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

Sweeny dynamics for the random-cluster model with small QQ

Zirui Peng Department of Modern Physics, University of Science and Technology of China, Hefei, Anhui 230026, China    Eren Metin Elçi School of Mathematical Sciences, Monash University, Clayton, VIC 3800, Australia    Youjin Deng yjdeng@ustc.edu.cn Department of Modern Physics, University of Science and Technology of China, Hefei, Anhui 230026, China Hefei National Laboratory, University of Science and Technology of China, Hefei 230088, China MinJiang Collaborative Center for Theoretical Physics, College of Physics and Electronic Information Engineering, Minjiang University, Fuzhou 350108, China    Hao Hu huhao@ahu.edu.cn School of Physics and Optoelectronic Engineering, Anhui University, Hefei, Anhui 230601, China
Abstract

The Sweeny algorithm for the QQ-state random-cluster model in two dimensions is shown to exhibit a rich mixture of critical dynamical scaling behaviors. As QQ decreases, the so-called critical speeding-up for non-local quantities becomes more and more pronounced. However, for some quantity of specific local pattern—e.g., the number of half faces on the square lattice, we observe that, as Q0Q\to 0, the integrated autocorrelation time τ\tau diverges as QζQ^{-\zeta}, with ζ1/2\zeta\simeq 1/2, leading to the non-ergodicity of the Sweeny method for Q0Q\to 0. Such QQ-dependent critical slowing-down, attributed to the peculiar form of the critical bond weight v=Qv=\sqrt{Q}, can be eliminated by a combination of the Sweeny and the Kawasaki algorithm. Moreover, by classifying the occupied bonds into bridge bonds and backbone bonds, and the empty bonds into internal-perimeter bonds and external-perimeter bonds, one can formulate an improved version of the Sweeny-Kawasaki method such that the autocorrelation time for any quantity is of order O(1)O(1).

I Introduction

The Hamiltonian of the QQ-state Potts model [1] reads =Jijδσi,σj\mathcal{H}=-J\sum_{\langle ij\rangle}\delta_{\sigma_{i},\sigma_{j}}, where J>0J>0 is the ferromagnetic coupling constant, ij\langle ij\rangle stands for a pair of nearest-neighboring sites, and each lattice site ii has a spin σi\sigma_{i}, taking one of the QQ states σi=1,2,,Q\sigma_{i}=1,2,...,Q. For Q=2Q=2, the Potts model corresponds to the famous Ising model. Given a spin configuration, under the Fortuin-Kasteleyn transformation [2, 3], bonds between nearest-neighboring sites with the same spin value are independently occupied with the probability p=1eJ/kTp=1-e^{-J/kT} (kk is the Boltzmann constant, TT is the temperature, and, for convenience, we set kT=1kT=1), and the resulting graph is a configuration of the random-cluster model [4]. For bond configurations AEA\subseteq{E} of a given graph G=(V,E)G=(V,E), the random-cluster model is defined by the partition function

Z=AEQk(A)v|A|.\displaystyle Z=\sum_{A\subseteq{E}}Q^{k(A)}v^{|A|}\;. (1)

Here AA is the set of “occupied bonds”, |A||A| is the number of occupied bonds, k(A)k(A) is the number of clusters (connected components), real positive numbers QQ and v=p/(1p)v=p/(1-p) are the cluster and bond weights, respectively. For integer Q2Q\geq 2, the random-cluster model serves as a graphical representation of the QQ-state Potts model. As Q1Q\to 1, the random-cluster model reduces to independent bond percolation [5].

In equilibrium, the weight ratio between two configurations are W(β)/W(α)=Qδk(A)vδ|A|W(\beta)/W(\alpha)=Q^{\delta k(A)}v^{\delta|A|}, where δk(A)\delta k(A) (δ|A|\delta|A|) represents the changes of k(A)k(A) (|A||A|) from configuration α\alpha to β\beta. As Q0Q\rightarrow 0, when v/Qv/Q is finite, a configuration consists of a spanning forest [6, 4]. This is because in this case W(β)/W(α)Qδ[k(A)+|A|]W(\beta)/W(\alpha)\sim Q^{\delta[k(A)+|A|]}, and spanning forests have the minimum k(A)+|A|k(A)+|A|. From a spanning forest α\alpha, occupying any internal-perimeter bond (defined later in Fig. 2) leads to a nonforest β\beta with δ[k(A)+|A|]=1\delta[k(A)+|A|]=1, such that W(β)/W(α)Q0W(\beta)/W(\alpha)\sim Q\rightarrow 0. In the case Q0Q\rightarrow 0 with v/Qσ(0<σ<1)v/Q^{\sigma}(0<\sigma<1) being finite, similarly the model consists of uniform spanning trees which have the smallest k(A)+σ|A|k(A)+\sigma|A| [6, 4].

Markov Chain Monte Carlo (MCMC) algorithms [7, 8] are effective in simulating the random-cluster model, such as the Swendsen-Wang-Chayes-Machta algorithm [9, 10] for Q>1Q>1 with non-local cluster updates, and the Sweeny algorithm [11] for Q>0Q>0 with local bond updates. To assess the efficiency of the MCMC algorithms, autocorrelation times are commonly used as a metric [12]. For any observable 𝒪{\cal O}, one first defines the autocorrelation function

ρ𝒪(t)=𝒪s𝒪s+t𝒪2var(𝒪).\displaystyle\rho_{\cal O}(t)=\frac{\langle{{\cal O}_{s}{\cal O}_{s+t}}\rangle-{\langle{\cal O}\rangle}^{2}}{{\rm var}({\cal O})}\;. (2)

Here expectations are taken after equilibrium is attained, and the time unit is one “hit”, in which only one bond is visited (a Monte Carlo sweep consists of |E||E| hits, where |E||E| is the number of all bonds). Then one defines the exponential autocorrelation time for an observable 𝒪\cal O

τexp,𝒪=lim supt±|t|log|ρ𝒪(t)|,\displaystyle\tau_{{\rm exp},\cal O}=\limsup_{t\rightarrow\pm\infty}\frac{|t|}{{\rm-log}|\rho_{\cal O}(t)|}\;, (3)

and the overall exponential autocorrelation time

τexp=sup𝒪τexp,𝒪.\displaystyle\tau_{{\rm exp}}=\sup_{\cal O}\tau_{{\rm exp},\cal O}\;. (4)

The integrated autocorrelation time for an observable 𝒪\cal O is defined as

τint,𝒪=12t=ρ𝒪(t).\displaystyle\tau_{{\rm int},\cal O}=\frac{1}{2}\sum_{t=-\infty}^{\infty}\rho_{\cal O}(t)\;. (5)

Observables typically have τexp,𝒪=τexp\tau_{{\rm exp},{\cal O}}=\tau_{\rm exp} and τint,𝒪τexp\tau_{{\rm int},\cal O}\leq\tau_{\rm exp}. The autocorrelation times τexp\tau_{\rm exp} and τint,𝒪\tau_{{\rm int},\cal O} play different roles in MCMC simulations [12]: τexp\tau_{{\rm exp}} is the relaxation time of the slowest mode in the system, and τint,𝒪\tau_{{\rm int},\cal O} controls the statistical error in the measurement of 𝒪\langle\cal O\rangle. Thus τint,𝒪\tau_{{\rm int},\cal O} mainly determines the efficiency of MCMC sampling. For a time series of total length MM, the sample mean

𝒪^=1Mt=1M𝒪t\displaystyle\widehat{{\cal O}}=\frac{1}{M}\sum_{t=1}^{M}{\cal O}_{t} (6)

has variance

var(𝒪^)2τint,𝒪var(𝒪)M,whenM.\displaystyle{\rm var}(\widehat{{\cal O}})\approx 2\tau_{{\rm int},{\cal O}}\frac{{\rm var}({\cal O})}{M},{\;\rm when\;}M\to\infty\;. (7)

The above Eq. (7) can be used to measure τint,𝒪\tau_{{\rm int},{\cal O}}.

In MCMC simulations, the dynamic processes usually undergo critical slowing-down [13]: the autocorrelation time diverges as τ/|E|ξz\tau/{|E|}\sim\xi^{z} when the critical point is being approached, where ξ\xi is the spatial correlation length and zz is called the dynamic exponent. For two-dimensional Ising model, Metropolis-Hastings [14, 15] and Glauber [16, 8] algorithms have z2z\approx 2.  To overcome the criticalcritical slowingslowing-downdown, the Swendsen-Wang-Chayes-Machta algorithm [9, 10] was proposed for the random-cluster model with Q>1Q>1. The latter algorithm has z0.2z\approx 0.2 for the two-dimensional Ising model, which represents a huge improvement. The gained efficiency is associated with nonlocal cluster updates.

Refer to caption
Figure 1: Autocorrelation function ρ𝒮2(t)\rho_{\mathcal{S}_{2}}(t) and ρ𝒩(t)\rho_{\mathcal{N}}(t) for the critical 2D random-cluster model at Q=0.005Q=0.005 and Q=0.5Q=0.5 on the square lattice. The unit of tt is one hit.

However, some local algorithms may still be very efficient, such as the worm algorithm [17] and the Sweeny algorithm [18]. The worm algorithm outperforms the cluster algorithms for the three-dimensional Ising model in measuring quantities such as the susceptibility and correlation length. With local single-bond updates, the Sweeny algorithm can surprisingly exhibit critical speeding-up for a suitable range of QQ [18]: the relaxation time of some observables decreases significantly at time scales much shorter than one sweep as the lattice size LL increases, and this phenomenon becomes even more pronounced as QQ becomes smaller. We illustrate the latter property of the Sweeny algorithm by observing the dynamical behavior of two quantities, i.e., the number of occupied bonds 𝒩=|A|\mathcal{N}=|A| and the second moment of cluster sizes 𝒮2=|𝒞|2\mathcal{S}_{2}=\sum{|\mathcal{C}|^{2}}. Hereafter, if not specified, all simulations were performed on the square lattice at the critical point vc(Q)=Qv_{c}(Q)=\sqrt{Q}. The average 𝒮2/|V|\langle\mathcal{S}_{2}\rangle/|V| is a susceptibility-like quantity of the random-cluster model [2, 3], with |V||V| being the volume. For Q2Q\geq 2, it is found [18] that the dynamic exponent zexpzint,𝒩z_{exp}\approx z_{{\rm int},\mathcal{N}}, and that both zint,𝒩z_{{\rm int},\mathcal{N}} and zint,𝒮2z_{{\rm int},\mathcal{S}_{2}} are only slightly larger than the lower bound α/ν\alpha/\nu [19], and could possibly be equal to it (here α\alpha and ν\nu are critical exponents for the specific heat and correlation length, respectively). For Q<2Q<2, the autocorrelation function of 𝒩\mathcal{N} follows an almost purely exponential function, as demonstrated for Q=0.005Q=0.005 and Q=0.5Q=0.5 in Fig. 1(a) and  1(c), respectively. However, the dynamic behavior of 𝒮2\mathcal{S}_{2} is richer [18]: (1) as shown in Fig. 1(b) and 1(d), the autocorrelation function ρ𝒮2(t)\rho_{\mathcal{S}_{2}}(t) decays very rapidly in much less than one sweep, as a prelude to the final exponential decay et/τexp\sim e^{-t/\tau_{exp}}; (2) the exponent zint,𝒮2z_{{\rm int},\mathcal{S}_{2}} is a negative number, and it decreases with the increasing of lattice size LL; (3) from Fig. 1(d) to 1(b), it is seen that the critical speeding-up is more pronounced for smaller QQ.

Refer to caption
Figure 2: Schematic diagrams for four kinds of bonds in the random-cluster model. They are categorized according to their occupation and connectivity: if bond xyxy is occupied and the two endpoints are connected (resp. not connected) via a path that does not include xyxy, then the occupied bond is called a backbonebackbone bondbond (resp. bridgebridge bondbond); if bond xyxy is empty and the two endpoints are connected (resp. not connected) via a path that does not include xyxy, then the empty bond is referred to as an internalinternal-perimeterperimeter bondbond (resp. externalexternal-perimeterperimeter bondbond).

In this article, we are interested in the dynamic behavior of the Sweeny algorithm for very small QQ values. A QQ-dependent critical slowing-down is found for the random-cluster model in two dimensions. The phenomenon can be observed in the dynamics of a specific local pattern on the square lattice, i.e., a unit square with configuration or , where solids lines represent occupied bonds. We call the local pattern as a half face. We find that, when Q0Q\rightarrow 0, the integrated autocorrelation time of the number of half faces in terms of sweeps (i.e., τint,𝒩HF/2L2\tau_{{\rm int},\mathcal{N}_{\rm HF}}/2L^{2}, and hereafter τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} is used for brevity) diverges as Q1/2Q^{-1/2}. This indicates that the Sweeny algorithm becomes not ergodic as Q0Q\to 0.

The above QQ-dependent critical slowing-down can be explained by the acceptance probability of updates in the Sweeny algorithm. As illustrated in Fig. 2, considering different occupation and connectivity, the bonds can be classified into four types — the backbonebackbone, bridgebridge, internalinternal-perimeterperimeter and externalexternal-perimeterperimeter bonds. At vc=Qv_{c}=\sqrt{Q}, when transforming an internal-perimeter bond into a backbone bond, one has δk(A)=0\delta k(A)=0 and δ|A|=1\delta|A|=1, which yields the weight ratio W(β)/W(α)=QW(\beta)/W(\alpha)=\sqrt{Q}; and transforming a bridge bond into an external-perimeter bond leads to δk(A)=1\delta k(A)=1 and δ|A|=1\delta|A|=-1, which also gives W(β)/W(α)=QW(\beta)/W(\alpha)=\sqrt{Q}. It is shown in Sec. II that these weight ratios are associated with vanishing acceptance probabilities Q~{}\sqrt{Q} as Q0Q\rightarrow 0. To overcome the QQ-dependent slowing-down, we first use the knowledge of four kinds of bonds in the detail balance condition and propose an improved Sweeny algorithm. The improved Sweeny algorithm largely alleviates the slowing-down but does not fully solve the problem. For example, at Q=0Q=0 where the graphs become spanning trees, the improved Sweeny algorithm is still invalid due to zero probability of deleting or adding a bond. We then employ the Kawasaki algorithm [20], which proposes the random exchange of an empty and an occupied bond. At Q=0Q=0 on spanning trees, the Kawasaki algorithm still has an LL-dependent divergence τ𝒩HF/L2L3/2\tau_{\mathcal{N}_{\rm HF}}/L^{2}\sim L^{3/2} which can also be related to the acceptance probability of updates. The latter divergence can be eliminated through an improved Kawasaki algorithm which uses the categories of bonds. Finally, we combine the use of the improved Sweeny and improved Kawasaki algorithms, and find no critical slowing-down for small QQ, i.e., the autocorrelation time τ/L2\tau/L^{2} is of order 𝒪(1){\cal O}(1), being independent of QQ.

The remainder of this paper is organized as follows. Section II describes the Sweeny algorithm and the QQ-dependent slowing-down. Section III presents an improved Sweeny algorithm which can greatly improve simulation efficiency. Section IV introduces an improvement of the Kawasaki algorithm and its combination with the improved Sweeny algorithm, which can completely eliminate the QQ-dependent slowing-down. Section V contains a discussion and conclusion.

II Sweeny algorithm and QQ-dependent slowing-down

1 Choose a bond xyxy uniformly at random
2 if xyxy is an external-perimeter bond then
3      Occupy xyxy with probability min{1,v/Q}\min\{1,v/Q\}
4 else if xyxy is an internal-perimeter bond then
5      Occupy xyxy with probability min{1,v}\min\{1,v\}
6 else if xyxy is a bridge bond then
7      Delete xyxy with probability min{1,Q/v}\min\{1,Q/v\}
8 else if xyxy is a backbone bond then
9      Delete xyxy with probability min{1,1/v}\min\{1,1/v\}
10 end if
Algorithm 1 Original Sweeny Algorithm

As mentioned in the Introduction, the Sweeny algorithm is a local single-bond update algorithm for the random-cluster model. Actually, it is the only known MCMC algorithm for the model with 0<Q<10<Q<1. The original Sweeny algorithm [11] is presented in Alg. 1 . A basic step of the Sweeny algorithm is proposing to change the status of a randomly selected bond, and accepting the proposal according to the Metropolis criterion.

For an efficient implementation of the Sweeny algorithm, the main challenge is in checking the connectivity between the two endpoints of a chosen bond. There are three typical methods with different asymptotic run-time scaling: breadth-first search, union-and-find, and dynamic connectivity algorithms [21, 22]. We use the method of simultaneous breadth-first searches starting at both endpoints of a bond xyxy, since it has very simple code and is effective enough for small and medium size LL [23]. We note that a simple modification of Alg. 1 can save some time by not requiring connectivity checking for a fraction of updates. The modification for Q1Q\leq 1 at vc=Qv_{c}=\sqrt{Q} is shown in Alg. 2. The basic idea is to consider together several probabilities in Alg. 1: if the generated random number is less than the minimum of these probabilities, the update must be accepted, and if it is greater than the maximum of these probabilities, the update must be rejected. For Q1Q\to 1, this small change in the algorithm can lead to a significant improvement of efficiency.

1 PminP_{\rm min}=Q\sqrt{Q}, PmaxP_{\rm max}=1
2 Choose a bond xyxy uniformly at random
3 Generate a random number P[0,1)P\in[0,1)
4 if P<=PminP<=P_{\rm min} then
5      Change the status of xyxy
6 else if xyxy is an external-perimeter bond or a backbone bond then
7      Change the status of xyxy
8 end if
Algorithm 2 Modified Sweeny Algorithm (for Q1Q\leq 1 at vc=Qv_{c}=\sqrt{Q})

As described in the Introduction, the Sweeny algorithm exhibits rich dynamics, including both critical slowing-down and critical speeding-up. However, its dynamics at small QQ values are not uncovered fully. For Alg. 1 and Alg. 2, one can observe that they encounter a problem when QQ is small: at vc=Qv_{c}=\sqrt{Q}, probabilities of some updates (occupying an internal-perimeter bond and deleting a bridge bond) are proportional to Q\sqrt{Q}, which causes the algorithm being not ergodic at Q0Q\to 0. The problem was not severe in previous studies, since the studied QQ values are not extremely small, and conventional observables such as 𝒩\mathcal{N} and 𝒮2\mathcal{S}_{2} could not detect any slowing-down effect due to the Q\sqrt{Q}-form updating probabilities. Actually on the square lattice in the limit Q0Q\to 0, a critical equilibrium configuration consists of a spanning tree, thus the above observables remain fixed as 𝒩=L21\mathcal{N}=L^{2}-1 and 𝒮2=L4\mathcal{S}_{2}=L^{4} for a given LL.

We find that the above problem can become dominant in sampling other quantities, such as the number of half faces 𝒩HF\mathcal{N}_{\rm HF}. Using the original Sweeny algorithm, we calculated τ𝒩HF\tau_{\mathcal{N}_{\rm HF}} as plotted in Fig. 3. It can be seen that τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} diverges as Q1/2Q^{-1/2}, independent of the system size LL. We refer to this as QQ-dependent critical slowing-down. This newly discovered phenomenon at small QQ enriches the dynamics of the Sweeny algorithm. It is important to note that, away from Q=0Q=0, the original Sweeny algorithm is not inherently flawed for small QQ. For example, when Q=106Q=10^{-6}, the autocorrelation time τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} is of order O(102)O(10^{2}), which means that 𝒩HF\mathcal{N}_{\rm HF} can still be measured rather efficiently.

Refer to caption
Figure 3: Autocorrelation time of the number of half faces for the original Sweeny algorithm. τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} diverges as Q1/2Q^{-1/2} when Q0Q\to 0.

To understand the QQ-dependent critical slowing-down, we associate the above factor Q1/2Q^{-1/2} to the updating probabilities in the Sweeny algorithm. In the following we measure the numbers of the aforementioned four kinds of bonds, and then give these updating probabilities.

On the square lattice in the limit Q0Q\to 0, the number of bridge bonds (𝒩BR\mathcal{N}_{\rm BR}) and of internal-perimeter bonds (𝒩IP\mathcal{N}_{\rm IP}) are 𝒩BRL21\mathcal{N}_{\rm BR}\to L^{2}-1 and 𝒩IP|E|𝒩BR\mathcal{N}_{\rm IP}\to|E|-\mathcal{N}_{\rm BR}, respectively. These numbers can be obtained from the characteristics of a spanning tree: since every node (i.e., site) corresponds to an occupied bond except the root, there are L21L^{2}-1 occupied bonds for a tree with L2L^{2} nodes; the L21L^{2}-1 occupied bonds are all bridge bonds, and other (empty) bonds are internal-perimeter bonds. For finite small QQ, changes of the numbers of four kinds of bonds with the decrease of QQ are shown in Fig. 4, with fitting results for two of them in Table 1. For L=16L=16, we find that for small QQ, the average number of backbone bonds (𝒩BB\mathcal{N}_{\rm BB}) and external-perimeter bonds (𝒩EP\mathcal{N}_{\rm EP}) are respectively

𝒩BB0.94L2Q,𝒩EPL2Q,\displaystyle\langle\mathcal{N}_{\rm BB}\rangle\simeq 0.94L^{2}\sqrt{Q}\;,\quad\langle\mathcal{N}_{\rm EP}\rangle\simeq L^{2}\sqrt{Q}\;, (8)

both being proportional to Q\sqrt{Q}, as plotted in Fig. 4(a). The average number of occupied bonds is found to be 𝒩BL21\langle\mathcal{N}_{B}\rangle\simeq L^{2}-1 for small QQ, which is consistent with 𝒩BRL21\mathcal{N}_{\rm BR}\to L^{2}-1 in the limit Q0Q\to 0. In the simulations, for convenience of a loop construction which will be described in Section III, we use periodic boundary conditions in one direction of the square lattice and open boundary conditions in the other direction. The total number of bonds on the lattice is |E|=2L2L|E|=2L^{2}-L. Thus, considering 𝒩B=𝒩BB+𝒩BR\mathcal{N}_{B}=\mathcal{N}_{\rm BB}+\mathcal{N}_{\rm BR} and |E|𝒩B=𝒩EP+𝒩IP|E|-\mathcal{N}_{B}=\mathcal{N}_{\rm EP}+\mathcal{N}_{\rm IP}, we derive

𝒩BRL20.94L2Q1L2(10.94Q),\displaystyle\langle\mathcal{N}_{\rm BR}\rangle\simeq L^{2}-0.94L^{2}\sqrt{Q}-1\simeq L^{2}(1-0.94\sqrt{Q})\;,
𝒩IPL2L2QL+1L2(1Q).\displaystyle\langle\mathcal{N}_{\rm IP}\rangle\simeq L^{2}-L^{2}\sqrt{Q}-L+1\simeq L^{2}(1-\sqrt{Q})\;. (9)

The above scaling behaviors are demonstrated by a plot of 1𝒩BR(IP)/L21-\langle\mathcal{N}_{\rm BR(IP)}\rangle/L^{2} as shown in Fig. 4(b). In the thermodynamic limit LL\rightarrow\infty, from the self-duality of the square lattice, the critical numbers of four kinds of bonds can be obtained [24] (an alternative derivation is given in the Appendix of the present paper) as

𝒩BB/L2=𝒩EP/L2=Q1+Q,\displaystyle\langle\mathcal{N}_{\rm BB}\rangle/L^{2}=\langle\mathcal{N}_{\rm EP}\rangle/L^{2}=\frac{\sqrt{Q}}{1+\sqrt{Q}}\;,
𝒩BR/L2=𝒩IP/L2=11+Q.\displaystyle\langle\mathcal{N}_{\rm BR}\rangle/L^{2}=\langle\mathcal{N}_{\rm IP}\rangle/L^{2}=\frac{1}{1+\sqrt{Q}}\;. (10)

In the limit Q0Q\rightarrow 0, the factor 0.940.94 in Eqs. (8) and (9) deviates from 11 in Eq. (10) due to finite-size effects.

Refer to caption
Figure 4: Average numbers of four types of bonds versus QQ at fixed L=16L=16. The Q1/2Q^{1/2}-behavior in (a) is confirmed by fits in Table 1, and the curves in (b) can be derived from Eq. (9) in the main text.

With the above bond numbers for small QQ, we can get the updating probabilities in the Sweeny algorithm as following: the probability of selecting a bridge bond or internal-perimeter bond is 1\sim 1, and the acceptance probability (Q/vQ/v or vv) is Q\sqrt{Q}, then the associated updating probability is Q\sim\sqrt{Q}; the probability of selecting a backbone bond or external-perimeter bond is Q\sim\sqrt{Q}, and the acceptance probability is 1, then the associated updating probability is also Q\sim\sqrt{Q}. Thus the Q1/2Q^{-1/2}-form critical slowing-down is associated with the Q\sqrt{Q}-form updating probabilities.

Table 1: Fitting results for 𝒩EP/L2\langle\mathcal{N}_{\rm EP}\rangle/L^{2} and 𝒩BB/L2\langle\mathcal{N}_{\rm BB}\rangle/L^{2} at L=16L=16. The data were fitted by the formula QyQ(a0+b1QyQ+b2Q2yQ)Q^{y_{Q}}(a_{0}+b_{1}Q^{y_{Q}}+b_{2}Q^{2y_{Q}}) using the least-square method, with an upper cutoff QmaxQ_{\rm max} on included values of QQ.
   QmaxQ_{\rm max}    yQy_{Q}    a0a_{0}    b1b_{1}    b2b_{2} χ2/DF\chi^{2}/{\rm DF}
0.015625 0.50001(8) 0.9959(8) -0.940(10) 0.762081(1) 10.2/11
𝒩EP/L2\langle\mathcal{N}_{\rm EP}\rangle/L^{2} 0.0078125 0.5000(1) 0.996(1) -0.95(2) 0.803890(1) 10.1/10
0.00390625 0.5000(1) 0.996(2) -0.95(3) 0.840794(1) 10.0/9
0.00195313 0.5002(2) 0.999(2) -1.02(6) 1.775890(1) 8.0/8
0.015625 0.49989(8) 0.9404(8) -0.969(10) 0.796009(1) 14.7/11
𝒩BB/L2\langle\mathcal{N}_{\rm BB}\rangle/L^{2} 0.0078125 0.4999(1) 0.941(1) -0.97(2) 0.815144(1) 14.6/10
0.00390625 0.5001(1) 0.942(2) -1.02(3) 1.284402(1) 11.3/9
0.00195313 0.4999(2) 0.940(2) -0.95(6) 0.401849(1) 9.5/8

III An improved Sweeny algorithm

To overcome the QQ-dependent critical slowing-down, we introduce an improved Sweeny algorithm as shown in Alg. 3. Instead of selecting a bond at random, the improved algorithm randomly proposes one of the following four operations: deleting a backbone bond (delete-backbone), deleting a bridge bond (delete-bridge), adding a backbone bond (add-backbone), and adding a bridge bond (add-bridge). Acceptance probabilities of the operations shall be given later. The algorithm consists of recording and dynamically updating the information of four kinds of bonds, which are directly used in proposing MCMC updates. To the end of this section, we shall see that the improved Sweeny algorithm can largely solve the QQ-dependent critical slowing-down.

1 Randomly choose an operation: delete-backbone, delete-bridge, add-backbone, or add-bridge
2 if delete-backbone then
3      Randomly pick up a backbone bond and delete it with probability PBB()P^{(-)}_{BB}
4 else if delete-bridge then
5      Randomly pick up a bridge bond and delete it with probability PBR()P^{(-)}_{BR}
6 else if add-backbone then
7      Randomly pick up an internal-perimeter bond and occupy it (i.e., add a backbone bond) with probability PBB(+)P^{(+)}_{BB}
8 else if add-bridge then
9      Randomly pick up an external-perimeter bond and occupy it (i.e., add a bridge bond) with probability PBR(+)P^{(+)}_{BR}
10 end if
Algorithm 3 Improved Sweeny Algorithm

For Alg. 3, a list is created to dynamically keep the information for each kind of bonds. The lists record the indices of corresponding bonds, and the numbers of four types of bonds, i.e., 𝒩BB,𝒩BR,𝒩EP,𝒩IP\mathcal{N}_{\rm BB},\mathcal{N}_{\rm BR},\mathcal{N}_{\rm EP},\mathcal{N}_{\rm IP}. The bonds are classified by using the Baxter-Kelland-Wu loops [25] on the medial lattice, as exemplified in Fig. 5. The loops are constructed by assigning each occupied bond two loop arcs to its two sides, and assigning each unoccupied bond two loop arcs which cross the bond. The types of bonds are determined as following: an occupied bond is a backbone (bridge) bond iff the loop arcs on its two sides do not (do) belong to the same loop, and an unoccupied bond is an external (internal) -perimeter bond if the two arcs crossing the bond do not (do) belong to the same loop. To avoid the case that an occupied bond is a backbone bond but has its two loop arcs in the same loop, the boundary conditions are chosen as periodic in one direction and open in the other direction such that the finite-LL square lattice is planar. During the update of a bond state, the connectivity checking and labelling of loop arcs in the loop configuration are performed by simultaneous breadth-first searches starting from two loop arcs of the bond [23, 21]. When a loop is split into two or two loops are merged into one during the update, the simultaneous searching time is only proportional to the size of the smaller loop.

Refer to caption
Figure 5: Example bond configuration and corresponding Baxter-Kelland-Wu loops on the square lattice. Red lines are bridge bonds, blue wavy lines are backbone bonds, green dashed lines represent internal-perimeter bonds and purple dotted lines stand for external-perimeter bonds.

The acceptance probabilities in Alg. 3 are derived as follows. Using α\alpha and β\beta to represent two configurations that can transform into each other in an update, the detailed balance condition has the form

π(α)Ppro(αβ)Pacc(αβ)\displaystyle\pi(\alpha){P_{pro}}(\alpha\rightarrow\beta){P_{acc}}(\alpha\rightarrow\beta) (11)
=\displaystyle= π(β)Ppro(βα)Pacc(βα),\displaystyle\pi(\beta){P_{pro}}(\beta\rightarrow\alpha){P_{acc}}(\beta\rightarrow\alpha)\;,

where π\pi is the configuration weight, PproP_{pro} and PaccP_{acc} are the proposal and acceptance probabilities, respectively. The detailed balance condition associated with the deletion or addition of a backbone bond can be written as

11𝒩BB(α)PBB()=1v1𝒩IP(β)PBB(+),\displaystyle 1\cdot\frac{1}{\mathcal{N}_{\rm BB}(\alpha)}\cdot P^{(-)}_{BB}=\frac{1}{v}\cdot\frac{1}{\mathcal{N}_{\rm IP}(\beta)}\cdot P^{(+)}_{BB}\;, (12)

and that associated with the deletion or addition of a bridge bond as

11𝒩BR(α)PBR()=Qv1𝒩EP(β)PBR(+).\displaystyle 1\cdot\frac{1}{\mathcal{N}_{\rm BR}(\alpha)}\cdot P^{(-)}_{BR}=\frac{Q}{v}\cdot\frac{1}{\mathcal{N}_{\rm EP}(\beta)}\cdot P^{(+)}_{BR}\;. (13)

Following the Metropolis-Hastings criterion, the acceptance probabilities are calculated as

PBB()=min{1,𝒩BB(α)𝒩IP(β)1v},\displaystyle P^{(-)}_{BB}=\min\{1,\frac{\mathcal{N}_{\rm BB}(\alpha)}{\mathcal{N}_{\rm IP}(\beta)}\cdot\frac{1}{v}\}\;, (14)
PBB(+)=min{1,𝒩IP(β)𝒩BB(α)v},\displaystyle P^{(+)}_{BB}=\min\{1,\frac{\mathcal{N}_{\rm IP}(\beta)}{\mathcal{N}_{\rm BB}(\alpha)}\cdot v\}\;, (15)
PBR()=min{1,𝒩BR(α)𝒩EP(β)Qv},\displaystyle P^{(-)}_{BR}=\min\{1,\frac{\mathcal{N}_{\rm BR}(\alpha)}{\mathcal{N}_{\rm EP}(\beta)}\cdot\frac{Q}{v}\}\;, (16)
PBR(+)=min{1,𝒩EP(β)𝒩BR(α)vQ}.\displaystyle P^{(+)}_{BR}=\min\{1,\frac{\mathcal{N}_{\rm EP}(\beta)}{\mathcal{N}_{\rm BR}(\alpha)}\cdot\frac{v}{Q}\}\;. (17)
Refer to caption
Figure 6: τ𝒩HF\tau_{\mathcal{N}_{\rm HF}} for the improved Sweeny algorithm.

Our results of τ𝒩HF\tau_{\mathcal{N}_{\rm HF}} at vc=Qv_{c}=\sqrt{Q} for the improved Sweeny algorithm is shown in Fig. 6. When fixing the size LL, it can be seen that, as QQ decreases, τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} is almost independent of QQ before a turning point Qc(L)Q_{c}(L), after which τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} tends to increase as Q1/2\sim Q^{-1/2}. The value of Qc(L)Q_{c}(L) decreases with increasing system size LL, and it is expected that Qc(L)0Q_{c}(L)\rightarrow 0 in the limit LL\rightarrow\infty. Thus, comparing with results of the original Sweeny algorithm (i.e., Fig. 3), the improved Sweeny algorithm greatly improves simulation efficiency. For Q=106Q=10^{-6} and L=32L=32, the autocorrelation time for the improved algorithm is of order O(10)O(10), and it decreases with LL, making the algorithm sufficient for most studies of the random-cluster model with small QQ.

The performance of the improved Sweeny algorithm can be understood from the acceptance probabilities: PBB(+)P^{(+)}_{BB} and PBR()P^{(-)}_{BR} are of order O(1)O(1) for QQc(L)Q\gg Q_{c}(L); PBB()P^{(-)}_{BB} and PBR(+)P^{(+)}_{BR} are of order O(1)O(1) for all QQ. The residual Q1/2Q^{-1/2} behavior is attributed to that PBB(+)P^{(+)}_{BB} and PBR()P^{(-)}_{BR} are of order O(Q)O(\sqrt{Q}) for QQc(L)Q\ll Q_{c}(L). Detailed analyses of these acceptance probabilities will be presented in Section IV. In the limit Q0Q\rightarrow 0, the improved Sweeny algorithm fails since PBR()=PBB(+)=0P^{(-)}_{BR}=P^{(+)}_{BB}=0, i.e., no bridge bond can be deleted from a uniform spanning tree and no backbone bond can be added to the tree.

IV An improved Kawasaki algorithm and its combination with the improved Sweeny algorithm

In the limit Q0Q\rightarrow 0, since both the original and improved Sweeny algorithms are invalid, one has to employ other methods. When Q0Q\rightarrow 0, though no bond can be erased or occupied, bonds may exchange their states. Thus, to simulate models with Q0Q\rightarrow 0, we can use the Kawasaki algorithm [20], in which a randomly selected empty bond is proposed to switch with a randomly selected occupied bond.

Refer to caption
Figure 7: Results for the uniform spanning trees on the square lattice. (a) The autocorrelation time of the Kawasaki algorithm τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} versus LL. Here each swap of two bond states is regarded as the time unit “hit”. (b) The average numbers of bonds in the intermediate state 𝒩EP\langle\mathcal{N}^{*}_{\rm EP}\rangle and 𝒩BB\langle\mathcal{N}^{*}_{\rm BB}\rangle versus LL. Deviations at small sizes represent finite-size corrections.

We applied the Kawasaki algorithm to the uniform spanning trees, which appear in the limit of Q0Q\rightarrow 0 with vc/Q1/2v_{c}/Q^{1/2} being finite, for the critical square-lattice model. The results are shown in Fig. 7(a), where it can be seen that τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} diverges as L3/2\sim L^{3/2}. We understand this LL-dependent slowing-down as follows.

A Kawasaki update can be performed in two steps: first deleting a randomly chosen occupied bond so that the spanning tree is split into two trees, then occupying a randomly chosen empty bond. We call the configuration between the two steps an intermediate state. For a uniform spanning tree, the overall update would be accepted if and only if the chosen empty bond is an external-perimeter bond in the intermediate state.

The number of external-perimeter bonds in the intermediate states 𝒩EP\mathcal{N}^{*}_{\rm EP} can be counted by observing changes of the loops: on a spanning tree, the deletion of an occupied bond decompose a large loop into two loops, and 𝒩EP\mathcal{N}^{*}_{\rm EP} is proportional to the size of the external perimeter associated with the smaller loop. The loop surrounding a cluster defines its hull, and the fraction of loop arcs in the hull corresponding to external-perimeter bonds is called the external perimeter of the cluster. The size of the smaller loop is determined as LdHx2\sim L^{d_{\rm H}-x_{2}} in Ref. [23], where dHd_{\rm H} is the fractal dimension of the hull and x2x_{2} is the two-arm exponent [26]. Analogously, we conjecture that the size of the external perimeter associated with the smaller loop is LdEPx2\sim L^{d_{\rm EP}-x_{2}}, where dEPd_{\rm EP} is the fractal dimension of the external perimeter. According to the literature [26, 27, 28, 23], the fractal dimensions take values dH=1+2/gd_{\rm H}=1+2/g and dEP=1+g/8d_{\rm EP}=1+g/8; the parameter gg and kk-arm exponents xkx_{k} are given by

Q=2+2cos(gπ/2),g[2,4],\displaystyle Q=2+2\cos(g\pi/2)\;,\;g\in[2,4]\;, (18)
xk=(g/8)k2(g4)2/(8g),k2.\displaystyle x_{k}=(g/8)k^{2}-(g-4)^{2}/(8g)\;,\;k\geq 2\;. (19)

For Q0Q\rightarrow 0, one gets g=2g=2, x2=3/4x_{2}=3/4, and dEP=5/4d_{\rm EP}=5/4. Then we expect 𝒩EPL5/43/4L1/2\langle\mathcal{N}^{*}_{\rm EP}\rangle\sim L^{5/4-3/4}\sim L^{1/2}, which is numerically confirmed in Fig. 7(b). Thus, the probability that a randomly selected empty bond is an external-perimeter bond is 𝒩EP/2L2L3/2\langle\mathcal{N}^{*}_{\rm EP}\rangle/2L^{2}\sim L^{-3/2}, which leads to τ𝒩HF/L2L3/2\tau_{\mathcal{N}_{\rm HF}}/L^{2}\sim L^{3/2} on the uniform spanning trees.

To eliminate the above LL-dependent slowing-down, similar to the idea in improving the Sweeny algorithm, we make use of the real-time information of four kinds of bonds and propose an improved Kawasaki algorithm, as shown in Alg. 4. This improved algorithm either exchange a bridge bond with an external-perimeter bond, or exchange a backbone bond with an internal-perimeter bond. Since the proposed update always exist and the acceptance probability is 11, the improved Kawasaki algorithm does not have the above LL-dependent slowing-down. This is confirmed by our simulation results which have τ𝒩HF/L20.55\tau_{\mathcal{N}_{\rm HF}}/L^{2}\simeq 0.55 for L=4L=4 to 128128.

1 Randomly choose an occupied bond and erase the bond, then we have an intermediate configuration
2 if The chosen bond is a bridge bond then
3      Randomly occupy an external-perimeter bond in the intermediate configuration
4 else if The chosen bond is a backbone bond then
5      Randomly occupy an internal-perimeter bond in the intermediate configuration
6 end if
Algorithm 4 Improved Kawasaki Algorithm

For nonzero small QQ, since updates in the (improved) Kawasaki algorithm cannot change the total number of occupied bonds, one has to employ the (improved) Sweeny algorithm. In Section III, for the improved Sweeny algorithm at fixed LL, we have shown the appearance of a QQ-dependent slowing-down after an turning point Qc(L)Q_{c}(L). This can be understood by observing the acceptance probabilities as follows. First, for updates of deleting a backbone bond or adding a bridge bond, substituting Eqs. (8) and (9) into Eqs. (14) and (17), one has

PBB()\displaystyle P^{(-)}_{BB} =min{1,𝒩BB(α)𝒩IP(α)+11Q}1,\displaystyle=\min\{1,\frac{\mathcal{N}_{\rm BB}(\alpha)}{\mathcal{N}_{\rm IP}(\alpha)+1}\cdot\frac{1}{\sqrt{Q}}\}\simeq 1\;, (20)
PBR(+)\displaystyle P^{(+)}_{BR} =min{1,𝒩EP(β)𝒩BR(β)+11Q}1,\displaystyle=\min\{1,\frac{\mathcal{N}_{\rm EP}(\beta)}{\mathcal{N}_{\rm BR}(\beta)+1}\cdot\frac{1}{\sqrt{Q}}\}\simeq 1\;, (21)

where 𝒩IP(α)+1=𝒩IP(β)\mathcal{N}_{\rm IP}(\alpha)+1=\mathcal{N}_{\rm IP}(\beta) and 𝒩BR(β)+1=𝒩BR(α)\mathcal{N}_{\rm BR}(\beta)+1=\mathcal{N}_{\rm BR}(\alpha) are used. Then, for updates of deleting a bridge bond or adding a backbone bond, more considerations are needed. For example, after deleting a bridge from configuration α\alpha, there is no simple relation between 𝒩EP(β)\mathcal{N}_{\rm EP}(\beta) and 𝒩EP(α)\mathcal{N}_{\rm EP}(\alpha). Instead, one has 𝒩EP(β)=𝒩EP\mathcal{N}_{\rm EP}(\beta)=\mathcal{N}^{*}_{\rm EP}, which scales as 𝒩EPLdEPx2\langle\mathcal{N}^{*}_{\rm EP}\rangle\sim L^{d_{\rm EP}-x_{2}} in the limit Q0Q\rightarrow 0 as previously discussed. For nonzero small QQ and large LL, we expect that Eq. (8) still holds, thus suppose that 𝒩EPL2Q+aLdEPx2\langle\mathcal{N}^{*}_{\rm EP}\rangle\sim L^{2}\sqrt{Q}+aL^{d_{\rm EP}-x_{2}}. Substituting this result and Eq. (9) into Eq. (16), we get the acceptance probability

PBR()=𝒩BR𝒩EPQQQ+aLdEPx22.\displaystyle P^{(-)}_{BR}=\frac{\mathcal{N}_{\rm BR}}{\mathcal{N}^{*}_{\rm EP}}\cdot\sqrt{Q}\sim\frac{\sqrt{Q}}{\sqrt{Q}+aL^{d_{\rm EP}-x_{2}-2}}\;. (22)

Thus, we expect that the autocorrelation time scales as the inverse of the above acceptance probability

τ𝒩HF/L2Q+aLdEPx22Q.\displaystyle\tau_{\mathcal{N}_{\rm HF}}/L^{2}\sim\frac{\sqrt{Q}+aL^{d_{\rm EP}-x_{2}-2}}{\sqrt{Q}}\;. (23)

From the above equation, for a fixed QQ, as LL increases, since dEPx22<0d_{\rm EP}-x_{2}-2<0, one has τ𝒩HF/L2𝒪(1)\tau_{\mathcal{N}_{\rm HF}}/L^{2}\to{\cal O}(1); and for a fixed LL, as QQ decreases, τ𝒩HF/L2Q1/2\tau_{\mathcal{N}_{\rm HF}}/L^{2}\to Q^{-1/2}. These imply the turning near Qc(L)Q_{c}(L) and explain the results of the improved Sweeny algorithm, i.e., Fig. 6.

Finally, since the acceptance probabilities for operations in the improved Kawasaki algorithm are also 11 for nonzero QQ, we can combine the improved Sweeny algorithm with the improved Kawasaki algorithm. Our simulation results for this combined method are shown in Fig. 8, where τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} versus LL is almost a straight line. Thus the improved Kawasaki method eliminates the Q1/2\sim Q^{-1/2} form slowing-down of the improved Sweeny algorithm at fixed LL, and the combined improved Sweeny-Kawasaki method serves as a complete solution for simulating the random-cluster model with small QQ.

Refer to caption
Figure 8: τ𝒩HF\tau_{\mathcal{N}_{\rm HF}} for the combined method of the improved Sweeny and improved Kawasaki algorithms. The LL-dependence of τ𝒩HF/L2\tau_{\mathcal{N}_{\rm HF}}/L^{2} is very small as demonstrated by data points at Q=106Q=10^{-6}. Here in our simulation a step consists of a single-bond update in the improved Sweeny algorithm and a two-bond swap in the improved Kawasaki algorithm, which is counted as two “hits”.

V Discussion and conclusion

To summarize, by introducing the number of half faces, we observe a QQ-dependent slowing-down in two dimensions, which further enriches the dynamics of the Sweeny algorithm for the random-cluster model. The slowing-down is understood by observing the updating probabilities, which contains the critical bond weight vc=Qv_{c}=\sqrt{Q} on the square lattice. By classifying the bonds into four types and dynamically updating their type information, we propose an improved Sweeny algorithm, which (at fixed LL) pushes the slowing-down to much smaller QQ-region. Since the QQ-dependent slowing-down renders the (improved) Sweeny algorithm being non-ergodic for Q0Q\rightarrow 0, we resort to the Kawasaki algorithm. The latter algorithm is found to exhibit an LL-dependent slowing-down, and this slowing-down can be fully suppressed in an improved version of the algorithm by making use of the bond-type classification. Finally, we formulate an improved Sweeny-Kawasaki method which combines advantages of the two improved algorithms and eliminates the above critical slowing-down.

We note that updates in the improved algorithms proceed with the aid of Baxter-Kelland-Wu loops, as it is convenient to classify the bonds using the loops, and it is also faster to perform connectivity checking in the loops than in the clusters. The computational cost of the algorithms primarily stems from the connectivity checking: simultaneous breadth-first search [23, 21] in the clusters has a runtime scaling of LdFx2\sim L^{d_{\rm F}-x_{2}}, while that in the loops has a runtime scaling of LdHx2\sim L^{d_{\rm H}-x_{2}} with dHdFd_{\rm H}\leq d_{\rm F}. For not too large LL, the improved Sweeny-Kawasaki method in this work should be adequate for simulating the random-cluster model. To speedup the simulations for large LL, it should be worth to study how to incorporate dynamic connectivity algorithms which have runtime scaling at most logarithmic in the system size [21, 22].

It is also noted that, solely for the uniform spanning trees, there exist algorithms which have excellent performance. These include the celebrated Wilson’s algorithm [29], which uniformly constructs a spanning tree from scratch by a loop-erased random walk, and a smart Kawasaki algorithm which uniformly generates spanning trees by using a fast link-cut tree data structure [30]. The latter is different from our improved Kawasaki algorithm, which constructs loops to classify the bonds and applies also to models with Q0Q\neq 0.

Furthermore, though our simulations were performed at vc=Qv_{c}=\sqrt{Q} on the square lattice, for v=Qσv=Q^{\sigma} with general values of 0<σ<10<\sigma<1, as Q0Q\rightarrow 0 there is also QQ-dependent slowing-down for the Sweeny algorithm. This is because in the latter case the updating probabilities tend to zero as QσQ^{\sigma} or Q1σQ^{1-\sigma}. The improved Sweeny and Kawasaki algorithms can also solve this QQ-dependent slowing-down, due to reasons similar to those for v=Qv=\sqrt{Q}. Also, the Q1/2Q^{-1/2}-form critical slowing-down for the random-cluster model is expected to exist on other 2D lattices. For example, on the triangular (𝕋\mathbb{T}) and honeycomb (\mathbb{H}) lattices, when Q4Q\leq 4, it is known [31, 1, 4] that

vc𝕋=2cos[23cos1(Q2)]1,\displaystyle v^{\mathbb{T}}_{c}=2\cos[\frac{2}{3}\cos^{-1}(\frac{\sqrt{Q}}{2})]-1\;, (24)
vc=2Qcos[13cos1(Q2)],\displaystyle v^{\mathbb{H}}_{c}=2\sqrt{Q}\cos[\frac{1}{3}\cos^{-1}(\frac{\sqrt{Q}}{2})]\;, (25)
vc𝕋vc=Q.\displaystyle v^{\mathbb{T}}_{c}\cdot v^{\mathbb{H}}_{c}=Q\;. (26)

These critical bond weights vc𝕋v^{\mathbb{T}}_{c} and vcv^{\mathbb{H}}_{c} also vanish as Q\sim\sqrt{Q} when Q0Q\rightarrow 0. However, when the spatial dimension d3d\geq 3, for small QQ the critical point behaves as vcQv_{c}\sim Q (hence the limit Q0Q\to 0 corresponds to the spanning-forest model) [32]. Thus, the QQ-dependent critical slowing-down does not occur for d3d\geq 3. Nevertheless, the improved Sweeny and Kawashaki algorithms may still help to improve the simulation efficiency.

Finally, the number of half faces proves to be very useful in demonstrating the slowing-down of the unimproved algorithms. Further work may explore more quantities which have similar behavior. Despite the mechanism discussed in this work, we cannot rule out any other property which may cause more severe slowing-down.

Acknowledgements.
This work has been supported by the National Natural Science Foundation of China (under grant No. 12275263 and No. 12375026), the Innovation Program for Quantum Science and Technology (under grant No. 2021ZD0301900), and the Natural Science Foundation of Fujian Province of China (under grant No. 2023J02032). We thank David B. Wilson for his early contributions on this work.

Appendix: Density of bonds in the thermodynamic limit

In this Appendix, we give an alternative derivation of the results in Eq. (10). On the square lattice in the thermodynamic limit LL\rightarrow\infty, the self-duality condition leads to the following relation between the critical numbers of bonds

𝒩BB+𝒩BR/2L2=𝒩IP+𝒩EP/2L2=1/2.\displaystyle\langle\mathcal{N}_{\rm BB}+\mathcal{N}_{\rm BR}\rangle/2L^{2}=\langle\mathcal{N}_{\rm IP}+\mathcal{N}_{\rm EP}\rangle/2L^{2}=1/2\;. (27)

In the Fortuin-Kasteleyn transformation, the density of occupied bonds can be related to the nearest-neighbor connectivity gng_{\rm n} (i.e., the probability that two nearest-neighbor sites belong to the same cluster) as [33]

𝒩BB+𝒩BR/2L2=pc[gn+(1gn)/Q],\displaystyle\langle\mathcal{N}_{\rm BB}+\mathcal{N}_{\rm BR}\rangle/2L^{2}=p_{c}[g_{\rm n}+(1-g_{\rm n})/Q]\;, (28)

where pc=Q/(1+Q)p_{c}=\sqrt{Q}/(1+\sqrt{Q}). Substituting Eq. (27) into Eq. (28), one gets [33]

gn=2+Q2(1+Q).\displaystyle g_{\rm n}=\frac{2+\sqrt{Q}}{2(1+\sqrt{Q})}\;. (29)

Letting PnP_{\rm n} be the probability that two nearest-neighbor sites are connected by a path not including the bond between the two sites, one can express the probability gng_{\rm n} as

gn=Pn+(1Pn)v/Q1+v/Q.\displaystyle g_{\rm n}=P_{\rm n}+(1-P_{\rm n})\frac{v/Q}{1+v/Q}\;. (30)

Substituting vc=Qv_{c}=\sqrt{Q} and Eq. (29) into Eq. (30) leads to Pn=1/2P_{\rm n}=1/2. Finally, one can express the critical densities of four kinds of bonds as

𝒩BB/2L2=Pnv1+v=Q2(1+Q),\displaystyle\langle\mathcal{N}_{\rm BB}\rangle/2L^{2}=P_{\rm n}\frac{{v}}{1+{v}}=\frac{\sqrt{Q}}{2(1+\sqrt{Q})}\;, (31)
𝒩BR/2L2=(1Pn)v/Q1+v/Q=12(1+Q),\displaystyle\langle\mathcal{N}_{\rm BR}\rangle/2L^{2}=(1-P_{\rm n})\frac{{v}/Q}{1+{v}/Q}=\frac{{1}}{2(1+\sqrt{Q})}\;, (32)
𝒩IP/2L2=Pn11+v=12(1+Q),\displaystyle\langle\mathcal{N}_{\rm IP}\rangle/2L^{2}=P_{\rm n}\frac{1}{1+{v}}=\frac{{1}}{2(1+\sqrt{Q})}\;, (33)
𝒩EP/2L2=(1Pn)11+v/Q=Q2(1+Q).\displaystyle\langle\mathcal{N}_{\rm EP}\rangle/2L^{2}=(1-P_{\rm n})\frac{1}{1+{v}/Q}=\frac{\sqrt{Q}}{2(1+\sqrt{Q})}\;. (34)

The above densities have been derived in Ref. [24], and they are consistent with 𝒩BB/2L2=𝒩BR/2L2=1/4\langle\mathcal{N}_{\rm BB}\rangle/2L^{2}=\langle\mathcal{N}_{\rm BR}\rangle/2L^{2}=1/4 for Q=1Q=1 in Ref. [34].

References

  • [1] F. Wu, Rev. Mod. Phys. 54, 235 (1982), and ibid. 55, 315 (1983) .
  • Kasteleyn and Fortuin [1969] P. W. Kasteleyn and C. M. Fortuin, J. Phys. Soc. Jpn. 26 (Suppl.), 11 (1969).
  • Fortuin and Kasteleyn [1972] C. M. Fortuin and P. W. Kasteleyn, Physica 57, 536 (1972).
  • Grimmett [2006] G. Grimmett, The random-cluster model (Springer, 2006).
  • [5] D. Stauffer and A. Aharony, Introduction to percolation theory, revised 2nd ed. (Taylor & Francis, 1994).
  • Jacobsen et al. [2005] J. L. Jacobsen, J. Salas, and A. D. Sokal, J. Stat. Phys. 119, 1153 (2005).
  • Newman and Barkema [1999] M. E. Newman and G. T. Barkema, Monte Carlo methods in statistical physics (Clarendon Press, 1999).
  • Landau and Binder [2005] D. Landau and K. Binder, A guide to Monte Carlo simulations in statistical physics, 2nd ed. (Cambridge University Press, 2005).
  • Swendsen and Wang [1987] R. H. Swendsen and J.-S. Wang, Phys. Rev. Lett. 58, 86 (1987).
  • Chayes and Machta [1998] L. Chayes and J. Machta, Physica 254, 477 (1998).
  • Sweeny [1983] M. Sweeny, Phys. Rev. B 27, 4445 (1983).
  • [12] A. D. Sokal, in Functional Integration: Basics and Applications, edited by C. Witt-Morette, P. Cartier and A. Folacci (Plenum, New York, 1997) pp. 131–192.
  • Hohenberg and Halperin [1977] P. C. Hohenberg and B. I. Halperin, Rev. Mod. Phys. 49, 435 (1977).
  • Metropolis et al. [1953] N. Metropolis, A. W. Rosenbluth, M. N. Rosenbluth, A. H. Teller, and E. Teller, J. Chem. Phys. 21, 1087 (1953).
  • Hastings [1970] W. K. Hastings, Biometrika 57, 97 (1970).
  • Glauber [1963] R. J. Glauber, J. Math. Phys. 4, 294 (1963).
  • Prokof’ev and Svistunov [2001] N. Prokof’ev and B. Svistunov, Phys. Rev. Lett. 87, 160601 (2001).
  • Deng et al. [2007a] Y. Deng, T. M. Garoni, and A. D. Sokal, Phys. Rev. Lett. 98, 230602 (2007a).
  • Li and Sokal [1989] X.-J. Li and A. D. Sokal, Phys. Rev. Lett. 63, 827 (1989).
  • Kawasaki [1966] K. Kawasaki, Phys. Rev. 145, 224 (1966).
  • Elçi and Weigel [2013] E. M. Elçi and M. Weigel, Phys. Rev. E 88, 033303 (2013).
  • Elçi [2015] E. M. Elçi, Algorithmic and geometric aspects of the random-cluster model, Ph.D. thesis, Coventry University (2015).
  • Deng et al. [2010] Y. Deng, W. Zhang, T. M. Garoni, A. D. Sokal, and A. Sportiello, Phys. Rev. E 81, 020102 (2010).
  • Elçi et al. [2016] E. M. Elçi, M. Weigel, and N. G. Fytas, Nucl. Phys. B 903, 19 (2016).
  • Baxter et al. [1976] R. J. Baxter, S. B. Kelland, and F. Y. Wu, J. Phys. A: Math. Gen 9, 397 (1976).
  • Saleur and Duplantier [1987] H. Saleur and B. Duplantier, Phys. Rev. Lett. 58, 2325 (1987).
  • Nienhuis [1984] B. Nienhuis, J. Stat. Phys. 34, 731 (1984).
  • Cardy [1998] J. Cardy, J. Phys. A: Math. Gen 31, L105 (1998).
  • Wilson [1996] D. B. Wilson, in Proceedings of the twenty-eighth annual ACM symposium on theory of computing (1996) pp. 296–303.
  • Russo et al. [2018] L. Russo, A. Teixeira, and A. Francisco, Algorithms 11, 53 (2018).
  • Baxter et al. [1978] R. J. Baxter, H. Temperley, and S. E. Ashley, Proc. R. Soc. Lond. A 358, 535 (1978).
  • Deng et al. [2007b] Y. Deng, T. M. Garoni, and A. D. Sokal, Phys. Rev. Lett. 98, 030602 (2007b).
  • Hu et al. [2014] H. Hu, H. W. J. Blöte, R. M. Ziff, and Y. Deng, Phys. Rev. E 90, 042106 (2014).
  • Xu et al. [2014] X. Xu, J. Wang, Z. Zhou, T. M. Garoni, and Y. Deng, Phys. Rev. E 89, 012120 (2014).