You are on page 1of 11

Langmuir 2001, 17, 1773-1783 1773

Molecular Simulation of a Dichain Surfactant/Water/


Carbon Dioxide System. 1. Structural Properties of
Aggregates
S. Salaniwal,†,‡ S. T. Cui,†,‡ H. D. Cochran,†,‡ and P. T. Cummings*,‡,§
Department of Chemical Engineering, University of Tennessee,
Knoxville, Tennessee 37996-2200, Departments of Chemical Engineering, Chemistry, and
Computer Science, University of Tennessee, Knoxville, Tennessee 37996-2200, and Chemical
Technology Division, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831-6224

Received April 12, 2000. In Final Form: October 11, 2000

Molecular dynamics simulation of a dichain surfactant + water + carbon dioxide (solvent) system is
performed to study the structural properties of reversed micelle-like surfactant aggregates formed in the
Downloaded via INDIAN INST OF TECHNOLOGY BOMBAY on February 3, 2021 at 12:33:37 (UTC).

system. The simulations use a detailed and realistic molecular model for the surfactant molecule and
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

explicit representation of the water and solvent molecules to enable quantitative comparisons with a prior
experimental (small-angle neutron scattering) study. The results of the simulation are found to be in
reasonable agreement with experimental values. The simulations show that the size and shape of the
surfactant aggregates depends on their water-to-surfactant ratio. A higher water-to-surfactant ratio results
in larger and more spherical aggregates. The two distinct tails of the surfactant molecule exhibit different
conformations in carbon dioxide indicating contrasting CO2-philic behavior. The perfluoroalkane tails
assume more extended conformation than the alkane tails. The microstructure of the aqueous core reveals
that the water molecules in the interfacial region are strongly oriented in response to the electric fields
of the anionic headgroups and sodium counterions, while water near the center of the core approaches
bulklike properties with the presence of a hydrogen-bonded network.

Introduction advantage of these techniques lies in their ability to use


Over the past few years, molecular simulation tech- simple molecular models for surfactant molecules (e.g.,
niques have been increasingly applied to investigate the bead-spring model used by Smit et al.)7-9 and other
morphology, dynamics, and rheology of surfactant ag- simplifications, such as constrained aggregate3-5 or lattice-
gregates (micelles) formed in aqueous system.1-14 The based simulations,13,14 to obtain a description of the size,
shape, surface roughness, and the internal structure of
* To whom correspondence may be addressed. micelles. In fact, in some instances,1,9 these studies have
† Department of Chemical Engineering, University of Tennessee.
even helped clarify interpretations of prior experimental
‡ Chemical Technology Division, Oak Ridge National Laboratory.
§ Departments of Chemical Engineering, Chemistry, and Com-
results. In certain conditions where experiments are
puter Science, University of Tennessee.
impractical or impossible (e.g., high temperatures and
(1) Karaborni, S.; et al. Simulating the Self-Assembly of Gemini pressures), molecular simulation techniques can be suc-
(Dimeric) Surfactants. Science 1994, 266, 254-256. cessfully applied to obtain the much-needed information
(2) Karaborni, S.; et al. Molecular Dynamics Simulations of Oil about the various properties of surfactant aggregates.
Solubilization in Surfactant Solutions. Langmuir 1993, 9, 1175-1178.
(3) Karaborni, S.; O’Connell, J. P. Molecular Dynamic Simulations Surprisingly, very few simulation studies15-17 have ad-
of Model Micelles. 3. Effects of Various Intermolecular Potentials. dressed the phenomenon of surfactant aggregation in
Langmuir 1990, 6, 905-911. nonpolar solvents. The phenomenon of surfactant ag-
(4) Karaborni, S.; O’Connell, J. P. Molecular Dynamic Simulations
of Model Micelles. 4. Effects of Chain Length and Headgroup Char- gregation in nonpolar solvents (e.g., benzene, toulene,
acteristics. J. Phys. Chem. 1990, 94, 2624-2631. cyclohexane, carbon tetrachloride, etc.) results in the
(5) Karaborni, S.; O’Connell, J. P. Molecular Dynamics Simulations formation of aggregates known as reversed (or inverted)
of Model Chain Molecules and Aggregates including Surfactants and
Micelles. Tenside, Surfactants, Deterg. 1993, 30, 235-242. micelles (RMs). Structurally, these aggregates are char-
(6) Karaborni, S.; Smit, B. Computer Simulations of surfactant acterized by the presence of a hydrophilic (often aqueous)
structures. Curr. Opin. Colloid Interface Sci. 1996, 1, 411-415. core that is surrounded by a hydrophobic corona made up
(7) Smit, B.; et al. Computer Simulations of Surfactant Self-Assembly.
Langmuir 1993, 9, 9-11. of surfactant tail groups. In essence, the structure of RMs
(8) Smit, B.; et al. Structure of a Water/Oil Interface in the Presence is simply reversed from their aqueous counterparts
of Micelles: A Computer Simulation Study. J. Phys. Chem. 1991, 95, because of the reversal in the roles of the hydrophilic and
6361-6368.
(9) Smit, B.; et al. Computer simulations of a water/oil interface in hydrophobic moieties of the surfactant molecules. These
the presence of micelles. Nature 1990, 348, 624-625. aggregates find numerous applications in fields such as,
(10) Jonsson, B.; Edholm, O.; Telleman, O. Molecular dynamics
simulations of a sodium octanoate micelle in aqueous solution. J. Chem.
Phys. 1986, 85 (4), 2259-2271. (14) Floriano, M. Antonio; Caponetti, E.; Panagiotopoulos, A. Mi-
(11) Shelley, J. C.; Sprik, M.; Klein, M. L. Molecular Dynamics cellization in Model Surfactant Systems. Langmuir 1999, 15, 3143-
Simulation of an Aqueous Sodium Octanoate Micelle using Polarizable 3151.
Surfactant Molecules. Langmuir 1993, 9, 916-926. (15) Linse, P. Molecular dynamics study of the aqueous core of a
(12) Watanabe, K.; Ferrario, M.; Klein, M. L. Molecular Dynamics reversed ionic micelle. J. Chem. Phys. 1989, 90 (9), 4992-5004.
Study of a Sodium Octanoate Micelle in Aqueous Solution. J. Phys. (16) Tobias, D. J.; Klein, M. L. Molecular Dynamics Simulations of
Chem. 1988, 92, 819-821. a Calcium Carbonate/Calcium Sulfonate Reverse Micelle. J. Phys. Chem.
(13) Mackie, A. D.; Panagiotopoulos, A. Z.; Szleifer, I. Aggregation 1996, 100, 6637-6648.
Behavior of a Lattice Model for Amphiphiles. Langmuir 1997, 13, 5022- (17) Brown, D.; Clarke, J. H. R. Molecular Dynamics Simulation of
5031. a Model Reverse Micelle. J. Phys. Chem. 1991, 92, 2881-2888.

10.1021/la000554f CCC: $20.00 © 2001 American Chemical Society


Published on Web 02/10/2001
1774 Langmuir, Vol. 17, No. 5, 2001 Salaniwal et al.

lubrication, emulsion polymerization reactions, drug Molecular simulation techniques can play an important
delivery, and enhanced oil recovery, to name a few. role in developing this much-needed knowledge base.
Recently, some experimental studies18-24 have inves- As an initial step in this direction, we performed
tigated the feasibility of using carbon dioxide (CO2) as a molecular dynamics (MD) simulation of a dichain surf-
nonpolar solvent medium for the formation of RMs. Carbon actant + water + carbon dioxide (solvent) system. The
dioxide is a nonpolar compound that is a chemically inert dichain surfactant molecule [(C7F15)(C7H15)CHSO4-Na+]
(nontoxic, nonflammable) and an environmentally benign (also known as a hybrid surfactant molecule) is one of the
alternative to potentially hazardous industrial solvents few CO2-philic surfactant molecules that has been shown29
currently in use. Also, its easy availability and low cost to form RMs with aqueous cores in CO2. Recently, Eastoe
make it a cheap raw material. Thus, formation of et al.35 performed a small-angle neutron scattering (SANS)
surfactant aggregates in CO2 offers the possibility of study on the dichain surfactant system and obtained
carrying out a wide range of industrial operations in a information regarding the structural properties (size and
medium with minimum environmental impact. Unfor- shape) of the RMs formed. In the simulations presented
tunately, as shown by Consani and Smith,25 most of the here, this particular surfactant molecule was chosen for
industrially available surfactant molecules are incapable two main reasons. First, in terms of the molecular
of forming stable RMs in CO2 because of their negligible complexity this is a relatively small molecule compared
solubility. This is attributed to the fact that most of these to other known CO2-philic surfactant molecules. Thus,
surfactants are suited for aqueous solvents while CO2 is from a computational perspective, it is relatively easy to
relatively nonpolar (CO2 has a weak quadrupole) because model this surfactant to a sufficient degree of detail and
of its low polarizibility and dielectric constant. This has study the aggregation behavior for a relatively large
motivated a number of experimental efforts26-34 focused system size. Second, the structural properties of the
on discovering suitable surfactant molecules. These surf- surfactant aggregates estimated from simulations can be
actants are termed “CO2-philic” surfactants because they quantitatively compared to the experimental results of
should exhibit favorable interactions with CO2 to enable Eastoe et al.35 In our previous paper,36 we presented
dispersion of the aggregates. Although these studies have preliminary results showing the spontaneous aggregation
been partially successful in identifying a few CO2-philic of the dichain surfactant and water molecules into RM-
surfactant molecules, little is known about the factors like aggregates in this ternary system. The results
determining CO2-philicity. Knowledge of these underlying presented there were for a small (exploratory) system that
principles can be quite helpful in guiding future efforts in was used to test the feasibility of these simulations. In
discovering or designing new and improved CO2-philic this paper, we present additional simulations on a much
surfactants that are suitable for industrial applications. larger system intended to reproduce the experimental
system.35 We also describe the detailed molecular models
(18) Hoefling, T. A.; Enick, R. M.; Beckman, E. J. Microemulsions in
Near-Critical and Supercritical CO2. J. Phys. Chem. 1991, 95, 7127- for the dichain surfactant molecule, water, and carbon
7129. dioxide adopted for this study and the simulation meth-
(19) Hoefling, T. A.; et al. The Incorporation of a Fluorianted Ether odology used to simulate this complex physical system.
Functionality into a Polymer or Surfactant to Enhance CO2-Solubility. Quantitative comparisons of the simulation results with
J. Supercrit. Fluids 1992, 5, 237-241.
(20) Newman, D. A.; et al. Phase Behavior of Fluoroether-Functional the experimental results of Eastoe et al.35 are presented.
Amphiphiles in Supercritical Carbon Dioxide. J. Supercrit. Fluids 1993, Also, microscopic structural details of the aqueous core
6, 205-210. are presented.
(21) Hoefling, T. A.; et al. Design and Synthesis of Highly CO2-soluble
Surfactants and Chelating Agents. Fluid Phase Equilib. 1993, 83, 203-
212.
(22) Jimenez, M. M.; Luque de Castro, M. D. Reverse-micelle POTENTIAL MODELS
formation: a strategy for enhancing CO2-supercritical fluid extraction
of polar analytes. Anal. Chim. Acta 1998, 358, 1-4. Since one of the objectives of the present work is to
(23) Iezzi, A.; et al. Gel formation in carbon dioxide-semifluorinated
alkane mixtures and phase equilibria of a carbon dioxide-perfluorinated make quantitative comparisons with the experimental
alkane mixture. Fluid Phase Equilib. 1989, 52, 307-317. work of Eastoe et al.,35 a high degree of realism is desirable
(24) Wikramanayake, R.; Enick, R.; Turberg, M. The phase behavior to accurately represent all the intermolecular interactions
and gel formation of binary mixtures of carbon dioxide and semi-
fluorinated alkanes. Fluid Phase Equilib. 1991, 70, 107-118.
present in the system. Thus, the simulations presented
(25) Consani, K. A.; Smith, R. D. Observations on the Solubility of in this work involve detailed and realistic molecular
Surfactants and Related Molecules in Carbon Dioxide at 50 °C. J. models for all the three chemical species (dichain surf-
Supercrit. Fluids 1990, 3, 51-65. actant, water, and carbon dioxide) present in the system.
(26) DeSimone, J. M.; Guan, Z.; Elsbernd, C. S. Synthesis of
Fluoropolymers in Supercritical Carbon Dioxide. Science 1992, 257, Although this makes the system computationally complex
945-947. and demanding, it minimizes any artificiality that might
(27) DeSimone, J. M.; et al. Dispersion Polymerizations in Super- be included in the system and affect the dynamics of
critical Carbon Dioxide. Science 1994, 265, 356.
(28) Johnston, K. P.; et al. Water-in-Carbon Dioxide Microemul-
surfactant aggregation and/or the morphology of the
sions: An Environment for Hydrophiles Including Proteins. Science aggregates formed. With these factors in consideration,
1996, 271, 624-626. the potential model for the hybrid surfactant molecule
(29) Harrison, K. L.; et al. Water-in-Carbon Dioxide Microemulsions used in this simulation study was constructed by as-
with a Fluorocarbon-Hydrocarbon Hybrid Surfactant. Langmuir 1994,
10, 3536-3541. sembling existing models for each of the four functionally
(30) Heitz, M. P.; et al. Water Core within Perfluoropolyether-Based distinct parts of the molecule, i.e., the alkane tail,
Microemulsions Formed in Supercritical Carbon Dioxide. J. Phys. Chem. perfluoroalkane tail, sulfate headgroup, and the sodium
1997, 101, 6707-6714.
(31) McClain, J. B.; et al. Design of Nonionic Surfactants for
counterion. This allows the surfactant model to capture
Supercritical Carbon Dioxide. Science 1996, 274, 2049-2052. most of the essential characteristics of the real dichain
(32) McFann, G. J.; Johnston, K. P.; Howdle, S. M. Solubilization in surfactant molecule.
Nonionic Reverse Micelles in Carbon Dioxide. AIChE J. 1994, 40, 543-
555.
(33) Fulton, J. L.; et al. Aggregation of Amphiphilic Molecules in (35) Eastoe, J.; et al. Droplet Structure in a Water-in-CO2 Micro-
Supercritical Carbon Dioxide: A Small-Angle X-ray Scattering Study. emulsion. Langmuir 1996, 12, 1423-1424.
Langmuir 1995, 11, 4241-4249. (36) Salaniwal, S.; et al. Self-Assembly of Reverse Micelles in Water/
(34) Cooper, A. I.; et al. Liquid-Liquid Extractions Using Dendrimer- Surfactant/Carbon Dioxide Systems by Molecular Simulation. Langmuir
Modified Carbon Dioxide. Nature 1997, 389, 368-371. 1999, 15 (16), 5188-5192.
Dichain Ternary System Langmuir, Vol. 17, No. 5, 2001 1775

For interactions between unlike sites (on the same or Table 1. Potential Parameters for theIntramolecular and
different molecules) we have used the conventional Intermolecular Degrees of Freedom for the Alkane Tail
Lorentz-Berthelot combining rules without any fitted and the Perfluoroalkane Tail
binary parameter. type of interaction alkane tail perfluoroalkane tail
Alkane Tail. The potential model for the alkane tail Lennard-Jones σCH3 ) σCH2 ) 3.93 Å σCF3 ) σCF2 ) 4.60 Å
of the surfactant molecule is the united atom model for interactions CH3/k ) 114 K CF3/k ) 79 K
alkanes proposed by Siepmann and co-workers37 (SKS CH2/k ) 47 K CF2/k ) 30 K
model). This model accurately predicts the experimentally bond-angle bending kb ) 62500 K/rad2 kb ) 62500 K/rad2
measured phase envelopes of n-alkanes. More importantly, θeq ) 114.0° θeq ) 114.6°
as shown by Cui et al.,38 the SKS model predicts the torsional a0 ) 1078.16 K a0 ) 959.4 K
a1 ) 355.03 K a1 ) -282.7 K
solubility of n-alkanes in CO2 that is also consistent with a2 ) -68.19 K a2 ) 1355.2 K
experimentally determined values. In this potential model, a3 ) 791.32 K a3 ) 6800.0 K
each of the methylene (CH2) and methyl (CH3) functional a4 ) -7875.3 K
groups is represented by a single spherical interaction a5 ) -14168.0 K
site (united atom) with its interaction center located at a6 ) 9213.7 K
the center of the carbon atom. The interaction sites are a7 ) 4123.7 K
connected via rigid bonds of length 1.54 Å. Interaction
sites on different surfactant molecules and those on the depth and size parameters are used to represent the CF2
same molecule separated by more than three bonds and CF3 functional groups of the tail (σCF3 ) σCF2 ) 4.6 Å,
interact via the well-known 12-6 Lennard-Jones (LJ) CF3/k ) 79 K, σCF2/k ) 30 K). The bond length between the
potential given by united atoms is assumed fixed at 1.54 Å. The nonbonded
interactions are described via a 12-6 LJ potential (eq 1)

[( ) ( ) ] while the bond-angle bending and torsional angle potential


σij 12 σij 6
ULJ(rij) ) 4ij - (1) account for the intramolecular interactions. The bond
rij rij
angle potential is similar to that of the alkane tail (eq 2)
with the only difference being the slightly larger (114.6°)
where ij is the well depth and σij is the zero point of the
equilibrium bond angle. The torsional angle potential for
potential between site i and site j. The LJ size parameters
the perfluoroalkane tail is given by
for the methyl and methylene functional groups are same
(σCH3 ) σCH2 ) 3.93 Å) while the well-depth energy 7
parameters are CH3/k ) 114 K and σCH2/k ) 47 K,
respectively. The intramolecular interactions include the
Utorsion(φ) ) a0 + ∑
i)1
ai (cos φ)i (4)
bond-angle bending potential to account for the bond
vibrational motion, and torsional angle potential to where φ is the dihedral angle. The important features of
account for the spatial orientation of the tails. The bond the torsional potential for perfluoroalkane are that it has
angle bending potential is described by a harmonic a twisted trans minimum at a dihedral angle of φ ∼ 17°,
function of the form a gauche minimum at φ ∼ 125°, and a second gauche
1 minimum at around φ ∼ 83°. Table 1 also lists the potential
Ubending(θi) ) k (θ - θeq)2 (2) parameters for the perfluoroalkane model.
2 θ i Sulfate Headgroup and Sodium Counterion. The
where θeq is the equilibrium angle between successive potential model for the sulfate headgroup of the surfactant
bonds and kθ is the force constant and is a measure of the molecule is a fully atomistic (each interaction site rep-
rigidity of the bond angle. The torsional angle potential resents an individual atom) model proposed by Cannon
for the alkane tail is given by39 and co-workers.41 The molecule has a rigid tetrahedral
structure (O-S-O bond angle of 109.4°) with the sulfur
3 atom at the axis and the oxygen atoms at the four apexes.
Utorsion(φ) ) ∑
i)0
ai cos(iφ) (3) The bond between the sulfur and oxygen atoms is fixed
at 1.489 Å. To account for electrostatic interactions, the
sulfur atom carries a charge of +2e while three oxygen
where φ is the torsional angle that is defined as the angle atoms each carry a charge of -1e. The fourth oxygen atom
between the normal to the two planes, each of which is does not carry any negative charge as it is attached to the
formed by three successive interaction sites. Table 1 lists CH group, which in turn is attached to the two tails. A
the potential parameters for the alkane model. sum of Coulombic (for electrostatic interactions) and LJ
Perfluoroalkane Tail. The potential model for the terms describes the intermolecular interactions

[( ) ( ) ]
perfluoroalkane tail used in this study is that proposed
by Cui et al.40 and is quite similar to that of the alkane qiqj σij 12 σij 6
tail described above. This model was chosen because it is
fitted to the phase envelope of n-perfluroralkanes and
U(rij) ) ∑i ∑j r + 4ij
rij
-
rij
(5)
ij
also does good de novo prediction of pefluoroalkane/CO2
phase equilibrium.38 United atoms with effective LJ well- where qi represents the Coulombic charge on the atom i.
A single LJ sphere with a unit positive charge represents
(37) Siepmann, J. I.; Karaborni, S.; Smit, B. Simulating the critical each sodium ion. Table 2 lists the potential parameters
behavior of complex fluids. Nature 1993, 365, 330-332.
(38) Cui, S. T.; Cochran, H. D.; Cummings, P. T. Vapor-liquid phase for sulfur and oxygen atoms and sodium ion used in this
coexistence of alkane carbon dioxide and perfluoralkane carbon dioxide study. According to this molecular model, each surfactant
mixtures. J. Phys. Chem. 1999, 103, 4485-4491. molecule consists of 21 interaction sites including one for
(39) Ploeg, P. van der; Berendsen, H. J. C. Molecular dynamics
simulation of a Bilayer membrane. J. Chem. Phys. 1982, 76, 3271- the Na+ ion.
3276.
(40) Cui, S. T.; et al. Intermolecular potentials and vapor-liquid- (41) Cannon, W. R.; Pettitt, B. M.; McCammon, J. A. Sulfate Anion
phase equilibria of perfluorinated alkanes. Fluid Phase Equilib. 1998, in Water: Model Structural, Thermodynamic, and Dynamic Properties.
146, 51-61. J. Phys. Chem. 1994, 98, 6225-6230.
1776 Langmuir, Vol. 17, No. 5, 2001 Salaniwal et al.

Table 2. Potential Parameters for the Sulfur and Oxygen


Atoms of the Headgroup and the Sodium Counterion
atom type σ (Å) /k (K) q (e)
sulfur 3.55 126 +2
oxygen 3.15 126 -1
sodium 2.667 37.65 +1

Table 3. Potential Parameters for the Carbon Dioxide


Potential Model
atom type σ (Å) /k (K) q (e)
carbon 2.757 28.129 +0.6512
oxygen 3.033 80.507 -0.3256

Table 4. Potential Parameters for the SPC/E Potential


Model for Water
atom type σ (Å) /k (K) q (e)
hydrogen 0.0 0.0 +0.4238
oxygen 3.166 78.2 -0.8476

Carbon Dioxide. The carbon dioxide (solvent) mol-


ecules are represented by a simple site-based potential
model proposed by Harris and Yung,42 which has been
shown to accurately reproduce the experimental vapor-
liquid coexistence curve and critical point. The model
consists of three LJ interaction sites with a Coulombic
charge centered on each atom. The molecule has a rigid
linear (O-C-O bond angle ) 180°) geometry with the
C-O bond length of 1.149 Å. The carbon atom carries a
charge of +0.6512e while the two oxygen atoms carry a
charge of -0.3256e each to make the molecule electrically
neutral. Table 3 lists the parameters of the potential model.
Water. For water molecules, the extended simple point
charge (SPC/E) potential model of Berendsen et al.43 is
used in this study. The model has a single LJ site for the
oxygen atom that also carries a charge of -0.8476e. The
two hydrogen atoms are simply represented by point
charges, each carrying a charge of +0.4238e. Rigid
molecular geometry is assumed with the O-H bond length
of unity and H-O-H bond angle equal to the tetrahedral
angle (109.4°). Table 4 lists the potential parameters for
this model. Figure 1. Snapshots of the starting configurations for the small
system used in this study: (a) aggregated configuration and (b)
Simulation Methodology scattered configuration. The color scheme of the various species
is as follows: light blue for perfluoroalkane tail, dark blue for
Molecular dynamics simulations of two different system alkane tail, yellow for sulfur, red for oxygen, white for hydrogen,
sizes were performed. The first system (referred to as the and gray for sodium.
small system) consists of 30 dichain surfactant, 132 water,
and 2452 CO2 molecules. This system (with approximately Eastoe et al.35 In this system, the CO2 is in the high-
8300 interaction sites) was chosen for exploratory calcu- temperature liquid state at 500 bar. The equations of
lations (with computational economy in mind) and to motion for each interaction site were solved using the
provide qualitative insight into the dynamics of aggrega- Rattle algorithm44 to constrain the rigid bond lengths in
tion and the morphology of the aggregates. Some pre- molecules. The time step used to solve the equations of
liminary results on this system were reported previously.36 motion was 1.483 fs.
The state conditions used with this system were, T ) 310 As discussed in the following paper of this series,
K and Fsolvent ) 0.482 g/cm3. These conditions were chosen aggregation of reverse micelles with aqueous cores in
to ensure that CO2 was in the supercritical state. The carbon dioxide from this system, with the scattered
critical point of CO2 is Tc ) 304.3 K, Pc ) 73.8 bar, and starting configuration, is rapid, on the order of 0.2 µs,
Fc ) 0.468 g/cm3. The second system (referred to as the compared with the 1-10 µs for a typical oil-in-water
large system) has 33 surfactant, 1175 water, and 12 800 micelle, because of the strong Coulombic forces present
CO2 molecules at T ) 298 K and Fsolvent ) 0.848 g/cm3. This in the system studied here in comparison with the much
system mimics the surfactant and water concentrations, weaker hydrophobic forces in the oil-in-water system.
temperature, and density used by Eastoe et al.35 in their The simulations were performed with two independent
SANS study of this ternary system. The system size initial conditions, viz., (a) an aggregated starting config-
corresponds to one “average-sized” aggregate observed by uration and (b) a scattered starting configuration, shown
in Figures 1 and 2 for the small system and the large
(42) Harris, J. G.; Yung, K. H. Carbon Dioxide’s Liquid-Vapor system, respectively. In the aggregated starting config-
Coexistence Curve and Critical Properties as Predicted by a Simple
Molecular Model. J. Phys. Chem. 1995, 99, 12021-12024. (44) Andersen, H. C. Rattle: a velocity verlet version of the shake
(43) Berendsen, H. J. C.; Grigera, J. R.; Straatsma, T. P. The Missing algorithm for molecular dynamics calculations. J. Comput. Phys. 1983,
Term in Effective Pair Potentails. J. Chem. Phys. 1987, 91, 6269-6271. 52, 24-34.
Dichain Ternary System Langmuir, Vol. 17, No. 5, 2001 1777

Figure 2. Snapshots of the starting configurations for the large


system used in this study: (a) aggregated configuration and (b)
scattered configuration. The color scheme of the various species
is as follows: light blue for perfluoroalkane tail, dark blue for
alkane tail, yellow for sulfur, red for oxygen, white for hydrogen,
and gray for sodium.

uration, all the water molecules were placed within a


spherical core at the center of a cubic simulation box. The Figure 3. Snapshots of the small system (aggregated starting
surfactant molecules with fully extended tails were placed configuration), large system (aggregated starting configuration),
on the surface of the core with their headgroups pointing and large system (scattered starting configuration) at (a) 1.036,
(b) 1.0, and (c) 2.0 ns. The color scheme of the various species
toward the center. Structurally, this configuration re- is as follows: light blue for perfluoroalkane tail, dark blue for
sembles a reversed micellar aggregate and hence the term alkane tail, yellow for sulfur, red for oxygen, white for hydrogen,
aggregated starting configuration. In contrast, the scat- and gray for sodium.
tered starting configuration has all the water molecules
scattered within the simulation box. These initial condi- pearance of RMs, i.e., aggregates consisting of aqueous
tions are two of the infinite configurations that can be cores surrounded by surfactant molecules with their
used as starting conditions and were chosen solely for the headgroups immersed in the core and the tails form a
sake of convenience. corona. Figure 3a shows the system consists of three
roughly similar-sized aggregates with their aqueous cores
Results and Discussion well shielded from the surrounding solvent. The surfactant
Figure 3 shows the snapshots of the small system (from aggregates formed in the small system with scattered
aggregated starting configuration), large system (from starting configuration (not shown) were also similar to
aggregated starting configuration), and large system (from those shown in Figure 3a. Parts b and c of Figure 3 show
scattered starting configuration) at 1.036, 1.0, and 2.0 ns, that the aggregates formed in the large systems have their
respectively. In all these snapshots, the CO2 molecules aqueous cores considerably exposed to the surrounding
are intentionally hidden so that the surfactant aggregates solvent. This difference between the small and the large
can be seen clearly. All the figures show the presence of system could be attributed to the higher water-to-
surfactant aggregates in the system that have the ap- surfactant molar ratio, Wo, in the large system (Wo ∼ 35
1778 Langmuir, Vol. 17, No. 5, 2001 Salaniwal et al.

for the large system; cf. Wo ∼ 4 for the small system). Table 5. Structural Properties of the Largest Aggregate
Figure 3b shows that during the time interval of 1.0 ns, Formed in the Large Simulation System (Aggregated
the aggregated starting configuration of the large system Starting Configuration)
evolves into a configuration with a large aggregate this work SANS experiments
consisting of most of the surfactant and water molecules water 916 ( 10
(the few remaining surfactant and water molecules were surfactant 26
distributed between three smaller aggregates). On the Wo 35.2 35
other hand, the scattered starting configuration of the Rg (Å) 15.4 ( 0.2 20.5 ( 1.0
large system evolves into a configuration consisting of Imax/Imin 1.3 ( 0.1
five small-sized surfactant aggregates. Since the overall η 0.12 ( 0.05
As (Å2) 115 ( 3 140
composition of the large system corresponds to one
average-sized aggregate observed by Eastoe et al.,35 the
expected final state in the large system is a single the values of these parameters reflects the effect of the
surfactant aggregate consisting of all the surfactant and spatial or temporal fluctuations in the shape of the aqueous
water molecules. Thus, it can be argued that Figure 3b core. Together these two parameters can reveal details of
is much closer to the expected final state compared to the actual shape (spherical, cylindrical, or disklike) of the
Figure 3c. On the basis of the times corresponding to these aggregates.
configurations, it can also be stated that the aggregated Table 5 provides the structural properties of the largest
starting configuration of the large system approaches the aggregate in Figure 3b along with the experimental values
expected final state in a shorter time compared to the measured by Eastoe et al.35 via SANS experiments. The
scattered starting configuration. In our next paper,45 we radius of gyration (Rg) of the aggregate formed in this
demonstrate that the scattered starting configuration of simulation study is 15.4 ( 0.3 Å. This is in reasonable
the large system follows a realistic trajectory and should agreement with the experimentally reported value of 20.5
also reach the expected final state in ∼200 ns. In essence, ( 1.0 Å, given that the time for which the simulation has
using the aggregated starting configuration as the initial been carried may be considerably shorter than that
condition for the large system simply reduces the time required to achieve a final equilibrium state. The shape
needed to approach the expected final state thereby parameters for the aggregate are Imax/Imin ) 1.2 ( 0.1 and
allowing the study of the structural properties of surfactant η ) 0.1 ( 0.05 (no experimental data available), which
aggregates in a reasonable amount of computer time. Thus, indicates that the aggregate is nearly spherical with small
the structural properties of the largest aggregate in Figure fluctuations. This is expected because of the high Wo in
3b can be quantitatively compared to the experimental this aggregate. The high value of Wo provides the aggregate
results of Eastoe et al.35 with an aqueous core of considerable volume (and surface
Aggregate Size and Shape. The general structural area) thereby facilitating a low packing density of surf-
properties of the RM-like aggregates formed in the actant molecules around the aqueous core. To characterize
simulations presented above can be characterized by the the shape of the aggregates formed in the experimental
shape and size of their aqueous cores. A good measure of investigation, Eastoe et al.35 determined the surface area
the size is the radius of gyration (Rg), which is a measure of the aqueous core per surfactant molecules (As), given
of the distribution of the mass of atomic groups or by the equation
molecules that constitute the aqueous core relative to its
center of mass. Since the core is essentially made up of As ) 4πRg2/Ns (8)
water molecules, sodium ions, and anionic headgroups,
the mass of only these groups (tail groups are not where Ns is the aggregation number (number of surfactant
considered) is considered in the estimation of Rg. The Rg molecules in the aggregate). In essence, this quantity is
of the aqueous core (to be compared with the core radius a measure of the packing density of the surfactant
fitted to the SANS data) can be calculated by the equation molecules in the aggregate and, thus, can be related to
the shape of its aqueous core. The value of As for the
Rg2 ) ∑i mi(ri - ro)2/∑i mi (6) aggregate estimated from the simulation is 115 Å2, which
is in approximate agreement with the experimentally
reported value of 140 Å2.35
where mi is the mass of atom i located at a radial distance The structural characteristics of the large aggregate in
ri from the center of mass of the aggregate core (ro). The Figure 3b merit some additional discussion. First, the
shape of the aqueous core can be quantified using two question must be addressed, “Is this aggregate structurally
different parameters, namely, the ratio of the largest to similar to the reverse micelles observed in the SANS
the smallest principal moments of inertia (Imax/Imin) and experiments?” The only way to answer this is as follows:
the eccentricity (η). The eccentricity is calculated by the (a) The models used have been tested and have yielded
equation results in reasonable agreement with the relevant ex-
perimental data in the literature, and (b) as shown in
Imin Table 5, the simulation results are in reasonable agree-
η)1- (7) ment with all of the experimentally measured structural
Iavg
properties of reverse micelles. A second issue then arises,
the large aggregate in Figure 3b does not look like the
where Iavg is the average of all the three components of the common picture of a reverse micelle; a large portion of the
principal moment of inertia. Both these parameters aqueous core is directly exposed to the CO2, and the
provide a measure of the deviation from a perfectly surfactant molecules are not uniformly distributed on the
spherical object (Imax/Imin ) 1, η ) 0). The uncertainty in interface. In fact, the experimental result (As ) 140 Å2)
indicates that even more of the aqueous core should be
(45) Salaniwal, S.; Cui, S. T.; Cochran, H. D.; Cummings, P. T. Self- exposed, and even though on average the surfactant
assembly in a dichain surfactant/water/carbon dioxide system via
molecular simulation. 2. Aggregation dynamics. Langmuir 2001, 17, distribution on the interface should be uniform, this does
1784. not imply that the instantaneous distribution should be
Dichain Ternary System Langmuir, Vol. 17, No. 5, 2001 1779

Table 6. Structural Properties of Surfactant Aggregates


Formed in the Small System (Aggregated Starting
Configuration)
aggregate no. 1 aggregate no. 2 aggregate no. 3
water 26 57 49
surfactant 10 9 11
Wo 2.6 6.3 4.4
Rg (Å) 6.2 ( 0.1 7.2 ( 0.2 6.9 ( 0.1
Imax/Imin 1.8 ( 0.1 1.7 ( 0.3 1.5 ( 0.1
η 0.28 ( 0.02 0.27 ( 0.08 0.13 ( 0.03

uniform. Perhaps the common picture of a reverse micelle


should be reconsidered.
For surfactant aggregates formed in the small system
(Figure 3a), the structural properties are influenced by
the low water-to-surfactant ratio (Wo ∼ 4) in this system.
Table 6 lists the Rg, Imax/Imin, and η values for the three
aggregates shown in Figure 3a. The Rg values for all the
three aggregates lie within the range of 6.2-7.2 Å
indicating nearly equal-sized aggregates. The Imax/Imin
values for all the three aggregates are significantly greater
than unity and the η values are significantly greater than
zero. Thus, the aggregates formed are highly nonspherical
compared to those formed in the large system. Clearly,
the low Wo in this system results in aggregates with much
smaller aqueous cores (smaller Rg), thereby resulting in
significantly higher packing density of surfactant mol-
ecules. Another factor that may also contribute to the high
nonsphericity in these aggregates is the ineffective
dielectric screening of ionic headgroups. The simulations
showed that most of the water molecules (∼94%) in the
small system lie within the hydration shell of anionic Figure 4. Structural properties of the aggregate formed in
headgroups and Na+ ions (in contrast to ∼62% for the the large system with aggregated starting configuration: (a)
large system) and thus very few water molecules were water and carbon dioxide; (b) microstructure of surfactant
available to provide the dielectric screening of ionic molecule.
charges. As a result the interactions between ionic charges
lead to the distortion of the interface. Similar results radial distance r from a reference point. The center of
regarding the nonsphericity of aggregates with low water- mass of the aqueous core represents a suitable reference
to-surfactant ratio (Wo ) 2) were obtained by Brown and point.
Clarke17 for a model reversed micelle. Interestingly, the Figure 4 shows the microscopic details of the largest
standard deviations in the values of Imax/Imin and η for the aggregate formed in the large system with aggregated
aggregates of Figure 3a, as shown in Table 6, are quite starting configuration (Figure 3b). Figure 4a shows the
low. As mentioned earlier, these deviations are a measure density profile of water and CO2 (solvent) molecules. The
of the fluctuations in the shape of the aggregates. The low density of water in the aqueous core is ∼1.0 g/cm3
value of these deviations suggests that the temporal indicating the presence of bulklike water. A broad (∼10-
fluctuations in the shape of these aggregates are quite 15 Å) interfacial region is observed, where the water
low. The fluctuations observed in the time evolution of density decreases while the CO2 density increases as the
Imax/Imin and η are more prominent in aggregate no. 2 distance from the center of mass of the increases. The
than in the other aggregates. This can be attributed to a solvent density approaches its bulk density of ∼0.85 g/cm3
slightly higher water-to-surfactant ratio in this particular outside the interfacial region. Negligible penetration of
aggregate (Wo ∼ 6) compared to the other two aggregates. the solvent molecule into the aqueous core of the aggregate
Aggregate Microstructure. Having established that is observed. This is consistent with the high ionic strength,
the structure of the largest aggregate formed in the large I ∼ 1.5, due to the ionic headgroups and the counterions
system (Figure 3a) with aggregated starting configuration in the small aqueous core. Figure 4b plots F(r) as a function
is in reasonable agreement with the experimental findings, of the radial distance (r) for the various groups of the
it is logical to explore the microstructure of this aggregate. surfactant molecules. The distribution of the headgroups
Molecular simulations provide an effective tool to obtain shows a peak at ∼17.2 Å indicating the presence of
this information. A primary measure of the microscopic headgroups at the surface of the aqueous core. The
structure is the singlet distribution function (singlet radial distribution for the Na+ ions is quite similar to that for
density) profiles of the various atomic groups or molecules the headgroups implying that the Na+ ions also reside at
constituting the aggregate. As the name suggests, the the surface of the aqueous core and may be ion-paired to
singlet distribution function is a time-averaged measure the anionic headgroups. The density distribution of the
of the distribution of an atom or molecule from a reference two tails suggests that the tails form the corona of the
point. It is calculated by the equation aggregate. The distribution of the perfluoroalkane tail is
shifted slightly away from the center of mass and is an
F(r) ) 〈N(r)〉/4πr2∆r (9) indication of its CO2-philic behavior.
Figure 5 shows the microscopic details of the aggregates
where 〈N(r)〉 is the time average of the number of an atom formed in the small system (Figure 3a). The figures show
or molecule that is found in a shell of thickness ∆r at a that the core of the aggregates is essentially made up of
1780 Langmuir, Vol. 17, No. 5, 2001 Salaniwal et al.

Figure 6. Time evolution of the % trans bonds (extent of fully


extended tails) for the perfluoroalkane tail and the alkane tail
of the C7-tailed hybrid surfactant.

Figure 5. Structural properties of the aggregate formed in


the small system with aggregated starting configuration: (a)
water and carbon dioxide; (b) microstructure of surfactant
molecule.

water molecules and the sodium ions. The headgroups lie Figure 7. Radial distribution of the functional groups of the
at the surface of the aqueous cores while the tails form perfluoroalkane tail and the alkane tail for the C7-tailed hybrid
the corona surrounding the cores. The nonzero radial surfactant.
density of the headgroups within the core (r < 6.2 Å) is
not due to the penetration of the headgroups but simply collapsed conformations in vacuo. This is an indication of
an artifact of the highly nonspherical shape of these the fact that the perfluoroalkane tails are more CO2-philic
aggregates (Table 6). Also, the presence of numerous peaks than the alkane tails.
and troughs in the radial density distribution of the sodium More detailed information on the spatial conformation
ions and anionic headgroups is a manifestation of the rigid of the two tails can be obtained by calculating the
structure of the aqueous core. The rigid structure of the distribution of the radial position of each segment
aqueous core also manifests itself in the considerably high (functional group) of the tails from the CH group, as shown
water density observed in Figure 5a. The radial density in Figure 7. The common feature of the two tails evident
distributions of the two tails are quite similar except for from the figure is the progressive broadening of the
a slight shift away from the center of mass observed for distribution for groups farther away from the headgroup.
the perfluoroalkane tail. As mentioned earlier, this is due The broadening is relatively larger for the alkane tail
to the CO2-philic nature of this tail. compared to that for the perfluoroalkane tail because of
Tail Conformation. A good measure of the CO2-philic the relative stiffness of the perfluoroalkane tail. Also, the
behavior of the two different tail groups is the degree to distributions of the first two segments are the same for
which these tails are fully extended. Figure 6 plots the the two tails. The broadening is attributed to the fact that
fraction of torsional angles in each tail that are in trans the farther away the segment is from the headgroup the
conformation (fully stretched) as a function of time for the more freedom it has to move. This mobility is due to the
small system with aggregated starting configuration. ability of torsional angles associated with these segments
Similar results were obtained for the large systems as to switch back and forth between trans and gauche
well. The perfluoroalkane tails have an average of 90 ( conformations. For the alkane tail the distribution of
2% trans conformation compared to 74 ( 5% for the alkane farther segments exhibits two distinct peaks reflecting a
tails. This implies that the perfluoroalkane tails are 1 to 3 ratio of gauche-to-trans conformation of the torsional
approximately 16% more extended than the alkane tails. angles associated with these segments. Thus, the per-
Interestingly, the behavior of the two tails is quite opposite fluoroalkane tails are more extended and less mobile than
in vacuo. The perfluoroalkane tail has a trans conformation the alkane tails.
of 80% in vacuo compared to 89% for the alkane tail. Thus, Structure of the Aqueous Core. As shown earlier,
in the presence of carbon dioxide as a solvent medium, the surfactant aggregates formed in this simulation study
the perfluoroalkane tails assume more extended confor- are characterized by the presence of aqueous cores. These
mations in vacuo, and the alkane tails assume more aqueous cores can provide a medium for solubilizing polar
Dichain Ternary System Langmuir, Vol. 17, No. 5, 2001 1781

Figure 8. Radial distribution function (rdf) for (a) Na-Na Figure 10. Interparticle-dipole vector angle angular distri-
pair and (b) Na-headgroup pair. bution function (adf) for water molecules lying within the
primary hydration shell of the ions.

hydration peak exists at 4.4 Å. Thus, there exist two


distinct hydration shells for the Na+ ion implying a strong
structuring of water molecules around the reference ion.
The average number of water molecules in the first
(primary) hydration ions is ∼4. The rdf for the X-W shows
a less prominent first peak at 4.1 Å and weak second peak
at 6.0 Å. The presence of a less prominent and broad first
peak in the X-W rdf may be attributed to the “correlation
hole effect”46 and is related to the tetrahedral geometry
of the surfactant headgroups. Since each headgroup has
three negatively charged oxygen atoms at a fixed distance
from each, a number of possibilities exist in which a
particular water molecule can exist around the reference
ion. As a result, the pair correlation function involves
averaging over distances that involve the van der Waal
diameter (σ) as well as the length scale corresponding to
Figure 9. Radial distribution function (rdf) for the (a) Na- fixed interatomic distances. Thus, the peak in the X-W
water pair and (b) headgroup-water pair. rdf is broader and lower. The average number of water
molecules in the first hydration shell of the ionic headgroup
and/or polymeric substances in CO2. Thus, understanding is ∼9, implying a higher degree of hydration of the
the detailed structure of water in these aqueous cores is headgroup compared to the Na+ ion because it is larger
of fundamental importance in understanding the solu- than the Na+ ion and has more charges.
bilization characteristics and behavior of solutes. An important characteristic of the water molecules in
To get an insight into the detailed structural charac- the aqueous core is the extent of their hydrogen bonding.
teristics of the aqueous cores, Figure 8 plots the radial In this study, a hydrogen bond is defined to exist if any
distribution function (rdf) for Na-Na and Na-X pairs (X one of the O-H bonds of a particular water molecule is
denotes the anionic headgroups) in the largest aggregate collinear with an oxygen atom of a neighboring water
shown in Figure 3b. As discussed earlier, the structural molecule. To analyze the orientation of the water mol-
properties of this surfactant aggregate are closest to the ecules, angular distribution functions (adf) are used.
RMs observed by Eastoe et al.35 via SANS experiments. Figure 10 shows the adf’s of the interparticle-dipole vector
Since the reference ion is located at the interface (Figure for water molecules belonging to the primary hydration
4b), its environment is anisotropic, even at small distances. shell of the Na+ ion and sulfate headgroup. The adf for the
As a result, the rdf decays to zero with increasing radial Na+ ion shows a monotonic decrease along the angular
distance instead of approaching unity as observed for range indicating the alignment of water in such a way
isotropic systems. Figure 8 shows that Na-X rdf has a that the dipole of water molecules is collinear with the
prominent first peak at 2.45 Å and a minimum at 3.2 Å. Na+ ions. This is due to the strong electrostatic interaction
The secondary peaks in the rdf at 3.9 and 4.5 Å are due between Na+ ion and the negatively charged oxygen atom
to the remaining two negatively charged oxygens on the of water molecules. The adf for the headgroup shows a
headgroup. There are about ∼0.75 Na+ ions within 3.2 Å peak at cos φ ∼ 0.6 (φ ∼ 53°) which is due to the fact that
of each headgroup, implying that most of the Na+ ions are the anionic headgroup interacts with positively charged
ion paired to the surfactant headgroup. The Na-Na rdf hydrogen atoms of the water molecules thereby aligning
shows a peak at 3.7 Å, a broad minimum at 5.0 Å, and a the water dipole at the angle of ∼53°. This clearly shows
second peak at 6.3 Å, indicating a weaker ion-ion that the strong electric fields from the Na+ ion and the
correlation, as expected. The location of the first peak at anionic headgroup tend to orient the water dipole vectors
3.7 Å is a manifestation of the fact that the Na+ ions are in their primary hydration shells. As expected, the
primarily located at the interfacial region. Figure 9 shows orientation effects are relatively stronger for the Na+ ions
the rdf for the Na-W (W denotes water molecules) and
X-W pairs. The rdf for the Na-W shows a distinct (46) Chandler, D. Introduction to Modern Statistical Mechanics;
hydration shell with a maximum at 2.5 Å. A second Oxford University Press: New York, 1987.
1782 Langmuir, Vol. 17, No. 5, 2001 Salaniwal et al.

distance of 3.3-6.0 Å from the reference ion (secondary


hydration shell) show very little orientation compared to
those in the primary hydration shell. The water molecules
lying at distances larger than 6.0 Å show negligible
orientation effects. Similar results were found for the water
molecules around the sulfate headgroups. Thus, on the
basis of this analysis the water molecules in the aqueous
core can be divided into two distinct populations, the
interfacial water which lies in the primary hydration shell
and bulklike water which lies outside the primary
hydration shell. Figure 11 shows the orientation of water
molecules present in the bulklike region of the aqueous
core. Figure 11a shows the adf for the dipole-dipole vector
angle for the water molecules. The figure shows most of
the water pairs have a dipole-dipole vector angle align-
ment greater than φ ∼ 0°. Thus, water molecules exhibit
a tendency to form hydrogen bonds (a hydrogen bond is
defined if one of the OH bond of a water molecule is
collinear with the oxygen atom of a neighboring water
molecule) in the bulklike region, as expected. Figure 11b
shows the adf for the interparticle-dipole vector angle.
The adf shows a distinct peak at cos φ ∼ 0.6 (φ ∼ 53°).
Figure 11c shows the adf for the interparticle-bond vector
angle and shows two peaks, one at cos φ ∼ -0.3 (φ ∼ 107°)
and the other at cos φ ∼ 1 (φ ∼ 0°). The two peaks
correspond to the two bond vectors in a water molecule.
The positions of the two peaks clearly indicate the
collinearity of OH bond on the water molecules with the
oxygen atom of the reference molecule.

Conclusions
Molecular dynamics simulations of two system sizes
(small system and large system) are presented. The small
system is used for exploratory calculations, while the
composition and the solvent condition of the large system
mimic the experimental system studied by Eastoe et al.,35
thus enabling quantitative comparison with experiments.
The simulations showed that the shape and size of the
aggregate formed in the large system (aggregated starting
configuration) are in reasonably good agreement with
experimental values. In contrast, the properties of the
aggregates formed in the small system are influenced by
the low water-to-surfactant ratio in this system. The
aggregates formed in the small system are quite small
and considerably nonspherical compared to those formed
in the large system.
In the present study, advantage was taken of the MD
simulation technique to explore the microstructure of the
aggregates formed. The radial density distribution of the
various molecules constituting the aggregate showed that
Figure 11. Angular distribution functions (adf’s) for (a) dipole- the aqueous core essentially consisted of water molecules
dipole vector angle, (b) interparticle-dipole vector angle, and
(c) interparticle-bond vector angle for water molecules within and Na+ ions. The sulfate headgroups are primarily
the bulklike region of the aqueous core. situated at the surface of the aqueous core. The radial
density distribution of the two tails indicated the presence
than the headgroups due to the delocalization of negative of a corona surrounding the aqueous core and thereby
charges on the anionic headgroup. The adf’s for the effectively shielding it from the nonpolar solvent. Neg-
interparticle-bond vector angle for water molecules (not ligible penetration of the solvent molecules in the aqueous
shown for brevity) residing in the primary hydration shell core is observed. For the aggregate in the large system,
also exhibit strong orientation effects of the two ionic the density of water near the center of the core is ∼1.0
species. The adf for the Na+ ion shows a peak at cos φ ∼ g/cm3, indicating the presence of bulklike water. A fairly
-0.5 (φ ∼ 120°), while the adf for the anionic headgroup broad (10-15 Å) interfacial region is observed. The adf’s
shows a peak at cos φ ∼ 1 (φ ∼ 0°). This is another indication of the water molecules in aqueous core revealed that the
that the water molecules directly associated to the two water molecules in the interfacial region are strongly
ionic species are oriented in response to the strong electric oriented in response to the electric fields of the anionic
fields of the ions. headgroups and Na+ ions, while those near the center of
For water molecules lying deeper within the core, beyond the core exhibit a hydrogen-bonded network. The tail
the primary hydration shell, decreases in the orientation conformations indicate that the perfluoroalkane tails are
effects are noted. The water molecules lying within the more extended than the alkane tails, which is opposite to
Dichain Ternary System Langmuir, Vol. 17, No. 5, 2001 1783

that observed in vacuo. Thus, the perfluoroalkane tails National Laboratory is managed by Lockheed Martin
exhibit more CO2-philic behavior than the alkane tails. Energy Research Corp. for the Department of Energy
under Contract Number DE-AC05-96OR2246. We are also
Acknowledgment. We are pleased to acknowledge pleased to acknowledge the allocation of computational
support of this work by the Chemical and Thermal Systems time on the Cray T3E at the National Energy Research
Division of the National Science Foundation under Grant Scientific Computing Center at Lawrence Berkeley Na-
Number CTS-9613555 at the University of Tennessee and tional Laboratory.
by the Chemical Sciences Division of the Department of
Energy at Oak Ridge National Laboratory. Oak Ridge LA000554F

You might also like