1: \documentclass[aps,pre,a4paper,amsmath,twocolumn,showpacs,floatfix]{revtex4}
2: \usepackage{graphicx}
3: %\usepackage{amsmath}
4: \usepackage{amssymb}
5: %\usepackage{amscd}
6: \bibliographystyle{apsrev}
7:
8: \begin{document}
9:
10: \title{Orientational ordering in hard rectangles: the role of three-body
11: correlations}
12: \author{Yuri Mart\'{\i}nez-Rat\'on}
13: \email{yuri@math.uc3m.es}
14:
15: \affiliation{Grupo Interdisciplinar de Sistemas Complejos (GISC),
16: Departamento de Matem\'aticas, Escuela Polit\'ecnica Superior,
17: Universidad Carlos III de Madrid,
18: Avenida de la Universidad 30, E-28911 Legan\'es, Madrid, Spain.
19: }
20:
21: \author{Enrique Velasco}
22: \email{enrique.velasco@uam.es}
23:
24: \affiliation{Departamento de F\'{\i}sica Te\'orica de la Materia Condensada
25: and Instituto de Ciencia de Materiales Nicol\'as Cabrera,
26: Universidad Aut\'onoma de Madrid, E-28049 Madrid, Spain.}
27:
28: \author{Luis Mederos}
29: \email{l.mederos@icmm.csic.es}
30:
31: \affiliation{Instituto de Ciencia de Materiales, Consejo Superior de
32: Investigaciones Cient\'{\i}ficas, E-28049 Cantoblanco, Madrid, Spain.}
33:
34:
35: \date{\today}
36:
37: \begin{abstract}
38: We investigate the effect of three-body correlations
39: on the phase behavior of hard rectangle two-dimensional fluids.
40: The third virial coefficient, $B_3$, is incorporated via an equation
41: of state that recovers scaled particle theory for parallel
42: hard rectangles. This coefficient, a functional of the
43: orientational distribution function, is calculated
44: by Monte Carlo integration, using an accurate parameterized distribution
45: function, for various particle aspect ratios in the range $1-25$.
46: A bifurcation analysis of the free energy calculated from the obtained
47: equation of state is applied to find the isotropic (I)-uniaxial nematic
48: (N$_u$) and isotropic-tetratic nematic (N$_t$) spinodals and to study the
49: order of these phase transitions. We find that the relative stability of the
50: N$_t$ phase with respect to the isotropic phase is enhanced by the
51: introduction of
52: $B_3$. Finally, we have calculated the complete phase diagram using a
53: variational procedure and compared the results with those obtained
54: from scaled particle theory and with Monte Carlo simulations carried
55: out for hard rectangles with various
56: aspect ratios. The predictions of our proposed equation of state as
57: regards the transition densities between the isotropic
58: and orientationally ordered
59: phases for small aspect ratios are in fair agreement with simulations.
60: Also, the critical aspect ratio
61: below which the N$_t$ phase becomes stable is predicted to increase
62: due to three-body correlations, although the corresponding value is
63: underestimated with respect to simulation.
64: \end{abstract}
65:
66: \pacs{64.70.Md,64.75.+g,61.20.Gy}
67: % 64.70.Md Transitions in liquid crystals
68: % 64.75.+g Solubility, segregation, and mixing; phase separation
69: % 61.20.Gy Theory and models of liquid structure
70:
71:
72: \maketitle
73:
74:
75: \section{Introduction}
76:
77: The (two-dimensional) hard-rectangle (HR) model has recently received some
78: attention due to the possibility that a dense film of these particles
79: exhibits spontaneous tetratic
80: order\cite{Schaklen,Martinez-Raton,Frenkel1,Donev}.
81: Additional interest originates from
82: the fact that some types of organic molecular semiconductors are made
83: of rectangularly shaped molecules; a notable example is the PTCDA molecule,
84: films of which have recently been studied quite intensely \cite{PTCDA}.
85: Even though
86: the interactions between these molecules involve high-order polar forces
87: (e.g. quadrupolar forces) it is of interest to investigate theoretically
88: the intrinsic order associated to purely excluded-volume effects with a view
89: to predicting structural and thermodynamic properties of the film by
90: incorporating other interactions via traditional perturbation theories.
91: The system we investigate in the present work mimics an incommensurate
92: film of these molecules with only excluded-volume interactions involved
93: and in the regime where molecules are free to move in the film
94: (i.e. fluid regime). Phases with two-dimensional crystalline order
95: will be left for future work.
96:
97: Recently monolayers of various macroscopically-sized particles have
98: been studied using mechanical vibrations on the monolayer to induce
99: motion \cite{Narayan}. Even though this is an athermal, non-equilibrium
100: system reaching steady-state configurations, these configurations
101: are mainly driven by packing effects and should give the trend as to
102: what types of order could be expected. In particular, tetratic order was
103: observed in particles with sufficiently sharp corners, resembling rectangles,
104: in contrast with particles such as discorectangles (projections of
105: spherocylinders on the plane) which only exhibit nematic ordering,
106: or basmati-rice grains, which in addition may have a smectic phase.
107:
108: Apart from the possible interest in modelling the behaviour of monolayers
109: made of molecules with technological interest, our investigations have the
110: additional, more fundamental aim of
111: elucidating the effect that three-body correlations have on the orientational
112: properties of two-dimensional fluids. Onsager showed that for three-dimensional
113: hard rod fluids in the limit of infinite aspect ratio (hard-needle limit),
114: $\kappa\to\infty$ (with
115: $\kappa\equiv L/\sigma$, $L$ and $\sigma$ being the length and width of the
116: constituent particles), the ratio between the third virial coefficient and
117: the second virial coefficient squared asymptotically vanishes, for the
118: isotropic fluid, as $B_3/B_2^2\sim(\sigma/L)\log(L/\sigma)$ \cite{Onsager}.
119: Taking this result into account, he used a second-order virial expansion
120: for the free energy as a functional of the orientational distribution function
121: and obtained predictions for the isotropic (I)-nematic (N) phase-transition
122: densities, exact in the above limit. By contrast, in two dimensions the
123: above ratio between virial coefficients has the approximate limiting value of
124: $0.514$ \cite{Tarjus}, implying that three-body correlations might play a very
125: important role in the isotropic fluid even in the hard-needle limit;
126: this is in sharp contrast with the three-dimensional case.
127: An investigation of the
128: effect of these high-order correlations on orientational ordering seems
129: therefore appropriate.
130:
131: Since Onsager theory does not account for higher-than-two body
132: correlations, an alternative theoretical approach is required.
133: Scaled particle theory (SPT), first developed for a mixture of hard spheres
134: \cite{Reiss} and later extended to anisotropic particles \cite{Cotter,
135: Lasher,Barboy}, includes as a main ingredient the exact analytic expression
136: for the second virial coefficient \cite{Isihara,Kihara}, but again the third
137: is approximated assuming $B_3/B^2_2\to 0$ in the hard-needle limit, an
138: assumption that is incorrect. Also, it has been shown that, for a variety
139: of particle shapes in two dimensions, the fourth and fifth virial
140: coefficients tend to negative
141: values in the same limit \cite{Tarjus,Rigby}. Thus it may very well occur
142: that in these cases the virial series exhibits poor convergence, and a
143: natural question arises: how does the phase behavior of anisotropic
144: hard convex two-dimensional fluids change when three- and higher-body
145: correlations, neither of which are included in the standard Onsager
146: and SPT approaches, are taken into account? One of the aims of
147: the present article is to shed some light on this question. For this purpose
148: we develop an equation of state (EOS) for HR which exactly
149: includes two- and three-body correlations in the nematic fluid and
150: recovers SPT in the case of perfectly aligned particles.
151:
152: Recent investigations of the HR system have used the SPT approach
153: \cite{Schaklen,Martinez-Raton} and Monte Carlo (MC) simulations
154: \cite{Frenkel1,Donev} to study its phase behavior. Aside from the
155: usual isotropic-uniaxial nematic (N$_u$) transition, these works
156: have shown that this peculiar system exhibits a continuous
157: transition between the isotropic phase and a tetratic nematic (N$_t$) phase.
158: The latter is an orientationally ordered phase but with $D_{4h}$ symmetry,
159: i.e., the system is invariant under rotation of $\pi/2$. The SPT predicts
160: that this phase is stable up to an aspect ratio $\kappa\approx 2.21$ and
161: that the packing fraction values of the I-N$_t$ transition
162: are around 0.85. This is
163: in disagreement \cite{Martinez-Raton} with the I-N$_t$ transition densities
164: obtained from simulation for $\kappa=1$ and $2$ \cite{Frenkel1,Donev}, which
165: predicts values around 0.7.
166: Considering the importance that high-body correlations may have
167: on the phase behavior of two-dimensional hard-convex bodies, it is
168: the second purpose of this article to apply our model (which
169: includes three-body correlations) to calculate the phase diagram of HR
170: and compare the results with those of SPT and MC simulations. The main features
171: of the phase diagram are calculated using bifurcation theory for the I-N
172: transition and also minimizing the nematic free energy functional
173: resulting from our proposed EOS.
174:
175: The article is organized as follows. In Section II we describe the
176: theoretical model for a general two-dimensional hard-convex fluid
177: as applied to the isotropic and nematic fluids. Section III is devoted to
178: the results obtained from the analysis of the theory, and comparison
179: is made with simulation results; also, the complete liquid-crystal
180: phase diagram is presented. Finally, some conclusions
181: are drawn in Section IV. The Appendix contains a detailed account of
182: the bifurcation analysis and the minimization method used to analyse
183: the phase behaviour of the model, together with some details on the
184: computer simulations.
185:
186: \section{Theory}
187: In the present section we introduce the theoretical formalism necessary
188: for the study of the phase behavior of the HR fluid. This includes the
189: derivation of the EOS for the isotropic and nematic fluids,
190: along with the corresponding free energy density. The phase behaviour of the
191: model, to be presented in the next section, is analysed by means of two
192: complementary techniques: a bifurcation analysis of the free energy density,
193: and a full minimization using an accurate functional form for the orientational
194: distribution function. Details on these techniques are given in the Appendix.
195:
196: \subsection{EOS for the isotropic fluid}
197:
198: In this section an equation of state for the isotropic phase, to be extended
199: later to the nematic phase, is proposed, on a somewhat ad-hoc, but at the
200: same well-founded, basis. The virial coefficients are intimately related
201: to geometric properties of the planar objects making up the fluid.
202: The second virial coefficient of planar hard particle has the analytic
203: form \cite{Kihara}
204: \begin{eqnarray}
205: B_2=v+\frac{
206: {\cal L}^2}{4\pi},
207: \label{segundo}
208: \end{eqnarray}
209: where $v$ and ${\cal L}$ are the area and perimeter of the particle.
210: Some approximate analytic expressions for third virial coefficients
211: of isotropic fluids made of three-dimensional bodies, as a function of
212: their volume, surface area and mean curvature, have been proposed
213: \cite{Boublik}. When compared with results from numerical calculations
214: \cite{Boublik}, some of these expressions are seen to constitute accurate
215: approximations. In two dimensions volume has to
216: be substituted by area, area by perimeter and mean curvature by a function
217: proportional to the perimeter (the latter is true for the most representative
218: two-dimensional convex bodies, i.e. rectangles, discorectangles,
219: and ellipses). Following some of the most successful approximations
220: \cite{Boublik} but translated to the two-dimensional case, we write the
221: following analytic expression for the third virial coefficient
222: \begin{eqnarray}
223: B_3=v^2+
224: \frac{\alpha}{4\pi}v{\cal L}^2+
225: \frac{\beta}{(4\pi)^2}{\cal L}^4
226: \label{tercero}
227: \end{eqnarray}
228: where the numerical coefficients $\alpha$ and $\beta$ are chosen in
229: such a way as to guarantee i) the correct asymptotic hard-needle limit, and
230: ii) a good comparison with well-known EOS for some isotropic fluids,
231: such as hard disks. For the latter we have ${\cal L}^2=4\pi v$ so that
232: the three terms in the right-hand side of Eqn. (\ref{tercero}) can be
233: unified into the single term $(1+\alpha+\beta)v^2$.
234: The SPT for hard disks is recovered by choosing $\alpha+\beta=2$, whereas
235: the SPT form for $B_3$ for a general anisotropic particle
236: is obtained from (\ref{tercero}) with $\alpha=2$ and $\beta=0$.
237: Note, from Eqns. (\ref{segundo}) and (\ref{tercero}), that
238: \begin{eqnarray}
239: B_2\sim \frac{{\cal L}^2}{4\pi},\quad
240: B_3\sim\frac{\beta}{(4\pi)^2}{\cal L}^4,\label{undos}
241: \end{eqnarray}
242: in the infinite aspect-ratio limit,
243: as the particle area is proportional to the product
244: of the two characteristic lengths of the particle
245: (the width $\sigma$ and the length $L$, the latter being
246: the larger one), while the perimeter is proportional to their sum.
247: Therefore we obtain the asymptotic limit
248: $B_3/B_2^2\to \beta$ when $L/\sigma\to\infty$ and
249: a sensible choice is $\beta=0.514$, the exact asymptotic value of
250: this ratio \cite{Tarjus}.
251:
252: The EOS is obtained by imposing two requirements:
253: (i) the divergence of pressure at
254: high packing fractions is of the form $\sim (1-\eta)^{-2}$ as stated by SPT,
255: and (ii) the second and third virial coefficients are obtained from the
256: exact virial expansion
257: \begin{eqnarray}
258: \beta P&=&\rho+\rho^2B_2+\rho^3B_3,\label{virial}
259: \label{lala}
260: \end{eqnarray}
261: (where $\rho$ is the density of particles). In other words, we require that
262: the third-order virial expansion of the interaction part of the EOS,
263: \begin{eqnarray}
264: \beta P_{\rm{exc}}v=\beta P v-\eta=\frac{a_2\eta^2+a_3 \eta^3}{(1-\eta)^2},
265: \end{eqnarray}
266: ($\eta=\rho v$ being the packing fraction)
267: coincides with the exact one (\ref{lala}).
268: This allows us to obtain $a_k$ ($k=1,2$) as $a_2=1+b_2$, and $a_3=b_3-2a_2-1$,
269: where the new coefficients
270: \begin{eqnarray}
271: b_k=\frac{B_k}{v^{k-1}}-1,\quad k=2,3,
272: \end{eqnarray}have been defined in terms of the virial coefficients $B_k$.
273: The resulting EOS has the form
274: \begin{eqnarray}
275: \beta Pv=\frac{\eta}{1-\eta}+\frac{\eta^2}{(1-\eta)^2}+
276: \frac{\eta^3}{(1-\eta)^2}\left(b_3-2b_2\right)
277: \label{note}
278: \end{eqnarray}
279: From this EOS the free energy density can be obtained: we first write
280: \begin{eqnarray}
281: \rho^2\frac{\partial \varphi}{\partial\rho}=\beta P(\rho),
282: \label{porco}
283: \end{eqnarray}
284: where $\varphi=\varphi_{\rm{id}}+\varphi_{\rm{ex}}$ is the free energy per
285: particle and $\varphi_{\rm{id}}$ and $\varphi_{\rm{ex}}$ the corresponding
286: ideal and excess contributions. Now using $\beta P$ from (\ref{note}),
287: Eqn. (\ref{porco}) can be integrated to give
288: \begin{eqnarray}
289: \varphi_{\rm{exc}}&=&-\ln(1-\eta)+\frac{\eta}{1-\eta} b_2+
290: (b_3-2b_2)\theta(\eta),\label{taken}\\
291: \theta(\eta)&=&\frac{\eta}{1-\eta}+\ln(1-\eta)
292: \label{exceso}
293: \end{eqnarray}
294:
295: The first two terms of (\ref{note}) and (\ref{taken}) are SPT-like terms.
296: Eq. (\ref{note}), with the exact second and third virial
297: coefficients for the particular case of parallel hard rectagles, recovers
298: the SPT result [this is easily obtained if we substitute the exact values
299: $B_2=2v$ ($b_2=1$) and $B_3=3v^2$ ($b_3=2$) in Eq. (\ref{note})].
300:
301: Inserting $B_2$ and the approximation for $B_3$ from
302: (\ref{segundo}) and (\ref{tercero}), respectively,
303: in Eqns. (\ref{note}) and (\ref{taken}),
304: we obtain our proposed EOS and the excess part of the free energy
305: per particle for the isotropic fluid as
306: \begin{eqnarray}
307: \beta Pv&=&\frac{\eta}{1-\eta}+\frac{\eta^2}{(1-\eta)^2}
308: \gamma\left[1+(\alpha-2+\beta\gamma)\eta\right],\label{state}\\
309: \varphi_{\rm{ex}}&=&-\ln(1-\eta)+\frac{\gamma\eta}{1-\eta}
310: +\gamma(\alpha-2+\beta\gamma)\theta(\eta),\nonumber\\
311: \end{eqnarray}
312: where
313: the anisometric parameter $\gamma={\cal L}^2/(4\pi v)$ was
314: defined. Note that
315: for $\alpha=2$, $\beta=0$, this equation recovers the SPT expression
316: for hard convex bodies.
317: From (\ref{state})
318: the following expression for the reduced virial coefficients is obtained:
319: \begin{eqnarray}
320: B_n^*\equiv\frac{B_n}{B_2^{n-1}}=\frac{1+\left[1+(\alpha-1)(n-2)\right]
321: \gamma+\beta(n-2)\gamma^2}{(1+\gamma)^{n-1}}\nonumber\\
322: \label{bast}
323: \end{eqnarray}
324:
325: \subsection{EOS for the nematic fluid}
326:
327: The EOS for the nematic fluid is now obtained from
328: Eqn. (\ref{note}) by substituting the virial coefficients of the
329: isotropic fluid $B_{n}$ ($n=2,3$) by their functional versions
330: $B_{n}[h]$ for the nematic fluid; here $h(\phi)$ is the orientational
331: distribution function.
332: The latter coefficients are obtained from the definitions of $B_2$ and $B_3$,
333: in terms of integrals over the Mayer function:
334: \begin{eqnarray}
335: &&B_k[h]=\frac{1}{k}\left[\prod_{l=1}^k\int d\phi_lh(\phi_l)\right]
336: {\cal K}(\phi_1,\dots,\phi_k),\\
337: &&{\cal K}(\phi_1,\phi_2)=-\int d{\bf r}f({\bf r},\phi_{12}),\\
338: &&{\cal K}(\phi_1,\phi_2,\phi_3)=-
339: \int d{\bf r}\int d{\bf r}'f({\bf r},\phi_{12})f({\bf r}',\phi_{23})
340: \nonumber \\&&\times f({\bf r}-{\bf r}',\phi_{13}),\label{kernel3}
341: \end{eqnarray}
342: where $\phi_{\alpha\beta}=\phi_{\alpha}-\phi_{\beta}$ is the relative
343: angle between axes of particles $\alpha$ and $\beta$, and
344: $f({\bf r},\phi_{\alpha\beta})$ the Mayer function. The corresponding
345: free-energy functional $\varphi[h]=\varphi_{\rm{id}}[h]+
346: \varphi_{\rm{ex}}[h]$ is obtained from Eqns. (\ref{taken}):
347: \begin{eqnarray}
348: \varphi_{\rm{exc}}[h]=-\ln(1-\eta)+\frac{\eta}{1-\eta} b_2[h]+
349: \left(b_3[h]-2b_2[h]\right)\theta(\eta)\nonumber\\
350: \label{taken1}
351: \end{eqnarray}
352: with the ideal part exactly calculated from
353: \begin{eqnarray}
354: \varphi_{\rm{id}}[h]=\ln \eta -1 +\int_0^{2\pi}d\phi h(\phi) \ln\left[
355: 2\pi h(\phi)\right].
356: \label{id}
357: \end{eqnarray}
358: The remaining virial
359: coefficients are approximated from Eq. (\ref{note}) by
360: \begin{eqnarray}
361: B_n[h]=v^{n-3}\left\{(n-2)B_3[h]-(n-3)B_2[h]v\right\}.
362: \end{eqnarray}
363:
364: The integral over spatial variables in the definition of $B_2[h]$
365: is known analytically for most convex bodies. In particular, for HR we have
366: \begin{eqnarray}
367: &&{\cal K}(\phi_1,\phi_2)=
368: \left(L^2+\sigma^2\right)|\sin \phi_{12}|+
369: 2L\sigma \left(1+|\cos\phi_{12}|\right),\nonumber \\
370: \end{eqnarray}
371: which is the excluded area between particles with relative orientation
372: $\phi_{12}$. However, the required double angular average over $h(\phi)$
373: has to be estimated numerically (we used Gaussian quadrature). Also, in
374: the case of $B_3[h]$, all integrals have to be calculated numerically
375: (using MC integration). For this purpose we found
376: it convenient to use a parameterized orientational distribution function
377: \begin{eqnarray}
378: h(\phi)=C \exp{\left(\sum_{\tau=1}^n\lambda_{\tau}
379: \cos(2\tau\phi),
380: \right)}
381: \end{eqnarray}
382: in terms of the $n$ parameters $\lambda_{\tau}$ ($\tau=1,\dots,n$).
383: $C$ is a normalization constant.
384: In practice two variational parameters ($n=2$) were used.
385: Details on how this calculation was realized in practice are relegated to
386: the Appendix.
387:
388: \section{Results}
389: In this section we present the main results obtained from the inclusion
390: of three-body correlations into the EOS for the isotropic and the nematic
391: fluids, as proposed in Section II. The results from bifurcation analysis
392: and from numerical minimization of the free-energy functional are presented
393: in Section III B. In the latter case we compare the results from the present
394: theory with those obtained from SPT and from simulations. But before
395: presenting the results, we show in Fig. \ref{config} a series of
396: particle snapshots extracted from our simulation runs. Details on the
397: simulations are given later on. The three configurations are representative
398: of an isotropic phase [Fig. \ref{config}(a)], a tetratic phase
399: [Fig. \ref{config}(b)], and a crystalline phase [Fig. \ref{config}(c)].
400: \begin{figure}
401: \mbox{\includegraphics*[width=2.9in, angle=0]{Fig1.eps}}
402: \caption{Typical particle configurations as obtained from MC simulations.
403: (a) isotropic phase; (b) tetratic phase; and (c) crystalline phase
404: with particles arranged in one of the possible configurations. See
405: text for details on the simulations.}
406: \label{config}
407: \end{figure}
408:
409:
410: \subsection{Isotropic fluid}
411:
412: Let us first compare the different approximations for $B_3$ and assess
413: their quality according to the degree of agreement with MC-integration
414: results for isotropic fluids made of different hard-convex bodies.
415: Our own MC calculations have been carried out for hard rectangles, while those
416: of Ref. \cite{Rigby} were focussed on hard discorectangles and hard ellipses.
417: All the results are plotted in Fig. \ref{fig1} along with
418: two different analytic approximations for the reduced virial coefficient
419: $B_3^*$. One of them (solid line) is calculated from Eqn. (\ref{bast}) with
420: $\beta=0.514$ [which gives the correct value of $B_3$ in the Onsager limit
421: \cite{Rigby} --see Eqn. (\ref{undos})], and setting $\alpha=1.611$, which
422: gives the correct third virial coefficient for hard disks
423: ($B_3^*\approx3.125$). The other (dotted) line is also calculated from
424: Eqn. (\ref{bast}), but choosing $\alpha=2$ and $\beta=1/8$,
425: which reduces to the proposal made by Boublik in Ref. \cite{Boublik1}.
426: As will be shown below, this proposal approximates the EOS for the
427: isotropic phase of hard convex bodies reasonably well.
428: Also plotted in Fig. \ref{fig1} with dashed lines are the best fits
429: calculated from
430: \begin{eqnarray}
431: B_3=\delta_{\rm{HB}} v^2+\frac{\alpha_{\rm{HB}}}{4\pi}v{\cal L}^2+
432: \frac{\beta_{\rm{HB}}}{(4\pi)^2}
433: {\cal L}^4
434: \label{misma}
435: \end{eqnarray}
436: with values for the coefficients $\delta_{\rm{HB}}$, $\alpha_{\rm{HB}}$,
437: and $\beta_{\rm{HB}}$ depending
438: on the particle geometry ($\rm{HB}\equiv HR,HDR,HE$), i.e. hard rectangles,
439: hard discorectangles, and hard ellipses. Note that Eqn. (\ref{misma}) is
440: the same as Eqn. (\ref{tercero}), but with a new numerical coefficient
441: $\delta_{\rm{HB}}$ as a prefactor of $v^2$. From Fig. \ref{fig1}
442: we can see that the present approximation ($\alpha=1.611,\beta=0.514$)
443: describes the behavior of $B_3$ as a function of the anisometric parameter
444: $\gamma$ much better than Boublik's proposal which, in the Onsager limit,
445: gives the (wrong) value $B_3^*=1/8$. Also, if one is to describe the correct
446: behavior of $B_3^*$ for different particle geometries in the whole range
447: of $\gamma$, it is necessary to take $\delta_{\rm{HB}}\neq 1$.
448:
449: Numerical values for the coefficients $B_4^*$ and $B_5^*$ have been
450: calculated in Ref. \cite{Rigby}, for three different particle geometries,
451: using MC integration. The results for $B_4^*$ are shown in Fig. \ref{fig2}.
452: Also plotted are our analytic proposal (solid line) and that of Boublik
453: (dotted line). For HR with small anisometry values our approximation is
454: better than that of Boublik, while the opposite occurs for high
455: anisometries. For hard ellipses, Boublik's approach describes reasonably
456: well the behavior of $B_4^*$ in the whole range of $\gamma$ (except for
457: very long particles which have negative values of $B_4^*$).
458:
459: \begin{figure}
460: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig2.eps}}
461: \caption{Reduced virial coefficient $B_3^*$
462: as a function of the anisometric parameter $\gamma$.
463: Simulation results are shown for
464: hard rectangles (squares),
465: discorectangles (asterisk), and ellipses (triangles). The solid and dotted
466: lines are the results from Eq. (\ref{bast})
467: with $(\alpha,\beta)=(1.611,0.514)$,
468: and $(\alpha,\beta)=(2,1/8)$ respectively. Also are shown with dashed lines
469: the best fits from Eq. (\ref{misma}).}
470: \label{fig1}
471: \end{figure}
472:
473: \begin{figure}
474: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig3.eps}}
475: \caption{Reduced virial coefficient $B_4^*$
476: as a function of the anisometric parameter $\gamma$.
477: All the lines and symbols label the same as in Fig. \ref{fig1}.}
478: \label{fig2}
479: \end{figure}
480:
481: The EOS obtained from the above approximations can be checked against
482: MC simulations of systems of HR particles. In order to realize this
483: comparison we have carried out constant-pressure MC simulations on
484: systems of HR with different aspect ratios in the range of pressures
485: where the isotropic fluid is the stable phase. The results for $\kappa=3$
486: and 9 are shown in Figs. \ref{fig3} (a) and (b), respectively. Simulations
487: were done on systems of $\sim 10^3$ particles, equilibrated along typically
488: $\sim 10^6$ MC steps, and averaging over $\sim 4\times 10^6$ MC steps.
489: The system was prepared in each case in a crystalline
490: low-density configuration with
491: perfectly aligned particles at low pressure. This configuration rapidly
492: turned into a disordered configuration, which was then equilibrated.
493: After averaging, the system was subject to a higher pressure and then
494: equilibrated, and the process was repeated increasing the pressure.
495: In this way the EOS in the entire region of isotropic stability was obtained.
496: For comparison we also show in Figs. \ref{fig3} (a) and (b)
497: the EOS corresponding to
498: the SPT (dashed line), Boublik proposal (dotted line),
499: our proposal (solid line) and the EOS [Eqn. (\ref{note})] with the virial
500: coefficient $B_3$ calculated from MC integration.
501: As can be seen the SPT and Boublik's proposal approximate better the
502: simulation results. However, given that both theories make wrong predictions
503: of the behavior of the third and fourth virial coefficients of HR as a
504: function of the anisometric parameter, these results are to be taken with
505: caution in the sense that they could be a mere coincidence. We can also see
506: from the figures that
507: our proposal overestimates the pressure. This kind of behavior is typical
508: of fluids composed of hard-core particles which exhibit poorly convergent
509: virial series; this seems to be the case for HR particles since
510: their fourth and fifth virial coefficients become negative for high
511: anisometries.
512:
513:
514: \begin{figure}
515: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig4.eps}}
516: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig5.eps}}
517: \caption{Results from MC simulations (filled circles) on a system of
518: 10800 HRs with $\kappa=3$ (a), and $\kappa=9$ (b). Dot-dashed line:
519: Eq. (\ref{note}) with $B_3$ calculated from MC integration, dashed line:
520: SPT, dotted line: Boublik proposal, and solid line: our proposal
521: [The $B3$ from Eq. (\ref{tercero}) with $\{\alpha=1.611,\beta=0.514\}$].}
522: \label{fig3}
523: \end{figure}
524:
525: \subsection{Bifurcation to nematic fluid}
526:
527: We implemented numerically the bifurcation-theory analysis, described
528: in detail in the Appendix, to calculate the spinodal instabilities from the
529: isotropic phase to the N$_u$ and N$_t$ phases, and elucidated
530: the order of these transitions within the same formalism. These results
531: were checked against a full minimisation of the free-energy functional
532: employing the methodology outlined in Section IIB, which in addition
533: enabled the N$_t-$N$_u$ spinodals, which cannot be easily calculated
534: using bifurcation theory, to be obtained. Also, in order to have
535: essentially exact results for the phase behaviour of this system,
536: we performed constant-pressure MC simulations on systems of $\sim 10^3-10^4$
537: HR particles and obtained the equations of state and orientational order
538: (details on these simulations are included in Section D of the Appendix).
539: All of these results are described in the following.
540:
541: The I-N$_u$ and I-N$_t$ spinodal lines
542: $\eta^*(\kappa)$ (the packing fraction at bifurcation as a function
543: of the aspect ratio $\kappa$) were calculated by solving Eqn. (39)
544: for $y=\eta/(1-\eta)$ (or $\eta$) for a discrete set of values of $\kappa$
545: (see Appendix). All the coefficients $b_3^{(k1,k2,k3)}$ that enter
546: this equation were calculated via MC integration; typically
547: $\sim 10^8$ MC steps were used to evaluate these coefficients. To elucidate
548: the order of transitions, we solved Eqn. (46) to find (i)
549: the value $\kappa_1$ at which the free energy difference
550: between N$_u$ or N$_t$ and isotropic phases
551: changes from negative to positive, which in turn
552: reflects the change of sign of
553: the coefficient $B^*$ (see Appendix), and (ii) the value $\kappa_2$ for
554: which the inverse of the isothermal compressibility of the $N_u$ or
555: $N_t$ phases [$\left(\kappa^{-1}_{\rm{N}}v\right)^*$] at the
556: bifurcation point becomes zero.
557: Again, all the coefficients
558: $b_3$ (the rescaled third virial coefficient) and $b_3^{(k1,k2,k3)}$,
559: necessary to solve Eqn. (46), were evaluated using MC integration with
560: the same number of steps as previously. The
561: quantities $B^*$ and $\left(\kappa^{-1}_{\rm{N}}v\right)^*$ are
562: shown as a function of $\kappa$ in Figs. 4 (a) and (b) for the
563: I-N$_u$ transition, and in Figs. 5 (a) and (b) for the I-N$_t$
564: transition.
565:
566: \begin{figure}
567: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig6.eps}}
568: \hspace*{0.02cm}
569: \mbox{\includegraphics*[width=3.35in, angle=0]{Fig7.eps}}
570: \caption{The coefficient $B^*$ (a), and the inverse of the
571: compressibility factor $\left(\varkappa_{\rm{N}}^{-1}v\right)^*$ (b) at the
572: I-N$_u$ bifurcation point as a function of the aspect ratio calculated
573: for a discrete set of values (open circles). The filled circles indicate
574: the value of $\kappa$ for which they become zero. Thus,
575: $\kappa^*=\kappa_2\approx 4.62$ is the true tricritical point.}
576: \label{fig4}
577: \end{figure}
578:
579:
580: \begin{figure}
581: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig8.eps}}
582: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig9.eps}}
583: \caption{The coefficient $B^*$ (a), and the inverse of the
584: compressibility factor $\left(\varkappa_{\rm{N}}^{-1}v\right)^*$ (b) at the
585: I-N$_t$ bifurcation point as a function of the aspect ratio calculated
586: for a discrete set of values (open circles). Both magnitudes are always
587: greater than zero, so the I-N$_t$ is always of second order. The arrow
588: indicates the maximum aspect ratio of N$_t$ phase stability.}
589: \label{fig5}
590: \end{figure}
591:
592: As can be seen from the figures, $\left(\varkappa_{\rm{N}}^{-1}v\right)^*$
593: first changes sign from positive to negative at $\kappa_2\approx 4.62$, and
594: then diverges at $\kappa\approx 4.11$, coinciding with $\kappa_1$, the
595: zero of $B^*$ [see Fig. \ref{fig4} (a)]. The latter has a pole at $\kappa=3.23$,
596: which is the intersection point between the I-N$_u$ and I-N$_t$ spinodals (for
597: larger values of $\kappa$ the I-N$_u$ spinodal lies below the I-N$_t$ spinodal).
598: This result can be understood from the definition of $B^*$
599: [see Eq. (\ref{otrab})], which diverges at $D^*=0$; this in turn coincides
600: with the condition $A^*=0$ if we change $k$ to $2k$ in Eqn. (\ref{laa})
601: to include tetratic symmetry. Thus $B^*$ as a function of $\kappa$ should
602: diverge at the point where the I-N$_u$ and I-N$_t$ spinodals intersect.
603: The values of $B^*$ and $\left(\varkappa_{\rm{N}}^{-1}v\right)^*$
604: as a function of $\kappa$, calculated this time at the I-N$_t$ spinodal,
605: are shown in Fig. \ref{fig5} (a) and (b). From the figure we can see
606: that they are always positive, in particular for values of $\kappa$ less
607: than $3.23$. Thus we can conclude that the I-N$_t$ transition is always of
608: second order. We also note the oscillatory behavior of
609: $B^*$ as a function of $\kappa$ [see Fig. \ref{fig5}(a)]; this feature
610: has been shown not to be a consequence of numerical errors inherent to
611: our MC integration: MC estimates of the coefficients involved in the definition
612: of $B^*$ were obtained by increasing the number of MC steps from $10^6$ to
613: $10^9$, and the oscillating behavior remained.
614:
615: \subsection{Phase diagram}
616:
617:
618: The resulting spinodal instabilities from the I to the orientationally
619: ordered phases, as calculated from the bifurcation analysis,
620: are shown in Fig. \ref{diagramafase}. Also shown in the same figure is the
621: complete phase diagram resulting from SPT, already calculated
622: in Ref. \cite{Martinez-Raton}. As can be seen from the figure, the inclusion
623: of three-body correlations considerably lowers the transition
624: densities between isotropic and the orientationally ordered phases. The new
625: results compare fairly well with those from MC simulations in the region of
626: low particle aspect ratio (our simulations for
627: $\kappa=$3, and simulations for $\kappa=1$
628: \cite{Frenkel1} and $\kappa=2$ \cite{Donev}), represented in the figure
629: by open squares. Another interesting
630: point to remark is that the critical value of $\kappa$
631: below which the N$_t$ phase is stable increases from $\kappa=2.62$ in SPT to
632: $\kappa=3.23$ in the new theory. Finally, the I-N$_u$
633: tricritical point occurs at $\kappa^*=\rm{max}(\kappa_1,\kappa_2)\approx
634: 4.62$ [see Fig. \ref{fig5} (a) and (b)],
635: which is lower than the SPT result ($\kappa^*=5.44$). Thus, we can conclude that
636: three-body correlations have the effect of lowering the transition densities,
637: increasing the stability of the N$_t$ phase and making the I-N$_u$ transition
638: weaker. The enhanced stability of the tetratic phase is in agreement with simulation
639: results.
640:
641: \begin{figure}
642: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig10.eps}}
643: \caption{Phase diagram of the HR fluid.
644: Continuous and first-order transitions are indicated by
645: dashed and solid lines, respectively. Dotted lines indicate extension
646: of I-N$_t$ line into region where N$_t$ is preempted by uniaxial nematic
647: phase. SPT transition lines and $B_3$ spinodals are indicated by
648: corresponding labels. Circles: minimisation of free-energy functional
649: in $B_3$ theory, giving first-order (filled circles) or second-order
650: (open circles) transitions. Open squares: simulation results for the
651: isotropic-to-nematic transition.}
652: \label{diagramafase}
653: \end{figure}
654:
655: It is also apparent from Fig. \ref{diagramafase} that the I-N$_u$ transition
656: predicted by the new theory for high values of $\kappa$ occurs
657: at packing fractions below those
658: predicted by SPT. In the Onsager limit, the reduced transition density
659: $\rho_r=\rho^* L^2$ for the I-N$_u$ transition can be calculated
660: within the third virial-coefficient approximation [see Eqn. (\ref{rho_r}) of
661: the Appendix].
662: This reduced density depends on the coefficient $\tau^*$ defined in Eqn.
663: (\ref{tau}), which can be calculated by extrapolating the data for
664: $\tau(\kappa)$ obtained from simulations. These data, shown in Fig.
665: \ref{fig6a} for high values of $\kappa$, are fitted very accurately by
666: means of a straight line that intersects the vertical axis at
667: $\tau^*\approx 0.314$.
668: Inserting this value in (\ref{rho_r}), we obtain $\rho_r^{\rm{B_3}}=3.15$,
669: which is less than the SPT result $\rho_r^{\rm{SPT}}=4.71$ and much less
670: than the MC simulation value, which has been estimated to be between
671: 7 and 7.5 \cite{Frenkel1}.
672: This disagreement is probably due to the poorly convergent character of the
673: virial series. As already pointed out, the fourth and fifth virial coefficients
674: are negative in this limit, so the proper inclusion of higher-order virial
675: coefficients is necessary in order to obtain an accurate approximation for the
676: I-N transition densities.
677:
678: Another interesting aspect of the phase diagram is the failure of the new theory
679: to reproduce the transition from the isotropic to the nematic phase in the
680: range of large aspect ratios explored by our simulations.
681: Note that, as will be discussed later, the simulations
682: cannot reach any definite conclusion as to the real nature (whether tetratic or uniaxial)
683: of the nematic phase, especially for large aspect ratio. The fact
684: that the isotropic-to-nematic transition line $\eta(\kappa)$ obtained from simulations
685: in the range $\kappa=1-9$ is a smoothly decreasing monotonic curve and that this
686: line is quite close to the spinodal line for the I-N$_t$ transition obtained from
687: the new theory in the same range of aspect ratios, may be indicating that the
688: stability of the uniaxial phase is largely overestimated by the new theory, but that
689: tetratic ordering is relatively well reproduced. This is simply a hypothesis not
690: based on any real evidence.
691:
692: \begin{figure}
693: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig11.eps}}
694: \caption{The coefficient $\tau(\kappa)$ [see Eq. (\ref{tau})] as a function of
695: $\kappa$ for a discrete set of $\kappa$'s (open circles) calculated
696: via Monte-Carlo integration. The straight line calculated from
697: mean square approximation intersects the ordinate at the value indicated
698: in the figure.}
699: \label{fig6a}
700: \end{figure}
701:
702: \subsection{Further results}
703:
704: In order to appreciate more deeply the differences between the SPT and the
705: new theory, we now compare the EOS for the isotropic and orientationally ordered
706: phases and the behaviour of the order parameters, $q_1,q_2$, with packing fraction.
707: The latter are defined by
708: \begin{eqnarray}
709: q_i=\int d\phi \cos{\left(2i\phi\right)} h(\phi),\hspace{0.7cm}i=1,2
710: \label{OPs}
711: \end{eqnarray}
712: with $q_1$ the uniaxial order parameter and $q_2$ the tetratic order parameter.
713: The comparison is done in Figs. \ref{fig7}(a-c) for the case $\kappa=3$ and in
714: Figs. \ref{fig8} (a-c) for $\kappa=9$. Also, the MC simulation
715: results are shown. In the case of the EOS, both theories severely overestimate
716: the pressure in the nematic regime when $\kappa=3$; however, the
717: transition point, as mentioned previously, is much better reproduced
718: when three-body correlations are included in the theory. For the longer
719: particles the pressures are better reproduced, but the location and
720: nature of the transition from the isotropic to the nematic phase are
721: not correct; as already mentioned, if the uniaxial nematic phase is
722: not taken into account, three-body correlations seem to be very important
723: in promoting tetratic order in the isotropic phase. These correlations alone,
724: when higher-order correlations are not considered, probably overemphasise
725: the relative stability of the uniaxial nematic phase with respect to the
726: tetratic phase in the case of long particles, and cause a premature
727: instability of the latter as particles become longer.
728:
729: Comparison of
730: the orientational distribution functions in the case $\kappa=3$ indicates
731: again the role of three-body correlations. Fig. \ref{fig8} shows the
732: corresponding function for the uniaxial nematic phase that coexists with
733: the tetratic (new theory) or isotropic (SPT) phase; even though the new
734: theory predicts a much lower transition density than SPT, tetratic ordering
735: is much more pronounced in the new theory since at this value of $\kappa$ the
736: tetratic phase is still stable.
737:
738: A point worth mentioning is the identification of the value of aspect ratio
739: where the tetratic phase is no longer stable. The nonequilibrium macroscopic
740: experiments by Narayan et al. \cite{Narayan} find substantial tetratic
741: correlations in cylindrical particles with aspect ratio $\kappa=12.6$. Our present
742: simulation data are not sufficiently detailed to give conclusive results. However,
743: data for $\kappa=7$ (not shown) and $\kappa=9$ seem to be compatible with N$_t$ stability:
744: the value of the uniaxial order parameter $q_1$ is compatible with zero
745: in the whole density range studied. In the case $\kappa=9$ [Fig. \ref{fig8}(b)],
746: however, there seems to be some tendency in the uniaxial order parameter
747: to increase from zero.
748:
749: Nevertheless, it is very difficult
750: within our present analysis, to settle this question. It is in fact
751: difficult to distinguish between the N$_t$ and N$_u$ nematic phases, since
752: the latter exhibits substantial tetratic correlations even for the
753: longer particles considered ($\kappa=9$).
754: Before further work is undertaken, all we can say conclusively from the
755: simulation results is that at some packing fraction the isotropic
756: phase begins to display substantial tetratic order in a rather abrupt manner;
757: whether this order corresponds to uniaxial or strictly tetratic nematic phases
758: is a matter that would require more simulation work using e.g. larger systems.
759: Our limited study is only intended to provide approximate phase boundaries
760: for systems of particles with various aspect ratios (note that previous
761: simulation
762: work on this and related systems \cite{Frenkel,Frenkel1,Donev} were more
763: detailed, but restricted to a particular aspect ratio).
764:
765: \begin{figure}
766: \mbox{\includegraphics*[width=3.in, angle=0]{Fig12.eps}}
767: \mbox{\includegraphics*[width=3.in, angle=0]{Fig13.eps}}
768: \mbox{\includegraphics*[width=3.in, angle=0]{Fig14.eps}}
769: \caption{Results from the SPT, the present model and computer simulation
770: for HR fluid with $\kappa=3$.
771: (a) equation of state resulting from SPT (dashed
772: line), our proposal (solid line) and MC simulation results
773: (filled circles). The arrow indicates the packing fraction of
774: the I-N transition as estimated by simulation, while the filled square
775: indicates the location of the I-N bifurcation point predicted by the present
776: model; (b) behaviour of the uniaxial (solid line) and tetratic (dashed line)
777: order parameters with packing fraction. Results from SPT and our model
778: are indicated by the corresponding label. Symbols are simulation
779: results for the order parameters (open symbols: uniaxial, filled
780: symbols: tetratic); (c) orientational distribution functions at the
781: coexistence packing fraction for the uniaxial nematic phase,
782: from SPT (dashed line) and the present model (solid line).}
783: \label{fig7}
784: \end{figure}
785:
786: \begin{figure}
787: \mbox{\includegraphics*[width=3.in, angle=0]{Fig15.eps}}
788: \mbox{\includegraphics*[width=3.in, angle=0]{Fig16.eps}}
789: \mbox{\includegraphics*[width=3.in, angle=0]{Fig17.eps}}
790: \caption{Same as in Fig. \ref{fig7} but for
791: $\kappa=9$. In (c) are shown the orientational distribution
792: functions calculated at packing fractions separated from
793: the bifurcation packing fraction a relative distance ($\Delta=0.01467$)
794: equal to that between the isotropic and nematic coexisting packing
795: fractions for $\kappa=3$.}
796: \label{fig8}
797: \end{figure}
798:
799: \begin{figure}
800: \mbox{\includegraphics*[width=3.5in, angle=0]{Fig18.eps}}
801: \caption{Orientational distribution function $h(\phi)$ from simulation for the
802: case $\kappa=9$ and packing fraction $\eta=0.618$. Symbols: simulation
803: data. Line: best fit to Eqn. (\ref{Varia}). Resulting values for the
804: order parameters are $q_1=0.153$ and $q_2=0.628$.}
805: \label{Ajuste}
806: \end{figure}
807:
808: \subsection{Dependence on EOS adopted}
809:
810: It should be noted that the packing fraction values of the I-N$_{u,t}$ phase transitions
811: depend sensitively on the approximation used for the EOS of the HR fluid. This can be easily shown
812: if we approximate the third virial coefficient
813: as a function of the second virial
814: coefficient using the relations
815: \begin{eqnarray}
816: \gamma=b_2,\quad \beta \gamma^2=b_3-\alpha b_2,
817: \label{acta}
818: \end{eqnarray}
819: which can be easily obtained from (\ref{segundo}) and (\ref{tercero})
820: and are only valid for the isotropic fluid. From (\ref{acta})
821: we obtain
822: \begin{eqnarray}
823: b_3=\alpha b_2+\beta b_2^2,
824: \label{approx}
825: \end{eqnarray}
826: which can be used as an approximation of the third
827: virial coefficient of the nematic fluid.
828: Thus, inserting the above expression
829: in Eq. (\ref{taken}), and carrying out the bifurcation analysis
830: described in Section IIC, we arrive at
831: \begin{eqnarray}
832: \Delta\varphi_k\approx\frac{h_k^2}{4}\left[1-\Psi(y)b_2^{(k,k,0)}\right],\quad k=1,2 \label{reason}\\
833: \Psi(y)=y+\left[\alpha-2+2\beta\frac{(\kappa+1)^2}{\pi\kappa}\right]\theta(y),
834: \end{eqnarray}
835: for the second order expansion of the free-energy difference between
836: the I and N phases at the bifurcation point; here the subindex $k=1,2$
837: labels the N$_u$ and N$_t$ phases, respectively. Solving equation
838: $\Delta\varphi_k=0$ for $\eta(\kappa)$, we find the spinodals
839: shown in Fig. \ref{fig6b} for different values of $(\alpha,\beta)$
840: corresponding to those of SPT, Boublik's proposal, and our proposal for the
841: EOS of the isotropic fluid. In the same figure the spinodal line resulting
842: from the EOS with the exact third virial coefficient is also plotted.
843: Comparing
844: the latter with those obtained from the different approximations
845: embodied in (\ref{approx}), we conclude that the location of the
846: I-N$_u$ tricritical point changes only if one uses the exact three-body
847: correlations. The reason for this behavior can be elucidated from Eqn.
848: (\ref{reason}): the values of $(\eta^*,\kappa^*)$ calculated from
849: $\Delta\varphi_1=\Delta\varphi_2=0$ gives us $\kappa^*=(3+\sqrt{5})/2$,
850: independent on the choice of $(\alpha,\beta)$ as the function $\Psi(y)$
851: does not depends on $k$.
852:
853: \begin{figure}
854: \mbox{\includegraphics*[width=3.4in, angle=0]{Fig19.eps}}
855: \caption{Spinodals of the transitions between the isotropic and orientational ordered phases. From top to
856: bottom are shown the results from the SPT, the EOS with the approximated $b_3$ (the Boublik, and our proposal), and from the EOS with the exact $b_3$. The dotted lines represent the position of the tricritical points
857: common to all the approximations and that corresponding to the exact $B_3$. Also are shown the simulation
858: results.}
859: \label{fig6b}
860: \end{figure}
861:
862: \section{Conclusions}
863: The main results presented in this article can be summarized as follows.
864: (i) The inclusion of many- (higher-than-two) body correlations in
865: two-dimensional systems of hard anisotropic bodies is of crucial importance
866: in order to adequately describe the phase behavior of these systems.
867: In two-dimensions two-body interactions are not enough to make quantitative
868: predictions of their phase behavior, a crucial difference with respect to
869: three-dimensional systems. This conclusion is supported by simulation
870: results. ii) We have proposed an EOS and a corresponding
871: free energy density functional for fluids of hard rectangles that
872: incorporates three-body correlations. While
873: predicting pressures for the isotropic and nematic fluids which are too high
874: when compared with simulation values, the theory gives values for the coexistence
875: densities of the I-N$_t$ transition that
876: compare fairly well with the simulation
877: results for small values of $\kappa$.
878: A shortcoming of the theory is that the third virial coefficient,
879: which is incorporated exactly, has to be evaluated numerically beforehand.
880: This is a practical, not fundamental, limitation of the theory, which
881: can be circumvented in all cases (i.e. for all different particle
882: geometries in two dimensions).
883:
884: A striking prediction of the theory is
885: the stability of a tetratic phase, in good quantitative agreement with
886: simulations for low particle aspect ratio. We conclude from this that
887: the four-fold correlations present in this phase are basically taken care
888: of by our third-virial coefficient based theory, but not by the usual
889: scaled-particle theory, which incorporates only two-body correlations.
890: More at variance with simulation results is the case of high aspect
891: ratios. In this regime the stability of the uniaxial nematic phase
892: is overestimated with respect to the isotropic fluid.
893:
894: As a final comment, we must say that no attention has been paid
895: to non-uniform phases (smectic, columnar or solid) in the present work.
896: Previous studies by our group \cite{Martinez-Raton}, based on a
897: density-functional theory which recovers SPT in the limit of spatially uniform
898: phases and combines Onsager and fundamental-measure theories,
899: indicate that the tetratic phase is preempted (in the sense of
900: bifurcation theory, i.e. spinodal lines) by a spatially ordered phase.
901: Inclusion of three-body correlations could severely affect
902: this result since these correlations may affect both phases differently.
903: In fact, simulations available so far support the conclusion that the
904: tetratic phase may be stabilised prior to crystallisation.
905: However, even in the case that it were possible to construct a density
906: functional, suitable for such non-uniform phases, and incorporating three-body
907: correlations, the effort involved in the minimization
908: with respect to the full density profile $\rho({\bf r},\phi)$ would
909: be rather huge. Work along this avenue is now in progress in our group.
910:
911: \begin{acknowledgments}
912: Y.M.-R. was supported by a Ram\'on y Cajal research contract from the
913: Ministerio de Ciencia y Tecnolog\'{\i}a (Spain).
914: This work is part of the research
915: Projects No. BFM2003-0180, FIS2005-05243-C02-01, FIS2005-05243-C02-02 and
916: FIS2004-05035-C03-02 of the Ministerio de Educaci\'on y Ciencia
917: (Spain), and S-0505/ESP-0299 of Comunidad Aut\'onoma de Madrid.
918: \end{acknowledgments}
919:
920: \section{Appendix}
921:
922: \subsection{Bifurcation analysis}
923: In this section we introduce the formalism
924: that allowed us to calculate the
925: spinodal instabilities and the order of the phase transitions
926: between the isotropic and orientational ordered phases. This formalism
927: is quite general, in the sense that it is independent of the geometry of
928: particles, and includes two- and three-particle correlations. As usual,
929: the analysis starts from an order-parameter expansion of the free-energy
930: difference ($\Delta\varphi$) between the bifurcated (orientationally ordered)
931: and the parent (isotropic) phases about the bifurcation point, and then the
932: evaluation of the inverse isothermal compressibility ($\varkappa^{-1}$)
933: of the bifurcated phase at the same point. The existence of a tricritical point,
934: at which the order of the transition changes from second to first order,
935: is predicted from the first change of sign (from positive to negative)
936: of $\Delta\varphi$ or $\varkappa^{-1}$. The starting point of the bifurcation
937: analysis is to assume that the orientational distribution function near the
938: I-N bifurcation point can be approximated as a Fourier series in small
939: amplitudes $h_k\sim \epsilon^k$ (where $\epsilon$ is an small parameter),
940: truncated at second order, i.e.
941: \begin{eqnarray}
942: h(\phi)\approx \frac{1}{\pi}\left(1+h_1\cos 2\phi +h_2\cos 4\phi\right).
943: \end{eqnarray}
944: Inserting this expression into (\ref{taken}) and (\ref{id}), we obtain the
945: difference between the nematic and isotropic free energies per particle as
946: \begin{eqnarray}
947: &&\Delta \varphi\equiv \varphi_{\rm{N}}-\varphi_{\rm{I}}
948: \approx Ah_1^2+Ch_1^2h_2+Dh_2^2+Eh_1^4,
949: \label{ener}
950: \end{eqnarray}
951: where the coefficients $A,C,D,E$ have the form
952: \begin{eqnarray}
953: A&=&\frac{1}{4}\left[1-yb_2^{(1,1)}-\theta(y)\left(b_3^{(1,1,0)}
954: -2b_2^{(1,1)}\right)\right],\label{laa}\\
955: C&=&-\frac{1}{8}\left[1+2\theta(y)b_3^{(1,1,2)}\right],\label{lac}\\
956: D&=&\frac{1}{4}\left[1-yb_2^{(2,2)}-\theta(y)\left(b_3^{(2,2,0)}
957: -2b_2^{(2,2)}\right)\right],\label{lad}\\
958: E&=&\frac{1}{32}\label{lae},
959: \end{eqnarray}
960: and we have defined $y=\eta/(1-\eta)$. $\theta(y)=y-\ln(1+y)$ is the same
961: function as in (\ref{exceso}), but here in terms of the new variable $y$.
962: The coefficients
963: \begin{eqnarray}
964: b_i^{(k_1,\dots,k_i)}&=&
965: -\frac{4}{\pi^iv^{i-1}}\left[\prod_{l=1}^i\int_0^{\pi}d\phi_l\cos 2k_l\phi_l
966: \right]\nonumber \\
967: &\times&{\cal K}(\phi_1,\dots,\phi_i),\quad i=2,3,\quad k_j\in \mathbb{N},
968: \end{eqnarray}
969: have also been defined, which originate from two- ($i=2$) and three- ($i=3$)
970: body correlations. For $i=2$ one can obtain analytic results which,
971: for the specific case of HR, give
972: \begin{eqnarray}
973: b_2^{(k,k)}=\frac{2}{(4k^2-1)\pi}\frac{\left(L+(-1)^k\sigma\right)^2}{v}
974: \label{anal}
975: \end{eqnarray}
976: If we set $\theta(y)=0$ in (\ref{laa})-(\ref{lad}) the SPT result is recovered.
977: Minimizing the free energy difference (\ref{ener}) with respect to $h_2$, we
978: obtain $h_2$ as a function of $h_1$ [$h_2=-Ch_1^2/(2D)$] which,
979: inserted in (\ref{ener}), results in
980: \begin{eqnarray}
981: \Delta\varphi&=&Ah_1^2+Bh_1^4,\label{efect}\\
982: B&=&E-\frac{C^2}{4D}\label{otrab}
983: \end{eqnarray}
984: where $A$ and $B$ are functions of the variable $y$.
985: Minimizing Eq. (\ref{efect}) with respect to $h_1$, and taking into account
986: the expansion of $y$ about its bifurcation value $y^*$, i.e.
987: $y\approx y^*+y^{(2)}h_1^2$, we arrive at
988: \begin{eqnarray}
989: \frac{\partial \Delta\varphi}{\partial h_1}=2h_1\left[
990: A^*+\left(2B^*+A_y^*y^{(2)}\right)h_1^2\right]=0,
991: \label{solv}
992: \end{eqnarray}
993: where $A_y$ is the first derivative of $A$ with respect to $y$, and
994: the asterisk on $A,B$, and $A_y$ means that these functions are evaluated at
995: the bifurcation point $y^*$. Solving (\ref{solv}) order by order, we
996: obtain two equations:
997: \begin{eqnarray}
998: A^*&=&0,\label{adin}\\
999: y^{(2)}&=&-\frac{2B^*}{A_y^*},
1000: \label{dva}
1001: \end{eqnarray}
1002: the first one allowing to find $y^*$, and hence the packing
1003: fraction $\eta^*(\kappa)$, as a function of the particle aspect ratio
1004: $\kappa=L/\sigma$, i.e the spinodal line of the I-N phase transition.
1005: Expanding (\ref{efect}) about the bifurcation point, and using
1006: (\ref{adin}) and (\ref{dva}), we obtain
1007: \begin{eqnarray}
1008: \Delta\varphi=-B^*h_1^4,
1009: \label{final}
1010: \end{eqnarray}
1011: which indicates that the I-N transition is of first order if $B^*<0$.
1012: Eqn. (\ref{final}) can be written in a different, more convenient form,
1013: with use of $h_1^2=(y-y^*)/y^{(2)}$ and Eqn. (\ref{dva}), which results in
1014: \begin{eqnarray}
1015: \varphi_{\rm{N}}=\varphi_{\rm{I}}^*-\frac{\left(A_y^*\right)^2}{4B^*}
1016: (y-y^*)^2.
1017: \label{otro}
1018: \end{eqnarray}
1019: Using the definition of the inverse isothermal compressibility,
1020: $\varkappa^{-1}=\rho \partial(\beta P)/\partial\rho$,
1021: in terms of the $y$ variable,
1022: \begin{eqnarray}
1023: \varkappa^{-1} v=y(1+y)\frac{\partial}{\partial y}\left(
1024: y^2\frac{\partial \varphi}{\partial y}\right),
1025: \end{eqnarray}
1026: together with Eqn. (\ref{otro}), we find
1027: \begin{eqnarray}
1028: \left(\varkappa^{-1}_{\rm{N}}v\right)^*=
1029: \left(\varkappa^{-1}_{\rm{I}}v\right)^*-\left(y^*\right)^3(1+y^*)
1030: \frac{\left(A_y^*\right)^2}{2B^*},
1031: \end{eqnarray}
1032: where
1033: \begin{eqnarray}
1034: \left(\varkappa_{\rm{I}}^{-1} v\right)^*&=&
1035: y^*\left(1+y^*\right)\left(1+2y^*b_2\right)\nonumber\\&+&
1036: \frac{\left(y^*\right)^3\left(3+2y^*\right)}{1+y^*}(b_3-2b_2).
1037: \end{eqnarray}
1038: The existence of a tricritical point, at which the I-N transition changes
1039: from second to first order as particles change from large to small
1040: aspect ratios, can be found for a value of the aspect ratio $\kappa^*$
1041: satifying $\kappa^*=
1042: \rm{max}\left(\kappa_1,\kappa_2\right)$, where $\kappa_j$ ($j=1,2$) are
1043: the solutions to the equations
1044: \begin{eqnarray}
1045: B^*(\kappa_1)=0,\quad \left(\varkappa_{\rm{N}}^{-1}v\right)^*(\kappa_2)=0
1046: \label{dosenuna}
1047: \end{eqnarray}
1048: The preceding analysis with $h_k\neq 0$ ($k=1,2$) corresponds
1049: to the bifurcation
1050: analysis of the transition between the isotropic and the uniaxial nematic
1051: phase N$_u$. If $h_1=0, h_2\neq 0$, the bifurcating phase is a tetratic nematic
1052: phase N$_t$. To carry out the bifurcation analysis for the I-N$_t$ transition,
1053: we can use exactly the same formalism, except that
1054: we have to make the substitutions $h_k\to h_{2k}$,
1055: and $b_i^{(k_1,\dots,k_i)}\to b_i^{(2k_1,\dots,2k_i)}$ in Eqns. (\ref{ener})-
1056: (\ref{lad}).
1057:
1058: Taking the Onsager limit $\kappa\to \infty$ in Eqn. (\ref{laa}) and considering
1059: that, in the asymptotic limit [see Eqn. (\ref{undos})], the coefficients
1060: $b_2^{(1,1)}$ and $b_3^{(1,1,0)}$ are of order $\kappa$ and $\kappa^2$,
1061: respectively, the condition (\ref{adin}) is equivalent to solving a
1062: second-order equation with respect to the reduced density
1063: $\rho_r\equiv \rho^* L^2$, with the solution
1064: \begin{eqnarray}
1065: \rho_r=\frac{1}{\tau^*}\left(\sqrt{1+3\pi\tau^*}-1\right),
1066: \label{rho_r}
1067: \end{eqnarray}
1068: where we have defined the coefficient
1069: \begin{eqnarray}
1070: \tau^*=\lim_{\kappa\to\infty} \tau(\kappa),\quad
1071: \tau(\kappa)=\frac{3\pi}{2}\frac{b_3^{(1,1,0)}(\kappa)}{\kappa^2}.
1072: \label{tau}
1073: \end{eqnarray}
1074: The limit $\tau^*\to 0$ of Eq. (\ref{rho_r}) recovers the SPT result
1075: $\rho_r=3\pi/2$. Also $\rho_r(\tau^*)$ as a function of $\tau^*$
1076: is a monotonically decreasing
1077: function whose domain and image are $[-1/3\pi,\infty)$ and
1078: $(0,3\pi]$, respectively. Thus if $\tau^*>0$ ($\tau^*<0$),
1079: the I-N transition in a
1080: two-dimensional hard-needle fluid occurs at a reduced density in the
1081: interval $(0,3\pi/2]$ ($[3\pi/2,3\pi)$).
1082:
1083: \subsection{Calculation of $B_3$}
1084:
1085: The third virial coefficient $B_3(\{\lambda_{\tau}\})$ was obtained
1086: by MC integration using, for the orientational distribution function,
1087: the form
1088: \begin{eqnarray}
1089: h(\phi)=C \exp{\left(\lambda_1\cos{2\phi}+\lambda_2\cos{4\phi}\right)}
1090: \label{Varia}
1091: \end{eqnarray}
1092: which contains two free parameters, $\lambda_1$ and $\lambda_2$.
1093: The MC data for $B_3$ were obtained for each value of $\kappa$ explored
1094: and for fixed values of $\lambda_1$ and $\lambda_2$. The technique followed
1095: was a generalization of the standard method for
1096: isotropic fluids \cite{Hoover}: each step involved generating angles for the
1097: three rectangles [see Eqn. (\ref{kernel3})] and positions for two of them
1098: [the first rectangle is placed at the origin, see Eqn. (\ref{kernel3})].
1099: Since the $B_3$ coefficient involves a single irreducible cluster integral
1100: where all three rectangles overlap, the positions of the second and third
1101: rectangles were chosen within their respective excluded volumes with the
1102: first rectangle to insure overlap, and only one overlap condition
1103: (second with third rectangles) had to be checked. Angles were
1104: generated using an acceptance-rejection method according to the
1105: angular distribution function $h(\phi)$ corresponding to the values of
1106: $\lambda_1$ and $\lambda_2$ (the method was checked by computing the
1107: second virial coefficient $B_2$ for the isotropic case, which can be
1108: compared with the exact result, and also the third virial coefficient
1109: for some special particle orientations where this coefficient is analytic;
1110: in this case the computed values of $B_2$ and $B_3$ for high values of
1111: $\lambda_1$ and $\lambda_2=0$ tended to the correct value). A single
1112: MC step involves generating one chain of rectangles, and
1113: $1-2\times 10^{7}$ MC steps were used in the calculations for each
1114: set of values of $\lambda_1,\lambda_2$.
1115:
1116: We generated numerical values of $B_3$ at a collection
1117: of mesh points on a rectangular region of the $\lambda_1-\lambda_2$ plane.
1118: The extension of this region was chosen according to the values of the packing
1119: fraction. In any case it contained the origin ($\lambda_1=\lambda_2=0$) to allow
1120: for the isotropic phase. In some cases the region $[-1,1]\times[-1,1]$
1121: did suffice; in others, higher minimum and maximum values for the parameters
1122: were needed, especially when a first-order transition was detected.
1123: The mesh interval was typically $\Delta\lambda=0.1$, with finer meshes
1124: when required. In order to use these data in a practical way, the data
1125: were fitted in two different ways. One involves constructing a polynomial
1126: $P_{N}(\lambda_1,\lambda_2)=\sum_{n=0}^N
1127: \sum_{m=0}^nc_{nm}\lambda_1^m\lambda_2^{n-m}$ by a least-square procedure.
1128: Symmetry considerations require some terms of this polynomial expansion
1129: not to appear, and the corresponding coefficients $c_{nm}$ were
1130: taken to be zero. The degree of the polynomial was typically
1131: in the range $8-10$. In the case of the I-N$_t$ transition, which only
1132: involves $q_2$ (and hence $\lambda_2$), calculations were also done using
1133: fits to a polynomial depending only on $\lambda_2$ (since necessarily
1134: $\lambda_1=0$). Results are consistent with the previous results based
1135: on a full fitting.
1136:
1137: For high packing fractions a fit to a function in the order parameters
1138: $(q_1,q_2)$ is more suitable since their values are close to one, whereas
1139: the $\lambda$ parameters grow without limit. However, the dependence of $B_3$
1140: on $(q_1,q_2)$ is strong. We found it useful to use a combination of
1141: polynomials in the $q$'s and factors of the form $(q_i\pm 1)^n$, with
1142: $n$ a power whose value is optimized in the fit.
1143:
1144: \subsection{Minimization of the free-energy functional}
1145:
1146: The minimizations were done using a variational scheme.
1147: An important question is how accurate is the variational function (\ref{Varia}).
1148: We can assess the quality of this function by comparing with
1149: simulation results for the distribution function $h(\phi)$. Fig. \ref{Ajuste}
1150: shows a distribution-function histogram obtained from a constant-pressure
1151: MC simulation, over $2\times 10^6$ steps, of a fluid with $\kappa=5$ at
1152: pressure $P\sigma^2/kT=1.4$.
1153: This is clearly a nematic phase with tetratic order (whether this
1154: corresponds to a uniaxial or purely tetratic phase is a different matter;
1155: extremely long runs are probably needed to fully equilibrate the system.
1156: For the present purpose this is of no importance). A least-square
1157: fitting to the variational function gives $\lambda_1=0.033$,
1158: $\lambda_2=1.217$ ($q_1=0.008$, $q_2=0.523$), which results in the
1159: function represented in the
1160: figure. Histograms exhibiting more structured orientational
1161: order can be similarly fitted with comparable accuracy. The function
1162: (\ref{Varia}) is therefore suitable as a variational function.
1163:
1164: The nematic order parameters can be related to the variational
1165: parameters via Eqn. (\ref{OPs}).
1166: There is a one-to-one correspondence between the sets
1167: $(q_1,q_2)$ and $(\lambda_1,\lambda_2)$.
1168: In our calculations the function
1169: $\varphi(q_1,q_2;\kappa,\eta)$ was then minimized with respect to
1170: $q_1,q_2$, using a standard Newton-Raphson technique.
1171: All transitions are obtained as second-order transitions, except
1172: in the approximate interval $3\alt\kappa\alt 5$ where discontinuous
1173: transitions were found. Of course these results are consistent with
1174: those from bifurcation theory, which is otherwise better suited for
1175: the calculation of the tricritical points since it is not tied to
1176: any variational scheme. The value of the functional minimization
1177: can be better appreciated in the case of the N$_t$-N$_u$ transition,
1178: which cannot easily be obtained using bifurcation theory.
1179:
1180: The results of the minimisation using the resulting
1181: fitted function for $B_3$ are very sensitive to
1182: details such as mesh interval and degree of fitting polynomial. A detailed study
1183: of the whole procedure, including fine tuning of the above parameters,
1184: was therefore necessary. The accuracy of the results is sufficient to
1185: locate the phase transitions with respect to packing fraction, and to
1186: discriminate between first- and second-order phase transitions when
1187: the system is far from the tricritical points, but the exact density
1188: gap in first-order transitions (which is otherwise small) could not
1189: be obtained with the present numerical implementation.
1190:
1191: \subsection{Some details on MC simulations}
1192:
1193: Our constant-pressure MC simulations were performed on systems of
1194: $\sim 10^3-10^4$ HR particles, using rectangular cells and periodic
1195: boundary conditions. The equation of state, orientational
1196: distribution function and nematic order parameters were obtained during the
1197: course of these simulations. The simulations were run typically over
1198: $\sim 10^7$ MC steps for equilibration and $\sim 2\times 10^7$
1199: MC steps for averaging (slow orientational dynamics, especially in the
1200: case of long particles, require longer runs than in the isotropic phase).
1201: The value of the pressure is fixed at some constant value. The average
1202: density (or packing fraction $\eta$) is obtained, which gives the equation of
1203: state $P(\eta)$.
1204:
1205: The orientational order is obtained from the eigenvalues and eigenvectors of
1206: the order matrix, defined for a given particle configuration as
1207: \begin{eqnarray}
1208: S_{ij}=\frac{1}{N}\sum_{k=1}^N \left(2\hat{n}_i^k\hat{n}_j^k-\delta_{ij}\right)
1209: \end{eqnarray}
1210: where $\hat{\bf n}^k$ is a unit vector along the long axis of the $k$th
1211: particle. These is to be averaged over MC configurations.
1212: The eigenvector associated with the largest eigenvalue gives the
1213: direction of the primary director, and this eigenvalue is the uniaxial
1214: order parameter, $q_1$, which can also be calculated from the average
1215: \begin{eqnarray}
1216: q_1=\frac{1}{N}\sum_{k=1}^N \left<\cos{2\left(\phi_k-\phi_0\right)}\right>
1217: \end{eqnarray}
1218: where $\phi_0$ is the polar angle of the director (which depends on the
1219: configuration) and $\phi_k$ the polar angle of the particle long axis, both
1220: with respect to some fixed direction in the plane.
1221: The tetratic order parameter can now be calculated from
1222: a similar equation, with the factor 2 in the cosine substituted by 4.
1223: Also, the orientational distribution function $h(\phi)$
1224: was calculated as a histogram, using the angle $\phi_0$ as the origin.
1225: From this the order parameters $q_i$ can also be obtained from Eqns.
1226: (\ref{OPs}). Yet another route is provided by the asymptotic value of
1227: the orientational correlation functions. No attempt was made at
1228: calculating these functions in the present work.
1229:
1230: The function $h(\phi)$ is always seen to have two maxima: a
1231: primary and a secondary maximum, separated by $\pi/2$; in the uniaxial
1232: nematic phase these maxima should have different heights, whereas in
1233: the tetratic phase their heights should be statistically equal.
1234: Due to many effects that
1235: affect the simulations, distinguishing these two
1236: situations is a rather delicate problem and our limited study did not
1237: allow identification of the true nature of the nematic phase.
1238: Questions such as effect of
1239: boundary conditions, system size, etc, may be of paramount importance
1240: in this analysis. For example, the rectangular periodic boundary conditions used
1241: in this work could be artificially promoting tetratic ordering in the system.
1242:
1243: As is characteristic of two-dimensional systems with continuous symmetries,
1244: the orientationally ordered phases of the present model seem to
1245: exhibit quasi-long-range order at long distances \cite{Donev}. This means, in
1246: particular, that the order parameters may show a strong system-size
1247: dependence. This point has not been considered at all, since our only
1248: aim was to establish approximate phase-stability boundaries that could
1249: serve as a test bed against which the (otherwise approximate) theories
1250: could be tested.
1251:
1252: Finally, since this was not the aim of this work, and also
1253: due to the difficulties involved in dealing with possibly multiply
1254: degenerate structures, both periodic and nonperiodic, crystalline
1255: configurations have not been studied in any detail; however, freezing into
1256: glassy states in compression runs were observed to occur (these states were
1257: characterised by extremely low particle diffusion) at high density. These
1258: densities at quite close to those at which melting into a nematic phase from
1259: crystalline configurations are observed to occur in expansion runs
1260: (starting from crystals with various types of
1261: packing --particles perfectly aligned on a rectangular lattice,
1262: square clusters on square lattices, various random tilings, etc.)
1263: A full discussion of this issue can be found in Ref. \cite{Donev}.
1264:
1265: \begin{references}
1266: \bibitem{Schaklen}H. Schlacken, H. -J. Mogel, and P. Schiller,
1267: Mol. Phys. {\bf 93}, 777 (1998).
1268: \bibitem{Martinez-Raton}Y. Mart\'{\i}nez-Rat\'on, E. Velasco, and
1269: L. Mederos, J. Chem. Phys. {\bf 122}, 064903 (2005).
1270: \bibitem{Frenkel1}K. W. Wojciechowski and D. Frenkel, Comp. Meth. Sci. Tech.
1271: {\bf 10}, 235 (2004).
1272: \bibitem{Donev}A. Donev, J. Burton, F. H. Stillinger, and
1273: S. Torquato, Phys. Rev. B {\bf 73}, 054109 (2006).
1274: \bibitem{PTCDA}M. M\"obus, N. Karl and T. Kobayashi, J. Crys. Growth {\bf 116},
1275: 495 (1992); L. Nony, R. Bennewitz, O. Pfeiffer, E. Gnecco, A. Baratoff,
1276: E. Meyer, T. Eguchi, A. Gourdon and C. Joachin, Nanotechnology {\bf 15},
1277: S91-S96 (2004); H. Proehl, T. Dienel, R. Nitsche and T. Fritz, Phys. Rev.
1278: Lett. {\bf 93}, 097403 (2004).
1279: \bibitem{Narayan} V. Narayan, N. Menon, and S. Ramaswamy,
1280: J. Stat. Mech. P01005 (2006)
1281: \bibitem{Onsager} L. Onsager, Ann. N. Y. Acad. Sci. {\bf 51}, 627 (1949).
1282: \bibitem{Tarjus} G. Tarjus, P. Viot, S. M.Ricci, and J. Talbot,
1283: Mol. Phys. {\bf 73}, 773 (1991).
1284: \bibitem{Reiss} H. Reiss, H. L. Frisch, and J. L. Lebowitz, J. Chem. Phys.
1285: {\bf 31}, 369 (1959).
1286: \bibitem{Cotter} M. A. Cotter and D. E. Martire, J. Chem. Phys.
1287: {\bf 52}, 1902 (1970); J. Chem. Phys. {\bf 53}, 4500 (1970); M. A.
1288: Cotter and D. C. Wacker, Phys. Rev. A {\bf 18}, 2669 (1978).
1289: \bibitem{Lasher} G. Lasher, J. Chem. Phys. {\bf 53}, 4141 (1970).
1290: \bibitem{Barboy} B. Barboy and W. Gelbart, J. Chem. Phys. {\bf 71},
1291: 3053 (1979).
1292: \bibitem{Isihara} A. Isihara, J. Chem. Phys. {\bf 18}, 1446 (1950).
1293: \bibitem{Kihara} T. Kihara, Rev. Mod. Phys. {\bf 25}, 831 (1953).
1294: \bibitem{Boublik} T. Boublik and I. Nezbeda, Coll. Czech. Chem.
1295: Comm. {\bf 51}, 2301 (1986).
1296: \bibitem{Boublik1}T. Boublik, Molec. Phys. {\bf 63}, 685 (1988).
1297: \bibitem{Rigby} M. Rigby, Mol. Phys. {\bf 78}, 21 (1993).
1298: \bibitem{Frenkel} D. Frenkel and R. Eppenga, Phys. Rev. A {\bf 31}, 1776 (1985).
1299: \bibitem{Hoover}F. H. Ree and W. G. J. Hoover, J. Chem Phys. {\bf 40}, 939 (1964); D. Frenkel, J. Phys. Chem. {\bf 91}, 4912 (1987).
1300: %\bibitem{hard_disks} E. Helfand, H.L. Frisch, and J.L. Lebowitz, J. Chem. Phys. {\bf 34}, 1037 (1961).
1301: %\bibitem{Dijkstra1} M. Dijkstra, R. van Roij, and R. Evans, Phys. Rev. Lett.
1302: %{\bf 81}, 2268 (1998); Phys. Rev. E {\bf 59}, 5744 (1998).
1303: %\bibitem{Almarza} N. G. Almarza and E. Enciso,
1304: %Phys. Rev. E {\bf 59}, 4426 (1998).
1305: %\bibitem{Buhot} A. Buhot and W. Krauth, Phys. Rev. Lett. {\bf 80}, 3787
1306: %(1997).
1307: %\bibitem{Shuri} Y. Mart\'{\i}nez-Rat\'on, J. A. Cuesta,
1308: %Phys. Rev. E {\bf 58}, R4080 (1998).
1309: %\bibitem{Mederos} E. Velasco, G. Navascu\'es and L. Mederos,
1310: %Phys. Rev. E {\bf 60}, 3158 (1999).
1311: %\bibitem{Lafuente} L. Lafuente and J. A. Cuesta, Phys. Rev. Lett.
1312: %{\bf 89}, 145701 (2002).
1313: %\bibitem{Biben} T. Biben, P. Blandon and D. Frenkel, J. Phys.: Condens. Matter
1314: %{\bf 8}, 10799 (1996).
1315: %%%%% mezcla de varillas y discos%%%%%%%%
1316: %\bibitem{Roij1}R. van Roij and B. Mulder, J. Phys. II France {\bf 4},
1317: %1763 (1994).
1318: %\bibitem{Wensink} H. H. Wensink, G. J. Vroege, and H. N. W.
1319: %Lekkerkerker, J. Chem. Phys. {\bf 115}, 7319 (2001).
1320: %\bibitem{Perera} S. Dubois and A. Perera, J. Chem. Phys. {\bf 116}, 6354
1321: %(2002); A. Perera, K. Cassou, F. Ple and S. Dubois,
1322: %Mol. Phys. {\bf 100}, 3409 (2002); A. Perera, J. Mol. Liq. {\bf 109},
1323: %73 (2004).
1324: %\bibitem{Varga1} S. Varga, A. Galindo, and G. Jackson, J. Chem. Phys.
1325: %{\bf 117}, 7207 (2002); A. Galindo, A. J.Haslam, S. Varga, G. Jackson,
1326: %A. G. Vanakaras, D. J. Photinos, and D. A. Dunmur, J. Chem. Phys.
1327: %{\bf 119}, 5216 (2003).
1328: %\bibitem{Martinez-Raton1}Y. Mart\'{\i}nez-Rat\'on and J. A. Cuesta,
1329: %J. Chem. Phys. {\bf 118}, 10164 (2003).
1330: %%%%%%% mezcla de esferas y varillas %%%%%%%
1331: %\bibitem{Schmidt}M.Schmidt and A. R. Denton, Phys. Rev. E {\bf 65},
1332: %021508 (2002).
1333: %%%%%%% mezcla de esferas y discos %%%%%%%%%
1334: %\bibitem{Oversteegen} S. M. Oversteegen and H. N. W. Lekkerkerker,
1335: %J.Chem. Phys. {\bf 120}, 2470 (2003).
1336: %%%%%%%% mezcla de varillas %%%%%%%
1337: %\bibitem{Roij2}R. van Roij, and B. Mulder, Phys. Rev. E {\bf 54},
1338: %6430 (1996); R. van Roij, B. Mulder and M. Dijkstra, Physica A {\bf 261},
1339: %374 (1998).
1340: %\bibitem{Dijkstra2}M. Dijkstra and R. van Roij, Phys. Rev. E {\bf 56},
1341: %5594 (1997).
1342: %\bibitem{Varga2} S. Varga, A. Galindo and G. Jackson, Mol. Phys. {\bf 101},
1343: %817 (2003).
1344: %%%%%% SPT for hard convex bodies %%%%%%%%
1345: %\bibitem{Talbot} J. Talbot, J. Chem. Phys. {\bf 106}, 4696 (1997).
1346: %\bibitem{Rosenfeld}Y. Rosenfeld, Phys.Rev. Lett. {\bf 63}, 980 (1989).
1347: %\bibitem{Tarazona} P. Tarazona, Phys. Rev. Lett. {\bf 84}, 694 (2000).
1348: %\bibitem{Tenne} R. Tenne and E. Bergmann, Phys. Rev. A {\bf 17},
1349: %2036 (1978); R. J. Bearman and R. M. Mazo, J. Chem. Phys. {\bf 88}, 1235
1350: %(1988); J. Chem. Phys. {\bf 91}, 1227 (1989); R. Mazo and R. J. Bearman,
1351: %J. Chem. Phys. {\bf 93}, 6694 (1990).
1352: %\bibitem{Mountain} R. Mountain and A. H. Harvey, J. Chem. Phys. {\bf 94},
1353: %2238 (1991); F. Saija and P.V. Giaquinta, J. Chem. Phys. {\bf 117},
1354: %5780 (2002).
1355: %\bibitem{Raveche}R. F. Kayser and H. Ravech\'e, Phys. Rev. A {\bf 17},
1356: %2067 (1978).
1357: %\bibitem{Schoot} P. van der Schoot, J. Chem. Phys. {\bf 106}, 2355 (1997).
1358: %\bibitem{Frenkel} M. A. Bates and D. Frenkel,
1359: %J. Chem. Phys. {\bf 112}, 10034 (2000).
1360: %\bibitem{Martinez-Raton2} Y.Mart\'{\i}nez-Rat\'on, E.Velasco, and L.
1361: %Mederos, J. Chem. Phys. {\bf 122}, 064903 (2005).
1362: %\bibitem{Sluckin} A. Poniewierski and T. J. Sluckin, Phys. Rev. A {\bf 43},
1363: %6837 (1991).
1364: %\bibitem{kappa}Note that in the one-component limit the
1365: %condition $H^*=0$ should be substituted by
1366: % $\left(\varkappa_T^{-1}\right)^*=0$
1367: %(where $\varkappa_T^{-1}=-V
1368: %\displaystyle{\left(\frac{\partial P}{\partial V}\right)_T}$ is
1369: %the inverse isothermal compressibility).
1370: %
1371: \end{references}
1372:
1373:
1374: \end{document}
1375:
1376:
1377: