrm{H}_{c}\to\widetilde{\mathrm{H}}_{c} is a poset isomorphism. It follows that rank(𝗌i(j))rank(𝗌i(j))=rank(𝗌i(1))rank(𝗌i(1))=rank(si)rank(si)\mathrm{rank}(\mathsf{s}_{i^{\prime}}^{(j^{\prime})})-\mathrm{rank}(\mathsf{s}_{i}^{(j)})=\mathrm{rank}(\mathsf{s}_{i}^{(1)})-\mathrm{rank}(\mathsf{s}_{i^{\prime}}^{(1)})=\mathrm{rank}(s_{i})-\mathrm{rank}(s_{i^{\prime}}). By definition, we have rank(𝗌i(j))=rank(si)+2j2\mathrm{rank}(\mathsf{s}_{i}^{(j)})=\mathrm{rank}(s_{i})+2j-2 and rank(𝗌i(j))=rank(si)+2j2\mathrm{rank}(\mathsf{s}_{i^{\prime}}^{(j^{\prime})})=\mathrm{rank}(s_{i^{\prime}})+2j^{\prime}-2, so

rank(si)+(2j2)(rank(si)+2j2)=rank(si)rank(si).\mathrm{rank}(s_{i^{\prime}})+(2j^{\prime}-2)-(\mathrm{rank}(s_{i})+2j-2)=\mathrm{rank}(s_{i})-\mathrm{rank}(s_{i^{\prime}}).

Rearranging this equation yields rank(si)=rank(si)+jj\mathrm{rank}(s_{i})=\mathrm{rank}(s_{i^{\prime}})+j^{\prime}-j. Consequently,

rank(𝗌i(j))=rank(si)+2j2=rank(si)+j+j2=rank(si)+h.\mathrm{rank}(\mathsf{s}_{i}^{(j)})=\mathrm{rank}(s_{i})+2j-2=\mathrm{rank}(s_{i^{\prime}})+j+j^{\prime}-2=\mathrm{rank}(s_{i^{\prime}})+h.

In particular, rank(𝗌i(j))h+1\mathrm{rank}(\mathsf{s}_{i}^{(j)})\geq h+1. This shows that every element of H~c\widetilde{\mathrm{H}}_{c} that represents sis_{i} has rank at least h+1h+1. As sis_{i} was arbitrary, it follows that all of the elements of H~c\widetilde{\mathrm{H}}_{c} have rank at least h+1h+1.

Suppose smXε(h1)s_{m}\in X_{\varepsilon(h-1)}, and let kk be the unique integer such that rank(𝗌m(k))=h1\mathrm{rank}(\mathsf{s}_{m}^{(k)})=h-1. It follows from the preceding paragraph that 𝗌m(k)H~c\mathsf{s}_{m}^{(k)}\not\in\widetilde{\mathrm{H}}_{c}, so 𝗌m(k)Hc\mathsf{s}_{m}^{(k)}\in\mathrm{H}_{c}. This shows that smZ(Hch1)s_{m}\in Z(\mathrm{H}_{c}^{h-1}). Hence, Xε(h1)Z(Hch1)X_{\varepsilon(h-1)}\subseteq Z(\mathrm{H}_{c}^{h-1}). ∎

Example 9.12.

Let us use Figure 6 to illustrate the proof of Lemma 9.11. Recall that, in this example, the map ψ\psi is given by ψ(sk)=s9k\psi(s_{k})=s_{9-k}. Let i=3i=3 so that i=6i^{\prime}=6. Then j=6j=6, j=5j^{\prime}=5, rank(si)=1\mathrm{rank}(s_{i})=1, and rank(si)=2\mathrm{rank}(s_{i^{\prime}})=2. Note that j+j2=9=hj+j^{\prime}-2=9=h and rank(𝗌i(j))=11=h+rank(si)\mathrm{rank}(\mathsf{s}_{i}^{(j)})=11=h+\mathrm{rank}(s_{i^{\prime}}). We have Z(Hch1)={s2,s4,s6,s8}=X2=Xε(h1)Z(\mathrm{H}_{c}^{h-1})=\{s_{2},s_{4},s_{6},s_{8}\}=X_{2}=X_{\varepsilon(h-1)}, as Lemma 9.11 claims.

Let c1=w(X1)=sX1sc_{1}=w_{\circ}(X_{1})=\prod_{s\in X_{1}}s and c2=w(X2)=sX2sc_{2}=w_{\circ}(X_{2})=\prod_{s\in X_{2}}s, and fix reduced words 𝖼1\mathsf{c}_{1} and 𝖼2\mathsf{c}_{2} for c1c_{1} and c2c_{2}, respectively. The element c×=c1c2c^{\times}=c_{1}c_{2} is called a bipartite Coxeter element of WW. (In type A, this definition of c×c^{\times} coincides with that of either c(n)×c^{\times}_{(n)} or c(n)××c^{\times\times}_{(n)} from Section 8.1.) Let 𝖼×=𝖼1𝖼2\mathsf{c}^{\times}=\mathsf{c}_{1}\mathsf{c}_{2}. Let 𝗎k\mathsf{u}_{k} be the word 𝖼1𝖼2𝖼1𝖼ε(k)\mathsf{c}_{1}\mathsf{c}_{2}\mathsf{c}_{1}\cdots\mathsf{c}_{\varepsilon(k)} that begins with 𝖼1\mathsf{c}_{1} and alternates between 𝖼1\mathsf{c}_{1} and 𝖼2\mathsf{c}_{2}, using k/2\left\lceil k/2\right\rceil copies of 𝖼1\mathsf{c}_{1} and k/2\left\lfloor k/2\right\rfloor copies of 𝖼2\mathsf{c}_{2}. In other words,

𝗎k={(𝖼×)k/2if k is even;(𝖼×)(k1)/2𝖼1if k is odd.\mathsf{u}_{k}=\begin{cases}(\mathsf{c}^{\times})^{k/2}&\mbox{if }k\mbox{ is even};\\ (\mathsf{c}^{\times})^{(k-1)/2}\mathsf{c}_{1}&\mbox{if }k\mbox{ is odd}.\end{cases}

It is known [Bou02, §\SV.6 Exercise 2] that 𝗌𝗈𝗋𝗍𝖼×(w)=𝗎h\mathsf{sort}_{\mathsf{c}^{\times}}(w_{\circ})=\mathsf{u}_{h}. Hence, for 1jkh1\leq j\leq k\leq h, the word 𝗎k\mathsf{u}_{k} is a reduced word for an element ukWu_{k}\in W, and Z(Heap(𝗎k)j)=Xε(j)Z(\mathrm{Heap}(\mathsf{u}_{k})^{j})=X_{\varepsilon(j)}.

Lemma 9.13.

For each k[h1]k\in[h-1], we have Des(uk)=Xε(k)\mathrm{Des}(u_{k})=X_{\varepsilon(k)}.

Proof.

The reduced word 𝗎k\mathsf{u}_{k} ends with 𝖼ε(k)\mathsf{c}_{\varepsilon(k)}, and Z(𝖼ε(k))=cε(k)Z(\mathsf{c}_{\varepsilon(k)})=c_{\varepsilon(k)}. Since all of the simple reflections in Xε(k)X_{\varepsilon(k)} commute with each other, we have Xε(k)Des(uk)X_{\varepsilon(k)}\subseteq\mathrm{Des}(u_{k}).

Now suppose by way of contradiction that sDes(uk)Xε(k+1)s\in\mathrm{Des}(u_{k})\cap X_{\varepsilon(k+1)}. Let v=uksv=u_{k}s, and note that (v)=(uk)1\ell(v)=\ell(u_{k})-1. Since 𝗎k+1=𝗎k𝖼ε(k+1)\mathsf{u}_{k+1}=\mathsf{u}_{k}\mathsf{c}_{\varepsilon(k+1)}, we have (uk+1)=(uk)+(cε(k+1))\ell(u_{k+1})=\ell(u_{k})+\ell(c_{\varepsilon(k+1)}). Let

v=sXε(k+1){s}s=scε(k+1).v^{\prime}=\prod_{s^{\prime}\in X_{\varepsilon(k+1)}\setminus\{s\}}s^{\prime}=sc_{\varepsilon(k+1)}.

Then (v)=(cε(k+1))1\ell(v^{\prime})=\ell(c_{\varepsilon(k+1)})-1. We have

uk+1=ukcε(k+1)=vssv=vv,u_{k+1}=u_{k}c_{\varepsilon(k+1)}=vssv^{\prime}=vv^{\prime},

which is a contradiction because

(v)+(v)=(uk)1+(cε(k+1))1=(uk+1)2<(uk+1).\ell(v)+\ell(v^{\prime})=\ell(u_{k})-1+\ell(c_{\varepsilon(k+1)})-1=\ell(u_{k+1})-2<\ell(u_{k+1}).\qed
Example 9.14.

Let W=D5W=D_{5}. Our convention is that S={s0,s1,s2,s3,s4}S=\{s_{0},s_{1},s_{2},s_{3},s_{4}\} and that the edges of ΓW\Gamma_{W} (each of which is unlabeled) are {s0,s2}\{s_{0},s_{2}\}, {s1,s2}\{s_{1},s_{2}\}, {s2,s3}\{s_{2},s_{3}\}, {s3,s4}\{s_{3},s_{4}\}. The Coxeter number of WW is h=8h=8. The map ψ\psi is given by ψ(s0)=s1\psi(s_{0})=s_{1}, ψ(s1)=s0\psi(s_{1})=s_{0}, and ψ(sk)=sk\psi(s_{k})=s_{k} for all k{2,3,4}k\in\{2,3,4\}. Let c=s0s2s1s3s4c=s_{0}s_{2}s_{1}s_{3}s_{4} and 𝖼=𝗌0𝗌2𝗌1𝗌3𝗌4\mathsf{c}=\mathsf{s}_{0}\mathsf{s}_{2}\mathsf{s}_{1}\mathsf{s}_{3}\mathsf{s}_{4}. The acyclic orientation of ΓW\Gamma_{W} corresponding to cc is

[Uncaptioned image].\includegraphics[height=29.42004pt]{CambrianPopPIC3}.

We have X1={s0,s1,s3}X_{1}=\{s_{0},s_{1},s_{3}\} and X2={s2,s4}X_{2}=\{s_{2},s_{4}\}. The region outlined in blue in Figure 7 is Heap(𝖼8)\mathrm{Heap}(\mathsf{c}^{8}) (drawn sideways, similarly to Figure 6). The region shaded in yellow is Hc\mathrm{H}_{c}. We have added the extra elements 𝗌1(0),𝗌3(0),𝗌4(0)\mathsf{s}_{1}^{(0)},\mathsf{s}_{3}^{(0)},\mathsf{s}_{4}^{(0)} (and appropriate edges) to the left of Heap(𝖼8)\mathrm{Heap}(\mathsf{c}^{8}) in order to realize Heap(𝗎5)\mathrm{Heap}(\mathsf{u}_{5}) as the region enclosed by the black rectangle. The yellow shaded region that lies inside of the black rectangle is Hc5\mathrm{H}_{c}^{\leq 5}; this illustrates how we can naturally identify Hc5\mathrm{H}_{c}^{\leq 5} with a subposet of Heap(𝗎5)\mathrm{Heap}(\mathsf{u}_{5}).

Refer to caption
Figure 7. An illustration of Example 9.14, in which W=D5W=D_{5}. The region outlined in blue is Heap(𝖼8)\mathrm{Heap}(\mathsf{c}^{8}). The heap Hc\mathrm{H}_{c} is shaded in yellow. The region enclosed by the black rectangle is Heap(𝗎5)\mathrm{Heap}(\mathsf{u}_{5}).
Lemma 9.15.

For each k[h1]k\in[h-1], we have Z(Hck)=Des(ζ(Hck))Z(\mathrm{H}_{c}^{k})=\mathrm{Des}(\zeta(\mathrm{H}_{c}^{\leq k})).

Proof.

We have Z(Hck)Xε(k)Z(\mathrm{H}_{c}^{k})\subseteq X_{\varepsilon(k)}, so all of the elements of Z(Hck)Z(\mathrm{H}_{c}^{k}) commute with each other. Therefore, ζ(Hck)=sZ(Hck)s\zeta(\mathrm{H}_{c}^{k})=\prod_{s\in Z(\mathrm{H}_{c}^{k})}s. Because Hck1\mathrm{H}_{c}^{\leq k-1} is an order ideal of Hck\mathrm{H}_{c}^{\leq k} whose complement (in Hck\mathrm{H}_{c}^{\leq k}) is Hck\mathrm{H}_{c}^{k}, we have ζ(Hck)=ζ(Hck1)ζ(Hck)\zeta(\mathrm{H}_{c}^{\leq k})=\zeta(\mathrm{H}_{c}^{\leq k-1})\zeta(\mathrm{H}_{c}^{k}), so Z(Hck)Des(ζ(Hck))Z(\mathrm{H}_{c}^{k})\subseteq\mathrm{Des}(\zeta(\mathrm{H}_{c}^{\leq k})). Now choose some rDes(ζ(Hck))r\in\mathrm{Des}(\zeta(\mathrm{H}_{c}^{\leq k})); we will show that rZ(Hck)r\in Z(\mathrm{H}_{c}^{k}).

Because Z(Hcm)Xε(m)Z(\mathrm{H}_{c}^{m})\subseteq X_{\varepsilon(m)} for every m[k]m\in[k], we can identify Hck\mathrm{H}_{c}^{\leq k} with a subposet of Heap(𝗎k)\mathrm{Heap}(\mathsf{u}_{k}) (as illustrated in Figure 7). We will also continue to use Equation 9.1 to view Hc\mathrm{H}_{c} as an order ideal of Heap(𝖼h)\mathrm{Heap}(\mathsf{c}^{h}) whose complement is H~c\widetilde{\mathrm{H}}_{c}. We claim that Hck\mathrm{H}_{c}^{\leq k} is an upper set of Heap(𝗎k)\mathrm{Heap}(\mathsf{u}_{k}). To see this, suppose 𝗌i(j)Hck\mathsf{s}_{i}^{(j)}\in\mathrm{H}_{c}^{\leq k} and 𝗌i(j)Heap(𝗎k)\mathsf{s}_{i^{\prime}}^{(j^{\prime})}\in\mathrm{Heap}(\mathsf{u}_{k}) are such that 𝗌i(j)<𝗌i(j)\mathsf{s}_{i}^{(j)}<\mathsf{s}_{i^{\prime}}^{(j^{\prime})} in Heap(𝗎k)\mathrm{Heap}(\mathsf{u}_{k}). Let si′′s_{i^{\prime\prime}} be a simple reflection in Xε(h1)X_{\varepsilon(h-1)} such that m(si,si′′)2m(s_{i^{\prime}},s_{i^{\prime\prime}})\neq 2. There is a unique integer j′′[h]j^{\prime\prime}\in[h] such that rank(𝗌i′′(j′′))=h1\mathrm{rank}(\mathsf{s}_{i^{\prime\prime}}^{(j^{\prime\prime})})=h-1. Then 𝗌i′′(j′′)\mathsf{s}_{i^{\prime\prime}}^{(j^{\prime\prime})} is maximal in Heap(𝗎h1)\mathrm{Heap}(\mathsf{u}_{h-1}). Since 𝗌i(j)\mathsf{s}_{i^{\prime}}^{(j^{\prime})} and 𝗌i′′(j′′)\mathsf{s}_{i^{\prime\prime}}^{(j^{\prime\prime})} are comparable (because m(si,si′′)2m(s_{i^{\prime}},s_{i^{\prime\prime}})\neq 2), we must have 𝗌i(j)<𝗌i′′(j′′)\mathsf{s}_{i^{\prime}}^{(j^{\prime})}<\mathsf{s}_{i^{\prime\prime}}^{(j^{\prime\prime})}. It follows from Lemma 9.11 that 𝗌i′′(j′′)Hch1\mathsf{s}_{i^{\prime\prime}}^{(j^{\prime\prime})}\in\mathrm{H}_{c}^{h-1}; in particular, 𝗌i′′(j′′)Hc\mathsf{s}_{i^{\prime\prime}}^{(j^{\prime\prime})}\in\mathrm{H}_{c}. Because Hc\mathrm{H}_{c} is an order ideal of Heap(𝖼h)\mathrm{Heap}(\mathsf{c}^{h}), it contains 𝗌i(j)\mathsf{s}_{i^{\prime}}^{(j^{\prime})}. But rank(𝗌i(j))k\mathrm{rank}(\mathsf{s}_{i^{\prime}}^{(j^{\prime})})\leq k, so 𝗌i(j)Hck\mathsf{s}_{i^{\prime}}^{(j^{\prime})}\in\mathrm{H}_{c}^{\leq k}. This proves that Hck\mathrm{H}_{c}^{\leq k} is an upper set of Heap(𝗎k)\mathrm{Heap}(\mathsf{u}_{k}). Because 𝗎k\mathsf{u}_{k} is a reduced word for uku_{k}, it follows that ζ(Hck)1uk1\zeta(\mathrm{H}_{c}^{\leq k})^{-1}\leq u_{k}^{-1} (in the weak order). Consequently, rDes(ζ(Hck))Des(uk)=Xε(k)r\in\mathrm{Des}(\zeta(\mathrm{H}_{c}^{\leq k}))\subseteq\mathrm{Des}(u_{k})=X_{\varepsilon(k)}.

Let 𝗋\mathsf{r} be the unique element of rank kk in Heap(𝗎k)\mathrm{Heap}(\mathsf{u}_{k}) that represents rr. Because rDes(ζ(Hck))r\in\mathrm{Des}(\zeta(\mathrm{H}_{c}^{\leq k})), there must be an element 𝗋Hck\mathsf{r}^{\prime}\in\mathrm{H}_{c}^{\leq k} that represents rr. Because rank(𝗋)k=rank(𝗋)\mathrm{rank}(\mathsf{r}^{\prime})\leq k=\mathrm{rank}(\mathsf{r}), we have 𝗋𝗋\mathsf{r}^{\prime}\leq\mathsf{r} in Heap(𝗎k)\mathrm{Heap}(\mathsf{u}_{k}). But Hck\mathrm{H}_{c}^{\leq k} is an upper set of Heap(𝗎k)\mathrm{Heap}(\mathsf{u}_{k}), so 𝗋Hck\mathsf{r}\in\mathrm{H}_{c}^{\leq k}. As 𝗋\mathsf{r} has rank kk, it must be in Hck\mathrm{H}_{c}^{k}. This proves that rZ(Hck)r\in Z(\mathrm{H}_{c}^{k}), as desired. ∎

Lemma 9.16.

Let cc be a Coxeter element of WW. We have

|𝒪Cambc(𝐳c)|=h.\left\lvert\mathcal{O}_{\mathrm{Camb}_{c}}(\mathbf{z}_{c})\right\rvert=h.
Proof.

As before, let X1X_{1} (resp. X2X_{2}) be the set of simple reflections ss such that the letters representing ss in 𝖼h\mathsf{c}^{h} have odd (resp. even) rank. For 0jh10\leq j\leq h-1, let vj=ζ(Hcj)v_{j}=\zeta(\mathrm{H}_{c}^{\leq j}). In particular, v0=ev_{0}=e, and vh1=𝐳cv_{h-1}=\mathbf{z}_{c}. Suppose k[h1]k\in[h-1]. We know that Z(Hck)Xε(k)Z(\mathrm{H}_{c}^{k})\subseteq X_{\varepsilon(k)}, so all of the elements of Z(Hck)Z(\mathrm{H}_{c}^{k}) commute with each other. Along with Lemma 9.15, this implies that

w(vk)=w(Des(ζ(Hck)))=w(Z(Hck))=sZ(Hck)s=ζ(Hck).w_{\circ}(v_{k})=w_{\circ}(\mathrm{Des}(\zeta(\mathrm{H}_{c}^{\leq k})))=w_{\circ}(Z(\mathrm{H}_{c}^{k}))=\prod_{s\in Z(\mathrm{H}_{c}^{k})}s=\zeta(\mathrm{H}_{c}^{k}).

We have vk=vk1ζ(Hck)v_{k}=v_{k-1}\zeta(\mathrm{H}_{c}^{k}), so it follows from Equation 6.1 that

popWeak(W)(vk)=vk1ζ(Hck)w(Des(vk))=vk1ζ(Hck)2=vk1.\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(v_{k})=v_{k-1}\zeta(\mathrm{H}_{c}^{k})w_{\circ}(\mathrm{Des}(v_{k}))=v_{k-1}\zeta(\mathrm{H}_{c}^{k})^{2}=v_{k-1}.

Because Hck1\mathrm{H}_{c}^{\leq k-1} is an order ideal of Hc\mathrm{H}_{c}, we know by Remark 9.9 that vk1v_{k-1} is in the spine of Cambc\mathrm{Camb}_{c}; in particular, it is in Cambc\mathrm{Camb}_{c}. Invoking Equation 6.3, we find that

popCambc(vk)=πc(popWeak(W)(vk))=πc(vk1)=vk1.\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(v_{k})=\pi_{\downarrow}^{c}(\mathrm{pop}^{\downarrow}_{\mathrm{Weak}(W)}(v_{k}))=\pi_{\downarrow}^{c}(v_{k-1})=v_{k-1}.

As this is true for all k[h1]k\in[h-1], we deduce that

𝒪Cambc(𝐳c)=𝒪Cambc(vh1)={vh1,vh2,,v0}.\mathcal{O}_{\mathrm{Camb}_{c}}(\mathbf{z}_{c})=\mathcal{O}_{\mathrm{Camb}_{c}}(v_{h-1})=\{v_{h-1},v_{h-2},\ldots,v_{0}\}.\qed

Combining Theorem 9.6 with Lemma 9.16 yields our last main result.

Theorem 9.17.

Let WW be a finite irreducible Coxeter group, and let cWc\in W be a Coxeter element. The maximum size of a forward orbit under popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}} is exactly the Coxeter number hh of WW.

9.4. Heaps and representation theory

We conclude this section with a brief explanation of how the heap of a cc-sortable word 𝖰\mathsf{Q} relates to representation theory of KQcKQ_{c}.

Let Λ\Lambda be an arbitrary finite-dimensional algebra. A morphism f:MNf\colon M\rightarrow N in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda is said to be irreducible if it is not an isomorphism and whenever L𝗆𝗈𝖽ΛL\in\operatorname{\mathsf{mod}}\Lambda, g:MLg\colon M\rightarrow L, and h:LNh\colon L\rightarrow N are such that f=hgf=h\circ g, we have that either gg or hh is an isomorphism. For M,N𝗆𝗈𝖽ΛM,N\in\operatorname{\mathsf{mod}}\Lambda, we denote by irr(M,N)\mathrm{irr}(M,N) the linear subspace of HomΛ(M,N)\mathrm{Hom}_{\Lambda}(M,N) spanned by the irreducible morphisms. The Auslander–Reiten (AR) quiver of Λ\Lambda is then the directed graph defined as follows.

  • The vertices of the AR quiver are the indecomposable objects in 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda.

  • For any two vertices M,NM,N, the number of arrows MNM\rightarrow N in the AR quiver coincides with the dimension of irr(M,N)\mathrm{irr}(M,N).

Now let 𝖰\mathsf{Q} be a word over SS. In [STW, Section 9.2], the Hasse diagram of Heap(𝖰)\mathrm{Heap}(\mathsf{Q}) is referred to as the combinatorial AR quiver of 𝖰\mathsf{Q}. To explain this, consider the special case when 𝖰=𝗌𝗈𝗋𝗍𝖼(w)\mathsf{Q}=\mathsf{sort}_{\mathsf{c}}(w_{\circ}). By [STW, Theorem 9.3.1], there is a bijection Ξ:Heap(𝗌𝗈𝗋𝗍𝖼(w))𝖻𝗋𝗂𝖼𝗄KQc\Xi\colon\mathrm{Heap}(\mathsf{sort}_{\mathsf{c}}(w_{\circ}))\rightarrow\operatorname{\mathsf{brick}}KQ_{c}. Under this bijection, we have 𝗌i(j)<𝗌i(j)\mathsf{s}_{i}^{(j)}{\,\,<\!\!\!\!\cdot\,\,\,}\mathsf{s}_{i^{\prime}}^{(j^{\prime})} if and only if there is an irreducible morphism from Ξ(𝗌i(j))\Xi\left(\mathsf{s}_{i}^{(j)}\right) to Ξ(𝗌i(j))\Xi\left(\mathsf{s}_{i^{\prime}}^{(j^{\prime})}\right). Thus, under the bijection Ξ\Xi, the Hasse diagram of Heap(𝗌𝗈𝗋𝗍𝖼(w))\mathrm{Heap}(\mathsf{sort}_{\mathsf{c}}(w_{\circ})) is precisely the AR quiver of KQcKQ_{c}. (In this setting, the dimension of irr(M,N)\mathrm{irr}(M,N) does not exceed 1.) In particular, the drawing of Heap(𝗌𝗈𝗋𝗍𝖼(w))\mathrm{Heap}(\mathsf{sort}_{\mathsf{c}}(w_{\circ})) shown as the shaded yellow part of Figure 6 agrees with the standard way of embedding the AR quiver of KQcKQ_{c} in the plane; see, e.g., [Sch14, Section 3.1].

It is also shown in [STW, Theorem 9.3.1] that, for all siSs_{i}\in S, the projective module P(i)P(i) satisfies P(i)=Ξ(𝗌i(1))P(i)=\Xi\left(\mathsf{s}_{i}^{(1)}\right). Because every nonzero morphism can be written as a composition of irreducible morphisms, the first copy of Qc=Heap(𝖼)Q_{c}=\mathrm{Heap}(\mathsf{c}) in Heap(𝗌𝗈𝗋𝗍𝖼(w))\mathrm{Heap}(\mathsf{sort}_{\mathsf{c}}(w_{\circ})) thus provides a visualization of the well-known fact that EndKQc(i=1nP(i))KQc\mathrm{End}_{KQ_{c}}\left(\bigoplus_{i=1}^{n}P(i)\right)\cong KQ_{c}.

10. Future Directions

10.1. Images

Consider the linear Coxeter element c=s1s2snc^{\to}=s_{1}s_{2}\cdots s_{n} of AnA_{n}. The Cambrian lattice Cambc\mathrm{Camb}_{c^{\to}} is the (n+1)(n+1)-st Tamari lattice. Hong [Hon22] proved that the size of the image of popCambc\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c^{\to}}} is the nn-th Motzkin number (i.e., the number of Motzkin paths of length nn). In Section 8, we determined the size of the image of popCambc×\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c^{\times}}}, where c×c^{\times} is a bipartite Coxeter element of AnA_{n}. Using these formulas, one can verify that |popCambc(Cambc)||popCambc×(Cambc×)||\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c^{\to}}}(\mathrm{Camb}_{c^{\to}})|\leq|\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c^{\times}}}(\mathrm{Camb}_{c^{\times}})|. Numerical evidence has led us to conjecture that the linear and bipartite Coxeter elements are, in some sense, extremal with regard to the sizes of the images of pop-stack operators.

Conjecture 10.1.

For every Coxeter element cc of AnA_{n}, we have

|popCambc(Cambc)||popCambc(Cambc)||popCambc×(Cambc×)|.|\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c^{\to}}}(\mathrm{Camb}_{c^{\to}})|\leq|\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c}}(\mathrm{Camb}_{c})|\leq|\mathrm{pop}^{\downarrow}_{\mathrm{Camb}_{c^{\times}}}(\mathrm{Camb}_{c^{\times}})|.

10.2. Forward orbits

Suppose LL is a finite lattice, and let υL=maxxL|𝒪L(x)|\upsilon_{L}=\max_{x\in L}|\mathcal{O}_{L}(x)|. Let

ΥL={xL|𝒪L(x)|=υL}\Upsilon_{L}=\{x\in L\mid|\mathcal{O}_{L}(x)|=\upsilon_{L}\}

be the set of elements of LL whose forward orbits under popL\mathrm{pop}^{\downarrow}_{L} attain the maximum possible size.

Let cc be a Coxeter element of a finite irreducible Coxeter group WW. We saw in Theorem 9.17 that υCambc\upsilon_{\mathrm{Camb}_{c}} is the Coxeter number hh, and we constructed an explicit element 𝐳cΥCambc{\bf z}_{c}\in\Upsilon_{\mathrm{Camb}_{c}}. However, we said nothing about the other elements ΥCambc\Upsilon_{\mathrm{Camb}_{c}}. In the case when W=AnW=A_{n} and cc is the linear Coxeter element cc^{\to} (i.e., Cambc\mathrm{Camb}_{c} is the (n+1)(n+1)-st Tamari lattice), it is known that |ΥCambc||\Upsilon_{\mathrm{Camb}_{c}}| is the (n1)(n-1)-st Catalan number [Def22a]. It would be interesting to understand ΥCambc\Upsilon_{\mathrm{Camb}_{c}} for other choices of cc. In particular, we have the following conjecture.

Conjecture 10.2.

Assume n3n\geq 3, and let c×c^{\times} be the bipartite Coxeter element of type AnA_{n} given by

c×=i[n]i oddsij[n]j evensj.c^{\times}=\prod_{\begin{subarray}{c}i\in[n]\\ i\text{ odd}\end{subarray}}s_{i}\prod_{\begin{subarray}{c}j\in[n]\\ j\text{ even}\end{subarray}}s_{j}.

The set of elements of Cambc×\mathrm{Camb}_{c^{\times}} whose forward orbits under the pop-stack operator have size hh is given by

ΥCambc×={{𝐳c}if n is odd;{𝐳c,𝐳cs1}if n is even.\Upsilon_{\mathrm{Camb}_{c^{\times}}}=\begin{cases}\{{\bf z}_{c}\}&\mbox{if }n\mbox{ is odd};\\ \{{\bf z}_{c},{\bf z}_{c}s_{1}\}&\mbox{if }n\mbox{ is even}.\end{cases}

The original use of the term pop-stack comes from the setting where LL is the weak order on AnA_{n}; in this case, Ungar proved that maxxWeak(An)|𝒪Weak(An)(x)|\max_{x\in\mathrm{Weak}(A_{n})}|\mathcal{O}_{\mathrm{Weak}(A_{n})}(x)| is n+1n+1 (which is the Coxeter number of AnA_{n}).

Question 10.3.

What can be said about |ΥWeak(An)||\Upsilon_{\mathrm{Weak}(A_{n})}|?

Defant [Def22b] proved that if WW is a finite irreducible Coxeter group with Coxeter number hh, then maxxW|𝒪Weak(W)(x)|=h\max_{x\in W}|\mathcal{O}_{\mathrm{Weak}(W)}(x)|=h. In Theorem 9.6, we found that maxxL|𝒪L(x)|h\max_{x\in L}|\mathcal{O}_{L}(x)|\leq h whenever LL is a lattice quotient of Weak(W)\mathrm{Weak}(W), and we saw in Theorem 9.17 that this inequality is an equality whenever LL is a Cambrian lattice. We are naturally led to ask the following questions.

Question 10.4.

Let WW be a finite irreducible Coxeter group with Coxeter number hh. For which lattice quotients LL of Weak(W)\mathrm{Weak}(W) is it the case that maxxL|𝒪L(x)|=h\max_{x\in L}|\mathcal{O}_{L}(x)|=h?

Question 10.5.

Let LL^{\prime} be a lattice quotient of a finite lattice LL. Is it necessarily the case that

maxxL|𝒪L(x)|maxxL|𝒪L(x)|?\max_{x^{\prime}\in L^{\prime}}|\mathcal{O}_{L^{\prime}}(x^{\prime})|\leq\max_{x\in L}|\mathcal{O}_{L}(x)|?

10.3. Generalizations

It would be interesting to see how much of our work on Cambrian lattices can be extended to more general families of lattices. For example, it could be interesting to study the pop-stack operators on mm-Cambrian lattices, which were introduced by Stump, Thomas, and Williams [STW].

Appendix A Proof of Theorem 4.6

In this appendix, we prove Theorem 4.6. More precisely, we prove only the first half of the theorem, as the second half is just the dual. We follow arguments similar to those appearing in [HI21a, Section 3] and [KY14, Section 7.2]. The main difference is that we work almost exclusively in the category 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda (with the exception of a portion of the proof of Lemma A.1 below), while the referenced papers work primarily in the bounded derived category.

In the arguments that follow, we recall that for any wide subcategory 𝒲𝗆𝗈𝖽Λ\mathcal{W}\subseteq\operatorname{\mathsf{mod}}\Lambda and any M,N𝒲M,N\in\mathcal{W}, we have HomΛ(M,N)=Hom𝒲(M,N)\mathrm{Hom}_{\Lambda}(M,N)=\mathrm{Hom}_{\mathcal{W}}(M,N) and ExtΛ1(M,N)=Ext𝒲1(M,N)\mathrm{Ext}^{1}_{\Lambda}(M,N)=\mathrm{Ext}^{1}_{\mathcal{W}}(M,N). On the other hand, we may have that ExtΛm(M,N)Ext𝒲m(M,N)\mathrm{Ext}^{m}_{\Lambda}(M,N)\supsetneq\mathrm{Ext}^{m}_{\mathcal{W}}(M,N) for m>1m>1.

Lemma A.1.

Let (𝒳,𝒴)sbpΛ(\mathcal{X},\mathcal{Y})\in\operatorname{\mathrm{sbp}}\Lambda be SM compatible. Let 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}, X𝒳𝒳X\in\mathcal{X}\setminus\mathcal{X}^{\prime}, and Z𝖥𝗂𝗅𝗍(𝒳)Z\in\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}).

  1. (1)

    The map HomΛ(XX,Z)ExtΛ1(X,Z)\mathrm{Hom}_{\Lambda}(X^{\prime}_{X},Z)\rightarrow\mathrm{Ext}^{1}_{\Lambda}(X,Z) induced by the short exact sequence ηX,𝒳\eta_{X,\mathcal{X}^{\prime}} from Equation 4.1 is bijective.

  2. (2)

    The map ExtΛ1(XX,Z)ExtΛ2(X,Z)\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{X},Z)\rightarrow\mathrm{Ext}^{2}_{\Lambda}(X,Z) induced by the short exact sequence ηX,𝒳\eta_{X,\mathcal{X}^{\prime}} from Equation 4.1 is injective.

Proof.

1 Since 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) is a wide subcategory, we have that ExtΛ1(X,Z)HomΛ(Ω𝒳X,Z)\mathrm{Ext}^{1}_{\Lambda}(X,Z)\cong\mathrm{Hom}_{\Lambda}(\Omega_{\mathcal{X}}X,Z). We have ker(γX,𝒳)𝖥𝗂𝗅𝗍(𝒳)\mathrm{ker}(\gamma_{X,\mathcal{X}^{\prime}})\in\operatorname{\mathsf{Filt}}(\mathcal{X}) since 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) is a wide subcategory. Now, 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}) is a Serre subcategory of 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}), so in particular, there is a torsion pair (𝒳,𝖥𝗂𝗅𝗍(𝒳))(\prescript{\perp}{}{\mathcal{X}^{\prime}},\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})) in 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}). This implies that ker(γX,𝒳)𝒳\mathrm{ker}(\gamma_{X,\mathcal{X}^{\prime}})\in\prescript{\perp}{}{\mathcal{X}^{\prime}}. From the exact sequence

0HomΛ(XX,Z)HomΛ(Ω𝒳X,Z)HomΛ(ker(γX,𝒳,Z),0\rightarrow\mathrm{Hom}_{\Lambda}(X^{\prime}_{X},Z)\rightarrow\mathrm{Hom}_{\Lambda}(\Omega_{\mathcal{X}}X,Z)\rightarrow\mathrm{Hom}_{\Lambda}(\mathrm{ker}(\gamma_{X,\mathcal{X}^{\prime}},Z),

we therefore conclude that the induced map HomΛ(XX,Z)HomΛ(Ω𝒳X,Z)ExtΛ1(X,Z)\mathrm{Hom}_{\Lambda}(X^{\prime}_{X},Z)\rightarrow\mathrm{Hom}_{\Lambda}(\Omega_{\mathcal{X}}X,Z)\cong\mathrm{Ext}^{1}_{\Lambda}(X,Z) is a bijection.

2 We first claim that HomΛ(Ω𝒳2X,Z)ExtΛ2(X,Z)\mathrm{Hom}_{\Lambda}(\Omega^{2}_{\mathcal{X}}X,Z)\subseteq\mathrm{Ext}^{2}_{\Lambda}(X,Z). We know from Theorem 3.7 that there exists 𝒳u𝗌𝖻𝗋𝗂𝖼𝗄Λ\mathcal{X}_{u}\in\operatorname{\mathsf{sbrick}}\Lambda such that (𝒳,𝒳u)(\mathcal{X},\mathcal{X}_{u}) is a 2-term simple-minded collection. Then, as in [KY14, Lemma 7.8], there exist a finite-dimensional algebra Γ\Gamma and a triangle functor

𝗋𝖾𝖺𝗅:𝒟b(𝗆𝗈𝖽Γ)𝒟b(𝗆𝗈𝖽Λ){\mathsf{real}\colon\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Gamma)\rightarrow\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda)}

such that the following hold:

  • 𝗆𝗈𝖽Γ=𝖥𝗂𝗅𝗍(𝒳𝒳u[1])\operatorname{\mathsf{mod}}\Gamma=\operatorname{\mathsf{Filt}}(\mathcal{X}\oplus\mathcal{X}_{u}[1]).

  • 𝗋𝖾𝖺𝗅\mathsf{real} acts as the identity on 𝗆𝗈𝖽Γ\operatorname{\mathsf{mod}}\Gamma.

  • For all M,N𝗆𝗈𝖽ΓM,N\in\operatorname{\mathsf{mod}}\Gamma, the induced map ExtΓ1(M,N)Hom𝒟b(𝗆𝗈𝖽Λ)(M,N[1])\mathrm{Ext}^{1}_{\Gamma}(M,N)\rightarrow\mathrm{Hom}_{\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda)}(M,N[1]) is a bijection, and the induced map ExtΓ2(M,N)Hom𝒟b(𝗆𝗈𝖽Λ)(M,N[2])\mathrm{Ext}^{2}_{\Gamma}(M,N)\rightarrow\mathrm{Hom}_{\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda)}(M,N[2]) is an injection.

Now, 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) is a Serre subcategory of 𝗆𝗈𝖽Γ\operatorname{\mathsf{mod}}\Gamma and a wide subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda. Thus, taking M=XM=X and N=ZN=Z above, we find that 𝗋𝖾𝖺𝗅\mathsf{real} induces an inclusion ExtΓ2(X,Z)ExtΛ2(X,Z)\mathrm{Ext}^{2}_{\Gamma}(X,Z)\subseteq\mathrm{Ext}^{2}_{\Lambda}(X,Z). To prove the claim, it therefore suffices to show that (HomΛ(Ω𝒳2X,Z)=)HomΓ(Ω𝒳2,Z)ExtΓ1(X,Z)(\mathrm{Hom}_{\Lambda}(\Omega^{2}_{\mathcal{X}}X,Z)=)\mathrm{Hom}_{\Gamma}(\Omega^{2}_{\mathcal{X}},Z)\subseteq\mathrm{Ext}^{1}_{\Gamma}(X,Z).

Consider the following commutative diagram, where the first row consists of projective covers and syzygies in 𝗆𝗈𝖽Γ\operatorname{\mathsf{mod}}\Gamma and the second row consists of projective covers and syzygies in 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}).

Ω2X{\Omega^{2}X}PX1{P_{X}^{1}}ΩX{\Omega X}PX0{P_{X}^{0}}X{X}Ω𝒳2X{\Omega^{2}_{\mathcal{X}}X}P𝒳,X1{P_{\mathcal{X},X}^{1}}Ω𝒳X{\Omega_{\mathcal{X}}X}P𝒳,X0{P_{\mathcal{X},X}^{0}}X.{X.}g\scriptstyle{g}f\scriptstyle{f}

Since 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) is a Serre subcategory of 𝗆𝗈𝖽Γ\operatorname{\mathsf{mod}}\Gamma, every module in 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) has the same top in both 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) and in 𝗆𝗈𝖽Γ\operatorname{\mathsf{mod}}\Gamma. Thus, the induced map PX0P𝒳,X0P_{X}^{0}\rightarrow P_{\mathcal{X},X}^{0}, and so also ff, must be surjective. Now for S𝒳S\in\mathcal{X}, we get an induced exact sequence

0HomΓ(P𝒳,X0,S)HomΓ(PX0,S)HomΓ(kerf,S)ExtΓ1(P𝒳,X0,S)=0,0\rightarrow\mathrm{Hom}_{\Gamma}(P_{\mathcal{X},X}^{0},S)\rightarrow\mathrm{Hom}_{\Gamma}(P_{X}^{0},S)\rightarrow\mathrm{Hom}_{\Gamma}(\mathrm{ker}f,S)\rightarrow\mathrm{Ext}^{1}_{\Gamma}(P_{\mathcal{X},X}^{0},S)=0,

where the last term is 0 because P𝒳,X0P_{\mathcal{X},X}^{0} is projective in 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}). Since PX0P_{X}^{0} and P𝒳,X0P_{\mathcal{X},X}^{0} have the same top, this implies that HomΓ(kerf,S)=0\mathrm{Hom}_{\Gamma}(\mathrm{ker}f,S)=0, and hence also that kerf𝒳\mathrm{ker}f\in\prescript{\perp}{}{\mathcal{X}}. Moreover, we see that the top of Ω𝒳X\Omega_{\mathcal{X}}X is a direct summand of the top of ΩX\Omega X, and thus that the induced map PX1P𝒳,X1P_{X}^{1}\twoheadrightarrow P_{\mathcal{X},X}^{1} is surjective. From the snake lemma, this means there is a surjective map ker(f)coker(g)\mathrm{ker}(f)\twoheadrightarrow\mathrm{coker}(g). On the other hand, coker(g)𝖥𝗂𝗅𝗍(𝒳)\mathrm{coker}(g)\in\operatorname{\mathsf{Filt}}(\mathcal{X}) since 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}) is a Serre subcategory of 𝗆𝗈𝖽Γ\operatorname{\mathsf{mod}}\Gamma. It follows that gg is surjective. In particular, the induced map HomΓ(Ω𝒳2X,Z)HomΓ(Ω2X,Z)=ExtΓ1(X,Z)\mathrm{Hom}_{\Gamma}(\Omega_{\mathcal{X}}^{2}X,Z)\rightarrow\mathrm{Hom}_{\Gamma}(\Omega^{2}X,Z)=\mathrm{Ext}^{1}_{\Gamma}(X,Z) is injective. This proves the claim.

Now consider the following commutative diagram of projective covers and syzygies in 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}):

Ω𝒳2X{\Omega_{\mathcal{X}}^{2}X}P𝒳,X1{P^{1}_{\mathcal{X},X}}Ω𝒳X{\Omega_{\mathcal{X}}X}Ω𝒳XX{\Omega_{\mathcal{X}}X^{\prime}_{X}}P𝒳,XX0{P_{\mathcal{X},X^{\prime}_{X}}^{0}}XX.{X^{\prime}_{X}.}αX,𝒳\scriptstyle{\alpha_{X,\mathcal{X}^{\prime}}}γX,𝒳\scriptstyle{\gamma_{X,\mathcal{X}^{\prime}}}

The induced map P𝒳,X1P𝒳,XX0P_{\mathcal{X},X}^{1}\rightarrow P_{\mathcal{X},X^{\prime}_{X}}^{0} is a (split) epimorphism. Consequently, we have a surjection ker(γX,𝒳)coker(αX,𝒳)\mathrm{ker}(\gamma_{X,\mathcal{X}^{\prime}})\twoheadrightarrow\mathrm{coker}(\alpha_{X,\mathcal{X}^{\prime}}). On the other hand, we have that ker(γX,𝒳)𝒳\mathrm{ker}(\gamma_{X,\mathcal{X}^{\prime}})\in\prescript{\perp}{}{\mathcal{X}^{\prime}}, as observed in the proof of 1. Similarly, we have coker(αX,𝒳)𝖥𝗂𝗅𝗍(𝒳)\mathrm{coker}(\alpha_{X,\mathcal{X}^{\prime}})\in\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}) since 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}) is a Serre subcategory of 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}). We conclude that coker(αX,𝒳)=0\mathrm{coker}(\alpha_{X,\mathcal{X}^{\prime}})=0. In particular, the induced map

Ext𝖥𝗂𝗅𝗍(𝒳)1(XX,Z)=HomΓ(Ω𝒳XX,Z)HomΓ(Ω𝒳2X,Z)=Ext𝖥𝗂𝗅𝗍(𝒳)2(X,Z)\mathrm{Ext}^{1}_{\operatorname{\mathsf{Filt}}(\mathcal{X})}(X^{\prime}_{X},Z)=\mathrm{Hom}_{\Gamma}(\Omega_{\mathcal{X}}X^{\prime}_{X},Z)\rightarrow\mathrm{Hom}_{\Gamma}(\Omega^{2}_{\mathcal{X}}X,Z)=\mathrm{Ext}^{2}_{\operatorname{\mathsf{Filt}}(\mathcal{X})}(X,Z)

is injective. Together with the claim, the fact that ExtΛ1(XX,Z)=Ext𝖥𝗂𝗅𝗍(𝒳)1(XX,Z)\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{X},Z)=\mathrm{Ext}^{1}_{\operatorname{\mathsf{Filt}}(\mathcal{X})}(X^{\prime}_{X},Z) then implies the result. ∎

Lemma A.2.

Let (𝒳,𝒴)sbpΛ(\mathcal{X},\mathcal{Y})\in\operatorname{\mathrm{sbp}}\Lambda be SM compatible. Let 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}, Y𝒴Y\in\mathcal{Y}, and Z𝖥𝗂𝗅𝗍(𝒳)Z\in\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}).

  1. (1)

    The map HomΛ(XY,Z)HomΛ(Y,Z)\mathrm{Hom}_{\Lambda}(X^{\prime}_{Y},Z)\rightarrow\mathrm{Hom}_{\Lambda}(Y,Z) induced by gY,𝒳g_{Y,\mathcal{X}^{\prime}} is bijective.

  2. (2)

    The map ExtΛ1(XY,Z)ExtΛ1(Y,Z)\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{Y},Z)\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Y,Z) induced by gY,𝒳g_{Y,\mathcal{X}^{\prime}} is injective.

Proof.

1 Let f:XYZf\colon X^{\prime}_{Y}\rightarrow Z be such that fgY,𝒳=0f\circ g_{Y,\mathcal{X}^{\prime}}=0. Equivalently, we have Im(gY,𝒳)ker(f)\mathrm{Im}(g_{Y,\mathcal{X}^{\prime}})\subseteq\mathrm{ker}(f). Note that ker(f)𝖥𝗂𝗅𝗍(𝒳)\mathrm{ker}(f)\in\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}) since 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}) is a wide subcategory. Since gY,𝒳g_{Y,\mathcal{X}^{\prime}} is a right minimal 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})-approximation, it follows that Y𝒳=kerfY_{\mathcal{X}}^{\prime}=\mathrm{ker}f; i.e., f=0f=0. This shows that the induced map is injective; the fact that it is surjective comes from the definition of an approximation.

2 Let ηExtΛ1(XY,Z)\eta\in\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{Y},Z) be such that ηgY,𝒳=0\eta\circ g_{Y,\mathcal{X}^{\prime}}=0. Then η\eta is a short exact sequence of the form

η:{\eta:}Z{Z}E{E}XY,{X^{\prime}_{Y},}q\scriptstyle{q}

so this means that there exists h:YEh\colon Y\rightarrow E such that gY,𝒳=qhg_{Y,\mathcal{X}^{\prime}}=q\circ h. Now, E𝖥𝗂𝗅𝗍(𝒳)E\in\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}), so there exists h:XYEh^{\prime}\colon X^{\prime}_{Y}\rightarrow E such that h=hgY,𝒳h=h^{\prime}\circ g_{Y,\mathcal{X}^{\prime}}. We therefore have

qh=qhgY,𝒳=IdXYgY,𝒳.q\circ h=q\circ h^{\prime}\circ g_{Y,\mathcal{X}^{\prime}}=\mathrm{Id}_{X^{\prime}_{Y}}\circ g_{Y,\mathcal{X}^{\prime}}.

It then follows from 1 that qh=IdXYq\circ h^{\prime}=\mathrm{Id}_{X^{\prime}_{Y}}; i.e., the exact sequence η\eta is split. ∎

Proposition A.3 (Theorem 4.6, 1).

Let (𝒳,𝒴)sbpΛ(\mathcal{X},\mathcal{Y})\in\operatorname{\mathrm{sbp}}\Lambda be SM compatible, and let 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}. Then μ𝒳(𝒳,𝒴)\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y}) is a semibrick pair.

Proof.

The proof is similar to that of [HI21a, Proposition 3.5]. For readability, let μ𝒳(𝒳,𝒴)d=𝒳1\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{d}=\mathcal{X}_{1} and μ𝒳(𝒳,𝒴)u=𝒴1\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})_{u}=\mathcal{Y}_{1}.

Claim 1: The set 𝒳2\mathcal{X}_{2} is a semibrick.

Let Z1,Z2𝒳2Z_{1},Z_{2}\in\mathcal{X}_{2}. First suppose that Z1𝒳Z_{1}\in\mathcal{X}^{\prime}. If Z2𝒳Z_{2}\in\mathcal{X}^{\prime}, then there is nothing to show since 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X} is a semibrick. If Z2=ker(gY,𝒳)Z_{2}=\mathrm{ker}(g_{Y,\mathcal{X}^{\prime}}) for some Y𝒴Y\in\mathcal{Y} such that gY,𝒳g_{Y,\mathcal{X}^{\prime}} is surjective, then HomΛ(Z1,Z2)=0\mathrm{Hom}_{\Lambda}(Z_{1},Z_{2})=0 because Z2YZ_{2}\subseteq Y and HomΛ(Z1,Y)=0\mathrm{Hom}_{\Lambda}(Z_{1},Y)=0.

Now suppose that there exists Y1𝒴Y_{1}\in\mathcal{Y} such that gY1,𝒳g_{Y_{1},\mathcal{X}^{\prime}} is surjective and Z1=ker(gY1,𝒳).Z_{1}=\mathrm{ker}(g_{Y_{1},\mathcal{X}^{\prime}}). We first consider the case Z2𝒳Z_{2}\in\mathcal{X}^{\prime}. Then we have a long exact sequence

0HomΛ(XY,Z2)HomΛ(Y,Z2)0HomΛ(Z,Z2)00\rightarrow\mathrm{Hom}_{\Lambda}(X^{\prime}_{Y},Z_{2})\rightarrow\mathrm{Hom}_{\Lambda}(Y,Z_{2})\xrightarrow{0}\mathrm{Hom}_{\Lambda}(Z,Z_{2})\xrightarrow{0}
0ExtΛ1(XY,Z2)ExtΛ1(Y,Z2),\xrightarrow{0}\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{Y},Z_{2})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Y,Z_{2}),

where we know the indicated maps are 0 by Lemma A.2.

It remains to consider the case where there exists Y2𝒴Y_{2}\in\mathcal{Y} such that gY2,𝒳g_{Y_{2},\mathcal{X}^{\prime}} is surjective and Z2=ker(gY2,𝒳)Z_{2}=\mathrm{ker}(g_{Y_{2},\mathcal{X}^{\prime}}). We then have a long exact sequence

0HomΛ(Z1,Z2)HomΛ(Z1,Y2)HomΛ(Z1,XY2)=0,0\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},Z_{2})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},Y_{2})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},X^{\prime}_{Y_{2}})=0,

where the last term is 0 by the case considered above. Similarly, we have a long exact sequence

0=HomΛ(Y1,Y2)HomΛ(Z1,Y2)ExtΛ1(XY1,Y2).0=\mathrm{Hom}_{\Lambda}(Y_{1},Y_{2})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},Y_{2})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{Y_{1}},Y_{2}).

Together, these sequences show that HomΛ(Z1,Z2)Hom(Y1,Y2)\mathrm{Hom}_{\Lambda}(Z_{1},Z_{2})\cong\mathrm{Hom}(Y_{1},Y_{2}). In particular, setting Z1=Z2Z_{1}=Z_{2}, we conclude that Z1Z_{1} is a brick. This proves the claim.

Claim 2: We have HomΛ(Z,X1)=0=ExtΛ1(Z,X1)\mathrm{Hom}_{\Lambda}(Z,X_{1})=0=\mathrm{Ext}^{1}_{\Lambda}(Z,X_{1}) for all X1𝒳X_{1}\in\mathcal{X}^{\prime} and Z𝒳1Z\in\mathcal{X}_{1}.

Again, there are two possibilities. Suppose first that there exists X2𝒳𝒳X_{2}\in\mathcal{X}\setminus\mathcal{X}^{\prime} such that Z=EX2,𝒳{Z=E_{X_{2},\mathcal{X}^{\prime}}}. Then we have a long exact sequence

0=HomΛ(X2,X1)HomΛ(Z,X1)0HomΛ(XX2,X1)0=\mathrm{Hom}_{\Lambda}(X_{2},X_{1})\rightarrow\mathrm{Hom}_{\Lambda}(Z,X_{1})\xrightarrow{0}\mathrm{Hom}_{\Lambda}(X^{\prime}_{X_{2}},X_{1})\rightarrow
ExtΛ1(X2,X1)0ExtΛ1(Z,X1)0ExtΛ1(XX2,X1)ExtΛ2(X2,X1),\rightarrow\mathrm{Ext}^{1}_{\Lambda}(X_{2},X_{1})\xrightarrow{0}\mathrm{Ext}^{1}_{\Lambda}(Z,X_{1})\xrightarrow{0}\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{X_{2}},X_{1})\rightarrow\mathrm{Ext}^{2}_{\Lambda}(X_{2},X_{1}),

where we know the indicated maps are 0 by Lemma A.1. This proves the claim in this case.

The other case to consider is that in which there exists Y𝒴Y\in\mathcal{Y} such that gY,𝒳g_{Y,\mathcal{X}^{\prime}} is injective and Z=coker(gY,𝒳)Z=\mathrm{coker}(g_{Y,\mathcal{X}^{\prime}}). In this case, we have a long exact sequence

0HomΛ(Z,X1)0HomΛ(XY,X1)HomΛ(Y,X1)00\rightarrow\mathrm{Hom}_{\Lambda}(Z,X_{1})\xrightarrow{0}\mathrm{Hom}_{\Lambda}(X^{\prime}_{Y},X_{1})\rightarrow\mathrm{Hom}_{\Lambda}(Y,X_{1})\xrightarrow{0}
0ExtΛ1(Z,X1)0ExtΛ1(XY,X1)ExtΛ1(Y,X1),\xrightarrow{0}\mathrm{Ext}^{1}_{\Lambda}(Z,X_{1})\xrightarrow{0}\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{Y},X_{1})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Y,X_{1}),

where again we know the indicated maps are 0 by Lemma A.2. This proves the claim.

Claim 3: We have HomΛ(Z,ker(gY1,𝒳))=0=ExtΛ1(Z,ker(gY1,𝒳))\mathrm{Hom}_{\Lambda}(Z,\mathrm{ker}(g_{Y_{1},\mathcal{X}^{\prime}}))=0=\mathrm{Ext}^{1}_{\Lambda}(Z,\mathrm{ker}(g_{Y_{1},\mathcal{X}^{\prime}})) for all Z𝒳1Z\in\mathcal{X}_{1} and Y1𝒴Y_{1}\in\mathcal{Y} such that gY1,𝒳g_{Y_{1},\mathcal{X}^{\prime}} is surjective.

Again, there are two possibilities. Suppose first that there exists X2𝒳𝒳X_{2}\in\mathcal{X}\setminus\mathcal{X}^{\prime} such that Z=EX2,𝒳{Z=E_{X_{2},\mathcal{X}^{\prime}}}. Then HomΛ(Z,ker(gY1,𝒳))=0\mathrm{Hom}_{\Lambda}(Z,\mathrm{ker}(g_{Y_{1},\mathcal{X}^{\prime}}))=0 because Z𝖥𝗂𝗅𝗍(𝒳)Z\in\operatorname{\mathsf{Filt}}(\mathcal{X}). Moreover, we have a long exact sequence

0=HomΛ(Z,𝒳Y1)ExtΛ1(Z,ker(gY1,𝒳))ExtΛ1(Z,Y1)=0,0=\mathrm{Hom}_{\Lambda}(Z,\mathcal{X}^{\prime}_{Y_{1}})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,\mathrm{ker}(g_{Y_{1},\mathcal{X}^{\prime}}))\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,Y_{1})=0,

where the first term is 0 by Claim 2 and the last term is 0 because Z𝖥𝗂𝗅𝗍(𝒳)Z\in\operatorname{\mathsf{Filt}}(\mathcal{X}). This proves the claim in this case.

The other case to consider is that in which there exists Y2𝒴Y_{2}\in\mathcal{Y} such that gY2,𝒳g_{Y_{2},\mathcal{X}^{\prime}} is injective and Z=coker(gY2,𝒳)Z=\mathrm{coker}(g_{Y_{2},\mathcal{X}^{\prime}}). The fact that HomΛ(Z,ker(gY1,𝒳))=0\mathrm{Hom}_{\Lambda}(Z,\mathrm{ker}(g_{Y_{1},\mathcal{X}^{\prime}}))=0 then follows from the fact that Y1𝒳Y_{1}\in\mathcal{X}^{\perp}. Moreover, we have an exact sequence

0=HomΛ(Z,XY1)ExtΛ1(Z,ker(gY1,𝒳))0=\mathrm{Hom}_{\Lambda}(Z,X^{\prime}_{Y_{1}})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,\mathrm{ker}(g_{Y_{1},\mathcal{X}^{\prime}}))\rightarrow
ExtΛ1(Z,Y1)ExtΛ1(Z,XY1)=0,\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,Y_{1})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,X^{\prime}_{Y_{1}})=0,

where the first and last terms are 0 by Claim 2. Likewise, we have an exact sequence

0=HomΛ(Y2,Y1)ExtΛ1(Z,Y1)ExtΛ1(XY2,Y1)=0,0=\mathrm{Hom}_{\Lambda}(Y_{2},Y_{1})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,Y_{1})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{Y_{2}},Y_{1})=0,

where the first and last terms are 0 since Y1Y2𝒳0,1Y_{1}\in Y_{2}^{\perp}\cap\mathcal{X}^{\perp_{0,1}}. Together, these two exact sequences yield that ExtΛ1(Z,ker(gY1,𝒳))=0\mathrm{Ext}^{1}_{\Lambda}(Z,\mathrm{ker}(g_{Y_{1},\mathcal{X}^{\prime}}))=0, as desired.

Claim 4: The set 𝒳1\mathcal{X}_{1} is a semibrick.

Let Z1,Z2𝒳1Z_{1},Z_{2}\in\mathcal{X}_{1}. First suppose that there exist X1,X2𝒳𝒳X_{1},X_{2}\in\mathcal{X}\setminus\mathcal{X}^{\prime} such that Z1=EX1,𝒳Z_{1}=E_{X_{1},\mathcal{X}^{\prime}} and Z2=EX2,𝒳Z_{2}=E_{X_{2},\mathcal{X}^{\prime}}. Then there is an exact sequence

0=HomΛ(Z1,XX2)HomΛ(Z1,Z2)HomΛ(Z1,X2)ExtΛ1(Z1,XX2)=0,0=\mathrm{Hom}_{\Lambda}(Z_{1},X^{\prime}_{X_{2}})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},Z_{2})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},X_{2})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z_{1},X^{\prime}_{X_{2}})=0,

where the first and last terms are 0 by Claim 2. We also have a long exact sequence

0Hom(X1,X2)Hom(Z1,X2)Hom(XX1,X2)=0.0\rightarrow\mathrm{Hom}(X_{1},X_{2})\rightarrow\mathrm{Hom}(Z_{1},X_{2})\rightarrow\mathrm{Hom}(X^{\prime}_{X_{1}},X_{2})=0.

Together, these sequences show that HomΛ(Z1,Z2)HomΛ(X1,X2)\mathrm{Hom}_{\Lambda}(Z_{1},Z_{2})\cong\mathrm{Hom}_{\Lambda}(X_{1},X_{2}). In particular, by setting Z1=Z2Z_{1}=Z_{2}, we conclude that Z1Z_{1} is a brick. This proves the claim in this case.

Now suppose there exists Y1𝒴Y_{1}\in\mathcal{Y} such that gY1,𝒳g_{Y_{1},\mathcal{X}^{\prime}} is injective and Z1=coker(gY1,𝒳)Z_{1}=\mathrm{coker}(g_{Y_{1},\mathcal{X}^{\prime}}) and that there exists X2𝒳𝒳X_{2}\in\mathcal{X}\setminus\mathcal{X}^{\prime} such that Z2=EX2,𝒳Z_{2}=E_{X_{2},\mathcal{X}^{\prime}}. Then HomΛ(Z1,X2)=0\mathrm{Hom}_{\Lambda}(Z_{1},X_{2})=0 since X2(𝒳)X_{2}\in(\mathcal{X}^{\prime})^{\perp}. We then have an exact sequence

0=HomΛ(Z1,XX2)HomΛ(Z1,Z2)HomΛ(Z1,X2)=0,0=\mathrm{Hom}_{\Lambda}(Z_{1},X^{\prime}_{X_{2}})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},Z_{2})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},X_{2})=0,

where the first term is 0 by Claim 2. Similarly, we have an exact sequence

0=HomΛ(Z2,XY1)HomΛ(Z2,Z1)ExtΛ1(Z2,Y1)=0,0=\mathrm{Hom}_{\Lambda}(Z_{2},X^{\prime}_{Y_{1}})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{2},Z_{1})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z_{2},Y_{1})=0,

where the first term is 0 by Claim 2 and the last term is 0 because Y𝒳0,1Y\in\mathcal{X}^{\perp_{0,1}}. We conclude that Z1Z2Z2Z_{1}\in Z_{2}^{\perp}\cap\prescript{\perp}{}{Z_{2}}.

It remains to consider the case where there exist Y1,Y2𝒴Y_{1},Y_{2}\in\mathcal{Y} such that gY1,𝒳g_{Y_{1},\mathcal{X}^{\prime}} and gY2,𝒳g_{Y_{2},\mathcal{X}^{\prime}} are injective, Z1=coker(gY1,𝒳)Z_{1}=\mathrm{coker}(g_{Y_{1},\mathcal{X}^{\prime}}), and Z2=coker(gY2,𝒳)Z_{2}=\mathrm{coker}(g_{Y_{2},\mathcal{X}^{\prime}}). As in earlier cases, we have an exact sequence

0=HomΛ(Z1,XY2)HomΛ(Z1,Z2)HomΛ(Z1,Y2[1])ExtΛ1(Z1,XY2)=0,0=\mathrm{Hom}_{\Lambda}(Z_{1},X^{\prime}_{Y_{2}})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},Z_{2})\rightarrow\mathrm{Hom}_{\Lambda}(Z_{1},Y_{2}[1])\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z_{1},X^{\prime}_{Y_{2}})=0,

where the first and last terms are 0 by Claim 2. Similarly, we have an exact sequence

0=Hom(XY1,Y2)HomΛ(Y1,Y2)ExtΛ1(Z1,Y2)ExtΛ1(XY1,Y2)=0,0=\mathrm{Hom}(X^{\prime}_{Y_{1}},Y_{2})\rightarrow\mathrm{Hom}_{\Lambda}(Y_{1},Y_{2})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z_{1},Y_{2})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{Y_{1}},Y_{2})=0,

where the first and last terms are 0 since Y𝒳0,1Y\in\mathcal{X}^{\perp_{0,1}}. We again conclude that HomΛ(Z1,Z2)HomΛ(Y1,Y2)\mathrm{Hom}_{\Lambda}(Z_{1},Z_{2})\cong\mathrm{Hom}_{\Lambda}(Y_{1},Y_{2}). This concludes the proof. ∎

Proposition A.4 (Theorem 4.6,25).

Let (𝒳,𝒴)sbpΛ(\mathcal{X},\mathcal{Y})\in\operatorname{\mathrm{sbp}}\Lambda be SM compatible, and let 𝒳𝒳\mathcal{X}^{\prime}\subseteq\mathcal{X}. Write μ𝒳(𝒳,𝒴)=(𝒳1,𝒴1)\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y})=(\mathcal{X}_{1},\mathcal{Y}_{1}).

  1. (1)

    If (𝒳1,𝒴1)(\mathcal{X}^{\prime}_{1},\mathcal{Y}^{\prime}_{1}) is an SM compatible semibrick pair such that 𝒳1𝒳1\mathcal{X}_{1}\subseteq\mathcal{X}^{\prime}_{1} and 𝒴1𝒴1\mathcal{Y}_{1}\subseteq\mathcal{Y}^{\prime}_{1}, then we have 𝒳μ𝒳(𝒳1,𝒴1)d\mathcal{X}\subseteq\mu_{\mathcal{X}^{\prime}}(\mathcal{X}^{\prime}_{1},\mathcal{Y}^{\prime}_{1})_{d} and 𝒴μ𝒳(𝒳1,𝒴1)u\mathcal{Y}\subseteq\mu_{\mathcal{X}^{\prime}}(\mathcal{X}^{\prime}_{1},\mathcal{Y}^{\prime}_{1})_{u}.

  2. (2)

    (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda if and only if (𝒳1,𝒴1)2-𝗌𝗆𝖼Λ(\mathcal{X}_{1},\mathcal{Y}_{1})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda.

  3. (3)

    (𝒳,𝒴)(\mathcal{X},\mathcal{Y}) is completable if and only if (𝒳1,𝒴1)(\mathcal{X}_{1},\mathcal{Y}_{1}) is completable.

  4. (4)

    If (𝒳,𝒴)2-𝗌𝗆𝖼Λ(\mathcal{X},\mathcal{Y})\in\operatorname{2\textnormal{-}\mathsf{smc}}\Lambda, then 𝖳(μ𝒳(𝒳,𝒴))=𝖳(𝒳,𝒴)𝒳\mathsf{T}(\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y}))=\mathsf{T}(\mathcal{X},\mathcal{Y})\cap\prescript{\perp}{}{\mathcal{X}^{\prime}}.

Proof.

1 First note that 𝒳𝒴1\mathcal{X}^{\prime}\subseteq\mathcal{Y}^{\prime}_{1} and therefore that 𝒳μ𝒳(𝒳1,𝒴1)d\mathcal{X}^{\prime}\subseteq\mu_{\mathcal{X}^{\prime}}(\mathcal{X}^{\prime}_{1},\mathcal{Y}^{\prime}_{1})_{d} by construction.

Now let X𝒳𝒳X\in\mathcal{X}\setminus\mathcal{X}^{\prime} so that EX,𝒳𝒴1E_{X,\mathcal{X}^{\prime}}\in\mathcal{Y}_{1}^{\prime}. Then Lemma A.1 tells us that the map XXEX,𝒳X^{\prime}_{X}\hookrightarrow E_{X,\mathcal{X}^{\prime}} is a minimal right 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})-approximation of EX,𝒳E_{X,\mathcal{X}^{\prime}} that is injective and has cokernel XX. It follows that Xμ𝒳(𝒳1,𝒴1)dX\in\mu_{\mathcal{X}^{\prime}}(\mathcal{X}^{\prime}_{1},\mathcal{Y}^{\prime}_{1})_{d}.

It remains to consider Y𝒴Y\in\mathcal{Y}. Suppose first that gY,𝒳g_{Y,\mathcal{X}^{\prime}} is injective. For any Z𝖥𝗂𝗅𝗍(𝒳){Z\in\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})}, the quotient map q:XYcoker(gY,𝒳)q\colon X^{\prime}_{Y}\rightarrow\mathrm{coker}(g_{Y,\mathcal{X}^{\prime}}) induces a bijection HomΛ(Z,XY2)HomΛ(Z,coker(gY,𝒳))\mathrm{Hom}_{\Lambda}(Z,X^{\prime}_{Y_{2}})\rightarrow\mathrm{Hom}_{\Lambda}(Z,\mathrm{coker}(g_{Y,\mathcal{X}^{\prime}})). Thus, qq is a minimal right 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})-approximation of coker(gY,𝒳)\mathrm{coker}(g_{Y,\mathcal{X}^{\prime}}) that is surjective and has kernel YY. We conclude that Yμ𝒳(𝒳1,𝒴1)uY\in\mu_{\mathcal{X}^{\prime}}(\mathcal{X}^{\prime}_{1},\mathcal{Y}^{\prime}_{1})_{u}.

It remains to consider the case when gY,𝒳g_{Y,\mathcal{X}^{\prime}} is surjective. Let Y=ker(gY,𝒳)Y^{\prime}=\mathrm{ker}(g_{Y,\mathcal{X}^{\prime}}). We consider the commutative diagram

Y{Y^{\prime}}Y{Y}XY{X^{\prime}_{Y}}Y{Y^{\prime}}I𝒴1,Y{I_{\mathcal{Y}^{\prime}_{1},Y^{\prime}}}Σ𝒴1Y,{\Sigma_{\mathcal{Y}^{\prime}_{1}}Y^{\prime},}gY,𝒳\scriptstyle{g_{Y,\mathcal{X}^{\prime}}}γ\scriptstyle{\gamma}

where the bottom row consists of the injective envelope and first cosyzygy of YY^{\prime} in 𝖥𝗂𝗅𝗍(𝒴1)\operatorname{\mathsf{Filt}}(\mathcal{Y}^{\prime}_{1}). The vertical maps are induced by the fact that the top row is an element of ExtΛ1(XY,Y)\mathrm{Ext}^{1}_{\Lambda}(X^{\prime}_{Y},Y^{\prime}), so in particular, the right square is a pullback. Now for Z𝖥𝗂𝗅𝗍(𝒳)Z\in\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime}), we have an exact sequence

0=HomΛ(Z,Y)HomΛ(Z,XY)ExtΛ1(Z,Y)ExtΛ1(Z,Y)=0,0=\mathrm{Hom}_{\Lambda}(Z,Y)\rightarrow\mathrm{Hom}_{\Lambda}(Z,X^{\prime}_{Y})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,Y^{\prime})\rightarrow\mathrm{Ext}^{1}_{\Lambda}(Z,Y)=0,

where the first and last terms are 0 since Y𝒳0,1Y\in\mathcal{X}^{\perp_{0,1}}. Moreover, we have a canonical isomorphism ExtΛ1(Z,Y)HomΛ(Z,Σ𝒴1Y)\mathrm{Ext}^{1}_{\Lambda}(Z,Y^{\prime})\cong\mathrm{Hom}_{\Lambda}(Z,\Sigma_{\mathcal{Y}^{\prime}_{1}}Y^{\prime}) since 𝖥𝗂𝗅𝗍(𝒳)𝖥𝗂𝗅𝗍(𝒴)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})\subseteq\operatorname{\mathsf{Filt}}(\mathcal{Y}^{\prime}) is a wide subcategory of 𝗆𝗈𝖽Λ\operatorname{\mathsf{mod}}\Lambda. We conclude that γ\gamma induces a bijection HomΛ(Z,XY)HomΛ(Z,Σ𝒴1Y)\mathrm{Hom}_{\Lambda}(Z,X^{\prime}_{Y})\cong\mathrm{Hom}_{\Lambda}(Z,\Sigma_{\mathcal{Y}^{\prime}_{1}}Y^{\prime}), so it is a minimal right 𝖥𝗂𝗅𝗍(𝒳)\operatorname{\mathsf{Filt}}(\mathcal{X}^{\prime})-approximation. Thus, the top row of our diagram coincides with the short exact sequence η𝒴1,Y\eta_{\mathcal{Y}^{\prime}_{1},Y^{\prime}} described above Definition 4.4. We conclude that Yμ𝒳(𝒳1,𝒴1)uY\in\mu_{\mathcal{X}^{\prime}}(\mathcal{X}^{\prime}_{1},\mathcal{Y}^{\prime}_{1})_{u}.

2 Let X𝒳𝒳X\in\mathcal{X}\setminus\mathcal{X}^{\prime}. Then the short exact sequence ηX,𝒳\eta_{X,\mathcal{X}^{\prime}} corresponds to a triangle

X[1]XXEX,𝒳XX[-1]\rightarrow X^{\prime}_{X}\rightarrow E_{X,\mathcal{X}^{\prime}}\rightarrow X

in 𝒟b(𝗆𝗈𝖽Λ)\operatorname{\mathcal{D}^{b}}(\operatorname{\mathsf{mod}}\Lambda). Thus, any triangulated subcategory that is closed under direct summands and contains (some shift of) 𝒳\mathcal{X}^{\prime} will contain XX if and only if it contains EX,𝒳E_{X,\mathcal{X}^{\prime}}. Similarly, for Y𝒴Y\in\mathcal{Y}, we have a triangle

YgY,𝒳XYZY[1],Y\xrightarrow{g_{Y,\mathcal{X}^{\prime}}}X^{\prime}_{Y}\rightarrow Z\rightarrow Y[1],

where either Z=coker(gY,𝒳)Z=\mathrm{coker}(g_{Y,\mathcal{X}^{\prime}}) or Z=ker(gY,𝒳)[1]Z=\mathrm{ker}(g_{Y,\mathcal{X}^{\prime}})[1]. Thus again, any triangulated subcategory that is closed under direct summands and contains 𝒳\mathcal{X}^{\prime} will contain YY if and only if it contains ZZ. This proves the result.

3 This is an immediate consequence of 1 and 3.

4 Recall that 𝖳(μ𝒳(𝒳,𝒴))=𝒴1\mathsf{T}(\mu_{\mathcal{X}^{\prime}}(\mathcal{X},\mathcal{Y}))=\prescript{\perp}{}{\mathcal{Y}_{1}} and 𝖳(𝒳,𝒴)=𝒴\mathsf{T}(\mathcal{X},\mathcal{Y})=\prescript{\perp}{}{\mathcal{Y}} by Theorem 3.7.

The fact that Y1𝒳\prescript{\perp}{}{Y_{1}}\subseteq\prescript{\perp}{}{\mathcal{X}^{\prime}} follows from the definitions. Hence, let M𝒳M\in\prescript{\perp}{}{\mathcal{X}^{\prime}}. Then for all Y𝒴Y\in\mathcal{Y} such that gY,𝒳g_{Y,\mathcal{X}^{\prime}} is injective, we necessarily have that MYM\in\prescript{\perp}{}{Y}. It therefore suffices to show that for Y𝒴Y\in\mathcal{Y} such that gY,𝒳g_{Y,\mathcal{X}^{\prime}} is surjective, we have that MYM\in\prescript{\perp}{}{Y} if and only if Mker(gY,𝒳)M\in\prescript{\perp}{}{\mathrm{ker}(g_{Y,\mathcal{X}^{\prime}})}. The inclusion Yker(gY,𝒳)\prescript{\perp}{}{Y}\subseteq\prescript{\perp}{}{\mathrm{ker}(g_{Y,\mathcal{X}^{\prime}})} holds in general, so suppose that Mker(gY,𝒳)M\in\prescript{\perp}{}{\mathrm{ker}(g_{Y,\mathcal{X}^{\prime}})}. Then the fact that M𝒳M\in\prescript{\perp}{}{\mathcal{X}^{\prime}} implies that MYM\in\prescript{\perp}{}{Y} as well. This concludes the proof. ∎

Acknowledgments

Colin Defant was supported by the National Science Foundation under Award No. 2201907 and by a Benjamin Peirce Fellowship at Harvard University. Eric Hanson was partially supported by Canada Research Chairs CRC-2021-00120 and by NSERC Discovery Grants RGPIN-2022-03960 and RGPIN/04465-2019. A portion of this work was completed while Eric Hanson was a postdoctoral fellow at l’Université du Québec à Montréal and l’Université de Sherbrooke. We are grateful to Nathan Williams for helpful conversations.

References

  • [AM17] A. Aihara and Y. Mizuno, Classifying tilting complexes over preprojective algebras of Dynkin type, Algebra Number Theory 11 (2017), no. 6, 1287–1315.
  • [AN09] S. Al-Nofayee, Simple objects in the heart of a tt-structure, J. Pure Appl. Algebra 213 (2009), no. 1, 54–59.
  • [AHI+] T. Aoki, A. Higashitani, O. Iyama, R. Kase, and Y. Mizuno, Fans and polytopes in tilting theory I: Foundations, arXiv:2203.15213.
  • [Asa20] S. Asai, Semibricks, Int. Math. Res. Not. IMRN 2020 (2020), no. 16, 4993–5054.
  • [AP22] S. Asai and C. Pfeifer, Wide subcategories and lattices of torsion classes, Algebr. Represent. Theory 25 (2022), 1611–1629.
  • [ABB+19] A. Asinowski, C. Banderier, S. Biley, B. Hackl, and S. Linusson, Pop-stack sorting and its image: permutations with overlapping runs, Acta Math. Univ. Comenianae 88 (2019), 395–402.
  • [ASS06] I. Assem, D. Simson, and A. Skowroński, Elements of the representation theory of associative algebras, volume 1: techniques of representation theory, London Math. Soc. Stud. Texts, vol. 65, Cambridge University Press, Cambridge, 2006.
  • [ARS95] M. Auslander, I. Reiten, and S. O. Smalø, Representation theory of Artin algebras, Cambridge Stud. Adv. Math., vol. 36, Cambridge University Press, Cambridge, 1995.
  • [Bar19] E. Barnard, The canonical join complex, Electron. J. Combin. 26 (2019).
  • [BCZ19] E. Barnard, A. Carroll, and S. Zhu, Minimal inclusions of torsion classes, Algebr. Comb. 2 (2019), no. 5, 879–901.
  • [BH] E. Barnard and E. J. Hanson, Exceptional sequences in semidistributive lattices and the poset topology of wide subcategories, arXiv:2209.11734.
  • [BH22] E. Barnard and E. J. Hanson, Pairwise compatibility for 2-simple minded collections II: preprojective algebras and full rank semibirck pairs, Ann. Comb. 26 (2022), 803–855.
  • [BR18] E. Barnard and N. Reading, Coxeter-biCatalan combinatorics, J. Algebraic Combin. 47 (2018), no. 2, 241–300.
  • [BTZ21] E. Barnard, G. Todorov, and S. Zhu, Dynamical combinatorics and torsion classes, J. Pure Appl. Algebra 225 (2021).
  • [BB05] A. Björner and F. Brenti, Combinatorics of coxeter groups, Graduate Texts in Mathematics, vol. 231, Springer, Berlin, Heidelberg, 2005.
  • [Bou02] N. Bourbaki, Lie groups and Lie algebras Chapters 4-6, Springer–Verlag, Berlin, Copyright Information Springer-Verlag Berlin, Heidelberg, 2002.
  • [BY13] T. Brüstle and D. Yang, Oriented exchange graphs, In (D. J. Benson, H. Krause, and A. Skowroński, eds.) Advances in Representation Theory of Algebras, EMS Series of Congress Reports, vol. 9, 2013.
  • [CS] Y. Choi and N. Sun, The image of the pop operator on various lattices, To appear in Adv. Appl. Math.
  • [CG19] A. Claesson and B. Á. Guðmundsson, Enumerating permutations sortable by kk passes through a pop-stack, Adv. Appl. Math. 108 (2019), 79–96.
  • [CGP21] A. Claesson, B. Á Guðmundsson, and J. Pantone, Counting pop-stacked permutations in polynomial time, Experiment. Math. (2021).
  • [Def22a] C. Defant, Meeting covered elements in ν\nu-tamari lattices, Adv. Appl. Math. 134 (2022).
  • [Def22b] C. Defant, Pop-stack-sorting for Coxeter groups, Comb. Theory 2 (2022).
  • [DKW] C. Defant, N. Kravitz, and N. Williams, The Ungar games, arXiv:2302.06552.
  • [DL] C. Defant and R. Li, Ungarian Markov chains, arXiv:2301.08206.
  • [DW23] C. Defant and N. Williams, Semidistrim lattices, Forum Math. Sigma 11 (2023), no. 50, 1–35.
  • [DIJ17] L. Demonet, O. Iyama, and G. Jasso, τ\tau-tilting finite algebras, bricks, and gg-vectors, Int. Math. Res. Not. IMRN 2019 (2017), no. 3, 852–892.
  • [DIR+23] L. Demonet, O. Iyama, N. Reading, I. Reiten, and H. Thomas, Lattice theory of torsion classes: beyond τ\tau-tilting theory, Trans. Amer. Math. Soc. Ser. B 10 (2023), 542–612.
  • [Dic66] S. E. Dickson, A torsion theory for abelian categories, Trans. Amer. Math. Soc. 121 (1966), no. 1, 223–235.
  • [DR75] V. Dlab and C. M. Ringel, On algebras of finite representation type, J. Algebra 33 (1975), 306–394.
  • [DR76] V. Dlab and C. M. Ringel, Indecomposable representations of graphs and algebras, Mem. Am. Math. Soc. 6 (1976).
  • [Eno21] H. Enomoto, Bruhat inversions in Weyl groups and torsion-free classes over preprojective algebras, Comm. Algebra 49 (2021), no. 5, 2156–2189.
  • [Eno22] H. Enomoto, Rigid modules and ICE-closed subcategories in quiver representations, J. Algebra 594 (2022), 364–388.
  • [Eno23] H. Enomoto, From the lattice of torsion classes to the posets of wide subcategories and ICE-closed subcategories, Algebr. Represent. Theory (2023).
  • [ES22] H. Enomoto and A. Sakai, ICE-closed subcategories and wide τ\tau-tilting modules, Math. Z. 300 (2022), 541–577.
  • [FJN95] R. Freese, J. Jezek, and J. Nation, Free lattices, Mathematical Surveys and Monographs, vol. 42, American Mathematical Society, 1995.
  • [GIMO19] A. Garver, K. Igusa, J. P. Matherne, and J. Ostroff, Combinatorics of exceptional sequences in type A, Electron. J. Combin. 26 (2019), no. 1.
  • [GM19] A. Garver and T. McConville, Lattice properties of oriented exchange graphs and torsion classes, Algebr. Represent. Theory 22 (2019), 43–78.
  • [GL91] W. Geigle and H. Lenzing, Perpendicular categories with applications to representations and sheaves, J. Algebra 144 (1991), no. 2, 273–343.
  • [Hana] E. J. Hanson, A facial order for torsion classes, arXiv:2305.06031.
  • [Hanb] E. J. Hanson, τ\tau-exceptional sequences and the shard intersection order in type A, arXiv:2303.11517.
  • [HI21a] E. J. Hanson and K. Igusa, Pairwise compatibility for 2-simple minded collections, J. Pure Appl. Algebra 225 (2021), no. 6.
  • [HI21b] E. J. Hanson and K. Igusa, τ\tau-cluster morphism categories and picture groups, Comm. Alg. 49 (2021), no. 10, 4376–4415.
  • [HW] W. Hara and M. Wemyss, Spherical objects in dimension two and three, arXiv:2206.11552.
  • [HY23] E. J. Hanson and X. You, Morphisms and extensions between bricks over preprojective algebras of type A, J. Algebra (2023).
  • [Hon22] L. Hong, The pop-stack-sorting operator on Tamari lattices, Adv. Appl. Math. 139 (2022).
  • [ITW] K. Igusa, G. Todorov, and J. Weyman, Picture groups of finite type and cohomology in type An{A}_{n}, arXiv:1609.02636.
  • [IT09] C. Ingalls and H. Thomas, Noncrossing partitions and representations of quivers, Compos. Math. 145 (2009), no. 6, 1533–1562.
  • [IRRT18] O. Iyama, N. Reading, I. Reiten, and H. Thomas, Lattice structure of Weyl groups via representation theory of preprojective algebras, Compos. Math. 154 (2018), no. 6, 1269–1305.
  • [IRTT15] O. Iyama, I. Reiten, H. Thomas, and G. Todorov, Lattice structure of torsion classes for path algebras, Bull. Lond. Math. Soc. 47 (2015), no. 4, 639–650.
  • [Jas15] G. Jasso, Reduction of τ\tau-tilting modules and torsion pairs, Int. Math. Res. Not. IMRN 2015 (2015), no. 16, 7190–7237.
  • [KQ15] A. King and Y. Qiu, Exchange graphs and Ext quivers, Adv. Math. 285 (2015), 1106–1154.
  • [KY14] S. Koenig and D. Yang, Silting objects, simple-minded collections, tt-structures and co-tt-structures for finite-dimensional algebras, Doc. Math. 19 (2014), 403–438.
  • [KS] M. Kontsevich and Y. Soibelman, Stability structures, motivic Donaldson-Thomas invariants and cluster transformations, arXiv:0811.2435.
  • [Kül17] J. Külshammer, Pro-species of algebras I: Basic properties, Algebr. Represent. Theory 20 (2017), no. 5, 1215–1238.
  • [LW] Y. Luo and J. Wei, Boolean lattices of torsion classes over finite-dimensional algebras, arXiv:2303.15802.
  • [MŠ17] F. Marks and J. Šťovíček, Torsion classes, wide subcategories, and localisations, Bull. Lond. Math. Soc. 49 (2017), no. 3.
  • [Miz14] Y. Mizuno, Classifying τ\tau-tilting modules over preprojective algebras of Dynkin type, Math. Z. 277 (2014), no. 3-4, 665–690.
  • [Miz22] Y. Mizuno, Arc diagrams and 2-term simple-minded collections of preprojective algebras of type A, J. Algebra 595 (2022), 444–478.
  • [MT20] Y. Mizuno and H. Thomas, Torsion pairs for quivers and the Weyl groups, Selecta Math. 3 (2020), no. 46.
  • [Müh19] H. Mühle, The core label order of a congruence-uniform lattice, Algebra Universalis 80 (2019).
  • [PS19] L. Pudwell and R. Smith, Two-stack-sorting with pop stacks, Australas. J. Combin. 74 (2019), 179–195.
  • [Rea04] N. Reading, Lattice congruences of the weak order, Order 21 (2004), 315–344.
  • [Rea06] N. Reading, Cambrian lattices, Adv. Math. 205 (2006), 313–353.
  • [Rea07] N. Reading, Clusters, Coxeter-sortable elements and noncrossing partitions, Trans. Amer. Math. Soc. 359 (2007), no. 12, 5931–5958.
  • [Rea11] N. Reading, Noncrossing partitions and the shard intersection order, J. Algebraic Combin. 33 (2011), 483–530.
  • [Rea15] N. Reading, Noncrossing arc diagrams and canonical join representations, SIAM J. Discrete Math. 29 (2015).
  • [Rea16] N. Reading, Lattice theory of poset of regions, In (G. Grätzer and F. Wehrung, eds.) Lattice Theory: Special Topics and Applications, vol. 2, Birkhäuser, 2016.
  • [RS11] N. Reading and D. E Speyer, Sortable elements in infinite Coxeter groups, Trans. Amer. Math. Soc. 363 (2011), no. 2, 699–761.
  • [RST21] N. Reading, D. E Speyer, and H. Thomas, The fundamental theorem of finite semidistributive lattices, Selecta Math. 27 (2021), no. 59.
  • [Ric02] J. Rickard, Equivalences of derived categories for symmetric algebras, J. Algebra 257 (2002), 460–481.
  • [Rin76] C. M. Ringel, Representations of KK-species and bimodules, J. Algebra 41 (1976), 269–302.
  • [The23] The Sage Developers, Sagemath, the Sage Mathematics Software System (Version 9.2), 2023, https://www.sagemath.org.
  • [Sak] A. Sakai, Classifying tt-structures via ICE-closed subcategories and a lattice of torsion classes, arXiv:2307.11347.
  • [Sch14] R. Schiffler, Quiver representations, CMS Books in Mathematics/Ouvrages de Mathématiques de la SMC, Springer, Cham, 2014.
  • [STW] C. Stump, H. Thomas, and N. Williams, Cataland: Why the Fuß, To appear in Mem. Amer. Math. Soc.
  • [Tho06] H. Thomas, An analogue of distributivity for ungraded lattices, Order 23 (2006), 249–269.
  • [Tho21] H. Thomas, An introduction to the lattice of torsion classes, Bull. Iranian Math. Soc. (2021).
  • [TW19a] H. Thomas and N. Williams, Independence posets, J. Comb. 10 (2019), no. 3, 545–578.
  • [TW19b] H. Thomas and N. Williams, Rowmotion in slow motion, Proc. Lon. Math. Soc. 119 (2019), no. 5, 1149–1178.
  • [Ung82] P. Ungar, 2N2{N} noncollinear points determine at least 2N2{N} directions, J. Combin. Theory Ser. A 33 (1982), 343–347.