1: %% 04 JAN 2006 :: revised version after referee reports
2: %% 24 NOV 2006 :: submitted to JCP
3:
4: %%\documentclass[prb,preprint,showpacs]{revtex4}
5: \documentclass[prb,twocolumn,showpacs]{revtex4}
6: \usepackage{graphicx,amsmath}
7:
8: %%%%%%%%%%%%%%%%%%%%
9: %% PRIVATE MACROS %%
10: %%%%%%%%%%%%%%%%%%%%
11:
12: \newcommand{\figwidth}{0.95\columnwidth}
13: \newcommand{\eq}[1]{Eq.(\ref{#1})}
14: \newcommand{\fig}[1]{Fig.~\ref{#1}}
15: \newcommand{\avg}[1]{ {\langle #1 \rangle} }
16: \newcommand{\olcite}[1]{Ref.~\onlinecite{#1}}
17:
18: \newcommand{\rhocr}{\rho_{\rm cr}}
19: \newcommand{\rhocra}{\rho_{\rm A,cr}}
20:
21: \newcommand{\na}{N_{\rm A}}
22: \newcommand{\nb}{N_{\rm B}}
23: \newcommand{\za}{z_{\rm A}}
24: \newcommand{\zb}{z_{\rm B}}
25:
26: \newcommand{\zbcr}{z_{\rm B,cr}}
27: \newcommand{\pc}{P_L(\na,\nb|\za,\zb)}
28:
29: \newcommand{\rl}{\rho_{\rm L}}
30: \newcommand{\rv}{\rho_{\rm V}}
31:
32: %% main results of this paper
33:
34: \newcommand{\ZBCR}{0.93791}
35: \newcommand{\RCR}{0.7486 \pm 0.0002}
36:
37: %%%%%%%%%%%%%%%%%%%%%
38: %% DOC STARTS HERE %%
39: %%%%%%%%%%%%%%%%%%%%%
40:
41: \begin{document}
42:
43: \title{Critical behavior of the Widom-Rowlinson mixture: coexistence
44: diameter and order parameter}
45:
46: \author{R. L. C. Vink}
47: \affiliation{Institut f\"ur Theoretische Physik II, Heinrich Heine
48: Universit\"at D\"usseldorf, Universit\"atsstra{\ss}e 1, 40225
49: D\"usseldorf, Germany}
50:
51: \date{\today}
52:
53: \begin{abstract}
54:
55: The critical behavior of the Widom-Rowlinson mixture [J.~Chem.~Phys.~{\bf
56: 52}, 1670 (1970)] is studied in $d=3$ dimensions by means of grand
57: canonical Monte Carlo simulations. The finite size scaling approach of
58: Kim, Fisher, and Luijten [Phys.~Rev.~Lett.~{\bf 91}, 065701 (2003)] is
59: used to extract the order parameter and the coexistence diameter. It is
60: demonstrated that the critical behavior of the diameter is dominated by a
61: singular term proportional to $t^{1-\alpha}$, with $t$ the relative
62: distance from the critical point, and $\alpha$ the critical exponent of
63: the specific heat. No sign of a term proportional to $t^{2\beta}$ could be
64: detected, with $\beta$ the critical exponent of the order parameter,
65: indicating that pressure-mixing in this model is small. The critical
66: density is measured to be $\rho \sigma^3 = \RCR$, with $\sigma$ the
67: particle diameter. The critical exponents $\alpha$ and $\beta$, as well as
68: the correlation length exponent $\nu$, are also measured and shown to
69: comply with $d=3$ Ising criticality.
70:
71: \end{abstract}
72:
73: %% 64.60.Fr Equilibrium properties near critical points, critical exponents
74: %% 05.70.Jk Critical point phenomena
75: %% 02.70.-c Computational techniques; simulations
76: %% 64.70.Fx Liquid-vapor transitions
77: %% 61.20.Ja Computer simulation of liquid structure
78: %% 64.75.+g Solubility, segregation, and mixing; phase separation
79:
80: \pacs{02.70.-c, 05.70.Jk, 64.70.Fx, 64.60.Fr}
81:
82: \maketitle
83:
84: \section{Introduction}
85:
86: The Widom-Rowlinson (WR) mixture \cite{widom.rowlinson:1970} is a simple
87: model of a fluid exhibiting phase separation. The model consists of A and
88: B particles that interact via simple pair potentials: the AA and BB pair
89: interaction is ideal, while AB pairs interact via a hard-core potential of
90: diameter $\sigma$. Upon increasing density, the WR mixture phase separates
91: into an A-rich and B-rich phase. In $d=3$ dimensions, computer simulations
92: agree that the corresponding universality class is that of the
93: three-dimensional (3D) Ising model \cite{shew.yethiraj:1996,
94: johnson.gould.ea:1997, gozdz:2005, buhot:2005}. There is, however, some
95: variation in the reported estimates of the critical density.
96:
97: In order to probe fluid criticality using computer simulation, high
98: quality data are required. The latter are typically generated using Monte
99: Carlo (MC) methods, and considerable effort has been devoted to develop
100: efficient MC schemes. In \olcite{johnson.gould.ea:1997}, for example, a MC
101: cluster move is described for the WR mixture that is (nearly) free of
102: critical slowing down. However, as pointed out in \olcite{buhot:2005},
103: this type of move cannot be used to obtain the coexistence curve, which
104: may be problematic if one is interested in measuring, say, the critical
105: exponent $\beta$ of the order parameter. In \olcite{buhot:2005},
106: therefore, a different cluster move is formulated, based on
107: \olcite{buhot.krauth:1998}, which not only gives access to the coexistence
108: curve, but is also rejection free. Recently, the latter approach was
109: generalized to continuous potentials \cite{liu.luijten:2004}.
110:
111: However, in addition to efficient MC sampling, of at least equal importance (if
112: not more) is the finite size scaling (FSS) algorithm used to extrapolate the
113: simulation data to the thermodynamic limit. FSS is essential because the
114: correlation length diverges at the critical point, and thus the true
115: thermodynamic limit is never captured in a finite simulation box, no matter how
116: efficiently it is simulated. For fluids, recently proposed {\it unbiased} FSS
117: algorithms formulated in the grand canonical ensemble seem particularly powerful
118: \cite{orkoulas.fisher.ea:2001, kim.fisher.ea:2003, kim.fisher:2004, kim:2005}.
119: The latter algorithms are unbiased in the sense that no prior knowledge of the
120: universality class is required: the critical point of the transition, as well as
121: some of the critical exponents, are an output. These unbiased algorithms were
122: used, for example, to resolve the universality class of the hard-core
123: square-well (HCSW) fluid and the restricted primitive electrolyte, both of which
124: were shown to exhibit 3D Ising critical behavior \cite{luijten.fisher.ea:2002,
125: kim.fisher.ea:2003}.
126:
127: Unfortunately, it is not obvious how the MC cluster moves for the WR mixture
128: generalize to the grand canonical ensemble. Grand canonical cluster moves for
129: mixtures seem less common, but some have been presented in the literature
130: \cite{orkoulas.panagiotopoulos:1994, vink.horbach:2004*1}. Of these, the MC move of
131: \olcite{vink.horbach:2004*1}, a generalization of its canonical variant
132: \cite{biben.bladon.ea:1996}, is readily applicable to the WR mixture. By using the
133: MC move of \olcite{vink.horbach:2004*1}, a FSS analysis of the WR mixture using the
134: above mentioned unbiased algorithms thus becomes possible. This, consequently, is
135: the aim of the present work. Of particular interest is the coexistence diameter,
136: whose critical behavior is governed by a very weak singularity that is challenging
137: to extract from simulation data. Note that, at the time of writing, the approach of
138: \olcite{kim.fisher.ea:2003} seems to be the only FSS algorithm available to extract
139: the coexistence diameter correctly from simulation data. A correct description of
140: the latter is required in order to reliably estimate the critical density
141: \cite{kim.fisher:2004*b}. Since the coexistence diameter of the WR model has not
142: received much attention in previous simulations, the present grand canonical
143: approach is certainly warranted.
144:
145: \section{Simulation method}
146:
147: In the grand canonical ensemble, the volume $V$ and the fugacity $\za$
148: ($\zb$) of species A (B) are fixed, while the particle numbers $\na$ and
149: $\nb$ fluctuate. The thermal wavelength is set to unity such that the
150: fugacity $z_\alpha$ directly reflects the number density $N_\alpha / V$ a
151: pure phase of $\alpha$ particles would have (recall that such a phase is
152: simply an ideal gas). In what follows, all remaining length scales are
153: expressed in terms of the hard-core diameter $\sigma$. The crucial
154: quantity is the finite-size grand canonical distribution $\pc$, defined as
155: the probability of observing a system containing $\na$ particles of
156: species A and $\nb$ particles of species B, at fugacities $\za$ and $\zb$,
157: with $L$ the lateral dimension of the cubic simulation box (the use of
158: periodic boundary conditions is assumed).
159:
160: The distribution is obtained numerically in a grand canonical MC
161: simulation via the insertion and removal of particles. To simulate
162: efficiently, the cluster move of \olcite{vink.horbach:2004*1} is used;
163: $\pc$ is then obtained simply by maintaining a histogram. To overcome the
164: free energy barrier separating the phases, a biased sampling scheme is
165: also implemented \cite{virnau.muller:2004}. Here, the simulation is
166: divided into distinct intervals (called windows) each spanning a single A
167: particle, while $\nb$ is allowed to fluctuate freely inside each window
168: (the choice for A or B is arbitrary). The windows are then sampled
169: separately and successively. CPU time is divided such that the number of
170: ``sweeps'' per window is the same for all windows. In this work, we say
171: that a sweep has passed when a given population of particles has
172: completely been replaced or updated by new ones. This is in contrast to
173: the more common approach of keeping the number of {\it attempted} MC moves
174: per window fixed. The latter approach, however, is less appropriate for
175: grand canonical simulations since the acceptance rate is typically density
176: dependent. Per window, approximately 1800 sweeps are generated. To obtain
177: a single distribution, an investment of around 7 CPU hours for a small
178: system ($L=8$), and 270 hours for a large system ($L=13$) is required. In
179: order to perform the subsequent FSS analysis, $\pc$ is measured for system
180: sizes $L=8-13$ at fugacities ranging from close to the critical point to
181: well into the coexistence region. Estimates of properties at intermediate
182: fugacities are obtained using the multiple histogram method
183: \cite{ferrenberg.swendsen:1989}.
184:
185: \section{Results}
186:
187: As mentioned before, the WR mixture exhibits phase separation into an
188: A-rich and B-rich phase. Note that the B-rich phase may equally well be
189: regarded as being poor in A species. In this sense, then, phase separation
190: is analogous to liquid-vapor coexistence: the A-rich phase being the
191: liquid, the A-poor phase being the vapor, and the fugacity of the B
192: particles being inverse temperature (again, the choice for A or B is
193: arbitrary). A natural definition of the order parameter is therefore
194: $\Delta \equiv (\rl-\rv)/2$, with $\rl$ the number density of A particles
195: in the A-rich phase, and $\rv$ the number density of A particles in the
196: A-poor phase. Close to the critical point, the order parameter is expected
197: to scale as $\Delta \propto t^\beta$, with $t = \zb/\zbcr - 1$ the
198: distance from the critical point, and $\zbcr$ the critical ``inverse
199: temperature''. Similarly, the coexistence diameter can be written as $D
200: \equiv (\rl+\rv)/2$. The critical behavior of the latter is given by
201: \cite{kim.fisher.ea:2003*b}
202: %%
203: \begin{equation}\label{eq:coex}
204: D = \rhocra \left( 1 + A_{2\beta} t^{2\beta}
205: + A_{1-\alpha} t^{1-\alpha} + A_1 t \right),
206: \end{equation}
207: %%
208: with $\rhocra$ the number density of A particles at the critical point,
209: and non-universal amplitudes $A_i$. For the 3D Ising universality class,
210: appropriate exponent values are $\beta \approx 0.326$ and $\alpha \approx
211: 0.109$ \cite{fisher.zinn:1998}.
212:
213: \subsection{Order parameter}
214:
215: To extract the order parameter, the FSS algorithm of \olcite{kim.fisher.ea:2003}
216: is used (for a more detailed description of the algorithm,
217: \olcite{kim.fisher:2004} is also highly recommended). The algorithm requires as
218: input the grand canonical distribution $\pc$ for at least three system sizes
219: $L$. Here, five system sizes $L=9,10,11,12,13$ are in fact used. Starting with
220: $\zb$ significantly above its critical value, the cumulant ratio $\avg{m^2}^2 /
221: \avg{m^4}$ is plotted as function of the average number density $\avg{\na}/V$,
222: with $m=\na-\avg{\na}$ (note that this plot is parameterized by the fugacity of
223: the other species $\za$). The resulting curve will reveal two minima, located at
224: $\rho^-$ and $\rho^+$, with respective values $Q^-$ and $Q^+$ at the minima.
225: Defining the quantities $Q_{\rm min} = (Q^+ + Q^-)/2$, $x = Q_{\rm min} \ln (4/e
226: Q_{\rm min} )$, and $y = (\rho^+ - \rho^-) / (2 \Delta)$, the points $(x,y)$
227: from the different system sizes should, in the limit far away from the critical
228: point, collapse onto the line $y=1+x/2$. Recall that $\Delta$ is the order
229: parameter in the thermodynamic limit at the considered fugacity $\zb$, precisely
230: the quantity of interest, which may thus be obtained by fitting until the best
231: collapse onto $1+x/2$ occurs. In the next step, $\zb$ is chosen closer to the
232: critical point, the points $(x,y)$ are calculated as before, but this time
233: $\Delta$ is chosen such that the new data set joins smoothly with the previous
234: one, yielding an estimate of the order parameter at the new fugacity. This
235: procedure is repeated all the way to the critical point, where $\Delta$
236: vanishes, leading to an estimate of the critical fugacity $\zbcr$. Moreover, the
237: procedure also yields $y$ as function of $x$. The latter scaling function is
238: universal within a universality class, and for the HCSW fluid can be found in
239: \olcite{kim.fisher.ea:2003}. Since the WR mixture belongs to the same
240: universality class, a similar curve should be found. The latter is verified in
241: \fig{scaling}, which shows $y$ as function of $x$ obtained in this work,
242: compared to the result of \olcite{kim.fisher.ea:2003}. The agreement is very
243: reasonable. From the vanishing of the scaling function, at $x_c = 0.280$, an
244: unbiased estimate of the critical fugacity $\zbcr = \ZBCR \pm 0.00004$ is
245: obtained. Note that $x_c$ is universal within a universality class. The estimate
246: reported here compares favorably to $x_c=0.286$ obtained for the HCSW fluid
247: \cite{kim.fisher.ea:2003}, and $x_c=0.296$ obtained for the 3D Ising model
248: \cite{tsypin.blote:2000}, providing additional confirmation that these systems
249: belong to the same universality class. Shown in \fig{order}, on double
250: logarithmic scales, is the order parameter $\Delta$ of the WR mixture as
251: function of the distance from the critical point $t$, where the above quoted
252: estimate of $\zbcr$ was used. The resolution of the present data is such that
253: $\Delta$ can be resolved down to $t \approx 5 \times 10^{-5}$. By fitting the
254: lowest few points in \fig{order} to $\Delta \propto t^\beta$, the critical
255: exponent is {\it measured} to be $\beta \approx 0.322 \pm 0.008$, which is
256: certainly compatible with the accepted 3D Ising value.
257:
258: \subsection{Coexistence diameter}
259:
260: To extract the coexistence diameter, the FSS algorithm of \olcite{kim:2005} is
261: used. The algorithm is similar in spirit to the previous one, in the sense that
262: it generates a scaling function $y=f(x)$, starting with data obtained well away
263: from the critical point, and then recursively working its way down toward
264: criticality. In \fig{sc-diam}, the scaling function of the diameter for the WR
265: mixture thus obtained is shown, where, as before, five system sizes $L=9-13$
266: were used. For $x \to 0$, this function is expected to approach $y=x/2$, which
267: indeed it does. In contrast to the order parameter, however, the scaling
268: function of the coexistence diameter is {\it not} universal \cite{kim:2005}.
269: Therefore, a direct comparison to scaling functions of other systems cannot, in
270: general, be carried out. Nevertheless, for systems with negligible pressure
271: mixing, such as the HCSW fluid and presumably also the WR mixture, the scaling
272: function is expected to be well described by the approximant \cite{kim:2005}
273: %%
274: \begin{eqnarray}\label{eq:app}
275: e_l(x) &=& C_l \left[ 1 - (1-\bar{x})^{1-\alpha} \right. \\
276: && \times \left. \frac{ 1+s_1 \bar{x} + s_2 \bar{x}^2 + s_3 \bar{x}^3 }{
277: 1 + t_1 \bar{x} + t_2 \bar{x}^2 + t_3 \bar{x}^3} \right], \nonumber
278: \end{eqnarray}
279: %%
280: with $\bar{x}=x/x_c$, $t_1 = s_1 - 1 + \alpha + x_c/2C_l$, and critical exponent
281: $\alpha \approx 0.109$. A fit to the WR data of \fig{sc-diam} shows that this is
282: indeed the case, with explicit parameter values $C_l = 0.429$, $x_c = 0.175$,
283: $s_1 = 4.50$, $s_2 = -5.72$, $s_3=0.12$, $t_1=3.81$, $t_2 = -9.08$ and
284: $t_3=4.25$. These values are remarkably consistent with estimates quoted in
285: \olcite{kim:2005} for the HCSW fluid. Note that $e_l(x)$ becomes singular
286: close to $x_c$, implied by the $(1 - \bar{x})^{1 - \alpha}$ factor in
287: \eq{eq:app}. The latter would yield a vertical tangent in the plot, at the arrow
288: in \fig{sc-diam}. The present simulation data, however, seem not to extend close
289: enough to the critical point to reach this regime.
290:
291: The critical behavior of the coexistence diameter $D$ is shown in
292: \fig{fit_diam}, where $\zbcr = \ZBCR$ obtained in the previous paragraph
293: was used. In order to facilitate the comparison to other work, $2 \times
294: D$ is actually plotted. Symmetry considerations ensure equal numbers of A
295: and B particles at criticality, such that the {\it overall} critical
296: number density equals $\rhocr = 2 \rhocra$, which is the quantity usually
297: quoted in the literature. A fit to the asymptotic expansion of
298: \eq{eq:coex} yields $\rhocr = \RCR$, where the error reflects the
299: variation stemming from the range over which the fit is performed
300: (repeating the entire analysis leaving out the smallest system size yields
301: a similar result). The corresponding amplitudes read as $A_{2\beta}
302: \approx 0$, $A_{1-\alpha} = 2.76 \pm 0.07$ and $A_1 \approx -1.27 \pm
303: 0.09$, implying that the singular behavior is dominated by $t^{1-\alpha}$.
304: This, in combination with the observation that the scaling function is
305: well described by $e_l(x)$, confirms that pressure mixing in the WR
306: mixture is small. The curvature of the diameter close to the critical
307: point thus reflects the $t^{1-\alpha}$ singularity. Since $1-\alpha$ is
308: close to unity, and the magnitudes of $A_{1-\alpha}$ and $A_1$ are
309: similar, the curvature is hard to see in \fig{fit_diam}. The singular
310: behavior of the diameter can be visualized nevertheless by plotting the
311: ``inverse temperature'' derivative $\kappa = 2 \, d D/d t$ instead. In
312: case of singular behavior, $\kappa$ is expected to diverge when $t \to 0$,
313: see \eq{eq:coex}. Though not very precise, this procedure even allows for
314: an unbiased measurement of the exponent $\alpha$. The result is summarized
315: in the inset of \fig{fit_diam}, which shows $\kappa$ as function of $t$,
316: where again $\zbcr = \ZBCR$ in $t$ was used. The divergence is clearly
317: visible. By fitting the data to the form $\kappa = a_1 t^{-\alpha} + a_2$,
318: with fit parameters $a_i$ and $\alpha$, the specific heat exponent is
319: measured to be $\alpha = 0.11 \pm 0.02$, which is surprisingly close to
320: the 3D Ising value.
321:
322: \subsection{Cumulant intersections}
323:
324: For the sake of completeness, and also to check the consistency of the results
325: obtained so far, the critical fugacity $\zbcr$ is measured again, but this time
326: around using the cumulant intersection approach \cite{binder:1981}. As was shown
327: by Binder \cite{binder:1981}, the (for example) first order cumulant $U_1 =
328: \avg{m^2} / \avg{|m|}^2$ becomes system-size independent at the critical point.
329: Plots of $U_1$ as function of $\zb$ for different system sizes are thus expected
330: to show a common intersection point, leading to an unbiased estimate of the
331: critical fugacity. Moreover, the cumulant value $Q_c$ at the intersection point
332: is universal, dependent only on the universality class. Shown in \fig{cumulant}
333: is the result of this procedure, on a rather fine scale. The resulting estimate
334: reads as $\zbcr = 0.9379 \pm 0.0004$, where the error reflects the scatter in
335: the intersection points. The latter is fully consistent with the previous, more
336: precise value, $\zbcr = \ZBCR \pm 0.00004$ (arrow in \fig{cumulant}). For the
337: critical value of the cumulant $Q_c \approx 1.223$ is obtained. This value
338: compares quite favorably to the estimate $Q_c = 1.2391 \pm 0.0014$ obtained in
339: large-scale simulations of the 3D Ising lattice model \cite{luijten:1999},
340: deviating from it by less than 2\%. The inset of \fig{cumulant} shows the slope
341: of the cumulant $Y_1 = | {\rm d} U_1 / {\rm d} \zb |$ at the critical fugacity,
342: as function of the system size $L$. It is expected that $Y_1 \propto L^{1/\nu}$
343: with $\nu$ the critical exponent of the correlation length. Although there is
344: some scatter in the intersection points, the cumulant slopes seem rather
345: constant over the range of \fig{cumulant}, and so it is expected that $\nu$ can
346: be obtained quite reliably nevertheless. Indeed, by performing a fit to the data
347: in the inset of \fig{cumulant}, the exponent is measured to be $\nu \approx
348: 0.630 \pm 0.005$, in excellent agreement with the accepted 3D Ising value
349: $\nu_{\rm Is} \approx 0.630$ \cite{fisher.zinn:1998}.
350:
351: \section{Discussion and Summary}
352:
353: In this work, the critical behavior of the WR mixture was investigated
354: using grand canonical MC simulations and unbiased FSS algorithms. As
355: expected, the universality class of the transition is that of the 3D Ising
356: model. This was demonstrated by direct measurements of the exponents $\alpha$,
357: $\beta$ and $\nu$. Substantial indirect evidence has also been provided,
358: by comparing the scaling function of the order parameter and the
359: coexistence diameter to those of the HCSW fluid, as well as via the
360: universality of $Q_c$ at the cumulant intersection point.
361:
362: The critical density obtained in this work $\RCR$ can be compared to other
363: simulations. Shew and Yethiraj report $0.762 \pm 0.016$ \cite{shew.yethiraj:1996}
364: using semigrand simulations \cite{deutsch.binder:1992, miguel.rio.ea:1995}, while
365: Johnson {\it et al.}~(JEA) obtained $0.748 \pm 0.002$ \cite{johnson.gould.ea:1997}.
366: More recent estimates are due to G\'o\'zd\'z, $0.759 \pm 0.019$ \cite{gozdz:2005}
367: using the Bruce-Wilding field mixing technique \cite{bruce.wilding:1992}, and Buhot
368: $0.7470 \pm 0.0008$ \cite{buhot:2005}. Of the above, JEA and Buhot are very close
369: to the value reported here, see \fig{fit_diam}. The estimate of G\'o\'zd\'z,
370: however, is higher. A possible explanation is the adapted FSS algorithm in the
371: latter, which seems to overestimate the critical density in some cases
372: \cite{kim.fisher:2004*b}. JEA also report an estimate of the critical fugacity
373: $\zbcr \approx 0.9403$ for the largest system size considered by them
374: \cite{johnson.gould.ea:1997}, but without a systematic FSS analysis of this
375: quantity. This overestimates the present value significantly. Interestingly, these
376: authors observe an {\it increase} of $\zbcr$ with system size, in disagreement with
377: the present work.
378:
379: Buhot, by using rejection-free cluster MC moves, is able to simulate
380: impressively large systems \cite{buhot:2005}, up to $L=100$, which exceeds
381: the typical system size of the present investigation by about one order of
382: magnitude. The improved accuracy of the critical density obtained in this
383: work may therefore seem surprising. It should be emphasized, however, that
384: critical phenomena are most conveniently studied in terms of a field
385: variable, such as temperature or, in the case of the WR mixture, the
386: fugacity. The ensemble used by Buhot, as well as the semigrand ensemble,
387: do not have access to the fugacity. Instead, in these ensembles, the
388: critical point is approached by varying the overall density
389: $\rho=(\na+\nb)/V$. This somewhat restricts the investigation of critical
390: phenomena because for every considered $\rho$, an explicit simulation
391: needs to be carried out. In grand canonical simulations, on the other
392: hand, one has access to the particle fugacities. This facilitates the {\it
393: extrapolation} of simulation data obtained at one set of fugacities to
394: different values via histogram reweighting
395: \cite{ferrenberg.swendsen:1989}. Clearly, the investigation of subtle
396: effects, such as the critical behavior of the coexistence diameter, is not
397: really feasible without such extrapolation methods. Note that the behavior
398: depicted in \fig{fit_diam} is not simply an ``artifact'' of the grand
399: canonical ensemble. In the semigrand ensemble, for example, the singular
400: behavior of the diameter leads to a renormalization of the critical
401: exponent $\beta$ \cite{fisher:1968}. Assuming negligible $A_{2\beta}$ in
402: \eq{eq:coex}, one obtains $\rho/\rhocr-1 \propto t^{1-\alpha}$ close to
403: the critical point. Combining this with the critical power law of the
404: order parameter $\Delta \propto t^\beta$ and eliminating $t$, yields
405: $\Delta \propto (\rho/\rhocr-1)^{\beta^\star}$, with renormalized exponent
406: $\beta^\star = \beta/(1-\alpha)$. The latter renormalized exponent has
407: been confirmed experimentally \cite{chen.payandeh:2000}, and should, in
408: principle, also show up in the WR mixture when, as mentioned above, the
409: critical point is approached by varying $\rho$.
410:
411: Needless to say, the WR mixture has also been studied by theoretical
412: means, using for example density functional theory \cite{schmidt:2001},
413: and integral equations \cite{shew.yethiraj:1996, yethiraj.stell:2000}.
414: These investigations, however, all pertain to the mean-field level. As
415: such, quantitative agreement with computer simulations close to
416: criticality is not to be expected, and a comparison is consequently not
417: carried out.
418:
419: \acknowledgments
420:
421: This work was supported by the {\it Deutsche Forschungsgemeinschaft} under
422: the SFB-TR6 (project section D3). I also thank K.~Binder for an initial
423: reading of the manuscript.
424:
425: \bibstyle{revtex}
426: \bibliography{mainz}
427:
428: %%%%%%%%%%%%%%%%%%%%%
429: %% FIGURE CAPTIONS %%
430: %%%%%%%%%%%%%%%%%%%%%
431:
432: \newpage
433: \section{Figure captions}
434:
435: FIG. 1: Scaling function of the order parameter. Following the convention of
436: \olcite{kim.fisher.ea:2003}, the scaling function is raised to a negative
437: exponent, with $\phi=1/\beta$ and $\beta=0.326$. The solid curve is the result
438: obtained in this work for the WR mixture; the dashed curve is the HCSW result of
439: \olcite{kim.fisher.ea:2003}. Also shown is the exact small $x$ limiting form $y
440: = 1 + x/2$.
441:
442: FIG. 2: Order parameter of the WR mixture as function of the
443: distance from the critical point. The dashed line has a slope
444: $\beta=0.326$, corresponding to the 3D Ising exponent.
445:
446: FIG. 3: Scaling function of the coexistence diameter for
447: the WR mixture. Open circles show simulation results obtained using the
448: FSS algorithm of \olcite{kim:2005}. The dashed curve is a fit to the
449: simulation data using the approximant of \eq{eq:app}. Also shown is the
450: exact small $x$ limiting form $y=x/2$.
451:
452: FIG. 4: Coexistence diameter of the WR mixture as
453: function of the distance from the critical point. Open circles are
454: simulation results obtained in this work using the FSS algorithm of
455: \olcite{kim:2005}. The dashed curve is a fit to \eq{eq:coex}. The black
456: dot marks the critical density obtained from the fit, where the vertical
457: line indicates the uncertainty. The arrows mark estimates of $\rhocr$
458: reported in \olcite{johnson.gould.ea:1997} (JEA) and \olcite{buhot:2005}
459: (Buhot), where the vertical lines again indicate the uncertainty. The
460: inset shows $\kappa$ as function of $t$. Open circles are simulation
461: results; the dashed curve, which essentially overlaps the simulation data,
462: is a three-parameter fit of the form $\kappa = a_1 t^{-\alpha} + a_2$,
463: with fit parameters $a_i$ and $\alpha$.
464:
465: FIG. 5: Cumulant analysis of the WR mixture. Shown is
466: the first order cumulant $U_1$ as function of the fugacity $\zb$ for
467: various system sizes $L$ as indicated. The inset shows the slope of the
468: cumulant $Y_1$ at the critical point as function of $L$. All data were
469: obtained along the symmetry locus $\za=\zb$.
470:
471: %%%%%%%%%%%%%
472: %% FIGURES %%
473: %%%%%%%%%%%%%
474:
475: %% FIG 1 %%
476: \newpage
477: \begin{figure}
478: \begin{center}
479: \includegraphics[clip=,width=\figwidth]{fig1}
480: \caption{\label{scaling}}
481: \end{center}
482: \end{figure}
483:
484: %% FIG 2 %%
485: \newpage
486: \begin{figure}
487: \begin{center}
488: \includegraphics[clip=,width=\figwidth]{fig2}
489: \caption{\label{order}}
490: \end{center}
491: \end{figure}
492:
493: %% FIG 3 %%
494: \newpage
495: \begin{figure}
496: \begin{center}
497: \includegraphics[clip=,width=\figwidth]{fig3}
498: \caption{\label{sc-diam}}
499: \end{center}
500: \end{figure}
501:
502: %% FIG 4 %%
503: \newpage
504: \begin{figure}
505: \begin{center}
506: \includegraphics[clip=,width=\figwidth]{fig4}
507: \caption{\label{fit_diam}}
508: \end{center}
509: \end{figure}
510:
511: %% FIG 5 %%
512: \newpage
513: \begin{figure}
514: \begin{center}
515: \includegraphics[clip=,width=\figwidth]{fig5}
516: \caption{\label{cumulant}}
517: \end{center}
518: \end{figure}
519:
520: \end{document}
521: