You are on page 1of 10

JOURNAL OF CHEMICAL PHYSICS VOLUME 117, NUMBER 13 1 OCTOBER 2002

Grand canonical Monte Carlo simulation of hydration forces between


nonorienting and orienting structureless walls
Tomohiro Hayashi, Alexander J. Pertsin, and Michael Grunze
Angewandte Physikalische Chemie, Universität Heidelberg, INF 253, D-69120 Heidelberg, Germany
共Received 9 May 2002; accepted 10 July 2002兲
The hydration forces between structureless model walls were calculated using the grand canonical
Monte Carlo technique. Several wall–water interaction potentials were tried, including both
orientation independent and strongly directional potentials which reflected the preference of water
for tetrahedral hydrogen bonding coordination. Primary attention was given to large wall-to-wall
separations 共4 nm and more兲, where the oscillations of the hydration force due to layering effects
decayed. The hydration force was found to be highly sensitive to the presence of
orientation-dependent terms in the wall–water interaction potential. Strongly directional potentials
led to hydrophobic attraction of the walls even when the wall–water interaction was substantially
stronger than the water–water interaction. The reason had to do with the orientational ordering
induced by the walls in the adjacent water layers. © 2002 American Institute of Physics.
关DOI: 10.1063/1.1504436兴

I. INTRODUCTION surfaces, the number of simulations concerned with hydra-


tion forces is very limited. The reason is the necessity of
The origin of forces occurring between various surfaces simulating an open confined water system which is allowed
in water and aqueous solutions has been discussed for a long to exchange molecules and is in chemical equilibrium with a
time, mainly in the context of the interactions in colloidal bulk water reservoir 共Fig. 1兲. The standard molecular-
and biological systems.1– 8 A first explanation of these forces dynamics 共MD兲 technique is not well suited for such simu-
was given by the Derjaguin, Landau, Verwey, and Overbeek lations because it requires an explicit simulation of both the
共DLVO兲 theory3 in terms of direct van der Waals attraction
confined water region and the bulk water reservoir. This im-
between the surfaces and screened electrostatic mean-field
poses severe restrictions on the size of the confined system,
repulsion between the ions adsorbed on or concentrated near
since it should be much smaller than the bulk water reservoir.
the surfaces. In more recent studies,9 the DLVO theory was
In this respect, the Monte Carlo 共MC兲 techniques based on
substantially improved by including the correlation contribu-
the implementation of the grand canonical 共GC兲 and isoten-
tion to the electrostatic interaction energy. As the experimen-
sion 共IT兲 ensembles have a great advantage because the bulk
tal techniques for measuring water-mediated forces have be-
water reservoir is present in such simulations implicitly.10–12
come available, forces of different nature have been found.
These non-DLVO forces, which are usually referred to as Nevertheless, these techniques are still of limited use be-
hydration forces, have nothing to do with the presence of cause the maintenance of a constant chemical potential
ions and they would occur even in ideally deionized and within the GC and IT ensembles is computationally expen-
nondissociable water. The source of hydration forces is the sive.
surface-induced changes in the structure and density of adja- Nearly all of the few simulations of hydration forces
cent water. Hydrophobic surfaces usually attract each other available in the literature are concerned with hydrophobic
in water, whereas hydrophilic ones show water-mediated attraction.13–16 Thus Wallqvist and Berne13 used the MD
repulsion.2 For this reason, the attractive and repulsive water- technique to simulate the hydration force between two large
mediated forces are frequently referred to as ‘‘hydrophobic hydrophobic ellipsoids interacting with water through a re-
attraction’’ and ‘‘hydrophilic repulsion,’’ respectively. A de- pulsive inverse power potential. The simulations were re-
tailed discussion of water-mediated forces and their impor- stricted to very short separations: The largest width of the slit
tance in colloid chemistry, biology, and other areas can be between the ellipsoids was about 10 Å. As the ellipsoids
found in numerous review articles.4 –7 were moved together, an oscillating hydration force was ob-
Inasmuch as hydration forces originate from microscopic served, until the constrained water between the ellipsoids
structural changes in the interfacial region, the specific mo- underwent capillary evaporation 共cavitation兲 leading to at-
lecular mechanisms behind these forces are difficult to elu- traction between the ellipsoids due to the pressure imbalance.
cidate experimentally. This imparts importance to the meth- The interpretation of hydrophobic attraction in terms of cap-
ods of computer simulation, which allow evaluation of illary evaporation 共or ‘‘drying’’兲 can also be found in recent
hydration forces based on the principles of statistical me- publications by Chandler and co-workers.17–19 By contrast,
chanics and an assumed form of the water–water and water– the ITMC simulation by Forsman et al.14 of water confined
surface interaction potentials. Despite a large body of litera- between two hard walls at separations ranging from 10 to 23
ture on computer simulation of water in contact with solid Å revealed a strong hydrophobic attraction due to a density

0021-9606/2002/117(13)/6271/10/$19.00 6271 © 2002 American Institute of Physics

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
6272 J. Chem. Phys., Vol. 117, No. 13, 1 October 2002 Hayashi, Pertsin, and Grunze

be tried to follow the effect of the surface–water interaction


strength on the hydration force. In addition to changes in the
hydration force, the behavior of various distribution func-
tions and order parameters will be monitored to see how the
changes in the water–surface interaction potential affect the
structure of the adjoining water layers. Of particular interest
will be the behavior of the average density of confined water.
Since the position of a water molecule close to a hydrophilic
surface is favorable, it can intuitively be expected that the
average water density between two such surfaces will be
always enhanced. Reciprocally, the average water density be-
tween two hydrophobic surfaces can be expected to be al-
ways depressed. Whether these expectations are justified or
not will be seen from the simulations described below.

II. TERMINOLOGY
FIG. 1. System of two parallel plates of area A, which are immersed in
water and repel each other with force Ap h ; p h is the hydration pressure and As far as the interaction of solid surfaces with water will
f is the external force applied in surface-force experiments to keep the sur- be concerned, we will inevitably have to employ the terms
faces at a given separation H apart.
‘‘hydrophilic’’ and ‘‘hydrophobic.’’ Despite the widespread
use of these terms in surface science, colloid chemistry, bi-
ology, and other areas, their meaning involves ambiguity be-
depression between the walls, with no cavitation observed. cause of the lack of a well-defined and commonly accepted
Similar results were obtained in an early study by Luzar criterion that distinguishes surfaces into hydrophilic and hy-
et al.20 for a simple one-site model of water confined be- drophobic. For a detailed analysis of the hydrophilic/
tween hard walls. hydrophobic terminology, we refer the reader to a recent re-
To the best of our knowledge, the only published com- view by Vogler23 and here we only make few comments
puter simulation of hydration forces on hydrophilic surfaces necessary to avoid confusion in the subsequent discussion.
is the one reported by Forsman et al.21 The water–surface The customary understanding of the terms hydrophilic
interaction potential comprised a short-range exponential at- and hydrophobic is associated with the strength of the inter-
tractive term and an inverse ninth power repulsion term. An action of a surface, colloid particle or solute with water. Thus
orientation dependent potential was also tried, which in- ‘‘hydrophilic 共hydrophobic兲 solute’’ is usually understood as
cluded an additional term proportional to the cosine of the a solute which attracts water molecules more 共less兲 strongly
angle between the molecular dipole moment and the surface than water molecules attract one another. A similar criterion
normal. The authors21 concluded that the hydration force was is frequently applied to surfaces: A surface is referred to as
mainly determined by the range of the water–surface poten- being hydrophilic 共hydrophobic兲 if it is capable 共incapable兲
tial: A strong repulsion was only observed when the potential of forming strong hydrogen bonds with water. It is clear,
decay length was greater than about half the molecular di- however, that the hydrogen bond strength is not the only
ameter. The inclusion of the orientation dependent term in factor responsible for the affinity of the surface for water.
the potential made the hydration interaction more repulsive, Equally important are the areal density, lateral arrangement,
though this effect was asserted to be of minor importance. orientation, and flexibility of the surface groups involved in
The general objective of this work is to gain a better the hydrogen bonding with water. Similar arguments apply to
understanding of the factors responsible for the sign and hydrophobic surfaces, whose effect upon the adjacent water
magnitude of hydration forces. As in all reported simulation may substantially differ from that of small hydrophobic
studies of hydration forces,13–16,20,21 we restrict ourselves to solutes.17
structureless model walls. In contrast to the previous work, A thermodynamically sound criterion for the hydrophi-
however, the focus of our simulations will be on large wall- licity of a surface is the condition that the water–surface
to-wall separations 共4 nm and more兲, where the oscillations interfacial tension, ␥ ws , is negative, i.e., that an increase in
of the hydration force have decayed and the sign of the hy- the area of the water–surface interface lowers the free energy
dration force reflects the thermodynamic affinity of the walls of the system. The criterion for hydrophobicity is opposite:
for water 共see the next section兲. Both hydrophobic and hy- ␥ ws⬎0. To relate ␥ ws to the hydration pressure p h consider
drophilic walls will be considered. Similar to the Forsman’s two parallel plates of an area A, which are immersed in a
et al. study,21 the walls will be described using both nonori- water reservoir of volume V and held in equilibrium at a
enting and orienting potentials. In the latter case, however, separation H apart by an external force f.24 The hydration
more realistic potentials, which reflect the preference of wa- pressure experienced by the plates can be expressed in terms
ter for tetrahedral hydrogen bonding coordination, will be of the tension of the water film between the plates ␥. If the
used, including potentials which model proton-acceptor sur- system is treated using the grand canonical ( ␮ VT⫽const)
faces and also surfaces bearing both proton acceptors and ensemble, then ␥ ⫽⌬⍀/A, where ⍀⫽U⫺TS⫺ ␮ N is the
proton donors in equal amounts.22 For each particular surface grand potential and ⌬ denotes the difference between the
type, several discrete values of the potential well depth will systems with and without plates. The hydration pressure is

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
J. Chem. Phys., Vol. 117, No. 13, 1 October 2002 Simulation of hydration forces between walls 6273

then given by differentiation, p h ⫽⫺( ⳵ ␥ / ⳵ H) ␮ T . 24 When work, the hydrophilicity/hydrophobicity dividing lines based
the surfaces are in contact (H⫽0), there is no water tension, on p h and ␪ are coincident.
␥ (0)⫽0, while at H→⬁, ␥ (H)→2 ␥ ws . If ␥ ws⬍0 and ␥ is
a monotonic function of separation, then p h ⬎0, i.e., the
plates repel each other. That is the hydrophilicity criterion III. MODELS
␥ ws⬍0 is equivalent to the condition that p h is repulsive. In all our simulations, the interactions of water mol-
Similarly, the hydrophobicity criterion ␥ ws⬎0 reduces to the ecules with themselves are described with the four-site
condition that p h is attractive. TIP4P model,26 which is perhaps the most reliable of rigid
At short separations, the monotonicity of ␥ (H) can be nonpolarizable water models. 共The effect of the assumed
violated due to water layering effects, which may lead to nonpolarizability on the behavior of water near solid surfaces
oscillations in p h . 3 In this case, however, the sign of p h can was studied by Wallqvist27 and found to be of minor signifi-
still be used as a hydrophilicity/hydrophobicity criterion pro- cance.兲 The summation of water–water interactions is per-
vided that H is large enough for the oscillations in p h to formed using a spherical cutoff 共SC兲 scheme, with a cutoff
decay.25 radius of 7.5 Å. Test simulations with the cutoff radius in-
Based on an analysis of the available surface force lit- creased to 8.5 Å gave very similar results: The deviations for
erature, Vogler2 suggested to include the reference to the sign the hydration pressure and average density were within the
of p h in the definition of hydrophilicity and hydrophobicity. statistical uncertainty of the calculations 共0.05 kbar and
Considering, however, that the true hydration contribution to 0.004 g cm⫺3, respectively兲. The preference of the SC sum-
the surface force may be masked by forces due to the pres- mation scheme over the minimal image and cylindrical cut-
ence of ions and impurities, p h can hardly serve as a practi- off ones in simulations of confined water was demonstrated
cal criterion for hydrophilicity/hydrophobicity. By contrast, it by Shelley and Patey.28 The SC scheme was shown to pro-
can well be used as such a criterion in computer simulations, vide nearly the same results as the Ewald method, while
which deal with ideally pure and ion-free water. being computationally much cheaper.
The hydration pressure can also be related to the water The water–wall interaction potentials are all constructed
contact angle ␪, based on the Young equation, ␥ w ␯ cos(␪) based on a 共9-3兲 inverse power function,

冋冉 冊 冉 冊 册
⫽␥s␯⫺␥ws , where the subscript ‘‘␯’’ refers to water vapor.
␧ zm 9
zm 3
When employing ␪ or the so-called ‘‘adhesion tension’’ u共 z 兲⫽ ⫺3 , 共1兲
␥ w ␯ cos(␪) as a hydrophilicity/hydrophobicity criterion, the 2 z z
dividing line between hydrophilic and hydrophobic is usually where z is the separation of the water oxygen atom from the
taken to be ␪ ⫽90° or ␥ w ␯ cos(␪)⫽0 共see Vogler’s review23 wall, ␧ is the potential well depth, and z m is the equilibrium
for discussion of alternative choices兲, which corresponds to separation, i.e., ␸ ⬘ (z m )⫽0. In all potentials, z m is fixed at
the null change in free energy upon immersing the surface in 2.97 Å,29 a typical O¯O separation in the O–H¯O hydro-
water. The offset of this dividing line from that based on ␥ ws gen bond. It is to be noted that the individual terms of the
or p h is equal to ␥ s ␯ . Since ␥ s ␯ is always positive, the 共9-3兲 potential, as derived from the 共12-6兲 Lennard-Jones
hydrophilicity/hydrophobicity criteria based on ␪ and potential by integration,3 are usually associated with the ex-
␥ w ␯ cos(␪) as if overestimate the surface’s hydrophilicity change repulsion and dispersion attraction, respectively. We
compared to the ␥ ws- and p h -based criteria: In the interval will, however, treat the 共9-3兲 potential in a purely formal
0⬍ ␥ w ␯ cos(␪)⬍␥s␯ , ␪ and ␥ w ␯ cos(␪) show that the surface way, without assigning the original physical meaning to its
is already hydrophilic, while it is still hydrophobic from the individual terms. In the interaction range important to our
viewpoint of ␥ ws and p h ( ␥ ws⬎0, p h ⬍0). In this interval, simulations, this potential can well be fitted by the sum of
the terms ‘‘hydrophilic repulsion’’ and ‘‘hydrophobic attrac- two exponential functions, with the largest deviation as small
tion’’ make no sense if the hydrophilicity/hydrophobicity of as 0.02␧. The decay length of the fitted exponential that de-
the surface is understood in terms of the water contact angle. scribes attraction proves to be 1.69 Å, which is larger than
The occurrence of an offset between the different half the molecular diameter of water. That is, based on the
hydrophilicity/hydrophobicity criteria can well be seen in the simulation results of Forsman et al.,21 we can expect that the
plot discussed by Vogler,2 where the characteristic decay 共9-3兲 potential will allow simulation of hydration repulsion,
length of surface force for partially silanized silicas was de- provided that the potential well depth ␧ is large enough to
picted as a function of ␪. In this plot, the changeover from model hydrogen bonding.
attractive to repulsive forces occurred at ␪ ⬵62° In describing nonorienting walls, isotropic water–wall
关 ␥ w ␯ cos(␪)⫽34 dyn cm⫺1 兴 . That is, in the range between potentials are used. For these walls, hereafter referred to as
62° and 90° the surface was hydrophobic with regard to ␪, walls of the I type, ␧ is treated as a constant. A total of four
while being hydrophobic with regard to p h . Similar ex- distinct values are tried for ␧. One value, ␧⫽␧ 0
amples were found by Besseling8 in his theoretical treatment ⫽0.46 kcal mole⫺1 , represents a typical interaction energy
of hydration forces. of a water molecule with hydrophobic paraffinlike surfaces.29
In the simulations described in the following sections, For walls capable of forming hydrogen bonds with water, ␧
the direct interaction between the constraining walls is ne- is assigned the meaning of the hydrogen bond energy be-
glected, so that no work is required to spread the walls in tween a water molecule and the wall, ␧⫽␧ H . Three discrete
vacuum from contact to infinity. Hence the term ␥ s ␯ in the values are tried for ␧ H : 6.24, 10, and 15 kcal mole⫺1. The
Young equation vanishes. That is, for the walls studied in our first value is the binding energy of a linear TIP4P water

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
6274 J. Chem. Phys., Vol. 117, No. 13, 1 October 2002 Hayashi, Pertsin, and Grunze

tion of the O–H bond or O–lone pair vector is partly down,


i.e., ␪ ⬎90°. In the transition range, 90°⬍ ␪ ⭐180°⫺⌰,
h( ␪ ) is represented by a cosine function and it rapidly in-
creases from 0 to 1 as the hydrogen bond geometry becomes
closer to linear. At ␪ ⬎180°⫺⌰, h( ␪ )⫽1, so that the hydro-
gen bond energy is independent of ␪ and takes its maximum
value, ␧ H . The width of the latter interval, ⌰, is an important
parameter which specifies the directionality of the hydrogen
bond and also the relationship between the binding energies
of single, double, and triple hydrogen bonds formed by the
water molecule with the wall. The presence of an orientation
independent range in ␧共␪兲 at 180°⫺⌰⬍ ␪ ⬍180° formally
reflects the flexibility of the surface groups, which allows
FIG. 2. Orientation dependence of the hydrogen bond potential, as specified them to adjust their orientations so as to meet water mol-
by h( ␪ ) in Eq. 共2兲. ecules at the most favorable angles for the formation of hy-
drogen bonds. The more flexible the surface groups, the
greater ⌰. For this reason, ⌰ will hereafter be referred to as
dimer,26 whereas the two higher values are selected some- the flexibility parameter. As ⌰ is decreased, the orientation
what arbitrarily just to follow the effect of the water–wall independent range in ␧共␪兲 becomes narrower and the orient-
bond strength on the hydration force. From the standpoint of ing effect of the wall is enhanced.
the water–wall interaction energy, all the three nonorienting For walls of the A type, the energies of the strongest
walls with ␧⫽␧ H can be regarded as hydrophilic. double and single hydrogen bonds can be easily shown to be
For orienting walls, ␧ is considered to be dependent on in the ratio 2:1 when ⌰⬎⬔HOH/2⬵52°. At ⌰⫽30°, this
the orientation of the water molecule. In constructing an ap- ratio reduces to 1.4:1, which is close to the ab initio elec-
propriate model for ␧, we assume that the preferred hydrogen tronic structure calculation results for the double and single
bonding coordination of the water molecule is tetrahedral. hydrogen bonds formed by water with the oxygen atoms of
With this choice, ␧ is represented as the sum of orientation- an oligo共ethylene oxide兲 chain.30 For walls of the AD type,
dependent contributions from the individual water protons the ratio 3:2:1 for the triple, double, and single hydrogen
and lone electron pairs: bond energies is attained when ⌰ is chosen to be greater than
⬃70°.
␧⫽␧ 0 ⫹␧ H 兺i h i共 ␪ i 兲 . 共2兲

In this equation h i ( ␪ i ) varies from 0 to 1 and describes the IV. SIMULATION METHOD
orientation dependence of the energy of the hydrogen bond
formed by the ith water proton or lone electron pair and the As already mentioned in the Introduction, there are two
surface; ␪ i is the angle between the O–H bond or O–lone basic MC techniques suitable for simulating open confined
pair vector and the outward-pointing vector normal to the fluid systems, as studied in surface-force experiments 共Fig.
wall surface (0⭐ ␪ i ⭐180°, see Fig. 2兲. It is assumed, for 1兲. In the GCMC technique, the chemical equilibrium of the
simplicity, that the angle between the two O–lone pair vec- confined system with the 共fictitious兲 bulk reservoir is main-
tors in the water molecule is the same as between the O–H tained by allowing density fluctuations through particle in-
bonds 共104.52° for the TIP4P model兲.26 The numerical val- sertions and deletions. The main problem involved in the use
ues of the potential parameters ␧ 0 and ␧ H in Eq. 共2兲 are taken of this technique is a too low probability of performing suc-
to be the same as discussed above for the non-orienting cessful insertions and deletions at densities typical of fluids.
walls. For the orienting walls, ␧ H takes on the meaning of In the past decade, however, this problem has been more or
the energy of a single hydrogen bond formed by a water less successfully solved using Swendsen–Wang filters,31 cav-
molecule with the wall. The parameter ␧ 0 is introduced in the ity bias,32 and related methods11 based on simple and com-
expression for ␧ to avoid penetration of the water molecule putationally inexpensive predictors for the success of inser-
into the wall when no hydrogen bonds are formed, i.e., when tion and deletion moves. The alternative to the GCMC
all h i ( ␪ i )⫽0. In modeling proton acceptor walls 共hereafter, technique is provided by ITMC simulations. In these simu-
the walls of the A type兲, the summation in Eq. 共2兲 is per- lations, the density fluctuations are implemented by allowing
formed over the two water protons, while for the walls bear- area fluctuations parallel to the confining walls at a fixed
ing both proton acceptors and proton donors 共hereafter, the lateral pressure P L . As the separation between the walls is
AD-type walls兲, the lone electron pairs are included as well. changed, the chemical potential is maintained constant by an
In this way, the ability of the wall to carry either proton appropriate adjustment of P L using the free-energy differ-
acceptors only or both proton acceptors and proton donors is ence method.12 The ITMC technique, however, cannot be
modeled. applied to simulation of fluids near structured solid sub-
The function h( ␪ ) in Eq. 共2兲 is in essence a somewhat strates because the area fluctuations in the fluid are inconsis-
smoothed on-or-off function. As seen from Fig. 2, where tent with the condition that the lateral dimensions of the
h( ␪ ) is depicted, the hydrogen bond forms only if the direc- simulation cell must be commensurate with the substrate lat-

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
J. Chem. Phys., Vol. 117, No. 13, 1 October 2002 Simulation of hydration forces between walls 6275

tice. Considering our parallel simulations of the behavior of p h ⫽ 具 f 典 /A⫺ p b , 共6兲


water near structured surfaces,33 we prefer the GCMC tech-
nique over the ITMC one. where p b is the bulk water pressure and 具f典 is the mean force
The computational procedure used in our GCMC simu- exerted by water molecules upon the lower surface,
lations is similar to that described in our previous work.33 To
improve the efficiency of insertions, the excluded volume
mapping method34 and a Swendsen–Wang filter31 based on
具 f 典⫽ 冓兺 冔 i
⳵u
⳵ z ic
. 共7兲

evaluation of the van der Waals energy of the system are In this equation, the angular brackets denote GC ensemble
employed. The usual length of the MC runs is 3⫻106 to 1 averaging, u is the wall–water potential, as given by Eq. 共1兲;
⫻107 passes, each comprising N moves, where N is the cur- the differentiation is performed with respect to the z coordi-
rent number of water molecules in the system. For a typical nate of the center of mass of the ith water molecule. With our
system size used, this corresponds to a total of 1.5⫻109 to choice of the z axis along the outward-pointing vector nor-
5⫻109 attempted configurations. mal to the lower confining surface, positive values of p h
The simulation box represents a rectangular prism with would correspond to attraction between the confining sur-
dimensions L x , L y , and L z . The lower confining wall is faces. To make our results consistent with the conventional
placed at z⫽0 parallel to the x-y plane. In the x and y di- choice of the sign of p h , 21 we will hereafter refer to the
mensions, the system is replicated periodically, with periods values of p h on the upper confining surface, so that positive
L x ⫽30 Å and L y ⫽26 Å. Along the z axis, the ‘‘gliding p h will correspond to repulsion, while negative p h to attrac-
plane’’ boundary 共GPB兲 conditions,33 which combine a mir- tion.
ror plane at z⫽L z with a half period translation along x, are The orientational structure of water at different separa-
used. That is, the space above the top face of the simulation tions from the wall is monitored by calculating the distribu-
cell is filled with a mirror image of the system shifted along tion of the angles, ␪ ␮ and ␪ OH , formed, respectively, by the
x by L x /2. 共The shift is necessary to avoid straightforward molecular dipole moments and O–H bonds with the z axis.
correlations between the displacements of a particle and its Also calculated are the orientational order parameter,
image across the mirror plane.兲 Because of the mirror plane S 10 共 z 兲 ⫽ 具 cos ␪ ␮ 共 z 兲 典 , 共8兲
at z⫽L z , the upper confining wall occurs at z⫽H⫽2L z .
The transformation of coordinates by the GPB conditions is and the integrated density-weighted order parameter,
the following:
x ⬘ ⫽x⫹kL x /2
⌶⫽ 冉冕 0
H/2
dz ␳ 共 z 兲 冊 冕
⫺1

0
H/2
dzS 10 共 z 兲 ␳ 共 z 兲 . 共9兲

共 k⫽1 for x⬍L x /2 and k⫽⫺1 otherwise兲 , 共3兲 While S 10 (z) is the average of cos ␪␮ over the molecules ly-
ing at distance z from the wall, ⌶ represents the average of
y ⬘ ⫽y, 共4兲 cos ␪␮ over all molecules between the wall and the midplane
z ⬘ ⫽2L z ⫺z. 共5兲 at z⫽H/2.
The average water density in the confined region is cal-
As a molecule leaves the simulation cell through the top culated in a straightforward way as ¯␳ ⫽ 具 N 典 /V c , where V c
face, its image, whose coordinates are given by Eqs. 共3兲–共5兲, ⫽L x L y L z is the volume of the simulation cell. In the calcu-
enter the cell through the other half of the top face. lations of the density profiles, orientational distributions, and
Although the GPB conditions save nearly half the CPU other quantities dependent on the distance from the wall, the
time, they may, in principle, affect the simulation results be- simulation box is divided into 100 slices of thickness ⌬z
cause of the artificial correlations introduced by Eqs. 共3兲–共5兲 ⫽L z /100 lying parallel to the wall. During the GCMC run,
in the configuration of the system near the midplane z the z-dependent quantities are averaged within each indi-
⫽L z . To assess the importance of these correlations, we vidual slice and then referred to the z coordinate of its center.
made a comparative simulation of two bulk water systems.
One system was simulated using the usual periodic boundary
conditions in all three dimensions. In the other system, the VI. RESULTS AND DISCUSSION
periodic boundary conditions were only applied along the x A. Bulk water
and y axes, while along the z axis the system was replicated
by applying the GPB conditions at z⫽0 and z⫽L z . The The knowledge of the chemical potential of bulk water is
three lengths of the simulation box were set equal to 30 Å. a prerequisite of simulating confined water in the chemical
The calculated thermodynamic and structural quantities equilibrium with the bulk water reservoir 共Fig. 1兲. For rigid
proved to be practically independent on whether the periodic water models, such as TIP4P, only the excess 共nonideal兲 part
or GPB conditions were used along the z axis. of the chemical potential, ␮ ⬘ , is required. An early MD
simulation by Hermans et al.35 using the thermodynamic in-
tegration method resulted—for the TIP4P model at room
V. QUANTITIES CALCULATED
temperature—in a value of ⫺5.3 kcal mole⫺1, in satisfactory
The set of thermodynamic and structural quantities cal- agreement with an experimental estimate of ⫺5.7
culated in our simulations is basically the same as described kcal mole⫺1 based on the known ratio between the molar
in our previous work.33 In addition, we calculate the hydra- volumes of water vapor and liquid. Note that Hermans
tion pressure et al.’s simulations35 were made with a very small periodic

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
6276 J. Chem. Phys., Vol. 117, No. 13, 1 October 2002 Hayashi, Pertsin, and Grunze

system (N⫽80) and a fairly short cutoff distance 共6 Å兲.


More recently, Shelley and Patey28 reported a value of ⫺6.0
kcal mole⫺1 as best reproducing the room-temperature bulk
water density when the TIP4P model and Ewald summation
method were employed.
To find ␮ ⬘ suited to the particular summation scheme,
cutoff distance, and typical system size used in our simula-
tions of confined water, we performed a series of bulk water
simulations as a function of ␮ ⬘ . In these simulations, the
periodic boundary conditions were used in all three dimen-
sions. The experimental bulk water density at room tempera-
ture, ␳ b ⫽0.997 g cm⫺3 , was reproduced at ␮ ⬘ ⫽
⫺6.10 kcal mole⫺1 , close to the Shelley and Patey result.28
However, the respective bulk pressure, as calculated from the
virial formula, proved to be 0.13⫾0.03 kbar, somewhat
higher than the atmospheric pressure. The desired pressure FIG. 3. Water density distributions for the hydrophobic I-type wall at two
was obtained at ␮ ⬘ ⫽⫺6.15 kcal mole⫺1 . The corresponding wall-to-wall separations H.
equilibrium density was 0.991 g cm⫺3, 0.6% lower than re-
quired. The observed small inconsistency between the calcu-
lated equilibrium pressure and density of TIP4P water is The water density distribution ␳ (z) near the ␧⫽␧ 0 wall
hardly surprising because the original fitting of the TIP4P is shown in Fig. 3 for the largest separation tried 共65 Å兲 and
model26 to the experimental properties of water employed a also for a separation of 58.6 Å just beyond H c . The vanish-
different cutoff distance, number of molecules in the simula- ing density at z⬍2 Å is associated with the sharply ascend-
tion cell, and statistical ensemble 共NPT兲 共Ref. 36兲 than our ing water–wall repulsion and it is observed for all the walls
simulations. In deciding between the two alternative values studied in this work. By contrast, the substantially depressed
found for ␮ ⬘ , our reasoning was that the ‘‘overpressure’’ of density in the range 3⬍z⬍6 Å is a distinguishing feature of
0.13 kbar could be readily compensated for through Eq. 共6兲, the ␧⫽␧ 0 I-type wall, which can be compared with the the-
unlike the shift in density, which might have an unpredict- oretical predictions by Lum et al.17 for water in contact with
able effect on the structural organization of water near con- hydrophobic walls. As H approaches H c , the density depres-
fining surfaces. So, the excess chemical potential of ⫺6.10 sion progresses. The average water density between the walls
kcal mole⫺1, which exactly reproduced ␳ b , and somewhat is also substantially depressed 共Table I兲.
overestimated p b was adopted in all subsequent calculations. Although the water–wall interaction potential for the
I-type surfaces contains no explicit orientation-dependent
terms, the ␧⫽␧ 0 wall does have a perceptible orienting ef-
B. Nonorienting walls fect on the neighboring water molecules. This can be seen
We begin the discussion of the I-type walls with the one from Fig. 4, which shows the distribution of angles formed
characterized by the most shallow potential (␧⫽␧ 0 ). The by the O–H bond vector and the z axis for different separa-
behavior of water between two such walls was studied in the tions from the wall. The orientation of a typical water mol-
separation range 30⭐H⭐65 Å. At the wall-to-wall separa- ecule at the wall is such that one of its O–H bonds is pref-
tions less than a critical separation H c ⫽58.4 Å, the confined erentially directed toward the surface ( ␪ OH⫽180°), in
water experienced capillary evaporation 共cavitation兲, typical agreement with the results of MD simulations by Lee and
of hydrophobic confining surfaces: Starting with a certain Rossky29 who used the same potential model for water and
length of the MC run, the number of water molecules rapidly the wall. With increasing separation, the molecules reorient
decreased until nearly all of the molecules left the confined so that one of their O–H vectors look preferentially outward
region. At separations in a ⬃0.5 Å range below H c , the state from the surface ( ␪ OH⫽0). The orientational ordering near
of the system was dependent on the starting configuration the ␧⫽␧ 0 wall was found to be short range and lost almost
used to initiate the MC run: Some configurations resulted in completely at separations greater than two molecular diam-
cavitation, while the other remained liquid up to 107 MC eters.
passes. It is therefore quite likely that H c would shift to
larger separations if the MC runs would be longer. Note that
the observed H c ⫽58.4 Å agrees with a theoretical mean- TABLE I. Average density, hydration pressure, and integrated order param-
field estimate of the limit of metastability reported by Lum eter for water confined between I-type walls.
et al.17 for water confined between two hard walls 共⬃50 Å兲.
␧, kcal mole⫺1 H, Å ¯␳ , g cm⫺3 p h , kbar ⌶
On the other hand, H c is about an order of magnitude larger
than the critical slit width found by Wallqvist and Berne13 for 0.46 65 0.898 ⫺0.51 0.011
two hydrophobic ellipsoids 共⬃6.4 Å兲. A likely reason is a 0.46 58.6 0.885 ⫺0.53 0.010
6.24 40 0.980 ⫺0.12 0.018
comparatively small size and large curvature of the ellip-
10 40 1.013 0.13 0.017
soids, as well as the absence of an attractive contribution in 15 40 1.040 0.28 0.018
the water–ellipsoid interaction potential.

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
J. Chem. Phys., Vol. 117, No. 13, 1 October 2002 Simulation of hydration forces between walls 6277

FIG. 4. Orientational distribution of O–H bonds for some selected separa- FIG. 6. Hydration pressure as a function of wall-to-wall separation for the
tions from I-type walls with ␧⫽0.46 and 6.24 kcal mole⫺1. I-, A- (⌰⫽30°), and AD-type walls with ␧ H ⫽6.24 kcal mole⫺1 .

The simulation results for p h at H⬎H c 共Table I兲 show a


strong hydrophobic attraction between the ␧⫽␧ 0 walls,
which can be associated with the reduction in water density. tational ordering perceptibly increases, which can be judged
That is, as the separation between the walls is increased, the from the increase of the density-weighted order parameter ⌶
driving force behind the hydrophobic attraction changes in Table I. A typical water molecule in the first hydration
from cavitation to density depression. Importantly, the mag- layer now has one of its O–H bonds directed outward the
nitude of p h due to the density depression substantially ex- wall 共Fig. 4兲.
ceeds the upper limiting value of p h due to cavitation, p h The behavior of p h as a function of H is shown in Fig. 6
⬍p b 共0.13 kbar in our case兲. This result can be roughly together with the data for the ␧ H ⫽6.24 kcal mole⫺1 walls of
understood in terms of the fact that the compressibility of the other types. Because of the high computational cost of
liquid water is very low, so that a small reduction in the these simulations, we restricted ourselves only to few se-
water density may result in a significant stretching of the lected separations and did not undertake calculation of the
confined water. The substantial exceeding of p h over p b also detailed profile p h (H). It can nevertheless be seen that p h
means that the treatment of the hydrophobic attraction solely ceases changing its sign at H⬇18 Å, long before the sepa-
in terms of cavitation or ‘‘drying’’ can hardly provide a com- ration of 40 Å used in most of our simulations. 共Note that at
prehensive description of this phenomenon. ␧ H ⭓6.24 kcal/mole, all the walls showed no cavitation at all
The increase in the wall–water interaction strength on separations tried.兲
going from the ␧⫽␧ 0 to ␧⫽␧ H walls is accompanied by From the negative sign of p h exhibited by the I-type
substantial changes in the structure and properties of con- ␧ H ⫽6.24 kcal mole⫺1 wall at large separations, we con-
fined water. Already at ␧ H ⫽6.24 kcal mole⫺1 , the water den- clude that it is hydrophobic. The origin of its hydrophobicity
sity distribution becomes sharply structured and shows two can be understood from a simple analysis of the individual
distinct hydration layers next to the wall 共Fig. 5兲. The orien-
components of the water film tension, ␥ ⫽(⌬U⫺T⌬S
⫺ ␮ ⌬N)/A. At the given ␧ H and H⫽40 Å, ¯␳ is fairly close
to ␳ b 共see Table I兲 and hence ⌬N vanishes. For the
ensemble-averaged potential energy, the simulations give
⫺11.4 kcal mole⫺1, which is substantially lower than that for
bulk water 共⫺10.1 kcal mole⫺1兲. Considering that these en-
ergies refer to the systems with nearly the same N, we obtain
that ⌬U⬍0. Thus we can conclude that the positive sign of ␥
can only result from a negative ⌬S. The obvious cause of the
entropy loss in the confined region is the surface-induced
ordering. The entropic factor was earlier found to be domi-
nant for water confined between hydrophobic ellipsoids13
and it turns out to be important in our case, too.
As ␧ H is increased to 10 and then 15 kcal mole⫺1, the
average potential energy decreases to ⫺12.3 and ⫺13.5
kcal mole⫺1, respectively. This more than compensates for
the loss in entropy with the result that p h goes repulsive and
FIG. 5. Water density distribution for the I-, A-, and AD-type walls with the wall becomes hydrophilic 共Table I兲. Simultaneously, ¯␳
␧ H ⫽6.24 kcal mole⫺1 . becomes greater than ␳ b , as intuitively expected.

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
6278 J. Chem. Phys., Vol. 117, No. 13, 1 October 2002 Hayashi, Pertsin, and Grunze

TABLE II. Average density, hydration pressure, and integrated order param-
eter for water confined between A-type walls (H⫽40 Å).

␧ H , kcal mole⫺1 ⌰, deg ¯␳ , g cm⫺3 p h , kbar ⌶

6.24 30 0.938 ⫺0.37 ⫺0.052


10 30 0.965 ⫺0.21 ⫺0.090
15 30 0.992 ⫺0.09 ⫺0.123
6.24 75 0.982 ⫺0.15 ⫺0.050
10 75 1.007 ⫺0.07 ⫺0.078
15 75 1.040 ⫺0.02 ⫺0.108

drophobic at all ␧ H tried. From Tables I and II, one can also
see that at a given ␧ H the average water density between the
A-type walls is higher than that between the I-type walls, so
FIG. 7. Orientational order parameters S 10 for the I-, A- (⌰⫽30°), and
that the more attractive hydration forces observed between
AD-type walls with ␧ H ⫽15 kcal mole⫺1 . the A-type walls cannot be explained in terms of density
depression.
A comparison of the I-and A-type walls in terms of the
internal energy shows that the latter are 0.7–1.1 kcal mole⫺1
C. Proton-acceptor walls
inferior. The reason can be understood from Fig. 8, which
The simulations of water near the walls of the A type presents the average interaction energy of a water molecule
were performed for two different flexibility parameters, ⌰ with its surroundings as a function of its separation from the
⫽30° and 75°. We first discuss the results for ⌰⫽30°, I- and A-type walls with ␧ H ⫽15 kcal mole⫺1 . Both the total
which correspond to a stiffer hydrogen bond and hence a interaction energy and its constituents due to the water–
stronger orienting effect of the wall on the neighboring water water and water–wall interactions are shown. One can see
molecules. Compared to the I-type wall with ␧ H that the residence of a water molecule near the A-type wall is
⫽6.24 kcal mole⫺1 , the water density distribution near the energetically less favorable, mainly because of an appre-
A-type wall with the same ␧ H is less structured 共Fig. 5兲. Both ciable loss in the water–water interaction energy. The energy
the first and second density maxima of ␳ (z) are substantially loss is particularly pronounced in the first hydration layer,
lower in intensity and the gap between them is partly filled. where the orientational ordering is strongest and each water
An analysis of the angular distribution of the molecular di-
pole moments and O–H vectors shows that the water mol-
ecules near the wall of the A type have two distinct preferred
orientations associated with the formation of double and
single hydrogen bonds with the wall. The presence of two
kinds of hydrogen bonds, which differ noticeably in strength,
leads to a wide variation in the wall–water bond lengths and
the associated smearing of ␳ (z) between the first and second
maxima.
The property of the A-type wall to bind water molecules
in a highly asymmetric way, with their O–H bonds directed
preferentially towards the wall, brings about strong and fairly
long-ranged orientational ordering. This can be appreciated
from Fig. 7, which compares the order parameters S 10 (z) for
water in contact with the I-, A-, and AD-type walls. In the
case of the I-type walls, which are characterized by oscilla-
tions of S 10 (z) about zero, ⌶ is fairly small, positive, and
practically independent of the strength of the water–wall in-
teractions, as specified by the magnitude of ␧ H 共Table I兲. By
contrast, water in contact with the A-type walls shows large
negative ⌶, strongly dependent on ␧ H 共Table II兲.
As noted in Sec. III, the A-type wall with ⌰⫽30° is
capable of forming double hydrogen bonds with a water mol-
ecule with a total binding energy of 1.4␧ H . That is, based on
the wall–water interaction strength, the A-type walls can be
expected to be more hydrophilic than the I-type walls with
the same ␧ H . In spite of this, however, the former are found FIG. 8. Average interaction energy of a water molecule with its surround-
to result in lower p h 共cf. Tables I and II兲. Moreover, the ings and the individual contributions to this energy for the I-type and A-type
A-type walls all show attractive p h and hence they are hy- (⌰⫽30°) walls with ␧ H ⫽15 kcal mole⫺1 .

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
J. Chem. Phys., Vol. 117, No. 13, 1 October 2002 Simulation of hydration forces between walls 6279

molecule has to sacrifice at least one hydrogen bond with its TABLE III. Average density, hydration pressure, and integrated order pa-
rameter for water confined between AD-type walls (H⫽40 Å, ⌰⫽75°).
neighbors in order to bind to the wall. The constraints im-
posed by the orientational ordering upon the freedom of the ␧ H , kcal mole⫺1 ¯␳ , g cm⫺3 p h , kbar ⌶
water molecules to form hydrogen bonds with each other
6.24 1.023 0.05 0.000
remain important in the second hydration layer and at farther
10 1.059 0.12 ⫺0.003
separations. In this separation range, no hydrogen bonds with 15 1.093 0.15 ⫺0.033
the wall can be formed, yet the loss of the water–water in-
teraction energy is quite perceptible.
The presence of sharply directional interactions at the The orientational order generated in water by an AD-
wall–water interface affects not only the water–water but type wall can be appreciated from Fig. 7 for the case ␧ H
also the water–wall contribution to the average interaction ⫽15 kcal mole⫺1 where the magnitude of S 10 is greatest. The
energy. From the respective curve in Fig. 8, one can see that ordering is seen to be particularly pronounced in the range
the interaction energy of the A-type wall with a typical water from 6 to 7 Å and from 8.5 to 9.5 Å, where S 10 assumes large
molecule in the first hydration layer is only slightly greater in negative values. These ranges, however, correspond to
magnitude than the energy of a single hydrogen bond (␧ H minima in the water density distribution ␳ (z), while the
⫽15 kcal mole⫺1 for the example shown in the figure兲. This positive portions of S 10 correspond to maxima of ␳ (z) 共Fig.
suggests that the too stiff hydrogen bond potential used in 5兲. As a consequence, the integrated density-weighted pa-
this simulation impedes the formation of double hydrogen rameter ⌶ proves to be fairly small 共Table III兲. As ␧ H is
bonds with the wall. All the above-mentioned factors make decreased, ⌶ vanishes.
the A-type walls energetically less favorable compared to the For all tried ␧ H , p h ⬎0 and hence the AD-type walls are
I-type walls. The stronger orientational ordering induced by all hydrophilic. The average density ¯␳ is greater than ␳ b for
the A-type walls should also make them unfavorable from all ␧ H 共Table III兲. An analysis of the energetics of the con-
the entropic standpoint, too. fined water region in terms of the water–water and water–
The increase of the flexibility parameter ⌰ from 30° to wall interactions shows that the AD-type walls are interme-
75° makes all O–H bond orientations with ␪ i ranging from diate in the water–water interaction energy between the I-
105° to 180° energetically equivalent. This reduces the ori- and A-type walls with the same ␧ H . However, the ability of
entational ordering 共see the changes in ⌶ in Table II兲 and the AD-type walls to form triple hydrogen bonds with water,
affords more orientational freedom for the water molecules with a total binding energy of 3 ␧ H , makes this type of wall
in forming hydrogen bonds with their neighbors and the much more favorable for water, compared to the I- and
wall. As a consequence, the wall becomes less hydrophobic A-type walls.
共or more hydrophilic兲 which manifests itself in an increase in
ph . VII. CONCLUDING REMARKS
Turning to the calculated average densities in Table II,
In this work we used the GCMC technique to study the
one can see that the changes in the density and hydration
hydration forces between structureless flat walls. Most atten-
pressure are symbate: With increasing ␧ H , both ¯␳ and p h
tion was given to large wall-to-wall separations, where the
increase. At the same time, the intuitive expectation that for
force oscillations due to the water layering effects decayed
hydrophobic confining walls ¯␳ should always be lower than
and the sign of the hydration pressure can be used as a cri-
␳ b proves to be fallacious: At ⌰⫽75° and ␧ H terion for the hydrophilicity/hydrophobicity of the walls. For
⭓10 kcal mole⫺1 , the system shows an enhanced ¯␳ along
the wall with the weakest interaction with water (␧
with an attractive 共hydrophobic兲 p h . A similar example has
⫽0.46 kcal mole⫺1 ), we observed capillary evaporation at
been reported by Besseling8 in his theoretical work.
H⭐58.4 Å and a strong attraction due to density depression
at larger H. In the latter case, the magnitude of p h substan-
tially exceeded p b 共the upper limiting pressure due to evapo-
D. Walls bearing both proton acceptors
ration兲, which cast some doubt on the treatment of capillary
and proton donors
evaporation as the main source of hydrophobic attraction.
With the accepted orientation of the lone pairs in the Our results for the hydrophilic walls are, in some re-
water molecule, an AD-type wall sees the molecule as pos- spects, opposite to the findings reported by Forsman et al.21
sessing a V d symmetry. This means, in particular, that the In contrast to the cited work,21 we found that the inclusion of
wall–water potential is invariant with respect to a fourfold orientation dependent terms in the water–wall interaction po-
inversion axis transformation of the water molecule and the tential added an attractive contribution to the hydration force
associated change in the sign of the molecular dipole mo- by making the confined region less favorable for water from
ment. The water–water potential, however, does not possess both the entropic and energetic points of view. While the loss
this symmetry. As a consequence, the whole system may in entropy can be well understood in terms of the surface-
well show a nonvanishing S 10 even though the wall itself does induced orientational ordering and confinement of the libra-
not distinguish between the symmetrically related configura- tional motion of water molecules, the reasons of the energy
tions with opposite dipole moments. 共Note that similar argu- loss are not so apparent. One is the reduction of the water–
ments apply to the I-type surfaces, which show noticeable S 10 water contribution to the internal energy due to the distur-
and ⌶ despite the fact that the I-type walls perceive the water bance of the hydrogen bonding network natural for water.
molecule as being isotropically symmetric.兲 Another reason has to do with constraints imposed by the

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
6280 J. Chem. Phys., Vol. 117, No. 13, 1 October 2002 Hayashi, Pertsin, and Grunze

15
directionality of the hydrogen bond upon the ability of water J. Forsman, B. Jönsson, and T. Åkesson, J. Phys. Chem. B 102, 5082
molecules to form multiple hydrogen bonds with the surface. 共1998兲.
16
J. Forsman, C. E. Woodward, and B. Jönsson, J. Colloid Interface Sci.
In the model potentials used in describing the water–surface
195, 264 共1997兲.
interaction, the orientation dependence of the hydrogen bond 17
K. Lum, D. Chandler, and J. D. Weeks, J. Phys. Chem. B 103, 4570
energy reflects not only the intrinsic trend of the hydrogen 共1999兲.
18
bond to linearity but also the flexibility of the surface groups D. M. Huang, P. L. Geissler, and D. Chandler, J. Phys. Chem. B 105, 6704
involved in the hydrogen bonding, i.e., the ability of these 共2001兲.
19
P. G. Bolhuis and D. Chandler, J. Chem. Phys. 113, 8154 共2001兲.
groups to adapt their orientations to the approaching water 20
A. Luzar, D. Bratko, and L. Blum, J. Chem. Phys. 86, 2955 共1987兲.
molecules. The confinement of this ability should obviously 21
J. Forsman, C. E. Woodward, and B. Jönsson, Langmuir 13, 5459 共1997兲.
be energetically unfavorable. 22
Surfaces bearing solely proton donors are of little practical importance as
far as strong hydrogen bonds such as O–H¯OH2 or N–H¯OH2 are
ACKNOWLEDGMENTS concerned. In both cases, the proton-donor group can also serve as a
This work was supported by the Deutsche Forschungs- proton acceptor for the water molecule, so that the respective surfaces
gemeinschaft, the Office for Naval Research, and the Fond cannot formally be regarded as purely proton donating.
23
E. A. Vogler, in Water in Biomaterials Surface Science, edited by M.
der Chemischen Industrie. The authors thank Professor E. A.
Morra 共Wiley, New York, 2001兲, pp. 149–182.
Vogler, Professor N. Spencer, Dr. P. Schmidt, and Dr. G. 24
R. Evans and U. M. B. Marconi, J. Chem. Phys. 86, 7138 共1987兲.
Hähner for critical reading of the manuscript and helpful 25
As shown by Besseling 共Ref. 8兲 in his mean-field lattice theory of hydra-
comments. tion forces, ␥ (H) may have an extremum regardless of the presence of
layering effects. Such a situation, which does not allow the use of p h as a
1
J. Israelachvili and H. Wennerström, Nature 共London兲 379, 219 共1996兲. hydrophilicity/hydrophobicity criterion at long separations, can, however,
2
E. A. Vogler, Adv. Colloid Interface Sci. 74, 69 共1998兲. be readily recognized both in real and computer experiments.
3
J. Israelachvili, Intermolecular and Surface Forces 共Academic, London, 26
W. L. Jorgensen, J. Chandrasekhar, J. D. Madura, R. W. Impey, and M. L.
1992兲. Klein, J. Chem. Phys. 79, 926 共1983兲.
4
J. Lyklema, Fundamentals of Interface and Colloid Science I: Fundamen- 27
A. Wallqvist, Chem. Phys. Lett. 165, 437 共1990兲.
tals 共Academic, London, 1991兲.
5
28
J. C. Shelley and G. N. Patey, Mol. Phys. 88, 385 共1996兲.
J. Lyklema, Fundamentals of Interface and Colloid Science II: Solid- 29
S. H. Lee and P. J. Rossky, J. Chem. Phys. 100, 3334 共1994兲.
Liquid Interfaces 共Academic, London, 1991兲. 30
R. L. C. Wang, H. J. Kreuzer, and M. Grunze, Phys. Chem. Chem. Phys.
6
R. P. Rand and V. A. Parsegian, Biochim. Biophys. Acta 778, 224 共1989兲.
7
V. A. Parsegian, Adv. Colloid Interface Sci. 16, 49 共1982兲. 2, 3613 共2000兲.
8
N. A. Besseling, Langmuir 13, 2113 共1997兲.
31
R. H. Swendsen and J.-S. Wang, Phys. Rev. Lett. 58, 86 共1987兲.
9
R. Podgornik, J. Chem. Phys. 91, 5840 共1989兲.
32
M. Mezei, Mol. Phys. 61, 565 共1987兲.
10
M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids 共Claren-
33
A. J. Pertsin and M. Grunze, Langmuir 16, 8829 共2000兲.
don, Oxford, 1993兲.
34
M. R. Stapleton and A. Panagiotopoulos, J. Chem. Phys. 92, 1285 共1990兲.
35
11
J. C. Shelley and G. N. Patey, J. Chem. Phys. 102, 7656 共1995兲. J. Hermans, A. Pathiaseril, and A. Anderson, J. Am. Chem. Soc. 110, 5982
12
B. Svensson and C. E. Woodward, J. Chem. Phys. 100, 4575 共1994兲. 共1988兲.
36
13
A. Wallqvist and B. J. Berne, J. Phys. Chem. 99, 2893 共1995兲. The difference in the statistical ensembles could introduce some discrep-
14 ancy in view of the relatively small system size used in deriving the TIP
J. Forsman, B. Jönsson, and C. E. Woodward, J. Phys. Chem. 100, 15005
共1996兲. 4P model (N⫽125) 共Ref. 26兲.

Downloaded 17 Oct 2002 to 129.206.85.25. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp

You might also like