Sei sulla pagina 1di 10

Article

pubs.acs.org/Macromolecules

Coassembly of Poly(ethylene oxide)-block-poly(methacrylic acid) and


NDodecylpyridinium Chloride in Aqueous Solutions Leading to
Ordered Micellar Assemblies within Copolymer Aggregates
Mariusz Uchman,*, Miroslav Stepanek, Sylvain Prevost,, Borislav Angelov, Jan Bednar,,#
Marie-Sousai Appavou,% Michael Gradzielski, and Karel Prochazka

Department of Physical and Macromolecular Chemistry, Faculty of Science, Charles University in Prague, Hlavova 2030, 128 40
Prague 2, Czech Republic

Stranski Laboratorium fur Physikalische und Theoretische Chemie, Technische Universitat Berlin, Strae des 17. Juni 124, 10623
Berlin, Germany

Soft Matter Department, Helmholtz-Zentrum Berlin, Hahn-Meitner-Platz 1, 14109 Berlin, Germany

Institute of Macromolecular Chemistry, Academy of Sciences of the Czech Republic, Heyrovsky Square 2, 16206 Prague 6, Czech
Republic

First Faculty of Medicine, Institute of Cellular Biology Albertov 4, Charles University in Prague, 128 01 Prague 2, Czech Republic
#
LIPhy UMR 5588, Univ. Grenoble 1/CNRS, Grenoble F-38041, France
%
Forschungszentrum Julich GmbH, IFF-JCNS, Lichtenbergerstrae 1, D-85747 Garching, Germany
S Supporting Information
*

ABSTRACT: Formation of polyelectrolytesurfactant (PE


S) complexes of poly(ethylene oxide)-block-poly(methacrylic
acid) (PEO705PMAA476) and N-dodecylpyridinium chloride
(DPCl) in aqueous solution was studied by static and dynamic
light scattering (SLS, DLS), small-angle neutron scattering
(SANS), small-angle X-ray scattering (SAXS), and cryogenic
transmission electron microscopy (cryo-TEM). While it was
found previously [Macromolecules 1997, 30, 3519] by
microcalorimetric titration that in a similar system (PEO176
PMAA186) crystallization of aliphatic tails of N-dodecylpyridinium bromide did not occur, in our system it was evidenced by SAXS that upon addition of DPCl to fully ionized PEO705
PMAA476 the ordered arrangement of the surfactant occurs in a certain range of PEO705PMAA476 concentrations and
surfactant-to-polyelectrolyte charge molar ratio (Z). Our data suggest a four-step process in the behavior of the PEO705
PMAA476/DPCl system: (i) coexistence of loose aggregates of electrostatically bound surfactants to PMAA block with free and
almost unperturbed copolymer coils at Z 1, (ii) formation of aggregates containing ill-dened cores formed by DPCl micelles
attached to coiled PMAA chains (beads-on-a-string nanoparticles) in the range around Z = 0.5, (iii) formation of compact core
shell nanoparticles with a core formed by densely packed ordered (crystalline) DPCl micelles and PEO shell starting slightly
before charge equimolarity (Z = 1), and (iv) the region of coexistence of the coreshell nanoparticles with free DPCl micelles in
excess above equimolarity (Z 1). In the region around Z = 0.5, the nanoparticles with nonordered cores coexist in a mixture
either with a fraction free chains and large swollen nanoparticles decorated by surfactant micelles (at lower Z) or with the core
shell nanoparticles (at higher Z). PES complexes were characterized in detail in terms of molar mass, size, shape, and internal
structure.

INTRODUCTION
Polyelectrolytesurfactant (PES) complexes have attracted
great interest of scientists and engineers not only because of a
curiosity driven fundamental research of their interesting
behavior but especially because of their wide industrial
applications ranging from cosmetics, detergents, and paints to
drug delivery and food technology.13 The coassembly of PE
S complexes is driven by a combination of electrostatic and
hydrophobic interactions, and both components aect the
phase behavior, interfacial properties, rheology, etc. To design
2012 American Chemical Society

materials with required properties, a number of factors have to


be taken into account, such as the polyelectrolyte4 and/or the
surfactant concentration, molecular weight,4 charge density,5
backbone rigidity,6 and degree of branching of the polyelectrolyte7 as well as the polarity of the headgroup and the length
of the aliphatic tail of the surfactant.8 The high number of
parameters that can be varied allows one, on one hand, to
Received: July 19, 2012
Published: August 7, 2012
6471

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Macromolecules

Article

complexes only, we performed a study of nanoparticles in a


broad range of surfactant/polymer stoichiometric ratios, Z,
which allows us to demonstrate dierences in the aggregation
behavior, as compared to the previously reported results.

control and tune the structure and stimuli-responsive properties


of prepared nanoparticles with a high eciency, but on the
other hand, it requires complex and lengthy fundamental
studies of potentially applicable systems.
In the past decades, most attention has been paid to PES
formed by homopolymer polyelectrolytes.13 Nevertheless,
several examples of studies on PES of block polyelectrolytes
have been published recently as well.1,6,915 It has been
reported that PES formed by double-hydrophilic block
copolymers containing polyelectrolyte blocks with oppositely
charged ionic surfactants assemble in both vesicles912 and
micelles.13,14 In those systems, the membrane of the vesicles or
the core of the micelles is composed of the insoluble PES
complex and the stabilizing water-soluble shell of the neutral
block. It is worth-mentioning that both the micelles with
disordered cores1619 and various ordered structures1923
containing polyelectrolyte blocks connected by densely packed
surfactant micelles were observed. Interestingly, very highly
ordered, e.g., macrolattice-forming materials, have been
prepared by the complexation of polyelectrolyte gels and
surfactants.3,24,25
Kabanov and co-workers1214 as the rst research team
investigated the formation of electrostatically stabilized
complexes between double hydrophilic copolymer poly(ethylene oxide)-b-poly(sodium methacrylate), of nearly
symmetric (PEO176PMAA188) and asymmetric (PEO210
PMAA35) monomeric unit composition with single-, double-,
and triple-tail cationic surfactants diering in the length of the
tail (C12, C16) as well as in chemical nature of the headgroup
and counterions. By using electron microscopy with selective
staining, the authors conrmed the formation of vesicles and
micelles of various sizes in aqueous solutions. On the basis of
isothermal titration calorimetry results, they concluded that
surfactant micelles crystallize in cores of PEO176PMAA188
cetylpyridinium bromide complexes; however, no ordered
structures were detected in analogous nanoparticles formed
with dodecylpyridinium bromide.14
Berret and co-workers19 found that the precipitated phase
formed as a result of the interaction of homopolyelectrolytes
with oppositely charged surfactants, poly(sodium acrylate) with
dodecyltrimethylammonium bromide, PANa420/DTAB, and
poly(trimethylammonium ethyl acrylate methyl sulfate) with
sodium dodecyl sulfate, PTEA41/SDS, respectively, contain
long-range order Pm3n cubic and hexagonal micellar structures.
However, once those polyelectrolytes (regardless of their
lengths) become blocks of double hydrophilic block copolymers containing a neutral poly(acrylamide) block, PAM, they
do not exhibit any crystalline order in cores of PANa-b-PAM/
DTAB, or PTEA-b-PAM/SDS complexes, because of additional
surfactantpolymer interactions.
In this paper, we investigate microstructures of surfactant
polyelectrolyte complexes of PEOPMAA with DPCl in detail.
Analogous complexes have already been studied by other
authors.1214 However, the use of a copolymer with a
signicantly higher molar mass (PEO705PMAA476) than that
studied earlier enables the formation of extended PES
complex domains in assembled structured complexes and, as
it will be shown later, modies appreciably the process of the
electrostatic self-assembly and properties of formed nanoparticles. The size and structure of the nanoparticles have been
studied by a combination of scattering (LS, SANS, and SAXS)
and microscopy (cryo-TEM) techniques. Because most studies
on this system published so far focused on stoichiometric

EXPERIMENTAL SECTION

Materials. Poly(ethylene oxide)-block-poly(methacrylic acid),


PEO705PMAA476, Mw/Mn = 1.5, was purchased form Polymer
Source, Inc., Dorval, Quebec, Canada. N-Dodecylpyridinium chloride
hydrate (98%), DPCl, was a product of Aldrich. Deuterium oxide
(99.95% D, Euriso-top, France) was used instead of water in samples
for SANS to increase the contrast with the hydrogenated material and
to lower the background essentially due to incoherent scattering
proportional to the amount of hydrogen. Sodium tetraborate (p.a.
99%) was purchased from Fluka and dissolved in water taken from
Millipore System; the nal buer concentration was 50 mM, pH 9.3.
Transmission Cryo-Electron Microscopy (Cryo-TEM). CryoTEM was used to visualize the morphology of the objects in solution.
This technique provides direct imaging of the hydrated sample without
perturbing the nanoparticles. The samples for cryo-TEM were
prepared as described earlier.26 3 L of the sample solution was
applied to an electron microscopy grid covered with perforated carbon
supporting lm (C-at 2/2-2C, Electron Microscopy Science) glowdischarged for 20 s with 10 mA current. Most of the sample was
removed by blotting (Whatman no. 1 lter paper) for 1 s, and the
grid was immediately plunged into liquid ethane held at 183 C. The
sample was then transferred without rewarming into a Tecnai Sphera
G20 electron microscope using a Gatan 626 cryo-specimen holder.
The images were recorded at 120 kV accelerating voltage and
microscope magnication ranging from 5000 to 14 500 using a
Gatan UltraScan 1000 slow scan CCD camera (giving a nal pixel size
from 2 to 0.7 nm) and low dose mode with the electron dose not
exceeding 1500 electrons per nm2. Typical value of applied underfocus
ranged between 1.5 and 2.7 m. The applied blotting conditions
resulted in the specimen thickness varying between 100 and ca. 300
nm.
Light Scattering. The light scattering setup (ALV, Langen,
Germany) consisted of a 22 mW HeNe laser, an ALV CGS/8F
goniometer, an ALV High QE APD detector, and an ALV 5000/EPP
multibit, multitau autocorrelator at 20 C. The copolymer
concentration in all solutions was c = 1 mg/mL.
Static light scattering (SLS) measurements of the corrected excess
scattering intensity, I(q), of the copolymer solutions, as a function of
the magnitude of the scattering vector q = (4n0/) sin(/2) ( being
the scattering angle, n0 = 1.332 the refractive index of the solvent, and
= 632.8 nm the wavelength of the incident light), were treated by
tting the data in the angular range from 30 to 90 to the Guinier
equation

ln

I(q)
1
= R g 2q2
I(0)
3

(1)

to obtain the forward scattering intensity, I(0), and the radius of


gyration, Rg.
Dynamic light scattering measurements were evaluated by tting the
measured normalized time autocorrelation function of the scattered
light intensity, g(2)(t), related to the electric eld autocorrelation
function, g(1)(t), by the Siegert relation,27 g(2)(t) = 1 + |g(1)(t)|2,
where is the coherence factor. A constrained regularization algorithm
(CONTIN) provided the distribution of relaxation times , A(), as
the inverse Laplace transform of g(1)(t) function

g(1)(t ) =

t
A( ) exp d

(2)

The A() distributions were recalculated to the distributions of


apparent hydrodynamic radii, RHapp, assuming the apparent diusion
coecient Dapp = 1/q2 and using the StokesEinstein formula, RHapp
= kBT/6Dapp, where kB is the Boltzmann constant, T the
temperature, and the solvent viscosity.
6472

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Macromolecules

Article

Alternatively, the autocorrelation functions, collected at scattering


angles from 30 to 90, were further tted to the second order
cumulant expansion

ln g(1)(t , q) = 1(q)t 2 +

2(q) 2
t
2

were isotropic, they were radially averaged, and spectra from the three
congurations were merged with no need for any arbitrary coecient.
Apparent Molecular Volumes and Scattering Length Densities.
SANS data depend on the contrast between media of dierent
scattering length densities, . The scattering length is an atomic
(nuclide)-dependent property and for a given compound depends on
its chemical composition. For the data analysis, the densities or
apparent volumes of the species are necessary, and the values used in
this work are presented in Table 1 (further details are given in the
Supporting Information).

(3)

where 1 and 2 respectively are the rst and the second moment of
the distribution function of relaxation rates. The diusion coecient of
the particles, D, was obtained by the linear extrapolation to zero q
values as

1(q)
q2

= D(1 + CR g 2q2)

Table 1. Apparent Volumes, Densities and Scattering Length


Densities of the Compounds Used in This Work
(4)

where C is the structure parameter dependent on the shape and degree


of polydispersity of the particles. The hydrodynamic radius of the
particles, RH, can be calculated from D using the StokesEinstein
formula.
Small-Angle X-ray Scattering (SAXS). Small-angle X-ray
scattering measurements with PEO705-PMAA476/DPCl samples with
the copolymer concentration, c = 5 mg/mL, were performed at the I22
beamline of Diamond Light Source (Didcot, U.K). In case of 1 mg/
mL PEO705PMAA476/DPCl samples, the measurements were done
at the ID02 beamline at the European Synchrotron Radiation Facility
in Grenoble, France.28 Both synchrotrons are of the third generation
and oer high X-ray ux (10121013 photons s1). Samples were
lled in borosilicate glass capillaries (Hilgenberg Gmbh, Germany,
diameter 1.5 mm, wall thickness 10 m), and the capillaries were
mounted in a holder in air at the room temperature (21 C) without a
special vacuum chamber. The I22 station was equipped with a gas wire
RAPID 2D detector,29,30 the area of which was divided into 512 512
pixels. The wavelength of the incident X-ray beam and the beam size at
the sample areas were 0.155 nm and 350 275 m2, respectively. The
accessible q-range was from 0.09 to 4.17 nm1. A metal attenuator was
installed in order to prevent the samples from damage by X-radiation.
The exposure time was about 1 s.
The ID02 beamline was equipped by a high-sensitivity CCD
FReLoN 4 M detector, the area of which was also divided into 512
512 pixels after 4 4 pixels binning. The X-ray wavelength was 0.1
nm, the beam size was 350 270 m2, and the accessible q-range was
from 0.07 to 3.2 nm1. Instead of using a metal attenuator to prevent
samples from X-ray radiation damage, the exposure time was decreased
to 10 ms by means of a high-precision X-ray beam shutter.
At both beam stations, the incident and transmitted beam intensities
were recorded simultaneously. Silver behenate was used for the qrange calibration and glassy carbon for the intensity normalization.
After the data acquisition, the obtained 2D images were integrated into
1D scattering curves by means of the Fit2D software.31 The
backgrounds, coming from the glass capillary and the solvent, were
measured and subtracted using conventional procedures. One and the
same capillary was used during the measurements with and without
solvent. This allowed for very precise subtraction of the background.
Small-Angle Neutron Scattering (SANS). SANS Apparatus.
SANS experiments were carried out on KWS-2 from the Julich Center
for Neutron Science (JCNS) at FRM-II in Munchen, Germany.
Samples were poured in quartz cuvettes (QX quality from Hellma) of
2 mm neutron pathway, thermostated at 25.0 C. Three congurations
were used with SD = 1.7 m, = 0.45 nm, SD = 7.7 m, = 0.45 nm,
and SD = 7.7 m, = 1.2 nm, where SD is the sample-to-detector
distance and the mean wavelength (fwhm = 20%); the nal q-range
obtained is 2.5 1023.2 nm1, corresponding to 2250 nm in the
real space (using Braggs law d = 2/q), where q is the magnitude of
the scattering vector, q = 4/ sin(/2), being the scattering angle.
The beam was collimated at 8 m from the sample in all cases. The
beam size at the sample position was 8 8 mm2. Data reduction was
performed on 2D patterns by means of BerSANS,32 using the
scattering by a 1.5 mm PMMA sheet to correct for pixel eciency,
taking a tabulated value for the absolute scale, and subtracting the
experimental intensity for the buer as background. Since the data

compound

,b g cm3

Vm,c nm3

104,d nm2

buer
DPCl
C12H25
C5H5NCl
MA
EO
PEO705PMAA476

0.9268 + 0.0478xa
0.8009
1.3966 + 0.0375x
2.07
1.1873
1.5670

0.4451 + 0.0386x
0.3511
0.0941 + 0.0386x
0.0681
0.0616
75.8

6.31
0.23 + 0.18x
0.39
2.540.02x
2.86
0.67
1.61

Fraction of binding counterions. bDensity. cApparent molecular


volume. dScattering length density. For PMAA the apparent volume
depends strongly on the nature of the counterions and the ionization
degree; in this work, the value was set and xed to the value for the
deprotonated polyelectrolyte.

Analysis of SANS Scattering Curves. Scattering data can be


expressed as a sum of two contributions, I(q) = Icoh(q) + Iinc, where
Icoh is the coherent scattering depending on the shape, size, and
repartition of domains in the sample and Iinc a constant background
due to incoherent scattering by atoms. For a single population of
monodisperse spheres, the coherent part can be separated into a
normalized form factor P(q) representing the shape, size, and internal
distribution of materials of the scatterers, a scaling proportional to
their number density, N, their average volume V, their average
contrast = part solvent, and a structure factor S(q) to account for
interactions:
Icoh(q) = N V 2 ()2 P(q)S(q)

(5)

with NV = , the volume fraction. When interactions can be neglected,


S(q) 1. Therefore, the forward scattering can be used to determine
the volume of a particle as long as interactions are negligible:

VI(0) =

I(0)
()2

(6)

If two or more populations of scatterers are present whose dimensions


dier by ca. 1 order of magnitude, the intensity can be approximated as
the sum of scattering contributions by each individual population
(cross-terms are neglected).
Because of the complexity of the system, a model-free analysis based
on integral structural parameters was carried out before model tting.
After obtaining realistic parameters from the model-free analysis, data
were tted with analytical models (see Results and Discussion). In the
case of the pure polymer, a model for the Gaussian coil is used
Ipolym(q) = I(0)

2
[exp( R g 2q2) (1 R g 2q2)]
R g 4q 4

(7)

Using eq 5, I(0) can be expressed as


I(0) =

cM w
NA2

()2

(8)

where Mw is the mass-averaged molar mass of the polymer, NA the


Avogadro constant, c the polymer concentration, and the polymer
apparent density.
6473

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Macromolecules

Article

RESULTS AND DISCUSSION


Light Scattering. Prior to detailed small-angle scattering
studies, the formation of PEO705PMAA476/DPCl particles was
studied by light scattering to gain a global overview on the selfassembling behavior. Figure 1 shows the CONTIN distribu-

Figure 2. Gyration radius, Rg (curve 1), and forward scattering


intensity of particles formed in PEO705PMAA476/DPCl solutions
normalized by that of the pure PEO705PMAA476 solution, I(0)/I0(0)
(curve 2), as functions of the stoichiometric ratio of DPCl-to-PMAA
units, Z. Inset: gyration-to-hydrodynamic radii ratio, , of particles
formed in PEO705PMAA476/DPCl solutions as a function of Z.

increases due to increasing association number. At Z ca. 0.6, a


critical charge ratio is attained at which the transition from
loose aggregates to coreshell particles occurs. The collapse of
the core-forming PMAA/DPCl blocks and the decrease of the
gyration radius as well as changes in the association number are
clearly seen in curves obtained by SLS. The minimum Rg is
observed close to the zero net charge at Z = 1. A slight increase
in the size of the coreshell particles with increasing Z for Z >
1 is probably caused by an increase of the number of surfactant
micelles in cores and by partial (secondary) aggregation of
PEO705PMAA476/DPCl particles as a result of their decreased
solubility in solutions containing excess surfactant molecules.
The dependence of the gyration radius of PEO705
PMAA476/DPCl particles on Z obtained by SLS is similar to
that of RH (calculated from their diusion coecients at q = 0
extrapolated by means of eq 4). However, the comparison of
both curves is worth mentioning because the ratio, = Rg/RH,
reects the inner structure and compactness of scattering
particles. The Z-dependent ratio, , is shown in the inset of
Figure 2. It gradually decreases from = 1.46, which indicates
the scattering from loose structures (for a monodisperse coil at
-conditions, coil = 1.50) to values around 0.8 which are typical
for hard spheres (sphere = 0.778) and other compact spherical
particles, e.g., block copolymer micelles. The observed trend
conrms the above-mentioned increasing compactness of the
aggregates with increasing Z.
The dependence of the forward scattering intensity on Z
(Figure 2, curve 2) is consistent with the model assuming the
transition from loose aggregates to coreshell particles: A steep
increase in I(0) close to Z = 0.6 reects a signicant rise of the
association number as soon as the coreshell aggregates appear
in the solution. However, for a correct interpretation of I(0)
values, it is necessary to keep in mind that even though the
forward scattering intensity is proportional to the molar mass of
PEO705PMAA476/DPCl particles, the proportionality constant
depends on Z due to changes in the refractive index increment
of the polymersurfactant complex.
Small-Angle Neutron Scattering (SANS). SANS curves
were measured for a series of solutions prepared by mixing (i)
stock solution of the polymer at 1 g L1 (13.9 M of the
copolymer, 6.6 mM of PMAA units) in 50 mM sodium
tetraborateD2O buer and (ii) 100 mM surfactant D2O
solution. The scattering curves in the Z-range from 0.15 to 3.6

Figure 1. DLS distributions of apparent hydrodynamic radii of


particles formed in PEO705PMAA476/DPCl solutions. Values of
stoichiometric ratio of DPCl to PMAA units, Z, are indicated above
the corresponding curves. Scattering angle, = 90.

tions of apparent hydrodynamic radii measured by DLS at =


90 for the pure PEO705PMAA476 and for several PEO705
PMAA476/DPCl mixtures in 0.05 M sodium tetraborate buer.
The polymer concentration is 1 mg/mL, and the composition
is given by the charge ratio of surfactant to MAA units: Z =
[DPCl]/[MAA]. In the case of the block copolymer solution,
the distribution is bimodal due to a small amount of strong
scatterers which coexist with individual PEO705PMAA476
chains in the solution (both observed relaxation modes
correspond to diusive motions as proven by linear dependences of the relaxation rate, vs q2). This observation is not
surprising because a spontaneous formation of large aggregates
in polymer solutions (usually somewhat vaguely described as
fairly concentrated metastable microphase-separated domains)
has been reported both for neutral polymers and polyelectrolytes.3337 Upon addition of DPCl, the surfactant interacts
with the copolymer and forms micelles which condense on its
chains. The solubility of the polymer decreases and new large
polymersurfactant aggregates form in the solution. At low Z,
they are in equilibrium with free polymer chains, but as they (i)
dominate the scattering behavior and (ii) their fraction
increases with increasing Z, an apparent unimodal distribution
of hydrodynamic radii is observed at Z ca. 0.4. The positions of
peak maxima shift and their half-widths change with Z due to
structural changes that will be discussed in subsequent
paragraphs.
Additional information on the structure of aggregates can be
obtained by static light scattering. The Z-dependences of the
gyration radius Rg of the copolymer and the aggregates and the
scattering intensity at the zero scattering angle normalized by
that for the pure PEO705PMAA476 solution are shown in
Figure 2. Up to Z ca. 0.6, the size of the particles steeply
6474

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Macromolecules

Article

loose aggregates. As the relative dierence between intensities


at low and high q decreased as compared with the previous
curve, we can conclude that the large and swollen aggregates
(seen in the low q region) for Z = 0.38 are on average more
compact than those at Z = 0.15. This is a reasonable conclusion
as the gradual compensation of the polyelectrolyte charge by
micelles attached to PMAA chains screens the electrostatic
forces, lowers the solubility, and causes partial collapse of
swollen aggregates and their possible reorganization.1
The curve for Z = 0.6 and intermediate q has a shape typical
for small globular aggregates, which indicates that the
interaction of polyelectrolyte chains with increasing content
of the surfactant micelles has already reduced their solubility
signicantly and promoted the formation of distinct globular
assemblies of arranged DPCl micelles held together by PMAA
and stabilized by PEO chains. We can now even see a trace of a
correlation peak at q ca. 1.6 nm1 which will be discussed later.
Nevertheless, the scattering intensity still increases steeply in
the low q range without any sign of leveling-o toward q = 0, in
contrast to systems with Z > 0.6. It is worth mentioning that all
curves for systems with Z in the range 0.150.6 rise steeply
with decreasing q in the low q-range without reaching a plateau,
which indicates that the aggregates (at least some of them) that
are present in studied systems are too large for the upper
length-scale accessible by SANS measurements (which was
about 250 nm in our case). To outline and summarize the
changes in the conformational behavior up to Z = 0.6, for Z =
0.15 and 0.38, we assume a coexistence of free chains with loose
aggregates, while for Z = 0.6, we suppose that the coreshell
micelles coexist with a fraction of voluminous aggregates. With
further increase in DPCl content, the curves in the low q range
adopt a scattering pattern with a kink around 0.1 nm1, which
corresponds to the presence of spherical aggregates of 45 nm
radius. This scattering pattern survives without signicant
changes up to high concentrations of DPCl. For Z > 0.6 a new
correlation peak appears at high q (q 1.6 nm1),
corresponding to a 3.9 nm characteristic spacing between
densely packed surfactant micelles in cores of polymer/
surfactant nanoparticles.
On the basis of the general discussion of the SANS curves
given above (and the more detailed analysis given below), we
can deduce the following scheme of structural changes
occurring in the studied system (Scheme 1).
Modeling the SANS Data. To conrm the proposed
qualitative scheme of the behavior in a broad range of Z and
to get a more quantitative picture, we performed an extensive
modeling of SANS curves based on models deduced from the
standard model free data analysis which is described in more
detail in the Supporting Information. The scattering intensity
for the pure polymer (Z = 0) is weak and the curve is noisy due
to a very low copolymer volume fraction (ca. 0.06%). The data
can be reasonably tted with a model for Gaussian coil (eq 7)
with I(0) = 0.06 cm1 and Rg = 15 nm. Using eq 8, a molar
mass of 40 kg mol1 is obtained which is appreciably less than
the value provided by the producer (72 kg mol1). Besides the
already mentioned low quality of the measured curve, the
discrepancy is most probably due to interchain interactions
which, despite low concentration and ecient electrostatic
screening, cannot be ruled out completely and may lead to the
formation of a low fraction of more or less stable or just
temporary high molar mass aggregates. As only a part of the
curve for high q could have been successfully tted by the
model for Gaussian coil, an accurate value of molar mass have

together with those for pure solutions of the copolymer and


surfactant micelles are shown in Figure 3. While the SANS

Figure 3. SANS data for the pure polymer (1 g L1), the pure
surfactant (100 mM), and mixtures of them (at constant polymer
concentration of 1 g L1), in D2Obuer at 25 C. The surfactant-topolymer molar ratio Z is given next to each spectrum (equimolarity at
Z = 1). Intensities of the data are incrementally shifted by a factor 8 for
better readability; data for the pure polymer (Z = 0) are directly at
scale. Continuous lines are best ts.

curve for the pure copolymer has very low intensity in the
whole q-range (and can be modeled reasonably by the curve for
polymer coils; see later), the addition of a small amount of
DPCl corresponding to 1 mM, i.e., to Z = 0.15 (well below cmc
which is 9.8 mM in 50 mM sodium tetraborate, as determined
by ITC),38 induces the intensity increase in the low q range
nearly by 2 orders of magnitude. The high q regime remains
less aected. This fact indicates the formation of a small
fraction of very large polymersurfactant aggregates coexisting
with free coils. When discussing and interpreting the data, it is
necessary to keep in mind that the excess scattering length
density of the surfactant is higher than that of the polymer
which partially explains the observed intensity increase at low q.
Nevertheless, it also means that the inhomogeneities monitored
by SANS are mainly the uctuations of structural patterns
created by surfactant micelles interacting with (and attached to)
polymer chains as will be supported by other data and their
careful analysis later.
An inspection of SANS data in the low Z range up to 0.6
reveals that the shapes and mainly the changes of individual
curves with Z are slightly surprising. The high q part of the
curve for Z = 0.38 resembles that for pure solution of surfactant
micelles (above cmc), suggesting that the scattering behavior at
short length scales is dominated by disordered surfactant
micelles attached to coiled chains. However, in a low q range,
the experimental intensity rises steeply, indicating the presence
of large patterns created by the surfactant micelles in large and
6475

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Macromolecules

Article

been used;39 for these micelles, the form factor and parameters
used were taken from the t of pure concentrated DPCl
micelles, with no modication.
Upon further addition of surfactant, the scattering from large
spherical objects is observed and a correlation peak appears at
high q. The strongly scattering coreshell nanoparticles with
cores formed by densely packed surfactant micelles are
modeled as follows: The form factor for densely packed
micelles is the coreshell sphere model (eq 7 in Supporting
Information) with Nagg = 47 (aggregation number of surfactant
molecules in the DPCl micelles) as obtained for pure DPCl
micelles (see below). Repulsive interactions between micelles
are evaluated using the simple hard-sphere potential40 (see
Supporting Information), assuming that electric charges are
essentially neutralized.
For the form factor for the whole nanoparticle, the core
shell model has been used (see Supporting Information). For Z
0.91, we assume that all polymer chains participate to the
formation of nanoparticles. The core is composed of densely
packed surfactant micelles that neutralize the charge at PMAA
chains and solvent molecules, and the shell is composed of
PEO moiety and the solvent. The NP core radius and the
number density of micelles in the core determine the
composition of the core, from which the contrast is calculated.
The experimental SANS curves are relatively smooth (as
compared with theoretical curves for the assumed monodisperse model species) due to slight polydispersity in sizes,
which is implemented in the model as a Gaussian distribution
of core radii.
Finally, for Z = 3.6 the SANS data indicate the presence of
free (excess) surfactant micelles in the solution. They are tted
by the same model as pure surfactant micelles in the buer (Z =
), i.e., biaxial oblate ellipsoids with Nagg = 47 and core halfaxes of 1 2 2 nm3 (see Supporting Information). The shell
thickness has been found to be 0.55 nm. The equivalent
spherical micelle (with same aggregation number) has a core
radius of 1.60 nm and a shell thickness of 0.58 nm; this
simplied model is used for densely packed micelles, while the
biaxial model is used for other micelles (either in clusters or as
excess micelles).
To summarize the analysis of the SANS curves, the data from
mixtures of PEO705PMAA476 and DPCl are modeled in the
absolute scale by the sum of several contributions (maximum
5). However, individual terms are relevant (i.e., they are
nonzero) only in some regions of Z, which are specied in
previous parts and are outlined in Scheme 1.

Scheme 1. Representation of the Suggested Self-Assembled


Structures in Solution as a Function of the Molar Ratio Za

(a) Pure polymer coils, (b) loose aggregates of the surfactant bound
electrostatically to the PMAA block, (c) beads-on-a-string nanoparticles with a core formed by low density nonordered micelles, (d)
coreshell nanoparticles with a core formed by densely packed
ordered micelles, and (e) coreshell nanoparticles coexisting with
excess free surfactant micelles.

not been reproduced, and we will further use the value given by
the manufacturer, i.e., PMAA: 41 kg mol1 and PEO: 31 kg
mol1. However, the modeling conrms that the copolymer
dissolves fairly well under given conditions (i.e., in a sodium
tetraborate buer at pH 9.3 when the carboxylic groups are
ionized and the electric charge is eciently screened by small
ions) and acquires the shape of an almost unperturbed
Gaussian coil and only a small fraction of chains aggregate.
With the rst addition of the surfactant (Z = 0.15), the curve
starts to change, but it does not show any distinct features so
far. It can be modeled assuming the presence of a very low
amount of large loose aggregates interconnected by surfactant
micelles (formed as a result of electrostatic polymersurfactant
interaction) with the size larger than that of pure DPCl, which
reects the fact that they are wrapped by parts of PMAA chains.
This yields on one side additional material to the micelle, but
more importantly leads to a pronounced electrostatic shielding
and thereby to a micellar growth as it would correspondingly
also be observed in a solution of higher ionic strength. The next
curve at Z = 0.38 supports this working hypothesis. Its shape
can be described by the scattering from large aggregates
stabilized by micelles with a larger size than those determined
previously (aggregation number 116 as compared with 47 for
pure DPCl). In agreement with other experimental data (e.g.,
cryo-TEM images), the curve for Z = 0.38 can be interpreted by
a model that assumes small clusters of micelles attached to
polymer chains (see also Scheme 1 and Supporting Information
Figure S3), and hence the beads-on-a-string structure factor has

I(q) = Iinc + Icluster(q) +

R [INP(q) + Im,NP(q)](Rc)
c

dR c + I xs(q)

(9)

where Iinc accounts for q-independent incoherent scattering


(and is xed to the value determined by a Porod analysis at
high q), Icluster(q) (used up to Z = 1) is the contribution from
micelles attached on polymer chains, modeled by small clusters
of coreshell ellipsoids (beads-on-a-string), where the only
parameter tted is the fraction of surfactant (the number of
micelles per string was tted once for Z = 0.38 and then kept
constant due to the lack of information resulting from the large
intensity contribution from nanoparticles for larger Z), INP(q) is
the contribution from the nanoparticles, modeled by coreshell
spheres with a homogeneous polydisperse core composed of
the PMAA moiety, solvent, and surfactant and a homogeneous
6476

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Macromolecules

Article

shell composed of solvated PEO, Im,NP(q) is the contribution


from densely packed DPCl micelles in the core of the
nanoparticles modeled by coreshell spheres interacting
through a simple hard-sphere potential, their size distribution
is described by SchulzZimm distribution function, (Rc), and
Ixs(q) (used for Z > 1) is a contribution from free micelles in
excess modeled with biaxial coreshell ellipsoids interacting
through electrostatic repulsions. The individual scattering
contributions can be expressed in the form (here we have
used the decoupling approximation for form and structure
factor (P(q) and S(q)) that is strictly valid only for the
monodisperse case but is conventionally also employed for
polydisperse cases of colloidal systems)42
Icluster(q) = NclusterVcluster 2cluster2P ellips(q)S bead(q)

(10)

INP(q) = NNPVNP 2NP2P CS,NP(q)

(11)

Im,NP(q) = Nm,NPVm,NP 2m,NP2P CS,m(q)SHS(q)

(12)

Im,xs(q) = Nm,xsVm,xs 2m,xs2P ellips(q)Scharged(q)

(13)

hydrodynamic radii obtained by light scattering. Indeed, the


actual shell thickness has little inuence on the SANS model
beyond a certain hydration. The number of polymer chains in
the shell is deduced from the total concentration of polymer
and from the number density of particles. The model curves are
shown together with experimental data on Figure 3, and the
decomposition of the model curve into individual contributions
is exemplied in Figure 4, which shows that the dierent tted

where N, V, and are the density numbers, volumes, and


average contrasts and indices cluster, NP, m,NP, and
m,xs respectively correspond to the clusters, nanoparticles,
micelles embedded in the nanoparticles, and excess surfactant
micelles; Sbead(q) is the beads-on-a-string structure factor,
SHS(q) is the hard-sphere structure factor for the densely
packed micelles, Scharged(q) is the structure factor of charged
spheres for the ellipsoidal micelles in excess, PCS,NP(q) and
PCS,m(q) respectively are the coreshell sphere form factors for
the nanoparticles and the micelles (here assumed spherical)
embedded in the nanoparticles, and Pellips(q) is the form factor
for the biaxial coreshell ellipsoidal micelles. The expression
for all structure factors and form factors can be found in the
Supporting Information.
Obviously, the hard-sphere structure factor is not perfect for
modeling the cores formed by densely packed micelles, which
are probably neither spherical nor neutral and their ordering
and packing increases with Z (see SAXS patterns later).
However, SANS is a relatively low-resolution technique, in
particular in this q-region with a rather large wavelength
smearing, so the details are suppressed anyway and we prefer as
simple and transparent model as possible. The number density
of micelles in the core of the nanoparticles was therefore xed
in all ts so that the bump due to the smeared peak of the
structure factor describes reasonably well the position of the
experimentally observed peak. The t is then performed in a q
range up to 1 nm1, i.e., before the inner structure of the core
starts to inuence experimental data.
There are nally only four parameters to be tted: the
average core radius Rc and standard deviation of the
nanoparticle size distribution s/Rc, the concentration of
surfactant participating to the nanoparticles, and to the micelles
outside the nanoparticles (corresponding to Ncluster or Nm,xs
depending on Z); the dierence between the known
concentration of surfactant and the sum of these two
concentrations from the ts is then interpreted as a cmc. We
recall that for Z 0.91 all the polymer molecules are assumed
to be part of the nanoparticles (this simplifying assumption
reects the charge compensation). The PEO shell thickness is
xed to a value of 25 nm so that the overall dimensions
determined by SANS are in a good agreement with

Figure 4. Decomposition of the t for Z = 1.1 into elementary


scatterers: the large and polydisperse coreshell sphere for nanoparticles, coreshell spheres with a structure factor for micelles
densely packed in the core of NP, and biaxial ellipsoids for micelles not
included in the core of the NP. Contributions are not in absolute scale
for readability.

parameters are obtained from dierent scattering features at


dierent q-ranges; the output parameters are listed in Table 2.
The outcome of the ts is consistent with the approximate
results provided by the model-free analysis, which indicate that
the ts are meaningful, in spite of the assumptions and
simplications required by the complexity of the systems. The
reliability of the parameters deduced from the ts is further
supported by their being physically realistic with continuous
evolution as a function of Z.
As the absolute scale has been used in all ts, the amount of
the surfactant participating to micelles is known and we can
evaluate the concentration of free surfactant by subtraction,
which is increasing with Z up to 7.4 mM (for Z = 3.6). The cmc
for pure surfactant in 50 mM sodium tetraborate determined by
ITC and is 9.8 mM,38 which corroborates further the SANS
analysis. The parameters resulting from the ts indicate a very
low density of PEO chains at the surface of the nanoparticles, in
the order of 60 nm2 per chain (increasing with Z from 30 up to
70 nm2).
In general, we can summarize that the detailed modeling
described above fully conrms the behavior of the system
depicted in Scheme 1.
SAXS Measurements. In order to gain complementary
information and higher q resolution, additional SAXS experiments were done. Because of their high resolution, they provide
deeper insight in the internal structure of PEO705PMAA476/
DPCl nanoparticles and enable to investigate packing of
surfactant micelles in the nanoparticle core. It was found that
the ordering of surfactant micelles in the core depends both on
the surfactant-to-polyelectrolyte molar ratio and on PEO705
PMAA476 concentration.
6477

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Macromolecules

Article

Table 2. Outcome from the Fits of SANS Data from CoreShell Nanoparticles
Z

Zca

Cm,xs,b mM

Cm,NP,c mM

Rc,d nm

s/Rce

Nagg,polyf

Nagg,micg

aPEO,h nm2

0.91
1.13
1.44
1.81
3.63

0.63
1.13
1.37
1.54
1.92

0.77
0.24
0.08
0.06
1.67

3.9
6.9
8.3
9.1
10.2

48
46
50
57
62

0.22
0.15
0.15
0.16
0.17

1060
482
513
682
706

6799
5540
7164
10692
13842

29
57
63
62
71

Surfactant-to-MMA molar ratio in the core of the nanoparticles. bConcentration of the surfactant outside the nanoparticle core (due to
interdependencies, the error is actually close to 0.1 mM). cConcentration of the surfactant in densely packed micelles in the NP (error up to 3 mM).
d
Average core radius (typical error: 0.03 nm). eRelative standard deviation of Rc. fAggregation number of polymer chains in one nanoparticle (error
up to 218 for Z = 1.81, in other cases much lower). gAggregation number of the micelles in one nanoparticle (error up to 16). hArea per polymer
chain at the core/shell interface of the nanoparticle (error up to 20 nm2 for Z = 1.81, in other cases much lower). Evaluation of uncertainties: see
Supporting Information.

Figure 5 shows SAXS curves for a series of PEO705


PMAA476/DPCl complexes in the q range 0.52.5 nm1 (the

(Figure S4). This result indicates that the liquid crystalline


structure of the core is kinetically frozen and remains preserved
after dilution of the sample.
Cryo-Transmission Electron Microscopy (Cryo-TEM).
Cryo-transmision electron microscopy was performed on NP
solutions in 50 mM sodium tetraborate D2O. The nanoparticles
were prepared at copolymer concentration 5 g L1 and DPClto-MAA ratios Z = 0.15, 0.38, and 1. Unfortunately, lower
concentrations of the copolymer (comparable with those used
for scattering experiments, ca. 1 g L1) did not provide good
quality cryo-TEM micrographs (Talmon et al. reported similar
problems for aggregates formed by short chain surfactants and
polyelectrolytes.)21 Figure 6 represents typical cryo-TEM
images of nanoparticles obtained at Z = 0.15 (a), 0.38 (b),
and 1 (c). At Z = 0.15 when the surfactant concentration is
lower than cmc of DPCl, the interaction of the surfactant with
PE induces the formation of micelles and their electrostatic
binding to charged PMAA block.41 A low fraction of
polydisperse aggregates form in which the elongated structures
could be found as indicated by arrows. For a further discussion,
it is important to realize that the polymer chains (both blocks)
are almost invisible, and only the surfactant micelles and their
assemblies are clearly seen on the images. As more surfactant is
added and its concentration approaches the cmc (Z = 0.38 and
cDPCl = 2.4 mM), fairly uniform beads-on-a-string structures are
formed Figure 6b. Since all experiments were done in a dilute
region, the elongated structures could be depicted as a few
spherical DPCl micelles electrostatically interacting with
PEO705PMAA476 chains. Those elongated structures grow
until Z ca. 0.6, when the electrostatic repulsion between
noncompensated negative charges on PMAA chains is
overbalanced by hydrophobic interactions and more compact
coreshell structures start to appear in the solution. When the
negative and positive charge is matched at Z = 1, a massive
aggregation occurs and quite dense coreshell nanoparticles
with an average diameter of 120 20 nm are formed (Figure
6c). The size estimated by cryo-TEM is in a good agreement
with LS and SANS results. The stabilizing PEO blocks are in
good solvent condition and form the micellar shell that is
invisible on the cryo-TEM images.

Figure 5. SAXS curves of PEO705PMAA476/DPCl complexes in the q


range 0.52.5 nm1. Copolymer concentrations were (a) 1 and (b) 5 g
L1. Surfactant-to-polyelectrolyte units molar ratios, , are indicated
above the corresponding curves.

region of the correlation peak revealed by SANS). At low


PEO705PMAA476 concentration (c = 1 g L1), the surfactant
micelles are disordered up to Z ca. 2, showing a single broad
correlation peak corresponding to the characteristic distance of
DPCl micelles in the core. For higher Z ratios, scattering curves
contain two overlapping peaks with the maxima at 1.85 and
1.62 nm1. The ratio of maxima positions is fairly close to
41/2:31/2, suggesting the fcc cubic packing of the DPCl micelles
in the core. That a signicantly higher value of Z than 1 is
required for seeing such well-dened peaks could be explained
such that the DPCl is carrying some of its Cl counterions into
the copolymer aggregates. Accordingly then, a larger number of
it can be become incorporated than foreseen for Z = 1 and once
the packing density of DPCl micelles in the copolymer
aggregate interior surpasses a certain concentration this
transition to a highly ordered state is observed. Increase of
the surfactant concentration increases uidity of the nanoparticles core and allows for better accommodation both
surfactants and polymer chains. The measurement at high
PEO705PMAA476 concentration (c = 5 g L1) revealed that the
transition of surfactant micelles in the core to the ordered state
occurs at Z ca. 1. In this case three scattering peaks appeared
with the maxima at 1.81, 1.66, and 1.49 nm1, i.e., close to
61/2:51/2:41/2, which corresponds to the Pm3n cubic packing. It
is noteworthy that the well-resolved scattering peaks of the
Pm3n cubic structure were still visible when the sample was
diluted down to the copolymer concentration of 0.3 g L1

CONCLUSIONS
In this paper, we have shown by combining information from
light scattering and SANS that mixing the copolymer PEO705
PMAA476 with DPCl in aqueous solution leads to the formation
of dierent states of self-assembled structures, depending on
the value of the charge ratio, Z, which can be subdivided into
four dierent regions:
6478

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Macromolecules

Article

Figure 6. Cryo-TEM images of nanoparticles formed at the stoichiometric ratio, Z, of DPCl to PMAA units (a) 0.15, (b) 0.38, and (c) 1 and the
copolymer concentration, c = 5 g L1; the arrow in (a) indicates the very rst elongated beads-on-a-string structures. Scale bars = 100 nm.

(i) at low Z (Z ca. 0.15), a low fraction of large elongated


copolymer aggregates form; the aggregates are swollen and the
chains are connected by surfactant oligomers, the formation of
which (below cmc) is induced by electrostatic interaction with
units of PMAA blocks; the aggregates coexist with free
copolymer coils; (ii) at higher Z (Z ca. 0.38), the fraction of
aggregates increases, but their size slightly diminishes due to the
screening and overall weakening of the electrostatic repulsion
between the ionized PMAA units by the surfactant micelles of
the opposite charge bound to the chains which partially
neutralize electric charges; the aggregates still coexist with free
copolymer chains which are partially decorated by surfactant
micelles and form quite random coils; (iii) around Z = 0.6, the
solubility of the copolymersurfactant aggregates formed on
PMAA blocks is low and because the solubility of PEO is
unaected, relatively compact coreshell aggregates start to
form; their cores are formed by densely packed and ordered
surfactant micelles attached to PMAA chains and are protected
by soluble PEO shells; at rst they coexist with a fraction of
large aggregates, but close to Z = 1.0, all polymer chains and
surfactant molecules are engaged in compact aggregates; (iv) at
Z 1, the coreshell aggregates partially solubilize DPCl
surfactant micelles (together with corresponding amount of Cl
counterions) and coexist with a certain excess of surfactant
molecules in the solution and at concentrations above the cmc
also with surfactant micelles.
For the coreshell nanoparticles formed around Z 1
(regions iii and iv), with a thin PEO shell and a core containing

PMAA interacting with oppositely charged surfactant micelles,


it was found by SAXS experiments that the ordering of DPCl
micelles in the core depends both on surfactant-to-polyelectrolyte molar ratio and the total concentration, becoming more
pronounced with increasing concentration. A fcc and Pm3n
cubic packing of the DPCl micelles in the core were found in
systems with Z > 2 and polymer concentration 1 g L1 and with
Z > 1 and 5 g L1, respectively.
In summary, this investigated PEO705PMAA476/DPCl
system shows a rich structural variety that can be controlled
in a systematic fashion by the mixing ratio characterized by Z
and by the total concentration. This renders it interesting as a
rather versatile system for purposes of solubilizing and
delivering active agents with them.

ASSOCIATED CONTENT

S Supporting Information
*

Further pieces of information on apparent volumes and


scattering length densities, hard-sphere structure factor, model
free analysis of SANS, data ts for nanoparticles and for pure
DPCl micelles in D2Obuer, and uncertainties on tted
parameters and SAXS data for dilution experiment. This
material is available free of charge via the Internet at http://
pubs.acs.org.

AUTHOR INFORMATION

Corresponding Author

*E-mail: mariuszuchman@go2.pl.
6479

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Macromolecules

Article

Notes

(25) Tararyshkin, D.; Kramarenko, E.; Khokhlov, A. R. J. Chem. Phys.


2007, 126, 164905.
(26) Dubochet, J.; Adrian, M.; Chang, J. J.; Homo, J. C.; Lepault, J.;
McDowall, A. W.; Schultz, P. Q. Rev. Biophys. 1988, 21, 129.
(27) Berne, B. J.; Pecora, R. Dynamic Light Scattering: With
Applications to Chemistry, Biology and Physics, reprinted ed.; Dover
Publications: Mineola, NY, 2000.
(28) Panine, P.; Finet, S.; Weiss, T. M.; Narayanan, T. Adv. Colloid
Interface Sci. 2006, 127, 9.
(29) Lewis, R. A.; Berry, A.; Hall, C. J.; Helsby, W. I.; Parker, B. T.
Nucl. Instrum. Methods Phys. Res., Sect. A 2000, 454, 165.
(30) http://www.diamond.ac.uk/Home/Beamlines/I22/tech/
detectors.html/.
(31) http://www.esrf.eu/computing/scientic/FIT2D/.
(32) Keiderling, U. Appl. Phys. A: Mater. Sci. Process. 2002, 74, 1455.
(33) Sedlak, M. Langmuir 1999, 15, 4045.
(34) Sedlak, M.; Konak, C .; Stepanek, P.; Jakes, J. Polymer 1987, 28,
873.
(35) Volk, N.; Vollmer, D.; Schmidt, M.; Oppermann, W.; Huber, K.
Adv. Polym. Sci. 2004, 166, 29.
(36) Groehn, F.; Antonietti, M. Macromolecules 2000, 33, 5938.
(37) Matejcek, P.; Stepanek, M.; Uchman, M.; Prochazka, K.;
Sprkova, M. Collect. Czech. Chem. Commun. 2006, 71, 723.
(38) Uchman, M.; Gradzielski, M.; Angelov, B.; Tosner, Z.;
Prochazka, K.; Stepanek, M., to be published.
(39) Burchard, W.; Kajiwara, K. Proc. R. Soc. London, A 1970, 316,
185.
(40) Ashcroft, N. W.; Lekner, J. Phys. Rev. 1966, 145, 83.
(41) Hansson, P. Langmuir 2001, 17, 4167.
(42) Chen, S. H. Annu. Rev. Phys. Chem. 1986, 37, 51.

The authors declare no competing nancial interest.

ACKNOWLEDGMENTS
The authors acknowledge the nancial support from the
Ministry of Education of the Czech Republic (long-term
Research Project No. MSM0021620857) and the Grant Agency
of the Czech Republic (Grants P208/10/0353, P208/12/P236,
P205/11/J043, and P106/12/0143) and German Academic
Exchange Service DAAD (Grants 2B08021 and D0804221
PPP-CZ-09-10, PKZ: 50016729). This research project has
been supported by the European Commission under the 7th
Framework Programme through the Key Action: Strengthening
the European Research Area, Research Infrastructures.
Contract: 226507 (NMI3). For allocation of beam time we
are grateful to ILL (Grenoble, France), ESFR (Grenoble,
France), and Diamond (Didcot, UK). B.A. acknowledges user
support from the Diamond Light Source (Didcot, Oxfordshire,
UK; Proposal SM3313, Beamline I22) and ESRF (Grenoble,
France, Proposal SC3113, Beamline ID02) and thanks Drs. S.
Filippov, P. Stepanek, N. Terrill, J. Gummel, and T. Narayanan
for cooperation and support.

REFERENCES

(1) Voets, I. K.; de Keizer, A.; Stuart, M. A. C. Adv. Colloid Interface


Sci. 2009, 147148, 300.
(2) Langevin, D. Adv. Colloid Interface Sci. 2009, 147148, 170.
(3) Zhou, S.; Chu, B. Adv. Mater. 2000, 12, 545.
(4) Li, Y.; Xia, J.; Dubin, P. L. Macromolecules 1994, 27, 7049.
(5) Hoffmann, I.; Heunemann, P.; Prevost, S.; Schweins, R.; Wagner,
N. J.; Gradzielski, M. Langmuir 2011, 27, 4386.
(6) Stepanek, M.; Matej cek, P.; Prochazka, K.; Filipov, S. K.;
Angelov, B.; Slouf, M.; Mountrichas, G.; Pispas, S. Langmuir 2011, 27,
5275.
(7) Balomenou, I.; Bokias, G. Langmuir 2005, 21, 9038.
(8) Mahendra, S. B.; Surekha, D. Adv. Colloid Interface Sci. 2006,
123126, 387.
(9) Pispas, S. Soft Matter 2011, 7, 474.
(10) Pispas, S. Soft Matter 2011, 7, 8697.
(11) Wang, Y.; Han, P.; Xu, H.; Wang, Z.; Zhang, X.; Kabanov, A. V.
Langmuir 2009, 26, 709.
(12) Kabanov, A. V.; Bronich, T. K.; Kabanov, V. A.; Yu, K.;
Eisenberg, A. J. Am. Chem. Soc. 1998, 120, 9941.
(13) Bronich, T. K.; Popov, A. M.; Eisenberg, A.; Kabanov, V. A.;
Kabanov, A. V. Langmuir 2000, 16, 481.
(14) Bronich, T. K.; Kabanov, A. V.; Kabanov, V. A.; Yu, K.;
Eisenberg, A. Macromolecules 1997, 30, 3519.
(15) Stepanek, M.; Skvarla, J.; Uchman, M.; Prochazka, K.; Angelov,
B.; Kovaci k, L.; Garamus, V. M.; Mantzaridis, Ch.; Pispas, S. Soft
Matter 2012, DOI: 10.1039/C2SM25588J.
(16) Berret, J.-F.; Herve, P.; Aguerre-Chariol, O.; Oberdisse, J. J. Phys.
Chem. B 2003, 107, 8111.
(17) Berret, J.-F.; Oberdisse, J. Physica B 2004, 350, 204.
(18) Courtois, J.; Berret, J.-F. Langmuir 2010, 26, 11750.
(19) Berret, J.-F.; Vigolo, B.; Eng, R.; Herve, P.; Grillo, I.; Yang, L.
Macromolecules 2004, 37, 4922.
(20) Nizri, G.; Magdassi, S.; Schmidt, J.; Cohen, Y.; Talmon, Y.
Langmuir 2004, 20, 4380.
(21) Nizri, G.; Makarsky, A.; Magdassi, S.; Talmon, Y. Langmuir
2009, 25, 1980.
(22) Kogej, K. J. Phys. Chem. B 2003, 107, 8003.
(23) Trabelsi, S.; Guillot, S.; Raspaud, E.; Delsanti, M.; Langevin, D.
Adv. Mater. 2006, 18, 2403.
(24) Yeh, F.; Solokov, E. L.; Khokhlov, A. R.; Chu, B. J. Am. Chem.
Soc. 1996, 118, 6615.
6480

dx.doi.org/10.1021/ma301510j | Macromolecules 2012, 45, 64716480

Potrebbero piacerti anche