summaryrefslogtreecommitdiff
path: root/monte-carlo.tex
blob: cf45bcd0205c5bd8ab7e5ba5ccdd9ec69956f3ee (plain)
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701

%  Created by Jaron Kent-Dobias on Thu Apr 20 12:50:56 EDT 2017.
%  Copyright (c) 2017 Jaron Kent-Dobias. All rights reserved.

\documentclass[aps,pre,reprint]{revtex4-1}

\usepackage{amsmath,amssymb,latexsym,mathtools}

% uncomment to label only equations that are referenced in the text
%\mathtoolsset{showonlyrefs=true}

% I want equation labels but don't want to type out `equation'
\def\[{\begin{equation}}
\def\]{\end{equation}}

% additional math
\def\tr{\mathop{\mathrm{Tr}}\nolimits}    % trace
\def\sgn{\mathop{\mathrm{sgn}}\nolimits}  % sign function
\def\dd{d}                                % differential
\def\Z{\mathbb Z}                         % integers
\def\R{\mathbb R}                         % reals

% physics conventions
\def\c{\mathrm c}               % critical label as in T_c
\def\H{\mathcal H}              % Hamiltonian
\def\J{Z}                       % site-site coupling
\def\B{B}                       % external field
\newcommand\set[1]{\mathbf #1}  % collection of N spins
\newcommand\avg[1]{\langle#1\rangle} % ensemble average

% dimensions
\def\dim{d}
\def\twodee{\textsc{2\dim} }
\def\threedee{\textsc{3\dim} }
\def\fourdee{\textsc{4\dim} }

% fancy partial derivative
\newcommand\pd[3][]{
  \ifthenelse{\isempty{#1}}
    {\def\tmp{}}
    {\def\tmp{^#1}}
  \frac{\partial\tmp#2}{\partial#3\tmp}
}

\begin{document}

\title{Cluster representations and the Wolff algorithm in arbitrary external fields}
\author{Jaron Kent-Dobias}
\author{James P.~Sethna}
\affiliation{Laboratory of Atomic and Solid State Physics, Cornell University, Ithaca, NY, USA}

\date\today

\begin{abstract}
  We introduce a natural way to extend celebrated spin-cluster Monte Carlo
  algorithms for fast thermal lattice simulations at criticality, like Wolff, to
  systems in arbitrary fields. The method relies on the generalization of the
  `ghost spin' representation to one with a `ghost transformation' that
  restores invariance to spin symmetries at the cost of an extra degree of
  freedom. The ordinary cluster-building process can then be run on the new
  representation. For several canonical systems, we show that this extension
  preserves the scaling of accelerated dynamics in the absence of a field.
\end{abstract}

\maketitle

Lattice models are important in the study of statistical physics and phase
transitions. Rarely exactly solvable, they are typically studied by
approximate and numerical methods. Monte Carlo techniques are a common way of
doing this, approximating thermodynamic quantities by sampling the
distribution of systems states. These Monte Carlo algorithms are better the
faster they arrive at a statistically independent sample. This typically
becomes a problem near critical points, where critical slowing down
\cite{wolff_critical_1990} results in power-law divergences of dynamic
timescales.

Celebrated cluster algorithms largely addressed this in the absence of
symmetry-breaking fields by using nonlocal updates \cite{janke_nonlocal_1998}
whose clusters undergo a percolation transition at the critical point of the
system \cite{coniglio_clusters_1980}.  These result in relatively small
dynamic exponents for many spin systems \cite{wolff_comparison_1989,
du_dynamic_2006, liu_dynamic_2014, wang_cluster_1990}, including the Ising,
$\mathrm O(n)$ \cite{wolff_collective_1989}, and Potts
\cite{swendsen_nonuniversal_1987, baillie_comparison_1991} models. These
algorithms rely on the natural symmetry of the systems in question under
symmetry operations on the spins.

Some success has been made in extending these algorithms to systems in certain
external fields by adding a `ghost site' \cite{coniglio_exact_1989} that
returns global rotation invariance to spin Hamiltonians at the cost of an
extra degree of freedom, allowing the method to be used in a subcategory of
interesting fields \cite{alexandrowicz_swendsen-wang_1989, wang_clusters_1989,
ray_metastability_1990}. Static fields have also been applied by including a
separate metropolis or heat bath update step after cluster formation
\cite{destri_swendsen-wang_1992, lauwers_critical_1989,
ala-nissila_numerical_1994}, and other categories of fields have been applied
using replica methods \cite{redner_graphical_1998, chayes_graphical_1998,
machta_replica-exchange_2000}.  Monte Carlo techniques that involve cluster
updates at fixed magnetization have been used to examine quantities at fixed
field by integrating the associated thermodynamic functions
\cite{martin-mayor_cluster_2009, martin-mayor_tethered_2011}.

We show that the scaling of correlation time near the critical point of
several models suggests that the `ghost' approach is a natural one, e.g., that
it extends the celebrated scaling of dynamics in these algorithms at zero
field to various non-symmetric perturbations. We also show, by a redefinition
of the spin--spin coupling in a generic class of spin systems,
\emph{arbitrary} external fields can be treated using cluster methods. Rather
than the introduction of a `ghost spin,\!' our representation relies on
introducing a `ghost transformation.\!'

\section{Clusters Without a Field}

We will pose the problem in a general way, but several specific examples can
be found in Table~\ref{table:models} for concreteness. Let $G=(V,E)$ be a graph, where the set of vertices $V=\{1,\ldots,N\}$
enumerates the sites of a lattice and the set of edges $E$ contains pairs of
neighboring sites. Let $R$ be a group acting on a set $X$, with the action of
group elements $r\in R$ on elements $s\in X$ denoted $r\cdot s$. $X$ is the
set of states accessible by a spin, and $R$ is the \emph{symmetry group} of
$X$. The set $X$ must admit a measure $\mu$ that is invariant under the action
of $R$, e.g., for any $A\subseteq X$ and $r\in R$, $\mu(r\cdot A)=\mu(A)$.
This trait is shared by the counting measure on any discrete set, or by any
group acting by isometries on a Riemannian manifold, such as $\mathrm O(n)$ on
$S^{n-1}$ in the $\mathrm O(n)$ model \cite{caracciolo_wolff-type_1993}.
Finally, a subset $R_2$ of elements in $R$ of order two must act transitively on
$X$. This property, while apparently obscure, is shared by any symmetric space
\cite{loos_symmetric_1969} or by any transitive, finitely generated isometry
group. In fact, all the examples listed here have spins spaces with natural
metrics whose symmetry group is their set of isometries.  We put one spin at
each site of the lattice described by $G$, so that the state of the entire
spin system is described by elements $\set s\in X\times\cdots\times X=X^N$. 

The Hamiltonian of this system is a function $\H:X^N\to\R$ defined by
\[
  \H(\set s)=-\!\!\!\!\sum_{\{i,j\}\in E}\!\!\!\!\J(s_i,s_j)-\sum_{i\in V}\B(s_i),
\]
where $\J:X\times X\to\R$ couples adjacent spins and $B:X\to\R$ is an external
field. $\J$ must be symmetric in its arguments and invariant under the action
of any element of $R$ applied to the entire lattice, that is, for any $r\in R$
and $s,t\in X$, $\J(r\cdot s,r\cdot t)=\J(s,t)$.  One may also allow $\J$ to also
be a function of edge---for modelling random-bond, long-range, or anisotropic
interactions---or allow $B$ to be a function of site---for applying arbitrary
boundary conditions or modelling random fields. The formal results of this
paper (that the algorithm obeys detailed balance and ergodicity) hold equally
well for these cases, but we will drop the additional index notation for
clarity. Statements about efficiency may not.

\begin{table*}[htpb]
  \begin{tabular}{l||ccccc}
          & Spins ($X$) & Symmetry ($R$)  &  Action ($g\cdot s$)  &
    Coupling ($\J(s,t)$) & Common Field ($B(s)$) \\
    \hline\hline
    Ising         & $\{-1,1\}$    & $\Z/2\Z$                 &  $0\cdot
    s\mapsto s$, $1\cdot s\mapsto -s$ & $st$ & $Hs$ \\
    $\mathrm O(n)$ & $S^{n-1}$ & $\mathrm O(n)$ & $M\cdot s\mapsto Ms$ & $s^{\mathrm T}t$ & $H^{\mathrm T}s$\\
    Potts & $\{1,\ldots,q\}$ & $\mathrm S_n$ & $(i_1,\ldots,i_q)\cdot s=i_s$ & $\delta(s,t)$ & $\sum_mH_m\delta(m,s)$\\
    Clock & $\Z/q\Z$ & $D_n$ & $r_m\cdot s=m+s$, $s_m\cdot
    s=-m-s$ & $\cos(2\pi\frac{s-t}q)$ & $\sum_mH_m\cos(2\pi\frac{s-m}q)$\\
    \textsc{Dgm} & $\Z$ & $D_{\mathrm{inf}}$ & $r_m\cdot s=m+s$, $s_m\cdot s=-m-s$
    & $(s-t)^2$ & $Hs^2$\\
  \end{tabular}
  \caption{Several examples of spin systems and the symmetry groups that act
    on them. Common choices for the spin--spin coupling in these systems and
    their external fields are also given. Other fields are possible, of course:
    for instance, some are interested in modulated fields $H\cos(2\pi k\theta(s))$ for
    integer $k$ and $\theta(s)$ giving the angle of $s$ to some axis applied
    to the $\mathrm O(2)$ model \cite{jose_renormalization_1977}.}
  \label{table:models}
\end{table*}

The goal of statistical mechanics is to compute expectation values of
observables $A:X^N\to\R$. Assuming the ergodic hypothesis holds (for systems
with broken-symmetry states, it does not), the expected value $\avg A$ of an
observable $A$ is its average over every state $\set s$ in the configuration
space $X^N$ weighted by the Boltzmann probability of that state appearing, or
\[
\avg A
  =\frac{\int_{X^N}A(\set s)e^{-\beta\H(\set s)}\,\dd\mu(\set s)}
  {\int_{X^N}e^{-\beta\H(\set s)}\,\dd\mu(\set s)},
\]
where for $Y_1\times\cdots\times Y_N=Y\subseteq X^N$ the product measure
$\mu(Y)=\mu(Y_1)\cdots\mu(Y_N)$ is the simple extension of the measure on $X$
to a measure on $X^N$. These values are estimated using Monte Carlo techniques
by constructing a finite sequence of states $\{\set{s_1},\ldots,\set{s_M}\}$
such that
\[
  \avg A\simeq\frac1M\sum_{i=1}^MA(\set{s_i}).
\]
Sufficient conditions for this average to converge to $\avg A$ as $M\to\infty$
are that the process that selects $\set{s_{i+1}}$ given the previous states be
Markovian (only depends on $\set{s_i}$), ergodic (any state can be accessed),
and obey detailed balance (the ratio of probabilities that $\set{s'}$ follows
$\set s$ and vice versa is equal to the ratio of weights for $\set s$ and
$\set{s'}$ in the ensemble).

While any of several related cluster algorithms can be described for this
system, we will focus on the Wolff algorithm \cite{wolff_collective_1989}. In
the absence of an external field, e.g., $B(s)=0$, the Wolff algorithm proceeds
in the following way.
\begin{enumerate}
  \item Pick a random site $m_0$ and add it to the stack.

  \item Select a rotation $r\in R_2$ distributed by $f(r\mid m_0,\set s)$.
    Often $f$ is taken as uniform on $R_2$, but it is sufficient for preserving
    detailed balance that $f$ be any function of the seed site $m_0$ and
    $Z(s,r\cdot s)$ for all $s\in\set s$. The flexibility offered by the
    choice of distribution will be useful in situations where the state space
    is infinite.
  \item While the stack isn't empty,
    \begin{enumerate}
      \item pop site $m$ from the stack.
      \item If site $m$ isn't marked, 
        \begin{enumerate}
          \item mark the site.
          \item For every $j$ such that $\{m,j\}\in E$, add site $j$ to the
            stack with probability
            \[
              p_r(s_m,s_j)=\min\{0,1-e^{\beta(\J(r\cdot s_m,s_j)-\J(s_m,s_j))}\}.
              \label{eq:bond_probability}
            \]
          \item Take $s_m\mapsto r\cdot s_m$.
      \end{enumerate}
    \end{enumerate}
\end{enumerate}
When the stack is exhausted, a cluster of connected spins will have been
rotated by the action of $r$. In order for this algorithm to be useful, it
must satisfy ergodicity and detailed balance. Ergodicity is satisfied since we
have ensured that the $R_2$ acts transitively on $X$, e.g., for any $s,t\in X$
there exists $r\in R_2$ such that $r\cdot s=t$. Since there is a nonzero
probability that only one spin is rotated and that spin can be rotated into
any state, ergodicity follows. The probability $P(\set s\to\set{s'})$ that the
configuration $\set s$ is brought to $\set s'$ by the flipping of a cluster
formed by accepting rotations of spins via bonds $C\subseteq E$ and rejecting
rotations via bonds $\partial C\subset E$ is related to the probability of the
reverse process $P(\set{s'}\to\set s)$ by
\begin{widetext}
\[
  \begin{aligned}
    \frac{P(\set s\to\set{s'})}{P(\set{s'}\to\set s)}
    &=\frac{f(r\mid m_0,\set s)}{f(r^{-1}\mid m_0,\set s')}
      \prod_{\{i,j\}\in }\frac{p_r(s_i,s_j)}{p_{r^{-1}}(s_i',s_j')}
      \prod_{\{i,j\}\in\partial C}\frac{1-p_r(s_i,s_j)}{1-p_{r^{-1}}(s'_i,s'_j)}
    =\!\!\!\prod_{\{i,j\}\in\partial C}e^{\beta(\J(s_i',s_j')-\J(s_i,s_j))}
    =\frac{e^{-\beta\H(\set{s'})}}{e^{-\beta\H(\set s)}},
  \end{aligned}
\]
\end{widetext}
whence detailed balance is also satisfied, using $r=r^{-1}$ and $Z(r\cdot
s',s')=Z(r\cdot s,s)$.

The Wolff algorithm is well known to be efficient in sampling many spin models
near and away from criticality, including the Ising, Potts, and $\mathrm O(n)$
models. In general, its efficiently will depend on the system at hand, e.g.,
the structure of the configurations $X$ and group $R$. A detailed discussion
of this dependence for a class of configuration spaces with continuous
symmetry groups can be found in \cite{caracciolo_generalized_1991,
caracciolo_wolff-type_1993}.

\section{Adding the field}

This algorithm relies on the fact that the coupling $\J$ depends only on
relative orientation of the spins---global reorientations do not affect the
Hamiltonian. The external field $B$ breaks this symmetry. Fortunately it can be
restored. Define a new graph $\tilde G=(\tilde V,\tilde E)$, where $\tilde
V=\{0,1,\ldots,N\}$ adds the new `ghost' site $0$ which is connected by
\[
  \tilde E=E\cup\big\{\{0,i\}\mid i\in V\big\}
\]
to all other sites.  Instead of assigning the ghost site a spin whose value
comes from $X$, we assign it values in the symmetry group $s_0\in R$, so that
the configuration space of the new model is $R\times X^N$. We introduce the
Hamiltonian $\tilde\H:R\times X^N\to\R$ defined by
\[
\begin{aligned}
  \tilde\H(s_0,\set s)
  &=-\!\!\!\!\sum_{\{i,j\}\in E}\!\!\!\!\J(s_i,s_j)
  -\sum_{i\in V}B(s_0^{-1}\cdot s_i)\\
  &=-\!\!\!\!\sum_{\{i,j\}\in\tilde E}\!\!\!\!\tilde\J(s_i,s_j),
\end{aligned}
\]
where the new coupling $\tilde\J:(R\cup X)\times(R\cup X)\to\R$ is defined for
$s,t\in R\cup X$ by
\[
  \tilde\J(s,t) =
  \begin{cases}
    \J(s,t) & \text{if $s,t\in X$} \\
    B(s^{-1}\cdot t) & \text{if $s\in R$} \\
    B(t^{-1}\cdot s) & \text{if $t\in R$}.
  \end{cases}
  \label{eq:new.z}
\]
The modified coupling is invariant under the action of group elements: for any
$r,s_0\in R$ and $s\in X$,
\[
\begin{aligned}
  \tilde\J(rs_0,r\cdot s)
  &=B((rs_0)^{-1}\cdot (r\cdot s))\\
  &=B(s_0^{-1}\cdot s)
  =\tilde\J(s_0,s)
\end{aligned}
\]
The invariance of $\tilde\J$ to rotations given other arguments follows from
the invariance properties of $\J$.

We have produced a system incorporating the field function $B$ whose
Hamiltonian is invariant under global rotations, but how does it relate to our
old system, whose properties we actually want to measure? If $A:X^N\to\R$ is
an observable of the original system, we construct an observable $\tilde
A:R\times X^N\to\R$ of the new system defined by
\[
  \tilde A(s_0,\set s)=A(s_0^{-1}\cdot\set s)
\]
whose expectation value in the new system equals that of the original
observable in the old system. First, note that $\tilde\H(1,\set s)=\H(\set
s)$. Since the Hamiltonian is invariant under global rotations, it follows
that for any $g\in R$, $\tilde\H(g,g\cdot\set s)=\H(\set s)$.  Using the
invariance properties of the measure on $X$ and introducing a measure $\rho$
on $R$, it follows that
\[
\begin{aligned}
  \avg{\tilde A}
  &=\frac{
    \int_R\int_{X^N}\tilde A(s_0,\set
    s)e^{-\beta\tilde\H(s_0,\set s)}\,\dd\mu(\set s)\,\dd\rho(s_0)
  } {
    \int_R\int_{X^N}e^{-\beta\tilde\H(s_0,\set s)}\,\dd\mu(\set s)\,\dd\rho(s_0)
  }\\
  &=\frac{
    \int_R\int_{X^N}A(s_0^{-1}\cdot\set
    s)e^{-\beta\tilde\H(s_0,\set s)}\,\dd\mu(\set s)\,\dd\rho(s_0)
  } {
    \int_R\int_{X^N}e^{-\beta\tilde\H(s_0,\set s)}\,\dd\mu(\set s)\,\dd\rho(s_0)
  }\\
  &=\frac{
    \int_R\int_{X^N}A(\set{s'})e^{-\beta\tilde\H(s_0,s_0\cdot\set{s'})}\dd\mu(s_0\cdot\set{s'})\,\dd\rho(s_0)
  } {
    \int_R\int_{X^N}e^{-\beta\tilde\H(s_0,s_0\cdot\set{s'})}\dd\mu(s_0\cdot\set{s'})\,\dd\rho(s_0)
  }\\
  &=\frac{
  \int_R\dd\rho(s_0)}{
  \int_R\dd\rho(s_0)}\frac{\int_{X^N}A(\set{s'})e^{-\beta\H(\set{s'})}\dd\mu(\set{s'})
  }{\int_{X^N}e^{-\beta\H(\set{s'})}\dd\mu(\set{s'})
  }
  =\avg A.
\end{aligned}
\]
Using this equivalence, spin systems in a field may be treated in the
following way.
\begin{enumerate}
  \item Add a site to your lattice adjacent to every other site.
  \item Initialize a `spin' at that site whose value is a representation of a
    member of the symmetry group of your ordinary spins.
  \item Carry out the ordinary Wolff cluster-flip procedure on this new
    lattice, substituting $\tilde\J$ as defined in \eqref{eq:new.z} for $\J$.
\end{enumerate}
Ensemble averages of observables $A$ can then be estimated by sampling the
value of $\tilde A$ on the new system. In contrast with the simpler ghost spin
representation, this form of the Hamiltonian might be considered the `ghost
transformation' representation.

\section{Examples}
\label{sec:examples}

Several specific examples from Table~\ref{table:models} are described in the
following.

\subsection{The Ising model}

In the Ising model spins are drawn from the set $\{1,-1\}$. Its symmetry group
is $C_2$, the cyclic group on two elements, which can be conveniently
represented by a multiplicative group with elements $\{1,-1\}$, exactly the
same as the spins themselves. The only nontrivial element is of order two, and
is selected every time in the algorithm.  Since the symmetry group and the
spins are described by the same elements, performing the algorithm on the
Ising model in a field is fully described by just using the `ghost spin'
representation.  This algorithm or algorithms based on the same decomposition
of the Hamiltonian have been applied by several researchers
\cite{alexandrowicz_swendsen-wang_1989, wang_clusters_1989,
ray_metastability_1990}. The algorithm has been implemented by one of the
authors in an existing interactive Ising simulator at
\texttt{https://mattbierbaum.github.io/ising.js}
\cite{bierbaum_ising.js_nodate}.

\subsection{The XY and other $\mathrm O(n)$ models}
\label{sec:examples:on}

In the $\mathrm O(n)$ model spins are described by vectors on the
$(n-1)$-sphere $S^{n-1}$. Its symmetry group is $O(n)$, $n\times n$ orthogonal
matrices, which act on the spins by matrix multiplication. The elements of
$O(n)$ of order two are reflections about hyperplanes through the origin and
$\pi$ rotations about any axis through the origin. Since the former generate
the entire group, reflections alone suffice to provide ergodicity. Sampling
those reflections uniformly works well at criticality. The `ghost spin'
version of the algorithm has been used to apply a simple vector field to the
$\mathrm O(3)$ model \cite{dimitrovic_finite-size_1991}. Other fields of
interest include $(n+1)$-dimensional spherical harmonics
\cite{jose_renormalization_1977} and cubic fields \cite{bruce_coupled_1975,
blankschtein_fluctuation-induced_1982}, which can be applied with the new
method. The method is quickly generalized to spins whose symmetry groups are
other compact Lie groups \cite{caracciolo_generalized_1991,
caracciolo_wolff-type_1993}.

At low temperature or high external vector field field selecting reflections
uniformly becomes inefficient because the excitations of the model are spin
waves, in which the magnetization only differs by a small amount between
neighboring spins. Under these conditions, most choices of reflection plane
will cause a change in energy so great that the whole system is always
flipped, resulting in many correlated samples. To ameliorate this, one can
draw reflections from a distribution that depends on how the seed spin is
transformed. We implement this in the following way. Say that the seed of the
cluster is $s$. Generate a vector $t$ taken uniformly from the space of unit
vectors orthogonal to $s$. Let the plane of reflection be that whose normal is
$n=s+\zeta t$, where $\zeta$ is drawn from a normal distribution of mean zero
and variance $\sigma$. It follows that the tangent of the angle between $s$
and the plane of reflection is also distributed normally with zero mean and
variance $\sigma$. Since the distribution of reflection planes only depends on
the angle between $s$ and the plane, and since that angle is invariant under
the reflection, this choice preserves detailed balance.

The choice of $\sigma$ can be inspired by mean field theory. At high field or
low temperature, spins are likely to both align with the field and each other
and the model is asymptotically equal to a simple Gaussian one, with in the
limit of large $L$ the expected square angle between neighbors being
\[
  \avg{\theta^2}\simeq\frac{(n-1)T}{D+H/2}.
\]
We take $\sigma=\sqrt{\avg{\theta^2}}/2$. Fig.~\ref{fig:generator_times} shows
the effect of making such a choice on autocorrelation times for a critical
\threedee \textsc{xy} ($\mathrm O(2)$) model. At small fields both methods
perform the same as zero field Wolff.  Intermediate field values see
efficiency gains for both methods. At large field the uniform sampling method
sees correlation times grow rapidly, while for the sampling method described
here the correlation time crosses over to a constant. A similar behavior holds
for the critical $\mathrm O(3)$ model, though in that case the constant value
the correlation time approaches at large field is larger than its minimum
value (see Fig.~\ref{fig:correlation_time-collapse}). This behavior isn't
particularly worrisome, since the very large field regime corresponds to
correlation lengths smaller than the lattice spacing and is well-described by
other algorithms.  More detailed discussion on correlation times and these
numeric experiments can be found in section \ref{sec:performance}.

\begin{figure}
  \include{fig_generator-times}
  \caption{
    The scaled autocorrelation time of the energy $\H$ for the Wolff algorithm
    on a $32\times32\times32$ \textsc{xy} model at its critical temperature as a
    function of applied vector field magnitude $|H|$. Red points correspond to
    reflections sampled uniformly, while the green points represent
    reflections sampled as described in section \ref{sec:examples:on}.
  }
  \label{fig:generator_times}
\end{figure}



\subsection{The Potts model}

In the $q$-state Potts model spins are described by elements of
$\{1,\ldots,q\}$. Its symmetry group is the symmetric group $\mathrm S_n$ of
permutations of its elements. The element $(i_1,\ldots,i_q)$ takes the spin
$s$ to $i_s$. There are potentially many elements of order two, but the
two-element swaps alone are sufficient to both generate the group and act
transitively on $\{1,\ldots,q\}$, providing ergodicity.

\subsection{Clock models}

In the $q$-state clock model spins are described by elements of $\Z/q\Z$, the
set of integers modulo $q$. Its symmetry group is the dihedral group
$D_q=\{r_0,\ldots,r_{q-1},s_0,\ldots,s_{q-1}\}$, the group of symmetries of a
regular $q$-gon. The element $r_n$ represents a rotation by $2\pi n/q$, and
the element $s_n$ represents a reflection composed with the rotation $r_n$.
The group acts on spins by permutation: $r_n\cdot m={n+m}\pmod q$ and
$s_n\cdot m={-(n+m)}\pmod q$. This is the natural action of the group on the
vertices of a regular polygon that have been numbered $0$ through $q-1$. The
elements of $D_q$ of order 2 are all reflections and $r_{q/2}$ if $q$ is even,
though the former can generate the latter. While reflections do not
necessarily generate the entire group, their action on $\Z/q\Z$ is transitive
and therefore the algorithm is ergodic.

\subsection{Roughening models}

Though not often thought of as a spin model, roughening of surfaces can be
described in this framework. Spins are described by integers $\Z$ and their
symmetry group is the infinite dihedral group $D_\infty=\{r_i,s_i\mid
i\in\Z\}$, whose action on the spin $j\in\Z$ is given by $r_i\cdot j=i+j$ and
$s_i\cdot j=-i-j$. The elements of order two are reflections $s_i$, whose
action on $\Z$ is transitive. The coupling can be any function of the absolute
difference $|i-j|$.  Because random choice of reflection will almost always
result in energy changes so large that the whole system is flipped, it is
better to select random reflections about integers or half-integers close to
the average state of the system.  A variant of the algorithm has been applied
without a field whose success relies both on this and another technique
\cite{evertz_stochastic_1991}. They note that detailed balance is still
satisfied if the bond probabilities \eqref{eq:bond_probability} is modified by
adding a constant $0<x\leq1$ with
\[
  p_r(s_m,s_j\mid x)=\min\{0,1-xe^{\beta(\J(r\cdot s_m,s_j)-\J(s_m,s_j))}\}.
\]
When $x<1$ transformations that do not change the energy of a bond can still
activate it in the cluster, which allows nontrival clusters to be seeded when
the height of the starting site is also the plane of reflection. This
modification is likely useful in general for systems with large yet discrete
state spaces.

\section{Performance}
\label{sec:performance}

No algorithm is worthwhile if it doesn't run efficiently. This algorithm,
being an extension of the Wolff algorithm into a new domain, should be
considered successful if it likewise extends the efficiency of the Wolff
algorithm into that domain. Some systems are not efficient under Wolff, and we
don't expect them to fare better when extended in a field. For instance, Ising
models with random fields or bonds technically can be treated with Wolff
\cite{dotsenko_cluster_1991}, but it is not efficient because the clusters
formed do scale naturally with the correlation length \cite{rieger_monte_1995,
redner_graphical_1998}. Other approaches, like replica methods, should be
relied on instead \cite{redner_graphical_1998, chayes_graphical_1998,
machta_replica-exchange_2000}. 

At a critical point, correlation time $\tau$ scales with system size
$L=N^{-D}$ as $\tau\sim L^z$. Cluster algorithms are celebrated for their
small dynamic exponents $z$. In the vicinity of an ordinary critical point,
the renormalization group predicts scaling behavior for the correlation time
as a function of temperature $t$ and field $h$ of the form
\[
  \tau=h^{-z\nu/\beta\delta}\mathcal T(ht^{-\beta\delta},hL^{\beta\delta/\nu}).
\]
If a given dynamics for a system at zero field results in scaling like $L^z$,
one should expect its natural extension in the presence of a field to scale
roughly like $h^{-z\nu/\beta\delta}$ and collapse appropriately as a function
of $hL^{\beta\delta/\nu}$.

We measured the autocorrelation time $\tau$ of the energy $\H$ for a variety of
models at critical temperature with many system sizes and canonical fields
(see Table~\ref{table:models} with $h=\beta H$) using standard methods for
obtaining the value and uncertainty from timeseries
\cite{ossola_dynamic_2004}. Since the computational effort expended in each
step of the algorithm depends linearly on the size of the associated cluster,
these values are then scaled by the average cluster size per site
$\avg{s_{\text{\sc 1c}}}/L^D$ to produce something proportional to machine
time. The resulting scaling behavior, plotted in
Fig.~\ref{fig:correlation_time-collapse}, is indeed consistent with an
extension to finite field of the behavior at zero field, with an eventual
finite-size crossover to constant autocorrelation time. This crossover isn't
always kind to the efficiency, e.g., in the $\mathrm O(3)$ model, but in the
large-field regime where the crossover happens the correlation length is on
the scale of the lattice spacing and better algorithms exist, like
Bortz--Kalos--Lebowitz for the Ising model \cite{bortz_new_1975}. Also plotted
are lines proportional to $h^{-z\nu/\beta\delta}$, which match the behavior of
the correlation times in the intermediate scaling region.

\begin{figure*}
  \include{fig_correlation-times}
  \caption{
    Scaling collapse of autocorrelation times $\tau$ for the energy $\H$
    scaled by the average cluster size as a function of external field for
    various models of Table~\ref{table:models}. Critical exponents are
    model-dependent. Colored lines and points depict values as measured by the
    extended algorithm. Solid black lines show a plot proportional to
    $h^{-z\nu/\beta\delta}$ for each model.
  }
  \label{fig:correlation_time-collapse}
\end{figure*}

Since the formation and flipping of clusters is the hallmark of Wolff
dynamics, another way to ensure that the dynamics with field scale like those
without is to analyze the distribution of cluster sizes. The success of the
algorithm at zero field is related to the fact that the clusters formed
undergo a percolation transition at models' critical point.  According to the
scaling theory of percolation \cite{stauffer_scaling_1979}, the distribution
of cluster sizes in a full Swendsen--Wang decomposition of the system scales
consistently near the critical point if it has the form
\[
  P_{\text{SW}}(s)=s^{-\tau}f(ts^\sigma,th^{-1/\beta\delta},tL^{1/\nu}).
\]
The distribution of cluster sizes in the Wolff algorithm can be computed from
this using the fact that the algorithm selects clusters with probability
proportional to their size, or
\[
  \begin{aligned}
    \avg{s_{\text{\sc 1c}}}&=\sum_ssP_{\text{\sc
    1c}}(s)=\sum_ss\frac sNP_{\text{SW}}(s)\\
    &=L^{\gamma/\nu}g(ht^{-\beta\delta},hL^{\beta\delta/\nu}).
  \end{aligned}
\]

For the Ising model, an additional scaling relation can be written. Since the
average cluster size is the average squared magnetization, it can be related
to the scaling functions of the magnetization and susceptibility per site by
(with $ht^{-\beta\delta}$ dependence dropped)
\[
  \begin{aligned}
    \avg{s_{\text{\sc 1c}}}
    &=L^{D}\avg{M^2}=\beta\avg\chi+L^{D}\avg{M}^2\\
    &=L^{\gamma/\nu}\big[(hL^{\beta\delta/\nu})^{-\gamma/\beta\delta}\beta \mathcal
      Y(hL^{\beta\delta/\nu},ht^{-\beta\delta})\\
      &\hspace{1em}+(hL^{\beta\delta/\nu})^{2/\delta}\mathcal
    M(hL^{\beta\delta/\nu},ht^{-\beta\delta})\big].
  \end{aligned}
\]
We therefore expect that, for the Ising model, $\avg{s_{\text{\sc
1c}}}L^{-\gamma/\nu}$ should go as $(hL^{\beta\delta/\nu})^{2/\delta}$ for
large argument. We further conjecture that this scaling behavior should hold
for other models whose critical points correspond with the percolation
transition of Wolff clusters.  This behavior is supported by our numeric work
along the critical isotherm for various Ising, Potts, and $\mathrm O(n)$
models, shown in Fig.~\ref{fig:cluster_scaling}. Fields for the Potts and
$\mathrm O(n)$ models take the form $B(s)=(h/\beta)\sum_m\cos(2\pi(s-m)/q)$
and $B(s)=(h/\beta)[1,0,\ldots,0]s$ respectively. As can be seen, the average
cluster size collapses for each model according to the scaling hypothesis, and
the large-field behavior likewise scales as we expect from the na\"ive Ising
conjecture.

\begin{figure*}
  \input{fig_clusters_ising2d}
  \caption{Collapses of rescaled average Wolff cluster size $\avg s_{\text{\sc
    1c}}L^{-\gamma/\nu}$ as a function of field scaling variable
    $hL^{\beta\delta/\nu}$ for a variety of models. Critical exponents
    $\gamma$, $\nu$, $\beta$, and $\delta$ are model-dependant. Colored lines
    and points depict values as measured by the extended algorithm. Solid
    black lines show a plot of $g(0,x)\propto x^{2/\delta}$ for each model.
  }
  \label{fig:cluster_scaling}
\end{figure*}

\section{Applying Nonlinear Fields to the xy Model}

Thus far our numeric work has quantified the performance of existing
techniques. Briefly, we demonstrate our general framework in a new way:
harmonic perturbations to the low-temperature {\sc xy}, or \twodee O(2),
model. We consider fields of the form $B_n(s)=h_n\cos(n\theta(s))$, where
$\theta$ is the angle made between $s$ and the $x$-axis, say. Corrections of
these types are expected to appear in realistic models of systems na\"ively
expected to exhibit Kosterlitz--Thouless critical behavior due to the presence
of the lattice or substrate. Whether these fields are relevant or irrelevant
in the renormalization group sense determines whether those systems spoil or
admit that critical behaviour. Among many fascinating
\cite{jose_renormalization_1977, kankaala_theory_1993,
ala-nissila_numerical_1994, dierker_consequences_1986, selinger_theory_1988}
results that emerge from systems with one or more of these fields applied,
it is predicted that $h_4$ is relevant while $h_6$ is not at some
sufficiently high temperatures below the Kosterlitz--Thouless point
\cite{jose_renormalization_1977}.

\begin{figure}
  \include{fig_harmonic-susceptibilities}
  \caption{Susceptibilities as a function of system size for a \twodee O(2)
    model at $T=0.7$ and with (top) fourfold symmetric and (bottom) sixfold
    symmetric perturbing fields. Different field strengths are shown in
    different colors.
  }
  \label{fig:harmonic-susceptibilities}
\end{figure}

We made a basic investigation of this result using our algorithm. Since we ran
the algorithm at fairly high fields we did not choose reflections though the
origin uniformly. Instead, we choose the planes of reflection first by
rotating our starting spin by $\pi m/n$ for $m$ uniformly taken from
$1,\ldots,n$ and generating a normal to the plane from that direction as
described in Section \ref{sec:examples:on}. The resulting susceptibilities as
a function of system size are shown for various field strengths in
Fig.~\ref{fig:harmonic-susceptibilities}. In the fourfold case, for each field
strength there is a system size at which the divergence in the susceptibility
is cut off, while for the sixfold case we measured no such cutoff, even up to
strong fields. This conforms to the expected result, that even in a strong field
the sixfold perturbations preserve the critical behavior. Previous work has
used Monte Carlo to investigate similar symmetry-breaking fields and used a hybrid
cluster--metropolis method \cite{ala-nissila_numerical_1994}. To our
knowledge, no application of a direct cluster method has been applied to this
problem before now.

\section{Conclusions}

We have taken several disparate extensions of cluster methods to spin models
in an external field and generalized them to work for any model of a broad
class.  The resulting representation involves the introduction of not a ghost
spin, but a ghost transformation. We provided evidence that algorithmic
extensions deriving from this method are the natural way to extend cluster
methods in the presence of a field, in the sense that they appear to reproduce
the scaling of dynamic properties in a field that would be expected from
renormalization group predictions.

In addition to uniting several extensions of cluster methods under a single
description, our approach allows the application of fields not possible under
prior methods. Instead of simply applying a spin-like field, this method
allows for the application of \emph{arbitrary functions} of the spins. For
instance, theoretical predictions for the effect of symmetry-breaking
perturbations on spin models can be tested numerically
\cite{jose_renormalization_1977, blankschtein_fluctuation-induced_1982,
bruce_coupled_1975, manuel_carmona_$n$-component_2000}.


\begin{acknowledgments}
  This work was supported by NSF grant NSF DMR-1719490.
\end{acknowledgments}

\bibliography{monte-carlo}


\end{document}