Topological Phases in the
Plaquette Random-Cluster Model
and Potts Lattice Gauge Theory
Abstract.
The -dimensional plaquette random-cluster model on a finite cubical complex is the random complex of -plaquettes with each configuration having probability proportional to
where is a real parameter and denotes the rank of the -homology group with coefficients in a specified coefficient field. When is prime and the coefficient field is , this model is coupled with the -dimensional -state Potts lattice gauge theory. We prove that the probability that an -cycle in 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 -dimensional plaquette random-cluster model on the -dimensional torus exhibits a sharp phase transition at the self-dual point in the sense of homological percolation. This implies a qualitative change in the generalized Swendsen–Wang dynamics from local to non-local behavior.
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
This model can then be extended to infinite graphs via limits on finite subgraphs when 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 -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 -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 -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 -complex associated to -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 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 counts the number of connected components regardless of the coefficient field..
In essence, Maritan and Omero defined a random -complex so
where the first Betti number counts the number of “independent loops” in 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 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 with yielding a random -complex that is not coupled with Potts lattice gauge theory for any choice of Nor is Potts lattice gauge theory coupled with a different generalized random-cluster model due to Chayes and Chayes [CC84] where 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 , , and the coefficient field Then, by setting and choosing 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 and taking the limit allows a comparison of -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 -dimensional plaquette random-cluster model on the -dimensional torus exhibits a sharp phase transition at the self-dual point marked by the the emergence of giant cycles which are non-trivial homology classes in the ambient torus. When 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 -dimensional plaquette random-cluster model, weights the probability of an -complex in terms of the -Betti number of the homology with coefficients in a field denoted We provide a definition of Betti numbers in Section 2 below.
Definition 1.
Let be a finite -dimensional cell complex, let and fix a field and parameters and The -dimensional plaquette random-cluster model on is the random -complex containing the full -skeleton of with the following distribution:
where is a normalizing constant, and and denote the number of -cells of and respectively.
As 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 -dimensional Potts lattice gauge theory assigns spins in an abelian group to the -dimensional faces of an oriented cell complex Denote by the set of functions from the oriented -cells of to so that cells with opposite orientations are mapped to inverse group elements. We follow the language of algebraic topology and call elements of 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 be a finite abelian group and let be a finite cubical complex. For let
(1) |
where is the coboundary operator and is the Kronecker delta function. -dimensional Potts lattice gauge theory on with coefficients in is the measure
where is a parameter (called the coupling constant or inverse temperature) and is a normalizing constant.
In this paper, we focus exclusively on the cases where is the additive group of integers modulo a prime number (which can be identified with the multiplicative group of -th complex roots of unity ). 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 and the multiplicative group of complex -th roots of unity though they are isomorphic. In the case where we will denote the Potts lattice gauge theory measure by or simply and call this measure -state Potts lattice gauge theory. We describe why the assumption that is prime is necessary in Section 1.2. We will be most interested in the Potts lattice gauge theory on and we will write 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 can be replaced with a complex matrix group In particular, -dimensional lattice gauge theory on with and , 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 and -state Potts lattice gauge theory coincide with (Ising) and “clock” lattice gauge theory respectively after an appropriate rescaling of the coupling constant (where denotes the multiplicative group of -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 -state Potts lattice gauge theory. These arise from evaluating a Potts state on a “loop” made of -cells. In topological language is called an -cycle.
Definition 3.
Let be an -cocycle. The generalized Wilson loop variable associated to is the random variable given by
where the superscript denotes that we are viewing the variable as a complex number. That is, if is the corresponding -th root of unity in
An important conjecture for the analogous quantities in Euclidean lattice gauge theory — called the Wilson area law — is that if is a -boundary and is the minimal bounding chain then the expectation of should decay as in some cases. For or 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 lattice gauge theory exhibits both “area law” and “perimeter law” phases [Gut80, FS82].
One-dimensional -state Potts lattice gauge theory on is thought to undergo a sharp phase transition as the coupling constant 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 and Then there are constants , and so that, for hyperrectangular -boundaries in
as all dimensions of are taken to In addition, if then
Recall that is defined as a complex number, so the expectation is complex as well. Here, the perimeter of an -boundary 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 consists of two vertices its perimeter is , and its area is the distance between and As such, the 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 when , but Laanait, Messager, Ruiz showed that such a transition occurs when 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 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 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 -dimensional -state Potts lattice gauge theory in for prime values of Towards that end, we prove an exact relationship between Wilson loop variables and a certain topological event.
An -chain is null-homologous in a cubical complex if for some -chain (in other words, the -dimensional homology group). Roughly speaking, when this means that is “bounded by a surface of plaquettes,” with the surfaces under consideration being dependent on the coefficient group For is null-homologous if and only if it is bounded by an orientable surface of plaquettes; when any surface of plaquettes will suffice. We denote the event that is null-homologous by with the choice of coefficients being understood in context. We have the following result.
Theorem 5.
Suppose is a prime integer, let and let be an -cycle in . Then
where is Potts lattice gauge theory and is the corresponding random cluster model.
While we state this theorem for the same proof works for any finite cubical complex. A precise definition of the infinite volume -dimensional plaquette random cluster model 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 is a prime integer and
Conjecture 6.
There exist constants and so that, for hyperrectangular -boundaries in
as all dimensions of are taken to Also, when .
When 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 -dimensional independent plaquette percolation in () 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 be a field, and let There exist positive, finite constants and so that, for hyperrectangular -boundaries in
(2) |
for all and so that
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 -plaquettes in [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 -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 -dimensional plaquette percolation in the dimensional torus. Here, we further generalize it to the -dimensional random-cluster model in the -dimensional torus.
Theorem 8.
Suppose and let be the induced map on homology with coefficients in a field with Denote by and the events that is non-trivial and surjective, respectively, and set Then
as
When is a prime integer and , 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 is distributed approximately as in a precise sense, where
Next, we prove that there are dual sharp phase transitions for and that are consistent with the critical probability for the random-cluster model in assuming a conjecture about the continuity of the critical probability in slabs. Let
Fix and let be the critical probability for the -dimensional random-cluster model with parameter on with free boundary conditions. This can be constructed by a limit of free random-cluster models on
Since it follows that is decreasing in Then let
Let be the critical threshold for the random-cluster model with parameter on These two critical values are conjectured to coincide.
Conjecture 9 ([Pis96]).
For all
Theorem 10.
Let Then the following statements hold:
If then
as
Furthermore, if then
as
In particular, if Conjecture 9 is true, then there are sharp thresholds at and respectively.
In general dimensions we demonstrate the existence of a sharp threshold function defined as follows: Let satisfy
and let and
Theorem 11.
Let be a prime integer and suppose that For every and
as
Moreover, for every and we have
and and have the following properties.
-
(a)
(Duality)
-
(b)
(Monotonicity in and ) for
1.2. Why does need to be prime?
Our results for -state Potts lattice gauge theory are for cases where the cohomology coefficients are a finite field where is a prime integer, instead of the more general additive group 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 where is dimension of the -vector space . If we were to use coefficients, we could try to replace with the rank of the -module (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,
This could be resolved by defining a dependent plaquette percolation model so that
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 -cycles and so that is homologous to Then an analogue of Theorem 5 fails when is defined to be the event that is bounded by a surface of plaquettes. However, we resolve this by redefining to depend on the choice of ; clearly in 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 -dimensional plaquettes (unit -dimensional cubes) for various 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 which is a union of all -plaquettes for which have corners in the integer lattice. The -dimensional torus also has a natural cubical complex structure obtained by dividing the cube into unit cubes and identifying opposite faces of the original cube. We call this cubical complex All examples we will consider in this paper are subcomplexes of either or
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 of linear combinations with coefficients in an abelian group (for a concrete example, take ) of -plaquettes called chains, and boundary operators, which map an -plaquette to a sum of its -faces. The continuous analogues of these concepts are, roughly speaking “-dimensional spaces on which one can integrate an -form” (called an -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 -plaquette is an edge and its boundary is the difference A two-plaquette is an oriented unit square with vertices and its boundary is More generally, let and let for and for Then is an -plaquette in and its boundary is given by
The reason that the sum is alternating in sign is so that the boundary operator satisfies the equation
(3) |

Of particular interest are the chains satisfying Such chains are called cycles, the space of which is denoted Equation 3 provides one source of -cycles, namely the boundaries of -plaquettes. We denote this space 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 seen in Figure 1. In the cubical complex structure we defined earlier, we have that consists of multiples of a single -cycle, namely the boundary of But now imagine that we make our lattice spacing instead of Now contains linear combinations of the boundaries of the four -plaquettes contained in One such combination can be seen in Figure 2. Though has not changed, our choice of lattice has made the cycle space significantly more complicated.

Therefore, in order to measure the shape of spaces in a way that does not depend on our choice of lattice, we define the -dimensional homology group as the quotient
Returning to the unit square every -cycle is a boundary regardless of the cubical complex structure on , so 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 by continuously deforming to a point, which has no nonzero -chains let alone -cycles.
For an example with nontrivial first homology, consider where is the topological boundary of i.e. the empty square formed by the union of the four sides of There is still the -cycle from before but there are no -plaquettes, so This example illustrates the often repeated informal description of as a measurement of the number of -dimensional holes of In the case of the cycles that remain are the loops that cannot be filled in by -plaquettes.
We can now define giant cycles in a subcomplex of the torus. If consists of a subset of the cubical cells of then the chain groups of are subgroups of the chain groups of The giant cycles of are the cycles that are not boundaries when considered as chains in In the case a giant cycle is a loop that spans the torus in some sense, and must therefore have length at least We interpret this as a finite volume analogue of an infinite path in the lattice. See Figure 3 for illustrations of two giant cycles.


2.2. Coefficients and Torsion
So far, the choice of coefficient group 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 for prime denoted 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 -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 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 Consider the second homology of both spaces. In both cases there are no -plaquettes and only one -plaquette so it suffices to check if the -chains generated by contain nontrivial cycles. In the torus,
so for any field In the Klein bottle,
which is nonzero over most fields, with a number of exceptions including Thus, and for any prime There is also a difference in first homology; one can compute and for 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 and higher. Homology that appears over some finite fields but not others is sometimes called torsion.

It is worth mentioning a special situation that arises for the -th homology group. It is not hard to check that where is the number of connected components of 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
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 -th Betti number of a cubical complex with coefficients in by
Also, let be the set of -plaquettes of and let be their number. The Euler characteristic of is then
The following theorem is a straightforward consequence of the rank–nullity theorem.
Theorem 12 (Euler–Poincaré formula).
For any cubical complex and any field
2.4. Cohomology
A small algebraic change in the definitions that give the homology groups leads to cohomology. The -th cochain group is defined as the group of -linear functions from to Since we are using field coefficients, this is the dual vector space to Then the coboundary operator is defined by
for The continuous analogues of an -cochain and and its coboundary are a differential form and its exterior derivative. Note that if is an -cochain and is an -chain, the evaluation corresponds to integration and the definition of the coboundary operator corresponds to Stokes’ Theorem.
Analogously to homology groups, we define the -th cocycle group to be the kernel of and the -th coboundary group to be the image of applied to Then -th cohomology group is the quotient
In the continuous setting, the analogue of the -dimensional cohomology group is the -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 computed using a cubical complex structure on [Bre13].
In some ways cohomology does not add more information than homology already provided. The universal coefficient theorem for cohomology tells us that when 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
for 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 copies of the circle In such a product space, the Künneth formula for homology (Section 3B of [Hat02]) tells us that there is an isomorphism of the form
Since and for all we see that
Furthermore, there is a set of generating -cycles consisting of the products of of the possible 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 is the complex obtained by shifting by in each coordinate direction, and the dual of is defined similarly. These pairs of dual complexes have the property that each -plaquette intersects exactly one -plaquette of 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 we define the dual complex to be the union of all plaquettes for which the dual plaquette is not included in When comes from the plaquette random-cluster model defined below, it will include the union of all lower dimensional plaquettes of (called the “-skeleton”) and no plaquettes of dimension higher than Note that in that case, automatically contains the -skeleton of Note that each statement in this section has an analogue for subcomplexes of a box in 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 deformation retracts to
Duality allows us to relates the homology of with that of Importantly, “global” cycles and “local” cycles behave differently. Let and be the maps on homology induced by the respective inclusions. Set and We think of as counting the local cycles (cycles of the plaquette system that are boundaries in the ambient torus) and as counting the global cycles. Furthermore, recall that and let
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 or of Euclidean space , 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 and any the following relations hold.
(4) |
(5) |
and,
(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
where denotes the relative homology group of with respect to a subset Combining this with the long exact sequence of relative homology (see Section 2.1 of [Hat02]), we obtain the following commutative diagram:
By Proposition 13 and the definition of the plaquette system,
so is surjective. Similarly,
At first glance, the dimensions in Equation 6 may seem off: the plaquette random-cluster model is weighted by the -th Betti number and is dual to a -dimensional model, but this equation relates its -th Betti number to the -st Betti number of its dual. However, the random-cluster model always contains all cells of dimension lower than (that is, the -skeleton). The presence or absence of -cells can only affect the Betti numbers in dimensions and 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 so that for any -dimensional subcomplex of containing the full -skeleton
(7) |
where denotes the number of -cells of
Proof.
By the Euler–Poincaré formula
(8) |
and the desired statement follows from the observations that only depends on and for and that also only depends on and when ∎
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 -dimensional plaquette random-cluster model has the distribution defined by
where 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 a configuration, and we refer to the elements of as open plaquettes and the elements of as closed plaquettes of respectively. We then define the associated complex to be the union of the -skeleton of and the open plaquettes of
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 and and a finite cubical complex. Then satisfies the FKG lattice condition and is thus positively associated, meaning that for any events that are increasing with respect to
Proof.
The key observation is that for any topological spaces and any
(9) |
To see this, consider a portion of the Mayer-Vietoris exact sequence (refer to Section 2.2 of [Hat02]):
The first isomorphism theorem gives
By exactness, and so putting everything together yields
Now since is strictly positive for it is enough to check the FKG lattice condition, which, written in terms of configurations, requires that
for any configurations Since
we only need to check that
where is the complex associated to (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 will be defined via limits of plaquette random-cluster models on finite boxes 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 can be defined as the i-random-cluster model on the subcomplex with vertices The wired boundary measure can be defined as the same measure conditioned on all plaquettes contained in being open. By an analogue of Theorem 4.19 of [Gri06], the weak limits of each of these measures exist.
Proposition 17.
Let and
-
(a)
The limits and exist.
-
(b)
and are automorphism invariant.
-
(c)
and are positively associated.
-
(d)
and are tail-trivial, that is, for any event that is independent of the states of any finite set of plaquettes,
-
(e)
For any nontrivial translation and are -ergodic, that is, any -invariant random variable is -almost-surely and -almost-surely constant.
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 and where
Beffara and Duminil-Copin [BDC12] used this relationship to prove that the critical probability is the self-dual point where namely
In this section, we consider a similar duality in higher dimensions within the torus, showing that the dual of the -dimensional plaquette random-cluster model is approximately a -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 be the inclusion map and let be the induced map on th homology. We say that an element is a giant cycle if and a local cycle if Recall that is informally the number of giant cycles in
First we define a \saybalanced version [BDC12] which satisfies exact duality:
The additional term 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 and the same is true for their dual models. As such, a sharp threshold for implies one for
Theorem 18.
The balanced plaquette random-cluster model satisfies
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
(10) |
Recall also that and contain the complete -skeleton and -skeleton respectively, so
(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
(12) |
Lastly, we recall the following property of
(13) |
As a corollary, we have the following duality relationship between the normalizing constants.
Corollary 19.
where is the constant defined in Proposition 15.
5. Relation to Potts Lattice Gauge Theory
We now elucidate the relationship between the -dimensional plaquette random-cluster and -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 be a finite cubical complex, be a prime, and Consider the coupling on defined by
where is the Kronecker delta function. Then has the following marginals.
-
•
The first marginal is the -state Potts lattice gauge theory with inverse temperature given by
where is the Hamiltonian for Potts lattice gauge theory:
-
•
The second marginal is the plaquette random-cluster model with parameters given by
Proof.
We include Hiraoka and Shirai’s proof adapted to our notation for completeness. The first marginal is calculated as
The second marginal is calculated as
The second to last line follows from the definition of a cocycle. The last line holds because the -skeleton of does not depend on so is fixed. ∎
Next, we compute the conditional measures defined by the coupling.
Proposition 21.
Let Then the conditional measures of are as follows.
-
•
Given the conditional measure is Bernoulli plaquette percolation with probability on the set of plaquettes that satisfy
-
•
Given the conditional measure is the uniform measure on -cocycles in
Proof.
From the definition of the coupling, under a plaquette is open with probability independently of other plaquettes when and always closed otherwise, giving the first conditional measure. The second conditional measure is determined by the observation that is supported on the set of satisfying for each 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 a spin configuration is updated by first sampling a random graph by performing percolation with probability on the edges between vertices with equal spins. Then, the spins are resampled uniformly on each component of In our language, this corresponds to sampling a uniformly random cocycle in 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 a random -complex is sampled by including plaquettes on which vanishes independently with probability An updated cochain is then found by selecting a uniformly random cocycle from 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 and 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 are trivial in the homology of the torus. On the other hand, when contains giant cocycles up to cohomology (each given the same weight), and a “non–local” move happens with probability approaching (as has 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 or by repeating the same proofs.. Interestingly, at the number of giant cocycles is somewhere between and with probability bounded away from and so the probability of non-local behavior is non-zero but lower than above 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 which is defined to be the usual Potts lattice gauge theory on the subcomplex induced by the vertices of (we use the superscript because this measure can be coupled with the free random-cluster measure as in Proposition 20). Next, we construct the limiting measure by taking 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 in a consistent manner.
Lemma 22.
For any configuration there is a basis of so that each -face of is in the support of finitely many elements of
Proof.
Take any basis and perform Gaussian elimination. ∎
We will call the minimal such basis with respect to lexicographical order
Corollary 23.
Let and
-
(a)
Let be distributed according to Conditional on let be i.i.d. random variables. Then the limit
exists, and the random cocycle
is distributed according to
-
(b)
Let be distributed according to Conditional on let be a random configuration in which each -plaquette is open with probability if independent of the states of other plaquettes and closed otherwise. Then is distributed according to
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 so that each is distributed according to and is distributed according to Moreover, for any -plaquette for large enough
Now let be i.i.d. random variables, and let
By the construction of is distributed according to and is eventually constant on any finite set of -cubes. Since is distributed as and we are done.
(b) Let be independent random variables. Also, let be distributed as coupled as before so that is eventually constant on any finite set of -cubes. Then, Proposition 21, is distributed according to Therefore is distributed according to ∎
In the remainder of this section, we will refer to “Potts lattice gauge theory on ” and “the plaquette random-cluster model on ” 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 be the normalizing constant for -state -dimensional Potts lattice gauge theory on Then, there are constants so that
where
Proof.
Note that the proofs can be adapted to relate the partition function for -dimensional Potts lattice gauge theory on a box in with free boundary conditions to that of -dimensional Potts lattice gauge theory on a box in with wired boundary conditions.
5.4. Generalized Wilson Loop Variables and Proof of Theorem 5
In this section, let either be a finite cubical complex or Let be an -dimensional cycle. It is natural to ask whether is a boundary in the -dimensional plaquette random-cluster model.
Definition 25.
Let to be the event that is null-homologous in for a specified coefficient field
The generalized Wilson loop variable is a closely related quantity for Potts lattice gauge theory.
Definition 26.
Let be an -cocycle. Recall that the generalized Wilson loop variable associated to is the random variable given by
We are particularly interested in the behavior of these random variables when is an -dimensional hyperrectangular boundary in — that is, the boundary of a hyperrectangle inside where for and or a congruent object obtained translation and/or by permuting the coordinates of 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 and Then for any so that in and any so that in we have
Proof.
First we show that If and there is nothing to prove. Now, assume For let be the dual of when is viewed as an element of Note that since is a cycle,
Then, because and are cohomologous, we can write
for some It follows that
Now as there is a chain so that Then since is a cocycle, we can calculate
∎
We now describe the distribution of Wilson loops as a function of the associated plaquette random-cluster model.
Proposition 28.
Let be a prime integer and let Then
and
Proof.
Fix and If then by Lemma 27.
Now assume that By Proposition 21, can be sampled by fixing a basis of and taking a random linear combination with independent coefficients. Thus, is uniformly distributed on an additive subgroup of Since the only such subgroups are and 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 so that Therefore, is distributed as so is distributed as ∎
The proposition has a few interesting corollaries.
Corollary 29.
Let
and
In particular, if is the event that is null-homologous in then
The last statement is Theorem 5.
Corollary 30.
Let Then conditioned on
almost surely. Conversely, conditioned on
and are independent.
Proof.
Conditioned on we can simply apply Proposition 28 to the Wilson loop variable for to get the desired result.
Now condition on the event First, we show independence of and when further conditioning on a fixed configuration If then and are trivially independent. Now assume Since holds, we can extend the set to a basis of again using the universal coefficent theorem for cohomology. Then by Proposition 21, we can sample by taking a random linear combination of elements of with independent coefficients. By construction, we then see that and are independent, and thus and are independent.
Now note that by the same argument as in Proposition 28. Then for any we can compute
Thus, and are independent conditioned on
∎
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 let
Note that if and and are sites of then is the two-point correlation between the sites.
Proposition 31.
where
Proof.
For simplicity, we write in the following calculation.
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 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 is stochastically decreasing in for fixed On the other hand, if we fix then is stochastically increasing in
Proof.
This is a consequence of the fact that adding an -plaquette can only reduce by one or leave unchanged.
Recall that by Equation 19, is the conditional probability that a plaquette is open given that it reduces by one when leaving the states of the other plaquettes unchanged. First fix Since is decreasing in and the conditional probability that a plaquette is open given that it does not kill an -cycle is constant in (equaling ), an application of Theorem 38 shows that is stochastically decreasing in Now fix Then is decreasing as a function of so again applying Theorem 38 to the conditional probabilities given by Equation 19 yields that is stochastically increasing in ∎
A similar statement for follows immediately since it is a local limit of random-cluster measures on finite cubical complexes.
Corollary 33.
is stochastically decreasing in for fixed If we fix is stochastically increasing in
Next, we require two more results of [ACC+83] for plaquette percolation.
Theorem 34 (Aizenman, Chayes, Chayes, Frölich, Russo [ACC+83]).
Consider -dimensional Bernoulli plaquette percolation on There exist finite, positive constants so that, for hyperrectangular -boundaries in
(14) |
Furthermore, there are constants so that
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 be a sequence of hyperrectangular -boundaries whose dimensions diverge with Then
converges, where denotes an infinite volume random-cluster measure, constructed with either free or wired boundary conditions.
Proof of Theorem 7.
6. Phase Transitions for Homological Percolation
In this section, we prove the existence of sharp phase transitions for the -state -dimensional plaquette random-cluster model on in the sense of homological percolation. First, we recall the definition of homological percolation. Let be an -dimensional subcomplex of and let be the map induced by inclusion. Then counts the number of giant cycles of A homological percolation phase transition occurs at if with high probability if and attains its maximum possible value, with high probability if
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 where is a giant -cycle.
Proposition 36.
Let and let be a subcomplex containing a giant -cycle. Then there exist Polyakov loops so that and
Proof.
Let be a giant -cycle. First, we show that a transverse -dimensional slice of contains a giant -cycle. Let and Subdivide the cubical complex to add an -plaquette at each nonempty intersection of a -plaquette of with or for each Call this subdivision and give and the resulting cubical complex structures.
Without loss of generality assume that is homologous to a sum of the standard generators of the torus, including at least one that is orthogonal to i.e. a standard generator that is a product of many 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 (remember that now has subdivided cells that were not originally in ), let
Note that each -plaquette of intersects in either an -plaquette or not at all. Also, let be constructed as with replaced by Observe that the supports of and are at distance at least apart.
Consider Since is disconnected, we can write it as a union of its connected components Let
Let
and
Then is an -chain with boundary so and are homologous. Then it only remains to show that they are nontrivial giant cycles, as we can then shift them by to obtain the desired Polyakov loops.
We claim that and are giant -cycles within and respectively. Since and are homologous, it is enough to show that at least one of them is a giant cycle. Suppose that neither are, so there are -chains so that and so is an -cycle. Moreover, since is contained in a subset of that deformations retracts to namely it is homologous within to an -cycle contained in Similarly, we set and can find contained in that is homologous to within Then we have
But the -tori and can only contain certain giant cycles described by the Künneth formula. In particular, they cannot contain orthogonal homology classes, contradicting the assumption about made earlier. ∎
Corollary 37.
Let let be an odd prime, and set
If then -dimensional then, with high probability, -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 we can take
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 be a finite set and let be probability measures on We say that is stochastically dominated by if there is a probability measure on with first and second marginals and so that
In this case we write 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 be a finite set and let be strictly positive probability measures on Suppose that for each pair with and each
Then
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 so that the following holds. Let and let be the set of subsets of Let be an increasing event. Let be a positive monotonic probability measure on Let and set If there exists a subgroup of the symmetric group on elements acting transitively on so that and are -invariant, then
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 and Let Then
-
•
there exists a so that for any sufficiently close to
-
•
for any there exists a so that for every
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 be a finite dimensional vector space and be a set. Let be the lattice of subspaces of Suppose is an increasing function, i.e. if then Let be a finite group which acts on both and whose action is compatible with That is, for each and Let be a -valued random variable with a -invariant distribution that is positively associated, meaning that increasing events with respect to are non-negatively correlated. Then if is an irreducible representation of , there are positive constants so that
where only depends on and only depends on
6.3. Proof of Theorem 8
We recall that is the event that the map on homology induced by the inclusion is nonzero and is the event that the same map is surjective. Note that by the intermediate value theorem we may define a threshold function so that for each We begin by showing a general threshold result for homological percolation. The proof is very similar to that in [DKS20].
Proposition 42.
Let be an odd prime and For any we have
and
as
Proof.
We prove the second statement first. Note that increasing events with respect to are positively correlated by Theorem 16 and that the group of symmetries of the torus contains an irreducible representation of (see Proposition 13 of [DKS20] for a proof). Then by Lemma 41,
where do not depend on In particular there is a so that for all
Now we show that has a sharp threshold. By Theorem 16, satisfies the FKG lattice condition and is thus monotonic [Gri04]. Then since the symmetries of act transitively on the plaquettes, we apply Theorem 39 to obtain
In particular, for close to By integrating this inequality, we have as
To obtain the first statement, we apply the same argument to the dual system. Since it follows that by Equation 5. Then since is continuous as a function of the above argument shows that as By another application of Equation 5, we then have as
∎
Corollary 43.
If then
as If then
as
It is now straightforward to prove that the -dimensional random-cluster model in exhibits a sharp phase transition at the self-dual point.
6.4. Proof of Theorem 10
Next, we will prove our theorem on the homological percolation transition for the -dimensional and -dimensional random-cluster models in Throughout this subsection, we fix and we borrow terminology from percolation on graphs. Let be a graph and let and be its vertex and edge sets respectively. Designate the edges of as open or closed according to some probability measure. For we write if is connected to by a path of open edges. For a vertex subset we write if is connected to by a path of open edges that only passes through vertices of
We will prove the statements of Theorem 10 for the case in each regime separately. The case 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 and let For a vertex of let the event that a giant -cycle passes through Since is contractible in Moreover, stochastically dominates so by Theorem 40 and translation invariance,
(15) |
for all vertices of
Let be the number of vertices in that are contained in a giant -cycle. Since
by Markov’s Inequality | ||||
by Equation 15 | ||||
as .
∎
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 -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 -cycle.
Let
Let be the boundary of in
Lemma 44.
Fix and There is a so that for any sufficiently close to and any sufficiently large,
The proof follows from the proof of Theorem 40 (Theorem 1.2 in [DCRT19]), replacing with 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
Lemma 45.
Fix and Let There is a constant not depending on so that
Proof.
Consider the random-cluster model Define three crossing events
and
For an event defined in relation to the origin let denote the event obtained by translating the origin to By Lemma 44 and symmetry, there is a not depending on so that
(16) |
Let and Denote the projection onto the first two coordinates by Our aim will be to create overlapping paths (i.e. paths with intersecting images under ) containing and 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 under Fix an arbitrary ordering of the finite paths in Assuming that occurs, let be the endpoint of the minimal path witnessing that event and let Then, the event
(17) |
ensures the existence of a path through with a single gap at which projects to a vertical crossing, as illustrated in in Figure 5.
Next, let be the event rotated radians counterclockwise in the first two coordinates about the point By symmetry, and that event ensures the existence of a path through with a single gap which projects to a horizontal crossing of

Now, assume that occurs so and are connected in We count how many edges we would need to add in order to connect the overlapping segments in , thus connecting to Any two points in with the same image under are at graph distance at most There is one point of overlap between and and one more from the symmetric event for In addition, we have a third point of overlap between the horizontal and vertical crossings. As such, we can connect to using at most additional open edges.
Let Given a configuration define
For an event we then define
Then by Theorem 3.45 of [Gri04] and the FKG inequality there is a constant so that
where we used Equations 16 and 17 in the final step. Since this bound does not depend on we are done.
∎
Proof of Theorem 10 (supercritical regime).
Let Then there is a such that Let We will construct a giant cycle in percolation with parameter by using Lemma 45 to connect the centers of four pairwise overlapping boxes in the torus, each of diameter If is not divisible by the starting and ending points of this constructed path may not exactly match. However, they will be at graph distance at most and are therefore connected with probability at least We may therefore assume that is divisible by in the remainder of the proof for simplicity. We apply Lemma 45 to copies of
that are centered at and to connect and For convenience, let denote the event
If the events and all occur, then there is an open path that is homotopic to the standard generator of contained in Thus,
Let denote the measure We then apply the FKG inequality and Lemma 47 to bound
Note that the final bound is uniform in Again using the FKG inequality, we obtain
Then by applying Theorem 39 as in Proposition 42, we have
as
∎
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 and
Proof.
We now turn to showing monotonicity in the critical probabilities. First we require a lemma comparing the -dimensional plaquette random-cluster percolation on a cubical complex to the random-cluster model on a subcomplex.
Lemma 47.
Fix Let be finite -dimensional cubical complexes with and recall that and are the -dimensional plaquette random-cluster measures with parameters on and respectively. Then stochastically dominates
Proof.
Let be an -cell. Let be a subcomplex of let and let Write and and similarly and Since has no -cells,
Then by the Euler–Poincaré formula we see that
(18) |
Recall that for any -cell the random subcomplex with distribution satisfies
(19) |
Finally, we compare percolation within subcomplexes of the torus in order to obtain the desired inequalities between critical probabilities.
Proposition 48.
For all
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
Our strategy will be to define a sequence of models between the random-cluster model on and 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 let be the associated subspace of giant cycles of in Then we say if there is a coupling of and so that
Note that implies
Let and For let be so that is the random-cluster model on and sets all plaquettes to be closed almost surely. By Lemma 47, we have
We now put a different measure on configurations that are closed outside Let be the set of -cells of contained in and let be the rest of the -cells of Let and be the number of open cells of and respectively. We set We then let be independent Bernoulli plaquette percolation with probability and declare all other plaquettes closed. More explicitly,
We think of this as doing Bernoulli percolation on with parameter and then a random-cluster percolation with free boundary conditions on For and a configuration on we clearly have
By Lemma 47, for we also have
The again applying Theorem 38, we have
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 be the set of -faces of that intersect, but are not contained in For an -face of let be the set of all perpendicular -faces that meet at an face. Then and are the -faces of an -face Also, for a perpendicular -face of let Let satisfy
For all pairs where and let be independent random variables. Then by construction, Bernoulli percolation with parameter on is equivalent to setting each cell to be open if and only if Given Bernoulli percolation on let
Let the have as open plaquettes the union of the open plaquettes and the plaquettes in (so all plaquettes of are all closed). Since the boundary of a plaquette in is necessarily the boundary of an open set of plaquettes,
Now let (i.e. the probability that a plaquette of is either open or in ) and take to be the random-cluster model on with parameter instead of Then we compare the probabilities in and in that a plaquette is open in the cases that its state does or does not affect In former case, the conditional probabilities are and respectively, and in the latter both are Thus, for any configuration on and any
Therefore we have
by Theorem 38, and so
Now, as in [DKS20], we may take so that Then for each we have
so
and thus
Combining this with Proposition 46 also gives ∎
Appendix A General Boundary Conditions
In this section we generalize the notion of boundary conditions. Let be the space of configurations of -plaquettes in For a subset of vertices let be the set of -plaquettes with all vertices contained in Then given let
Intuitively, the boundary condition should describe how the states of external plaquettes affect the random-cluster measure within In the classical model, this is done by keeping track of which vertices of are connected externally. In higher dimensions we will need slightly more information, but the idea is the same. Let be the complex consisting of the union of the -skeleton of and the open plaquettes of contained in Let be the -skeleton of Then we construct a cubical complex (not necessarily a subcomplex of ) by taking and attaching a cubical complex so that
-
•
-
•
The map induced by the inclusion is surjective.
-
•
The kernel of is the same as the kernel of the map induced by the inclusion
Such an can be constructed by taking and filling the -cycles that are not homologous to cycles in
Now we can define the plaquette random-cluster model on with boundary condition as follows:
In particular, the free boundary measure and the wired boundary measure are obtained by taking to be the all closed and all open configurations respectively.
Lemma 49.
Let and Then for every is positively associated.
Proof.
The proof is analogous to the proof of Theorem 4.14 of [Gri06]. Consider the plaquette random-cluster model This satisfies the FKG lattice condition and is thus strongly postively associated. Then since is conditioned on the plaquettes of being open, it follows that 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 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 . 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 lattice gauge theories implies confinement in 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 expansions for -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. and -state potts lattice gauge theories: Phase diagrams, first-order transitions, functions, and 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 and . 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. topological excitations in Yang-Mills theories: duality and confinement. Nuclear Physics B, 144(1):195–218, 1978.