cond-mat0507394/BW.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %    INSTITUTE OF PHYSICS PUBLISHING                                   %
3: %                                                                      %
4: %   `Preparing an article for publication in an Institute of Physics   %
5: %    Publishing journal using LaTeX'                                   %
6: %                                                                      %
7: %    LaTeX source code `ioplau2e.tex' used to generate `author         %
8: %    guidelines', the documentation explaining and demonstrating use   %
9: %    of the Institute of Physics Publishing LaTeX preprint files       %
10: %    `iopart.cls, iopart12.clo and iopart10.clo'.                      %
11: %                                                                      %
12: %    `ioplau2e.tex' itself uses LaTeX with `iopart.cls'                %
13: %                                                                      %
14: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
15: %
16: %
17: % First we have a character check
18: %
19: % ! exclamation mark    " double quote  
20: % # hash                ` opening quote (grave)
21: % & ampersand           ' closing quote (acute)
22: % $ dollar              % percent       
23: % ( open parenthesis    ) close paren.  
24: % - hyphen              = equals sign
25: % | vertical bar        ~ tilde         
26: % @ at sign             _ underscore
27: % { open curly brace    } close curly   
28: % [ open square         ] close square bracket
29: % + plus sign           ; semi-colon    
30: % * asterisk            : colon
31: % < open angle bracket  > close angle   
32: % , comma               . full stop
33: % ? question mark       / forward slash 
34: % \ backslash           ^ circumflex
35: %
36: % ABCDEFGHIJKLMNOPQRSTUVWXYZ 
37: % abcdefghijklmnopqrstuvwxyz 
38: % 1234567890
39: %
40: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
41: %
42: \documentclass[12pt]{iopart}
43: % Uncomment next line if AMS fonts required
44: \usepackage{iopams,epsf,graphicx}  
45: \begin{document}
46: 
47: %\title[Author guidelines for IOPP journals]{Monte Carlo study of the Pure and Dilute Baxter-Wu model}
48: \title{Monte Carlo study of the Pure and Dilute Baxter-Wu model}
49: %\twocolumn
50: 
51: \author{Nir Schreiber%\dag 
52:  \ and Joan Adler %Robertson\ddag  
53: %\footnote[3]{To
54: %whom correspondence should be addressed (romneya.robertson@iop.org)}
55: }
56: 
57: \address{Physics Department, Technion, Israel Institute of Technology, 
58: Haifa, Israel, 32000  }%Lines break automatically or can be forced with \\
59: %\dag\ Production Editor, Institute of Physics 
60: %Publishing, Dirac
61: %House, Temple Back, Bristol BS1 6BE, UK}
62: 
63: %\address{\ddag\ Electronic Services Specialist, Institute of Physics Publishing, 
64: %Dirac House, Temple Back, Bristol BS1 6BE, UK}
65: 
66: 
67: \begin{abstract}
68: We studied the pure and  dilute Baxter-Wu (BW) models 
69: using the Wang-Landau (WL)
70: sampling method to calculate the Density-Of-States (DOS).
71: We first used the exact result for the DOS of the Ising model to test our code. Then
72: we calculated the DOS of the dilute Ising model to obtain a phase diagram,
73: in good agreement with previous studies.
74: We calculated the energy distribution, together  
75: with its first, second and fourth moments,
76: to give the specific heat and the 
77: energy fourth order cumulant, better known as the Binder parameter, for the pure BW model. 
78: For small samples, the energy distribution displayed a doubly peaked shape, and
79:  finite size scaling analysis 
80: %%, however,
81: %%%this scaling form, due to Lee and Kosterlitz is related to the doubly peaked distribution. Therefore I 
82: %%% changed ``however'' to ``in accordance''. 
83:  showed the expected reciprocal scaling of the positions of the peaks with $L$.
84: The energy distribution yielded the expected BW $\alpha=2/3$ critical exponent for the specific heat. The Binder parameter minimum appeared to scale with lattice size $L$ 
85: with an exponent $\theta_B$ equal to the specific heat exponent. Its location (temperature) 
86: showed a large correction-to-scaling term $\theta_1=0.248\pm0.025$. For the dilute BW model we found a clear crossover to a single peak 
87: in the energy distribution even for small sizes  and 
88:  the expected
89: $\alpha=0$ was recovered. 
90: 
91: \end{abstract}
92: 
93: %Uncomment for PACS numbers title message
94: %\pacs{00.00, 20.00, 42.10}
95: \pacs{05.10.Ln,05.50.+q,02.70.Rr}
96: 
97: % Uncomment for Submitted to journal title message
98: \submitto{\JPA}
99: 
100: % Comment out if separate title page not required
101: \maketitle
102: 
103: \section{Introduction}
104: The two-dimensional 
105: Ising model has received such widespread attention as the paradigm
106: system for phase transitions, that one sometimes says that a certain
107: system is the ``Ising model'' of a class of problems.
108: While it is clearly special,  it is not the only two dimensional model
109: of phase transitions with an exact expression for its free energy. 
110: Its critical behavior is, in fact, rather atypical 
111: relative to many other
112: two-dimensional systems
113: and even to the three-dimensional Ising model, especially
114: in the specific heat,
115: where its critical exponent, $\alpha$, is zero.  The nature of the
116:  corrections-to-scaling in the spin 1/2 Ising model
117: is also very different to that of many other
118:  interesting systems.
119: 
120: Another spin system, now known as the
121:  Baxter Wu  (BW) model, 
122:   was solved  by R.J. Baxter
123:  and F.Y. Wu~\cite{BW1,BW2}.
124: Spins $\sigma_i=\pm1$,
125: are situated on the triangular lattice
126: and interact via a three spin interaction, 
127: \begin{equation}
128: {\cal{H}}=-J\sum_{i,j,k}\sigma_i\sigma_j\sigma_k,
129: \end{equation}
130: where $i,j$ and $k$ are the vertices of a
131:  triangle as shown in Fig.~\ref{fig:BW}. $J>0$ is the ferromagnetic
132: coupling between nearest neighbor spins.
133: \begin{figure}[h]
134: \begin{center}
135: \includegraphics[width=5.00cm]{1.eps}
136: \end{center}
137: \caption{\label{fig:BW}The energy of a given configuration is the sum of all interacting 
138: triangles formed by nearest neighbor spins}
139: \end{figure}
140: The BW model exhibits a second order phase
141: transition with its 
142: critical temperature ($T_c$) given by 
143: $2J/kT_c=\ln(1+\sqrt 2)=2.26918...$, 
144: (the same numerical value  as for 
145:   the Ising model on the square lattice). The specific heat critical exponent is equal to the correlation length
146:  exponent, $\alpha=\nu=2/3$. 
147: Series-expansion results~\cite{BaxterSykesWatts}, gave the conjectured 
148: magnetization exponent of $\beta=1/12$ and
149: a susceptibility  exponent of $\gamma\approx1.17$~\cite{Griffiths}. 
150: The latter confirmed the prediction of $\gamma=7/6$ from the well known
151:  scaling relation
152: $\alpha+2\beta+\gamma=2$~\cite{Essam,Rushbrooke}.
153:  Real Space Renormalization Group methods 
154: have also been used~\cite{Braathen,Imbro,Nijs} to study the pure model,
155: and the critical eigenvalues obtained gave critical exponents 
156: consistent with series-expansion and exact results.
157:  An exact form for  BW corrections-to-scaling  was found 
158: by Joyce~\cite{Joyce}
159:  who conjectured
160: that the
161:  spontaneous magnetization varied as 
162: $M=t^\beta\left(f_0(t) +t^{2/3}f_1(t)+\cdots\right)$
163: with analytic functions $f_0, \, f_1$ of the distance
164: $t=(T-T_c)/T_c$. Adler and Stauffer confirmed this with
165: series and Metropolis Monte Carlo estimates~\cite{adst}. 
166: 
167: Dilute Ising models are also somewhat famous, but for rather 
168: different reasons, as they have been the source of a great deal of controversy.
169: Presumably because of the anomalous specific heat structure in the 
170: pure case, numerical work in the dilute regime, especially 
171: near the pure limit is painful, and although 
172: a majority of authors (see e.g. Roder $et \ al$,~\cite{Roder}) have
173: claimed that the controversy is resolved in favor 
174: of SSL theory~\cite{Shalaev,Shankar,Ludwig}, 
175: more study is useful.
176: 
177: 
178: The annealed dilute BW model was studied by Kinzel, Domany and Aharony\cite{aharony}
179: who showed from this exploration 
180: that its dominant critical behavior is in the 
181: universality class of the four state Potts model, although the Potts 
182: model has logarithmic correction terms for this case. Domany and Riedel~\cite{Domany}, argued the same for 
183: the pure BW model by means of symmetries of the Landau-Ginzburg-Wilson Hamiltonian.
184:  (Note that there 
185: are first order fixed points in the neighborhood of these models).
186: The quenched dilute
187:  BW model was studied by Landau and Novotny~\cite{Novotny}, who found
188:  a substantial change in the critical behavior
189: of the specific heat~\cite{Harris} for an impurity concentration of
190:  $1-x=0.1$. They also 
191: conjectured that the zero temperature threshold
192: concentration above which no long-range order could be seen was
193:  about $x_c\simeq0.71$. (See also results from a cluster-algorithm study  
194: of this system~\cite{Evertz}).
195: More recent calculations~\cite{fried,adler} showed that the value of $x_c$ is 
196:  even higher ($x_c\simeq0.755$), and is bounded by $x_c^{\rm low}=0.710\pm0.001$ and
197:  $x_c^{\rm high}=0.784\pm0.004$. This is substantially above the value for Ising models where $x_c$ is simply
198:  the percolation threshold of the corresponding lattice, which is rarely above 0.5 .
199: 
200: 
201: 
202: Recently Wang and Landau~\cite{WL1,WL2} proposed a very efficient
203: algorithm for calculating the density-of-states (DOS), (i.e. the degeneracy of any level in energy space),
204: $g(E)$,
205: for Ising models and some related systems.
206: To explore
207:   the issues of both pure and dilute BW models further,
208:  and to see how its different
209: particulars of large $\alpha$ and corrections to scaling emerge from
210: the calculation of the DOS,
211: we have chosen to apply the WL algorithm to both the
212: pure and quenched dilute BW
213:  models, and to study the behavior of the energy distribution and related moments~\cite{Challa,histogram2}
214:  using the simulated DOS.
215:  The DOS of the pure and dilute Ising models was studied for comparison purposes.
216: 
217: In the next section we discuss the WL algorithm.
218:  In section~\ref{sec:Ising} we
219:  present a comparison of an exact calculation of the DOS for the
220: Ising model~\cite{Beale}
221: with simulations using WL and give some results for the dilute Ising case.
222:  In section~\ref{sec:BW_pure}
223:  we give in detail our results for the pure BW model,
224: and in section ~\ref{sec:BW_dilute}
225: the results for the dilute BW model are presented.
226: Finally we discuss the implications of our results in
227: section~\ref{sec:discussion}.
228: 
229: \section{ The simulation method}
230: 
231: 
232: Conventional Monte-Carlo (MC) methods~\cite{Metropolis,Swendsen,Wolff}
233: generate the canonical energy distribution at a given
234: temperature $T_0$. It is usually narrowly peaked around this temperature.
235: The need to perform
236: multiple simulations in order to obtain thermodynamics
237: in a large range of temperatures requires a large
238: computational effort.
239: Other methods based on histogram accumulation
240: ~\cite{histogram2,histogram1}
241: approximate the distribution by the energy histogram at $T_0$.
242: This distribution can then be reweighted
243: to give statistics at another temperature. The reweighted
244:  distributions, however, are also
245: restricted to a very narrow range of temperatures and suffer
246:  from large statistical errors in their tails for
247: temperatures far from $T_0$.
248: The broad histogram method~\cite{deOliveira}
249: calculates the DOS through the consideration of the average number of visits to
250: any two adjacent energy levels. Lee~\cite{Lee} offered the entropic sampling 
251: method using the observation that if the transition probability between any
252: two energy levels is proportional to the ratio between the DOS
253: of these levels, then a crude estimate to the DOS can be given 
254: when sampling at infinite temperature.
255: 
256: Wang and Landau improved Lee's method by introducing a modification factor
257: which together with generating a ``flat'' histogram 
258: (we have used the condition $|H(E)-\langle H\rangle|/\langle H\rangle \le0.05$ for any $E$),
259: carefully controls the updating of the DOS.  
260: By dividing the energy space into
261:  different segments, and
262: performing an independent random walk in each segment,
263: one can generate very accurately, in a reasonable amount of CPU time,
264: the DOS of the whole energy space,
265: thus obtaining the canonical distribution at any desired temperature.
266: \section{Ising model results}{\label{sec:Ising}}
267: \subsection{The pure Ising model}
268: We began by validating the accuracy of our implementation of
269:  the WL algorithm against exact
270:  results for the Ising model on the square lattice with no impurities.
271: A detailed comparison was made for the case of $L=32$.
272: The partition function for the Ising model  on a lattice of length
273: $L$ can be written as a low
274: temperature expansion
275: \begin{equation}
276: {\cal{Z}}_N=e^{2KN}\sum_{\ell}g_{\ell}x^{2\ell},
277: \label{Z_lowT}
278: \end{equation}
279: where $N=L\times{L}$ is the number of spins, $K=J/kT$ is
280: the reduced inverse temperature,
281: and $x=e^{-2K}$ is the low temperature variable. Each energy level can be labeled (relative to the ground state energy $-2JN$) by $E_{\ell}=4J\ell$ $(\ell=0,2,3,...N-2,N)$,
282: so that $g_{\ell}$ is its corresponding DOS.
283: Beale~\cite{Beale} used an  extension of Onsager's solution~\cite{Onsager}, to give
284:  the exact expression
285: for the partition function on a finite lattice~\cite{Kaufman},
286: and extracted the DOS coefficients from the
287: expansion~(\ref{Z_lowT}). When we plotted our results 
288: on top of Beale's expression for the case of $L=32$ we saw no
289: deviations between the exact  and the simulated data within the resolution
290: of the figure. The relative error between the exact and simulated data was
291: also plotted and was found to be three orders of magnitude smaller than 
292: the calculated DOS and two orders of magnitude larger from
293: the systematic error due to the choice of the final modification factor
294: $f_{\rm final}=0.001$. This showed that the choice of this 
295: quite large $f_{\rm final}$ was sufficient, so that only a relatively small
296: number of iterations was required for all the simulations performed
297: throughout this work.
298: 
299: Further results from the pure Ising simulations will be introduced for comparison purposes in section~\ref{sec:BW_pure}.
300: 
301: \subsection{The dilute Ising model}
302: We continued the validation process by studying the dilute Ising model.
303: at a lattice size of $L=22$.
304: The Hamiltonian for the dilute Ising model may be written as
305: \begin{equation}
306: {\cal{H}}=-J\sum_{i,j}\epsilon_i\epsilon_j\sigma_i\sigma_j,
307: \end{equation}
308: where the random disorder variables $\epsilon_i$ take the values 0 and 1,
309: such that their configurational average is equal to a dilution of  $0<x<1$.
310: We considered the position of the specific heat maxima, $T_{C_{\rm{max}}}$, for different nominal concentrations
311:  centered around the values $x=0.8,0,9$ and $x=0.95$, as
312:  indicated in Fig.~\ref{fig:Idilute}.
313: It is clearly seen that for large concentrations $(x\geq0.9)$ the
314:  circles tend to a continuously critical
315: line, slightly shifted from the solid line. The shift is a finite
316: size effect due to the use of a small sample.
317: For smaller concentrations there is a large dispersion of the circles
318: and the data are 
319: %%les
320: %%% data is ``worse'' for that reason, hence, less->more, or just leave it....?
321: less reliable.
322: As shown in Fig.~\ref{fig:CIdilute}, the specific heat maximum becomes
323: broader with decreasing concentration and is hard to locate precisely.
324:  The reason for this is that when lowering the concentration %$x$ 
325: isolated clusters which rarely interact
326: with each are formed, and hence
327: energy fluctuations become smaller. For a concentration of $x=0.75$ 
328: %%they are also->these fluctuations
329:  these fluctuations are also nearly constant
330: and therefore no pronounced peak can be identified. It should be noted that 
331: presumably, when much larger samples would be used, a pronounced peak should be clearly
332: seen for concentration even lower than $x=0.75$ (see, for example~\cite{Selke}).
333: In the absence of analytic results, the location of our points
334: close to earlier estimates validates both our dilute code and our analysis methods.
335: \begin{figure}[h]
336: \begin{center}
337: \includegraphics[height=9.0cm,angle=-90]{2.ps}%{T_p2.ps}
338: \caption{Critical line $T_c(x)$ in the $T$-$x$ plane of the dilute Ising model. Small energy fluctuations for $x\leq0.8$ make it hard to reliably determine $T_{C_{\rm{max}}}$.
339: The asterisks represent results from MC Renormalization Group calculations~\cite{deSouza} and the solid line
340: is the prediction $T_c(x)=\{\tanh^{-1}[e^{-1.45(x-x_c)}]\}^{-1}$, converging to the value of $x_c=p_c=0.593$~\cite{Stauffer}.}
341: \label{fig:Idilute}
342: \includegraphics[height=9.0cm,angle=-90]{3.ps}%{c22_1.ps}{c22.ps}
343: \caption{The specific heat of the dilute Ising model for different concentrations on a $L=22$ lattice.
344: For $x=0.75$ there is no pronounced peak present}
345: \label{fig:CIdilute}
346: \end{center}
347: \end{figure}
348: 
349: \section{The pure Baxter-Wu model}{\label{sec:BW_pure}
350: We calculated the DOS for the BW model
351: using lattice sizes $L$ ranging from 6 to 120,
352: with periodic boundary conditions being imposed. For each lattice size the data was collected separately for each energy segment and then was combined to give the density of states for the entire energy landscape. We averaged over nine different runs for $L=30$, and saw that the fluctuations were
353: three orders of magnitude smaller then the measured quantity ($\ln g$), so that we neglected these fluctuations and for each
354: lattice size we executed a single run per segment only. By symmetry, for any state with negative energy, there exist a state with positive energy, so that it was sufficient to carry out the random walk only for non positive energies. (A similar argument holds for the Ising model). Plots of the internal energy, specific heat, free energy and entropy are given
355: in Fig.~\ref{fig:54results}.
356: \begin{figure}[h]
357: \begin{center}
358: \includegraphics[height=9.0cm,angle=-90]{4.ps}
359: \caption{Calculation of thermodynamic functions for the pure BW model on an $L=54$ lattice: (a) Internal Energy, (b) Specific Heat, (c) Entropy and (d) Free energy. The specific heat displays a very clear pronounced peak at the transition point.}
360: \label{fig:54results}
361: \end{center}
362: \end{figure}
363: 
364: Early simulations~\cite{Novotny} showed the formation and motion of domains around the ferromagnetic and ferrimagnetic ground
365: states, due to the special connectivity of the BW model, causing low frequency large energy fluctuations. These fluctuations made the impression that the system was in
366: a metastable state, thus indicating a first order transition.
367: In Fig~\ref{fig:H} we examined the energy distribution at $T_{C_{\rm{max}}}$ and found a doubly peaked curve
368: (see ref.~\cite{Kosterlitz}).
369: The system appears to fluctuate between these two peaks denoted by $E_-$, corresponding to an "ordered" state (more negative) energy, incorporating
370: small clusters, and $E_+$, corresponding to a "disordered" state energy incorporating large clustering.
371: A plot of the distribution for the Ising model both at $T_{C_{\rm{max}}}^{\rm{Ising}}$ and at $T_{C_{\rm{max}}}^{\rm{BW}}$
372: shows clearly sharp single peaks centered approximately at
373: the critical energy $U_c=-\sqrt2J$ (Fig.~\ref{fig:dos}).
374: This supports the uniqueness of the distributions in Fig~\ref{fig:H}.
375: \begin{figure}[h]
376: \begin{center}
377: \centerline{\includegraphics[height=9.00cm,angle=-90]{5.ps}}%histograms4.ps{histogram.ps}
378: \caption{Critical distribution calculated at $T_{C_{\rm{max}}}$ for the pure BW model. The lattice sizes are denoted by arrows. The $L=120$ data suffers from the
379: systematic errors resulting from the DOS calculations for large systems.}
380: \label{fig:H}
381: \includegraphics[height=9.0cm,angle=-90]{6.ps}%{p54_BW_Ising2.ps}
382:  \caption{Energy distributions at $T_{C_{\rm{max}}}^{\rm{Ising}}=2.28948 J/k_B$, 
383: at $T_{C_{\rm{max}}}^{\rm{BW}}=2.27549 J/k_B$ and at the exact transition point $T_c$ on the same lattice with $L=54$. The numbers in parenthesis
384: denote: (1) Ising at $T_{C_{\rm{max}}}^{\rm{BW}}$, (2) BW at $T_c$,
385: (3) Ising at $T_{C_{\rm{max}}}^{\rm{Ising}}$, (4) BW at $T_{C_{\rm{max}}}^{\rm{BW}}$.
386: Note the distribution at the exact transition point (2) with the ratio of $r\simeq4$ between the pronounced peak on the left and the "hump" on the right~\cite{histogram2}.
387: The asterisk denotes their common critical energy $U_c=-\sqrt2J$.}
388: \label{fig:dos}
389: \end{center}
390: \end{figure}
391: The positions (energies) of the peaks are found to scale
392: with $L^{-1}$~\cite{Kosterlitz} as seen in Fig.~\ref{fig:E1}, 
393: and are expected to eventually intersect for a large enough sample.
394: \begin{figure}[h]
395: \begin{center}
396: \includegraphics[height=9.00cm,angle=-90]{7.ps}
397: \caption{Variation of the energy distribution's two maxima positions with
398: $L^{-1}$. The ``disordered'' energies are denoted by ($\triangle$) and the
399: ``ordered'' energies by ($\circ$).}
400: \label{fig:E1}
401: \end{center}
402: \end{figure}
403: 
404: A comparison between the DOS of the Ising model and the BW model (Fig.~\ref{fig:DOS}) shows a significant difference between the two models.
405: Although they have approximately the same number of different energy levels ($N-1$ for Ising and $N-3$ for BW), the function $\ln g$, appears to be concave everywhere on the interval $[-2,0]$ for the Ising model, while, for the BW graph this may not be so. This suggests an explanation for the appearance of the two peaks
406: which is also consistent with the fact that they have the same height:
407: The condition that the distribution will have $extrema$ is satisfied by
408: \begin{equation}
409:  d(\ln g)/dE=1/kT.
410: \label{eq:Tc}
411: \end{equation}
412: If, at  $T_{C_{\rm max}}$,
413:  Eq.~(\ref{eq:Tc}) has locally,
414: a solution $f_1(E)=E/kT_{C_{\rm max}}+C_1$ tangent to $\ln g$ at $E_-$ and $E_+$,
415: and another solution $f_1(E)=E/kT_{C_{\rm max}}+C_2$ tangent to $\ln g$ at $U_c(L)$ (at the shifted
416: critical energy, or the minimum between the peaks), then the distribution will have two peaks with
417: equal height satisfying 
418: \begin{equation}
419: p(E_-)=p(E_+)=e^{C_1},
420: \end{equation}
421: as seen in Fig.~\ref{fig:H}.
422: This is essentially a finite size effect and 
423: should be recovered by a large enough sample, to give an "Ising like" concave everywhere DOS function, and a single peaked distribution as its consequence.
424: 
425: We further calculated the specific heat for each lattice size and then plotted its maximal value $C_{\rm{max}}$ versus $L$.
426: We see in Fig.~\ref{fig:C} a very nice agreement between the calculated data and the second order ansatz
427: $C_{\rm max}(L)\propto L^{\alpha/\nu}$, with $\alpha/\nu=1$, even for very small lattices $(L=6)$.
428: 
429: \begin{figure}[h]
430: \begin{center}
431: \includegraphics[height=9.0cm,angle=-90]{8.ps}%{DOS.ps}{54dos_ising.ps}
432: \caption{DOS of BW and Ising models on the $L$=54 lattice. The function $\ln g$ appears to be concave "everywhere" in $[-2,0]$ for the Ising model, while this may not be so in the BW case.
433: A plot of $\ln g(E)$ versus $E$ for a larger BW system $(120\times120)$ is
434: given in the inset.}
435: \label{fig:DOS}
436: \includegraphics[height=9.0cm,angle=-90]{9.ps}%{C_max_alpha_over_nu.ps}%{C_alpha_over_ni1.ps}
437: \caption{Scaling of the specific heat maxima with the length $L$ for the pure BW model. The predicted
438: $C_{\rm{max}}(L)\propto L$ behavior is indicated by a solid line.}
439: \label{fig:C}
440: \end{center}
441: \end{figure}
442: 
443: \begin{figure}[h]
444: \begin{center}
445: \includegraphics[height=9.0cm,angle=-90]{10.ps}%{Binder.ps}
446: \caption{The Binder parameter for the pure BW model versus temperature, for various lattice sizes, from top $(L=120)$ to bottom $(L=60)$ in descending order.
447: The Binder parameter is seen in the figure to display an inverse peak whose depth
448: decreases as the system size increases. 
449: The infinite volume upper bound $B_{\rm{min}}^\infty$ was estimated using first 
450: order scaling theory to $B_{\rm{min}}(L)$.}
451: \label{fig:B}
452: \includegraphics[height=9.0cm,angle=-90]{11.ps}%{Binder_ising1.ps}
453: \caption{Temperature variation of the Binder parameter for the Ising model, from top $(L=60)$ to bottom $(L=24)$ in descending order. The data in the inset is given for the
454: same lattices on a larger scale. The data for $L=54$ is calculated using simulated DOS. All other data is exact.}
455: \label{fig:B_ising}
456: \end{center}
457: \end{figure}
458: 
459: Another quantity of interest was the so called Binder parameter~\cite{Binder1,Binder2}
460: \begin{equation}
461: B=1-\frac{\langle E^4\rangle}{3\langle E^2\rangle^2},
462: \end{equation}
463: where $\langle...\rangle$ stands for the canonical thermal average.
464: When we calculated the Binder parameter,(whose plot as a function of temperature is given in Fig.~\ref{fig:B}),
465: we saw a sharp inverse peak that usually occurs in first order transitions~\cite{Challa,histogram2}. Another manifestation of the strong finite size effects is the very precise (though quite unreliable) estimate of
466: $T_c=2.2696\pm0.0004$ to the transition point we obtained, when performing first order finite
467: size scaling theory to the position of $B_{\rm min}$, $T_{B_{\rm min}}$. 
468: 
469: \begin{figure}[h]
470: \begin{center}
471: \includegraphics[height=9.0cm,angle=-90]{13.ps}
472: \end{center}
473: \caption{Scaling of $T_{B_{\rm{min}}}$ with the inverse volume of the system for the pure BW model.}
474: \label{fig:Tc_Bmin}
475: \end{figure}
476: 
477: Obviously, since the transition is continuous and therefore no ordered and disordered states coexist at the transition point, the critical probability distribution in the infinite volume limit is expected to be single peaked, causing
478: $B_{\rm min}$ to eventually vanish with some exponent and the Binder parameter to take the trivial value
479: of 2/3 also at the critical point.
480: It was therefore convenient to repeat finite size scaling for $B_{\rm min}$
481: according to \begin{equation}
482: B_{\rm min}=\frac{2}{3}-{\cal{B}}_0L^{-\theta_B/\nu},
483: \label{eq:Bnir}
484: \end{equation}
485: where $\theta_B$ is an exponent yet to be determined. In Fig.~\ref{fig:Bmin_Cmax} we see the variation of the inverse distances $t_{C_{\rm max}}^{-1}\equiv\left(T_{C_{\rm max}}-T_c\right)^{-1}$ and
486: $t_{B_{\rm min}}^{-1}\equiv\left(T_{B_{\rm min}}-T_c\right)^{-1}$, correspond to the positions of the specific heat maxima and Binder parameter minima, respectively, with $L$. A least square fit gave a slope of $1.529\pm0.039$ for the specific heat temperature and $1.748\pm0.025$ for the Binder parameter temperature.
487: %%From-> In accordance with~\cite{histogram1}
488:  In accordance with~\cite{histogram1}
489: \begin{eqnarray}
490: T_{C_{\rm max}}=T_c+A_0L^{-1/\nu}\left(1+A_1L^{-\omega_1}+\cdots\right),
491: \end{eqnarray}
492: we use the analogy
493: \begin{equation}
494: T_{B_{\rm min}}=T_c+B_0L^{-1/\nu}\left(1+B_1L^{-\theta_1}+\cdots\right) \label{eq:T_B_min},
495: \end{equation}
496: where $\omega_1$ and $\theta_1$ are correction exponents and $A_0,A_1,B_0$ and $B_1$ are amplitudes determined from simulations. It is therefore evident that
497: $T_{B_{\rm min}}$ displays a large correction-to-scaling term ($\theta_1\simeq0.25$),
498: in contrary to the resulting $1/\nu$ scaling from the $T_{C_{\rm max}}$ fit, which is in fair agreement with the exact 3/2 value, and which is also consistent with the scaling of $C_{\rm max}$. It is also evident, however, from Fig.~\ref{fig:Cmax_Bmin} and Eq.~(\ref{eq:Bnir}), that $\alpha$ and $\theta_B$ have the same value. Similar exact and simulational calculations of the Binder parameter for the
499: Ising model on the same temperature scale (Fig.~\ref{fig:B_ising}) showed much
500: broader and less deep minima at $T_{B_{\rm min}}$, suggesting that these minima vanish with an exponent $\theta_B$ larger than the BW exponent.
501: 
502: 
503: 
504: \begin{figure}[h]
505: \begin{center}
506: \includegraphics[height=9.0cm,angle=-90]{14.ps}%{Bmin_Cmax.ps}%{Tc_L-Tc1.ps}
507: \caption{Scaling of the inverse distances $t_{C_{\rm max}}^{-1}$ and $t_{B_{\rm min}}^{-1}$, with $L$, for the BW model. The larger slope of the Binder parameter position's fit, may be a result of the large correction term $\theta_1$
508: (see Eq.~\ref{eq:T_B_min}). The specific heat data was shifted for ease of reading.}
509: \label{fig:Bmin_Cmax}
510: \includegraphics[height=9.0cm,angle=-90]{15.ps}%{Cmax_Bmin.ps}%{Tc1.ps}%{Tc_L-Tc1.ps}{Cmax_Bmin1.ps}
511: \caption{Scaling of the quantity $\left(2/3-B_{\rm min}\right)^{-1}$ and
512: the specific heat maximum $C_{\rm max}$, with $L$, for the BW model.}
513: \label{fig:Cmax_Bmin}
514: \end{center}
515: \end{figure}
516: 
517: \section{The dilute BW model}{\label{sec:BW_dilute}}
518: Let us consider now the ferromagnetic BW model with quenched impurities. The Hamiltonian is given by
519: \begin{equation}
520: {\cal{H}}=-J\sum_{i,j,k}\epsilon_i\epsilon_j\epsilon_k\sigma_i\sigma_j\sigma_k.
521: \end{equation}
522: We studied systems with lengths $L$ between 18 and 36. We kept concentrations of $x=0.8$ for $L=18$ and of $x=0.9, 0.95$ and $x=0.97$
523: for $L=33$, fixed, and let them vary around $x=0.9$ for $L=33$. The data for $L=24$ was calculated for concentrations varied around
524: different values from $x=0.85$ to $x=0.97$.
525: In Fig.~\ref{fig:DOS_36_pure_dilute} we compare the DOS of the pure and dilute BW models. The apparent crossover to a manifestly clear second order transition may give rise again to a concave everywhere form of $\ln g$, already seen for the Ising model
526: in Fig.~\ref{fig:DOS}.
527: The energy levels differ now only in the amount of $2J$ and can take even or odd values for the same lattice size, depending on
528: the vacancy distribution.
529: We then performed a calculation similar to that made above for the dilute Ising model, of $T_{C_{\rm{max}}}$, to obtain
530: the $T_c(x)$ critical line on a lattice
531: with $L=24$, and then fitted the high concentration data into a continuous (dotted) line (Fig.~\ref{fig:BWdilute}).  All the data except for the $L=18$ with a vacancy concentration
532: of 0.2, which was, as for the dilute Ising model, unreliable because of 
533: relatively high dilution, fell very well on the dotted line. This may suggest that the
534: critical behavior is rather universal for large enough concentrations, because due
535: to the special connectivity of the BW model, one would expect smaller energy fluctuations
536: and therefore a larger scatter of data for large enough vacancy concentrations, whilst the dilute BW data seems to agree with the
537: dilute Ising data for concentrations of $x\geq0.9$. Of course, in order to make
538: definitive statements about universality, larger samples would be needed than 
539: those used here.
540: \begin{figure}%[h]
541: \begin{center}
542: \includegraphics[height=9.0cm,angle=-90]{16.ps}
543: \caption
544: {DOS for a pure (upper curve) and for a dilute (lower curve) BW model with $x=0.9$, on an $L=36$ lattice.}
545: \label{fig:DOS_36_pure_dilute}
546: \end{center}
547: \end{figure}
548: \begin{figure}[h]
549: \begin{center}
550: \includegraphics[height=9.0cm,angle=-90]{18.ps}%{Tc_p_BW_18_30_33_1.ps}%{Tc_p_BW_18_30_33.ps}
551: \caption{Normalized critical temperature for fixed concentrations
552: on different lattice sizes for the dilute BW model.}
553: \label{fig:BWdilute}
554: \end{center}
555: \end{figure}
556: 
557: We performed a rough finite size scaling for the specific heat maxima at a concentration of $x=0.9$,
558: using the three points measured for $L=33$ that were averaged and the other data collected for
559: fixed concentrations.
560: Novotny and Landau~\cite{Novotny} predicted $\alpha/\nu\simeq0$ for a concentration of 0.9.
561: Our results, presented in Fig.~\ref{fig:nu=0} also indicate, at least qualitatively, a significant change in $\alpha$.
562: Since spatial correlations become smaller and hence $\nu$ becomes smaller,  the value of $\alpha$ substantially decreases, thus indicating an "Ising like" singularity at the finite lattice transition point. Moreover, the Harris criterion for the diluted case is hereby confirmed.
563: \begin{figure}
564: \begin{center}
565: \includegraphics[height=9.0cm,angle=-90]{19.ps}%{alpha_over_ni_0.9_log_scale1.ps}% {alpha_over_ni_0.9_2.ps}
566: \caption{Finite size scaling of $C_{\rm max}$ with $L$ for the pure and dilute BW model. The data for the dilute model
567: reveals an $\alpha$ exponent close to zero.}\label{fig:nu=0}
568: \end{center}
569: \end{figure}
570: Another question of interest was the influence of vacancies on the nature of the transition.
571: In order to make a statement regarding this question we plotted in Fig.~\ref{fig:pBWdilute} the energy distribution
572: for different concentrations. We see clearly and unsurprisingly that lowering the concentration causes the doubly peaked distribution
573: to vanish and become a singled peaked one with a narrower width centered away from $U_c$. It may then be plausible to say that
574: in contrast to energy fluctuations which become negligible at sufficiently low concentrations,
575: magnetic fluctuations increase with increasing dilution and the transition is manifestly
576: second order.
577: \begin{figure}
578: \begin{center}
579: \includegraphics[height=9.0cm,angle=-90]{20.ps}%{dilute_histogram.ps}%{p24_trial1.ps}{dilute_his.ps}
580: \caption{Critical energy distribution for various concentrations and critical temperatures, calculated on a $L=24$ lattice. 
581: The numbers in parenthesis denote: (1) $x=0.85$; $T_c=1.74926$, (2)  $x=0.95$; $T_c=1.93379$,
582: (3)  $x=0.97$; $T_c=2.07671$, and (4) $x=1$ (pure); $T_c=2.29164$.
583:  The distribution is seen to become sharper and narrower when the concentration is reduced.}
584: \label{fig:pBWdilute}
585: \end{center}
586: \end{figure}
587: \section{Conclusions}{\label{sec:discussion}}
588: Our simulations have shown that the WL sampling is a very accurate algorithm.
589: The thermodynamic quantities resulting from the calculated $g(E)$, which yield reasonable quality critical data, provide good evidence for this.
590: 
591: Our results show that the pure Baxter-Wu model is strongly influenced by finite size effects and corrections to
592: scaling. The scaling of the specific heat maxima is in excellent agreement with the second order form $C_{\rm max}\propto L^{\alpha/\nu}$, even for small lattices, and no correction terms are observed. The Binder parameter, however, displays large
593: minima for small samples, thus incorrectly could be thought of as a "first order" scaling field. 
594: It is an "irrelevant" field in the sense that it gives no additional information about the universal exponent
595: $\nu$, but rather vanishes with an exponent $\theta_B$.
596: This exponent is also evident in the Ising model and is presumably larger for this model. The vanishing inverse peak in both models states that the energy distribution approaches a delta function in the thermodynamic limit, although it is essentially non-Gaussian.
597: The doubly peaked shape of the latter is rather peculiar. One would usually expect a single peaked distribution which 
598: becomes narrower, the closer to criticality one is. This shape is essentially a finite size effect due to the large fluctuations between the ferromagnetic and ferrimagnetic clusters formed in the vicinity of the transition
599: point, and will eventually vanish in the thermodynamic limit. 
600: The WL method is also very successful when applied for the dilute BW model even for small lattices, 
601: both in terms of the critical isotherm in temperature-concentration plane for a weak dilution, and probability distribution.
602: A crossover to a single peaked critical distribution is clearly seen when decreasing the concentration of spins,
603: and a single peaked distribution is evident at a concentration of $x=0.85$. This is a result of the formation of isolated
604: domains causing relatively small energy fluctuations around the critical energy. 
605: 
606: It would be interesting in the future to use larger lattices to confirm our explanations of the finite 
607: size problems, The relatively high accuracy of the WL method for small dilute systems could be applied in the future to study 
608: disorder in other models.
609: 
610: \section*{Acknowledgments}
611: %\begin{Acknowledgments}
612: We thank Prof. D. Stauffer, Prof. D.P. Landau, Prof. M.E. Fisher and Prof. W. Janke for useful comments and suggestions.
613: We would like to thank E. Warszawski and I. Klich for helpful discussions.
614: We thank the BSF for generous support 
615: throughout this project. The financial support of the Technion is also 
616: gratefully acknowledged. \\ 
617: %\end{Acknowledgments}
618: 
619: \begin{thebibliography}{99}
620: \bibitem{BW1} R.J. Baxter and F.Y. Wu, Phys. Rev. Lett. {\bf 31}, 1294 (1973)
621: \bibitem{BW2} R.J. Baxter and F.Y. Wu, Aust. J. Phys. {\bf 27}, 357 (1974)
622: \bibitem{BaxterSykesWatts} M.G. Watts, J. Phys A {\bf 7}, L85 (1974);
623: M.F. Sykes and M.G. Watts. {\em ibid} {\bf 8}, 1469 (1975);
624: R.J. Baxter, M.F. Sykes and M.G. Watts, {\em ibid} {\bf 8} 245 (1975)
625: \bibitem{Griffiths} H.P. Griffiths and D.W. Wood, J. Phys. C {\bf 6}, 2533 (1973);
626: D.W. Wood and H.P. Griffiths, {\em ibid} {\bf 7}, 1417 (1974)
627: \bibitem{Essam} J.W. Essam and M.E. Fisher, J. Chem. Phys. {\bf 38}, 802 (1963)
628: \bibitem{Rushbrooke} G.S. Rushbrooke, J. Chem. Phys. {\bf 39}, 842 (1963)
629: \bibitem{Braathen} H.J. Braathen and P.C. Hemmer, Phys. Norv. {\bf 8}, 69 (1975)
630: \bibitem{Imbro} D. Imbro and P.C. Hemmer, Phys. Lett. {\bf 57A}, 297 (1976)
631: \bibitem{Nijs} M.P.M. den Nijs, A.M.M. Pruisken and J.M.J. Van Leeuwen, Physica {\bf 84A}, 539 (1976)
632: \bibitem{Joyce} G.S. Joyce, Proc. R. Soc. Lond. A {\bf 345} 277 (1975)
633: \bibitem{adst} J. Adler and D. Stauffer, Physica A {\bf 181} 396 (1992)
634: \bibitem{Roder} A. Roder, J. Adler and W. Janke, Phys. Rev. Lett. {\bf 80}, 4697 (1998)
635: \bibitem{Shalaev} B.N. Shalaev, Sov. Phys. Solid State {\bf 26}, 1811 (1984)
636: \bibitem{Shankar} R. Shankar, Phys. Rev. Lett. {\bf 58}, 2466 (1987)
637: \bibitem{Ludwig} A.W.W. Ludwig, Phys. Rev. Lett. {\bf 61}, 2388 (1988)
638: \bibitem{aharony} W. Kinzel, E. Domany and A. Aharony, J. Phys. A {\bf 14} L417 (1981)
639: \bibitem{Domany} E. Domany and E. K. Riedel, J. Appl. Phys. {\bf 49}, 1315 (1978)
640: \bibitem{Novotny} M.A. Novotny and D.P. Landau, Phys. Rev. B {\bf 24}, 1468 (1981)
641: \bibitem{Harris} A.B. Harris, J. Phys. C {\bf 7}, 1671 (1974)
642: \bibitem{Evertz} M.A. Novotny and H.G. Evertz, {\em Computer Simulations in Condenced Matter Physics 6}, 188 (1993)
643: \bibitem{fried} H. Fried, J. Phys. A {\bf 25}, 2545 (1992)
644: \bibitem{adler} J. Adler, Physica A {\bf 177} 45 (1992)
645: \bibitem{WL1} F. Wang and D.P. Landau, Phys. Rev. Lett. {\bf 86}, 2050 (2001) 
646: \bibitem{WL2} F. Wang and D.P. Landau, Phys. Rev. B {\bf 64}, 056101 (2001)
647: \bibitem{Challa} Murty S.S. Challa, D.P. Landau and K.Binder, Phys. Rev. B {\bf 34}, 1841 (1986)
648: \bibitem{histogram2} W. Janke, Phys. Rev. B {\bf 47}, 14757 (1993)
649: \bibitem{Beale} Paul D. Beale, Phys. Rev. Lett. {\bf 76}, 78 (1996)
650: \bibitem{Metropolis} N. Metropolis, A.W. Rosenbluth, M.N. Rosenbluth, A.M. Teller and E. Teller, J. Chem Phys. {\bf 21}, 1087 (1953)
651: \bibitem{Swendsen} R.H. Swendsen and J.S. Wang, Phys. Rev. Lett. {\bf 58}, 86 (1987)
652: \bibitem{Wolff} U. Wolff,  Phys. Rev. Lett. {\bf 62}, 361 (1989)
653: \bibitem{histogram1} A.M. Ferrenberg and D.P. Landau,
654: Phys. Rev. B {\bf 44}, 5081 (1991)
655: \bibitem{deOliveira} P.M.C. de Oliveira, T.J.P. Penna, and H.J. Hermann, Eur. Phys. J. B {\bf 1}, 205 (1998)
656: \bibitem{Lee} J. Lee, Phys. Rev. Lett. {\bf 68}, 9 (1992)
657: \bibitem{Onsager} L. Onsager, Phys. Rev. {\bf 65}, 117 (1944)
658: \bibitem{Kaufman} B. Kaufman, Phys. Rev. {\bf 76}, 1232 (1949)
659: \bibitem{Selke} W. Selke, L.N. Shchur and O.A. Vasilyev, Physica A {\bf 259}, 338 (1998)
660: \bibitem{deSouza} A.J.F. de Souza and F.G. Moreira, Europhys. Lett. {\bf 17}, 491 (1992)
661: \bibitem{Stauffer} D. Stauffer and A. Aharony, {\em Introduction to Percolation Theory} (1994)
662: \bibitem{Kosterlitz} J. Lee and J.M. Kosterlirz, Phys. Rev. Lett. {\bf 65}, 137 (1992)
663: \bibitem{Binder1} K. Binder, Phys. Rev. Lett. {\bf 47}, 693 (1981)
664: \bibitem{Binder2} K. Binder, Z. Phys. B {\bf 43}, 119 (1981)
665: 
666: 
667: 
668: \end{thebibliography}
669: 
670: 
671: 
672: 
673: 
674: \end{document}
675: 
676: