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

Quivers with potentials associated to triangulations of closed surfaces with at most two punctures

Jan Geuenich Jan Geuenich
Fakultät für Mathematik
Universität Bielefeld
Germany
jgeuenich@math.uni-bielefeld.de
Daniel Labardini-Fragoso Daniel Labardini-Fragoso
Instituto de Matemáticas
Universidad Nacional Autónoma de México
Mexico
labardini@im.unam.mx
 and  José Luis Miranda-Olvera José Luis Miranda-Olvera
Department of Mathematical Sciences
Carnegie Mellon University
USA
joseluismiranda@cmu.edu
Abstract.

We tackle the classification problem of non-degenerate potentials for quivers arising from triangulations of surfaces in the cases left open by Geiss-Labardini-Schröer. Namely, for once-punctured closed surfaces of positive genus, we show that the quiver of any triangulation admits infinitely many non-degenerate potentials that are pairwise not weakly right-equivalent; we do so by showing that the potentials obtained by adding the 3-cycles coming from triangles and a fixed power of the cycle surrounding the puncture are well behaved under flips and QP-mutations. For twice-punctured closed surfaces of positive genus, we prove that the quiver of any triangulation admits exactly one non-degenerate potential up to weak right-equivalence, thus confirming the veracity of a conjecture of the aforementioned authors.

Key words and phrases:
Surface, marked points, punctures, triangulation, flip, quiver, potential, mutation, non-degenerate potential
2010 Mathematics Subject Classification:
Primary 16P10, 16G20; Secondary 13F60, 57N05, 05E99

1. Introduction

Albeit technical in nature, the problem of classifying all non-degenerate potentials on a given 2-acyclic quiver is relevant in different interesting, seemingly unrelated, contexts. In cluster algebra theory, having only one weak right-equivalence class means, very roughly speaking, that Derksen-Weyman-Zelevinsky’s representation-theoretic approach to the corresponding cluster algebra can be performed in essentially only one way.

The classification problem of non-degenerate potentials plays a role also in algebraic geometry and in symplectic geometry (more precisely, in the subjects of Bridgeland stability conditions and Fukaya categories). In [1, Theorem 9.9], the uniqueness of non-degenerate potentials on the quivers arising from positive genus closed surfaces with at least three punctures is used by Tom Bridgeland and Ivan Smith to prove that there is a short exact sequence

1\textstyle{1\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒮𝓅𝒽(𝒟(Σ,𝕄))\textstyle{\mathpzc{Sph}_{\triangle}(\mathcal{D}(\Sigma,\mathbb{M}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒜𝓊𝓉(𝒟(Σ,𝕄))\textstyle{\mathpzc{Aut}_{\triangle}(\mathcal{D}(\Sigma,\mathbb{M}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MCG±(Σ,𝕄)\textstyle{\operatorname{MCG}^{\pm}(\Sigma,\mathbb{M})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1,\textstyle{1,}

where 𝒟(Σ,𝕄)\mathcal{D}(\Sigma,\mathbb{M}) is the 3-Calabi-Yau triangulated category associated to (Σ,𝕄)(\Sigma,\mathbb{M}), defined as the full subcategory that the dg-modules with finite-dimensional cohomology determine inside the derived category of the Ginzburg dg-algebra of the quiver with potential of any111That 𝒟(Σ,𝕄)\mathcal{D}(\Sigma,\mathbb{M}) is independent of the tagged triangulation used follows after combining results of Keller-Yang [6] and Labardini [9], see [10, Section 5]. tagged triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}), the group 𝒜𝓊𝓉(𝒟(Σ,𝕄))\mathpzc{Aut}_{\triangle}(\mathcal{D}(\Sigma,\mathbb{M})) is the quotient of the group of auto-equivalences of 𝒟(Σ,𝕄)\mathcal{D}(\Sigma,\mathbb{M}) that preserve the distinguished connected component Tilt(𝒟(Σ,𝕄))\operatorname{Tilt}_{\triangle}(\mathcal{D}(\Sigma,\mathbb{M})) by the subgroup of auto-equivalences that act trivially on Tilt(𝒟(Σ,𝕄))\operatorname{Tilt}_{\triangle}(\mathcal{D}(\Sigma,\mathbb{M})), 𝒮𝓅𝒽(𝒟(Σ,𝕄))\mathpzc{Sph}_{\triangle}(\mathcal{D}(\Sigma,\mathbb{M})) is the subgroup of 𝒜𝓊𝓉(𝒟(Σ,𝕄))\mathpzc{Aut}_{\triangle}(\mathcal{D}(\Sigma,\mathbb{M})) generated by (the quotient images of) the twist functors at the simple objects of an heart 𝒜Tilt(𝒟(Σ,𝕄))\mathcal{A}\in\operatorname{Tilt}_{\triangle}(\mathcal{D}(\Sigma,\mathbb{M})), and MCG±(Σ,𝕄)=MCG(Σ,𝕄)2𝕄\operatorname{MCG}^{\pm}(\Sigma,\mathbb{M})=\operatorname{MCG}(\Sigma,\mathbb{M})\ltimes\mathbb{Z}_{2}^{\mathbb{M}} is the signed mapping class group.

In [15, Theorem 1.1], Ivan Smith shows that if (Σ,𝕄)(\Sigma,\mathbb{M}) is a positive genus closed surface with at least three punctures (i.e., |𝕄|3|\mathbb{M}|\geq 3), then there is a linear fully faithful embedding of the 3-Calabi-Yau triangulated category 𝒟(Σ,𝕄)\mathcal{D}(\Sigma,\mathbb{M}) into a Fukaya category of a 3-fold that fibers over Σ\Sigma (with poles of a quadratic differential removed from Σ\Sigma). He explains that the reason behind the hypothesis |𝕄|3|\mathbb{M}|\geq 3 in [15, Theorem 1.1] arises from the fact that for positive genus closed surfaces with at least three punctures, the quiver of any triangulation has exactly one non-degenerate potential up to weak right-equivalence (a fact shown by Geiss-Labardini-Schröer [4]). See [15, Sections 1.3 and 2.2].

Together with results from his work [1] with Bridgeland, the embeddings from the previous paragraph allow Smith to obtain non-trivial computations of spaces of stability conditions on Fukaya categories of symplectic six-manifolds.

In this paper we prove the following:

Theorem 1.1.
  1. (1)

    For once-punctured closed surfaces of positive genus, the quiver of any triangulation admits infinitely many non-degenerate potentials that are pairwise not weakly right-equivalent, provided the underlying field has characteristic zero.

  2. (2)

    For twice-punctured closed surfaces of positive genus, the quiver of any triangulation admits exactly one non-degenerate potential up to weak right-equivalence, provided the underlying field is algebraically closed.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a once-punctured closed surface, nn a positive integer and xKx\in K any scalar. For a triangulation τ\tau of (Σ,𝕄)(\Sigma,\mathbb{M}) let S(τ,x,n)S(\tau,x,n) be the potential obtained by adding the 3-cycles of Q(τ)Q(\tau) arising from triangles of τ\tau and the xx-multiple of the nthn^{\operatorname{th}} power of the cycle of Q(τ)Q(\tau) that runs around the puncture of (Σ,𝕄)(\Sigma,\mathbb{M}). The following result of independent interest plays a central role in our proof of part (1) of Theorem 1.1:

Theorem 1.2.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a once-punctured closed surface. If τ\tau and σ\sigma are triangulations of (Σ,𝕄)(\Sigma,\mathbb{M}) related by the flip of an arc kτk\in\tau, then the quivers with potential (Q(τ),S(τ,x,n))(Q(\tau),S(\tau,x,n)) and (Q(σ),S(σ,x,n))(Q(\sigma),S(\sigma,x,n)) are related by the mutation of quivers with potential μk\mu_{k}.

That the quivers associated to triangulations of once-punctured closed surfaces of positive genus admit more than one weak-right-equivalence class of non-degenerate potentials has always been expected, since the works [2] of Derksen-Weyman-Zelevinsky and [7, 8] of Labardini exhibit non-degenerate potentials for the Markov quiver222The Markov quiver arises as the quiver associated to any triangulation of the once-punctured torus. that are not weakly right-equivalent.

In [4, Theorem 8.4], Geiss-Labardini-Schröer proved that every quiver associated to some triangulation of a positive-genus closed surface with at least three punctures admits exactly one weak-right-equivalence class of non-degenerate potentials, and conjectured that the same result holds in the case of two punctures. The reason why their proof fails for twice-punctured closed surface is that these do not admit triangulations all of whose arcs connect distinct punctures. The fact that such triangulations do exist for closed surfaces with at least three punctures plays an essential role in the proof of [4, Theorem 8.4].

The structure of the paper is straightforward: in Section 2 we prove a few facts (some of them quite technical) about the form of cycles and non-degenerate potentials for quivers arising from combinatorially nice triangulations of surfaces with empty boundary (see conditions (2.1) and (2.2)). Section 3 is devoted to proving part (1) of Theorem 1.1, whereas Section 4 is devoted to showing part (2).

2. Preliminaries

Let KK be any field. For a quiver QQ, the vertex span is the KK-algebra RR defined as the KK-vector space with basis {ej|jQ0}\{e_{j}\ |\ j\in Q_{0}\}, with multiplication defined as the KK-bilinear extension of the rule

eiej:=δi,jejfor all i,jQ0,e_{i}e_{j}:=\delta_{i,j}e_{j}\hskip 28.45274pt\text{for all $i,j\in Q_{0}$},

where δi,jK\delta_{i,j}\in K is the Kronecker delta of ii and jj. Thus, RR is (a KK-algebra isomorphic to) KQ0K^{Q_{0}} with both sum and multiplication defined componentwise. The complete path algebra of QQ is the KK-vector space

KQ:=0A(),K\langle\langle Q\rangle\rangle:=\prod_{\ell\in\mathbb{Z}_{\geq 0}}A^{(\ell)},

where A(0):=RA^{(0)}:=R, and for >0\ell>0, A()A^{(\ell)} is the KK-vector space with basis all the paths of length \ell on QQ. The multiplication of KQK\langle\langle Q\rangle\rangle is defined in terms of the concatenation of paths.

The vertex span RR is obviously a subring of KQK\langle\langle Q\rangle\rangle (actually, a KK-subalgebra), but it is often not a central subring. Despite this, any ring automorphism φ:KQKQ\varphi:K\langle\langle Q\rangle\rangle\rightarrow K\langle\langle Q\rangle\rangle such that φ|R=11R\varphi|_{R}=11_{R} will be said to be an RR-algebra automorphism of KQK\langle\langle Q\rangle\rangle.

Definition 2.1.

Let QQ be a quiver and S,WKQS,W\in K\langle\langle Q\rangle\rangle be potentials on QQ. We will say that:

  1. (1)

    two cycles a1aa_{1}\cdots a_{\ell} and b1bmb_{1}\cdots b_{m} on QQ are rotationally equivalent if a1a=b1bma_{1}\cdots a_{\ell}=b_{1}\cdots b_{m} or a1a=bkbmb1bk1a_{1}\cdots a_{\ell}=b_{k}\cdots b_{m}b_{1}\cdots b_{k-1} for some k{2,,m}k\in\{2,\ldots,m\};

  2. (2)

    SS and WW are rotationally disjoint if no cycle appearing in SS is rotationally equivalent to a cycle appearing in WW;

  3. (3)

    SS and WW are cyclically equivalent if, with respect to the 𝔪\mathfrak{m}-adic topology of KQK\langle\langle Q\rangle\rangle, the element SWS-W belongs to the topological closure of the vector subspace of KQK\langle\langle Q\rangle\rangle spanned by all elements of the form a1aa2aa1a_{1}\cdots a_{\ell}-a_{2}\cdots a_{\ell}a_{1} with a1aa_{1}\cdots a_{\ell} running through the set of all cycles on QQ; notation: ScycWS\sim_{\operatorname{cyc}}W;

  4. (4)

    SS and WW are right-equivalent if there exists a right equivalence from SS to WW, i. e., an RR-algebra automorphism φ:KQKQ\varphi:K\langle\langle Q\rangle\rangle\rightarrow K\langle\langle Q\rangle\rangle that acts as the identity on the set of idempotents {ej|jQ0}\{e_{j}\ |\ j\in Q_{0}\} and satisfies φ(S)cycW\varphi(S)\sim_{\operatorname{cyc}}W; notation: Sr.e.WS\sim_{\operatorname{r.e.}}W;

  5. (5)

    SS and WW are weakly right-equivalent if SS and λW\lambda W are right-equivalent for some non-zero scalar λK\lambda\in K.

Throughout the paper, (Σ,𝕄)(\Sigma,\mathbb{M}) will be a punctured closed surface of positive genus. That is, Σ\Sigma will be a compact, connected, oriented two-dimensional real differentiable manifold with positive genus and empty boundary, and 𝕄\mathbb{M} will be a non-empty finite subset of Σ\Sigma.

It is very easy to show that there exists at least one triangulation τ\tau of (Σ,𝕄)(\Sigma,\mathbb{M}) such that

(2.1) Every puncture has valency at least 44 with respect to τ\tau;
(2.2) for any two arcs ii and jj of τ\tau, the quiver Q(τ)Q(\tau) has at most one arrow from jj to ii.

Throughout the paper, we will permanently suppose that τ\tau satisfies (2.1) and (2.2).

Following Ladkani [11] we define two maps f,g:Q(τ)1Q(τ)1f,g:Q(\tau)_{1}\rightarrow Q(\tau)_{1} as follows. Each triangle \triangle of τ\tau gives rise to an oriented 3-cycle αβγ\alpha_{\triangle}\beta_{\triangle}\gamma_{\triangle} on Q(τ)Q(\tau). We set f(α)=γf(\alpha_{\triangle})=\gamma_{\triangle}, f(β)=αf(\beta_{\triangle})=\alpha_{\triangle} and f(γ)=βf(\gamma_{\triangle})=\beta_{\triangle}. Now, given any arrow α\alpha of Q(τ)Q(\tau), the quiver Q(τ)Q(\tau) has exactly two arrows starting at the terminal vertex of α\alpha. One of these two arrows is f(α)f(\alpha). We define g(α)g(\alpha) to be the other arrow.

Note that the map ff (resp. gg) splits the arrow set of Q(τ)Q(\tau) into ff-orbits (resp. gg-orbits). The set of ff-orbits is in one-to-one correspondence with the set of triangles of τ\tau. All ff-orbits have exactly three elements. The set of gg-orbits is in one-to-one correspondence with the set of punctures of (Σ,𝕄)(\Sigma,\mathbb{M}). For every arrow α\alpha of Q(τ)Q(\tau), we denote by mαm_{\alpha} the size of the gg-orbit of α\alpha (mα4m_{\alpha}\geq 4 by (2.1)). Note that gmα1(α)gmα2(α)g(α)αg^{m_{\alpha}-1}(\alpha)g^{m_{\alpha}-2}(\alpha)\cdots g(\alpha)\alpha is a cycle surrounding the puncture pp corresponding to the gg-orbit of α\alpha, we denote this cycle as 𝒢(α)\mathcal{G}(\alpha) or 𝒢(p)\mathcal{G}(p). Whereas, for every arrow β\beta of Q(τ)Q(\tau) and any non-negative integer rr, we use the notation G(r,β)G(r,\beta) to denote the path gr1(β)gr2(β)g(β)βg^{r-1}(\beta)g^{r-2}(\beta)\cdots g(\beta)\beta. Similarly, we use the notation F(r,β)F(r,\beta) to denote the path fr1(β)fr2(β)f(β)βf^{r-1}(\beta)f^{r-2}(\beta)\cdots f(\beta)\beta.

Let 𝐱=(xp)p𝕄\mathbf{x}=(x_{p})_{p\in\mathbb{M}} be a choice of a non-zero scalar xpKx_{p}\in K for each puncture p𝕄p\in\mathbb{M}. For ideal triangulations which satisfy (2.1) and (2.2) the potential S(τ,𝐱)S(\tau,\mathbf{x}) defined by the second author [9] takes a simple form, namely,

S(τ,𝐱)=T(τ)+pxp𝒢(p),S(\tau,\mathbf{x})=T(\tau)+\sum_{p\in\mathbb{P}}x_{p}\mathcal{G}(p),

with T(τ)cycαΓ(f2(α)f(α)α)T(\tau)\sim_{\operatorname{cyc}}\sum_{\alpha\in\Gamma}(f^{2}(\alpha)f(\alpha)\alpha) for any fixed set Γ\Gamma containing exactly one arrow from each triangle of τ\tau.

Lemma 2.2 (Types of cycles).

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a punctured surface with empty boundary, and let τ\tau be a triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}) that satisfies (2.1) and (2.2). Then every cycle in Q(τ)Q(\tau) is rotationally equivalent to a cycle of one of the following types:

  • (ff-cycles)

    (f2(α)f(α)α)n(f^{2}(\alpha)f(\alpha)\alpha)^{n} for some n1n\geq 1;

  • (gg-cycles)

    (gmβ1(β)gmβ2(β)g(β)β)n(g^{m_{\beta}-1}(\beta)g^{m_{\beta}-2}(\beta)\cdots g(\beta)\beta)^{n} for some n1n\geq 1;

  • (fgfg-cycles)

    f2(a)f(a)λf^{2}(a)f(a)\lambda for some arrow aa and some path λ\lambda, such that λ=g1f(a)λ\lambda=g^{-1}f(a)\lambda^{\prime} with λ\lambda^{\prime} of positive length.

Proof.

Let ξ=α1αr\xi=\alpha_{1}\cdots\alpha_{r} be any cycle on Q(τ)Q(\tau). Denote αr+1=α1\alpha_{r+1}=\alpha_{1}, and notice that for every =1,,r\ell=1,\cdots,r, we have either α=f(α+1)\alpha_{\ell}=f(\alpha_{\ell+1}) or α=g(α+1)\alpha_{\ell}=g(\alpha_{\ell+1}). Let 𝐬ξ\mathbf{s}_{\xi} be the length-rr sequence of ffs and ggs that has an ff at the th\ell^{\operatorname{th}} place if α=f(α+1)\alpha_{\ell}=f(\alpha_{\ell+1}) and a gg otherwise.

If 𝐬ξ\mathbf{s}_{\xi} consists only of ffs, then ξ\xi is rotationally equivalent to (f2(α)f(α)α)n(f^{2}(\alpha)f(\alpha)\alpha)^{n} for some arrow α\alpha and some n1n\geq 1. Furthemore, if 𝐬ξ\mathbf{s}_{\xi} consists only of ggs, then ξ\xi is rotationally equivalent to (gmβ1(β)gmβ2(β)g(β)β)n(g^{m_{\beta}-1}(\beta)g^{m_{\beta}-2}(\beta)\cdots g(\beta)\beta)^{n} for some arrow β\beta and some n1n\geq 1. Therefore, if 𝐬ξ\mathbf{s}_{\xi} involves only ffs or only ggs, then ξ\xi is an ff-cycle or a gg-cycle.

Suppose that at least one ff and at least one gg appear in 𝐬ξ\mathbf{s}_{\xi}. Rotating ξ\xi if necessary, we can assume that 𝐬ξ\mathbf{s}_{\xi} starts with an ff followed by a gg, i.e., 𝐬ξ=(f,g,)\mathbf{s}_{\xi}=(f,g,\cdots). This means that if we set a:=f1(α2)a:=f^{-1}(\alpha_{2}), then α1=f2(a)\alpha_{1}=f^{2}(a), α2=f(a)\alpha_{2}=f(a) and α3=g1f(a)\alpha_{3}=g^{-1}f(a). By (2.2) aa is the only arrow in Q(τ)1Q(\tau)_{1} such that α1α2a\alpha_{1}\alpha_{2}a is a cycle. Since α3=g1f(a)a\alpha_{3}=g^{-1}f(a)\neq a, this implies ξ=f2(a)f(a)g1f(a)λ\xi=f^{2}(a)f(a)g^{-1}f(a)\lambda^{\prime} with λ\lambda^{\prime} of positive length. ∎

Remark 2.3.

As in the case of cycles, every path falls within exactly one of three types of paths: ff-paths, gg-paths, and fgfg-paths.

By Lemma 2.2, up to cyclical equivalence we can write every potential SS in Q(τ)Q(\tau) as S=Sf+Sg+SfgS=S_{f}+S_{g}+S_{fg}, where

Sf\displaystyle S_{f} =\displaystyle= n=1z,n(f2(α)f(α)α)n,\displaystyle\sum_{\triangle}\sum_{n=1}^{\infty}z_{\triangle,n}(f^{2}(\alpha_{\triangle})f(\alpha_{\triangle})\alpha_{\triangle})^{n},
Sg\displaystyle S_{g} =\displaystyle= pn=1νp,n(𝒢(p))n,\displaystyle\sum_{p\in\mathbb{P}}\sum_{n=1}^{\infty}\nu_{p,n}(\mathcal{G}(p))^{n},
Sfg\displaystyle S_{fg} =\displaystyle= aQ(τ)1f2(a)f(a)ωa,\displaystyle\sum_{a\in Q(\tau)_{1}}f^{2}(a)f(a)\omega_{a},

with each z,n,νp,nKz_{\triangle,n},\nu_{p,n}\in K, and ωa\omega_{a} a possibly infinite linear combination of paths of the form g1f(a)λg^{-1}f(a)\lambda^{\prime} for each aQ(τ)a\in Q(\tau).

Lemma 2.4.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a punctured surface with empty boundary, and let τ\tau be a triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}) that satisfies (2.1) and (2.2). Every non-degenerate potential SS on Q(τ)Q(\tau) is right-equivalent to a potential of the form T(τ)+UT(\tau)+U for some UU rotationally disjoint from T(τ)T(\tau).

Proof.

By (2.2) the hypotheses of [4, Corollary 2.5] are satisfied, so if SS is a non-degenerate potential, then every ff-cycle f2(α)f(α)αf^{2}(\alpha)f(\alpha)\alpha appears in SS. So,

S\displaystyle S cycz,1f2(α)f(α)α+U,\displaystyle\sim_{\operatorname{cyc}}\sum_{\triangle}z_{\triangle,1}f^{2}(\alpha_{\triangle})f(\alpha_{\triangle})\alpha_{\triangle}+U^{\prime},

with all z,10z_{\triangle,1}\neq 0 and UU^{\prime} rotationally disjoint from T(τ)T(\tau).

We define an RR-algebra automorphism φ:KQ(τ)KQ(τ)\varphi:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle by means of the rule

φ(α)=1z,1α.\varphi(\alpha_{\triangle})=\frac{1}{z_{\triangle,1}}\alpha_{\triangle}.

We see that φ(S)cycT(τ)+U\varphi(S)\sim_{\operatorname{cyc}}T(\tau)+U, for some potential UU rotationally disjoint from T(τ)T(\tau). ∎

Lemma 2.5 (Replacing ff-potentials and fgfg-potentials by longer ones).

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a punctured surface with empty boundary, and let τ\tau be a triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}) that satisfies (2.1) and (2.2). Let ϕ\phi be one of the symbols ff and fgfg, and let ν\nu be the other symbol, so that {ϕ,ν}={f,fg}\{\phi,\nu\}=\{f,fg\} as sets of symbols. If W,AKQ(τ)W,A\in K\langle\langle Q(\tau)\rangle\rangle are potentials rotationally disjoint from T(τ)T(\tau), and if Aϕ0A_{\phi}\neq 0, then there exists a potential BKQ(τ)B\in K\langle\langle Q(\tau)\rangle\rangle which is rotationally disjoint from T(τ)T(\tau) and satisfies the following four conditions:

short(Bϕ)\displaystyle\operatorname{short}(B_{\phi}) >\displaystyle> short(Aϕ);\displaystyle\operatorname{short}(A_{\phi});
short(Bg)\displaystyle\operatorname{short}(B_{g}) \displaystyle\geq min(short(Ag),short(Aϕ)+1);\displaystyle\min(\operatorname{short}(A_{g}),\operatorname{short}(A_{\phi})+1);
short(Bν)\displaystyle\operatorname{short}(B_{\nu}) \displaystyle\geq min(short(Aν),short(Aϕ)+1);\displaystyle\min(\operatorname{short}(A_{\nu}),\operatorname{short}(A_{\phi})+1);
(Q(τ),T(τ)+W+A)\displaystyle(Q(\tau),T(\tau)+W+A) r.e.\displaystyle\sim_{\operatorname{r.e.}} (Q(τ),T(τ)+W+B).\displaystyle(Q(\tau),T(\tau)+W+B).
Proof.

Let us deal with the case ϕ=f\phi=f. Write

Af=nshort(Af)3z,n(f2(α)f(α)α)nA_{f}=\sum_{\triangle}\sum_{n\geq\frac{\operatorname{short}(A_{f})}{3}}z_{\triangle,n}\left(f^{2}(\alpha_{\triangle})f(\alpha_{\triangle})\alpha_{\triangle}\right)^{n}

and define an RR-algebra homomorphism φ:KQ(τ)KQ(τ)\varphi:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle by means of the rule

φ(α)=αnshort(Af)3z,nα(f2(α)f(α)α)n1.\varphi(\alpha_{\triangle})=\alpha_{\triangle}-\sum_{n\geq\frac{\operatorname{short}(A_{f})}{3}}z_{\triangle,n}\alpha_{\triangle}\left(f^{2}(\alpha_{\triangle})f(\alpha_{\triangle})\alpha_{\triangle}\right)^{n-1}.

Then φ\varphi is a unitriangular automorphism of depth short(Af)3\operatorname{short}(A_{f})-3, and

φ(T(τ)+W+A)=T(τ)Af+W+A+(φ(W+A)(W+A)).\varphi(T(\tau)+W+A)=T(\tau)-A_{f}+W+A+(\varphi(W+A)-(W+A)).

Consequently, if we set B=Ag+Afg+(φ(W+A)(W+A))B=A_{g}+A_{fg}+(\varphi(W+A)-(W+A)), then:

  • φ(T(τ)+W+A)=T(τ)+W+B\varphi(T(\tau)+W+A)=T(\tau)+W+B;

  • short(φ(W+A)(W+A))0pt(φ)+short(W+A)short(Af)3+4=short(Af)+1\operatorname{short}(\varphi(W+A)-(W+A))\geq 0pt(\varphi)+\operatorname{short}(W+A)\geq\operatorname{short}(A_{f})-3+4=\operatorname{short}(A_{f})+1;

  • short(Bf)=short((φ(W+A)(W+A))f)short(Af)+1\operatorname{short}(B_{f})=\operatorname{short}((\varphi(W+A)-(W+A))_{f})\geq\operatorname{short}(A_{f})+1;

  • short(Bg)min(short(Ag),short((φ(W+A)(W+A))g)min(short(Ag),short(Af)+1)\operatorname{short}(B_{g})\geq\min(\operatorname{short}(A_{g}),\operatorname{short}((\varphi(W+A)-(W+A))_{g})\geq\min(\operatorname{short}(A_{g}),\operatorname{short}(A_{f})+1); and

  • short(Bfg)min(short(Afg),short((φ(W+A)(W+A))fg))min(short(Afg),short(Af)+1)\operatorname{short}(B_{fg})\geq\min(\operatorname{short}(A_{fg}),\operatorname{short}((\varphi(W+A)-(W+A))_{fg}))\geq\min(\operatorname{short}(A_{fg}),\operatorname{short}(A_{f})+1).

Now we deal with the case ϕ=fg\phi=fg. Write

Afg=aQ(τ)1f2(a)f(a)ωa,A_{fg}=\sum_{a\in Q(\tau)_{1}}f^{2}(a)f(a)\omega_{a},

with ωaeh(a)KQ(τ)et(a)\omega_{a}\in e_{h(a)}K\langle\langle Q(\tau)\rangle\rangle e_{t(a)} for each aQ(τ)1a\in Q(\tau)_{1} and define an RR-algebra homomorphism φ:KQ(τ)KQ(τ)\varphi:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle by means of the rule φ(a)=aωa\varphi(a)=a-\omega_{a} for aQ(τ)1a\in Q(\tau)_{1}. Then φ\varphi is a unitriangular automorphism of depth short(Afg)3\operatorname{short}(A_{fg})-3, and

φ(T(τ)+W+A)\displaystyle\varphi(T(\tau)+W+A) =\displaystyle= (f2(α)ωf2(α))(f(α)ωf(α))(αωα)\displaystyle\sum_{\triangle}\left(f^{2}(\alpha_{\triangle})-\omega_{f^{2}(\alpha_{\triangle})}\right)\left(f(\alpha_{\triangle})-\omega_{f(\alpha_{\triangle})}\right)\left(\alpha_{\triangle}-\omega_{\alpha_{\triangle}}\right)
+W+A+(φ(W+A)(W+A))\displaystyle+W+A+\left(\varphi(W+A)-(W+A)\right)
cyc\displaystyle\sim_{\operatorname{cyc}} T(τ)+W+Af+Ag+(φ(W+A)(W+A))\displaystyle T(\tau)+W+A_{f}+A_{g}+\left(\varphi(W+A)-(W+A)\right)
+aQ(τ)1f2(a)ωf(a)ωaωf2(a)ωf(a)ωa.\displaystyle+\sum_{a\in Q(\tau)_{1}}f^{2}(a)\omega_{f(a)}\omega_{a}-\sum_{\triangle}\omega_{f^{2}(a_{\triangle})}\omega_{f(a_{\triangle})}\omega_{a_{\triangle}}.

Consequently, if we set B=Af+Ag+(φ(W+A)(W+A))+aQ(τ)1f2(a)ωf(a)ωaωf2(a)ωf(a)ωaB=A_{f}+A_{g}+\left(\varphi(W+A)-(W+A)\right)+\sum_{a\in Q(\tau)_{1}}f^{2}(a)\omega_{f(a)}\omega_{a}-\sum_{\triangle}\omega_{f^{2}(a_{\triangle})}\omega_{f(a_{\triangle})}\omega_{a_{\triangle}}, then:

  • φ(T(τ)+W+A)cycT(τ)+W+B\varphi(T(\tau)+W+A)\sim_{\operatorname{cyc}}T(\tau)+W+B;

  • short(φ(W+A)(W+A))0pt(φ)+short(W+A)short(Afg)3+4=short(Afg)+1\operatorname{short}\left(\varphi(W+A)-(W+A)\right)\geq 0pt(\varphi)+\operatorname{short}(W+A)\geq\operatorname{short}(A_{fg})-3+4=\operatorname{short}(A_{fg})+1;

  • short(aQ(τ)1f2(a)ωf(a)ωa)2short(Afg)3short(Afg)+43=short(Afg)+1\operatorname{short}\left(\sum_{a\in Q(\tau)_{1}}f^{2}(a)\omega_{f(a)}\omega_{a}\right)\geq 2\operatorname{short}(A_{fg})-3\geq\operatorname{short}(A_{fg})+4-3=\operatorname{short}(A_{fg})+1 (since short(Afg)4\operatorname{short}(A_{fg})\geq 4);

  • short(ωf2(a)ωf(a)ωa)3short(Afg)6short(Afg)+86short(Afg)+1\operatorname{short}\left(\sum_{\triangle}\omega_{f^{2}(a_{\triangle})}\omega_{f(a_{\triangle})}\omega_{a_{\triangle}}\right)\geq 3\operatorname{short}(A_{fg})-6\geq\operatorname{short}(A_{fg})+8-6\geq\operatorname{short}(A_{fg})+1;

  • short(Bfg)min(short(φ(W+A)(W+A)),\operatorname{short}(B_{fg})\geq\min\left(\operatorname{short}\left(\varphi(W+A)-(W+A)\right),\right. short(aQ(τ)1f2(a)ωf(a)ωa),\left.\operatorname{short}\left(\sum_{a\in Q(\tau)_{1}}f^{2}(a)\omega_{f(a)}\omega_{a}\right),\right.
    short(ωf2(a)ωf(a)ωa))short(Afg)+1\left.\operatorname{short}\left(\sum_{\triangle}\omega_{f^{2}(a_{\triangle})}\omega_{f(a_{\triangle})}\omega_{a_{\triangle}}\right)\right)\geq\operatorname{short}(A_{fg})+1;

  • short(Bg)min(short(Ag),\operatorname{short}(B_{g})\geq\min\left(\operatorname{short}(A_{g}),\right. short(φ(W+A)(W+A)),\operatorname{short}\left(\varphi(W+A)-(W+A)\right), short(aQ(τ)1f2(a)ωf(a)ωa),\operatorname{short}\left(\sum_{a\in Q(\tau)_{1}}f^{2}(a)\omega_{f(a)}\omega_{a}\right),
    short(ωf2(a)ωf(a)ωa))min(short(Ag),short(Afg)+1)\left.\operatorname{short}\left(\sum_{\triangle}\omega_{f^{2}(a_{\triangle})}\omega_{f(a_{\triangle})}\omega_{a_{\triangle}}\right)\right)\geq\min\left(\operatorname{short}(A_{g}),\operatorname{short}(A_{fg})+1\right); and

  • short(Bf)min(short(Ag),\operatorname{short}(B_{f})\geq\min\left(\operatorname{short}(A_{g}),\right. short(φ(W+A)(W+A)),\operatorname{short}\left(\varphi(W+A)-(W+A)\right), short(aQ(τ)1f2(a)ωf(a)ωa),\operatorname{short}\left(\sum_{a\in Q(\tau)_{1}}f^{2}(a)\omega_{f(a)}\omega_{a}\right),
    short(ωf2(a)ωf(a)ωa))min(short(Af),short(Afg)+1)\left.\operatorname{short}\left(\sum_{\triangle}\omega_{f^{2}(a_{\triangle})}\omega_{f(a_{\triangle})}\omega_{a_{\triangle}}\right)\right)\geq\min\left(\operatorname{short}(A_{f}),\operatorname{short}(A_{fg})+1\right).

Lemma 2.5 is proved. ∎

Proposition 2.6 (Replacing potentials by sums of powers of gg-cycles).

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a punctured surface with empty boundary, and let τ\tau be a triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}) satisfying (2.1) and (2.2). If U,ZKQ(τ)U,Z\in K\langle\langle Q(\tau)\rangle\rangle are potentials rotationally disjoint from T(τ)T(\tau), then there exist a unitriangular automorphism φ:KQ(τ)KQ(τ)\varphi:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle of depth at least short(U)3\operatorname{short}(U)-3 and a potential WKQ(τ)W\in K\langle\langle Q(\tau)\rangle\rangle involving only positive powers of gg-cycles, such that short(W)short(U)\operatorname{short}(W)\geq\operatorname{short}(U) and φ\varphi is a right-equivalence (Q(τ),T(τ)+Z+U)(Q(τ),T(τ)+Z+W)(Q(\tau),T(\tau)+Z+U)\rightarrow(Q(\tau),T(\tau)+Z+W).

Proof.

Set W0=UgW_{0}=U_{g} and U0=UUgU_{0}=U-U_{g}. We obviously have short(U0),short(W0)short(U)\operatorname{short}(U_{0}),\operatorname{short}(W_{0})\geq\operatorname{short}(U).

Claim 1.

There exist sequences (Un)n1(U_{n})_{n\geq 1} and (Wn)n1(W_{n})_{n\geq 1} of potentials on the quiver Q(τ)Q(\tau), and a sequence (φn)n1(\varphi_{n})_{n\geq 1} of unitriangular automorphisms of KQ(τ)K\langle\langle Q(\tau)\rangle\rangle, such that the following properties are satisfied for every n1n\geq 1:

  • φn\varphi_{n} is a right-equivalence (Q(τ),T(τ)+Z+Un1+Wn1)(Q(τ),T(τ)+Z+Un+Wn)(Q(\tau),T(\tau)+Z+U_{n-1}+W_{n-1})\rightarrow(Q(\tau),T(\tau)+Z+U_{n}+W_{n});

  • 0pt(φn)=short(Un1)30pt(\varphi_{n})=\operatorname{short}(U_{n-1})-3;

  • each of UnU_{n} and WnW_{n} is rotationally disjoint from T(τ)T(\tau), UnU_{n} does not involve powers of gg-cycles and WnW_{n} involves only powers of gg-cycles;

  • short(WnWn1)short(Un1)+1\operatorname{short}(W_{n}-W_{n-1})\geq\operatorname{short}(U_{n-1})+1;

  • short(Un+1)short(Un1)+1\operatorname{short}(U_{n+1})\geq\operatorname{short}(U_{n-1})+1.

Proof of Claim 1.

We shall produce the three sequences (Un)n1(U_{n})_{n\geq 1}, (Wn)n1(W_{n})_{n\geq 1} and (φn)n1(\varphi_{n})_{n\geq 1} recursively. Fix a positive integer nn. If Un1=0U_{n-1}=0, we set UnU_{n} to be Un1U_{n-1}, WnW_{n} to be Wn1W_{n-1} and φn\varphi_{n} to be the identity of KQ(τ)K\langle\langle Q(\tau)\rangle\rangle. Otherwise, let ϕn1,νn1{f,fg}\phi_{n-1},\nu_{n-1}\in\{f,fg\} be symbols such that {ϕn1,νn1}={f,fg}\{\phi_{n-1},\nu_{n-1}\}=\{f,fg\} and short((Un1)ϕn1)short((Un1)νn1)\operatorname{short}((U_{n-1})_{\phi_{n-1}})\leq\operatorname{short}((U_{n-1})_{\nu_{n-1}}). By the proof of Lemma 2.5, there exist a potential VnKQ(τ)V_{n}\in K\langle\langle Q(\tau)\rangle\rangle rotationally disjoint from T(τ)T(\tau) and a unitriangular automorphism φn:KQ(τ)KQ(τ)\varphi_{n}:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle such that

  • 0pt(φn)=short(Un1)30pt(\varphi_{n})=\operatorname{short}(U_{n-1})-3;

  • φn\varphi_{n} is a right-equivalence (Q(τ),T(τ)+Z+Wn1+Un1)(Q(τ),T(τ)+Z+Wn1+Vn)(Q(\tau),T(\tau)+Z+W_{n-1}+U_{n-1})\rightarrow(Q(\tau),T(\tau)+Z+W_{n-1}+V_{n});

  • short((Vn)ϕn1)>short((Un1)ϕn1)\operatorname{short}((V_{n})_{\phi_{n-1}})>\operatorname{short}((U_{n-1})_{\phi_{n-1}});

  • short((Vn)g)min(short((Un1)g),short((Un1)ϕn1)+1)=short((Un1)ϕn1)+1\operatorname{short}((V_{n})_{g})\geq\min(\operatorname{short}((U_{n-1})_{g}),\operatorname{short}((U_{n-1})_{\phi_{n-1}})+1)=\operatorname{short}((U_{n-1})_{\phi_{n-1}})+1;

  • short((Vn)νn1)min(short((Un1)νn1),short((Un1)ϕn1)+1)\operatorname{short}((V_{n})_{\nu_{n-1}})\geq\min(\operatorname{short}((U_{n-1})_{\nu_{n-1}}),\operatorname{short}((U_{n-1})_{\phi_{n-1}})+1).

We set Un=Vn(Vn)gU_{n}=V_{n}-(V_{n})_{g} and Wn=Wn1+(Vn)gW_{n}=W_{n-1}+(V_{n})_{g}. It is clear that the first four properties stated in the claim are satisfied. For the fifth property, note that if ϕn=ϕn1\phi_{n}=\phi_{n-1}, then short((Un)ϕn)>short((Un1)ϕn1)\operatorname{short}((U_{n})_{\phi_{n}})>\operatorname{short}((U_{n-1})_{\phi_{n-1}}), whereas if ϕnϕn1\phi_{n}\neq\phi_{n-1}, then

short((Un+1)ϕn)>short((Un)ϕn)short(Un1)and\operatorname{short}((U_{n+1})_{\phi_{n}})>\operatorname{short}((U_{n})_{\phi_{n}})\geq\operatorname{short}(U_{n-1})\ \ \ \text{and}
short((Un+1)ϕn1)min(short((Un)ϕn1),short((Un)ϕn)+1)\operatorname{short}((U_{n+1})_{\phi_{n-1}})\geq\min(\operatorname{short}((U_{n})_{\phi_{n-1}}),\operatorname{short}((U_{n})_{\phi_{n}})+1)\geq
min(short((Un1)ϕn1)+1,min(short((Un1)νn1),short((Un1)ϕn1)+1)+1)>short((Un1)ϕn1).\min(\operatorname{short}((U_{n-1})_{\phi_{n-1}})+1,\min(\operatorname{short}((U_{n-1})_{\nu_{n-1}}),\operatorname{short}((U_{n-1})_{\phi_{n-1}})+1)+1)>\operatorname{short}((U_{n-1})_{\phi_{n-1}}).

These facts, together with the observation that for each n0n\geq 0 we have short((Un)ϕn)=short(Un)\operatorname{short}((U_{n})_{\phi_{n}})=\operatorname{short}(U_{n}), allow us to deduce that short(Un+1)short(Un1)+1\operatorname{short}(U_{n+1})\geq\operatorname{short}(U_{n-1})+1 for all n1n\geq 1. ∎

From the claim, we see that

limnshort(Un)=,limnshort(WnWn1)= and limn0pt(φn)=.\lim_{n\to\infty}\operatorname{short}(U_{n})=\infty,~{}\lim_{n\to\infty}\operatorname{short}(W_{n}-W_{n-1})=\infty\mbox{ and }\lim_{n\to\infty}0pt(\varphi_{n})=\infty.

Hence, if we set W=limnWnW=\lim_{n\to\infty}W_{n}, then φ:=limnφnφ1\varphi:=\lim_{n\to\infty}\varphi_{n}\circ\ldots\circ\varphi_{1} is a right-equivalence (Q(τ),T(τ)+Z+U)(Q(τ),T(τ)+Z+W)(Q(\tau),T(\tau)+Z+U)\rightarrow(Q(\tau),T(\tau)+Z+W). Proposition 2.6 follows. ∎

Lemma 2.7.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a punctured surface with empty boundary, let τ\tau be a triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}) satisfying (2.1) and (2.2), and let 𝐱=(xp)p𝕄\mathbf{x}=(x_{p})_{p\in\mathbb{M}} be any choice of non-zero scalars. Suppose that mm and tt are positive integers and U,WKQ(τ)U,W\in K\langle\langle Q(\tau)\rangle\rangle are potentials rotationally disjoint from S(τ,𝐱)S(\tau,\mathbf{x}) that satisfy the following properties:

  1. (1)

    short(U)m\operatorname{short}(U)\geq m;

  2. (2)

    2short(W)3>m2\operatorname{short}(W)-3>m;

  3. (3)

    W=λf(a)aG(t,gt(a))cW=\lambda f(a)aG(t,g^{-t}(a))c for some non-zero scalar λK\lambda\in K, some arrow aa and some path cc.

Then there exists a unitriangular RR-algebra automorphism ζ:KQ(τ)KQ(τ)\zeta:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle of depth short(W)3\operatorname{short}(W)-3 that serves as a right-equivalence between the QPs (Q(τ),S(τ,𝐱)+U+W)(Q(\tau),S(\tau,\mathbf{x})+U+W) and (Q(τ),S(τ,𝐱)+U+U+W)(Q(\tau),S(\tau,\mathbf{x})+U+U^{\prime}+W^{\prime}) for some potentials U,WKQ(τ)U^{\prime},W^{\prime}\in K\langle\langle Q(\tau)\rangle\rangle that satisfy:

  1. (1)

    short(U)>m\operatorname{short}(U^{\prime})>m;

  2. (2)

    short(W)>short(W)\operatorname{short}(W^{\prime})>\operatorname{short}(W);

  3. (3)

    W=λf(b)bG(t1,g(t1)(b))cW^{\prime}=\lambda^{\prime}f(b)bG(t-1,g^{-(t-1)}(b))c^{\prime} for some non-zero scalar λ\lambda^{\prime}, some arrow bb and some path cc^{\prime}.

Proof.

Let ζ:KQ(τ)KQ(τ)\zeta:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle be the RR-algebra homomorphism given by the rule

ζ(f1(a))=f1(a)λG(t,gt(a))c.\zeta(f^{-1}(a))=f^{-1}(a)-\lambda G(t,g^{-t}(a))c.

Since τ\tau satisfies (2.2), short(W)3\operatorname{short}(W)-3 is a positive integer by Lemma 2.2, and hence ζ\zeta is actually a unitriangular automorphism of KQ(τ)K\langle\langle Q(\tau)\rangle\rangle. The depth of ζ\zeta is obviously short(W)3\operatorname{short}(W)-3.

The arrow f1(a)f^{-1}(a) connects two arcs of τ\tau. Let pf1(a)p_{f^{-1}(a)} be the puncture at which these arcs are incident. Direct computation shows that

ζ(S(τ,𝐱)+U+W)\displaystyle\zeta(S(\tau,\mathbf{x})+U+W) cyc\displaystyle\sim_{\operatorname{cyc}} S(τ,𝐱)Wλxpf1(a)G(mf1(a)1,gf1(a))G(t,gt(a))c\displaystyle S(\tau,\mathbf{x})-W-\lambda x_{p_{f^{-1}(a)}}G(m_{f^{-1}(a)}-1,gf^{-1}(a))G(t,g^{-t}(a))c
+U+W+(ζ(U+W)(U+W))\displaystyle+U+W+\left(\zeta(U+W)-(U+W)\right)
=\displaystyle= S(τ,𝐱)λxpf1(a)G(mf1(a)2,g2f1(a))gf1(a)g1(a)G(t1,gt(a))c\displaystyle S(\tau,\mathbf{x})-\lambda x_{p_{f^{-1}(a)}}G(m_{f^{-1}(a)}-2,g^{2}f^{-1}(a))gf^{-1}(a)g^{-1}(a)G(t-1,g^{-t}(a))c
+U+(ζ(U+W)(U+W))\displaystyle+U+\left(\zeta(U+W)-(U+W)\right)
cyc\displaystyle\sim_{\operatorname{cyc}} S(τ,𝐱)λxpf1(a)gf1(a)g1(a)G(t1,gta)cG(mf1(a)2,g2f1(a))\displaystyle S(\tau,\mathbf{x})-\lambda x_{p_{f^{-1}(a)}}gf^{-1}(a)g^{-1}(a)G(t-1,g^{-t}a)cG(m_{f^{-1}(a)}-2,g^{2}f^{-1}(a))
+U+(ζ(U+W)(U+W)).\displaystyle+U+\left(\zeta(U+W)-(U+W)\right).

So, the lemma follows if we remember that gf1(a)=fg1(a)gf^{-1}(a)=fg^{-1}(a) and set

U\displaystyle U^{\prime} :=\displaystyle:= ζ(U+W)(U+W),\displaystyle\zeta(U+W)-(U+W),
λ\displaystyle\lambda^{\prime} :=\displaystyle:= λxpf1(a),\displaystyle-\lambda x_{p_{f^{-1}(a)}},
b\displaystyle b :=\displaystyle:= g1(a),\displaystyle g^{-1}(a),
c\displaystyle c^{\prime} :=\displaystyle:= cG(mf1(a)2,g2f1(a))\displaystyle cG(m_{f^{-1}(a)}-2,g^{2}f^{-1}(a))
andW\displaystyle\text{and}\ W^{\prime} :=\displaystyle:= λf(b)bG(t1,g(t1)(b))c.\displaystyle\lambda^{\prime}f(b)bG(t-1,g^{-(t-1)}(b))c^{\prime}.

Indeed, property (3) is obviously satisfied, whereas the inequalities short(U)m\operatorname{short}(U)\geq m, 0pt(ζ)>00pt(\zeta)>0 and 2short(W)3>m2\operatorname{short}(W)-3>m imply that

short(U)\displaystyle\operatorname{short}(U^{\prime}) \displaystyle\geq min(short(ζ(U)U),short(ζ(W)W))\displaystyle\min(\operatorname{short}(\zeta(U)-U),\operatorname{short}(\zeta(W)-W))
\displaystyle\geq min(0pt(ζ)+short(U),0pt(ζ)+short(W))\displaystyle\min(0pt(\zeta)+\operatorname{short}(U),0pt(\zeta)+\operatorname{short}(W))
=\displaystyle= min(0pt(ζ)+short(U),2short(W)3)\displaystyle\min(0pt(\zeta)+\operatorname{short}(U),2\operatorname{short}(W)-3)
>\displaystyle> m.\displaystyle m.

Furthermore, we also have

short(W)=mf1(a)2+short(W)1>short(W),\operatorname{short}(W^{\prime})=m_{f^{-1}(a)}-2+\operatorname{short}(W)-1>\operatorname{short}(W),

where the inequality follows from the fact that τ\tau satisfies (2.1). ∎

Corollary 2.8 (Replacing certain cycles by sums of long gg-cycles).

Under the same hypotheses of Lemma 2.7, if the path cc is assumed to be an arrow, then there exists a unitriangular RR-algebra automorphism Π:KQ(τ)KQ(τ)\Pi:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle of depth at least min(m3,short(W)3)\min(m-3,\operatorname{short}(W)-3) that serves as a right-equivalence between the QPs (Q(τ),S(τ,𝐱)+U+W)(Q(\tau),S(\tau,\mathbf{x})+U+W) and (Q(τ),S(τ,𝐱)+U+ξ)(Q(\tau),S(\tau,\mathbf{x})+U+\xi) for some potential ξ\xi that involves only positive powers of gg-cycles and satisfies short(ξ)>m\operatorname{short}(\xi)>m.

Proof.

This corollary follows from an inductive use of Lemma 2.7. Set U0=UU_{0}=U, W0=WW_{0}=W, a0=aa_{0}=a, c0=cc_{0}=c and λ0=λ\lambda_{0}=\lambda. Using Lemma 2.7, we obtain a unitriangular automorphism ζ1:KQ(τ)KQ(τ)\zeta_{1}:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle, potentials Z1,W1KQ(τ)Z_{1},W_{1}\in K\langle\langle Q(\tau)\rangle\rangle, an arrow a1a_{1}, a path c1c_{1} and a non-zero scalar λ1\lambda_{1}, such that:

  1. (1)

    0pt(ζ1)=short(W0)30pt(\zeta_{1})=\operatorname{short}(W_{0})-3;

  2. (2)

    ζ1\zeta_{1} is a right-equivalence (Q(τ),S(τ,𝐱)+U0+W0)(Q(τ),S(τ,𝐱)+U0+Z1+W1)(Q(\tau),S(\tau,\mathbf{x})+U_{0}+W_{0})\rightarrow(Q(\tau),S(\tau,\mathbf{x})+U_{0}+Z_{1}+W_{1});

  3. (3)

    short(Z1)>m\operatorname{short}(Z_{1})>m and short(W1)short(W0)+1\operatorname{short}(W_{1})\geq\operatorname{short}(W_{0})+1;

  4. (4)

    W1=λ1f(a1)a1G(t1,g(t1)(a1))c1W_{1}=\lambda_{1}f(a_{1})a_{1}G(t-1,g^{-(t-1)}(a_{1}))c_{1}.

Setting U1=U0+Z1U_{1}=U_{0}+Z_{1}, we see that U1U_{1}, W1W_{1}, a1a_{1}, c1c_{1} and λ1\lambda_{1} satisfy the hypotheses of Lemma 2.7 for the integers mm and t1t-1.

Assuming that for i{0,,t1}i\in\{0,\ldots,t-1\} we have UiU_{i}, WiW_{i}, aia_{i}, cic_{i} and λi\lambda_{i} satisfying the hypotheses of Lemma 2.7 for the integers mm and tit-i, we can produce a unitriangular automorphism ζi+1:KQ(τ)KQ(τ)\zeta_{i+1}:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle, potentials Zi+1,Wi+1KQ(τ)Z_{i+1},W_{i+1}\in K\langle\langle Q(\tau)\rangle\rangle, an arrow ai+1a_{i+1}, a path ci+1c_{i+1} and a non-zero scalar λi+1\lambda_{i+1}, such that:

  1. (1)

    0pt(ζi+1)=short(Wi)30pt(\zeta_{i+1})=\operatorname{short}(W_{i})-3;

  2. (2)

    ζi+1\zeta_{i+1} is a right-equivalence (Q(τ),S(τ,𝐱)+Ui+Wi)(Q(τ),S(τ,𝐱)+Ui+Zi+1+Wi+1)(Q(\tau),S(\tau,\mathbf{x})+U_{i}+W_{i})\rightarrow(Q(\tau),S(\tau,\mathbf{x})+U_{i}+Z_{i+1}+W_{i+1});

  3. (3)

    short(Zi+1)>m\operatorname{short}(Z_{i+1})>m and short(Wi+1)short(Wi)+1\operatorname{short}(W_{i+1})\geq\operatorname{short}(W_{i})+1;

  4. (4)

    Wi+1=λi+1f(ai+1)ai+1G(ti1,g(ti1)(ai+1))ci+1W_{i+1}=\lambda_{i+1}f(a_{i+1})a_{i+1}G(t-i-1,g^{-(t-i-1)}(a_{i+1}))c_{i+1}.

Setting Ui+1=Ui+Zi+1U_{i+1}=U_{i}+Z_{i+1}, we see that Ui+1U_{i+1}, Wi+1W_{i+1}, ai+1a_{i+1}, ci+1c_{i+1} and λi+1\lambda_{i+1} satisfy the hypotheses of Lemma 2.7 for the integers mm and t(i+1)t-(i+1).

The composition ζ=ζtζt1ζ1\zeta=\zeta_{t}\circ\zeta_{t-1}\circ\ldots\circ\zeta_{1} is a unitriangular automorphism of KQ(τ)K\langle\langle Q(\tau)\rangle\rangle that has depth at least short(W)3\operatorname{short}(W)-3 and serves as a right-equivalence (Q(τ),S(τ,𝐱)+U+W)(Q(τ),S(τ,𝐱)+Ut+Wt)(Q(\tau),S(\tau,\mathbf{x})+U+W)\rightarrow(Q(\tau),S(\tau,\mathbf{x})+U_{t}+W_{t}). Notice that Ut=U+i=1tZiU_{t}=U+\sum_{i=1}^{t}Z_{i}, that short(i=1tZi)>m\operatorname{short}\left(\sum_{i=1}^{t}Z_{i}\right)>m, and that short(Wt)short(W)+t=2short(W)3>m\operatorname{short}(W_{t})\geq\operatorname{short}(W)+t=2\operatorname{short}(W)-3>m.

By Proposition 2.6, there exists a unitriangular automorphism φ:KQ(τ)KQ(τ)\varphi:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle of depth greater than m3m-3 that makes (Q(τ),S(τ,𝐱)+U+i=1tZi+Wt)(Q(\tau),S(\tau,\mathbf{x})+U+\sum_{i=1}^{t}Z_{i}+W_{t}) right-equivalent to (Q(τ),S(τ,𝐱)+U+ξ)(Q(\tau),S(\tau,\mathbf{x})+U+\xi) for some potential ξKQ(τ)\xi\in K\langle\langle Q(\tau)\rangle\rangle that involves only powers of gg-cycles and satisfies short(ξ)short(i=1tZi+Wt)>m\operatorname{short}(\xi)\geq\operatorname{short}\left(\sum_{i=1}^{t}Z_{i}+W_{t}\right)>m.

From the two previous paragraphs we deduce that the automorphism Π:=φζ\Pi:=\varphi\circ\zeta satisfies the desired conclusion of Corollary 2.8. ∎

3. Once-punctured surfaces

In [7] and [9], the second author showed that the potentials S(τ,𝐱)S(\tau,\mathbf{x}) are well behaved with respect to flips and mutations, in the sense that if two triangulations are related by a flip, then the associated QPs are related by the corresponding QP-mutation. In this section, we show that for once-punctured closed surfaces the same result is true for a wider class of potentials. Namely, given a triangulation τ\tau of a once-punctured close surface of positive genus (Σ,𝕄)(\Sigma,\mathbb{M}), a scalar x0x\neq 0 and a positive integer nn, we define a potential S(τ,x,n)S(\tau,x,n) as

S(τ,x,n)=T(τ)+x𝒢(p)n,\displaystyle S(\tau,x,n)=T(\tau)+x\mathcal{G}(p)^{n},

where pp is the only puncture in (Σ,𝕄)(\Sigma,\mathbb{M}).

Theorem 3.1.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a once-punctured closed surface of positive genus, nn be any positive integer, let xKx\in K be any scalar. If τ\tau and σ\sigma are triangulations of (Σ,𝕄)(\Sigma,\mathbb{M}) that are related by the flip of an arc kτk\in\tau, then the QPs μk(Q(τ),S(τ,x,n))\mu_{k}(Q(\tau),S(\tau,x,n)) and (Q(σ),S(σ,x,n))(Q(\sigma),S(\sigma,x,n)) are right-equivalent.

Proof.

Let ai,bi,ci,i=1,2a_{i},b_{i},c_{i},~{}i=1,2, be the arrows in the two triangles with one side kk as in the figure below.

a2a_{2}c2c_{2}b2b_{2}a1a_{1}b1b_{1}c1c_{1}kk
Figure 1. The two triangles with one side kk.

Up to rotation we can write 𝒢(p)=a1Aa2B\mathcal{G}(p)=a_{1}Aa_{2}B. Notice that b2c1,b1c2b_{2}c_{1},b_{1}c_{2} are factors of 𝒢(p)\mathcal{G}(p), but b1c1b_{1}c_{1} and b2c2b_{2}c_{2} are not. The potential μ~k(S(τ,x,n))\widetilde{\mu}_{k}(S(\tau,x,n)) is cyclically equivalent to

[T(τ)]+x([a1Aa2B])n+c1b2[b2c1]+c2b1[b1c2]+c1b1[b1c1]+c2b2[b2c2]\displaystyle[T(\tau)]+x([a_{1}Aa_{2}B])^{n}+c_{1}^{*}b_{2}^{*}[b_{2}c_{1}]+c_{2}^{*}b_{1}^{*}[b_{1}c_{2}]+c_{1}^{*}b_{1}^{*}[b_{1}c_{1}]+c_{2}^{*}b_{2}^{*}[b_{2}c_{2}] =\displaystyle=
T(σ)+a1[b1c1]+a2[b2c2]+c1b1[b1c1]+c2b2[b2c2]+x(a1[A]a2[B])n,\displaystyle T(\sigma)+a_{1}[b_{1}c_{1}]+a_{2}[b_{2}c_{2}]+c_{1}^{*}b_{1}^{*}[b_{1}c_{1}]+c_{2}^{*}b_{2}^{*}[b_{2}c_{2}]+x(a_{1}[A]a_{2}[B])^{n},

where the paths [A],[B][A],[B] are the result of replacing b2c1,b1c2b_{2}c_{1},b_{1}c_{2} in A,BA,B by [b2c1],[b1c2][b_{2}c_{1}],[b_{1}c_{2}], respectively.

a1a_{1}a2a_{2}BBAAQ(τ)Q(\tau)b1b_{1}^{*}c1c_{1}^{*}c2c_{2}^{*}b2b_{2}^{*}[B][B][A][A]Q(σ)Q(\sigma)
Figure 2. The cycle on Q(τ)Q(\tau) and Q(σ)Q(\sigma) surrounding the puncture.

We define RR-algebra homomorphisms φ1,φ2:KQ(τ)KQ(τ)\varphi_{1},\varphi_{2}:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle by means of the rules

φ1(a1)=\displaystyle\varphi_{1}(a_{1})= a1c1b1;\displaystyle a_{1}-c_{1}^{*}b_{1}^{*};
φ2([b1c1])=\displaystyle\varphi_{2}([b_{1}c_{1}])= [b1c1]xj=0n1(1)j[A]a2[B]((a1c1b1)[A]a2[B])nj1(c1b1[A]a2[B])j.\displaystyle[b_{1}c_{1}]-x\sum_{j=0}^{n-1}(-1)^{j}[A]a_{2}[B]((a_{1}-c_{1}^{*}b_{1}^{*})[A]a_{2}[B])^{n-j-1}(c_{1}^{*}b_{1}^{*}[A]a_{2}[B])^{j}.

Applying φ1\varphi_{1} to μ~k(S(τ,x,n))\widetilde{\mu}_{k}(S(\tau,x,n)) we get

φ1(μ~k(S(τ,x,n)))cyc\displaystyle\varphi_{1}(\widetilde{\mu}_{k}(S(\tau,x,n)))\sim_{\operatorname{cyc}} T(σ)+a1[b1c1]+a2[b2c2]+c2b2[b2c2]+x((a1c1b1)[A]a2[B])n\displaystyle T(\sigma)+a_{1}[b_{1}c_{1}]+a_{2}[b_{2}c_{2}]+c_{2}^{*}b_{2}^{*}[b_{2}c_{2}]+x((a_{1}-c_{1}^{*}b_{1}^{*})[A]a_{2}[B])^{n}
cyc\displaystyle\sim_{\operatorname{cyc}} T(σ)+a1[b1c1]+a2[b2c2]+c2b2[b2c2]+x(1)n(c1b1[A]a2[B])n\displaystyle T(\sigma)+a_{1}[b_{1}c_{1}]+a_{2}[b_{2}c_{2}]+c_{2}^{*}b_{2}^{*}[b_{2}c_{2}]+x(-1)^{n}(c_{1}^{*}b_{1}^{*}[A]a_{2}[B])^{n}
+xj=0n1(1)ja1[A]a2[B]((a1c1b1)[A]a2[B])nj1(c1b1[A]a2[B])j.\displaystyle+x\sum_{j=0}^{n-1}(-1)^{j}a_{1}[A]a_{2}[B]((a_{1}-c_{1}^{*}b_{1}^{*})[A]a_{2}[B])^{n-j-1}(c_{1}^{*}b_{1}^{*}[A]a_{2}[B])^{j}.

The potential φ2φ1(μ~k(S(τ,x,n)))\varphi_{2}\varphi_{1}(\widetilde{\mu}_{k}(S(\tau,x,n))) is cyclically equivalent to

φ2φ1(μ~k(S(τ,x,n)))cyc\displaystyle\varphi_{2}\varphi_{1}(\widetilde{\mu}_{k}(S(\tau,x,n)))\sim_{\operatorname{cyc}} T(σ)+a1[b1c1]+a2[b2c2]+c2b2[b2c2]+x(1)n(c1b1[A]a2[B])n.\displaystyle T(\sigma)+a_{1}[b_{1}c_{1}]+a_{2}[b_{2}c_{2}]+c_{2}^{*}b_{2}^{*}[b_{2}c_{2}]+x(-1)^{n}(c_{1}^{*}b_{1}^{*}[A]a_{2}[B])^{n}.

In an analogous way, we define RR-algebra homomorphisms φ3,φ4:KQ(τ)KQ(τ)\varphi_{3},\varphi_{4}:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle by means of the rules

φ3(a2)=\displaystyle\varphi_{3}(a_{2})= a2c2b2;\displaystyle a_{2}-c_{2}^{*}b_{2}^{*};
φ4([b2c2])=\displaystyle\varphi_{4}([b_{2}c_{2}])= [b2c2]x(1)nj=0n1(1)j[B]c1b1[A]((a2c2b2)[B]c1b1[A])nj1(c2b2[B]c1b1[A])j.\displaystyle[b_{2}c_{2}]-x(-1)^{n}\sum_{j=0}^{n-1}(-1)^{j}[B]c_{1}^{*}b_{1}^{*}[A]((a_{2}-c_{2}^{*}b_{2}^{*})[B]c_{1}^{*}b_{1}^{*}[A])^{n-j-1}(c_{2}^{*}b_{2}^{*}[B]c_{1}^{*}b_{1}^{*}[A])^{j}.

We obtain

φ4φ3φ2φ1(μ~k(S(τ,x,n)))cyc\displaystyle\varphi_{4}\varphi_{3}\varphi_{2}\varphi_{1}(\widetilde{\mu}_{k}(S(\tau,x,n)))\sim_{\operatorname{cyc}} T(σ)+a1[b1c1]+a2[b2c2]+x(c1b1[A]c2b2[B])n\displaystyle T(\sigma)+a_{1}[b_{1}c_{1}]+a_{2}[b_{2}c_{2}]+x(c_{1}^{*}b_{1}^{*}[A]c_{2}^{*}b_{2}^{*}[B])^{n}
cyc\displaystyle\sim_{\operatorname{cyc}} S(σ,x,n)+a1[b1c1]+a2[b2c2].\displaystyle S(\sigma,x,n)+a_{1}[b_{1}c_{1}]+a_{2}[b_{2}c_{2}].

Therefore, the QPs μk(Q(τ),S(τ,x,n))\mu_{k}(Q(\tau),S(\tau,x,n)) and (Q(σ),S(σ,x,n))(Q(\sigma),S(\sigma,x,n)) are right-equivalent. ∎

Remark 3.2.
  1. (1)

    For once-punctured closed surfaces, Theorem 3.1 constitutes a generalization of the second author’s [7, Theorem 30] and [9, Theorem 8.1].

  2. (2)

    It was observed by Ladkani [12, Proposition 3.1] that the proof of [7, Theorem 30] can be applied without change to produce a proof of Theorem 3.1 above for x=0x=0.

  3. (3)

    In his Master thesis [5], the first author of this paper proved Theorem 3.1 for the once-punctured torus and x0x\neq 0.

  4. (4)

    Motivated by the first author’s Master thesis, the third author proved Theorem 3.1 in his Undergraduate thesis [13].

Proposition 3.3.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a once-punctured closed surface of positive genus, τ\tau a triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}), and xKx\in K a non-zero scalar. If the characteristic of the field KK is zero, then

dimK(𝒫(Q(τ),S(τ,x,n)))<andlimndimK(𝒫(Q(τ),S(τ,x,n)))=.\dim_{K}(\mathcal{P}(Q(\tau),S(\tau,x,n)))<\infty\ \ \ \text{and}\ \ \ \lim_{n\to\infty}\dim_{K}(\mathcal{P}(Q(\tau),S(\tau,x,n)))=\infty.
Proof.

For the proof of finite-dimensionality we follow ideas suggested by Ladkani in his proof of [11, Proposition 4.2], whereas our proof that the limits of the dimensions is \infty follows ideas that appear in the first author’s Master thesis.

First, note that when we compute the cyclic derivative of S(τ,x,n)S(\tau,x,n) with respect to an arrow α\alpha, we get

(3.1) α(S(τ,x,n))=f2(α)f(α)+xnG(nmα1,g(α)).\partial_{\alpha}(S(\tau,x,n))=f^{2}(\alpha)f(\alpha)+xnG(nm_{\alpha}-1,g(\alpha)).

So, f2(α)f(α)f^{2}(\alpha)f(\alpha) and xnG(nmα1,g(α))-xnG(nm_{\alpha}-1,g(\alpha)) become equal in the Jacobian algebra 𝒫(Q(τ),S(τ,x,n))\mathcal{P}(Q(\tau),S(\tau,x,n)).

Every fgfg-path of length three has the form f2(α)f(α)g1f(α)f^{2}(\alpha)f(\alpha)g^{-1}f(\alpha) or gf2(α)f2(α)f(α)gf^{2}(\alpha)f^{2}(\alpha)f(\alpha) for some arrow α\alpha, and it is hence equal to xnG(nmα1,g(α))g1f(α)=xnG(nmα3,g3(α))g2(α)g(α)f1g(α)-xnG(nm_{\alpha}-1,g(\alpha))g^{-1}f(\alpha)=-xnG(nm_{\alpha}-3,g^{3}(\alpha))g^{2}(\alpha)g(\alpha)f^{-1}g(\alpha) or xngf2(α)G(nmα1,g(α))=xnfg1(α)g1(α)g2(α)G(nmα3,g(α))-xngf^{2}(\alpha)G(nm_{\alpha}-1,g(\alpha))=-xnfg^{-1}(\alpha)g^{-1}(\alpha)g^{-2}(\alpha)G(nm_{\alpha}-3,g(\alpha)) in 𝒫(Q(τ),S(τ,x,n))\mathcal{P}(Q(\tau),S(\tau,x,n)). Thus every fgfg-path of length three is equal in 𝒫(Q(τ),S(τ,x,n))\mathcal{P}(Q(\tau),S(\tau,x,n)) to another fgfg-path of length greater than three. In the same vein, an easy inductive argument shows that, in the Jacobian algebra, every fgfg-path is to an arbitrarily long fgfg-path, and therefore equal to 0𝒫(Q(τ),S(τ,x,n))0\in\mathcal{P}(Q(\tau),S(\tau,x,n)).

Any ff-path F(r,f(β))=F(r2,β)f2(β)f(β)F(r,f(\beta))=F(r-2,\beta)f^{2}(\beta)f(\beta) of length rr greater than three, is equal to the fgfg-path xnF(r2,β)G(nmβ1,g(β))-xnF(r-2,\beta)G(nm_{\beta}-1,g(\beta)) in 𝒫(Q(τ),S(τ,x,n))\mathcal{P}(Q(\tau),S(\tau,x,n)), and in this way, to 0. Furthermore, any gg-path of the form G(r,g(β))=G(rnmβ+1,β)G(nmβ1,g(β))G(r,g(\beta))=G(r-nm_{\beta}+1,\beta)G(nm_{\beta}-1,g(\beta)) with length greater than nmβnm_{\beta}, is equal to the fgfg-path x1n1G(rnmβ+1,β)f2(β)f(β)x^{-1}n^{-1}G(r-nm_{\beta}+1,\beta)f^{2}(\beta)f(\beta), hence equal to 0 in the Jacobian algebra. Notice that here, we have used that KK is a field of characteristic zero.

Thus far, we have shown that every path of length greater than nmαnm_{\alpha} is equivalent to 0 in the Jacobian algebra 𝒫(Q(τ),S(τ,x,n))\mathcal{P}(Q(\tau),S(\tau,x,n)), and therefore the latter has finite dimension.

On the other hand, as the cyclic derivative of S(τ,x,n)S(\tau,x,n) with respect to any arrow α\alpha is equal to the sum of an ff-path of length two and a scalar multiple of a gg-path of length nmα1nm_{\alpha}-1 (3.1), and since no gg-path is a multiple of any ff-path of length greater than one, we conclude that for any a,bKQ(τ)a,b\in K\langle\langle Q(\tau)\rangle\rangle, no gg-path of length smaller than nmα1nm_{\alpha}-1 appears in the expression of the element aα(S(τ,x,n))ba\partial_{\alpha}(S(\tau,x,n))b as a possibly infinite sum of paths on the quiver Q(τ)Q(\tau). From this, it follows that no finite linear combination of gg-paths of lengths smaller than nmα1nm_{\alpha}-1 can be written as a limit of finite sums of elements of the form aα(S(τ,x,n))ba\partial_{\alpha}(S(\tau,x,n))b, i.e., the set of gg-paths of length smaller than nmα1nm_{\alpha}-1 is linearly independent in the Jacobian algebra 𝒫(Q(τ),S(τ,x,n))\mathcal{P}(Q(\tau),S(\tau,x,n)). Therefore, dimK(𝒫(Q(τ),S(τ,x,n)))nmα2\dim_{K}(\mathcal{P}(Q(\tau),S(\tau,x,n)))\geq nm_{\alpha}-2. ∎

Corollary 3.4.

Over a field of characteristic zero, the quiver of any triangulation of a once-punctured closed surface of positive genus admits infinitely many non-degenerate potentials up to weak right-equivalence.

Remark 3.5.
  1. (1)

    In the case of the once-punctured torus, Proposition 3.3, was proved by the first author in his Master thesis [5].

  2. (2)

    In his Undergraduate thesis [13], the third author has computed an actual KK-vector space basis of 𝒫(Q(τ),S(τ,x,n))\mathcal{P}(Q(\tau),S(\tau,x,n)) for each n1n\geq 1, showing in particular that different values of nn never yield Jacobian algebras with the same dimension. This implies that different values of nn always yield potentials that are not weakly right-equivalent.

4. Twice-punctured surfaces

In this section we prove part (2) of Theorem 1.1, namely:

Theorem 4.1.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a twice-punctured closed surface of positive genus, and let τ\tau be any (tagged) triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}). Over an algebraically closed field, any two non-degenerate potentials on the quiver Q(τ)Q(\tau) are weakly right-equivalent.

Since any two ideal triangulations of (Σ,𝕄)(\Sigma,\mathbb{M}) are related by a finite sequence of flips (see [14]), the first paragraphs of the proof of [4, Lemma 8.5] imply that the mere exhibition of a single triangulation τ\tau of (Σ,𝕄)(\Sigma,\mathbb{M}), with Q(τ)Q(\tau) having only one weak right equivalence class of non-degenerate potentials, suffices in order to prove Theorem 4.1.

Example 4.2.

Figure 3 sketches a triangulation τ\tau of a positive-genus twice-punctured surface with empty boundary. The triangulation is easily seen to satisfy (2.1) and (2.2). Note that the puncture pp has valency 8g8g and the other puncture qq has valency 4g4g.

112211222g12g-12g2g2g2gqqpppppppppppppppp2g+12g+12g+22g+22g+42g+42g+32g+36g6g6g16g-16g36g-36g26g-2
Figure 3. A triangulation τ\tau of a twice-punctured closed surface (Σ,𝕄)(\Sigma,\mathbb{M}) of positive-genus.
112211222g12g-12g2g2g2gqqppppppppppppppppa1a_{1}b1b_{1}c1c_{1}a3a_{3}c3c_{3}b3b_{3}a4g1a_{4g-1}b4g1b_{4g-1}c4g1c_{4g-1}a2a_{2}c2c_{2}b2b_{2}a4a_{4}b4b_{4}c4c_{4}a4ga_{4g}b4gb_{4g}c4gc_{4g}a4g2a_{4g-2}c4g2c_{4g-2}b4g2b_{4g-2}
Figure 4. The associated quiver Q(τ)Q(\tau) to the triangulation τ\tau.
Lemma 4.3.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a twice-punctured closed surface of positive genus, and let τ\tau be the triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}) depicted in Figure 3. If VKQ(τ)V\in K\langle\langle Q(\tau)\rangle\rangle is a potential involving only 2\geq 2-powers of gg-cycles, then (Q(τ),S(τ,𝐱)+V)(Q(\tau),S(\tau,\mathbf{x})+V) is right-equivalent to (Q(τ),S(τ,𝐱))(Q(\tau),S(\tau,\mathbf{x})) for any choice 𝐱=(xp,xq)\mathbf{x}=(x_{p},x_{q}) of non-zero scalars.

Proof.

Let gg be the genus of (Σ,𝕄)(\Sigma,\mathbb{M}). Then

Vcycn=2νp,n(𝒢(p))n+n=2νq,n(𝒢(q))nV\sim_{\operatorname{cyc}}\sum_{n=2}^{\infty}\nu_{p,n}(\mathcal{G}(p))^{n}+\sum_{n=2}^{\infty}\nu_{q,n}(\mathcal{G}(q))^{n}

for some scalars νp,n\nu_{p,n} and νq,n\nu_{q,n} for n2n\geq 2. Note that short(V)2valτ(q)=8g\operatorname{short}(V)\geq 2\operatorname{val}_{\tau}(q)=8g.

Claim 2.

There exist a sequence (Vm)m=8g(V_{m})_{m=8g}^{\infty} of potentials on Q(τ)Q(\tau), and a sequence (φm)m=8g(\varphi_{m})_{m=8g}^{\infty} of unitriangular RR-algebra automorphisms of KQ(τ)K\langle\langle Q(\tau)\rangle\rangle, satisfying the following properties:

  1. (1)

    V8g=VV_{8g}=V;

  2. (2)

    limm0pt(φm)=\lim_{m\to\infty}0pt(\varphi_{m})=\infty;

  3. (3)

    for every m8gm\geq 8g:

    1. (a)

      φm\varphi_{m} is a right-equivalence (Q(τ),S(τ,𝐱)+Vm)(Q(τ),S(τ,𝐱)+Vm+1)(Q(\tau),S(\tau,\mathbf{x})+V_{m})\rightarrow(Q(\tau),S(\tau,\mathbf{x})+V_{m+1});

    2. (b)

      VmV_{m} involves only 2\geq 2-powers of gg-cycles;

    3. (c)

      short(Vm)m\operatorname{short}(V_{m})\geq m.

Proof of Claim 2.

Start by setting V8g=VV_{8g}=V. Let apa_{p} (resp. aqa_{q}) be an arrow lying in the gg-orbit that surrounds pp (resp. qq). Suppose that for a fixed value of m8gm\geq 8g we have already defined a potential VmV_{m} involving only 2\geq 2-powers of gg-cycles and satisfying short(Vm)m\operatorname{short}(V_{m})\geq m. We shall use VmV_{m} to define Vm+1V_{m+1} and φm\varphi_{m}. Write:

Vmcycn=2λp,n(𝒢(ap))n+n=2λq,n(𝒢(aq))nV_{m}\sim_{\operatorname{cyc}}\sum_{n=2}^{\infty}\lambda_{p,n}(\mathcal{G}(a_{p}))^{n}+\sum_{n=2}^{\infty}\lambda_{q,n}(\mathcal{G}(a_{q}))^{n}

with λp,n,λq,nK\lambda_{p,n},\lambda_{q,n}\in K for n2n\geq 2. Set rp,mr_{p,m} (resp. rq,mr_{q,m}) to be the first value of nn for which λp,n0\lambda_{p,n}\neq 0 (resp. λq,n0\lambda_{q,n}\neq 0) if such an nn exists, and \infty if such an nn does not exist. Note that short(Vm)=min(8grp,m,4grq,n)8g\operatorname{short}(V_{m})=\min(8gr_{p,m},4gr_{q,n})\geq 8g.

Define an RR-algebra homomorphism Υp,m:KQ(τ)KQ(τ)\Upsilon_{p,m}:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle by means of the rule

Υp,m\displaystyle\Upsilon_{p,m} :\displaystyle: apapλp,rp,mxpap(𝒢(ap))rp,m1.\displaystyle a_{p}\mapsto a_{p}-\frac{\lambda_{p,r_{p,m}}}{x_{p}}a_{p}(\mathcal{G}(a_{p}))^{r_{p,m}-1}.

Since rp,m1>0r_{p,m}-1>0, Υp,n\Upsilon_{p,n} is a unitriangular automorphism, its depth is 8g(rp,m1)8g(r_{p,m}-1). Direct computation shows that

Υp,m(S(τ,𝐱)+Vm)\displaystyle\Upsilon_{p,m}(S(\tau,\mathbf{x})+V_{m}) cyc\displaystyle\sim_{\operatorname{cyc}} S(τ,𝐱)+U+W,\displaystyle S(\tau,\mathbf{x})+U+W,

where

U\displaystyle U =\displaystyle= λp,rp,m(𝒢(ap))rp,m+Υp,m(n=rp,mλp,n(𝒢(ap))n)+n=rq,mλq,n(𝒢(aq))n,\displaystyle-\lambda_{p,r_{p,m}}(\mathcal{G}(a_{p}))^{r_{p,m}}+\Upsilon_{p,m}\left(\sum_{n=r_{p,m}}^{\infty}\lambda_{p,n}(\mathcal{G}(a_{p}))^{n}\right)+\sum_{n=r_{q,m}}^{\infty}\lambda_{q,n}(\mathcal{G}(a_{q}))^{n},
W\displaystyle W =\displaystyle= λp,rp,mxpf(ap)ap(𝒢(ap))rp,m1f2(ap).\displaystyle-\frac{\lambda_{p,r_{p,m}}}{x_{p}}f(a_{p})a_{p}(\mathcal{G}(a_{p}))^{r_{p,m}-1}f^{2}(a_{p}).

Note that short(U)m\operatorname{short}(U)\geq m and 2short(W)3=28g(rp,m1)+38grp,m+3>8grp,mm2\operatorname{short}(W)-3=2*8g(r_{p,m}-1)+3\geq 8gr_{p,m}+3>8gr_{p,m}\geq m. So, applying Corollary 2.8, we see that there exists a unitriangular RR-algebra automorphism Πp,m\Pi_{p,m} of KQ(τ)K\langle\langle Q(\tau)\rangle\rangle that has depth at least min(m3,8g(rp,m1))\min(m-3,8g(r_{p,m}-1)) and serves as a right-equivalence between S(τ,𝐱)+U+WS(\tau,\mathbf{x})+U+W and S(τ,𝐱)+U+ξS(\tau,\mathbf{x})+U+\xi for some potential ξ\xi that involves only positive powers of gg-cycles and satisfies short(ξ)>m8g\operatorname{short}(\xi)>m\geq 8g. These last inequalities imply that, actually, ξ\xi involves only 2\geq 2-powers of gg-cycles.

Now, we can definitely write

(4.1) Ucycn=rp,m+1κp,n(𝒢(ap))n+n=rq,mλq,n(𝒢(aq))nU\sim_{\operatorname{cyc}}\sum_{n=r_{p,m}+1}^{\infty}\kappa_{p,n}(\mathcal{G}(a_{p}))^{n}+\sum_{n=r_{q,m}}^{\infty}\lambda_{q,n}(\mathcal{G}(a_{q}))^{n}

for some scalars κp,nK\kappa_{p,n}\in K. Define an RR-algebra homomorphism Υq,m:KQ(τ)KQ(τ)\Upsilon_{q,m}:K\langle\langle Q(\tau)\rangle\rangle\rightarrow K\langle\langle Q(\tau)\rangle\rangle by means of the rule

Υq,m\displaystyle\Upsilon_{q,m} :\displaystyle: aqaqλq,nxqaq(𝒢(aq))rq,m1.\displaystyle a_{q}\mapsto a_{q}-\frac{\lambda_{q,n}}{x_{q}}a_{q}(\mathcal{G}(a_{q}))^{r_{q,m}-1}.

Since rq,m1>0r_{q,m}-1>0, Υq,m\Upsilon_{q,m} is a unitriangular automorphism, its depth is 4g(rq,m1)4g(r_{q,m}-1). Direct computation shows that

Υq,m(S(τ,𝐱)+U+ξ)\displaystyle\Upsilon_{q,m}(S(\tau,\mathbf{x})+U+\xi) cyc\displaystyle\sim_{\operatorname{cyc}} S(τ,𝐱)+U+W,\displaystyle S(\tau,\mathbf{x})+U^{\prime}+W^{\prime},

where

U\displaystyle U^{\prime} =\displaystyle= λq,rq,m(𝒢(aq))rq,m+n=rp,m+1κp,n(𝒢(ap))n+Υq,n(n=rq,mλq,n(𝒢(aq))n)+Υq,m(ξ),\displaystyle-\lambda_{q,r_{q,m}}(\mathcal{G}(a_{q}))^{r_{q,m}}+\sum_{n=r_{p,m}+1}^{\infty}\kappa_{p,n}(\mathcal{G}(a_{p}))^{n}+\Upsilon_{q,n}\left(\sum_{n=r_{q,m}}^{\infty}\lambda_{q,n}(\mathcal{G}(a_{q}))^{n}\right)+\Upsilon_{q,m}(\xi),
W\displaystyle W^{\prime} =\displaystyle= λq,rq,mxqf(aq)aq(𝒢(aq))rq,m1f2(aq).\displaystyle-\frac{\lambda_{q,r_{q,m}}}{x_{q}}f(a_{q})a_{q}(\mathcal{G}(a_{q}))^{r_{q,m}-1}f^{2}(a_{q}).

Note that short(U)>m\operatorname{short}(U^{\prime})>m and 2short(W)3=24g(rq,m1)+34grq,m+3>4grq,mm2\operatorname{short}(W^{\prime})-3=2*4g(r_{q,m}-1)+3\geq 4gr_{q,m}+3>4gr_{q,m}\geq m. So, applying Corollary 2.8, we see that there exists a unitriangular RR-algebra automorphism Πq,m\Pi_{q,m} of KQ(τ)K\langle\langle Q(\tau)\rangle\rangle that has depth at least min(m3,4g(rq,m1))\min(m-3,4g(r_{q,m}-1)) and serves as a right-equivalence between S(τ,𝐱)+U+WS(\tau,\mathbf{x})+U^{\prime}+W^{\prime} and S(τ,𝐱)+U+ξS(\tau,\mathbf{x})+U^{\prime}+\xi^{\prime} for some potential ξ\xi^{\prime} that involves only positive powers of gg-cycles and satisfies short(ξ)>m8g\operatorname{short}(\xi^{\prime})>m\geq 8g. These last inequalities imply that, actually, ξ\xi^{\prime} involves only 2\geq 2-powers of gg-cycles.

It is clear that UU^{\prime} involves only positive powers of gg-cycles; this powers are actually greater than 1 because short(U)>m8g\operatorname{short}(U^{\prime})>m\geq 8g. So, if we set Vm+1=U+ξV_{m+1}=U^{\prime}+\xi^{\prime} and φm=Πq,mΥq,mΠp,mΥp,m\varphi_{m}=\Pi_{q,m}\Upsilon_{q,m}\Pi_{p,m}\Upsilon_{p,m}, we see that φm\varphi_{m} is a right-equivalence (Q(τ),S(τ,𝐱)+Vm)(Q(τ),S(τ,𝐱)+Vm+1)(Q(\tau),S(\tau,\mathbf{x})+V_{m})\rightarrow(Q(\tau),S(\tau,\mathbf{x})+V_{m+1}), that Vm+1V_{m+1} involves only 2\geq 2-powers of gg-cycles, and that short(Vm+1)m+1\operatorname{short}(V_{m+1})\geq m+1.

From the previous paragraph we deduce that the sequences (Vm)m8g(V_{m})_{m\geq 8g} and (φm)m8g(\varphi_{m})_{m\geq 8g} satisfy the third condition stated in Claim 2. Moreover, since mshort(Vm)=min(8grp,m,4grq,m)m\leq\operatorname{short}(V_{m})=\min(8gr_{p,m},4gr_{q,m}) for every m8gm\geq 8g, we deduce that limmrp,m==limmrq,m\lim_{m\to\infty}r_{p,m}=\infty=\lim_{m\to\infty}r_{q,m}. This and the inequalities

0pt(φm)\displaystyle 0pt(\varphi_{m}) \displaystyle\geq min(0pt(Πq,m),0pt(Υq,m),0pt(Πp,m),0pt(Υp,m))\displaystyle\min(0pt(\Pi_{q,m}),0pt(\Upsilon_{q,m}),0pt(\Pi_{p,m}),0pt(\Upsilon_{p,m}))
\displaystyle\geq min(min(m3,4g(rq,m1)),4g(rq,m1),min(m3,8g(rp,m1)),8g(rp,m1))\displaystyle\min(\min(m-3,4g(r_{q,m}-1)),4g(r_{q,m}-1),\min(m-3,8g(r_{p,m}-1)),8g(r_{p,m}-1))

imply that limm0pt(φm)=\lim_{m\to\infty}0pt(\varphi_{m})=\infty.

Our Claim 2 is proved. ∎

Lemma 4.3 follows from an obvious combination of Claim 2 and [9, Lemma 2.4]. ∎

Proposition 4.4.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) be a twice-punctured closed surface of positive genus, and let τ\tau be the triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}) depicted in Figure 3. If WKQ(τ)W\in K\langle\langle Q(\tau)\rangle\rangle is a potential that involves only positive powers of gg-cycles and such that (Q(τ),T(τ)+W)(Q(\tau),T(\tau)+W) is a non-degenerate QP, then WW involves each of the gg-cycles that arise from the two punctures pp and qq of (Σ,𝕄)(\Sigma,\mathbb{M}), that is, T(τ)+W=S(τ,𝐱)+VT(\tau)+W=S(\tau,\mathbf{x})+V for some choice 𝐱=(xp,xq)\mathbf{x}=(x_{p},x_{q}) of non-zero scalars and some potential VV involving only 2\geq 2-powers of gg-cycles.

Proof.

With the notation of Figures 3 and 4, let us write

W\displaystyle W =\displaystyle= ya1a2a4g+A+z(j=0g1b4(gj)c4(gj)2b4(gj)3c4(gj)1b4(gj)2c4(gj)b4(gj)1c4(gj)3)\displaystyle ya_{1}a_{2}\ldots a_{4g}+A+z\left(\prod_{j=0}^{g-1}b_{4(g-j)}c_{4(g-j)-2}b_{4(g-j)-3}c_{4(g-j)-1}b_{4(g-j)-2}c_{4(g-j)}b_{4(g-j)-1}c_{4(g-j)-3}\right)
+B,with\displaystyle+B,\text{with}
A\displaystyle A =\displaystyle= n=2yn(a1a2a4g)nand\displaystyle\sum_{n=2}^{\infty}y_{n}(a_{1}a_{2}\ldots a_{4g})^{n}\ \text{and}
B\displaystyle B =\displaystyle= n=2zn(j=0g1b4(gj)c4(gj)2b4(gj)3c4(gj)1b4(gj)2c4(gj)b4(gj)1c4(gj)3)n.\displaystyle\sum_{n=2}^{\infty}z_{n}\left(\prod_{j=0}^{g-1}b_{4(g-j)}c_{4(g-j)-2}b_{4(g-j)-3}c_{4(g-j)-1}b_{4(g-j)-2}c_{4(g-j)}b_{4(g-j)-1}c_{4(g-j)-3}\right)^{n}.

If we set I={2g+1,2g+2,,6g1,6g}I=\{2g+1,2g+2,\ldots,6g-1,6g\}, then (Q(τ),T(τ)+W)(Q(\tau),T(\tau)+W) and II satisfy the hypotheses of [4, Proposition 2.4], and we deduce that y0y\neq 0.

Note that for every k{1,,2g1}k\in\{1,\ldots,2g-1\}, the quiver μ~kμ~k1μ~2μ~1(Q(τ))\widetilde{\mu}_{k}\widetilde{\mu}_{k-1}\ldots\widetilde{\mu}_{2}\widetilde{\mu}_{1}(Q(\tau)) does not have 2-cycles incident to the vertex labelled k+1k+1. Therefore, the QP μ2gμ2g1μ2μ1(Q(τ),T(τ)+W)\mu_{2g}\mu_{2g-1}\ldots\mu_{2}\mu_{1}(Q(\tau),T(\tau)+W) is right-equivalent to the reduced part of the QP μ~2gμ~2g1μ~2μ~1(Q(τ),T(τ)+W)\widetilde{\mu}_{2g}\widetilde{\mu}_{2g-1}\ldots\widetilde{\mu}_{2}\widetilde{\mu}_{1}(Q(\tau),T(\tau)+W), whose underlying quiver and potential are μ~2gμ~1(Q(τ))\widetilde{\mu}_{2g}\ldots\widetilde{\mu}_{1}(Q(\tau)) and

μ~2gμ~1(T(τ)+W)\displaystyle\widetilde{\mu}_{2g}\ldots\widetilde{\mu}_{1}(T(\tau)+W) =\displaystyle= (j=14gaj[bjcj])+ya1a4g+A+[B]\displaystyle\left(\sum_{j=1}^{4g}a_{j}[b_{j}c_{j}]\right)+ya_{1}\ldots a_{4g}+A+[B]
+\displaystyle+ z(j=0g1[b4(gj)c4(gj)2][b4(gj)3c4(gj)1][b4(gj)2c4(gj)][b4(gj)1c4(gj)3])\displaystyle z\left(\prod_{j=0}^{g-1}[b_{4(g-j)}c_{4(g-j)-2}][b_{4(g-j)-3}c_{4(g-j)-1}][b_{4(g-j)-2}c_{4(g-j)}][b_{4(g-j)-1}c_{4(g-j)-3}]\right)
+\displaystyle+ (j=12gcjbj[bjcj]+cj+2bj[bjcj+2]+cjbj+2[bj+2cj]+cj+2bj+2[bj+2cj+2]).\displaystyle\left(\sum_{j=1}^{2g}c_{j}^{*}b_{j}^{*}[b_{j}c_{j}]+c_{j+2}^{*}b_{j}^{*}[b_{j}c_{j+2}]+c_{j}^{*}b_{j+2}^{*}[b_{j+2}c_{j}]+c_{j+2}^{*}b_{j+2}^{*}[b_{j+2}c_{j+2}]\right).

Consider the QP (μ~2gμ~1(Q(τ)),S¯)(\widetilde{\mu}_{2g}\ldots\widetilde{\mu}_{1}(Q(\tau)),\overline{S}), where

S¯\displaystyle\overline{S} =\displaystyle= (j=14gaj[bjcj])+ya1a4g+A+(j=12gcjbj[bjcj]+cj+2bj+2[bj+2cj+2]),\displaystyle\left(\sum_{j=1}^{4g}a_{j}[b_{j}c_{j}]\right)+ya_{1}\ldots a_{4g}+A+\left(\sum_{j=1}^{2g}c_{j}^{*}b_{j}^{*}[b_{j}c_{j}]+c_{j+2}^{*}b_{j+2}^{*}[b_{j+2}c_{j+2}]\right),

and let (Q,S)(Q,S) be its reduced part, computed according to the limit process with which Derksen-Weyman-Zelevinsky [2, Theorem 4.6] prove their Splitting Theorem. Note the presence of the sum j=14gaj[bjcj]\sum_{j=1}^{4g}a_{j}[b_{j}c_{j}] in S¯\overline{S}. Then Q=Q(σ)Q=Q(\sigma), where σ\sigma is a triangulation that can be obtained from τ\tau by applying an orientation-preserving homeomorphism of (Σ,𝕄)(\Sigma,\mathbb{M}) that exchanges pp and qq (thus τ\tau and σ\sigma have the same shape, sketched in Figure 4; see also Example 4.5 below). Moreover, since no arrow of the form aja_{j} or [bjcj][b_{j}c_{j}] appears in any of the terms of the potential

W\displaystyle W^{\prime} :=\displaystyle:= z(j=0g1[b4(gj)c4(gj)2][b4(gj)3c4(gj)1][b4(gj)2c4(gj)][b4(gj)1c4(gj)3])+[B]\displaystyle z\left(\prod_{j=0}^{g-1}[b_{4(g-j)}c_{4(g-j)-2}][b_{4(g-j)-3}c_{4(g-j)-1}][b_{4(g-j)-2}c_{4(g-j)}][b_{4(g-j)-1}c_{4(g-j)-3}]\right)+[B]
+(j=12gcj+2bj[bjcj+2]+cjbj+2[bj+2cj]),\displaystyle+\left(\sum_{j=1}^{2g}c_{j+2}^{*}b_{j}^{*}[b_{j}c_{j+2}]+c_{j}^{*}b_{j+2}^{*}[b_{j+2}c_{j}]\right),

the QP (Q(σ),S+W)(Q(\sigma),S+W^{\prime}) is a reduced part of (μ~2gμ~1(Q(τ)),μ~2gμ~1(T(τ)+W))(\widetilde{\mu}_{2g}\ldots\widetilde{\mu}_{1}(Q(\tau)),\widetilde{\mu}_{2g}\ldots\widetilde{\mu}_{1}(T(\tau)+W)) and hence is (right-equivalent to) the mutation μ2gμ1(Q(τ),T(τ)+W){\mu}_{2g}\ldots{\mu}_{1}(Q(\tau),T(\tau)+W). Furthermore, from the fact that no arrow of the form [bjc][b_{j}c_{\ell}] with jj\neq\ell appears in any of the terms of S¯\overline{S} we deduce that the coefficient in SS of any of the rotations of the cycle

(j=0g1[b4(gj)c4(gj)2][b4(gj)3c4(gj)1][b4(gj)2c4(gj)][b4(gj)1c4(gj)3])\left(\prod_{j=0}^{g-1}[b_{4(g-j)}c_{4(g-j)-2}][b_{4(g-j)-3}c_{4(g-j)-1}][b_{4(g-j)-2}c_{4(g-j)}][b_{4(g-j)-1}c_{4(g-j)-3}]\right)

is 0. Therefore, the coefficient of this cycle in S+WS+W^{\prime} is zz (and its proper rotations do not appear).

The non-degeneracy of (Q(τ),T(τ)+W)(Q(\tau),T(\tau)+W) implies the non-degeneracy of (Q(σ),S+W)(Q(\sigma),S+W^{\prime}). Furthermore, it is easy to see that if we set I={2g+1,2g+2,,6g1,6g}I=\{2g+1,2g+2,\ldots,6g-1,6g\}, then (Q(σ),S+W)(Q(\sigma),S+W^{\prime}) and II satisfy the hypotheses of [4, Proposition 2.4], from which we deduce that z0z\neq 0. This finishes the proof of Proposition 4.4. ∎

Example 4.5.

Figure 5 sketches the flip sequence in the proof of Proposition 4.4 in the case of a twice-punctured torus. Note that the first and last triangulations have the same shape.

qqqqqqqqqqqqqqqqqqppppppppppppppppppppppppppppppppqqqqqqqqqqqqqqqqqqppppppppppppppppppppppppppppppppqqqqqqqqqqqqqqqqqqpppppppppppppppppppppppppppppppp
Figure 5. Proving Proposition 4.4 for the twice-punctured torus.
Proof of Theorem 4.1.

Let (Σ,𝕄)(\Sigma,\mathbb{M}) a be twice-punctured closed surface of positive genus, and let τ\tau be a triangulation of (Σ,𝕄)(\Sigma,\mathbb{M}) satisfying (2.1) and (2.2). By Lemma 2.4, every non-degenerate potential on Q(τ)Q(\tau) is right-equivalent to a potential of the form T(τ)+UT(\tau)+U for some UU which is rotationally disjoint from T(τ)T(\tau). By Proposition 2.6, T(τ)+UT(\tau)+U is right-equivalent to T(τ)+WT(\tau)+W for some potential that involves only positive powers of gg-cycles. Theorem 4.1 now follows from Proposition 4.4, Lemma 4.3 and [4, Lemma 8.5]. ∎

Acknowledgements

We thank Christof Geiss and Jan Schröer for many helpful discussions.

The three authors were supported by the second author’s grant PAPIIT-IA102215. The first two authors were supported by the second author’s grant CONACyT-238754 as well. DLF received support from a Cátedra Marcos Moshinsky and the grant PAPIIT-IN112519.

References

  • [1] Tom Bridgeland, Ivan Smith. Quadratic differentials as stability conditions. Publications mathématiques de l’IHÉS, June 2015, Volume 121, Issue 1, 155–278. arXiv:1302.7030
  • [2] Harm Derksen, Jerzy Weyman, Andrei Zelevinsky. Quivers with potentials and their representations I: Mutations. Selecta Math. 14 (2008), no. 1, 59–119. arXiv:0704.0649
  • [3] Sergey Fomin, Michael Shapiro, Dylan Thurston. Cluster algebras and triangulated surfaces, part I: Cluster complexes. Acta Mathematica 201 (2008), 83-146. arXiv:math.RA/0608367
  • [4] Christof Geiss, D. Labardini-Fragoso, Jan Schröer. The representation type of Jacobian algebras. Advances in Mathematics, Vol. 290 (2016), 364-452. doi:10.1016/j.aim.2015.09.038   arXiv:1308.0478
  • [5] J. Geuenich. Quivers with potentials on three vertices. Master thesis. Universität Bonn. 2013.
  • [6] Bernhard Keller, Dong Yang. Derived equivalences from mutations of quiver with potential. Adv. Math. 226: 2118–2168, (2011).
  • [7] D. Labardini-Fragoso. Quivers with potentials associated to triangulated surfaces. Proc. London Mathematical Society (2009) 98 (3): 797-839. arXiv:0803.1328
  • [8] Daniel Labardini-Fragoso. Quivers with potentials associated to triangulated surfaces, part II: Arc representations. arXiv:0909.4100
  • [9] D. Labardini-Fragoso. Quivers with potentials associated to triangulated surfaces, part IV: Removing boundary assumptions. Selecta Mathematica (New series), Vol. 22 (2016), Issue 1, 145–189. DOI: 10.1007/s00029-015-0188-8 arXiv:1206.1798
  • [10] D. Labardini-Fragoso. On triangulations, quivers with potentials and mutations. Contemporary Mathematics (AMS), Vol. 657 “Mexican Mathematicians Abroad: Recent Contributions” (2016), 103–127. http://dx.doi.org/10.1090/conm/657/13092   arXiv:1302.1936
  • [11] Sefi Ladkani. On Jacobian algebras from closed surfaces. arXiv:1207.3778
  • [12] Sefi Ladkani. On cluster algebras from once punctured closed surfaces. arXiv:1310.4454
  • [13] J. L. Miranda-Olvera. Carcajes con potencial no degenerados asociados a triangulaciones de superficies: Existencia y unicidad. Undergraduate thesis. Facultad de Ciencias, Universidad Nacional Autónoma de México. 2016.
  • [14] Lee Mosher. Tiling the projective foliation space of a punctured surface. Trans. Amer. Math. Soc. 306 (1988), 1–70.
  • [15] Ivan Smith. Quiver algebras as Fukaya categories. Geom. Topol. 19 (2015), no. 5, 2557–2617. arXiv:1309.0452