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

On the properties of acyclic sign-skew-symmetric cluster algebras

Siyang Liu Siyang Liu
School of Mathematics, Hangzhou Normal University, Hangzhou 311121, P.R.China
[email protected]
Abstract.

We study the tropical dualities and properties of exchange graphs for the totally sign-skew-symmetric cluster algebra under a condition. We prove that the condition always holds for acyclic cluster algebras, then all results hold for the acyclic case.


1. Introduction

Cluster algebras, introduced by Fomin and Zelevinsky in [FZ02] for providing a combinatorial framework to study total positivity and canonical bases associated by Lusztig to algebraic groups, are certain commutative algebras generated by certain combinatorially defined generators (called cluster variables) which are grouped into overlapping clusters. Since cluster algebras were invented, many links between cluster algebras and other subjects have been discovered. For example, cluster algebras have close relations with Riemann surfaces and triangulations, Zamolodchikov’s periodicity properties concerning YY-systems, representation theory, Donaldson-Thomas invariant theory and scattering diagrams.

The initial data to construct a cluster algebra is a totally sign-skew-symmetric matrix in the first paper in the series of foundational papers of cluster theory. Since then, most authors prefer to study skew-symmetrizable cluster algebras and many important properties and conjectures were proved in the skew-symmetrizable case. Even though there are still many unkonwns for totally sign-skew-symmetric cluster algebras, we believe that their properties behave like the skew-symmetrizable case.

The most important property of cluster algebras is the Laurent phenomenon. Fomin and Zelevinsky conjectured the cluster variables written as Laurent polynomials of a cluster always have the positive coefficients, which is called the positivity conjecture. This conjecture was proved by Lee and Schiffler for the symmetric case [LS15], by Gross, Hacking, Keel and Kontsevich for the skew-symmetrizable case [GHKK18], Huang and Li for the acyclic sign-skew-symmetric cluster algebras. Recently, Li and Pan proved it for all totally sign-skew-symmetric case [LP].

In the fourth paper of the foundational papers [FZ07], Fomin and Zelevinsky introduced a lot of combinatorial objects including cc-vectors and gg-vectors (also CC-matrices and GG-matrices) to study cluster algebras. The well-known sign coherence conjecture was also proposed in [FZ07]. The dualities of GG-matrix and CC-matrix were systematically studied by Nakanishi and Zelevinsky in [NZ12]. In their paper, they shall assume the sign coherence conjecture holds for cc-vectors for all skew-symmetrizable cluster algebras. This was confirmed by Derksen, Weyman, and Zelevinsky in the skew-symmetric case [DWZ10] and by Gross, Hacking, Keel, and Kontsevich in the skew-symmetrizable case [GHKK18]. Thus the dualities obtained by Nakanishi and Zelevinsky in [NZ12] are true for skew-symmetrizable cluster algebras. Let 𝒜\mathcal{A} be a skew-symmetrizable cluster algebra of rank nn. Roughly speaking, they proved the following results for skew-symmetrizable cluster algebras:

(1.1) Every GG-matrix is an invertible matrix of some CC-matrix;
(1.2) The invertible matrix of every CC-matrix is also some CC-matrix;
(1.3) Every GG-matrix has the row sign-coherent property;
(1.4) The dual mutations of GG-matrix and CC-matrix hold.

These claims were proved by Nakanishi and Zelevinsky [NZ12] (see also [Nak]) by assuming cc-vectors are sign coherent. Even though the sign coherence conjecture of cc-vectors was proved for the acyclic case by Huang and Li [HL18], and in full generality by Li and Pan [LP], the proofs in [NZ12] largely depends on the skew-symmetrizable property.

Reading studied the property of GG-fan G\mathcal{F}_{G} consisting of all GG-cones (non-negative spans of the column vectors of GG-matrices) and their faces in [Rea14]. For the skew-symmetrizable cluster algebras, he also defined the mutation fan, and he proved the following result.

(1.5) G\mathcal{F}_{G} is a fan. Indeed it is a subfan of the mutation fan.

The bijection of gg-vectors and cluster variables is a corollary of the positivity conjecture and the above claims as proved in [Nak].

(1.6) Two cluster variables are the same \Longleftrightarrow their corresponding gg-vectors are the same.

Fomin and Zelevinsky also made a series of conjectures on the properties of exchange graphs of 𝒜\mathcal{A} as follows.

(1.7) Every seed is uniquely defined by its cluster.
(1.8) Two clusters are adjacent \Longleftrightarrow they have exactly n1n-1 common cluster variables.
(1.9) Each seed with principal coefficients is determined by the corresponding CC-matrix..

These conjectures were proved for skew-symmetrizable cluster algebras by Gekhtman, Shapiro, and Vainshtein [GSV08], Cao and Li [CL18], and Cao, Huang, and Li [CHL18].

In this paper, we show that under an condition (the Assumption in Section 4.2), the claims (1.11.9) hold for totally sign-skew-symmetric cluster algebras. In particular, they hold for acyclic sign-skew-symmetric cluster algebras.

Theorem 1.1 (Proposition 4.7, 4.8, Theorem 4.16, 4.23, 5.8, 5.10 ).

Let BB be a totally sign-skew-symmetric matrix satisfying the Assumption. Consider the corresponding matrix pattern and cluster algebra, then the above claims (1.11.9) hold. In particular, claims (1.11.9) hold for acyclic cluster algebras.

In the next section, we recall basic theory of cluster algebras. In Section 3, we describe behaviors about cluster algebras under folding theory. In Section 4, we prove the properties of GG-matrices for totally sign-skew-symmetric cluster algebras that satisfy a certain condition (which we conjecture it is always true). In the final section, we study the properties of exchange graphs.

2. Preliminaries

Let II be a countable set. We say an integer matrix BMatI×I()B\in Mat_{I\times I}(\mathbb{Z}) is locally finite if for each iIi\in I, there are only finitely many nonzero entries bijb_{ij} and bjib_{ji}, and say BB is sign-skew-symmetric if bijbji0b_{ij}b_{ji}\leq 0 and bijbji=0bij=bji=0b_{ij}b_{ji}=0\Leftrightarrow b_{ij}=b_{ji}=0 for all i,jIi,j\in I. A skew-symmetrizable matrix BB admits a diagonal matrix D=diag(di,iI)D=\text{diag}(d_{i},i\in I) with positive integer diagonal entries such that dibij=djbjid_{i}b_{ij}=-d_{j}b_{ji}, which is clearly sign-skew-symmetric. A quiver is a directed graph without loops or 22-cycles, and since a locally finite skew-symmetric matrix naturally corresponds to a locally finite quiver, we do not distinguish them in the following sections.

A semifield \mathbb{P} is a set equipped with a structure of abelian multiplicative group and a structure of abelian additive semigroup (the addition is usually denoted by \oplus) such that the addition \oplus is distributive with respect to the multiplication in \mathbb{P}. The tropical semifield =trop(zj,jJ)\mathbb{P}=\mathrm{{trop}}(z_{j},j\in J) with an index set JJ is a multiplicative group generated freely by the elements zj,jJz_{j},j\in J, whose addition \oplus is given by

jJzjajjJzjbj=jJzjmin(aj,bj).\prod_{j\in J}z_{j}^{a_{j}}\oplus\prod_{j\in J}z_{j}^{b_{j}}=\prod_{j\in J}z_{j}^{\mathrm{min}(a_{j},b_{j})}.

Let \mathbb{QP} be the quotient field of the integral domain \mathbb{ZP}.

Definition 2.1.

A (labeled) seed is a triplet Σ=(𝐱,𝐲,B)\Sigma=(\mathbf{x},\mathbf{y},B), where

  1. -

    𝐱=(xi,iI)\mathbf{x}=(x_{i},i\in I) is a countable set of indeterminates over \mathbb{QP} such that =(xi,iI)\mathcal{F}=\mathbb{QP}(x_{i},i\in I) is a purely transcendental field extension of \mathbb{QP}, and 𝐱\mathbf{x} is called a cluster, xi,iIx_{i},i\in I are called cluster variables;

  2. -

    𝐲=(yi,iI)\mathbf{y}=(y_{i},i\in I) is a subset of the semifield \mathbb{P} indexed by II;

  3. -

    BMatI×I()B\in Mat_{I\times I}(\mathbb{Z}) is a locally finite sign-skew-symmetric matrix, and it is called the exchange matrix.

For kIk\in I, define another triplet (𝐱,𝐲,B)=μk(𝐱,𝐲,B)(\mathbf{x^{\prime}},\mathbf{y^{\prime}},B^{\prime})=\mu_{k}(\mathbf{x},\mathbf{y},B) which is called the mutation of (𝐱,𝐲,B)(\mathbf{x},\mathbf{y},B) at kk and obtained by the following rules:

  1. -

    𝐱=(xi,iI)\mathbf{x^{\prime}}=(x^{\prime}_{i},i\in I) is given by

    xi={xi,ik,ykxi[bik]++xi[bik]+(yk1)xk,i=k.x^{\prime}_{i}=\begin{cases}x_{i},&\text{$i\neq k$},\\ \frac{y_{k}\prod x_{i}^{[b_{ik}]_{+}}+\prod x_{i}^{[-b_{ik}]_{+}}}{(y_{k}\oplus 1)x_{k}},&\text{$i=k$.}\end{cases}
  2. -

    𝐲=(yi,iI)\mathbf{y^{\prime}}=(y^{\prime}_{i},i\in I) is given by

    yi={yk1,i=k,yiyk[bki]+(yk1)bki,ik.y^{\prime}_{i}=\begin{cases}y_{k}^{-1},&\text{$i=k$,}\\ y_{i}y_{k}^{[b_{ki}]_{+}}(y_{k}\oplus 1)^{-b_{ki}},&\text{$i\neq k$.}\end{cases}
  3. -

    B=(bij)B^{\prime}=(b^{\prime}_{ij}) is given by

    bij={bij,if i=k or j=k,bij+sign(bik)[bikbkj]+,otherwise.b^{\prime}_{ij}=\begin{cases}-b_{ij},&\text{if $i=k$ or $j=k$,}\\ b_{ij}+\mathrm{sign}(b_{ik})[b_{ik}b_{kj}]_{+},&\text{otherwise.}\end{cases}

where [a]+=max{a,0}[a]_{+}=\mathrm{max}\{a,0\}.

Say two labeled seeds Σ\Sigma and Σ\Sigma^{\prime} define the same unlabeled seed if there is a bijection σ\sigma of II such that xi=xσ(i)x^{\prime}_{i}=x_{\sigma(i)}, yi=yσ(i)y^{\prime}_{i}=y_{\sigma(i)}, bij=bσ(i)σ(j)b^{\prime}_{ij}=b_{\sigma(i)\sigma(j)} for all i,jIi,j\in I.

If every matrix obtained from a locally finite sign-skew-symmetric matrix by an arbitrary finite sequence of mutations is also sign-skew-symmetric, we call it a totally sign-skew-symmetric matrix. Clearly, skew-symmetrizable matrices are totally mutable. For a sign-skew-symmetric matrix B=(bij)i,jIB=(b_{ij})_{i,j\in I}, we may define a directed graph whose vertices are indexed by II, and there is an arrow from ii to jj if bij>0b_{ij}>0. We say BB is acyclic if the corresponding directed graph has no oriented cycles. Huang and Li also proved acyclic sign-skew-symmetric matrices are totally mutable [HL18].

Let 𝕋\mathbb{T} be an |I||I|-regular tree and its valencies emitting from each common vertex are labeled by II. A cluster pattern is an |I||I|-regular tree 𝕋\mathbb{T} such that for each vertex t𝕋t\in\mathbb{T}, there is a seed Σt=(𝐱t,𝐲t,Bt)\Sigma_{t}=(\mathbf{x}_{t},\mathbf{y}_{t},B_{t}) and for each edge labeled by kIk\in I, two seeds in the endpoints are obtained from each other by seed mutation at kk. Note that a cluster pattern is determined by a totally sign-skew-symmetric matrix. The corresponding exchange graph is the graph whose vertices are unlabeled seeds and two unlabeled seeds are connected by an edge if they are related by a single mutation. The cluster algebra 𝒜=𝒜(𝐱,𝐲,B)\mathcal{A}=\mathcal{A}(\mathbf{x},\mathbf{y},B) associated to a totally sign-skew-symmetric matrix BB is the \mathbb{ZP}-subalgebra of \mathcal{F} generated by all cluster variables on 𝕋\mathbb{T} corresponding to the matrix BB. The cluster algebra 𝒜=𝒜(𝐱,𝐲,B)\mathcal{A}=\mathcal{A}(\mathbf{x},\mathbf{y},B) is said to have principal coefficients, if the semifield is the tropical semifield given by =trop(yi,iI)\mathbb{P}=\mathrm{trop}(y_{i},i\in I).

Fomin and Zelevinsky proved the following Laurent property for cluster algebras, and they also conjectured cluster variables have non-negative Laurent expressions.

Theorem 2.2 (The Laurent phenomenon, [FZ02, FZ07]).

Let 𝕋\mathbb{T} be a cluster pattern. Then for arbitrary two vertices t,t0𝕋t,t_{0}\in\mathbb{T} and any jIj\in I, we have that xj,t[xi,t0±1,iI]x_{j,t}\in\mathbb{ZP}[x_{i,t_{0}}^{\pm 1},i\in I]. Moreover,

  1. (1)

    when =trop(zi,iJ)\mathbb{P}=trop(z_{i},i\in J) is a tropical semifield, we have that xj,t[zp,xi,t0±1,pJ,iI]x_{j,t}\in\mathbb{Z}[z_{p},x_{i,t_{0}}^{\pm 1},p\in J,i\in I].

  2. (2)

    when =trop(yi,t0,iI)\mathbb{P}=trop(y_{i,t_{0}},i\in I) is the principal case, each cluster variable xj,tx_{j,t} is homogeneous with respect to a given |I|\mathbb{Z}^{|I|}-grading in [yi,t0,xi,t0±1,iI]\mathbb{Z}[y_{i,t_{0}},x_{i,t_{0}}^{\pm 1},i\in I], which is given by

    deg(xi,t0)=𝐞i,deg(yi)=𝐛i,deg(x_{i,t_{0}})=\mathbf{e}_{i},\,\,\,\,deg(y_{i})=-\mathbf{b}_{i},

    where 𝐞i,iI\mathbf{e}_{i},i\in I are the standard basis vectors in |I|\mathbb{Z}^{|I|}, and 𝐛i,t0\mathbf{b}_{i,t_{0}} is the ii-th column vector of Bt0B_{t_{0}}.

In the following discussion, we shall fix a seed, which we call the initial seed and is usually placed at t0𝕋t_{0}\in\mathbb{T}, to obtain other datum from it.

Let 𝒜\mathcal{A} be a cluster algebra with principal coefficients and initial seed Σ=Σt0\Sigma=\Sigma_{t_{0}}. Then by Theorem 2.2, we have that xj,t[yi,xi±1,iI]x_{j,t}\in\mathbb{Z}[y_{i},x_{i}^{\pm 1},i\in I]. The polynomial Fj,t:=xj,t|xi=1,iIF_{j,t}:=x_{j,t}|_{x_{i}=1,i\in I} is called the FF-polynomial (with the initial seed at t0t_{0}). Since yj,ttrop(yi,iI)y_{j,t}\in\text{trop}(y_{i},i\in I), we have that yj,ty_{j,t} is of the form yj,t=iIyicij,ty_{j,t}=\prod_{i\in I}y_{i}^{c_{ij,t}} for some integers cij,t,iIc_{ij,t},i\in I. The column vector cj,t=(cij,t,iI)Tc_{j,t}=(c_{ij,t},i\in I)^{T} is called the cc-vectors and the matrix Ctt0=(cj,t)jIC^{t_{0}}_{t}=(c_{j,t})_{j\in I} is called the CC-matrix at tt. The column vector gj,tg_{j,t}, the degree vector of xj,tx_{j,t} with respect to the grading given in Theorem 2.2, is called the gg-vector, and the matrix Gtt0=(gj,t)jIG^{t_{0}}_{t}=(g_{j,t})_{j\in I} is called the GG-matrix at tt. Note that the definitions of FF-polynomials, CC-matrices and GG-matrices depend on the choice of the initial seed. Also in [FZ07], Fomin and Zelevinsky obtained the mutation rules for FF-polynomials, CC-matrices and GG-matrices.

For each edge tktt\frac{k}{\quad\quad}t^{\prime} on 𝕋\mathbb{T}, we have that

  1. -

    Fi,t={Fi,t,if ik;(Fk,t)1(jIyjcjk,tFj,t[bjk,t]++jIyjcjk,tFj,t[bjk,t]+),otherwise.F_{i,t^{\prime}}=\begin{cases}F_{i,t},&\text{if $i\neq k$;}\\ (F_{k,t})^{-1}(\prod_{j\in I}y_{j}^{c_{jk,t}}F_{j,t}^{[b_{jk,t}]_{+}}+\prod_{j\in I}y_{j}^{-c_{jk,t}}F_{j,t}^{[-b_{jk,t}]_{+}}),&\text{otherwise.}\end{cases}

  2. -

    cij,t={cik,t,if j=k;cij,t+sign(cik,t)[cik,tbjk,t]+,otherwise.c_{ij,t^{\prime}}=\begin{cases}-c_{ik,t},&\text{if $j=k$;}\\ c_{ij,t}+\mathrm{sign}(c_{ik,t})[c_{ik,t}b_{jk,t}]_{+},&\text{otherwise.}\end{cases}

  3. -

    gij,t={gij,t,if jk;gik,t+ΣsIgis,t[bsk,t]+ΣsIbis,t[csk,t]+,otherwise.g_{ij,t^{\prime}}=\begin{cases}g_{ij,t},&\text{if $j\neq k$;}\\ -g_{ik,t}+\Sigma_{s\in I}g_{is,t}[-b_{sk,t}]_{+}-\Sigma_{s\in I}b_{is,t}[-c_{sk,t}]_{+},&\text{otherwise.}\end{cases}

A nonzero vector is said to be sign-coherent if it has either all non-negative coordinates or all non-positive coordinates. Fomin and Zelevinsky conjectured every cc-vectors is sign coherent and each row vector of every GG-matrix is sign-coherent. The sign coherence conjecture were solved for skew-symmetrizable cluster algebras in [GHKK18] and for acyclic sign-skew-symmetric cluster algebras in [HL18, CHL22].

3. Folding theory

Recall that a quiver is a directed graph (may be infinite) which has no loops or 22-cycles. It is well-known that a quiver is naturally associated to a skew-symmetric matrix.

Definition 3.1.

Let QQ be a locally finite quiver with frozen vertices set FF, and Γ\Gamma be a group acting on the vertex set Q0Q_{0} of QQ. The vertices in Q0FQ_{0}\setminus F are called mutable vertices. Assume that B=(bij)i,jQ0B=(b_{ij})_{i,j\in Q_{0}} is the matrix corresponding to QQ, we say QQ is Γ\Gamma-admissible, if the following condition are satisfied:

  1. (i)

    ii is mutable \iff g(i)g(i) is mutable, for all iQ0i\in Q_{0} and gΓg\in\Gamma;

  2. (ii)

    bij=bg(i)g(j)b_{ij}=b_{g(i)g(j)}, for all i,jQ0i,j\in Q_{0}, and gΓg\in\Gamma;

  3. (iii)

    big(i)=0b_{ig(i)}=0, for all mutable iQ0i\in Q_{0}, and gΓg\in\Gamma;

  4. (iv)

    bijbg(i)j0b_{ij}b_{g(i)j}\geq 0, for all iQ0i\in Q_{0}, jQ0Fj\in Q_{0}\setminus F, and gΓg\in\Gamma.

Definition 3.2 (Orbit mutation).

Assume that QQ is Γ\Gamma-admissible, and B=(bij)i,jQ0B=(b_{ij})_{i,j\in Q_{0}} its corresponding matrix. For each mutable kQ0k\in Q_{0}, the matrix B=(bij)i,jQ0B^{\prime}=(b^{\prime}_{ij})_{i,j\in Q_{0}} of orbit mutation Q=μ[i](Q)Q^{\prime}=\mu_{[i]}(Q) of QQ at [k][k] is given by the following rule:

bij={bij,if i[k] or j[k];bij+p[k]|bip|bpj+bip|bpj|2,otherwise.b^{\prime}_{ij}=\begin{cases}-b_{ij},&\text{if $i\in[k]$ or $j\in[k]$};\\ b_{ij}+\sum\limits_{p\in[k]}\frac{|b_{ip}|b_{pj}+b_{ip}|b_{pj}|}{2},&\text{otherwise.}\end{cases}

The following Lemma follows easily from the definition.

Lemma 3.3.

Suppose that there are two actions of Γ\Gamma on the vertices set Q0Q_{0} making QQ Γ\Gamma-admissible, and these two actions on the mutable vertices are the same. Then for each mutable vertex kQ0k\in Q_{0}, the orbit mutations of QQ at [k][k] with respect to the two actions of Γ\Gamma are the same.

Assume that QQ is Γ\Gamma-admissible, denote [i][i] the Γ\Gamma-orbit for each iQ0i\in Q_{0}, I¯\bar{I} the set of Γ\Gamma-orbits, and μ[i]:=j[i]μj\mu_{[i]}:=\prod_{j\in[i]}\mu_{j} the orbit mutation. The matrix BΓ=(b[i][j])[i],[j]I¯B^{\Gamma}=(b_{[i][j]})_{[i],[j]\in\bar{I}} is defined by the following rule:

b[i][j]=i[i]bij.b_{[i][j]}=\sum_{i^{\prime}\in[i]}b_{i^{\prime}j}.
Definition 3.4.

The locally finite quiver QQ is said to be globally foldable with respect to the group Γ\Gamma, if QQ is Γ\Gamma-admissible, and for any sequence of orbits [i1],,[ik][i_{1}],\dots,[i_{k}], the quiver μikμi1(Q)\mu_{i_{k}}\dots\mu_{i_{1}}(Q) is also Γ\Gamma-admissible. In this case, we say (Q,Γ)(Q,\Gamma) is an unfolding of BΓB^{\Gamma}.

The following Lemma follows easily from the definition.

Lemma 3.5 ([FWZ]).

Let QQ be a locally finite quiver globally foldable with respect to an action of a group Γ\Gamma. Let Q¯\bar{Q} be a quiver constructed from QQ by introducing new frozen vertices together with some arrows connecting them to the mutable vertices in QQ. Extend the action of Γ\Gamma from QQ to Q¯\bar{Q} by making Γ\Gamma fix every newly added vertex. Then the quiver Q¯\bar{Q} is globally foldable with respect to Γ\Gamma.

Notice that for a locally finite quiver QQ, the cluster algebra 𝒜(𝐱,Q)\mathcal{A}(\mathbf{x},Q), FF-polynomials, cc-vectors, and gg-vectors could be defined naturally. If QQ is globally foldable with respect to Γ\Gamma and for any mutable iQ0i\in Q_{0}, then define μ[i](xj)\mu_{[i]}(x_{j}) by the rules: μ[i](xj)=μj(xj)\mu_{[i]}(x_{j})=\mu_{j}(x_{j}) if j[i]j\in[i], and μ[i](xj)=xj\mu_{[i]}(x_{j})=x_{j} if j[i]j\notin[i].

Lemma 3.6 ([HL18]).

Assume that the locally finite quiver QQ is globally foldable with respect to the group Γ\Gamma. Then we have the following results: for any mutable vertices i1,,isi_{1},\dots,i_{s}

  1. (1)

    the cluster variables of 𝒜(μ[is]μ[i1](𝐱,Q))\mathcal{A}(\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\mathbf{x},Q)) is the same as the cluster variables of 𝒜(𝐱,Q)\mathcal{A}(\mathbf{x},Q);

  2. (2)

    any finite variables in μ[is]μ[i1](𝐱,Q)\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\mathbf{x},Q) is contained in a cluster of 𝒜(𝐱,Q)\mathcal{A}(\mathbf{x},Q);

  3. (3)

    any variable in μ[is]μ[i1](𝐱,Q)\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\mathbf{x},Q) is a cluster variable of 𝒜(𝐱,Q)\mathcal{A}(\mathbf{x},Q);

  4. (4)

    any monomial with variables in μ[is]μ[i1](𝐱,Q)\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\mathbf{x},Q) is a cluster monomial of 𝒜(𝐱,Q)\mathcal{A}(\mathbf{x},Q);

  5. (5)

    each FF-polynomial has the constant term 11.

Proof.

The statements (1)(4)(1)-(4) are proved in [HL18] for the case BΓB^{\Gamma} is acyclic. It is easy to see that the proof works for all globally foldable quiver QQ. The claim (5)(5) is proved in [Lemma 7.12, [HL18]]. ∎

Let QQ be a locally finite quiver without frozen vertices. The corresponding framed quiver Q~\tilde{Q} is obtained from QQ by adding frozen vertices Q0:={i,iQ0}Q^{\prime}_{0}:=\{i^{\prime},i\in Q_{0}\} and arrows Q1:={ii,iQ0}Q^{\prime}_{1}:=\{i^{\prime}\rightarrow i,i\in Q_{0}\}. If QQ is globally foldable with respect to Γ\Gamma, by Lemma 3.5, Q~\tilde{Q} is globally foldable with respect to Γ\Gamma with Γ\Gamma fixing frozen vertices.

Lemma 3.7.

If the locally finite quiver QQ without frozen vertices is globally foldable with respect to Γ\Gamma, then for any vertices i1,,isQ0i_{1},\dots,i_{s}\in Q_{0}, there are no two frozen vertices i,ji^{\prime},j^{\prime} and a mutable vertex kk such that there are arrows of the form ikji^{\prime}\rightarrow k\rightarrow j^{\prime} in μ[is]μ[i1](Q~)\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{Q}).

Proof.

By Lemma 3.5, we know that Q~\tilde{Q} is globally foldable with respect to Γ\Gamma with Γ\Gamma fixing frozen vertices, then μ[is]μ[i1](Q~)\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{Q}) is well-defined.

If the claim is not true, let us consider the cluster algebra 𝒜(𝐱~,Q~)\mathcal{A}(\tilde{\mathbf{x}},\tilde{Q}), here 𝐱~=(𝐱,𝐲)\tilde{\mathbf{x}}=(\mathbf{x},\mathbf{y}) and 𝐲=(yi,iQ0)\mathbf{y}=(y_{i},i\in Q_{0}) are frozen variables. Suppose that (𝐱~,Q~)=μ[is]μ[i1](𝐱~,Q~)(\tilde{\mathbf{x}}^{\prime},\tilde{Q}^{\prime})=\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{\mathbf{x}},\tilde{Q}). By Lemma 3.6, the variables in 𝐱~\tilde{\mathbf{x}}^{\prime} and the variable

x′′=ikyiikxi+kjyjkjxjxkx^{\prime\prime}=\frac{\prod\limits_{i^{\prime}\rightarrow k}y_{i}\prod\limits_{i\rightarrow k}x^{\prime}_{i}+\prod\limits_{k\rightarrow j^{\prime}}y_{j}\prod\limits_{k\rightarrow j}x^{\prime}_{j}}{x^{\prime}_{k}}

are cluster variables in 𝒜(𝐱~,Q~)\mathcal{A}(\tilde{\mathbf{x}},\tilde{Q}). Under the specialization of xi=1x_{i}=1 for all iQ0i\in Q_{0}, we have that

F′′=ikyiikFi+kjyjkjFjFk.F^{\prime\prime}=\frac{\prod\limits_{i^{\prime}\rightarrow k}y_{i}\prod\limits_{i\rightarrow k}F^{\prime}_{i}+\prod\limits_{k\rightarrow j^{\prime}}y_{j}\prod\limits_{k\rightarrow j}F^{\prime}_{j}}{F^{\prime}_{k}}.

By Lemma 3.6, the left side will be 11 and the right side will be 0 by letting yi=0y_{i}=0 for all iQ0i\in Q_{0}. This is a contradiction. ∎

Lemma 3.8.

Let QQ be a globally foldable quiver with respect to Γ\Gamma, and Q~\tilde{Q} its framed quiver. Extend the action of Γ\Gamma on Q0Q_{0} to the action of Γ\Gamma on Q~0=Q0\tilde{Q}_{0}=Q_{0} by g.(i):=g(i)g.(i^{\prime}):=g(i)^{\prime} for each iQ0i\in Q_{0}. then μ[is]μ[i1](Q~)\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{Q}) is Γ\Gamma-admissible for any sequence of orbit mutations with i1,,isQ0i_{1},\dots,i_{s}\in Q_{0}.

Proof.

We prove the claim by the induction on the length of orbit mutation sequences.

It is not hard to see that Q~\tilde{Q} is also Γ\Gamma-admissible. Assume that μ[is]μ[i1](Q~)\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{Q}) is Γ\Gamma-admissible, we only need to show that μ[k]μ[is]μ[i1](Q~)\mu_{[k]}\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{Q}) is also Γ\Gamma-admissible for every kQ0k\in Q_{0}.

For simplicity, we may abuse notations and let (BCTCM)\begin{pmatrix}B&-C^{T}\\ C&M\end{pmatrix} denote the matrix of μ[is]μ[i1](Q~)\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{Q}) and (B(C)TCM)\begin{pmatrix}B^{\prime}&-(C^{\prime})^{T}\\ C^{\prime}&M^{\prime}\end{pmatrix} to denote the matrix of μ[k]μ[is]μ[i1](Q~)\mu_{[k]}\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{Q}), where C=(cij)i,jQ0C=(c_{i^{\prime}j})_{i,j\in Q_{0}} and C=(cij)i,jQ0C^{\prime}=(c^{\prime}_{i^{\prime}j})_{i,j\in Q_{0}}. Since μ[is]μ[i1](Q~)\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{Q}) is Γ\Gamma-admissible, by Lemma 3.3 and Lemma 3.7, CC and CC^{\prime} have the column sign-coherent property and M=M=0M=M^{\prime}=0. Thus we have that cijcg(i)j0c^{\prime}_{i^{\prime}j}c^{\prime}_{g(i^{\prime})j}\geq 0 for i,jQ0i,j\in Q_{0}.

Note that CC and CC^{\prime} are related by the following relations by Definition :

cij={cij,j[k]cij+p[k]cipbpj,j[k],p[k] such that cip>0,bpj>0;cij+p[k]cipbpj,j[k],p[k] such that cip<0,bpj<0;cij,otherwise.c^{\prime}_{i^{\prime}j}=\begin{cases}-c_{i^{\prime}j},&\text{$j\in[k]$}\\ c_{i^{\prime}j}+\sum\limits_{p\in[k]}c_{i^{\prime}p}b_{pj},&\text{$j\notin[k],\,\,\exists\,p\in[k]$ such that $c_{i^{\prime}p}>0,b_{pj}>0;$}\\ c_{i^{\prime}j}+\sum\limits_{p\in[k]}c_{i^{\prime}p}b_{pj},&\text{$j\notin[k],\,\,\exists\,p\in[k]$ such that $c_{i^{\prime}p}<0,b_{pj}<0;$}\\ c_{i^{\prime}j},&\text{otherwise.}\end{cases}

It is easy to check that cg(i)g(j)=cijc^{\prime}_{g(i^{\prime})g(j)}=c^{\prime}_{i^{\prime}j} for all i,jQ0i,j\in Q_{0}. If j[k]j\in[k], then g(j)kg(j)\in k, and hence cg(i)g(j)=cg(i)g(j)=cij=cijc^{\prime}_{g(i^{\prime})g(j)}=-c_{g(i^{\prime})g(j)}=-c_{i^{\prime}j}=c^{\prime}_{i^{\prime}j}. If j[k],j\notin[k], and p[k]\exists\,p\in[k], such that cip>0,bpj>0c_{i^{\prime}p}>0,b_{pj}>0, then g(j)[k]g(j)\notin[k], and g(p)kg(p)\in k satisfies that cg(i)g(p)=cip>0c_{g(i^{\prime})g(p)}=c_{i^{\prime}p}>0, bg(p)g(j)=bpj>0b_{g(p)g(j)}=b_{pj}>0. Thus we have that

cij=cij+p[k]cipbpj=cg(i)g(j)+p[k]cg(i)g(p)bg(p)g(j)=cg(i)g(j)+p[k]cg(i)pbpg(j)=cg(i)g(j).c^{\prime}_{i^{\prime}j}=c_{i^{\prime}j}+\sum\limits_{p\in[k]}c_{i^{\prime}p}b_{pj}=c_{g(i^{\prime})g(j)}+\sum\limits_{p\in[k]}c_{g(i^{\prime})g(p)}b_{g(p)g(j)}=c_{g(i^{\prime})g(j)}+\sum\limits_{p\in[k]}c_{g(i^{\prime})p}b_{pg(j)}=c^{\prime}_{g(i^{\prime})g(j)}.

The other cases can be checked similarly.

This proves that μ[k]μ[is]μ[i1](Q~)\mu_{[k]}\mu_{[i_{s}]}\dots\mu_{[i_{1}]}(\tilde{Q}) is also Γ\Gamma-admissible. ∎

Theorem 3.9.

Let QQ be a locally finite globally foldable quiver with respect to Γ\Gamma. Then the positivity conjecture holds for the cluster algebra 𝒜(BΓ)\mathcal{A}(B^{\Gamma}), i.e., for every cluster variable xx and every cluster 𝐱\mathbf{x}, we have that xx is a Laurent polynomial with variables in 𝐱\mathbf{x} and positive coefficients.

Moreover, each FF-polynomial has the constant term 11, and thus the cc-vectors have the sign-coherent property.

Proof.

It is enough to prove positivity conjecture for cluster algebras with principal coefficients. This follows from Lemma 3.8, Lemma 3.6 and the fact that positivity conjecture holds for skew-symmetric cluster algebras of infinite rank. ∎

Theorem 3.10 ([HL18]).

There exists an unfolding for every sign-skew-symmetric matrix mutation equivalent to an acyclic sign-skew-symmetric matrix.

Combining Theorem 3.9 and Theorem 3.10, we have the following results.

Theorem 3.11 (cf. [HL18]).

The positivity conjecture holds for the sign-skew-symmetric cluster algebras with the initial seed mutation equivalent to an acyclic seed.

Corollary 3.12.

If B=(bij)n×nB=(b_{ij})_{n\times n} is a sign-skew-symmetric matrix which is mutation equivalent to an acyclic sign-skew-symmetric matrix, then for any sequence k1,,ks[1,n]k_{1},\dots,k_{s}\in[1,n], the matrix μksμk1(BInIn0)\mu_{k_{s}}\dots\mu_{k_{1}}\begin{pmatrix}B&-I_{n}\\ I_{n}&0\end{pmatrix} is sign-skew-symmetric.

Proof.

By Theorem 3.10, there is an unfolding (Q,Γ)(Q,\Gamma) for BB. By Lemma 3.8, there is an unfolding for the matrix (BInIn0)\begin{pmatrix}B&-I_{n}\\ I_{n}&0\end{pmatrix}.

Conjecture 3.13.

Corollary 3.12 holds for all totally sign-skew-symmetric matrices.

4. The properties of GG-matrices

The tropical dualities for GG-matrices and CC-matrices were studied by Nakanishi and Zelevinky for the skew-symmetrizable cluster algebras [NZ12]. Notice that Nakanishi and Zelevinsky’s proof depends on the sign coherence of cc-vectors. In this section, we consider the dual phenomenon for the totally sign-skew-symmetric case. In this section and the next section, we always assume that |I|=n<|I|=n<\infty.

As we mentioned in the Introduction, Fomin and Zelevinsky proposed a series of conjectures on GG-matrices and CC-matrices in their paper [FZ07]. Let us recall them.

Conjecture 4.1.

For the totally sign-skew-symmetric matrix pattern with Bt0=BB_{t_{0}}=B, the following statements are true.

  1. (i)

    For each t𝕋nt\in\mathbb{T}_{n}, the CC-matrix CtC_{t} has the column sign-coherent property.

  2. (ii)

    For each t𝕋nt\in\mathbb{T}_{n}, the gg-vectors g1,t,,gn,tg_{1,t},\dots,g_{n,t} form a \mathbb{Z}-basis of the lattice n\mathbb{Z}^{n}.

  3. (iii)

    For each t𝕋nt\in\mathbb{T}_{n}, the GG-matrix GtG_{t} has the row sign-coherent property.

  4. (iv)

    Let t0kt1t_{0}\frac{k}{\quad\,}t_{1} be two adjacent vertices in 𝕋n\mathbb{T}_{n}, and let B1=μk(B)B^{1}=\mu_{k}(B). Then, for any t𝕋nt\in\mathbb{T}_{n} and 𝐚n\mathbf{a}\in\mathbb{N}^{n}, then gg-vectors g𝐚,tB,t0:=l=1nalgl,tB,t0=(g1,,gn)Tg_{\mathbf{a},t}^{B,t_{0}}:=\sum_{l=1}^{n}a_{l}g_{l,t}^{B,t_{0}}=(g_{1},\dots,g_{n})^{T} and g𝐚,tB1,t1:=l=1nalgl,tB1,t1=(g1,,gn)Tg_{\mathbf{a},t}^{B^{1},t_{1}}:=\sum_{l=1}^{n}a_{l}g_{l,t}^{B^{1},t_{1}}=(g^{\prime}_{1},\dots,g^{\prime}_{n})^{T} are related as follows:

    gj={gkj=k;gj+[bjk]+gkbjk0min(gk,0)jk.g^{\prime}_{j}=\begin{cases}-g_{k}\quad&j=k;\\ g_{j}+[b_{jk}]_{+}g_{k}-b_{jk}^{0}\mathrm{min}(g_{k},0)\quad&j\neq k.\end{cases}

In the skew-symmetrizable case, these conjectures are proved by a series of papers [DWZ10, GHKK18]. In the totally sign-skew-symmetric case, the Conjectures (i-iv) were proved by Huang and Li [HL18] and by Cao, Huang and Li [CHL22] for the acyclic sign-skew-symmetric case using the folding theory and categorification, and the Conjecture 4.1 (i) was proved by Li and Pan in full generality. In the following sections, we show that under an assumption Conjecture (ii-iv) are true. In particular, for the acyclic sign-skew-symmetric matrix, we obtain these results. Moreover, we show the duality between GG-matrices and CC-matrices, which has been only proved for the skew-symmetrizable case by Nakanishi and Zelevinsky [NZ12]. The essential idea of the proofs is developed by Nakanishi and Zelevinsky.

4.1. Mutations revisited

The mutations of GG-matrices and CC-matrices are given in the Section 2. Now we re-write them in the matrix form. Let 𝕋n\mathbb{T}_{n} be the nn-regular tree and BB be a totally sign-skew-symmetric matrix. Then we have an exchange pattern of matrices such that Bt0=BB_{t_{0}}=B. Now for each vertex t𝕋nt\in\mathbb{T}_{n}, let us replace BtB_{t} by a triple (Bt,Ct,Gt)(B_{t},C_{t},G_{t}), where CtC_{t} and GtG_{t} are respectively the corresponding GG-matrix and CC-matrices at tt with respect to the initial vertex t0t_{0}. Then we obtain a new pattern and we call it the matrix pattern.

Let us introduce some notations we will need in the following. Assume that A=(aij)n×nA=(a_{ij})_{n\times n} is a real matrix, and k[1,n]k\in[1,n].

  1. (1)

    [A]+[A]_{+} denotes the n×nn\times n matrix obtained from AA by replacing aija_{ij}by max{aij,0}max\{a_{ij},0\} for all i,ji,j.

  2. (2)

    AkA^{k\cdot} denotes the n×nn\times n matrix obtained from AA by replacing aija_{ij} for iki\neq k with 0.

  3. (3)

    AkA^{\cdot k} denotes the n×nn\times n matrix obtained from AA by replacing aija_{ij} for jkj\neq k with 0.

  4. (4)

    JkJ_{k} denotes the n×nn\times n matrix obtained from the identity matrix InI_{n} by replacing the kk-th diagonal element 11 with 1-1.

Proposition 4.2 ([FZ07, NZ12]).

Let B=Bt0B=B_{t_{0}} be a totally sign-skew-symmetric matrix. In the matrix pattern, for each tktt\frac{k}{\quad\,}t^{\prime} in 𝕋n\mathbb{T}_{n}, we have that

Ct=Ct(Jk+[εBt]+k)+[εCt]+kBt,C_{t^{\prime}}=C_{t}(J_{k}+[\varepsilon B_{t}]_{+}^{k\cdot})+[-\varepsilon C_{t}]_{+}^{\cdot k}B_{t},
Gt=Gt(Jk+[εBt]+k)Bt0[εCt]k,G_{t^{\prime}}=G_{t}(J_{k}+[-\varepsilon B_{t}]_{+}^{\cdot k})-B_{t_{0}}[-\varepsilon C_{t}]^{\cdot k},

where ε{1,1}\varepsilon\in\{1,-1\}.

We shall notice that the so-called the first duality always holds for the totally sign-skew-symmetric matrix pattern.

Proposition 4.3 ([FZ07]).

For each t𝕋nt\in\mathbb{T}_{n}, we have the following equality:

GtBt=Bt0Ct.G_{t}B_{t}=B_{t_{0}}C_{t}.
Proof.

This equality was proved by Fomin and Zelevinsky in [FZ07] using the fact that gg-vectors are the n\mathbb{Z}^{n}-grading of cluster variables coming from cluster algebras with principal coefficients. Later, Nakanishi proved it using the Proposition 4.2 and induction on the distance between t0t_{0} and tt on 𝕋n\mathbb{T}_{n} in [Nak]. Their proofs all hold for the totally sign-skew-symmetric case. ∎

In Li and Pan’s latest paper [LP], they proved the following theorem which affirms the column sign-coherence for all CC-matrices arising from totally sign-skew-symmetric matrices.

Theorem 4.4 ([LP]).

Let BB be a totally sign-skew-symmetric matrix and 𝒜(B)\mathcal{A}(B) be the corresponding cluster algebras with principal coefficients. Then every FF-polynomial has a constant term 11. Consequently, the CC-matrices are column sign-coherent.

Since the column sign-coherence holds for CC-matrices, we can use εk(C)\varepsilon_{k}(C) to denote the sign of the kk-th column of the CC-matrix CC. Then we have the following result.

Proposition 4.5.

Let B=Bt0B=B_{t_{0}} be a totally sign-skew-symmetric matrix. In the matrix pattern, for each tktt\frac{k}{\quad\,}t^{\prime} in 𝕋n\mathbb{T}_{n}, we have that

Ct=Ct(Jk+[εBt]+k),C_{t^{\prime}}=C_{t}(J_{k}+[\varepsilon B_{t}]_{+}^{k\cdot}),
Gt=Gt(Jk+[εBt]+k),G_{t^{\prime}}=G_{t}(J_{k}+[-\varepsilon B_{t}]_{+}^{\cdot k}),

where ε=εk(Ct)\varepsilon=\varepsilon_{k}(C_{t}).

Proof.

By the sign-coherence of the CC-matrix, we know that [ε(Ct)Ct]+k=0[-\varepsilon(C_{t})C_{t}]_{+}^{\cdot k}=0. Then this proposition is an easy corollary of the Proposition 4.2. ∎

Now we can see that Conjecture 4.1 (ii) always holds for totally sign-skew-symmetric matrix pattern.

Theorem 4.6.

For each t𝕋nt\in\mathbb{T}_{n}, we have that

  1. (i)

    the gg-vectors g1,t,,gn,tg_{1,t},\dots,g_{n,t} form a \mathbb{Z}-basis of the lattice n\mathbb{Z}^{n},

  2. (ii)

    the cc-vectors c1,t,,cn,tc_{1,t},\dots,c_{n,t} form a \mathbb{Z}-basis of the lattice n\mathbb{Z}^{n}.

Moreover, we have that |Gt|=|Ct|{1,1}|G_{t}|=|C_{t}|\in\{1,-1\}.

Proof.

Let us prove |Gt|=|Ct|{1,1}|G_{t}|=|C_{t}|\in\{1,-1\} by induction on the distance d(t0,t)d(t_{0},t) between t0t_{0} and tt on 𝕋n\mathbb{T}_{n}.

For d(t0,t)=0d(t_{0},t)=0, then t=t0t=t_{0}, we have that Gt=In=CtG_{t}=I_{n}=C_{t} and thus |Gt|=|Ct|=1|G_{t}|=|C_{t}|=1. Assume that it holds for d(t0,t)=dd(t_{0},t)=d and let tktt\frac{k}{\quad\,}t^{\prime} be an edge on 𝕋n\mathbb{T}_{n}. Then by Proposition 4.5, we have that

Ct=Ct(Jk+[εBt]+k),C_{t^{\prime}}=C_{t}(J_{k}+[\varepsilon B_{t}]_{+}^{k\cdot}),
Gt=Gt(Jk+[εBt]+k),G_{t^{\prime}}=G_{t}(J_{k}+[-\varepsilon B_{t}]_{+}^{\cdot k}),

where ε=εk(Ct)\varepsilon=\varepsilon_{k}(C_{t}). Note that

(Jk+[εBt]+k)2=In=(Jk+[εBt]+k)2,(J_{k}+[\varepsilon B_{t}]_{+}^{k\cdot})^{2}=I_{n}=(J_{k}+[-\varepsilon B_{t}]_{+}^{\cdot k})^{2},

and

|Jk+[εBt]+k|=1=|Jk+[εBt]+k|,|J_{k}+[\varepsilon B_{t}]_{+}^{k\cdot}|=-1=|J_{k}+[-\varepsilon B_{t}]_{+}^{\cdot k}|,

and |Gt|=|Ct|{1,1}|G_{t}|=|C_{t}|\in\{1,-1\}. We have that GtG_{t^{\prime}} and CtC_{t^{\prime}} are invertible in Mn()M_{n}(\mathbb{Z}), and |Gt|=|Ct|{1,1}|G_{t^{\prime}}|=|C_{t^{\prime}}|\in\{1,-1\}. ∎

4.2. Sign-coherence and tropical dualities

Let BB be a totally sign-skew-symmetric matrix and (Bt,Ct,Gt)t𝕋n(B_{t},C_{t},G_{t})_{t\in\mathbb{T}_{n}} be the corresponding matrix pattern with the initial vertex t0t_{0}. It is clear that BTB^{T} is also totally sign-skew-symmetric. The corresponding GG-matrices and CC-matrices for BTB^{T} are denoted by G¯\bar{G} and C¯\bar{C} respectively. Note that B~=BT\tilde{B}=-B^{T} is also a totally sign-skew-symmetric matrix. Let (B~t,C~t,G~t)t𝕋n(\tilde{B}_{t},\tilde{C}_{t},\tilde{G}_{t})_{t\in\mathbb{T}_{n}} denote the matrix pattern corresponding to the matrix B~=BT\tilde{B}=-B^{T} with the initial vertex t0t_{0}. It is easy to see that B~t=BtT\tilde{B}_{t}=-B_{t}^{T} for all t𝕋nt\in\mathbb{T}_{n}.

Assumption for BB: The signs of the columns of CC-matrices CC and C~\tilde{C} are always the same for arbitrary initial vertex for the cluster algebra 𝒜(B)\mathcal{A}(B) and for the cluster algebra 𝒜(B)\mathcal{A}(-B).

Note that Conjecture 3.13 implies the Assumption, thus Assumption holds for acyclic sign-skew-symmetric matrices.

Proposition 4.7.

Suppose the Assumption holds. For any t0,t𝕋nt_{0},t\in\mathbb{T}_{n}, we have that

(Gtt0)TC~tt0=In.(G_{t}^{t_{0}})^{T}\tilde{C}_{t}^{t_{0}}=I_{n}.
Proof.

Let us prove it by induction on the distance d(t0,t)d(t_{0},t) of t0t_{0} and tt on 𝕋n\mathbb{T}_{n}.

It is obvious when d(t0,t)=0d(t_{0},t)=0, since in this case Gtt0=C~tt0=InG_{t}^{t_{0}}=\tilde{C}_{t}^{t_{0}}=I_{n} Suppose that d(t0,t)=sd(t_{0},t)=s. Let us consider the following sequence:

t0tktt_{0}\frac{}{\quad\quad\quad\quad\quad\quad}t\frac{k}{\quad\,}t^{\prime}

By the Proposition 4.5, we have

(Gt)T=(Gt(Jk+[εk(Ct)Bt]+k))T=(Jk+[εk(Ct)(Bt)T]+k)GtT,(G_{t^{\prime}})^{T}=(G_{t}(J_{k}+[-\varepsilon_{k}(C_{t})B_{t}]_{+}^{\cdot k}))^{T}=(J_{k}+[-\varepsilon_{k}(C_{t})(B_{t})^{T}]_{+}^{k\cdot})G_{t}^{T},
C~t=C~t(Jk+[εk(C~t)(Bt)T]+k).\tilde{C}_{t^{\prime}}=\tilde{C}_{t}(J_{k}+[-\varepsilon_{k}(\tilde{C}_{t})(B_{t})^{T}]_{+}^{k\cdot}).

Then

(Gt)TC~t=(Jk+[εk(Ct)(Bt)T]+k)GtTC~t(Jk+[εk(C~t)(Bt)T]+k)=(Jk+[εk(Ct)(Bt)T]+k)(Jk+[εk(C~t)(Bt)T]+k)=In.\begin{split}(G_{t^{\prime}})^{T}\tilde{C}_{t^{\prime}}&=(J_{k}+[-\varepsilon_{k}(C_{t})(B_{t})^{T}]_{+}^{k\cdot})G_{t}^{T}\tilde{C}_{t}(J_{k}+[-\varepsilon_{k}(\tilde{C}_{t})(B_{t})^{T}]_{+}^{k\cdot})\\ &=(J_{k}+[-\varepsilon_{k}(C_{t})(B_{t})^{T}]_{+}^{k\cdot})(J_{k}+[-\varepsilon_{k}(\tilde{C}_{t})(B_{t})^{T}]_{+}^{k\cdot})\\ &=I_{n}.\end{split}

Here the third equality holds because the Assumption ensures that εk(Ct)=εk(C~t).\varepsilon_{k}(C_{t})=\varepsilon_{k}(\tilde{C}_{t}). Then we finish the proof. ∎

The row sign-coherence for GG-matrices was proved for the skew-symmetric case by Derksen, Weyman and Zelevinsky in [DWZ10], for the skew-symmetrizable case by Gross, Hacking, Keel and Kontsevich in [GHKK18], and for the acyclic sign-skew-symmetric case in [CHL22]. Next, we will prove the row sign-coherence for GG-matrices under the Assumption. The idea of the following proof comes from the proofs in [NZ12] with a little modification.

Proposition 4.8.

Suppose the Assumption holds. Then we have the following results:

  1. (a)

    For any t0,t𝕋nt_{0},t\in\mathbb{T}_{n}, we have

    (4.1) Ctt0=(G¯t0t)T,C_{t}^{t_{0}}=(\bar{G}_{t_{0}}^{t})^{T},
    (4.2) Gtt0=(C¯t0t)T.G_{t}^{t_{0}}=(\bar{C}_{t_{0}}^{t})^{T}.

    This implies that the row sign-coherence holds for GG-matrices, and we use εk(G)\varepsilon_{k}(G) to denote the sign of the kk-th row of the GG-matrix GG.

  2. (b)

    For any t0,t1,t𝕋nt_{0},t_{1},t\in\mathbb{T}_{n} such that t0kt1t_{0}\frac{k}{\quad\,}t_{1} are two adjacent vertices on 𝕋n\mathbb{T}_{n}, we have that

    (4.3) Ctt1=(Jk+[εk(Gtt0)Bt0]+k)Ctt0,C_{t}^{t_{1}}=(J_{k}+[-\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}]^{k\cdot}_{+})C^{t_{0}}_{t},
    (4.4) Gtt1=(Jk+[εk(Gtt0)Bt0]+k)Gtt0.G_{t}^{t_{1}}=(J_{k}+[\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}]^{\cdot k}_{+})G^{t_{0}}_{t}.

    Here εk(Gtt0)\varepsilon_{k}(G^{t_{0}}_{t}) is well-defined when (a) holds.

  3. (c)

    For any t0,t𝕋nt_{0},t\in\mathbb{T}_{n} and i,j[1,n]i,j\in[1,n], we have that the ii-th column of Ctt0C_{t}^{t_{0}} is ±ej\pm e_{j} if and only if the the ii-th column of C~tt0\tilde{C}_{t}^{t_{0}} is ±ej\pm e_{j}.

  4. (d)

    For any t0,t𝕋nt_{0},t\in\mathbb{T}_{n} and i,j[1,n]i,j\in[1,n], we have that the signs of the ii-th rows of Gtt0G_{t}^{t_{0}} and Gtt0G_{t}^{t_{0}} are the same, and the ii-th row of Gtt0G_{t}^{t_{0}} is ±ej\pm e_{j} if and only if the the ii-th row of G~tt0\tilde{G}_{t}^{t_{0}} is ±ej\pm e_{j}.

Proof.

We will prove these statements by induction on the distance d(t0,t)d(t_{0},t) of t0t_{0} and tt on 𝕋n\mathbb{T}_{n}.

Let us write the following claims:

  1. (a)s(a)_{s}.

    The claim (a)(a) holds for any t0,t𝕋nt_{0},t\in\mathbb{T}_{n} such that d(t0,t)sd(t_{0},t)\leq s.

  2. (b)s(b)_{s}.

    The claim (b)(b) holds for any t0,t1,t𝕋nt_{0},t_{1},t\in\mathbb{T}_{n} such that d(t0,t)sd(t_{0},t)\leq s, and t1kt0t_{1}\frac{k}{\quad\,}t_{0} are two adjacent vertices on 𝕋\mathbb{T} labeled by any k[1,n]k\in[1,n].

  3. (c)s(c)_{s}.

    The claim (c)(c) holds for any t0,t𝕋nt_{0},t\in\mathbb{T}_{n} such that d(t0,t)sd(t_{0},t)\leq s.

  4. (d)s(d)_{s}.

    The claim (d)(d) holds for any t0,t𝕋nt_{0},t\in\mathbb{T}_{n} such that d(t0,t)sd(t_{0},t)\leq s.

Let us prove these claims in the following order.

(4.5) (a)0(b)0(c)0(d)0(a)1(b)1(a)_{0}\implies(b)_{0}\implies(c)_{0}\implies(d)_{0}\implies(a)_{1}\implies(b)_{1}\implies\dots

Firstly, we show that (a)0,(b)0,(c)0(a)_{0},(b)_{0},(c)_{0} and (d)0(d)_{0} hold. Since s=0s=0 implies that t=t0t=t_{0}, thus we have that

Ctt0=G¯t0t=Gtt0=C¯t0t=C~tt0=G~tt0=In,C_{t}^{t_{0}}=\bar{G}_{t_{0}}^{t}=G_{t}^{t_{0}}=\bar{C}_{t_{0}}^{t}=\tilde{C}_{t}^{t_{0}}=\tilde{G}_{t}^{t_{0}}=I_{n},

and therefore (a)0,(c)0,(d)0(a)_{0},(c)_{0},(d)_{0} hold naturally. For (b)0(b)_{0}, we need to show that

Ct0t1=Jk+[Bt0]+k,C_{t_{0}}^{t_{1}}=J_{k}+[-B_{t_{0}}]^{k\cdot}_{+},
Gt0t1=Jk+[Bt0]+k.G_{t_{0}}^{t_{1}}=J_{k}+[B_{t_{0}}]^{\cdot k}_{+}.

By the Proposition 4.5, we have that

Ct0t1=Ct1t1(Jk+[Bt1]+k)=Jk+[Bt0]+k,C_{t_{0}}^{t_{1}}=C_{t_{1}}^{t_{1}}(J_{k}+[B_{t_{1}}]_{+}^{k\cdot})=J_{k}+[-B_{t_{0}}]^{k\cdot}_{+},
Gt0t1=Gt1t1(Jk+[Bt1]+k)=Jk+[Bt0]+k.G_{t_{0}}^{t_{1}}=G_{t_{1}}^{t_{1}}(J_{k}+[-B_{t_{1}}]_{+}^{\cdot k})=J_{k}+[B_{t_{0}}]^{\cdot k}_{+}.

Thus (b)0(b)_{0} holds.

Assume that the claims in 4.5 up to (d)s(d)_{s}, then we show (a)s+1(a)_{s+1}. Suppose that d(t0,t)=sd(t_{0},t)=s. Let us consider the following sequence:

t1kt0tt_{1}\frac{k}{\quad\,}t_{0}\frac{}{\quad\quad\quad\quad\quad\quad}t

We need to prove that

Ctt1=(G¯t1t)T,Gtt1=(C¯t1t)T.C_{t}^{t_{1}}=(\bar{G}_{t_{1}}^{t})^{T},\quad\quad G_{t}^{t_{1}}=(\bar{C}_{t_{1}}^{t})^{T}.

By (a)s(a)_{s} and (b)s(b)_{s}, we have that

Ctt0=(G¯t0t)T,Gtt0=(C¯t0t)T,C_{t}^{t_{0}}=(\bar{G}_{t_{0}}^{t})^{T},\quad\quad G_{t}^{t_{0}}=(\bar{C}_{t_{0}}^{t})^{T},
Ctt1=(Jk+[εk(Gtt0)Bt0]+k)Ctt0,C_{t}^{t_{1}}=(J_{k}+[-\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}]^{k\cdot}_{+})C^{t_{0}}_{t},
Gtt1=(Jk+[εk(Gtt0)Bt0]+k)Gtt0.G_{t}^{t_{1}}=(J_{k}+[\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}]^{k\cdot}_{+})G^{t_{0}}_{t}.

Note that εk(Gtt0)=ε(Ct0t)\varepsilon_{k}(G^{t_{0}}_{t})=\varepsilon(C^{t}_{t_{0}}). Then take the transpose, we have

(Ctt1)T=(Ctt0)T(Jk+[εk(Gtt0)Bt0T]+k)=G¯t0t(Jk+[εk(Ct0t)Bt0T]+k)=G¯t1t.\begin{split}(C_{t}^{t_{1}})^{T}&=(C^{t_{0}}_{t})^{T}(J_{k}+[-\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}^{T}]^{\cdot k}_{+})\\ &=\bar{G}_{t_{0}}^{t}(J_{k}+[-\varepsilon_{k}(C^{t}_{t_{0}})B_{t_{0}}^{T}]^{\cdot k}_{+})\\ &=\bar{G}_{t_{1}}^{t}.\end{split}

The third equality follows from the Proposition 4.5. And similarly, we have that

(Gtt1)T=(Gtt0)T(Jk+[εk(Gtt0)Bt0T]+k)=C¯t0t(Jk+[εk(Ct0t)Bt0T]+k)=C¯t1t.\begin{split}(G_{t}^{t_{1}})^{T}&=(G^{t_{0}}_{t})^{T}(J_{k}+[\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}^{T}]^{k\cdot}_{+})\\ &=\bar{C}_{t_{0}}^{t}(J_{k}+[\varepsilon_{k}(C^{t}_{t_{0}})B_{t_{0}}^{T}]^{k\cdot}_{+})\\ &=\bar{C}_{t_{1}}^{t}.\end{split}

Hence (a)s+1(a)_{s+1} holds.

Now assume that the claims in 4.5 up to (a)s+1(a)_{s+1}, then we show (b)s+1(b)_{s+1}. Assume that d(t0,t)=sd(t_{0},t)=s. Let us consider the following sequence:

t1kt0tltt_{1}\frac{k}{\quad\,}t_{0}\frac{}{\quad\quad\quad\quad\quad\quad}t\frac{l}{\quad\,}t^{\prime}

We need to prove that

(4.6) Ctt1=(Jk+[εk(Gtt0)Bt0]+k)Ctt0,C_{t^{\prime}}^{t_{1}}=(J_{k}+[-\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})B_{t_{0}}]^{k\cdot}_{+})C^{t_{0}}_{t^{\prime}},
(4.7) Gtt1=(Jk+[εk(Gtt0)Bt0]+k)Gtt0.G_{t^{\prime}}^{t_{1}}=(J_{k}+[\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})B_{t_{0}}]^{\cdot k}_{+})G^{t_{0}}_{t^{\prime}}.

To prove the equality 4.6, we need to prove that

Ctt1=(Jk+[εk(Gtt0)Bt0]+k)Ctt0(Jl+[εl(Ctt0Bt)]+l).C_{t^{\prime}}^{t_{1}}=(J_{k}+[-\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})B_{t_{0}}]^{k\cdot}_{+})C^{t_{0}}_{t}(J_{l}+[\varepsilon_{l}(C_{t}^{t_{0}}B_{t})]_{+}^{l\cdot}).

Note that we have that

Ctt1=Ctt1(Jl+[εl(Ctt1)Bt]+l)=(Jk+[εk(Gtt0)Bt0]+k)Ctt0(Jl+[εl(Ctt1)Bt]+l).\begin{split}C_{t^{\prime}}^{t_{1}}&=C_{t}^{t_{1}}(J_{l}+[\varepsilon_{l}(C^{t_{1}}_{t})B_{t}]^{l\cdot}_{+})\\ &=(J_{k}+[-\varepsilon_{k}(G_{t}^{t_{0}})B_{t_{0}}]_{+}^{k\cdot})C_{t}^{t_{0}}(J_{l}+[\varepsilon_{l}(C^{t_{1}}_{t})B_{t}]^{l\cdot}_{+}).\end{split}

The second equality follows from (b)s(b)_{s}. So it is enough to prove that

(Jk+[εk(Gtt0)Bt0]+k)Ctt0(Jl+[εl(Ctt0Bt)]+l)=(Jk+[εk(Gtt0)Bt0]+k)Ctt0(Jl+[εl(Ctt1)Bt]+l).\begin{split}&(J_{k}+[-\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})B_{t_{0}}]^{k\cdot}_{+})C^{t_{0}}_{t}(J_{l}+[\varepsilon_{l}(C_{t}^{t_{0}}B_{t})]_{+}^{l\cdot})\\ =&(J_{k}+[-\varepsilon_{k}(G_{t}^{t_{0}})B_{t_{0}}]_{+}^{k\cdot})C_{t}^{t_{0}}(J_{l}+[\varepsilon_{l}(C^{t_{1}}_{t})B_{t}]^{l\cdot}_{+}).\end{split}

Now let us discuss the signs in detail. We shall notice that by the mutation of GG-matrices, Gtt0G^{t_{0}}_{t^{\prime}} and Gtt0G^{t_{0}}_{t} differ only in the ll-th column, and thus we have the following result:

  1. (1)

    εk(Gtt0)=εk(Gtt0)\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})=\varepsilon_{k}(G^{t_{0}}_{t}) if and only if there is a nonzero entry in the kk-th row of Gtt0G^{t_{0}}_{t} outside the entry at (k,l)(k,l),

  2. (2)

    εk(Gtt0)=εk(Gtt0)\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})=-\varepsilon_{k}(G^{t_{0}}_{t}) if and only if the kk-th row of Gtt0G^{t_{0}}_{t} is ±el\pm e_{l}.

On the other hand, by (b)s(b)_{s}, we know that Ctt0C_{t}^{t_{0}} and Ctt1C_{t}^{t_{1}} differ only in the kk-th row, and thus we have the following result:

  1. (3)

    εl(Ctt1)=εl(Ctt0)\varepsilon_{l}(C^{t_{1}}_{t})=\varepsilon_{l}(C^{t_{0}}_{t}) if and only if there is a nonzero entry in the ll-th column of Ctt0C^{t_{0}}_{t} outside the entry at (l,k)(l,k),

  2. (4)

    εl(Ctt1)=εl(Ctt0)\varepsilon_{l}(C^{t_{1}}_{t})=-\varepsilon_{l}(C^{t_{0}}_{t}) if and only if the ll-th column of Ctt0C^{t_{0}}_{t} is ±ek\pm e_{k}.

If εk(Gtt0)=εk(Gtt0)\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})=\varepsilon_{k}(G^{t_{0}}_{t}), we need to show that εl(Ctt1)=εl(Ctt0)\varepsilon_{l}(C^{t_{1}}_{t})=\varepsilon_{l}(C^{t_{0}}_{t}). Indeed, if εl(Ctt1)=εl(Ctt0)\varepsilon_{l}(C^{t_{1}}_{t})=-\varepsilon_{l}(C^{t_{0}}_{t}), then the ll-th column of Ctt0C^{t_{0}}_{t} is ±ek\pm e_{k}, and by (c)s(c)_{s}, we know that the ll-th column of C~tt0\tilde{C}^{t_{0}}_{t} is ±ek\pm e_{k}. By Proposition 4.7, (Gtt0)TC~tt0=In(G^{t_{0}}_{t})^{T}\tilde{C}^{t_{0}}_{t}=I_{n}, this implies that the kk-th row of Gtt0G^{t_{0}}_{t} is ±el\pm e_{l}, and hence εk(Gtt0)=εk(Gtt0)\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})=\varepsilon_{k}(G^{t_{0}}_{t}), which is a contradiction. Thus if εk(Gtt0)=εk(Gtt0)\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})=\varepsilon_{k}(G^{t_{0}}_{t}), the equality 4.6 holds.

If εk(Gtt0)=εk(Gtt0)\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})=-\varepsilon_{k}(G^{t_{0}}_{t}), then the kk-th row of Gtt0G^{t_{0}}_{t} is ±el\pm e_{l}. Thus by Proposition 4.7 and the Assumption, we know that ll-th column of Ctt0C^{t_{0}}_{t} is ±ek\pm e_{k}. Notice that in this case, we have εk(Gtt0)=εl(Ctt0)\varepsilon_{k}(G^{t_{0}}_{t})=\varepsilon_{l}(C^{t_{0}}_{t}), and hence εl(Ctt1)=εl(Ctt0)\varepsilon_{l}(C^{t_{1}}_{t})=-\varepsilon_{l}(C^{t_{0}}_{t}). Thus we may let εk(Gtt0)=εl(Ctt0)=ε{1,1}\varepsilon_{k}(G^{t_{0}}_{t})=\varepsilon_{l}(C^{t_{0}}_{t})=\varepsilon\in\{1,-1\}, and εk(Gtt0)=εl(Ctt1)=ε\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})=\varepsilon_{l}(C^{t_{1}}_{t})=-\varepsilon. To prove the equality 4.6, it suffices to show that

(Jk+[εBt0]+k)Ctt0(Jl+[εBt)]+l)=(Jk+[εBt0]+k)Ctt0(Jl+[εBt]+l),\begin{split}&(J_{k}+[\varepsilon B_{t_{0}}]^{k\cdot}_{+})C^{t_{0}}_{t}(J_{l}+[\varepsilon B_{t})]_{+}^{l\cdot})\\ =&(J_{k}+[-\varepsilon B_{t_{0}}]_{+}^{k\cdot})C_{t}^{t_{0}}(J_{l}+[-\varepsilon B_{t}]^{l\cdot}_{+}),\end{split}

which is equivalent to prove that

Ctt0(Jl+[εBt)]+l)(Jl+[εBt]+l)=(Jk+[εBt0]+k)(Jk+[εBt0]+k)Ctt0.\begin{split}&C^{t_{0}}_{t}(J_{l}+[\varepsilon B_{t})]_{+}^{l\cdot})(J_{l}+[-\varepsilon B_{t}]^{l\cdot}_{+})\\ =&(J_{k}+[\varepsilon B_{t_{0}}]^{k\cdot}_{+})(J_{k}+[-\varepsilon B_{t_{0}}]_{+}^{k\cdot})C_{t}^{t_{0}}.\end{split}

Then compute the above equality, we only need to show that

(Bt0Ctt0)k=Ctt0(Bt)l.(B_{t_{0}}C_{t}^{t_{0}})^{k\cdot}=C_{t}^{t_{0}}(B_{t})^{l\cdot}.

Note that in this case we have that

(Ctt0)l=εEkl=(Gtt0)k.(C_{t}^{t_{0}})^{\cdot l}=\varepsilon E_{kl}=(G_{t}^{t_{0}})^{k\cdot}.

Thus we have that

Ctt0(Bt)l=(Ctt0)lBt=(Gtt0)kBt=(Gtt0Bt)k.C_{t}^{t_{0}}(B_{t})^{l\cdot}=(C_{t}^{t_{0}})^{\cdot l}B_{t}=(G_{t}^{t_{0}})^{k\cdot}B_{t}=(G_{t}^{t_{0}}B_{t})^{k\cdot}.

Here the first equality and the third equality follow from the facts that for any n×nn\times n matrices PP and QQ, and any k[1,n]k\in[1,n], we have that

PkQ=PQk,PkQ=(PQ)k.P^{\cdot k}Q=PQ^{k\cdot},\quad\quad P^{k\cdot}Q=(PQ)^{k\cdot}.

By Proposition 4.3, we have that (Bt0Ctt0)k=(Gtt0Bt)k(B_{t_{0}}C_{t}^{t_{0}})^{k\cdot}=(G_{t}^{t_{0}}B_{t})^{k\cdot}, and (Bt0Ctt0)k=Ctt0(Bt)l(B_{t_{0}}C_{t}^{t_{0}})^{k\cdot}=C_{t}^{t_{0}}(B_{t})^{l\cdot}. This proves the equality 4.6 in this case.

Let us prove the equality 4.7. Similarly, it is enough to show that

(Jk+[εk(Gtt0)Bt0]+k)Gtt0(Jl+[εl(Ctt0Bt)]+l)=(Jk+[εk(Gtt0)Bt0]+k)Ctt0(Jl+[εl(Ctt1)Bt]+l).\begin{split}&(J_{k}+[\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})B_{t_{0}}]^{\cdot k}_{+})G^{t_{0}}_{t}(J_{l}+[-\varepsilon_{l}(C_{t}^{t_{0}}B_{t})]_{+}^{\cdot l})\\ =&(J_{k}+[\varepsilon_{k}(G_{t}^{t_{0}})B_{t_{0}}]_{+}^{\cdot k})C_{t}^{t_{0}}(J_{l}+[-\varepsilon_{l}(C^{t_{1}}_{t})B_{t}]^{\cdot l}_{+}).\end{split}

After similar discussion as above, the case of εk(Gtt0)=εk(Gtt0)\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})=\varepsilon_{k}(G^{t_{0}}_{t}) is similar. When εk(Gtt0)=εk(Gtt0)\varepsilon_{k}(G^{t_{0}}_{t^{\prime}})=-\varepsilon_{k}(G^{t_{0}}_{t}), as above, we only need to show that

(Bt0)kGtt0=(Gtt0Bt)l.(B_{t_{0}})^{\cdot k}G_{t}^{t_{0}}=(G_{t}^{t_{0}}B_{t})^{\cdot l}.

Similarly, we have that

(Bt0)kGtt0=Bt0(Gtt0)k=Bt0(Ctt0)l=(Bt0Ctt0)l=(Gtt0Bt)l.(B_{t_{0}})^{\cdot k}G_{t}^{t_{0}}=B_{t_{0}}(G_{t}^{t_{0}})^{k\cdot}=B_{t_{0}}(C_{t}^{t_{0}})^{\cdot l}=(B_{t_{0}}C_{t}^{t_{0}})^{\cdot l}=(G_{t}^{t_{0}}B_{t})^{\cdot l}.

Therefore, the equality 4.7 holds in this case. Then (b)s+1(b)_{s+1} holds.

Now assume that the claims in 4.5 up to (b)s+1(b)_{s+1}, then we show (c)s+1(c)_{s+1}. Suppose that d(t0,t)=s+1d(t_{0},t)=s+1. (We shall emphasis that here we assume that d(t0,t)=s+1d(t_{0},t)=s+1, while in the previous proof we often assume that d(t0,t)=sd(t_{0},t)=s.) For any i,j[1,n]i,j\in[1,n], we need to prove that the ii-th column of Ctt0C_{t}^{t_{0}} is ±ej\pm e_{j} if and only if the ii-th column of C~tt0\tilde{C}_{t}^{t_{0}} is ±ej\pm e_{j}. Consider the following sequence

tjjt0tt_{j}\frac{j}{\quad\,}t_{0}\frac{}{\quad\quad\quad\quad\quad\quad}t

Since d(t0,t)=s+1d(t_{0},t)=s+1 and (b)s+1(b)_{s+1} holds, we have that

(4.8) Ctt1=(Jk+[εk(Gtt0)Bt0]+k)Ctt0,C_{t}^{t_{1}}=(J_{k}+[-\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}]^{k\cdot}_{+})C^{t_{0}}_{t},
(4.9) C~tt1=(Jk+[εk(G~tt0)B~t0]+k)C~tt0.\tilde{C}_{t}^{t_{1}}=(J_{k}+[-\varepsilon_{k}(\tilde{G}^{t_{0}}_{t})\tilde{B}_{t_{0}}]^{k\cdot}_{+})\tilde{C}^{t_{0}}_{t}.

Then we have the following facts:

  1. (1)

    By the equality 4.8, we know that the ii-th column of Ctt0C_{t}^{t_{0}} is ±ej\pm e_{j} if and only if εi(Ctt0)=εi(Cttj)\varepsilon_{i}(C_{t}^{t_{0}})=-\varepsilon_{i}(C_{t}^{t_{j}}).

  2. (2)

    By the equality 4.9, we know that the ii-th column of C~tt0\tilde{C}_{t}^{t_{0}} is ±ej\pm e_{j} if and only if εi(C~tt0)=εi(C~ttj)\varepsilon_{i}(\tilde{C}_{t}^{t_{0}})=-\varepsilon_{i}(\tilde{C}_{t}^{t_{j}}).

  3. (3)

    By the Assumption, we know that εi(Ctt0)=εi(C~tt0)\varepsilon_{i}(C_{t}^{t_{0}})=\varepsilon_{i}(\tilde{C}_{t}^{t_{0}}) and εi(Cttj)=εi(C~ttj).\varepsilon_{i}(C_{t}^{t_{j}})=\varepsilon_{i}(\tilde{C}_{t}^{t_{j}}).

Therefore the ii-th column of Ctt0C_{t}^{t_{0}} is ±ej\pm e_{j} if and only if the ii-th column of C~tt0\tilde{C}_{t}^{t_{0}} is ±ej\pm e_{j}.

Now assume that the claims in 4.5 up to (c)s+1(c)_{s+1}, then we show (d)s+1(d)_{s+1}.

Suppose that d(t0,t)=s+1d(t_{0},t)=s+1. By (a)s+1(a)_{s+1}, we have that

Gtt0=(Ct0BtT,t)T,G~tt0=(Ct0Bt,t)T.G_{t}^{t_{0}}=(C_{t_{0}}^{B^{T}_{t},t})^{T},\quad\tilde{G}_{t}^{t_{0}}=(C_{t_{0}}^{-B_{t},t})^{T}.

By the Assumption, we know that the signs of the columns of the CC-matrices Ct0BtT,tC_{t_{0}}^{B^{T}_{t},t} and Ct0Bt,tC_{t_{0}}^{-B_{t},t} are the same. Thus the signs of the rows of the GG-matrices Gtt0G_{t}^{t_{0}} and G~tt0\tilde{G}_{t}^{t_{0}} are the same.

Note that the ii-th row of Gtt0G_{t}^{t_{0}} is ±ej\pm e_{j} if and only if the jj-th column of C~tt0\tilde{C}_{t}^{t_{0}} is ±ei\pm e_{i}, and the ii-th row of G~tt0\tilde{G}_{t}^{t_{0}} is ±ej\pm e_{j} if and only if the jj-th column of Ctt0C_{t}^{t_{0}} is ±ei\pm e_{i}. While by (c)s+1(c)_{s+1}, the jj-th column of C~tt0\tilde{C}_{t}^{t_{0}} is ±ei\pm e_{i} if and only if the jj-th column of Ctt0C_{t}^{t_{0}} is ±ei\pm e_{i}. Hence we have that the ii-th row of Gtt0G_{t}^{t_{0}} is ±ej\pm e_{j} if and only if the the ii-th row of G~tt0\tilde{G}_{t}^{t_{0}} is ±ej\pm e_{j}. This finishes the proof. ∎

Proposition 4.9.

Suppose that the Assumption holds. Let t0kt1t_{0}\frac{k}{\quad\,}t_{1} be two adjacent vertices in 𝕋n\mathbb{T}_{n}, and let B1=μk(B)B^{1}=\mu_{k}(B). Then, for any t𝕋nt\in\mathbb{T}_{n} and 𝐚n\mathbf{a}\in\mathbb{N}^{n}, then gg-vectors g𝐚,tB,t0:=l=1nalgl,tB,t0=(g1,,gn)Tg_{\mathbf{a},t}^{B,t_{0}}:=\sum_{l=1}^{n}a_{l}g_{l,t}^{B,t_{0}}=(g_{1},\dots,g_{n})^{T} and g𝐚,tB1,t1:=l=1nalgl,tB1,t1=(g1,,gn)Tg_{\mathbf{a},t}^{B^{1},t_{1}}:=\sum_{l=1}^{n}a_{l}g_{l,t}^{B^{1},t_{1}}=(g^{\prime}_{1},\dots,g^{\prime}_{n})^{T} are related as follows:

(4.10) gj={gkj=k;gj+[bjk]+gkbjkmin(gk,0)jk.g^{\prime}_{j}=\begin{cases}-g_{k}\quad&j=k;\\ g_{j}+[b_{jk}]_{+}g_{k}-b_{jk}min(g_{k},0)\quad&j\neq k.\end{cases}
Proof.

We may write the equation 4.10 in the matrix form as follows:

Gtt1=(Jk+[εB]+k)Gtt0+B[εGtt0]+k,G^{t_{1}}_{t}=(J_{k}+[\varepsilon B]_{+}^{\cdot k})G_{t}^{t_{0}}+B[-\varepsilon G_{t}^{t_{0}}]_{+}^{k\cdot},

where ε{1,1}\varepsilon\in\{1,-1\}. Thanks to Proposition 4.8 (a), we may let ε=εk(Gtt0)\varepsilon=\varepsilon_{k}(G_{t}^{t_{0}}), and by Proposition 4.8 (d), the equation 4.10 holds. ∎

Then we have the following results for acyclic cluster algebras.

Theorem 4.10.

If BB is mutation to an acyclic sign-skew-symmetric matrix. Then we have that

  1. (1)

    For each t𝕋nt\in\mathbb{T}_{n} and k[1,n]k\in[1,n], εk(Ct)=εk(C~t)\varepsilon_{k}(C_{t})=\varepsilon_{k}(\tilde{C}_{t}).

  2. (2)

    For any t𝕋nt\in\mathbb{T}_{n} and i,j[1,n]i,j\in[1,n], we have that the ii-th column of Ctt0C_{t}^{t_{0}} is ±ej\pm e_{j} if and only if the the ii-th column of C~tt0\tilde{C}_{t}^{t_{0}} is ±ej\pm e_{j}.

  3. (3)

    For any t0,t𝕋nt_{0},t\in\mathbb{T}_{n}, (Gtt0)TC~tt0=In.(G_{t}^{t_{0}})^{T}\tilde{C}_{t}^{t_{0}}=I_{n}.

  4. (4)

    For any t0,t1,t𝕋nt_{0},t_{1},t\in\mathbb{T}_{n} such that t0kt1t_{0}\frac{k}{\quad\,}t_{1} are two adjacent vertices on 𝕋n\mathbb{T}_{n}, we have that

    (4.11) Ctt1=(Jk+[εk(Gtt0)Bt0]+k)Ctt0,C_{t}^{t_{1}}=(J_{k}+[-\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}]^{k\cdot}_{+})C^{t_{0}}_{t},
    (4.12) Gtt1=(Jk+[εk(Gtt0)Bt0]+k)Gtt0.G_{t}^{t_{1}}=(J_{k}+[\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}]^{\cdot k}_{+})G^{t_{0}}_{t}.
  5. (5)

    For each t𝕋nt\in\mathbb{T}_{n} and i,j,k[1,n]i,j,k\in[1,n], εk(Gt)=εk(G~t)\varepsilon_{k}(G_{t})=\varepsilon_{k}(\tilde{G}_{t}), and the ii-th row of Gtt0G_{t}^{t_{0}} is ±ej\pm e_{j} if and only if the the ii-th row of G~tt0\tilde{G}_{t}^{t_{0}} is ±ej\pm e_{j}.

4.3. GG-fan and the mutation fan

In this subsection, we always assume that the Assumption holds. Let us recall some basic notions on cones and fans.

Definition 4.11.

A subset CnC\subset\mathbb{R}^{n} is called a convex cone if avCav\in C and v+vCv+v^{\prime}\in C for any a>0a\in\mathbb{R}_{>0} and v,vCv,v^{\prime}\in C. In particular, CC is called a convex polyhedral cone if there exists a finite number of elements v1,,vsCv_{1},\dots,v_{s}\in C such that

C=0v1++0vs:={a1v1++asvs|ai0,  1is}.C=\mathbb{R}_{\geq 0}v_{1}+\dots+\mathbb{R}_{\geq 0}v_{s}:=\{a_{1}v_{1}+\dots+a_{s}v_{s}\,|\,a_{i}\geq 0,\,\,1\leq i\leq s\}.

The convex polyhedral cone CC is called a simplicial cone, if v1,,vsv_{1},\dots,v_{s} can be chosen to be \mathbb{R}-linearly independent.

A convex cone is rational if it is non-negative \mathbb{R}-linear span of finitely many integer vectors, or equivalently, it can be defined by finitely many weak linear inequalities with integer coefficients.

Definition 4.12.

A subset FF of a convex set CC is a face if FF is convex and if any line segment LCL\subset C whose interior intersects FF has LFL\subset F.

Definition 4.13.

A fan \mathcal{F} in n\mathbb{R}^{n} is a collection of closed convex cones, such that

  1. (1)

    any face of a cone in \mathcal{F} is also a cone in \mathcal{F};

  2. (2)

    the intersection of any two cones in \mathcal{F} is a face of each.

In the Definition 4.13, we do not require that the fan is a finite set.

Definition 4.14.

For a totally sign-skew-symmetric matrix pattern {(Bt,Ct,Gt)}t𝕋n\{(B_{t},C_{t},G_{t})\}_{t\in\mathbb{T}_{n}}, the cone

σ(Gt):=0g1,t++0gn,t\sigma(G_{t}):=\mathbb{R}_{\geq 0}g_{1,t}+\dots+\mathbb{R}_{\geq 0}g_{n,t}

is called a GG-cone. The interior of σ(Gt)\sigma(G_{t}) is denoted by σ¯(Gt)\bar{\sigma}(G_{t}).

Since the GG-matrix GtG_{t} is invertible, we have that

σ¯(Gt):=>0g1,t++>0gn,t.\bar{\sigma}(G_{t}):=\mathbb{R}_{>0}g_{1,t}+\dots+\mathbb{R}_{>0}g_{n,t}.

For a totally sign-skew-symmetric matrix pattern {(Bt,Ct,Gt)}t𝕋n\{(B_{t},C_{t},G_{t})\}_{t\in\mathbb{T}_{n}} with the initial vertex t0t_{0}, let us introduce a notation as follows:

Gt0:={σ(Gtt0)and its faces,t𝕋n}.\mathcal{F}_{G^{t_{0}}}:=\{\sigma(G_{t}^{t_{0}})\,\,\text{and its faces},t\in\mathbb{T}_{n}\}.

For k[1,n]k\in[1,n], there are two half spaces of n\mathbb{R}^{n} defined as follows:

ek+:={w=(w1,,wn)n|wk0},ek:={w=(w1,,wn)n|wk0}.e_{k}^{+}:=\{w=(w_{1},\dots,w_{n})\in\mathbb{R}^{n}\,|\,w_{k}\geq 0\},\quad e_{k}^{-}:=\{w=(w_{1},\dots,w_{n})\in\mathbb{R}^{n}\,|\,w_{k}\leq 0\}.

Let t0kt1t_{0}\frac{k}{\quad\,}t_{1} be two adjacent vertices in 𝕋n\mathbb{T}_{n}. Define the piece-linear map ηt0t1\eta_{t_{0}}^{t_{1}} as follows:

ηt0t1:nnw{(Jk+[Bt0]+k)wwek+;(Jk+[Bt0]+k)wwek.\begin{split}\eta_{t_{0}}^{t_{1}}:&\mathbb{R}^{n}\rightarrow\quad\quad\mathbb{R}^{n}\\ &w\mapsto\begin{cases}(J_{k}+[B_{t_{0}}]_{+}^{\cdot k})w\quad&w\in e_{k}^{+};\\ (J_{k}+[-B_{t_{0}}]_{+}^{\cdot k})w\quad&w\in e_{k}^{-}.\end{cases}\end{split}

Note that ηt0t1\eta_{t_{0}}^{t_{1}} is well-defined, since if wk=0w_{k}=0, we have that (Jk+[Bt0]+k)w=w=(Jk+[Bt0]+k)w(J_{k}+[B_{t_{0}}]_{+}^{\cdot k})w=w=(J_{k}+[-B_{t_{0}}]_{+}^{\cdot k})w. By the Proposition 4.8, we have that for any t𝕋nt\in\mathbb{T}_{n},

Gtt1=(Jk+[εk(Gtt0)Bt0]+k)Gtt0,G_{t}^{t_{1}}=(J_{k}+[\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}]^{\cdot k}_{+})G^{t_{0}}_{t},

which means that

gi,tt1=(Jk+[εk(Gtt0)Bt0]+k)gi,tt0,i[1,n].g_{i,t}^{t_{1}}=(J_{k}+[\varepsilon_{k}(G^{t_{0}}_{t})B_{t_{0}}]^{\cdot k}_{+})g^{t_{0}}_{i,t},\forall i\in[1,n].

Thus we have that

ηt0t1(gi,tt0)=gi,tt1\eta_{t_{0}}^{t_{1}}(g^{t_{0}}_{i,t})=g_{i,t}^{t_{1}}

for any i[1,n]i\in[1,n] and t𝕋nt\in\mathbb{T}_{n}.

Proposition 4.15 (cf, [Nak, Rea14]).

The following statements hold.

  1. (1)

    ηt0t1ηt1t0=id=ηt1t0ηt0t1\eta_{t_{0}}^{t_{1}}\eta_{t_{1}}^{t_{0}}=id=\eta_{t_{1}}^{t_{0}}\eta_{t_{0}}^{t_{1}}.

  2. (2)

    ηt0t1(σ(Gtt0))=σ(Gtt1)\eta_{t_{0}}^{t_{1}}(\sigma(G_{t}^{t_{0}}))=\sigma(G_{t}^{t_{1}}).

  3. (3)

    the map

    ηt0t1:Gt0Gt0gi,tt0gi,tt1\begin{split}\eta_{t_{0}}^{t_{1}}:&\mathcal{F}_{G^{t_{0}}}\rightarrow\mathcal{F}_{G^{t_{0}}}\\ &g^{t_{0}}_{i,t}\mapsto g_{i,t}^{t_{1}}\end{split}

    is a bijection between Gt0\mathcal{F}_{G^{t_{0}}} and Gt0\mathcal{F}_{G^{t_{0}}} preserving the intersection and the inclusion of cones.

Proof.

By Proposition 4.8, we know that the row sign-coherence holds for GG-matrices, thus we have that either σ(Gtt0)ek+\sigma(G_{t}^{t_{0}})\subset e_{k}^{+} or σ(Gtt0)ek\sigma(G_{t}^{t_{0}})\subset e_{k}^{-}. Thus ηt0t1\eta_{t_{0}}^{t_{1}} is linear on σ(Gtt0)\sigma(G_{t}^{t_{0}}) and note that ηt0t1(gi,tt0)=gi,tt1\eta_{t_{0}}^{t_{1}}(g^{t_{0}}_{i,t})=g_{i,t}^{t_{1}}, we have that

ηt0t1(σ(Gtt0))=σ(Gtt1).\eta_{t_{0}}^{t_{1}}(\sigma(G_{t}^{t_{0}}))=\sigma(G_{t}^{t_{1}}).

Theorem 4.16 (cf.[Nak, Rea14]).

Suppose that the Assumption holds. Then the following set

Gt0:={σ(Gt)and its faces,t𝕋n}\mathcal{F}_{G^{t_{0}}}:=\{\sigma(G_{t})\,\,\text{and its faces},t\in\mathbb{T}_{n}\}

is a fan.

Proof.

The proof is similar to the proof of [Theorem 3.17, [Nak]]. ∎

Corollary 4.17.

For any t𝕋nt\in\mathbb{T}_{n} and i[1,n]i\in[1,n], if the gg-vector gi,tg_{i,t} is non-negative, then gi,t=ejg_{i,t}=e_{j} for some jj, and consequently, the cluster variable xi,tx_{i,t} satisfies that xi,t=xj,t0x_{i,t}=x_{j,t_{0}}.

Definition 4.18.

Let BB be a totally sign-skew-symmetric matrix, and let {(Bt,Ct,Gt)}t𝕋n\{(B_{t},C_{t},G_{t})\}_{t\in\mathbb{T}_{n}} be the corresponding matrix pattern. Two vectors w,wnw,w^{\prime}\in\mathbb{R}^{n} are said to be sign-equivalent (with respect to BB), if

sign(ηt0t(w))=sign(ηt0t(w))sign(\eta_{t_{0}}^{t}(w))=sign(\eta_{t_{0}}^{t}(w^{\prime}))

for any t𝕋nt\in\mathbb{T}_{n}. Here for wnw\in\mathbb{R}^{n}, sign(w):=(sign(w1),,sign(wn))sign(w):=(sign(w_{1}),\dots,sign(w_{n})).

Sign-equivalence defines an equivalence relation. We call the equivalence classes BB-classes.

Reading proved that the closure of any BB-class are convex cones.

Lemma 4.19 ([Rea14]).

The closures of BB-classes are convex cones.

We call them BB-cones.

Definition 4.20.

Let (B)\mathcal{MF}(B) be the collection consisting of all BB-cones, together with all faces of BB-cones. This collection (B)\mathcal{MF}(B) is called the mutation fan for BB.

Theorem 4.21 ([Rea14]).

The mutation fan (B)\mathcal{MF}(B) is a complete fan.

Next, following Reading, we show that the GG-fan is a subfan of the mutation fan. Note that we always assume that the Assumption holds for BB appearing in this section.

Definition 4.22.

Two full dimensional cones are adjacent if they have a common face of codimension 11 and they have disjoint interiors.

We say two full dimensional cones CC and CC^{\prime} in a fan \mathcal{F} are transitively adjacent if there is a sequence C=C0,C1,,Ck=CC=C_{0},C_{1},\dots,C_{k}=C^{\prime} of full-dimensional cones in \mathcal{F} such that Ci=1C_{i=1} and CiC_{i} are adjacent for all i[1,k]i\in[1,k]. The full dimensional cones in (B)\mathcal{MF}(B) that are transitively adjacent to nonnegative cone 0n\mathbb{R}^{n}_{\geq 0} in (B)\mathcal{MF}(B) are the maximal cones of a subfan ¯(B)\bar{\mathcal{MF}}(B) of (B)\mathcal{MF}(B) .

Theorem 4.23 (cf. [Rea14]).

For a totally sign-skew-symmetric matrix BB, the following fans are the same:

¯(B)=Gt0.\bar{\mathcal{MF}}(B)=\mathcal{F}_{G^{t_{0}}}.
Proof.

The proof is similar to the proof in [Rea14]. ∎

5. Exchange graphs

Recall that Fomin and Zelevinsky proposed the following conjecture:

Conjecture 5.1.

Every seed is uniquely defined by its cluster; thus, the vertices of the exchange graph can be identified with the clusters, up to a permutation of cluster variables.

Conjecture 5.2.

Two clusters are adjacent in the exchange graph if and only if they have exactly n1n-1 common cluster variables.

Conjecture 5.3.

Let 𝒜¯\bar{\mathcal{A}} be the cluster algebra with principal coefficients. Then each cluster (𝐱¯t,𝐲¯t,Bt)(\bar{\mathbf{x}}_{t},\bar{\mathbf{y}}_{t},B_{t}) is uniquely determined by the corresponding CC-matrix CtC_{t}.

These two conjectures were proved by Gekhtman, Shapiro, and Vainshtein for the skew-symmetrizable cluster algebras of geometric type [GSV08], and they had shown that if Conjecture 5.1 holds, then the Conjecture 5.2 holds. Cao and Li proved Conjecture 5.1 for generalized cluster algebras with any coefficients [CL18]. Cao, Huang and Li proved Conjecture 5.3 for the skew-symmetrizable cluster algebras [CHL22]. In this section, we prove these three conjectures for cluster algebras whose exchange matrices satisfying the Assumption and the positivity conjecture. These conditions hold for acyclic sign-skew-symmetric cluster algebras. We also give a proof of Conjecture 5.1 and Conjecture 5.2 for cluster algebras of odd rank and with indecomposable exchange matrix without the Assumption condition.

The following theorem is powerful, since it establishes the one-to-one correspondence between gg-vectors and cluster variables with any coefficients.

Theorem 5.4.

Let 𝒜\mathcal{A} be a totally sign-skew-symmetric cluster algebra with coefficients coming from an semifield \mathbb{P} satisfying the Assumption. Then for any t0𝕋nt_{0}\in\mathbb{T}_{n}, we have the following bijective map

{g-vectorsgi,tt0,i[1,n],t𝕋n}{cluster variables xi,t in 𝒜 ,i[1,n],t𝕋n}gi,tt0xi,t.\begin{split}\{\text{$g$-vectors}\,\,g_{i,t}^{t_{0}},\,\,i\in[1,n],\,\,t\in\mathbb{T}_{n}\}&\longrightarrow\{\text{cluster variables $x_{i,t}$ in $\mathcal{A}$ },\,\,i\in[1,n],\,\,t\in\mathbb{T}_{n}\}\\ g_{i,t}^{t_{0}}&\mapsto x_{i,t}.\end{split}
Proof.

See the proof of [Theorem 8.2, [Nak]] for the details. ∎

Notice that Theorem 5.4 was also proved in full generality for cluster algebras with principal coefficients by Li and Pan in their recent paper [LP].

Definition 5.5.

Let 𝒜\mathcal{A} be a totally sign-skew-symmetric cluster algebra with coefficients coming from an arbitrary semifield \mathbb{P}. The cluster complex 𝒞(𝒜)\mathcal{C}(\mathcal{A}) is the simplicial complex whose vertices are cluster variables and whose simplicies are non-empty subsets of clusters.

We have the following results:

Corollary 5.6.

Let 𝒜\mathcal{A} be a totally sign-skew-symmetric cluster algebra with coefficients coming from an arbitrary semifield \mathbb{P}. Then for any t0𝕋nt_{0}\in\mathbb{T}_{n}, we have an isomorphism of simplicial complices:

𝒞(𝒜)Gt0xi,tgi,tt0.\begin{split}\mathcal{C}(\mathcal{A})&\longrightarrow\mathcal{F}_{G^{t_{0}}}\\ x_{i,t}&\leftrightarrow g_{i,t}^{t_{0}}.\end{split}

Here Gt0\mathcal{F}_{G^{t_{0}}} is the corresponding GG-fan.

Before we prove the main results, let us make some preparations. Recall the following formula for the universal coefficients for cluster algebras with totally sign-skew-symmetric matrix.

Theorem 5.7 ([FZ07]).

Let 𝒜\mathcal{A} be a totally sign-skew-symmetric cluster algebra with coefficients coming from the universal semifield =sf(Y1,,Yn)\mathbb{P}=\mathbb{Q}_{sf}(Y_{1},\dots,Y_{n}). Then we have that for any j[1,n]j\in[1,n] and t𝕋nt\in\mathbb{T}_{n},

Yj,t=(i=1nYicij,t)i=1nFi,t(𝐘)bij,t,Y_{j,t}=(\prod_{i=1}^{n}Y_{i}^{c_{ij,t}})\prod_{i=1}^{n}F_{i,t}(\mathbf{Y})^{b_{ij,t}},

where Fi,tF_{i,t} are FF-polynomials.

For a permutation σSn\sigma\in S_{n} and a seed (𝐱t,𝐲t,Bt)(\mathbf{x}_{t},\mathbf{y}_{t},B_{t}), the action of σ\sigma on the seed (𝐱t,𝐲t,Bt)(\mathbf{x}_{t},\mathbf{y}_{t},B_{t}), is denoted by σ(𝐱t,𝐲t,Bt)\sigma(\mathbf{x}_{t},\mathbf{y}_{t},B_{t}) and given by

σ(𝐱t):=(xσ(1),t,,xσ(n),t),σ(𝐲t):=(yσ(1),t,,yσ(n),t),σ(Bt):=(bσ(i)σ(j),t).\sigma(\mathbf{x}_{t}):=(x_{\sigma(1),t},\dots,x_{\sigma(n),t}),\,\sigma(\mathbf{y}_{t}):=(y_{\sigma(1),t},\dots,y_{\sigma(n),t}),\sigma(B_{t}):=(b_{\sigma(i)\sigma(j),t}).
Theorem 5.8.

Let 𝒜\mathcal{A} be a totally sign-skew-symmetric cluster algebra with coefficients coming from an semifield \mathbb{P}. Suppose that the Assumption holds for the exchange matrix of 𝒜\mathcal{A}. Then we have that:

  1. (1)

    Every seed is uniquely defined by its cluster, i.e., if there is a permutation σSn\sigma\in S_{n} such that σ(𝐱t)=𝐱t\sigma(\mathbf{x}_{t^{\prime}})=\mathbf{x}_{t} for some t,t𝕋nt,t^{\prime}\in\mathbb{T}_{n}, then

    (𝐱t,𝐲t,Bt)=σ(𝐱t,𝐲t,Bt).(\mathbf{x}_{t},\mathbf{y}_{t},B_{t})=\sigma(\mathbf{x}_{t^{\prime}},\mathbf{y}_{t^{\prime}},B_{t^{\prime}}).
  2. (2)

    Two clusters are adjacent in the exchange graph if and only if they have exactly n1n-1 common cluster variables.

Proof.

(1) Assume that σ(𝐱t)=𝐱t\sigma(\mathbf{x}_{t^{\prime}})=\mathbf{x}_{t} for some t,t𝕋nt,t^{\prime}\in\mathbb{T}_{n}. We may let t=t0t^{\prime}=t_{0} for simplicity. Then by Theorem 5.4, we have that there is a bijective map

{g-vectorsgi,tt0,i[1,n],t𝕋n}{cluster variables xi,t in 𝒜 ,i[1,n],t𝕋n}gi,tt0xi,t.\begin{split}\{\text{$g$-vectors}\,\,g_{i,t}^{t_{0}},\,\,i\in[1,n],\,\,t\in\mathbb{T}_{n}\}&\longrightarrow\{\text{cluster variables $x_{i,t}$ in $\mathcal{A}$ },\,\,i\in[1,n],\,\,t\in\mathbb{T}_{n}\}\\ g_{i,t}^{t_{0}}&\mapsto x_{i,t}.\end{split}

Let us consider the cluster algebra 𝒜¯(𝐱¯,𝐲¯,B)\bar{\mathcal{A}}(\bar{\mathbf{x}},\bar{\mathbf{y}},B) with principal coefficients simultaneously.

Since σ(𝐱t0)=𝐱t\sigma(\mathbf{x}_{t_{0}})=\mathbf{x}_{t}, thus the GG-matrix Gtt0G^{t_{0}}_{t} is a permutation matrix PP (every column and every row of PP only have one element 11.) This implies that the FF-polynomials Fi,tt0F_{i,t}^{t_{0}} are 11, and the matrix C~tt0\tilde{C}_{t}^{t_{0}} is also PP, as PTP=InP^{T}P=I_{n} and (Gtt0)TC~tt0=In(G^{t_{0}}_{t})^{T}\tilde{C}_{t}^{t_{0}}=I_{n}. By the Proposition 4.8, we know that Ctt0=C~tt0=P{C}_{t}^{t_{0}}=\tilde{C}_{t}^{t_{0}}=P.

By the Proposition 4.3, we have that

Bt=(Gtt0)1Bt0Ctt0=PTBt0P=σ(Bt0).B_{t}=(G^{t_{0}}_{t})^{-1}B_{t_{0}}{C}_{t}^{t_{0}}=P^{T}B_{t_{0}}P=\sigma(B_{t_{0}}).

By the Theorem 5.7, we have that Yj,t=Yσ(j)Y_{j,t}=Y_{\sigma(j)} for each j[1,n]j\in[1,n], and thus yj,t=yσ(j)y_{j,t}=y_{\sigma(j)} for each j[1,n]j\in[1,n], i.e., 𝐲t=σ(𝐲t0).\mathbf{y}_{t}=\sigma(\mathbf{y}_{t_{0}}).

(2) It is a Corollary of (1). ∎

Next, we abandon the Assumption, and prove that Theorem 5.8 holds for all totally sign-skew-symmetric cluster algebras of odd rank.

Theorem 5.9.

Let 𝒜\mathcal{A} be a totally sign-skew-symmetric cluster algebra of odd rank with coefficients coming from an semifield \mathbb{P} and with indecomposable exchange matrix. Then we have that:

  1. (1)

    Every seed is uniquely defined by its cluster, i.e., if there is a permutation σSn\sigma\in S_{n} such that σ(𝐱t)=𝐱t\sigma(\mathbf{x}_{t^{\prime}})=\mathbf{x}_{t} for some t,t𝕋nt,t^{\prime}\in\mathbb{T}_{n}, then

    (𝐱t,𝐲t,Bt)=σ(𝐱t,𝐲t,Bt).(\mathbf{x}_{t},\mathbf{y}_{t},B_{t})=\sigma(\mathbf{x}_{t^{\prime}},\mathbf{y}_{t^{\prime}},B_{t^{\prime}}).
  2. (2)

    Two clusters are adjacent in the exchange graph if and only if they have exactly n1n-1 common cluster variables.

Proof.

Let t=t0t^{\prime}=t_{0} for simplicity. Without loss of generality, we may assume that σ=id\sigma=id.

We also consider the cluster algebra 𝒜¯(𝐱¯,𝐲¯,B)\bar{\mathcal{A}}(\bar{\mathbf{x}},\bar{\mathbf{y}},B) with principal coefficients simultaneously. By Theorem 5.4, we have that

σ(𝐱t0)=𝐱tgi,tt0=gi,t0t0,iσ(𝐱¯t0)=𝐱¯t\sigma(\mathbf{x}_{t_{0}})=\mathbf{x}_{t}\implies g_{i,t}^{t_{0}}=g_{i,t_{0}}^{t_{0}},\forall i\implies\sigma(\mathbf{\bar{x}}_{t_{0}})=\mathbf{\bar{x}}_{t}

We also know that the GG-matrix Gtt0G^{t_{0}}_{t} is the identity matrix InI_{n}. For any k[1,n]k\in[1,n], we have that

μk(xk,t)=yi[cik,t]+xi[bik,t]++yi[cik,t]+xi[bik,t]+xk=yi[cik,t]+xi[bik,t]++yi[cik,t]+xi[bik,t0]+yi[cik,t0]+xi[bik,t0]++yi[cik,t0]+xi[bik,t0]+μk(xk,t0).\begin{split}\mu_{k}(x_{k,t})&=\frac{\prod y_{i}^{[c_{ik,t}]_{+}}\prod x_{i}^{[b_{ik,t}]_{+}}+\prod y_{i}^{[-c_{ik,t}]_{+}}\prod x_{i}^{[-b_{ik,t}]_{+}}}{x_{k}}\\ &=\frac{\prod y_{i}^{[c_{ik,t}]_{+}}\prod x_{i}^{[b_{ik,t}]_{+}}+\prod y_{i}^{[-c_{ik,t}]_{+}}\prod x_{i}^{[-b_{ik,t_{0}}]_{+}}}{\prod y_{i}^{[c_{ik,t_{0}}]_{+}}\prod x_{i}^{[b_{ik,t_{0}}]_{+}}+\prod y_{i}^{[-c_{ik,t_{0}}]_{+}}\prod x_{i}^{[-b_{ik,t_{0}}]_{+}}}\mu_{k}(x_{k,t_{0}}).\end{split}

Since μk(xk,t)\mu_{k}(x_{k,t}) is a Laurent polynomial of x1,,μk(x1,t0),,xnx_{1},\dots,\mu_{k}(x_{1,t_{0}}),\dots,x_{n} with coefficients in [y1,y2,,yn]\mathbb{N}[y_{1},y_{2},\dots,y_{n}], thus we have that for each kk, there is a number εk{1,1}\varepsilon_{k}\in\{1,-1\} such that

𝐛k,t=εk𝐛k,t0,𝐜k,t=εk𝐜k,t0.\mathbf{b}_{k,t}=\varepsilon_{k}\mathbf{b}_{k,t_{0}},\quad\mathbf{c}_{k,t}=\varepsilon_{k}\mathbf{c}_{k,t_{0}}.

Since the exchange matrix is indecomposable and sign-skew-symmetric, we have that ε1==εn{1,1}\varepsilon_{1}=\dots=\varepsilon_{n}\in\{1,-1\}. Thus we have that either

Bt=Bt0,Ct=Ct0,B_{t}=-B_{t_{0}},\quad C_{t}=-C_{t_{0}},

or

Bt=Bt0,Ct=Ct0.B_{t}=B_{t_{0}},\quad C_{t}=C_{t_{0}}.

Since the cluster algebra is of odd rank, i.e., nn is odd. Then if Ct=Ct0C_{t}=-C_{t_{0}}, we have that |Ct|=(1)n=1|C_{t}|=(-1)^{n}=-1. Note that we have proved that Gt=InG_{t}=I_{n}, thus |Gt|=1|G_{t}|=1. While by Proposition 4.6, we have that |Gt|=|Ct||G_{t}|=|C_{t}|. Thus |Ct|=1|C_{t}|=-1 is a contradiction. Thus we have that Bt=Bt0B_{t}=B_{t_{0}}, Ct=Ct0=InC_{t}=C_{t_{0}}=I_{n}.

By Theorem 5.7 and the fact that 𝐱¯t0=𝐱¯t\mathbf{\bar{x}}_{t_{0}}=\mathbf{\bar{x}}_{t}, we know that 𝐲t=𝐲t0\mathbf{y}_{t}=\mathbf{y}_{t_{0}}. This finishes the proof. ∎

Theorem 5.10.

Let 𝒜¯\bar{\mathcal{A}} be the cluster algebra with principal coefficients such that the Assumption holds for the exchange matrix. Then each cluster (𝐱¯t,𝐲¯t,Bt)(\bar{\mathbf{x}}_{t},\bar{\mathbf{y}}_{t},B_{t}) is uniquely determined by the corresponding CC-matrix CtC_{t}.

Proof.

If there are two vertices t,tt,t^{\prime} on 𝕋n\mathbb{T}_{n} such that Ct=CtC_{t}=C_{t^{\prime}}. Let us show that (𝐱¯t,𝐲¯t,Bt)=(𝐱¯t,𝐲¯t,Bt)(\bar{\mathbf{x}}_{t},\bar{\mathbf{y}}_{t},B_{t})=(\bar{\mathbf{x}}_{t^{\prime}},\bar{\mathbf{y}}_{t^{\prime}},B_{t^{\prime}}). Note that by Proposition 4.7, Ct=CtG~t=G~tC_{t}=C_{t^{\prime}}\implies\tilde{G}_{t}=\tilde{G}_{t^{\prime}}. While by Theorem 5.4, G~t=G~t𝐱~¯t=𝐱~¯t.\tilde{G}_{t}=\tilde{G}_{t^{\prime}}\implies\bar{\tilde{{\mathbf{x}}}}_{t^{\prime}}=\bar{\tilde{{\mathbf{x}}}}_{t^{\prime}}. By Theorem 5.8, we have that C~t=C~t\tilde{C}_{t}=\tilde{C}_{t^{\prime}}. Thus Gt=GtG_{t}=G_{t^{\prime}} and 𝐱¯t=𝐱¯t.\bar{{\mathbf{x}}}_{t^{\prime}}=\bar{{\mathbf{x}}}_{t^{\prime}}. By Theorem 5.4, we have that (𝐱¯t,𝐲¯t,Bt)=(𝐱¯t,𝐲¯t,Bt)(\bar{\mathbf{x}}_{t},\bar{\mathbf{y}}_{t},B_{t})=(\bar{\mathbf{x}}_{t^{\prime}},\bar{\mathbf{y}}_{t^{\prime}},B_{t^{\prime}}). ∎

Corollary 5.11.

If the totally sign-skew-symmetric matrix satisfies the Assumption, then the GG-matrix and the CC-matrix are uniquely determined by each other, and they determine the seeds uniquely.

As a corollary, we obtain the following results.

Theorem 5.12.
  1. (1)

    Let 𝒜\mathcal{A} be an acyclic sign-skew-symmetric cluster algebra with coefficients coming from an semifield \mathbb{P}. Then

    1. (a)

      Every seed is uniquely defined by its cluster, i.e., if there is a permutation σSn\sigma\in S_{n} such that σ(𝐱t)=𝐱t\sigma(\mathbf{x}_{t^{\prime}})=\mathbf{x}_{t} for some t,t𝕋nt,t^{\prime}\in\mathbb{T}_{n}, then

      (𝐱t,𝐲t,Bt)=σ(𝐱t,𝐲t,Bt).(\mathbf{x}_{t},\mathbf{y}_{t},B_{t})=\sigma(\mathbf{x}_{t^{\prime}},\mathbf{y}_{t^{\prime}},B_{t^{\prime}}).
    2. (b)

      Two clusters are adjacent in the exchange graph if and only if they have exactly n1n-1 common cluster variables.

  2. (2)

    Let 𝒜¯\bar{\mathcal{A}} be an acyclic sign-skew-symmetric cluster algebras with principal coefficients. Then each cluster (𝐱¯t,𝐲¯t,Bt)(\bar{\mathbf{x}}_{t},\bar{\mathbf{y}}_{t},B_{t}) is uniquely determined by the corresponding CC-matrix CtC_{t}.

Acknowledgements:  This work was supported by the National Natural Science Foundation of China (Grants No.12071422).

References

  • [CHL18] P. Cao, M. Huang, and F. Li. A conjecture on cc-matrices of cluster algebras. Nagoya Mathematical Journal, pages 1–10, 2018.
  • [CHL22] P. Cao, M. Huang, and F. Li. Categorification of sign-skew-symmetric cluster algebras and some conjectures on g-vectors. Algebr. Represent. Theory, 25(6):1685–1698, 2022.
  • [CL18] P. Cao and F. Li. Some conjectures on generalized cluster algebras via the cluster formula and dd-matrix pattern. Journal of Algebra, 493(57-78), 2018.
  • [DWZ10] H. Derksen, J. Weyman, and A. Zelevinsky. Quivers with potentials and their representations ii: applications to cluster algebras. J. Amer. Math. Soc, 23(749-790), 2010.
  • [FWZ] S. Fomin, L. Williams, and A. Zelevinsky. Introduction to cluster algebras chapter 4-5. arxiv: 1707.07190.
  • [FZ02] S. Fomin and A. Zelevinsky. Cluster algebras i: Foundations. J. Amer. Math. Soc., 15(2):497–529, 2002.
  • [FZ07] S. Fomin and A. Zelevinsky. Cluster algebras iv: Coefficients. Compos. Math., 143(1):112–164, 2007.
  • [GHKK18] M. Gross, P. Hacking, S. Keel, and M. Kontsevich. Canonical bases for cluster algebras. J. Amer. Math. Soc., 31:497–608, 2018.
  • [GSV08] M. Gekhtman, M. Shapiro, and A. Vainshtein. On the properties of the exchange graph of a cluster algebra. Math. Res. Lett., 15(2):321–330, 2008.
  • [HL18] M. Huang and F. Li. Unfolding of acyclic sign-skew-symmetric cluster algebras and applications to positivity and ff-polynomials. Advances in Mathematics, 340:221–283, 2018.
  • [LP] F. Li and J. Pan. Recurrence formula, positivity and polytope basis in cluster algebras via newton polytopes. arXiv:2201.01440.
  • [LS15] K. Lee and R. Schiffler. Positivity for cluster algebras. Ann. Math., 2(1):73–125, 2015.
  • [Nak] T. Nakanishi. Cluster patterns and scattering diagrams. arxiv: 2103.16309.
  • [NZ12] T. Nakanishi and A. Zelevinsky. On tropical dualities in cluster algebras. Contemp. Math, 565:217–226, 2012.
  • [Rea14] N. Reading. Universal geometric cluster algebras. Mathematische Zeitschrift, 277:499–547, 2014.