Skip to main content
Proceedings. Mathematical, Physical, and Engineering Sciences logoLink to Proceedings. Mathematical, Physical, and Engineering Sciences
. 2020 Aug 19;476(2240):20200215. doi: 10.1098/rspa.2020.0215

Statistics of geometric clusters in Potts model: statistical mechanics approach

P N Timonin 1,
PMCID: PMC7482199  PMID: 32922154

Abstract

The percolation of Potts spins with equal values in Potts model on graphs (networks) is considered. The general method for finding the Potts clusters' size distributions is developed. It allows full description of percolation transition when a giant cluster of equal-valued Potts spins appears. The method is applied to the short-ranged q-state ferromagnetic Potts model on the Bethe lattices with the arbitrary coordination number z. The analytical results for the field-temperature percolation phase diagram of geometric spin clusters and their size distribution are obtained. The last appears to be proportional to that of the classical non-correlated bond percolation with the bond probability, which depends on temperature and Potts model parameters.

Keywords: percolation theory, Potts model, geometric clusters

1. Introduction

The thermodynamics of q-state Potts model is thoroughly studied on a vast set of graph and lattices [13]. In many cases, the analytical and numerical results are obtained for its phase diagram, order parameters and critical indexes both for ferromagnetic and antiferromagnetic case. This interest stems from the existence of a number of the model's physical realizations that range from percolation, resistors networks and epidemic spreading [3] to community detection [4]. For these applications, a more detailed picture and the phase transition mechanisms can be elucidated in the studies of statistics of the like-valued (geometric) spin clusters of Potts model. Geometric clusters distribution is, in principle, an observable quantity and its knowledge can be important for the physical realizations of this model. Also, percolation of geometric Potts clusters is very important from the theoretical point of view providing the example of correlated percolation [5].

The influence of geometric clusters percolation in the ferromagnetic spin models on their thermodynamics attracted much attention very early [69]. At first, it was suggested that the ferromagnetic transition results from the appearance of infinite percolation cluster of the like-valued spins. However, very soon it was found that such percolation transition may not coincide with the ferromagnetic one [6,7] and this stimulated further studies of such correlated percolation. Coniglio & Klein [9] noted that Ising models on arbitrary lattice can have percolation transition right at the ferromagnetic critical point and it has the same Ising scaling indexes. But in this case, the so-called Fortuin–Kasteleyn (FK) clusters percolate. FK clusters [10] are constructed as subsets of geometric clusters via random placement of bonds in them with the temperature-dependent probability p = 1 − eJ/T. The percolation of FK clusters is also studied in Potts model on the Euclidean lattices [11,12] and on the complete graph [13]. Here, the properties of Kertész line Tp(H) below which such clusters percolate are determined. Meanwhile the study of true observable geometric like-valued clusters without FK condition deserves special attention.

This is notoriously hard problem, which is usually tackled numerically via Monte-Carlo technique [1418] or via enumeration of clusters in random bonds [19] and sites [20,21] patterns. Yet such numerics is very time-consuming for sufficiently large samples. The exact analytical results can be obtained with the replica method only for some simple graphs (e.g. various chains and ladders) and without it for the graphs with the tree-like structure.

One can overcome the computational problems via casting percolation into statistical mechanics framework as has been done for the classical bond percolation [10,22]. This may help to diminish the numerical efforts due to a number of exact methods developed for partition function calculation such as transfer matrix technique, renormalization group and Monte Carlo simulations. Here, we suggest the statistical mechanics approach for the determination of size distribution of geometric clusters for Potts model on arbitrary graph.

In §2, the general method is described for the Potts model on a graph, in §§3 and 4, the results of its implementation on the Bethe lattice ferromagnetic Potts model are presented and §5 is devoted to the discussion and conclusions.

2. General formalism

Consider a graph with N sites numbered by integers i = 1,…, N, and set of bonds E between some sites.

Potts distribution on graph (non-normalized) is

ρ(σ,h)=exp{Ki<jEδ(σi,σj)+hi[1δ(σi,0)]},σi={0,1,,q1},K=JT,h=HT 2.1

and

g(σi,σj,τi,τj,α)=1+δ(σi,σα)δ(σj,σα)[δ(τi,τj)1],τi={0,1,,q~1}, 2.2
σα=α,α=0,1g(σi,σj,τi,τj,α)={δ(τi,τj)if σi=α and σj=α1if σiα or σjα.

Thus, in a given spin configuration all bonds in clusters with all sites having Potts variable σ equal to α, σ = α (α − clusters, in short) are endowed with factor δ(τi, τj). Hence, variables τi have the same values at all sites of every α − cluster. Consider partition function

Zq~(ζ,h,α)=Trσ,τU(σ,τ,ζ,h,α),U(σ,τ,ζ,h,α)=ρ(σ,h)i,jEg(σi,σj,τi,τj,α)i=1Nζ(1δτi,α).

In FK cluster representation [10], it has the form

Zq~(ζ,h,α)=Trσ{ρ(σ,h)[τ=0q~1ζ(1δτ,α)]βαNβ(σ)s=1N[τ=0q~1ζs(1δτ,α)]Ns(α)(σ)}=Trσ{ρ(σ,h)[1+(q~1)ζ]βαNβ(σ)s=1N[1+(q~1)ζs]Ns(α)(σ)}, 2.3

where Ns(α)(σ) is the number of clusters with s sites, each having σ = α in configuration σ. Apparently,

βαNβ(σ)NP=1Nα(σ)NP.

At q~1 small, we have approximately

Zq~(ζ,h,α)Z1(ζ,h,α)[1+(q~1)s=1NζsNs(α)(σ)NP+(q~1)ζ(1Nα(σ)NP)]. 2.4

Furthermore, we note that at q~=1 partition function (2.4) becomes the ordinary Potts one on a graph considered

Z1(ζ,h,α)=Trσρ(σ,h)ZPotts(h),

and the skew brackets with sub-index P denote ordinary Potts averages

A(σ)PZP1(h)Trσ[ρ(σ,h)A(σ)].

Let us define the thermodynamic potential

Fq~(ζ,h,α)=limNN1lnZq~(ζ,h,α). 2.5

Note that at q~=1 Fq~(ζ,h,α) becomes proportional to the ordinary Potts potential

F1(ζ,h,α)=βFPotts(h).

Then their difference

Φq~(ζ,h,α)=Fq~(ζ,h,α)F1(h) 2.6

define the generating function for clusters' size distribution

Gα(ζ,h)=s=1ζsνs(α),νs(α)=limNNs(α)(σ)NP. 2.7

Here, νs(α) is the average number (per site) of α − clusters with s sites (composed of sites with σi = α). Indeed, we have from (2.4)

Gα(ζ,h)=q~Φq~(ζ,h,α)|q~=1ζ(1cα),cα=limNNα(σ)NP. 2.8

Now we only need to find cα- the average fraction of sites with σ = α. We can express them via average Potts spin

limNN1i=1NσiP=σiP=β=0q1βcβ.

Due to the permutation symmetry of Potts spins with σi ≠ 0 cα = c1 for all α > 0 so

σP=c1β=0q1β=c12q(q1).

By definition, we also have β=0q1cβ=c0+(q1)c1=1. Thus we have

c0=12qσiP,c1=2q(q1)σiP.

Note also that cα can be expressed via Potts ‘magnetization’ conjugate to field h

mP=F1h=1δ(σi,0)P=1c0=(q1)c1. 2.9

Once we have Gα(ζ, h) from (2.8) we can find the average number of finite α − clusters

ncl,α=s=1νs(α)=Gα(1,h),

and the average number of sites in them

nαsites=s=1sνs(α)=ζGα(ζ,h)|ζ=1.

Let Pα to be the fraction of α − sites belonging to the infinite percolation α − cluster then

Pα=cαnαsites,

Thus we get a full description of the Potts clusters percolation on a graph if we manage to calculate the potential Φq~(ζ,h,α) (2.6).

3. Application to a Bethe lattice

The Bethe lattice is a popular example of a hierarchical graph for which a wealth of analytical results on phase transitions in spin and percolation models were obtained [3,6,8,23,24]. Here, we apply the above method to find analytically the geometric clusters' size distribution for Potts model on a general Bethe lattice with a coordination number z. The potential (2.5) for it can be found with the partial partition function for the Caley tree of l-th order Vα,l(σ, τ) summed over all spins except the root ones (σ, τ) [23]

Fq~(ζ,h,α)=limNN1lnZq~(ζ,h,α)=2z2limllnTrσ,τeh[1δσ,0]ζ1δτ,αVα,lz(σ,τ), 3.1

Vα,l(σ, τ) obeys the recursion relations

Vα,l+1(σ,τ)=σeKδ(σ,σ)+h[1δ(σ,0)]τg(σ,τ,σ,τ,α)ζ1δ(τ,α)Wα,l(σ,τ),Wα,l(σ,τ)=[Vα,l(σ,τ)]z1, 3.2

so with its stationary values at l → ∞ Vα,∞(σ, τ) ≡ Vα(σ, τ) we get the potential for α = 0, 1 using its symmetry under permutations of σ ≠ 0 and τ ≠ α

Fq~(ζ,h,α)=2z2ln{Vα(0,α)Wα(0,0)+(q1)ehVα(1,α)Wα(1,α)+(q~1)ζ[Vα(0,1α)Wα(0,1α)+(q1)ehVα(1,1α)Wα(1,1α)]} 3.3

and

Wα(σ,τ)[Vα(σ,τ)]z1.

Summing over spins in (2.8), we get the stationary values’ equations for α = 0, 1

Vα(σ,τ)=Wα(0,α)+(q~1)ζWα(0,1α)+(q1)eh[Wα(1,α)+(q~1)ζWα(1,1α)]+(eK1)eh[1δ(σ,0)][Wα(σ,α)+(q~1)ζWα(σ,1α)]+eKδ(σ,α)eh[1δ(α,0)]{ζ1δ(τ,α)Wα(α,τ)[Wα(α,α)+(q~1)ζWα(α,1α)]}. 3.4

Apparently, Vα(σ, α) − Vα(σ, 1 − α) = eKδ(σ, α)eh[1−δ(σ,0)][Wα(α, α) − ζWα(α, 1 − α)]

(a). Case α = 0

Introducing variables

v=W0(0,0)W0(1,0)andu=W0(0,1)W0(1,0)

we can express potential Fq~(ζ,h,0) via just these two variables, see Appendix A.

u and v obey the equations (A 1) that follow from (3.4). The solutions to them at small εq~1 can be represented as

ueh(y+εζy1),veh(x+εζx1). 3.5

From (A 1), the equations for x and y follow

x=eh[eKx+q1x+eK+q2]z1 3.6
y=eh[q1+eKζyx+eK+q2]z1, 3.7

while x1 is expressed via x and y

x1=(z1)xx(eK+q1)(1eK)y(eKx+q1)(eKx+q1)(x+eK+q2)+(z1)(eK+q1)(1eK)x. 3.8

Substituting (3.6)–(3.8) into (A 2), we get in the first order in εq~1

Fq~(h,ζ,α=0)F1(h,α=0)+εG0(w)+εζ(1c0),

where

G0(w)=c0(q1)w+((2z)/2)eKw2x(eKx+q1),w=ζy, 3.9
F1(α=0,x)=h+(z1)ln[x+eK+q2]+2z2ln[eKx2+2(q1)x+(q1)(eK+q2)],w=ζy. 3.10
c0(x)=1mP(x)=x(eKx+q1)eKx2+2(q1)x+(q1)(eK+q2) 3.11
andmP(x)=F1(x,α=0)h=F1(x,α=0)xxh=(q1)(x+eK+q2)eKx2+2(q1)x+(q1)(eK+q2), 3.12

w = w(ζ) obeys the equation

w=ehζ[q1+eKwx+eK+q2]z1, 3.13

so

wζ=wζq1+eKwq1+(2z)eKw, 3.14

and we have

n0sites=ζG0[w(ζ)]|ζ=1=wG0[w]wζ|ζ=1=c0w1(eKw1+q1)x(eKx+q1).ncl,0=G0(w1), 3.15

w1 obeys the equation

w1=eh[q1+eKw1x+eK+q2]z1. 3.16

For the capacity of giant 0-cluster, we have

P0=c0n0sites=c0c0w1(eKw1+q1)x(eKx+q1)=c0(xw1)[q1+eK(x+w1)]x(eKx+q1). 3.17

Thus we have obtained the thermodynamic parameters ((3.10)–(3.12)) and the percolation ones (3.10), (3.15), (3.17) as functions of x (3.6), w (3.13) and w1 (3.16). Finding stable solutions of (3.6), (3.13), (3,16) and substituting them into the expressions for the thermodynamic and the percolation parameters we can obtain their values at all T/J = K−1 and H/J = hK−1. Note also that above equations present the parametric representations of the thermodynamic and the percolation parameters as functions of T/J and H/J which greatly simplifies their graphing.

Most simply 0-clusters' size distribution νs(0) can be obtained from (3.9), (3.13) using the change of integration variable and the integration by parts

νs(0)=|ζ|=cdζ2πiG0[w(ζ)]ζs+1=|w|=ρdw2πisζs(w)dG0(w)dw, 3.18

and the relation

|w|=ρdwwm(w+R)n=wm1(m1)!(w+R)n|w=0=Rm1Rn(m1)!=(nm1)Rnm+1.

So we get

νs(0)=c0zstps1(1p)t(s(z1)s1), 3.19

where

p=eKxeKx+q1,t=s(z2)+2. 3.20

Here, t is the number of empty bonds in the perimeter of s-site 0-cluster and b = s − 1 is the number of bonds inside it. Thus νs(0) in (3.19) coincides up to pre-factor c0 with the sizes’ distribution of clusters that appear in the process of the independent random placement the bonds with probability p [23], i.e. in the classical bond percolation.

(b). Case α ≠ 0

Similarly to the previous case, we introduce the variables

u~=W1(1,0)W1(0,0)andv~=W1(1,1)W1(0,0),

which obey the equations

u~={[1+(q~1)ζ]+(q2)eh[v~+(q~1)ζu~]+eKehζu~eK[1+(q~1)ζ]+(q1)eh[v~+(q~1)ζu~]}z1v~={[1+(q~1)ζ]+(q2)eh[v~+(q~1)ζu~]+eKehv~eK[1+(q~1)ζ]+(q1)eh[v~+(q~1)ζu~]}z1.

Again, the thermodynamic potential Fq~(h,ζ,α=1) can be expressed as function of these two variables.

At small εq~1

u~eh(y~+εζy~1),v~eh(x~+εζx~1),

where x~ and y~ obey the equations

x~=eh{1+(q2+eK)x~eK+(q1)x~}z1andy~=eh{1+(q2)x+eKζy~eK+(q1)x~}z1 3.21

Using similar procedure, we obtain

Fq~(h,ζ,α=1)F1(x~,α=1)+εG1(ζ)+εζ(1c1),
F1(α=1,x~)=(z1)ln[eK+(q1)x~]+2z2ln[eK+2(q1)x~+(eK+q2)(q1)x~2], 3.22
G1(ζ)=ζ(c11+c0)+G~1[w~(ζ)]=ζ(2q)c1+G~1[w~(ζ)],
G~1(w~)=mP(x~)w~1+(q2)x~+2z2eKw~x~[1+(q2+eK)x~],
mP(x~)=F1h=F1(x~,α=1)x~x~h=(q1)x~[1+(q2+eK)x~](q1)(eK+q2)x~2+2(q1)x~+eK, 3.23
c1(x~)=mP(x~)q1=x~[1+(q2+eK)x~](q1)(eK+q2)x~2+2(q1)x~+eK,
w~=ζy~,w~=ζeh(1+(q2)x~+eKw~eK+(q1)x~)z1w~ζ=w~ζ1+(q2)x~+eKw~1+(q2)x~+(2z)eKw~,
w~1=eh(1+(q2)x~+eKw~1eK+(q1)x~)z1, 3.24
ncl,1=G1(ζ=1)=(2q)c1+G~1(w~1),
n1sites=ζG1(ζ)|ζ=1=(2q)c1+w~G~1(w~)|w~=w~1w~ζ|ζ=1, 3.25
n1sites=(2q)c1+mP(x~)w~1x~1+(q2)x~+eKw~11+(q2+eK)x~ 3.26
P1=c1n1sites=(q1)c1mP(x~)w~1x~1+(q2)x~+eKw~11+(q2+eK)x~=mPx~w~1x~1+(q2)x~+eK(w~1+x~)1+(q2+eK)x~. 3.27

As before, we need to solve the equations for x~ (3.21) and w~1 (3.24) for stable solutions and substitute them to get the final physical results from (3.25) to (3.26).

For 1-clusters′ size distribution, we get similarly

νs>1(1)=mPzst(s(z1)s1)p~s1(1p~)t,ν1(1)=c1[2q+(q1)(1p~)z] 3.28

and

p~=eKx~1+(q2+eK)x~,t=s(z2)+2. 3.29

The same as νs(0) in (3.19), νs(1) coincides up to pre-factor mP = 1 − c0 with the sizes' distribution of clusters in the classical bond percolation with probability p~ (3.29).

4. Thermodynamic and percolation phase diagram

Let us first consider the relation between the expressions for Potts thermodynamic variables that are obtained in §§3a and 3b. Here, we have two distinct sets of formulae for the same parameters – magnetization mP, cα and F1(α). Actually, they are the same as it should be. To see this, we note that equation x1(h)=x~(h) follows from (3.6) and (3.21).

Then inspecting (3.12) and (3.23), we find that they transform one to another under the change x(h) → x−1(h)

mP[α=0,x(h)]=mP[α=1,x1(h)],

so they have equal values as x1(h)=x~(h). The same goes for cα. For the thermodynamic potential, we have from (3.10) and (3.22)

F1(α=0,x)=h+(z1)ln(x+eK+q2)+2z2ln[eKx2+2(q1)x+(q1)(eK+q2)]=(z1)ln(eKx+q1)lnx+2z2ln[eKx2+2(q1)x+(q1)(eK+q2)]=(z1)ln[eK+(q1)x1]+2z2ln[eK+2(q1)x1+(q1)(eK+q2)x2]=F1(α=1,x1).

Here, we used the equation of state (3.6) according to which

h(x)=(z1)lneKx+q1eK+x+q2lnx. 4.1

Hence

F1(α=0,x)=F1(α=1,x~)=βFPotts,

so we can use both equivalent representations to describe thermodynamics of the Potts model.

Furthermore, we choose the α = 0 representation to consider the influence of thermodynamics on the percolation of the geometric α − clusters.

Stable solutions to the equation of state (3.6) must obey the condition

ehddx[eKx+q1eK+x+q2]z1=(z1)x(eK1)(eK+q1)(eKx+q1)(eK+x+q2)<1. 4.2

This is quadratic inequality with respect to x. It fulfils at all x if

R(z,q,T)=(eJ/T1)(eJ/T+q1)[(z2)2(eJ/T1)(eJ/T+q1)4(z1)(q1)]<0. 4.3

When R(z, q, T) > 0, that is at T < Tt(z, q)

Tt(z,q)=J/lnq2z24(z1)(q2)2(z2)(q2)2(z2), 4.4

(4.2) holds for

x(z,q,T)<x<x+(z,q,T) 4.5

and

xs(z,q,T)=12eJ/T[(z2)(eJ/T1)(eJ/T+q1)2(q1)+sR(z,q,T)]. 4.6

There are two stable solutions to the equation of state (3.6) at T < Tt(z, q) that obey the condition (4.5). On the H-T plane the region where they coexist lies at

H(z,q,T)<H<H+(z,q,T), 4.7

where the expressions for Hs(z, q, T) follow from (4.1)

Hs(z,q,T)/T=h[xs(z,q,T)]=(z1)ln[eJ/Txs(z,q,T)+q1eJ/T+xs(z,q,T)+q2]lnxs(z,q,T) 4.8

It follows from (4.8) that Hs(z, q, T = 0) = s(z − 2).

At the field

Htr(z,q,T)=T(z1)ln(eKxtr+q1xtr+eK+q2)Tlnxtr=T2(z2)ln(q1)T2zln[1+(q2)eJ/T] 4.9

the first-order transition takes place between the equilibrium phase with mP<(1/2) (having the lowest potential at H < Htr(z, q, T)) into another equilibrium at H > Htr(z, q, T) phase with mP>1/2.

Equation (4.9) follows from the Maxwell rule [24], see Appendix B.

The dashed line in figure 1 divides the H-T plane into two regions in which mP<1/2 and mP>1/2 in equilibrium states. Note that at all T this line is described by (4.9) but only atT < Tt(z, q) it designates the first-order transition.

Figure 1.

Figure 1.

(a) H-T phase diagram for Potts model on Bethe lattice with q=3, z=5, Tt=1.546 J, Ht=−0.02 J, Ttr=1.51 J, Tc=1.443 J. Dotted lines denote coexistence region (4.7), dashed line—first-order transition, above upper full (red) line giant 1-cluster exist, under lower (blue) line giant 0-cluster exists. (b) the vicinity of tricritical point expanded. (Online version in colour.)

At T = Tt(z, q)

H(z,q,T)=H+(z,q,T)=Htr(z,q,T),

so all three lines meet here. Their meeting point is called the tricritical point, its coordinates on H-T plane are

{Tt(z,q),Ht(z,q)=Ht[z,q,Tt(z,q)]}. 4.10

Beside the equilibrium phases in the region (4.7), there are two metastable ones - with mP>1/2 at H < Htr(z, q, T) and with mP<1/2 at H > Htr(z, q, T). At the boundaries of the coexistence region (4.7) susceptibility χ=mP/h diverges in these states.

The transition line (4.9) always crosses the line H = 0 and there always exists the first-order transition at H = 0 at which mP drops down at

Ttr(z,q)=J/lnq2(q1)z2/z1. 4.11

Note also that the lower boundary of the coexistence region (4.7) always touch the line H = 0 at T = Tc(z, q),

H[z,q,Tc(z,q)]=0,Tc(z,q)=J/lnq+z2z2. 4.12

This coexistence region (4.7) is shown in figure 1. Figure 2 shows the field dependence of magnetization mP and susceptibility χ=mP/h . Figure 3 shows the temperature dependency of mP at H = 0.

Figure 2.

Figure 2.

Field dependence of mP = ∂FPotts/∂h at T=J and T=2 J. Dashed lines correspond to metastable states. (Online version in colour.)

Figure 3.

Figure 3.

Temperature dependencies of mP = ∂FPotts/∂h (full line) and χ = ∂mP/∂h (dashed line) at H = 0. (Online version in colour.)

We should note that above formulae ((4.4)–(4.12)) coincide with that of [25] for ferromagnetic Potts model on Bethe lattice after due account of the relation of our parameters to that of [25]. In this paper, authors add to the Potts Hamiltonian the term Hiδ(σi,1) so our h=H/T=H and the magnetization in this paper M=(βFPotts/H)=δ(σi,1) is equal to our c0. Note also that mathematical aspects of Potts model thermodynamics on various graphs and lattices are discussed in [26,27].

The stable solutions to the percolation equation of state for the 0-clusters (3.16) must obey the condition eh(d/dw1)[q1+eKw1/x+eK+q2]z1=w1eK(z1/q1+eKw1)<1 or w1<(q1/eK(z2))xp0. On the other hand, the infinite 0-cluster does not exist in phases where w1 > x, see (3.17). Thus, non-percolating phase exists when x < w1 < xp0 and percolation transition takes place at x = xp0 or at

Hp0=J(z2)+T{(z1)ln[(q1)(z1)(z2)[1+eJ/T(q2)]+e2J/T(q1)]lnq1z2}, 4.13

In the stable thermodynamic phases (equilibrium and metastable) (x/h)<0 so the non-percolating condition x < xp0 means that the infinite 0-cluster is absent at Hp0 < H while it is present at H < Hp0.

Meanwhile the stable solutions to the percolation equation of state for the 1-clusters (3.24) must obey the condition eh(/w~1)(1+(q2)x~+eKw~1/eK+(q1)x~)z1<1 or w~1<(1+(q2)x~/eK(z2)), while infinite 1-cluster is absent when x<w~1, see (3.27). Hence, the non-percolating phase exists when x~<w~1<(1+(q2)x~/eK(z2)). This inequality needs the following condition to be fulfilled

x~<1eK(z2)+2qxp1,

so the percolation transition takes place at x~=xp1 or at

Hp1=(z2)J+T(z1)lnz2+(2q)eJ/T+(q1)e2J/Tz1Tln[z2+(2q)eJ/T]. 4.14

In the stable thermodynamic phases x~/h=x2(x/h)>0 so the non-percolating condition x~<xp1 means that the infinite 1-cluster is absent at H < Hp1 while it is present at Hp1 < H.

Note, that when z < q (4.14) is valid only at T<J/ln(q2/z2)T1, above this temperature the infinite 1-cluster does not exist (figure 4).

Figure 4.

Figure 4.

Phase diagram at q < z. q = 6, z = 5. The giant 1-cluster is absent above T1 = 3.476 shown by dashed vertical line. (Online version in colour.)

Here, we should remind that in the region denoted as ‘Giant 1-cluster’ the other giant α − clusters exist with α > 1 due to the permutation symmetry of partition function (2.3).

5. Discussion and conclusion

We show that the calculation of specific double-Potts partition function is the useful method to study the percolation of geometric Potts clusters. With it, the percolation phase diagram and the size distribution of α − clusters in ferromagnetic Potts model on the Bethe lattice are found analytically. The size distribution in such correlated percolation appears to be proportional to that of the classical non-correlated bond percolation with the bond occupation probability depending on Potts model parameters (K, h) same as for the Ising clusters on this lattice [12,23]. Accordingly, the α − clusters percolation has the same classical critical indexes.

It seems that this result is not solely the property of the Bethe lattice model as the last is the good mean-field approximation for graphs and lattices with z ≫ 1outside the critical region. Probably, in the mean-field percolation region of a wide class of the Potts and Ising ferromagnetic models the correlations amount to the formation of independent pairs of nearest neighbour like-valued spins.

Beside providing the bond occupation probability, the influence of the model on the percolation of its geometric clusters is rather scarce. We may note that at T = 0 giant clusters exist strictly outside coexistence region. Meanwhile, at finite T both Hp0 and Hp1 cross the line of the first-order phase transition, cf. Figure 1b, which means that the thermodynamic transition is not a consequence of the percolation of geometric clusters.

The present approach can be useful for the numerical studies of Potts clusters' percolation on Euclidean lattices [11,12]. Thus, for rough estimate of α − clusters size generation function the usual Monte Carlo simulations can be used to obtain Zq~(ζ) for several integer q~ and to interpolate it to q~=1. To get more precise results one should extend the expression (2,3) for Zq~(ζ) to real q~. This can be done, for example, within the transfer matrix representation of Zq~(ζ), see [28]. Note also, that present method can be easily modified for other types of Potts clusters.

Acknowledgement

The author gratefully acknowledges the fruitful discussions and support of G.Y. Chitov, N. Ter-Oganessyan and V.P. Sakhnenko.

Appendix A

Using the relations that follow from (3.4)

V0(0,0)=W0(1,0){(q1)eh[1+(q~1)ζ]+eKv},V0(0,1)=W0(1,0){(q1)eh[1+(q~1)ζ]+eKζu},V0(1,0)=W0(1,0){v+(q~1)ζu+eh(eK+q2)[1+(q~1)ζ]},V0(1,1)=V0(1,0),W1,02z={v+(q~1)ζu+eh(eK+q2)[1+(q~1)ζ]}z1,

We get the following equations for u and v

v={(q1)eh[1+(q~1)ζ]+eKvv+(q~1)ζu+eh(eK+q2)[1+(q~1)ζ]}z1andu={(q1)eh[1+(q~1)ζ]+eKζuv+(q~1)ζu+eh(eK+q2)[1+(q~1)ζ]}z1. A 1

Then, we can express potential for α = 0 via just two variables, u and v as follows:

Fq~(h,ζ,α=0)=(z1)ln{v+(q~1)ζu+eh(eK+q2)[1+(q~1)ζ]}+2z2ln{v[(q1)eh[1+(q~1)ζ]+eKv]+(q1)eh[v+(q~1)ζu+eh(eK+q2)[1+(q~1)ζ]]+(q~1)ζ[u[(q1)eh[1+(q~1)ζ]+eKζu]+(q1)eh[v+(q~1)ζu+eh(eK+q2)[1+(q~1)ζ]]]} A 2

Appendix B

The equation (4.9) follows from the Maxwell rule [13]

mP(x1)mP(x2)dmh[x(m)]htr[mP(x2)mP(x1)]=0. B 1

Here, x(m) is the inverse function to mP(x) so h[x(m)] = h(x) and htr is the transition field such that

htr=h(x2)=h(x1). B 2

Introducing the function

Φ(m)=1/2mdm~[h(m~)htr], B 3

we can represent (B 1) as

Φ[mP(x2)]=Φ[mP(x1)].

Integrating (B 3) by parts and using (B 2), we have

Φ[mP(xn)]=mP(xn)[h(xn)htr]12[h(m=12)htr]h(m=1/2)h(xn)dm~h(m~)m~=12[h(m=12)htr]h(m=1/2)htrdm~h(m~)m~.

Thus Φ[mP(x2)] = Φ[mP(x1)] = 0 when htr=h(m=1/2). Hence, the first-order transition takes place at the field

htr=h(xtr)=(z1)lneKxtr+q1xtr+eK+q2lnxtr,

where xtr is defined through the equation mP(xtr)=1/2. We easily find that

xtr=(q1)[1+(q2)eK],

which gives Htr(z, q, T) = h(xtr) in (4.9).

Data accessibility

This article has no additional data.

Competing interests

I declare I have no competing interests.

Funding

Research was financially supported by the Ministry of Science and Higher Education of the Russian Federation (State assignment in the field of scientific activity, Southern Federal University, 2020).

References

  • 1.Wu FY. 1982. The Potts model. Rev. Mod. Phys. 54, 235–268. ( 10.1103/RevModPhys.54.235) [DOI] [Google Scholar]
  • 2.Baxter RJ. 1982. Exactly solved model in statistical mechanics. London, UK: Academic Press. [Google Scholar]
  • 3.Tsallis C, de Magalhaes ACN. 1996. Pure and random Potts-like models: real-space renormalization-group approach. Phys. Rep. 268, 305–430. ( 10.1016/0370-1573(95)00064-X) [DOI] [Google Scholar]
  • 4.Li H-J, Wang Y, Wu L-Y, Zhang J, Zhang X-S. 2012. Potts model based on a Markov process computation solves the community structure problem effectively. Phys. Rev. E 86, 016109 ( 10.1103/PhysRevE.86.016109) [DOI] [PubMed] [Google Scholar]
  • 5.Coniglio A, Fierro A. 2009. In Encyclopedia of complexity and systems science, part 3, 1596, New York, NY: Berlin, Germany: Springer. [Google Scholar]
  • 6.Nappi CR, Peruggi F, Russo L. 1977. Percolation points and critical point in the Ising model. J. Phys. A: Math. Gen. 10, 205–218. ( 10.1088/0305-4470/10/2/010) [DOI] [Google Scholar]
  • 7.Coniglio A. 1976. Some cluster-size and percolation problems for interacting spins. Phys. Rev. B 13, 2194–2207. ( 10.1103/PhysRevB.13.2194) [DOI] [Google Scholar]
  • 8.Murata KK. 1979. Hamiltonian formulation of site percolation in a lattice gas. J. Phys. A: Math. Gen. 12, 81–89. ( 10.1088/0305-4470/12/1/020) [DOI] [Google Scholar]
  • 9.Coniglio A, Klein W. 1980. Clusters and Ising critical droplets: a renormalisation group approach. J. Phys. A: Math. Gen. 13, 2775–2780. ( 10.1088/0305-4470/13/8/025) [DOI] [Google Scholar]
  • 10.Fortuin CM, Kasteleyn PW. 1972. On the random-cluster model. Physica 57, 536–564. ( 10.1016/0031-8914(72)90045-6) [DOI] [Google Scholar]
  • 11.Blanchard Ph, Gandolfo D, Laanait L, Ruiz J, Satz H. 2008. On the Kertész line: thermodynamic versus geometric criticality. J. Phys. A: Math. Gen. 41, 085001 ( 10.1088/1751-8113/41/8/085001) [DOI] [Google Scholar]
  • 12.Ruiz J, Wouts M. 2008. On the Kertész line: some rigorous bounds. J. Math. Phys. 49, 053303 ( 10.1063/1.2924322) [DOI] [Google Scholar]
  • 13.Blanchard Ph, Gandolfo D, Ruiz J, Wouts M. 2008. Thermodynamic vs. topological phase transitions: Cusp in the Kertész line. Europhys. Lett. 82, 50003 ( 10.1209/0295-5075/82/50003) [DOI] [Google Scholar]
  • 14.Saberi AA. 2009. Thermal behavior of spin clusters and interfaces in the two-dimensional Ising model on a square lattice. J. Stat. Mech. 2009, P07030 ( 10.1088/1742-5468/2009/07/p07030) [DOI] [Google Scholar]
  • 15.Saberi AA, Dashti-Naserabadi H. 2010. Three-dimensional Ising model, percolation theory and conformal invariance. Europhys. Lett. 92, 67005 ( 10.1209/0295-5075/92/67005) [DOI] [Google Scholar]
  • 16.Noh JD, Lee HK, Park H. 2011. Scaling of cluster heterogeneity in percolation transitions. Phys. Rev. E 84, 010101 ( 10.1103/PhysRevE.84.010101) [DOI] [PubMed] [Google Scholar]
  • 17.Jo WS, Yi SD, Baek SK, Kim BJ. 2012. Cluster-size heterogeneity in the two-dimensional Ising model. Phys. Rev. E 86, 032103 ( 10.1103/physreve.86.032103) [DOI] [PubMed] [Google Scholar]
  • 18.de la Rocha AR, de Oliveira PMC, Arenzon JJ. 2015. Domain-size heterogeneity in the Ising model: geometrical and thermal transitions. Phys. Rev. E 91, 042113 ( 10.1103/PhysRevE.91.042113) [DOI] [PubMed] [Google Scholar]
  • 19.Stanley HE. 1979. A polychromatic correlated-site percolation problem with possible relevance to the unusual behaviour of supercooled H2O and D2O. J. Phys. A: Math. Gen. 12, L211–L337. ( 10.1088/0305-4470/12/12/003) [DOI] [Google Scholar]
  • 20.Kogut PM, Leath PL. 1982. High-density site percolation on real lattices. J. Phys. C: Solid State Phys. 15, 4225–4233. ( 10.1088/0022-3719/15/20/008) [DOI] [Google Scholar]
  • 21.Branco NS, de Queiroz SLA, dos Santos RR. 1986. Critical exponents for high density and bootstrap percolation. J. Phys. C: Solid State Phys. 19, 1909–1921. ( 10.1088/0022-3719/19/12/006) [DOI] [Google Scholar]
  • 22.Stephen MJ. 1977. Site-cluster distributions and equation of state for the bond percolation model. Phys. Rev. B 15, 5674–5680. ( 10.1103/PhysRevB.15.5674) [DOI] [Google Scholar]
  • 23.Timonin PN. 2019. Statistics of geometric clusters in the Ising model on a Bethe lattice. Physica A 527, 121402 ( 10.1016/j.physa.2019.121402) [DOI] [Google Scholar]
  • 24.Lebowitz JL, Penrose O. 1966. Rigorous treatment of the Van Der Waals-Maxwell Theory of the liquid-vapor transition. J. Math. Phys. 7, 98–113. ( 10.1063/1.1704821) [DOI] [Google Scholar]
  • 25.Peruggi F, di Liberto F, Monroy G. 1983. The Potts model on Bethe lattices. I. General results. J. Phys. A: Math. Gen. 16, 811–827. ( 10.1088/0305-4470/16/4/018) [DOI] [Google Scholar]
  • 26.Rozikov U. 2013. Gibbs measures on Cayley trees. Singapore: World Scientific. [Google Scholar]
  • 27.Külske C, Rozikov UA, Khakimov RM. 2014. Description of the translation-invariant splitting Gibbs measures for the Potts Model on a Cayley Tree. J. Stat. Phys. 156, 189–200. ( 10.1007/s10955-014-0986-y) [DOI] [Google Scholar]
  • 28.Jacobsen JL, Cardy J. 1998. Critical behaviour of random-bond Potts models: a transfer matrix study. Nuclear Phys. B 515, 701–742. ( 10.1016/S0550-3213(98)00024-8) [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

This article has no additional data.


Articles from Proceedings. Mathematical, Physical, and Engineering Sciences are provided here courtesy of The Royal Society

RESOURCES