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

Topological Phases in the
Plaquette Random-Cluster Model
and Potts Lattice Gauge Theory

Paul Duncan [email protected] Einstein Institute of Mathematics, Hebrew University of Jerusalem, Jerusalem 91904, Israel  and  Benjamin Schweinhart [email protected] Department of Mathematical Sciences, George Mason University, Fairfax, VA 22030, USA
Abstract.

The ii-dimensional plaquette random-cluster model on a finite cubical complex is the random complex of ii-plaquettes with each configuration having probability proportional to

p# of plaquettes(1p)# of complementary plaquettesq𝐛i1,p^{\text{\# of plaquettes}}\left(1-p\right)^{\text{\# of complementary plaquettes}}q^{\mathbf{b}_{i-1}}\,,

where q1q\geq 1 is a real parameter and 𝐛i1\mathbf{b}_{i-1} denotes the rank of the (i1)(i-1)-homology group with coefficients in a specified coefficient field. When qq is prime and the coefficient field is 𝔽q\mathbb{F}_{q}, this model is coupled with the (i1)(i-1)-dimensional qq-state Potts lattice gauge theory. We prove that the probability that an (i1)(i-1)-cycle in d\mathbb{Z}^{d} is null-homologous in the plaquette random-cluster model equals the expectation of the corresponding generalized Wilson loop variable. This provides the first rigorous justification for a claim of Aizenman, Chayes, Chayes, Frölich, and Russo that there is an exact relationship between Wilson loop variables and the event that a loop is bounded by a surface in an interacting system of plaquettes. We also prove that the ii-dimensional plaquette random-cluster model on the 2i2i-dimensional torus exhibits a sharp phase transition at the self-dual point psdq1+qp_{\mathrm{sd}}\coloneqq\frac{\sqrt{q}}{1+\sqrt{q}} in the sense of homological percolation. This implies a qualitative change in the generalized Swendsen–Wang dynamics from local to non-local behavior.

P.D. gratefully acknowledges the support of NSF-DMS #1547357.

1. Introduction

The random-cluster model [FK72, Gri06], also known as the Fortuin–Kasteleyn model, is a random subgraph of a finite graph where the probability of a given configuration is proportional to

p# of edges(1p)# of complementary edgesq# of connected components.p^{\text{\# of edges}}\left(1-p\right)^{\text{\# of complementary edges}}q^{\text{\# of connected components}}.

This model can then be extended to infinite graphs via limits on finite subgraphs when q1.q\geq 1. The random cluster model is an important tool for the study of the Potts model of interacting spins on the vertices of a graph. Together, they are the subject of an expansive literature in mathematical physics, probability, and statistical physics. The relationship between the models is exemplified by the connection–correlation theorem, which expresses the correlation between the spins of two vertices in the Potts model in terms of the probability that they are in the same connected component in the corresponding random cluster model.

One-dimensional qq-state Potts lattice gauge theory is a higher-dimensional analogue of the Potts model, where spins are assigned to edges of a graph rather than to vertices. It was as introduced in the physics literature [Weg71, KPSS80] as a relatively approachable analogue of Euclidean lattice gauge theory. Lattice gauge theory is in turn a discretized models of Euclidean Yang–Mills theory [Wil74, Wil04, Cha16]. The most important observables for these models are the Wilson loop variables, which are the sum of the spins around a loop of edges (or, equivalently, the evaluation of the co-chain on an 11-cycle)111We view a Wilson loop variable as a sum rather than as a product to match the notation in algebraic topology. This point will be discussed more below.

In their celebrated paper on independent plaquette percolation, Aizenman, Chayes, Chayes, Frölich, and Russo posited the existence of an exact relationship between Wilson loop variables and the event that a loop is bounded by a surface in an interacting system of plaquettes, or random 22-complex [ACC+83], in some sense generalizing the connection-correlation theorem. Inspired by the random cluster model and its relationship with the Potts model, Maritan and Omero [MO82] defined such a random 22-complex associated to 11-dimensional Potts lattice gauge theory. Their goals were to to make rigorous the relationship between lattice gauge theory and the behavior of random surfaces and to interpret the q1q\rightarrow 1 limit of Potts lattice gauge theory in terms of surfaces of independent plaquettes (plaquette percolation). However, their model was defined vaguely in terms of closed surfaces of plaquettes rather than homology. This imprecision led to inconsistencies that were found by were found by Aizenman and Frölich  [AF84]: the Betti numbers depend on the coefficients222Note that 𝐛0\mathbf{b}_{0} counts the number of connected components regardless of the coefficient field..

In essence, Maritan and Omero defined a random 22-complex PP so

(P)p# of plaquettes(1p)# of complementary plaquettesq𝐛1(P;𝔽2),\mathbb{P}\left(P\right)\propto p^{\text{\# of plaquettes}}\left(1-p\right)^{\text{\# of complementary plaquettes}}q^{\mathbf{b}_{1}\left(P;\,\mathbb{F}_{2}\right)}\,,

where the first Betti number 𝐛1(P;𝔽2)\mathbf{b}_{1}\left(P;\,\mathbb{F}_{2}\right) counts the number of “independent loops” in P,P, and is related to the number of “independent closed surfaces” by the Euler–Poincaré formula. Upon close inspection, this model is not coupled with Potts lattice gauge theory when q2.q\neq 2. An earlier paper of Ginsparg, Goldschmidt, and Zuber [GGZ80] performed a series expansion for Potts lattice gauge theory which suggests this construction, and also an alternative model where only oriented surfaces are taken into account. This has the effect of replacing 𝐛1(P;𝔽2)\mathbf{b}_{1}\left(P;\,\mathbb{F}_{2}\right) with 𝐛1(P;),\mathbf{b}_{1}\left(P;\,\mathbb{Z}\right), yielding a random 22-complex that is not coupled with Potts lattice gauge theory for any choice of q.q. Nor is Potts lattice gauge theory coupled with a different generalized random-cluster model due to Chayes and Chayes [CC84] where 𝐛1(P;𝔽2)\mathbf{b}_{1}\left(P;\,\mathbb{F}_{2}\right) is replaced by the the number of strongly connected components of the set of plaquettes. We show here that the “correct” approach is to view the plaquette random-cluster of Hiraoka and Shirai [HS16] as a family with three parameters, namely pp, qq, and the coefficient field 𝔽.\mathbb{F}. Then, by setting 𝔽=𝔽q\mathbb{F}=\mathbb{F}_{q} and choosing pp appropriately, we can express Wilson loop variables in terms of the probability that the loop is bounded by a surface of plaquettes, where the coefficient field delimits the set of admissible surfaces. Moreover, fixing 𝔽q\mathbb{F}_{q} and taking the q1q\rightarrow 1 limit allows a comparison of qq-state Potts lattice gauge theory with plaquette percolation.

It is conjectured that the expectations of Wilson loop variables in Potts lattice gauge theory exhibit a sharp phase transition from a “perimeter law” to an “area law,” in which they decay exponentially in the perimeter and area of the loop respectively. Our results show that this is equivalent to a corresponding conjecture for the event that the loop is null-homologous in the plaquette random cluster model. While these conjectures remains out of reach, we are able to prove a qualitative analogue in terms of homological percolation on the torus. That is, we show that the ii-dimensional plaquette random-cluster model on the 2i2i-dimensional torus exhibits a sharp phase transition at the self-dual point psdq1+qp_{\mathrm{sd}}\coloneqq\frac{\sqrt{q}}{1+\sqrt{q}} marked by the the emergence of giant cycles which are non-trivial homology classes in the ambient torus. When qq is an odd prime, these cycles are ruled by Polyakov loops that are constant in the corresponding Potts lattice gauge theory. The critical point is alternatively characterized by a qualitative change in the generalized Swendsen–Wang dynamics from local to non-local behavior.

1.1. Definitions, Conjectures, and Results

The generalized random cluster model we study here was introduced by Hiraoka and Shirai [HS16]. Their model, which we call the ii-dimensional plaquette random-cluster model, weights the probability of an ii-complex PP in terms of the (i1)(i-1)-Betti number of the homology with coefficients in a field 𝔽,\mathbb{F}, denoted 𝐛i1(P;𝔽).\mathbf{b}_{i-1}\left(P;\,\mathbb{F}\right). We provide a definition of Betti numbers in Section 2 below.

Definition 1.

Let XX be a finite dd-dimensional cell complex, let 0<i<d,0<i<d, and fix a field 𝔽\mathbb{F} and parameters p[0,1]p\in\left[0,1\right] and q(0,).q\in\left(0,\infty\right). The ii-dimensional plaquette random-cluster model on XX is the random ii-complex PP containing the full (i1)\left(i-1\right)-skeleton of XX with the following distribution:

μX(P)=μX,p,q,i,𝔽(P)1Zp|P|(1p)|Xi||P|q𝐛i1(P;𝔽),\displaystyle\mu_{X}\left(P\right)=\mu_{X,p,q,i,\mathbb{F}}\left(P\right)\coloneqq\frac{1}{Z}p^{\left|P\right|}\left(1-p\right)^{\left|X^{i}\right|-\left|P\right|}q^{\mathbf{b}_{i-1}\left(P;\,\mathbb{F}\right)}\,,

where Z=Z(X,p,q,i,𝔽)Z=Z\left(X,p,q,i,\mathbb{F}\right) is a normalizing constant, and |Xi|\left|X^{i}\right| and |P|\left|P\right| denote the number of ii-cells of XX and P,P, respectively.

As 𝐛0\mathbf{b}_{0} is the number of connected components, this is indeed a generalization of the classical random-cluster model. In their paper introducing the plaquette random-cluster model, Hiraoka and Shirai’s main focus is proving an expression for the partition function in terms of a generalized Tutte polynomial. They also show positive association and construct a coupling between the model and a generalization of the Potts model. However, they do not mention that this generalized Potts model had been previously defined as Potts lattice gauge theory [HS16].

Generalized ii-dimensional Potts lattice gauge theory assigns spins in an abelian group 𝒢\mathcal{G} to the ii-dimensional faces of an oriented cell complex X.X. Denote by Ci(X;𝒢)C^{i}\left(X;\,\mathcal{G}\right) the set of functions from the oriented ii-cells of XX to 𝒢\mathcal{G} so that cells with opposite orientations are mapped to inverse group elements. We follow the language of algebraic topology and call elements of Ci(X;𝒢)C^{i}\left(X;\,\mathcal{G}\right) cochains. In other works on lattice gauge theory they are sometimes termed discrete differential forms, for reasons that we will describe in Section 2 below.

Definition 2.

Let 𝒢\mathcal{G} be a finite abelian group and let XX be a finite cubical complex. For fCi(X;𝒢)f\in C^{i}\left(X;\,\mathcal{G}\right) let

H(f)=σK(δf(σ),1),H\left(f\right)=-\sum_{\sigma}K\left(\delta f\left(\sigma\right),1\right)\,, (1)

where δ\delta is the coboundary operator δf(σ)=f(σ)\delta f\left(\sigma\right)=f\left(\partial\sigma\right) and KK is the Kronecker delta function. (i1)(i-1)-dimensional Potts lattice gauge theory on XX with coefficients in 𝒢\mathcal{G} is the measure

νX,β,𝒢,i1,d(f)=1𝒵eβH(f)\nu_{X,\beta,\mathcal{G},i-1,d}\left(f\right)=\frac{1}{\mathcal{Z}}e^{-\beta H\left(f\right)}

where β\beta is a parameter (called the coupling constant or inverse temperature) and 𝒵=𝒵(X,β,𝒢,i1,d)\mathcal{Z}=\mathcal{Z}\left(X,\beta,\mathcal{G},i-1,d\right) is a normalizing constant.

In this paper, we focus exclusively on the cases where 𝒢=𝔽q\mathcal{G}=\mathbb{F}_{q} is the additive group of integers modulo a prime number qq (which can be identified with the multiplicative group of qq-th complex roots of unity (q)\mathbb{Z}\left(q\right)). In keeping with conventions from algebraic topology, we use additive notation rather than the multiplicative notation of the definition of Potts lattice gauge theory. In particular, we will distinguish between the additive group of integers modulo nn n\mathbb{Z}_{n} and the multiplicative group of complex nn-th roots of unity (n),\mathbb{Z}(n), though they are isomorphic. In the case where 𝒢=𝔽q,\mathcal{G}=\mathbb{F}_{q}, we will denote the Potts lattice gauge theory measure by νβ,q,i1,d,\nu_{\beta,q,i-1,d}, or simply ν,\nu, and call this measure qq-state Potts lattice gauge theory. We describe why the assumption that qq is prime is necessary in Section 1.2. We will be most interested in the Potts lattice gauge theory on d,\mathbb{Z}^{d}, and we will write ν𝐟=νd,β,𝒢,i1,d𝐟\nu^{\mathbf{f}}=\nu^{\mathbf{f}}_{\mathbb{Z}^{d},\beta,\mathcal{G},i-1,d} to denote the limiting Potts lattice gauge theory with free boundary conditions constructed in Section 5.2.

Potts lattice gauge theory is motivated by its similarities with Euclidean lattice gauge theory. Lattice gauge theory is in turn a discretized model of Euclidean Yang–Mills theory [Wil74, Wil04, Cha16]. Very briefly, Euclidean lattice gauge theory is defined so that 𝔽q\mathbb{F}_{q} can be replaced with a complex matrix group 𝒢.\mathcal{G}. In particular, 11-dimensional lattice gauge theory on 4\mathbb{Z}^{4} with 𝒢=U(1),\mathcal{G}=U(1), 𝒢=SU(2),\mathcal{G}=SU(2), and 𝒢=SU(3)\mathcal{G}=SU(3), models the electromagnetic, weak nuclear, and strong nuclear forces, respectively. While Potts lattice gauge theory is not itself physical, it has been studied in the physics literature as it is more tractable and is thought to present some of the same behavior as more physically relevant cases [GGZ80, KS82, MO82, AF84, LMR89]. The special cases of 22 and 33-state Potts lattice gauge theory coincide with (2)\mathbb{Z}(2) (Ising) and (3)\mathbb{Z}(3) “clock” lattice gauge theory respectively after an appropriate rescaling of the coupling constant β\beta (where (n)\mathbb{Z}(n) denotes the multiplicative group of nn-th complex roots of unity). These models have also been studied in the physical and the mathematical literatures as relatively approachable examples of lattice gauge theory [Alt78, Yon78, Frö79, MP79, MMS79, KPSS80, WHK+07, Weg17, Cha20].

The most important observables in lattice gauge theory are arguably Wilson loop variables, which we define for qq-state Potts lattice gauge theory. These arise from evaluating a Potts state ff on a “loop” γ\gamma made of (i1)(i-1)-cells. In topological language γCi1(X;𝒢)\gamma\in C_{i-1}\left(X;\;\mathcal{G}\right) is called an (i1)(i-1)-cycle.

Definition 3.

Let ff be an (i1)(i-1)-cocycle. The generalized Wilson loop variable associated to γ\gamma is the random variable Wγ:Ci1(X;𝔽q)W_{\gamma}:C^{i-1}\left(X;\,\mathbb{F}_{q}\right)\rightarrow\mathbb{C} given by

Wγ(f)=(f(γ)),W_{\gamma}\left(f\right)=\left(f\left(\gamma\right)\right)^{\mathbb{C}}\,,

where the \mathbb{C} superscript denotes that we are viewing the variable as a complex number. That is, if g𝔽q,g\in\mathbb{F}_{q}, gg^{\mathbb{C}} is the corresponding qq-th root of unity in .\mathbb{C}.

An important conjecture for the analogous quantities in Euclidean lattice gauge theory — called the Wilson area law — is that if γ=ρ\gamma=\partial\rho is a 11-boundary and ρ\rho is the minimal bounding chain then the expectation of Wρ(ω)W_{\partial\rho}\left(\omega\right) should decay as ec|ρ|e^{-c\left|\rho\right|} in some cases. For 𝒢=SU(2)\mathcal{G}=SU(2) or SU(3),SU(3), this conjecture is thought to be related to the phenomenon of quark confinement: that charged particles for the weak or strong nuclear forces are not seen in isolation, unlike for for the electromagnetic force [Wil74, Cha21] where the corresponding U(1)U\left(1\right) lattice gauge theory exhibits both “area law” and “perimeter law” phases [Gut80, FS82].

One-dimensional qq-state Potts lattice gauge theory on 4\mathbb{Z}^{4} is thought to undergo a sharp phase transition as the coupling constant β\beta increases, from a “perimeter law” to an“area law” regime. We state a more general form of this “sharpness” conjecture here.

Conjecture 4.

Fix integers q2q\geq 2 and 0i<d.0\leq i<d. Then there are constants 0<βc(q)<0<\beta_{c}\left(q\right)<\infty, and 0<c1(β,q),c2(β,q)<0<c_{1}(\beta,q),c_{2}(\beta,q)<\infty so that, for hyperrectangular (i1)(i-1)-boundaries γ\gamma in d,\mathbb{Z}^{d},

log(𝔼ν𝐟(Wγ))Area(γ)\displaystyle-\frac{\log\left(\mathbb{E}_{\nu^{\mathbf{f}}}(W_{\gamma})\right)}{\mathrm{Area}(\gamma)}\rightarrow c4(β,q)\displaystyle c_{\ref{const:conj1}}(\beta,q)\qquad β<βc(q)\displaystyle\beta<\beta_{c}(q)
log(𝔼ν𝐟(Wγ))Per(γ)\displaystyle-\frac{\log\left(\mathbb{E}_{\nu^{\mathbf{f}}}(W_{\gamma})\right)}{\mathrm{Per}(\gamma)}\rightarrow c4(β,q)\displaystyle c_{\ref{const:conj2}}(\beta,q)\qquad β>βc(q),\displaystyle\beta>\beta_{c}(q)\,,

as all dimensions of γ\gamma are taken to .\infty. In addition, if d=2i,d=2i, then βc(q)=βsd(q)=log(1+q).\beta_{c}\left(q\right)=\beta_{\mathrm{sd}}(q)=\log\left(1+\sqrt{q}\right).

Recall that WγW_{\gamma} is defined as a complex number, so the expectation is complex as well. Here, the perimeter of an (i1)(i-1)-boundary γ\gamma is the number of plaquettes in its support and its area is the number of plaquettes supported in the minimal bounding chain. Observe that when i=1,i=1, γ\gamma consists of two vertices {v,w},\left\{v,w\right\}, its perimeter is 22, and its area is the distance between vv and w.w. As such, the i=1i=1 case of the conjecture is the the sharpness of the phase transition for the Potts model, as proven by Duminil-Copin, Raoufi, and Tassion [DCRT19].

This conjecture is not known for any specific value of qq when i>1i>1, but Laanait, Messager, Ruiz showed that such a transition occurs when qq is sufficiently large [LMR89]. In addition, classical series expansions have been used to prove the existence of perimeter law and area law regimes for sufficiently extreme values of β.\beta. Specifically, Osterwalder and Seiler employed a “high–temperature expansion” to show existence of an area law regime for any lattice gauge theory [OS78, Sei82], and the argument should work for Potts lattice gauge theory as well. On the other hand, a perimeter law regime is demonstrated by a “low–temperature expansion,” a technique that has recently been used by Chatterjee [Cha20], Cao [Cao20], and Forsström, Lenells, and Viklund [FLV21] to prove precise estimates for the asymptotics of Wilson loop variables for lattice gauge theory when β\beta is large. We use the coupling of the plaquette random cluster model and Potts lattice gauge theory to give an alternate proof of the existence of perimeter law and area law regimes for ii-dimensional qq-state Potts lattice gauge theory in d,\mathbb{Z}^{d}, for prime values of q.q. Towards that end, we prove an exact relationship between Wilson loop variables and a certain topological event.

An ii-chain γ\gamma is null-homologous in a cubical complex PP if γ=ρ\gamma=\partial\rho for some (i+1)(i+1)-chain ρCi+1(P;𝒢)\rho\in C_{i+1}\left(P;\,\mathcal{G}\right) (in other words, γ=0\gamma=0 the ii-dimensional homology group). Roughly speaking, when i=1i=1 this means that γ\gamma is “bounded by a surface of plaquettes,” with the surfaces under consideration being dependent on the coefficient group 𝒢.\mathcal{G}. For 𝒢=,\mathcal{G}=\mathbb{Z}, γ\gamma is null-homologous if and only if it is bounded by an orientable surface of plaquettes; when 𝒢=𝔽2\mathcal{G}=\mathbb{F}_{2} any surface of plaquettes will suffice. We denote the event that γ\gamma is null-homologous by Vγ,V_{\gamma}, with the choice of coefficients being understood in context. We have the following result.

Theorem 5.

Suppose q2q\geq 2 is a prime integer, let 0<i<d1,0<i<d-1, and let γ\gamma be an (i1)(i-1)-cycle in d\mathbb{Z}^{d}. Then

𝔼ν𝐟(Wγ)=μd(Vγ),\mathbb{E}_{\nu^{\mathbf{f}}}\left(W_{\gamma}\right)=\mu_{\mathbb{Z}^{d}}\left(V_{\gamma}\right)\,,

where ν𝐟=νd,β,q,i1,d𝐟\nu^{\mathbf{f}}=\nu_{\mathbb{Z}^{d},\beta,q,i-1,d}^{\mathbf{f}} is Potts lattice gauge theory and μd=μd,1eβ,q,i\mu_{\mathbb{Z}^{d}}=\mu_{\mathbb{Z}^{d},1-e^{-\beta},q,i} is the corresponding random cluster model.

While we state this theorem for d,\mathbb{Z}^{d}, the same proof works for any finite cubical complex. A precise definition of the infinite volume ii-dimensional plaquette random cluster model μd=μd,p,q,i\mu_{\mathbb{Z}^{d}}=\mu_{\mathbb{Z}^{d},p,q,i} is given in Section 4.2. As a corollary, we have that Conjecture 4 is equivalent to the special case of the following conjecture for the plaquette random cluster model when qq is a prime integer and 𝔽=q.\mathbb{F}=\mathbb{Z}_{q}.

Conjecture 6.

There exist constants pc(q)(0,1)p_{c}\left(q\right)\in\left(0,1\right) and 0<c3(p,q),c4(p,q)<0<c_{3}\left(p,q\right),c_{4}\left(p,q\right)<\infty so that, for hyperrectangular (i1)(i-1)-boundaries γ\gamma in d,\mathbb{Z}^{d},

log(μd(Vγ))Area(γ)\displaystyle-\frac{\log\left(\mu_{\mathbb{Z}^{d}}(V_{\gamma})\right)}{\mathrm{Area}(\gamma)}\rightarrow c6(p,q)\displaystyle c_{\ref{const:conj3}}(p,q)\qquad p<pc(q)\displaystyle p<p_{c}(q)
log(μd(Vγ))Per(γ)\displaystyle-\frac{\log\left(\mu_{\mathbb{Z}^{d}}(V_{\gamma})\right)}{\mathrm{Per}(\gamma)}\rightarrow c6(p,q)\displaystyle c_{\ref{const:conj4}}(p,q)\qquad p>pc(q),\displaystyle p>p_{c}(q)\,,

as all dimensions of γ\gamma are taken to .\infty. Also, when d=2i,d=2i, pc(q)=q1+qp_{c}\left(q\right)=\frac{\sqrt{q}}{1+\sqrt{q}}.

When i=1,i=1, this conjecture is the sharpness of the phase transition of the classical random cluster model, as proven in [DCRT19]. In addition, the special case of the conjecture for 22-dimensional independent plaquette percolation in 3\mathbb{Z}^{3} (i=2,d=3,q=1i=2,d=3,q=1) is a theorem of Aizenman, Chayes, Chayes, Frölich, and Russo, who demonstrated that this phase transition is dual to the bond percolation transition333Their paper phrases this result in terms of possibly distinct critical probabilities, but these have since been shown to coincide [Men86, GM90].. Their proof works for general coefficient fields.

Another consequence of Theorem 5, together with a comparison with plaquette percolation, is a new proof of the existence of area law and perimeter law regimes for Potts lattice gauge theory. We state a more general result for the plaquette random cluster model.

Theorem 7.

Let 𝔽\mathbb{F} be a field, q[1,)q\in\left[1,\infty\right) and let 0<i<d.0<i<d. There exist positive, finite constants c5=c7(p,q,i,d,𝔽),c6=c7(p,q,i,d,𝔽)c_{5}=c_{\ref{const:3}}(p,q,i,d,\mathbb{F}),c_{6}=c_{\ref{const:4}}(p,q,i,d,\mathbb{F}) and 0<p1p2<10<p_{1}\leq p_{2}<1 so that, for hyperrectangular (i1)(i-1)-boundaries γ\gamma in d,\mathbb{Z}^{d},

exp(c7Area(γ))μd(Vγ)exp(c7Per(γ))\exp(-c_{\ref{const:3}}\mathrm{Area}(\gamma))\leq\mu_{\mathbb{Z}^{d}}(V_{\gamma})\leq\exp(-c_{\ref{const:4}}\mathrm{Per}(\gamma)) (2)

for all p(0,1)p\in\left(0,1\right) and so that

log(μd(Vγ))Per(γ))=Θ(1)\displaystyle-\frac{\log\left(\mu_{\mathbb{Z}^{d}}(V_{\gamma})\right)}{\mathrm{Per}(\gamma))}\;\;=\;\;\Theta\left(1\right) if p<p1\displaystyle\quad\text{if }p<p_{1}
log(μd(Vγ))Area(γ)c7\displaystyle-\frac{\log\left(\mu_{\mathbb{Z}^{d}}(V_{\gamma})\right)}{\mathrm{Area}(\gamma)}\;\;\rightarrow\;\;c_{\ref{const:3}} if p>p2.\displaystyle\quad\text{if }p>p_{2}\,.

Note that the constants could depend on the boundary conditions used to construct the infinite volume plaquette random cluster model.

The perimeter law phase can be interpreted as a “hypersurface–dominated regime” in a quantitative sense, where there are many large hypersurfaces of ii-plaquettes in d\mathbb{Z}^{d} [ACC+83]. For plaquette percolation on the torus, Duncan, Kahle, and Schweinhart [DKS20] showed that this phase coincides with a “hypersurface–dominated regime” in the qualitative sense of homological percolation, that is, one marked by the appearance of ii-dimensional cycles in the random cubical complex that are representatives of nonzero homology classes in the ambient torus (see Section 2 for more detail). Unlike the theorem of [ACC+83], this homological percolation result generalizes to to ii-dimensional plaquette percolation in the 2i2i dimensional torus. Here, we further generalize it to the ii-dimensional random-cluster model in the 2i2i-dimensional torus.

Theorem 8.

Suppose d=2i,d=2i, and let ϕ:Hi(P;𝔽)Hi(𝕋Nd;𝔽)\phi_{*}:H_{i}\left(P;\,\mathbb{F}\right)\to H_{i}\left(\mathbb{T}^{d}_{N};\,\mathbb{F}\right) be the induced map on homology with coefficients in a field 𝔽\mathbb{F} with char(F)2.\mathrm{char}\left(F\right)\neq 2. Denote by AA and SS the events that ϕ\phi_{*} is non-trivial and surjective, respectively, and set psd=q1+q.p_{\mathrm{sd}}=\frac{\sqrt{q}}{1+\sqrt{q}}. Then

{μ𝕋Nd(A)0p<psd(q)μ𝕋Nd(S)1p>psd(q)\begin{cases}\mu_{\mathbb{T}^{d}_{N}}\left(A\right)\rightarrow 0&p<p_{\mathrm{sd}}(q)\\ \mu_{\mathbb{T}^{d}_{N}}\left(S\right)\rightarrow 1&p>p_{\mathrm{sd}}(q)\\ \end{cases}

as N.N\rightarrow\infty.

When qq is a prime integer and i=2i=2, the phase transition is marked by the emergence of “giant sheets” on which the Wilson loop variables are constant within homology classes. The critical point is alternatively characterized by a qualitative change in the generalized Swendsen–Wang dynamics from local to non-local behavior, as defined in Section 5.1 below.

We also generalize the other two main theorems of [DKS20] from plaquette percolation to the plaquette random-cluster model. The proofs of these results and the previous one rely heavily on the duality properties of the plaquette random-cluster model; in Theorem 18 we show that the dual of μ𝕋Nd,p,q,i\mu_{\mathbb{T}^{d}_{N},p,q,i} is distributed approximately as μ𝕋Nd,p,q,di\mu_{\mathbb{T}^{d}_{N},p^{*},q,d-i} in a precise sense, where

p=p(p,q)=(1p)q(1p)q+p.p^{*}=p^{*}(p,q)=\frac{\left(1-p\right)q}{\left(1-p\right)q+p}\,.

Next, we prove that there are dual sharp phase transitions for i=1i=1 and i=d1i=d-1 that are consistent with the critical probability for the random-cluster model in d,\mathbb{Z}^{d}, assuming a conjecture about the continuity of the critical probability in slabs. Let

Sk2×{k,k+1,,k}d2d.S_{k}\coloneqq\mathbb{Z}^{2}\times\left\{-k,-k+1,\ldots,k\right\}^{d-2}\subset\mathbb{Z}^{d}\,.

Fix q1q\geq 1 and let pc(Sk)p_{c}\left(S_{k}\right) be the critical probability for the 11-dimensional random-cluster model with parameter qq on SkS_{k} with free boundary conditions. This can be constructed by a limit of free random-cluster models on

Sk,l{l,l+1,,l}2×{k,k+1,,k}d2.S_{k,l}\coloneqq\left\{-l,-l+1,\ldots,l\right\}^{2}\times\left\{-k,-k+1,\ldots,k\right\}^{d-2}\,.

Since Sk,lSk+1,l,S_{k,l}\subset S_{k+1,l}, it follows that pc(Sk)p_{c}\left(S_{k}\right) is decreasing in k.k. Then let

pcslab=limkpc(Sk).p_{c}^{\mathrm{slab}}=\lim_{k\to\infty}p_{c}\left(S_{k}\right)\,.

Let p^c=p^c(q,d)\hat{p}_{c}=\hat{p}_{c}(q,d) be the critical threshold for the random-cluster model with parameter qq on d.\mathbb{Z}^{d}. These two critical values are conjectured to coincide.

Conjecture 9 ([Pis96]).

For all q1,q\geq 1,

pcslab=p^c.p_{c}^{\mathrm{slab}}=\hat{p}_{c}\,.
Theorem 10.

Let q1.q\geq 1. Then the following statements hold:

If i=1i=1 then

{μ𝕋Nd(A)0p<p^cμ𝕋Nd(S)1p>pcslab\begin{cases}\mu_{\mathbb{T}^{d}_{N}}\left(A\right)\rightarrow 0&p<\hat{p}_{c}\\ \mu_{\mathbb{T}^{d}_{N}}\left(S\right)\rightarrow 1&p>p_{c}^{\mathrm{slab}}\\ \end{cases}

as N.N\rightarrow\infty.

Furthermore, if i=d1i=d-1 then

{μ𝕋Nd(A)0p<(p^c)μ𝕋Nd(S)1p>(pcslab)\begin{cases}\mu_{\mathbb{T}^{d}_{N}}\left(A\right)\rightarrow 0&p<(\hat{p}_{c})^{*}\\ \mu_{\mathbb{T}^{d}_{N}}\left(S\right)\rightarrow 1&p>(p_{c}^{\mathrm{slab}})^{*}\\ \end{cases}

as N.N\rightarrow\infty.

In particular, if Conjecture 9 is true, then there are sharp thresholds at p^c=pcslab\hat{p}_{c}=p_{c}^{\mathrm{slab}} and (p^c)=(pcslab)\left(\hat{p}_{c}\right)^{*}=\left(p_{c}^{\mathrm{slab}}\right)^{*} respectively.

In general dimensions we demonstrate the existence of a sharp threshold function defined as follows: Let λ=λ(N,q,i,d)\lambda=\lambda\left(N,q,i,d\right) satisfy

μ𝕋Nd,λ,q,i(A)=1/2\displaystyle\mu_{\mathbb{T}^{d}_{N},\lambda,q,i}\left(A\right)=1/2

and let pl=pl(q,i,d)lim infNλ(N,q,i,d)p_{l}=p_{l}\left(q,i,d\right)\coloneqq\liminf_{N\to\infty}\lambda\left(N,q,i,d\right) and pu=pu(q,i,d)lim supNλ(N,q,i,d).p_{u}=p_{u}\left(q,i,d\right)\coloneqq\limsup_{N\to\infty}\lambda\left(N,q,i,d\right).

Theorem 11.

Let q1q\geq 1 be a prime integer and suppose that char(F)2.\mathrm{char}\left(F\right)\neq 2. For every d2,d\geq 2, 1id1,1\leq i\leq d-1, and ϵ>0\epsilon>0

{μ𝕋Nd,λϵ,q,i(A)0μ𝕋Nd,λ+ϵ,q,i(S)1\begin{cases}\mu_{\mathbb{T}^{d}_{N},\lambda-\epsilon,q,i}\left(A\right)\rightarrow 0\\ \mu_{\mathbb{T}^{d}_{N},\lambda+\epsilon,q,i}\left(S\right)\rightarrow 1\\ \end{cases}

as N.N\rightarrow\infty.

Moreover, for every d2d\geq 2 and 1id11\leq i\leq d-1 we have

0<plpu<1,0<p_{l}\leq p_{u}<1\,,

and plp_{l} and pup_{u} have the following properties.

  1. (a)

    (Duality) pu(q,i,d)=(pl(q,di,d)).p_{u}\left(q,i,d\right)=\left(p_{l}\left(q,d-i,d\right)\right)^{*}.

  2. (b)

    (Monotonicity in ii and dd) pu(q,i,d)<pu(q,i,d1)<pu(q,i+1,d)p_{u}\left(q,i,d\right)<p_{u}\left(q,i,d-1\right)<p_{u}\left(q,i+1,d\right) for 0<i<d1.0<i<d-1.

1.2. Why does qq need to be prime?

Our results for qq-state Potts lattice gauge theory are for cases where the cohomology coefficients are a finite field 𝔽q\mathbb{F}_{q} where qq is a prime integer, instead of the more general additive group n.\mathbb{Z}_{n}. This is not just for convenience, though it is more convenient to work with homology and cohomology with field coefficients since the associated groups are then vector spaces. There is one key place where this assumption is necessary, as noticed by Aizenman and Frölich [AF84]. It arises trying to couple the plaquette random-cluster model with Potts lattice gauge theory. Recall that the definition of the random-cluster model involves the term q𝐛i1(P),q^{\mathbf{b}_{i-1}\left(P\right)}, where 𝐛i1(P)\mathbf{b}_{i-1}\left(P\right) is dimension of the 𝔽\mathbb{F}-vector space Hi1(P;𝔽)H_{i-1}\left(P;\,\mathbb{F}\right). If we were to use n\mathbb{Z}_{n} coefficients, we could try to replace 𝐛i1(P)\mathbf{b}_{i-1}\left(P\right) with the rank of the n\mathbb{Z}_{n}-module Hi(P;n)H_{i}\left(P;\,\mathbb{Z}_{n}\right) (that is, the size of the minimal generating set). However, this does not lead to the same model as the marginal of the generalized Edwards–Sokal coupling defined in Section 5. The argument in Proposition 20 breaks down because, in general,

|Hi(P;n)|nrank(Hi(P;n)).\left|H_{i}\left(P;\,\mathbb{Z}_{n}\right)\right|\neq n^{\mathrm{rank}\left(H_{i}\left(P;\,\mathbb{Z}_{n}\right)\right)}\,.

This could be resolved by defining a dependent plaquette percolation model so that

(P)p|P|(1p)|Xi||P||Hi(P;n)|.\mathbb{P}\left(P\right)\propto p^{\left|P\right|}\left(1-p\right)^{\left|X^{i}\right|-\left|P\right|}\left|H_{i}\left(P;\,\mathbb{Z}_{n}\right)\right|\,.

We expect some of our results to carry over to this model, but we defer them to a later paper.

Aizenman and Frölich identified a second issue, where they constructed a plaquette system in which there exist (i1)(i-1)-cycles γ\gamma and γ\gamma^{\prime} so that [γ]\left[\gamma\right] is homologous to q[γ].q\left[\gamma^{\prime}\right]. Then an analogue of Theorem 5 fails when VγV_{\gamma} is defined to be the event that γ\gamma is bounded by a surface of plaquettes. However, we resolve this by redefining VγV_{\gamma} to depend on the choice of qq; clearly [γ]=0\left[\gamma\right]=0 in Hi1(P;q)H_{i-1}\left(P;\;\mathbb{Z}_{q}\right) in this example.

Finally, in the proof of the homological percolation phase transition, Lemma 41 requires that the homology groups are vector spaces.

1.3. Outline

We outline the remainder of the paper. First, in Section 2 reviews the definitions of homology and cohomology for cubical complexes. These are used in Section 3 to show useful topological duality results. Next, we define the plaquette random-cluster model in Section 4 and prove several of its basic properties. These include positive association (which was previously shown by Hiraoka and Shirai [HS16]), the existence of the infinite volume limit, and duality. Section 5 covers the relationship between the plaquette random-cluster model and Potts lattice gauge theory: the coupling between them, the definition of the plaquette Swendsen–Wang algorithm, and proofs of Theorems 5 and 7. Finally, we prove our results for homological percolation — Theorems 8, 10, and 11 — in Section 6.

2. Homology and Cohomology

We give a brief review of homology and cohomology, two fundamental invariants studied in algebraic topology. These theories exhibit a number of dualities which are closely related to classical dualities between lattice spin models and — as we will demonstrate below — allow them to be generalized to higher dimensions. Though this paper is not self-contained with respect to topological background, we hope to provide enough context for a reader with minimal previous knowledge of the area to make sense of our results. More detail can be found in [Hat02] (the standard mathematical reference) or [DW82] which provides an exposition specific to the context of lattice spin models but does not cover all the material we need here. Those familiar with differential forms but not algebraic topology may find some of these concepts familiar, as they are important in a continuous version of the discrete theories developed here.

In this paper we will deal exclusively with spaces called cubical complexes, which are composed of ii-dimensional plaquettes (unit ii-dimensional cubes) for various i.i. Two references with details specific to this setting are [KMM04, Sav16]. However, many of the definitions and results in this paper can easily be adapted to more general cell complexes. A first example of a cubical complex is the integer lattice cubical complex d,\mathbb{Z}^{d}, which is a union of all ii-plaquettes for 0id0\leq i\leq d which have corners in the integer lattice. The dd-dimensional torus also has a natural cubical complex structure obtained by dividing the cube [0,N]d\left[0,N\right]^{d} into unit cubes and identifying opposite faces of the original cube. We call this cubical complex 𝕋Nd.\mathbb{T}_{N}^{d}. All examples we will consider in this paper are subcomplexes of either d\mathbb{Z}^{d} or 𝕋Nd.\mathbb{T}_{N}^{d}.

2.1. Homology

In order to convert vague geometric questions about “the number of holes” of a topological space into concrete algebraic quantities, homology theory defines a space Ci(X;𝒢)C_{i}\left(X;\,\mathcal{G}\right) of linear combinations with coefficients in an abelian group 𝒢\mathcal{G} (for a concrete example, take 𝒢=\mathcal{G}=\mathbb{Z}) of ii-plaquettes called chains, and boundary operators, which map an ii-plaquette to a sum of its (i1)(i-1)-faces. The continuous analogues of these concepts are, roughly speaking “ii-dimensional spaces on which one can integrate an ii-form” (called an ii-current) and the geometric boundary of a space.

For reasons that become apparent below, it is important to give each term in the boundary of a plaquette a sign corresponding to the orientation of a plaquette. The formula is relatively simple in low dimensions. A zero-plaquette is a vertex and its boundary is zero. A 11-plaquette is an edge (v1,v2)\left(v_{1},v_{2}\right) and its boundary is the difference v2v1.v_{2}-v_{1}. A two-plaquette is an oriented unit square with vertices (v1,v2,v3,v4)\left(v_{1},v_{2},v_{3},v_{4}\right) and its boundary is (v1,v2)+(v2,v3)+(v3,v4)(v1,v4).\left(v_{1},v_{2}\right)+\left(v_{2},v_{3}\right)+\left(v_{3},v_{4}\right)-\left(v_{1},v_{4}\right). More generally, let 1k1<k2<<kid1\leq k_{1}<k_{2}<\ldots<k_{i}\leq d and let Ij=[0,1]I_{j}=\left[0,1\right] for j{k1,k2,,ki}j\in\left\{k_{1},k_{2},\ldots,k_{i}\right\} and Ij={0}I_{j}=\left\{0\right\} for j[d]k1,k2,,ki.j\in\left[d\right]\setminus{k_{1},k_{2},\ldots,k_{i}}. Then σ=1jdIj\sigma=\prod_{1\leq j\leq d}I_{j} is an ii-plaquette in d,\mathbb{R}^{d}, and its boundary is given by

σ=l=0i(1)l1(1j<klIj×{1}×kl<mdIm1j<klIj×{0}×kl<mdIm).\partial\sigma=\sum_{l=0}^{i}\left(-1\right)^{l-1}\left(\prod_{1\leq j<k_{l}}I_{j}\times\left\{1\right\}\times\prod_{k_{l}<m\leq d}I_{m}-\prod_{1\leq j<k_{l}}I_{j}\times\left\{0\right\}\times\prod_{k_{l}<m\leq d}I_{m}\right)\,.

The reason that the sum is alternating in sign is so that the boundary operator satisfies the equation

=0.\partial\circ\partial=0\,. (3)
Refer to caption
Figure 1. The boundary map for a two dimensional plaquette.

Of particular interest are the chains αCi(X;𝒢)\alpha\in C_{i}\left(X;\,\mathcal{G}\right) satisfying α=0.\partial\alpha=0. Such chains are called cycles, the space of which is denoted Zi(X;𝒢).Z_{i}\left(X;\,\mathcal{G}\right). Equation 3 provides one source of ii-cycles, namely the boundaries of (i+1)(i+1)-plaquettes. We denote this space Bi(X;𝒢).B_{i}\left(X;\,\mathcal{G}\right). It turns out that the most interesting cycles are the ones that are not boundaries.

To illustrate why this is the case, consider the unit square Q=[0,1]22,Q=\left[0,1\right]^{2}\subset\mathbb{R}^{2}, seen in Figure 1. In the cubical complex structure we defined earlier, we have that Z1(Q;𝒢)𝒢Z_{1}\left(Q;\,\mathcal{G}\right)\simeq\mathcal{G} consists of multiples of a single 11-cycle, namely the boundary of Q.Q. But now imagine that we make our lattice spacing 1/21/2 instead of 1.1. Now Z1(Q;𝒢)Z_{1}\left(Q;\,\mathcal{G}\right) contains linear combinations of the boundaries of the four 22-plaquettes contained in Q.Q. One such combination can be seen in Figure 2. Though QQ has not changed, our choice of lattice has made the cycle space significantly more complicated.

Refer to caption
Figure 2. The boundary of a union of 4 plaquettes computed with the linearity of the boundary operator.

Therefore, in order to measure the shape of spaces in a way that does not depend on our choice of lattice, we define the ii-dimensional homology group as the quotient

Hi(X;𝒢)=Zi(X;𝒢)/Bi(X;𝒢).H_{i}\left(X;\,\mathcal{G}\right)=Z_{i}\left(X;\,\mathcal{G}\right)/B_{i}\left(X;\,\mathcal{G}\right)\,.

Returning to the unit square Q,Q, every 11-cycle is a boundary regardless of the cubical complex structure on QQ, so H1(Q;𝒢)=0.H_{1}\left(Q;\,\mathcal{G}\right)=0. In general, it turns out that homology groups do not depend on the cubical complex structure of a space, and are invariant to continuous deformations of the space. In particular, one can show that H1(Q;𝒢)=0H_{1}\left(Q;\,\mathcal{G}\right)=0 by continuously deforming QQ to a point, which has no nonzero 11-chains let alone 11-cycles.

For an example with nontrivial first homology, consider H1(Q;𝒢),H_{1}\left(\partial Q;\,\mathcal{G}\right), where Q\partial Q is the topological boundary of Q,Q, i.e. the empty square formed by the union of the four sides of Q.Q. There is still the 11-cycle from before but there are no 22-plaquettes, so H1(Q;𝒢)0.H_{1}\left(\partial Q;\,\mathcal{G}\right)\neq 0. This example illustrates the often repeated informal description of Hi(X)H_{i}\left(X\right) as a measurement of the number of ii-dimensional holes of X.X. In the case of H1,H_{1}, the cycles that remain are the loops that cannot be filled in by 22-plaquettes.

We can now define giant cycles in a subcomplex of the torus. If X𝕋NdX\subset\mathbb{T}^{d}_{N} consists of a subset of the cubical cells of 𝕋Nd,\mathbb{T}^{d}_{N}, then the chain groups of XX are subgroups of the chain groups of 𝕋Nd.\mathbb{T}^{d}_{N}. The giant cycles of XX are the cycles that are not boundaries when considered as chains in 𝕋Nd.\mathbb{T}^{d}_{N}. In the case i=1,i=1, a giant cycle is a loop that spans the torus 𝕋Nd\mathbb{T}^{d}_{N} in some sense, and must therefore have length at least N.N. We interpret this as a finite volume analogue of an infinite path in the lattice. See Figure 3 for illustrations of two giant cycles.

Refer to caption
Refer to caption
Figure 3. Examples of giant cycles: (left) a 11-dimensional giant cycle in 𝕋2\mathbb{T}^{2} and (right) a 22-dimensional giant cycle in 𝕋3,\mathbb{T}^{3}, shown in cubes with periodic boundary conditions. Reproduced with permission from [DKS20].

2.2. Coefficients and Torsion

So far, the choice of coefficient group 𝒢\mathcal{G} has not been important to our computations in this section, though we have mentioned their importance to the random-cluster model in the introduction. In this paper, we will almost exclusively work with homology over the finite field of integers mod qq for prime q,q, denoted 𝔽q.\mathbb{F}_{q}. We will see later that this is the correct choice to study Potts lattice gauge theory, and it has the advantage of simplifying the algebra required since a homology group with field coefficients is a vector space.

For an example where different coefficients produce different results, consider the identifications of the sides of the 22-plaquette in Figure 4. In the first case we roll up the plaquette into a cylinder and then put the opposite ends of the cylinder together to form a torus 𝕋.\mathbb{T}. In the second, we also form a cylinder but put the opposite ends of the cylinder together in the reverse orientation to form a Klein bottle 𝕂.\mathbb{K}. Consider the second homology of both spaces. In both cases there are no 33-plaquettes and only one 22-plaquette σ,\sigma, so it suffices to check if the 22-chains generated by σ\sigma contain nontrivial cycles. In the torus,

σ=e1+e2e1e2=0,\partial\sigma=e_{1}+e_{2}-e_{1}-e_{2}=0\,,

so H2(𝕋;𝔽)𝔽H_{2}\left(\mathbb{T};\,\mathbb{F}\right)\simeq\mathbb{F} for any field 𝔽.\mathbb{F}. In the Klein bottle,

σ=e1+e2+e1e2=2e1,\partial\sigma=e_{1}+e_{2}+e_{1}-e_{2}=2e_{1}\,,

which is nonzero over most fields, with a number of exceptions including 𝔽2.\mathbb{F}_{2}. Thus, H2(𝕂;𝔽2)𝔽2,H_{2}\left(\mathbb{K};\,\mathbb{F}_{2}\right)\simeq\mathbb{F}_{2}, and H2(𝕂;𝔽q)=0H_{2}\left(\mathbb{K};\,\mathbb{F}_{q}\right)=0 for any prime q2.q\neq 2. There is also a difference in first homology; one can compute H1(𝕂;𝔽2)𝔽2𝔽2H_{1}\left(\mathbb{K};\,\mathbb{F}_{2}\right)\simeq\mathbb{F}_{2}\oplus\mathbb{F}_{2} and H1(𝕂;𝔽q)𝔽qH_{1}\left(\mathbb{K};\,\mathbb{F}_{q}\right)\simeq\mathbb{F}_{q} for q2.q\neq 2. Of course, there are no single plaquettes with sides identified this way in the integer lattice, but non-orientable surfaces such as the Klein bottle appear as unions of plaquettes in dimensions 44 and higher. Homology that appears over some finite fields but not others is sometimes called torsion.

Refer to caption
Figure 4. A torus 𝕋\mathbb{T} and a Klein bottle 𝕂,\mathbb{K}, the resulting shapes of two possible identifications of the opposite sides of a plaquette. The left plaquette is a cycle with any field coefficients and the right is only a cycle with 𝔽2\mathbb{F}_{2} coefficients.

It is worth mentioning a special situation that arises for the 0-th homology group. It is not hard to check that H0(X;𝔽)=𝔽cX,H_{0}\left(X;\,\mathbb{F}\right)=\mathbb{F}^{{c_{X}}}, where cXc_{X} is the number of connected components of X.X. Some standard results in algebraic topology have simpler statements when all of the homology groups of a space consisting of a single point are zero. This is not true in the setup that we have described so far, since there is one connected component. To fix this, define the reduced homology groups

H~k(X;𝔽)={Hk(X;𝔽)k1𝔽cX1k=0.\tilde{H}_{k}\left(X;\,\mathbb{F}\right)=\begin{cases}H_{k}\left(X;\,\mathbb{F}\right)&k\geq 1\\ \mathbb{F}^{c_{X}-1}&k=0\end{cases}\,.

2.3. The Euler–Poincaré formula

The Euler–Poincaré formula is an important result that relates the Euler characteristic of a cubical complex to the ranks of its homology groups with coefficients in any field. Define the ii-th Betti number of a cubical complex XX with coefficients in 𝔽\mathbb{F} by

𝐛i(X)=𝐛i(X;𝔽)dimHi(X;𝔽).\mathbf{b}_{i}\left(X\right)=\mathbf{b}_{i}\left(X;\,\mathbb{F}\right)\coloneqq\dim H_{i}\left(X;\,\mathbb{F}\right).

Also, let XiX^{i} be the set of ii-plaquettes of XX and let |Xi|\left|X^{i}\right| be their number. The Euler characteristic of XX is then

χ(X)=i=0dim(X)(1)i|Xi|.\chi\left(X\right)=\sum_{i=0}^{\mathrm{dim}\left(X\right)}\left(-1\right)^{i}\left|X^{i}\right|\,.

The following theorem is a straightforward consequence of the rank–nullity theorem.

Theorem 12 (Euler–Poincaré formula).

For any cubical complex XX and any field 𝔽\mathbb{F}

χ(X)=i=0dim(X)(1)i𝐛i(X;𝔽).\chi\left(X\right)=\sum_{i=0}^{\mathrm{dim}\left(X\right)}\left(-1\right)^{i}\mathbf{b}_{i}\left(X;\,\mathbb{F}\right)\,.

2.4. Cohomology

A small algebraic change in the definitions that give the homology groups leads to cohomology. The ii-th cochain group Ci(X;𝔽)C^{i}\left(X;\,\mathbb{F}\right) is defined as the group of 𝔽\mathbb{F}-linear functions from Ci(X;𝔽)C_{i}\left(X;\,\mathbb{F}\right) to 𝔽.\mathbb{F}. Since we are using field coefficients, this is the dual vector space to Ci(X;𝔽).C_{i}\left(X;\,\mathbb{F}\right). Then the coboundary operator δ:Ci(X;𝔽)Ci+1(X;𝔽)\delta:C^{i}\left(X;\,\mathbb{F}\right)\to C^{i+1}\left(X;\,\mathbb{F}\right) is defined by

δf(α)=f(α)\delta f\left(\alpha\right)=f\left(\partial\alpha\right)

for fCi(X;𝔽),αCi+1(X;𝔽).f\in C^{i}\left(X;\,\mathbb{F}\right),\alpha\in C_{i+1}\left(X;\,\mathbb{F}\right). The continuous analogues of an ii-cochain and and its coboundary are a differential form and its exterior derivative. Note that if ff is an ii-cochain and α\alpha is an ii-chain, the evaluation f(α)f\left(\alpha\right) corresponds to integration and the definition of the coboundary operator corresponds to Stokes’ Theorem.

Analogously to homology groups, we define the ii-th cocycle group Zi(X;𝔽)Z^{i}\left(X;\,\mathbb{F}\right) to be the kernel of δ\delta and the ii-th coboundary group Bi(X;𝔽)B^{i}\left(X;\,\mathbb{F}\right) to be the image of δ\delta applied to Ci+1(X;F).C^{i+1}\left(X;\,F\right). Then ii-th cohomology group is the quotient

Hi(X;𝔽)=Zi(X;𝔽)/Bi(X;𝔽).H^{i}\left(X;\,\mathbb{F}\right)=Z^{i}\left(X;\,\mathbb{F}\right)/B^{i}\left(X;\,\mathbb{F}\right)\,.

In the continuous setting, the analogue of the ii-dimensional cohomology group is the ii-dimensional de Rham cohomology of closed forms modulo exact forms. This is more than just as analogy: de Rham’s theorem states that the de Rham cohomology of a smooth manifold is isomorphic to Hi(M,)H^{i}\left(M,\mathbb{R}\right) computed using a cubical complex structure on MM [Bre13].

In some ways cohomology does not add more information than homology already provided. The universal coefficient theorem for cohomology tells us that Hi(X;𝔽)Hi(X;𝔽)H_{i}\left(X;\,\mathbb{F}\right)\simeq H^{i}\left(X;\,\mathbb{F}\right) when 𝔽\mathbb{F} is a field (see Corollary 3.3 of [Hat02]). However, this algebraic equivalence does not preclude problems lending themselves more naturally to one perspective or the other. For example, it is arguably more intuitive to think of giant cycles as homological in nature, due to their interpretation as giant “hypersurfaces” spanning the surface. On the other hand, we argue that lattice gauge theory should be thought of as a random cochain and Wilson loop variables as cohomological quantities. It is also worth noting that although the homology and cohomology groups themselves may be isomorphic, there is a relationship between the different dimensional cohomology groups in the form of a multiplication of cocycles called the cup product that does not always have a homological analogue (see Section 3.2 of [Hat02]).

Homology and cohomology come together in global duality theorems, of which there are many versions. We primarily use variations of Alexander duality, the original form of which says that for a sufficiently “nice” subspace XSd,X\subset S^{d},

H~i(X;𝔽)H~di1(SdX;𝔽)\tilde{H}_{i}\left(X;\,\mathbb{F}\right)\simeq\tilde{H}^{d-i-1}\left(S^{d}\setminus{X};\,\mathbb{F}\right)

for 1id1.1\leq i\leq d-1. In the simplest case, this says that the number of bounded components of a “nice” subset of the plane equals the number of “minimal contours” of its complement. Since the complement of a set of plaquettes is a thickening of the dual system (as shown in Proposition 13 below), we are able to use techniques related to the proof of Alexander duality to relate both the giant cycles and local cycles of the two sets (Theorem 14 below). This relationship is quantified in terms of the dimension of the relevant subspaces of the homology groups, and is a main ingredient in the proof that the dual complex is close to a random-cluster model.

We will now briefly discuss the relevant information about the topology of the torus specifically. Although in percolation theory the torus is most often thought of as a cube with periodic boundary conditions, it is useful topologically to think of it as the product of dd copies of the circle S1.S^{1}. In such a product space, the Künneth formula for homology (Section 3B of [Hat02]) tells us that there is an isomorphism of the form

i+j=kHi(X;𝔽)×Hj(Y;𝔽)Hk(X×Y;𝔽).\bigoplus_{i+j=k}H_{i}\left(X;\,\mathbb{F}\right)\times H_{j}\left(Y;\,\mathbb{F}\right)\simeq H_{k}\left(X\times Y;\,\mathbb{F}\right)\,.

Since H1(S1;𝔽)H0(S1;𝔽)𝔽H_{1}\left(S^{1};\,\mathbb{F}\right)\simeq H_{0}\left(S^{1};\,\mathbb{F}\right)\simeq\mathbb{F} and Hi(S1;𝔽)=0H_{i}\left(S^{1};\,\mathbb{F}\right)=0 for all i2,i\geq 2, we see that

Hi(𝕋;𝔽)𝔽(di).H_{i}\left(\mathbb{T};\,\mathbb{F}\right)\simeq\mathbb{F}^{\binom{d}{i}}\,.

Furthermore, there is a set of generating ii-cycles consisting of the products of ii of the dd possible S1S^{1} factors.

3. Topology and Duality

Before defining the plaquette random-cluster model, we will prove some useful duality results that are true for any cubical complex on the torus. Some of these results are from our previous paper on the Bernoulli plaquette model. The dual of the cubical complex 𝕋Nd\mathbb{T}^{d}_{N} is the complex (𝕋Nd)\left(\mathbb{T}^{d}_{N}\right)^{\bullet} obtained by shifting by 1/21/2 in each coordinate direction, and the dual (d)\left(\mathbb{Z}^{d}\right)^{\bullet} of d\mathbb{Z}^{d} is defined similarly. These pairs of dual complexes have the property that each ii-plaquette intersects exactly one (di)(d-i)-plaquette of (𝕋Nd),\left(\mathbb{T}^{d}_{N}\right)^{\bullet}, giving a matching of plaquettes and dual plaquettes. As in classical bond percolation in the plane, this induces a matching between subcomplexes of each. Namely, given a cubical complex P=P(ω),P=P\left(\omega\right), we define the dual complex PP^{\bullet} to be the union of all plaquettes for which the dual plaquette is not included in ω.\omega. When PP comes from the plaquette random-cluster model defined below, it will include the union of all lower dimensional plaquettes of 𝕋Nd\mathbb{T}^{d}_{N} (called the “(i1)(i-1)-skeleton”) and no plaquettes of dimension higher than i.i. Note that in that case, PP^{\bullet} automatically contains the (di1)(d-i-1)-skeleton of (𝕋Nd).\left(\mathbb{T}^{d}_{N}\right)^{\bullet}. Note that each statement in this section has an analogue for subcomplexes of a box in d\mathbb{Z}^{d} with free boundary, where the dual complex is a subcomplex of a box with wired boundary.

As mentioned in our review of homology and cohomology, our use of topological duality relies on decomposing the torus into a union of disjoint subspaces. The plaquette system and its its dual are not complementary, but the dual system deformation retracts (collapses) onto the complement of the plaquette system. By standard results in algebraic topology, this implies that the dual has the same topological invariants as the complement (see Chapter 0 of [Hat02]).

Proposition 13 (Duncan, Kahle, and Schweinhart).

For any cubical complex P𝕋Nd,P\subseteq\mathbb{T}^{d}_{N}, 𝕋dP\mathbb{T}^{d}\setminus P deformation retracts to P.P^{\bullet}.

Duality allows us to relates the homology of PP with that of P.P^{\bullet}. Importantly, “global” cycles and “local” cycles behave differently. Let ϕ:Hk(P;𝔽)Hk(𝕋d;𝔽)\phi_{*}:H_{k}\left(P;\,\mathbb{F}\right)\rightarrow H_{k}\left(\mathbb{T}^{d};\,\mathbb{F}\right) and ψ:Hk(P;𝔽)Hk(𝕋d;𝔽)\psi_{*}:H_{k}\left(P^{\bullet};\,\mathbb{F}\right)\rightarrow H_{k}\left(\mathbb{T}^{d};\,\mathbb{F}\right) be the maps on homology induced by the respective inclusions. Set ak=dimkerϕk,ak=dimkerψk,bk=rankϕk,a_{k}=\dim\ker\phi_{k*},a_{k}^{\bullet}=\dim\ker\psi_{k*},b_{k}=\operatorname{rank}\phi_{k*}, and bk=rankψk.b_{k}^{\bullet}=\operatorname{rank}\psi_{k*}. We think of aka_{k} as counting the local cycles (cycles of the plaquette system that are boundaries in the ambient torus) and bkb_{k} as counting the global cycles. Furthermore, recall that 𝐛k=dimHk(P)\mathbf{b}_{k}=\dim H_{k}\left(P\right) and let 𝐛k=Hk(P).\mathbf{b}^{\bullet}_{k}=H_{k}\left(P^{\bullet}\right).

The following is a special case of the principle of Alexander duality, which is that the topology of a “nice” subset of a manifold should be related to that of its complement. Traditionally, Alexander duality is stated for a subset of the sphere SdS^{d} or of Euclidean space d\mathbb{R}^{d}, where there are no global cycles. The proof of the following relations for the torus makes use of the same techniques in the usual proof for those cases. These techniques, which include long exact sequences and relative homology, are beyond the level of the brief introduction in Section 2. We refer an interested reader to [Hat02] for background.

Theorem 14 (Alexander Duality).

For any cubical complex P𝕋NdP\subseteq\mathbb{T}^{d}_{N} and any 1kd11\leq k\leq d-1 the following relations hold.

ak+bk=𝐛k,ak+bk=𝐛k,\displaystyle a_{k}+b_{k}=\mathbf{b}_{k},\quad a^{\bullet}_{k}+b^{\bullet}_{k}=\mathbf{b}^{\bullet}_{k}\,, (4)
bk+bk=rankHk(𝕋d)=(ki),\displaystyle b_{k}+b_{k}^{\bullet}=\operatorname{rank}H_{k}\left(\mathbb{T}^{d}\right)=\binom{k}{i}\,, (5)

and,

ak=adk1.\displaystyle a_{k}=a_{d-k-1}^{\bullet}\,. (6)
Proof.

The first two equations are an immediate consequence of the rank–nullity theorem, and the third is Lemma 10 of [DKS20]. To prove the fourth, we require additional results from algebraic topology. Theorem 3.44 of [Hat02] gives the isomorphism

Hdi1(P)Hi+1(𝕋Nd,𝕋NdP),\displaystyle H^{d-i-1}\left(P\right)\cong H_{i+1}\left(\mathbb{T}^{d}_{N},\mathbb{T}^{d}_{N}\setminus P\right)\,,

where Hi(X,Y)H_{i}\left(X,Y\right) denotes the relative homology group of XX with respect to a subset Y.Y. Combining this with the long exact sequence of relative homology (see Section 2.1 of [Hat02]), we obtain the following commutative diagram:

Hi+1(𝕋Nd)Hi+1(𝕋Nd,𝕋NdP)Hi(𝕋NdP)Hi(𝕋Nd)Hdi1(𝕋Nd)Hdi1(P)φχϵ.\leavevmode\hbox to240.61pt{\vbox to40.66pt{\pgfpicture\makeatletter\hbox{\hskip 120.3061pt\lower-21.87753pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-123.3061pt}{-21.7777pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 21.06227pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-18.06227pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{i+1}\left(\mathbb{T}^{d}_{N}\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 21.06227pt\hfil&\hfil\hskip 49.82413pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-34.82414pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{i+1}\left(\mathbb{T}^{d}_{N},\mathbb{T}^{d}_{N}\setminus P^{\bullet}\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 37.82414pt\hfil&\hfil\hskip 39.78899pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-24.789pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{i}\left(\mathbb{T}^{d}_{N}\setminus P^{\bullet}\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 27.789pt\hfil&\hfil\hskip 29.48448pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-14.4845pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{i}\left(\mathbb{T}^{d}_{N}\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 17.4845pt\hfil\cr\vskip 8.99994pt\cr\hfil\hskip 22.20851pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-19.20851pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H^{d-i-1}\left(\mathbb{T}^{d}_{N}\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 22.20851pt\hfil&\hfil\hskip 34.44455pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-19.44456pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H^{d-i-1}\left(P^{\bullet}\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 22.44456pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{{ {\pgfsys@beginscope \pgfsys@setdash{}{0.0pt}\pgfsys@roundcap\pgfsys@roundjoin{} {}{}{} {}{}{} \pgfsys@moveto{-2.07988pt}{2.39986pt}\pgfsys@curveto{-1.69989pt}{0.95992pt}{-0.85313pt}{0.27998pt}{0.0pt}{0.0pt}\pgfsys@curveto{-0.85313pt}{-0.27998pt}{-1.69989pt}{-0.95992pt}{-2.07988pt}{-2.39986pt}\pgfsys@stroke\pgfsys@endscope}} }{}{}{{}}{}{}{{}}\pgfsys@moveto{-101.0976pt}{-1.20001pt}\pgfsys@lineto{-101.0976pt}{-9.4pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-101.0976pt}{-9.59998pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{{}{}}}{{}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-108.70036pt}{-7.24998pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\cong}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-79.83533pt}{6.99998pt}\pgfsys@lineto{-67.48906pt}{6.99998pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-67.28908pt}{6.99998pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-75.7518pt}{10.71384pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\varphi}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-29.06496pt}{-1.20001pt}\pgfsys@lineto{-29.06496pt}{-9.4pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-29.06496pt}{-9.59998pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{{}{}}}{{}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-36.66772pt}{-7.24998pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\cong}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{8.95918pt}{6.99998pt}\pgfsys@lineto{20.15921pt}{6.99998pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{20.35919pt}{6.99998pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{12.56926pt}{10.71384pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\chi}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{76.53717pt}{6.99998pt}\pgfsys@lineto{87.7372pt}{6.99998pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{87.93718pt}{6.99998pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{80.9165pt}{9.35275pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\epsilon}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-78.68909pt}{-19.2777pt}\pgfsys@lineto{-52.10948pt}{-19.2777pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-51.9095pt}{-19.2777pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}\,.

By Proposition 13 and the definition of the plaquette system,

Hi+1(𝕋NdP)Hi+1(P)0,\displaystyle H_{i+1}\left(\mathbb{T}^{d}_{N}\setminus P^{\bullet}\right)\cong H_{i+1}\left(P\right)\cong 0\,,

so φ\varphi is surjective. Similarly,

Hi(𝕋NdP)Hi(P).\displaystyle H_{i}\left(\mathbb{T}^{d}_{N}\setminus P^{\bullet}\right)\cong H_{i}\left(P\right)\,.

Then since ϵ\epsilon is the map on homology induced by the inclusion (𝕋NdP)𝕋Nd,\left(\mathbb{T}^{d}_{N}\setminus P^{\bullet}\right)\hookrightarrow\mathbb{T}^{d}_{N}, its image is isomorphic to the space of giant cycles of P.P. Thus, χ\chi restricts to an isomorphism between vector spaces of dimension 𝐛di1bdi1\mathbf{b}^{\bullet}_{d-i-1}-b^{\bullet}_{d-i-1} and 𝐛ibi\mathbf{b}_{i}-b_{i} respectively, and Equation 6 follows from Equation 4. ∎

At first glance, the dimensions in Equation 6 may seem off: the plaquette random-cluster model is weighted by the ii-th Betti number and is dual to a (di)(d-i)-dimensional model, but this equation relates its ii-th Betti number to the (di1)(d-i-1)-st Betti number of its dual. However, the random-cluster model always contains all cells of dimension lower than j<ij<i (that is, the (i1)(i-1)-skeleton). The presence or absence of ii-cells can only affect the Betti numbers in dimensions ii and (i1),(i-1), so these are the only variable Betti numbers in the random-cluster model. This, together with the Euler–Poincaré formula implies the following proposition.

Proposition 15.

There is a constant c7=c15(N,i,d)c_{7}=c_{\ref{const:5}}\left(N,i,d\right) so that for any ii-dimensional subcomplex PP of 𝕋Nd\mathbb{T}^{d}_{N} containing the full (i1)(i-1)-skeleton

𝐛i𝐛i1=η(P)+c15,\displaystyle\mathbf{b}_{i}-\mathbf{b}_{i-1}=\eta\left(P\right)+c_{\ref{const:5}}\,, (7)

where η(P)\eta\left(P\right) denotes the number of ii-cells of P.P.

Proof.

By the Euler–Poincaré formula

j=0d(1)j𝐛j=j=0i1(1)j|FNj|+η(ω)\displaystyle\sum_{j=0}^{d}\left(-1\right)^{j}\mathbf{b}_{j}=\sum_{j=0}^{i-1}\left(-1\right)^{j}\left|F_{N}^{j}\right|+\eta\left(\omega\right) (8)

and the desired statement follows from the observations that |FNj|\left|F_{N}^{j}\right| only depends on NN and jj for 1ji1,1\leq j\leq i-1, and that 𝐛j\mathbf{b}_{j} also only depends on NN and jj when 1ji2.1\leq j\leq i-2.

4. The Plaquette Random-Cluster Model

In this section we will investigate the properties of the plaquette random-cluster model. For the reader’s convenience, we recall that the ii-dimensional plaquette random-cluster model has the distribution defined by

μX(P)=μX,p,q,i(P)1Zpη(P)(1p)|Xi|η(P)q𝐛i1(P),\displaystyle\mu_{X}\left(P\right)=\mu_{X,p,q,i}\left(P\right)\coloneqq\frac{1}{Z}p^{\eta\left(P\right)}\left(1-p\right)^{\left|X^{i}\right|-\eta\left(P\right)}q^{\mathbf{b}_{i-1}\left(P\right)}\,,

where Z=Z(X,p,q,i)Z=Z\left(X,p,q,i\right) is a normalizing constant [HS16].

Similarly to classical percolation, the random complex can alternatively be viewed as a a function designating plaquettes as open or closed. Certain probabilistic arguments are easier to present in this way, so we will introduce notation for this perspective here as well. We call a function ω:Xi{0,1}\omega:X^{i}\to\left\{0,1\right\} a configuration, and we refer to the elements of ω1{1}\omega^{-1}\left\{1\right\} as open plaquettes and the elements of ω1{0}\omega^{-1}\left\{0\right\} as closed plaquettes of ω,\omega, respectively. We then define the associated complex P=P(ω)P=P\left(\omega\right) to be the union of the (i1)(i-1)-skeleton of XX and the open plaquettes of ω.\omega.

4.1. Positive Association

Perhaps the most important property of the random-cluster model is the fact that it satisfies the FKG inequality. Hiraoka and Shirai showed that the plaquette random-cluster model also has this property, which we will use extensively. Since their proof is short, we reproduce it here for completeness.

Theorem 16 (Hiraoka and Shirai).

Let p(0,1),p\in(0,1), and q1,q\geq 1, and XX a finite cubical complex. Then μX\mu_{X} satisfies the FKG lattice condition and is thus positively associated, meaning that for any events E,FE,F that are increasing with respect to ω,\omega,

μX(EF)μX(E)μX(F).\mu_{X}\left(E\cap F\right)\geq\mu_{X}\left(E\right)\mu_{X}\left(F\right).
Proof.

The key observation is that for any topological spaces A,BA,B and any k0,k\in\mathbb{Z}_{\geq 0},

𝐛k(AB)+𝐛k(AB)𝐛k(A)+𝐛k(B).\mathbf{b}_{k}\left(A\cap B\right)+\mathbf{b}_{k}\left(A\cup B\right)\geq\mathbf{b}_{k}\left(A\right)+\mathbf{b}_{k}\left(B\right)\,. (9)

To see this, consider a portion of the Mayer-Vietoris exact sequence (refer to Section 2.2 of [Hat02]):

Hk(AB)Hk(A)Hk(B)Hk(AB)Hk1(AB)φχ.\leavevmode\hbox to316.76pt{\vbox to17.42pt{\pgfpicture\makeatletter\hbox{\hskip 158.38168pt\lower-8.75955pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-158.38168pt}{-8.65971pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 27.65167pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-23.34613pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{k}\left(A\cap B\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 27.65167pt\hfil&\hfil\hskip 61.09334pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-32.78783pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{k}\left(A\right)\oplus H_{k}\left(B\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 37.09337pt\hfil&\hfil\hskip 51.65164pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-23.34613pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{k}\left(A\cup B\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 27.65167pt\hfil&\hfil\hskip 53.98499pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-25.67947pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${H_{k-1}\left(A\cap B\right)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 29.98502pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-102.87834pt}{-6.15971pt}\pgfsys@lineto{-79.67833pt}{-6.15971pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-79.47835pt}{-6.15971pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-93.36794pt}{-2.44586pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\varphi}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-4.69164pt}{-6.15971pt}\pgfsys@lineto{18.50838pt}{-6.15971pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{18.70836pt}{-6.15971pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{4.91844pt}{-2.44586pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\chi}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{74.61168pt}{-6.15971pt}\pgfsys@lineto{97.81169pt}{-6.15971pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{98.01167pt}{-6.15971pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{84.55351pt}{-3.80695pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\partial}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}\,.

The first isomorphism theorem gives

𝐛k(AB)=rankφ+dimkerφ,\displaystyle\mathbf{b}_{k}\left(A\cap B\right)=\operatorname{rank}\varphi+\dim\ker\varphi\,,
𝐛k(A)+𝐛k(B)=rankχ+dimkerχ,\displaystyle\mathbf{b}_{k}\left(A\right)+\mathbf{b}_{k}\left(B\right)=\operatorname{rank}\chi+\dim\ker\chi\,,
𝐛k(AB)=rank+dimker.\displaystyle\mathbf{b}_{k}\left(A\cup B\right)=\operatorname{rank}\partial+\dim\ker\partial\,.

By exactness, rankφ=dimkerχ\operatorname{rank}\varphi=\dim\ker\chi and rankχ=dimker,\operatorname{rank}\chi=\dim\ker\partial, so putting everything together yields

𝐛k(AB)+𝐛k(AB)𝐛k(A)𝐛k(B)=dimkerφ+rank0.\mathbf{b}_{k}\left(A\cap B\right)+\mathbf{b}_{k}\left(A\cup B\right)-\mathbf{b}_{k}\left(A\right)-\mathbf{b}_{k}\left(B\right)=\dim\ker\varphi+\operatorname{rank}\partial\geq 0\,.

Now since μX\mu_{X} is strictly positive for p(0,1),p\in(0,1), it is enough to check the FKG lattice condition, which, written in terms of configurations, requires that

μX(ωω)μX(ωω)μX(ω)μX(ω)\mu_{X}\left(\omega\vee\omega^{\prime}\right)\mu_{X}\left(\omega\wedge\omega^{\prime}\right)\geq\mu_{X}\left(\omega\right)\mu_{X}\left(\omega^{\prime}\right)

for any configurations ω,ω.\omega,\omega^{\prime}. Since

η(ωω)+η(ωω)=η(ω)+η(ω),\eta\left(\omega\vee\omega^{\prime}\right)+\eta\left(\omega\wedge\omega^{\prime}\right)=\eta\left(\omega\right)+\eta\left(\omega^{\prime}\right)\,,

we only need to check that

𝐛i1(PP)+𝐛i1(PP)𝐛i1(P)+𝐛i1(P),\mathbf{b}_{i-1}\left(P\cap P^{\prime}\right)+\mathbf{b}_{i-1}\left(P\cup P^{\prime}\right)\geq\mathbf{b}_{i-1}\left(P\right)+\mathbf{b}_{i-1}\left(P^{\prime}\right)\,,

where PP^{\prime} is the complex associated to ω\omega^{\prime} (where we are using the definition of the plaquette random-cluster model). This inequality is a special case of Equation 9.

4.2. The Plaquette Random-Cluster Model in Infinite Volume

In this section we will show that, like the classical random-cluster model, the plaquette random-cluster model can be extended to an infinite volume setting. Specifically, the plaquette random-cluster model on d\mathbb{Z}^{d} will be defined via limits of plaquette random-cluster models on finite boxes Λn[n,n]dd.\Lambda_{n}\coloneqq[-n,n]^{d}\cap\mathbb{Z}^{d}. The proof given in Chapter 4 of [Gri06] only requires minor changes to extend to higher dimensions, which are described here.

We will need to consider both free and wired boundary conditions for the plaquette random-cluster model in a subspace. A more general discussion of boundary conditions is given in Appendix A. The free boundary measure μΛn𝐟=μΛn,p,q,i𝐟\mu_{\Lambda_{n}}^{\mathbf{f}}=\mu_{\Lambda_{n},p,q,i}^{\mathbf{f}} can be defined as the i-random-cluster model on the subcomplex with vertices Λn.\Lambda_{n}. The wired boundary measure μΛn𝐰=μΛn,p,q,i𝐰\mu_{\Lambda_{n}}^{\mathbf{w}}=\mu_{\Lambda_{n},p,q,i}^{\mathbf{w}} can be defined as the same measure conditioned on all plaquettes contained in [n,n]d\partial[-n,n]^{d} being open. By an analogue of Theorem 4.19 of [Gri06], the weak limits of each of these measures exist.

Proposition 17.

Let p[0,1]p\in\left[0,1\right] and q1.q\geq 1.

  1. (a)

    The limits μd𝐟limnμΛn𝐟\mu^{\mathbf{f}}_{\mathbb{Z}^{d}}\coloneqq\lim_{n\to\infty}\mu_{\Lambda_{n}}^{\mathbf{f}} and μd𝐰limnμΛn𝐰\mu_{\mathbb{Z}^{d}}^{\mathbf{w}}\coloneqq\lim_{n\to\infty}\mu_{\Lambda_{n}}^{\mathbf{w}} exist.

  2. (b)

    μd𝐟\mu^{\mathbf{f}}_{\mathbb{Z}^{d}} and μd𝐰\mu^{\mathbf{w}}_{\mathbb{Z}^{d}} are automorphism invariant.

  3. (c)

    μd𝐟\mu^{\mathbf{f}}_{\mathbb{Z}^{d}} and μd𝐰\mu^{\mathbf{w}}_{\mathbb{Z}^{d}} are positively associated.

  4. (d)

    μd𝐟\mu^{\mathbf{f}}_{\mathbb{Z}^{d}} and μd𝐰\mu^{\mathbf{w}}_{\mathbb{Z}^{d}} are tail-trivial, that is, for any event AA that is independent of the states of any finite set of plaquettes, μd𝐟(A),μd𝐰(A){0,1}.\mu^{\mathbf{f}}_{\mathbb{Z}^{d}}\left(A\right),\mu^{\mathbf{w}}_{\mathbb{Z}^{d}}\left(A\right)\in\left\{0,1\right\}.

  5. (e)

    For any nontrivial translation 𝐭,\mathbf{t}, μd𝐟\mu^{\mathbf{f}}_{\mathbb{Z}^{d}} and μd𝐰\mu^{\mathbf{w}}_{\mathbb{Z}^{d}} are 𝐭\mathbf{t}-ergodic, that is, any 𝐭\mathbf{t}-invariant random variable is μd𝐟\mu^{\mathbf{f}}_{\mathbb{Z}^{d}}-almost-surely and μd𝐰\mu^{\mathbf{w}}_{\mathbb{Z}^{d}}-almost-surely constant.

Proof.

(a), (b), (d) Each proof is the same as in Theorem 4.19 of [Gri06]

(c) Both measures are limits of positively associated measures by Lemma 49, so they are positively associated by Proposition 4.10 of [Gri06].

(e) The proof is the same as in Corollary 4.23 of [Gri06]. ∎

4.3. Duality for the Random-Cluster Model

In the classical random-cluster model in the plane, the dual graph is also distributed as a random-cluster model with parameters qq and pp^{*} where

p(1p)q(1p)q+p.p^{*}\coloneqq\frac{\left(1-p\right)q}{\left(1-p\right)q+p}\,.

Beffara and Duminil-Copin [BDC12] used this relationship to prove that the critical probability is the self-dual point where p=p,p=p^{*}, namely

psdq1+q.\displaystyle p_{\mathrm{sd}}\coloneqq\frac{\sqrt{q}}{1+\sqrt{q}}\,.

In this section, we consider a similar duality in higher dimensions within the torus, showing that the dual of the ii-dimensional plaquette random-cluster model is approximately a (di)(d-i)-dimensional random-cluster model. We will see below that this recovers the duality properties of Potts lattice gauge theory as observed in [KPSS80], but in a way that perhaps provides more geometric intuition. Our proof makes heavy use of the topological duality results in Section 3, especially Theorem 14. We will use the notation from that theorem.

Since it will be important to distinguish between giant and local cycles, we recall their definitions. Let ϕ:PX\phi:P\hookrightarrow X be the inclusion map and let ϕ:Hi(P)Hi(X)\phi_{*}:H_{i}\left(P\right)\to H_{i}\left(X\right) be the induced map on iith homology. We say that an element αHi(P)\alpha\in H_{i}\left(P\right) is a giant cycle if I(α)0I_{*}\left(\alpha\right)\neq 0 and a local cycle if i(α)=0.i_{*}\left(\alpha\right)=0. Recall that bi(P)=rankϕb_{i}\left(P\right)=\operatorname{rank}\phi_{*} is informally the number of giant cycles in P.P.

First we define a \saybalanced version [BDC12] which satisfies exact duality:

μ~𝕋Nd(P)(q)bi(P)Z~pη(P)(1p)|FNi|η(P)q𝐛i1(P).\displaystyle\tilde{\mu}_{\mathbb{T}^{d}_{N}}\left(P\right)\coloneqq\frac{\left(\sqrt{q}\right)^{-b_{i}\left(P\right)}}{\tilde{Z}}p^{\eta\left(P\right)}\left(1-p\right)^{\left|F^{i}_{N}\right|-\eta\left(P\right)}q^{\mathbf{b}_{i-1}\left(P\right)}\,.

The additional term (q)bi(P)\left(\sqrt{q}\right)^{-b_{i}\left(P\right)} term “corrects” for the different behavior of local and global cycles under duality. The two models are absolutely continuous with respect to each other with a Radon-Nikodym derivatives bounded above and below by functions of qq and the same is true for their dual models. As such, a sharp threshold for μ~𝕋Nd\tilde{\mu}_{\mathbb{T}^{d}_{N}} implies one for μ~𝕋Nd.\tilde{\mu}_{\mathbb{T}^{d}_{N}}.

Theorem 18.

The balanced plaquette random-cluster model satisfies

μ~𝕋Nd,p,q,i(P)=dμ~𝕋Nd,p,q,di(P).\tilde{\mu}_{\mathbb{T}^{d}_{N},p,q,i}\left(P\right)\buildrel d\over{=}\tilde{\mu}_{\mathbb{T}^{d}_{N},p^{*},q,d-i}\left(P^{\bullet}\right)\,.
Proof.

The idea of the proof is the same as in the classical random-cluster model. We need only take care to keep track of the giant cycles and local cycles separately, since they behave differently under duality. Note that

η(P)+η(P)=|FNi|.\displaystyle\eta\left(P\right)+\eta\left(P^{\bullet}\right)=\left|F_{N}^{i}\right|\,. (10)

Recall also that PP and PP^{\bullet} contain the complete (i1)(i-1)-skeleton and (di1)(d-i-1)-skeleton respectively, so

bi1=(di1),bdi1=(ddi1).\displaystyle b_{i-1}=\binom{d}{i-1},\qquad b_{d-i-1}=\binom{d}{d-i-1}\,. (11)

It is not crucial to the argument, but we can simplify the upcoming calculation slightly to note from the bijection between plaquettes and dual plaquettes that

|FNi|=|FNdi|.\displaystyle\left|F^{i}_{N}\right|=\left|F^{d-i}_{N}\right|\,. (12)

Lastly, we recall the following property of p,p^{*},

pp(1p)(1p)=q.\displaystyle\frac{pp^{*}}{\left(1-p\right)\left(1-p^{*}\right)}=q\,. (13)

For convenience, in the following calculation we will denote C|FNi|=|FNdi|C\coloneqq\left|F^{i}_{N}\right|=\left|F^{d-i}_{N}\right| (equal by Equation 12). Now we compute μ~𝕋Nd,p,q,i(P)=\tilde{\mu}_{\mathbb{T}^{d}_{N},p,q,i}\left(P\right)=

(q)biZ~pη(P)(1p)|FNi|η(P)q𝐛i1\displaystyle\phantom{=}\frac{\left(\sqrt{q}\right)^{-b_{i}}}{\tilde{Z}}p^{\eta\left(P\right)}\left(1-p\right)^{\left|F^{i}_{N}\right|-\eta\left(P\right)}q^{\mathbf{b}_{i-1}}
=(1p)|FNi|Z~(q)bi(p1p)η(P)q𝐛i1\displaystyle=\frac{\left(1-p\right)^{\left|F^{i}_{N}\right|}}{\tilde{Z}}\left(\sqrt{q}\right)^{-b_{i}}\left(\frac{p}{1-p}\right)^{\eta\left(P\right)}q^{\mathbf{b}_{i-1}}
=qc15(1p)|FNi|Z~(q)bi(pq(1p))η(P)q𝐛i\displaystyle=\frac{q^{c_{\ref{const:5}}}\left(1-p\right)^{\left|F^{i}_{N}\right|}}{\tilde{Z}}\left(\sqrt{q}\right)^{-b_{i}}\left(\frac{p}{q\left(1-p\right)}\right)^{\eta\left(P\right)}q^{\mathbf{b}_{i}}\; by (7)
=qc15(1p)|FNi|Z~(q)bi(q(1p)p)η(P)qai+bi\displaystyle=\frac{q^{c_{\ref{const:5}}}\left(1-p\right)^{\left|F^{i}_{N}\right|}}{\tilde{Z}}\left(\sqrt{q}\right)^{-b_{i}}\left(\frac{q(1-p)}{p}\right)^{-\eta\left(P\right)}q^{a_{i}+b_{i}} by (4)
=qc15(1p)|FNi|Z~(q)bi(p1p)η(P)qai+bi\displaystyle=\frac{q^{c_{\ref{const:5}}}\left(1-p\right)^{\left|F^{i}_{N}\right|}}{\tilde{Z}}\left(\sqrt{q}\right)^{-b_{i}}\left(\frac{p^{*}}{1-p^{*}}\right)^{-\eta\left(P\right)}q^{a_{i}+b_{i}} by (13)
=qc15+(di)/2(1p)|FNi|Z~(q)bdi(p1p)η(P)qadi1bdi\displaystyle=\frac{q^{c_{\ref{const:5}}+\binom{d}{i}/2}\left(1-p\right)^{\left|F^{i}_{N}\right|}}{\tilde{Z}}\left(\sqrt{q}\right)^{b^{\bullet}_{d-i}}\left(\frac{p^{*}}{1-p^{*}}\right)^{-\eta\left(P\right)}q^{a_{d-i-1}^{\bullet}-b_{d-i}^{\bullet}} by (5), (6)
=qc15+(di)/2(1p)|FNi|Z~(q)bdi(p1p)η(P)|FNi|qadi1bdi\displaystyle=\frac{q^{c_{\ref{const:5}}+\binom{d}{i}/2}\left(1-p\right)^{\left|F^{i}_{N}\right|}}{\tilde{Z}}\left(\sqrt{q}\right)^{b^{\bullet}_{d-i}}\left(\frac{p^{*}}{1-p^{*}}\right)^{\eta\left(P^{\bullet}\right)-\left|F^{i}_{N}\right|}q^{a_{d-i-1}^{\bullet}-b_{d-i}^{\bullet}} by (10)
=qc15+(di)/2(1p)|FNi|qbdi1Z~(q)bdi(p1p)η(P)|FNi|q𝐛di1\displaystyle=\frac{q^{c_{\ref{const:5}}+\binom{d}{i}/2}\left(1-p\right)^{\left|F^{i}_{N}\right|}}{q^{b^{\bullet}_{d-i-1}}\tilde{Z}}\left(\sqrt{q}\right)^{-b^{\bullet}_{d-i}}\left(\frac{p^{*}}{1-p^{*}}\right)^{\eta\left(P^{\bullet}\right)-\left|F^{i}_{N}\right|}q^{\mathbf{b}_{d-i-1}^{\bullet}} by (4), (5)
=qc15+(di)/2(1p)|FNi|q(ddi1)p|FNi|Z~(q)bdi(p)η(P)(1p)η(P)|FNdi|q𝐛di1\displaystyle=\frac{q^{c_{\ref{const:5}}+\binom{d}{i}/2}\left(1-p\right)^{\left|F^{i}_{N}\right|}}{q^{\binom{d}{d-i-1}}p^{\left|F^{i}_{N}\right|}\tilde{Z}}\left(\sqrt{q}\right)^{-b^{\bullet}_{d-i}}\frac{\left(p^{*}\right)^{\eta\left(P^{\bullet}\right)}}{\left(1-p^{*}\right)^{\eta\left(P^{\bullet}\right)-\left|F^{d-i}_{N}\right|}}q^{\mathbf{b}_{d-i-1}^{\bullet}} by (11)
(q)bdiZ~(p)η(P)(1p)|FNdi|η(P)q𝐛di1\displaystyle\coloneqq\frac{\left(\sqrt{q}\right)^{-b^{\bullet}_{d-i}}}{\tilde{Z}^{\bullet}}\left(p^{*}\right)^{\eta\left(P^{\bullet}\right)}\left(1-p^{*}\right)^{\left|F^{d-i}_{N}\right|-\eta\left(P^{\bullet}\right)}q^{\mathbf{b}_{d-i-1}^{\bullet}}
=μ~𝕋Nd,p,q,di(P),\displaystyle=\tilde{\mu}_{\mathbb{T}^{d}_{N},p^{*},q,d-i}\left(P^{\bullet}\right)\,,

where we used the definition of the (di)(d-i)-dimensional random-cluster model in the last step. ∎

As a corollary, we have the following duality relationship between the normalizing constants.

Corollary 19.
Z(p,q,di,N)=qc+(di)/2(ddi1)(1p)|FNi|Z(p,q,i,N),Z(p^{*},q,d-i,N)=q^{c+\binom{d}{i}/2-\binom{d}{d-i-1}}\left(1-p\right)^{\left|F^{i}_{N}\right|}Z(p,q,i,N)\,,

where cc is the constant c15c_{\ref{const:5}} defined in Proposition 15.

5. Relation to Potts Lattice Gauge Theory

We now elucidate the relationship between the ii-dimensional plaquette random-cluster and (i1)(i-1)-dimensional Potts lattice gauge theory. In particular, we find a topological interpretation of the generalized Wilson loop variables. First, we review the coupling between these models, which generalizes the Edwards–Sokal coupling between the classical random-cluster model and the Potts model [FK72, ES88].

Proposition 20 (Hiraoka and Shirai [HS16]).

Let XX be a finite cubical complex, qq be a prime, p[0,1),p\in[0,1), and p=1eβ.p=1-e^{-\beta}. Consider the coupling on Ci1(X)×{0,1}XiC^{i-1}\left(X\right)\times\left\{0,1\right\}^{X^{i}} defined by

κ(f,P(ω))σXi[(1p)K(ω(σ),0)+pK(ω(σ),1)K(δf(σ),0)],\kappa\left(f,P\left(\omega\right)\right)\propto\prod_{\sigma\in X^{i}}\left[\left(1-p\right)K\left(\omega\left(\sigma\right),0\right)+pK\left(\omega\left(\sigma\right),1\right)K\left(\delta f\left(\sigma\right),0\right)\right]\,,

where K(x,y)K(x,y) is the Kronecker delta function. Then κ\kappa has the following marginals.

  • The first marginal is the qq-state Potts lattice gauge theory with inverse temperature β\beta given by

    Pκ(f,P)eβH(f),\sum_{P}\kappa\left(f,P\right)\propto e^{-\beta H(f)}\,,

    where H(f)H(f) is the Hamiltonian for Potts lattice gauge theory:

    H(f)=σK(δf(σ),1).H\left(f\right)=\sum_{\sigma}K\left(\delta f\left(\sigma\right),1\right)\,.
  • The second marginal is the plaquette random-cluster model with parameters p,qp,q given by

    fCi1(X)κ(f,P)pη(P)(1p)|Xi|η(P)q𝐛i1(P;𝔽q).\sum_{f\in C^{i-1}\left(X\right)}\kappa\left(f,P\right)\propto p^{\eta\left(P\right)}\left(1-p\right)^{\left|X^{i}\right|-\eta\left(P\right)}q^{\mathbf{b}_{i-1}\left(P;\,\mathbb{F}_{q}\right)}\,.
Proof.

We include Hiraoka and Shirai’s proof adapted to our notation for completeness. The first marginal is calculated as

κ1(f)\displaystyle\kappa_{1}\left(f\right) ω{0,1}Xiκ(f,P(ω))\displaystyle\coloneqq\sum_{\omega\in\left\{0,1\right\}^{X^{i}}}\kappa\left(f,P\left(\omega\right)\right)
ω{0,1}XiσXi[(1p)K(ω(σ),0)+pK(ω(σ),1)K(δf(σ),0)]\displaystyle\propto\sum_{\omega\in\left\{0,1\right\}^{X^{i}}}\prod_{\sigma\in X^{i}}\left[\left(1-p\right)K\left(\omega\left(\sigma\right),0\right)+pK\left(\omega\left(\sigma\right),1\right)K\left(\delta f\left(\sigma\right),0\right)\right]
=σXi[(1p)+pK(δf(σ),0)]\displaystyle=\prod_{\sigma\in X^{i}}\left[\left(1-p\right)+pK\left(\delta f\left(\sigma\right),0\right)\right]
=σXi[eβ+(1eβ)K(δf(σ),0)]\displaystyle=\prod_{\sigma\in X^{i}}\left[e^{-\beta}+\left(1-e^{-\beta}\right)K\left(\delta f\left(\sigma\right),0\right)\right]
=eβ|Xi|σXi[1+(eβ1)K(δf(σ),0)]\displaystyle=e^{-\beta\left|X^{i}\right|}\prod_{\sigma\in X^{i}}\left[1+\left(e^{\beta}-1\right)K\left(\delta f\left(\sigma\right),0\right)\right]
=eβ|Xi|eβH(f)\displaystyle=e^{-\beta\left|X^{i}\right|}e^{-\beta H(f)}
eβH(f).\displaystyle\propto e^{-\beta H(f)}\,.

The second marginal is calculated as

κ2(P)\displaystyle\kappa_{2}\left(P\right) fCi1(X)κ(f,P(ω))\displaystyle\coloneqq\sum_{f\in C^{i-1}\left(X\right)}\kappa\left(f,P\left(\omega\right)\right)
fCi1(X)σXi[(1p)K(ω(σ),0)+pK(ω(σ),1)K(δf(σ),0)]\displaystyle\propto\sum_{f\in C^{i-1}\left(X\right)}\prod_{\sigma\in X^{i}}\left[\left(1-p\right)K\left(\omega\left(\sigma\right),0\right)+pK\left(\omega\left(\sigma\right),1\right)K\left(\delta f\left(\sigma\right),0\right)\right]
=(1p)|Xi|η(P)pη(P)fCi1(X)σXiω(σ)=1K(δf(σ),0)\displaystyle=\left(1-p\right)^{\left|X^{i}\right|-\eta\left(P\right)}p^{\eta\left(P\right)}\sum_{f\in C^{i-1}\left(X\right)}\prod_{\begin{subarray}{c}\sigma\in X^{i}\\ \omega\left(\sigma\right)=1\end{subarray}}K\left(\delta f\left(\sigma\right),0\right)
=(1p)|Xi|η(P)pη(P)qdimZi1(P;𝔽q)\displaystyle=\left(1-p\right)^{\left|X^{i}\right|-\eta\left(P\right)}p^{\eta\left(P\right)}q^{\dim Z^{i-1}\left(P;\,\mathbb{F}_{q}\right)}
(1p)|Xi|η(P)pη(P)q𝐛i1(P;𝔽q).\displaystyle\propto\left(1-p\right)^{\left|X^{i}\right|-\eta\left(P\right)}p^{\eta\left(P\right)}q^{\mathbf{b}_{i-1}\left(P;\,\mathbb{F}_{q}\right)}\,.

The second to last line follows from the definition of a cocycle. The last line holds because the (i1)(i-1)-skeleton of PP does not depend on ω,\omega, so Bi1(P;𝔽q)B^{i-1}\left(P;\,\mathbb{F}_{q}\right) is fixed. ∎

Next, we compute the conditional measures defined by the coupling.

Proposition 21.

Let p=1eβ.p=1-e^{-\beta}. Then the conditional measures of κ\kappa are as follows.

  • Given f,f, the conditional measure κ(f)\kappa\left(\cdot\mid f\right) is Bernoulli plaquette percolation with probability pp on the set of plaquettes σ\sigma that satisfy δf(σ)=0.\delta f\left(\sigma\right)=0.

  • Given P,P, the conditional measure κ(P)\kappa\left(\cdot\mid P\right) is the uniform measure on (i1)\left(i-1\right)-cocycles in Zi1(P;𝔽q).Z^{i-1}\left(P;\,\mathbb{F}_{q}\right).

Proof.

From the definition of the coupling, under κ(s)\kappa\left(\cdot\mid s\right) a plaquette σ\sigma is open with probability pp independently of other plaquettes when δf(σ)=0\delta f\left(\sigma\right)=0 and always closed otherwise, giving the first conditional measure. The second conditional measure is determined by the observation that κ(P)\kappa\left(\cdot\mid P\right) is supported on the set of fCi1(X)f\in C^{i-1}\left(X\right) satisfying δf(σ)=0\delta f\left(\sigma\right)=0 for each σP,\sigma\in P, and that each such cochain has the same weight. ∎

In the next section, we discuss how to use this proposition to sample Potts lattice gauge theory.

5.1. The Plaquette Swendsen–Wang Algorithm

The classical Swendsen–Wang algorithm was the first non-local Monte Carlo algorithm for the Potts model, which alternately samples from the marginals in the combined random-cluster representation [SW87, ES88]. That is, given a coupling constant β,\beta, a spin configuration fC0(X;𝔽q)f\in C^{0}\left(X;\,\mathbb{F}_{q}\right) is updated by first sampling a random graph GG by performing percolation with probability p=11βp=1-1^{-\beta} on the edges between vertices with equal spins. Then, the spins are resampled uniformly on each component of G.G. In our language, this corresponds to sampling a uniformly random cocycle in Z0(G;𝔽q).Z^{0}\left(G;\,\mathbb{F}_{q}\right). The Swendsen–Wang algorithm is observed to converge significantly faster than Glauber dynamics, especially at and near criticality [SW87].

We generalize the Swendsen–Wang algorithm to sample Potts lattice gauge theory using the coupled representation in Proposition 20. That is, given a cochain fCi1(X;𝔽q),f\in C^{i-1}\left(X;\,\mathbb{F}_{q}\right), a random ii-complex PP is sampled by including plaquettes on which δf\delta f vanishes independently with probability p=1eβ.p=1-e^{-\beta}. An updated cochain is then found by selecting a uniformly random cocycle from Zi1(P;𝔽q).Z^{i-1}\left(P;\,\mathbb{F}_{q}\right). It is easily seen that the resulting Markov chain is ergodic and satisfies detailed balance, with the stationary distribution being Potts lattice gauge theory. An implementation of this algorithm will be described in [DS].

When d=2id=2i and qq is an odd prime, we can interpret Theorem 8 in terms of the qualitative behavior of the Swendsen–Wang algorithm on the torus at the self-dual point. Below the threshold, the behavior of the algorithm is “local” in the sense that all co-cycles in Zi(P;𝔽q)Z^{i}\left(P;\,\mathbb{F}_{q}\right) are trivial in the homology of the torus. On the other hand, when p>pc,p>p_{c}, Zi(P;𝔽q)Z^{i}\left(P;\,\mathbb{F}_{q}\right) contains (di)\binom{d}{i} giant cocycles up to cohomology (each given the same weight), and a “non–local” move happens with probability approaching 1q(di)1-q^{-\binom{d}{i}} (as PP has (di)\binom{d}{i} giant cocycles, with high probability)444While we state our theorems in terms of giant cycles, the analogous results for giant cocycles follow immediately from them, either by the functoriality of Hom(,𝔽q)\mathrm{Hom}\left(\cdot,\mathbb{F}_{q}\right) or by repeating the same proofs.. Interestingly, at p=pc,p=p_{c}, the number of giant cocycles is somewhere between 0 and (di)\binom{d}{i} with probability bounded away from 0 and 1,1, so the probability of non-local behavior is non-zero but lower than above pc.p_{c}. Theorems 10 and 11 have a similar interpretation.

5.2. Potts Lattice Gauge Theory in Infinite Volume

We can use the infinite volume random-cluster measure to define the corresponding infinite volume Potts lattice gauge theory. We start with finite volume measure νΛn𝐟=νΛn,β,q,i1,d𝐟,\nu^{\mathbf{f}}_{\Lambda_{n}}=\nu^{\mathbf{f}}_{\Lambda_{n},\beta,q,i-1,d}, which is defined to be the usual Potts lattice gauge theory on the subcomplex induced by the vertices of Λn\Lambda_{n} (we use the superscript 𝐟\mathbf{f} because this measure can be coupled with the free random-cluster measure μΛn𝐟\mu_{\Lambda_{n}}^{\mathbf{f}} as in Proposition 20). Next, we construct the limiting measure ν𝐟=νβ,q,i1,d𝐟\nu^{\mathbf{f}}=\nu^{\mathbf{f}}_{\beta,q,i-1,d} by taking n.n\to\infty. A similar construction can be done with wired boundary conditions, but we will omit the details here. We start with an easy lemma that will allow us to choose random cocycles in increasing subcomplexes of PP in a consistent manner.

Lemma 22.

For any configuration ωΩ,\omega\in\Omega, there is a basis \mathcal{B} of Zi1(P(ω),F)Z^{i-1}\left(P\left(\omega\right),F\right) so that each (i1)(i-1)-face of d\mathbb{Z}^{d} is in the support of finitely many elements of .\mathcal{B}.

Proof.

Take any basis and perform Gaussian elimination. ∎

We will call the minimal such basis with respect to lexicographical order ω\mathcal{B}_{\omega}

Corollary 23.

Let p[0,1),q{2,3,},p\in[0,1),q\in\left\{2,3,\ldots\right\}, and p=1eβ.p=1-e^{-\beta}.

  1. (a)

    Let ω\omega be distributed according to μd𝐟.\mu^{\mathbf{f}}_{\mathbb{Z}^{d}}. Conditional on ω,\omega, let {Ag:gω}\left\{A_{g}:g\in\mathcal{B}_{\omega}\right\} be i.i.d. Unif(𝔽q)\mathrm{Unif}\left(\mathbb{F}_{q}\right) random variables. Then the limit

    ν𝐟limnνΛn𝐟\nu^{\mathbf{f}}\coloneqq\lim_{n\to\infty}\nu^{\mathbf{f}}_{\Lambda_{n}}

    exists, and the random cocycle

    f=gAggf=\sum_{g\in\mathcal{B}}A_{g}g

    is distributed according to ν𝐟.\nu^{\mathbf{f}}.

  2. (b)

    Let ff be distributed according to ν𝐟.\nu^{\mathbf{f}}. Conditional on f,f, let ω\omega be a random configuration in which each ii-plaquette σ\sigma is open with probability pp if f(σ)=0f\left(\partial\sigma\right)=0 independent of the states of other plaquettes and closed otherwise. Then ω\omega is distributed according to μd𝐟.\mu^{\mathbf{f}}_{\mathbb{Z}^{d}}.

Proof.

We proceed similarly to proof of Theorem 4.91 of [Gri06].

(a) By the proof of Theorem 4.19 of [Gri06], there is an increasing set of configurations ωn\omega_{n} so that each ωn\omega_{n} is distributed according to μΛn𝐟\mu_{\Lambda_{n}}^{\mathbf{f}} and limnωn\lim_{n\to\infty}\omega_{n} is distributed according to μd𝐟.\mu_{\mathbb{Z}^{d}}^{\mathbf{f}}. Moreover, for any ii-plaquette σ,\sigma, ωn(σ)=ω(σ)\omega_{n}\left(\sigma\right)=\omega\left(\sigma\right) for large enough n.n.

Now let {Ag:gCi1(d;𝔽q)}\left\{A_{g}:g\in C^{i-1}\left(\mathbb{Z}^{d};\,\mathbb{F}_{q}\right)\right\} be i.i.d. Unif(𝔽q)\mathrm{Unif}\left(\mathbb{F}_{q}\right) random variables, and let

fngωnAgg.f_{n}\coloneqq\sum_{g\in\mathcal{B}_{\omega_{n}}}A_{g}g\,.

By the construction of ωn,\mathcal{B}_{\omega_{n}}, fnf_{n} is distributed according to ν𝐟\nu^{\mathbf{f}} and is eventually constant on any finite set of (i1)(i-1)-cubes. Since ωnω,\mathcal{B}_{\omega_{n}}\to\mathcal{B}_{\omega}, limnfn\lim_{n\to\infty}f_{n} is distributed as ν𝐟\nu^{\mathbf{f}} and we are done.

(b) Let {Bσ:σFi}\left\{B_{\sigma}:\sigma\in F^{i}\right\} be independent Ber(p)\mathrm{Ber}\left(p\right) random variables. Also, let fnf_{n} be distributed as νΛn𝐟,\nu^{\mathbf{f}}_{\Lambda_{n}}, coupled as before so that fnf_{n} is eventually constant on any finite set of (i1)(i-1)-cubes. Then, Proposition 21, ωn(σ)BσK(fn(σ),1)\omega_{n}\left(\sigma\right)\coloneqq B_{\sigma}K\left(f_{n}\left(\partial\sigma\right),1\right) is distributed according to μΛn𝐟.\mu_{\Lambda_{n}}^{\mathbf{f}}. Therefore ωlimnωn\omega\coloneqq\lim_{n\to\infty}\omega_{n} is distributed according to μd𝐟=limnμΛn𝐟.\mu_{\mathbb{Z}^{d}}^{\mathbf{f}}=\lim_{n\to\infty}\mu_{\Lambda_{n}}^{\mathbf{f}}.

In the remainder of this section, we will refer to “Potts lattice gauge theory on d\mathbb{Z}^{d}” and “the plaquette random-cluster model on d\mathbb{Z}^{d}” without explicit reference to the boundary conditions used to construct the infinite volume limit. Note that the constants defined below may depend on this choice.

5.3. Duality for Potts Lattice Gauge Theory

Another corollary of the coupling between Potts lattice gauge theory and the plaquette random-cluster model is a new proof of duality for the partition functions of Potts lattice gauge theory. This duality was first observed in the paper that introduced Potts lattice gauge theory [KPSS80].

Corollary 24.

Let 𝒵(𝕋Nd,q,β,i)\mathcal{Z}\left(\mathbb{T}^{d}_{N},q,\beta,i\right) be the normalizing constant for qq-state i1i-1-dimensional Potts lattice gauge theory on 𝕋Nd.\mathbb{T}^{d}_{N}. Then, there are constants 0<c8,c9<0<c_{8},c_{9}<\infty so that

c24𝒵(𝕋Nd,q,β,i)𝒵(𝕋Nd,q,β,di)c24𝒵(𝕋Nd,q,β,i)c_{\ref{const:Z1}}\mathcal{Z}\left(\mathbb{T}^{d}_{N},q,\beta,i\right)\leq\mathcal{Z}\left(\mathbb{T}^{d}_{N},q,\beta^{*},d-i\right)\leq c_{\ref{const:Z2}}\mathcal{Z}\left(\mathbb{T}^{d}_{N},q,\beta,i\right)

where

β=β(β,q)=log(eβ+q1eβ1).\beta^{*}=\beta^{*}\left(\beta,q\right)=\log\left(\frac{e^{\beta}+q-1}{e^{\beta}-1}\right)\,.
Proof.

This follows from Corollary 19, the relationship between the balanced and unbalanced plaquette random cluster models on the torus, and the computations in the proof of Proposition 20. ∎

Note that the proofs can be adapted to relate the partition function for i1i-1-dimensional Potts lattice gauge theory on a box in d\mathbb{Z}^{d} with free boundary conditions to that of (di1)(d-i-1)-dimensional Potts lattice gauge theory on a box in d\mathbb{Z}^{d} with wired boundary conditions.

5.4. Generalized Wilson Loop Variables and Proof of Theorem 5

In this section, let XX either be a finite cubical complex or d.\mathbb{Z}^{d}. Let γZi1(X,F)\gamma\in Z_{i-1}\left(X,F\right) be an (i1)(i-1)-dimensional cycle. It is natural to ask whether γ\gamma is a boundary in the ii-dimensional plaquette random-cluster model.

Definition 25.

Let VγV_{\gamma} to be the event that γ\gamma is null-homologous in Hi(P;𝔽)H_{i}\left(P;\,\mathbb{F}\right) for a specified coefficient field 𝔽.\mathbb{F}.

The generalized Wilson loop variable is a closely related quantity for Potts lattice gauge theory.

Definition 26.

Let ff be an (i1)(i-1)-cocycle. Recall that the generalized Wilson loop variable associated to γ\gamma is the random variable Wγ:Ci1(X;𝔽q)W_{\gamma}:C^{i-1}\left(X;\,\mathbb{F}_{q}\right)\rightarrow\mathbb{C} given by

Wγ(f)=(f(γ)).W_{\gamma}\left(f\right)=\left(f\left(\gamma\right)\right)^{\mathbb{C}}\,.

We are particularly interested in the behavior of these random variables when γ\gamma is an (i1)(i-1)-dimensional hyperrectangular boundary in d\mathbb{Z}^{d} — that is, the boundary of a hyperrectangle [0,n1]××[0,ni]×𝐱\left[0,n_{1}\right]\times\ldots\times\left[0,n_{i}\right]\times\mathbf{x} inside i\mathbb{R}^{i} where njn_{j}\in\mathbb{Z} for j=1,,ij=1,\ldots,i and 𝐱di,\mathbf{x}\in\mathbb{Z}^{d-i}, or a congruent object obtained translation and/or by permuting the coordinates of d.\mathbb{Z}^{d}. The next several results show a close connection between Wilson loop variables and the topology of the plaquette random-cluster model.

We first show that Wilson loop variables are perfectly correlated if they are homologous, and are otherwise independent. For completeness, we begin by proving an elementary topological lemma which will be useful in a couple places.

Lemma 27.

Let γZi1(X;𝔽q)\gamma\in Z_{i-1}\left(X;\,\mathbb{F}_{q}\right) and fZi1(X;𝔽q).f\in Z^{i-1}\left(X;\,\mathbb{F}_{q}\right). Then for any fZi1(X;𝔽q)f^{\prime}\in Z^{i-1}\left(X;\,\mathbb{F}_{q}\right) so that [f]=[f]\left[f\right]=\left[f^{\prime}\right] in Hi1(X;𝔽q)H^{i-1}\left(X;\,\mathbb{F}_{q}\right) and any γZi1(X;𝔽q)\gamma^{\prime}\in Z_{i-1}\left(X;\,\mathbb{F}_{q}\right) so that [γ]=[γ]\left[\gamma\right]=\left[\gamma^{\prime}\right] in Hi1(X;𝔽q),H_{i-1}\left(X;\,\mathbb{F}_{q}\right), we have f(γ)=f(γ).f\left(\gamma\right)=f^{\prime}\left(\gamma^{\prime}\right).

Proof.

First we show that f(γ)=f(γ).f\left(\gamma\right)=f^{\prime}\left(\gamma\right). If i=1,i=1, f=ff=f^{\prime} and there is nothing to prove. Now, assume i2.i\geq 2. For hFNi2,h\in F^{i-2}_{N}, let hCi2(X;𝔽q)h^{*}\in C^{i-2}\left(X;\,\mathbb{F}_{q}\right) be the dual of hh when hh is viewed as an element of Ci2(X;𝔽q).C_{i-2}\left(X;\,\mathbb{F}_{q}\right). Note that since γ\gamma is a cycle,

δh(γ)=h(γ)=0.\delta h^{*}\left(\gamma\right)=h^{*}\left(\partial\gamma\right)=0\,.

Then, because ff and ff^{\prime} are cohomologous, we can write

ff=hXi2chδhf-f^{\prime}=\sum_{h\in X^{i-2}}c_{h}\delta h^{*}

for some {ch}𝔽qXi2.\left\{c_{h}\right\}\in\mathbb{F}_{q}^{X^{i-2}}. It follows that

f(γ)=f(γ)+hXi2chδh(γ)=f(γ).f\left(\gamma\right)=f^{\prime}\left(\gamma\right)+\sum_{h\in X^{i-2}}c_{h}\delta h^{*}\left(\gamma\right)=f^{\prime}\left(\gamma\right)\,.

Now as [γ]=[γ],\left[\gamma\right]=\left[\gamma^{\prime}\right], there is a chain ρCi(X;𝔽q)\rho\in C_{i}\left(X;\,\mathbb{F}_{q}\right) so that γ=ρ+γ.\gamma=\partial\rho+\gamma^{\prime}. Then since ff^{\prime} is a cocycle, we can calculate

f(γ)\displaystyle f^{\prime}\left(\gamma\right) =f(ρ)+f(γ)\displaystyle=f^{\prime}\left(\partial\rho\right)+f^{\prime}\left(\gamma^{\prime}\right)
=δf(ρ)+f(γ)\displaystyle=\delta f^{\prime}\left(\rho\right)+f^{\prime}\left(\gamma^{\prime}\right)
=f(γ).\displaystyle=f^{\prime}\left(\gamma^{\prime}\right)\,.

We now describe the distribution of Wilson loops as a function of the associated plaquette random-cluster model.

Proposition 28.

Let qq be a prime integer and let γZi1(X;𝔽q).\gamma\in Z_{i-1}\left(X;\,\mathbb{F}_{q}\right). Then

(WγVγ)1\left(W_{\gamma}\mid V_{\gamma}\right)\equiv 1

and

(WγVγc)Unif(𝔽q).\left(W_{\gamma}\mid V_{\gamma}^{c}\right)\sim\mathrm{Unif}\left(\mathbb{F}_{q}\right)^{\mathbb{C}}\,.
Proof.

Fix ω\omega and γ.\gamma. If 0=[γ]Hi1(P;𝔽q),0=\left[\gamma\right]\in H_{i-1}\left(P;\,\mathbb{F}_{q}\right), then f(γ)=0f\left(\gamma\right)=0 by Lemma 27.

Now assume that 0[γ]Hi1(P;𝔽q).0\neq\left[\gamma\right]\in H_{i-1}\left(P;\,\mathbb{F}_{q}\right). By Proposition 21, ν(fω)\nu\left(f\mid\omega\right) can be sampled by fixing a basis of Hi1(P)H^{i-1}\left(P\right) and taking a random linear combination with independent Unif(𝔽q)\mathrm{Unif}\left(\mathbb{F}_{q}\right) coefficients. Thus, f(γ)f\left(\gamma\right) is uniformly distributed on an additive subgroup of 𝔽q.\mathbb{F}_{q}. Since the only such subgroups are 𝔽q\mathbb{F}_{q} and {0},\left\{0\right\}, we only need to rule out the latter. By the universal coefficient theorem for cohomology (Corollary 3.3 of [Hat02]), there is a dual element 0[γ]Hi1(P;𝔽q)0\neq\left[\gamma\right]^{*}\in H^{i-1}\left(P;\,\mathbb{F}_{q}\right) so that [γ]([γ])0.\left[\gamma\right]^{*}\left(\left[\gamma\right]\right)\neq 0. Therefore, f(γ)f\left(\gamma\right) is distributed as Unif(𝔽q),\mathrm{Unif}\left(\mathbb{F}_{q}\right), so WγW_{\gamma} is distributed as Unif(𝔽q).\mathrm{Unif}\left(\mathbb{F}_{q}\right)^{\mathbb{C}}.

The proposition has a few interesting corollaries.

Corollary 29.

Let γZi1(X;𝔽q).\gamma\in Z_{i-1}\left(X;\,\mathbb{F}_{q}\right).

𝔼ν(WγVγ)=1\mathbb{E}_{\nu}\left(W_{\gamma}\mid V_{\gamma}\right)=1

and

𝔼ν(WγVγc)=0.\mathbb{E}_{\nu}\left(W_{\gamma}\mid V_{\gamma}^{c}\right)=0\,.

In particular, if VγV_{\gamma} is the event that γ\gamma is null-homologous in PP then

𝔼ν(Wγ)=μX(Vγ).\mathbb{E}_{\nu}\left(W_{\gamma}\right)=\mu_{X}\left(V_{\gamma}\right)\,.

The last statement is Theorem 5.

Corollary 30.

Let γ,γZi1(X;𝔽q).\gamma,\gamma^{\prime}\in Z_{i-1}\left(X;\,\mathbb{F}_{q}\right). Then conditioned on

[γ]=[γ]Hi1(P;𝔽q),\left[\gamma\right]=\left[\gamma^{\prime}\right]\in H_{i-1}\left(P;\,\mathbb{F}_{q}\right)\,,

Wγ=WγW_{\gamma}=W_{\gamma^{\prime}} almost surely. Conversely, conditioned on

[γ]span([γ])Hi1(P;𝔽q),\left[\gamma\right]\notin\mathrm{span}\left(\left[\gamma^{\prime}\right]\right)\subset H_{i-1}\left(P;\,\mathbb{F}_{q}\right)\,,

WγW_{\gamma} and WγW_{\gamma^{\prime}} are independent.

Proof.

Conditioned on [γ]=[γ],\left[\gamma\right]=\left[\gamma^{\prime}\right], we can simply apply Proposition 28 to the Wilson loop variable for γγ\gamma-\gamma^{\prime} to get the desired result.

Now condition on the event Dγ,γ{[γ]span([γ])}.D_{\gamma,\gamma^{\prime}}\coloneqq\left\{\left[\gamma\right]\notin\mathrm{span}\left(\left[\gamma^{\prime}\right]\right)\right\}. First, we show independence of WγW_{\gamma} and WγW_{\gamma^{\prime}} when further conditioning on a fixed configuration ξ.\xi. If [γ]=0Hi1(P(ξ),𝔽q),\left[\gamma^{\prime}\right]=0\in H_{i-1}\left(P\left(\xi\right),\mathbb{F}_{q}\right), then WωW_{\omega} and WωW_{\omega^{\prime}} are trivially independent. Now assume [γ]0.\left[\gamma^{\prime}\right]\neq 0. Since Dγ,γD_{\gamma,\gamma^{\prime}} holds, we can extend the set {[γ],[γ]}\left\{\left[\gamma\right]^{*},\left[\gamma^{\prime}\right]^{*}\right\} to a basis \mathcal{B} of Hi1(Pξ,𝔽q)H^{i-1}\left(P_{\xi},\mathbb{F}_{q}\right) again using the universal coefficent theorem for cohomology. Then by Proposition 21, we can sample ν(fω=ξ)\nu\left(f\mid\omega=\xi\right) by taking a random linear combination of elements of \mathcal{B} with independent Unif(𝔽q)\mathrm{Unif}\left(\mathbb{F}_{q}\right) coefficients. By construction, we then see that f(ω)f\left(\omega\right) and f(ω)f\left(\omega^{\prime}\right) are independent, and thus WγW_{\gamma} and WγW_{\gamma^{\prime}} are independent.

Now note that (WγDγ,γ,ω=ξ)Unif(𝔽q)\left(W_{\gamma}\mid D_{\gamma,\gamma^{\prime}},\omega=\xi\right)\sim\mathrm{Unif}\left(\mathbb{F}_{q}\right)^{\mathbb{C}} by the same argument as in Proposition 28. Then for any ζ1,ζ2(𝔽q),\zeta_{1},\zeta_{2}\in\left(\mathbb{F}_{q}\right)^{\mathbb{C}}, we can compute

ν(Wγ=ζ1Wγ=ζ2,Dγ,γ)\displaystyle\nu\left(W_{\gamma}=\zeta_{1}\mid W_{\gamma^{\prime}}=\zeta_{2},D_{\gamma,\gamma^{\prime}}\right)
=ξν(Wγ=ζ1ω=ξ,Wγ=ζ2,Dγ,γ)ν(ω=ξWγ=ζ2,Dγ,γ)\displaystyle\quad=\sum_{\xi}\nu\left(W_{\gamma}=\zeta_{1}\mid\omega=\xi,W_{\gamma^{\prime}}=\zeta_{2},D_{\gamma,\gamma^{\prime}}\right)\nu\left(\omega=\xi\mid W_{\gamma^{\prime}}=\zeta_{2},D_{\gamma,\gamma^{\prime}}\right)
=ξ(Unif(𝔽q)=ζ)ν(ω=ξWγ=ζ2,Dγ,γ)\displaystyle\quad=\sum_{\xi}\mathbb{P}\left(\mathrm{Unif}\left(\mathbb{F}_{q}\right)^{\mathbb{C}}=\zeta\right)\nu\left(\omega=\xi\mid W_{\gamma^{\prime}}=\zeta_{2},D_{\gamma,\gamma^{\prime}}\right)
=(Unif(𝔽q)=ζ1)\displaystyle\quad=\mathbb{P}\left(\mathrm{Unif}\left(\mathbb{F}_{q}\right)^{\mathbb{C}}=\zeta_{1}\right)
=ν(Wγ=ζ1Dγ,γ).\displaystyle\quad=\nu\left(W_{\gamma}=\zeta_{1}\mid D_{\gamma,\gamma^{\prime}}\right)\,.

Thus, WγW_{\gamma} and WγW_{\gamma^{\prime}} are independent conditioned on Dγ,γ.D_{\gamma,\gamma^{\prime}}.

Finally, we prove a direct generalization of the relationship between the two-point correlation function of the classical Potts model and connection probabilities in the corresponding random-cluster model (Theorem 1.16 in [Gri06]). For any γZi1(X,𝔽q),\gamma\in Z_{i-1}\left(X,\mathbb{F}_{q}\right), let

τβ,q(γ)=ν(Wγ=1)1q.\tau_{\beta,q}\left(\gamma\right)=\nu\left(W_{\gamma}=1\right)-\frac{1}{q}\,.

Note that if i=1i=1 and v0v_{0} and v1v_{1} are sites of X,X, then τβ,q(v1v0)\tau_{\beta,q}\left(v_{1}-v_{0}\right) is the two-point correlation between the sites.

Proposition 31.
τβ,q(γ)=(11q)μX(Vγ),\tau_{\beta,q}\left(\gamma\right)=\left(1-\frac{1}{q}\right)\mu_{X}\left(V_{\gamma}\right)\,,

where p=1eβ.p=1-e^{-\beta}.

Proof.

For simplicity, we write μ=μX\mu=\mu_{X} in the following calculation.

τβ,q(γ)\displaystyle\tau_{\beta,q}\left(\gamma\right) =\displaystyle= ν(Wγ=1)1q\displaystyle\nu\left(W_{\gamma}=1\right)-\frac{1}{q}
=\displaystyle= (ν(Wγ=1Vγ)1q)μ(Vγ)+(ν(Wγ=1Vγc)1q)μ(Vγc)\displaystyle\left(\nu\left(W_{\gamma}=1\mid V_{\gamma}\right)-\frac{1}{q}\right)\mu\left(V_{\gamma}\right)+\left(\nu\left(W_{\gamma}=1\mid V_{\gamma}^{c}\right)-\frac{1}{q}\right)\mu\left(V_{\gamma}^{c}\right)
=\displaystyle= (11q)μ(Vγ),\displaystyle\left(1-\frac{1}{q}\right)\mu\left(V_{\gamma}\right)\,,

where we used Proposition 28 in the last step. ∎

5.5. Proof of Theorem 7

By Theorem 5, a sharp phase transition for Wilson loop variables is equivalent to one for the topological variables VγV_{\gamma} in the corresponding random-cluster model. Such a result is unknown for two-dimensional percolation in four dimension, let alone the random-cluster model, though we prove a “qualitative” analogue in Section 6 below. We can, however, show a partial result by comparison to plaquette percolation. First, we prove a stochastic domination result for the plaquette random-cluster model which is a direct generalization of the corresponding classical result [FK72].

Lemma 32.

For any finite cubical complex X,X, μX\mu_{X} is stochastically decreasing in q1q\geq 1 for fixed p.p. On the other hand, if we fix p^=p/q1p+p/q\hat{p}=\frac{p/q}{1-p+p/q} then μX,p^,q,i\mu_{X,\hat{p},q,i} is stochastically increasing in q1.q\geq 1.

Proof.

This is a consequence of the fact that adding an ii-plaquette can only reduce 𝐛i1\mathbf{b}_{i-1} by one or leave 𝐛i1\mathbf{b}_{i-1} unchanged.

Recall that by Equation 19, p^\hat{p} is the conditional probability that a plaquette is open given that it reduces 𝐛i1\mathbf{b}_{i-1} by one when leaving the states of the other plaquettes unchanged. First fix p.p. Since p^\hat{p} is decreasing in qq and the conditional probability that a plaquette is open given that it does not kill an (i1)(i-1)-cycle is constant in qq (equaling pp), an application of Theorem 38 shows that μX\mu_{X} is stochastically decreasing in q.q. Now fix p^.\hat{p}. Then pp is decreasing as a function of q,q, so again applying Theorem 38 to the conditional probabilities given by Equation 19 yields that μX\mu_{X} is stochastically increasing in q.q.

A similar statement for μd\mu_{\mathbb{Z}^{d}} follows immediately since it is a local limit of random-cluster measures on finite cubical complexes.

Corollary 33.

μd\mu_{\mathbb{Z}^{d}} is stochastically decreasing in q1q\geq 1 for fixed p.p. If we fix p^=p/q1p+p/q,\hat{p}=\frac{p/q}{1-p+p/q}, μd,p^,q,i\mu_{\mathbb{Z}^{d},\hat{p},q,i} is stochastically increasing in q1.q\geq 1.

Next, we require two more results of [ACC+83] for plaquette percolation.

Theorem 34 (Aizenman, Chayes, Chayes, Frölich, Russo [ACC+83]).

Consider ii-dimensional Bernoulli plaquette percolation on d.\mathbb{Z}^{d}. There exist finite, positive constants c^1=c^1(p,d,i),c^2=c^2(p,d,i)>0\hat{c}_{1}=\hat{c}_{1}(p,d,i),\hat{c}_{2}=\hat{c}_{2}(p,d,i)>0 so that, for hyperrectangular (i1)(i-1)-boundaries in d,\mathbb{Z}^{d},

exp(c^1Area(γ)p(Vγ)exp(c^2Per(γ)).\exp(-\hat{c}_{1}\mathrm{Area}(\gamma)\leq\mathbb{P}_{p}(V_{\gamma})\leq\exp(-\hat{c}_{2}\mathrm{Per}(\gamma))\,. (14)

Furthermore, there are constants 0<p~1(d,i)p~2(d,i)<10<\tilde{p}_{1}\left(d,i\right)\leq\tilde{p}_{2}\left(d,i\right)<1 so that

log(p(Vγ))Area(γ)c^1\displaystyle-\frac{\log\left(\mathbb{P}_{p}(V_{\gamma})\right)}{\mathrm{Area}(\gamma)}\;\;\rightarrow\;\;\hat{c}_{1} if p<p~1(d,i)\displaystyle\quad\text{if }p<\tilde{p}_{1}\left(d,i\right)
log(p(Vγ))Per(γ))=Θ(1)\displaystyle-\frac{\log\left(\mathbb{P}_{p}(V_{\gamma})\right)}{\mathrm{Per}(\gamma))}\;\;=\;\;\Theta\left(1\right) if p>p~2(d,i).\displaystyle\quad\text{if }p>\tilde{p}_{2}\left(d,i\right)\,.

The next lemma for the plaquette random-cluster model follows from the same tiling argument used to show Proposition 2.4 of [ACC+83], which only requires positive association.

Lemma 35.

Let {γn}\left\{\gamma_{n}\right\} be a sequence of hyperrectangular (i1)(i-1)-boundaries whose dimensions diverge with n.n. Then

limnμd(Vγn)Area(γn)\lim_{n\rightarrow\infty}\frac{\mu_{\mathbb{Z}^{d}}\left(V_{\gamma_{n}}\right)}{\mathrm{Area}(\gamma_{n})}

converges, where μd\mu_{\mathbb{Z}^{d}} denotes an infinite volume random-cluster measure, constructed with either free or wired boundary conditions.

Proof of Theorem 7.

Let p1(β)=1eβp_{1}\left(\beta\right)=1-e^{-\beta} and p2(β)=p1(β)/q1p1(β)+p1(β)/qp_{2}\left(\beta\right)=\frac{p_{1}\left(\beta\right)/q}{1-p_{1}\left(\beta\right)+p_{1}\left(\beta\right)/q}. It follows from Corollaries 29 and 33 that

μd,p1(β),q,i𝐟(Vγ)𝔼ν𝐟(Wγ)μd,p2(β),q,i𝐟(Vγ).\mu^{\mathbf{f}}_{\mathbb{Z}^{d},p_{1}\left(\beta\right),q,i}\left(V_{\gamma}\right)\leq\mathbb{E}_{\nu^{\mathbf{f}}}\left(W_{\gamma}\right)\leq\mu^{\mathbf{f}}_{\mathbb{Z}^{d},p_{2}\left(\beta\right),q,i}\left(V_{\gamma}\right)\,.

Then the inequalities

exp(c7Area(γ))𝔼ν𝐟(Wγ)exp(c7Per(γ)).\exp(-c_{\ref{const:3}}\mathrm{Area}(\gamma))\leq\mathbb{E}_{\nu^{\mathbf{f}}}(W_{\gamma})\leq\exp(-c_{\ref{const:4}}\mathrm{Per}(\gamma))\,.

follow from the corresponding statements in Equation 14. In addition, we may set β1=log(1p^)\beta_{1}=-\log\left(1-\hat{p}\right) where

p^=p~1(i,d)qp~1(i,d)(q1)+1\hat{p}=\frac{\tilde{p}_{1}\left(i,d\right)q}{\tilde{p}_{1}\left(i,d\right)(q-1)+1}

and p~1(i,d)\tilde{p}_{1}\left(i,d\right) is given by Lemma 34. Finally, fix β2=log(1p~2(i,d))\beta_{2}=-\log\left(1-\tilde{p}_{2}\left(i,d\right)\right) and note that the existence of the constant c7-c_{\ref{const:3}} follows from Lemma 35. Thus, we have the desired statement:

log(𝔼ν𝐟(Wγ))Area(γ)c7\displaystyle-\frac{\log\left(\mathbb{E}_{\nu^{\mathbf{f}}}(W_{\gamma})\right)}{\mathrm{Area}(\gamma)}\;\;\rightarrow\;\;c_{\ref{const:3}} if β<β1\displaystyle\quad\text{if }\beta<\beta_{1}
log(𝔼ν𝐟(Wγ))Per(γ))=Θ(1)\displaystyle-\frac{\log\left(\mathbb{E}_{\nu^{\mathbf{f}}}(W_{\gamma})\right)}{\mathrm{Per}(\gamma))}\;\;=\;\;\Theta\left(1\right) if β>β2.\displaystyle\quad\text{if }\beta>\beta_{2}\,.

6. Phase Transitions for Homological Percolation

In this section, we prove the existence of sharp phase transitions for the qq-state ii-dimensional plaquette random-cluster model on 𝕋d\mathbb{T}^{d} in the sense of homological percolation. First, we recall the definition of homological percolation. Let PP be an ii-dimensional subcomplex of 𝕋Nd\mathbb{T}^{d}_{N} and let ϕ:Hi(P;𝔽)Hi(𝕋d;𝔽)\phi_{*}:H_{i}\left(P;\,\mathbb{F}\right)\rightarrow H_{i}\left(\mathbb{T}^{d};\,\mathbb{F}\right) be the map induced by inclusion. Then bi=rankϕb_{i}=\operatorname{rank}\phi_{*} counts the number of giant cycles of P.P. A homological percolation phase transition occurs at pcp_{c} if bi=0b_{i}=0 with high probability if p<pc,p<p_{c}, and bib_{i} attains its maximum possible value, (di),\binom{d}{i}, with high probability if p>pc.p>p_{c}.

6.1. Interpretation for Potts Lattice Gauge Theory

The homological percolation transition has two interpretations for Potts lattice gauge theory. The first is in terms of the behavior of the plaquette Swendsen–Wang algorithm, and was discussed in Section 5.1. The other is related to the behavior of generalized Polyakov loop variables, random variables of the form Wγ(f)W_{\gamma}\left(f\right) where γ\gamma is a giant (i1)(i-1)-cycle.

Proposition 36.

Let i1i\geq 1 and let X𝕋NdX\subset\mathbb{T}^{d}_{N} be a subcomplex containing a giant ii-cycle. Then there exist Polyakov loops γ1,γ2Zi1(X;𝔽q)\gamma_{1},\gamma_{2}\in Z_{i-1}\left(X;\,\mathbb{F}_{q}\right) so that d𝕋nd(supp(γ1),supp(γ2))N/21d_{\mathbb{T}^{d}_{n}}\left(\mathrm{supp}\left(\gamma_{1}\right),\mathrm{supp}\left(\gamma_{2}\right)\right)\geq N/2-1 and [γ1]=[γ2]Hi1(X;𝔽q).\left[\gamma_{1}\right]=\left[\gamma_{2}\right]\in H_{i-1}\left(X;\,\mathbb{F}_{q}\right).

Proof.

Let ρHi(X;𝔽q)\rho\in H_{i}\left(X;\,\mathbb{F}_{q}\right) be a giant ii-cycle. First, we show that a transverse (d1)(d-1)-dimensional slice of ρ\rho contains a giant (i1)(i-1)-cycle. Let T1=𝕋Nd{x1=1/2}T_{1}=\mathbb{T}^{d}_{N}\cap\left\{x_{1}=1/2\right\} and T2=𝕋Nd{x1=N/2+1/2}.T_{2}=\mathbb{T}^{d}_{N}\cap\left\{x_{1}=\lfloor N/2\rfloor+1/2\right\}. Subdivide the cubical complex 𝕋Nd\mathbb{T}_{N}^{d} to add an (k1)(k-1)-plaquette at each nonempty intersection of a kk-plaquette of 𝕋Nd\mathbb{T}^{d}_{N} with T1T_{1} or T2T_{2} for each kd.k\leq d. Call this subdivision T,T, and give T1,T_{1}, T2,T_{2}, and XX the resulting cubical complex structures.

Without loss of generality assume that ρ\rho is homologous to a sum of the standard generators of the torus, including at least one that is orthogonal to T,T, i.e. a standard generator that is a product of ii many S1S^{1} factors, one of which is in the first coordinate direction (see the discussion of the homology of the torus at the end of Section 2). If ρ=σXicσσ\rho=\sum_{\sigma\in X^{i}}c_{\sigma}\sigma (remember that XiX^{i} now has subdivided cells that were not originally in 𝕋Nd)\mathbb{T}^{d}_{N})), let

γ1=σXicσ(σT1).\gamma_{1}=\sum_{\sigma\in X^{i}}c_{\sigma}\left(\sigma\cap T_{1}\right)\,.

Note that each ii-plaquette of XiX^{i} intersects T1T_{1} in either an (i1)(i-1)-plaquette or not at all. Also, let γ2\gamma_{2} be constructed as γ1\gamma_{1} with T1T_{1} replaced by T2.T_{2}. Observe that the supports of γ1\gamma_{1} and γ2\gamma_{2} are at distance at least N/21N/2-1 apart.

Consider D𝕋Nd(TT).D\coloneqq\mathbb{T}^{d}_{N}\setminus\left(T\cup T^{\prime}\right). Since DD is disconnected, we can write it as a union of its connected components D=D1D2.D=D_{1}\cup D_{2}. Let

Let

ρ1σD1cσσ\rho_{1}\coloneqq\sum_{\sigma\cap D_{1}\neq\emptyset}c_{\sigma}\sigma

and

ρ2σD2cσσ.\rho_{2}\coloneqq\sum_{\sigma\cap D_{2}\neq\emptyset}c_{\sigma}\sigma\,.

Then ρ1\rho_{1} is an ii-chain with boundary γ1+γ2,\gamma_{1}+\gamma_{2}, so γ1\gamma_{1} and γ2\gamma_{2} are homologous. Then it only remains to show that they are nontrivial giant cycles, as we can then shift them by (1/2,0,,0)\left(-1/2,0,\ldots,0\right) to obtain the desired Polyakov loops.

We claim that γ1\gamma_{1} and γ2\gamma_{2} are giant (i1)(i-1)-cycles within T1T_{1} and T2T_{2} respectively. Since γ1\gamma_{1} and γ2\gamma_{2} are homologous, it is enough to show that at least one of them is a giant cycle. Suppose that neither are, so there are ii-chains α1,α2Hi(𝕋Nd;𝔽q)\alpha_{1},\alpha_{2}\in H_{i}\left(\mathbb{T}^{d}_{N};\,\mathbb{F}_{q}\right) so that γ1=α1\gamma_{1}=\partial\alpha_{1} and γ2=α2.\gamma_{2}=\partial\alpha_{2}. so β1(ρ1)α1α2\beta_{1}\coloneqq\left(\rho_{1}\right)-\alpha_{1}-\alpha_{2} is an ii-cycle. Moreover, since β1\beta_{1} is contained in a subset of 𝕋Nd\mathbb{T}^{d}_{N} that deformations retracts to T1,T_{1}, namely T1D1T2,T_{1}\cup D_{1}\cup T_{2}, it is homologous within 𝕋Nd\mathbb{T}^{d}_{N} to an ii-cycle β^1\hat{\beta}_{1} contained in T1.T_{1}. Similarly, we set β2(ρE2)α1α2,\beta_{2}\coloneqq\left(\rho\cap E_{2}\right)-\alpha_{1}-\alpha_{2}, and can find β^2\hat{\beta}_{2} contained in TT that is homologous to β2\beta_{2} within 𝕋Nd.\mathbb{T}^{d}_{N}. Then we have

[ρ]=[β^1+β^2]Hi(𝕋Nd;𝔽q).\left[\rho\right]=\left[\hat{\beta}_{1}+\hat{\beta}_{2}\right]\in H_{i}\left(\mathbb{T}^{d}_{N};\,\mathbb{F}_{q}\right)\,.

But the (d1)(d-1)-tori T1T_{1} and T2T_{2} can only contain certain giant cycles described by the Künneth formula. In particular, they cannot contain orthogonal homology classes, contradicting the assumption about ρ\rho made earlier. ∎

The following is now a corollary of Theorems 8 and  11.

Corollary 37.

Let 1i<d,1\leq i<d, let qq be an odd prime, and set

βsurf=βsurf(q,d,N)log(1λ(q,d,N)).\beta_{\mathrm{surf}}=\beta_{\mathrm{surf}}\left(q,d,N\right)\coloneqq-\log\left(1-\lambda\left(q,d,N\right)\right)\,.

If β<βsurf,\beta<\beta_{\mathrm{surf}}, then (i1)(i-1)-dimensional then, with high probability, qq-state Potts lattice gauge theory has Polyakov loops which all take the same value and rule a giant cycle of the corresponding plaquette random cluster model. When d=2i,d=2i, we can take βsurf=βsd=log(1+q).\beta_{\mathrm{surf}}=\beta_{\mathrm{sd}}=\log\left(1+\sqrt{q}\right).

6.2. Some Probabilistic Tools

Though the random-cluster model is more difficult to work with than its Bernoulli counterpart, useful tools have been developed that work in both settings, particularly in recent years. We will use several of these results, both for graph and for higher dimensional cubical complexes.

When comparing different critical probabilities, it is essential to be able to compare the probabilities of events in different percolation models. This is easy to do when we can show that one model has strictly more open edges or plaquettes in a probabilistic sense. Let EE be a finite set and let μ1,μ2\mu_{1},\mu_{2} be probability measures on Ω={0,1}E.\Omega=\left\{0,1\right\}^{E}. We say that μ1\mu_{1} is stochastically dominated by μ2\mu_{2} if there is a probability measure κ\kappa on Ω×Ω\Omega\times\Omega with first and second marginals μ1\mu_{1} and μ2\mu_{2} so that

κ({(ω1,ω2):ω1ω2})=1.\kappa\left(\left\{\left(\omega_{1},\omega_{2}\right):\omega_{1}\leq\omega_{2}\right\}\right)=1\,.

In this case we write μ1stμ2.\mu_{1}\leq_{\mathrm{st}}\mu_{2}. We will use a version of Holley’s theorem that gives a criterion for stochastic domination in terms of conditional probabilities. This can be found as Theorem 2.3 of [Gri06].

Theorem 38 (Holley).

Let EE be a finite set and let μ1,μ2\mu_{1},\mu_{2} be strictly positive probability measures on Ω={0,1}E.\Omega=\left\{0,1\right\}^{E}. Suppose that for each pair ξ,ζΩ\xi,\zeta\in\Omega with ξζ\xi\leq\zeta and each e0E,e_{0}\in E,

μ1(ω(e0)=1:ω(e)=ξ(e) for all eE{e0})\displaystyle\mu_{1}\left(\omega\left(e_{0}\right)=1:\omega\left(e\right)=\xi\left(e\right)\text{ for all }e\in E\setminus\left\{e_{0}\right\}\right)
μ2(ω(e0)=1:ω(e)=ζ(e) for all eE{e0}).\displaystyle\qquad\leq\mu_{2}\left(\omega\left(e_{0}\right)=1:\omega\left(e\right)=\zeta\left(e\right)\text{ for all }e\in E\setminus\left\{e_{0}\right\}\right)\,.

Then μ1stμ2.\mu_{1}\leq_{\mathrm{st}}\mu_{2}.

We will also employ a sharp threshold theorem for monotonic measures due to Graham and Grimmett [GG06].

Theorem 39 (Graham and Grimmett).

There exists a constant 0<c10<0<c_{10}<\infty so that the following holds. Let N1,N\geq 1, I={1,,N},I=\left\{1,\ldots,N\right\}, Ω={0,1}N,\Omega=\left\{0,1\right\}^{N}, and let \mathcal{F} be the set of subsets of Ω.\Omega. Let AA\in\mathcal{F} be an increasing event. Let μ\mu be a positive monotonic probability measure on (Ω,).(\Omega,\mathscr{F}). Let Xi=ω(i)X_{i}=\omega\left(i\right) and set p=μ(Xi=1).p=\mu\left(X_{i}=1\right). If there exists a subgroup 𝒜\mathcal{A} of the symmetric group on NN elements ΠN\Pi_{N} acting transitively on II so that μ\mu and AA are 𝒜\mathcal{A}-invariant, then

ddpμp(A)c39μp(X1)(1μp(X1))p(1p)min{μp(A),1μp(A)}logN.\displaystyle\frac{d}{dp}\mu_{p}\left(A\right)\geq\frac{c_{\ref{const:6}}\mu_{p}\left(X_{1}\right)\left(1-\mu_{p}\left(X_{1}\right)\right)}{p\left(1-p\right)}\min\left\{\mu_{p}\left(A\right),1-\mu_{p}\left(A\right)\right\}\log N\,.

In the 1-dimensional case, we will use the recent breakthrough results of Duminil-Copin, Raoufi, and Tassion characterizing the subcritical and supercritical regimes of the random-cluster model.

Theorem 40 (Duminil-Copin, Raoufi, Tassion).

Fix d2d\geq 2 and q1.q\geq 1. Let θ(p)=μd,p,q,1𝐰(0).\theta(p)=\mu^{\mathbf{w}}_{\mathbb{Z}^{d},p,q,1}\left(0\leftrightarrow\infty\right). Then

  • there exists a c11>0c_{11}>0 so that θ(p)c40(pp^c)\theta\left(p\right)\geq c_{\ref{const:7}}\left(p-\hat{p}_{c}\right) for any ppcp\geq p_{c} sufficiently close to pc;p_{c};

  • for any p<p^c,p<\hat{p}_{c}, there exists a cpc_{p} so that for every n0,n\geq 0,

    μΛn,p,q,1𝐰(0Λn)exp(cpn).\mu^{\mathbf{w}}_{\Lambda_{n},p,q,1}\left(0\leftrightarrow\partial\Lambda_{n}\right)\leq\exp\left(-c_{p}n\right)\,.

Lastly, we will use a result of Kahle and the authors on vector-valued random variables whose range is an irreducible representation of the symmetry group of the probability space.

Lemma 41 (Duncan, Kahle, and Schweinhart).

Let VV be a finite dimensional vector space and YY be a set. Let 𝒜\mathcal{A} be the lattice of subspaces of V.V. Suppose h:𝒫(Y)𝒜h:\mathcal{P}\left(Y\right)\to\mathcal{A} is an increasing function, i.e. if ABA\subset B then h(A)h(B).h\left(A\right)\subset h\left(B\right). Let 𝒢\mathcal{G} be a finite group which acts on both YY and VV whose action is compatible with h.h. That is, for each g𝒢g\in\mathcal{G} and vV,v\in V, g(h(v))=h(gv).g\left(h\left(v\right)\right)=h\left(gv\right). Let XX be a 𝒫(Y)\mathcal{P}\left(Y\right)-valued random variable with a 𝒢\mathcal{G}-invariant distribution that is positively associated, meaning that increasing events with respect to XX are non-negatively correlated. Then if VV is an irreducible representation of 𝒢\mathcal{G}, there are positive constants c12,c13c_{12},c_{13} so that

p(h(X)=V)c41p(h(X)0)c41,\mathbb{P}_{p}\left(h(X)=V\right)\geq c_{\ref{const:8}}\mathbb{P}_{p}\left(h(X)\neq 0\right)^{c_{\ref{const:9}}}\,,

where c41c_{\ref{const:8}} only depends on 𝒢\mathcal{G} and c41c_{\ref{const:9}} only depends on dimV.\dim V.

6.3. Proof of Theorem 8

We recall that AA is the event that the map on homology induced by the inclusion P𝕋dP\hookrightarrow\mathbb{T}^{d} is nonzero and SS is the event that the same map is surjective. Note that by the intermediate value theorem we may define a threshold function λ=λ(N)\lambda=\lambda\left(N\right) so that μ𝕋Nd,λ,q,i(A)=1/2\mu_{\mathbb{T}^{d}_{N},\lambda,q,i}\left(A\right)=1/2 for each N.N. We begin by showing a general threshold result for homological percolation. The proof is very similar to that in [DKS20].

Proposition 42.

Let qq be an odd prime and 1id1.1\leq i\neq d-1. For any ϵ>0,\epsilon>0, we have

μ𝕋Nd,λϵ,q,i(A)0\displaystyle\mu_{\mathbb{T}^{d}_{N},\lambda-\epsilon,q,i}\left(A\right)\to 0

and

μ𝕋Nd,λ+ϵ,q,i(S)1\displaystyle\mu_{\mathbb{T}^{d}_{N},\lambda+\epsilon,q,i}\left(S\right)\to 1

as N.N\to\infty.

Proof.

We prove the second statement first. Note that increasing events with respect to μ𝕋Nd\mu_{\mathbb{T}^{d}_{N}} are positively correlated by Theorem 16 and that the group of symmetries of the torus contains an irreducible representation of Hi(𝕋Nd)H_{i}\left(\mathbb{T}^{d}_{N}\right) (see Proposition 13 of [DKS20] for a proof). Then by Lemma 41,

μ𝕋Nd(S)c41μ𝕋Nd(A)c41,\displaystyle\mu_{\mathbb{T}^{d}_{N}}\left(S\right)\geq c_{\ref{const:8}}\mu_{\mathbb{T}^{d}_{N}}\left(A\right)^{c_{\ref{const:9}}}\,,

where c41,c41c_{\ref{const:8}},c_{\ref{const:9}} do not depend on N.N. In particular there is a δ>0\delta>0 so that μ𝕋Nd,λ,q,i(S)δ\mu_{\mathbb{T}^{d}_{N},\lambda,q,i}\left(S\right)\geq\delta for all N.N.

Now we show that μ𝕋Nd(S)\mu_{\mathbb{T}^{d}_{N}}\left(S\right) has a sharp threshold. By Theorem 16, μ𝕋Nd\mu_{\mathbb{T}^{d}_{N}} satisfies the FKG lattice condition and is thus monotonic [Gri04]. Then since the symmetries of 𝕋Nd\mathbb{T}^{d}_{N} act transitively on the plaquettes, we apply Theorem 39 to obtain

ddpμ𝕋Nd(S)\displaystyle\frac{d}{dp}\mu_{\mathbb{T}^{d}_{N}}\left(S\right) c39qmin{μ𝕋Nd(S),1μ𝕋Nd(S)}log|FNi|.\displaystyle\geq\frac{c_{\ref{const:6}}}{q}\min\left\{\mu_{\mathbb{T}^{d}_{N}}\left(S\right),1-\mu_{\mathbb{T}^{d}_{N}}\left(S\right)\right\}\log\left|F^{i}_{N}\right|\,.

In particular, for pp close to λ,\lambda, ddpμ𝕋Nd,p,q,i(S)c39δqlog|FNi|.\frac{d}{dp}\mu_{\mathbb{T}^{d}_{N},p,q,i}\left(S\right)\geq\frac{c_{\ref{const:6}}\delta}{q}\log\left|F^{i}_{N}\right|. By integrating this inequality, we have μ𝕋Nd,λ+ϵ,q,i(S)1\mu_{\mathbb{T}^{d}_{N},\lambda+\epsilon,q,i}\left(S\right)\to 1 as N.N\to\infty.

To obtain the first statement, we apply the same argument to the dual system. Since μ𝕋Nd,p,q,i(A)=1/2,\mu_{\mathbb{T}^{d}_{N},p,q,i}\left(A\right)=1/2, it follows that μ𝕋Nd,(λ),q,i(A)1/2\mu_{\mathbb{T}^{d}_{N},\left(\lambda\right)^{*},q,i}\left(A\right)\geq 1/2 by Equation 5. Then since pp^{*} is continuous as a function of p,p, the above argument shows that μ𝕋Nd,(λϵ),q,i(S)1\mu_{\mathbb{T}^{d}_{N},\left(\lambda-\epsilon\right)^{*},q,i}\left(S\right)\to 1 as N.N\to\infty. By another application of Equation 5, we then have μ𝕋Nd,λϵ,q,i(A)0\mu_{\mathbb{T}^{d}_{N},\lambda-\epsilon,q,i}\left(A\right)\to 0 as N.N\to\infty.

Corollary 43.

If p0>pu(q,i,d),p_{0}>p_{u}\left(q,i,d\right), then

μ𝕋Nd,p0,q,i(S)1\displaystyle\mu_{\mathbb{T}^{d}_{N},p_{0},q,i}\left(S\right)\to 1

as N.N\to\infty. If p0<pl(q,i,d),p_{0}<p_{l}\left(q,i,d\right), then

μ𝕋Nd,p0,q,i(A)0\displaystyle\mu_{\mathbb{T}^{d}_{N},p_{0},q,i}\left(A\right)\to 0

as N.N\to\infty.

It is now straightforward to prove that the ii-dimensional random-cluster model in 𝕋2d\mathbb{T}^{2d} exhibits a sharp phase transition at the self-dual point.

Proof.

By self duality and Equation 5,

μ𝕋Nd,psd,q,i(A)1/2.\mu_{\mathbb{T}^{d}_{N},p_{\mathrm{sd}},q,i}\left(A\right)\geq 1/2\,.

In particular, pupsd.p_{u}\leq p_{\mathrm{sd}}. Then by monotonicity and Corollary 43,

μ𝕋Nd(S)1\mu_{\mathbb{T}^{d}_{N}}\left(S\right)\to 1

as NN\to\infty for all p>psd.p>p_{\mathrm{sd}}. Since pp^{*} is decreasing as a function of pp with fixed point psd,p_{\mathrm{sd}}, applying Equation 5 again gives

μ𝕋Nd(A)1\mu_{\mathbb{T}^{d}_{N}}\left(A\right)\to 1

as NN\to\infty for all p<psd.p<p_{\mathrm{sd}}.

6.4. Proof of Theorem 10

Next, we will prove our theorem on the homological percolation transition for the 11-dimensional and (d1)(d-1)-dimensional random-cluster models in 𝕋d.\mathbb{T}^{d}. Throughout this subsection, we fix i=1,i=1, and we borrow terminology from percolation on graphs. Let GG be a graph and let V(G)V\left(G\right) and E(G)E\left(G\right) be its vertex and edge sets respectively. Designate the edges of E(G)E\left(G\right) as open or closed according to some probability measure. For v,wV(G),v,w\in V\left(G\right), we write vwv\leftrightarrow w if vv is connected to ww by a path of open edges. For a vertex subset VV(G),V^{\prime}\subset V\left(G\right), we write vVwv\leftrightarrow_{V^{\prime}}w if vv is connected to ww by a path of open edges that only passes through vertices of V.V^{\prime}.

We will prove the statements of Theorem 10 for the case i=1i=1 in each regime separately. The case i=d1i=d-1 then follows immediately from Equation 5. The subcritical case is a straightforward application of the exponential decay of the size of components.

Proof of Theorem 10 (subcritical regime).

Let p<p^cp<\hat{p}_{c} and let M=N/2.M=\lfloor N/2\rfloor. For a vertex xx of 𝕋Nd,\mathbb{T}_{N}^{d}, let AxA_{x} the event that a giant 11-cycle passes through x.x. Since [M,M]d=ΛM[-M,M]^{d}=\Lambda_{M} is contractible in 𝕋Nd,\mathbb{T}^{d}_{N}, A0{0𝕋NdΛM}.A_{0}\subset\left\{0\leftrightarrow_{\mathbb{T}^{d}_{N}}\partial\Lambda_{M}\right\}. Moreover, μΛM𝐰\mu^{\mathbf{w}}_{\Lambda_{M}} stochastically dominates μ𝕋Nd|ΛM,\mu_{\mathbb{T}^{d}_{N}}|_{\Lambda_{M}}, so by Theorem 40 and translation invariance,

μ𝕋Nd(Ax)μΛM𝐰(0ΛM)exp(cpM)\mu_{\mathbb{T}^{d}_{N}}\left(A_{x}\right)\leq\mu^{\mathbf{w}}_{\Lambda_{M}}\left(0\leftrightarrow\Lambda_{M}\right)\leq\exp\left(c_{p}M\right) (15)

for all vertices xx of 𝕋Nd.\mathbb{T}_{N}^{d}.

Let XX be the number of vertices in 𝕋Nd\mathbb{T}_{N}^{d} that are contained in a giant 11-cycle. Since A={X1},A=\left\{X\geq 1\right\},

μ𝕋Nd(A)\displaystyle\mu_{\mathbb{T}^{d}_{N}}\left(A\right) =μ𝕋Nd(X1)\displaystyle=\;\;\mu_{\mathbb{T}^{d}_{N}}\left(X\geq 1\right)
𝔼μ(X)\displaystyle\leq\;\;\mathbb{E}_{\mu}\left(X\right) by Markov’s Inequality
=x𝕋Ndμ𝕋Nd(Ax)\displaystyle=\;\;\sum_{x\in\mathbb{T}_{N}^{d}}\mu_{\mathbb{T}^{d}_{N}}\left(A_{x}\right)
NdecpM\displaystyle\leq\;\;N^{d}e^{-c_{p}M} by Equation  15
=NdecpN/20\displaystyle=\;\;N^{d}e^{-c_{p}\lfloor N/2\rfloor}\to 0

as NN\rightarrow\infty.

The supercritical case is more involved and will require assuming Conjecture 9 — the continuity of the critical probabilities for the existence of an infinite component for the 11-dimensional random-cluster model in slabs — in order to obtain a sharp threshold. Our strategy will be to \saypin together paths from supercritical percolation in large slabs in order to form a giant 11-cycle.

Let

Λn,k[n,n]2×[k,k]d2dSk.\Lambda_{n,k}\coloneqq\left[-n,n\right]^{2}\times\left[-k,k\right]^{d-2}\cap\mathbb{Z}^{d}\subset S_{k}\,.

Let Dn,k{vΛn,k:vu for some uSkΛn,k}D_{n,k}\coloneqq\left\{v\in\Lambda_{n,k}:v\sim u\text{ for some }u\in S_{k}\setminus\Lambda_{n,k}\right\} be the boundary of Λn,k\Lambda_{n,k} in Sk.S_{k}.

Lemma 44.

Fix q1,d2,q\geq 1,d\geq 2, and k1.k\geq 1. There is a c14>0c_{14}>0 so that for any p>pc(Sk)p>p_{c}\left(S_{k}\right) sufficiently close to pc(Sk)p_{c}\left(S_{k}\right) and any nn sufficiently large,

μΛn,k𝐟(0Dn,k)c44(ppc(Sk)).\mu_{\Lambda_{n,k}}^{\mathbf{f}}\left(0\leftrightarrow D_{n,k}\right)\geq c_{\ref{const:10}}\left(p-p_{c}\left(S_{k}\right)\right)\,.

The proof follows from the proof of Theorem 40 (Theorem 1.2 in [DCRT19]), replacing Λn\Lambda_{n} with Λn,k.\Lambda_{n,k}. With Lemma 44 in hand, we can give a lower bound for the probability that two vertices are connected within a box contained in the slab Sk.S_{k}.

Lemma 45.

Fix q1,d2,k1,q\geq 1,d\geq 2,k\geq 1, and p>pc(Sk).p>p_{c}\left(S_{k}\right). Let Λ=[3n,3n]×[2n,4n]×[k,k]d2d.\Lambda=[-3n,3n]\times[-2n,4n]\times[-k,k]^{d-2}\cap\mathbb{Z}^{d}. There is a constant c15>0c_{15}>0 not depending on nn so that

μΛ𝐟((n,0,,0)(n,0,,0))c45.\mu_{\Lambda}^{\mathbf{f}}\left(\left(-n,0,\ldots,0\right)\leftrightarrow\left(n,0,\ldots,0\right)\right)\geq c_{\ref{const:11}}\,.
Proof.

Consider the random-cluster model μΛ𝐟.\mu_{\Lambda}^{\mathbf{f}}. Define three crossing events

V+{(𝟎)Λ2n{(j,2n,x3,,xd):0j2n,kx3,,xdk}},V_{+}\coloneqq\left\{\left(\bm{0}\right)\leftrightarrow_{\Lambda_{2n}}\left\{\left(j,2n,x_{3},\ldots,x_{d}\right):0\leq j\leq 2n,-k\leq x_{3},\ldots,x_{d}\leq k\right\}\right\}\,,
V{(𝟎)Λ2n{(j,2n,x3,,xd):0j2n,kx3,,xdk}},V_{-}\coloneqq\left\{\left(\bm{0}\right)\leftrightarrow_{\Lambda_{2n}}\left\{\left(-j,2n,x_{3},\ldots,x_{d}\right):0\leq j\leq 2n,-k\leq x_{3},\ldots,x_{d}\leq k\right\}\right\}\,,

and

U+{(𝟎)Λ2n{(j,2n,x3,,xd):0j2n,kx3,,xdk}}.U_{+}\coloneqq\left\{\left(\bm{0}\right)\leftrightarrow_{\Lambda_{2n}}\left\{\left(j,-2n,x_{3},\ldots,x_{d}\right):0\leq j\leq 2n,-k\leq x_{3},\ldots,x_{d}\leq k\right\}\right\}\,.

For an event A(𝟎)A(\bm{0}) defined in relation to the origin 𝟎,\bm{0}, let A(𝒙)A(\bm{x}) denote the event A+𝒙A+\bm{x} obtained by translating the origin to 𝒙.\bm{x}. By Lemma 44 and symmetry, there is a c16>0c_{16}>0 not depending on nn so that

μΛ𝐟(V+)=μΛ𝐟(V)=μΛ𝐟(U+)=c6.4.\mu_{\Lambda}^{\mathbf{f}}\left(V_{+}\right)=\mu_{\Lambda}^{\mathbf{f}}\left(V_{-}\right)=\mu_{\Lambda}^{\mathbf{f}}\left(U_{+}\right)=c_{\ref{const:12}}\,. (16)

Let 𝒗=(n,0,,0)\bm{v}=\left(-n,0,\ldots,0\right) and 𝒘=(n,0,,0).\bm{w}=\left(n,0,\ldots,0\right). Denote the projection onto the first two coordinates by π12:d2.\pi_{12}:\mathbb{Z}^{d}\to\mathbb{Z}^{2}. Our aim will be to create overlapping paths (i.e. paths with intersecting images under π12\pi_{12}) containing vv and w,w, which are then close enough to be connected with positive probability.

First, we will construct two paths that project to a horizontal crossing of of [2n,4n]2\left[-2n,4n\right]^{2} under π12.\pi_{12}. Fix an arbitrary ordering of the finite paths in d.\mathbb{Z}^{d}. Assuming that V+(𝒗)V_{+}\left(\bm{v}\right) occurs, let 𝒗^\hat{\bm{v}} be the endpoint of the minimal path witnessing that event and let 𝒗=π12(𝒗^)×(0,,0).\bm{v}^{\prime}=\pi_{12}\left(\hat{\bm{v}}\right)\times\left(0,\ldots,0\right). Then, the event

H=U+(𝒗)V+(𝒗)V(𝒗),H=U_{+}\left(\bm{v}\right)\cap V_{+}\left(\bm{v}\right)\cap V_{-}\left(\bm{v}^{\prime}\right)\,, (17)

ensures the existence of a path through 𝒗\bm{v} with a single gap at 𝒗\bm{v}^{\prime} which projects to a vertical crossing, as illustrated in in Figure 5.

Next, let II be the event HH rotated π/2\pi/2 radians counterclockwise in the first two coordinates about the point (0,n,0,,0).\left(0,n,0,\ldots,0\right). By symmetry, μΛ𝐟(I)=μΛ𝐟(H),\mu_{\Lambda}^{\mathbf{f}}\left(I\right)=\mu_{\Lambda}^{\mathbf{f}}\left(H\right), and that event ensures the existence of a path through 𝒘\bm{w} with a single gap which projects to a horizontal crossing of [2n,4n]2.\left[-2n,4n\right]^{2}.

Refer to caption
Figure 5. The two paths constructed in Lemma 45. The blue path is a projection of a witness for the event HH and the yellow path is a projection of a witness for I.I. The circled blue dot is 𝒗\bm{v} and the circled orange dot is 𝒘.\bm{w}. Both paths are shown in Λ\Lambda with a grid spacing of 2n2n for clarity.

Now, assume that HIH\cap I occurs so π12(𝒗)\pi_{12}\left(\bm{v}\right) and π12(𝒘)\pi_{12}\left(\bm{w}\right) are connected in π12(P).\pi_{12}\left(P\right). We count how many edges we would need to add in order to connect the overlapping segments in PP, thus connecting 𝒗\bm{v} to 𝒘.\bm{w}. Any two points in SkS_{k} with the same image under π12\pi_{12} are at graph distance at most 2k(d2).2k\left(d-2\right). There is one point of overlap between 𝒗^\hat{\bm{v}} and 𝒗,\bm{v}^{\prime}, and one more from the symmetric event for 𝒘.\bm{w}. In addition, we have a third point of overlap between the horizontal and vertical crossings. As such, we can connect 𝒗\bm{v} to 𝒘\bm{w} using at most 6k(d2)6k\left(d-2\right) additional open edges.

Let p>r>pc(Sk).p>r>p_{c}\left(S_{k}\right). Given a configuration ω,\omega, define

Sm(ω)={ω:eE(G)|ω(e)ω(e)|m}.S^{m}(\omega)=\left\{\omega^{\prime}:\sum_{e\in E\left(G\right)}\left|\omega\left(e\right)-\omega^{\prime}\left(e\right)\right|\leq m\right\}\,.

For an event E,E, we then define

Em(E)={ω:Sm(ω)E}.E^{m}(E)=\left\{\omega:S^{m}(\omega)\cap E\neq\emptyset\right\}\,.

Then by Theorem 3.45 of [Gri04] and the FKG inequality there is a constant c17c_{17} so that

μΛ𝐟(𝒗𝒘)\displaystyle\mu_{\Lambda}^{\mathbf{f}}\left(\bm{v}\leftrightarrow\bm{w}\right)
c6.46k(d2)μΛ𝐟(E(6k(d2))(𝒗𝒘))\displaystyle\qquad\geq c_{\ref{const:13}}^{6k\left(d-2\right)}\mu_{\Lambda}^{\mathbf{f}}\left(E^{\left(6k\left(d-2\right)\right)}\left(\bm{v}\leftrightarrow\bm{w}\right)\right)
c6.46k(d2)μΛ𝐟(HI)\displaystyle\qquad\geq c_{\ref{const:13}}^{6k\left(d-2\right)}\mu_{\Lambda}^{\mathbf{f}}\left(H\cap I\right)
c6.46k(d2)c6.46k(d2),\displaystyle\qquad\geq c_{\ref{const:12}}^{6k\left(d-2\right)}c_{\ref{const:13}}^{6k\left(d-2\right)}\,,

where we used Equations 16 and 17 in the final step. Since this bound does not depend on nn we are done.

Proof of Theorem 10 (supercritical regime).

Let p>p>pcslab.p>p^{\prime}>p_{c}^{\mathrm{slab}}. Then there is a kk such that p>pc(Sk).p^{\prime}>p_{c}\left(S_{k}\right). Let N2k.N\geq 2k. We will construct a giant cycle in percolation with parameter pp^{\prime} by using Lemma 45 to connect the centers of four pairwise overlapping boxes in the torus, each of diameter N/2.\lfloor N/2\rfloor. If NN is not divisible by 2,2, the starting and ending points of this constructed path may not exactly match. However, they will be at graph distance at most 1,1, and are therefore connected with probability at least pq.\frac{p^{\prime}}{q}. We may therefore assume that NN is divisible by 22 in the remainder of the proof for simplicity. We apply Lemma 45 to copies of

Λ[3N/8,3N/8]×[N/4,2N/4]×[k,k]d2d\Lambda\coloneqq[-3N/8,3N/8]\times[-N/4,2N/4]\times[-k,k]^{d-2}\cap\mathbb{Z}^{d}

that are centered at u0=(N/8,N/8,,0),u_{0}=\left(N/8,N/8,\ldots,0\right), u1=(3N/8,N/8,,0),u_{1}=\left(3N/8,N/8,\ldots,0\right), u2=(5N/8,N/8,,0),u_{2}=\left(5N/8,N/8,\ldots,0\right), and u3=(7N/8,N/8,,0),u_{3}=\left(7N/8,N/8,\ldots,0\right), to connect v0=𝟎,v_{0}=\bm{0}, v1=(N/4,0,,0),v_{1}=\left(N/4,0,\ldots,0\right), v2=(N/2,0,,0),v_{2}=\left(N/2,0,\ldots,0\right), and v3=(3N/4,0,,0).v_{3}=\left(3N/4,0,\ldots,0\right). For convenience, let B(u,v,w)B(u,v,w) denote the event uλ(w).u\leftrightarrow_{\lambda(w)}.

If the events B(v0,v1,u0),B(v_{0},v_{1},u_{0}), B(v1,v2,u1),B(v_{1},v_{2},u_{1}), B(v2,v3,u2),B(v_{2},v_{3},u_{2}), and B(v3,v0,u3)B(v_{3},v_{0},u_{3}) all occur, then there is an open path that is homotopic to the standard generator of H1(𝕋Nd)H_{1}\left(\mathbb{T}^{d}_{N}\right) contained in {x2=x3==xd=0}.\left\{x_{2}=x_{3}=\ldots=x_{d}=0\right\}. Thus,

HB(v0,v1,u0)B(v1,v2,u1)B(v2,v3,u2)B(v3,v0,u3).H\supseteq B(v_{0},v_{1},u_{0})\cap B(v_{1},v_{2},u_{1})\cap B(v_{2},v_{3},u_{2})\cap B(v_{3},v_{0},u_{3})\,.

Let μu0\mu_{u_{0}} denote the measure μΛ(u0),p𝐟.\mu_{\Lambda\left(u_{0}\right),p^{\prime}}^{\mathbf{f}}. We then apply the FKG inequality and Lemma 47 to bound

μ𝕋Nd,p(H)\displaystyle\mu_{\mathbb{T}^{d}_{N},p^{\prime}}\left(H\right) μ𝕋Nd,p(B(v0,v1,u0)B(v1,v2,u1)B(v2,v3,u2)B(v3,v0,u3))\displaystyle\geq\mu_{\mathbb{T}^{d}_{N},p^{\prime}}\left(B(v_{0},v_{1},u_{0})\cap B(v_{1},v_{2},u_{1})\cap B(v_{2},v_{3},u_{2})\cap B(v_{3},v_{0},u_{3})\right)
μu0(v0v1)μu1(v1v2)μu2(v2v3)μu3(v3v4)\displaystyle\geq\mu_{u_{0}}\left(v_{0}\leftrightarrow v_{1}\right)\mu_{u_{1}}\left(v_{1}\leftrightarrow v_{2}\right)\mu_{u_{2}}\left(v_{2}\leftrightarrow v_{3}\right)\mu_{u_{3}}\left(v_{3}\leftrightarrow v_{4}\right)
c454.\displaystyle\geq c_{\ref{const:11}}^{4}\,.

Note that the final bound is uniform in N.N. Again using the FKG inequality, we obtain

μ𝕋Nd,p(S)μ𝕋Nd,p(H)dc454d.\mu_{\mathbb{T}^{d}_{N},p^{\prime}}\left(S\right)\geq\mu_{\mathbb{T}^{d}_{N},p^{\prime}}\left(H\right)^{d}\geq\geq c_{\ref{const:11}}^{4d}\,.

Then by applying Theorem 39 as in Proposition 42, we have

μ𝕋Nd,p(S)1\mu_{\mathbb{T}^{d}_{N},p}\left(S\right)\to 1

as N.N\to\infty.

6.5. Proof of Theorem 11

We conclude the paper by finishing our proof of our theorem which establishes the existence of threshold functions for homological percolation in general, and enumerates some of their properties. All that remains is to prove the duality and monotonicity statements. The former is topological, and does not require modification from the proof of the analogous statement in [DKS20].

Proposition 46.

For any d2d\geq 2 and 1id1,1\leq i\leq d-1,

pu(q,i,d)=(pl(q,di,d)).p_{u}\left(q,i,d\right)=\left(p_{l}\left(q,d-i,d\right)\right)^{*}\,.
Proof.

This follows from Theorem  18, Equation 5, and Corollary 43 in a similar manner to Proposition 18 of [DKS20]. ∎

We now turn to showing monotonicity in the critical probabilities. First we require a lemma comparing the ii-dimensional plaquette random-cluster percolation on a cubical complex to the random-cluster model on a subcomplex.

Lemma 47.

Fix q1.q\geq 1. Let X,YX,Y be finite ii-dimensional cubical complexes with XYX\subset Y and recall that μX\mu_{X} and μY\mu_{Y} are the ii-dimensional plaquette random-cluster measures with parameters p,qp,q on XX and YY respectively. Then μY|X\mu_{Y}|_{X} stochastically dominates μX.\mu_{X}.

Proof.

Let σX\sigma\subset X be an ii-cell. Let SS be a subcomplex of Y,Y, let YωY_{\omega} and let R=SX.R=S\cap X. Write Sσ=SσS^{\sigma}=S\cup\sigma and Sσ=SσS_{\sigma}=S\setminus\sigma and similarly Rσ=RσR^{\sigma}=R\cup\sigma and Rσ=Rσ.R_{\sigma}=R\setminus\sigma. Since YY has no (i+1)(i+1)-cells,

𝐛i(Rσ)𝐛i(Rσ)𝐛i(Sσ)𝐛i(Sσ).\mathbf{b}_{i}\left(R^{\sigma}\right)-\mathbf{b}_{i}\left(R_{\sigma}\right)\leq\mathbf{b}_{i}\left(S^{\sigma}\right)-\mathbf{b}_{i}\left(S_{\sigma}\right)\,.

Then by the Euler–Poincaré formula we see that

𝐛i1(Rσ)𝐛i1(Rσ)𝐛i1(Sσ)𝐛i1(Sσ).\displaystyle\mathbf{b}_{i-1}\left(R_{\sigma}\right)-\mathbf{b}_{i-1}\left(R^{\sigma}\right)\geq\mathbf{b}_{i-1}\left(S_{\sigma}\right)-\mathbf{b}_{i-1}\left(S^{\sigma}\right)\,. (18)

Recall that for any ii-cell σZ,\sigma\subset Z, the random subcomplex TT with distribution μZ\mu_{Z} satisfies

μZ(σT𝐛i1(Tσ)𝐛i1(Tσ))={p𝐛i1(Tσ)𝐛i1(Tσ)=0p/q1p+p/q𝐛i1(Tσ)𝐛i1(Tσ)=1\displaystyle\mu_{Z}\left(\sigma\in T\mid\mathbf{b}_{i-1}\left(T_{\sigma}\right)-\mathbf{b}_{i-1}\left(T^{\sigma}\right)\right)=\begin{cases}p&\mathbf{b}_{i-1}\left(T_{\sigma}\right)-\mathbf{b}_{i-1}\left(T^{\sigma}\right)=0\\ \frac{p/q}{1-p+p/q}&\mathbf{b}_{i-1}\left(T_{\sigma}\right)-\mathbf{b}_{i-1}\left(T^{\sigma}\right)=1\end{cases} (19)

Then by positive association and Equations 18 and  19, we may apply Theorem 38 to conlcude that μX\mu_{X} stochastically dominates μY|X.\mu_{Y}|_{X}.

Finally, we compare percolation within subcomplexes of the torus in order to obtain the desired inequalities between critical probabilities.

Proposition 48.

For all 1id1,1\leq i\leq d-1,

pu(q,i,d)<pu(q,i,d1)<pu(q,i+1,d).p_{u}\left(q,i,d\right)<p_{u}\left(q,i,d-1\right)<p_{u}\left(q,i+1,d\right)\,.
Proof.

The topological properties of any given configuration of plaquettes are identical to those discussed Proposition 24 of [DKS20]. We will therefore only modify the probabilistic arguments as necessary.

Our first goal is to show

pu(q,i,d)<pu(q,i,d1).p_{u}\left(q,i,d\right)<p_{u}\left(q,i,d-1\right).

Our strategy will be to define a sequence of models between the random-cluster model on 𝕋Nd\mathbb{T}^{d}_{N} and 𝕋Nd1\mathbb{T}^{d-1}_{N} in which the giant cycle space of each model stochastically dominates the giant cycle space of the one before. More precisely, for a configuration of plaquettes ω,\omega, let G(ω)G\left(\omega\right) be the associated subspace of giant cycles of P(ω)P\left(\omega\right) in Hi(𝕋Nd).H_{i}\left(\mathbb{T}^{d}_{N}\right). Then we say μ1Gμ2\mu_{1}\leq_{G}\mu_{2} if there is a coupling κ\kappa of μ1\mu_{1} and μ2\mu_{2} so that

κ({(ω1,ω2):G(ω1)G(ω2)})=1.\kappa\left(\left\{\left(\omega_{1},\omega_{2}\right):G\left(\omega_{1}\right)\subseteq G\left(\omega_{2}\right)\right\}\right)=1\,.

Note that μ1stμ2\mu_{1}\leq_{\mathrm{st}}\mu_{2} implies μ1Gμ2.\mu_{1}\leq_{G}\mu_{2}.

Let 𝒯0=𝕋Nd,\mathcal{T}_{0}=\mathbb{T}^{d}_{N}, 𝒯1=𝕋Nd{x1[0,1]},\mathcal{T}_{1}=\mathbb{T}^{d}_{N}\cap\left\{x_{1}\in[0,1]\right\}, and 𝒯2=𝕋Nd{x1=0}.\mathcal{T}_{2}=\mathbb{T}^{d}_{N}\cap\left\{x_{1}=0\right\}. For j=0,1,2,j=0,1,2, let μ𝒯j\mu_{\mathcal{T}_{j}} be so that μ𝒯j|𝒯j\mu_{\mathcal{T}_{j}}|_{\mathcal{T}_{j}} is the random-cluster model on 𝒯j\mathcal{T}_{j} and μ𝒯j|𝕋Nd𝒯j\mu_{\mathcal{T}_{j}}|_{\mathbb{T}^{d}_{N}\setminus\mathcal{T}_{j}} sets all plaquettes to be closed almost surely. By Lemma 47, we have

μ𝒯0stμ𝒯1.\mu_{\mathcal{T}_{0}}\geq_{\mathrm{st}}\mu_{\mathcal{T}_{1}}\,.

We now put a different measure μ𝒯1\mu_{\mathcal{T}_{1}}^{\prime} on configurations that are closed outside 𝒯1.\mathcal{T}_{1}. Let F2F_{2} be the set of ii-cells of 𝒯1\mathcal{T}_{1} contained in 𝒯2\mathcal{T}_{2} and let F1F_{1} be the rest of the ii-cells of 𝒯1.\mathcal{T}_{1}. Let η1\eta_{1} and η2\eta_{2} be the number of open cells of F1F_{1} and F2F_{2} respectively. We set μ𝒯1|𝒯2=μ𝒯2|𝒯2.\mu_{\mathcal{T}_{1}}^{\prime}|_{\mathcal{T}_{2}}=\mu_{\mathcal{T}_{2}}|_{\mathcal{T}_{2}}. We then let μ𝒯1|𝒯1𝒯2\mu_{\mathcal{T}_{1}}^{\prime}|_{\mathcal{T}_{1}\setminus\mathcal{T}_{2}} be independent Bernoulli plaquette percolation with probability p/qp/q and declare all other plaquettes closed. More explicitly,

μ𝒯1(ω)1Zpη2(ω)(1p)|F2|η2(ω)q𝐛i1(Pω|𝒯2)(pq)η1(1pq)|F1|η1.\displaystyle\mu_{\mathcal{T}_{1}}^{\prime}\left(\omega\right)\coloneqq\frac{1}{Z}p^{\eta_{2}\left(\omega\right)}\left(1-p\right)^{\left|F_{2}\right|-\eta_{2}\left(\omega\right)}q^{\mathbf{b}_{i-1}\left(P_{\omega|_{\mathcal{T}_{2}}}\right)}\left(\frac{p}{q}\right)^{\eta_{1}}\left(1-\frac{p}{q}\right)^{\left|F_{1}\right|-\eta_{1}}\,.

We think of this as doing Bernoulli percolation on F1F_{1} with parameter p/qp/q and then a random-cluster percolation with free boundary conditions on F2.F_{2}. For σF1\sigma\in F_{1} and a configuration ξ\xi on 𝒯1,\mathcal{T}_{1}, we clearly have

μ𝒯1(ω(σ)=1ω(τ)=ξ(τ) for all τ𝒯1σ)\displaystyle\mu_{\mathcal{T}_{1}}^{\prime}\left(\omega\left(\sigma\right)=1\mid\omega\left(\tau\right)=\xi\left(\tau\right)\text{ for all }\tau\in\mathcal{T}_{1}\setminus\sigma\right)
μ𝒯1(ω(σ)=1ω(τ)=ξ(τ) for all τ𝒯1σ).\displaystyle\qquad\leq\mu_{\mathcal{T}_{1}}\left(\omega\left(\sigma\right)=1\mid\omega\left(\tau\right)=\xi\left(\tau\right)\text{ for all }\tau\in\mathcal{T}_{1}\setminus\sigma\right)\,.

By Lemma 47, for σF2\sigma\in F_{2} we also have

μ𝒯1(ω(σ)=1ω(τ)=ξ(τ) for all τ𝒯1σ)\displaystyle\mu_{\mathcal{T}_{1}}^{\prime}\left(\omega\left(\sigma\right)=1\mid\omega\left(\tau\right)=\xi\left(\tau\right)\text{ for all }\tau\in\mathcal{T}_{1}\setminus\sigma\right)
μ𝒯1(ω(σ)=1ω(τ)=ξ(τ) for all τ𝒯1σ).\displaystyle\qquad\leq\mu_{\mathcal{T}_{1}}\left(\omega\left(\sigma\right)=1\mid\omega\left(\tau\right)=\xi\left(\tau\right)\text{ for all }\tau\in\mathcal{T}_{1}\setminus\sigma\right)\,.

The again applying Theorem 38, we have

μ𝒯1stμ𝒯1.\mu_{\mathcal{T}_{1}}\geq_{\mathrm{st}}\mu_{\mathcal{T}_{1}}^{\prime}\,.

We now perform a splitting of the state of a plaquette into several Bernoulli variables similar to one found in Proposition 24 of [DKS20]. We adapt some of the definitions used there. Let SS be the set of ii-faces of 𝒯1\mathcal{T}_{1} that intersect, but are not contained in 𝒯2.\mathcal{T}_{2}. For an ii-face vv of 𝒯2,\mathcal{T}_{2}, let J(v)J(v) be the set of all perpendicular ii-faces that meet vv at an (i1)(i-1) face. Then v,v, v+𝒆1,v+\bm{e}_{1}, and J(v)J(v) are the ii-faces of an (i+1)(i+1)-face w(v).w(v). Also, for a perpendicular ii-face uu of S,S, let K(u)={v:uJ(v)}.K(u)=\left\{v:u\in J(v)\right\}. Let pSp_{S} satisfy

pq=1(1pS)2(di).\frac{p}{q}=1-\left(1-p_{S}\right)^{2\left(d-i\right)}\,.

For all pairs (v,u)\left(v,u\right) where v𝒯2v\in\mathcal{T}_{2} and uJ(v),u\in J(v), let κ(v,u)\kappa\left(v,u\right) be independent Ber(pS)\mathrm{Ber}\left(p_{S}\right) random variables. Then by construction, Bernoulli percolation with parameter p/qp/q on SS is equivalent to setting each cell vSv\in S to be open if and only if uJ(v)κ(v,u)>0.\sum_{u\in J(v)}\kappa\left(v,u\right)>0. Given Bernoulli p/qp/q percolation on F1,F_{1}, let

L={v𝒯2:κ(v,u)=1 for each uJ(v) and v+𝒆1 is open}.L=\left\{v\in\mathcal{T}_{2}:\kappa\left(v,u\right)=1\text{ for each }u\in J\left(v\right)\text{ and $v+\bm{e}_{1}$ is open}\right\}\,.

Let the μ𝒯1′′\mu_{\mathcal{T}_{1}}^{\prime\prime} have as open plaquettes the union of the open plaquettes μ𝒯1|𝒯2\mu_{\mathcal{T}_{1}}^{\prime}|_{\mathcal{T}_{2}} and the plaquettes in LL(so all plaquettes of F1F_{1} are all closed). Since the boundary of a plaquette in LL is necessarily the boundary of an open set of plaquettes,

μ𝒯1Gμ𝒯1′′.\mu_{\mathcal{T}_{1}}^{\prime}\geq_{G}\mu_{\mathcal{T}_{1}}^{\prime\prime}.

Now let p=p+(1p)pqpS2ip^{\prime}=p+(1-p)\frac{p}{q}p_{S}^{2i} (i.e. the probability that a plaquette of 𝒯2\mathcal{T}_{2} is either open or in LL) and take μ𝒯2,p\mu_{\mathcal{T}_{2},p^{\prime}} to be the random-cluster model on 𝒯2\mathcal{T}_{2} with parameter pp^{\prime} instead of p.p. Then we compare the probabilities in μ𝒯1′′\mu_{\mathcal{T}_{1}}^{\prime\prime} and in μ𝒯2,p\mu_{\mathcal{T}_{2},p^{\prime}} that a plaquette is open in the cases that its state does or does not affect 𝐛i1.\mathbf{b}_{i-1}. In former case, the conditional probabilities are pq+(1p)pqpS2i\frac{p}{q}+(1-p)\frac{p}{q}p_{S}^{2i} and pq\frac{p^{\prime}}{q} respectively, and in the latter both are p.p^{\prime}. Thus, for any configuration ξ\xi on 𝒯1\mathcal{T}_{1} and any σF2,\sigma\in F_{2},

μ𝒯1′′(ω(σ)=1ω(τ)=ξ(τ) for all τ𝒯1σ)\displaystyle\mu_{\mathcal{T}_{1}}^{\prime\prime}\left(\omega\left(\sigma\right)=1\mid\omega\left(\tau\right)=\xi\left(\tau\right)\text{ for all }\tau\in\mathcal{T}_{1}\setminus\sigma\right)
μ𝒯1,p(ω(σ)=1ω(τ)=ξ(τ) for all τ𝒯1σ).\displaystyle\quad\geq\mu_{\mathcal{T}_{1},p^{\prime}}\left(\omega\left(\sigma\right)=1\mid\omega\left(\tau\right)=\xi\left(\tau\right)\text{ for all }\tau\in\mathcal{T}_{1}\setminus\sigma\right).

Therefore we have

μ𝒯1′′stμ𝒯2,p\mu_{\mathcal{T}_{1}}^{\prime\prime}\geq_{\mathrm{st}}\mu_{\mathcal{T}_{2},p^{\prime}}

by Theorem 38, and so

μ𝒯0Gμ𝒯2,p,\mu_{\mathcal{T}_{0}}\geq_{G}\mu_{\mathcal{T}_{2},p^{\prime}}\,,

Now, as in [DKS20], we may take pp so that p<pu(q,i,d1)<p.p<p_{u}\left(q,i,d-1\right)<p^{\prime}. Then for each NN we have

μ𝕋Nd,p,q,i(A)μ𝕋Nd1,p,q,i(A),\mu_{\mathbb{T}^{d}_{N},p,q,i}\left(A\right)\geq\mu_{\mathbb{T}^{d-1}_{N},p^{\prime},q,i}\left(A\right)\,,

so

lim infμ𝕋Nd,p,q,i(A)lim infμ𝕋Nd1,p,q,i(A)1/2,\liminf\mu_{\mathbb{T}^{d}_{N},p,q,i}\left(A\right)\geq\liminf\mu_{\mathbb{T}^{d-1}_{N},p^{\prime},q,i}\left(A\right)\geq 1/2\,,

and thus

pu(q,i,d)p<pu(q,i,d1).p_{u}\left(q,i,d\right)\leq p<p_{u}\left(q,i,d-1\right)\,.

Combining this with Proposition 46 also gives pu(q,i,d1)<pu(q,i+1,d).p_{u}\left(q,i,d-1\right)<p_{u}\left(q,i+1,d\right).

Appendix A General Boundary Conditions

In this section we generalize the notion of boundary conditions. Let Ω\Omega be the space of configurations of ii-plaquettes in d.\mathbb{Z}^{d}. For a subset of vertices Vd,V\subset\mathbb{Z}^{d}, let FVkF^{k}_{V} be the set of kk-plaquettes with all vertices contained in V.V. Then given ξΩ,\xi\in\Omega, let

ΩΛnξ={ωΩ:ω(σ)=ξ(σ) for all σFdkFΛn1k}.\Omega_{\Lambda_{n}}^{\xi}=\left\{\omega\in\Omega:\omega\left(\sigma\right)=\xi\left(\sigma\right)\text{ for all }\sigma\in F^{k}_{\mathbb{Z}^{d}}\setminus F^{k}_{\Lambda_{n-1}}\right\}\,.

Intuitively, the boundary condition should describe how the states of external plaquettes affect the random-cluster measure within Λn.\Lambda_{n}. In the classical model, this is done by keeping track of which vertices of Λn\partial\Lambda_{n} are connected externally. In higher dimensions we will need slightly more information, but the idea is the same. Let Pξ,VP_{\xi,V} be the complex consisting of the union of the (i1)(i-1)-skeleton of d\mathbb{Z}^{d} and the open plaquettes of ξ\xi contained in FVi.F^{i}_{V}. Let Dni1D_{n}^{i-1} be the (i1)(i-1)-skeleton of Λn.\partial\Lambda_{n}. Then we construct a cubical complex Qω,ξQ_{\omega,\xi} (not necessarily a subcomplex of d\mathbb{Z}^{d}) by taking Pω,ΛnP_{\omega,\Lambda_{n}} and attaching a cubical complex AξA_{\xi} so that

  • AξPω,ΛnDni1.A_{\xi}\cap P_{\omega,\Lambda_{n}}\subset D_{n}^{i-1}.

  • The map φA:Hi1(Dni1;𝔽)Hi1(Aξ;𝔽)\varphi_{A}:H_{i-1}\left(D_{n}^{i-1};\,\mathbb{F}\right)\to H_{i-1}\left(A_{\xi};\,\mathbb{F}\right) induced by the inclusion Dni1AξD_{n}^{i-1}\hookrightarrow A_{\xi} is surjective.

  • The kernel of φA\varphi_{A} is the same as the kernel of the map Hi1(Dni1;𝔽)φP:Hi1(Pξ,dΛn1;𝔽)H_{i-1}\left(D_{n}^{i-1};\,\mathbb{F}\right)\to\varphi_{P}:H_{i-1}\left(P_{\xi,\mathbb{Z}^{d}\setminus\Lambda_{n-1}};\,\mathbb{F}\right) induced by the inclusion Dni1Pξ,dΛn1.D_{n}^{i-1}\hookrightarrow P_{\xi,\mathbb{Z}^{d}\setminus\Lambda_{n-1}}.

Such an AξA_{\xi} can be constructed by taking PξP_{\xi} and filling the (i1)(i-1)-cycles that are not homologous to cycles in Dni1.D_{n}^{i-1}.

Now we can define the plaquette random-cluster model on Λn\Lambda_{n} with boundary condition ξ\xi as follows:

μΛnξ(ω)={1ZΛnξ[σFΛnipω(σ)(1p)1ω(σ)]q𝐛i1(Qω,ξ;𝔽)ωΩΛnξ0otherwise.\mu^{\xi}_{\Lambda_{n}}\left(\omega\right)=\begin{cases}\frac{1}{Z^{\xi}_{\Lambda_{n}}}\left[\prod_{\sigma\in F^{i}_{\Lambda_{n}}}p^{\omega\left(\sigma\right)}\left(1-p\right)^{1-\omega\left(\sigma\right)}\right]q^{\mathbf{b}_{i-1}\left(Q_{\omega,\xi};\,\mathbb{F}\right)}&\omega\in\Omega_{\Lambda_{n}}^{\xi}\\ 0&\text{otherwise.}\end{cases}

In particular, the free boundary measure μΛn𝐟\mu_{\Lambda_{n}}^{\mathbf{f}} and the wired boundary measure μΛn𝐰\mu_{\Lambda_{n}}^{\mathbf{w}} are obtained by taking ξ\xi to be the all closed and all open configurations respectively.

Lemma 49.

Let p[0,1],q1,p\in\left[0,1\right],q\geq 1, and n.n\in\mathbb{N}. Then for every ξΩ,\xi\in\Omega, μΛnξ\mu^{\xi}_{\Lambda_{n}} is positively associated.

Proof.

The proof is analogous to the proof of Theorem 4.14 of  [Gri06]. Consider the plaquette random-cluster model μΛnAξ.\mu_{\Lambda_{n}\cup A_{\xi}}. This satisfies the FKG lattice condition and is thus strongly postively associated. Then since μΛnξ\mu^{\xi}_{\Lambda_{n}} is μΛnAξ\mu_{\Lambda_{n}\cup A_{\xi}} conditioned on the plaquettes of AξA_{\xi} being open, it follows that μΛnξ\mu^{\xi}_{\Lambda_{n}} is positively associated. ∎

Acknowledgments

We would like to thank Sky Cao, Matthew Kahle, and David Sivakoff for interesting and useful discussions.

References

  • [ACC+83] M. Aizenman, J. T. Chayes, L. Chayes, J. Fröhlich, and L. Russo. On a sharp transition from area law to perimeter law in a system of random surfaces. Comm. Math. Phys., 92(1):19–69, 1983.
  • [AF84] Michael Aizenman and Jürg Fröhlich. Topological anomalies in the n dependence of the n-states Potts lattice gauge theory. Nuclear Physics B, 235(1):1–18, 1984.
  • [Alt78] CP Korthals Altes. Duality for z(n)z(n) gauge theories. Nuclear Physics B, 142(3):315–326, 1978.
  • [BDC12] Vincent Beffara and Hugo Duminil-Copin. The self-dual point of the two-dimensional random-cluster model is critical for q1q\geq 1. Probability Theory and Related Fields, 153(3-4):511–542, 2012.
  • [Bre13] Glen E Bredon. Topology and geometry, volume 139. Springer Science & Business Media, 2013.
  • [Cao20] Sky Cao. Wilson loop expectations in lattice gauge theories with finite gauge groups. Communications in Mathematical Physics, 380(3):1439–1505, 2020.
  • [CC84] JT Chayes and L Chayes. The correct extension of the Fortuin-Kasteleyn result to plaquette percolation. Nuclear Physics B, 235(1):19–23, 1984.
  • [Cha16] Sourav Chatterjee. Yang–Mills for probabilists. In International Conference in Honor of the 75th Birthday of SRS Varadhan, pages 1–16. Springer, 2016.
  • [Cha20] Sourav Chatterjee. Wilson loops in Ising lattice gauge theory. Communications in Mathematical Physics, 377(1):307–340, 2020.
  • [Cha21] Sourav Chatterjee. A probabilistic mechanism for quark confinement. Communications in Mathematical Physics, 385(2):1007–1039, 2021.
  • [DCRT19] Hugo Duminil-Copin, Aran Raoufi, and Vincent Tassion. Sharp phase transition for the random-cluster and Potts models via decision trees. Annals of Mathematics, 189(1):75–99, 2019.
  • [DKS20] Paul Duncan, Matthew Kahle, and Benjamin Schweinhart. Homological percolation on a torus: plaquettes and permutohedra. arXiv preprint arXiv:2011.11903, 2020.
  • [DS] Paul Duncan and Benjamin Schweinhart. The plaquette Swendsen–Wang algorithm for Potts lattice gauge theory. In preparation.
  • [DW82] K Drühl and H Wagner. Algebraic formulation of duality transformations for abelian lattice models. Annals of Physics, 141(2):225–253, 1982.
  • [ES88] Robert G Edwards and Alan D Sokal. Generalization of the Fortuin-Kasteleyn-Swendsen-Wang representation and Monte Carlo algorithm. Physical review D, 38(6):2009, 1988.
  • [FK72] Cornelius Marius Fortuin and Piet W Kasteleyn. On the random-cluster model: I. Introduction and relation to other models. Physica, 57(4):536–564, 1972.
  • [FLV21] Malin Palö Forsström, Jonatan Lenells, and Fredrik Viklund. Wilson loops in finite abelian lattice gauge theories. to appear in Annales de l’Institut Henri Poincaré (B) Probabilités et Statistiques, 2021.
  • [Frö79] Jurg Fröhlich. Confinement in znz_{n} lattice gauge theories implies confinement in SU(n){S}{U}(n) lattice Higgs theories. Physics Letters B, 83(2):195–198, 1979.
  • [FS82] Jürg Fröhlich and Thomas Spencer. Massless phases and symmetry restoration in abelian gauge theories and spin systems. Communications in Mathematical Physics, 83(3):411–454, 1982.
  • [GG06] Benjamin T Graham and Geoffrey R Grimmett. Influence and sharp-threshold theorems for monotonic measures. The Annals of Probability, 34(5):1726–1745, 2006.
  • [GGZ80] Paul Ginsparg, Yadin Y Goldschmidt, and Jean-Bernard Zuber. Large qq expansions for qq-state gauge-matter Potts models in Lagrangian form. Nuclear Physics B, 170(3):409–432, 1980.
  • [GM90] Geoffrey Richard Grimmett and John M Marstrand. The supercritical phase of percolation is well behaved. Proceedings of the Royal Society of London. Series A: Mathematical and Physical Sciences, 430(1879):439–457, 1990.
  • [Gri04] Geoffrey Grimmett. The random-cluster model. In Probability on discrete structures, pages 73–123. Springer, 2004.
  • [Gri06] Geoffrey R Grimmett. The random-cluster model, volume 333. Springer Science & Business Media, 2006.
  • [Gut80] Alan H Guth. Existence proof of a nonconfining phase in four-dimensional u (1) lattice gauge theory. Physical Review D, 21(8):2291, 1980.
  • [Hat02] Allen Hatcher. Algebraic Topology. Cambridge University Press, 2002.
  • [HS16] Yasuaki Hiraoka and Tomoyuki Shirai. Tutte polynomials and random-cluster models in Bernoulli cell complexes (stochastic analysis on large scale interacting systems). RIMS Kokyuroku Bessatsu, 59:289–304, 2016.
  • [KMM04] Tomasz Kaczynski, Konstantin Michael Mischaikow, and Marian Mrozek. Computational homology. Springer, 2004.
  • [KPSS80] JB Kogut, RB Pearson, J Shigemitsu, and DK Sinclair. znz_{n} and nn-state potts lattice gauge theories: Phase diagrams, first-order transitions, β\beta functions, and 1/n1/n expansions. Physical Review D, 22(10):2447, 1980.
  • [KS82] Roman Koteckỳ and SB Shlosman. First-order phase transitions in large entropy lattice models. Communications in Mathematical Physics, 83(4):493–515, 1982.
  • [LMR89] L Laanait, A Messager, and J Ruiz. Discontinuity of the Wilson string tension in the 4-dimensional lattice pure gauge Potts model. Communications in mathematical physics, 126(1):103–131, 1989.
  • [Men86] Mikhail V Menshikov. Coincidence of critical points in percolation problems. In Soviet Mathematics Doklady, volume 33, pages 856–859, 1986.
  • [MMS79] R Marra and S Miracle-Sole. On the statistical mechanics of the gauge invariant Ising model. Communications in Mathematical Physics, 67(3):233–240, 1979.
  • [MO82] A Maritan and C Omero. On the gauge Potts model and the plaquette percolation problem. Nuclear Physics B, 210(4):553–566, 1982.
  • [MP79] G Mack and VB Petkova. Comparison of lattice gauge theories with gauge groups z2z_{2} and su(2)su(2). Annals of Physics, 123(2):442–467, 1979.
  • [OS78] Konrad Osterwalder and Erhard Seiler. Gauge field theories on a lattice. Annals of Physics, 110(2):440–471, 1978.
  • [Pis96] Agoston Pisztora. Surface order large deviations for Ising, Potts and percolation models. Probability Theory and Related Fields, 104(4):427–466, 1996.
  • [Sav16] Peter Saveliev. Topology illustrated. Peter Saveliev, 2016.
  • [Sei82] Erhard Seiler. Gauge theories as a problem of constructive quantum field theory and statistical mechanics. Springer Berlin Heidelberg, 1982.
  • [SW87] Robert H Swendsen and Jian-Sheng Wang. Nonuniversal critical dynamics in monte carlo simulations. Physical review letters, 58(2):86, 1987.
  • [Weg71] Franz J Wegner. Duality in generalized Ising models and phase transitions without local order parameters. Journal of Mathematical Physics, 12(10):2259–2272, 1971.
  • [Weg17] Franz J Wegner. Duality in generalized Ising models. Topological Aspects of Condensed Matter Physics: Lecture Notes of the Les Houches Summer School: Volume 103, August 2014, 103:219, 2017.
  • [WHK+07] Andreas Wipf, Thomas Heinzl, Tobias Kaestner, Christian Wozar, et al. Generalized Potts-models and their relevance for gauge theories. SIGMA. Symmetry, Integrability and Geometry: Methods and Applications, 3:006, 2007.
  • [Wil74] Kenneth G Wilson. Confinement of quarks. Physical review D, 10(8):2445, 1974.
  • [Wil04] Kenneth G Wilson. The origins of lattice gauge theory. arXiv preprint hep-lat/0412043, 2004.
  • [Yon78] Tamiaki Yoneya. z(n)z(n) topological excitations in Yang-Mills theories: duality and confinement. Nuclear Physics B, 144(1):195–218, 1978.