Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

This article was downloaded by: [Adler, S]

On: 3 February 2011


Access details: Access Details: [subscription number 791040946]
Publisher Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-
41 Mortimer Street, London W1T 3JH, UK

Molecular Physics
Publication details, including instructions for authors and subscription information:
http://www.informaworld.com/smpp/title~content=t713395160

Simulations of phase transitions and free energies for ionic systems


Antti-Pekka Hynninena; Athanassios Z. Panagiotopoulosa
a
Department of Chemical Engineering and, Institute for the Science and Technology of Materials,
Princeton University, Princeton, NJ 08540, USA

First published on: 01 December 2010

To cite this Article Hynninen, Antti-Pekka and Panagiotopoulos, Athanassios Z.(2008) 'Simulations of phase transitions
and free energies for ionic systems', Molecular Physics, 106: 16, 2039 — 2051, First published on: 01 December 2010
(iFirst)
To link to this Article: DOI: 10.1080/00268970802112160
URL: http://dx.doi.org/10.1080/00268970802112160

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

This article may be used for research, teaching and private study purposes. Any substantial or
systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or
distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses
should be independently verified with primary sources. The publisher shall not be liable for any loss,
actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly
or indirectly in connection with or arising out of the use of this material.
Molecular Physics
Vol. 106, Nos. 16–18, 20 August–20 September 2008, 2039–2051

INVITED TOPICAL REVIEW


Simulations of phase transitions and free energies for ionic systems
Antti-Pekka Hynninen and Athanassios Z. Panagiotopoulos*
Department of Chemical Engineering and Institute for the Science and Technology of Materials, Princeton University,
Princeton, NJ 08540, USA
(Received 26 February 2008; final version received 7 April 2008)

A review of simulation studies of phase equilibria and free energies for systems dominated by coulombic
interactions is presented. Phase transitions occur for ionic systems in the strong-coupling limit realized in low-
dielectric constant solvents, at low temperatures, or for high charge valences. The majority of simulation results
to date are for primitive models that treat the solvent as a uniform dielectric continuum. Transitions involving
fluid and solid phases for such models have been studied extensively in the past decade. There is now strong
evidence that the vapour–liquid transition is in the Ising universality class. For highly charged colloids the
vapour–liquid transition becomes metastable with respect to the fluid–solid transition and the behaviour matches
that of charged hard plates. Phase transitions of charged chains illustrate sensitivity of the phase behaviour to the
charge pattern. Studies of salt solubilities using models with explicit solvent suggest that reasonable agreement
with experiment can be achieved with existing force fields, but there is considerable room for improvement. Areas
Downloaded By: [Adler, S] At: 15:56 3 February 2011

of future research needs are briefly discussed.


Keywords: ionic systems; Monte Carlo; phase transitions; primitive model; charged colloids

1. Introduction They are relevant for systems as diverse as colloids,


A detailed understanding of interactions between micelles, red-blood cells, bundling of DNA, and
charged objects in solution is important for environ- aggregation of F-actin, microtubules, and viruses.
mental science, biology and applied chemistry. While originally controversial, this is now an accepted
Technologically relevant systems for which coulombic idea strongly supported by experimental evidence [5].
interactions are important include molten and Significant progress has been made in recent years
dissolved salts, charge-stabilized colloids and micellar in theoretical understanding and atomistic simulation
solutions of ionic surfactants. Proteins and nucleic of the structure, thermodynamics and phase transitions
acids carry charges along their backbones and are of strongly interacting ionic systems. Nevertheless,
strongly influenced by the presence of ions in solution. atomistic simulations in the presence of explicit solvent
Multivalent salts and charged proteins are potent still have difficulty reaching the length and time scales
agents for DNA aggregation and precipitation [1] and relevant for phase transitions and achieving adequate
exert a strong influence on the structure and dynamics sampling of phase space. Much of the work on phase
of highly charged polyelectrolyte solutions [2]. The transitions in ionic systems has been performed on
classic Poisson–Boltzmann approach provides excel- coarse grained, ‘primitive models’ that contain no
lent results at low electrolyte concentrations in aqueous explicit solvent. Even for these simplified models,
solutions, but breaks down for higher concentrations, simulations of Coulomb-dominated systems can
multivalent ions and low dielectric-constant solvents encounter significant sampling difficulties. For a
[3,4]. Phase and conformational transitions in ionic simple 1:1 primitive model electrolyte at conditions
systems take place in the ‘strong-interaction’ limit at near the liquid-vapour critical point, the thermal
which the Poisson–Boltzmann approach is not valid. A energy is only 1/20th of the energy of two ions at
key aspect of strongly interacting ionic systems is the contact. As a result, the systems are strongly associat-
existence of effective attractions between like-charged ing and hard to equilibrate. Early simulation work
objects in solution. These attractions result from produced results that (in retrospect) turned out to
correlations between ions and counterions and man- suffer from inadequate sampling and equilibration
ifest themselves in low-dielectric constant environ- problems, even though most of the qualitative features
ments, low temperatures or for highly charged salts. of the observed behaviour were correct. Partly as a

*Corresponding author. Email: azp@princeton.edu

ISSN 0026–8976 print/ISSN 1362–3028 online


ß 2008 Taylor & Francis
DOI: 10.1080/00268970802112160
http://www.informaworld.com
2040 A.-P. Hynninen and A.Z. Panagiotopoulos

result of the increased availability of consistent reduced system of units so that temperature is normal-
simulation data, improved theories of systems domi- ized by the energy of interaction of two ions at contact,
nated by coulombic interactions have become available
4""0 kT
in the past few years [6–10]. Unlike earlier theoretical T ¼ ð2Þ
approaches [11,12], these predict the correct trends q2
with respect to size and valence asymmetries. Recent where k is Boltzmann’s constant. Another common
simulation studies have also contributed towards measure of the strength of interactions is the Bjerrum
clarification of the character of ionic criticality, a length,
question of great theoretical importance [13–15] and
the subject of many experimental investigations B ¼ e2 =""0 kB T ¼ =T  ð3Þ
[e.g. 16–18]. Recent reviews of theoretical tools for
The total number density in the system is
charged systems [19] and ionic criticality [20] are
available. A review of the statics and dynamics of Nþ þ N 3
 ¼  ð4Þ
charged soft matter [21] covers in detail coulomb- V
dominated interactions between charged surfaces, where Nþ ¼ N is the number of positive and negative
while an earlier review [22] discusses effective interac- particles and V is the system volume.
tions between double layers and initial studies of phase Because of its simplicity, the RPM has been
transitions of charged colloidal systems. Older work on extensively studied by both simulations and theoretical
structure and thermodynamics of polyelectrolytes has methods. Vorontsov-Velyaminov et al. [25,26] were the
been reviewed in [23].
Downloaded By: [Adler, S] At: 15:56 3 February 2011

first to propose that the RPM has a liquid–vapour


The objective of the present review, which expands transition and critical point; using constant-pressure
and updates a shorter article [24], is to summarize the Monte Carlo simulations; they obtained Tc  0.095
state-of-the-art on simulation studies of the phase and c  0.17. Stell, Wu and Larsen [27] used equation
behaviour of systems dominated by coulombic inter- of state data from simulations and theoretical approx-
actions. In Section 2, we review the current status of imations of the free energy to obtain Tc  0.085 and
vapour-liquid and solid–fluid equilibria for the c  0.01. While the estimated values of the critical
restricted primitive model of ionic solutions. Other parameters differ significantly from ‘modern’ values
spherically symmetric ionic systems, including charge- given in Table 1, these early studies established firmly
and size- asymmetric primitive models and charged
that the system has a low-temperature vapour-liquid
colloids are the focus of Section 3. Section 4 is devoted
transition.
to charged chains and networks and Section 5 to ionic In the early 1990s, Panagiotopoulos [28] obtained
liquid phase equilibria and the modelling of solubilities an estimate of the critical point as Tc ¼ 0.056 and
and activity coefficients in salt solutions. Finally, c ¼ 0.04 using Gibbs ensemble Monte Carlo [29] with
Section 6 summarizes the current status and touches single-ion transfers. Orkoulas and Panagiotopoulos
upon unresolved questions and research needs. [30] introduced biased pair transfers used by many
subsequent investigations and obtained Tc ¼ 0.053 and
c ¼ 0.025. It is now known that the Gibbs ensemble is
2. The restricted primitive model
not well suited to high precision phase coexistence
A simple description of systems dominated by calculations in the immediate vicinity of critical points
coulombic interactions is provided by the ‘restricted [31]. Clearly, significant uncertainties remained regard-
primitive model’ (RPM) that consists of charged hard ing the location of the critical point of the RPM ten
spheres of equal diameter  in a uniform continuum of years ago.
dielectric constant ". Half of the spheres carry a
positive and half a negative charge of identical
magnitude, q. The energy of interaction between two Table 1. Critical parameters for the gas–liquid transition of
particles is the RPM.
(
þ1 rij <  Ref. Tc c
Uij ¼ qi qj ð1Þ
rij   44 0.04933  0.00005 0.075  0.001
4""0 rij
88 0.04917  0.00002 0.080  0.005
where rij is the distance between spheres i and j, qi and 39 0.0490  0.0003 0.070  0.005
40 0.0492  0.0003 0.062  0.005
qj their respective charges and "0 the dielectric
41 0.0489  0.0003 0.076  0.003
permittivity of vacuum. It is customary to choose a
Molecular Physics 2041

A methodological advance in calculations of the two most recent estimates [44,38] disagree by more
critical points of fluids that do not have particle–hole than their combined error bars but the difference is
symmetry was the mixed-field finite-size scaling only 0.3% of the value of Tc . The other simulations
approach of Bruce and Wilding [32,33]. The method listed in the Table (with more generous error bars)
is based on matching results for the scaled order overlap these two. This difference may be due to the
parameter probability distribution to the universal use of the Bruce–Wilding approach in [38] versus an
curve appropriate for the universality class of the unbiased extrapolation with no assumption of uni-
system under study. For bulk fluids, the appropriate versality class in [44]. The situation for the critical
scaling variable is a linear combination of density and density is less satisfactory, with large relative uncer-
energy, assuming that there is no pressure mixing in the tainties and several differences outside the combined
scaling fields [34]. This method has been used by most simulation uncertainties. Kim and Fisher [34] sug-
simulation studies of the vapour–liquid critical point of gested that the use of the Bruce–Wilding method with
the RPM in the past decade. The method works well in no pressure mixing could lead to unreliable estimates
combination with the grand canonical histogram of the critical density, but is not likely to result in
reweighting approach of Ferrenberg and Swendsen major errors in the critical temperature. Clearly, the
[35], which allows extrapolation of results from a precise determination of the critical density for the
simulation run to a range of chemical potentials and RPM remains a topic for future work.
temperatures in its vicinity. It has already been mentioned that several recent
The critical parameters of the RPM were obtained studies used a discretized version of the RPM. In this
using the Bruce–Wilding approach by Caillol et al. approach, calculations are performed on a simple cubic
Downloaded By: [Adler, S] At: 15:56 3 February 2011

[36–38] in hyperspherical boundary conditions and lattice of spacing l less than the particle diameter ,
Orkoulas and Panagiotopoulos [39] in standard cubic with pre-computed interactions between all lattice sites
boundary conditions. Yan and de Pablo [40] used it in for computational efficiency [49,50]. Parameter  ¼ /l
combination with hyper-parallel tempering Monte controls how closely the model approximates contin-
Carlo that allows for multiple replicas of the simula- uous space. For  ¼ 1, the lattice-discretized RPM has
tion at different conditions of temperature and a tricritical point and an order–disorder transition, as
chemical potential. Panagiotopoulos [41] obtained the first observed by Dickman and Stell [51]. The lattice
critical parameters of the continuum RPM by extra- RPM model with  ¼ 2 was also found [49] to have
polating results on the finely discretized lattice analog phase behaviour qualitatively different from the
to the RPM to infinitely fine discretization. continuum model. However, for   3 the phase
The Bruce–Wilding approach used in these studies behaviour is qualitatively identical to the continuum;
of RPM criticality assumes that the system is in the critical point and coexistence curves match continuum
three-dimensional Ising universality class. At most, one data within a few percent. Lattice discretization effects
may claim consistency with Ising critical behaviour. An have been subsequently analyzed in detail using
unbiased finite-size extrapolation method using the theoretical approaches [52–55] and simulations [41,44].
grand canonical ensemble [42] was applied to a Several systems closely related to the RPM have
discretized version of the RPM by Luijten et al. [43]. also been studied with respect to their vapour–liquid
Kim and Fisher [44] examined the discretization phase behaviour. The first is charged hard dumbbells
dependence of the critical behaviour and extrapolated or tethered dimers, approximating the RPM
to the continuum limit. The unequivocal conclusion because there is strong ion pairing in the (untethered)
from these two studies was that the RPM is in the RPM at its critical temperature. For the contact
three-dimensional Ising universality class, as also tethered dimer of equal-size opposite charges, the
argued on theoretical grounds [45]. Some earlier critical temperature is almost the same as for
studies of the heat capacity in the canonical (NVT) the unconstrained RPM, but the critical density
ensemble [46,47] did not observe a peak near the is significantly higher: Tc ¼ 0:04911  0:00003,
critical point and interpreted this as an indication of c ¼ 0:101  0:003 [56]. Tethered-ion systems of vary-
non-Ising behaviour. However, these observations ing tether length were studied in [57]. Interestingly, as
were later shown to be due to the suppression of the tether length approaches infinity, care must be
density fluctuations in the canonical ensemble [48]. exercised to prevent an ‘entropy catastrophe’ and
Table 1 gives a summary of recent results for the obtain a proper thermodynamic limit. The dependence
critical parameters of the RPM. There is generally of the critical parameters on tether length was found to
good agreement between simulations that used differ- be smooth but non-monotonic. The other related
ent boundary conditions and sampling algorithms. The system is that of soft charged spheres studied in [58],
critical temperature has relatively small uncertainties; for which Tc  0:047, c  0:070.
2042 A.-P. Hynninen and A.Z. Panagiotopoulos
Downloaded By: [Adler, S] At: 15:56 3 February 2011

Figure 1. (color online). Unit cells of (a) CsCl or bcc, (b) CuAu, and (c) tetragonal fcc structures. The dark (red) and light
(yellow) colors denote opposite charges. Reprinted with permission from [62]. Copyright ß (2006) American Physical Society.

The RPM has several transitions involving solid 0.4


phases, in addition to the liquid–vapour transition. Fcc
0.35
Early simulations of fluid–solid and solid–solid disordered
equilibria for the RPM by Smit et al. [59] and Vega 0.3
et al. [60], identified a liquid–bcc transition at low 0.25
temperatures and a liquid–fcc transition at high Tetragonal
T*

temperatures. The bcc crystal is made out of two 0.2 Fluid


species of oppositely charged spheres, and therefore 0.15 CuAu
corresponds to the atomic structure of CsCl, see
Figure 1(a). Bresme et al. [61] amended the phase 0.1
CsCl
diagram by identifying an additional tetragonal fcc 0.05
solid (see Figure 1c) that becomes stable at high
0
colloid densities. Recently, Hynninen et al. [62] found 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
a novel stable crystal phase that has the symmetry of Figure 2. Phase diagram of the RPM in the packing fraction-
the CuAu binary alloy, shown in Figure 1(b). reduced temperature T* plane. Reprinted with permission
Interestingly, this crystal structure was first suggested from [62]. Copyright ß (2006) American Physical Society.
by experimental observations on a binary mixture of
oppositely charged colloidal spheres. Indeed, accord-
ing to calculations by Hynninen et al., the phase to a liquid–bcc phase transition [63]. The phase
behaviour of the RPM is qualitatively similar to that diagram of the RPM is shown in Figure 2 in the
of oppositely charged colloids. Such a correspondence packing fraction , reduced temperature T* repre-
is expected since the pair potential in the system of sentation. The gray areas denote coexistence regions
oppositely charged colloids is given by a Coulomb of the first-order phase transitions where tie lines are
potential multiplied
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi by a screening factor er, where horizontal. The CsCl–CuAu transition is a weak first-
 ¼ 8B salt is the Debye screening length and salt order martensitic phase transition, i.e. the transition
is the salt concentration. As  ! 0, the two systems occurs by a continuous deformation of the crystal
become identical. The main qualitative difference unit cell.
between the two systems is that at   4.5/, the The RPM phase diagram was determined using
gas–liquid critical point is metastable with respect thermodynamic integration, a method that calculates
Molecular Physics 2043

the Helmholtz free energy relative to the reference state counterion limit using thermodynamic scaling Monte
whose free energy is known, enabling the calculation of Carlo. Linse [72,73] proposed a scaling law for the
the full free energy. This approach was first proposed critical temperature of systems with charge ratios up to
by Kirkwood in 1935 [64] and was later extended to 80:1, also at the point counterion limit. A study of
include solids by Frenkel and Ladd [65]. In the fluid charge-asymmetric electrolytes with charge ratios up to
phase, the reference state is the hard-sphere fluid. In 10:1 was performed in [74], using reservoir grand
the case of solid phases the reference state is the canonical simulations. Some discrepancies were
Einstein crystal where each particle is tied to its lattice observed between the results of this study and the
position by a harmonic spring. The phase boundaries scaling law of Linse that remain to be resolved.
are drawn using the Helmholtz free energies and a However, it is clear from these studies that salt-free
common tangent construction, which guarantees ther- colloidal systems phase separate through effective like-
modynamic stability, i.e. equality of temperature, ion attractions at sufficiently high couplings.
pressure, and chemical potential. A similar computa- The existence of like-charge attraction at low
tional approach has been used for some of the work on coupling (e.g. micrometre colloids in aqueous solvent
solid phases of charged colloids described in the with monovalent ions) has been a controversial topic
following section. since the experiments that among other things showed
a broad gas-solid coexistence [75,76], gas–liquid con-
densation [77] (see, however, [78]), large stable voids
3. Generalized primitive models and charged colloids [79,80], and anomalously long-lived dense clusters [81].
These experiments, which have been proven difficult to
In the RPM model, all ions have the same size and
Downloaded By: [Adler, S] At: 15:56 3 February 2011

replicate, provided the motivation for numerous


absolute charge and only interact through coulombic
theoretical [82–84] and simulation studies. According
forces and hard-sphere repulsions. This section sum-
to the well-established DLVO theory, no attractions
marizes simulation studies of ionic systems for which
are possible at low coupling. However, the theoretical
these constraints have been relaxed, resulting in
descriptions propose a situation where so called
generalized primitive models.
‘volume terms’, that include all non-pairwise additive
Phase diagrams of size-asymmetric electrolytes
many-body terms, can give rise to a gas-liquid or gas-
were obtained by Romero-Enrique et al. [66] using a solid phase coexistence. Many of the volume theories
fine-lattice discretization approach and by Yan and have been criticized for their approximations, most
de Pablo [67,68] using a multidimensional parallel important of which is the linearization of the Poisson–
tempering algorithm. The key finding from these Boltzmann equation. In fact, as Lu and Denton [85]
studies was that the critical temperature and critical recently showed, the other major approximation, the
density are at a maximum for the size-symmetric case use of the first-order thermodynamic perturbation
and then drop off as the size asymmetry increases. This theory, is of lesser importance. In a recent theoretical
is in contrast to predictions of commonly used theories treatment [86], the linearization approximation was
for ionic fluids [11,12]. Newer theories have been partly avoided by solving the nonlinear Poisson–
developed [6–10] that predict the correct trends of the Boltzmann equation numerically close to the colloid
critical parameters with size asymmetry. surface, where the linearization is most problematic,
Multivalent electrolytes have been studied much and applying the linear Poisson–Boltzmann for dis-
less than monovalent ones, despite their importance in tances further from the surface. This volume theory
micellar, colloidal and biological systems. Detailed also predicted broad coexistence regions at low
studies of 2:1 and 3:1 electrolytes for a broad range of coupling, however, consequently the parameter space
size ratios were performed by Panagiotopoulos and where phase separation occurred was considerably
Fisher [69] and Yan and de Pablo [70]. The main smaller than the one obtained using the purely linear
finding from these studies is that the critical tempera- Poisson–Boltzmann theory in [82].
ture and density show a maximum for multivalent ions Realistic intermolecular potentials for ionic systems
of increasing size as the valency is increased. include soft repulsions and dispersion forces in
Panagiotopoulos and Fisher [69] suggest that for addition to the coulombic interactions [87].
small multivalent ions with large monovalent counter- Simulation studies of phase transitions for realistic
ions the vapour–liquid phase transition is likely to salt models of salts have been performed by Guissani
disappear because of the formation of an open and Guillot for NaCl [88] and NH4Cl [89].
network. Thermodynamic integration methods were used to
Reščič and Linse [71] have investigated the critical obtain the melting point for the Tosi-Fumi-Tosi model
parameters for a 10:1 electrolyte at the point of NaCl [90] in good agreement to experimental data.
2044 A.-P. Hynninen and A.Z. Panagiotopoulos

Studies of charged Yukawa systems are also available (0.74) packed. Since the gas phase is dilute, the
[91–93]. In addition to the melting point, a calculation particles rarely interact and one can justify a simula-
of the homogeneous nucleation rate for the Tosi-Fumi tion with only few colloids. Similarly, since the solid
model of NaCl [94] was performed at deeply subcooled phase is very dense, the colloids are almost fixed to
temperatures using forward flux [95] and transition their ideal lattice sites and one can imagine approx-
interface [96] sampling methods. The nucleation rate imating them with fixed colloids. Hence, we proposed a
extrapolated to temperatures closer to melting was minimal system approximation where the gas phase is
several orders of magnitude lower than experimental replaced by a single colloid fixed at the center of
measurements, possibly because of model deficiencies the box and the fcc phase is replaced by four colloids
or the limitations of the extrapolation procedure. fixed at their ideal fcc lattice positions. In both cases,
In charged nanoparticle systems, where there are the counterions are mobile and the free energy is
competing interactions to the electrostatic ones calculated in the same manner as in the full system.
operating at shorter distances, several unusual Since periodic boundary conditions are used, the
morphologies are possible. For example, when attrac- colloids interact with the periodic images of themselves
tive short-range interactions exist between like-charge and of the counterions even in the case of a single
ions, fascinating aggregation patterns have been colloid. We tested the validity of the gas phase
observed in two dimensions [97]. Simple systems in approximation by using four mobile colloids and
three dimensions in which attractions and repulsions found that the results became closer to the full
operate among all ions have been studied [98–100]. For system, but the difference was small. In Figure 3(c)
the case of nearest-neighbor exclusions [101] a charge- and (d), the results of the minimal system approxima-
Downloaded By: [Adler, S] At: 15:56 3 February 2011

ordered ‘striped’ phase was found to be stable. tion are denoted by the filled symbols. As can be seen,
Real charged chain or colloidal systems almost the approximate critical temperature is pretty close to
always have varying amount of salt or other ionic the true value and the approximation becomes better
species. Simulations of phase transitions in charged with increasing charge asymmetry. Using the minimal
colloidal (macroion) systems with added salt have been system approximation, we calculated the critical point
performed by Hynninen et al. [102]. The critical for charge-asymmetries 200:1, 1000:1, and 2000:1. In
parameters were determined for macroion to counter- Figure 4, we plot the critical temperature as a function
ion charge-asymmetries of 2:1, 3:1 and 10:1. This study
of the charge-asymmetry Q:1. The solid line is a power-
found that binary electrolyte mixtures are type-I
law fit to the data given by
mixtures, with the two-components mixing
continuously. Tc ¼ 0:27Q0:48 ð5Þ
In recent work [103], we have examined the global
phase diagram of charged colloids with charge The dashed line is the scaling proposed by Linse et al.,
asymmetries up to 2000:1. The full phase diagrams Tc  Q0:57 . While our data is based on (minimal
were calculated for asymmetries 2:1, 10:1, 20:1, and system) free energy calculations, Linse et al. based their
50:1, and are shown in Figure 3. These phase diagrams results on the calculation of the structure factor for
were calculated using the Helmholtz free energy 10  Q  80. The difference in the two results could
methods mentioned near the end of Section 2. An stem from the fact that in the structure factor analysis,
important observation was that the gas–liquid phase one cannot rigorously tell if the chosen state points are
separation becomes metastable with respect to gas- thermodynamically stable or in phase coexistence,
solid phase separation at charge-asymmetry 20:1 and while in our method thermodynamic stability is
beyond. In order to have meaningful critical para- guaranteed. Interestingly, the approximately
meters for cases where the gas-liquid critical point is Tc  Q1=2 scaling behaviour in Equation (5) can be
metastable, we defined a ‘critical temperature’ as the derived analytically using known results for planar
inflection point of the gas–solid phase separation line. charged plates. For two infinite parallel plates with
The advantage of this definition is that the inflection constant (and similar) surface charge and point-like
point coincides with the gas–liquid critical point counterions, in the low-coupling regime, the plates
continuously as charge-asymmetry is lowered below repel each other as is expected based on the Poisson–
20:1. In Figure 3(c) and (d), the inflection points are Boltzmann theory. In the high coupling regime, the
denoted by the squares. From Figure 3 we see that the assumptions of the Poisson–Boltzmann theory break
dominant part of the phase diagram for charge down and plates start attracting each other. In the
asymmetries greater than 20:1 is a broad gas–fcc asymptotic high coupling limit, this phenomenon is
coexistence region, where the gas phase is very captured by the so-called strong coupling theory by
dilute and the fcc solid phase is almost maximally Moreira and Netz [104]. The strength of
Molecular Physics 2045
Downloaded By: [Adler, S] At: 15:56 3 February 2011

Figure 3. (color online). Phase diagrams for charged colloidal suspension with charge ratios: (a) 2:1, (b) 10:1, (c) 20:1, and (d)
50:1. Open symbols represent a full calculation of the free energy and closed symbols an approximation based on a minimal
system. The stars are critical points and the triangles are triple points. Reprinted with permission from [103]. Copyright ß (2007)
American Physical Society.

the electrostatic correlations is described by a coupling The advanced volume-term theory presented in [86]
parameter shows metastable gas–liquid phase coexistence and a
q2  B stable gas–solid coexistence at Q ¼ 2000, T* ¼ 20 and no
¼ ¼ 2Q=ðT  Þ2 , ð6Þ added salt (a ¼ 0), in the region where the

where ¼ 1/2qBs is the Gouy–Chapman length and Poisson–Boltzmann equation is valid. However, our
s ¼ Qq/ 2 is the surface charge density of the colloid. simulation results at Q ¼ 2000 show that no broad
Figure 5 shows the phase diagram of charged planar coexistence is expected for T*  10. Furthermore,
plates as a function of reduced plate separation separate (minimal system) free energy calculations
d~ ¼ d= and coupling strength  as obtained by performed at T* ¼ 20 did not show any signs of
Moreira and Netz [104]. At   c  17 the system instabilities. Direct comparison is somewhat hampered
undergoes a first-order transition, where the stable by the fact that the theory relies on small amount of
state is coexistence between densely packed plates and added salt to be present in the system to produce optimal
plates at infinite separation. Respectively, these two results. However, this first comparison between simula-
states correspond to our densely packed fcc solid phase tions and theory shows that there is room for improve-
and low-density gas phase. Solving T* from equation ment in the accuracy of the volume-term theories.
above and using  ¼ c, we obtain
 1=2
2

Tc,plate ¼ Q1=2 ¼ 0:343Q1=2 ð7Þ 4. Charged chains and ionic gels
c
The fact that the spherical particles and plates have the Polyelectrolytes and polyampholytes are polymers that
same critical scaling law nicely illustrates the common contain charged groups along their backbones or in
origin of like-charge attraction in these two systems. side chains. They can be found naturally in the form of
2046 A.-P. Hynninen and A.Z. Panagiotopoulos

10 properties such as the osmotic pressure are harder to


obtain accurately [110]. Phase transitions and critical
8
points are especially challenging because they occur at
conditions for which charge correlation effects are
important.
6
* There have been relatively few studies of phase
Tc
transitions in charged chains. Orkoulas et al. [111]
4 studied fully charged lattice chains of chain lengths up
to 24 beads, along with monomeric counterions. The
2 critical density was found to be approximately inde-
pendent of chain length, in sharp contrast to homo-
0
polymers for which the critical density approaches zero
10 100 1000 as the chain length increases. Theoretical predictions for
Q critical parameters in salt-free polyelectrolyte solutions
Figure 4. (color online). Critical temperature as a function of [112] are in only qualitative agreement with simulations.
colloid charge-asymmetry. The squares are the critical Cheong and Panagiotopoulos [113] studied polyam-
temperatures. Reprinted with permission from [103]. pholyte lattice chains, both fully charged and end-
Copyright ß (2007) American Physical Society. charged. For fully charged chains, they determined that
the critical parameters are sensitive functions of the
charge sequence. End-charged chains with intervening
Downloaded By: [Adler, S] At: 15:56 3 February 2011

neutral ‘spacer’ beads have decreasing critical tempera-


tures and densities with increasing the spacer length.
These features are qualitatively reproduced by the
charged hard-sphere chain model of Jiang et al. [114],
but significant quantitative discrepancies remain
between simulation and theoretical results.
Polyelectrolyte networks are of interest as possible
drug delivery systems and pH- or concentration-
sensitive actuators. Of particular interest to the present
review are studies of discontinuous volumetric transi-
tions in polyelectrolyte gels due to coulombic interac-
tions, rather than transitions due to polymer-solvent
interactions. Linse and coworkers [115–117] have used
Figure 5. MC phase diagram of charged planar plates as a
function of reduced plate separation d~ and coupling strength Monte Carlo methods to investigate such phenomena
. Reprinted with permission from [104]. Copyright ß (2001) in perfect primitive polyelectrolyte networks of dia-
American Physical Society. mond topology with explicit counterions. Instabilities
indicative of first-order transitions between collapsed
and expanded gels are found as the electrostatic
proteins and nucleic acids, or they can be made coupling or additional solvent-mediated short-range
synthetically. They are involved in industrial processes attractions are increased. The presence of salt [117]
such as wastewater treatment or oil recovery and are reduced the volume of the swollen state but did not
found in many consumer products. An important fully suppress the transition. The salt content of the gel
natural polyelectrolyte is DNA, the main carrier of was found to be greater than the prediction of ideal
genetic information. It is well known that multivalent Donnan equilibrium. A related model has been studied
ions induce attractions between DNA strands using molecular dynamics simulations by de Pablo and
mediated by strong correlation effects [1]. coworkers [118,119], who provide a detailed discussion
Experiments have shown precipitation and subsequent of the contributions to the network osmotic pressure
redissolution of DNA with addition of polyamine salts due to chain elasticity and electrostatic contributions.
[105]. Properties of polyelectrolytes in solution have
been extensively studied by simulation [e.g. 106–108]
and theoretical methods [e.g. 109], with generally good 5. Ionic liquids and aqueous salt solutions
agreement for single-chain conformational properties Ionic liquids are organic salts that have melting points
at low coupling strengths. Collective (multichain) near or below room temperature [120,121]. They have
Molecular Physics 2047

received considerable attention in recent years as NaCl/H2O binary systems were more recently studied
potential ‘green’ solvents for separations [122] and by Sanz and Vega [137] for a range of temperatures.
chemical reactions. Their primary advantage is their For the KF/H2O system, discrepancies between model
low to negligible volatility and their ability to dissolve and experiments were shown to increase rapidly at
both polar and non-polar species. The properties of lower temperatures, but for NaCl good agreement was
ionic liquids are strongly influenced by electrostatic seen. An expanded-ensemble fractional particle
interactions, even though specific short-range ion-ion approach for getting electrolyte solution free energies
and ion-solvent forces also play a role [123]. was recently proposed and applied to the NaCl/H2O
Many simulation studies of ionic liquids aim to system [138].
obtain single-phase thermodynamic and transport The presence of salts modifies the solubility of
properties such as their density or viscosity [124]. other components in water. A recent study for the
Phase equilibrium properties of ionic liquids that have solubility of methane in aqueous NaCl solutions used
been studied by simulation fall under two main the Widom test particle approach to obtain the CH4
categories. The first is the determination of solid- chemical potential at infinite dilution [139]. The
fluid equilibria for the pure components. The melting CH4–H2O interaction parameters were obtained in an
(or freezing) temperature is one of the key parameters earlier study [140] by matching experimental solubility
for many practical applications and its accurate data in binary system. For the aqueous salt system, a
calculation is impossible in direct simulations because salting out effect is correctly predicted, but its
of strong hysteresis effects. Accurate freezing point magnitude is about twice as large as that seen
calculations for ionic liquids were first performed only experimentally, a difference attributed in [139] to the
Downloaded By: [Adler, S] At: 15:56 3 February 2011

recently [125], using a specially developed thermo- neglect of polarization effects of the ions in solution.
dynamic integration approach [126]. The second Interestingly, earlier studies on CO2 solubility in NaCl
category of simulations of phase equilibria involving solutions [141] and mixed solvent electrolyte systems
ionic liquids is the determination of small-molecule [142] using Gibbs ensemble simulations [29] also
solubilities in ionic liquids. This can be done more observed an overprediction of salt effects on the
easily, as it does not involve the free energy of the ionic solubilities of other components.
liquid itself. Direct particle insertions [127], thermo- Deviations of the free energy of mixing from ideal
dynamic integration and Gibbs ensemble Monte Carlo behaviour are often described in liquid solutions in
methods have been used to study the solubility of terms of activity coefficients. For salt solutions, activity
water and volatile gases such as CO2 in ionic liquids coefficients are well described at low concentrations
[128–131]. using the Debye–Hückel limiting law, but deviations
Aqueous solutions of electrolytes are of central become important above 0.1 M for 1:1 electrolytes and
importance in geochemical, electrochemical, many at even lower concentrations for multivalent salts [143].
separation processes and for biological systems. Many models have been proposed to describe activity
Atomistic explicit-solvent simulations of aqueous salt coefficients in concentrated solutions of electrolytes
solutions with explicit solvent have been used to obtain [144], but these generally rely on a number of
structural and dynamic properties [e.g.132–134] and phenomenological parameters that need to be fitted to
potentials-of-mean-force (PMFs) [135] in salt solu- experimental data. Several investigators have used
tions. Once again, obtaining free energies (or activity PMFs derived from atomistic simulations to obtain
coefficients) and solubilities from simulations of activity coefficients for use in implicit-solvent calcula-
explicit-solvent models is cumbersome and computa- tions [145,146], but quantitative agreement for solutions
tionally expensive, but is becoming more common as with salt concentrations greater than 1.0 M is generally
both computational capabilities and simulation meth- not attained.
odologies advance. The calculation of solubilities of Recent work in our group [147] has combined
ionic salts in water requires the calculation of the dielectric saturation and hydration forces to obtain
chemical potential in solution and in the solid phase. effective potentials that describe quantitatively activity
The solubility of KF in water was studied by Ferrario coefficients over a broad range of concentrations. An
et al. using thermodynamic integration for the example for the effective potential obtained for NaCl
individual ions in solution and integration to the at ambient conditions is shown in Figure 6, with the
Einstein crystal for the solid [136]. Using previously corresponding activity coefficients shown in Figure 7.
available non-polarizable force fields, the saturation Interestingly, the activity coefficients for NaCl
concentration at 320 K was obtained from the simula- obtained from free energy calculations of atomistic
tions to be 26 M, in reasonable agreement to models in [137] were in only qualitative agreement to
the experimental value of 17 M. The KF/H2O and experimental data at intermediate concentrations.
2048 A.-P. Hynninen and A.Z. Panagiotopoulos

transition has been confirmed to belong to the Ising


universality class. For highly charge asymmetric
colloids, the phase behaviour scales in a fashion
identical to that of attractions between charged flat
plates.
While significant progress has been made in
simulations of phase equilibria for primitive models,
relatively few studies of realistic potential models for
pure ionic liquids and aqueous salt solutions are
available. Much work remains to be done in obtaining
accurate force fields for ionic liquids and dissolved
salts. It is not clear at the present time if ionic
polarizability needs to be included in these force fields
explicitly to attain quantitative agreement to experi-
ment for broad sets of conditions. An accurate
polarizable force field for water has been proposed
Figure 6. Effective interaction potential between Naþ and recently [148] but its performance for ionic solution
Cl ions and comparison to fixed dielectric constant model calculations has not been explored. Effective potentials
(" ¼ 78). Reprinted with permission from [147]. Copyright ß for aqueous salt solutions have been obtained to model
(2007) American Institute of Physics. solution nonidealities over broad concentration ranges,
Downloaded By: [Adler, S] At: 15:56 3 February 2011

but their transferability with respect to salt valence and


temperature remains an open issue.
An important consideration for calculating inter-
actions in ionic systems relates to the surface boundary
conditions. Common assumptions are constant-charge,
constant potential, or full charge regulation. The first
case corresponds to the surfaces having a fixed charge,
independent of local environment and surface separa-
tion, which is physically unrealistic. The second case
can be imposed for a conducting (electrode) surface.
For insulating surfaces or colloidal particles, a more
realistic assumption is often that there is chemical
equilibrium of ionizable groups on the surface but
most simulation studies to date correspond to the fixed
charge condition. The incorporation of ionizable
groups in Monte Carlo simulations can be done
through a ‘Reactive Canonical Monte Carlo’ approach
[149,150] that has also been used for the study of pH
Figure 7. Mean ionic activity coefficient for NaCl as a effects in proteins [151].
function of molality at 298 K. Simulations with an implicit
model from [147] are shown as circles. Experimental data are
shows as continuous lines; dashed lines are the Debye-Hückel
limiting law (DH) and a fixed dielectric constant model
Acknowledgements
(" ¼ 78). Reprinted with permission from [147]. Copyright ß The authors acknowledge financial support for this work by
(2007) American Institute of Physics. the National Science Foundation, through the Princeton
Center for Complex Materials (NSF-MRSEC Program,
grant DMR- 0213706) and by the Department of Energy,
Office of Basic Energy Sciences (grant DE-FG201ER15121).
6. Summary and future outlook
Phase transitions of strongly coupled ionic systems
References
have been extensively studied over the past decade. For
primitive models that describe the solvent as a uniform [1] E. Allahyarov, G. Gompper, and H. Löwen, Phys. Rev. E
dielectric continuum, there are now quantitative phase 69, 041904 (2004).
diagrams including fluid and solid phases for arbitrary [2] Y. Zhang, J.F. Douglas, B.D. Ermi, et al., J. Chem. Phys.
charge and size asymmetries. The liquid-vapour phase 114, 3299 (2001).
Molecular Physics 2049

[3] L. Guldbrand, B. Jonsson, H. Wennerstrom, et al., [34] Y.C. Kim and M.E. Fisher, J. Phys. Chem. B 108, 6750
J. Chem. Phys. 80, 2221 (1984). (2004).
[4] M. Boström, D.R.M. Williams, and B.W. Ninham, [35] A.M. Ferrenberg and R.H. Swendsen, Phys. Rev. Lett.
Phys. Rev. Lett. 87, 168103 (2001). 61, 2635 (1988).
[5] O. Zohar, I. Leizerson, and U. Sivan, Phys. Rev. Lett [36] J.M. Caillol, D. Levesque, and J.J. Weis, Phys. Rev.
96, 177802 (2006). Lett. 77, 4039 (1996).
[6] Y.V. Kalyuzhnyi, M. Holovko, and V. Vlachy, J. Stat. [37] J.M. Caillol, D. Levesque, and J.J. Weis, J. Chem. Phys.
Phys. 100, 243 (2000). 107, 1565 (1997).
[7] D.M. Zuckerman, M.E. Fisher, and S. Bekiranov, Phys. [38] J.M. Caillol, D. Levesque, and J.J. Weis, J. Chem. Phys.
Rev. E 64, 011206 (2001). 119, 10794 (2002).
[8] M.N. Artyomov, V. Kobelev, and A.B. Kolomeisky, [39] G. Orkoulas and A.Z. Panagiotopoulos, J. Chem. Phys.
J. Chem. Phys. 118, 6394 (2003). 110, 1581 (1999).
[9] O. Patsahan, I. Mryglod, and T. Patsahan, J. Phys. [40] Q.L. Yan and J.J. de Pablo, J. Chem. Phys. 111, 9509
Condens. Matter 18, 10223 (2006). (1999).
[10] A. Ciach, W.T. Gozdz, and G. Stell, Phys. Rev. E 75, [41] A.Z. Panagiotopoulos, J. Chem. Phys. 116, 3007 (2002).
051505 (2007). [42] G. Orkoulas, M.E. Fisher, and A.Z. Panagiotopoulos,
[11] E. González-Tovar, Mol. Phys. 97, 1203 (1999). Phys. Rev. E 63, 051507 (2001).
[12] F.O. Raineri, J.P. Routh, and G. Stell, J. Phys. IV [43] E. Luijten, M.E. Fisher, and A.Z. Panagiotopoulos,
France 10, 99 (2000). Phys. Rev. Lett. 88, 185701 (2002).
[13] M.E. Fisher, J. Stat. Phys. 75, 1 (1994). [44] Y.C. Kim and M.E. Fisher, Phys. Rev. Lett. 92, 185703
[14] J.M.H. Levelt Sengers and J.A. Given, Mol. Phys. 80, (2004).
899 (1993). [45] G. Stell, Phys., Rev. A 45, 7628 (1992).
Downloaded By: [Adler, S] At: 15:56 3 February 2011

[15] G. Stell, J. Stat. Phys. 78, 197 (1995). [46] J. Valleau and G. Torrie, J. Chem. Phys. 108, 5169
[16] R.R. Singh and K. S., Pitzer, J. Chem. Phys. 92, 6775 (1998).
(1990). [47] J. Valleau and G. Torrie, J. Chem. Phys. 117, 3305
[17] S. Wiegand, J.M.H. Levelt Sengers, K.J. Zhang, et al., (2002).
J. Chem. Phys. 106, 2777 (1997). [48] C.D. Daub, P.J. Camp, and G.N. Patey, J. Chem. Phys.
[18] M. Kleemeier, S. Wiegand, W. Schröer, et al., J. Chem. 121, 8956 (2004).
Phys. 110, 3085 (1999). [49] A.Z. Panagiotopoulos and S.K. Kumar, Phys. Rev.
[19] Y. Levin, Brazilian J. Phys. 34, 1158 (2004). Lett. 83, 2981 (1999).
[20] H. Weingärtner and W. Schröer, Adv. Chem. Phys. 116, [50] A.Z. Panagiotopoulos, J. Chem. Phys. 112, 7132 (2000).
1 (2001). [51] G. Stell, in New Approaches to Problems in Liquid State
[21] H. Boroudjerdi, Y.W. Kim, A. Naji, et al., Phys. Rep. Theory, edited by C. Caccamo, J.P. Hansen and G. Stell,
416, 129 (2005). NATO ASI Series (Kluwer Academic Publishers,
[22] J.P. Hansen and H. Löwen, Ann. Rev. Phys. Chem. 51, Dordrecht, 1999).
209 (2000). [52] S. Moghaddam, Y.C. Kim, and M.E. Fisher, J. Phys.
[23] J.L. Barrat and J.F. Joanny, Adv. Chem. Phys. 94, 1 Chem. B 109, 6824 (2005).
(1996). [53] A. Ciach and G. Stell, Phys. Rev. E 70, 016114 (2004).
[24] A.Z. Panagiotopoulos, J. Phys. Condens. Matter 17, [54] A. Ciach, Phys., Rev. E 70, 046103 (2004).
S3205 (2005). [55] M.N. Artyomov and A.B. Kolomeisky, Mol. Phys. 103,
[25] P.N. Vorontsov-Velyaminov, A.M. Elyashevich, 2863 (2005).
L.A. Morgenshtern, et al., High Temp. (USSR) 8, 261 [56] C.D. Daub, G.N. Patey, and P.J. Camp, J. Chem. Phys.
(1970). 119, 7952 (2003).
[26] P.N. Vorontsov-Velyaminov and V.P. Chasovskikh, [57] J.M. Romero-Enrique, L.F. Rull, and
High Temp. (USSR) 13, 1071 (1975). A.Z. Panagiotopoulos, Phys. Rev. E 59, 041204 (2002).
[27] G. Stell, K.C. Wu, and B. Larsen, Phys. Rev. Lett. 37, [58] G. Ganzenmüller and P.J. Camp, J. Chem. Phys. 126,
1369 (1976). 191104 (2007).
[28] A.Z. Panagiotopoulos, Fluid Phase Equil. 76, 97 (1992); [59] B. Smit, K. Esselink, and D. Frenkel, Mol. Phys. 87, 159
erratum in Fluid Phase Equil. 92, 313 (1994). (1996).
[29] A.Z. Panagiotopoulos, Mol. Phys. 61, 813 (1987). [60] C. Vega, F. Bresme, and J.L.F. Abascal, Phys. Rev. E
[30] G. Orkoulas and A.Z. Panagiotopoulos, J. Chem. Phys. 54, 2746 (1996).
101, 1452 (1994). [61] F. Bresme, C. Vega, and J.L.F. Abascal, Phys. Rev.
[31] A.Z. Panagiotopoulos, J. Phys.: Condens. Matter 12, Lett. 85, 3217 (2000).
R25 (2000). [62] A.-P. Hynninen, M.E. Leunissen, A. van Blaarderen,
[32] A.D. Bruce and N.B. Wilding, Phys. Rev. Lett. 68, 193 et al., Phys. Rev. Lett. 96, 018303 (2006).
(1992). [63] A. Fortini, A.-P. Hynninen, and M. Dijkstra, J. Chem.
[33] N.B. Wilding and A.D. Bruce, J. Phys.: Condens. Phys. 125, 094502 (2006).
Matter 4, 3087 (1992). [64] J.G. Kirkwood, J. Chem. Phys. 3, 300 (1935).
2050 A.-P. Hynninen and A.Z. Panagiotopoulos

[65] D. Frenkel and A.J.C. Ladd, J. Chem. Phys. 81, 3188 [96] T.S. van Erp, D. Moroni, and P.G. Bolhuis, J. Chem.
(1984). Phys. 118, 7762 (2003).
[66] J.M. Romero-Enrique, G. Orkoulas, A.Z. Panagiotopoulos, [97] S.M. Loverde, Y.S. Velichko, and M.O. de la Cruz, J.
et al., Phys. Rev. Lett. 85, 4558 (2000). Chem. Phys. 124, 144702 (2006).
[67] Q.L. Yan and J.J. de Pablo, J. Chem. Phys. 114, 1727 [98] V. Kobelev, A.B. Kolomeisky, and
(2001). A.Z. Panagiotopoulos, Phys. Rev. E 68, 066110 (2003).
[68] Q.L. Yan and J.J. de Pablo, Phys. Rev. Lett. 86, 2054 [99] A. Diehl and A.Z. Panagiotopoulos, J. Chem. Phys.
(2001). 118, 4993 (2003).
[69] A.Z. Panagiotopoulos and M.E. Fisher, Phys. Rev. Lett. [100] A. Diehl and A.Z. Panagiotopoulos, Phys. Rev. E 71,
88, 045701 (2002). 046118 (2005).
[70] Q.L. Yan and J.J. de Pablo, Phys. Rev. Lett. 88, 095504 [101] A. Diehl and A.Z. Panagiotopoulos, J. Chem. Phys.
(2002). 124, 194509 (2006).
[71] J. Reščič and P. Linse, J. Chem. Phys. 114, 10131 [102] A.-P. Hynninen, M. Dijkstra, and
(2001). A.Z. Panagiotopoulos, J. Chem. Phys. 123, 084903
[72] P. Linse, J. Chem. Phys. 113, 4359 (2000). (2005).
[73] P. Linse, Trans. Phil. Soc. London A 359, 853 [103] A.-P. Hynninen and A.Z. Panagiotopoulos, Phys. Rev.
(2001). Lett. 98, 198301 (2007).
[74] D.W. Cheong and A.Z. Panagiotopoulos, J. Chem. [104] A.G. Moreira and R.R. Netz, Phys. Rev. Lett. 87,
Phys. 119, 8526 (2003). 078301 (2001).
[75] N. Ise, M. Sugimura, K. Ito, et al., J. Chem. Phys. 73, [105] J. Pelta, D. Durand, J. Doucet, et al., Biophys. J. 71, 48
536 (1983). (1996).
[76] N. Ise and M.V. Smalley, Phys. Rev. B 50, 16722 (1994). [106] M. Deserno, C. Holm, and S. May, Macromolecules
Downloaded By: [Adler, S] At: 15:56 3 February 2011

[77] B.V.R. Tata, M. Rajalakshmi, and A.K. Arora, Phys. 33, 199 (2000).
Rev. Lett. 69, 3778 (1992). [107] S. Liu and M. Muthukumar, J. Chem. Phys. 116, 9975
[78] T. Palberg and M. Würth, Phys. Rev. Lett. 72, 786 (2002).
(1994). [108] A. Jusufi, C.N. Likos, and H. Lowen, J. Chem. Phys.
[79] K. Ito, H. Yoshida, and N. Ise, Science 263, 66 116, 11011 (2002).
(1994). [109] A. Bizjak, et al., J. Chem. Phys. 125, 214907
[80] B.V.R. Tata, E. Yamahara, P.V. Rajamani, et al., Phys. (2006).
Rev. Lett. 78, 2660 (1997). [110] R. Chang and A. Yethiraj, Macromolecules 38, 607
[81] A.E. Larsen and D.G. Grier, Nature (London) 385, 230 (2005).
(1997). [111] G. Orkoulas, S.K. Kumar, and A.Z. Panagiotopoulos,
[82] R. Van Roij and J.P. Hansen, Phys. Rev. Lett. 79, 3082 Phys. Rev. Lett. 90, 048303 (2003).
(1997). [112] J.W. Jiang, L. Blum, O. Bernard, et al., Mol. Phys. 99,
[83] P.B. Warren, J. Chem. Phys. 112, 4683 (2000). 1121 (2001).
[84] Y. Levin, E. Trizac, and L. Bocquet, J. Phys.: Condens. [113] D.W. Cheong and A.Z. Panagiotopoulos, Mol. Phys.
Matter 15, S3523 (2003). 103, 3031 (2005).
[85] B. Lu and A.R. Denton, Phys. Rev. E 75, 061403 (2007). [114] J.W. Jiang, J. Feng, H. Liu, et al., J. Chem. Phys.
[86] B. Zoetekouw and R. van Roij, Phys. Rev. Lett. 97, 124, 144908 (2006).
258302 (2006). [115] S. Schneider and P. Linse, J. Phys. Chem. B 107, 8030
[87] F.G. Fumi and M.P. Tosi, Phys. Chem. Solids 25, 31 (2003).
(1964). [116] S. Schneider and P. Linse, Macromolecules 37, 3850
[88] Y. Guissani and B. Guillot, J. Chem. Phys. 101, 490 (2004).
(1994). [117] S. Edgecombe, S. Schneider, and P. Linse,
[89] Y. Guissani and B. Guillot, J. Chem. Phys. 116, 2058 Macromolecules 37, 10089 (2004).
(2002). [118] Q.L. Yan and J.J. de Pablo, Phys. Rev. Lett. 91,
[90] J. Anwar, D. Frenkel, and M.G. Noro, J. Chem. Phys. 018301 (2003).
118, 728 (2003). [119] D.W. Yin, Q.L. Yan, and J.J. de Pablo, J. Chem. Phys.
[91] T. Kristóf, D. Boda, J. Liszi, et al., Mol. Phys. 101, 1611 123, 174909 (2005).
(2003). [120] J.F. Brennecke and E.J. Maginn, Amer. Inst. Chem.
[92] J.B. Caballero, A.M. Puertas, A. Fernandez-Barbero, Eng. J. 47, 2384 (2001).
et al., J. Chem. Phys. 121, 2428 (2004). [121] R.D. Rogers and K.R. Seddon, Science 302, 792
[93] T. Kristóf, D. Boda, and D. Henderson, J. Chem. Phys. (2003).
120, 2846 (2004). [122] X. Han and D.W. Armstrong, Acc. Chem. Res. 40,
[94] C. Valeriani, E. Sanz, and D. Frenkel, J. Chem. Phys. 1079 (2007).
122, 194501 (2005). [123] R.M. Lynden-Bell, M.G. Del Pópolo, T.G.A. Youngs,
[95] R.J. Allen, P.B. Warren, and P.R. ten Wolde, Phys. et al., Acct. Chem. Res. 40, 1138 (2007).
Rev. Lett. 94, 018104 (2005). [124] E.J. Maginn, Acct. Chem. Res. 40, 1200 (2007).
Molecular Physics 2051

[125] S. Jayaraman and E.J. Maginn, J. Chem. Phys. 127, [139] H. Docherty, et al., J. Phys. Chem. B 111, 8993
214504 (2007). (2007).
[126] D.M. Eike, J.F. Brennecke, and E.J. Maginn, J. Chem. [140] H. Docherty, et al., J. Chem. Phys. 125, 074510
Phys. 122, 014115 (2005). (2006).
[127] B. Widom, J. Chem. Phys. 39, 2808 (1963). [141] J. Vorholz, et al., Fluid Phase Equilibria 226, 237
[128] R.M. Lynden-Bell, N.A. Atamas, A. Vasilyuk, et al., (2004).
Molec. Phys. 100, 3225 (2002). [142] H.J. Strauch and P.T. Cummings, Fluid Phase
[129] J.K. Shah and E.J. Maginn, Fluid Phase Equilibria Equilibria 83, 213 (1993).
222, 195 (2004). [143] P.W. Atkins, Physical Chemistry, 4th ed. (W.H.
[130] J.K. Shah and E.J. Maginn, J. Phys. Chem. B 109, Freeman, New York, 1990), x10.2.
10395 (2005). [144] J.W. Tester and M. Modell, Thermodynamics and its
[131] I. Urukova, J. Vorholz, and G. Maurer, J. Phys. Chem. Applications, 3rd ed. (Prentice-Hall, Upper Saddle
B 109, 12154 (2005). River, NJ, 1996).
[132] D.E. Smith and L.X. Dang, J. Chem. Phys. 100, 3757 [145] B. Hess, C. Holm, and N. van der Vegt, Phys. Rev.
(1994). Lett. 96, 147801 (2006).
[133] H. Du, J.C. Rasaiah, and J.D. Miller, J. Phys. Chem. B [146] S. Gavryushov and P. Linse, J. Phys. Chem. B 110,
111, 209 (2007). 10878 (2006).
[134] J. Alejandre and J.P. Hansen, Phys. Rev. E 76, 061505 [147] P.J. Lenart, A. Jusufi, and A.Z. Panagiotopoulos,
(2007). J. Chem. Phys. 126, 044509 (2007).
[135] A.P. Lyubartsev and S. Marcelja, Phys. Rev. E 65, [148] P. Paricaud, et al., J. Chem. Phys. 122, 244511
041202 (2002). (2005).
[136] M. Ferrario, G. Ciccotti, E. Spohr, et al., J. Chem. [149] J. K.Johnson, A.Z. Panagiotopoulos, and
Downloaded By: [Adler, S] At: 15:56 3 February 2011

Phys. 117, 4947 (2002). K.E. Gubbins, Mol. Phys. 81, 717 (1994).
[137] E. Sanz and C. Vega, J. Chem. Phys. 126, 014507 [150] W.R. Smith and B. Triska, J. Chem. Phys. 100, 3019
(2007). (1994).
[138] M. Lisal, W.R. Smith, and J. Kolafa, J. Phys. Chem. B [151] J. Mongan, D.A. Case, and J.A. McCammon,
109, 12956 (2005). J. Comput. Chem. 25, 2038 (2004).

You might also like