cond-mat0608547/loop.tex
1: \documentclass[aps,pre,preprint,showpacs,floatfix]{revtex4}
2: \usepackage{graphicx,bm,epsfig}
3: \topmargin=-2.0mm
4: 
5: \newcommand{\zy}[1]{\begin{center}
6:   \fbox{\parbox{0.8\hsize}{\normalsize #1}}
7: \end{center}}
8: 
9: \begin{document}
10: \title{Geometric properties of two-dimensional O($n$) loop configurations}
11: \author{Chengxiang Ding$^{1}$, Xiaofeng Qian$^{2}$, 
12: Youjin Deng$^{3}$, Wenan Guo$^{1}$
13: and Henk W.J.~Bl\"ote$^{2,4}$ 
14: }
15: \affiliation{$^{1}$ Physics Department, Beijing Normal University, 
16: Beijing 100875, P.R.China\\
17: $^{2}$ Instituut Lorentz, Universiteit Leiden,
18: Postbus 9506, 2300 RA Leiden, The Netherlands \\
19: $^{3}$ New York University, USA\\
20: $^{4}$Faculty of Applied Sciences, Delft University of
21: Technology,  P.O. Box 5046, 2600 GA Delft, The Netherlands \\
22: }
23: \date{\today}
24: \pacs{64.60.Ak, 64.60.Fr, 64.60.Kw, 75.10.Hk}
25: \begin{abstract}
26: We study the fractal geometry of O($n$) loop configurations
27: in two dimensions by means of scaling and a Monte Carlo method, and
28: compare the results with predictions based on the Coulomb gas technique.
29: The Monte Carlo algorithm is applicable to models with
30: noninteger $n$ and uses local updates. Although these updates typically
31: lead to nonlocal modifications of loop connectivities, the number of 
32: operations required per update is only of order one. The Monte Carlo
33: algorithm is applied to the O($n$) model for several values of $n$, 
34: including noninteger ones. We thus determine scaling exponents that
35: describe the fractal nature of O($n$) loops at criticality. The results
36: of the numerical analysis agree with the theoretical predictions.
37: \end{abstract}
38: \maketitle
39: 
40: \section{Introduction}
41: The O($n$) model consists of $n$-component spins  
42: $\vec{s_{i}}=( s^{1}_{i},s^{2}_{i}, \ldots s^{n}_{i})$ on a lattice, with
43: isotropic, i.e., O($n$) invariant couplings. The common form of the reduced
44: Hamiltonian of the O($n$) spin model is
45: \begin{equation}\label{spinhamilton}
46: H=-\frac{J}{k_B T} \sum_{\langle i, j\rangle } \vec{s_{i}} \cdot \vec{s_{j}} 
47: \label{Ham}
48: \end{equation}
49: where the indices $i$ and $j$ represent lattice sites, and the sum is
50: over all nearest neighbor pairs; $J$ is the coupling constant, $k_B$
51: is Boltzmann's constant, and $T$ is the temperature.
52: Thus, the partition function of the model is
53: \begin{equation}
54: Z_{\rm spin}=\int \prod_{<i,j>} \exp \left(\frac{J}{k_B T}  \,
55: \vec{s_{i}} \cdot \vec{s_{j}} \right) \prod_{k} d \vec{s_{k}}
56: \end{equation}
57: where the spins are normalized such that 
58: This model includes as special cases the Ising, the XY and the Heisenberg 
59: model, for $n=1$, $2$ and $3$, respectively. 
60: 
61: In the high-temperature limit, the bond weight
62: $w(\vec{s_{i}} \cdot \vec{s_{j}})$ reduces in first order to
63: $(1+x\vec{s_i}\cdot \vec{s_j})$  with $x=J/(k_BT)$. Thus the partition
64: function takes the form
65: \begin{equation}
66: Z_{\rm spin}=\int\prod_{\langle i, j\rangle }
67: (1+x\vec{s_i}\cdot \vec{s_j}) \prod_{k}d\vec{s_{k}}
68: \label{spinpartition}
69: \end{equation}
70: The bond weight still satisfies the O($n$) symmetry implied by 
71: Eq.~(\ref{Ham}). According to the assumption of universality, the
72: universality class of a phase transition is determined by only very
73: few parameters including the symmetry of the spin-spin interactions.
74: It is thus reasonable to expect \cite{N1} that the reduced Hamiltonian
75: that corresponds with Eq.~(\ref{spinpartition}), namely 
76: $H=-\sum_{<i,j>}\ln(1+x\vec{s_{i}}\cdot\vec{s_{j}})$
77: with $x=J/(k_BT)$ not necessarily small, still belongs to the
78: O($n$) universality class in two dimensions.
79: 
80: The O($n$) model (\ref{spinpartition}) on the honeycomb lattice
81: can be mapped on the O($n$) loop model \cite{DMNS} on the same lattice,
82: with a partition sum
83: \begin{equation}
84: Z_{\rm loop}=\sum_{G}x^{N_{b}}n^{N_{l}}
85: \label{graphpartition}
86: \end{equation}
87: where the graph $G$ covers $N_{b}$ bonds of the lattice,  and consists of 
88: $N_{l}$ closed, non-intersecting loops.
89: 
90: In the language of the $O(n)$ loop model, $x$ is the weight of a bond
91: visited by a loop, and $n$ is the loop weight, and no longer restricted
92: to be an integer.
93: 
94: The research of O($n$) models is a subject of a considerable history,
95: in which a prominent place is occupied by the exact results \cite{N1}
96: for the O($n$) loop model on the honeycomb lattice.
97: These results include the critical points for $-2\leq n\leq 2$, and the
98: temperature and the magnetic exponent. 
99: 
100: Also the geometric description of fluctuations at and near criticality 
101: has a long history, which goes back to the formulation of phase 
102: transitions in terms of the droplet model \cite{Fisher}.
103: For the $q$-state Potts model (for a review, see Ref.~\onlinecite{Wu}),
104: it was found that the fractal dimension of Kasteleyn-Fortuin (KF)
105: clusters \cite{KF} is equal to the magnetic scaling exponent $y_h$.
106: More generally, geometric Potts clusters can be defined by connecting 
107: neighboring, equal Potts spins by a bond percolation process. Several
108: new critical exponents were found by Coulomb gas and other 
109: methods \cite{CP, BKN, DBN,Con}. These exponents describe the geometric
110: properties and the renormalization flow of this model.
111: 
112: For the O($n$) loop model, the fractal dimension $d_l$ of the critical
113: loop gas has, to our knowledge, not yet been derived explicitly.
114: However, a clue can be obtained from the relation \cite{DBN,JS1} 
115: between the exponents describing random clusters of the tricritical
116: Potts model, and those describing Potts clusters in the critical Potts
117: model. From the equivalence of the critical O($n$)
118: model and the tricritical $q=n^2$-state Potts model \cite{N1}, one 
119: can thus associate the fractal dimension $d_l$ of the critical O($n$)
120: loop gas to the hull fractal dimension of critical Potts clusters.
121: The latter dimension was conjectured by Vanderzande \cite{V}. 
122: 
123: In the present paper, we focus on a fundamental non-thermodynamic scaling 
124: dimension behind the geometric properties of O($n$) loops which follows
125: from the known correspondence with the Coulomb gas \cite{N}, and relate it
126: to some exponents describing such properties. These exponents are exact
127: and include the fractal dimension $d_l$ of O($n$) loops at criticality. 
128: {}From a similar calculation, Duplantier and Saleur \cite{SD, DS} derived
129: the fractal dimension $d_a$ of the interior of these loops. 
130: 
131: While theoretical predictions are available, thus far there is no
132: numerical evidence in support of these, except for the Ising case $n=1$
133: or $q=2$ \cite{JS1}.  One of the reasons behind this situation may be
134: that the O($n$) partition sum contains the number of loops.
135: In the existing Monte Carlo algorithm for the loop model \cite{Wdyn},
136: the acceptance probability of a local update thus depends (for $n \neq 1$)
137: on the change of the number of loops due to the Monte Carlo move.
138: The determination of this number is a nonlocal task and requires a
139: number of operations that increases algebraically with the system size
140: $L$. Critical slowing down can make this situation even worse, so that
141: this way of simulation is restricted to rather small system sizes.
142: 
143: Until now, a sufficiently efficient Monte Carlo algorithm for
144: the O($n$) loop model has not been described.
145: Therefore, in this work, we develop a new Monte Carlo algorithm, which
146: is applicable to models with noninteger $n>1$, and uses local updates.
147: Although these updates typically lead to nonlocal modifications of the
148: loop connectivities, the number of operations required per update is
149: only of order one, and essentially independent of the system size. 
150: 
151: We then apply the algorithm to the critical O($n$) loop model and 
152: determine exponents of some geometric observables. 
153: The results agree with the theoretical predictions. 
154: 
155: The outline of the rest of this paper is as follows. In Sec. \ref{CG}
156: we show how a fundamental non-thermodynamic scaling dimension behind
157: some geometric properties of O($n$) loops can be derived exactly from
158: a mapping on the Coulomb gas, and how it relates to exponents describing
159: some geometric observables. Section \ref{MC} introduces the Monte Carlo
160: algorithm. In Sec. \ref{Sim} we apply the algorithm to the critical
161: O($n$) loop model and determine exponents of some geometric observables.
162: 
163: \section{Coulomb gas derivation and scaling formulas}
164: \label{CG}
165: It is well known that geometric and fractal properties of O($n$) loops
166: and various types of critical clusters can be analyzed by means of
167: a mapping on the Coulomb gas \cite{Kad,N}. A number of exact scaling
168: dimensions was obtained by this technique, see e.g.,
169: Refs.~\onlinecite{dN,DS,SD,BKN,N}. Here we base ourselves on these
170: analyses, which rely on a reformulation of correlation functions 
171: $g(r)$ in the model of interest in terms of the Coulomb gas.
172: The dimensions $X(e,m)$ associated with such correlation
173: functions are described by pairs of electric and magnetic charges,
174: $(e_0,e_r)$ and $(m_0,m_r)$, separated by a distance $r$:
175: \begin{equation}
176: X(e,m) = - \frac{e_0 e_r}{2 g} - \frac{m_0 m_r g}{2}
177: \label{Xem}
178: \end{equation}
179: where $g$ is the coupling constant of the Coulomb gas. For the critical
180: O($n$) model it is given by
181: \begin{equation}
182: g= 1+ \frac{1}{\pi} {\rm arccos} \frac{n}{2}
183: \label{g}
184: \end{equation}
185: where we use a normalization that is in agreement with earlier literature
186: but different from that used in
187: Ref.~\onlinecite{N} (our $g$ is four times smaller, and the charges
188: differ by a factor two such that the $X(e,m)$ are the same).
189: 
190: Let us now consider the correlation function at criticality describing
191: the probability that two lattice edges separated by a distance $r$ are
192: part of the same loop. It decays with an exponent $2X_l$ where $X_l$
193: is the O($n$) loop scaling dimension. The exponent $X_l$ is described
194: by a pair of magnetic charges $m_0=-m_r= 1$ and a pair of electric
195: charges $e_0=e_r=1-g$.  This leads to
196: \begin{equation}
197: X_l = 1-\frac{1}{2g}
198: \label{xl}
199: \end{equation}
200: This dimension $X_l$ is the renormalization exponent 
201: behind geometrical and fractal properties of O($n$) loops, just as the
202: renormalization exponents $X_t$ and $X_h$ determine the thermodynamic
203: singularities.
204: 
205: In another application of the Coulomb gas technique we can explore
206: corrections to scaling associated \cite{BKN} with the exponent
207: $X_{2l}$ that describes the decay of the probability that two O($n$) 
208: loops collide in two points separated by a distance $r$. The value of
209: this exponent is determined by electric charges as above and magnetic
210: charges $m_0=-m_r= 2$:
211: \begin{equation}
212: X_{2l} = 1-\frac{1}{2g}  + \frac{3g}{2}
213: \label{x2l}
214: \end{equation}
215: This exponent becomes marginal at $n=2$ and is thus expected to
216: lead to poor convergence of finite-size data near $n=2$.
217: 
218: The physical relevance of $X_l$ can be demonstrated by means of scaling
219: arguments. The probability $g_l(r)$ that two points at a distance $r$
220: lie on the same loop is, as given above, $g_l(r)\simeq a r^{-2X_l}$.
221: Let, at criticality, the fractal dimension of the loops be $d_l$.
222: Thus, under a rescaling by a linear factor $b$, the length $l$ of the
223: loop decreases by a factor $b^{d_l}$, and its density increases by a 
224: factor $b^{2-d_l}$. This determines the correlation in the rescaled
225: system as $g_l(r/b)\simeq a b^{4-2d_l} r^{-2X_l}$ which is, as specified
226: above, to be compared with $g_l(r/b)\simeq a b^{2X_l} r^{-2X_l}$.
227: It thus follows that the fractal dimension of loops is 
228: \begin{equation}
229: d_l = 2- X_{l}
230: \label{dl}
231: \end{equation}
232: 
233: Let $P_l(l)$ be the density of loops of length $l$. It is natural that,
234: at criticality, $P_l(l)$ depends algebraically on $l$, with an exponent
235: denoted as $p_l$: 
236: \begin{equation}
237: P_l(l) \propto l^{p_l} \, .
238: \label{ml}
239: \end{equation}
240: Under a rescaling by a linear factor $b$, the loop
241: density is affected for two reasons: first, the loops decrease in length
242: by a factor $b^{d_l}=b^{2- X_{l}}$; and second, the density increases
243: by a factor $b^2$ because the volume is reduced. Consistency requires
244: that $P_l(l)dl=b^{-2} P_l(l b^{2- X_{l}}) d(l b^{2- X_{l}})$ or
245: $P_l(l)=b^{- X_{l}} P_l(l b^{2- X_{l}})$.
246: The requirement $P_l(l) \propto l^{-p_l}$ yields
247: $l^{-p_l} = (l b^{-2+ X_{l}})^{-p_l} b^{-2+ X_{l}-2}$. Matching the exponents
248: shows that
249: \begin{equation}
250: p_l = 1+\frac{2}{2- X_{l}}
251: \label{pl}
252: \end{equation}
253: Since scaling implies that the divergence of the expectation value of the
254: linear size of the largest loop goes as $(x_c-x)^{-\nu}=(x_c-x)^{-1/y_t}$
255: when the critical point is approached, and the actual length $l_{\rm max}(x)$ 
256: as expressed in lattice edges behaves as a power $2-X_l$ of the linear
257: size, it follows that the largest loop length diverges as
258: \begin{equation}
259: l_{\rm max}(x) \propto  (x_c-x)^{(X_l-2)/y_t}
260: \label{lm}
261: \end{equation}
262: 
263: For $L \to \infty$ and $x>x_c$, there exists an infinite spanning loop.
264: Under a rescaling by
265: a linear scale factor $b$ its density increases by a factor $b^{X_l}$
266: while, as usual, the temperature field $t\propto x-x_c$ scales as
267: $t \to t' = b^{y_t} t$. The fraction $s_l(x-x_c)$ of the edges covered
268: by the spanning loop scales as $s_l(b^{y_t}(x_c-x))=b^{X_l} s_l((x_c-x)$.
269: The choice $b=(x_c-x)^{-1/y_t}$ leads to a constant on the left hand
270: side of this equation, and after substitution in the right hand side,
271: the scaling behavior follows as
272: \begin{equation}
273: s_l(x_c-x)\propto (x_c-x)^{X_l/y_t} 
274: \label{lm1}
275: \end{equation}
276: The finite-size dependence of the similar fraction $s_L$ of a system 
277: with finite size $L$ at criticality can simply be found by rescaling
278: the system to a given size, say 1. This leads to
279: \begin{equation}
280: s_L  \propto L^{-X_l}
281: \label{ss}
282: \end{equation} 
283: Including a correction to scaling, we may modify this into
284: \begin{equation}
285: s_L =a L^{-X_l}(1+b L ^{y_i}+\cdots)
286: \label{ssc}
287: \end{equation}
288: where $y_i=2-X_{2l}$ is a candidate for the leading correction
289: exponent \cite{BKN}, and $a$ and $b$ are unknown amplitudes.
290: 
291: In analogy with magnetic systems, a susceptibility-like quantity $\chi_l$
292: can be defined on the basis of the distribution of the loop sizes as
293: \begin{equation}
294: \chi_l\equiv \sum_{l=1}^{l_{\rm max}} P_l(l) l^2
295: \label{chid}
296: \end{equation}
297: According to the aforementioned scaling behavior, the largest loop
298: in a critical system of finite size $L$ has a length scale
299: $l_{\rm max}\propto L^{2-X_l}$. Thus
300: \begin{equation}
301: \chi_l(L) \propto \sum_{l=1}^{L^{2-X_l}} l^{2-p_l} 
302: \label{chil}
303: \end{equation}
304: Substitution of $p_l = 1+ 2 /(2- X_{l})$ yields
305: \begin{equation}
306: \chi_l(L) \propto L^{2-2X_l}
307: \label{chis}
308: \end{equation}
309: Again, we can include corrections to scaling 
310: \begin{equation}
311: \chi_l(L) = c L^{2-2X_l}(1 + d L^{y_i} +\cdots)
312: \label{chisc}
313: \end{equation}
314: with $y_i$ as mentioned above, and $c$ and $d$ are unknown constants. 
315: 
316: Another correlation function of interest describes the probability that
317: two sites on the dual triangular lattice separated by a distance $r$
318: sit inside the same loop (i.e., not separated by any loop of the model).
319: The exponent $X_a$ describing the decay of this function at criticality
320: was derived by Duplantier and Saleur \cite{DS} as $X_a=1-g/2-3/(8g)$.
321: The fractal dimension of the interior of O($n$) loops is therefore 
322: $d_a=2-X_a=1+g/2+3/(8g)$. The area inside a loop  does not include
323: the area inside loops enclosed by that loop.  The exponent $X_a$ thus
324: also determines the finite-size scaling of the spanning loop. We are 
325: therefore interested in the fraction $s_a$ of the dual lattice sites 
326: that sit inside the spanning loop.
327: Scaling indicates that this quantity is subject to the following
328: finite-size behavior:
329: \begin{equation}
330: s_a(L) \propto L^{-X_a}
331: \label{sis}
332: \end{equation}
333: We furthermore define
334: another susceptibility-like quantity $\chi_a$ on the basis of the 
335: distribution $P_a(a)$ of the area $a$ (expressed in the number of
336: enclosed sites of the dual lattice) of the interior of the loops as
337: \begin{equation}
338: \chi_a(L) \equiv  L^{-2} \sum_a P_a(a) a^2
339: \label{xis}
340: \end{equation}
341:  which is expected to scale as
342: \begin{equation}
343: \chi_a(L) \propto L^{2-2X_a}
344: \label{chiis}
345: \end{equation}
346: 
347: \section{Monte Carlo Algorithm}
348: \label{MC}
349: In the existing Monte Carlo algorithm for the loop model \cite{Wdyn},
350: the local updates require time-consuming nonlocal operations as 
351: explained above. To get rid of these we adopt the following procedure.
352: 
353: As a first step of such an algorithm for the simulation of the O($n$)
354: loop model on the
355: honeycomb lattice, we represent the loop configuration by means of 
356: Ising spins on the dual lattice, which is triangular. The loops are just
357: the interfaces between neighboring spins of a different sign.
358: We restrict ourselves to systems with periodic boundary conditions, so
359: that the interfaces indeed form a system of closed, nonintersecting loops 
360: loops on the honeycomb lattice.
361: This is illustrated in Fig.~\ref{cfg1} using a loop configuration which
362: consists of 2 loops and 16 bonds, shown as bold lines. This graph
363: contributes a weight $x^{16}n^{2}$ to the partition function. The loops
364: can of course be represented by two opposite Ising spin configuration,
365: but this degeneracy has no further consequences for our line of argument.
366: \begin{figure}
367: \centering
368: \includegraphics[scale=0.6]{cfg.eps}
369: \caption{Representation of a loop configuration with Ising spins}
370: \label{cfg1}
371: \end{figure}
372: 
373: We now show how one can update the loop configuration by means of local
374: Metropolis-type updates of the Ising spins representing the loop
375: configuration. It works only for $n \geq 1$.
376: One step of importance sampling is realized by the following operations,
377: which are to be repeated cyclically:
378: \begin{enumerate}
379: \item
380: For each loop, assign its color to be either 'green' with probability
381: $1/n$ or 'red' with the remaining probability $1-1/n$.
382: \item
383: Randomly select an Ising spin on the dual lattice. 
384: \item
385: Check if the spin is adjacent to a red loop segment. If so, do
386: nothing; if not, update the spin using the Metropolis probabilities,
387: with Ising couplings $K$ determined as $e^{-2K}=x$.
388: \item
389: Repeat steps 2 and 3 until the number of update attempts is equal to
390: the number of sites of the dual lattice.
391: \item
392: Perform a sweep through the Ising system to find all loops.
393: \item
394: Repeat steps 1 to 5 a fixed number $n_s-1$ times.
395: \item
396: Sample the data of interest from the loop configuration.
397: \end{enumerate}
398: A Monte Carlo run consists of many of these cycles, each of which thus
399: includes $n_s$ Metropolis sweeps, new random assignments of loop colors,
400: and the data sampling procedure.
401: Each of these steps satisfies the conditions of detailed balance,
402: so that the algorithm should indeed generate configurations in accordance 
403: with Boltzmann statistics. Tests confirm that the simulation results
404: agree with those of the existing algorithm \cite{Wdyn}.
405: Since this algorithm assigns colors to the loops, we refer to it
406: as "coloring algorithm".
407: 
408: \section{Simulation}
409: \label{Sim}
410: With the help of the algorithm described above, we simulated the O($n$)
411: model on the honeycomb lattice at the exactly known \cite{N1} critical
412: points $x_c=(2+(2-n)^{1/2})^{-1/2}$ in order to check the theoretical
413: predictions described in Sec.~\ref{CG}.
414: 
415: We first applied our algorithm to the case $n=1.5$, using $12$ system
416: sizes in the range $8 \leq L \leq 128$,  where $L$ is the linear size
417: of the dual triangular lattice, and periodic boundary conditions.
418: 
419: For each system size, a run was executed with a length of $4\times 10^7$
420: Monte Carlo cycles after equilibration of the system. Each cycle 
421: included $n_s=5$ sweeps and loop formation steps, and one sampling as 
422: described above.  
423: Statistical errors were estimated by means of binning in 2000 partial
424: results.
425: 
426: The data sampling included the density $P_l(l)$ of loops of length $l$.
427: The results for the system of size $L=128$ are shown in Fig.~\ref{ml15}. 
428: It displays a substantial interval of algebraic decay, as described by
429: Eq.~(\ref{ml}).
430: 
431: The fractions $s_l$ and $s_a$, and the susceptibility-like quantities
432: $\chi_l$ and $\chi_a$ were also sampled. The finite-size dependence
433: of these quantities at criticality is shown in Table \ref{tablen15}.
434: 
435: \begin{table}[table1]
436: \caption{numerical data for $s_l(L), \chi_{l}(L)$ and $s_a(L),\chi_a(L)$ for 
437: different system sizes $L$ at the critical point of  O($n=1.5$) model}
438: \label{tablen15}
439: \begin{tabular}{l|lr|lr|lr|lr}
440: \hline
441: $L $     & $s_l$    &    &$\chi_{l}$&   & $s_a$  &     &$\chi_a$&     \\
442: \hline
443:     8    &  0.12907 &(2) &  0.8648&(1)  & 0.81028&(3)  &   2.166&(1)  \\
444:    16    &  0.08611 &(2) &  2.1984&(3)  & 0.76386&(5)  &   8.439&(6)  \\
445:    24    &  0.06785 &(3) &  3.4141&(6)  & 0.73829&(8)  &  18.39 &(2)  \\
446:    32    &  0.05718 &(3) &  4.5578&(8)  & 0.72108&(9)  &  31.66 &(5)  \\
447:    40    &  0.05006 &(4) &  5.6466&(11) & 0.70812&(11) &  48.15 &(9)  \\
448:    48    &  0.04504 &(4) &  6.6993&(14) & 0.69729&(13) &  68.11 &(14) \\
449:    56    &  0.04107 &(4) &  7.7199&(23) & 0.68869&(16) &  90.80 &(24) \\
450:    64    &  0.03793 &(4) &  8.7142&(25) & 0.68130&(15) & 116.3  &(3)  \\
451:    80    &  0.03324 &(4) & 10.635 & (3) & 0.66899&(19) & 176.6  &(6)  \\
452:    96    &  0.02987 &(4) & 12.477 & (6) & 0.65895&(22) & 249.0  &(10) \\
453:   112    &  0.02723 &(5) & 14.273 & (6) & 0.65095&(26) & 330.7  &(14) \\
454:   128    &  0.02513 &(5) & 16.027 & (8) & 0.6441 & (3) & 422.4  &(23) \\
455: \hline
456: \end{tabular}
457: \end{table}
458: 
459: These quantities are well described by power laws as a function of the
460: lattice size $L$  for sufficiently large $L$. This is just as expected 
461: on the basis of finite-size scaling as expressed by Eqs.~(\ref{ss}),
462: (\ref{chis}), (\ref{sis}) and (\ref{chiis}). This behavior is illustrated
463: in Figs.~\ref{s1.5}, \ref{chi1.5}, \ref{si1.5} and \ref{chii1.5}.
464: 
465: Using the nonlinear Levenberg-Marquardt least-squares algorithm, we 
466: fitted for the exponents and amplitudes
467: according to the finite-size-scaling formulas including correction terms,
468: as given in Eqs.~(\ref{ssc}), (\ref{chisc}), (\ref{sis}) and (\ref{chiis}). 
469: Thus we can numerically determine $X_l$, $X_a$, and thereby the
470: fractal dimension $d_l$ of the spanning loop, and the 
471: fractal dimension $d_a$ of interior of the spanning loop.
472: Comparing the residual $\chi^2$ of the fits with the number of degrees
473: of freedom, we find satisfactory fits including all system sizes $L\geq 8$.
474: We obtain $X_l=0.593 \,(2)$ from the fit of $s_l(L)$ 
475: and $X_l=0.595 \,(2)$ from the fit of $\chi(L)$.
476: These  results are consistent with the theoretical value
477: $X_l \approx 0.593513601$.
478: The fit of $s_a(L)$ yields $X_a=0.080 \,(1)$. From the fit of $\chi_a(L)$, 
479: we obtain $2-2 X_a=1.84\,(1)$, or $X_a=0.080\,(5)$. These results agree 
480: well with the theoretical prediction $X_a \approx 0.0801085234$.
481: 
482: In these fits, the correction-to-scaling exponent $y_i$ was left free.
483: The fits suggest that the exponent of the dominant correction to
484: scaling does not assume the expected value $2-X_{2l}=-0.43859$, but
485: instead $y_i=-0.75 \pm 0.10$.
486: On the other hand, we have also performed similar simulations of the O($1$) 
487: model at the critical point. The fractions $s_l$ and $s_a$, and the
488: susceptibility-like quantities $\chi_l, \chi_a$ were sampled for
489: several system sizes. The results are shown in Tables \ref{tablen1}. 
490: Least-squares fit results agree well with the theoretical prediction, 
491: as listed in Table \ref{fitresult}. 
492: For the O(1) model, only the data for $\chi_l$ allowed a reasonably accurate
493: estimate of the leading correction-to-scaling exponent. In this case,
494: we find $y_i=-0.628\,(7)$, in a good agreement with the expected value
495: $2-X_{2l}=-0.625$. We thus have fixed $y_i$ at this value in the other fits.
496: 
497: Furthermore we performed similar simulations of the O($n$) models with
498: $n=\sqrt{2}$, $n=\sqrt{3}$ and $n=2$ at their critical points as given in
499: Ref.~\onlinecite{N1}.  The  fractions $s_l$ and $s_a$, and the
500: susceptibility-like quantities $\chi_l, \chi_a$ were sampled for
501: several system sizes. The results are shown in Tables \ref{tablesqrt2},
502: \ref{tablensqrt3} and \ref{tablen2}.
503: Also these quantities appear to depend algebraically on the system size
504: $L$, in agreement with the finite-size scaling equations
505: (\ref{ss}), (\ref{chis}), (\ref{sis}) and (\ref{chiis}). 
506: 
507: \begin{table}[htbp]
508: \caption{Numerical data for $s_l(L)$, $\chi_{l}(L)$, $s_a(L)$ and
509: $\chi_a(L)$ for several system sizes $L$ at the critical point of the
510: O($n=1$) model.}
511: \label{tablen1}
512: \begin{center}
513: \begin{tabular} {l|lr|lr|lr|lr}
514: \hline
515:     $L $ & $s_l$   &    &$\chi_{l}$&   &    $s_a$  &    & $\chi_a$&       \\
516: \hline
517:     8    &  0.09126&(2) &  0.43532&(8) &   0.88355 &(2) &   1.1563& (6)   \\
518:    16    &  0.06015&(2) &  1.1859 &(2) &   0.85090 &(5) &   4.580 & (5)   \\
519:    24    &  0.04678&(2) &  1.8531 &(3) &   0.83257 &(7) &  10.108 &(16)   \\
520:    32    &  0.03911&(2) &  2.4658 &(4) &   0.81987 &(9) &  17.682 &(39)   \\
521:    40    &  0.03405&(3) &  3.0413 &(6) &   0.81015 &(12)&  27.27  & (8)   \\
522:    48    &  0.03040&(3) &  3.5901 &(8) &   0.80234 &(14)&  38.73  &(13)   \\
523:    56    &  0.02754&(3) &  4.1149 &(12)&   0.79617 &(15)&  51.68  &(19)   \\
524:    64    &  0.02539&(4) &  4.6235 &(14)&   0.79038 &(20)&  66.98  &(32)   \\
525:    80    &  0.02214&(5) &  5.5903 &(24)&   0.78077 &(27)& 104.0   & (7)   \\
526:    96    &  0.01968&(5) &  6.5121 &(33)&   0.77384 &(28)& 145.3   &(11)   \\
527:   112    &  0.01781&(6) &  7.4027 &(42)&   0.76799 &(35)& 193.0   &(18)   \\
528:   128    &  0.01645&(6) &  8.2442&(43) &   0.76213 &(42)& 253.2&(26)    \\
529: \hline
530: \end{tabular}
531: \end{center}
532: \end{table}
533: 
534: 
535: \begin{table}[htbp]
536: \caption{Numerical data for $s_l(L)$, $\chi_{l}(L)$, $s_a(L)$ and 
537: $\chi_a(L)$ for several system sizes $L$ at the critical point of the
538: O($n=\sqrt{2}$) model.}
539: \label{tablesqrt2}
540: \begin{center}
541: \begin{tabular} {l|lr|lr|lr|lr}
542: \hline
543:     $L $ & $s_l$    &   & $\chi_{l}$&  & $s_a$      &   & $\chi_a$& \\
544: \hline
545:     8    &  0.12186 &(2) &   0.7813 &(1) &  0.82429 &(3)&    1.953&(1)    \\
546:    16    &  0.08104 &(2) &   1.9989 &(3) &  0.78004 &(5)&    7.651&(7)    \\
547:    24    &  0.06361 &(3) &   3.1027 &(5) &  0.75586 &(8)&   16.65 &(2)    \\
548:    32    &  0.05358 &(3) &   4.1362 &(7) &  0.7392  &(1)&   28.81 &(5)    \\
549:    40    &  0.04693 &(4) &   5.1183 &(11)&  0.7264  &(1)&   44.12 &(9)\\
550:    48    &  0.04206 &(4) &   6.0630 &(16)&  0.7164  &(2)&   62.14 &(18)    \\
551:    56    &  0.03835 &(5) &   6.9786 &(18)&  0.7081  &(2)&   82.86 &(22)    \\
552:    64    &  0.03536 &(4) &   7.8623 &(22)&  0.7011  &(2)&  106.5  &(3)    \\
553:    80    &  0.03096 &(5) &   9.580  & (4)&  0.6890  &(2)&  162.4  &(6)    \\
554:    96    &  0.02769 &(5) &  11.222  & (5)&  0.6800  &(3)&  227.0  &(12)    \\
555:   112    &  0.02532 &(6) &  12.825  & (7)&  0.6716  &(3)&  304.4  &(18)    \\
556:   128    &  0.02338 &(6) &  14.370  & (9)&  0.6649  &(4)&  390.7  &(25)    \\
557: \hline
558: \end{tabular}
559: \end{center}
560: \end{table}
561: 
562: \begin{table}[htbp]
563: \caption{Numerical data for $s_l(L)$, $\chi_{l}(L)$, $s_a(L)$ and
564: $\chi_a(L)$ of the critical O($n=\sqrt{3}$) model for several system
565: sizes $L$.}
566: \label{tablensqrt3}
567: \begin{center}
568: \begin{tabular} {l|lr|lr|lr|lr}
569: \hline
570:     $L $ & $s_l$    &    &$\chi_{l}$&     & $s_a$    &    & $\chi_a$ &      \\
571: \hline
572:     8    &  0.15197 &(2) &   1.1247 &(2)  &  0.76686 &(3) &    2.891 &(1)   \\
573:    16    &  0.10286 &(2) &   2.8312 &(3)  &  0.71434 &(5) &   11.150 &(6)   \\
574:    24    &  0.08169 &(3) &   4.4245 &(7)  &  0.68606 &(6) &   24.00  &(2)   \\
575:    32    &  0.06929 &(4) &   5.9453 &(12) &  0.66677 &(9) &   41.16  &(5)   \\
576:    40    &  0.06111 &(4) &   7.4130 &(17) &  0.6522  &(1) &   62.44  &(9)   \\
577:    48    &  0.05500 &(4) &   8.8463 &(23) &  0.6408  &(1) &   87.2 1 &(13)  \\
578:    56    &  0.05035 &(4) &  10.241  & (3) &  0.6314  &(1) &  115.7   &(2)   \\
579:    64    &  0.04665 &(4) &  11.611  & (4) &  0.6233  &(1) &  148.0   &(3)   \\
580:    80    &  0.04112 &(5) &  14.272  & (6) &  0.6098  &(2) &  223.7   &(6)   \\
581:    96    &  0.03701 &(4) &  16.885  & (7) &  0.5993  &(2) &  311.0   &(9)   \\
582:   112    &  0.03387 &(4) &  19.407  & (9) &  0.5906  &(2) &  411.4   &(12)  \\
583:   128    &  0.03153 &(5) &  21.865  &(11) &  0.5824  &(2) &  531.2   &(19)  \\
584: \hline
585: \end{tabular}
586: \end{center}
587: \end{table}
588: 
589: \begin{table}[htbp]
590: \caption{Numerical data for $s_l(L)$, $\chi_{l}(L)$, $s_a(L)$ and
591: $\chi_a(L)$ of the critical O($n=2$) model for several system
592: sizes $L$.}
593: \label{tablen2}
594: \begin{center}
595: \begin{tabular} {l|lr|lr|lr|lr}
596: \hline
597:     $L $ & $s_l$  &    & $\chi_{l}$&  & $s_a$   &    &$\chi_a$&       \\
598: \hline
599:     8  & 0.21534  &(3) & 1.7038 &(3)  & 0.66465 &(3) &  5.082 &(1)    \\
600:    16  & 0.15273  &(2) & 4.4449 &(7)  & 0.60441 &(3) & 18.981 &(6)    \\
601:    24  & 0.12479  &(2) & 7.2121 &(11) & 0.57283 &(4) & 39.965 &(16)   \\
602:    32  & 0.10812  &(2) & 9.9856 &(17) & 0.55170 &(4) & 67.330 &(29)   \\
603:    40  & 0.09674  &(2) &12.762  &(2)  & 0.53600 &(5) &101.00  &(5)    \\
604:    48  & 0.08832  &(2) &15.544  &(3)  & 0.52358 &(5) &139.39  &(7)    \\
605:    56  & 0.08177  &(3) &18.328  &(4)  & 0.51336 &(6) &183.47  &(11)    \\
606:    64  & 0.07650  &(2) &21.106  &(4)  & 0.50468 &(5) &232.70  &(14)    \\
607:    80  & 0.06844  &(3) &26.676  &(7)  & 0.49057 &(7) &345.66  &(30)    \\
608:    96  & 0.06243  &(2) &32.251  &(9)  & 0.47944 &(6) &476.81  &(37)    \\
609:   112  & 0.05783  &(3) &37.828  &(11) & 0.47013 &(7) &626.4   &(5)    \\
610:   128  & 0.05410  &(3) &43.392  &(13) & 0.46223 &(7) &793.1   &(7)    \\
611:   192  & 0.04418  &(3) &65.680  &(26) & 0.43916 &(10)&1620    &(2)    \\
612: 
613: \hline
614: \end{tabular}
615: \end{center}
616: \end{table}
617: 
618: Using finite size scaling and the Levenberg-Marquardt algorithm, we 
619: determined the exponents $X_l$ and $X_a$. The results
620: are summarized in Table \ref{fitresult}.
621: 
622: \begin{table}[htbp]
623: \caption{Results for the exponents $X_l$ and $X_a$ for five O($n$)
624: models, as obtained from least-squares fits as described in the text.}
625: \label{fitresult}
626: \begin{center}
627: \begin{tabular} {l|lr|l|lr|lr|l|lr}
628: \hline
629: $n$&$X_l$       &  &$X_l$   &$2-2X_l$& &$X_a$    &   &$X_a$   & $2-2X_a$ &\\
630: &from $s_l$  &  &theory&from $\chi_l$& &from $s_a$&&theory &from $\chi_a$&\\
631: \hline
632: $1$      &0.625 &(1) & 5/8   & 0.73  &(2)&0.0518 &(3)&0.05208 &1.89 &(1)\\
633: $\sqrt2$ &0.599 &(1) & 3/5   & 0.796 &(5)&0.075  &(1)& 3/40   &1.86 &(1)\\
634: $1.5   $ &0.595 &(3) &0.5935 & 0.810 &(4)&0.080  &(1)&0.08011 &1.84 &(1)\\
635: $\sqrt3$ &0.571 &(1) &0.5714 & 0.856 &(1)&0.095  &(1)&0.0952  &1.81 &(1)\\
636: $2$      &0.4997&(3) & 1/2   & 0.996 &(5)&0.1249 &(4)&1/8     &1.747&(2)\\ 
637: \hline
638: \end{tabular}
639: \end{center}
640: \end{table}
641: \section{Discussion}
642: \label{Dis}
643: The time-consuming character of simulations of loop models is due to the
644: nonlocal character of the loops. We have reduced this problem by splitting 
645: the loop weight $n$ in two parts, namely $n-1$ and $1$. Proper summation
646: over both contributions is done by assigning color variables to the
647: loops; a sum on all color variables is included in the partition sum.
648: The algorithm treats these color variables as dynamical variables which
649: are updated by the Monte Carlo process.
650: The idea to use an additional color variable for each loop
651: was already used in a context unrelated to Monte Carlo methods,
652: e.g. in Ref.~\onlinecite{sq}.  The presence of loops of weight 1
653: in such a configuration then leads, at least locally, to a system
654: that behaves precisely as an Ising configuration. Thus, the system may
655: be locally updated by means of Metropolis-like Monte Carlo steps.
656: Care should be taken to leave the loops of weight $n-1$ unchanged,
657: because it would violate the condition of detailed balance.
658: The procedure given in Sec.~\ref{MC} satisfies this constraint.
659: 
660: The development of this coloring algorithm was motivated by the
661: possibility to further explore the physics of O($n$) models.
662: In the course of this work, we realized that it should be possible to
663: construct an even more efficient algorithm of a `cluster' nature.
664: Algorithms of this type will be presented elsewhere \cite{DGBca}.
665: The `coloring' trick is only useful for $n>1$. For $0<n<1$ the
666: existing algorithm \cite{Wdyn} is, although relatively inefficient,
667: still applicable.
668: 
669: For the interpretation of the simulation results, it is relevant that
670: we are using periodic boundary conditions, and that the mapping between
671: loop and Ising configurations imposes a condition of `evenness' on the 
672: loop configurations: a path spanning the periodic boundaries must 
673: have an even number of intersections with a loop. Therefore, the
674: statistical ensemble generated by the algorithm does not coincide
675: with that of Eq.~(\ref{graphpartition}). The difference is related
676: to the boundary conditions and is expected to modify the finite-size
677: behavior, but should vanish in the thermodynamic limit.
678: 
679: The present work is restricted to the `even' loop configurations.
680: It is, however, possible to simulate `odd' loop configurations by 
681: introducing a `seam' on the dual lattice, a vertical column of
682: horizontal antiferromagnetic Ising bonds spanning the system.
683: For these antiferromagnetic bonds we use the rule that there is a loop
684: segment if and only if the two associated dual spins are of the same
685: sign. Horizontal and vertical seams can be introduced independently, as
686: prescribed by the class of loop configurations that is to be sampled.
687: 
688: \acknowledgments
689: We wish to thank Prof. B. Nienhuis for many valuable discussions.
690: This work was supported by by the National Science Foundation of China
691: under Grant \#10105001, and by the Lorentz Fund (Leiden University).
692: 
693: \newpage
694: 
695: \begin{figure}
696: \centering
697: \includegraphics{m1.5_128.eps}
698: \caption{Distribution $P_l$ of loops of length $l$ on logarithmic
699: scales, for the critical O($n$) model with $n=1.5$ and size $L=128$.
700: The dashed line shows a power law decay with exponent -2.42198318,
701: which is the theoretical asymptotic value for the infinite system. }  
702: \label{ml15}
703: \end{figure}
704: 
705: \begin{figure}
706: \centering
707: \includegraphics{s1.5.eps}
708: \caption{Fraction $s_l$ of lattice edges covered by the spanning loop,
709: versus system size $L$ for the critical O($n$) model with $n=1.5$,
710: on logarithmic scales. The curve is added as a guide to the eye, and 
711: estimated error bars are smaller than the size of the symbols.}
712: \label{s1.5}
713: \end{figure}
714: 
715: \begin{figure}
716: \centering
717: \includegraphics{chi1.5.eps}
718: \caption{Susceptibility-like quantity $\chi_l$ versus system size $L$
719: for the critical O($n$) model with $n=1.5$ on logarithmic scales. The
720: curve is added as a guide to the eye, and estimated error bars
721: are smaller than the size of the symbols.}
722: 
723: \label{chi1.5}
724: \end{figure}
725: 
726: \begin{figure}
727: \centering
728: \includegraphics{si1.5.eps}
729: \caption{Fraction $s_a$  of the number of dual lattice sites
730: inside the spanning loop, versus system size $L$ for the critical
731: O($n$) model with $n=1.5$, on logarithmic scales. The curve is added
732: as a guide to the eye, and estimated error bars are smaller than the
733: size of the symbols.}
734: \label{si1.5}
735: \end{figure}
736: 
737: \begin{figure}
738: \centering
739: \includegraphics{chii1.5.eps}
740: \caption{Susceptibility-like quantity $\chi_a$ versus system size $L$
741: for the critical O($n$) model with $n=1.5$ on logarithmic scales. The
742: curve is added as a guide to the eye, and estimated error bars
743: are smaller than the size of the symbols.}
744: \label{chii1.5}
745: \end{figure}
746: 
747: 
748: \begin{thebibliography}{reference}
749: \bibitem{N1}
750: B. Nienhuis, Phys. Rev. Lett. {\bf 49}, 1062 (1982); 
751: J. Stat. Phys. {\bf 34}, 731 (1984).
752: \bibitem{DMNS}
753: E. Domany, D. Mukamel, B. Nienhuis and A. Schwimmer, Nucl. Phys. B {\bf 190},
754: 279 (1981).
755: \bibitem{Fisher}
756: M.E. Fisher, Physics (N.Y.) {\bf 3}, 25 (1967).
757: \bibitem{Wu}
758: F.Y. Wu, Rev. Mod. Phys. {\bf 54}, 235 (1982).
759: \bibitem{KF}
760: P.W. Kasteleyn and C.M. Fortuin , J. Phys. Soc. Jpn. {\bf 46} (Suppl.), 
761: 11 (1969); C.M. Fortuin and P.W. Kasteleyn, Physica (Amsterdam) {\bf 57}, 
762: 536 (1972).
763: \bibitem{CP}
764: A. Coniglio and F. Peruggi, J. Phys. A {\bf 15}, 1873 (1982).
765: \bibitem{Con}
766: A. Coniglio, Phys. Rev. Lett. {\bf 62}, 3054 (1989).
767: \bibitem{BKN}
768: H.W.J. Bl\"ote, Y.M.M. Knops, and B. Nienhuis,
769: Phys. Rev. Lett. {\bf 23}, 3440 (1992).
770: \bibitem{DBN}
771: Y. Deng, H.W.J. Bl\"ote and B. Nienhuis, Phys. Rev. E {\bf 69}, 026123 (2004).
772: \bibitem{V}
773: C. Vanderzande, J. Phys. A {\bf 25}, L75(1992).
774: \bibitem{JS1}
775: W. Janke and A.M.J. Schakel, Nucl. Phys. B {\bf 700}, 385 (2004).
776: \bibitem{SD}
777: H. Saleur and B. Duplantier, Phys. Rev. Lett. {\bf 58}, 2325 (1987).
778: \bibitem{DS}
779: B. Duplantier and H. Saleur, Phys. Rev. Lett. {\bf 63}, 2536 (1989).
780: \bibitem{Wdyn}
781: W.A. Guo and H.W.J. Bl\"ote, unpublished (2001).
782: \bibitem{Kad}
783: L.P. Kadanoff, J. Phys. A {\bf 11}, 1399 (1978).
784: \bibitem{N}
785: B. Nienhuis, in {\it Phase Transitions and Critical Phenomena}, Vol. {\bf 11},
786: edited by C. Domb and J.L. Lebowitz (Academic Press, London, 1987).
787: \bibitem{dN}
788: M.P.M. den Nijs, J. Phys. A {\bf 17}, L295 (1984).
789: \bibitem{sq}
790: H.W.J. Bl\"ote and B. Nienhuis, J. Phys. A {\bf 22}, 1415 (1989).
791: \bibitem{DGBca}
792: Y. Deng, W.A. Guo, T. Garoni, A.D. Sokal and H.W.J. Bl\"ote, to be
793: published (2006).
794: \end{thebibliography}
795: \end{document}
796: