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

Constrained percolation, Ising model and XOR Ising model on planar lattices

Zhongyang Li Department of Mathematics, University of Connecticut, Storrs, Connecticut 06269-3009, USA zhongyang.li@uconn.edu https://mathzhongyangli.wordpress.com
Abstract.

We study site percolation models on planar lattices including the [m,4,n,4][m,4,n,4] lattice and the square tilings on the Euclidean plane (2{\mathbb{R}}^{2}) or the hyperbolic plane (2\mathbb{H}^{2}), satisfying certain local constraints on degree-4 faces. These models are closely related to Ising models and XOR Ising models (product of two i.i.d Ising models) on regular tilings of 2{\mathbb{R}}^{2} or 2\mathbb{H}^{2}. In particular, we obtain a description of the numbers of infinite “++” and “-” clusters of the ferromagnetic Ising model on a vertex-transitive triangular tiling of 2{\mathbb{H}}^{2} for different boundary conditions and coupling constants. Our results show the possibility that such an Ising configuration has infinitely many infinite “++” and “-” clusters, while its random cluster representation has no infinite open clusters. Percolation properties of corresponding XOR Ising models are also discussed.

1. Introduction

A constrained percolation model is a probability measure on subgraphs of a lattice satisfying certain local constraints. Each subgraph is called a configuration. These models are abstract mathematical models for ubiquitous phenomena in nature, and have been interesting topics in mathematical and scientific research for long. Examples of constrained percolation models include the dimer model (see [28]), the 1-2 model (see [18]), the six-vertex model (or 6V model, see [2, 8, 29]), and general vertex models (see [48, 46, 34]). The study of these models may give deep insights to understand many natural phenomena, such as structure of matter, phase transition, limit shape, and critical behavior.

We are interested in the classical percolation problem in a constrained model: under which probability measure does there exist an infinite connected set (infinite cluster) in which every vertex is present in the random configuration, or equivalently, included in the randomly-chosen subgraph? Such a question has been studied extensively in the unconstrained case - in particular the i.i.d Bernoulli percolation - see, for instance, [24, 31, 23, 3, 15, 16]. The major difference between the constrained percolation and the unconstrained percolation lies in the fact that imposing local constraints usually makes stochastic monotonicity, which is a crucial property when studying the unconstrained model, invalid. Therefore new techniques need to be developed to study constrained percolation models.

Some constrained percolation models, including the 1-2 model, the periodic plane dimer model, certain 6V models, are exactly solvable; see [30, 36, 35, 37, 19, 8]. The integrability properties of these models make it possible to compute the correlations. When the parameters associated to the probability measure vary, different behaviors of the local correlations imply a phase transition from a microscopic point of view. If we consider phase transitions from a macroscopic, or geometric point of view, different approaches may be applied to study the existence of infinite clusters for a large class of constrained percolation models.

In [27], we studied a constrained percolation model on the 2{\mathbb{Z}}^{2} lattice, and showed that if the underlying probability measure satisfies mild assumptions like symmetry, ergodicity and translation-invariance, then with probability 0 the number of infinite clusters is nonzero and finite. The technique makes use of the planarity and amenability of the 2D square grid 2{\mathbb{Z}}^{2}. As an application, we obtained percolation properties for the XOR Ising model (a random spin configuration on a graph in which each spin is the product of two spins from two i.i.d Ising models, see [49]) on 2{\mathbb{Z}}^{2}, with the help of the combinatorial correspondence between the XOR Ising model and the dimer model proved in [13, 9]. In this paper, we further develop the technique to study constrained percolation models on a number of planar lattices, which may be amenable or non-amenable, including the [m,4,n,4][m,4,n,4] lattice and the square tilings of the hyperbolic plane; see [11] for an introduction to hyperbolic geometry.

The XOR Ising model was first introduced in [49] with interesting conformal invariance properties at criticality. For positive integers m,n3m,n\geq 3, the [m,4,n,4][m,4,n,4] lattice is a vertex-transitive planar graph in which each vertex is incident to 4 faces with degrees m,4,n,4m,4,n,4 in cyclic order. The constrained percolation model on the [m,4,n,4][m,4,n,4] lattice is of special interest because there is a measure-preserving correspondence between its configurations and the XOR Ising configurations on the mm-regular lattice or the nn-regular lattice. The Euclidean-plane version of such a correspondence was introduced in [9]. When 1m+1n<12\frac{1}{m}+\frac{1}{n}<\frac{1}{2}, the [m,4,n,4][m,4,n,4] lattice is no longer amenable but can be embedded into the hyperbolic plane. Although phase transitions and conformal invariance for statistical mechanical models in the Euclidean plane have been studied extensively, statistical mechanical models, including the Ising model and the related random cluster model, have been fascinating problems for mathematicians and physicists for a long time, however, a lot of things remain unknown. For example, it is well-known that for statistical mechanical models in the hyperbolic plane, there is an “intermediate” phase between the non-percolation phase and unique-percolation phase, which usually does not exist for statistical mechanical models in the Euclidean plane; a lot of descriptions of the “intermediate” phase seem to be “qualitative” while not “quantitative” - for which values of the parameters does the model have such an “intermediate” phase? Indeed, the general results we obtain in this paper can be used to prove further results concerning percolation properties of the XOR Ising model on the hexagonal and the triangular lattices, as well as on regular tilings of the hyperbolic plane.

The specific geometric properties of non-amenable graphs make it an interesting problem to study percolation models on such graphs; and a set of techniques have been developed in the past few decades; see [5, 4, 3, 21, 39, 44, 22, 42, 50, 45, 20, 40, 38] for an incomplete list. In this paper, we also study the general automorphism-invariant percolation models on transitive planar graphs.

One of the most classical percolation models is the i.i.d Bernoulli site percolation on a graph, in which the vertices are open (resp. closed) with probability pp (resp. 1p1-p) independently, where p[0,1]p\in[0,1]. The critical probability pcp_{c} is the supremum of pp’s such that almost surely there are no infinite open clusters. A graph G=(V,E)G=(V,E) is a vertex-transitive graph if there exists a subgroup Γ𝐀𝐮𝐭(G)\Gamma\subseteq\mathbf{Aut}(G) of the automorphism group GG such that for any two vertices v,wVv,w\in V, there exist γΓ\gamma\in\Gamma satisfying γv=w\gamma v=w. The number of ends of a connected graph is the supremum over its finite subgraphs of the number of infinite components that remain after removing the subgraph.

Our results may be related to the following two conjectures. More precisely, we prove the following conjectures for some special vertex-transitive planar graphs.

Conjecture 1.1.

(Conjecture 7 of [5]) Suppose that GG is a planar, connected graph, and the minimal vertex degree in GG is at least 7. In an i.i.d Bernoulli site percolation on GG, at every pp in the range (pc,1pc)(p_{c},1-p_{c}), there are infinitely many infinite open clusters in the i.i.d Bernoulli site percolation on GG. Moreover, we conjecture that pc<12p_{c}<\frac{1}{2}, and the above interval is nonempty.

In Example 2.3, we explain why 1.1 is true for the i.i.d Bernoulli site percolation on vertex-transitive triangular tilings of the hyperbolic plane where each vertex has degree n7n\geq 7.

Conjecture 1.2.

(Conjecture 8 of [5]) Let GG be a planar, connected graph. Let p=12p=\frac{1}{2} be the probability that a vertex is open and assume that a.s. percolation occurs in the site percolation on GG. Then almost surely there are infinitely many infinite clusters.

Our Proposition 10.8 implies that 1.2 is true for automorphism-invariant site percolation (not necessarily independent, or insertion tolerant) on vertex-transitive triangular tilings of the hyperbolic plane where each vertex has degree n7n\geq 7 if the underlying measure is ergodic and invariant under switching state-1 vertices and state-0 vertices.

We then apply our results concerning the general automorphism-invariant percolation models on transitive planar graphs to study the infinite “++”-clusters and “-”-clusters for the Ising model on vertex-transitive triangular tilings of the hyperbolic plane where each vertex has degree n7n\geq 7, and describe the behaviors of such clusters with respect to varying coupling constants under the free boundary condition and the wired boundary condition. A surprising result we obtain is that it is possible that the random cluster representation of the Ising model has no infinite open clusters, while the Ising model has infinitely many infinite “++”-clusters and infinitely many infinite “-”-clusters - in contrast with the Ising percolation and its random cluster representation on the 2d square grid 2{\mathbb{Z}}^{2} (see [12, 25, 17]) where the Ising model has an infinite “++” or “-”-cluster if and only if its random cluster representation has an infinite open cluster.

The main tools to prove these results are the planar duality of graphs, ergodicity and symmetry of probability measures, as well as properties of amenablity and non-amenablity. One characteristic of the constrained percolation obtained from a natural correspondence with the XOR Ising model, which is not shared with the unconstrained percolation, is that given such a constraint, there are two sets of “contours” separating clusters of vertices of different states. These two sets of contours lie on two planar graphs dual to each other, and the present edges in these two different sets of contours never cross. As a result, there are four types of infinite components in our constrained percolation model: infinite “0”-cluster, infinite “1”-cluster, infinite planar contour and infinite dual contour. The geometric configurations of these infinite components, together with the ergodicity and symmetry of the probability measure, lead to interesting properties that are particular and unique to the constrained percolation model.

The organization of the paper is as follows.

In Section 2, we introduce the [m,4,n,4][m,4,n,4] lattice and state the result concerning constrained percolation models on the [m,4,n,4][m,4,n,4] lattice. In Section 3, we state the main results concerning infinite clusters in the Ising model on regular triangular tilings of the hyperbolic plane, and, in particular, provide a description of the numbers of infinite “++” and “-” clusters of the ferromagnetic Ising model with the free boundary condition, the “++” boundary condition or the “-” boundary condition on such a lattice for different values of coupling constants. In Section 4, we state the main results concerning infinite clusters in the XOR Ising model on regular triangular tilings of the hyperbolic plane and its dual graph. In Section 5, we state the result proved in this paper concerning the percolation properties of the XOR Ising model on the hexagonal lattice and the triangular lattice. In Section 6, we introduce the square tilings of the hyperbolic plane, state and prove the main result concerning constrained percolation models on such a lattice.

The remaining sections are devoted to prove the theorems stated in preceding sections. In Section 7, we prove Theorem 2.2. In Section 8, we prove Theorem 2.4. In Section 9, we prove Theorem 2.5. In Section 10, we discuss the applications of the techniques developed in the proof of Theorem 2.2 to prove results concerning unconstrained site percolation on vertex-transitive, triangular tilings of the hyperbolic plane in preparation of proving Theorems 3.3, 4.1 and 4.2. In Section 11, we prove Theorem 3.3. In Section 12, we prove Theorem 4.1 and Theorem 4.2. In Section 13, we prove Theorems 5.1 and 5.2. In Appendix A, we prove combinatorial results concerning contours and clusters in preparation to prove the main theorems.

2. Constrained percolation on the [m,4,n,4][m,4,n,4] lattice

In this section, we state the main result proved in this paper for the constrained percolation models on the [m,4,n,4][m,4,n,4] lattice. We shall start with a formal definition of the [m,4,n,4][m,4,n,4] lattice.

Let m,nm,n be positive integers satisfying

(1) m3,n3\displaystyle m\geq 3,\qquad n\geq 3
(2) 1m+1n12.\displaystyle\frac{1}{m}+\frac{1}{n}\leq\frac{1}{2}.

The [m,4,n,4][m,4,n,4] lattice is a vertex-transitive graph which can be embedded into the Euclidean plane or the hyperbolic plane such that each vertex is incident to 4 faces with degrees m,4,n,4m,4,n,4 in cyclic order. When 1m+1n=12\frac{1}{m}+\frac{1}{n}=\frac{1}{2}, the graph is amenable and can be embedded into the Euclidean plane. When 1m+1n<12\frac{1}{m}+\frac{1}{n}<\frac{1}{2}, the graph is non-amenable and can be embedded into the hyperbolic plane ([43]). Note that when m=n=4m=n=4, the graph is the square grid embedded into the 2D Euclidean plane. See Figure 1 for an illustration of the [3,4,6,4] lattice, Figure 2 for the [3,4,7,4] lattice, and Figure 3 for the [6,4,6,4] lattice.

Refer to caption
Figure 1. The [3,4,6,4] lattice, the auxiliary hexagonal lattice and triangular lattice. Black lines represent the [3,4,6,4] lattice; dashed red lines represent the triangular lattice; dashed blue lines represent the hexagonal lattice.
Refer to caption
Figure 2. The [3,4,7,4] lattice (picture from the wikipedia)
Refer to caption
Figure 3. The [6,4,6,4] lattice represented by blue lines (picture from http://epinet.anu.edu.au/)

Let G=(V,E)G=(V,E) be an [m,4,n,4][m,4,n,4] lattice. We color all the faces of degree mm or nn with white and all the other faces with black, such that any two faces sharing an edge have different colors. We consider the site percolation on VV satisfying the following constraint (see Figure 4):

  • around each black face, there are six allowed configurations (0000)(0000), (1111)(1111), (0011)(0011), (1100)(1100), (0110)(0110), (1001)(1001), where the digits from the left to the right correspond to vertices in clockwise order around the black face, starting from the lower left corner. See Figure 4.

Refer to caption
(a) 0000
Refer to caption
(b) 0011
Refer to caption
(c) 0110
Refer to caption
(d) 1111
Refer to caption
(e) 1100
Refer to caption
(f) 1001
Figure 4. Local configurations of the constrained percolation around a black square. Red and blue lines mark contours separating 0’s and 1’s (in 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} respectively). Yellow (resp. green) disks represent 0’s (resp. 1’s).

Let Ω{0,1}V\Omega\subset\{0,1\}^{V} be the probability space consisting of all the site configurations on GG satisfying the constraint above. To the [m,4,n,4][m,4,n,4] lattice GG, we associate two auxiliary lattices 𝕃1=(V(𝕃1),E(𝕃1)){\mathbb{L}}_{1}=(V({\mathbb{L}}_{1}),E({\mathbb{L}}_{1})) and 𝕃2=(V(𝕃2),E(𝕃2)){\mathbb{L}}_{2}=(V({\mathbb{L}}_{2}),E({\mathbb{L}}_{2})) as follows. Each vertex of 𝕃1{\mathbb{L}}_{1} (resp. 𝕃2{\mathbb{L}}_{2}) is located at the center of each degree-mm face (resp. degree-nn face) of GG. Two vertices of 𝕃1{\mathbb{L}}_{1} (resp. 𝕃2{\mathbb{L}}_{2}) are joined by an edge of 𝕃1{\mathbb{L}}_{1} (resp. 𝕃2{\mathbb{L}}_{2}) if and only if the two corresponding mm-faces (resp. nn-faces) of GG are adjacent to the same square face of GG through a pair of opposite edges (edges of a square face that do not share a vertex), respectively.

We say an edge eE(𝕃1)E(𝕃2)e\in E({\mathbb{L}}_{1})\cup E({\mathbb{L}}_{2}) crosses a square face of the [m,4,n,4][m,4,n,4] lattice if the edge ee crosses a pair of opposite edges of the square face. Note that

  1. i

    𝕃1{\mathbb{L}}_{1} (resp. 𝕃2{\mathbb{L}}_{2}) is a planar lattice in which each face has degree nn (resp. mm) and each vertex has degree mm (resp. nn).

  2. ii

    𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} are planar dual to each other.

  3. iii

    Each edge in E(𝕃1)E(𝕃2)E({\mathbb{L}}_{1})\cup E({\mathbb{L}}_{2}) crosses a unique square face of the [m,4,n,4][m,4,n,4] lattice. When m4m\neq 4 and n4n\neq 4, each square face of the [m,4,n,4][m,4,n,4] lattice is crossed by a unique edge e1E(𝕃1)e_{1}\in E({\mathbb{L}}_{1}) and a unique edge e2E(𝕃2)e_{2}\in E({\mathbb{L}}_{2}); and moreover, e1e_{1} and e2e_{2} are dual to each other.

When m=nm=n, both 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} are lattices in which each face has degree nn and each vertex has degree nn. When m=3m=3 and n=6n=6, 𝕃1{\mathbb{L}}_{1} is the hexagonal lattice and 𝕃2{\mathbb{L}}_{2} is the triangular lattice; see Figure 1. When m=3m=3 and n=7n=7, 𝕃2{\mathbb{L}}_{2} is a vertex-transitive triangular tiling of the hyperbolic plane, in which each vertex has degree 3; see the left graph of Figure 5; while 𝕃1{\mathbb{L}}_{1} is the [7,7,7][7,7,7] lattice on the hyperbolic plane, see the right graph of Figure 5.

Refer to caption
Refer to caption
Figure 5. The [3,3,3,3,3,3,3] lattice on the left and the [7,7,7] lattice on the right

Let Φ{0,1}E(𝕃1)E(𝕃2)\Phi\subset\{0,1\}^{E({\mathbb{L}}_{1})\cup E({\mathbb{L}}_{2})} be the set of contour configurations satisfying the condition that each vertex of V(𝕃1)V({\mathbb{L}}_{1}) and V(𝕃2)V({\mathbb{L}}_{2}) is incident to an even number of present edges, and present edges in E(𝕃1)E({\mathbb{L}}_{1}) and E(𝕃2)E({\mathbb{L}}_{2}) never cross. Any constrained percolation configuration ωΩ\omega\in\Omega is mapped to a contour configuration ϕ(ω)Φ\phi(\omega)\in\Phi, where an edge ee in E(𝕃1)E({\mathbb{L}}_{1}) or E(𝕃2)E({\mathbb{L}}_{2}) is present (i.e., have state 1) if and only if the following condition holds

  • Let SS be the square face of GG crossed by ee. Let v1,v2,v3,v4v_{1},v_{2},v_{3},v_{4} be the four vertices of SS, such that v1v_{1} and v2v_{2} are on one side of ee and v3v_{3}, v4v_{4} are on the other side of ee. Then v1v_{1} and v2v_{2} have the same state, v3v_{3} and v4v_{4} have the same state, and v1v_{1} and v3v_{3} have different states.

See Figure 6 for a contour configuration obtained from a constrained percolation configuration on the [3,4,6,4][3,4,6,4] lattice. Note that the mapping ϕ:ΩΦ\phi:\Omega\rightarrow\Phi is 2-to-1 since ϕ(ω)=ϕ(1ω)\phi(\omega)=\phi(1-\omega).

A contour is a connected component of present edges in a contour configuration in Φ\Phi. A contour may be finite or infinite depending on the number of edges in the contour. Since present edges of a contour configuration in E(𝕃1)E({\mathbb{L}}_{1}) and in E(𝕃2)E({\mathbb{L}}_{2}) never cross, either all the edges in a contour are edges of 𝕃1{\mathbb{L}}_{1}, or all the edges in a contour are edges in 𝕃2{\mathbb{L}}_{2}. We call a contour primal contour (resp. dual contour) if all the edges in the contour are edges of 𝕃1{\mathbb{L}}_{1} (resp. 𝕃2{\mathbb{L}}_{2}).

Refer to caption
Figure 6. A constrained percolation configuration on the [3,4,6,4] lattice. Red lines represent contours on the triangular lattice. Blue lines represent contours on the hexagonal lattice.

Let Γ\Gamma be the automorphism group Aut(G)\mathrm{Aut}(G) of the graph GG. For i{1,2}i\in\{1,2\}, let Γi\Gamma_{i} be the automorphism group Aut(𝕃i)\mathrm{Aut}({\mathbb{L}}_{i}) of the graph 𝕃i{\mathbb{L}}_{i}. Let μ\mu be a probability measure on Ω\Omega. The measure μ\mu naturally induces a measure on Φ\Phi under the 2-to-1 mapping ϕ\phi; with a little abuse of notation, we still use μ\mu to denote this induced measure on contour configurations Φ\Phi. We may assume that μ\mu satisfies the following conditions:

  1. (A1)

    μ\mu is Γ\Gamma-invariant;

  2. (A2)

    μ\mu is Γi\Gamma_{i}-ergodic for i=1,2i=1,2; i.e. any Γi\Gamma_{i}-invariant event has μ\mu-probability 0 or 1;

  3. (A3)

    μ\mu is symmetric: let θ:ΩΩ\theta:\Omega\rightarrow\Omega be the map defined by θ(ω)(v)=1ω(v)\theta(\omega)(v)=1-\omega(v), for each vertex vVv\in V, then μ\mu is invariant under θ\theta, that is, for any event AA, μ(A)=μ(θ(A))\mu(A)=\mu(\theta(A)).

Let Φ1\Phi_{1} (resp. Φ2\Phi_{2}) be the set of all contour configurations on 𝕃1{\mathbb{L}}_{1} (resp. 𝕃2{\mathbb{L}}_{2}) satisfying the condition that each vertex of 𝕃1{\mathbb{L}}_{1} (resp. 𝕃2{\mathbb{L}}_{2}) has an even number of incident present edges. For each contour configuration ψΦ\psi\in\Phi, we have ψ=ψ1ψ2\psi=\psi_{1}\cup\psi_{2}, where ψ1Φ1\psi_{1}\in\Phi_{1} and ψ2Φ2\psi_{2}\in\Phi_{2}; moreover, ψ1ψ2=\psi_{1}\cap\psi_{2}=\emptyset.

Let ν1\nu_{1} (resp. ν2\nu_{2}) be the marginal distribution of μ\mu on Φ1\Phi_{1} (resp. Φ2\Phi_{2}). When 1m+1n=12\frac{1}{m}+\frac{1}{n}=\frac{1}{2}, the [m,4,n,4][m,4,n,4] lattice is amenable. It is not hard to see that if 1m+1n=12\frac{1}{m}+\frac{1}{n}=\frac{1}{2}, then (m,n){(4,4),(3,6),(6,3)}(m,n)\in\{(4,4),(3,6),(6,3)\}. When m=n=4m=n=4, the [m,4,n,4][m,4,n,4] lattice is the 2D square grid, on which the constrained percolation models was discussed in [27].

Now we consider the case when (m,n)=(3,6)(m,n)=(3,6). As discussed before, in this case 𝕃1{\mathbb{L}}_{1} is the hexagonal lattice {\mathbb{H}}, and 𝕃2{\mathbb{L}}_{2} is the triangular lattice 𝕋{\mathbb{T}}. We first define the finite energy condition of a random contour configuration on a planar graph.

Definition 2.1.

Let G=(V,E)G=(V,E) be a vertex-transitive, planar graph. Let Φ\Phi be the set of all contour configurations on GG, in which each contour configuration is a subset of edges such that each vertex is incident to an even number of present edges. Let ν\nu be a probability measure on Φ\Phi. We say ν\nu has finite energy if for any face SS of GG, let SE\partial S\subset E consist of of all the sides of the polygon SS. Define ϕS\phi_{S} to be the configuration obtained by switching the states of each element of S\partial S, i.e. ϕS(e)=1ϕ(e)\phi_{S}(e)=1-\phi(e) if eSe\in\partial S, and ϕS(e)=ϕ(e)\phi_{S}(e)=\phi(e) otherwise. Let EE be an event, and define

(3) ES={ϕS:ϕE}.\displaystyle E_{S}=\{\phi_{S}:\phi\in E\}.

Then ν(ES)>0\nu(E_{S})>0 whenever ν(E)>0\nu(E)>0.

We may assume that ν1\nu_{1} or ν2\nu_{2} has finite energy as follows.

  1. (A4)

    ν1\nu_{1} has finite energy.

  2. (A5)

    ν2\nu_{2} has finite energy.

See Figures 7 and 8 for illustrations of the configuration-changing process on the hexagonal lattice {\mathbb{H}} and the triangular lattice 𝕋{\mathbb{T}}, respectively.

Refer to caption
Refer to caption
Figure 7. Change of contour configurations in 𝕃1={\mathbb{L}}_{1}={\mathbb{H}}
Refer to caption
Refer to caption
Figure 8. Change of contour configurations in 𝕃2=𝕋{\mathbb{L}}_{2}={\mathbb{T}}

For a random contour configuration ψΦ1\psi\in\Phi_{1} (resp. ψΦ2\psi\in\Phi_{2}) whose distribution is the marginal distribution ν1\nu_{1} (resp. ν2\nu_{2}) of μ\mu on Φ1\Phi_{1} (resp. Φ2\Phi_{2}), ψ\psi induces a random constrained configuration ωϕ1(ψ)\omega\in\phi^{-1}(\psi) as follows. Let v0v_{0} be a fixed vertex of GG. Assume that ω(v0)=1\omega(v_{0})=1 with probability 12\frac{1}{2}, and ω(v0)=0\omega(v_{0})=0 with probability 12\frac{1}{2}, and is independent of ψ\psi. For two vertices v1,v2v_{1},v_{2} of GG joined by an edge ee, v1v_{1} and v2v_{2} have different states if and only if ee crosses a present edge in ψ\psi. Let λ1\lambda_{1} (resp. λ2\lambda_{2}) be the distribution of ω\omega. We may further make the following assumptions

  1. (A6)

    λ1\lambda_{1} is Γ1\Gamma_{1}-ergodic;

  2. (A7)

    λ2\lambda_{2} is Γ2\Gamma_{2}-ergodic.

Also we may sometimes assume that

  1. (A8)

    μ\mu is Γ1\Gamma_{1}-invariant.

The main theorems of this section are stated as follows.

Theorem 2.2.

Let GG be the [3,4,n,4][3,4,n,4] lattice with n7n\geq 7. Let s0s_{0} (resp. s1s_{1}) be the number of infinite 0-clusters (resp. 1-clusters). Let t1t_{1} (resp. t2t_{2}) be the number of infinite 𝕃1{\mathbb{L}}_{1}-contours (resp. 𝕃2{\mathbb{L}}_{2}-contours).

  1. I

    Let μ\mu be a probability measure on Ω\Omega satisfying (A2),(A3),(A7),(A8). Then μ\mu-a.s. (s0,s1,t1)=(,,)(s_{0},s_{1},t_{1})=(\infty,\infty,\infty).

  2. II

    Let μ\mu be a probability measure on Ω\Omega satisfying (A2),(A3),(A6),(A7),(A8). Then μ\mu-a.s. (s0,s1,t1,t2){(,,,1),(,,,)}(s_{0},s_{1},t_{1},t_{2})\in\{(\infty,\infty,\infty,1),(\infty,\infty,\infty,\infty)\}.

The case m=3,n7m=3,n\geq 7 is of special interest, because in this case 𝕃1{\mathbb{L}}_{1} is a cubic graph (each vertex has degree 3), and 𝕃2{\mathbb{L}}_{2} is a triangular tiling of the hyperbolic plane. As a result, any infinite contour on 𝕃1{\mathbb{L}}_{1} must be a doubly infinite self-avoiding path. An application of Theorem 2.2 is illustrated in the following example.

Example 2.3.

Consider the i.i.d. Bernoulli site percolation on the regular tiling 𝕃2{\mathbb{L}}_{2} of the hyperbolic plane with triangles, such that each vertex has degree n7n\geq 7. Assume that each vertex of 𝕃2{\mathbb{L}}_{2} takes value 1 with probability 12\frac{1}{2}. The corresponding contour configuration on the dual graph 𝕃1{\mathbb{L}}_{1} to the site percolation on 𝕃2{\mathbb{L}}_{2} induces a constrained configuration in the [3,4,n,4][3,4,n,4] lattice satisfying (A8),(A2),(A3),(A7). Then by Theorem 2.2 μ\mu-a.s. (s0,s1)=(,)(s_{0},s_{1})=(\infty,\infty).

Theorem 2.4.

Let GG be the [m,4,n,4][m,4,n,4] lattice such that

(4) m3,n3;\displaystyle m\geq 3,\qquad n\geq 3;
1m+1n=12\displaystyle\frac{1}{m}+\frac{1}{n}=\frac{1}{2}

Let μ\mu be a probability measure on Ω\Omega. Then

  1. I

    if μ\mu satisfies (A1)-(A6), then almost surely there are no infinite contours in 𝕃2{\mathbb{L}}_{2};

  2. II

    if μ\mu satisfies (A1)-(A5) and (A7), then almost surely there are no infinite contours in 𝕃1{\mathbb{L}}_{1};

  3. III

    if μ\mu satisfies (A1)-(A7), almost surely there are neither infinite contours nor infinite clusters.

When m=nm=n, the two lattice 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} are isomorphic to each other, this allows us to use symmetry to obtain the following theorem.

Theorem 2.5.

Let GG be the [m,4,n,4][m,4,n,4] lattice satisfying

m=n5.\displaystyle m=n\geq 5.

Let μ\mu be a probability measure on Ω\Omega. Let s0s_{0} (resp. s1s_{1}) be the number of infinite 0-clusters (resp. 1-clusters), and let t1t_{1} (resp. t2t_{2}) be the number of infinite 𝕃1{\mathbb{L}}_{1}-contours (resp. 𝕃2{\mathbb{L}}_{2}-contours). If μ\mu satisfies (A1)-(A3); then μ\mu-a.s. (s0,s1,t1,t2)=(,,,)(s_{0},s_{1},t_{1},t_{2})=(\infty,\infty,\infty,\infty).

Theorem 2.2 is proved in Section 7. Theorem 2.4 is proved in Section 8, and Theorem 2.5 is proved in Section 9.

3. Ising model on transitive, triangular tilings of the hyperbolic plane

In this section, we state the main result concerning the percolation properties of the Ising model on transitive, triangular tilings of the hyperbolic plane. These results, as given in Theorem 3.3, will be proved in Section 11.

The random cluster representation of an Ising model on a transitive, triangular tiling of the hyperbolic plane can be defined as in [20]. Here we briefly summarize basic facts about the Fortuin-Kasteleyn random cluster model, which is a probability measure on bond configurations of a graph, and the related Potts model. See [17] for more information.

The random cluster measure RC:=RCp,qG0RC:=RC_{p,q}^{G_{0}} on a finite graph G0=(V0,E0)G_{0}=(V_{0},E_{0}) with parameters p[0,1]p\in[0,1] and q1q\geq 1 is the probability measure on {0,1}E0\{0,1\}^{E_{0}} which to each ξ{0,1}E0\xi\in\{0,1\}^{E_{0}} assigns probability

(5) RC(ξ):qk(ξ)eE0pξ(e)(1p)1ξ(e).\displaystyle RC(\xi):\propto q^{k(\xi)}\prod_{e\in E_{0}}p^{\xi(e)}(1-p)^{1-\xi(e)}.

where k(ξ)k(\xi) is the number of connected components in ξ\xi.

Let G=(V,E)G=(V,E) be an infinite, locally finite, connected graph. For each q[1,)q\in[1,\infty) and each p(0,1)p\in(0,1), let WRCp,qGWRC_{p,q}^{G} be the random cluster measure with the wired boundary condition, and let FRCp,qGFRC_{p,q}^{G} be the random cluster measure with the free boundary condition. More precisely, WRCp,qGWRC_{p,q}^{G} (resp. FRCp,qGFRC_{p,q}^{G}) is the weak limit of RCRC’s defined by (5) on larger and larger finite subgraphs approximating GG, where we assume that all the edges outside each finite subgraph are present (resp. absent).

The Gibbs measure μ+\mu^{+} (resp. μ\mu^{-}) for the Ising model on GG with coupling constant J0J\geq 0 on each edge and “++”-boundary conditions (resp. “-”-boundary conditions) can be obtained by considering a random configuration of present and absent edges according to the law WRCp,2GWRC_{p,2}^{G}, p=1e2Jp=1-e^{-2J}, and assigning to all the vertices in each infinite cluster the state “++” (resp. “-”), and to all the vertices in each finite cluster a state from {+,}\{+,-\}, chosen uniformly at random for each cluster and independently for different clusters.

The Gibbs measure μf\mu^{f} for the Ising model on GG with coupling constant J0J\geq 0 on each edge and free boundary conditions can be obtained by considering a random configuration of present and absent edges according to the law FRCp,2GFRC_{p,2}^{G}, p=1e2Jp=1-e^{-2J}, and assigning to all the vertices in each cluster a state from {+,}\{+,-\}, chosen uniformly at random for each cluster and independently for different clusters.

When there is no confusion, we may write FRCp,qGFRC_{p,q}^{G} and WRCp,qGWRC_{p,q}^{G} as FRCp,qFRC_{p,q} and WRCp,qWRC_{p,q} for simplicity. Assume that GG is transitive. Then measures FRCp,qFRC_{p,q} and WRCp,qWRC_{p,q} are Aut(G)\mathrm{Aut}(G)-invariant, and Aut(G)\mathrm{Aut}(G)-ergodic; see the explanations on Page 295 of [45].

Now we introduce the following definitions.

Definition 3.1.

A transitive graph G=(V,E)G=(V,E) is unimodular, if there exists a subgroup ΓAut(G)\Gamma\subseteq\mathrm{Aut}(G) acting transitively on GG, such that for any two vertices u,vVu,v\in V, we have

|Stabu(v)|=|Stabv(u)|;\displaystyle|\mathrm{Stab}_{u}(v)|=|\mathrm{Stab}_{v}(u)|;

where StabuΓ\mathrm{Stab}_{u}\subseteq\Gamma is the stabilizer of uu, i.e.

(6) Stabu={γΓ:γu=u};\displaystyle\mathrm{Stab}_{u}=\{\gamma\in\Gamma:\gamma u=u\};

and |||\cdot| is the cardinality of a set.

Definition 3.2.

A graph G=(V,E)G=(V,E) is called amenable, if its edge isoperimetric constant

(7) ıE(G):=infKV,|K|<|EK||K|=0.\displaystyle\imath_{E}(G):=\mathrm{inf}_{K\subset V,|K|<\infty}\frac{|\partial_{E}K|}{|K|}=0.

where EK\partial_{E}K is the set of edges with exactly one endpoint in KK and one endpoint not in KK.

If the edge isoperimetric constant is strictly positive, the graph is called nonamenable.

If we further assume that GG is unimodular, nonamenable and planar, it is known that there exists pc,qw,pc,qf,pu,qw,pc,qf[0,1]p_{c,q}^{w},\ p_{c,q}^{f},\ p_{u,q}^{w},\ p_{c,q}^{f}\in[0,1], such that FRCp,qFRC_{p,q}-a.s. the number of infinite clusters equals

(11) {0forppc,qfforp(pc,qf,pu,qf)1forp>pu,qf;\displaystyle\left\{\begin{array}[]{cc}0&\mathrm{for}\ p\leq p_{c,q}^{f}\\ \infty&\mathrm{for}\ p\in(p_{c,q}^{f},p_{u,q}^{f})\\ 1&\mathrm{for}\ p>p_{u,q}^{f};\end{array}\right.

and WRCp,qWRC_{p,q}-a.s. the number of infinite clusters equals

(15) {0forp<pc,qwforp(pc,qw,pu,qw)1forppu,qw.;\displaystyle\left\{\begin{array}[]{cc}0&\mathrm{for}\ p<p_{c,q}^{w}\\ \infty&\mathrm{for}\ p\in(p_{c,q}^{w},p_{u,q}^{w})\\ 1&\mathrm{for}\ p\geq p_{u,q}^{w}.\end{array}\right.;

see expressions (17),(18), Theorem 3.1 and Corollary 3.7 of [20].

It is well known that for the i.i.d Bernoulli percolation on a infinite, connected, locally finite transitive graph GG, there exist pc,pup_{c},p_{u} such that

  1. (i)

    0<pcpu10<p_{c}\leq p_{u}\leq 1;

  2. (ii)

    for p[0,pc)p\in[0,p_{c}) there is no infinite cluster a.s.

  3. (iii)

    for p(pc,pu)p\in(p_{c},p_{u}) there are infinitely many infinite clusters, a.s.

  4. (iv)

    for p(pu,1]p\in(p_{u},1], there is exactly one infinite cluster, a.s.

The monotonicity in pp of the uniqueness of the infinite cluster was proved in [21, 44]. Combining with Theorem 7.5 of [38] (proved first in [41]), we obtain statements (i)-(iv) above. It is proved that pc=pup_{c}=p_{u} for amenable transitive graphs (see [5]); and conjectured that pc<pup_{c}<p_{u} for transitive non-amenable graphs. The conjecture pc<pup_{c}<p_{u} was proved for transitive hyperbolic planar graphs (see [6]) and non-amenable Cayley graphs with small spectral radii (see [42, 45, 47]) or large girths (see [40]).

Theorem 3.3.

Let 𝕃2{\mathbb{L}}_{2} be a triangulation of the hyperbolic plane such that each vertex has degree n7n\geq 7. Consider the Ising model with spins located on vertices of 𝕃2{\mathbb{L}}_{2} and coupling constant JJ\in{\mathbb{R}} on each edge. Let pcp_{c}, pup_{u} be critical i.i.d Bernoulli site percolation probabilities on 𝕃2{\mathbb{L}}_{2} as defined by (i)-(iv) above.

  1. I

    Let h>0h>0 satisfy

    (16) eheh+eh=pc\displaystyle\frac{e^{-h}}{e^{h}+e^{-h}}=p_{c}

    Let μ+\mu^{+} (resp. μ\mu^{-}. μf\mu^{f}) be the infinite-volume Ising Gibbs measure with “++”-boundary conditions (resp. “-” boundary conditions, free boundary conditions). If

    (17) n|J|<h,\displaystyle n|J|<h,

    then μ\mu-a.s. there are infinitely many infinite “++”-clusters, infinitely many infinite “-”-clusters and infinitely many infinite contours, where μ\mu is an arbitrary Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariant Gibbs measure for the Ising model on 𝕃2{\mathbb{L}}_{2} with coupling constant JJ.

  2. II

    Assume J0J\geq 0. If one of the following conditions

    1. (a)

      μf\mu^{f} is Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-ergodic;

    2. (b)

      infu,vV(𝕃2)σuσvμf=0\mathrm{inf}_{u,v\in V({\mathbb{L}}_{2})}\langle\sigma_{u}\sigma_{v}\rangle_{\mu^{f}}=0, where σu\sigma_{u} and σv\sigma_{v} are two spins associated to vertices u,vV(𝕃2)u,v\in V({\mathbb{L}}_{2}) in the Ising model;

    3. (c)

      0J<12ln(11pu,2f)0\leq J<\frac{1}{2}\ln\left(\frac{1}{1-p_{u,2}^{f}}\right), where pu,2fp_{u,2}^{f} is the critical probability for the existence of a unique infinite open cluster of the corresponding random cluster representation of the Ising model on 𝕃2{\mathbb{L}}_{2}, with free boundary conditions as given in (11);

    4. (d)

      0J<12ln(11pu,1)0\leq J<\frac{1}{2}\ln\left(\frac{1}{1-p_{u,1}}\right), where pu,1p_{u,1} is the critical probability for the existence of a unique infinite open cluster for the i.i.d Bernoulli bond percolation on 𝕃2{\mathbb{L}}_{2};

    holds, then μf\mu^{f}-a.s. there are infinitely many infinite “++”-clusters and infinitely many infinite “-”-clusters. Indeed, we have (d)(c)(b)(a)(d)\Rightarrow(c)\Rightarrow(b)\Rightarrow(a).

  3. III

    Assume

    (18) J12ln(11pu,2w).\displaystyle J\geq\frac{1}{2}\ln\left(\frac{1}{1-p_{u,2}^{w}}\right).

    Let 𝒜+{\mathcal{A}}_{+} be the event that there is a unique infinite “++”-cluster, no infinite “-”-clusters and no infinite contours; and let 𝒜{\mathcal{A}}_{-} be the event that there is a unique infinite “-”-cluster, no infinite “++”-clusters and no infinite contours. then

    (19) μ+(𝒜+)\displaystyle\mu^{+}({\mathcal{A}}_{+}) =\displaystyle= 1.\displaystyle 1.
    (20) μ(𝒜)\displaystyle\mu^{-}({\mathcal{A}}_{-}) =\displaystyle= 1.\displaystyle 1.
  4. IV

    If

    (21) J>12ln(11pu,2f)\displaystyle J>\frac{1}{2}\ln\left(\frac{1}{1-p_{u,2}^{f}}\right)

    then (19) and (20) hold, and moreover,

    (22) μf(𝒜+)=μf(𝒜)=12.\displaystyle\mu^{f}({\mathcal{A}}_{+})=\mu^{f}({\mathcal{A}}_{-})=\frac{1}{2}.

From Theorem 3.3, we can also obtain the following corollary:

Corollary 3.4.

Let 𝕃2{\mathbb{L}}_{2} be a triangulation of the hyperbolic plane such that each vertex has degree n7n\geq 7. Consider the Ising model with spins located on vertices of 𝕃2{\mathbb{L}}_{2} and coupling constant J0.J\geq 0. on each edge. If

(23) J<12ln(11pc,2w),\displaystyle J<\frac{1}{2}\ln\left(\frac{1}{1-p_{c,2}^{w}}\right),

then for any Gibbs measure μ\mu for the Ising model on 𝕃2{\mathbb{L}}_{2} with coupling constant JJ, μ\mu-a.s. there are infinitely many infinite “++”-clusters, infinitely many infinite “-”-clusters and infinitely many infinite contours. Here pc,2wp_{c,2}^{w} is defined as in (15).

We can see that when the conditions of Corollary 3.4 are satisfied, almost surely there are no infinite open clusters in the corresponding random cluster representation of the Ising model, however, the conclusion of the corollary says that there are infinitely many infinite “++”-clusters and infinitely many infinite “-”-clusters in the Ising model.

4. XOR Ising model on transitive, triangular tilings of the hyperbolic plane

In this section, we state the main result concerning the percolation properties of the XOR Ising model on transitive, triangular tilings of the hyperbolic plane. These results, as given in Theorems 4.1 and 4.2, will be proved in Section 12 as applications of Theorem 2.2.

Throughout this section, we let 𝕃1{\mathbb{L}}_{1} be the [n,n,n][n,n,n] regular tiling of the hyperbolic plane, such that each face has degree n7n\geq 7, and each vertex has degree 3. Let 𝕃2{\mathbb{L}}_{2} be the planar dual graph of 𝕃1{\mathbb{L}}_{1}. More precisely, 𝕃2{\mathbb{L}}_{2} is the vertex-transitive triangular tiling of the hyperbolic plane such that each vertex has degree n7n\geq 7. An XOR Ising model on 𝕃2{\mathbb{L}}_{2} is a probability measure on σXOR{±1}V(𝕃2)\sigma_{XOR}\in\{\pm 1\}^{V({\mathbb{L}}_{2})}, such that

σXOR(v)=σ1(v)σ2(v),vV(𝕃2),\displaystyle\sigma_{XOR}(v)=\sigma_{1}(v)\sigma_{2}(v),\qquad\forall v\in V({\mathbb{L}}_{2}),

where σ1\sigma_{1}, σ2\sigma_{2} are two i.i.d. Ising models with spins located on V(𝕃2)V({\mathbb{L}}_{2}). A contour configuration of the XOR Ising configuration on 𝕃2{\mathbb{L}}_{2} is a subset of edges of 𝕃1{\mathbb{L}}_{1} in which each edge has a dual edge in E(𝕃2)E({\mathbb{L}}_{2}) joining two vertices u,vV(𝕃2)u,v\in V({\mathbb{L}}_{2}) satisfying σXOR(u)=σXOR(v)\sigma_{XOR}(u)=-\sigma_{XOR}(v). A connected component in a contour configuration is called a contour. Obviously each vertex of 𝕃1{\mathbb{L}}_{1} has 0 or 2 incident present edges in a contour configuration of an XOR Ising configuration, since 𝕃1{\mathbb{L}}_{1} has vertex degree 3. Each contour of an XOR Ising configuration on 𝕃2{\mathbb{L}}_{2} is either a self-avoiding cycle or a doubly-infinite self-avoiding path.

We can similarly define an XOR Ising model with spins located on vertices of 𝕃1{\mathbb{L}}_{1}, and its contours to be even-degree subgraphs of 𝕃2{\mathbb{L}}_{2}.

Theorem 4.1.

Let σ1\sigma_{1}, σ2\sigma_{2} be two i.i.d. Ising models with spins located on vertices of 𝕃2{\mathbb{L}}_{2}, coupling constant J[0,)J\in[0,\infty) and free boundary conditions. Let μ1f\mu_{1}^{f} (resp. μ2f\mu_{2}^{f}) be the distribution of σ1\sigma_{1} (resp. σ2\sigma_{2}). Assume that one of the following cases occurs

  1. I

    If μ1f×μ2f\mu_{1}^{f}\times\mu_{2}^{f} is Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-ergodic; or

  2. II

    lim inf|ij|σ1,iσ1,jμ1,f=0\liminf_{|i-j|\rightarrow\infty}\langle\sigma_{1,i}\sigma_{1,j}\rangle_{\mu_{1,f}}=0, where σ1,i\sigma_{1,i} and σ1,j\sigma_{1,j} are two spins in the Ising model σ1\sigma_{1} with distance |ij||i-j|; or

  3. III

    JJ satisfies Condition (c) of Theorem 3.3 II.

  4. IV

    JJ sasifies Condition (d) of Theorem 3.3 II.

then μ1f×μ2f\mu_{1}^{f}\times\mu_{2}^{f}-a.s. there are infinitely many infinite “++”-clusters and infinitely many infinite “-”-clusters.

Theorem 4.2.

Let σ1\sigma_{1}, σ2\sigma_{2} be two i.i.d. Ising models with spins located on vertices of 𝕃1{\mathbb{L}}_{1}, and coupling constant K0K\geq 0. For i=1,2i=1,2, let μi,+\mu_{i,+} (resp. μi,\mu_{i,-}) be the distribution of σi\sigma_{i} with “++”-boundary conditions (resp. “-”-boundary conditions). Let J0J\geq 0 be given by

(24) e2J=1e2K1+e2K,\displaystyle e^{-2J}=\frac{1-e^{-2K}}{1+e^{-2K}},

and let tt be the number of infinite contours. Let μ++\mu_{++} (resp. μ\mu_{--}, μ+\mu_{+-}) be the product measure of μ1,+\mu_{1,+} and μ2,+\mu_{2,+} (resp. μ1,\mu_{1,-} and μ2,\mu_{2,-}, μ1,+\mu_{1,+} and μ2,\mu_{2,-}). Assume JJ satisfies the assumption of Theorem 4.1, then we have

μ++(t{0,})=μ(t{0,})=μ+(t{0,})=1.\displaystyle\mu_{++}(t\in\{0,\infty\})=\mu_{--}(t\in\{0,\infty\})=\mu_{+-}(t\in\{0,\infty\})=1.

5. XOR Ising models on the hexagonal and triangular lattices

In this section, we define the XOR Ising models on the hexagonal and triangular lattices, and state the main results proved in this paper concerning the percolation properties of these models.

Let σ1\sigma_{1}, σ2\sigma_{2} be two i.i.d. ferromagnetic Ising models with spins located on vertices of the hexagonal lattice =(V,E){\mathbb{H}}=(V_{{\mathbb{H}}},E_{{\mathbb{H}}}). The hexagonal lattice has edges in three different directions. Assume that both σ1\sigma_{1} and σ2\sigma_{2} have nonnegative coupling constants JaJ_{a}, JbJ_{b}, JcJ_{c} on edges of {\mathbb{H}} with the three different directions, respectively. Assume also that the distributions of both σ1\sigma_{1} and σ2\sigma_{2} are weak limits of Gibbs measures under periodic boundary conditions. Recall that the XOR Ising model σXOR(v)=σ1(v)σ2(v)\sigma_{XOR}(v)=\sigma_{1}(v)\sigma_{2}(v), for vVv\in V_{{\mathbb{H}}}.

A contour configuration for an XOR Ising configuration, σXOR\sigma_{XOR}, defined on {\mathbb{H}} (resp. 𝕋{\mathbb{T}}), is a subset of {0,1}E(𝕋)\{0,1\}^{E({\mathbb{T}})} (resp. {0,1}E()\{0,1\}^{E({\mathbb{H}})}), whose state-1-edges (present edges) are edges of 𝕋{\mathbb{T}} (resp. {\mathbb{H}}) separating neighboring vertices of {\mathbb{H}} (resp. 𝕋{\mathbb{T}}) with different states in σXOR\sigma_{XOR}. (Note that {\mathbb{H}} and 𝕋{\mathbb{T}} are planar duals of each other.) Contour configurations of the XOR Ising model were first studied in [49], in which the scaling limits of contours of the critical XOR Ising model are conjectured to be level lines of Gaussian free field. It is proved in [9] that the contours of the XOR Ising model on a plane graph correspond to level lines of height functions of the dimer model on a decorated graph, inspired by the correspondence between Ising model and bipartite dimer model in [13]. We will study the percolation properties of the XOR Ising model on {\mathbb{H}} and 𝕋{\mathbb{T}}, as an application of the main theorems proved in this paper for the general constrained percolation process.

Let

(25) f(x,y,z)=e2x+e2y+e2z+e2(x+y)+e2(x+z)+e2(y+z)e2(x+y+z)1.\displaystyle f(x,y,z)=e^{-2x}+e^{-2y}+e^{-2z}+e^{-2(x+y)}+e^{-2(x+z)}+e^{-2(y+z)}-e^{-2(x+y+z)}-1.
(26) g(x,y,z)=e2x+e2y+e2ze2(x+y+z).\displaystyle g(x,y,z)=e^{2x}+e^{2y}+e^{2z}-e^{2(x+y+z)}.

We say the XOR Ising model on {\mathbb{H}} with coupling constants (Ja,Jb,Jc)(J_{a},J_{b},J_{c}) is in the high-temperature state (resp. low-temperature state, critical state) if f(Ja,Jb,Jc)>0f(J_{a},J_{b},J_{c})>0 (resp. f(Ja,Jb,Jc)<0f(J_{a},J_{b},J_{c})<0, f(Ja,Jb,Jc)=0f(J_{a},J_{b},J_{c})=0). Note that f(Ja,Jb,Jc)=0f(J_{a},J_{b},J_{c})=0 is the well-known condition that an Ising model on the 2D hexagonal lattice {\mathbb{H}} is critical; see, for example, [35] for a rigorous proof. The XOR Ising model σ1σ2\sigma_{1}\cdot\sigma_{2} on {\mathbb{H}} is in the high-temperature state (resp. low-temperature state, critical state), if and only if both σ1\sigma_{1} and σ2\sigma_{2} are in the high-temperature state (resp. low-temperature state, critical state).

Let 𝕋=(V𝕋,E𝕋){\mathbb{T}}=(V_{{\mathbb{T}}},E_{{\mathbb{T}}}) be the dual triangular lattice of {\mathbb{H}}. We also consider the XOR Ising model with spins located on V𝕋V_{{\mathbb{T}}}. Assume that the coupling constants on edges with 3 different directions are KaK_{a}, KbK_{b} and KcK_{c}, respectively, such that Ka,Kb,Kc0K_{a},K_{b},K_{c}\geq 0. Also for i{a,b,c}i\in\{a,b,c\}, assume that KiK_{i} is the coupling constant on an edge of 𝕋{\mathbb{T}} dual to an edge of {\mathbb{H}} with coupling constant JiJ_{i}. We say the XOR Ising model on the triangular lattice is in the low-temperature state (resp. high-temperature state, critical state) if g(Ka,Kb,Kc)<0g(K_{a},K_{b},K_{c})<0 (resp. g(Ka,Kb,Kc)>0g(K_{a},K_{b},K_{c})>0, g(Ka,Kb,Kc)=0g(K_{a},K_{b},K_{c})=0). Again these come from the known fact that if g(Ka,Kb,Kc)<0g(K_{a},K_{b},K_{c})<0 (resp. g(Ka,Kb,Kc)>0g(K_{a},K_{b},K_{c})>0, g(Ka,Kb,Kc)=0g(K_{a},K_{b},K_{c})=0), both Ising models, each of which is a factor of the XOR Ising model, are in the low-temperature state (resp. high-temperature state, critical state).

Similar to the square grid case, in the high temperature state, the Ising model on the hexagonal lattice or the triangular lattice has a unique Gibbs measure, and the spontaneous magnetization vanishes; while in the low temperature state, the Gibbs measures are not unique and the spontaneous magnetization is strictly positive under the “++”-boundary condition. See [32, 1, 35, 14].

If

(27) e2Kτ=1e2Jτ1+e2Jτ,forτ=a,b,c,\displaystyle e^{-2K_{\tau}}=\frac{1-e^{-2J_{\tau}}}{1+e^{-2J_{\tau}}},\qquad\mathrm{for}\ \tau=a,b,c,

then the XOR Ising model on {\mathbb{H}} with coupling constants (Ja,Jb,Jc)(J_{a},J_{b},J_{c}) is in the low-temperature state (resp. high-temperature state, critical state) if and only if the XOR Ising model on the triangular lattice with coupling constants (Ka,Kb,Kc)(K_{a},K_{b},K_{c}) is in the high-temperature state (resp. low-temperature state, critical state).

We define clusters and contours with respect to an XOR Ising configuration on {\mathbb{H}} or 𝕋{\mathbb{T}} in the usual way. Then we have the following theorems.

Theorem 5.1.

Consider the critical XOR Ising model on {\mathbb{H}} or 𝕋{\mathbb{T}}. Then

  1. I

    almost surely there are no infinite clusters;

  2. II

    almost surely there are no infinite contours.

Theorem 5.2.

In the low-temperature XOR Ising model on {\mathbb{H}} or on 𝕋{\mathbb{T}}, almost surely there are no infinite contours.

Theorems 5.1 and 5.2 are proved in Section 13.

6. Square tilings of the hyperbolic plane

In this section, we introduce the square tilings of the hyperbolic plane, and then state and prove properties of the constrained percolation models on such graphs. We first discuss known results about percolation on non-amenable graphs that will be used to prove main theorems of the paper.

The following lemma is proved in [6, 4].

Lemma 6.1.

Let G be a quasi-transitive, non-amenable, planar graph with one end, and let ω\omega be an invariant percolation on G. Then a.s. the number of infinite 1-clusters of ω\omega is 0, 1, or \infty.

Proof.

See Lemma 3.5 of [6]. ∎

Lemma 6.2.

(Threshold for bond percolation on non-amenable graphs) Let G=(V,E)G=(V,E) be a non-amenable graph. Let ΓAut(G)\Gamma\subseteq\mathrm{Aut}(G) be a closed unimodular quasi-transitive subgroup, and let o1,,oLo_{1},\ldots,o_{L} be a complete set of representatives in VV of the orbits of GG. For 1iL1\leq i\leq L, let Staboi\mathrm{Stab}_{o_{i}} is defined as in (6) and

ηi:\displaystyle\eta_{i}: =\displaystyle= |Staboi|.\displaystyle|\mathrm{Stab}_{o_{i}}|.

Let \mathbb{P} be a bond percolation on GG whose distribution is Γ\Gamma-invariant. Let DiD_{i} be the random degree of oio_{i} in the percolation subgraph, and let did_{i} be the degree of oio_{i} in GG. Write p,vp_{\infty,v} for the probability that vVv\in V is in an infinite component. Let p,ip_{\infty,i} be the probability that oio_{i} is in an infinite cluster. Then

(28) i=1L(diα(G))p,iηij=1L𝔼Djα(G)ηj\displaystyle\sum_{i=1}^{L}\frac{(d_{i}-\alpha(G))p_{\infty,i}}{\eta_{i}}\geq\sum_{j=1}^{L}\frac{\mathbb{E}D_{j}-\alpha(G)}{\eta_{j}}

where α(G)\alpha(G) is a constant depending on the structure of the graph GG defined by

αK:\displaystyle\alpha_{K}: =\displaystyle= 1|K|xKdegK(x)\displaystyle\frac{1}{|K|}\sum_{x\in K}\deg_{K}(x)
α(G):\displaystyle\alpha(G): =\displaystyle= sup{αK:KGisfinite}\displaystyle\sup\{\alpha_{K}:K\subset G\ \mathrm{is\ finite}\}

In particular, if the right-hand side of (28) is positive, then there is an infinite component in the percolation subgraph with positive probability.

Proof.

See Theorem 4.1 of [4]. ∎

Let G=(V,E)G=(V,E) be a graph corresponding to a square tiling of the hyperbolic plane. Assume that

  1. I

    each face of GG has 4 edges; and

  2. II

    each vertex of GG is incident to 2n2n faces, where n3n\geq 3.

See Figure 9 for an example of such a graph when n=3n=3.

Refer to caption
Figure 9. The [4,4,4,4,4,4] lattice: each face has degree 4, and each vertex have degree 6

We can color all the faces of GG by black and white such that black faces can share edges only with white faces and vice versa. Let G=(V,E)G=(V,E) denote the graph embedded into the hyperbolic plane as described above.

We consider the site configurations in {0,1}V\{0,1\}^{V}. We impose the following constraint on site configurations

  • Around each black face, there are six allowed configurations (0000)(0000), (1111)(1111), (0011)(0011), (1100)(1100), (0110)(0110), (1001)(1001), where the digits from the left to the right correspond to vertices in clockwise order around the black face, starting from the lower left corner. See Figure 4.

Let Ω{0,1}V\Omega\subset\{0,1\}^{V} be the set of all configurations satisfying the constraint above. We use Ω\Omega to denote the sample space throughout this paper, however, Ω\Omega have different meanings in different sections.

Note that GG is a vertex-transitive graph. Since each face of GG has an even number of edges, GG itself is a bipartite graph - we can color the vertices of GG by red and green such that red vertices are adjacent only to green vertices and vice versa. We assign an integer in 1,2,,n1,2,\ldots,n to each white face of GG according to the following rules

  1. I

    around each red vertex of GG, white faces are assigned integers 1,2,,n1,2,\ldots,n clockwise; and

  2. II

    around each green vertex of GG, white faces are assigned integers 1,2,,n1,2,\ldots,n counterclockwise; and

  3. III

    any two white faces adjacent to the same black face along two opposite edges have the same assigned integer.

See Figure 10 for an example of assignments of integers 1,2,31,2,3 to the white faces of the [4,4,4,4,4,4][4,4,4,4,4,4] lattice.

Refer to caption
Figure 10. Labels of white faces of the [4,4,4,4,4,4] lattice

For 1in1\leq i\leq n, we construct a graph 𝕃i{\mathbb{L}}_{i} as follows. The vertex set of 𝕃i{\mathbb{L}}_{i} consists of all the white faces of GG whose assigned integers are ii. Two vertices of 𝕃i{\mathbb{L}}_{i} are joined by an edge of 𝕃i{\mathbb{L}}_{i} if and only if they correspond to two white faces of GG adjacent to the same black face along two opposite edges. We have the following proposition regarding the connected components of 𝕃i{\mathbb{L}}_{i}

Proposition 6.3.

When n3n\geq 3, each component of 𝕃i{\mathbb{L}}_{i} (1in1\leq i\leq n) is a regular tree of degree 4. For any integer ii satisfying 1in1\leq i\leq n, the edges of 𝕃i1{\mathbb{L}}_{i-1} (if i=1i=1, 𝕃i1:=𝕃n{\mathbb{L}}_{i-1}:={\mathbb{L}}_{n}) and 𝕃i{\mathbb{L}}_{i} cross; the edges of 𝕃i+1{\mathbb{L}}_{i+1} (if i=ni=n, 𝕃i+1:=𝕃1{\mathbb{L}}_{i+1}:={\mathbb{L}}_{1}) and 𝕃i{\mathbb{L}}_{i} cross.

Proof.

We consider a doubly infinite sequence of edges in GG consisting of edges ,e1,e0,e1,e2,,\ldots,e_{-1},e_{0},e_{1},e_{2},\ldots, such that

  • For each kk\in{\mathbb{Z}}, eke_{k} and ek+1e_{k+1} share a vertex vv, such that there are exactly (n1)(n-1) edges incident to vv to the left of eke_{k} and ek+1e_{k+1}, and (n1)(n-1) edges incident to vv to the right of eke_{k} and ek+1e_{k+1}.

Then ,e1,e0,e1,\ldots,e_{-1},e_{0},e_{1},\ldots form a doubly infinite self-avoiding path in GG because its left side and right side are symmetric. Indeed, if the path crosses itself, starting from e0e_{0}, we move the path along both the positive direction e1,e2,e_{1},e_{2},\ldots and the negative direction e1,e2,e_{-1},e_{-2},\ldots, until the first time the movements along the two directions meet, and form a cycle 𝒞ab:=ea,ea+1,,e0,,eb1,eb\mathcal{C}_{ab}:=e_{-a},e_{-a+1},\ldots,e_{0},\ldots,e_{b-1},e_{b}, where a,b{0,1,2,}a,b\in\{0,1,2,\ldots\}. Then G𝒞abG\setminus\mathcal{C}_{ab} has a finite component and an infinite component; moving from eae_{-a} to ebe_{b} along 𝒞ab\mathcal{C}_{ab}, the finite component is either on the left or on the right, but this is a contradiction to the fact that on the left and right side of ,e1,e0,e1,e2,\ldots,e_{-1},e_{0},e_{1},e_{2},\ldots, GG is symmetric. We call the infinite self-avoiding path obtained this way a central path.

Assume there is a cycle in 𝕃i{\mathbb{L}}_{i} for some 1in1\leq i\leq n, then we can find a face in 𝕃i{\mathbb{L}}_{i}. Let (u,v)(u,v) be an edge of 𝕃i{\mathbb{L}}_{i}. Moving from uu to vv, at vv there are 3 other incident edges except the edge (u,v)(u,v); since the graph is embedded in the hyperbolic plane, we may label the three incident edges at vv other than (u,v)(u,v) by the left edge, the middle edge, and the right edge, in such a way that starting from the edge (u,v)(u,v) and moving around vv clockwise along a small circle, one will cross the left edge first, then the middle edge, and finally the right edge. If we can find a face in 𝕃i{\mathbb{L}}_{i}, then the face can be found by always moving along the right edge at each vertex for finitely many times, and the face is on the right of an oriented cycle obtained this way. But when n3n\geq 3, this is not possible since any oriented path in 𝕃i{\mathbb{L}}_{i} obtained by always moving along the right edge at each vertex has a central path on its right, which is infinite.

Note that each black face of GG has two pairs of opposite edges. There exists 1in1\leq i\leq n, such that along one pair of opposite edges the black face is adjacent to two white faces labeled by ii, and along the other pair of opposite edges the black face is adjacent to two white faces labeled by (i+1)(i+1) (if i=ni=n, then i+1=1i+1=1). Then from the construction of 𝕃i{\mathbb{L}}_{i}’s we can see that an edge of 𝕃i{\mathbb{L}}_{i} and an edge of 𝕃i+1{\mathbb{L}}_{i+1} cross at the black face of GG. ∎

Any constrained percolation configuration in Ω\Omega gives rise to a contour configuration on i=1n𝕃i\cup_{i=1}^{n}{\mathbb{L}}_{i}. An edge ee in i=1n𝕃i\cup_{i=1}^{n}{\mathbb{L}}_{i} is present in the contour configuration if and only if it crosses a black face bb in GG, such that the states of the vertices of bb on the two sides separated by ee in the configuration are different, and any two vertices of bb on the same side of ee have the same state. This is a contour configuration satisfying the condition that each vertex in i=1n𝕃i\cup_{i=1}^{n}{\mathbb{L}}_{i} has an even number of incident present edges. For any 1i<jn1\leq i<j\leq n, present edges in 𝕃i{\mathbb{L}}_{i} and 𝕃j{\mathbb{L}}_{j} can never cross.

A cluster is a maximal connected set of vertices in GG in which every vertex has the same state in a constrained percolation configuration. A contour is a maximal connected set of edges in i=1n𝕃i\cup_{i=1}^{n}{\mathbb{L}}_{i} in which every edge is present in the contour configuration. Note that each contour must be a connected subgraph of 𝕃i{\mathbb{L}}_{i}, for some 1in1\leq i\leq n. Hence by Proposition 6.3, each contour must be a tree. Since each vertex in a contour has an even number of incident present edges in the contour, each contour must be an infinite tree.

Let μ\mu be a probability measure on Ω\Omega. We may assume that μ\mu satisfies the following conditions

  1. (D1)

    μ\mu is Aut(G)\mathrm{Aut}(G)-invariant;

  2. (D2)

    μ\mu is Aut(𝕃i)\mathrm{Aut}({\mathbb{L}}_{i})-ergodic, for 1in1\leq i\leq n;

  3. (D3)

    μ\mu is symmetric, i.e. let θ:ΩΩ\theta:\Omega\rightarrow\Omega be the map defined by θ(ω)(v)=1ω(v)\theta(\omega)(v)=1-\omega(v), for each vVv\in V, then μ\mu is invariant under θ\theta, that is, for any event AA, μ(A)=μ(θ(A))\mu(A)=\mu(\theta(A)).

Note that when n3n\geq 3, the graph GG is a non-amenable group. Recall that the number of ends of a connected graph is the supremum over its finite subgraphs of the number of infinite components that remain after removing the subgraph.

Here is the main theorem concerning the properties of constrained percolations on the square tilings of the hyperbolic plane.

Theorem 6.4.
  1. (a)

    Let μ\mu be a probability measure on Ω\Omega satisfying (D1). Let n0n_{0} (resp. n1n_{1}) be the number of infinite 0-clusters (resp. 1-clusters). Then μ\mu-a.s. (n0,n1){(0,1),(1,0),(1,),(,1),(,)}(n_{0},n_{1})\in\{(0,1),(1,0),(1,\infty),(\infty,1),(\infty,\infty)\}.

  2. (b)

    Let ν\nu be a probability measure on Ω\Omega satisfying (D1) - (D3). Then ν\nu-a.s. there are infinitely many infinite 0-clusters and infinitely many infinite 1-clusters.

In order to prove 6.4, we first prove a few lemmas.

Lemma 6.5.

In a contour configuration in i=1n𝕃i\cup_{i=1}^{n}{\mathbb{L}}_{i} as described above, any contour must be an infinite tree (a tree consisting of infinite many edges of i=1n𝕃i\cup_{i=1}^{n}{\mathbb{L}}_{i}) in which each vertex has degree 2 or 4.

Proof.

This lemma is straightforward from the facts that each contour is a connected subgraph of 𝕃i{\mathbb{L}}_{i} for some i{1,,n}i\in\{1,\ldots,n\}; each component of 𝕃i{\mathbb{L}}_{i} 1in1\leq i\leq n is a regular tree of degree 4, and each vertex in a contour has an even number of incident present edges. ∎

Lemma 6.6 below is proved in [4] and [3] using the mass transport principle.

Lemma 6.6.

Let G be a nonamenable graph whose automorphism group has a closed subgroup acting transitively and unimodularly on GG, and let ω\omega be an invariant percolation on G which has a single component a.s. Then pc(ω)<1p_{c}(\omega)<1 a.s., where pc()p_{c}(\cdot) is the critical i.i.d. Bernoulli percolation probability on a graph.

Proof.

See Theorem 1.5 of [4]. ∎


Proof of Theorem 6.4 First we show that Part (a) of the theorem together with Assumptions (D2), (D3) implies Part (b). Let ν\nu be a probability measure on Ω\Omega satisfying (D1) - (D3). By Assumption (D2) and (D3), there exists a positive integer kk (possibly infinite), such that ν(n0=n1=k)=1\nu(n_{0}=n_{1}=k)=1. Then Part (b) follows from Part (a).

Now we prove Part (a). Obviously (n0,n1){(0,1),(1,0)}(n_{0},n_{1})\in\{(0,1),(1,0)\} if there are no contours. Now assume that contours do exist. By Lemma 6.1, n0,n1{0,1,}n_{0},n_{1}\in\{0,1,\infty\}. By Lemma A.12, n0,n1{1,}n_{0},n_{1}\in\{1,\infty\}. Let ϕ\phi be the contour configuration. If there are infinitely many contours in ϕ\phi, or there exists a contour of ϕ\phi in which infinitely many vertices have degree 4, then 2ϕ{\mathbb{H}}^{2}\setminus\phi has infinitely many unbounded components. By Lemma A.13, n0+n1=n_{0}+n_{1}=\infty. Therefore {n0,n1}{(1,),(,1),(,)}\{n_{0},n_{1}\}\in\{(1,\infty),(\infty,1),(\infty,\infty)\} in this case.

Now consider the case that the number of contours is finite and nonzero, and on each contour only finitely many vertices have degree 4. Fix an ii satisfying 1in1\leq i\leq n, and conditional on the event that the number of contours on 𝕃i{\mathbb{L}}_{i} is finite and nonzero. Choose a contour τ\tau on 𝕃i{\mathbb{L}}_{i} uniformly at random; then τ\tau forms an invariant bond percolation on 𝕃i{\mathbb{L}}_{i} which has a single component. By Lemma 6.6, almost surely τ\tau has infinitely many vertices with degree 4 - since otherwise pc(τ)=1p_{c}(\tau)=1. Therefore this case does not occur a.s. \hfill\Box

7. Proof of Theorem 2.2

In this section, we prove Theorem 2.2. The idea of the proof is to consider all the possible values of (s0,s1,t1,t2)(s_{0},s_{1},t_{1},t_{2}) and exclude those with probability 0 to occur using the symmetry and ergodicity of the probability measure. In Lemma 7.1, we exclude the case t1=t2=1t_{1}=t_{2}=1; the proof is based on constructing a superimpostion G^\hat{G} of the lattice 𝕃1{\mathbb{L}}_{1} and its dual lattice 𝕃2{\mathbb{L}}_{2}; and the union of contour configurations on 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} form an invariant bond percolation on G^\hat{G}, in which the number of infinite clusters can only be 0,1,0,1,\infty by Lemma 6.1 a.s.; however, if t1=t2=1t_{1}=t_{2}=1, since the contour configurations on 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} do not cross each other, the number of infinite clusters in the union would be 22. Proposition 7.2 excludes the case when (s0,s1,t1,t2)=(0,0,0,0)(s_{0},s_{1},t_{1},t_{2})=(0,0,0,0), the proof applies planarity to obtain an infinite sequence of contours, one surrounding another, and then obtain a contradiction with non-amenability. Lemma 7.3 excludes the case (t1,t2)=(0,k)(t_{1},t_{2})=(0,k) and (t1,t2)=(k,0)(t_{1},t_{2})=(k,0) for 1k1\leq k\leq\infty by ergodicity, symmetry and planarity. Lemma 7.4 excludes the case that (s0,s1)=(1,1)(s_{0},s_{1})=(1,1) again by constructing an invariant bound percolation on GG with 2 infinite clusters and obtaining a contradiction to Lemma 6.1. In the proof of Theorem 2.2, we use symmetry, ergodicity, Lemma 6.1 and Lemma 7.4 to obtain that a.s. (s0,s1){(0,0),(,)}(s_{0},s_{1})\in\{(0,0),(\infty,\infty)\}; to rule out the case (s0,s1)=(0,0)(s_{0},s_{1})=(0,0), we apply Lemma 6.1 again to show that if (s0,s1)=(0,0)(s_{0},s_{1})=(0,0), then (s0,s1){(0,0,0),(0,0,1),(0,0,)}(s_{0},s_{1})\in\{(0,0,0),(0,0,1),(0,0,\infty)\}, we then show that each of the cases has probability 0 to occur by applying Lemmas 7.1 and 7.3 and Proposition 7.2.

We start with Lemma 7.1.

Lemma 7.1.

Let GG be the [m,4,n,4][m,4,n,4] lattice satisfying (1) m3,n3m\geq 3,n\geq 3 and

(29) 1m+1n<12.\displaystyle\frac{1}{m}+\frac{1}{n}<\frac{1}{2}.

Let μ\mu be a probability measure on Ω\Omega satisfying (A1). Let t1t_{1} (resp. t2t_{2}) be the number of infinite 𝕃1{\mathbb{L}}_{1}-contours (resp. 𝕃2{\mathbb{L}}_{2}-contours). Then

μ((t1,t2)=(1,1))=0.\displaystyle\mu((t_{1},t_{2})=(1,1))=0.
Proof.

The proof is inspired by the proof of Corollary 3.6 of [6].

We embed 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} in the hyperbolic plane in such a way that every edge ee intersects its dual edge ee^{*} at one point vev_{e}, and there are no other intersections of 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2}. We define a new graph G^=(V^,E^)\hat{G}=(\hat{V},\hat{E}), where V^=V(𝕃1)V(𝕃2){ve,eE(𝕃1)}\hat{V}=V({\mathbb{L}}_{1})\cup V({\mathbb{L}}_{2})\cup\{v_{e},e\in E({\mathbb{L}}_{1})\}, and an edge in E^\hat{E} is either a half-edge of E(𝕃1)E({\mathbb{L}}_{1}) joining a vertex in V(𝕃1)V({\mathbb{L}}_{1}) and a vertex in {ve,eE(𝕃1)}\cup\{v_{e},e\in E({\mathbb{L}}_{1})\}, or a half-edge of E(𝕃2)E({\mathbb{L}}_{2}) joining a vertex in V(𝕃2)V({\mathbb{L}}_{2}) and a vertex in {ve,eE(𝕃1)}\cup\{v_{e},e\in E({\mathbb{L}}_{1})\}.

For i{1,2}i\in\{1,2\}, let ϕiΦi\phi_{i}\in\Phi_{i} be the random contour configuration restricted on 𝕃i{\mathbb{L}}_{i}. Let

ϕ^:={[v,ve]E^:vV(𝕃1),eϕ1}{[v,ve]E^:vV(𝕃2),eϕ2}\displaystyle\hat{\phi}:=\{[v,v_{e}]\in\hat{E}:v\in V({\mathbb{L}}_{1}),e\in\phi_{1}\}\cup\{[v_{*},v_{e}]\in\hat{E}:v_{*}\in V({\mathbb{L}}_{2}),e_{*}\in\phi_{2}\}

We say ϕ^\hat{\phi} is a contour configuration on G^\hat{G}, and each connected component of ϕ^\hat{\phi} is called a contour. Then ϕ^\hat{\phi} is an invariant bond percolation on the quasi-transitive, non-amenable, planar, one-ended graph G^\hat{G}. Note that the number of infinite components of ϕ^\hat{\phi} is the number of infinite contours of ϕ1\phi_{1} plus the number of infinite contours of ϕ2\phi_{2}. If there is a positive probability that (t0,t1)=(1,1)(t_{0},t_{1})=(1,1), then the number of infinite components in ϕ^\hat{\phi} is 2. This contradicts Lemma 6.1, which says that the number of infinite components in the invariant percolation ϕ^\hat{\phi} on the quasi-transitive, one-ended, nonamenable, planar graph G^\hat{G} can only be 0,10,1 or \infty. ∎

Proposition 7.2.

Let GG be the [m,4,n,4][m,4,n,4] lattice with m,nm,n satisfying (1) m3,n3m\geq 3,\ n\geq 3 and (29) 1m+1n<12\frac{1}{m}+\frac{1}{n}<\frac{1}{2}. Let ωΩ\omega\in\Omega be a Γ\Gamma-invariant, Γ1\Gamma_{1}-ergodic constrained percolation on GG. Let s0s_{0} (resp. s1s_{1}) be the number of infinite 0-clusters (resp. 1-clusters) in ω\omega, and let t1t_{1} (resp. t2t_{2}) be the number of infinite 𝕃1{\mathbb{L}}_{1}-contours (resp. infinite 𝕃2{\mathbb{L}}_{2}-contours) in ω\omega. Then almost surely (s0,s1,t1,t2)(0,0,0,0)(s_{0},s_{1},t_{1},t_{2})\neq(0,0,0,0).

Proof.

The proof is inspired by Lemma 3.3 of [6]. Let G^=(V^,E^)\hat{G}=(\hat{V},\hat{E}), ϕ^\hat{\phi} be defined as in the proof of Lemma 7.1. Note that when m,nm,n satisfy (1) and (29), G^\hat{G} is a quasi-transitive, non-amenable, planar and one-ended graph; and that the [m,4,n,4][m,4,n,4] lattice is exactly the dual graph of G^\hat{G}. It is also known that quasi-transitive planar graphs with one end are unimodular; see [38].

Define a generalized contour in a contour configuration ϕ^\hat{\phi} of G^\hat{G} to be either a single vertex in V^\hat{V} which has no incident present edges in ϕ^\hat{\phi}, or a contour in ϕ^\hat{\phi}. This way each vertex vV^v\in\hat{V} has a unique generalized contour in ϕ^\hat{\phi} passing through the vertex vv.

Suppose that (s0,s1,t1,t2)=(0,0,0,0)(s_{0},s_{1},t_{1},t_{2})=(0,0,0,0) a.s. Then a.s. given a generalized contour CC of ϕ^\hat{\phi}, there is a cluster CC^{\prime} of ω\omega surrounding it. Similarly, for every cluster CC in ω\omega, there is a contour CC^{\prime} in ϕ^\hat{\phi} that surrounds it. Let 𝒞0\mathcal{C}_{0} denote the set of all generalized contours of ϕ^\hat{\phi}. We set

𝒞j+1:={C′′:C𝒞j};\displaystyle\mathcal{C}_{j+1}:=\{C^{\prime\prime}:C\in\mathcal{C}_{j}\};

in which CC is a generalized contour, CC^{\prime} is a cluster, and C′′C^{\prime\prime} is a contour. For C𝒞0C\in\mathcal{C}_{0} and vV^v\in\hat{V}, let r(C):=sup{j:C𝒞j}r(C):=\sup\{j:C\in\mathcal{C}_{j}\}, and define r(v):=r(C)r(v):=r(C) if CC is the generalized contour of vv in ϕ^\hat{\phi}. Intuitively, we may consider r(C)r(C) as the maximal length of sequences of nesting contours, in which CC is the outermost contour.

Then there exist i{1,2}i\in\{1,2\} and a sequence of finite contours C1,C2,,Cn,C_{1},C_{2},\ldots,C_{n},\ldots in 𝕃i{\mathbb{L}}_{i}, such that Cn+1C_{n+1} surrounds CnC_{n}, and

limnr(Cn)=.\displaystyle\lim_{n\rightarrow\infty}r(C_{n})=\infty.

For each rr let ωr\omega^{r} be the set of edges in E(𝕃i)E({\mathbb{L}}_{i}) whose both endpoints u,vV(𝕃i)u,v\in V({\mathbb{L}}_{i}) satisfy r(v)rr(v)\leq r and r(u)rr(u)\leq r. Then ωr\omega^{r} is an invariant bond percolation and for any vV(𝕃i)v\in V({\mathbb{L}}_{i}),

deg𝕃iv=𝐄limr[degωrv]lim infr𝐄[degωrv]lim supr𝐄[degωrv]deg𝕃iv.\displaystyle\deg_{{\mathbb{L}}_{i}}v=\mathbf{E}\lim_{r\rightarrow\infty}[\deg_{\omega^{r}}v]\leq\liminf_{r\rightarrow\infty}\mathbf{E}[\deg_{\omega^{r}}v]\leq\limsup_{r\rightarrow\infty}\mathbf{E}[\deg_{\omega^{r}}v]\deg_{{\mathbb{L}}_{i}}v.

Note that 𝕃i{\mathbb{L}}_{i} is a transitive, non-amenable graph. We have

α(𝕃i)=deg𝕃ivıE(𝕃i)<deg𝕃iv\displaystyle\alpha({\mathbb{L}}_{i})=\mathrm{deg}_{{\mathbb{L}}_{i}}v-\imath_{E}({\mathbb{L}}_{i})<\mathrm{deg}_{{\mathbb{L}}_{i}}v

where ıE(𝕃i)\imath_{E}({\mathbb{L}}_{i}) is the edge isoperimetric constant defined as in (7), and α(𝕃i)\alpha({\mathbb{L}}_{i}) is defined in Lemma 6.2. By Lemma 6.2, the right hand side of (28) is strictly positive for sufficiently large rr; we deduce that ωr\omega^{r} has infinite components with positive probability for all sufficiently large rr.

However, since (s0,s1,t1,t2)=(0,0,0,0)(s_{0},s_{1},t_{1},t_{2})=(0,0,0,0), by the arguments above each vertex in 𝕃i{\mathbb{L}}_{i} is surrounded by infinitely many finite contours in 𝕃i{\mathbb{L}}_{i}. This implies that for any rr\in\mathbb{N}, for any vertex vV(𝕃i)v\in V({\mathbb{L}}_{i}), there exists a finite contour CC surrounding vv, such that r(C)>rr(C)>r, and therefore Cωr=C\cap\omega^{r}=\emptyset. As a result, the components in ωr\omega^{r} including vv is finite. Then the proposition follows from the contradiction. ∎

Lemma 7.3.

Let GG be the [m,4,n,4][m,4,n,4] lattice with m,nm,n satisfying (1), (29). Let (s0,s1,t1,t2)(s_{0},s_{1},t_{1},t_{2}) be given as in Theorem 2.2.

  1. I

    Let μ\mu be a probability measure on Ω\Omega satisfying (A2)(A7). Then

    μ((t1,t2)=(0,k))=0.\displaystyle\mu((t_{1},t_{2})=(0,k))=0.

    for any integer 1k1\leq k\leq\infty.

  2. II

    Let μ\mu be a probability measure on Ω\Omega satisfying (A2)(A6). Then

    μ((t1,t2)=(k,0))=0.\displaystyle\mu((t_{1},t_{2})=(k,0))=0.

    for any integer 1k1\leq k\leq\infty.

Proof.

We prove Part I here; Part II can be proved using exactly the same technique. By (A2) μ\mu is Γi\Gamma_{i} ergodic, either μ((t1,t2)=(0,k))=0\mu((t_{1},t_{2})=(0,k))=0 or μ((t1,t2)=(0,k))=1\mu((t_{1},t_{2})=(0,k))=1. Assume that μ((t1,t2)=(0,k))=1\mu((t_{1},t_{2})=(0,k))=1; we shall obtain a contradiction. Since there exists an infinite 𝕃2{\mathbb{L}}_{2}-contour; hence there exists an infinite cluster in λ2\lambda_{2} containing the infinite 𝕃2{\mathbb{L}}_{2}-contour. By (A7) λ2\lambda_{2} is Γ2\Gamma_{2}-ergodic, and the symmetry of λ2\lambda_{2}, there exist an infinite 0-cluster and an infinite 1-cluster in λ2\lambda_{2} a.s.. Note that the configuration in λ2{0,1}V(𝕃2)\lambda_{2}\in\{0,1\}^{V({\mathbb{L}}_{2})} naturally induces a configuration ωΩ\omega\in\Omega by the condition that the contour configurations corresponding to λ2\lambda_{2} and ω\omega are the same. We can see that if in λ2\lambda_{2} there exist both an infinite 0-cluster and an infinite 1-cluster, then in the induced constrained configuration ωΩ\omega\in\Omega, there is both an infinite 0-cluster and an infinite 1-cluster. By Lemma A.3, there exist an infinite 𝕃1{\mathbb{L}}_{1}-contour. But this is a contradiction to the fact that t1=0t_{1}=0. ∎

Lemma 7.4.

Let G=(V,E)G=(V,E) be the [m,4,n,4][m,4,n,4] lattice with m,nm,n satisfying (1) m3,n3m\geq 3,n\geq 3 and (29) 1m+1n<12\frac{1}{m}+\frac{1}{n}<\frac{1}{2}. Let μ\mu be a probability measure on Ω\Omega satisfying (A2), (A8). Let (s0,s1,t1,t2)(s_{0},s_{1},t_{1},t_{2}) be given as in Theorem 2.2. Then

μ((s0,s1)=(1,1))=0.\displaystyle\mu((s_{0},s_{1})=(1,1))=0.
Proof.

By (A2) μ\mu is Γi\Gamma_{i}-ergodic, either μ((s0,s1)=(1,1))=0\mu((s_{0},s_{1})=(1,1))=0 or μ((s0,s1)=(1,1))=1\mu((s_{0},s_{1})=(1,1))=1. Assume that μ((s0,s1)=(1,1))=1\mu((s_{0},s_{1})=(1,1))=1; we shall obtain a contradiction.

Let ωΩ\omega\in\Omega. We first construct a bond configuration ωb{0,1}E\omega_{b}\in\{0,1\}^{E} by letting an edge eEe\in E to be present if and only if it joins two edges in ω\omega with the same state; i.e. either both its endpoints have state 0; or both its endpoints have state 1. It is easy to check that the (0 or 1) clusters in ω\omega are exactly the components in ωb\omega_{b}. Then ωb\omega_{b} forms a Γ1\Gamma_{1}-invariant percolation on GG. If (s0,s1)=(1,1)(s_{0},s_{1})=(1,1), then ωb\omega_{b} has exactly two infinite components. But this is a contradiction to Lemma 6.1. ∎

Proof of Theorem 2.2 I. Assume that μ\mu is a probability measure on Ω\Omega satisfying (A2),(A3),(A7),(A8).

Let (s0,s1,t1,t2)(s_{0},s_{1},t_{1},t_{2}) be given as in the theorem. By Lemma 6.1, we have μ\mu-a.s. s0{0,1,}s_{0}\in\{0,1,\infty\}, s1{0,1,}s_{1}\in\{0,1,\infty\} and t1{0,1,}t_{1}\in\{0,1,\infty\}. By (A2) μ\mu is Γi\Gamma_{i}-ergodic and (A3) μ\mu is symmetric with respect to interchanging state “0” and state “1”, we have μ\mu-a.s. (s0,s1){(0,0),(1,1),(,)}(s_{0},s_{1})\in\{(0,0),(1,1),(\infty,\infty)\}. Hence we need to rule out the case that (s0,s1)=(1,1)(s_{0},s_{1})=(1,1) and the case that (s0,s1)=(0,0)(s_{0},s_{1})=(0,0). Almost surely we have (s0,s1)(1,1)(s_{0},s_{1})\neq(1,1) by Lemma 7.4. Now we show that almost surely (s0,s1)(0,0)(s_{0},s_{1})\neq(0,0).

We claim that μ\mu-a.s. t1{0,}t_{1}\in\{0,\infty\}. Assume that μ\mu-a.s. t1=1t_{1}=1, we shall obtain a contradiction. Let τ\tau be the unique infinite 𝕃1{\mathbb{L}}_{1}-contour. Then τ\tau forms an invariant bond percolation on 𝕃1{\mathbb{L}}_{1} which has a single component a.s.. By Lemma 6.6, pc(τ)<1p_{c}(\tau)<1 a.s. However, τ\tau is an even-degree subgraph of 𝕃1{\mathbb{L}}_{1} and 𝕃1{\mathbb{L}}_{1} has vertex-degree 3; as a result, τ\tau must be a doubly-infinite self-avoiding path. This is a contradiction to the fact that pc(τ)<1p_{c}(\tau)<1. Therefore we have either μ\mu-a.s. t1=0t_{1}=0 or μ\mu-a.s. t1=t_{1}=\infty.

If μ\mu-a.s. t1=t_{1}=\infty, let ϕ\phi be the contour configuration on 𝕃1𝕃2{\mathbb{L}}_{1}\cup{\mathbb{L}}_{2} corresponding to the constrained percolation configuration. Since each infinite contour in ϕ\phi is a doubly-infinite self-avoiding path, if there are infinitely many infinite contours, then 2ϕ{\mathbb{H}}^{2}\setminus\phi has infinitely many unbounded components. Note also that there exists an infinite cluster in each infinite component of 2ϕ{\mathbb{H}}^{2}\setminus\phi; hence μ\mu-a.s. (s0,s1,t1)=(,,)(s_{0},s_{1},t_{1})=(\infty,\infty,\infty) in this case.

Now consider the case that μ\mu-a.s. t1=0t_{1}=0.

We assume that μ\mu-a.s. (s0,s1,t1)=(0,0,0)(s_{0},s_{1},t_{1})=(0,0,0) and shall again obtain a contradiction. By Proposition 7.2, a.s. (s0,s1,t1,t2)(0,0,0,0)(s_{0},s_{1},t_{1},t_{2})\neq(0,0,0,0). Moreover, it is impossible to have (s0,s1,t1,t2)=(0,0,0,)(s_{0},s_{1},t_{1},t_{2})=(0,0,0,\infty) since if t2=t_{2}=\infty, then there are infinitely many infinite clusters. By Lemma 7.3, a.s. (s0,s1,t1,t2)(0,0,0,1)(s_{0},s_{1},t_{1},t_{2})\neq(0,0,0,1). Therefore μ((s0,s1,t1)=(0,0,0))=0\mu((s_{0},s_{1},t_{1})=(0,0,0))=0.

We next assume that μ\mu-a.s. (s0,s1,t1)=(,,0)(s_{0},s_{1},t_{1})=(\infty,\infty,0). By Lemma 7.3, μ\mu-a.s. (s0,s1,t1,t2)=(,,0,0)(s_{0},s_{1},t_{1},t_{2})=(\infty,\infty,0,0). Since there exists an infinite 0-cluster and an infinite 1-cluster, by Lemma A.3, there exists an infinite contour, and s0+s1>0s_{0}+s_{1}>0. The contradiction implies that μ((s0,s1,t1)=(,,0))=0\mu((s_{0},s_{1},t_{1})=(\infty,\infty,0))=0. This completes the proof of Part I of Theorem 2.2. \hfill\Box


Proof of Theorem 2.2 II. Assume that μ\mu is a probability measure on Ω\Omega satisfying (A2),(A3),(A6),(A7),(A8). By Theorem 2.2 I, μ\mu-a.s. (s0,s1,t1)=(,,)(s_{0},s_{1},t_{1})=(\infty,\infty,\infty). Part II of Theorem 2.2 then follows from Lemma 7.3. \hfill\Box


8. Proof of Theorem 2.4

In this section, we prove Theorem 2.4.

We first prove that Parts (a) and (b) implies Part (c). If μ\mu satisfies (A1)-(A7), then by (a) and (b), μ\mu-a.s. there are neither infinite primal contours nor infinite dual contours. Therefore μ\mu-a.s. there are no infinite contours.

Let 0\mathcal{E}_{0} (resp. 1\mathcal{E}_{1}) be the event that there exists an infinite 0-cluster (resp. infinite 1-cluster). Assume that μ(01)>0\mu(\mathcal{E}_{0}\cup\mathcal{E}_{1})>0. Then by (A2) μ\mu is Γi\Gamma_{i} ergodic,

(30) μ(01)=1.\displaystyle\mu(\mathcal{E}_{0}\cup\mathcal{E}_{1})=1.

By (A3) μ\mu is symmetric with respect to exchanging state “0” and state “1”, μ(0)=μ(1)\mu(\mathcal{E}_{0})=\mu(\mathcal{E}_{1}). By (A2), either μ(0)=μ(1)=1\mu(\mathcal{E}_{0})=\mu(\mathcal{E}_{1})=1 or μ(0)=μ(1)=0\mu(\mathcal{E}_{0})=\mu(\mathcal{E}_{1})=0. By (30), we have μ(0)=μ(1)=1\mu(\mathcal{E}_{0})=\mu(\mathcal{E}_{1})=1. By Lemma A.3, μ\mu-a.s. there exists an infinite contour. But this is a contradiction to the fact that μ\mu-a.s. there are no infinite contours. Therefore μ\mu-a.s. there are no infinite clusters.

Next we prove (a) and (b). Note that the [m,4,n,4][m,4,n,4] lattice GG is amenable if and only if

(31) 1m+1n=12.\displaystyle\frac{1}{m}+\frac{1}{n}=\frac{1}{2}.

When m,nm,n are positive integers greater than or equal to 3, the only pairs of (m,n)(m,n) satisfying (31) are (m,n)=(4,4)(m,n)=(4,4), (m,n)=(3,6)(m,n)=(3,6) and (m,n)=(6,3)(m,n)=(6,3). When (m,n)=(4,4)(m,n)=(4,4), GG is the square grid embedded into 2{\mathbb{R}}^{2}. In this case (a) and (b) were proved in [27]. Then cases (m,n)=(3,6)(m,n)=(3,6) and (m,n)=(6,3)(m,n)=(6,3) can be proved in the same way. We write down the proof of the case when (m,n)=(3,6)(m,n)=(3,6) here.

When (m,n)=(3,6)(m,n)=(3,6), 𝕃1{\mathbb{L}}_{1} is the hexagonal lattice =(V(),E()){\mathbb{H}}=(V({\mathbb{H}}),E({\mathbb{H}})) and 𝕃2{\mathbb{L}}_{2} is the triangular lattice 𝕋=(V(𝕋),E(𝕋)){\mathbb{T}}=(V({\mathbb{T}}),E({\mathbb{T}})).

Lemma 8.1.

Assume that (m,n)=(3,6)(m,n)=(3,6). When μ\mu satisfies (A1), (A2) and (A5), almost surely there exists at most one infinite contour in 𝕋{\mathbb{T}}.

Proof.

Let \mathbb{N} be the set of all nonnegative integers. Let 𝒩\mathcal{N} be the number of infinite contours in 𝕋{\mathbb{T}}. By (A2), there exists k0{}k_{0}\in\mathbb{N}\cup\{\infty\}, s.t. μ(𝒩=k0)=1\mu(\mathcal{N}=k_{0})=1.

By [4] (see also Exercise 7.24 of [38]), (A1) and the fact that the triangular lattice 𝕋{\mathbb{T}} is transitive and amenable, μ\mu-a.s. no infinite contours has more than 2 ends.

The triangular lattice 𝕋{\mathbb{T}} can be obtained from a square grid 𝕊\mathbb{S} by adding a diagonal in each square face of 𝕊\mathbb{S}.

Let BnB_{n} be an n×nn\times n box of 𝕊\mathbb{S}. Let B~n\widetilde{B}_{n} be the corresponding box in 𝕋{\mathbb{T}}, i.e. B~n\widetilde{B}_{n} can be obtained from BnB_{n} by adding a diagonal edge on each square face of BnB_{n}.

Let ϕ\phi (resp. ϕ~\widetilde{\phi}) be a contour configuration on 𝕊\mathbb{S} (resp. 𝕋{\mathbb{T}}), such that ϕ\phi and ϕ~\widetilde{\phi} satisfy the following conditions (note that the vertices in Bn\partial B_{n} and B~n\partial\widetilde{B}_{n} are in 1-1 correspondence)

  • for each vertex vBnv\in\partial B_{n}, no edges incident to vv outside B~n\widetilde{B}_{n} are present in ϕ~\widetilde{\phi} if and only if no edges incident to vv outside BnB_{n} are present in ϕ\phi;

  • for each vertex vBnv\in\partial B_{n}, if there are incident present edges of vv in ϕ~\widetilde{\phi} outside B~n\widetilde{B}_{n}, then the parity of the number of incident present edges of vv outside B~n\widetilde{B}_{n} in ϕ~\widetilde{\phi} is the same as the parity of the number of incident present edges of vv outside BnB_{n} in ϕ\phi; i.e. either both numbers are even or both are odd.

Let n2n\geq 2. Given ϕ\phi, we can find a configuration ξ\xi in BnB_{n}, such that [ϕBn]ξ[\phi\setminus B_{n}]\cup\xi is a contour configuration on 𝕊\mathbb{S} (i.e. each vertex of 𝕊\mathbb{S} has an even number of incident present edges in [ϕBn]ξ[\phi\setminus B_{n}]\cup\xi), and all the incident present edges of Bn\partial B_{n} outside BnB_{n} are in the same contour; see Lemma 4.2 of [27]. If ϕ~\widetilde{\phi} and ϕ\phi satisfy the conditions described above, then [ϕ~B~n]ξ[\widetilde{\phi}\setminus\widetilde{B}_{n}]\cup\xi is a contour configuration on 𝕋{\mathbb{T}}, and all the incident present edges of B~n\partial\widetilde{B}_{n} outside B~n\widetilde{B}_{n} are in the same contour.

Note that ξ\xi can be obtained from ϕ~B~n\widetilde{\phi}\cap\widetilde{B}_{n} by changing configurations on finitely many triangles in B~\widetilde{B} as described in (A6). That is because any contour configuration on 𝕋{\mathbb{T}} naturally induces two site configurations ω\omega, 1ω1-\omega, in {0,1}V()\{0,1\}^{V({\mathbb{H}})}, such that two adjacent vertices in {\mathbb{H}} have different states if and only if the edge in 𝕋{\mathbb{T}} separating the two vertices are present in the contour configuration. Any two site configurations in {0,1}V()\{0,1\}^{V({\mathbb{H}})} differ only in Bn~\widetilde{B_{n}} can be obtained from each other by changing states on finitely many vertices in V()B~nV({\mathbb{H}})\cap\widetilde{B}_{n}. Changing the state at a vertex in V()V({\mathbb{H}}) corresponds to changing the states on all the edges of the dual triangle face including the vertex in the contour configuration of 𝕋{\mathbb{T}}.

We claim that k0{0,1}k_{0}\in\{0,1\}. Indeed, if 2k0<2\leq k_{0}<\infty, we can find a box B~n\widetilde{B}_{n} in 𝕋{\mathbb{T}}, such that B~n\widetilde{B}_{n} intersects all the k0k_{0} infinite contours. Then we can change configurations on finitely many triangles in B~n\widetilde{B}_{n}, such that after the configuration change, there is exactly one infinite contour. By the finite energy assumption (A5), with positive probability, there exists exactly one infinite contour, but this is a contradiction to μ(𝒩=k0)=1\mu(\mathcal{N}=k_{0})=1, where 2k0<2\leq k_{0}<\infty.

If k0=k_{0}=\infty, we can find a box B~m\widetilde{B}_{m} in 𝕋{\mathbb{T}}, such that B~m\widetilde{B}_{m} intersects at least 3 infinite contours. Then we can change configurations on finitely many triangles in B~n\tilde{B}_{n}, such that after the configuration change, all the infinite contours intersecting B~m\tilde{B}_{m} merge into one infinite contour, which has at least 3 ends. By (A5), with positive probability there exists an infinite contour with more than 2 ends. But this is a contradiction to the fact that almost surely no infinite contours have more than two ends. ∎

Lemma 8.2.

Assume that (m,n)=(3,6)(m,n)=(3,6). When μ\mu satisfies (A1) and (A4), almost surely there exists at most one infinite contour in {\mathbb{H}}.

Proof.

Recall that a contour is a connected set of edges in which each vertex has an even number of incident present edges. Since the hexagonal lattice {\mathbb{H}} is a cubic graph, i.e. each vertex has 3 incident edges; each vertex in a contour of {\mathbb{H}} has 2 incident present edges. As a result, each contour in {\mathbb{H}} is either a self-avoiding cycle or a doubly-infinite self-avoiding path. In particular, each infinite contour in {\mathbb{H}} is a doubly-infinite self-avoiding path.

Each contour configuration in {\mathbb{H}} naturally induces two site configurations ω,1ω\omega,1-\omega in {0,1}V(𝕋)\{0,1\}^{V({\mathbb{T}})}, in which two adjacent vertices of 𝕋{\mathbb{T}} have the same state if and only if the dual edge in {\mathbb{H}} separating the two vertices is absent in the contour configuration. The finite energy assumption (A4) implies the finite energy in the induced site configuration in {0,1}V(𝕋)\{0,1\}^{V({\mathbb{T}})}; see [10] for a definition. When μ\mu satisfies (A1) (A4), by the result in [10], almost surely there exists at most one infinite 1-cluster and at most one infinite 0-cluster. In particular, there exist at most two infinite clusters. However, if in {\mathbb{H}} there are more than one infinite contour, then there are at least two doubly-infinite self-avoiding paths in {\mathbb{H}}. As a result, the number of infinite clusters in the induced site configuration on 𝕋{\mathbb{T}} is at least 3. The contradiction implies the lemma. ∎

Lemma 8.3.

Let ωΩ\omega\in\Omega be a constrained percolation configuration on the [3,4,6,4][3,4,6,4] lattice GG. Let ψ=ϕ(ω)Φ\psi=\phi(\omega)\in\Phi be the corresponding contour configuration in E()E(𝕋)E({\mathbb{H}})\cup E({\mathbb{T}}). Assume that ψ=ψ1ψ2\psi=\psi_{1}\cup\psi_{2}, where ψ1\psi_{1} (resp. ψ2\psi_{2}) is the contour configuration in {\mathbb{H}} (resp. 𝕋{\mathbb{T}}). If there is an infinite contour in ψ1\psi_{1} (resp. ψ2\psi_{2}), then there is an infinite cluster in ϕ1(ψ2)\phi^{-1}(\psi_{2}) (resp. ϕ1(ψ1)\phi^{-1}(\psi_{1})).

Proof.

Assume that there is an infinite contour CC in {\mathbb{H}} (resp. 𝕋{\mathbb{T}}). Let VCV_{C} be the set consisting of all the vertices of GG such that

  • vVCv\in V_{C} if and only if vv is a vertex of a face of GG crossed by an edge present in the contour CC.

Let FF be a square face of GG crossed by CC; then all the vertices in FF are in the same cluster of ϕ1(ψ2)\phi^{-1}(\psi_{2}) (resp. ϕ1(ψ1)\phi^{-1}(\psi_{1})). That is because ψ1ψ2=\psi_{1}\cap\psi_{2}=\emptyset, if FF is crossed by Cψ1C\subseteq\psi_{1} (resp. Cψ2C\subseteq\psi_{2}), then Fψ2=F\cap\psi_{2}=\emptyset (resp. Fψ1=F\cap\psi_{1}=\emptyset).

Let FF^{\prime} be a triangle (resp. hexagon) face of GG crossed by CC; then all the vertices in FF are also in the same cluster of ϕ1(ψ2)\phi^{-1}(\psi_{2}) (resp. ϕ1(ψ1)\phi^{-1}(\psi_{1})). That is because the boundary edges of FF^{\prime} cannot be crossed by edges of 𝕋{\mathbb{T}} (resp. {\mathbb{H}}) at all.

We claim that all the vertices in VCV_{C} are in the same cluster in ϕ1(ψ2)\phi^{-1}(\psi_{2}) (resp. ϕ1(ψ1)\phi^{-1}(\psi_{1})). Indeed, for any two vertices u,vVCu,v\in V_{C}, we can find a sequence of faces F0,F1,,FkF_{0},F_{1},\ldots,F_{k}, such that

  • uF0u\in F_{0} and vFkv\in F_{k}; and

  • for 0ik0\leq i\leq k, FiF_{i} is crossed by CC;

  • for 1jk1\leq j\leq k, FjF_{j} and Fj1F_{j-1} share a vertex.

Then uu and vv are in the same cluster in ϕ1(ψ2)\phi^{-1}(\psi_{2}) (resp. ϕ1(ψ1)\phi^{-1}(\psi_{1})) since all the vertices in i=0nFi\cup_{i=0}^{n}F_{i} are in the same cluster in ϕ1(ψ2)\phi^{-1}(\psi_{2}) (resp. ϕ1(ψ1)\phi^{-1}(\psi_{1})). Moreover, |VC|=|V_{C}|=\infty since CC is an infinite contour. Therefore, ϕ1(ψ2)\phi^{-1}(\psi_{2}) (resp. ϕ1(ψ1)\phi^{-1}(\psi_{1})) has an infinite cluster. ∎

Parts (a) and (b) can be proved using similar techniques; we write down the proof of (a) here.

Let μ\mu be a probability measure on Ω\Omega satisfying (A1)-(A6). Assume that with strictly positive probability, there exist infinite contours in 𝕋{\mathbb{T}}. Then by (A2), μ\mu-a.s. there exist infinite contours in 𝕋{\mathbb{T}}. By Lemma 8.1, μ\mu-a.s. there exists exactly one infinite contour C1C_{1} in 𝕋{\mathbb{T}}. By Lemma 8.3, μ\mu-a.s. there exist infinite clusters in ϕ1(ψ1)\phi^{-1}(\psi_{1}). Let 0\mathcal{F}_{0} (resp. 1\mathcal{F}_{1}) be the event that there exists an infinite 0-cluster (resp. infinite 1-cluster) in ϕ1(ψ1)\phi^{-1}(\psi_{1}), then

(32) λ1(01)=1,\displaystyle\lambda_{1}(\mathcal{F}_{0}\cup\mathcal{F}_{1})=1,

where the probability measure λ1\lambda_{1} is defined before (A6). By (A6), and the symmetry of λ1\lambda_{1}, either λ1(0)=λ1(1)=0\lambda_{1}(\mathcal{F}_{0})=\lambda_{1}(\mathcal{F}_{1})=0, or λ1(0)=λ1(1)=1\lambda_{1}(\mathcal{F}_{0})=\lambda_{1}(\mathcal{F}_{1})=1. By (32), we have λ1(01)=1\lambda_{1}(\mathcal{F}_{0}\cap\mathcal{F}_{1})=1. By Lemma A.3, μ\mu-a.s. there are infinite contours in {\mathbb{H}}. By Lemma 8.2, μ\mu-a.s. there is exactly one infinite contour C2C_{2} in {\mathbb{H}}.

Hence there is exactly one infinite contour C2C_{2} in {\mathbb{H}} and exactly one infinite contour C1C_{1} in 𝕋{\mathbb{T}}. By Lemma A.7, there exists an infinite cluster incident to both C1C_{1} and C2C_{2}.

Let 0\mathcal{H}_{0} (resp. 1\mathcal{H}_{1}) be the event that there exists an infinite 0-cluster (resp. infinite 1-cluster) in ω\omega incident to both the infinite contour in {\mathbb{H}} and the infinite contour in 𝕋{\mathbb{T}}. Then

(33) μ(01)=1.\displaystyle\mu(\mathcal{H}_{0}\cup\mathcal{H}_{1})=1.

By (A3), μ(0)=μ(1)\mu(\mathcal{H}_{0})=\mu(\mathcal{H}_{1}). By (A2), either μ(0)=μ(1)=0\mu(\mathcal{H}_{0})=\mu(\mathcal{H}_{1})=0, or μ(0)=μ(1)=1\mu(\mathcal{H}_{0})=\mu(\mathcal{H}_{1})=1. By (33), μ(0)=μ(1)=1\mu(\mathcal{H}_{0})=\mu(\mathcal{H}_{1})=1, therefore μ(01)=1\mu(\mathcal{H}_{0}\cap\mathcal{H}_{1})=1, i.e. there exist an infinite 0-cluster ξ0\xi_{0} and an infinite 1-cluster ξ1\xi_{1}, such that ξ0\xi_{0} is incident to both C1C_{1} and C2C_{2}, and ξ1\xi_{1} is incident to both C1C_{1} and C2C_{2}. But this is a contradiction to Lemma A.10. Therefore we conclude that μ\mu-a.s. there are no infinite contours in 𝕋{\mathbb{T}}; this completes the proof of Part (a).

9. Proof of Theorem 2.5

In this section, we prove Theorem 2.5.

Lemma 9.1.

Let GG be an [m,4,m,4][m,4,m,4] lattice with m5m\geq 5. Let μ\mu be a probability measure on Ω\Omega satisfying (A1)-(A3). Then the distribution of infinite clusters can only be one of the following 2 cases.

  1. I

    There are no infinite clusters μ\mu-a.s.

  2. II

    There are infinitely many infinite 1-clusters and infinitely many infinite 0-clusters μ\mu-a.s.

Lemma 9.1 can be obtained from Lemma 7.4; it may also be proved using different arguments below.

Proof.

Let s0s_{0} (resp. s1s_{1}) be the number of infinite 0-clusters (resp. 1-clusters). By Lemma 6.1 and (A2) (A3), there exist k{0,1,}k\in\{0,1,\infty\}, such that μ((s0,s1)=(k,k))=1\mu((s_{0},s_{1})=(k,k))=1. It suffices to show that k1k\neq 1.

Let 𝒜{\mathcal{A}} be the event that (s0,s1)=(1,1)(s_{0},s_{1})=(1,1). Assume that μ(𝒜)=1\mu({\mathcal{A}})=1, we will obtain a contradiction.

As explained before the constrained site configurations on GG correspond to contour configurations in 𝕃1𝕃2{\mathbb{L}}_{1}\cup{\mathbb{L}}_{2}.

Since μ\mu-a.s. there exists exactly one infinite 0-cluster and exactly one infinite 1-cluster simultaneously, then by Lemma A.8, μ\mu-a.s. there exists an infinite primal or dual contour incident to both the infinite 0-cluster and the infinite 1-cluster. Let 𝒟1{\mathcal{D}}_{1} (resp. 𝒟2{\mathcal{D}}_{2}) be the event that there exists an infinite primal (resp. dual) contour in 𝕃1{\mathbb{L}}_{1} (resp. 𝕃2{\mathbb{L}}_{2}), incident to both the infinite 0-cluster and the infinite 1-cluster. So we have

(34) μ(𝒟1𝒟2)=1.\displaystyle\mu({\mathcal{D}}_{1}\cup{\mathcal{D}}_{2})=1.

By (A1) μ\mu is Γi\Gamma_{i} invariant, we have

(35) μ(𝒟1)=μ(𝒟2).\displaystyle\mu({\mathcal{D}}_{1})=\mu({\mathcal{D}}_{2}).

Moreover, by (A2), we have either

(36) μ(𝒟1)=0orμ(𝒟1)=1.\displaystyle\mu({\mathcal{D}}_{1})=0\ \mathrm{or}\ \mu({\mathcal{D}}_{1})=1.

Combining (34), (35) and (36), we have

(37) μ(𝒟1𝒟2)=1.\displaystyle\mu({\mathcal{D}}_{1}\cap{\mathcal{D}}_{2})=1.

Thus, by (37), we have exactly one infinite 1-cluster on GG, denoted by ξ1\xi_{1} and exactly one infinite 0-cluster on GG, denoted by ξ0\xi_{0}. There is an infinite primal contour incident to both ξ0\xi_{0} and ξ1\xi_{1}, denoted by C1C_{1}; as well as an infinite dual contour incident to both ξ0\xi_{0} and ξ1\xi_{1}, denoted by C2C_{2}. But this is impossible by Lemma A.10. The contradiction implies the lemma. ∎

Lemma 9.2.

Let μ\mu be a probability measure on Ω\Omega satisfying (A1)-(A3). Then the distribution of infinite contours can only be one of the following 2 cases.

  1. I

    There are neither infinite primal contours nor infinite dual contours μ\mu-a.s..

  2. II

    There are infinitely many infinite primal contours and infinitely many infinite dual contours μ\mu-a.s..

Proof.

The primal (resp. dual) contours form an invariant bond percolation on 𝕃1{\mathbb{L}}_{1} (resp. 𝕃2{\mathbb{L}}_{2}) under μ\mu. Let t1t_{1} (resp. t2t_{2}) be the number of infinite primal (resp. dual) contours. By Lemma 6.1 and (A1)-(A3), only 3 cases may occur:

  1. i.

    μ\mu-a.s. (t1,t2)=(0,0)(t_{1},t_{2})=(0,0);

  2. ii.

    μ\mu-a.s. (t1,t2)=(,)(t_{1},t_{2})=(\infty,\infty);

  3. iii.

    μ\mu-a.s. (t1,t2)=(1,1)(t_{1},t_{2})=(1,1).

It remains to exclude in Case iii.. Assume that Case iii. occurs. Let C1C_{1} (resp. C2C_{2}) be the unique infinite primal (resp. dual) contour. By Lemma A.7, there exists an infinite cluster incident to both C1C_{1} and C2C_{2}. Moreover, by (A2)-(A3), μ\mu-a.s. there exists an infinite 0-cluster incident to both C1C_{1} and C2C_{2}, as well as an infinite 1-cluster incident to both C1C_{1} and C2C_{2}. But this is impossible by Lemma A.10. ∎


Proof of Theorem 2.5. By Lemmas 9.1, 9.2 and 7.2, it suffices to show that there exists an infinite cluster μ\mu-a.s. if and only if there exists an infinite contour μ\mu-a.s. if μ\mu satisfies (A1)-(A3).

First assume that there exists an infinite cluster μ\mu-a.s. By (A2)-(A3), μ\mu-a.s. there exist both an infinite 0-cluster and an infinite 1-cluster. By Lemma A.3, μ\mu-a.s. there exists an infinite contour.

Now assume that there exists an infinite contour μ\mu-a.s. By (A1)-(A2), μ\mu-a.s. there exist both an infinite primal contour and an infinite dual contour. By Lemma A.2, μ\mu-a.s. there exists an infinite cluster. \hfill\Box

10. Percolation on transitive, triangular tilings of hyperbolic plane

In this section, we discuss the applications of the techniques developed in the proof of Theorem 2.2 to prove results concerning unconstrained site percolation on vertex-transitive, triangular tilings of the hyperbolic plane. We also discuss results about Bernoulli percolation on such graphs. These results will be used to prove Theorems 3.3, 4.1 and 4.2.

Lemma 10.1.

Let 𝕃2{\mathbb{L}}_{2} a vertex-transitive, regular tiling of the hyperbolic plane with triangles, such that each vertex has degree n7n\geq 7. Consider an Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariant site percolation ω\omega on 𝕃2{\mathbb{L}}_{2} with distribution μ\mu. Assume that μ\mu is Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-ergodic. Let s0s_{0} (resp. s1s_{1}) be the number of infinite 0-clusters (resp. infinite 1-clusters) in the percolation. Then

μ((s0,s1)=(0,0))=0.\displaystyle\mu((s_{0},s_{1})=(0,0))=0.
Proof.

Since the event {(s0,s1)=(0,0)}\{(s_{0},s_{1})=(0,0)\} is Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariant, and μ\mu is Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-ergodic, either μ((s0,s1)=(0,0))=0\mu((s_{0},s_{1})=(0,0))=0 or μ((s0,s1)=(0,0))=1\mu((s_{0},s_{1})=(0,0))=1. Assume that μ((s0,s1)=(0,0))=1\mu((s_{0},s_{1})=(0,0))=1; we shall obtain a contradiction.

Note that the dual graph 𝕃1{\mathbb{L}}_{1} of 𝕃2{\mathbb{L}}_{2} is a vertex-transitive, non-amenable, planar graph in which each vertex has degree 3. A contour configuration ϕ(ω)E(𝕃1)\phi(\omega)\subset E({\mathbb{L}}_{1}) is a subset of edges of 𝕃1{\mathbb{L}}_{1} in which each present edge has a dual edge in E(𝕃2)E({\mathbb{L}}_{2}) joining exactly one vertex with state 0 and one vertex with state 1 in ω\omega. As usual, a contour is a maximal connected component of present edges in a contour configuration. Each contour configuration, by definition, must be an even-degree subgraph of 𝕃1{\mathbb{L}}_{1}. Given that 𝕃1{\mathbb{L}}_{1} has vertex-degree 3, each vertex in 𝕃1{\mathbb{L}}_{1} is incident to zero or two present edges in a contour configuration. Let tt be the number of infinite contours. Each infinite contour on 𝕃1{\mathbb{L}}_{1} must be a doubly infinite self-avoiding path.

If t1t\geq 1, let CC_{\infty} be an infinite contour. Then 2C{\mathbb{H}}^{2}\setminus C_{\infty} has exactly two unbounded components, since CC_{\infty} is a doubly-infinite self-avoiding path. Let VV_{\infty} be the set of all vertices of 𝕃2{\mathbb{L}}_{2} lying on a face crossed by CC_{\infty}. Given CC_{\infty} a fixed orientation. Let V+V_{\infty}^{+} (resp. VV_{\infty}^{-}) be the subset of VV_{\infty} consisting of all the vertices in VV_{\infty} on the left hand side (resp. right hand side) of CC_{\infty} when traversing CC_{\infty} along the given orientation. Then exactly one of V+V_{\infty}^{+} and VV_{\infty}^{-} is part of an infinite 1-cluster, and the other is part of an infinite 0-cluster. Therefore we have s01s_{0}\geq 1 and s11s_{1}\geq 1 if t1t\geq 1.

The rest of the proof is an adaptation of the proof of Proposition 7.2 to different graphs. Define a generalized contour in a contour configuration ϕE(𝕃1)\phi\subset E({\mathbb{L}}_{1}) to be either a single vertex in V(𝕃1)V({\mathbb{L}}_{1}) which has no incident present edges in ϕ\phi, or a contour in ϕ\phi. This way for each vertex vV(𝕃1)v\in V({\mathbb{L}}_{1}), and each contour configuration ϕE(𝕃1)\phi\subset E({\mathbb{L}}_{1}), there is a unique generalized contour passing through vv. By the arguments in the last paragraph, if (s0,s1)=(0,0)(s_{0},s_{1})=(0,0), then t=0t=0.

Now consider the case when (s0,s1,t)=(0,0,0)(s_{0},s_{1},t)=(0,0,0). Given a cluster CC in ω\omega, there is a unique contour CC^{\prime} of ϕ(ω)\phi(\omega) surrounding ξ\xi. Similarly, for every generalized contour CC^{\prime}, there is a cluster C′′C^{\prime\prime} that surrounds CC^{\prime}. Let 𝒞0\mathcal{C}_{0} denote the set of all generalized contours of ω\omega. We set

𝒞j+1:={C′′:K𝒞j}.\displaystyle\mathcal{C}_{j+1}:=\{C^{\prime\prime}:K\in\mathcal{C}_{j}\}.

For C𝒞0C\in\mathcal{C}_{0} and vV(𝕃1)v\in V({\mathbb{L}}_{1}), let r(C):=sup{j:C𝒞j}r(C):=\sup\{j:C\in\mathcal{C}_{j}\}, and define r(v):=r(C)r(v):=r(C) if CC is the generalized contour of vv in ϕ(ω)\phi(\omega). For each rr let ωr\omega^{r} be the set of edges in E(𝕃1)E({\mathbb{L}}_{1}) whose both end-vertices u,vV(𝕃1)u,v\in V({\mathbb{L}}_{1}) satisfy r(v)rr(v)\leq r and r(u)rr(u)\leq r. Then ωr\omega^{r} is an invariant bond percolation and

limr𝐄[degωrv]=3.\displaystyle\lim_{r\rightarrow\infty}\mathbf{E}[\deg_{\omega^{r}}v]=3.

By Lemma 6.2, we deduce that ωr\omega^{r} has infinite components with positive probability for all sufficiently large rr.

However, since (s0,s1,t)=(0,0,0)(s_{0},s_{1},t)=(0,0,0), by the arguments above each vertex in vV(𝕃1)v\in V({\mathbb{L}}_{1}) is surrounded by infinitely many contours. This implies that for any rr\in\mathbb{N}, for any vertex vV(𝕃1)v\in V({\mathbb{L}}_{1}), there exists a contour CC surrounding vv, such that r(C)>rr(C)>r, and therefore Cωr=C\cap\omega^{r}=\emptyset. As a result, the components in ωr\omega^{r} including vv is finite. The contradiction implies the lemma. ∎

Lemma 10.2.

Let 𝕃2{\mathbb{L}}_{2} be the regular tiling of the hyperbolic plane with triangles, such that each vertex has degree n7n\geq 7. Consider an Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariant site percolation ω\omega on 𝕃2{\mathbb{L}}_{2} with distribution μ\mu. Let tt be the number of infinite contours. Then μ\mu-a.s. t{0,}t\in\{0,\infty\}

Proof.

By Lemma 6.1, μ\mu-a.s. t{0,1,}t\in\{0,1,\infty\}. Let 𝒜{\mathcal{A}} be the event that t=1t=1. Assume μ(𝒜)>0\mu({\mathcal{A}})>0, we shall obtain a contradiction. Conditional on the event 𝒜{\mathcal{A}}, let τ\tau be the unique infinite contour. Since τ\tau is a infinite, connected, even-degree subgraph of 𝕃1{\mathbb{L}}_{1}, and each vertex of 𝕃1{\mathbb{L}}_{1} has degree 3, τ\tau must be a doubly infinite self-avoiding path. Then pc(τ)=1p_{c}(\tau)=1. But this is a contradiction to Lemma 6.6. Then the lemma follows. ∎

Lemma 10.3.

Let 𝕃2{\mathbb{L}}_{2} be the regular tiling of the hyperbolic plane with triangles, such that each vertex has degree n7n\geq 7. Consider an Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariant site percolation ω\omega on 𝕃2{\mathbb{L}}_{2} with distribution μ\mu. Assume that μ\mu is Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-ergodic. Let s0s_{0} (resp. s1s_{1}) be the number of infinite 0-clusters (resp. infinite 1-clusters) in the percolation. Then

μ((s0,s1)=(1,1))=0.\displaystyle\mu((s_{0},s_{1})=(1,1))=0.
Proof.

We may construct a [3,4,n,4][3,4,n,4] lattice embedded into the hyperbolic plane 2{\mathbb{H}}^{2}, such that the [3,4,n,4][3,4,n,4] lattice, 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} satisfy the conditions as described in Section 2. Then each percolation configuration ω\omega in {0,1}V(𝕃2)\{0,1\}^{V({\mathbb{L}}_{2})} induces a constrained configuration ω~Ω\widetilde{\omega}\in\Omega by the condition that ω\omega and ω~\widetilde{\omega} has the same contour configuration; and that ω(v)=1\omega(v)=1 for vV(𝕃2)v\in V({\mathbb{L}}_{2}) if and only if all the vertices of the [3,4,n,4][3,4,n,4] lattice in the degree-nn face of 𝕃1{\mathbb{L}}_{1} containing vv have state 1 in ω~\widetilde{\omega}. Then the lemma follows from Lemma 7.4. ∎

Lemma 10.4.

Let 𝕃2{\mathbb{L}}_{2} be the regular tiling of the hyperbolic plane with triangles, such that each vertex has degree n7n\geq 7. Consider a site percolation ω\omega on 𝕃2{\mathbb{L}}_{2} with distribution μ\mu. Let s0s_{0} (resp. s1s_{1}) be the number of infinite 0-clusters (resp. infinite 1-clusters) in the percolation. Then it is not possible that

(s0,s1)=(0,k);for 2k.\displaystyle(s_{0},s_{1})=(0,k);\ \mathrm{for}\ 2\leq k\leq\infty.
Proof.

Assume that s0=0s_{0}=0, and that there exist at least two distinct infinite 1-clusters C1C_{1} and C2C_{2}. Let \ell be a path consisting of edges of 𝕃2{\mathbb{L}}_{2} joining a vertex xC1x\in C_{1} and a vertex yC2y\in C_{2}. If \ell does not cross infinite contours, then we can find another path \ell^{\prime} joining xx and yy such that \ell^{\prime} does not cross contours at all. Then C1=C2C_{1}=C_{2}. The contradiction implies that there exists at least one infinite contour in 𝕃1{\mathbb{L}}_{1}. Since each infinite contour in 𝕃1{\mathbb{L}}_{1} is a doubly infinite self-avoiding path, if there exists an infinite contour, then there exist at least one infinite 0-cluster and at least one infinite 1-cluster. But this is a contradiction to the fact that s0=0s_{0}=0. ∎

Definition 10.5.

Let G=(V,E)G=(V,E) be a graph. Given a set A2VA\in 2^{V}, and a vertex vVv\in V, denote ΠvA=A{v}\Pi_{v}A=A\cup\{v\}. For 𝒜2V{\mathcal{A}}\subset 2^{V}, we write Πv𝒜={ΠvA:A𝒜}\Pi_{v}{\mathcal{A}}=\{\Pi_{v}A:A\in\mathcal{A}\}. A site percolation process (𝐏,ω)(\mathbf{P},\omega) on GG is insertion-tolerant if 𝐏(Πv𝒜)>0\mathbf{P}(\Pi_{v}{\mathcal{A}})>0 for every vVv\in V and every event 𝒜2V{\mathcal{A}}\subset 2^{V} satisfying 𝐏(𝒜)>0\mathbf{P}({\mathcal{A}})>0.

We can similarly define an insertion tolerant bond percolation by replacing a vertex with an edge in the above definition. A bond percolation is deletion tolerant if 𝐏[Π¬e𝒜]>0\mathbf{P}[\Pi_{\neg e}{\mathcal{A}}]>0 whenever eE(G)e\in E(G) and 𝐏(𝒜)>0\mathbf{P}({\mathcal{A}})>0, where Π¬eω=ω{e}\Pi_{\neg e}\omega=\omega\setminus\{e\}.

Lemma 10.6.

Let GG be a graph with a transitive, unimodular, closed automorphism group ΓAut(G)\Gamma\subset\mathrm{Aut}(G). If (𝐏,ω)(\mathbf{P},\omega) is a Γ\Gamma-invariant, insertion-tolerant percolation process on GG with infinitely many infinite clusters a.s., then a.s. every infinite cluster has infinitely many ends.

Proof.

See Proposition 3.10 of [39]; see also [21] and [22]. ∎

Lemma 10.7.

Let 𝕃2{\mathbb{L}}_{2} be the regular tiling of the hyperbolic plane with triangles, such that each vertex has degree n7n\geq 7. Consider an Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariant, insertion-tolerant site percolation ω\omega on 𝕃2{\mathbb{L}}_{2} with distribution μ\mu. Assume that μ\mu is Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-ergodic. Let s0s_{0} (resp. s1s_{1}) be the number of infinite 0-clusters (resp. infinite 1-clusters) in the percolation. Then

μ((s0,s1)=(1,))=0.\displaystyle\mu((s_{0},s_{1})=(1,\infty))=0.
Proof.

Assume that μ(s0,s1)=(1,)=1\mu(s_{0},s_{1})=(1,\infty)=1; we shall obtain a contradiction. By Lemma 10.6, a.s. every infinite 1-cluster has infinitely many ends. Then the lemma follows from Lemma A.11. ∎

Proposition 10.8.

Let 𝕃2{\mathbb{L}}_{2} be the regular tiling of the hyperbolic plane with triangles, such that each vertex has degree n7n\geq 7. Consider an Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariant, Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-ergodic, insertion-tolerant site percolation ω\omega on 𝕃2{\mathbb{L}}_{2} with distribution μ\mu. Let s0,s1,ts_{0},s_{1},t be given as in Lemmas 10.1 and 10.2, then

(s0,s1,t){(0,1,0),(1,0,0),(,,)}a.s.\displaystyle(s_{0},s_{1},t)\in\{(0,1,0),(1,0,0),(\infty,\infty,\infty)\}\ a.s.
Proof.

By Lemma 6.1, we have μ\mu-a.s. s0,s1,t{0,1,}s_{0},s_{1},t\in\{0,1,\infty\}. By Lemmas 10.1, 10.3, 10.4 and 10.7, we have μ\mu-a.s. (s0,s1){(0,1),(1,0),(,)}(s_{0},s_{1})\in\{(0,1),(1,0),(\infty,\infty)\}. By Lemma 10.2, μ\mu-a.s. t{0,}t\in\{0,\infty\}. Moreover, since each infinite contour must be a doubly infinite self-avoiding path, if t=t=\infty, then s0+s1=s_{0}+s_{1}=\infty. This implies that if s0+s1=1s_{0}+s_{1}=1, then t=0t=0, a.s. Moreover, if s0+s12s_{0}+s_{1}\geq 2, then t1t\geq 1. Then the proposition follows. ∎

Example 10.9.

Consider Example 2.3. We have

1n1pc<12<pu=1pcnn1\displaystyle\frac{1}{n-1}\leq p_{c}<\frac{1}{2}<p_{u}=1-p_{c}\leq\frac{n}{n-1}

by Theorems 1.1, 1.2. and 1.3 of [6]. By Proposition 10.8, we have

  • if p[0,pc]p\in[0,p_{c}], a.s. (s0,s1,t)=(1,0,0)(s_{0},s_{1},t)=(1,0,0);

  • if p(pc,pu)p\in(p_{c},p_{u}), a.s. (s0,s1,t)=(,,)(s_{0},s_{1},t)=(\infty,\infty,\infty);

  • if p[pu,1]p\in[p_{u},1], a.s. (s0,s1,t)=(0,1,0)(s_{0},s_{1},t)=(0,1,0).

11. Proof of Theorem 3.3 and Corollary 3.4

In this section, we prove Theorem 3.3 and Corollary 3.4. The proof of Theorem 3.3 I. is based on a stochastic domination between the i.i.d. Bernoulli site percolation and the Ising model on the same graph, which correspond to random cluster models with q=1q=1 and q=2q=2, respectively. The proof of of Theorem 3.3 II(a) is based on the ergodicity and symmetry of the measure μf\mu^{f}, as well as Proposition 10.8, which lists all the possible numbers of infinite 0-clusters and infinite-1 clusters which have strictly positive probability to occur. Theorem 3.3 II (b)(c)(d) then follow from the fact that any one of the conditions (b)(c) and (d) implies (a). We then prove Corollary 3.4 by an inequality between pc,2wp_{c,2}^{w} and pu,2fp_{u,2}^{f}, and Theorem 3.3 II(c). Theorem 3.3 III is proved by the well-known combinatorial correspondence between the Ising model and its random-cluster representation. Theorem 3.3 IV then follows from the decomposition of μf\mu^{f} as a convex combination of μ+\mu^{+} and μ\mu^{-} and an inequality between pu,2fp_{u,2}^{f} and pu,2wp_{u,2}^{w}.

We shall first review the stochastic domination we use to prove these results.

Definition 11.1.

(Stochastic Domination) Let G=(V,E)G=(V,E) be a graph. Let Ω={0,1}E\Omega=\{0,1\}^{E} (resp. Ω={0,1}V\Omega=\{0,1\}^{V}). Then the configuration space Ω\Omega is a partially ordered set with partial order given by ω1ω2\omega_{1}\leq\omega_{2} if ω1(e)ω2(e)\omega_{1}(e)\leq\omega_{2}(e) for all eEe\in E (resp. ω1(v)ω2(v)\omega_{1}(v)\leq\omega_{2}(v) for all vVv\in V). A random variable X:ΩX:\Omega\rightarrow{\mathbb{R}} is called increasing if X(ω1)X(ω2)X(\omega_{1})\leq X(\omega_{2}) whenever ω1ω2\omega_{1}\leq\omega_{2}. An event AΩA\subset\Omega is called increasing (respectively, decreasing) if its indicator function 1A1_{A} is increasing (respectively, decreasing). Given two probability measures μ1\mu_{1}, μ2\mu_{2} on Ω\Omega, we write μ1μ2\mu_{1}\prec\mu_{2}, and we say that μ2\mu_{2} stochastically dominates μ1\mu_{1}, if μ1(A)μ2(A)\mu_{1}(A)\leq\mu_{2}(A) for all increasing events AΩA\subset\Omega.

Lemma 11.2.

(Holley inequality) Let G=(V,E)G=(V,E) be a finite graph. Let Ω={0,1}E\Omega=\{0,1\}^{E} (resp. Ω={0,1}V\Omega=\{0,1\}^{V}). Let μ1\mu_{1} and μ2\mu_{2} be strictly positive probability measures on Ω\Omega such that

(38) μ2(max{ω1,ω2})μ1(min(ω1,ω2))μ1(ω1)μ2(ω2),ω1,ω2Ω,\displaystyle\mu_{2}(\max\{\omega_{1},\omega_{2}\})\mu_{1}(\min(\omega_{1},\omega_{2}))\geq\mu_{1}(\omega_{1})\mu_{2}(\omega_{2}),\qquad\omega_{1},\omega_{2}\in\Omega,

Then

μ1μ2.\displaystyle\mu_{1}\prec\mu_{2}.
Proof.

See Theorem (2.1) of [17]; see also [26]. ∎

Lemma 11.3.

Let G=(V,E)G=(V,E) be a finite graph. Let Ω={0,1}E\Omega=\{0,1\}^{E}. Let μ1\mu_{1} and μ2\mu_{2} be strictly positive probability measures on Ω\Omega such that

(39) μ2(ω{e})μ1(ω{e})μ1(ω{e})μ2(ω{e}),ωΩ,eE.\displaystyle\mu_{2}(\omega\cup\{e\})\mu_{1}(\omega\setminus\{e\})\geq\mu_{1}(\omega\cup\{e\})\mu_{2}(\omega\setminus\{e\}),\qquad\omega\in\Omega,e\in E.

Here ω\omega is interpreted as the subset of EE consisting of all the edges with state 1. If either μ1\mu_{1} or μ2\mu_{2} satisfies

(40) μ(ω{e,f})μ(ω{e,f})μ(ωe{f})μ(ω{f}{e}),\displaystyle\mu(\omega\cup\{e,f\})\mu(\omega\setminus\{e,f\})\geq\mu(\omega\cup{e}\setminus\{f\})\mu(\omega\cup\{f\}\setminus\{e\}),

then (38) holds.

Proof.

See Theorem 2.6 of [17]. ∎

For a planar graph GG, let GG_{*} be its planar dual graph. The following lemmas concerning planar duality, are proved in [6, 20].

Lemma 11.4.

Let GG be a planar nonamenable quasi-transitive graph, and let p,p(0,1)p,p_{*}\in(0,1) satisfy

(41) p=(1p)qp+(1p)q\displaystyle p_{*}=\frac{(1-p)q}{p+(1-p)q}

In the natural coupling of FRCp,qGFRC_{p,q}^{G} and WRCp,qGWRC_{p_{*},q}^{G_{*}} as dual measures (i.e. a dual edge is present if and only if the corresponding primal edge is absent), the number of infinite clusters with respect to each is a.s. one of the following: (0,1)(0,1), (1,0)(1,0), (,)(\infty,\infty).

Proof.

See Proposition 3.5 of [20], which is proved using the same technique as the proof of Theorem 3.7 of [6]. ∎

Lemma 11.5.

Let

h(x):=x1x.\displaystyle h(x):=\frac{x}{1-x}.

For any planar non-amenable quasi-transitive graph GG,

h(pc,qw(G))h(pu,qf(G))=h(pu,qw(G))h(pc,qf(G))=1,\displaystyle h(p_{c,q}^{w}(G))h(p_{u,q}^{f}(G_{*}))=h(p_{u,q}^{w}(G_{*}))h(p_{c,q}^{f}(G))=1,
0<pc,qw(G)pc,qf(G)<1,and 0<pu,qw(G)pu,qf(G)<1\displaystyle 0<p_{c,q}^{w}(G)\leq p_{c,q}^{f}(G)<1,\ \mathrm{and}\ 0<p_{u,q}^{w}(G)\leq p_{u,q}^{f}(G)<1
Proof.

See Corollary 3.6 of [20]. ∎

We start the proof of Theorem 3.3 with the following lemma.

Lemma 11.6.

Let 𝕃2{\mathbb{L}}_{2} be a regular tiling of the hyperbolic plane such that each vertex has degree nn and each face has degree mm. Let q1q\geq 1. Assume that WRCp,q𝕃2WRC^{{\mathbb{L}}_{2}}_{p,q}-a.s. there is a unique infinite open cluster for the random cluster model on 𝕃2{\mathbb{L}}_{2}. Let τ\tau be the unique infinite open cluster in the random cluster configuration ω\omega. We define a site percolation configuration ξ\xi on V(𝕃2)V({\mathbb{L}}_{2}), by letting all the vertices in τ\tau have state 1, and all the other vertices have state 0. Then a.s. ξ\xi has no infinite 0-cluster.

Proof.

Let 𝒜0{\mathcal{A}}_{0} be the event that ξ\xi has an infinite 0-cluster. By ergodicity of WRCp,q𝕃2WRC_{p,q}^{{\mathbb{L}}_{2}}, either WRCp,q𝕃2(𝒜0)=0WRC_{p,q}^{{\mathbb{L}}_{2}}({\mathcal{A}}_{0})=0 or WRCp,q𝕃2(𝒜0)=1WRC_{p,q}^{{\mathbb{L}}_{2}}({\mathcal{A}}_{0})=1. Assume that WRCp,q𝕃2(𝒜0)=1WRC_{p,q}^{{\mathbb{L}}_{2}}({\mathcal{A}}_{0})=1; we shall obtain a contradiction.

The dual configuration to the random cluster configuration on 𝕃2{\mathbb{L}}_{2} is a bond configuration on 𝕃1{\mathbb{L}}_{1} such that an edge in 𝕃1{\mathbb{L}}_{1} is present if and only if its dual edge in 𝕃2{\mathbb{L}}_{2} is absent in the random cluster configuration of 𝕃2{\mathbb{L}}_{2}. Note also that the free boundary condition is dual to the wired boundary condition by the relation between dual configurations described above. Moreover, if the random cluster configuration on 𝕃2{\mathbb{L}}_{2} has distribution WRCp,q𝕃2WRC_{p,q}^{{\mathbb{L}}_{2}}, then its dual configuration on 𝕃1{\mathbb{L}}_{1} has distribution FRCp,q𝕃1FRC_{p_{*},q}^{{\mathbb{L}}_{1}}; where p,pp,p_{*} satisfy (41).

Let kk be the number of infinite clusters in the bond configuration in 𝕃2{\mathbb{L}}_{2}, and let kk^{*} be the number of infinite clusters in the corresponding dual configuration in 𝕃1{\mathbb{L}}_{1}. By Lemma 11.4, a.s. (k,k){(0,1),(1,0),(,)}(k,k^{*})\in\{(0,1),(1,0),(\infty,\infty)\}. (Indeed this is true for any Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariant, insertion tolerant and deletion tolerant bond configuration.) Hence if WRCp,q𝕃2WRC_{p,q}^{{\mathbb{L}}_{2}}-a.s. there is a unique infinite open cluster, then FRCp,q𝕃1FRC_{p_{*},q}^{{\mathbb{L}}_{1}}-a.s. there is no infinite cluster in the corresponding dual configuration.

Since τ\tau is an infinite cluster, there exists an infinite 1-cluster in ξ\xi by construction. If there exists an infinite 0-cluster in ξ\xi as well, by Lemma A.3, there exists an infinite contour CC consisting of edges of 𝕃1{\mathbb{L}}_{1} in which each edge has a dual edge joining a vertex of V(𝕃2)V({\mathbb{L}}_{2}) with state 1 in ξ\xi and a vertex of V(𝕃2)V({\mathbb{L}}_{2}) with state 0 in ξ\xi. Moreover, all the edges in CC must be present in the dual configuration of ω\omega, since every edge in CC is dual to an edge of V(𝕃2)V({\mathbb{L}}_{2}) not open in ω\omega. Then we have k=1k=1 and k1k^{*}\geq 1. But this is a contradiction to Lemma 11.4. Hence a.s. there are no infinite 0-clusters in ξ\xi. ∎

Lemma 11.7.

Let 𝕃2{\mathbb{L}}_{2} be a regular tiling of the hyperbolic plane such that each vertex has degree nn and each face has degree mm. Then for the graph 𝕃2{\mathbb{L}}_{2},

pu,1pu,2wpu,2f\displaystyle p_{u,1}\leq p_{u,2}^{w}\leq p_{u,2}^{f}
Proof.

The fact that pu,2wpu,2fp_{u,2}^{w}\leq p_{u,2}^{f} follows from Lemma 11.5.

Now we prove that pu,1pu,2wp_{u,1}\leq p_{u,2}^{w}. Note that the following stochastic monotonicity result holds:

(42) WRCp,2WRCp,1=FRCp,1,\displaystyle WRC_{p,2}\prec WRC_{p,1}=FRC_{p,1},

by (4.1) of [45].

Let 1\mathcal{F}_{1} be the event that there exists a unique infinite cluster in the random cluster configuration in 𝕃2{\mathbb{L}}_{2}. By ergodicity of WRCp,2WRC_{p,2} and WRCp,1WRC_{p,1} and the monotonicity of WRCp,2(1)WRC_{p,2}(\mathcal{F}_{1}) and WRCp,1(1)WRC_{p,1}(\mathcal{F}_{1}) with respect to pp, to show that pu,1pu,2wp_{u,1}\leq p_{u,2}^{w}, it suffices to show that whenever WRCp,2(1)=1WRC_{p,2}(\mathcal{F}_{1})=1, then WRCp,1(1)=1WRC_{p,1}(\mathcal{F}_{1})=1.

Let 1,01\mathcal{F}_{1,0}\subset\mathcal{F}_{1} be the event consisting of all the configurations in which both of the following two cases occur

  • there exists a unique infinite cluster τ\tau in the random cluster configuration on 𝕃2{\mathbb{L}}_{2}; and

  • let ξ{0,1}V(𝕃2)\xi\in\{0,1\}^{V({\mathbb{L}}_{2})} be the site configuration obtained by assigning the state 1 to all the vertices in τ\tau, and the state 0 to all the vertices not in τ\tau; then there exists no infinite 0-cluster in ξ\xi.

By Lemma 11.6, if WRCp,2(1)=1WRC_{p,2}(\mathcal{F}_{1})=1, then WRCp,2(1,0)=1WRC_{p,2}(\mathcal{F}_{1,0})=1. Since 1,0\mathcal{F}_{1,0} is an increasing event, by (42) we have WRCp,2(1,0)=1WRC_{p,2}(\mathcal{F}_{1,0})=1, then WRCp,1(1,0)=1WRC_{p,1}(\mathcal{F}_{1,0})=1. Since 1,01\mathcal{F}_{1,0}\subset\mathcal{F}_{1}, we have WRCp,1(1)=1WRC_{p,1}(\mathcal{F}_{1})=1. This completes the proof. ∎

Lemma 11.8.

Let 𝕃2{\mathbb{L}}_{2} be a vertex-transitive, triangular tiling of the hyperbolic plane such that each vertex has degree n7n\geq 7. Consider the following Conditions (a),(b),(c) and (d) in Part II of Theorem 3.3. We have

(d)(c)(b)(a);\displaystyle(d)\Rightarrow(c)\Rightarrow(b)\Rightarrow(a);

where ABA\Rightarrow B means that if AA holds, then BB holds.

Proof.

The statement (b)(a)(b)\Rightarrow(a) follows from Theorem 4.1 of [45].

The fact that (c)(a)(c)\Rightarrow(a) follows from Theorem 3.2 (v) of [20]; while the fact that (c)(b)(c)\Rightarrow(b) follows from Theorem 4.1 and Lemma 6.4 of [39].

The fact that (d)(c)(d)\Rightarrow(c) follows from Lemma 11.7. ∎

11.1. Proof of Part I of Theorem 3.3.

First note that if (16) hold, then

(43) eheh+eh=pu.\displaystyle\frac{e^{h}}{e^{h}+e^{-h}}=p_{u}.

by the fact that pc+pu=1p_{c}+p_{u}=1.

Let ν1\nu_{1} (resp. ν2\nu_{2}) be the probability measure for the i.i.d. Bernoulli site percolation on 𝕃2{\mathbb{L}}_{2} in which each vertex takes the value “++” with probability p1p_{1} (resp. p2p_{2}) satisfying

enJenJ+enJ<p1<pu\displaystyle\frac{e^{nJ}}{e^{nJ}+e^{-nJ}}<p_{1}<p_{u}
pc<p2<enJenJ+enJ\displaystyle p_{c}<p_{2}<\frac{e^{-nJ}}{e^{nJ}+e^{-nJ}}

and the value “-” with probability 1p11-p_{1} (resp. 1p21-p_{2}). Such p1p_{1} and p2p_{2} exist by (17).

Fix a triangle face F0F_{0} of 𝕃2{\mathbb{L}}_{2}. Let BR=(V(BR),E(BR))B_{R}=(V(B_{R}),E(B_{R})) be the finite subgraph of 𝕃2{\mathbb{L}}_{2} consisting of all the faces of 𝕃2{\mathbb{L}}_{2} whose graph distance to F0F_{0} is at most RR. Let ν1,R\nu_{1,R} (resp. ν2,R\nu_{2,R}) be the restriction of ν1\nu_{1} (resp. ν2\nu_{2}) on BRB_{R}. Let μR+\mu_{R}^{+} (resp. μR)\mu_{R}^{-}) be the probability measure for the Ising model on BRB_{R} with respect to the coupling constant JJ and the “+” boundary condition (resp.  the “-” boundary condition). Let ω1\omega_{1}, ω2\omega_{2} be two configurations in {1,1}V(BR)\{-1,1\}^{V(B_{R})}. Then by Lemmas 11.2 and 11.3, we can check the F.K.G. lattice conditions below

ν1,R(max{ω1,ω2})μR+(min{ω1,ω2})ν1,R(ω1)μR+(ω2)\displaystyle\nu_{1,R}(\max\{\omega_{1},\omega_{2}\})\mu_{R}^{+}(\min\{\omega_{1},\omega_{2}\})\geq\nu_{1,R}(\omega_{1})\mu_{R}^{+}(\omega_{2})
μR(max{ω1,ω2})ν2,R(min{ω1,ω2})μR(ω1)ν2,R(ω1).\displaystyle\mu_{R}^{-}(\max\{\omega_{1},\omega_{2}\})\nu_{2,R}(\min\{\omega_{1},\omega_{2}\})\geq\mu_{R}^{-}(\omega_{1})\nu_{2,R}(\omega_{1}).

Then we obtain the following stochastic domination result:

ν2,RμRμR+ν1,R.\displaystyle\nu_{2,R}\prec\mu_{R}^{-}\prec\mu_{R}^{+}\prec\nu_{1,R}.

Letting RR\rightarrow\infty, for any Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariant Gibbs measure μ\mu for the Ising model on 𝕃2{\mathbb{L}}_{2} with coupling constant JJ, we have

ν2μμμ+ν1.\displaystyle\nu_{2}\prec\mu^{-}\prec\mu\prec\mu^{+}\prec\nu_{1}.

Since ν2\nu_{2}-a.s. there are infinite “++”-clusters, μ\mu-a.s. there are infinite “++”-clusters. Similarly, μ\mu-a.s. there are infinite “-”-clusters, since ν1\nu_{1}-a.s. there are infinite “-”-clusters. By Proposition 10.8, we conclude that when (17) hold, μ\mu-a.s. there are infinitely many infinite “++”-clusters, infinitely many infinite “-”-clusters and infinitely many infinite contours. This completes the proof of Part I.

11.2. Proof of Part II of Theorem 3.3.

We first assume that μf\mu^{f} is Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-ergodic. Since μf\mu^{f} is also symmetric in switching “++” and “-” states, μf\mu^{f}-a.s. the number of infinite “++”-clusters and the number of infinite “-”-clusters are equal. Then the conclusion follows from Proposition 10.8, which implies that if the number of infinite “++”-clusters and the number of infinite “-”-clusters are equal a.s., then both numbers are infinite.

By Lemma 11.8, if one of the conditions (a), (b), (c) and (d) in Lemma 11.8 holds, then μf\mu^{f} almost surely there are infinitely many infinite “++”-clusters and infinitely many infinite “-”-clusters.

11.3. Proof of Corollary 3.4.

By Proposition 3.2 (i) of [20], if (23) holds, then there is a unique infinite-volume Gibbs measure for the Ising model on 𝕃2{\mathbb{L}}_{2} with coupling constant JJ. Since

pc,2wpc,2fpu,2f,\displaystyle p_{c,2}^{w}\leq p_{c,2}^{f}\leq p_{u,2}^{f},

(23) implies Condition (c). Then Theorem 3.3 II(c) implies μf\mu^{f} a.s. there are neither infinite “++”-clusters nor infinite “-”-cluster. The corollary now follows from the uniqueness of the infinite-volume Gibbs measure.

11.4. Proof of Part III of Theorem 3.3.

Since p=1e2Jp=1-e^{-2J}, when (18) holds, we have ppu,2wp\geq p_{u,2}^{w}. By Corollary 3.7 of [20] (see also (15)), there exists a unique infinite open cluster τ\tau in the random cluster representation of the Ising model with wired boundary conditions WRCp,2WRC_{p,2}-a.s.

By the correspondence of random-cluster configurations and Ising configurations, each infinite cluster in the random cluster representation must be a subset of an infinite cluster in the Ising model. Hence if WRCp,2WRC_{p,2}-a.s. there is a unique infinite open cluster, μ+\mu^{+} a.s. there exists an infinite “++”-cluster in the Ising model, and μ\mu^{-}-a.s. there exists an infinite “-”-cluster in the Ising model.

Let τ\tau be the unique infinite open cluster in the random cluster configuration ω\omega. We define a site percolation configuration ξ\xi on V(𝕃2)V({\mathbb{L}}_{2}), by letting all the vertices in τ\tau have state 1, and all the other vertices have state 0. By Lemma 11.6, a.s. ξ\xi has no-infinite 0-clusters. Again by the correspondence of the random cluster configuration and the Ising configuration and proposition 10.8, we obtain μ+(𝒜+)=1\mu^{+}({\mathcal{A}}_{+})=1, and μ(𝒜)=1\mu^{-}({\mathcal{A}}_{-})=1.

11.5. Proof of Part IV of Theorem 3.3.

The identities (19) and (20) follows Part I and the fact that

pu,2wpu,2f;\displaystyle p_{u,2}^{w}\leq p_{u,2}^{f};

and (22) follows from the fact that when (21) hold,

μf=μ++μ2;\displaystyle\mu^{f}=\frac{\mu^{+}+\mu^{-}}{2};

The decomposition of μf\mu^{f} as convex combination of the extremal measures μ+\mu^{+} and μ\mu^{-} it is a classical results in the case of regular lattices and for our lattices it was proven in Theorem 4.2 of [45] and expressions (17) (18) of [20].

12. Proof of Theorems 4.1 and 4.2

In this section, we prove Theorems 4.1 and 4.2.


Proof of theorem 4.1. We first prove Part I of the theorem. Let s+s_{+} (resp. ss_{-}) be the number of infinite “++”-clusters (resp. infinite “-”-clusters). By ergodicity, Aut(𝕃2)\mathrm{Aut}({\mathbb{L}}_{2})-invariance and symmetry in “++” and “-” of μ1,f×μ2,f\mu_{1,f}\times\mu_{2,f}, as well as Lemma 6.1, one of the following cases occurs:

  1. (i)

    μ1f×μ2f((s+,s)=(0,0))=1\mu_{1}^{f}\times\mu_{2}^{f}((s_{+},s_{-})=(0,0))=1; or

  2. (ii)

    μ1f×μ2f((s+,s)=(1,1))=1\mu_{1}^{f}\times\mu_{2}^{f}((s_{+},s_{-})=(1,1))=1; or

  3. (iii)

    μ1f×μ2f((s+,s)=(,))=1\mu_{1}^{f}\times\mu_{2}^{f}((s_{+},s_{-})=(\infty,\infty))=1

Case (i) is impossible to occur by Lemma 10.1. Case (ii) is impossible to occur by Lemma 10.3. This completes the proof of Part I.

Now we show that Assumption II implies Assumption I. This follows from Theorem 4.1 of [44], and the fact that the XOR Ising measure is the product measure of two i.i.d Ising models.

The facts that Assumption IV implies Assumption III and Assumption III implies Assumption II follows from Lemma 11.8. This completes the proof of the theorem. \hfill\Box


Before proving Theorem 4.2, we shall first introduce the following definition and proposition in [39] (see Theorem 3.3, Remark 3.4).

Definition 12.1.

Let G=(V,E)G=(V,E) be a graph and Γ\Gamma a transitive group acting on GG. Suppose that XX is either VV, EE or VEV\cup E. Let QQ be a measurable space and Ω:=2V×QX\Omega:=2^{V}\times Q^{X}. A probability measure 𝐏\mathbf{P} on Ω\Omega will be called a site percolation with scenery on GG. The projection onto 2V2^{V} is the underlying percolation and the projection onto QXQ^{X} is the scenery. If (ω,q)Ω(\omega,q)\in\Omega, we set Πv(ω,q)=(Πvω,q)\Pi_{v}(\omega,q)=(\Pi_{v}\omega,q). We say the percolation with scenery 𝐏\mathbf{P} is insertion-tolerant if 𝐏(Πv)>0\mathbf{P}(\Pi_{v}\mathcal{B})>0 for every measurable Ω\mathcal{B}\subset\Omega with positive measure. We say that 𝐏\mathbf{P} has indistinguishable infinite clusters if for every 𝒜2V×2V×QX\mathcal{A}\subset 2^{V}\times 2^{V}\times Q^{X} that is invariant under diagonal actions of Γ\Gamma, for 𝐏\mathbf{P}-a.e. (ω,q)(\omega,q), either all infinite clusters CC of ω\omega satisfy (C,ω,q)𝒜(C,\omega,q)\in\mathcal{A}, or they all satisfy (C,ω,q)𝒜(C,\omega,q)\notin\mathcal{A}.

Proposition 12.2.

Let 𝐏\mathbf{P} be a site percolation with scenery on a graph G=(V,E)G=(V,E) with state space Ω:=2V×QX\Omega:=2^{V}\times Q^{X}, where QQ is a measurable space and XX is either VV, EE or VEV\cup E. If 𝐏\mathbf{P} is Γ\Gamma-invariant and insertion tolerant, then 𝐏\mathbf{P} has indistinguishable infinite clusters.

Proof of Theorem 4.2 Let Λ=(VΛ,EΛ)\Lambda=(V_{\Lambda},E_{\Lambda}) be a subgraph of 𝕃2{\mathbb{L}}_{2} consisting of faces of 𝕃2{\mathbb{L}}_{2}. Let Λ=(VΛ,EΛ)\Lambda_{*}=(V_{\Lambda_{*}},E_{\Lambda_{*}}) be the dual graph of Λ\Lambda, such that there is a vertex in VΛV_{\Lambda} corresponding to each triangle face in Λ\Lambda, as well as the unbounded face; the edges in EΛE_{\Lambda} and EΛE_{\Lambda_{*}} are in 1-1 correspondence by duality.

Consider an XOR Ising model on Λ\Lambda with respect to two i.i.d. Ising models σ3\sigma_{3}, σ4\sigma_{4} with free boundary conditions and coupling constants J0J\geq 0 satisfying the assumption of Theorem 4.1. The partition function of the XOR Ising model can be computed by

ZΛ,f=σ3,σ4{±1}VΛ(u,v)EΛeJ(σ3,uσ3,v+σ4,uσ4,v).\displaystyle Z_{\Lambda,f}=\sum_{\sigma_{3},\sigma_{4}\in\{\pm 1\}^{V_{\Lambda}}}\prod_{(u,v)\in E_{\Lambda}}e^{J(\sigma_{3,u}\sigma_{3,v}+\sigma_{4,u}\sigma_{4,v})}.

Following the same computations as in [9], we obtain

(44) ZΛ,f=C1P𝒫,P𝒫,PP=(2e2J1+e4J)|P|(1e4J1+e4J)|P|.\displaystyle Z_{\Lambda,f}=C_{1}\sum_{P_{*}\in\mathcal{P}_{*},P\in\mathcal{P},P\cap P_{*}=\emptyset}\left(\frac{2e^{-2J}}{1+e^{-4J}}\right)^{|P_{*}|}\left(\frac{1-e^{-4J}}{1+e^{-4J}}\right)^{|P|}.

where 𝒫\mathcal{P}_{*} (resp. 𝒫\mathcal{P}) consists of all the contour configurations on EΛE_{\Lambda_{*}} (resp. EΛE_{\Lambda}) such that each vertex of VΛV_{\Lambda_{*}} (resp. VΛV_{\Lambda}) has an even number of incident present edges, and C1=2|VΛ||EΛ|+2(e2je2J)|EΛ|C_{1}=2^{|V_{\Lambda}|-|E_{\Lambda}|+2}(e^{2j}-e^{-2J})^{|E_{\Lambda}|} is a constant.

When J,KJ,K satisfies (24), we have

2e2J1+e4J\displaystyle\frac{2e^{-2J}}{1+e^{-4J}} =\displaystyle= 1e4K1+e4K;\displaystyle\frac{1-e^{-4K}}{1+e^{-4K}};
2e2K1+e4K\displaystyle\frac{2e^{-2K}}{1+e^{-4K}} =\displaystyle= 1e4J1+e4J.\displaystyle\frac{1-e^{-4J}}{1+e^{-4J}}.

Thus the partition function ZΛ,fZ_{\Lambda,f}, up to a multiplicative constant, is the same as the partition function of the XOR Ising model on Λ\Lambda_{*} with coupling constant KK.

Recall that there is exactly one vertex vVΛv_{\infty}\in V_{\Lambda_{*}} corresponding to the unbounded face in Λ\Lambda. The XOR Ising model σXOR=σ1σ2\sigma_{XOR}=\sigma_{1}\sigma_{2} on Λ\Lambda_{*}, corresponds to an XOR Ising model on Λ{v}\Lambda_{*}\setminus\{v_{\infty}\} (which is a subgraph of 𝕃1{\mathbb{L}}_{1}) with the boundary condition that all the boundary vertices have the same state in σ1\sigma_{1} and all the boundary vertices have the same state in σ2\sigma_{2}. Hence the boundary condition must be a mixture of ++++, ++-, +-+ and --. However, each one of the 4 possible boundary conditions gives the same distribution of contours in the XOR Ising model. From the expression (44), we can see that there is a natural probability measure on the set of contours Φ={(P,P):P𝒫,P𝒫,PP=}\Phi=\{(P,P_{*}):P\in\mathcal{P},P_{*}\in\mathcal{P}_{*},P\cap P_{*}=\emptyset\}, such that the probability of each pair of contours (P,P)Φ(P,P_{*})\in\Phi is proportional to (2e2J1+e4J)|P|(1e4J1+e4J)|P|\left(\frac{2e^{-2J}}{1+e^{-4J}}\right)^{|P_{*}|}\left(\frac{1-e^{-4J}}{1+e^{-4J}}\right)^{|P|}, and the marginal distribution on 𝒫\mathcal{P} is the distribution of contours for the XOR Ising model on 𝕃2{\mathbb{L}}_{2} with coupling constant JJ and free boundary conditions, while the marginal distribution on 𝒫\mathcal{P}_{*} is the distribution of contours for the XOR Ising model on 𝕃1{\mathbb{L}}_{1} with coupling constant KK and ++++ boundary conditions.

We let Λ\Lambda and Λ{v}\Lambda_{*}\setminus\{v_{\infty}\} increase and approximate the graph 𝕃2{\mathbb{L}}_{2} and 𝕃1{\mathbb{L}}_{1}, respectively. If with a positive μ++\mu_{++} probability, there exists exactly one infinite contour CC in 𝕃2{\mathbb{L}}_{2}, then μf\mu_{f}-a.s. there exists an infinite cluster in 𝕃2{\mathbb{L}}_{2} containing CC, since contours in 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} are disjoint. Consider the XOR Ising spin configuration as a site percolation on 𝕃2{\mathbb{L}}_{2}, with scenery given by contour configurations of 𝕃2{\mathbb{L}}_{2} within the “++” clusters of 𝕃2{\mathbb{L}}_{2}. In the notation of Definition 12.1, Q={0,1}Q=\{0,1\}, and X=E(𝕃2)X=E({\mathbb{L}}_{2}). An edge in E(𝕃2)E({\mathbb{L}}_{2}) is present (has state “1”) if and only if it is in a “++”-cluster of the XOR Ising configuration on 𝕃2{\mathbb{L}}_{2} and present in the contour configuration of 𝕃2{\mathbb{L}}_{2}. This way we obtain an automorphism-invariant and insertion-tolerant percolation with scenery. Let 𝒜2V(𝕃2)×2V(𝕃2)×2E(𝕃2)\mathcal{A}\subset 2^{V}({\mathbb{L}}_{2})\times 2^{V}({\mathbb{L}}_{2})\times 2^{E}({\mathbb{L}}_{2}) be the triple (C,ω,q)(C,\omega,q) such that

  • ω\omega is an XOR Ising spin configuration on 𝕃2{\mathbb{L}}_{2}; and

  • CC is an infinite “++”-cluster; and

  • qq is the 𝕃2{\mathbb{L}}_{2}-contour configuration within “++”-clusters of ω\omega; and

  • CC contains an infinite contour in qq.

We can see that 𝒜\mathcal{A} is invariant under diagonal actions of automorphisms. By Theorem 4.1, μf\mu_{f}-a.s. there exists infinitely many infinite “++”-clusters in 𝕃2{\mathbb{L}}_{2}. By Proposition 12.2, either all the infinite clusters are in 𝒜\mathcal{A}, or no infinite clusters are in 𝒜\mathcal{A}. Similar arguments applies for “-”-clusters of 𝕃2{\mathbb{L}}_{2}. Hence almost surely the number of infinite contours in 𝕃2{\mathbb{L}}_{2} is 0 or \infty. Since the distribution of infinite contours in 𝕃2{\mathbb{L}}_{2} is exactly that of contours for the XOR Ising model on 𝕃1{\mathbb{L}}_{1} with coupling constant KK and ++++ (or ++-, +-+, --) boundary condition, the theorem follows. \hfill\Box

13. Proof of Theorems 5.1 and 5.2

In this section, we prove Theorems 5.1 and 5.2.

Consider the XOR Ising model with spins located on vertices of the hexagonal lattice {\mathbb{H}} with coupling constants JaJ_{a}, JbJ_{b}, JcJ_{c} on horizontal, NW/SE, NE/SW edges, see Figure 11 for an embedding of {\mathbb{H}} into the plane such that all the edges are either horizontal, NW/SE or NE/SW.

Refer to caption
(a) {\mathbb{H}} and 𝕋{\mathbb{T}}
Refer to caption
(b) 𝕋{\mathbb{T}} and {\mathbb{H}}^{\prime}
Figure 11. Hexagonal lattice {\mathbb{H}} (represented by black lines), dual triangular lattice 𝕋{\mathbb{T}} (represented by red lines) and hexagonal lattice {\mathbb{H}}^{\prime} obtained from the star-triangle transformation (represented by blue lines).

A [4,6,12] lattice is a graph that can be embedded into the Euclidean plane 2{\mathbb{R}}^{2} such that each vertex is incident to 3 faces with degrees 4, 6, and 12, respectively. See Figure 12. We shall explain the relation between perfect matchings on the [4,6,12] lattice and constrained percolation configurations in the [3,4,6,4] lattice as discussed in Section 2.

Refer to caption
Figure 12. [3,4,6,4] lattice (represented by dashed lines) and the [4,6,12] lattice 𝔸\mathbb{A} (represented by black lines).

A perfect matching, or a dimer configuration on a [4,6,12] lattice is a subset of edges such that each vertex of the [4,6,12] lattice is incident to exactly one edge in the subset.

A Type-I edge of a [4,6,12] lattice is an edge of a square face. Any other edge of the [4,6,12] lattice is a Type-II edge. We say two Type-II edges e1,e2e_{1},e_{2} are adjacent if there exists a Type-I edge e3e_{3}, such that both e1e_{1} and e2e_{2} share a vertex with e3e_{3} in the [4,6,12] lattice. A subset of Type-II edges is connected if for any two edges ee and ff in the subset, there exist a sequence of Type-II edges e0(=e),e1,,en(=f)e_{0}(=e),\ e_{1},\ \ldots,\ e_{n}(=f) in the subset, such that eie_{i} and ei1e_{i-1} are adjacent, for 1in1\leq i\leq n. A Type-II cluster is a maximal connected set of present Type-II edges in a perfect matching.

The [3,4,6,4] lattice is constructed as follows. A vertex of the [3,4,6,4] lattice is placed at the midpoint of each Type-II edge of the [4,6,12] lattice. Two vertices of the [3,4,6,4] lattices are joined by an edge if and only if they are midpoints of two adjacent Type-II edges. See Figure 12. The restriction of any dimer configuration on the [4,6,12] lattice to Type-II edges naturally correspond to a constrained percolation configuration on the [3,4,6,4] lattice in Ω\Omega. A Type-II edge is present in a dimer configuration if and only if its midpoint has state “1” in the corresponding constrained percolation configuration. It is straightforward to check that this way we obtain a bijection between restrictions to Type-II edges of dimer configurations on the [4,6,12] lattice and constrained percolation configurations on the [3,4,6,4] lattice in Ω\Omega. Recall that constrained percolation configurations on the [3,4,6,4] lattice induces contour configurations on the hexagonal lattice {\mathbb{H}} and the triangular lattice 𝕋{\mathbb{T}} by a 2-to-1 mapping ϕ:ΩΦ\phi:\Omega\rightarrow\Phi. See Figure 6.

From the connection of the [4,6,12] lattice and the [3,4,6,4] lattice, as well as the connection of the [3,4,6,4] lattice with the hexagonal lattice {\mathbb{H}} and the dual triangular lattice 𝕋{\mathbb{T}} as described in Section 2, we can see that each square face of the [4,6,12][4,6,12] lattice is crossed by a unique edge of {\mathbb{H}} and a unique edge of 𝕋{\mathbb{T}}. Each vertex of {\mathbb{H}} is located at the center of a hexagon face of the [4,6,12] lattice, and each vertex of 𝕋{\mathbb{T}} is located in the center of a degree-12 face of the [4,6,12] lattice.

Note that the {\mathbb{H}} is a bipartite graph; i.e. all the vertices can be colored black and white such that the vertices of the same color are not adjacent. Let Γ\Gamma be the translation group of the hexagonal lattice generated by translations along two different directions, such that the set of black vertices and the set of white vertices form two distinct orbits under the action of Γ\Gamma. Note that Γ\Gamma acts transitively on the dual triangular lattice 𝕋{\mathbb{T}}.

In order to define a probability measure for perfect matchings on the [4,6,12] lattice, we introduce edge weights. We assign weight 1 to each Type-II edge, and weight wew_{e} to the Type-I edge ee. Assume that the edge weights of the [4,6,12] lattice satisfy the following conditions.

  1. (B1)

    The edge weights are Γ\Gamma-invariant.

  2. (B2)

    If e1e_{1}, e2e_{2} are two parallel Type-I edges around the same square face, then we1=we2w_{e_{1}}=w_{e_{2}}.

  3. (B3)

    If e1e_{1}, e2e_{2} are two perpendicular Type-I edges around the same square face, then we12+we22=1w_{e_{1}}^{2}+w_{e_{2}}^{2}=1.

The reason we assume (B1) is to define a Γ\Gamma-translation-invariant probability measure. The reason we assume (B2) and (B3) is to define a probability measure for dimer configurations of the [4,6,12] lattice, which, under the connection described above to constrained percolation configurations in Ω\Omega, will induce a probability measure on Ω\Omega satisfying the symmetry assumption (A3).

Under (B1)–(B3), the edge weights are described by three independent parameters. We may sometimes assume that the parameters satisfy the identity below, which reduces the number of independent parameters to two.

  1. (B4)

    Let

    (45) h(x,y,z)=x+y+z+xy+xz+yzxyz1.\displaystyle h(x,y,z)=x+y+z+xy+xz+yz-xyz-1.

    For each edge ee of the hexagonal lattice {\mathbb{H}}, let e1e_{1} (resp. e2e_{2}) be a Type-I edge of the [4,6,12] lattice parallel (resp. perpendicular) to ee. Let

    te=1we1we2,\displaystyle t_{e}=\frac{1-w_{e_{1}}}{w_{e_{2}}},

    where we1w_{e_{1}} (resp. we2w_{e_{2}}) is the edge weight of e1e_{1} (resp. e2e_{2}) for dimer configurations on the [4,6,12] lattice. Under the assumption (B1), tet_{e} is uniquely defined independent of the e1e_{1}, e2e_{2} chosen - as long as e1e_{1} is parallel to ee and e2e_{2} is perpendicular to ee. Let ea,eb,ece_{a},e_{b},e_{c} be three edges of {\mathbb{H}} with distinct orientations in the embedding of {\mathbb{H}} into 2{\mathbb{R}}^{2}. Then h(tea,teb,tec)=0h(t_{e_{a}},t_{e_{b}},t_{e_{c}})=0.

In [30], the authors define a probability measure for any bi-periodic, bipartite, 2-dimensional lattice. Specializing to our case, let μn,D\mu_{n,D} be the probability measure of dimer configurations on a toroidal n×nn\times n [4,6,12] lattice 𝔸n\mathbb{A}_{n} (see [30] for details). Let n\mathcal{M}_{n} be the set of all perfect matchings on 𝔸n\mathbb{A}_{n}, and let MnM\in\mathcal{M}_{n} be dimer configuration, then

(46) μn,D(M)=eMweMneMwe,\displaystyle\mu_{n,D}(M)=\frac{\prod_{e\in M}w_{e}}{\sum_{M\in\mathcal{M}_{n}}\prod_{e\in M}w_{e}},

where wew_{e} is the weight of the edge ee. As nn\rightarrow\infty, μn,D\mu_{n,D} converges weakly to a translation-invariant measure μD\mu_{D} (see [30]).

Theorem 13.1.

For the dimer model on the [4,6,12] lattice.

  1. I

    If the edge weights satisfy (B1)-(B4), μD\mu_{D} almost surely there are neither infinite Type-II clusters nor infinite contours.

  2. II

    If the edge weights satisfy (B1)-(B3), μD\mu_{D} almost surely there exists at most one infinite contour.

Proof.

Let μ~D\widetilde{\mu}_{D} be the marginal distribution of μD\mu_{D} restricted on Type-II edges. Recall that the restriction of dimer configurations to Type-II edges on the [4,6,12] lattice, are in 1-1 correspondence with constrained percolation configurations on the [3,4,6,4][3,4,6,4] lattice in Ω\Omega, as described before. See also Figure 12.

Also recall that μ~D\widetilde{\mu}_{D} is the weak limit of measures on larger and larger tori; since the edge weights are translation-invariant, the measures on tori are translation-invariant. Hence μ~D\widetilde{\mu}_{D} satisfies (A1) if edge weights satisfies (B1).

The measure μ~D\widetilde{\mu}_{D} is both translation-invariant and mixing (see [30]), hence μ~D\widetilde{\mu}_{D} is totally ergodic and satisfies (A2).

If the edge weights satisfy (B2) and (B3), the measures on tori are symmetric under θ\theta. Hence μ~D\widetilde{\mu}_{D} satisfies (A3).

By the results in [9], the marginal distribution of contours in {\mathbb{H}} (resp. 𝕋{\mathbb{T}}) under μ~D\widetilde{\mu}_{D} is the same as the distribution of contours of an XOR Ising model σXOR,𝕋\sigma_{XOR,{\mathbb{T}}} (resp. σXOR,\sigma_{XOR,{\mathbb{H}}}) with spins located on vertices of 𝕋{\mathbb{T}} (resp. {\mathbb{H}}), if the dimer edge weights and Ising coupling constants satisfy the following conditions:

  • each Type-II edge has weight 1;

  • each Type-I edge parallel to an edge of ee with coupling constants JeJ_{e} has weight wew_{e} such that we=1e4Je1+e4Jew_{e}=\frac{1-e^{-4J_{e}}}{1+e^{-4J_{e}}};

  • each Type-I edge perpendicular to an edge ee with coupling constants JeJ_{e} has weight wew_{e} such that we=2e2Je1+e4Jew_{e}=\frac{2e^{-2J_{e}}}{1+e^{-4J_{e}}}.

Moreover, when the edge weights satisfy (B1)-(B3), σXOR,𝕋\sigma_{XOR,{\mathbb{T}}} and σXOR,\sigma_{XOR,{\mathbb{H}}} are dual to each other, i.e. the coupling constants JτJ_{\tau} and KτK_{\tau} on a pair of dual edges e,e𝕋e\in{\mathbb{H}},e^{*}\in{\mathbb{T}} satisfy (27). The finite energy assumptions (A4) and (A5) follows from the finite energy of σXOR,\sigma_{XOR,{\mathbb{H}}} and σXOR,𝕋\sigma_{XOR,{\mathbb{T}}}.

Since σXOR,\sigma_{XOR,{\mathbb{H}}} and σXOR,𝕋\sigma_{XOR,{\mathbb{T}}} are dual to each other, one of the following cases might occur

  1. I

    σXOR,\sigma_{XOR,{\mathbb{H}}} is in the low-temperature state, and σXOR,𝕋\sigma_{XOR,{\mathbb{T}}} is in the high-temperature state;

  2. II

    σXOR,\sigma_{XOR,{\mathbb{H}}} is in the high-temperature state, and σXOR,𝕋\sigma_{XOR,{\mathbb{T}}} is in the low-temperature state;

  3. III

    both σXOR,\sigma_{XOR,{\mathbb{H}}} and σXOR,𝕋\sigma_{XOR,{\mathbb{T}}} are in the critical state.

See Section 5 for definitions of the low-temperature state, high-temperature state and critical state for XOR Ising models.

In Case III, both (A6) and (A7) are satisfied because of the ergodicity of measures for the critical XOR Ising model on {\mathbb{H}} and 𝕋{\mathbb{T}}; see Lemma 13.2. Moreover, Case III occurs if and only if the edge weights satisfy (B4). Hence when the edge weights satisfy (B1)-(B4), μ~D\widetilde{\mu}_{D} satisfy (A1)-(A7). Theorem 13.1 I follows from Theorem 2.4II(c).

Note that the measures for the high-temperature XOR Ising models on {\mathbb{H}} and 𝕋{\mathbb{T}} are also ergodic; see Lemma 13.3. Therefore, in each case of I, II and III, at least one of (A6) and (A7) is satisfied. Then Theorem 13.1 II follows from Theorem 2.4 II(a)(b) and Lemmas 8.1 and 8.2.

In order to prove the Theorems 5.1 and 5.2, we prove the following lemmas.

Lemma 13.2.

The measure for the critical XOR Ising model on {\mathbb{H}} (resp. 𝕋{\mathbb{T}}), obtained as the weak limit of measures on tori, is ergodic.

Proof.

Let ρ=σ+12\rho=\frac{\sigma+1}{2}, where σ:V{±1}\sigma:V_{{\mathbb{H}}}\rightarrow\{\pm 1\} is the spin configuration for an Ising model on {\mathbb{H}}. Following the same arguments as in the proof of Lemma 10.2 in [27], it suffices to show that for the critical Ising model on {\mathbb{H}} (resp. 𝕋{\mathbb{T}}), we have

lim|uv||ρ(u)ρ(v)ρ(u)ρ(v)|=0,\displaystyle\lim_{|u-v|\rightarrow\infty}|\langle\rho(u)\rho(v)\rangle-\langle\rho(u)\rangle\langle\rho(v)\rangle|=0,

which is equivalent to show that

(47) lim|uv|σ(u)σ(v)=0.\displaystyle\lim_{|u-v|\rightarrow\infty}\langle\sigma(u)\sigma(v)\rangle=0.

Note that σ(u)σ(v)\langle\sigma(u)\sigma(v)\rangle is an even spin correlation function (i.e. the expectation of the product of spins on an even number of vertices), and hence for all the infinite-volume, translation-invariant Gibbs measures of the Ising model on {\mathbb{H}} (resp. 𝕋{\mathbb{T}}) corresponding to the given coupling constant, σuσv\langle\sigma_{u}\sigma_{v}\rangle has a unique value; see [33].

Consider the FK random cluster representation of the Ising model with q=2q=2, the two-point spin correlation σ(u)σ(v)\langle\sigma(u)\sigma(v)\rangle is exactly the connectivity probability of uu and vv in the random cluster model, up to a multiplicative constant; see Chapter 1.4 of [17]. Therefore in order to show (47), it suffices to show that the connectivity probabilities of two vertices in the corresponding random cluster model, as the distances of the two vertices go to infinity, converge to zero.

By Theorem 4 of [7], we infer that the connectivity probabilities of two vertices in the q=2q=2 random cluster model corresponding to the critical Ising model on the triangular lattices with coupling constants Ka,Kb,KcK_{a},K_{b},K_{c} satisfying g(Ka,Kb,Kc)=0g(K_{a},K_{b},K_{c})=0 converge to zero as the distances of the two vertices go to infinity.

Note that the hexagonal lattice is a bipartite graph, i.e., all the vertices can be colored black and white such that vertices of the same color can never be adjacent. Recall that the star-triangle transformation is a replacement of each black vertex of {\mathbb{H}}, as well as its incident edges, into a triangle. The resulting graph is a triangular lattice 𝕋{\mathbb{T}}^{\prime}; see the right graph of Figure 11. The parameters of the random cluster model on {\mathbb{H}} and the random cluster model 𝕋{\mathbb{T}}^{\prime} satisfy certain identities, such that the probabilities of connections of any two adjacent vertices in 𝕋{\mathbb{T}}^{\prime} (which are also vertices in {\mathbb{H}}) internal to each triangle face in 𝕋{\mathbb{T}}^{\prime} which has a black vertex of {\mathbb{H}} in the center are the same for the random cluster model on {\mathbb{H}} and the random cluster model on 𝕋{\mathbb{T}}^{\prime}; see Page 160-161 of [17].

Using a star-triangle transformation (see the right graph of Figure 11), and (6.69) of [17], we deduce that the connectivity probabilities of two vertices in the q=2q=2 random cluster model on the hexagonal lattice corresponding to the critical Ising model on {\mathbb{H}} with coupling constants Ja,Jb,JcJ_{a},J_{b},J_{c} satisfying f(Ja,Jb,Jc)=0f(J_{a},J_{b},J_{c})=0 converge to zero as the distances of the two vertices go to infinity. Note that the weak limit of of measures with free boundary conditions is known to exist and translation-invariant, see Theorem (4.19) of [17]. By the uniqueness of σ(u)σ(v)\langle\sigma(u)\sigma(v)\rangle under all the translation-invariant measures, we obtain that (47) holds under the measure obtained as the weak limit of measures with periodic boundary conditions. ∎

Lemma 13.3.

The measure for the high-temperature XOR Ising model on {\mathbb{H}} (resp. 𝕋{\mathbb{T}}), obtained as the weak limit of measures on tori, is ergodic.

Proof.

The identity (47) holds under the measure for the high-temperature Ising model on {\mathbb{H}} (resp. 𝕋{\mathbb{T}}); see [35, 14]. ∎

Proof of Theorem 5.1. We first show that in the critical XOR Ising model on {\mathbb{H}} or 𝕋{\mathbb{T}}, almost surely there are no infinite contours. It is proved in [9] that the contours of XOR Ising model with spins located on {\mathbb{H}} (resp. 𝕋{\mathbb{T}}) have the same distribution as contours in 𝕋{\mathbb{T}} (resp. {\mathbb{H}}) for the Type-II clusters of dimer configurations on the [4,6,12][4,6,12] lattice, if the coupling constants of the XOR Ising model and the edge weights of the [4,6,12][4,6,12] lattice satisfy certain conditions. It is not hard to check that when the coupling constants of the XOR Ising model on 𝕋{\mathbb{T}} (resp. {\mathbb{H}}) are critical, then the edge weights of the corresponding dimer model on the [4,6,12] lattice satisfy (B1)-(B4). By Theorem 13.1 I, almost surely there are no infinite contours in the critical XOR Ising model on 𝕋{\mathbb{T}} or {\mathbb{H}}.

Now we prove that almost surely there are no infinite clusters for the critical XOR Ising model on {\mathbb{H}} or 𝕋{\mathbb{T}}. We write down the proof for the critical XOR Ising model on {\mathbb{H}} here, the case for the XOR Ising model on 𝕋{\mathbb{T}} can be proved in a similar way.

Let 𝒜{\mathcal{A}} be the event that there exists an infinite cluster for the XOR Ising model on {\mathbb{H}}. Assume that μ(𝒜)>0\mu({\mathcal{A}})>0; we will obtain a contradiction. By translation-invariance of 𝒜{\mathcal{A}} and Lemma 13.2, if μ(𝒜)>0\mu({\mathcal{A}})>0 then μ(𝒜)=1\mu({\mathcal{A}})=1. Let 𝒜1{\mathcal{A}}_{1} (resp. 𝒜2{\mathcal{A}}_{2}) be the event that there exists an infinite “++”-cluster (resp. “-”-cluster) for the XOR Ising model on {\mathbb{H}}, then

(48) μ(𝒜1𝒜2)=1.\displaystyle\mu({\mathcal{A}}_{1}\cup{\mathcal{A}}_{2})=1.

By symmetry μ(𝒜1)=μ(𝒜2)\mu({\mathcal{A}}_{1})=\mu({\mathcal{A}}_{2}). By translation-invariance of 𝒜1{\mathcal{A}}_{1}, 𝒜2{\mathcal{A}}_{2} and Lemma 13.2, either μ(𝒜1)=μ(𝒜2)=1\mu({\mathcal{A}}_{1})=\mu({\mathcal{A}}_{2})=1, or μ(𝒜1)=μ(𝒜2)=0\mu({\mathcal{A}}_{1})=\mu({\mathcal{A}}_{2})=0. By (48), we have μ(𝒜1)=μ(𝒜2)=1\mu({\mathcal{A}}_{1})=\mu({\mathcal{A}}_{2})=1, hence μ(𝒜1𝒜2)=1\mu({\mathcal{A}}_{1}\cap{\mathcal{A}}_{2})=1, i.e. μ\mu-a.s. there exist both an infinite “++”-cluster and an infinite “-”-cluster in the critical XOR Ising configuration on {\mathbb{H}}.

Let ϕ𝕋\phi_{{\mathbb{T}}} be the contour configuration associated to the critical XOR Ising configuration on {\mathbb{H}}. Let ωϕ1(ϕ𝕋)\omega\in\phi^{-1}(\phi_{{\mathbb{T}}}) be a constrained percolation configuration on the [3,4,6,4][3,4,6,4] lattice whose contour configuration is ϕ𝕋\phi_{{\mathbb{T}}}. It is not hard to check that in ω\omega there exist both an infinite 1-cluster and an infinite 0-cluster if in the original XOR Ising model on {\mathbb{H}}, there exist both an infinite “++”-cluster and an infinite “-”-cluster. By Lemma A.3, μ\mu-a.s. there exists an infinite contour in ϕ𝕋\phi_{{\mathbb{T}}}. The contradiction implies that μ\mu-a.s. there are no infinite clusters in the critical XOR Ising model on {\mathbb{H}}. \hfill\Box


Proof of Theorem 5.2. By the correspondence between contours in an XOR Ising model with spins located on vertices of {\mathbb{H}} (resp. 𝕋{\mathbb{T}}) and contours on 𝕋{\mathbb{T}} (resp. {\mathbb{H}}) for the Type-II clusters of dimer configurations on the [4,6,12][4,6,12] lattice, as proved in [9], as well as correspondence between Type-II clusters of dimer configurations on the [4,6,12] lattice and clusters of constrained configurations on the [3,4,6,4][3,4,6,4] lattice and Theorem 2.4, it suffices to show that the probability measure for the low-temperature XOR Ising model on {\mathbb{H}} (resp. 𝕋{\mathbb{T}}) satisfies (A1)-(A5) and (A7) (resp. (A1)-(A6)), for 𝕃1={\mathbb{L}}_{1}={\mathbb{H}}.

It is straightforward to verify (A1)-(A5). The assumption (A6) (resp. (A7)) follows from Lemma 13.3. \hfill\Box

Appendix A Deterministic Results about Contours and Clusters

In this section, we prove deterministic results concerning contours and clusters for the constrained percolation model on the [m,4,n,4][m,4,n,4] lattice in preparation to prove the main theorems.

Let PP be the underlying plane into which the [m,4,n,4][m,4,n,4] lattice is embedded. Recall that when 1m+1n=12\frac{1}{m}+\frac{1}{n}=\frac{1}{2}, PP is the Euclidean plane 2{\mathbb{R}}^{2} and the graph GG is amenable. When 1m+1n<12\frac{1}{m}+\frac{1}{n}<\frac{1}{2}, PP is the hyperbolic plane 2{\mathbb{H}}^{2} and the graph GG is non-amenable.

We consider an embedding of the [m,4,n,4][m,4,n,4] lattice into PP in such a way that each edge has length 1. Let ϕΦ\phi\in\Phi be a contour configuration, and let CC be a contour in ϕ\phi. To each component of PϕP\setminus\phi, we associate an interface, which is a closed set consisting of all the points in the component whose distance to CC is 14\frac{1}{4}. Here by distance, we mean either Euclidean distance or hyperbolic distance depending on whether PP is 2{\mathbb{R}}^{2} or 2{\mathbb{H}}^{2}. Obviously each interface is either a self-avoiding cycle or a doubly infinite self-avoiding walk. See Figure 13 for an example of interfaces on the [3,4,6,4][3,4,6,4] lattice.

Refer to caption
Figure 13. Contour configuration and interfaces. Blue lines represent contours of {\mathbb{H}}. Red lines represent contours of 𝕋{\mathbb{T}}. Green lines represent interfaces.

Note that when 1m+1n<12\frac{1}{m}+\frac{1}{n}<\frac{1}{2} and min{m,n}3\min\{m,n\}\geq 3, both 𝕃1{\mathbb{L}}_{1} and 𝕃2{\mathbb{L}}_{2} are vertex-transitive, non-amenable, planar graphs with one end.

Lemma A.1.

Whenever we have an interface II, let FIF_{I} be the set consisting of all the vertices of GG whose (Euclidean or hyperbolic) distance to the interface is 14\frac{1}{4}. Then all the vertices of FIF_{I} are in the same cluster. If II is a doubly-infinite self-avoiding path, then FIF_{I} is part of an infinite cluster.

Proof.

Recall that the interface II is either a self-avoiding cycle or a doubly-infinite self-avoiding walk. Give II a fixed direction. Moving along II following the fixed direction, let {Sj}jJ\{S_{j}\}_{j\in J} be the set of faces crossed by II in order, where JJ\subseteq{\mathbb{Z}} is a set of integers, such that for any j1<j2j_{1}<j_{2}, j1,j2Jj_{1},j_{2}\in J, II crosses Sj1S_{j_{1}} first, and then crosses Sj2S_{j_{2}}. Note that it is possible to have Sj1=Sj2S_{j_{1}}=S_{j_{2}} for j1j2j_{1}\neq j_{2}.

For any two vertices u,vFIu,v\in F_{I}, we can find a sequence of indices j1<j2<<jkJj_{1}<j_{2}<\ldots<j_{k}\in J, and a sequence of vertices of GG,

(49) u=vj1,1,vj1,2(=vj2,1),vj2,2(=vj3,1)==vjk,2=v,\displaystyle u=v_{j_{1},1},v_{j_{1},2}(=v_{j_{2},1}),v_{j_{2},2}(=v_{j_{3},1})=\ldots=v_{j_{k},2}=v,

such that for any 1ik1\leq i\leq k, vji,1v_{j_{i},1} and vji,2v_{j_{i},2} are two vertices (which may not be distinct) in FISjiF_{I}\cap\partial S_{j_{i}}, and there exists a path vji,1vji,2\ell_{v_{j_{i},1}v_{j_{i},2}} connecting vji,1v_{j_{i},1} and vji,2Sjiv_{j_{i},2}\subset\partial S_{j_{i}}, and vji,1vji,2I=\ell_{v_{j_{i},1}v_{j_{i},2}}\cap I=\emptyset. Note that any two consecutive vertices in (49) are in the same cluster, therefore uu and vv are in the same cluster.

If II is a doubly-infinite self-avoiding path, then II crosses infinitely many faces. Each face crossed by II has at least one boundary vertex in FIF_{I}. Each vertex of FIF_{I} is a boundary vertex of at most 4 faces. Therefore |FI|=|F_{I}|=\infty. Since all the vertices in FIF_{I} are in the same cluster, FIF_{I} is part of an infinite cluster. ∎

In the following lemma, contours may be primal or dual as usual.

Lemma A.2.

If there exist at least two infinite contours, then there exists an infinite 0-cluster or an infinite 1-cluster. Moreover, if C1C_{1} and C2C_{2} are two infinite contours, then there exists an infinite cluster incident to C1C_{1}.

Proof.

If there exist at least two infinite contours, then we can find two distinct infinite contours C1C_{1} and C2C_{2}, two points xC1x\in C_{1} and yC2y\in C_{2} and a self-avoiding path pxyp_{xy}, consisting of edges of GG and two half-edges, one starting at xx and the other ending at yy, and connecting xx and yy, such that pxyp_{xy} does not intersect any infinite contours except at xx and at yy. Indeed, we may take any path intersecting two distinct contours, and then take a minimal subpath with this property.

Let vVv\in V be the first vertex of GG along pxyp_{xy} starting from xx. Let uu be the point along the line segment [v,x][v,x] lying on an interface of C1C_{1}. Let u\ell_{u} be the interface of C1C_{1} containing uu. Then u\ell_{u} is either a doubly-infinite self-avoiding path or a self-avoiding cycle.

We consider these two cases separately. Firstly, if u\ell_{u} is a doubly-infinite self-avoiding path, then we claim that vv is in an infinite (0 or 1-)cluster of the constrained site configuration on GG. Indeed, this follows from Lemma A.1.

Secondly, if u\ell_{u} is a self-avoiding cycle, then PuP\setminus\ell_{u} has two components, QvQ_{v} and QvQ_{v}^{\prime}, where QvQ_{v} is the component including vv. Since u\ell_{u} is a cycle, exactly one of QvQ_{v} and QvQ_{v}^{\prime} is bounded, the other is unbounded. Since C1QvC_{1}\subseteq Q_{v}^{\prime}, and C1C_{1} is an infinite contour, we deduce that QvQ_{v}^{\prime} is unbounded, and QvQ_{v} is bounded. Since yuy\notin\ell_{u}, either yQvy\in Q_{v}, or yQvy\in Q_{v}^{\prime}. If yQvy\in Q_{v}^{\prime}, then any path, consisting of edges of GG and one half-edge incident to yy, connecting vv and yy must cross C1C_{1}. In particular, pxyp_{xy} crosses C1C_{1} not only at xx, but also at some point other than xx. This contradicts the definition of pxyp_{xy}. Hence yQvy\in Q_{v}. Since C1C2=C_{1}\cap C_{2}=\emptyset, this implies C2QvC_{2}\subseteq Q_{v}; because if C2QvC_{2}\cap Q_{v}^{\prime}\neq\emptyset, then C2C1C_{2}\cap C_{1}\neq\emptyset. But C2QvC_{2}\subseteq Q_{v} is impossible since C2C_{2} is infinite and QvQ_{v} is bounded. Hence this second case is impossible.

Therefore we conclude that if there exist at least two infinite contours, then there exists an infinite (0 or 1)-cluster. ∎

Lemma A.3.

Let xVx\in V be in the infinite 0-cluster, let yVy\in V be in the infinite 1-cluster, and let xy\ell_{xy} be a path, consisting of edges of GG and connecting xx and yy. Then xy\ell_{xy} has an odd number of crossings with infinite contours in total.

In particular, if there exist both an infinite 0-cluster and an infinite 1-cluster in a constrained percolation configuration ωΩ\omega\in\Omega, then there exists an infinite contour in ϕ(ω)Φ\phi(\omega)\in\Phi.

Proof.

Same as Lemma 2.8 of [27]. ∎

Lemma A.4.

Let CC_{\infty} be an infinite contour. Then each infinite component of GCG\setminus C_{\infty} contains an infinite cluster that is incident to CC_{\infty}.

Proof.

The lemma can be proved using similar technique as in Lemma 2.7 of [27]. ∎

Lemma A.5.

Let ωΩ\omega\in\Omega. Assume that there is exactly one infinite 0-cluster and exactly one infinite 1-cluster in ω\omega. Assume that there exist a vertex xx in the infinite 0-cluster, a vertex yy in the infinite 1-cluster, and a path xy\ell_{xy}, consisting of edges of GG and joining xx and yy, such that xy\ell_{xy} crosses exactly one infinite contour, CC_{\infty}. Then CC_{\infty} is incident to both the infinite 0-cluster and the infinite 1-cluster.

Proof.

By Lemma A.4, there is an infinite cluster in each infinite component of GCG\setminus C_{\infty}. Since there are exactly two infinite clusters, GCG\setminus C_{\infty} has at most 2 infinite components. Since each infinite cluster lies in some infinite component of GCG\setminus C_{\infty}, the number of infinite components of GCG\setminus C_{\infty} is at least one.

If GCG\setminus C_{\infty} has exactly two infinite components, then we can construct two infinite connected set of vertices in the two infinite components of GCG\setminus C_{\infty}, as a consequence of Lemma A.4, denoted by W1W_{1} and W2W_{2}, such that CC_{\infty} is incident to both W1W_{1} and W2W_{2}. Moreover, W1W_{1} and W2W_{2} are exactly part of the infinite 0-cluster and part of the infinite 1-cluster. Therefore CC_{\infty} is incident to both the infinite 0-cluster and the infinite 1-cluster.

If GCG\setminus C_{\infty} has exactly one infinite component, denoted by RR, then both the infinite 0-cluster and the infinite 1-cluster lie in RR, and in particular x,yRx,y\in R. We can find a path xy\ell^{\prime}_{xy}, connecting xx and yy, using edges of GG, such that xy\ell^{\prime}_{xy} does not cross CC_{\infty} at all. We can change path from xy\ell^{\prime}_{xy} to xy\ell_{xy} by choosing finitely many faces S1S_{1},S2S_{2},…, SkS_{k} of GG; along the boundary of each face, make every present edge in the path absent and every absent edge in the path present; and we perform this procedure for S1,,SkS_{1},\ldots,S_{k} one by one. Such a the path modification procedure does not change the parity of the number of crossings of the path with CC_{\infty}. Hence we infer that xy\ell_{xy} intersects CC_{\infty} an even number of times. But this is a contradiction to Lemma A.3 which says xy\ell_{xy} crosses CC_{\infty} an odd number of times, since we assume that xy\ell_{xy} crosses exactly one infinite contour CC_{\infty}. ∎

Lemma A.6.

Assume that ξ\xi is an infinite cluster, and CC is an infinite contour. Assume that xx is a vertex of GG in ξ\xi, and let yCy\in C be the midpoint of an edge of GG. Assume that there exists a path pxyp_{xy} connecting xx and yy, consisting of edges of GG and a half-edge incident to yy, such that pxyp_{xy} crosses no infinite contours except at yy. Let zz be the first vertex of GG along pxyp_{xy} starting from yy. Then zξz\in\xi.

Proof.

The proof is an adaptation of the proof of Lemma 2.9 in [27]. Lemma 2.9 in [27] applies when the graph GG is a square grid embedded in the Euclidean plane 2{\mathbb{R}}^{2}. This lemma applies when the graph GG is a general [m,4,n,4][m,4,n,4] lattice embedded in either the Euclidean plane or the hyperbolic plane.

Since pxyp_{xy} crosses no infinite contours except at yy, let C1,,CmC_{1},\ldots,C_{m} be all the finite contours crossing pxyp_{xy}. We claim that Pi=1mCiP\setminus\cup_{i=1}^{m}C_{i} has a unique unbounded component, which contains both xx and yy. Indeed, since xξx\in\xi and yCy\in C; neither the infinite cluster ξ\xi nor the infinite contour CC can lie in a bounded component of Pi=1mCiP\setminus\cup_{i=1}^{m}C_{i}.

Let II be the intersection of the union of the interfaces of C1,,CmC_{1},\ldots,C_{m} with the unique unbounded component of Pi=1mCiP\setminus\cup_{i=1}^{m}C_{i}. Since each CiC_{i}, 1im1\leq i\leq m, is a finite contour, each interface of CiC_{i} is finite. In particular, II consists of finitely many disjoint self-avoiding cycles, denoted by D1,,DtD_{1},\ldots,D_{t}. For 1it1\leq i\leq t, PDiP\setminus D_{i} has exactly one unbounded component, and one bounded component. Moreover, for iji\neq j, DiD_{i} and DjD_{j} come from interfaces of distinct contours.

Let BiB_{i} be the bounded component of PDiP\setminus D_{i}. We claim that each BiB_{i} is simply-connected, and BiBj=B_{i}\cap B_{j}=\emptyset, for iji\neq j. Indeed, BiB_{i} is simply connected, since the boundary of BiB_{i}, DiD_{i} is a self-avoiding cycle, whose embedding on the plane is a simple closed curve, for 1it1\leq i\leq t. Let 1i<jt1\leq i<j\leq t. Since DiD_{i} and DjD_{j} are disjoint, either BiBj=B_{i}\cap B_{j}=\emptyset, or one of BiB_{i} and BjB_{j} is a proper subset of the other. Without loss of generality, assume BiB_{i} is a proper subset of BjB_{j}. Then DiD_{i} is a proper subset of BjB_{j}. Hence DiD_{i} is in a bounded component of Pk=1mCkP\setminus\cup_{k=1}^{m}C_{k}, which contradicts the definition of DiD_{i}.

Let RiR_{i} be the set of faces FF of GG, for which BiFB_{i}\cap F\neq\emptyset. Let B~i=FRiF\widetilde{B}_{i}=\cup_{F\in R_{i}}F. Note that for 1it1\leq i\leq t, each B~i\widetilde{B}_{i} is a simply-connected, closed set. Let BiB_{i}^{\prime} be the interior of B~i\widetilde{B}_{i}. Then each BiB_{i}^{\prime} is a simply-connected, open set; moreover, BiBj=B_{i}^{\prime}\cap B_{j}^{\prime}=\emptyset, if iji\neq j. This follows from the fact that for iji\neq j, DiD_{i} and DjD_{j} come from interfaces of distinct contours, and the fact that BiBj=B_{i}\cap B_{j}=\emptyset, for iji\neq j.

Let B=i=1tBiB^{\prime}=\cup_{i=1}^{t}B^{\prime}_{i}. Then BB^{\prime} is open, and x,y,zPBx,y,z\in P\setminus B^{\prime}, although xx and zz may be on the boundary of BB^{\prime}.

There is a path pxy[pxy(PB)]Bp_{xy}^{\prime}\subseteq[p_{xy}\cap(P\setminus B^{\prime})]\cup\partial B^{\prime}, connecting xx and yy, where B\partial B^{\prime} is the boundary of BB^{\prime}. More precisely, pxyp_{xy} is divided by B\partial B^{\prime} into segments; on each segment of pxyp_{xy} in PBP\setminus B^{\prime}, pxyp^{\prime}_{xy} follows the path of pxyp_{xy}; for each segment of pxyp_{xy} in BB^{\prime}, pxyp^{\prime}_{xy} follows the boundary of BB^{\prime} to connect the two endpoints of the segment. This is possible since BB^{\prime} consists of bounded, disjoint, simply-connected, open sets BiB_{i}^{\prime}, for 1it1\leq i\leq t, and both xx and vv are in the complement of BB^{\prime} in PP.

All the vertices along pxyp^{\prime}_{xy} are in the same cluster. In particular, this implies that xx and zz are in the same infinite cluster ξ\xi. ∎

Lemma A.7.

If there exist exactly two infinite contours, then there exists an infinite cluster incident to both infinite contours.

Proof.

Let C1C_{1}, C2C_{2} be the two infinite contours. Since there exist only two infinite contours, we can find two points xC1x\in C_{1}, yC2y\in C_{2}, and a self-avoiding path pxyp_{xy}, consisting of edges of GG and two half-edges, one starting at xx and the other ending at yy, and connecting xx and yy, such that pxyp_{xy} does not intersect infinite contours except at xx and at yy.

Let vVv\in V be the first vertex of GG along pxyp_{xy} starting from xx. By the proof of Lemma A.2, vv is in an infinite cluster ξ\xi incident to C1C_{1}. Let zz be the first vertex of GG along pxyp_{xy}. By Lemma A.6, zξz\in\xi, and therefore ξ\xi is an infinite cluster incident to both C1C_{1} and C2C_{2}. ∎

Lemma A.8.

Let ωΩ\omega\in\Omega. If there is exactly one infinite 0-cluster and exactly one infinite 1-cluster in ω\omega, then there exists an infinite contour that is incident to both the infinite 0-cluster and the infinite 1-cluster in ω\omega.

Proof.

Let xx be a vertex in the infinite 0-cluster, and let yy be a vertex in the infinite 1-cluster. Let xy\ell_{xy} be a path joining xx and yy and consisting of edges of GG.

By Lemma A.3, xy\ell_{xy} must cross infinite contours an odd number of times. By Lemma A.5, if xy\ell_{xy} crosses exactly one infinite contour, CC_{\infty}, then CC_{\infty} is incident to both the infinite 0-cluster and the infinite 1-cluster, and so the lemma is proved in this case.

Suppose that there exist more than one infinite contour crossing xy\ell_{xy}. Let C1C_{1} and C2C_{2} be two distinct infinite contours crossing xy\ell_{xy}.

Let uC1xyu\in C_{1}\cap\ell_{xy} and vC2xyv\in C_{2}\cap\ell_{xy} (Here we interpret the contours and the paths as their embeddings to PP, so that u,vu,v are points in PP), such that the portion of xy\ell_{xy} between uu and vv, puvp_{uv}, does not cross any infinite contours except at uu and at vv. As in the proof of Lemma A.2, let u1u_{1} be the first vertex of GG along puvp_{uv}, starting from uu; and let v1v_{1} be the first vertex of GG along puvp_{uv} starting from vv. Let u2u_{2} (resp. v2v_{2}) be the point along the line segment [u,u1][u,u_{1}] (resp. [v,v1][v,v_{1}]) lying on an interface. Following the procedure in the proof of Lemma A.2, we can find an infinite cluster ξ1\xi_{1}, such that u1ξ1u_{1}\in\xi_{1}. The following cases might happen:

  1. I

    xξ1x\notin\xi_{1} and yξ1y\notin\xi_{1};

  2. II

    xξ1x\notin\xi_{1} and yξ1y\in\xi_{1};

  3. III

    xξ1x\in\xi_{1} and yξ1y\notin\xi_{1};

  4. IV

    xξ1x\in\xi_{1} and yξ1y\in\xi_{1}.

First of all, Case IV is impossible because we assume xx and yy are in two distinct infinite clusters. Secondly, if Case I is true, then there exist at least three infinite clusters, which is a contradiction to our assumption.

Case II and Case III can be handled using similar arguments, and we write down the proof of Case II here.

If Case II is true, first note that yξ1y\in\xi_{1} implies that C1C_{1} is incident to the infinite 1-cluster. Let zz be the first point in C1xyC_{1}\cap\ell_{xy} (again interpret edges as line segments), when traveling along xy\ell_{xy} starting from xx. Let pxzp_{xz} be the portion of xy\ell_{xy} between xx and zz.

Next, we will prove the following claim by induction on the number of complete edges of GG along pxzp_{xz} (in contrast to the half edge along pxzp_{xz} with an endpoint zz).

Claim A.9.

Under Case II, there is an infinite contour incident to both the infinite 0-cluster and the infinite 1-cluster.

Assume that the number of complete edges of GG along pxzp_{xz} is nn, where n=0,1,2,n=0,1,2,\ldots.

First of all, consider the case when n=0n=0. This implies that C1C_{1} is incident to the infinite 0-cluster at xx. Recall that C1C_{1} is also incident to the infinite 1-cluster at yy, and so A.9 is proved.

We make the following induction hypothesis:

  • Claim A.9 holds for nkn\leq k, where k0k\geq 0.

Now we consider the case when n=k+1n=k+1. The interior points of pxzp_{xz} are all points along pxzp_{xz} except xx and zz. We consider two cases:

  1. (a)

    at interior points, pxzp_{xz} crosses only finite contours but not infinite contours;

  2. (b)

    at interior points, pxzp_{xz} crosses infinite contours.

We claim that if Case (a) occurs, then C1C_{1} is incident to both the infinite 0-cluster and the infinite 1-cluster. It suffices to show that C1C_{1} is incident to the infinite 0-cluster.

Let z1z_{1} be the first vertex in VV along pxzp_{xz} starting from zz. According to Lemma A.6, both xx and z1z_{1} are in the infinite 0-cluster. We infer that C1C_{1} is incident to the infinite 0-cluster at xx, if pxzp_{xz} intersects only finite contours at interior points.

Now we consider Case (b). Let C3C_{3} be an infinite contour crossing pxzp_{xz} at interior points. Obviously, C3C_{3} and C1C_{1} are distinct, because C1C_{1} crosses pxzp_{xz} only at zz. Let ww be the last point in C3pxzC_{3}\cap p_{xz}, when traveling along pxzp_{xz}, starting from xx, and let pwzp_{wz} be the portion of pxzp_{xz} between ww and zz. Assume pwzp_{wz} does not cross infinite contours at interior points.

Let w1w_{1} be the first vertex of GG along pwzp_{wz}, starting from ww, and let w2w_{2} be the midpoint of ww and w1w_{1}. According to the proof of Lemma A.2, we can find an infinite cluster ξ3\xi_{3} including w1w_{1}. The following cases might happen:

  1. i

    xξ3x\notin\xi_{3}, and yξ3y\notin\xi_{3};

  2. ii

    xξ3x\in\xi_{3}, and yξ3y\notin\xi_{3};

  3. iii

    xξ3x\notin\xi_{3}, and yξ3y\in\xi_{3};

  4. iv

    xξ3x\in\xi_{3}, and yξ3y\in\xi_{3}.

First of all, Case iv is impossible because we assume xx and yy are in two distinct infinite clusters. Secondly, if Case i is true, then there exist at least three infinite clusters, which is a contradiction to the assumption that there exists exactly one infinite 0-cluster and one infinite 1-cluster.

If Case ii is true, then C3C_{3} is incident to the infinite 0-cluster including xx. Since w1ξ3w_{1}\in\xi_{3}, and pwzp_{wz} does not cross infinite contours except at ww and zz, by Lemma A.6, we infer that zξ3z\in\xi_{3}, and ξ3\xi_{3} is exactly the infinite 0-cluster including xx. We conclude that C1C_{1} is incident to the infinite 0-cluster including xx as well, and Claim A.9 is proved.

If Case iii is true, then C3C_{3} is incident to the infinite 1-cluster including yy. Let tt be the first vertex in pxzC3p_{xz}\cap C_{3}, when traveling from pxzp_{xz}, starting at xx, and let pxtp_{xt} be the portion of pxzp_{xz} between xx and tt. We explore the path pxtp_{xt} as we have done for pxzp_{xz}. Since the length of pxzp_{xz} is finite, and the number of full edges of GG along pxtp_{xt} is less than that of pxzp_{xz} by at least 1, we apply the induction hypothesis with C1C_{1} replaced by C3C_{3}, C2C_{2} replaced by C1C_{1}, ξ1\xi_{1} replaced by ξ3\xi_{3}, pxzp_{xz} replaced by pxtp_{xt}, and we conclude that there exists an infinite contour adjacent to both the infinite 0-cluster and infinite 1-cluster. ∎

Lemma A.10.

Let C1C_{1} and C2C_{2} be two infinite contours, and let ξ0\xi_{0} and ξ1\xi_{1} be two infinite clusters. The following two cases cannot occur simultaneously.

  • ξ0\xi_{0} is incident to both C1C_{1} and C2C_{2};

  • ξ1\xi_{1} is incident to both C1C_{1} and C2C_{2}.

Proof.

Same arguments as Lemma 6.3 of [27]. ∎

Lemma A.11.

Let 𝕃2{\mathbb{L}}_{2} be the regular tiling of the hyperbolic plane with triangles, such that each vertex has degree n7n\geq 7. Let ω{0,1}V(𝕃2)\omega\in\{0,1\}^{V({\mathbb{L}}_{2})} be a site percolation configuration on 𝕃2{\mathbb{L}}_{2}. Assume that there exists an infinite 1-cluster ηω\eta\subseteq\omega with infinitely many ends. Then there exist at least two infinite 0-clusters in ω\omega.

Proof.

Since ηω\eta\subseteq\omega has infinitely many ends, there exists a finite box BB of 𝕃1{\mathbb{L}}_{1}, such that ηB\eta\setminus B has at least two distinct infinite components. Let XX, YY be two distinct infinite components of ηB\eta\setminus B. Define the boundary of XX (resp. Y), X\partial X (resp. Y\partial Y) to be the set of all edges in 𝕃1{\mathbb{L}}_{1}, such that each dual edge has exactly one endpoint in XX (resp. Y), and one endpoint in 𝕃2B{\mathbb{L}}_{2}\setminus B and not in XX (resp. not in YY). Then X\partial X and Y\partial Y are part of contours - each vertex of 𝕃1{\mathbb{L}}_{1} in X\partial X and Y\partial Y has degree 1 or 2, whose degree-1 vertices are along B\partial B (here B\partial B consists of all the edges of 𝕃1{\mathbb{L}}_{1} on the boundary of the finite box BB).

Each component of X\partial X or Y\partial Y (here we assume that points on B\partial B are not included in X\partial X or Y\partial Y) must be one of the following three cases:

  1. I

    a finite component; or

  2. II

    a doubly infinite self-avoiding path which does not intersect B\partial B;

  3. III

    a singly-infinite self-avoiding path starting from a vertex along B\partial B.

Let B1BB_{1}\supset B be a box of 𝕃2{\mathbb{L}}_{2} containing BB. Then the embedding of B1\partial B_{1} into 2{\mathbb{H}}^{2} is a simple closed curve consisting of edges in 𝕃2{\mathbb{L}}_{2}. Since XX and YY are two infinite components of ηB\eta\setminus B, we deduce that B1X\partial B_{1}\cap X\neq\emptyset and B1Y\partial B_{1}\cap Y\neq\emptyset. Since B1\partial B_{1} is a closed curve, there exist x1,x2B1Xx_{1},x_{2}\in\partial B_{1}\cap X and y1,y2B1Yy_{1},y_{2}\in\partial B_{1}\cap Y, such that there are segments px1y1B1p_{x_{1}y_{1}}\subset\partial B_{1} joining x1x_{1} and y1y_{1}, and px2y2B1p_{x_{2}y_{2}}\subset\partial B_{1} joining x2x_{2} and y2y_{2} such that px1y1p_{x_{1}y_{1}} and px2y2p_{x_{2}y_{2}} does not intersect each other except possibly at x1,x2,y1,y2x_{1},x_{2},y_{1},y_{2}; px1y1p_{x_{1}y_{1}} does not intersect XYX\cup Y except at x1,y1x_{1},y_{1}; and that px2y2p_{x_{2}y_{2}} does not intersect XYX\cup Y except at x2,y2x_{2},y_{2}.

Then we claim that both px1y1p_{x_{1}y_{1}} and px2y2p_{x_{2}y_{2}} cross a component of X\partial X of Type III, and a component of Y\partial Y of Type III. To see why that is true, assume that px1y1p_{x_{1}y_{1}} crosses only components of X\partial X of Type I or II, then we can find a path qx1y1q_{x_{1}y_{1}} consisting of edges in 𝕃2B{\mathbb{L}}_{2}\setminus B and joining x1x_{1} and y1y_{1} such that qx1y1q_{x_{1}y_{1}} does not cross X\partial X at all. Then YY and XX are the same component of ηB\eta\setminus B. The contradiction implies that px1y1p_{x_{1}y_{1}} must cross a component of X\partial X of Type III. Similarly, px1y1p_{x_{1}y_{1}} must cross a component of Y\partial Y of Type III; px2y2p_{x_{2}y_{2}} must cross a component of X\partial X of Type III, and a component of Y\partial Y of Type III.

Let 1\ell_{1} (resp. 2\ell_{2}) be a component of X\partial X of Type III crossed by px1y1p_{x_{1}y_{1}} (resp. px2y2p_{x_{2}y_{2}}). Let V1V_{1} (resp. V2V_{2}) consist of all the vertices of 𝕃2{\mathbb{L}}_{2} on a triangle face crossed by 1\ell_{1} (resp. 2\ell_{2}) but not in XX. Then V1V_{1} is part of an infinite 0-cluster ξ\xi, and V2V_{2} is part of an infinite 0-cluster ζ\zeta; and moreover, ξ\xi and ζ\zeta are distinct since both px1y1p_{x_{1}y_{1}} and px2y2p_{x_{2}y_{2}} also cross components of Y\partial Y of Type III. This completes the proof. ∎

Lemma A.12.

Let G=(V,E)G=(V,E) be a square tiling of the hyperbolic plane satisfying I and II. Let ωΩ\omega\in\Omega be a constrained percolation configuration on GG. If there exists a contour in the corresponding contour configuration of ω\omega, then there exist at least one infinite 1-cluster and at least one infinite 0-cluster in ω\omega.

Proof.

Let CC be a contour in the corresponding contour configurations. By Lemma 6.5, CC is an infinite tree in which each vertex is incident to 2 or 4 edges. Since CC has no cycles, the complement 2C{\mathbb{H}}^{2}\setminus C of CC in the hyperbolic plane 2{\mathbb{H}}^{2} has no bounded components.

We claim that each unbounded component of 2C{\mathbb{H}}^{2}\setminus C contains at least one infinite cluster. Let Λ\Lambda be an unbounded component of 2C{\mathbb{H}}^{2}\setminus C. Let VΛ,CVV_{\Lambda,C}\subset V consist of all the vertices in Λ\Lambda that are also in a face of GG intersecting CC. Then all the vertices in VΛ,CV_{\Lambda,C} are in the same cluster of ω\omega and |VΛ,C|=|V_{\Lambda,C}|=\infty. Hence there exists an infinite cluster in ω\omega containing every vertex in VΛ,CV_{\Lambda,C}.

Let eCe\in C be an edge crossing a pair of opposite edges of a black face bb of GG. Let v1,v2,v3,v4v_{1},v_{2},v_{3},v_{4} be the 4 vertices of bb. Assume that v1v_{1} and v2v_{2} are on one side of ee while v3v_{3} and v4v_{4} are on the other side of ee. By the arguments above, v1v_{1} and v2v_{2} are in an infinite cluster ξ1\xi_{1} of ω\omega; similarly, v3v_{3} and v4v_{4} are in an infinite cluster ξ2\xi_{2} of ω\omega. Moreover, eCe\in C implies that v1v_{1} and v3v_{3} have different state; and therefore exactly one of ξ1\xi_{1} and ξ2\xi_{2} is an infinite 0-cluster, and the other is an infinite 1-cluster. ∎

Lemma A.13.

Let G=(V,E)G=(V,E) be a square tiling of the hyperbolic plane satisfying I and II. Let ωΩ\omega\in\Omega be a constrained percolation configuration on GG, and let ϕ\phi be the corresponding contour configuration of ω\omega. Then each component of 2ϕ{\mathbb{H}}^{2}\setminus\phi contains an infinite cluster in ω\omega.

Proof.

By Lemma 6.5, ϕ\phi is the disjoint union of infinite trees, in which each vertex has degree 2 or 4. Since ϕ\phi contains no cycles, each component of 2ϕ{\mathbb{H}}^{2}\setminus\phi is unbounded.

Let Λ\Lambda be an unbounded component of 2ϕ{\mathbb{H}}^{2}\setminus\phi. Let VΛ,ϕVV_{\Lambda,\phi}\subset V consist of all the vertices in Λ\Lambda that are also in a face of GG intersecting ϕ\phi. Then all the vertices in VΛ,ϕV_{\Lambda,\phi} are in the same cluster of ω\omega and |VΛ,ϕ|=|V_{\Lambda,\phi}|=\infty. Hence there exists an infinite cluster in ω\omega containing every vertex in VΛ,ϕV_{\Lambda,\phi}. Since every vertex in VΛ,ϕV_{\Lambda,\phi} is in Λ\Lambda, and any cluster intersecting Λ\Lambda is completely in Λ\Lambda, we conclude that Λ\Lambda contains an infinite cluster of ω\omega. ∎


Acknowledgements. ZL thanks Alexander Holroyd for stimulating discussions in the preparation of the paper, and Geoffrey Grimmett and Russ Lyons for comments. ZL is grateful for anonymous reviewers’ careful reading of the paper and valuable suggestions to improve the readability. ZL’s research is supported by Simons Foundation grant 351813 and National Science Foundation grant 1608896.

References

  • [1] M. Aizenman, Translation invariance and instability of phase coexistence in the two-dimensional Ising system, Comm. Math. Phys. 73 (1980), 83–94.
  • [2] R. J. Baxter, Exactly solved models in statistical mechanics, Academic Press, 2008.
  • [3] I. Benjamini, R. Lyons, Y. Peres, and O. Schramm, Critical percolation on any non-amenable group has no infinite clusters, Ann. Probab. 27 (1999), 1347–1356.
  • [4] by same author, Group-invariant percolation on graphs, Geom. funct. anal. 9 (1999), 29–66.
  • [5] I. Benjamini and O. Schramm, Percolation beyond d\mathbb{Z}^{d} many questions and a few answers, Electron. Commun. Probab. 1 (1996), 71–82.
  • [6] by same author, Percolation in the hyperbolic plane, Journal of the American Mathematical Society 14 (2000), 487–507.
  • [7] B. Bollobás and O. Riordan, Percolation on dual lattices with k-fold symmetry, Random Structures and Algorithms 32 (2008), 463–472.
  • [8] A. Borodin, I. Corwin, and V. Gorin, Stochastic six-vertex model, Duke Math. J. 165 (2016), 563–624.
  • [9] C. Boutillier and B. de Tilière, Height representation of XOR-Ising loops via bipartite dimers, Electron. J. Probab. 19 (2014), 1–33.
  • [10] R.M. Burton and M. Keane, Density and uniqueness in percolation, Commun. Math. Phys. 121 (1989), 501–505.
  • [11] J.W. Cannon, W. J. Floyd, R. Kenyon, and Parry W.R., Hyperbolic geometry, Flavors of Geometry (Math. Sci. Res. Inst. Publ., 31), Cambridge University Press, Cambridge, 1997, pp. 59–115.
  • [12] A. Coniglio, CR. Nappi, F. Peruggi, and R. Russo, Percolation and phase transitions in the Ising model, Commun. Math. Phys. 51 (1976), 315–323.
  • [13] J. Dubédat, Exact bosonization of the Ising model, arXiv:1112.4399.
  • [14] H. Duminil-Copin and D. Cimasoni, The critical temperature for the Ising model on planar doubly periodic graphs, Electron. J. Probab. 44 (2013), 1–18.
  • [15] R. Fitzner and R. van der Hofstad, Mean-field behavior for nearest-neighbor percolation in d>10d>10, arXiv:1506.07977v2.
  • [16] G. Grimmett, Percolation, Springer, 1999.
  • [17] by same author, The random cluster model, Springer-Verlag, 2006.
  • [18] G. Grimmett and Z. Li, The 1-2 model, Contemporary Mathematics 696 (2017), 139–152.
  • [19] by same author, Critical surface of the 1-2 model, International Mathematics Research Notices 21 (2018), 6617–6672.
  • [20] O. Häggstöm, J. Jonasson, and R. Lyons, Explicit isoperimetric constants and phase-transitions in the random cluster model, Ann. Probab. 30 (2002), 443–473.
  • [21] O. Häggstöm and Y. Peres, Monotonicity of uniqueness for percolation on cayley graphs: All infinite clusters are born simultaneously, Probab. Theory and Related Fields 113 (1999), 273–285.
  • [22] O. Häggstöm, Y. Peres, and R.H. Schonmann, Percolation on transitive graphs as a coalescent process: relentless merging followed by simultaneous uniqueness, Perplexing Probability Problems :Papers in Honor of H. Kesten (M. Bramson and R. Durrett, eds.), Birkhäuster, Boston, 1999, pp. 69–90.
  • [23] T. Hara and G. Slade, Mean-field behavior and the lace expansion, Probability and Phase Transition (G. R. Grimmett, ed.), Kluwer, Dordrecht, 1994, pp. 87–122.
  • [24] T. E. Harris, A lower bound for the critical probability in a certain percolation process, Proc. Cambridge Philos. Soc. 56 (1960), 13–20.
  • [25] Y. Higuchi, A sharp transition for the two-dimensional Ising percolation, Probab. Theory Relat. Fields 97 (1993), 489–514.
  • [26] R. Holley, Remarks on the FKG inequalities, Commun. Math. Phys. 36 (1974), 227–231.
  • [27] A. Holroyd and Z. Li, Constrained percolation in two dimensions, ANNALES DE L’INSTITUT HENRI POINCARÉ D (2020), arXiv:1510.03943v2.
  • [28] R. Kenyon, Lectures on dimers, Statistical Mechanics IAS/Park City Math.Ser.,16 (2009), 191–230.
  • [29] R. Kenyon, J. Miller, S. Sheffield, and D.B. Wilson, The six-vertex model and Schramm-Loewner evolution, arXiv:1605.06471v1.
  • [30] R. Kenyon, A. Okounkov, and S. Sheffield, Dimers and ameoba, Ann. Math. 163 (2006), 1019–1056.
  • [31] H. Kesten, The critical probability of bond percolation on the square lattice equals 12\frac{1}{2}, Comm. Math. Phys. 74 (1980), 41–59.
  • [32] J. Lebowitz, On the uniqueness of the equilibrium state for Ising spin systems, Commun. Math. Phys. 25 (1972), 276–282.
  • [33] by same author, Coexistence of phases in Ising ferromagnets, J. Statist. Phys. 16 (1977), 453–461.
  • [34] Z. Li, Local statistics of realizable vertex models, Commun. Math. Phys. 304 (2011), 723–763.
  • [35] by same author, Critical temperature of periodic Ising models, Comm. Math. Phys. 315 (2012), 337–381.
  • [36] by same author, 1-2 model, dimers and clusters, Electr. J. Probab. 19 (2014), 1–28.
  • [37] by same author, Uniqueness of the infinite homogeneous cluster in the 1-2 model, Electron. Commun. Probab. 19 (2014), 1–8.
  • [38] R. Lyons and Y. Peres, Probability on trees and networks, Cambridge University Press, 2016, Available at http://pages.iu.edu/~rdlyons/.
  • [39] R. Lyons and O. Schramm, Indistinguishability of percolation clusters, Ann. Probab. 27 (1999), 1809–1836.
  • [40] A. Nachmias and Y. Peres, Non-amenable graphs of high girth have pc<pup_{c}<p_{u} and mean field exponents, Electron. Commun. Probab. 17 (2012), 1–8.
  • [41] C.M. Newman and L.S. Schulman, Infinite clusters in percolation models, J. Statist. Phys. 26 (1981), 613–628.
  • [42] I. Pak and T. Smirnova Nagnibeda, On non-uniqueness of percolation in non-amenable cayley graphs, C.R.Acad.Sci.Paris 330 (2000), 495–550.
  • [43] D. Renault, The vertex-transitive TLF-planar graphs, Discrete Mathematics 303 (2009), 2815–2833.
  • [44] R.H. Schonmann, Stability of infinite clusters in supercritical percolation, Probab. Theory and Related Fields 113 (1999), 287–300.
  • [45] by same author, Multiplicity of phase transitions and mean-field criticality on highly non-amenable graphs, Commun. Math. Phys. 219 (2001), 271–322.
  • [46] M. Schwartz and J. Bruck, Constrained codes as network of relations, Information theory, IEEE transactions (2008).
  • [47] A. Thom, A remark about the spectral radius, Int. Math. Res. Notices 10 (2015), 2856–2864.
  • [48] L. G. Valiant, Holographic algorithms (extended abstract), Proc. 45th IEEE Symposium on Foundations of Computer Science (2004), 306–315.
  • [49] D. B. Wilson, XOR-Ising loops and the Gaussian free field, arXiv:1102.3782.
  • [50] C. Wu, Ising model on hyperbolic graphs II, J Stats. Phys. 100 (2000), 893–904.