1: %\documentclass[preprint,prb,groupedaddress,showpacs,preprintnumbers,amsmath,amssymb]{revtex4}
2: \documentclass[twocolumn,prb,groupedaddress,showpacs,preprintnumbers,amsmath,amssymb,floatfix]{revtex4}
3: \usepackage{graphicx}
4:
5: \begin{document}
6:
7: \title{Why do ultrasoft repulsive particles cluster and crystallize?\\
8: Analytical results from density functional theory}
9:
10: \author{Christos N.\ Likos}
11: \email{likos@thphy.uni-duesseldorf.de}
12: \affiliation{Institut f\"ur Theoretische Physik II: Weiche Materie,
13: Heinrich-Heine-Universit\"at D{\"u}sseldorf,
14: Universit{\"a}tsstra{\ss}e 1, D-40225 D\"{u}sseldorf,
15: Germany}
16:
17: \author{Bianca M.\ Mladek}
18: \affiliation{Center for Computational Materials Science and Institut f\"ur
19: Theoretische Physik, Technische Universit\"at Wien, Wiedner Hauptstra{\ss}e
20: 8-10, A-1040 Wien, Austria}
21:
22: \author{Dieter Gottwald}
23: \affiliation{Center for Computational Materials Science and Institut f\"ur
24: Theoretische Physik, Technische Universit\"at Wien, Wiedner Hauptstra{\ss}e
25: 8-10, A-1040 Wien, Austria}
26:
27: \author{Gerhard Kahl}
28: \affiliation{Center for Computational Materials Science and Institut f\"ur
29: Theoretische Physik, Technische Universit\"at Wien, Wiedner Hauptstra{\ss}e
30: 8-10, A-1040 Wien, Austria}
31:
32: \date{\today}
33:
34: \begin{abstract}
35: We demonstrate the accuracy of the hypernetted chain closure and of
36: the mean-field approximation for the calculation of the
37: fluid-state properties of systems interacting by means of bounded
38: and positive-definite pair potentials with oscillating Fourier
39: transforms. Subsequently, we prove the validity of a bilinear,
40: random-phase density functional for arbitrary inhomogeneous phases
41: of the same systems. On the basis of this functional, we calculate
42: analytically the freezing parameters of the latter. We demonstrate
43: explicitly that the stable crystals feature a lattice constant that
44: is independent of density and
45: whose value is dictated by the position of the negative minimum of
46: the Fourier transform of the pair potential.
47: This property is equivalent with the
48: existence of clusters, whose population scales proportionally to
49: the density. We establish that
50: regardless of the form of the interaction
51: potential and of the location on the freezing line, all
52: cluster crystals have a universal Lindemann ratio $L_{\rm f} =
53: 0.189$ at freezing.
54: We further make an explicit link between the
55: aforementioned density functional and the harmonic theory of crystals.
56: This allows us to establish an equivalence between the
57: emergence of clusters and the existence of negative
58: Fourier components of the interaction potential.
59: Finally, we make a connection between the
60: class of models at hand and the system of infinite-dimensional
61: hard spheres, when the limits of interaction steepness and
62: space dimension are both taken to infinity in a particularly
63: described fashion.
64: \end{abstract}
65:
66: \pacs{61.20.-p, 64.70.Dv, 82.70.-y, 61.46.Bc}
67:
68: \maketitle
69:
70: \section{Introduction}
71: \label{intro:sec}
72:
73: Cluster formation in complex fluids is a topic that has
74: attracted considerable attention
75: recently.\cite{sear,fs1,fs2,bartlett1,bartlett2,lr1,lr2,jpcm,epl}
76: The general belief is that a short-range attraction in the
77: pair interaction potential is necessary to initiate
78: aggregation and a long-range repulsive tail in order to
79: limit cluster growth and prevent phase separation.
80: An alternative scenario for cluster formation pertains to systems whose
81: constituent particles interact by means of
82: purely repulsive potentials. Cluster formation in this case
83: is counterintuitive at first sight: why should particles form clusters
84: if there is no attraction acting between them? The answer lies
85: in an additional property of the effective repulsion, namely
86: that of being {\it bounded}, thus allowing full particle overlaps.
87: Though surprising and seemingly unphysical at first, bounded
88: interactions are fully legitimate and natural as effective
89: potentials\cite{likos:pr:01}
90: between polymeric macromolecular aggregates of low internal
91: monomer concentration, such as
92: polymers,\cite{krueger:89, hall:94, louis:prl:00}
93: dendrimers,\cite{ingo:jcp:04,ballik:ac:04}
94: microgels,\cite{denton:03,gottwald:04,gottwald:05}
95: or coarse-grained block copolymers.\cite{pierleoni:prl:06,hansen:molphys:06}
96: The growing
97: interest in this type of effective interactions is also underlined
98: by the recent mathematical proof of the existence of crystalline
99: ground states for such systems.\cite{suto:prl:05,suto:prb:06}
100:
101: Cluster formation in the fluid {\it and} in the crystal phases
102: was explicitly seen in the system of penetrable spheres,
103: following
104: early simulation results\cite{klein:94} and
105: subsequent cell-model calculations.\cite{pensph:98}
106: Cluster formation was
107: attributed there to the tendency of particles to
108: create free space by
109: forming full overlaps. The conditions under which ultrasoft
110: and purely repulsive particles form clusters have been conjectured
111: a few years ago\cite{criterion} and explicitly confirmed
112: by computer simulation very recently.\cite{bianca:prl:06}
113: The key lies in the behavior of
114: the Fourier transform of the effective interaction potential:
115: for clusters to form, it must contain negative parts, forming
116: thus the class of $Q^{\pm}$-interactions. The complementary
117: class of potentials with purely nonnegative Fourier transforms,
118: $Q^{+}$, does not lead to clustering but to remelting at high
119: densities.\cite{stillinger:physica,likos:gauss,saija1, saija2, saija3, archer}
120: An intriguing feature of the crystals
121: formed by $Q^{\pm}$-systems is the independence of the
122: lattice constant on density,\cite{criterion,bianca:prl:06}
123: a feature that reflects the flexibility of soft matter
124: systems in achieving forms of self-organization unknown
125: to atomic ones.\cite{daan:nature,daan:web}
126: The same characteristic has recently been seen also
127: in slightly modified models that contain a short-range
128: hard core.\cite{primoz:07}
129: In this work, we provide an analytical
130: solution of the crystallization problem and of the properties
131: of the ensuing solids within the framework of an accurate
132: density functional approach. We explicitly demonstrate the
133: persistence of a single length scale at all densities and
134: for all members of the $Q^{\pm}$-class and offer thus broad
135: physical insight into the mechanisms driving the stability
136: of the clustered crystals. We further establish some universal
137: structural properties of all $Q^{\pm}$-systems both in the
138: fluid and in the solid state, justifying the use of the
139: mean-field density functional on which this work rests.
140: We make a connection between our results and the harmonic
141: theory of solids in the Einstein-approximation.
142: Finally, we establish a connection with suitably-defined
143: infinite dimensional models of hard spheres.
144:
145: The rest of this paper is organized as follows:
146: in Sec.\ \ref{dft:sec} we derive an accurate density
147: functional by starting with the uniform phase and establishing
148: the behavior of the direct correlation functions of the fluid
149: with density and temperature. Based on this density functional,
150: we perform an analytical calculation
151: of the freezing characteristics of the $Q^{\pm}$-systems by
152: employing a weak approximation in Sec.\ \ref{analytical:sec}.
153: The accuracy of this approximation is successfully tested against full
154: numerical minimization of the functional in Sec.\ \ref{compare:sec}.
155: In Sec.\ \ref{harmonic:sec} the equivalence between the density
156: functional and the theory of harmonic crystals is demonstrated,
157: whereas in Sec.\ \ref{invpower:sec} a connection is made with
158: inverse-power potentials. Finally, in Sec.\ \ref{summary:sec}
159: we summarize and draw our conclusions. Some intermediate, technical
160: results that would interrupt the flow of the text are relegated
161: in the Appendix.
162:
163: \section{Density functional theory}
164: \label{dft:sec}
165:
166: \subsection{Definition of the model}
167: \label{model:sec}
168:
169: In this work, we focus our interest on systems of spherosymmetric
170: particles
171: interacting by means of {\it bounded}
172: pair interactions $v(r)$ of the form:
173: \begin{equation}
174: v(r) = \epsilon\phi(r/\sigma),
175: \label{gen_poten:eq}
176: \end{equation}
177: with an energy scale $\epsilon$ and a length scale $\sigma$,
178: and which fulfill the Ruelle conditions for stability.\cite{Ruelle:book} In
179: Eq.\ (\ref{gen_poten:eq}) above, $\phi(x)$ is some dimensionless function
180: of a dimensionless variable and $r$ denotes the distance between
181: the spherosymmetric particles. In the context of soft matter physics,
182: $v(r)$ is an effective potential between, e.g., the centers of
183: mass of macromolecular entities, such as polymer chains or
184: dendrimers.\cite{ballik:ac:04} As the centers of mass
185: of the aggregates can fully overlap without this
186: incurring an infinitely prohibitive cost in (free) energy, the
187: condition of boundedness is fulfilled:
188: \begin{equation}
189: 0 \leq v(r) < K,
190: \label{bounded:eq}
191: \end{equation}
192: with some constant $K$.
193:
194: The interaction range is set by
195: $\sigma$, typically the physical size
196: (e.g., the gyration radius) of the macromolecular
197: aggregates that feature $v(r)$ as their effective interaction.
198: In addition to being bounded, the second requirement to be
199: fulfilled by the function $\phi(x)$ is that it decay sufficiently
200: fast to zero as $x \to \infty$, so that its Fourier transform
201: $\tilde\phi(y)$ exists. In three spatial dimensions, we have
202: \begin{equation}
203: \tilde\phi(y) = 4\pi
204: \int_0^{\infty}{\rm d}x\,\frac{\sin(y x)}{yx} x^2 \phi(x)
205: \label{ftphi:eq}
206: \end{equation}
207: and, correspondingly,
208: \begin{equation}
209: \tilde v(k) = \epsilon\sigma^3\tilde\phi(k\sigma)
210: \label{ftv:eq}
211: \end{equation}
212: for the Fourier transform $\tilde v(k)$ of the potential, evaluated
213: at wavenumber $k$. Our focus in this work is on systems for which
214: $\tilde v(k)$ is oscillatory, i.e., $v(r)$ features positive
215: and negative Fourier components, classifying it as a
216: $Q^{\pm}$-potential.\cite{criterion} Though this work is general,
217: for the purposes of demonstration of our results, we consider
218: a particular realization of $Q^{\pm}$-potentials, namely the
219: generalized exponential model of exponent $m$, (GEM-$m$):
220: \begin{equation}
221: v(r) = \epsilon\exp[-(r/\sigma)^m],
222: \label{gemn:eq}
223: \end{equation}
224: with $m=4$. It can be shown that all members of the GEM-$m$ family
225: with $m > 2$ belong to the $Q^{\pm}$-class, see Appendix A.
226:
227: A short account
228: of the freezing and clustering behavior of the GEM-4 model has been
229: recently published.\cite{bianca:prl:06}
230: Another prominent member of the family
231: is the $m=\infty$ model, which corresponds to
232: penetrable spheres\cite{klein:94,pensph:98,fernaud}
233: with a finite overlap energy penalty
234: $\epsilon$. Indeed, the explicitly
235: calculated phase behavior of these two show strong resemblances,
236: with the phase diagram of both being dominated by the phenomenon
237: of formation of clusters of overlapping particles
238: and the subsequent ordering of the same in
239: periodic crystalline arrangements.\cite{pensph:98,fernaud}
240: In this work, we provide a generic,
241: accurate, and analytically tractable theory of
242: inhomogeneous phases of $Q^{\pm}$-systems.
243:
244: \subsection{The uniform fluid}
245: \label{uniform:sec}
246:
247: Let us start from the simpler system of a homogeneous fluid,
248: consisting of $N$ spherosymmetric particles enclosed in a
249: macroscopic volume $V$.
250: The structure and thermodynamics of the system are determined by the
251: density $\rho = N/V$ and the absolute temperature $T$ or, better,
252: their dimensionless counterparts:
253: \begin{eqnarray}
254: \nonumber
255: \rho^{*} \equiv \rho\sigma^3
256: \end{eqnarray}
257: and
258: \begin{eqnarray}
259: \nonumber
260: T^{*} \equiv \frac{k_{\rm B}T}{\epsilon},
261: \end{eqnarray}
262: with Boltzmann's constant $k_{\rm B}$. As usual, we also introduce
263: for future convenience the inverse temperature $\beta = (k_{\rm B}T)^{-1}$.
264: We seek for appropriate and accurate closures to the Ornstein-Zernike
265: relation\cite{hansen:book}
266: \begin{equation}
267: h(r) = c(r) + \rho\int{\rm d}^3r'c(|{\bf r}-{\bf r'}|)h(r'),
268: \label{oz:eq}
269: \end{equation}
270: connecting the total correlation function $h(r)$ to the direct
271: correlation function $c(r)$ of the uniform fluid. One possibility
272: is offered by the hypernetted chain closure (HNC) that reads
273: as\cite{hansen:book}
274: \begin{equation}
275: h(r) = \exp\left[-\beta v(r) + h(r) - c(r)\right] - 1.
276: \label{hnc:eq}
277: \end{equation}
278: An additional closure, the mean-field approximation (MFA),
279: was also employed and will be discussed later.
280:
281: Our solution by means of approximate closures was accompanied
282: by extensive $NVT$-Monte Carlo (MC) simulations. We measured the radial
283: distribution function $g(r) \equiv h(r) + 1$ as well as the
284: structure factor $S(k) = 1 + \rho {\tilde h}(k)$,
285: where ${\tilde h}(k)$ is the Fourier transform of $h(r)$, to provide an assessment
286: of the accuracy of the approximate theories. We typically simulated
287: ensembles of up to $3\,000$ particles for a total of $150\,000$
288: Monte Carlo steps. Measurements were taken
289: in every tenth step after equilibration.
290:
291: \begin{figure}%[hbt]
292: \begin{center}
293: \begin{minipage}[t]{8.3cm}
294: \includegraphics[width=8.2cm, clip]
295: {./gofr.t0.5.eps}
296: \end{minipage}
297: \begin{minipage}[t]{8.3cm}
298: \includegraphics[width=8.2cm, clip]
299: {./gofr.t1.eps}
300: \end{minipage}
301: \begin{minipage}[t]{8.3cm}
302: \includegraphics[width=8.2cm, clip]
303: {./gofr.t2.eps}
304: \end{minipage}
305: \end{center}
306: \caption{Radial distribution functions $g(r)$ of the GEM-4 model
307: as obtained by Monte Carlo simulation (points), the HNC-closure
308: (solid lines) and the MFA (dashed lines), at various temperatures
309: and densities. For clarity, the curves on every panel have been
310: shifted upwards by certain amounts, which are given below in
311: square brackets, following
312: the value indicating the density $\rho^{*}$.
313: (a) $T^{*} = 0.5$ and densities, from bottom to top:
314: $\rho^{*} = 0.5$ [0]; $\rho^{*} = 1.0$ [0.2]; $\rho^{*} = 1.5$ [0.4];
315: $\rho^{*} = 2.0$ [0.6]; $\rho^{*} = 2.5$ [0.8].
316: (b) $T^{*} = 1.0$ and densities, from bottom to top:
317: $\rho^{*} = 1.0$ [0]; $\rho^{*} = 2.0$ [0.2]; $\rho^{*} = 3.0$ [0.4];
318: $\rho^{*} = 4.0$ [0.6]; $\rho^{*} = 5.0$ [0.8].
319: (c) $T^{*} = 2.0$ and densities, from bottom to top:
320: $\rho^{*} = 2.0$ [0]; $\rho^{*} = 4.0$ [0.2]; $\rho^{*} = 6.0$ [0.4];
321: $\rho^{*} = 8.0$ [0.6]; $\rho^{*} = 10.0$ [0.8].}
322: \label{gofr:fig}
323: \end{figure}
324:
325: In Fig.\ \ref{gofr:fig} we show comparisons for the function
326: $g(r)$ as obtained from the MC simulations and from the HNC closure
327: for a variety of temperatures and densities. It can be seen that
328: agreement between the two is obtained, to a degree of
329: quality that is excellent. Tiny deviations
330: between the HNC and MC results appear only at the highest density
331: and for a small region around $r = 0$ for
332: low temperatures, $T^{*} = 0.5$.
333: Otherwise, the system at hand
334: is described by the HNC with an extremely high accuracy
335: and for a very broad range of temperatures and densities.
336: We note that, although in Fig.\ \ref{gofr:fig} we restrict ourselves
337: to temperatures $T^{*} \leq 2.0$, the quality of the HNC remains
338: unaffected also at higher temperatures.\cite{criterion}
339:
340: In attempting to gain some insight into the remarkable ability of
341: the HNC to describe the fluid structure at such a high level of
342: accuracy,
343: it is useful to recast this
344: closure in density-functional language. Following the famous
345: Percus idea,\cite{hansen:book,percus:62,percus:64,denton:pra:91,yethiraj:jcp:01}
346: the quantity $\rho g(r)$ can be identified
347: with the nonuniform density $\rho(r)$ of an inhomogeneous fluid that
348: results when a single particle is kept fixed at the origin, exerting
349: an `external' potential $V_{\rm ext}(r) = v(r)$ on the rest of the
350: system. Following standard procedures from density functional
351: theory, we find that the sought-for density profile $\rho(r)$
352: is given by
353: \begin{equation}
354: \rho(r) = \Lambda^{-3}
355: \exp\left\{\beta\mu - \beta v(r)
356: - \frac{\delta \beta F_{\rm ex}[\rho]}{\delta \rho(r)}\right\},
357: \label{profile_dft:eq}
358: \end{equation}
359: where $\Lambda$ is the thermal de Broglie wavelength and $\mu$
360: the chemical potential associated with average density $\rho$
361: and temperature $T$. Moreover, $F_{\rm ex}[\rho]$ is the
362: intrinsic excess free energy, a {\it unique} functional of the
363: density $\rho(r)$. As such, $F_{\rm ex}[\rho]$ can be expanded
364: in a functional Taylor series around its value for a uniform liquid
365: of some (arbitrary) reference density $\rho_0$. For the problem
366: at hand, $\rho_0 = \rho$ is a natural choice and we obtain\cite{singh}
367: %\begin{widetext}
368: \begin{eqnarray}
369: \nonumber
370: \beta F_{\rm ex}[\rho] = \beta F_{\rm ex}(\rho)
371: & - & \sum_{n=1}^{\infty}\frac{1}{n!}\int\int\cdots\int
372: {\rm d}^3r_1{\rm d}^3r_2\ldots{\rm d}^3r_n
373: \\
374: \nonumber
375: &\times&
376: c_0^{(n)}({\bf r}_1,{\bf r}_2,\ldots,{\bf r}_n;\rho)
377: \\
378: &\times&
379: \Delta\rho({\bf r}_1)\Delta\rho({\bf r}_2)\ldots\Delta\rho({\bf r}_n),
380: \label{taylor_expand:eq}
381: \end{eqnarray}
382: %\end{widetext}
383: where $F_{\rm ex}(\rho)$ denotes the excess free energy of the
384: {\it homogeneous} fluid, as opposed to that of the inhomogeneous
385: fluid, $F_{\rm ex}[\rho]$, and
386: \begin{equation}
387: \Delta\rho({\bf x}) \equiv \rho({\bf x}) - \rho.
388: \label{deltarho:eq}
389: \end{equation}
390: The basis of the expansion given by Eq.\ (\ref{taylor_expand:eq}) above is
391: the fact that $F_{\rm ex}[\rho]$ is the generating functional for
392: the hierarchy of the direct correlation functions (dcf's) $c_0^{(n)}$.
393: In particular, $-\beta^{-1}c_0^{(n)}$
394: is the $n$-th functional derivative of the excess
395: free energy with respect to the density field, evaluated at the
396: uniform density $\rho$:\cite{singh,evans:78}
397: \begin{equation}
398: c_0^{(n)}({\bf r}_1,{\bf r}_2,\ldots,{\bf r}_n;\rho) =
399: -\frac{\delta^n \beta F_{\rm ex}[\rho]}
400: {\delta\rho({\bf r}_1)\delta\rho({\bf r}_2)\ldots
401: \delta\rho({\bf r}_n)}\Bigg|_{\rho(.) = \rho}.
402: \label{cofrn:eq}
403: \end{equation}
404: As the functional derivatives are evaluated for a uniform system,
405: translational and rotational invariance reduce the number of variables
406: on which the $n$-th order
407: dcf's $c_0^{(n)}$ depend; in fact,
408: $c_0^{(1)}({\bf r}_1;\rho)$ is a position-independent constant and
409: equals $-\beta\mu_{\rm ex}$, where $\mu_{\rm ex}$ is the excess
410: chemical potential.\cite{evans:78}
411: Similarly, $c_0^{(2)}({\bf r}_1,{\bf r}_2;\rho)$
412: is simply the
413: Ornstein-Zernike direct correlation function $c(|{\bf r}_1-{\bf r}_2|)$
414: entering in Eqs.\ (\ref{oz:eq}) and (\ref{hnc:eq}) above. The HNC
415: closure
416: is equivalent to jointly solving Eqs.\
417: (\ref{oz:eq}) and (\ref{profile_dft:eq}) by employing
418: an approximate excess free energy functional $F_{\rm ex}[\rho]$,
419: arising by a truncation of the expansion of Eq.\ (\ref{taylor_expand:eq})
420: at $n = 2$, i.e., discarding all terms with $n \geq 3$; this
421: is the famous Ramakrishnan-Yussouff second-order
422: approximation.\cite{ry:79,haymet:81}
423: Indeed,
424: the so-called bridge function $b(r)$ can be written as an
425: expansion over integrals involving as kernels all the
426: $c_0^{(n)}$ with $n \geq 3$ and the HNC amounts to setting
427: the bridge function equal to zero.\cite{hansen:book,denton:pra:91,denton:pra:90}
428:
429: Whereas in many cases, such as the
430: one-component plasma,\cite{ocp1,ocp2}
431: and other systems with long-range interactions,
432: the
433: HNC is simply an adequate or, at best, a very good approximation,
434: for the case at hand the degree of agreement between the HNC and
435: simulation is indeed extremely high.
436: What is particularly important is that the accuracy of the HNC persists
437: for a very wide range of densities, at all temperatures considered.
438: This fact has far-reaching consequences, because it means that the
439: corresponding profiles $\rho({\bf x})$ that enter the
440: multiple integrals in Eq.\ (\ref{taylor_expand:eq}) vary enormously
441: depending on the uniform density considered.
442: Thus, it is tempting to conjecture that for the systems under
443: consideration (soft, penetrable particles at $T^{*} \gtrsim 1$),
444: not simply the integrals with $n \geq 3$ vanish but rather the
445: kernels themselves. In other words,
446: \begin{equation}
447: c_0^{(n)}({\bf r}_1,{\bf r}_2,\ldots,{\bf r}_n;\rho) \cong 0,
448: \qquad (n \geq 3).
449: \label{approx:eq}
450: \end{equation}
451:
452: \begin{figure}%[hbt]
453: \begin{center}
454: \begin{minipage}[t]{8.3cm}
455: \includegraphics[width=8.2cm, clip]
456: {./cofr.t0.5.eps}
457: \end{minipage}
458: \begin{minipage}[t]{8.3cm}
459: \includegraphics[width=8.2cm, clip]
460: {./cofr.t1.eps}
461: \end{minipage}
462: \begin{minipage}[t]{8.3cm}
463: \includegraphics[width=8.2cm, clip]
464: {./cofr.t2.eps}
465: \end{minipage}
466: \end{center}
467: \caption{The main plots show the difference between the
468: direct correlation function $c(r)$ calculated in the HNC and its
469: MFA-approximation, $c(r) = -\beta v(r)$, for a GEM-4 model,
470: at various densities indicated in the legends.
471: The insets show the MFA approximation for the same quantity,
472: which is density-independent. Each panel
473: corresponds to a different temperature: (a) $T^{*} = 0.5$;
474: (b) $T^{*} = 1.0$; (c) $T^{*} = 2.0$. These are exactly
475: the same parameter combinations as the ones for which
476: $g(r)$ is shown in Fig.\ \ref{gofr:fig}.}
477: \label{cofr:fig}
478: \end{figure}
479:
480: The behavior of the higher-order dcf's
481: is related to the density-derivative of lower-order ones through
482: certain sum rules, to be discussed below. Hence, it is pertinent
483: to examine the density dependence of the dcf $c(r)$
484: of the HNC.
485: In Fig.\ \ref{cofr:fig} we show the difference between
486: the dcf $c_{\rm HNC}(r)$ and the
487: mean-field approximation (MFA) to the same quantity:
488: \begin{equation}
489: c_{\rm MFA}(r) = -\beta v(r).
490: \label{cmfa:eq}
491: \end{equation}
492: Eq.\ (\ref{cmfa:eq}) above is {\it meaningless} if the pair potential
493: $v(r)$ diverges as $r \to 0$, because $c(r)$ has to remain finite
494: at all $r$, as follows from exact diagrammatic expansions of the
495: same.\cite{hansen:book} In fact, the form $c(r) \sim -\beta v(r)$ denotes
496: the large-$r$ asymptotic behavior of $c(r)$.
497: In our case, however, where
498: $v(r)$ lacks a hard core, the MFA-form for $c(r)$ cannot be
499: a priori rejected on fundamental grounds;
500: the quantity $-\beta v(r)$ remains bounded as $r \to0$. In fact, as can be seen
501: in Fig.\ \ref{cofr:fig}, the deviations between the MFA
502: and the HNC-closure are very small for $T^* \gtrsim 1$.
503: In addition, the differences between $c_{\rm HNC}(r)$ and
504: $c_{\rm MFA}(r)$ drop,
505: both in absolute and in
506: relative terms, as temperature grows, see the trend in
507: Figs.\ \ref{cofr:fig}(a)-(c). At fixed temperature, the evolution
508: of the difference with density is nonmonotonic: it first drops
509: as density grows and then it starts growing again at the highest
510: densities shown in the three panels of Fig.\ \ref{cofr:fig}.
511:
512: Motivated by these findings, we
513: employ now a second closure, namely the above-mentioned
514: MFA, Eq.\ (\ref{cmfa:eq}). Introducing the latter into the
515: Ornstein-Zernike relation, Eq.\ (\ref{oz:eq}), we obtain the
516: MFA-results for the radial distribution function $g(r)$ that
517: are also shown in Fig.\ \ref{gofr:fig} with dashed lines. Before
518: proceeding to a critical comparison between the $g(r)$ obtained
519: from the two closures, it is useful to make a clear connection
520: between the MFA and the HNC.
521:
522: As mentioned above,
523: the members of the sequence of the $n$-th order direct correlation functions
524: are not independent from one another; rather, they are constrained
525: to satisfy a corresponding hierarchy of sum rules,
526: namely:\cite{denton:pra:91,barrat:prl:87,barrat:molphys:88,denton:pra:89}
527: \begin{eqnarray}
528: \nonumber
529: \int{\rm d}^3r_k c_0^{(n+1)}({\bf r}_1,\ldots,{\bf r}_{k-1},{\bf r}_{k},
530: {\bf r}_{k+1},\ldots,{\bf r}_{n+1};\rho) =
531: \\
532: \frac{\partial c_0^{(n)}({\bf r}_1,\ldots,{\bf r}_{k-1},
533: {\bf r}_{k+1},\ldots,{\bf r}_{n+1};\rho)}{\partial\rho}.
534: \label{inductive:eq}
535: \end{eqnarray}
536: In particular, for $n=2$, we have
537: \begin{eqnarray}
538: \nonumber
539: \int{\rm d}^3r' c_0^{(3)}(r,r',|{\bf r} - {\bf r'}|;\rho) & = &
540: \\
541: \nonumber
542: \int{\rm d}^3r' c_0^{(3)}(r',r,|{\bf r'} - {\bf r}|;\rho) & = &
543: \\
544: \frac{\partial c(r;\rho)}{\partial \rho},
545: \label{sumrule:eq}
546: \end{eqnarray}
547: where we have used the translational and
548: rotational invariance of the fluid phase to reduce the
549: number of arguments of the dcf's and we
550: show explicitly the generic dependence of $c(r)$ on $\rho$.
551: In the
552: MFA, one assumes $c(r) = -\beta v(r)$, with the immediate consequence
553: \begin{equation}
554: \frac{\partial c(r)}{\partial \rho} = 0.
555: \label{deriv:eq}
556: \end{equation}
557: Eqs.\ (\ref{sumrule:eq}) and (\ref{deriv:eq}) imply that the integral
558: of $c_0^{(3)}$ with respect to any of its arguments must {\it vanish}
559: for {\it arbitrary} density. As $c_0^{(3)}$ has a complex dependence on
560: its arguments, this is a strong indication that $c_0^{(3)}$ itself
561: vanishes. In fact, both for the Barrat-Hansen-Pastore factorization
562: approximation for this quantity\cite{barrat:prl:87,barrat:molphys:88}
563: and for the alternative,
564: Denton-Ashcroft $k$-space factorization
565: of the same,\cite{denton:pra:89} the
566: vanishing of the right-hand side of Eq.\ (\ref{sumrule:eq}) implies that
567: $c_0^{(3)} = 0$. Now, if $c_0^{(3)}$ vanishes, so does also its
568: density derivative and use of sum rule (\ref{inductive:eq}) for
569: $n=3$ implies $c_0^{(4)} = 0$. Successive use of the same for
570: higher $n$-values leads then to the conclusion that in the MFA:
571: \begin{equation}
572: c_0^{(n)}({\bf r}_1,{\bf r}_2,\ldots,{\bf r}_n;\rho) = 0,\qquad (n\geq 3).
573: \label{kill:eq}
574: \end{equation}
575:
576: We can now see that the accuracy of the HNC stems from the fact
577: that for these systems we can write
578: \begin{equation}
579: c(r) = -\beta v(r) + \varepsilon(r;\rho),
580: \label{epsilon:eq}
581: \end{equation}
582: where $\varepsilon(r;\rho)$ is a small function at all densities $\rho$,
583: offering concomitantly a very small, and {\it the only}, contribution
584: to the quantity $\partial c(r)/\partial\rho$. This implies that
585: $c_0^{(3)}$ itself is negligible by means of Eq.\ (\ref{sumrule:eq}).
586: Repeated use of Eq.\ (\ref{inductive:eq}) leads then to
587: Eqs.\ (\ref{approx:eq}) and shows that the contributions from the
588: $n > 2$ terms, that are ignored in the HNC, are indeed negligible.
589: The HNC is, thus, very
590: accurate, due to the strong mean-field character
591: of the fluids at hand.\cite{likos:pusey} The deviations between the HNC and the MFA
592: come through the function $\varepsilon(r;\rho)$ above.
593:
594: Let us now return to the discussion of the results for $g(r)$ and
595: $c(r)$ and the relative quality of the two closures at various
596: thermodynamic points. Referring first to Fig.\ \ref{gofr:fig}(c), we
597: see that at $T^{*} = 2.0$ both the HNC and the MFA perform
598: equally well.
599: The agreement between the two (and between the MFA and MC)
600: worsens somewhat at $T^{*} = 1.0$ and even more at $T^{*} = 0.5$. The
601: MFA is, thus, an approximation valid for $T^{*} \gtrsim 1$, in agreement with
602: previous results.\cite{criterion} The reason lies in the
603: accuracy of the low-density limit of the MFA. In general, as
604: $\rho\to 0$, $c(r)$ tends to the Mayer
605: function $f(r) = \exp[-\beta v(r)] - 1$. If $T^{*} \gtrsim 1$, one
606: may expand the exponential to linear order and obtain
607: $f(r) \cong -\beta v(r)$, so that the MFA can be fulfilled.
608:
609: In the HNC, it is implicitly assumed that all dcf's
610: with $n \geq 3$ vanish. In the MFA, this is also the case. The two
611: closures differ in one important point, though: in the HNC, the
612: second-order direct correlation function is {\it not} prescribed
613: but rather {\it determined}, so that both the `test-particle equation',
614: Eq.\ (\ref{profile_dft:eq}), and the Ornstein-Zernike relation,
615: Eq.\ (\ref{oz:eq}),
616: are fulfilled. In the MFA, it is a priori assumed that
617: $c(r) = -\beta v(r)$, which is introduced into the Ornstein-Zernike
618: relation and thus $h(r)$ is determined. This is one particular
619: way of obtaining $g(r)$ in the MFA, called the
620: Ornstein-Zernike route. Alternatively, one could follow the
621: test-particle route in solving Eq.\ (\ref{profile_dft:eq}) in conjunction
622: with Eq.\ (\ref{taylor_expand:eq}) and the MFA-approximation,
623: Eq.\ (\ref{cmfa:eq}). In this case, the resulting expression for
624: the total correlation function $h(r)$ in the MFA
625: reads as
626: \begin{equation}
627: h(r) = \exp[-\beta v(r) -\beta\rho(h*v)(r)] - 1,
628: \label{testpart:eq}
629: \end{equation}
630: with $*$ denoting the convolution. Previous studies
631: with ultrasoft systems have shown that the test-particle
632: $h(r)$ from the MFA is closer to the HNC-result than the
633: MFA result obtained from the Ornstein-Zernike route.\cite{ingo:06}
634: The discrepancy
635: between the two is a measure of the approximate character of the MFA;
636: were the theory to be exact, all routes would give the same result.
637: As a way to quantify the approximations involved in the MFA,
638: let us attempt to {\it impose} consistency between the test-particle
639: and Ornstein-Zernike routes. Since $c(r) = -\beta v(r)$ in this
640: closure, the exponent in Eq.\ (\ref{testpart:eq})
641: above is just the right-hand side
642: of the Ornstein-Zernike relation, Eq.\ (\ref{oz:eq}). Thus, if we
643: insist that the latter is fulfilled, we obtain the constraint
644: \begin{equation}
645: h(r) = \exp[h(r)] - 1,
646: \label{hofr:eq}
647: \end{equation}
648: which is strictly satisfied only for $h(r) = 0$. However, as long
649: as $|h(r)| < 1$, one can linearize the exponential and an identity
650: follows; the internal inconsistency of the MFA is of quadratic
651: order in $h(r)$ and it follows that
652: the MFA provides an accurate closure for the
653: systems at hand, as long as $|h(r)|$ remains small. This explains
654: the deviations between MFA and MC seen at small $r$ for the
655: highest density at $T^{*} = 2.0$, Fig.\ \ref{gofr:fig}(c). The same
656: effect
657: can also be seen in Fig.\ \ref{cofr:fig}(c) as a growth of the
658: discrepancy between $c_{\rm HNC}(r)$ and $c_{\rm MFA}(r)$ at small
659: $r$-values for the highest density shown. In absolute terms,
660: however, this discrepancy remains very small. Note also that
661: discrepancies at small $r$-values become strongly suppressed
662: upon taking a Fourier transform, due to the additional geometrical
663: $r^2$-factor involved in the three-dimensional integration.
664:
665: It can therefore be seen that the MFA and the HNC are closely related
666: to one another: the HNC is so successful due to the strong mean-field
667: character of the systems under consideration. This fact has also
668: been established and extensively discussed for the case of the
669: Gaussian model,\cite{likos:gauss,louis:mfa} i.e., the $m=2$ member of the
670: GEM-$m$ class. Once more, the HNC and the MFA there are very accurate for
671: high densities and/or temperatures, where $h(r) \cong 0$ and the
672: system's behavior develops similarities with an `incompressible ideal gas',\cite{likos:gauss}
673: in
674: full agreement with the remarks presented above. Subsequently,
675: the MFA and HNC closures have been also successfully applied to the
676: study of structure and thermodynamics of
677: binary soft mixtures.\cite{louis:mfa,archer1,archer2,archer3,archer4,finken}
678:
679: \begin{figure}
680: \begin{center}
681: \includegraphics[width=8.5cm, clip]
682: {./sofk.t1.eps}
683: \end{center}
684: \caption{The structure factor $S(k)$ of the GEM-4 model at
685: temperature $T^{*} = 1.0$ and various densities, indicated
686: below. For clarity, the curves have been shifted vertically by
687: amounts shown in the square brackets, following the numbers
688: that indicate the values of the density $\rho^{*}$.
689: From bottom to top: $\rho^{*} = 1.0$ [0]; $\rho^{*} = 2.0$ [0.5];
690: $\rho^{*} = 3.0$ [1.0]; $\rho^{*} = 4.0$ [1.5]; $\rho^{*} = 5.0$ [2.0].
691: The points are results from Monte Carlo simulations and the
692: dashed lines from the HNC. As the HNC- and MFA-curves run very
693: close to each other, we show the MFA-result by the solid curve
694: only for the highest density, $\rho^{*} = 5.0$.}
695: \label{sofk:fig}
696: \end{figure}
697:
698: A crucial
699: difference between the Gaussian model, which belongs to the
700: $Q^{+}$-class, and members of the $Q^{\pm}$-class, which are
701: the subject of the present work, lies in the consequences of the
702: MFA-closure on the structure factor $S(k)$ of the system. Since
703: $c(r) = -\beta v(r)$, the Ornstein-Zernike relation leads to
704: the expression
705: \begin{equation}
706: S(k) = \frac{1}{1+\beta\rho\tilde v(k)}.
707: \label{sofk:eq}
708: \end{equation}
709: Whereas for $Q^{+}$-potentials $S(k)$ is devoid of pronounced
710: peaks that exceed the asymptotic value $S(k\to\infty) = 1$,
711: for $Q^{\pm}$-systems a local maximum of $S(k)$ appears at the value
712: $k_*$ for which $\tilde v(k)$ attains its negative minimum.
713: In Fig.\ \ref{sofk:fig} we show representative results for
714: the system at hand, where it can also be seen that the
715: HNC and the MFA yield practically indistinguishable results.
716: In full agreement with the MC simulations, the location of the main peak of
717: $S(k)$ is {\it density-independent}, a feature unknown for
718: usual fluids, having its origin in the fact that $c(r)$ itself
719: is density independent.
720:
721: Associated with this is the development of
722: a $\lambda$-line,\cite{archer4,finken,stell,ciach}
723: also known as {\it Kirkwood instability},\cite{kirkwood:51}
724: on which the denominator
725: in Eq.\ (\ref{sofk:eq}) vanishes at $k_*$ and
726: thus $S(k_*) \to \infty$. The locus
727: of points $(\rho_{\lambda},T_{\lambda})$ on the
728: density-temperature plane for
729: which this divergence takes place is, evidently,
730: given by
731: \begin{equation}
732: \frac{T^*_{\lambda}}{\rho^*_{\lambda}} = -{\tilde \phi(k_*\sigma)}.
733: \label{lambda:eq}
734: \end{equation}
735: In the region $\rho \geq \rho_{\lambda}$
736: (equivalently: $T \leq T_{\lambda}$) on the $(\rho,T)$-plane,
737: the MFA predicts that the fluid is absolutely unstable, since
738: the structure factor there has multiple divergences and also develops
739: negative parts. This holds {\it only} for $Q^{\pm}$-systems; for
740: $Q^{+}$-ones the very same line of argumentation leads to the
741: opposite conclusion, namely that the fluid is the phase of stability
742: at high densities and/or temperatures. The latter conclusion has already
743: been reached by Stillinger and
744: coworkers\cite{still1,still2,still3,still4,still5}
745: in their pioneering work of the
746: Gaussian model in the mid-1970's, and explicitly
747: confirmed by extensive theory and computer simulations many
748: years later.\cite{stillinger:physica,likos:gauss,saija1, saija2, saija3, archer}
749: However, Stillinger's original argument was based on duality relations
750: that are strictly fulfilled only for the Gaussian model, whereas
751: the MFA-arguments are quite general.
752:
753: \subsection{Nonuniform fluids}
754:
755: Having established the validity of the MFA for vast domains in the
756: phase diagram of the systems under consideration as far as the
757: {\it uniform} fluids are concerned, we now turn our attention
758: to nonuniform ones. Apart from an obvious general interest in the
759: properties of nonuniform fluids, the necessity to consider
760: deviations from homogeneity in the density for $Q^{\pm}$-models
761: is dictated by the $\lambda$-instability mentioned above: the
762: theory of the uniform fluid
763: contains its own breakdown, thus the system
764: has to undergo a phase transformation to a phase with a
765: spontaneously broken translational symmetry. Whether this
766: transformation takes place {\it exactly} on the instability line
767: or already at densities $\rho < \rho_{\lambda}$
768: (or temperatures $T > T_{\lambda}$) and which is
769: the stable phase are some of the questions that have to be
770: addressed. Density-functional theory of inhomogeneous systems
771: is the appropriate theoretical tool in this direction.
772:
773: Let us consider a path $\rho_{\chi}({\bf r})$ in the space
774: of density functions, which is characterized
775: by a single parameter $\chi$; this path starts at some reference
776: density $\rho_{\rm r}({\bf r})$ and terminates at another
777: density $\rho({\bf r})$.
778: The uniqueness of the excess free energy functional and its
779: dependence on the inhomogeneous density field $\rho({\bf r})$
780: allow us to
781: integrate
782: $\partial F_{\rm ex}[\rho]/\partial\chi$
783: along this path,
784: obtaining $F_{\rm ex}[\rho]$, provided that
785: $F_{\rm ex}[\rho_{\rm r}]$ is known. A convenient
786: parametrization reads as
787: $\rho_{\chi}({\bf r}) = \rho_{\rm r}({\bf r})
788: +\chi[\rho({\bf r})-\rho_{\rm r}({\bf r})]$,
789: with $\chi = 0$ corresponding to
790: $\rho_{\rm r}({\bf r})$ and $\chi = 1$ to $\rho({\bf r})$.
791: The excess free energy of the final
792: state can be expressed as\cite{evans:78}
793: \begin{eqnarray}
794: \nonumber
795: \beta F_{\rm ex}[\rho] & = &
796: \beta F_{\rm ex}[\rho_{\rm r}]
797: \\
798: & - & \int_0^{1} {\rm d}\chi
799: \int{\rm d}^3r c^{(1)}({\bf r};[\rho_{\chi}])
800: \Delta\rho({\bf r}),
801: \label{chi_path:eq}
802: \end{eqnarray}
803: where $c^{(1)}({\bf r};[\rho_{\chi}])$ denotes the first
804: functional derivative of the quantity $-\beta F_{\rm ex}[\rho]$
805: evaluated at the inhomogeneous density $\rho_{\chi}({\bf r})$,
806: and $\Delta\rho({\bf r}) = \rho({\bf r})-\rho_{\rm r}({\bf r})$.
807: Since $c^{(1)}({\bf r};[\rho_{\chi}])$
808: is in its own turn a unique functional of the density profile,
809: repeated use of the same argument leads to a functional Taylor
810: expansion of the excess free energy around that of a reference
811: system,
812: an expansion that extends to infinite order.
813: For the {\it particular} choice of a {\it uniform} reference system,
814: $\rho_{\rm r}({\bf r}) = \rho$,
815: we obtain then the Taylor series of Eq.\ (\ref{taylor_expand:eq}).
816: In general, however, the reference system does not have to have
817: the same average density as the final one, hence the uniform
818: density $\rho$ there must be replaced by a more general quantity
819: $\rho_0$.
820:
821: The usefulness of Eq.\ (\ref{taylor_expand:eq}) in calculating
822: the free energies of extremely nonuniform phases, such as crystals,
823: is limited both on principal and on practical grounds. Fundamentally,
824: there is {\it no small parameter} guiding such an expansion,
825: since the differences between the nonuniform density of a crystal
826: and that of a fluid are enormous; the former has extreme variations
827: between lattice- and interstitial regions. Hence, the very
828: convergence of the series is in doubt.\cite{singh}
829: In practice,
830: the direct correlation functions for $n = 3$ are very cumbersome to
831: calculate\cite{gerhard:00} and those for $n \geq 4$ are practically
832: unknown.\cite{denton:pra:91} The solution is either to arbitrarily
833: terminate the series at second order\cite{ry:79} or to seek for
834: nonperturbative functionals.\cite{singh,likos:prl:92,likos:jcp:93,wda1,wda2,mwda,henderson}
835: In our case, however, things are
836: different because, for the systems we consider,
837: we have given evidence that the dcf's
838: of order $n > 2$ are extremely small and we take them
839: at this point as vanishing. Then,
840: the functional Taylor
841: expansion of the free energy $F_{\rm ex}[\rho]$ terminates
842: (to the extent that the approximation holds) at second order.
843: The Taylor series becomes a finite sum and
844: convergence is not an issue any more.
845:
846: Let us, accordingly, expand $F_{\rm ex}[\rho]$
847: around an {\it arbitrary}, homogeneous reference fluid of density $\rho_0 = N_0/V$,
848: taking into account that the volume $V$ is fixed but the system
849: with density $\rho({\bf r})$ contains $N$ particles, whereas
850: the reference fluid contains $N_0$ particles and, in general,
851: $N \neq N_0$:
852: \begin{eqnarray}
853: \nonumber
854: \beta F_{\rm ex}[\rho] & = & \beta F_{\rm ex}(\rho_0)
855: - c_0^{(1)}(\rho_0)\int{\rm d}^3r[\rho({\bf r}) - \rho_0]
856: \\
857: \nonumber
858: & - &
859: \frac{1}{2}\int\int{\rm d}^3r{\rm d}^3r' c_0^{(2)}(|{\bf r}-{\bf r'}|;\rho_0)
860: \\
861: & \times &
862: [\rho({\bf r}) - \rho_0][\rho({\bf r'}) - \rho_0].
863: \label{taylor:eq}
864: \end{eqnarray}
865:
866: Using $c_0^{(2)}(r; \rho_0) =
867: c_0^{(2)}(r) \equiv c(r) = -\beta v(r)$ and the
868: sum rule (\ref{inductive:eq}) for $n=1$ together with the vanishing
869: of the excess chemical potential at zero density, we readily obtain
870: \begin{equation}
871: c_0^{(1)}(\rho_0) = -\beta {\tilde v}(0) \rho_0.
872: \label{c1:eq}
873: \end{equation}
874: Formally substituting in Eq.\ (\ref{taylor:eq}),
875: $\rho_0 \to 0$ and $\rho({\bf r}) \to \rho_0$ and making use of
876: the fact that the excess free energy of a system vanishes with the density,
877: we also obtain the dependence
878: of the fluid excess free energy on the density:
879: \begin{equation}
880: \beta F_{\rm ex}(\rho_0) = \frac{N_0}{2}\beta{\tilde v}(0)\rho_0,
881: \label{frex_liq:eq}
882: \end{equation}
883: with the particle number $N_0$ {\it of the reference fluid} and
884: the Fourier transform ${\tilde v}(k)$ of the interaction potential.
885: Introducing Eqs.\ (\ref{c1:eq}) and (\ref{frex_liq:eq})
886: into the Taylor expansion, Eq.\ (\ref{taylor:eq})
887: above, we obtain:
888: \begin{eqnarray}
889: \nonumber
890: \beta F_{\rm ex}[\rho] & = & \frac{N_0}{2}\beta{\tilde v}(0)\rho_0
891: \\
892: \nonumber
893: & + & \beta{\tilde v}(0)\rho_0
894: (N - N_0)
895: \\
896: \nonumber
897: & + & \frac{\beta}{2}\int\int{\rm d}^3r{\rm d}^3r'v(|{\bf r} - {\bf r'}|)
898: \rho({\bf r})\rho({\bf r'})
899: \\
900: \nonumber
901: & - & \beta\rho_0\int\int{\rm d}^3r{\rm d}^3r'v(|{\bf r} - {\bf r'}|)
902: \rho({\bf r})
903: \\
904: & + & \frac{N_0}{2}\beta{\tilde v}(0)\rho_0.
905: \label{expand:eq}
906: \end{eqnarray}
907: Introducing
908: ${\bf x} \equiv {\bf r}-{\bf r'}$ the
909: fourth term above becomes:
910: \begin{equation}
911: - \beta\rho_0\int{\rm d}^3r\,\rho({\bf r})\int{\rm d}^3x\,v(|{\bf x}|)
912: = -\beta N \rho_0 {\tilde v}(0).
913: \label{third:eq}
914: \end{equation}
915: Now the sum of the 1st, 2nd, 4th and 5th term in Eq.\ (\ref{expand:eq}) yields:
916: \begin{eqnarray}
917: \nonumber
918: \frac{N_0}{2}\beta{\tilde v}(0)\rho_0 + \beta{\tilde v}(0)\rho_0(N-N_0)
919: \\
920: \nonumber
921: -\beta N \rho_0 {\tilde v}(0) + \frac{N_0}{2}\beta{\tilde v}(0)\rho_0
922: \\
923: \nonumber
924: = \beta{\tilde v}(0)\left[\frac{N_0}{2}\rho_0 + \rho_0(N-N_0)
925: -N\rho_0 + \frac{N_0}{2}\rho_0\right]
926: \\
927: = \beta{\tilde v}(0)\left[N_0\rho_0 + \rho_0(N-N_0) - N\rho_0\right] = 0.
928: \end{eqnarray}
929: This is a remarkable cancelation because then only the 3rd term in
930: Eq. (\ref{expand:eq}) survives and we obtain:
931: \begin{equation}
932: F_{\rm ex}[\rho] = \frac{1}{2}\int\int{\rm d}^3r{\rm d}^3r'
933: v(|{\bf r} - {\bf r'}|)\rho({\bf r})\rho({\bf r'}),
934: \label{dftmfa:eq}
935: \end{equation}
936: which is our desired result.\cite{criterion,likos:gauss,louis:mfa}
937:
938: The derivation above
939: demonstrates that the excess free energy
940: of {\it any} inhomogeneous phase
941: for our ultrasoft fluids is given by Eq.\ (\ref{dftmfa:eq}),
942: irrespectively of the density of the reference fluid $\rho_0$.
943: This is particularly important because, usually, functional Taylor
944: expansions are carried out around a reference fluid whose density lies
945: close to the average density of the inhomogeneous system (crystal,
946: in our case). However, in our systems this is impossible. The
947: crystals occur predominantly in domains of the phase diagram in which the reference
948: fluid is meaningless, because they are on the high density
949: side of the
950: $\lambda$-line. It is therefore important to be able to
951: justify the use of the functional and to avoid the inherent
952: contradiction of expanding around an unstable fluid.
953: In practice, of course, the higher-order dcf's do not
954: exactly vanish,
955: hence deviations from result (\ref{dftmfa:eq}) are expected to occur,
956: in particular at low temperatures and densities. Nevertheless,
957: the comparisons with simulations, e.g., in
958: Refs.\ [\onlinecite{criterion}], [\onlinecite{bianca:prl:06}] and [\onlinecite{archer4}] fully
959: justify our approximation.
960:
961: A mathematical proof of the mean-field character for fluids
962: with infinitely long-range and infinitesimally strong
963: repulsions
964: has existed since the late 1970's, see
965: Refs.\ [\onlinecite{grewe1}] and [\onlinecite{grewe2}]. However, even far away
966: from fulfillment of this limit, and for conditions that are
967: quite realistic for soft matter systems, the mean-field
968: behavior continues to be valid.\cite{criterion,likos:gauss,louis:mfa}
969: The mean-field result of Eq.\ (\ref{dftmfa:eq}) has been put
970: forward for the Gaussian model at high densities,\cite{likos:gauss}
971: on the basis of physical argumentation: in the absence of
972: diverging excluded-volume interactions, at sufficiently high
973: densities any given particle sees an ocean of others -- the
974: classical mean-field picture. The mean-field character of the
975: Gaussian model for moderate to high temperatures was
976: demonstrated independently in Ref.\ [\onlinecite{louis:mfa}].
977: Here, we have provided a more
978: rigorous justification of its validity, based on the vanishing
979: of high-order direct correlation functions in the fluid.
980: It must also be noted that the mean-field approximation has
981: recently been applied to a system with a broad shoulder and
982: a much shorter hard-core interaction, providing good agreement
983: with simulation results\cite{primoz:07} and allowing for the
984: formulation of a generalized clustering criterion for
985: the inhomogeneous phases.
986:
987: An astonishing similarity
988: exists between the mean-field functional of Eq.\ (\ref{dftmfa:eq})
989: and an exact result derived for
990: infinite-dimensional hard spheres. Indeed, for
991: this case Frisch
992: {\it et al.}\cite{frisch:prl:85,frisch:jcp:86,frisch:pra:87,frisch:pre:99} as
993: well as Bagchi and Rice\cite{bagchi:jcp:88} have shown that
994: \begin{equation}
995: \beta F_{\rm ex}[\rho] = -\frac{1}{2}\int\int{\rm d}^D r{\rm d}^D r'
996: f(|{\bf r} - {\bf r'}|)\rho({\bf r})\rho({\bf r'}),
997: \label{mayer:eq}
998: \end{equation}
999: where $D \to \infty$ and $f(|{\bf r} - {\bf r'}|)$ is the Mayer function
1000: of the infinite dimensional hard spheres. Again, one has a bilinear
1001: excess functional whose integration kernel does not depend on the
1002: density; in this case, this is minus the (bounded) Mayer function whereas for
1003: mean-field fluids, it is the interaction potential itself,
1004: divided by the thermal energy $k_{\rm B}T$. In fact, the Mayer function
1005: and the direct correlation function coincide for infinite-dimensional
1006: systems and higher-order contributions vanish there as
1007: well,\cite{bagchi:jcp:88} making the analogy with our three-dimensional,
1008: ultrasoft systems complete. Accordingly, infinite dimensional
1009: hard spheres have an instability at some finite $k$
1010: at the density
1011: $\rho_{\star}$, given by $\rho_{\star}{\tilde f}(k) = 1$. This so-called
1012: {\it Kirkwood instability}\cite{kirkwood:51}
1013: is of the same nature as our $\lambda$-line
1014: but hard hyperspheres are athermal, so it occurs at a single
1015: point on the density axis and not a line on the density-temperature
1016: plane.
1017: Following Kirkwood's work,\cite{kirkwood:51} it was
1018: therefore argued\cite{frisch:pra:87,bagchi:jcp:88} that hard hyperspheres
1019: might have a second-order freezing transition at the
1020: density $\rho_{\star}$ expressed as
1021: \begin{equation}
1022: \rho_{\star} = 0.239(e/8)^{D/2}\exp[z_0(D/2)^{1/3}]D^{1/6},
1023: \label{rhostar:eq}
1024: \end{equation}
1025: where the limit $D \to \infty$ must be taken and $z_0 = 1.8558$ is the
1026: value of the minimum of the Bessel function $J_{D/2}(z)$ as $D \to \infty$.
1027: Note that $\rho_{\star} \to 0$ as $D \to \infty$.
1028: Later on, Frisch and Percus argued that
1029: most likely the Kirkwood instability is never encountered
1030: because it is preempted by a first-order freezing
1031: transition.\cite{frisch:pre:99}
1032: In what follows, we will show analytically that this
1033: is also the case for our systems, which might provide a finite-dimensional
1034: realization of the above-mentioned mathematical limit.
1035:
1036: \section{Analytical calculation of the freezing properties}
1037: \label{analytical:sec}
1038:
1039: As mentioned above, an obvious candidate for a spatially modulated
1040: phase is a periodic crystal. The purpose of this section is
1041: to employ density functional theory in order to calculate the
1042: freezing properties. Under some weak, simplifying assumptions,
1043: the problem can be solved analytically.
1044:
1045: Adding the ideal contribution to the
1046: excess functional of Eq.\ (\ref{dftmfa:eq}),
1047: the free energy of any spatially modulated phase is obtained as
1048: %\begin{widetext}
1049: \begin{eqnarray}
1050: \nonumber
1051: F[\rho] & = & F_{\rm id}[\rho] + F_{\rm ex}[\rho]
1052: \\
1053: \nonumber
1054: & = & k_{\rm B}T\int{\rm d}^3r\left[\ln\left[\rho({\bf r})\Lambda^3\right]
1055: -1\right]
1056: \\
1057: & + & \frac{1}{2}\int\int{\rm d}^3r{\rm d}^3r'
1058: v(|{\bf r}-{\bf r'}|)\rho({\bf r})\rho({\bf r'}).
1059: \label{totalfren:eq}
1060: \end{eqnarray}
1061: %\end{widetext}
1062: As we are interested in crystalline phases, we parametrize the
1063: density profile as a sum of Gaussians centered around the lattice
1064: sites $\{{\bf R}\}$, forming a Bravais lattice. In sharp contrast
1065: with systems interacting by hard, diverging potentials, however,
1066: the assumption of one particle per lattice site has to be dropped.
1067: Indeed, it will be seen that $Q^{\pm}$ systems employ the strategy
1068: of optimizing their lattice constant by adjusting the number
1069: of particles per lattice site, $n_c$, at any given density
1070: $\rho$ and temperature $T$. Accordingly, we normalize the
1071: profiles to $n_c$ and write
1072: \begin{eqnarray}
1073: \nonumber
1074: \rho({\bf r}) & = & n_c\left(\frac{\alpha}{\pi}\right)^{3/2}
1075: \sum_{{\bf R}}e^{-\alpha\left({\bf r}-{\bf R}\right)^2}
1076: \\
1077: & = & \sum_{{\bf R}}\rho_l({\bf r}-{\bf R}),
1078: \label{gaussian_real:eq}
1079: \end{eqnarray}
1080: where the occupation variable $n_c$ and the
1081: localization parameter $\alpha$ have to be determined variationally,
1082: and the lattice site density $\rho_l({\bf r})$ is expressed as
1083: \begin{equation}
1084: \rho_l({\bf r}) = n_c\left(\frac{\alpha}{\pi}\right)^{3/2}
1085: e^{-\alpha r^2}.
1086: \label{orbital:eq}
1087: \end{equation}
1088: Contrary to crystals of single occupancy, thus, the number of
1089: particles $N$ and the number of sites $N_s$ of the Bravais lattice
1090: do not coincide. In particular, it holds
1091: \begin{equation}
1092: \frac{N}{N_s} = n_c,
1093: \label{nc:eq}
1094: \end{equation}
1095: and we are interested in multiple site occupancies, i.e., $n_c > 1$ or
1096: even $n_c \gg 1$; it will be shown that this clustering scenario indeed
1097: minimizes the crystal's free energy.
1098:
1099: It is advantageous, at this point, to express the periodic
1100: density profile of Eq.\ (\ref{gaussian_real:eq}) as a Fourier series,
1101: introducing the Fourier components $\rho_{\bf K}$ of the same:
1102: \begin{equation}
1103: \rho({\bf r}) = \sum_{{\bf K}}
1104: e^{{\rm i}{\bf K}\cdot{\bf r}}\rho_{\bf K},
1105: \label{denfour:eq}
1106: \end{equation}
1107: where the set $\{{\bf K}\}$ contains all reciprocal lattice vectors
1108: (RLVs) of the Bravais lattice formed by the set $\{{\bf R}\}$.
1109: Accordingly, the inverse of (\ref{denfour:eq}) reads as:\cite{ashcroft}
1110: \begin{eqnarray}
1111: \nonumber
1112: \rho_{\bf K} & = & \frac{1}{v_c}\int_{{\mathcal C}}
1113: {\rm d}^3r e^{{\rm i}{\bf K}\cdot{\bf r}}\rho({\bf r})
1114: \\
1115: & = & \frac{1}{v_c}
1116: \int {\rm d}^3r e^{{\rm i}{\bf K}\cdot{\bf r}}\rho_l({\bf r}),
1117: \label{deninfour:eq}
1118: \end{eqnarray}
1119: where the first integral extends over the elementary
1120: unit cell ${\mathcal C}$ of the crystal and $v_c = V/N_s$ is the
1121: volume of ${\mathcal C}$, containing a single lattice site.
1122: The second integral extends over all space, where
1123: use of the periodicity of $\rho({\bf r})$ and
1124: its expression as a sum over lattice site densities,
1125: Eq.\ (\ref{gaussian_real:eq}), has been made. Using
1126: Eq.\ (\ref{orbital:eq}) we obtain
1127: \begin{eqnarray}
1128: \nonumber
1129: \rho_{\bf K} & = & \frac{n_c}{v_c}e^{-K^2/(4\alpha)}
1130: \\
1131: & = & \rho e^{-K^2/(4\alpha)}.
1132: \label{rhok:eq}
1133: \end{eqnarray}
1134: Note that the site occupancy $n_c$ does not appear explicitly
1135: in the functional form of the Fourier components of $\rho({\bf r})$,
1136: a feature that may seem paradoxical at first sight. However, for
1137: fixed density $\rho$ and any crystal type, the lattice constant
1138: and thus also the reciprocal lattice vectors ${\bf K}$ are affected
1139: by the possibility of clustering, thus the dependence on $n_c$
1140: remains, albeit in an implicit fashion. With the density being
1141: expressed in reciprocal space, the excess free energy takes a
1142: simple form that reads as
1143: \begin{equation}
1144: \frac{F_{\rm ex}}{N} = \frac{\rho}{2}\sum_{{\bf K}}\tilde v(K)
1145: e^{-K^2/(2\alpha)}.
1146: \label{fexk:eq}
1147: \end{equation}
1148:
1149: The ideal term, $F_{\rm id}[\rho]$, can also be approximated
1150: analytically, provided that the Gaussians centered
1151: at different lattice sites do not overlap. Let $a$ denote the
1152: lattice constant of any particular crystal. Then, for
1153: $\alpha a^2 \gg 1$, the ideal free energy of the crystal takes
1154: the form
1155: \begin{equation}
1156: \frac{\beta F_{\rm id}}{N} =
1157: \ln n_c + \frac{3}{2}\ln\left(\frac{\alpha\sigma^2}{\pi}\right)
1158: -\frac{5}{2} + 3\ln\left(\frac{\Lambda}{\sigma}\right),
1159: \label{fida:eq}
1160: \end{equation}
1161: where the trivial last term will be dropped in what follows, since
1162: is also appears in the expression of the free energy of the fluid
1163: and does not affect any phase boundaries. Putting together
1164: Eqs.\ (\ref{fexk:eq}) and (\ref{fida:eq}), we obtain a variational
1165: free energy per particle, $\tilde f$, for the crystal, that reads as
1166: %\begin{widetext}
1167: \begin{eqnarray}
1168: \nonumber
1169: \frac{F_{\rm id}+F_{\rm ex}}{N\epsilon} & \equiv &
1170: \tilde f(n_c,\alpha^*;T^*,\rho^*)
1171: \\
1172: \nonumber
1173: & = &
1174: T^*\left[\ln n_c + \frac{3}{2}\ln\left(\frac{\alpha^*}{\pi}\right)
1175: -\frac{5}{2}\right]
1176: \\
1177: & + & \frac{\rho^*}{2}\sum_{\bf Y}\tilde\phi(Y)
1178: e^{-Y^2/(2\alpha^*)},
1179: \label{fren_var:eq}
1180: \end{eqnarray}
1181: %\end{widetext}
1182: where $\alpha^* \equiv \alpha\sigma^2$ and ${\bf Y} \equiv {\bf K}\sigma$.
1183: In the list of arguments of $\tilde f$ the first two are
1184: variational parameters whereas the last two denote simply its
1185: dependence on temperature and density. The free energy per particle,
1186: $f_{\rm sol}(T^*,\rho^*) \equiv F/(N\epsilon)$ of the crystal is obtained
1187: by minimization of $\tilde f$, i.e.,
1188: \begin{equation}
1189: f_{\rm sol}(T^*,\rho^*) = \min_{\{n_c,\alpha^*\}}
1190: \tilde f(n_c,\alpha^*;T^*,\rho^*).
1191: \label{minimize:eq}
1192: \end{equation}
1193:
1194: In carrying out the minimization, it proves
1195: useful to measure the localization length of the Gaussian
1196: profile, $\ell \equiv 1/\sqrt{\alpha}$, in units of the
1197: lattice constant $a$ instead of units of $\sigma$.
1198: To perform this change, we first express the average
1199: density $\rho$ of the crystal in terms of $n_c$ and $a$ as
1200: \begin{equation}
1201: \rho = \frac{z n_c}{a^3},
1202: \label{den_a:eq}
1203: \end{equation}
1204: where $z$ is a lattice-dependent numerical coefficient of order
1205: unity. Introduce now the quantity
1206: $\alpha a^2 = \gamma^{-1}$. Using Eq.\ (\ref{den_a:eq}) above,
1207: we obtain
1208: \begin{equation}
1209: \alpha^*
1210: = \gamma^{-1}\left(\frac{\rho^*}{zn_c}\right)^{2/3}.
1211: \label{amin_nc:eq}
1212: \end{equation}
1213: This change of variables is just a mathematical transformation that
1214: simplifies the mathematics to follow; all results to be derived
1215: maintain their validity also in the original representation.
1216: For a further discussion of this point, see also Appendix B.
1217:
1218: Next we make the simplifying approximation to
1219: ignore in the sum over reciprocal lattice vectors on the right-hand-side
1220: of Eq.\ (\ref{fren_var:eq}) above all the RLVs beyond the first
1221: shell, whose length is $Y_1 = K_1\sigma$. This is justified
1222: already because of the
1223: exponentially damping factors
1224: $\exp[-Y^2/(2\alpha^*)]$ in the sum. In addition, the
1225: coefficients
1226: $|\tilde\phi(Y)|$ themselves decay to
1227: zero as $Y \to \infty$, with an asymptotic behavior that
1228: depends on the form of $\phi(x)$ in real space. The
1229: length of the first
1230: shell of RLVs of any Bravais lattice of lattice constant $a$
1231: scales as $K_1 = \zeta/a$, with some positive, lattice-dependent
1232: numerical constant $\zeta$ of order unity. Together with
1233: Eqs.\ (\ref{den_a:eq}) this implies that the length of the
1234: first RLV depends on the aggregation number $n_c$ as
1235: \begin{equation}
1236: Y_1(n_c) = \zeta\left(\frac{\rho^*}{z n_c}\right)^{1/3},
1237: \label{y1nc:eq}
1238: \end{equation}
1239: and using Eq.\ (\ref{amin_nc:eq}), we see that the ratio $Y_1^2/(2\alpha^*)$
1240: takes a form that depends {\it solely} on the parameter $\gamma$,
1241: namely
1242: \begin{equation}
1243: \frac{Y_1^2(n_c)}{2\alpha^*} = \frac{\gamma\zeta^2}{2}.
1244: \label{d:eq}
1245: \end{equation}
1246:
1247: Introducing Eqs.\ (\ref{amin_nc:eq}) and ({\ref{d:eq}}) into
1248: (\ref{fren_var:eq}), we obtain another
1249: functional form for the variational free energy,
1250: $\bar f(n_c,\gamma;T^*,\rho^*)$, expressed
1251: in the new variables. It can be seen that
1252: upon making the transformation
1253: (\ref{amin_nc:eq}),
1254: the term
1255: $3/2\ln[(\alpha^*/\pi)]$
1256: delivers a contribution {\it minus} $\ln n_c$
1257: that exactly cancels the same term with a positive sign on
1258: the right-hand side of Eq.\ (\ref{fren_var:eq}).
1259: Accordingly, the
1260: {\it only} remaining quantity of the variational free energy
1261: that still depends on $n_c$ is the length of the first
1262: nonvanishing RLV, $Y_1$, whose $n_c$ dependence is expressed
1263: by Eq.\ (\ref{y1nc:eq}) above.
1264: Putting everything together,
1265: we obtain
1266: %\begin{widetext}
1267: \begin{eqnarray}
1268: \nonumber
1269: \bar f(n_c,\gamma;T^*,\rho^*) & = & T^*
1270: \left[\ln \rho^* - 1 - \frac{3}{2}\left[\ln(\gamma\pi) - 1\right] - \ln z
1271: \right]
1272: \\
1273: \nonumber
1274: & + & \frac{1}{2}\rho^*\tilde\phi(0)
1275: \\
1276: & + & \frac{\xi_1\rho^*}{2}\tilde\phi\left(Y_1(n_c)\right)e^{-\gamma\zeta^2/2},
1277: \label{fbar:eq}
1278: \end{eqnarray}
1279: %\end{widetext}
1280: where $\xi_1$ is the coordination number of the {\it reciprocal} lattice.
1281: Minimizing $\bar f$ with respect to $n_c$ is trivial and using
1282: Eq.\ (\ref{d:eq}) we obtain
1283: \begin{equation}
1284: \frac{\partial {\bar f}}{\partial n_c} = 0
1285: \Rightarrow \tilde \phi'(Y_1)Y_1^4 = 0,
1286: \label{partialnc:eq}
1287: \end{equation}
1288: where the prime denotes the derivative with respect to the argument.
1289: Evidently, $Y_1$ coincides with $y_*\equiv k_*\sigma$, the
1290: dimensionless wavenumber for which
1291: the dimensionless Fourier transform of the interaction potential
1292: attains its negative minimum. The other mathematical solution
1293: of (\ref{partialnc:eq}), $Y_1 = 0$, can
1294: be rejected because it yields nonpositive second derivatives or,
1295: on physical grounds, because it corresponds to a crystal with
1296: $n_c \to \infty$,
1297: whose occurrence would violate the thermodynamic stability
1298: of the system.
1299: Regarding second derivatives,
1300: it can be easily shown that
1301: \begin{equation}
1302: \frac{\partial^2 \bar f}{\partial n_c^2}\Bigg|_{Y_1 = y_*} > 0,
1303: \label{second_nc:eq}
1304: \end{equation}
1305: and
1306: \begin{equation}
1307: \frac{\partial^2 \bar f}{\partial\gamma\partial n_c}\Bigg|_{Y_1 = y_*} = 0,
1308: \label{second_ncgamma:eq}
1309: \end{equation}
1310: irrespective of $\gamma$.
1311:
1312: Having shown the coincidence of $Y_1$ with $y_*$, we set
1313: $\tilde\phi(Y_1) = \tilde\phi(y_*) < 0$ in Eq.\ (\ref{fbar:eq})
1314: above. Further, we notice that the term $T^*[\ln\rho^*-1]$ on the
1315: right-hand side of Eq.\ (\ref{fbar:eq}) gives the
1316: ideal free energy of a uniform fluid of density
1317: $\rho^*$ and the term $\rho^*\tilde\phi(0)/2$ the excess part
1318: of the same, see
1319: Eq.\ (\ref{frex_liq:eq}).
1320: Subtracting, thus, the total fluid free
1321: energy per particle, $f_{\rm fl}(T^*,\rho^*)$, we introduce
1322: the difference $\Delta {\bar f} \equiv \bar f - f_{\rm fl}$,
1323: which reads as
1324: \begin{eqnarray}
1325: \nonumber
1326: \Delta \bar f(n_c(y_*),\gamma;T^*,\rho^*) =
1327: & - & \frac{3T^*}{2}\left[\ln(\gamma\pi) + 1 + \frac{2\ln z}{3}\right]
1328: \\
1329: & + & \frac{\xi_1\rho^*}{2}\tilde\phi(y_*)e^{-\gamma\zeta^2/2}.
1330: \label{deltaf:eq}
1331: \end{eqnarray}
1332:
1333: The requirement of no overlap between Gaussians
1334: centered on different lattice sites restricts $\gamma$ to be small;
1335: a very generous upper limit is $\gamma \leq 0.05$. For such
1336: small values of $\gamma$,
1337: the first term on the right-hand side of Eq.\ (\ref{deltaf:eq}) above
1338: is positive.
1339: This positivity
1340: expresses the entropic cost
1341: of localization that a crystal pays, compared to the fluid in which
1342: the delocalized
1343: particles possess translational entropy. This cost must be compensated
1344: by a gain in the excess term, which is only possible if $\tilde\phi(y_*) < 0$.
1345: An additional degree of freedom is offered by the candidate crystal
1346: structures.
1347: The excess free energy is minimized by the
1348: direct Bravais lattice whose reciprocal lattice has
1349: the maximum possible coordination number $\xi_1$.
1350: The most highly coordinated periodic arrangement
1351: of sites is fcc, for which $\xi_1 = 12$. Therefore,
1352: in the framework of this approximation, the stable lattice
1353: is bcc. It must be emphasized, though, that these results hold
1354: as long as only the first shell of RLVs is kept in the excess
1355: free energy. Inclusion of higher-order shells can, under
1356: suitable thermodynamic conditions, stabilize fcc in favor of bcc.
1357: We will return to this point later.
1358:
1359: Choosing now $a$ as the edge-length of the {\it conventional}
1360: lattice cell of the bcc-lattice, we have $z=2$ and
1361: $\zeta = 2\sqrt{2}\pi$. Evidently,
1362: the lattice constant of the crystal is
1363: density-independent,
1364: $a/\sigma = (2\sqrt{2}\pi)/y_*$, contrary to the case
1365: of usual crystals, for which $a \propto \rho^{-1/3}$.
1366: The density-independence of $a$ is achieved
1367: by the creation of clusters that consist of $n_c$ particles,
1368: each of them occupying a
1369: lattice site. The proportionality
1370: relation connecting $n_c$ and $\rho^*$ follows from
1371: Eq.\ (\ref{y1nc:eq}) and reads as
1372: \begin{equation}
1373: n_c = \frac{8\sqrt{2}\pi^3}{y_*^3}\rho^*.
1374: \label{proport:eq}
1375: \end{equation}
1376:
1377: It remains to minimize $\bar f$ (equivalently, $\Delta\bar f$)
1378: with respect to $\gamma$ to determine the free energy of the crystal.
1379: We are interested, in particular, in estimating the `freezing
1380: line', determined by the equality of free energies of the fluid
1381: and the solid.\cite{foot1}
1382: Accordingly,
1383: we search for the simultaneous solution of the equations
1384: \begin{eqnarray}
1385: \frac{\partial {\bar f}}{\partial \gamma} & = & 0,
1386: \\
1387: \Delta {\bar f} & = & 0,
1388: \end{eqnarray}
1389: resulting into
1390: \begin{equation}
1391: \frac{3T^*}{2\gamma} +
1392: \frac{\xi_1\zeta^2\rho^*}{4}\tilde\phi(y_*)e^{-\gamma\zeta^2/2}=0,
1393: \label{dfdg:eq}
1394: \end{equation}
1395: and
1396: \begin{equation}
1397: \frac{\xi_1\rho^*}{2}\tilde\phi(y_*)e^{-\gamma\zeta^2/2}=
1398: \frac{3T^*}{2}\left[\ln(\gamma\pi)+1+\frac{2\ln z}{3}\right].
1399: \label{dfzero:eq}
1400: \end{equation}
1401: Substituting (\ref{dfzero:eq}) into (\ref{dfdg:eq}) and using $z = 2$ and
1402: $\zeta = 2\sqrt{2}\pi$, we obtain an implicit equation for $\gamma$
1403: that reads as
1404: \begin{equation}
1405: \gamma^{-1} = -4\pi^2
1406: \left[\ln\left(\gamma\pi\right)+1+\frac{2\ln 2}{3}\right],
1407: \label{gamma:eq}
1408: \end{equation}
1409: and has two solutions, $\gamma_1 \cong 0.018$ and $\gamma_2 \cong 0.038$.
1410: Due to (\ref{second_nc:eq}) and (\ref{second_ncgamma:eq}),
1411: the sign of the determinant of the Hessian matrix at the extremum
1412: is set by the sign of $\partial^2{\bar f}/\partial \gamma^2$;
1413: Using Eqs.\ (\ref{dfdg:eq}), (\ref{dfzero:eq}),
1414: and (\ref{gamma:eq}) we obtain
1415: \begin{equation}
1416: \frac{\partial^2{\bar f}}{\partial \gamma^2} =
1417: \frac{3T^*}{2\gamma}\left(\gamma^{-1}-4\pi^2\right),
1418: \label{second_gamma:eq}
1419: \end{equation}
1420: which is positive for $\gamma = \gamma_1$ but negative for
1421: $\gamma = \gamma_2$. Only the first solution corresponds to
1422: a minimum and thus to freezing,
1423: whereas the second is a saddle point.
1424: Within the limits of the first-RLV-shell approximation,
1425: the crystals formed by $Q^{\pm}$-potentials feature thus a
1426: {\it universal localization parameter} at freezing: irrespective
1427: of the location on the freezing line and
1428: even {\it of the interaction potential itself}, the localization
1429: length $\ell$ at freezing is a fixed fraction of the lattice constant
1430: and the parameter $\gamma = (\alpha a^2)^{-1}$ attains along the
1431: entire crystallization line the value
1432: \begin{equation}
1433: \gamma_{\rm f} \cong 0.018.
1434: \label{gammaf:eq}
1435: \end{equation}
1436:
1437: We can
1438: understand the physics behind the
1439: constancy of the ratio $\ell/a$ by examining anew the
1440: variational form of the free energy, Eq.\ (\ref{fren_var:eq}).
1441: Suppose we have a fixed density $\rho^*$ and we vary $n_c$, seeking
1442: to achieve a minimum of $\tilde f$. An increase in $n_c$ implies
1443: an increase in the lattice constant $a$ by virtue of Eq.\ (\ref{den_a:eq}).
1444: The density profile takes advantage of the additional space
1445: created between neighboring sites and becomes more delocalized. This
1446: increase of the spreading of the profile brings with it an
1447: entropic gain which exactly compensates the corresponding loss
1448: from the accumulation of particles on a single site, expressed
1449: by the term $\ln n_c$ in Eq.\ (\ref{fren_var:eq}). Expressing
1450: $\ell$ in units of $a$, i.e., working with the variable $\gamma$
1451: instead with the original one, $\alpha^*$, brings the additional
1452: advantage that $\gamma_{\rm f}$ becomes independent of the
1453: pair potential. The corresponding value of $\alpha^*$ at freezing,
1454: $\alpha^*_{\rm f}$, can be obtained from Eqs.\ (\ref{den_a:eq})
1455: and (\ref{proport:eq}),
1456: and reads for the bcc-lattice as
1457: \begin{equation}
1458: \alpha^*_{\rm f} = \frac{y_*^2}{8\pi^2\gamma_{\rm f}}.
1459: \label{alphaf:eq}
1460: \end{equation}
1461: Here, a dependence on the pair interaction appears through
1462: the value of $y_*$.
1463:
1464: Complementary to the localization parameter,
1465: we can consider the Lindemann ratio $L$ at freezing,\cite{lindemann}
1466: taking into account that for the bcc lattice the nearest neighbor
1467: distance is $d = a\sqrt{3}/2$. Employing
1468: the Gaussian density parametrization,
1469: we find $\langle r^2 \rangle = 3/(2\alpha)$ and thus
1470: $L \equiv \sqrt{\langle r^2 \rangle}/d = \sqrt{2\gamma}$.
1471: Using (\ref{gammaf:eq}), the
1472: Lindemann ratio at freezing, $L_{\rm f}$, is determined as
1473: \begin{equation}
1474: L_{\rm f} \cong 0.189.
1475: \label{lind:eq}
1476: \end{equation}
1477: This value is considerably larger than the typical value of $0.10$ usually
1478: quoted for systems with harshly repulsive particles, such as,
1479: e.g., the bcc alcali metals and the fcc metals Al, Cu, Ag, and
1480: Au [\onlinecite{shapiro}], but close to the value 0.160 found
1481: by Stillinger and Weber\cite{still3} for the Gaussian core model.
1482: The particles in the cluster crystal are quite more delocalized
1483: than the ones for singly-occupied solids. The clustering strategy
1484: enhances the stability of the crystal with respect to oscillations
1485: about the equilibrium lattice positions.
1486:
1487: The locus of freezing points $(T^*_{\rm f}, \rho^*_{\rm f})$
1488: is easily obtained by Eqs.\ (\ref{dfzero:eq})
1489: and (\ref{gamma:eq}) and
1490: takes the form of of a straight line:
1491: \begin{equation}
1492: \frac{T^*_{\rm f}}{\rho^*_{\rm f}} = 16\pi^2\gamma_{\rm f}
1493: |\tilde\phi(y_*)|e^{-4\pi^2\gamma_{\rm f}}
1494: \cong 1.393\, |\tilde\phi(y_*)|.
1495: \label{freeze:eq}
1496: \end{equation}
1497: Contrary to the Lindemann ratio, which is independent of
1498: the pair potential, the freezing line does depend on the interaction
1499: potential between the particles. Yet, this dependence is a
1500: particular one, as it rests exclusively on the absolute
1501: value of the Fourier transform at the minimum,
1502: $|\tilde\phi(y_*)|$ and is simply proportional to it.
1503: Comparing with the location of the $\lambda$-line from
1504: Eq.\ (\ref{lambda:eq}),
1505: $T^*_{\lambda}/\rho^*_{\lambda} = |\tilde\phi(y_*)|$,
1506: we find that crystallization
1507: {\it preempts} the occurrence
1508: of the instability: indeed, at fixed $T^*$, $\rho^*_{\rm f}
1509: < \rho^*_{\lambda}$ or, equivalently, at fixed $\rho^*$,
1510: $T^*_{\rm f} > T^*_{\lambda}$; see also Fig.\ \ref{phdg:fig}.
1511: The transition is first-order, as witnessed by the jumps
1512: of the values of $\alpha$ and $\rho_{\bf K}$ at the
1513: transition, which
1514: are nonzero for the crystal but vanish in the fluid.
1515: This is analogous to the conjectured preemption of the Kirkwood instability
1516: for infinite-dimensional hard spheres by a first-order freezing
1517: transition.\cite{frisch:pre:99}
1518:
1519: The freezing properties of $Q^{\pm}$-potentials are, thus,
1520: quite unusual and at the same time quite simple: the lattice
1521: constant is fixed due to a clustering mechanism that drives
1522: the aggregation number $n_c$ proportional to the density.
1523: The constant of proportionality depends solely on the wavenumber
1524: $y_*$ for which the Fourier transform of the pair interaction
1525: has a negative minimum, Eq.\ (\ref{proport:eq}).
1526: The freezing line is a straight line
1527: whose slope depends only on the value of the Fourier transform
1528: of the potential at the minimum, Eq.\ (\ref{freeze:eq}). The Lindemann ratio
1529: at freezing is a universal number, independent of interaction
1530: potential and thermodynamic state.
1531:
1532: Whereas the Lindemann ratio is employed as a measure of the
1533: propensity of a crystal to melt, the height of the peak of the structure
1534: factor of the fluid is looked upon as a measure of the tendency
1535: of the fluid to crystallize. The Hansen-Verlet
1536: criterion\cite{hv1,hv2}
1537: states
1538: that crystallization takes place when this quantity exceeds
1539: the value 2.85. For the systems at hand, the maximum of $S(y)$
1540: lies at $y_*$, as is clear from Eq.\ (\ref{sofk:eq}).
1541: Using Eq.\ (\ref{freeze:eq})
1542: for the location of the freezing line, we obtain the value
1543: $S_{\rm f}(y_*)$ on the freezing line as
1544: \begin{equation}
1545: S_{\rm f}(y_*) = \left[1 + \frac{\tilde\phi(y_*)}{1.393\,|\tilde\phi(y_*)|}
1546: \right]^{-1} \cong 3.542.
1547: \end{equation}
1548: This value is considerably larger than the Hansen-Verlet threshold.\cite{hv1,hv2}
1549: In the fluid phase,
1550: $Q^{\pm}$-systems can therefore sustain a higher
1551: degree of spatial correlation before they crystallize
1552: than particles with diverging interactions do. This
1553: property lies in the fact that some contribution to
1554: the peak height comes from correlations from {\it within} the clusters
1555: that form in the fluid; the formation of clusters already in
1556: the uniform phase is witnessed by the maxima of $g(r)$ at $r=0$
1557: seen in Fig.\ \ref{gofr:fig} and also explicitly
1558: visualized in our previous simulations of the model.\cite{bianca:prl:06}
1559: These, however,
1560: do not contribute to intercluster ordering that leads to
1561: crystallization.
1562: At any rate,
1563: the Hansen-Verlet peak height is also a
1564: universal quantity for all $Q^{\pm}$ systems, in the
1565: framework of the current approximation. Moreover, both for the Lindemann
1566: and for the Hansen-Verlet
1567: criteria, the $Q^{\pm}$ systems are more robust than usual ones,
1568: since they allow for stable fluids with peak heights that exceed
1569: $2.85$ by 25\%
1570: and for stable crystals with Lindemann ratios that
1571: exceed $0.10$ by almost 90\%.
1572:
1573: \section{Comparison with numerical minimization}
1574: \label{compare:sec}
1575:
1576: The density functional of Eq.\ (\ref{totalfren:eq}) is very accurate
1577: for the bounded ultrasoft potentials
1578: at hand.\cite{criterion,likos:gauss,louis:mfa,archer1,archer2,archer3,archer4,finken} The modeling of
1579: the inhomogeneous density as a sum of Gaussians is
1580: an approximation but, again, an accurate one, as has been shown
1581: by comparing with simulation results,\cite{bianca:prl:06}
1582: see also section
1583: \ref{harmonic:sec} of this work. The analytical results
1584: derived in the preceding section rest on one additional approximation,
1585: namely on ignoring the RLVs beyond the first shell. Here, we
1586: want to compare with a full minimization of the functional
1587: (\ref{totalfren:eq}) under the modeling of the density via
1588: (\ref{gaussian_real:eq}), so as to test the accuracy of the
1589: hitherto drawn conclusions on clustering and crystallization.
1590:
1591: \begin{figure}
1592: \begin{center}
1593: \includegraphics[width=8.5cm, clip]
1594: {./phdg.eps}
1595: \end{center}
1596: \caption{The phase diagram of the GEM-4 model, obtained by full
1597: minimization of the density functional (\ref{totalfren:eq}), under
1598: the Gaussian parametrization of the density, Eq.\ (\ref{gaussian_real:eq}),
1599: redrawn from Ref.\ [\onlinecite{bianca:prl:06}]. On the same plot, we show by the
1600: dotted line the
1601: approximate analytical result
1602: for the freezing line, Eq.\ (\ref{freeze:eq}), as well as the
1603: $\lambda$-line of the system, Eq.\ (\ref{lambda:eq}).}
1604: \label{phdg:fig}
1605: \end{figure}
1606:
1607: We work with the concrete GEM-4 system, for which the minimization
1608: of the density functional has been carried out
1609: and the phase diagram has been calculated in Ref.\ [\onlinecite{bianca:prl:06}].
1610: In Fig.\ \ref{phdg:fig} we show the phase diagram obtained by the
1611: full minimization, compared with the freezing line from the
1612: analytical approximation, Eq.\ (\ref{freeze:eq}), for this system.
1613: It can be seen that the latter is a very good approximation
1614: to the full result, its quality improving
1615: slowly as the temperature grows; the analytical approximation
1616: consistently overestimates the region of stability of the crystal.
1617: Moreover, whereas the approximation only predicts a stable bcc
1618: crystal, the high-density phase of the system is fcc. Although
1619: bcc indeed is, above the triple temperature, the stable crystal
1620: immediately post-freezing, it is succeeded at higher densities by
1621: a fcc lattice, which our analytical theory fails to predict.
1622:
1623: \begin{figure}
1624: \begin{center}
1625: \includegraphics[width=8.5cm, clip]
1626: {./phiofq.eps}
1627: \end{center}
1628: \caption{Inset: the Fourier transform $\tilde\phi(y)$ of the GEM-4
1629: potential. Main plot: a zoom at the region of $\tilde\phi(y)$ in which
1630: the first few nonvanishing RLV shells of the cluster-crystals of Fig.\ \ref{phdg:fig}
1631: lie. The arrows denote the positions of the shells and the numbers
1632: in square brackets the numbers of distinct RLVs within each shell.
1633: These positions are the result of the full minimization of the
1634: density functional. Downwards pointing arrows pertain to the direct bcc
1635: lattice and upwards pointing arrows to the direct fcc one. In agreement
1636: with clustering predictions, the positions of the RLVs are
1637: density-independent.}
1638: \label{phiofq:fig}
1639: \end{figure}
1640:
1641: All these discrepancies can be easily understood by looking at the
1642: effects of ignoring the higher RLV shells from the summation
1643: in the excess free energy, Eq.\ (\ref{fren_var:eq}).
1644: Consider first exclusively the bcc
1645: lattice. In Fig.\ \ref{phiofq:fig} we show the locations of
1646: the bcc-RLVs, as obtained from the full minimization, by the
1647: downwards pointing arrows. It can be seen that the first shell
1648: is indeed located very closely to $y_*$, as the analytical
1649: solution predicts. However, the next two RLV shells do have
1650: contributions and, due to their location on the hump of
1651: $\tilde\phi(y)$, the latter is positive. By ignoring them
1652: in performing the analytical solution, we are artificially lowering the
1653: free energy of the crystal, increasing thereby its domain
1654: of stability.
1655:
1656: The occurrence of a fcc-lattice that beats the bcc at high
1657: densities is only slightly more complicated to understand.
1658: A first remark is that the parameter $\alpha^*$
1659: increases proportionally to $\rho^*/T^*$, see the following
1660: section. Hence, the Gaussian factors
1661: from RLVs beyond the first shell, $\exp[-Y_i^2/(2\alpha^*)]$, $i \geq 2$,
1662: gain weight in the sum as density grows. The cutoff for the RLV-sum is
1663: now provided rather by the short-range nature of
1664: $\tilde\phi(y)$ than by the exponential factors.
1665: Due to the increased importance of the contributions from the $i \geq 2$-terms
1666: in the excess free energy sum, the relative location of higher RLVs
1667: becomes crucial and can
1668: tip the balance in favor of fcc, although the bcc-lattice
1669: has a {\it higher} number of RLVs in its first shell than the fcc.
1670: In Fig.\ \ref{phiofq:fig} we see that this is precisely what
1671: happens: the second RLV shell of the fcc is located fairly
1672: close to the first. In the full minimization, both of them
1673: arrange their positions so as to lie close enough to
1674: $y_*$. Now, a total of fourteen 1st- and 2nd-shell RLVs
1675: of the fcc can beat the twelve 1st-shell RLVs of the bcc
1676: and bring about a structural phase transformation from the
1677: latter to the former.
1678:
1679: The relative importance of the first and second
1680: neighbors is quantified by the ratio
1681: \begin{equation}
1682: \Delta = \frac{\xi_2}{\xi_1}\frac{|\tilde\phi(Y_2)|}{|\tilde\phi(Y_1)|}
1683: \exp[-(Y_2^2-Y_1^2)/(2\alpha^*)],
1684: \label{ratio:eq}
1685: \end{equation}
1686: where $\xi_2$ is the number of RLVs in the second shell.
1687: If $\tilde\phi (y)$
1688: does not decay sufficiently fast to zero as $y$ grows,
1689: then the fcc lattice might even win over the bcc everywhere,
1690: since then $\Delta$ could be considerable even for values of
1691: $\alpha^*$ close to freezing, which are not terribly high.
1692: In fact, the penetrable sphere model (GEM-$m$ with $m \to \infty$)
1693: does not possess, on these grounds,
1694: a stable bcc phase at all.\cite{pensph:98}
1695: The prediction of the analytical theory
1696: on bcc stability has to be taken with care and is conditional
1697: to $\Delta$ being sufficiently small. A quantitative criterion
1698: on the smallness of $\Delta$ is model specific and cannot be
1699: given in general.
1700: The determination of the stable phases of the GEM-$m$ family
1701: and their dependence on $m$ can be achieved by employing
1702: genetic algorithms\cite{dieter:jcp:05} and will be presented
1703: elsewhere.\cite{bianca:long}
1704:
1705: Notwithstanding the quantitative discrepancies between
1706: the simplified, analytically tractable version of DFT and the
1707: full one, which are small in the first place,
1708: the central conclusion of the former remains intact:
1709: the RLVs of the crystals are density-independent. Whereas
1710: the analytical approximation predicts that the
1711: length of the first RLV shell
1712: coincides with $y_*$, the numerical minimization brings about
1713: small deviations from this prediction. However, by reading off
1714: the relevant values from Fig.\ \ref{phiofq:fig}, we obtain
1715: $Y_1 = 5.625$ for the bcc and $Y_1 = 5.441$ for the fcc-lattice
1716: of the GEM-4 model.
1717: Comparing with the ideal value
1718: $y_* = 5.573$, we find that the deviation between them is only
1719: a few percent. Clustering takes place, so that the lattice
1720: constants of both lattices remain fixed, a characteristic
1721: that was also explicitly confirmed by computer simulations
1722: of the model.\cite{bianca:prl:06}
1723:
1724: \section{Connection to harmonic theory of crystals}
1725: \label{harmonic:sec}
1726:
1727: The use of a Gaussian parametrization for the
1728: one-particle density profiles, Eq.\ (\ref{gaussian_real:eq}),
1729: is a standard modeling of the latter for periodic crystals.
1730: This functional form is closely related to the harmonic theory of
1731: crystals.\cite{ashcroft} Each particle performs
1732: oscillations around its lattice site, experiencing thereby
1733: an effective, one-particle site potential, $V_{\rm site}({\bf s})$
1734: that is quadratic in the displacement $s$, for small values
1735: $s/a$ [\onlinecite{foot2}].
1736: Here, we will explicitly demonstrate that the Gaussian form
1737: with the localization parameter predicted from density functional
1738: theory coincides with the results obtained by performing
1739: a harmonic expansion of the said site potential.
1740:
1741: The formation of clustered crystals is a generic property
1742: of all $Q^{\pm}$-systems, since the $\lambda$-instability
1743: is common to all of them; the form of the clusters that
1744: occupy the lattice sites, however, can be quite complex,
1745: depending on the details of the interaction. The Gaussian
1746: parametrization (\ref{gaussian_real:eq}) implies that for
1747: each of the $n_c$ particles of the cluster, the lattice
1748: site ${\bf R}$ is an equilibrium position. In other words,
1749: the particular clusters we consider here are internally
1750: structureless. Clusters with a well-defined internal order
1751: have been found when an additional hard core of small extent is
1752: introduced.\cite{primoz:07}
1753: A necessary
1754: requirement for the lack of internal order is that the Laplacian
1755: of the interaction potential $v(r)$ be finite at
1756: $r=0$, as will be shown shortly.
1757: On these grounds, we impose
1758: from the outset on the interaction potential
1759: the additional requirement:
1760: \begin{equation}
1761: \nabla^2 v(r) = \frac{1}{r^2}\left(r^2 v'(r)\right)'
1762: < \infty\;\;{\rm for}\;\; r \to 0,
1763: \label{condition1:eq}
1764: \end{equation}
1765: where the primes denote the derivative with respect to $r$.
1766: Eq.\ (\ref{condition1:eq}) implies that $v'(r)$ must be
1767: {\it at least} linear in $r$ as $r \to 0$. Concomitantly,
1768: $v''(r)$ must be at least $O(1)$ as $r \to 0$. As a
1769: consequence, we have
1770: \begin{equation}
1771: v''(r) - \frac{v'(r)}{r} \to 0\;\;{\rm for}\;\; r \to 0.
1772: \label{condition:eq}
1773: \end{equation}
1774: It can be easily checked that (\ref{condition:eq}) is satisfied by
1775: all members of the GEM-$m$ class for $m > 2$. It is also satisfied
1776: by the $m=2$ member, i.e., the Gaussian model, which does
1777: {\it not} display clustering because it belongs to the $Q^{+}$-class.
1778: This is, however, no contradiction. As mentioned above,
1779: the condition (\ref{condition:eq}) is {\it necessary} for the
1780: formation of structureless clusters and not a sufficient one.
1781: For $Q^{\pm}$-potentials for which (\ref{condition:eq}) is not
1782: fulfilled (such as the Fermi distribution models of
1783: Ref.\ [\onlinecite{criterion}]), this does not mean that
1784: clusters do not form; it rather points to the fact that
1785: they possess some degree of internal order.
1786:
1787: The clustered crystals can be considered as Bravais lattices with
1788: a $n_c$-point basis. Accordingly, their phonon spectrum will
1789: feature 3 acoustic modes and $3(n_c - 1)$ optical modes, for which
1790: the oscillation frequency $\omega(k)$ remains finite as $k \to 0$.
1791: We are interested in the case $n_c \gg 1$, i.e., deep in the
1792: region of stability of the crystal, where the clusters have
1793: a very high occupation number. Consequently, the phonon spectra
1794: and the particle displacements will be dominated by the optical
1795: branches. Further, we simplify the problem by choosing, in the
1796: spirit of the Einstein model of the crystal,\cite{ashcroft} one
1797: specific optical phonon with $k = 0$
1798: as a representative for the whole spectrum. This mode corresponds
1799: to the relative partial displacement of two sublattices:
1800: one with $n_c - 1$ particles on each site and one with just
1801: the remaining one particle per site. The two sublattices
1802: coincide at the equilibrium position and maintain their
1803: shape throughout the oscillation mode, consistent with the
1804: fact of an infinite-wavelength mode, $k=0$. Accordingly, the
1805: site potential felt by any one of the particles of the
1806: single-occupied sublattice, $V_{\rm site}({\bf s})$, can
1807: be expressed as
1808: \begin{equation}
1809: V_{\rm site}({\bf s}) = (n_c - 1)\left[v(s) + \sum_{{\bf R} \ne 0}
1810: v(|{\bf s} - {\bf R}|)
1811: \right],
1812: \label{vsite:eq}
1813: \end{equation}
1814: where ${\bf s}$ is the relative displacement of the two sublattices.
1815: For brevity, we also define
1816: \begin{equation}
1817: W_{\rm site}({\bf s}) \equiv \frac{V_{\rm site}({\bf s})}{n_c-1}.
1818: \label{wsite:eq}
1819: \end{equation}
1820:
1821: The Taylor expansion of a scalar function $f$
1822: around a reference point ${\bf r}$ reads as
1823: \begin{equation}
1824: f({\bf r} + {\bf s}) = f({\bf r}) +
1825: {\bf s}\cdot\nabla f({\bf r}) +
1826: \frac{1}{2}\left({\bf s}\cdot\nabla\right)^2 f({\bf r}) + \cdots
1827: \label{taylor3:eq}
1828: \end{equation}
1829: Setting ${\bf r} \to 0$ and $f \to W_{\rm site}$, we obtain the
1830: quadratic expansion of the site potential; the constant $V_{\rm site}(0)$
1831: is unimportant. For the linear term, we have
1832: \begin{equation}
1833: \nabla W_{\rm site}({\bf r}=0) = v'(r){\bf {\hat r}}\Big|_{{\bf r} = 0}
1834: + \sum_{{\bf R} \ne 0} v'(R){\bf {\hat R}},
1835: \label{linear:eq}
1836: \end{equation}
1837: where ${\bf {\hat r}}$ and ${\bf {\hat R}}$ are unit vectors.
1838: The sum in (\ref{linear:eq}) vanishes due to lattice inversion
1839: symmetry; the first term also, since $v'(0) = 0$. Thus, the term
1840: linear in ${\bf s}$ in the expansion of $V_{\rm site}({\bf s})$
1841: vanishes, consistently with the fact that ${\bf s} = 0$ is
1842: an equilibrium position.
1843:
1844: We now introduce
1845: Cartesian coordinates and write
1846: ${\bf r} = (x,y,z)$, ${\bf s} = (s_x,s_y,s_z)$, and
1847: ${\bf R} = (R_x, R_y, R_z)$. The quadratic term in
1848: (\ref{taylor3:eq}) takes the explicit form
1849: \begin{eqnarray}
1850: \nonumber
1851: \frac{1}{2}\left({\bf s}\cdot\nabla\right)^2 & = &
1852: s_xs_y\frac{\partial^2}{\partial x\partial y} +
1853: s_ys_z\frac{\partial^2}{\partial y\partial z} +
1854: s_zs_x\frac{\partial^2}{\partial z\partial x}
1855: \\
1856: & + &
1857: \frac{1}{2}\left(s_x^2\frac{\partial^2}{\partial x^2} +
1858: s_y^2\frac{\partial^2}{\partial y^2} +
1859: s_z^2\frac{\partial^2}{\partial z^2}\right).
1860: \label{explicit:eq}
1861: \end{eqnarray}
1862: Let us consider first the mixed derivative acting on $W_{\rm site}({\bf r})$,
1863: evaluated at ${\bf r} = 0$. Using Eqs.\ (\ref{wsite:eq})
1864: and (\ref{vsite:eq}), we obtain
1865: \begin{eqnarray}
1866: \nonumber
1867: \frac{\partial^2 W_{\rm site}({\bf r})}{\partial x\partial y}\Big|_{{\bf r} = 0}
1868: & = &
1869: \left[\frac{xy}{r^2}\left(v''(r)-\frac{v'(r)}{r}\right)\right]_{{\bf r} = 0}
1870: \\
1871: & + & \sum_{{\bf R}\ne 0}\frac{R_xR_y}{R^2}
1872: \left[v''(R)-\frac{v'(R)}{R}\right].
1873: \label{mixed:eq}
1874: \end{eqnarray}
1875: The first term on the right-hand side of (\ref{mixed:eq}) vanishes by
1876: virtue of (\ref{condition:eq}). The second one also vanishes due to
1877: the cubic symmetry of the lattice. Clearly, the other two terms in
1878: (\ref{explicit:eq})
1879: with mixed derivatives vanish as well. For the remaining terms,
1880: the cubic symmetry of
1881: the lattice implies that all three
1882: second partial derivatives of $W_{\rm site}$ at ${\bf r} = 0$
1883: are equal:
1884: \begin{eqnarray}
1885: \nonumber
1886: \frac{\partial^2 W_{\rm site}({\bf r})}{\partial x^2}\Big|_{{\bf r} = 0}
1887: & = &
1888: \frac{\partial^2 W_{\rm site}({\bf r})}{\partial y^2}\Big|_{{\bf r} = 0}
1889: =
1890: \frac{\partial^2 W_{\rm site}({\bf r})}{\partial z^2}\Big|_{{\bf r} = 0}
1891: \\
1892: & = & \frac{1}{3}\nabla^2W_{\rm site}({\bf r})\Big|_{{\bf r} = 0}.
1893: \label{laplace:eq}
1894: \end{eqnarray}
1895:
1896: Gathering the results, we obtain the expansion of $V_{\rm site}({\bf s})$
1897: to quadratic order in $s$ as
1898: \begin{equation}
1899: V_{\rm site}({\bf s}) = V_{\rm site}(0) + \left[\frac{(n_c - 1)}{6}
1900: \sum_{\bf R}\nabla^2v(R)\right]s^2,
1901: \label{isotropic:eq}
1902: \end{equation}
1903: which is isotropic in $s$, as should for a crystal of cubic symmetry.
1904: The one-particle motion is therefore harmonic; as we consider
1905: $n_c \gg 1$, we set $n_c - 1 \cong n_c$ in (\ref{isotropic:eq})
1906: and we obtain the effective, one-particle Hamiltonian
1907: ${\mathcal H}_1$ in the form:
1908: \begin{equation}
1909: {\mathcal H}_1 = \frac{p^2}{2m} + \kappa s^2,
1910: \label{hamilton:eq}
1911: \end{equation}
1912: with
1913: \begin{equation}
1914: \kappa = \frac{n_c}{6}\sum_{\bf R}\nabla^2 v(R)
1915: \label{kappa:eq}
1916: \end{equation}
1917: and the momentum $p$ and mass $m$ of the particle. The density
1918: profile $\rho_1({\bf r})$ of this single-particle problem is
1919: easily calculated as
1920: $\langle \delta({\bf r} - {\bf s})\rangle_{{\mathcal H}_1}$, yielding
1921: \begin{equation}
1922: \rho_1({\bf r}) = \left(\frac{\beta\kappa}{\pi}\right)^{3/2}
1923: e^{-\beta\kappa r^2}.
1924: \label{rho1:eq}
1925: \end{equation}
1926: This is indeed a Gaussian of a single
1927: particle, with a localization parameter
1928: $\alpha_{\rm h} = \beta\kappa$; the total density on a given
1929: site will be then just $n_c\rho_1({\bf r})$, in agreement with the
1930: functional form put forward in Eq.\ (\ref{gaussian_real:eq}).
1931:
1932: It is useful to consider in detail the form of the
1933: localization parameter $\alpha_{\rm h}$ predicted by the
1934: harmonic theory. The parameter $\kappa$ is expressed as
1935: a sum of the values of $\psi(r) = \nabla^2 v(r)$ over
1936: the periodic set $\{{\bf R}\}$. For every function
1937: $\psi(r)$ that possesses a Fourier transform
1938: $\tilde\psi(k)$, it holds\cite{ashcroft}
1939: \begin{equation}
1940: \sum_{{\bf R}} \psi(R) = \rho_s\sum_{{\bf K}}
1941: \tilde\psi(K),
1942: \label{parseval:eq}
1943: \end{equation}
1944: where $\{{\bf K}\}$ is the set of RLVs of $\{{\bf R}\}$ and
1945: $\rho_s = N_s/V$ is the density of {\it lattice sites} of $\{{\bf R}\}$.
1946: From (\ref{nc:eq}),
1947: $\rho_s n_c = \rho$.
1948: Taking into account that the Fourier transform of $\nabla^2 v(r)$
1949: is $-k^2\tilde v(k)$, we obtain for the localization parameter
1950: of the harmonic theory the result
1951: \begin{equation}
1952: \alpha_{\rm h} = -\frac{\rho}{6k_{\rm B}T}\sum_{{\bf K}}K^2\tilde v(K).
1953: \label{alphah:eq}
1954: \end{equation}
1955: The localization parameter must be, evidently, positive.
1956: Eq.\ (\ref{alphah:eq}) manifests the impossibility for cluster
1957: formation if the Fourier transform of the pair potential is
1958: nonnegative, i.e., for $Q^{+}$ interactions. In the
1959: preceding section, we showed within the DFT formalism that if
1960: the potential is $Q^{\pm}$, this implies the formation of
1961: cluster crystals. Harmonic theory allows us to make the
1962: opposite statement as well: if the potential is {\it not}
1963: $Q^{\pm}$, then there can be {\it no clustered crystals}.
1964: Therefore, an {\it equivalence} between the $Q^{\pm}$
1965: character of the interaction and the formation of clustered
1966: crystals can be established. Moreover, Eq.\ (\ref{alphah:eq})
1967: offers an additional indication as to why the RLV where
1968: $\tilde v(K)$ is most negative is selected as the shortest
1969: nonvanishing one by the clustered
1970: crystals: this is the best strategy in order to keep the localization
1971: parameter positive.
1972:
1973: Harmonic theory provides, therefore,
1974: an insight into the necessity of locating the first shell
1975: of the RLVs at $k_*$ from a different point of view than
1976: density functional theory does.
1977: The choice $K_1 = k_*$ guarantees that the particles
1978: inhabiting neighboring clusters provide the restoring forces that
1979: push any given particle back towards its equilibrium position.
1980: The density functional treatment of the preceding sections
1981: establishes that
1982: the lattice constant is chosen by $Q^{\pm}$-systems in such
1983: a way that the sum of {\it intracluster} and
1984: {\it intercluster} interactions, together with the entropic
1985: penalty for the aggregation of $n_c$ particles is optimized.\cite{bianca:long}
1986: The unlimited growth of $n_c$ is avoided by the requirement of
1987: mechanical stability of the crystal. Indeed, for too high
1988: $n_c$-values, the lattice constant would concomitantly grow,
1989: so that the resulting
1990: restoring forces working against the thermal fluctuations,
1991: would become too weak to sustain the particles at their
1992: equilibrium positions.
1993:
1994: Let us, finally, compare the result (\ref{alphah:eq}) for
1995: the localization parameter with the prediction from DFT.
1996: We consider the high-density crystal phase, for which
1997: $\alpha^*$ is very large, so that the simplification that only
1998: the first shell of RLVs can be kept in (\ref{fren_var:eq}) must be dropped.
1999: Setting $\partial\tilde f/\partial\alpha^* = 0$ there, we
2000: obtain
2001: \begin{equation}
2002: \frac{3T^*}{2\alpha^*} + \frac{\rho^*}{\left(2\alpha^*\right)^2}
2003: \sum_{{\bf Y}}Y^2\tilde\phi(Y)e^{-Y^2/(2\alpha^*)} = 0.
2004: \label{dftall:eq}
2005: \end{equation}
2006: The function $\tilde\phi(y)$ is short-ranged in
2007: reciprocal space, thus the sum in (\ref{dftall:eq})
2008: can be effectively truncated at some finite upper cutoff
2009: $Y_c$. Then, there exists a sufficiently large
2010: density $\rho^*$ beyond which the
2011: parameter $\alpha^*$ is so large that
2012: $Y^2/(2\alpha^*) \ll 1$ for all $Y \leq Y_c$ included in
2013: the summation. Accordingly, we can approximate all exponential factors
2014: with uniti in (\ref{dftall:eq}), obtaining an algebraic equation
2015: for $\alpha^*$. Reverting back to dimensional quantities, its
2016: solution reads as
2017: \begin{equation}
2018: \alpha = -\frac{\rho}{6k_{\rm B}T}\sum_{{\bf K}}
2019: K^2\tilde v(K),
2020: \label{alphasdft:eq}
2021: \end{equation}
2022: and is {\it identical} with the result from the harmonic theory,
2023: Eq.\ (\ref{alphah:eq}). Thus, density functional theory and
2024: harmonic theory become identical to each other at the limit of
2025: high localization. This finding completes and generalizes the result of
2026: Archer,\cite{archer} who established a close relationship between
2027: the mean-field DFT and the Einstein model for the form of the
2028: variational free energy functional of the system.
2029:
2030: Consistently with our assumptions,
2031: $\alpha$ indeed grows with density. In fact, since the set of
2032: RLVs in the sum of (\ref{alphasdft:eq}) is fixed, it can be
2033: seen that $\alpha$ is simply proportional to $\rho/T$.
2034: This peculiar feature of
2035: the class of systems we consider
2036: is not limited to the localization parameter, and its significance
2037: is discussed in the following section.
2038:
2039: \section{Connection with inverse-power potentials}
2040: \label{invpower:sec}
2041:
2042: As can be easily confirmed by the form of the density functional
2043: of Eq.\ (\ref{totalfren:eq}),
2044: the mean-field nature of the class of ultrasoft systems considered
2045: here (both $Q^{+}$- and $Q^{\pm}$-potentials) implies that
2046: the structure and thermodynamics of the systems is fully determined
2047: by the ratio $\rho^{*}/T^{*}$ between density and temperature and
2048: not separately by $\rho^{*}$ and $T^{*}$. This is a particular
2049: type of scaling between the two relevant thermodynamic variables,
2050: reminiscent of the situation for systems interacting by means of
2051: inverse-power-law potentials $v(r)$ having the form:
2052: \begin{equation}
2053: v(r) = \epsilon\left(\frac{\sigma}{r}\right)^n.
2054: \label{powerlaw:eq}
2055: \end{equation}
2056: For such systems, it can be shown that their
2057: statistical mechanics is governed by a
2058: a single coupling constant $\Gamma_D(n)$ expressed as\cite{weeks:prb:81}
2059: \begin{equation}
2060: \Gamma_D(n) =
2061: \frac{\rho^{*}}{\left({T^{*}}\right)^{D/n}},
2062: \label{gamman:eq}
2063: \end{equation}
2064: where $D$ is the space dimension. It would appear that inverse-powers
2065: $n=D$ satisfy precisely the same scaling as mean-field systems do
2066: but there is a condition to be
2067: fulfilled: inverse-power systems are stable against
2068: explosion, {\it provided} that $n > D$; this can be most easily seen
2069: by considering the expression for the excess internal energy
2070: per particle, $u(\rho,T)$, given by\cite{hansen:book}
2071: \begin{equation}
2072: u(\rho,T) = \frac{2\pi^{D/2}\rho\epsilon\sigma^n}{[(D/2)-1]!}
2073: \int_0^{\infty}r^{D-1-n}g(r;\rho,T){\rm d}r.
2074: \label{inten:eq}
2075: \end{equation}
2076: As $g(r) \to 1$ for $r \to \infty$, we see that the integral
2077: in (\ref{inten:eq}) converges only if $n > D$; a logarithmic
2078: divergence results for $n=D$. A `uniform neutralizing background'
2079: has to be formally introduced for $n \leq D$,
2080: to obtain stable
2081: pseudo one-component systems, such as
2082: the one-component plasma.\cite{ocp1,ocp2,lieb}
2083: Since we aim at staying with genuine one-component systems
2084: throughout, we must strictly maintain $n > D$.
2085:
2086: Instead of taking the limit $n \to D$, we consider therefore a
2087: different procedure by setting
2088: \begin{equation}
2089: n = D + \delta
2090: \end{equation}
2091: with some arbitrary, finite $\delta > 0$. Then the coupling constant
2092: $\Gamma_D(n)$ becomes
2093: \begin{equation}
2094: \Gamma_D(D+\delta) = \rho^*\left(T^{*}\right)^{-\left(1+\delta/D\right)^{-1}}.
2095: \label{gammam:eq}
2096: \end{equation}
2097: Now take $D \to \infty$ in this prescribed fashion, obtaining:
2098: \begin{equation}
2099: \lim_{D\to\infty}\Gamma_D(D+\delta) = \frac{{\rho^{*}}}{{T^{*}}},
2100: \label{gammainf:eq}
2101: \end{equation}
2102: which has precisely the same form as the coupling constant of
2103: our systems. In taking the
2104: limit $D \to \infty$, the exponent $n = D + \delta$ of the
2105: inverse-power potential diverges as well. It can be easily
2106: seen that in this case, the interaction $v(r)$ of
2107: Eq.\ (\ref{powerlaw:eq}) becomes a hard-sphere potential of
2108: diameter $\sigma$. In other words, the procedure
2109: prescribed above brings us once more to infinite-dimensional
2110: hard spheres, because also the inverse-power interaction becomes
2111: arbitrarily steep.
2112: This is a very different way of taking the limit
2113: than in Refs.\ [\onlinecite{frisch:pra:87,frisch:pre:99,bagchi:jcp:88}]:
2114: there, the interaction is hard
2115: in the first place and subsequently the limit
2116: $D \to \infty$ is taken,
2117: whereas here, interaction and dimension of space change
2118: together, in a well-prescribed fashion.
2119:
2120: The fact that the
2121: statistical mechanics of ultrasoft fluids in three dimensions
2122: is determined by the same dimensionless parameter as that
2123: of a particular realization of hard spheres in infinite
2124: dimensions is intriguing. In a sense, ultrasoft systems
2125: are effectively high-dimensional, since they allow for
2126: extremely high densities, for which every particle interacts
2127: with an exceedingly high number of neighbors. They might,
2128: in this sense, provide for three-dimensional approximate
2129: realizations of infinite-dimensional models.
2130: This is yet another relation to infinite-dimensional
2131: systems, in addition to the one discussed at the end
2132: of Sec.\ \ref{dft:sec}.
2133: Whether there
2134: exists a deeper mathematical connection between the two classes,
2135: remains a problem for the future.
2136:
2137: \section{Summary and concluding remarks}
2138: \label{summary:sec}
2139:
2140: We have provided a detailed analysis of the properties of
2141: bounded, ultrasoft systems, with emphasis on the $Q^{\pm}$-class
2142: of interaction potentials. After having demonstrated the
2143: suppression of the contributions from the high-order direct correlation
2144: functions of the fluid phases (of order 3 and higher), we
2145: established as a consequence the accurate mean-field density
2146: functional for arbitrary inhomogeneous phases. Though this
2147: functional has been introduced and successfully used
2148: in the recent past both in
2149: statics\cite{criterion,bianca:prl:06,likos:gauss,louis:mfa,archer1,archer2,
2150: archer3,archer4,finken}
2151: and in dynamics,\cite{rex:molphys:06}
2152: a sound justification of its basis on the properties of the
2153: uniform phase was still lacking.
2154:
2155: The persistence of a {\it single}, finite length scale
2156: for the lattice constants of the ensuing
2157: solids of $Q^{\pm}$-systems has been understood by a detailed
2158: analysis of the structure of the free energy functional.
2159: In the fluid, the same length scale appears since the position
2160: of maximum of the liquid structure factor is independent of
2161: density.
2162: The negative minimum of the interaction potential in Fourier
2163: space sets this unique scale and forces in the crystal the
2164: formation of clusters, whose population scales proportionally
2165: with density. The analytical solution of an approximation of
2166: the density functional is checked to be accurate when
2167: confronted with the full numerical minimization of the latter.
2168: Universal Lindemann ratios and Hansen-Verlet values at crystallization
2169: are predicted to hold for all these systems, which
2170: differ substantially from those for hard matter systems.
2171: The analytical
2172: derivation of these results provides useful insight into the
2173: robustness of these structural values for an enormous variety
2174: of interactions.
2175:
2176: Though the assumption of bounded interactions has been made
2177: throughout, recent results\cite{primoz:07}
2178: indicate that both cluster formation and the persistence of
2179: the length scale survive when a short-range diverging core is
2180: superimposed on the ultrasoft potential, provided the range of
2181: the hard core does not exceed, roughly, 20\% of the overall
2182: interaction range.\cite{primoz:07} The morphology of the resulting
2183: clusters is more complex, as full overlaps are explicitly
2184: forbidden; even the macroscopic phases are affected, with
2185: crystals, lamellae, inverted lamellae and `inverted crystals'
2186: showing up at increasing densities. The generalization of our
2187: density functional theory to such situations and the modeling
2188: of the nontrivial, internal cluster morphology is a challenge
2189: for the future. Here, a mixed density functional, employing
2190: a hard-sphere and a mean-field part of the direct correlation
2191: function seems to be a promising way to proceed.\cite{sear,lr1}
2192: Finally, the study of the vitrification, dynamical arrest and
2193: hopping processes in concentrated $Q^{\pm}$-systems is another
2194: problem of current interest. The recent `computer synthesis'
2195: of model, amphiphilic dendrimers that do display precisely
2196: the form of $Q^{\pm}$-interactions discussed in this work,\cite{bianca:dendris}
2197: offers concrete suggestions for the experimental realization
2198: of the hitherto theoretically predicted phenomena.
2199:
2200: \section*{ACKNOWLEDGMENTS}
2201: We are grateful to Andrew Archer for helpful discussions and
2202: a critical reading of the manuscript and to Andras S{\"u}t{\H o}
2203: for helpful comments.
2204: This work has been supported by
2205: the {\"O}sterreichische Forschungsfond under
2206: Project No.\ P17823-N08, as well as by the
2207: Deutsche Forschungsgemeinschaft within the
2208: Collaborative Research Center SFB-TR6,
2209: ``Physics of Colloidal Dispersions in External Fields'',
2210: Project Section C3.
2211:
2212: \appendix
2213: \section{Proof that the GEM-$m$ models with $m > 2$ are
2214: $Q^{\pm}$-potentials}
2215:
2216: Consider the inverse Fourier transform of the
2217: spherically symmetric, bounded pair
2218: potential $v(r)$, reading as
2219: \begin{equation}
2220: v(r) = \frac{1}{2\pi^2}\int_0^{\infty}k^2\tilde v(k)\frac{\sin kr}{kr}{\rm d}k
2221: \label{invft:eq}
2222: \end{equation}
2223: From (\ref{invft:eq}), it is straightforward to
2224: show that the second derivative of $v(r)$ at $r=0$ takes
2225: the form
2226: \begin{equation}
2227: v''(r=0) = -\frac{1}{6\pi^2}\int_0^{\infty}k^4\tilde v(k){\rm d}k.
2228: \label{vpp:eq}
2229: \end{equation}
2230: Evidently,
2231: if $v''(r=0) \geq 0$, then $\tilde v(k)$ {\it must}
2232: have negative parts and hence $v(r)$ is $Q^{\pm}$. For the
2233: GEM-$m$ family, it is easy to show that $v''(r=0) = 0$ for
2234: $m > 2$, thus these members are indeed $Q^{\pm}$, as stated in the
2235: main text. Double-Gaussian potentials of the form
2236: \begin{equation}
2237: v(r) = |\epsilon_1|e^{-(r/\sigma_1)^2} - |\epsilon_2|e^{-(r/\sigma_2)^2},
2238: \label{dgauss:eq}
2239: \end{equation}
2240: with $|\epsilon_1| > |\epsilon_2|$, $\sigma_1 > \sigma_2$,
2241: which feature a local {\it minimum}
2242: at $r = 0$ are also $Q^{\pm}$, for the same reason. Notice,
2243: however, that $v''(r=0) \geq 0$ is a sufficient, not a necessary
2244: condition for membership in the $Q^{\pm}$-class. Thus, there exist
2245: $Q^{\pm}$-potentials for which $v''(r=0) < 0$.
2246:
2247: \section{Proof of the equivalence between the variational
2248: free energies $\tilde f$ and $\bar f$.}
2249:
2250: The introduction of the new variable $\gamma$ instead of
2251: $\alpha^*$, Eq.\ (\ref{amin_nc:eq}), and the subsequent new
2252: form $\bar f$ of the variational free energy, Eq.\ (\ref{fbar:eq}),
2253: are just a matter of convenience, which makes the minimization
2254: procedure more transparent. A free gift of the
2255: variable transformation is also the ensuing diagonal
2256: form of the Hessian matrix at the extremum. Fully equivalent
2257: results are obtained, of course, by working with the original
2258: variational free energy, $\tilde f$, Eq.\ (\ref{fren_var:eq}).
2259: Here we explicitly demonstrate this equivalence.
2260:
2261: Keeping, consistently, only the first shell of RLVs with length
2262: $Y_1$, $\tilde f$ takes the form:
2263: \begin{eqnarray}
2264: \nonumber
2265: \tilde f(n_c,\alpha^*;T^*,\rho^*)
2266: & = &
2267: T^*\left[\ln n_c + \frac{3}{2}\ln\left(\frac{\alpha^*}{\pi}\right)
2268: - \frac{5}{2}\right]
2269: \\
2270: \nonumber
2271: & + & \frac{\rho^*}{2}\tilde\phi(0)
2272: \\
2273: & + &
2274: \frac{\xi_1\rho^*}{2}\tilde\phi(Y_1)
2275: e^{-Y_1^2/(2\alpha^*)},
2276: \label{fren_new:eq}
2277: \end{eqnarray}
2278: where $Y_1$ and $n_c$ are related via Eq.\ (\ref{d:eq}). Minimizations
2279: of $\tilde f$ with respect to $n_c$ and $\alpha^*$ yield,
2280: respectively:
2281: \begin{equation}
2282: T^* -\frac{\xi_1\rho^*Y_1}{6}\left[\frac{\partial\tilde\phi(Y_1)}{\partial Y_1}
2283: -\frac{Y_1}{\alpha^*}\tilde\phi(Y_1)\right]e^{-Y_1^2/(2\alpha^*)} = 0,
2284: \label{delnc:eq}
2285: \end{equation}
2286: and
2287: \begin{equation}
2288: T^* + \frac{\xi_1\rho^*}{6\alpha^*}
2289: \tilde\phi(Y_1)Y_1^2e^{-Y_1^2/(2\alpha^*)} = 0.
2290: \label{dela:eq}
2291: \end{equation}
2292: Subtracting the last two equations from one another we obtain
2293: \begin{equation}
2294: \frac{\xi_1\rho^*}{6}\tilde\phi'(Y_1)Y_1
2295: e^{-Y_1^2/(2\alpha^*)} = 0,
2296: \end{equation}
2297: implying $Y_1 = y_*$, as in the main text (once more, $Y_1 = 0$
2298: is a formal solution that must
2299: be rejected on the same grounds mentioned in the text.) From this property,
2300: Eq.\ (\ref{proport:eq})
2301: immediately follows. Introducing
2302: $\Delta \tilde f \equiv \tilde f - f_{\rm liq}$, we can
2303: determine $\alpha^*_{\rm f}$
2304: on the freezing line by requiring the simultaneous satisfaction of
2305: the minimization conditions above and of $\Delta\tilde f = 0$. The latter
2306: equation yields
2307: \begin{eqnarray}
2308: \nonumber
2309: T^*\left[\ln\left(\frac{n_c}{\rho^*}\right)
2310: + \frac{3}{2}\left[\ln\left(\frac{\alpha^*_{\rm f}}{\pi}\right) - 1\right]\right]
2311: \\
2312: = - \frac{\xi_1\rho^*}{2}\tilde\phi(y_*)e^{-y_*^2/(2\alpha^*_{\rm f})},
2313: \label{delftildezero:eq}
2314: \end{eqnarray}
2315: which, together with Eq.\ (\ref{dela:eq}), yields, after some
2316: algebra
2317: \begin{equation}
2318: \ln\left[
2319: \left(\frac{\alpha^*_{\rm f}}{\pi}\right)
2320: \left(\frac{n_c}{\rho^*}\right)^{2/3}\right] - 1
2321: = \frac{2\alpha^*_{\rm f}}{y_*^2}.
2322: \label{onemore:eq}
2323: \end{equation}
2324: Using Eqs.\ (\ref{amin_nc:eq}) and (\ref{y1nc:eq}), we
2325: obtain the equation for the $\gamma_{\rm f}$-parameter at freezing as
2326: \begin{equation}
2327: -\left[
2328: \ln\left(z^{2/3}\gamma_{\rm f}\pi \right) + 1\right] =
2329: \frac{2}{\gamma_{\rm f}\zeta^2},
2330: \label{gamma_new:eq}
2331: \end{equation}
2332: which, upon setting $z=2$ and $\zeta = 2\sqrt{2}\pi$, yields
2333: Eq.\ (\ref{gamma:eq}) of the main text. Alternatively, we can
2334: introduce the
2335: variable $t \equiv \alpha^*_{\rm f}/y_*^2$ and rewrite Eq.\ (\ref{onemore:eq})
2336: as
2337: \begin{equation}
2338: \ln\left[\left(8\sqrt{2}\right)^{2/3}\pi t\right] - 1 = 2t,
2339: \label{omega:eq}
2340: \end{equation}
2341: which delivers $t \cong 0.704$ as a
2342: solution or, equivalently, $t = (8\pi^2\gamma_{\rm f})^{-1}$,
2343: in agreement with Eqs.\ (\ref{gammaf:eq}) and (\ref{alphaf:eq}) of the main text.
2344:
2345:
2346: \begin{thebibliography}{99}
2347:
2348: \bibitem{sear} R. Sear and W. M. Gelbart, J. Chem. Phys. {\bf 110},
2349: 4582 (1999).
2350:
2351: \bibitem{fs1} F. Sciortino, S. Mossa, E. Zaccarelli, and
2352: P. Tartaglia, Phys. Rev. Lett. {\bf 93}, 055701 (2004).
2353:
2354: \bibitem{fs2} S. Mossa, F. Sciortino, P. Tartaglia,
2355: and E. Zaccarelli, Langmuir {\bf 20}, 10756 (2004).
2356:
2357: \bibitem{bartlett1} A. I. Campbell, V. J. Anderson, J. S. van Duijneveldt,
2358: and P. Bartlett, Phys. Rev. Lett. {\bf 94}, 208301 (2005).
2359:
2360: \bibitem{bartlett2} R. Sanchez and P. Bartlett, J. Phys.: Condensed Matter
2361: {\bf 17}, S3351 (2005).
2362:
2363: \bibitem{lr1} A. Imperio and L. Reatto, J. Phys.: Condens. Matter
2364: {\bf 16}, S3769 (2004).
2365:
2366: \bibitem{lr2} A. Imperio and L. Reatto, J. Chem. Phys. {\bf 124},
2367: 164712 (2006).
2368:
2369: \bibitem{jpcm} C. N. Likos, C. Mayer, E. Stiakakis, and G. Petekidis,
2370: J. Phys.: Condens. Matter {\bf 17}, S3363 (2005).
2371:
2372: \bibitem{epl} E. Stiakakis, G. Petekidis, D. Vlassopoulos,
2373: C. N. Likos, H. Iatrou, N. Hadjichristidis, and J. Roovers,
2374: Europhys. Lett. {\bf 72}, 664 (2005).
2375:
2376: \bibitem{likos:pr:01} C. N. Likos, Phys. Rep. {\bf 348}, 267 (2001).
2377:
2378: \bibitem{krueger:89} B. Kr{\"u}ger, L. Sch{\"a}fer, and
2379: A. Baumg{\"a}rtner, J. Phys. (France) {\bf 50}, 319 (1989).
2380:
2381: \bibitem{hall:94} J. Dautenhahn and C. K. Hall, Macromolecules {\bf 27},
2382: 5399 (1994).
2383:
2384: \bibitem{louis:prl:00} A. A. Louis, P. G. Bolhuis, J.-P. Hansen,
2385: and E. J. Meijer, Phys. Rev. Lett. {\bf 85}, 2522 (2000).
2386:
2387: \bibitem{ingo:jcp:04} I. O. G{\"o}tze, H. M. Harreis, and
2388: C. N. Likos, J. Chem. Phys. {\bf 120}, 7761 (2004).
2389:
2390: \bibitem{ballik:ac:04} M. Ballauff and C. N. Likos, Angew. Chemie
2391: Intl. English Ed. {\bf 43}, 2998 (2004).
2392:
2393: \bibitem{denton:03} A. R. Denton, Phys. Rev. E {\bf 67}, 011804 (2003);
2394: Erratum, ibid.\ {\bf 68}, 049904 (2003).
2395:
2396: \bibitem{gottwald:04} D. Gottwald, C. N. Likos, G. Kahl, and H. L{\"o}wen,
2397: Phys. Rev. Lett. {\bf 92}, 068301 (2004).
2398:
2399: \bibitem{gottwald:05} D. Gottwald, C. N. Likos, G. Kahl, and H. L{\"o}wen,
2400: J. Chem. Phys. {\bf 122}, 074903 (2005).
2401:
2402: \bibitem{pierleoni:prl:06} C. Pierleoni, C. Addison, J.-P. Hansen,
2403: and V. Krakoviack, Phys. Rev. Lett. {\bf 96}, 128302 (2006).
2404:
2405: \bibitem{hansen:molphys:06} J.-P. Hansen and C. Pearson, Mol. Phys. {\bf 104},
2406: 3389 (2006).
2407:
2408: \bibitem{suto:prl:05} A. S{\"u}t{\H o}, Phys. Rev. Lett. {\bf 95}, 265501
2409: (2005).
2410:
2411: \bibitem{suto:prb:06} A. S{\"u}t{\H o}, Phys. Rev. B {\bf 74}, 104117 (2006).
2412:
2413: \bibitem{klein:94} W. Klein, H. Gould, R. A. Ramos, I. Clejan,
2414: and A. I. Melcuk, Physica (Amsterdam) {\bf 205A}, 738 (1994).
2415:
2416: \bibitem{pensph:98} C. N. Likos, M. Watzlawek, and H. L{\"o}wen,
2417: Phys. Rev. E {\bf 58}, 3135 (1998).
2418:
2419: \bibitem{criterion} C. N. Likos, A. Lang, M. Watzlawek, and H. L{\"o}wen,
2420: Phys. Rev. E {\bf 63}, 031206 (2001).
2421:
2422: \bibitem{bianca:prl:06} B. M. Mladek, D. Gottwald, G. Kahl, M. Neumann,
2423: and C. N. Likos, Phys. Rev. Lett. {\bf 96}, 045701 (2006); Erratum,
2424: {\it ibid.}\ {\bf 97}, 019901 (2006).
2425:
2426: \bibitem{stillinger:physica} F. H. Stillinger and D. K. Stillinger,
2427: Physica (Amsterdam) {\bf 244A}, 358 (1997).
2428:
2429: \bibitem{likos:gauss} A. Lang, C. N. Likos, M. Watzlawek, and H. L{\"o}wen,
2430: J. Phys.: Condens. Matter {\bf 12}, 5087 (2000).
2431:
2432: \bibitem{saija1} S. Prestipino, F. Saija, and P. V. Giaquinta,
2433: Phys. Rev. E {\bf 71}, 050102 (2005),
2434:
2435: \bibitem{saija2} F. Saija and P. V. Giaquinta, Chem. Phys. Chem. {\bf 6},
2436: 1768 (2005)
2437:
2438: \bibitem{saija3} S. Prestipino, F. Saija, and P. V. Giaquinta,
2439: J. Chem. Phys. {\bf 123}, 144110 (2005).
2440:
2441: \bibitem{archer} A. J. Archer, Phys. Rev. E {\bf 72}, 051501 (2006).
2442:
2443: \bibitem{daan:nature} D. Frenkel, Nature (London) {\bf 440}, 4-5 (02 March 2006).
2444:
2445: \bibitem{daan:web} See the extensive commentary by D. Frenkel in:
2446: Journal Club for Condensed Matter Physics, selection of March 2006,
2447: link
2448: http://www.bell-labs.com/jc-cond-mat/march/march\_2006.html.
2449:
2450: \bibitem{primoz:07} M.~A.~Glaser, G.~M.~Grason, R.~D.~Kamien,
2451: A.~Ko{\v s}mrlj, C.~D.~Santagelo, and P.~Ziherl, preprint,
2452: cond-mat/0609570.
2453:
2454: \bibitem{Ruelle:book} D. Ruelle, {\it Statistical Mechanics}
2455: (Benjamin, New York, 1969).
2456:
2457: \bibitem{fernaud} M.-J. Fernaud, E. Lomba, and L. L. Lee,
2458: J. Chem. Phys. {\bf 112}, 810 (2000).
2459:
2460: \bibitem{hansen:book} J.-P. Hansen and I. R. McDonald,
2461: {\it Theory of Simple Liquids}, 3rd ed. (Elsevier, Amsterdam, 2006).
2462:
2463: \bibitem{percus:62} J. K. Percus, Phys. Rev. Lett. {\bf 8}, 462 (1962).
2464:
2465: \bibitem{percus:64} J. K. Percus, in {\it The Equilibrium Theory of
2466: Classical Fluids}, edited by H. L. Fritsch and J. L. Lebowitz
2467: (Benjamin, New York, 1964).
2468:
2469: \bibitem{denton:pra:91} A. R. Denton and N. W. Aschroft,
2470: Phys. Rev. A {\bf 44}, 1219 (1991).
2471:
2472: \bibitem{yethiraj:jcp:01} A. Yethiraj, H. Fynewever, and
2473: C.-Y. Shew, J. Chem. Phys. {\bf 114}, 4323 (2001).
2474:
2475: \bibitem{singh} Y. Singh, Phys. Rep. {\bf 207}, 351 (1991).
2476:
2477: \bibitem{evans:78} R. Evans, Adv. Phys. {\bf 28}, 143 (1979).
2478:
2479: \bibitem{ry:79} T. V. Ramakrishnan and M. Yussouff, Phys. Rev. B {\bf 19},
2480: 2775 (1979).
2481:
2482: \bibitem{haymet:81} A. D. J. Haymet and D. W. Oxtoby, J. Chem. Phys. {\bf 74},
2483: 2559 (1981).
2484:
2485: \bibitem{denton:pra:90} A. R. Denton and N. W. Ashcroft, Phys. Rev. A {\bf 41},
2486: 2224 (1990).
2487:
2488: \bibitem{ocp1} M. Baus and J.-P. Hansen, Phys. Rep. {\bf 59}, 1 (1980).
2489:
2490: \bibitem{ocp2} S. Ichimaru, H. Iyetomi, and S. Tanaka,
2491: Phys. Rep. {\bf 149}, 91 (1987).
2492:
2493: \bibitem{barrat:prl:87} J.-L. Barrat, J.-P. Hansen, and G. Pastore,
2494: Phys. Rev. Lett. {\bf 58}, 2075 (1987).
2495:
2496: \bibitem{barrat:molphys:88} J.-L. Barrat, J.-P. Hansen, and G. Pastore,
2497: Mol. Phys. {\bf 63}, 747 (1988).
2498:
2499: \bibitem{denton:pra:89} A. R. Denton and N. W. Ashcroft, Phys. Rev. A {\bf 39},
2500: 426 (1989).
2501:
2502: \bibitem{likos:pusey} C. N. Likos, N. Hoffmann, H. L{\"o}wen, and
2503: A. A. Louis, J. Phys.: Condens. Matter {\bf 14}, 7681 (2002).
2504:
2505: \bibitem{ingo:06} I. O. G{\"o}tze, A. J. Archer, and C. N. Likos,
2506: J. Chem. Phys. {\bf 124}, 084901 (2006).
2507:
2508: \bibitem{louis:mfa} A. A. Louis, P. G. Bolhuis, and J.-P. Hansen,
2509: Phys. Rev. E {\bf 62}, 7961 (2000).
2510:
2511: \bibitem{archer1} A. J. Archer and R. Evans, Phys. Rev. E {\bf 64},
2512: 041501 (2001).
2513:
2514: \bibitem{archer2} A. J. Archer and R. Evans, J. Phys.: Condens. Matter
2515: {\bf 14}, 1131 (2002).
2516:
2517: \bibitem{archer3} A. J. Archer, C. N. Likos, and R. Evans,
2518: J. Phys.: Condens. Matter {\bf 14}, 12031 (2002).
2519:
2520: \bibitem{archer4} A. J. Archer, C. N. Likos, and R. Evans,
2521: J. Phys.: Condens. Matter {\bf 16}, L297 (2004).
2522:
2523: \bibitem{finken} R. Finken, J.-P. Hansen, and A. A. Louis,
2524: J. Phys. A {\bf 37}, 577 (2004).
2525:
2526: \bibitem{stell} G. Stell, J. Stat. Phys. {\bf 78}, 197 (1995).
2527:
2528: \bibitem{ciach} A. Ciach, W. T. G{\' o}{\' z}d{\' z}, and R. Evans,
2529: J. Chem. Phys. {\bf 118}, 9726 (2003).
2530:
2531: \bibitem{kirkwood:51} J. G. Kirkwood, in {\it Phase Transitions in Solids},
2532: ed. by R. Smoluchowski, J. E. Meyer, and A. Weyl (Wiley, New York, 1951).
2533:
2534: \bibitem{still1} F. H. Stillinger, J. Chem. Phys. {\bf 65}, 3968 (1976).
2535:
2536: \bibitem{still2} F. H. Stillinger and T. A. Weber, J. Chem. Phys. {\bf 68},
2537: 3837 (1978).
2538:
2539: \bibitem{still3} F. H. Stillinger and T. A. Weber, Phys. Rev. B {\bf 22},
2540: 3790 (1980).
2541:
2542: \bibitem{still4} F. H. Stillinger, J. Chem. Phys. {\bf 70}, 4067 (1979).
2543:
2544: \bibitem{still5} F. H. Stillinger, Phys. Rev. B {\bf 20}, 299 (1979).
2545:
2546: \bibitem{gerhard:00} S. Jorge, G. Kahl, E. Lomba, and J.~L.~F.~Abascal,
2547: J. Chem. Phys. {\bf 113}, 3302 (2000).
2548:
2549: \bibitem{likos:prl:92} C. N. Likos and N. W. Ashcroft, Phys. Rev. Lett. {\bf 69},
2550: 316 (1992); Erratum, ibid., {\bf 69}, 3134 (1992).
2551:
2552: \bibitem{likos:jcp:93} C. N. Likos and N. W. Ashcroft, J. Chem. Phys. {\bf 99},
2553: 9090 (1993).
2554:
2555: \bibitem{wda1} W. A. Curtin and N. W. Ashcroft, Phys. Rev. A {\bf 32}, 2909
2556: (1985).
2557:
2558: \bibitem{wda2} W. A. Curtin and N. W. Ashcroft, Phys. Rev. Lett. {\bf 56},
2559: 2775 (1986); Erratum, ibid., {\bf 57}, 1192 (1986).
2560:
2561: \bibitem{mwda} A. R. Denton and N. W. Ashcroft, Phys. Rev. A {\bf 39}, 4701
2562: (1989).
2563:
2564: \bibitem{henderson} R. Evans, in {\it Fundamentals of Inhomogeneous Fluids},
2565: ed. by D. Henderson (Marcel Dekker, New York 1992).
2566:
2567: \bibitem{grewe1} N. Grewe and W. Klein, J. Math. Phys. {\bf 18}, 1729 (1977).
2568:
2569: \bibitem{grewe2} N. Grewe and W. Klein, J. Math. Phys. {\bf 18}, 1735 (1977).
2570:
2571: \bibitem{frisch:prl:85} H. L. Frisch, N. Rivier, and D. Wyler,
2572: Phys. Rev. Lett. {\bf 54}, 2061 (1985).
2573:
2574: \bibitem{frisch:jcp:86} W. Klein and H. L. Frisch, J. Chem. Phys. {\bf 84},
2575: 968 (1986).
2576:
2577: \bibitem{frisch:pra:87} H. L. Frisch and J. K. Percus, Phys. Rev. A {\bf 35},
2578: 4696 (1987).
2579:
2580: \bibitem{frisch:pre:99} H. L. Frisch and J. K. Percus, Phys. Rev. E {\bf 60},
2581: 2942 (1999).
2582:
2583: \bibitem{bagchi:jcp:88} B. Bagchi and S. A. Rice, J. Chem. Phys. {\bf 88},
2584: 1177 (1987).
2585:
2586: \bibitem{ashcroft} N. W. Ashcroft and N. D. Mermin, {\it Solid State
2587: Physics} (Holt Saunders, Philadelphia, 1976).
2588:
2589: \bibitem{foot1}
2590: Strictly speaking, there is no freezing
2591: {\it line} on the density-temperature plane but rather a domain
2592: of coexistence, since it is equality of the grand potentials that
2593: determines phase coexistence. Nevertheless, the condition of
2594: equality of the Helmholtz free energies determines a line that
2595: runs between the melting- and crystallization lines and therefore
2596: serves as a reliable locator of the freezing transition.
2597:
2598: \bibitem{lindemann} F. A. Lindemann, Phys. Z. {\bf 11}, 609 (1910).
2599:
2600: \bibitem{shapiro} J. N. Shapiro, Phys. Rev. B {\bf 1}, 3982 (1970).
2601:
2602: \bibitem{hv1} J.-P. Hansen and L. Verlet, Phys. Rev. {\bf 184}, 151 (1969).
2603:
2604: \bibitem{hv2} J.-P. Hansen and D. Schiff, Mol. Phys. {\bf 25}, 1281 (1973).
2605:
2606: \bibitem{dieter:jcp:05} D. Gottwald, G. Kahl, and C. N. Likos,
2607: J. Chem. Phys. {\bf 122}, 204503 (2005).
2608:
2609: \bibitem{bianca:long} B. M. Mladek, D. Gottwald, G. Kahl, M. Neumann,
2610: and C. N. Likos, in preparation (2007).
2611:
2612: \bibitem{foot2}
2613: A very important exception
2614: is the hard-sphere potential, for which no harmonic expansion
2615: can be made. The fact that hard-sphere crystals do still
2616: have density profiles that are very well modeled by Gaussians,
2617: calls for a different explanation there. It can be argued that
2618: the huge number of uncorrelated collisions with the neighbors,
2619: together with the central limit theorem, are responsible for
2620: the Gaussianity of density profiles in hard-sphere solids.
2621:
2622: \bibitem{weeks:prb:81} J.~D.\ Weeks, Phys.\ Rev.\ B {\bf 24}, 1530 (1981).
2623:
2624: \bibitem{lieb} E. Lieb, Rev. Mod. Phys. {\bf 48}, 553 (1976).
2625:
2626: \bibitem{rex:molphys:06} M. Rex, C. N. Likos, H. L{\"o}wen, and
2627: J. Dzubiella, Mol. Phys. {\bf 104}, 527 (2006) and references therein.
2628:
2629: \bibitem{bianca:dendris} B. M. Mladek, G. Kahl, and C. N. Likos,
2630: in preparation (2007).
2631:
2632: \end{thebibliography}
2633: \end{document}
2634:
2635:
2636: