Você está na página 1de 10

Article

pubs.acs.org/Langmuir

Nature of CTAB/Water/Chloroform Reverse Micelles at Above- and


Subzero Temperatures Studied by NMR and Molecular Dynamics
Simulations
Lubica Klcova, Eva Muchova, Peter Sebej, Petr Slavcek, and Petr Klan*,

Department of Chemistry and RECETOX, Masaryk University, Kamenice 5, 625 00 Brno, Czech Republic
Department of Physical Chemistry, University of Chemistry and Technology, Technicka 5, 16628 Prague 6, Czech Republic

S Supporting Information
*

ABSTRACT: The nature and stability of cetyltrimethylammonium bromide (CTAB) reverse micelles in chloroform formed above the critical
micellar concentration at above- and subzero temperatures were examined
by NMR and molecular dynamics simulations. The experiments showed
that the supercooled micellar water pool becomes unstable upon cooling to
relatively high temperatures (253 K), and smaller micelles are formed.
Upon freezing at lower temperatures (233 K), micelles become completely
frozen and remain intact in the solution. With an average hydrodynamic
radius of approximately 1.3 nm, we estimate that the water pool contains
approximately 50 water molecules, which is well below the onset of ice
crystal formation. To support the experimental results, molecular dynamics
simulations were used to model the structure of CTAB/water/chloroform reverse micelles of dierent sizes. The MD simulations
show that the reverse micelles contain a water pool with bromide anions residing on its surface and their shape is nonspherical,
especially in the case of larger water pools. Upon fast freezing, the mobility of the water molecules is suppressed, and the pool
becomes more spherical.

INTRODUCTION
Water in all forms is one of the most abundant molecules on
Earth and its role is widely appreciated, yet there are still
surprising gaps in our knowledge on its behavior. One of the
questions attracting continuously wide interest is the nature of
liquid water below the normal melting point. Bulk water can
remain in the liquid state below its melting point, but it cannot
be supercooled below the homogeneous nucleation temperature (at around TH = 235 K).1 However, the liquid state can
be maintained even below this temperature in nite-size water
particles.2 This nanoconned, deeply supercooled water
exhibits surprising properties. For example, it has been
suggested that deeply supercooled liquid water has a density
minimum.3
Conned water can be found in protein pockets or in
synthetic nanopores. It has been hypothesized that water
compartmentalization represents a mechanism for the cryogenic protection of organisms.4 Finite-size water particles, such
as nanometer-sized water aerosols, were shown to be critical for
the nucleation processes in the atmosphere.5
Reverse micelles, self-organized assemblies of amphiphilic
molecules in nonpolar solvents, serve as a useful model for
conned water in the condensed phase.6 The polar heads of
amphiphiles are oriented toward the water cores and their
hydrophobic chains form the outer shells. The nature and
behavior of reverse micelles at ambient temperatures have been
a subject of many investigations. Recently, we have validated
2015 American Chemical Society

Eickes association model for CTAB/water/chloroform reverse


micelles, according to which micelles are formed by a structural
reorganization of linear associates within the apparent critical
micelle concentration (cmc).7
Various factors can aect the lowest temperature to which
water can be cooled before freezing to ice. The volume and
spatial connement8 as well as the presence of charged
interfaces 9 are among the most important parameters
determining the supercooling/crystallization process.10 Only a
few studies have been performed to study AOT reverse micelles
or proteins at subzero temperatures. Flynn and co-workers have
reported that, at relatively low subzero temperatures, a
supercooled water core of anionic AOT reverse micelles stays
liquid and micelles become unstable,11 which is manifested by
water shedding, that is the loss of water from the core until
equilibrium is established.11,12 It was hypothesized that the
entropically favorable encapsulation does not always contribute
enough to the Gibbs energy to keep reverse micelles intact.11
These authors have also demonstrated that the water loading
(w; the ratio of the molar concentrations cwater(core) and
camphiphile(micelle)) of reverse micelles is inuenced by the ionic
strength of the aqueous phase. Kevan and co-workers have
reported that water shedding can be prevented by shock
Received: May 14, 2015
Revised: June 29, 2015
Published: July 21, 2015
8284

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Article

Langmuir
cooling of a micellar system to 77 K.13 Nucci and Vanderkooi
have shown that water cores of a reverse AOT micellar system
in n-pentane can freeze at 238 K independently on the waterto-surfactant molar ratio (x).14 Bright and co-workers have
studied AOT reverse micelles in n-heptane using uorescence
probes at subzero temperatures.12 Their steady-state uorescence emission and uorescence anisotropy studies indicated
that freezing occurs within the water pool at temperatures
between 263 and 213 K and is aected by x. It has also been
reported by Dokter and co-workers that fast freezing to 180 K
does not aect the structure of reverse AOT micelles for water
loadings smaller than 3.5.15 The structure of the cores of these
small micelles, containing approximately 150 water molecules,
resembled an amorphous form of ice. Recently, Suzuki and Yui
demonstrated that loss of conned water in AOT reverse
micelles upon freezing can be prevented by a combination of
rapid cooling and a small sample cell size that allows for the
crystallization of water pools with larger radii (over 2.1 nm).16
Attenuated total reection infrared spectroscopy (ATR-IR)
measurements revealed that the frozen pool exhibits features
similar to the spectrum of metastable cubic ice (Ic), and also
that an icewater coexistence phase is formed during melting.17
The interactions of conned micellar water with the ionic
amphiphile headgroups and the presence of counterions aect
the structure of the micellar cores and their dynamics.18 For
example, interfacial water molecules in AOT reverse micelles
form rigid hydrogen bonds with the anionic sulfonate
headgroups,19 whereas hydrogen bonding of interfacial water
in cetyltrimethylammonium bromide (CTAB) reverse micelles
is less directional due to the large polarizable cloud of the
bromide ion.20 The ammonium headgroup has a negligible
primary hydration capacity, as observed in didodecyldimethylammonium bromide-water systems where no interfacial
water was detected.21 The behavior of cationic reverse micelles
at subzero temperatures has hardly been studied yet. Only one
work reports that CTAB/n-hexanol/alkane reverse micelles are
stable above 263 K.22
This work is a follow-up of our previous study,7 in which we
determined the boundary conditions of the stability of CTAB
reverse micelles in chloroform. Here, we investigate the nature,
size and dynamics of the water pool of such micelles in the
temperature range of 303233 K using 1H NMR spectroscopy.
The NMR experiments are supported by molecular dynamics
(MD) simulations, which already proved to be useful for
investigating the structure of reverse micelle at ambient
conditions;2326 yet only little work has been done for reverse
micelles at subzero temperatures. Here, we use MD simulations
to reveal how the structural and dynamical properties of the
CTAB/water/chloroform micelles vary with temperature.

whole time of equilibration). For experiments lasting over 1 day, the


NMR tubes were sealed.
1
H NMR spectra were obtained on 300 and 500 MHz
spectrometers at dierent temperatures. The initial and nal spectra
of an individual sample were always acquired at 30 C. The
temperature of a sample was equilibrated for at least 3 min after the
given temperature in the NMR probe was set.
1
H NMR was also used to evaluate the diusion coecients of
reverse micelles in chloroform. The 2D DOSY NMR experiments
were carried out on a 500 MHz spectrometer at 303 K, employing a
PABBO probe-head equipped with the z-gradients. The experiments
consisted of several measurements of a pulsed gradient stimulated
echo (PGSTE) sequence with a longitudinal eddy current delay, two
bipolar gradient pulse pairs of smoothed-square shape and two
additional spoiling gradients, placed in the delay, and the Z-period of
the diusion time in the pulse sequence. The resulting data were
obtained as a set of the 1D measurements diering in the gradient
strength value, which were linearly modied from 2 to 95% of the
maximum gradient strength (a gradient calibration constant) in 32
gradient steps, causing a gradual attenuation of the signal. The
longitudinal relaxation time was 18.8 s for water in the case of gradient
calibration experiments and to 3 s for CTAB in the case of diusion
coecient measurements. The gradient system was calibrated by
measuring the diusion coecient of H2O in D2O at 298 K using a
previously reported value of 1.872 109 m2 s1.27 The corrected
gradient calibration constant was determined before each measurement and was found to be in the range of 5.01 to 5.43 G cm1. The
intensity and/or peak area decay curves for CTAB protons with
chemical shifts of 3.49, 1.74, 1.25, or 0.88 ppm were tted
independently. The diusion coecient was calculated using the
equation: S(G)/S(0) = exp[22G2D( /3 /2)],28 where
S(G) is the signal at a gradient amplitude G, S(0) is the signal at zero
gradient, is the gyromagnetic ratio of a proton, is a duration of the
magnetic eld gradient pulse, is the diusion time, and is the time
between the two gradient pulses in the bipolar gradient pair. The
diusion delay was set to 50 ms. The gradient pulse length of /2
was set in the range of 1.62 ms, whereas the spoil gradient pulse
length was set in the range of 0.61 ms. The mean self-diusion
coecient of the reverse micelles was determined as an average of 19
values of the diusion coecients measured for 5 independent
samples.
The Size of Reverse Micelles. In order to obtain qualitative
information about the size of reverse micelles, the StokesEinstein
equation was used, and the hydrodynamic radius RH was calculated
according to RH = kBT/6D, where is the solvent viscosity. A rather
low volume fraction of the dispersed phase (0.0425) was obtained
by the equation = (Vwater+ nCTABvCTAB)/Vtotal, where Vwater is the
volume of water, nCTAB is the number of moles of CTAB, and vCTAB is
the molar volume of CTAB (363 mL mol1 at room temperature29). It
was not necessary to make a correction for collisions.30 The value of
RH is an upper limit because the obstruction eect has not been taken
into account; it is very small (within the measurement error of the
diusion coecient) for the volume fraction used.
For the present purpose, water assemblies in the micellar core were
assumed to be spherical, monodispersed, and separated from the
homogeneous organic solvent by a monolayer of surfactant molecules,
and all water in the sample was uniformly distributed inside the reverse
NMR
NMR
micelles (for wobs
water, core = cwater, core/cCTAB = x = 3.4). These assumptions
allowed us to calculate the aggregation number Ns (the number of
molecules present in a micelle once the cmc is reached) of the
prepared reverse micelles together with the amount of water molecules
Nw inside their water pools. The hydrodynamic radius RH of a reverse
micelle (the volume of one CTAB molecule is Vs= 0.469 nm3,31 and
the volume of one water molecule is Vw = 0.03 nm332) were calculated
from the following system of two linear equations and two variables,
Ns and Nw: VRM = Vs + Vw = NsVs + NwVw, where VRM = (4/3)RH3 is
the volume of one reverse micelle, and Ns = Nw/w. The amount of
water molecules Nw was then expressed as Nw = (4/3)RH3w/(Vs +
Vww).

MATERIALS AND METHODS

Materials. Cetyltrimethylammonium bromide (CTAB; >99.0%)


and chloroform-d (99.8%) were used as purchased. Chloroform-d was
stored in amber bottles over ame-dried molecular sieves (3 ) with a
piece of silver foil as a stabilizer.
1
H NMR and Diusion Coecient Measurements. The
chloroform-d solutions were prepared by direct weighing of CTAB
(cCTAB = 0.1 mol dm3) into NMR tubes. Water was subsequently
added to this solution to adjust the water-to-surfactant molar ratio, x =
analytical
to 3.4.7 The mixture was then vigorously agitated. A
canalytical
water /cCTAB
sealed capillary tube containing a chloroform-d solution of
tetramethylsilane or dichloromethane was added as an internal
standard. All samples were equilibrated for 14 h prior to the
measurements (the NMR signal of water remained unchanged during
8285

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Article

Langmuir

intramicellar exchange is faster50 than the time resolution of the


NMR spectroscopy.
For x > 4, the phase separation is observed during the sample
preparation before equilibrium is reached, which is manifested
by the presence of a separated aqueous layer and the signal of
bulklike water at water,bulk 4.7 ppm.51 In order to evaluate
time needed for equilibration in an unperturbed mixture, a
sample (cCTAB = 0.1 mol L1 in dry chloroform-d) with 4 L of
water initially added to the surface of a CTAB solution was
prepared without shaking. Subsequently, the complete
incorporation of water into the reverse CTAB micelles was
accomplished in approximately 200 h. The incorporation rate
constants were calculated from the water loading wobs
water,core
values obtained from the relative water and CTAB concenNMR
NMR 7
obs
trations, wobs
water,core = cwater,core/cCTAB, when wwater,core increased
from 0 to the maximum value of 3.1. Biexponential tting of the
data then provided two rate constants of k = 1.0 104 and 8.1
106 s1 (Figure S2).
In the next step, the temperature of an equilibrated sample
was decreased from 303 to 253 K in 10 min and was kept there
for 2 h. The initial NMR signal of core water at water,core 3.7
ppm at 303 K was slightly shifted upon cooling to 253 K to 4
ppm (Figure 1, black empty circles). The observed water

Using simple geometric considerations, the radius of the water pool


Rw could be expressed as Rw = [3Ns(wVw + VBr)/4]1/3,32 where Rw is
given in nm, and VBr represents the partial molecular volume of the
surfactant counterion Br (VBr = 0.0493 nm3 at 25 C33). The
spherical surface area a, occupied by each cationic surfactant ion at
the water pool surface, was then obtained according to a/nm2 =
4Rw2/Ns.
Molecular Dynamics Simulations. Molecular dynamics simulations were performed with the GROMACS 4.5.3 code.34 A standard
nonpolarizable force eld, comprising point charges and Lennard
Jones potentials, was used to account for intermolecular interactions.
The rigid SPC/E model for water35 was applied. Chloroform was
modeled using the OPLS nonbonded parameters36 ( = 0.3800 nm,
= 0.3269 kJ mol1, and charge 0.420 e for the CH group; = 0.3470
nm, = 1.2560 kJ mol1, and charge 0.140 e for the Cl group). The
force eld of the CTAB molecule was constructed with the
GROMOS-87 with corrections to nonbonded parameters for
atoms,37 with the geometries and charges taken from the DFT/
BLYP/6-31g* calculation using a CHelpG population analysis (see the
Supporting Information, Table S1). Nonbonding parameters for the
bromide ion were taken from the literature.38
The simulations box was prepared as follows. The bromide ions and
CTAB cationic heads were randomly distributed on the surface of ice
particles. The cluster was optimized in the gas phase and subsequently
immersed in a box containing 512 chloroform molecules. The clusters
were then equilibrated for 20 ns.
The productions runs were 20 ns long, using a time-step of 1.5 fs.
Most of the simulations were performed at 300 K; however, the
behavior of the reverse micelles was also studied at 273 and 200 K.
Using the previously described force eld, chloroform remains liquid at
all studied temperatures. Periodic boundary conditions were applied
with the forces being cuto at 19 . The simulation was run in the
NpT ensemble. The Berendsen thermostat with the time constant of
0.5 ps and the Berendsen barostat with the time constant of 1 ps were
used.
It is well-known that the lack of polarization can seriously aect the
structure of dissolved electrolytes, particularly those with a low charge
density.39 A simple remedy for this problem was recently introduced
within the concept of electronic continuum correction (ECC).40 To
account for the screening eect of the electronic polarization, the
charges should be scaled by a factor of 1/el, where el is the optical
part of the dielectric constant. The value of this parameter is almost
equal for most solvents, for example, 1.78 for water (at 298 K) and
2.09 for chloroform.41 In our simulations, the value of 2.09 for
chloroform as a solvent was used. Note that the water charges were
not scaled as their charges are tted to the experiment and have to be
considered as eective charges. Simulations with scaled atomic charges
were frequently used in dierent elds, such as the simulations of ionic
liquids,42 electrolyte solutions,43 or interfaces.44,45 The ECC framework brings a justication for this scaling.46 The MSMS (Maximal
Speed Molecular Surface) code was used to calculate the surface and
volume of the water pool in the reverse micelle.47

Figure 1. Dependence of wobs


water,core (black full circles; right ordinate),
wobs
water,bulk (red full triangles; right ordinate), water,core (black empty
circles; left ordinate), and water,bulk (red empty triangles; left ordinate)
(bottom graph) on temperature (blue solid line; upper graph) and
time (abscissa) for the CTAB/water/chloroform-d solution (canalytical
CTAB =
100 mmol dm3) with the initial x = 3.4. The dashed lines are shown
to guide the eye.

loading wobs
water,core, calculated from the relative water and CTAB
NMR
NMR 7
concentrations, wobs
water,core = cwater,core/cCTAB, dropped concomitantly from 3.0 to 0.5 (Figure 1, black full circles).
Simultaneously, the NMR signal of bulklike water,52 water,bulk
= 5.26.0 ppm, appeared (Figure 1, red empty triangles). The
corresponding bulklike water concentration, wwater,bulk =
NMR
cNMR
water,bulk/cCTAB = 0.26, gradually (in 50 min) disappeared
(Figures S3 and S4), whereas several minor water signals
appeared (Figures S5 and S6). After warming of the sample to
303 K, the chemical shift water,core decreased, and the water
loading wobs
water,core increased to 1.0, which was lower by a factor
of 3 compared to the initial value determined before the
cooling cycle started. When the sample temperature was kept at
303 K without agitation for 50 h, the water loading slowly
rose to wobs
water,core = 2.8 biexponentially with rate constants of the
same magnitude as those measured in the model experiment

RESULTS AND DISCUSSION


Temperature Range: 303253 K. Chloroform-d solutions of CTAB reverse micelles with a concentration of 100
mmol dm3, which is well above the critical micellar
concentration of 40 mM,7 and the water-to-surfactant molar
ratio of x = 3.4, were prepared in NMR tubes, and the solutions
were left to equilibrate at 303 K for 14 h. The corresponding
1
H NMR spectrum contained a single averaged water signal at
water,core 3.7 ppm of the micellar core7 (Figure S1) which is
consistent with NMR observations that water encapsulated in
the reverse micelles exists as a single pseudophase48 below the
phase separation limit of the water-to-surfactant molar ratio of x
4 at 303 K.7 Such an averaging of the water signal may be
caused by the connement of core water molecules49 or their
8286

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Article

Langmuir
described above (4.9 104 s1 and 2.0 106 s1; Figure S2).
These values dier by factors of 4 (the faster component is
faster, the slower component is slower) compared to those of a
model experiment (see above), but they are still in the same
order of magnitude. We hypothesize that the formation of
initial linear associates7 corresponds to the fast process
observed, whereas the subsequent formation of micelles is
characterized by the slow component. It is obvious that such
rate constants should depend on the experimental conditions,
such as the size of phase-separated water droplets, the CTAB
concentration, and temperature. It is thus probable that water
expelled from micelles oats in the form of very small droplets
throughout the sample volume and is incorporated in the
micelles faster than phase-separated water. It is to be noted that
the micelles precipitated upon cooling to 253 K in some
experiments, which was evidenced by disappearance of the
CTAB NMR signal.
At temperatures below 273 K, a supercooled state can be
formed in small volumes of encapsulated water, characterized
by a hydrogen bonding rearrangement and manifested, for
example, by an increase in the water,core values.53 In addition,
due to decreased mobility of water molecules in the
supercooled state, several water populations inside the reverse
micelles may be observable because their lifetime exceeds the
time window of an NMR measurement11 (a specic icewater
coexistence has also been observed during melting of frozen
AOT micelles by IR spectroscopy17). A small shift of water,core
from 3.7 to 4.1 ppm (Figure 1) can be related in part to these
phenomena as this value dropped again upon warming the
sample back to 303 K. A signicant decrease of the observed
water loading wobs
water,core upon cooling to 253 K by a factor of 6
must be attributed to water shedding, the loss of water content
from a reverse micelle under perturbing conditions,11,12 which
could eventually be circumvented at larger water pool radii.17
This water loss was observed already at x = 3.4 which is the
value at which AOT reverse micelles are usually stable while
their core stays liquid (above 218238 K).12,14,15 The stability
of cationic CTAB reverse micelles when the water pool is not
frozen must thus be lower compared to that of the analogous
AOT micelles. The lack of strong headgroup hydration21 can be
related to a lower enthalpic stabilization of the CTAB reverse
micelles at subzero temperatures and the specic electronic9
and steric properties of the CTAB headgroups and the bromide
counterion.20 It is thus possible that the amphiphile heads serve
as a nucleating agent.
Expelled water gradually segregated at 253 K, which was
observed as a bulklike water signal by NMR (Figure 1, red
empty triangles), and changed to free ice crystals oating in the
sample upon freezing, being apparent to the bare eye but
invisible by NMR (loss of the signal). This caused an
irreversible change of the water loading noticeable after the
temperature quickly increased above 273 K. Only a small part
of the encapsulated, probably supercooled water, remained
visible by NMR at 253 K (wobs
water,core = 0.5). Freezing of the
remaining encapsulated water inside the reverse micelle is only
one possibility of the NMR signal loss (water shedding or
precipitation of micelles could be other reasons). If micelles
kept the initial water loading upon cooling to 253 K, a water
pool signal should be observed immediately after the
temperature was increased over 0 C. As only 30% of the
initial core water was recovered upon heating, micelles had to
shed water to become smaller and more stable at the given
conditions. Indeed, the subsequent equilibration required tens

of hours at 303 K, and its kinetics (Figure S2) corresponded to


that of a model equilibration experiment for x = 3.4 (Figure
S2). The experimental temperature of 253 K is still higher than
that of homogeneous nucleation (225232 K) found in the
case AOT reverse micelles possessing a water pool size equal to
1.23.4 nm,3 and comparable to the heterogeneous nucleation
in the presence of a nucleating agent (241266 K) when the
water pool size is 1.22 nm.3 In the next step, we thus
investigated the dynamics of CTAB micelles upon rapid
freezing at lower temperatures.
Temperature Range: 303233 K. The solutions of
reverse CTAB micelles in chloroform-d (canalytical
= 100 mmol
CTAB
dm3, x = 3.4) in NMR tubes equilibrated for 14 h were
exposed to a temperature cycle, in which the temperature
oscillated between 303 and 233 K (Figure 2). The water,core

Figure 2. Dependence of wobs


water,core (black full circles; right ordinate),
the CTAB amount (aCTAB, red full squares; normalized; right ordinate)
and water,core (black empty circles; left ordinate) (bottom graph) as a
function of temperature (blue solid line, upper graph) and time
= 100
(abscissa) for a CTAB/water/chloroform-d solution (canalytical
CTAB
mmol dm3) with the initial x = 3.4. The dashed lines are shown to
guide the eye.

signal at 3.4 ppm (w 3.4; Figure 2, black empty circle) almost


disappeared (w 0) upon relatively fast (3 min) initial cooling
to 233 K. Several very weak signals (w 0.04) appeared, but
the signal of CTAB almost disappeared (Figure S7). Water
loading and the observed chemical shift reverted back to nearly
the initial values of wobs
water,core 3.3 and water,core = 3.2 ppm upon
the rst warming step to 303 K. Subsequently, slightly reduced
values of wobs
water,core 3.0 and water,core = 3.0 ppm (90%
recovery) were observed in the end of the second cooling/
warming cycle at 303 K; the signal of CTAB exhibited the same
temperature change as that observed during the rst cycle.
Disappearance of the NMR signals of both water and CTAB
molecules inside reverse micelles at 233 K, nearly a full and
repetitious recovery of the water loading, as well as all other
important NMR parameters determined upon warming to 303
K and essentially undetected water shedding imply that, in
contrast to cooling at 253 K, micelles were fully frozen at 233 K
and retained their initial size. A lower hydration of the
ammonium headgroup21 and a higher orientational mobility of
counterion-bound water in CTAB reverse micelles20 are
probably the major factors that inuence the rigidity of
CTAB reverse micelles at low temperatures. In addition, a
higher viscosity of chloroform at 233 K might also contribute to
their stabilization. The presence of multiple populations of
water of negligible concentrations that coexisted at 233 K,
8287

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Article

Langmuir

Figure 3. Typical snapshots for a micelle containing (a) 50 water molecules and (b) 216 water molecules. Atoms of water, bromide and CTAB are
represented as spheres whereas chloroform covalent bonds are shown as lines. Nitrogen atoms of CTAB are depicted in blue, carbon atoms in cyan,
and bromide ions in pink. Radial distribution functions for water and the bromide ion with respect to the center of mass of the water pool are
demonstrated for micelles containing either (c) 50 or (d) 216 water molecules at 300 K.

characterized by = 1.85 (in the range typical for dissolved


water in chloroform7) and = 7.8 and 7.9 ppm (possibly a
supercooled aqueous phase), appeared and subsequently
disappeared upon warming. An analogous stabilizing eect at
lower temperatures, enhanced by an increased ionic strength of
the water pool, has been observed in the case of AOT reverse
micelles by Flynn and co-workers,11 or when the water pool
radius of AOT micelles was larger than 1 nm to allow ice
formation.16 Therefore, the size of CTAB micelles was
evaluated for comparative purposes.
Size of Micelles and Properties of the Water Core. The
diusion coecients of reverse-micellar aggregates in equili= 100
brated CTAB/water/chloroform-d solutions (canalytical
CTAB
mmol dm3, x = 3.4) obtained using the 2D DOSY NMR
experiments were found to be (3.58 0.80) 1010 m2 s1 as
an average over the measurements of ve independent samples
at 303 K. The corresponding calculated hydrodynamic radius,
RH, was then (1.26 0.28) nm.54 When the sample underwent
two cooling/warming cycles (233/263 K), RH dropped to (1.06
0.28) nm that is attributed to partial water shedding.
Using geometrical considerations, the initial reverse micelles
contained (50 30) water molecules in the water pool with an
aggregation number of (15 8). The water pool radius is then
Rw = (0.82 0.20) nm and the spherical surface area a = (0.55
0.15) nm2. For example, the magnitude of RH, determined by
SAXS, SANS, and DLS, for typical AOT reverse micelles with w
5 in i-octane or n-heptane is approximately 2.5 nm;55 RH =
0.9 nm was estimated in the case of CTAB/n-hexanol/water
microemulsion with a volume fraction = 0.4 and w = 7.2
using the pulsed-gradient stimulated-echo NMR experiments.30
Thus, RH for our CTAB/water/chloroform system is of the
same magnitude. The higher values of RH reported by us

previously7 were not observed under the experimental


conditions in this work.
Since the mass fraction of CTAB was equal to 2.4 wt %, a
mean Ns value of 15 was calculated, and for the RH deviations,
Ns values of 7 and 27 were found. Similar Ns values in the range
of 520 have been determined for 015 wt % of CTAB in
CTAB/n-hexanol/water reverse micelles.56 The calculated
CTAB headgroup area of 0.55 0.15 nm2 and Rw = 0.82
nm are fairly comparable to the parameters obtained for other
systems investigated. For example, the parameters Ns = 65, Rw
= 1.46 nm, and a = 0.41 nm2 were calculated for CTAB reverse
micelles in chloroform/i-octane (2:1, v/v) in the case of 0.75
mol dm3 CTAB (w = 5) in the presence of a uorescence
probe at 298 K,32 whereas the headgroup area of CTAB in
normal (oil-in-water) micelles was determined to be 0.64
nm2.57
In our system, the values of Rw = 0.82 nm and RH = 1.26 nm
correspond to a surfactant shell of a thickness of 0.44 nm,
which is considerably lower than the length of a fully extended
hydrocarbon chain of 2.2 nm.57 It suggests that the surfactant
chains could be twisted toward the surface of the water pool;
however, the MD simulations do not support this interpretation. If CTAB reverse micelles with a hydrodynamic radius of
1.26 nm contain only 50 water molecules, this number is
insucient to form ice upon freezing;8 instead, water could be
vitried as the viscosity increases.58 It is thus possible that
amorphous frozen water is formed inside the CTAB reverse
micelles upon cooling to 233 K, although the Rw value of 0.82
nm is only slightly lower than that limiting the water pool
crystallization (Rw > 1 nm) to form cubic ice Ic in AOT
micelles.16 We therefore decided to perform molecular
8288

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Article

Langmuir
dynamics simulations to better understand the structural
parameters and dynamics of CTAB micelles in chloroform.
Molecular Dynamics Simulations. We considered
micelles with two sizes of the water pool containing either 50
or 216 water molecules. The smaller system corresponds to the
size determined by our experiment; the larger one was chosen
to investigate how the properties of the system evolve with the
micelle size. The same fraction x = 3.4 as that in the
experiments was kept constant during all the simulations
(specically, the smaller system contained 50 water molecules,
15 bromide ions and 15 CTAB molecules; the larger system
contained 216 water molecules, 63 bromide ions and 63 CTAB
molecules). Typical snapshots for the smallest and largest
micelle are shown in Figure 3a and b. Visual inspection of these
structures demonstrates that some hydrocarbon chains point
toward the bulk solution whereas some chains are not fully
extended and approach the water pool. On average, the alkyl
chains are however rather straight (see also the Supporting
Information, Figure S8). We also observed that particularly the
system with a smaller water pool is not fully encapsulated by
the surfactant. The sparse coverage of the water pool by CTAB
can explain the observed smaller eective length of the micelles.
A critical step in the simulations is a proper choice of the
force eld. This can be exemplied on distributions of the
bromide ions in the water pool. Figure 3c and d show the radial
distribution functions of water and the bromide ions with
respect to the center of mass of the water pool.
We immediately observed a dramatic dierence between the
nonpolarizable force eld and the ECC based calculations. The
radial distribution functions for water as a solvent and the
bromide ion almost coincided for both investigated sizes. This
indicates that the bromide ions are fully dissolved in the water
pool. On the other hand, we clearly observed a surface excess of
the bromide ions for the ECC calculations. In fact, the bromide
ions were fully expelled from the water pool and the observed
nonzero intensity at r distances close to zero resulted from a
nonspherical shape of the water pool. The strong surface
preference for heavier halide solvation upon adding the
polarization eects was previously observed for nite size
clusters and aerosols59 as well as the airwater interfaces.60
More relevant in our context is the increase of an interfacial
halide anion concentration observed at the interface between
the aqueous phase and hydrophobic surfaces.44 In this case, the
eect calculated within the ECC model agreed with explicitly
polarizable calculations; the authors also showed that the ECC
calculations agree with experimental data. The ECC model thus
seems to be suitable for our micellar systems. Based on the
ECC simulations, we conclude that the water pool is formed by
a compact hydrogen-bonded network of the water molecules.
The bromide ions are placed on the surface of the cluster,
surrounded by the cetyltrimethylammonium heads and chloroform.
The radial distribution functions were also consistent with
the experimentally determined radius of the micelles (Rw =
(0.82 0.20) nm that corresponds to the reduced water
density for the smaller cluster). In the subsequent step, we
asked what the structure of the reverse micelles is in particular,
what the shape of the water pool is. The sphericity of the water
pool is visualized here by displaying semiaxes a, b, and c that are
related to the principal moments of inertia:
I1 =

I2 =

1
M (a 2 + c 2 )
5

I3 =

1
M (b 2 + c 2 )
5

The three semiaxes would be identical for a spherical object.


Another quantitative criterion is based on calculating the
solvent-excluded surface (SES) and the corresponding volume.
We can dene a ratio s as

s=

SSES
Ssphere

where SSES is the SES and Ssphere is the surface calculated from
the SES-related volume considering a spherical shape of the
object. This ratio decreases for a more spherical object, with a
lower surface-to-volume ratio. Figure 4 displays both of these

Figure 4. (a) Time evolution of the semiaxes (a, b, c) along the


trajectories for micelles containing 50 (brown, dark green, green) or
216 water molecules (red, orange, yellow) at 300 K. (b) Time
evolution of s for reverse micelles containing 50 (red) or 216 (black)
water molecules at 300 K and for pure water clusters containing 50
(green) or 216 (brown) water molecules at 300 K. Semiaxes a,b,c are
shown in nm units.

criteria for the two sizes of the water pools investigated here.
Inspection of the semiaxes (Figure 4a) shows that the water
pool is rather nonspherical, with a roughly prolate shape. A
similar picture is obtained from the s parameter. Let us rst
consider pure water clusters in chloroform. Since water is only
poorly soluble in chloroform, we can expect that pure water
clusters immersed in chloroform tend to form spherical
particles. Figure 4b demonstrates a relatively small value of
the s parameter irrespective of the cluster size, indicating a
spherical shape of the particle. Once a micelle is formed, it loses
the spherical shape. This eect is even more pronounced for
larger micelles as can be seen in Figure 4: the s parameter is

1
M (a 2 + b 2 )
5
8289

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Article

Langmuir

function of time. Note that for the condensed phase, the slope
of the MSD should be proportional to the diusion coecient.
Figure 6 shows the MSD of dierent molecular units of the
micelle as a function of time for micelles containing 50 water
molecules at dierent temperatures.

signicantly higher for a larger micelle (black curve in Figure


4b), and the semiaxes do not coincide as in the smaller micelles.
Clearly, an onset of the solubilization of the water cluster is
observed here. It should be mentioned that nonspherical shapes
of larger micelles are found also for other reverse micelle
systems.25 Interestingly, a highly nonspherical shape of water
particles is observed also in the gas phase water particles
formed in a supersonic expansion.61 As the micelles formed in
this study are rather small, we made the approximation that
their shape is spherical while processing the experimental data.
Let us now explore how the structure of the micelle changes
with decreasing temperature. In the simulation, cooling will
always be ultrafast and we can thus observe only a
nonequilibrium situation. We compared the properties of
water/CTAB/chloroform micelles at the temperatures of 300
and 200 K. The latter temperature is slightly below the melting
temperature of chloroform; however, the system stayed liquid
with the force eld used in this study. Figure 5 shows the

Figure 5. (a) Semiaxes (a, b, c) for micelles containing 50 waters at


300 K (red, orange, yellow) and 200 K (blue, violet, maroon). (b)
Time evolution of s for micelles containing 50 water molecules at 300
K (black) and 200 K (green). Semiaxes a,b,c are shown in nm units.

Figure 6. MSD for three dierent temperatures: 200 K (blue), 273 K


(green), and 300 K (red) as a function of time for micelles containing
50 water molecules for (a) water, (b) chloroform, and (c) terminal
CH3 groups. Data for water and terminal CH3 were obtained from a
progressive t, for chloroform from an untted trajectory.

structural parameters characterizing the shape of the water


pool, the semiaxes a, b, c and the ratio s. From both parameters,
we observe (i) a slower dynamics of the structural uctuations
at lower temperatures and (ii) more spherical shapes of the
water pools at the lower temperature. The latter nding
suggests a tendency for a phase separation of water under these
conditions.
The central question is what is the state of the water pool at a
lower temperature? The chloroform stays liquid; yet, it is
unclear whether the small water particles immersed in
chloroform freeze or the thermal agitation of the surrounding
chloroform keeps it in a liquid state. Water clusters might form
a cubic ice16 even though it is uncertain whether such small
particles can form true ice structures.8 Nevertheless, the water
pool can still form an amorphous solid phase.
The character of the cluster can be revealed by looking at the
mean square displacement (MSD) of the water molecules as a

Note that, for water molecules, the MSD grows linearly only
for a short time before MSD turns over and saturates to give a
nite value. This reects a nite size of the water particles
immersed in chloroform. The water molecule can diuse freely
only within the water pool, the maximum displacement is thus
restricted by the size of the water cluster. For a liquid, the value
at which MSD saturates can be considered as the measure of
the eective size of the water pool. We observed that (R2)1/2
is approximately 1 nm for a micelle with 50 water molecules. At
300 K, the water pool is liquid and the diusion of water
molecules is possible. The initial slope is related to the diusion
coecient of water in the water pool. We found no qualitative
dierence upon decreasing the temperature from 300 to 273 K.
However, it should be mentioned that the ice described with
8290

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Langmuir
the SPC/E water model melts well below this temperature.
Freezing the cluster to 200 K brings a qualitative change. The
observed dynamics is much slower which is particularly
apparent in the beginning. On the other hand, the mobility
of the terminal CH3 group decreases much less with the
decreasing temperature, whereas chloroform still stays liquid
even at such low temperatures.
The molecular dynamics simulations have thus demonstrated
that upon fast freezing, the mobility of the water molecule, and
to a lesser extent also of the CTAB, is much suppressed. This
may have caused the disappearance of the NMR signal in our
experiments with rapid cooling. The water pool simultaneously
becomes more spherical which suggests a tendency for a phase
separation. However, a phase separation and possible water
shedding are well beyond the time scale of our MD simulations.

ACKNOWLEDGMENTS

REFERENCES

(1) Caupin, F. Escaping the No Mans Land: Recent Experiments on


Metastable Liquid Water. J. Non-Cryst. Solids 2015, 407, 441448.
(2) Alba-Simionesco, C.; Coasne, B.; Dosseh, G.; Dudziak, G.;
Gubbins, K. E.; Radhakrishnan, R.; Sliwinska-Bartkowiak, M. Effects of
Confinement on Freezing and Melting. J. Phys.: Condens. Matter 2006,
18, R15R68.
(3) Liu, J.; Nicholson, C. E.; Cooper, S. J. Direct Measurement of
Critical Nucleus Size in Confined Volumes. Langmuir 2007, 23, 7286
7292.
(4) Yeh, Y.; Feeney, R. E. Antifreeze Proteins: Structures and
Mechanisms of Function. Chem. Rev. 1996, 96, 601617.
(5) Kulmala, M.; Kontkanen, J.; Junninen, H.; Lehtipalo, K.;
Manninen, H. E.; Nieminen, T.; Petaja, T.; Sipila, M.;
Schobesberger, S.; Rantala, P.; Franchin, A.; Jokinen, T.; Jarvinen,
E.; Aijala, M.; Kangasluoma, J.; Hakala, J.; Aalto, P. P.; Paasonen, P.;
Mikkila, J.; Vanhanen, J.; Aalto, J.; Hakola, H.; Makkonen, U.;
Ruuskanen, T.; Mauldin, R. L.; Duplissy, J.; Vehkamaki, H.; Back, J.;
Kortelainen, A.; Riipinen, I.; Kurten, T.; Johnston, M. V.; Smith, J. N.;
Ehn, M.; Mentel, T. F.; Lehtinen, K. E. J.; Laaksonen, A.; Kerminen, V.
M.; Worsnop, D. R. Direct Observations of Atmospheric Aerosol
Nucleation. Science 2013, 339, 943946.
(6) Levinger, N. E. Water in Confinement. Science 2002, 298, 1722
1723.
(7) Klicova, L.; Sebej, P.; Stacko, P.; Filippov, S. K.; Bogomolova, A.;
Padilla, M.; Klan, P. CTAB/Water/Chloroform Reverse Micelles: A
Closed or Open Association Model? Langmuir 2012, 28, 15185
15192.
(8) Pradzynski, C. C.; Forck, R. M.; Zeuch, T.; Slavicek, P.; Buck, U.
A Fully Size-Resolved Perspective on the Crystallization of Water
Clusters. Science 2012, 337, 15291532.
(9) Ehre, D.; Lavert, E.; Lahav, M.; Lubomirsky, I. Water Freezes
Differently on Positively and Negatively Charged Surfaces of
Pyroelectric Materials. Science 2010, 327, 672675.
(10) Broto, F.; Clausse, D. A Study of the Freezing of Supercooled
Water Dispersed within Emulsions by Differential Scanning
Calorimetry. J. Phys. C: Solid State Phys. 1976, 9, 42514257.
(11) Simorellis, A. K.; Van Horn, W. D.; Flynn, P. F. Dynamics of
Low Temperature Induced Water Shedding from AOT Reverse
Micelles. J. Am. Chem. Soc. 2006, 128, 50825090.
(12) Munson, C. A.; Baker, G. A.; Baker, S. N.; Bright, F. V. Effects of
Subzero Temperatures on Fluorescent Probes Sequestered within
Aerosol-OT Reverse Micelles. Langmuir 2004, 20, 15511557.
(13) Baglioni, P.; Nakamura, H.; Kevan, L. Electron Spin Echo
Modulation Study of AOT Reverse Micelles. J. Phys. Chem. 1991, 95,
38563859.
(14) Nucci, N. V.; Vanderkooi, J. M. Temperature Dependence of
Hydrogen Bonding and Freezing Behavior of Water in Reverse
Micelles. J. Phys. Chem. B 2005, 109, 1830118309.
(15) Dokter, A. M.; Petersen, C.; Woutersen, S.; Bakker, H. J.
Vibrational Dynamics of Ice in Reverse Micelles. J. Chem. Phys. 2008,
128, 044509.

CONCLUSION
In this work, we investigated the behavior of CTAB/water/
chloroform reverse micelles (w 3.4) at subzero temperatures
using NMR. The experiments provided evidence that supercooled water inside the micellar water pool became unstable at
253 K, shed water, and new smaller micelles (w 1) were
formed. A part of the encapsulated water was frozen at 253 K,
which is well above the temperature of homogeneous
nucleation, indicating that the positively charged surface
might induce a heterogeneous nucleation of the water pool.
The stability of cationic CTAB reverse micelles is thus lower
than the stability of anionic AOT reverse micelles, probably due
to a lower enthalpic stabilization caused by a lower surface
hydration of the CTAB ammonium headgroups. By cooling to
233 K, the reverse micelles became completely frozen and
stayed intact in the solution. Such reverse micelles with an
average hydrodynamic radius of 1.26 nm at 30 C contain 50
water molecules in the water core, which is well below the
onset of an ice crystal formation, thus the frozen water should
be in an amorphous form.
The structure of the CTAB/water/chloroform reverse
micelles was further studied by molecular dynamics simulations.
It turned out that the micelles are formed by an aqueous pool
with bromide anions residing on its surface. Such micelles tend
to form nonspherical structures, especially for larger water
pools. Upon fast freezing, the mobility of the water molecules is
suppressed. This is true also for the terminal CH3 group of the
CTAB molecule, however, with a slower decrease of the
diusion coecient with decreasing temperature. Larger cooled
CTAB/water/chloroform micelles tend to become more
spherical.
ASSOCIATED CONTENT

S Supporting Information
*

NMR experiments; parameters of the force eld. The


Supporting Information is available free of charge on the
ACS Publications website at DOI: 10.1021/acs.langmuir.5b01776.

Support for this work was provided by the Czech Science


Foundation (15-12386S). The RECETOX research infrastructure was supported by the projects of the Czech Ministry
of Education (LO1214) and (LM2011028). P.S. was supported
by the project Employment of Best Young Scientists for
International Cooperation Empowerment (CZ.1.07/2.3.00/
30.0037) conanced from European Social Fund and the state
budget of the Czech Republic. The authors express their thanks
to Lukas Maier and Vaclav Havel for their help with the NMR.
Peter Stacko, Zdenek Moravec, and Jakob Wirz are acknowledged for fruitful discussions.

Article

AUTHOR INFORMATION

Corresponding Author

*E-mail: klan@sci.muni.cz. Phone: +420-54949-4856. Fax:


+420-54949-2443.
Notes

The authors declare no competing nancial interest.


8291

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Article

Langmuir

(37) van Gunsteren, W. F.; Berendsen. Groningen Molecular


Simulation (GROMOS) Library Manual; Biomos: Groningen, Netherlands, 1987; pp. 1221
(38) Dang, L. X. Computational Study of Ion Binding to the Liquid
Interface of Water. J. Phys. Chem. B 2002, 106, 1038810394.
(39) Kann, Z. R.; Skinner, J. L. A Scaled-Ionic-Charge Simulation
Model that Reproduces Enhanced and Suppressed Water Diffusion in
Aqueous Salt Solutions. J. Chem. Phys. 2014, 141, 104507.
(40) Leontyev, I.; Stuchebrukhov, A. Accounting for Electronic
Polarization in Non-polarizable Force Fields. Phys. Chem. Chem. Phys.
2011, 13, 26132626.
(41) Lear, B. J.; Glover, S. D.; Salsman, J. C.; Londergan, C. H.;
Kubiak, C. P. Solvent Dynamical Control of Ultrafast Ground State
Electron Transfer: Implications for Class IIIII Mixed Valency. J. Am.
Chem. Soc. 2007, 129, 1277212779.
(42) Fileti, E. E.; Chaban, V. V. The Scaled-Charge Additive Force
Field for Amino Acid Based Ionic Liquids. Chem. Phys. Lett. 2014, 616,
205211.
(43) Pluharova, E.; Mason, P. E.; Jungwirth, P. Ion Pairing in
Aqueous Lithium Salt Solutions with Monovalent and Divalent
Counter-Anions. J. Phys. Chem. A 2013, 117, 1176611773.
(44) Vazdar, M.; Pluharova, E.; Mason, P. E.; Vacha, R.; Jungwirth, P.
Ions at Hydrophobic Aqueous Interfaces: Molecular Dynamics with
Effective Polarization. J. Phys. Chem. Lett. 2012, 3, 20872091.
(45) Otten, D. E.; Shaffer, P. R.; Geissler, P. L.; Saykally, R. J.
Elucidating the Mechanism of Selective Ion Adsorption to the Liquid
Water Surface. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 701705.
(46) Leontyev, I. V.; Stuchebrukhov, A. A. Polarizable Molecular
Interactions in Condensed Phase and their Equivalent Nonpolarizable
Models. J. Chem. Phys. 2014, 141, 014103.
(47) Sanner, M. F.; Olson, A. J.; Spehner, J. C. Reduced Surface: An
Efficient Way to Compute Molecular Surfaces. Biopolymers 1996, 38,
305320.
(48) Shinoda, K.; Hutchinson, E. Pseudo-Phase Separation Model for
Thermodynamic Calculations on Micellar Solutions. J. Phys. Chem.
1962, 66, 577582.
(49) El Seoud, O. A.; Pires, P. A. R. FTIR and 1H NMR Studies on
the Structure of Water Solubilized by Reverse Aggregates of
Dodecyltrimethylammonium Bromide, Didodecyldimethylammonium
Bromide, and their Mixtures in Organic Solvents. In Surface and
Interfacial Forces - From Fundamentals to Applications; Auernhammer,
G. K., Butt, H. J., Vollmer, D., Eds.; Springer: New York, 2008; Vol.
134, pp 101110.
(50) Zhang, J.; Bright, F. V. Nanosecond Reorganization of Water
within the Interior of Reversed Micelles Revealed by FrequencyDomain Fluorescence Spectroscopy. J. Phys. Chem. 1991, 95, 7900
7907.
(51) Gottlieb, H. E.; Kotlyar, V.; Nudelman, A. NMR Chemical
Shifts of Common Laboratory Solvents as Trace Impurities. J. Org.
Chem. 1997, 62, 75127515.
(52) Cho, H.; Shepson, P. B.; Barrie, L. A.; Cowin, J. P.; Zaveri, R.
NMR Investigation of the Quasi-Brine Layer in Ice/Brine Mixtures. J.
Phys. Chem. B 2002, 106, 1122611232.
(53) Mallamace, F.; Corsaro, C.; Broccio, M.; Branca, C.; GonzalezSegredo, N.; Spooren, J.; Chen, S. H.; Stanley, H. E. NMR Evidence of
a Sharp Change in a Measure of Local Order in Deeply Supercooled
Confined Water. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 12725
12729.
(54) The diusion coecients for the 40 and 8 mM CTAB
concentrations were found in the range of 6.86.3 1010 m2 s1,
which corresponds to very small RH (0.620.68 nm, respectively).
(55) Liu, J. C.; Li, G. Z.; Han, B. X. Characteristics of AOT
Microemulsion Structure: A Small Angle X-Ray Scattering Study. Chin.
Chem. Lett. 2001, 12, 10231026.
(56) Rodenas, E.; Valiente, M. The Determination of Some Physical
Properties of Reverse CTAB Micelles in 1- Hexanol. Colloids Surf.
1992, 62, 289295.

(16) Suzuki, A.; Yui, H. Crystallization of Confined Water Pools with


Radii Greater than 1 nm in AOT Reverse Micelles. Langmuir 2014, 30,
72747282.
(17) Suzuki, A.; Yui, H. Spectroscopic Study of the Melting and
Reconstruction of Sodium Bis(2-ethylhexyl) Sulfosuccinate (AOT)
Reverse Micelles from their Frozen States. J. Colloid Interface Sci. 2015,
443, 188196.
(18) Fayer, M. D. Dynamics of Water Interacting with Interfaces,
Molecules, and Ions. Acc. Chem. Res. 2012, 45, 314.
(19) El Seoud, O. A.; Blasko, A.; Bunton, C. A. Proton NMR Studies
on the Structure of Water at Interfaces of Aqueous Micelles. 4. Effects
of Cationic and Zwitterionic Headgroups. Ber. Bunsen-Ges. Phys. Chem.
1995, 99, 12141220.
(20) Dokter, A. M.; Woutersen, S.; Bakker, H. J. Ultrafast Dynamics
of Water in Cationic Micelles. J. Chem. Phys. 2007, 126, 124507.
(21) Schulz, P. C. DSC Analysis of the State of Water in SurfactantBased Microstructures. J. Therm. Anal. Calorim. 1998, 51, 135149.
(22) Lefebvre, B. G.; Liu, W. X.; Peterson, R. W.; Valentine, K. G.;
Wand, A. J. NMR Spectroscopy of Proteins Encapsulated in a
Positively Charged Surfactant. J. Magn. Reson. 2005, 175, 158162.
(23) Faeder, J.; Albert, M. V.; Ladanyi, B. M. Molecular Dynamics
Simulations of the Interior of Aqueous Reverse Micelles: A
Comparison between Sodium and Potassium Counterions. Langmuir
2003, 19, 25142520.
(24) Abel, S.; Sterpone, F.; Bandyopadhyay, S.; Marchi, M. Molecular
Modeling and Simulations of AOT-Water Reverse Micelles in
Isooctane: Structural and Dynamic Properties. J. Phys. Chem. B
2004, 108, 1945819466.
(25) Mills, A. J.; Wilkie, J.; Britton, M. M. NMR and Molecular
Dynamics Study of the Size, Shape, and Composition of Reverse
Micelles in a Cetyltrimethylammonium Bromide (CTAB)/n-Hexane/
Pentanol/Water Microemulsion. J. Phys. Chem. B 2014, 118, 10767
10775.
(26) Cata, G. F.; Rojas, H. C.; Gramatges, A. P.; Zicovich-Wilson, C.
M.; Alvarez, L. J.; Searle, C. Initial Structure of Cetyltrimethylammonium Bromide Micelles in Aqueous Solution from Molecular
Dynamics Simulations. Soft Matter 2011, 7, 85088515.
(27) Holz, M.; Weingartner, H. Calibration in Accurate Spin-Echo
Self-Diffusion Measurements Using 1H and Less Common Nuclei. J.
Magn. Reson. 1991, 92, 115125.
(28) Johnson, C. S. Diffusion Ordered Nuclear Magnetic Resonance
Spectroscopy: Principles and Applications. Prog. Nucl. Magn. Reson.
Spectrosc. 1999, 34, 203256.
(29) Reisshusson, F.; Luzzati, V. The Structure of the Micellar
Solutions of Some Amphiphilic Compounds in Pure Water as
Determined by Absolute Small-Angle X-Ray Scattering Techniques.
J. Phys. Chem. 1964, 68, 35043511.
(30) Law, S. J.; Britton, M. M. Sizing of Reverse Micelles in
Microemulsions using NMR Measurements of Diffusion. Langmuir
2012, 28, 1169911706.
(31) Zhao, F.; Du, Y. K.; Xu, J. K.; Liu, S. F. Determination of
Surfactant Molecular Volume by Atomic Force Microscopy. Colloid J.
2006, 68, 784787.
(32) Lang, J.; Mascolo, G.; Zana, R.; Luisi, P. L. Structure and
Dynamics of Cetyltrimethylammonium Bromide Water-in-Oil Microemulsions. J. Phys. Chem. 1990, 94, 30693074.
(33) Millero, F. J. Molal Volumes of Electrolytes. Chem. Rev. 1971,
71, 147176.
(34) Van der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A.
E.; Berendsen, H. J. C. GROMACS: Fast, Flexible, and Free. J.
Comput. Chem. 2005, 26, 17011718.
(35) Berendsen, H. J. C.; Grigera, J. R.; Straatsma, T. P. The Missing
Term in Effective Pair Potentials. J. Phys. Chem. 1987, 91, 62696271.
(36) Jorgensen, W. L.; Maxwell, D. S.; TiradoRives, J. Development
and Testing of the OPLS All-Atom Force Field on Conformational
Energetics and Properties of Organic Liquids. J. Am. Chem. Soc. 1996,
118, 1122511236.
8292

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Article

Langmuir
(57) Warr, G. G.; Sen, R.; Evans, D. F.; Trend, J. E. Microemulsion
Formation and Phase Behavior of Dialkydimethylammonium Bromide
Surfactants. J. Phys. Chem. 1988, 92, 774783.
(58) Demontis, P.; Gulin-Gonzalez, J.; Jobic, H.; Masia, M.; Sale, R.;
Suffritti, G. B. Dynamical Properties of Confined Water Nanoclusters.
Simulation Study of Hydrated Zeolite NaA: Structural and Vibrational
Properties. ACS Nano 2008, 2, 16031614.
(59) Caleman, C.; Hub, J. S.; van Maaren, P. J.; van der Spoel, D.
Atomistic Simulation of Ion Solvation in Water Explains Surface
Preference of Halides. Proc. Natl. Acad. Sci. U. S. A. 2011, 108, 6838
6842.
(60) Jungwirth, P.; Tobias, D. J. Specific Ion Effects at the Air/Water
Interface. Chem. Rev. 2006, 106, 12591281.
(61) Lengyel, J.; Pysanenko, A.; Poterya, V.; Slavicek, P.; Farnik, M.;
Kocisek, J.; Fedor, J. Phys. Rev. Lett. 2014, 112, 113401.

8293

DOI: 10.1021/acs.langmuir.5b01776
Langmuir 2015, 31, 82848293

Você também pode gostar