You are on page 1of 9

32 Biomacromolecules 2009, 10, 32–40

Micelles and Polymersomes Obtained by Self-Assembly of


Dextran and Polystyrene Based Block Copolymers
Clément Houga,† Joanna Giermanska,‡ Sébastien Lecommandoux,† Redouane Borsali,§
Daniel Taton,† Yves Gnanou,*,† and Jean-François Le Meins*,†
Université de Bordeaux, Laboratoire de Chimie des Polymères Organiques UMR5629, ENSCPB-CNRS,
16 avenue Pey Berland, 33607, Pessac cedex, France, Université de Bordeaux, Centre de Recherche
Paul Pascal CNRS-UPR 8641, Avenue Albert Schweitzer, 33600 Pessac, France, and Centre de
Recherche sur les Macromolécules Végétales CERMAV and Joseph Fourier University,
BP53, 38041, Grenoble Cedex 9, France
Received March 25, 2008; Revised Manuscript Received November 10, 2008

The self-assembly of dextran-block-polystyrene (dex-b-PS) block copolymers was investigated in solution. The
hydrophobic PS weight fraction in these block copolymers ranges from 7 to 92% w/w, whereas the average
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

number molar mass of dextran was kept constant at 6600 gmol-1. Self-assembly by direct dissolution in water
could be performed only for block copolymers with a low hydrophobic content (7% w/w), whereas mixtures of
Downloaded via UNIV OF SAO PAULO on October 22, 2022 at 19:58:34 (UTC).

tetrahydrofuran and dimethylsulfoxide were required for higher PS content, before transferring the structures into
water. Core-shell micelles, ovoı̈ds, and vesicles could be identified upon characterization by light and neutrons
scattering, atomic force microscopy, and transmission electron microscopy. Most of the morphologies observed
were not expected considering the chemical composition of the block copolymers. Finally, the size and shape of
these nanoparticles were fixed upon cross-linking the dextran block through reaction of the hydroxyl groups with
divinylsulfone. The role of the dextran conformation on the self-assembly process is discussed.

Introduction micelles, while those with f < 25% self-assemble into inverted
nanostructures. Block copolymers with hydrophilic weight
The self-assembly of block copolymers in a selective solvent fraction around 35 ( 10% are predicted to form polymeric
provides a great variety of morphologies such as spherical and vesicles also referred to as polymersomes.
wormlike micelles, vesicles, and numerous other microstruc-
Nanostructures formed by self-assembly in solution of block
tures. Such a process is the result of two driving forces that
copolymers make them attractive materials in various applica-
interplay oppositely: on the one hand, long-range repulsive
tions, in particular, as nanocarriers for the encapsulation and
interactions between incompatible blocks, and on the other hand,
controlled release of drugs. For example, polyester-based block
short-range attractive interactions due to the chemical bond
copolymers are the subject of numerous studies12-17 aiming at
linking the two blocks, which leads to microphase rather than
the development of biomaterials or drug delivery systems.
macrophase separation. The morphologies observed at thermo-
dynamic equilibrium result from the minimization of the free In addition to providing abundant sources of raw materials,
energy of the self-assembled structures and are controlled by naturally occurring oligosaccharides and polysaccharides exhibit
molecular parameters such as the chemical nature of the blocks, attractive properties such as biodegradability, biocompatibility,
the volume fraction of each block, the molar mass and the texturing, or gelifying properties. Harnessing these properties
overall architecture of the copolymer.1 Solution parameters such in nanostructures obtained from the self-assembly of block
as polymer concentration, temperature, solvent quality, pH, and copolymers could be of great interest for numerous applications
ionic strength2 also play a crucial role on the self-assembly. from cosmetics to medicine. Pioneer work in this field has been
Finally, the dissolution process is another key parameter that performed in the early 1980s by Ziegast and Pfannemüller who
impacts morphological aspects of the nanoparticles formed.3-6 coupled an oligosaccharide functionalized with an aldonolactone
The self-assembly phenomenon in solution is therefore ex- at its reducing end to an R,ω-diamino poly(ethylene oxide)
tremely complex; numerous theoretical works have been devoted (PEO).18 In recent years, other routes, though scarce, have been
to the subject7-10 as well as experimental studies, which have proposed to develop new “hybrid” polysaccharide-based block
helped to clarify the micellization process and the way to copolymers. A first strategy is based on the coupling of an
generate stable morphologies.3,5,11 amino-terminated synthetic polymer to the terminal aldehyde
Discher and Eisenberg11 proposed a general empirical law of an oligosaccharide,19-23 possibly followed by the enzymatic
which stipulates that block copolymers possessing a hydrophilic growth of the oligosaccharide block.21,24,25 A second route starts
weight fraction (f) >45% are expected to form spherical from the chemical modification of the reducing ends to
subsequently grow the synthetic block by means of a controlled/
* To whom correspondence should be addressed. Tel.: 00 33 5 57 00 36 living polymerization. In this method, sugar derivatives or
96. Fax: 00 33 5 57 40 00 84 87. E-mail: lemeins@enscpb.fr (J.F.L.M.); cyclodextrine are derivatized into macroinitiators of controlled
gnanou@enscpb.fr (Y.G.).

radical polymerization process26,27 to create “end sugar” func-
Laboratoire de Chimie des Polymères Organiques UMR5629. tionalized synthetic polymers. Recently, our group has reported

Centre de Recherche Paul Pascal CNRS-UPR 8641.
§
Centre de Recherche sur les Macromolécules Végétales CERMAV and a methodology to synthesize dextran-block-polystyrene (dex-
Joseph Fourier University. b-PS) diblock copolymer by atom transfer radical polymerization
10.1021/bm800778n CCC: $40.75  2009 American Chemical Society
Published on Web 12/11/2008
Micelles and Polymersomes Biomacromolecules, Vol. 10, No. 1, 2009 33

Table 1. Molecular Characteristics of the Block Copolymers Used used were of technical grade. Divinyl sulfone (Aldrich) was used
in this Study without purification. The following nomenclature was used for the
sample Mn SECa DPn NMRb PDIa ΦPS %w/w different block copolymers: dexa-b-PSb, where a and b stand for the
degree of polymerization of dextran and polystyrene, respectively.
dex40-b-PS5 17500 5 1.4 7
dex40-b-PS270 82200 270 1.7 81
Preparation of the Self-Assembled Nanoparticles. The direct
dex40-b-PS775 160000 775 1.9 92 dissolution of dex40-b-PS5 in water was achieved after stirring for two
a
days at 90 °C, followed by sonication. The direct dissolution in water
Determined by SEC in THF (calibration with PS standards). b Overall
composition determined by 1H spectroscopy, knowing the molar weight
for systems with a larger PS content was just not possible. As no
of commercial dextran (Mn ) 6600 g · mol-1). common solvent for both PS and dextran could be found, a mixture of
THF and DMSO was used to dissolve these diblock copolymers, DMSO
being selective for the dextran block and THF for the PS one. This
(ATRP).28 From a synthetic viewpoint, most of reported works
method is slightly different from the procedure generally used that
take advantage of the terminal aldehyde of the oligosaccharide
consists in the dissolution of the block copolymer in a common solvent
that is in equilibrium with a hemiacetal function, except in a
for both blocks, the self-assembly being induced by adding a selective
few references.21,24 However, the number of saccharide based solvent for one of the blocks.2,6,33-35 All dex-b-PS block copolymer
block copolymers systems that have been thoroughly investi- samples could be dissolved in DMSO/THF mixtures in a composition
gated is limited and among existing studies only a few contain range from 50 to 75% in DMSO for dex40-b-PS270 and from 30 to 45%
informations about their solution behavior. For instance, Liu in DMSO for dex40-b-PS775 (dotted lines in Figure 3). In these
and Zhang20 have shown by transmission electron microscopy compositions, no clear scattered signal was detected, indicating that
(TEM) and dynamic light scattering (DLS) that dextran-b-poly- the block copolymers behave as unimers. Then, the solvent composition
(ε-caprolactone) self-assembles in water into polydisperse was changed by addition of either DMSO or THF to induce self-
micellar structures with an average diameter of 100 nm. pH- assembly and engineer different morphologies.
responsive carboxymethyl dextran-b-poly(ethylene oxide) were Cross-Linking Procedure of the Nanoparticles. Divinyl sulfone
also shown to self-assemble in aggregates with an average size (DVS) was added under nitrogen to dex40-b-PS270 dissolved in a DMSO/
of 100 nm in acidic media by intra- and intermolecular hydrogen THF mixture with a composition of 40/60 to cross-link dextran in the
bonding.22 The self-assembly of PEO-b-amylose in chloroform core of the nanoparticles or in a DMSO/THF mixture with a composi-
results in the formation of reverse micelles in which the amylose tion of 90/10 to cross-link dextran in the shell of the nanoparticles. A
block is believed to adopt a helical structure.29 The behavior of large excess (>10 equiv) of DVS was used relative to the hydroxyl
sugar end-functionalized PS (from glucose to maltohexaose) has groups present in the dextran chains. The reaction was performed at
also been studied in toluene.30 Reverse micelles were obtained room temperature during 24 h.
with an aggregation number varying with the overall composi- Atomic Force Microscopy (AFM). Samples for AFM analysis were
tion of the block copolymer. In a recent addition, the self- prepared by evaporation at room temperature on substrates starting from
assembly of amylose-b-PS in THF revealed that nonequilibrium water solutions. Practically, for aqueous media, 20 µL of a dilute
aggregates with different sizes were obtained,31 whereas crew solution (∼5 mg/L) was dropped on a 1 × 1 cm2 freshly cleaved mica
cut micelles with a diameter ranging from 10 to 30 nm were substrate. Samples were analyzed after complete evaporation of water
observed in water. In the latter case, the amylose corona was at room temperature. For solution in THF, 20 µL of a dilute solution
found to exhibit a rod-like structure. It is important to note that (∼0.5 mg/L) was spin-coated on freshly cleaved HOPG. For solution
depending on their chemical nature, polysaccharide chains can in DMSO, 20 µL (∼5 mg/mL) were dropped on HOPG surfaces and
present a rod conformation (Xanthan, amylose,...) or a coil the evaporation of the solvent was performed under dynamic vacuum.
AFM images were recorded in air with a dimension microscope (Digital
conformation like, for instance, pullulan or dextran, although
Instruments, Santa Barbara, CA) operated in tapping mode. The probes
the latter one behaves as a rod for molar mass below 2000
were commercially available silicon tips with a spring constant of 40
g/mol.32 This coil or rod character can have a marked influence
N/m, a resonance frequency lying in the 270-320 kHz range and a
on the self-assembly process, as it will be discussed later in the
radius of curvature in the 10-15 nm range. In this work, both the
paper.
topography and the phase signal images were recorded with the highest
From these findings, it appears difficult to establish general sampling resolution available, that is, 512 × 512 data points.
principles with respect to the self-assembly of polysaccharide Dynamic and Static Light Scattering. Dynamic and static light
based block copolymers. For instance, the influence of the helical scattering measurements were performed using an ALV laser goniom-
structure of a polysaccharide such as amylose on the self- eter, which consists of a 22 mW HeNe linear polarized laser operating
assembly process is not yet clearly understood. In the present at a wavelength of 632.8 nm and an ALV-5000/EPP multiple τ digital
report, we describe the self-assembly properties of a series of correlator with 125 ns initial sampling time. The copolymer solutions
dex-b-PS block copolymers in organic solvent but also in water. were maintained at a constant temperature of 25.0 ( 0.1 °C in all
The degree of polymerization (DP) of the polysaccharide was experiments. All the scattering measurements were carried out from
kept constant (∼40) but that of PS blocks was varied from 5 to 40 to 120° by steps of 10°. Data were collected using ALV Correlator
770. The synthesis of these hybrid block copolymers has been Control software, and the counting time was fixed at 300 s for each
already described in a previous report.28 The influence of the angle. Dynamic light scattering measurements were evaluated by fitting
block copolymer composition on the micellar structures formed of the measured normalized time autocorrelation function of the
has been thoroughly investigated by means of dynamic and static scattered light intensity. The data were fitted with the help of the
light scattering, atomic force microscopy and freeze fracture constrained regularization algorithm (CONTIN), which provides the
transmission electronic microscopy. We discuss at the end of distribution of relaxation times τ, A(τ), as the inverse Laplace transform
this work the possible influence of the dextran block on the of g(1)(t) function
self-assembly process.


g(1)(t) ) A(τ) exp(-t ⁄ τ)dτ (1)
0
Experimental Section Apparent diffusion coefficients D were obtained by plotting the
Materials. Molecular features of dex-b-PS diblock copolymers relaxation frequency, Γ (Γ ) τ-1) versus q2, where q is the wavevector
previously synthesized28 are reported in Table 1. All organic solvents defined as
34 Biomacromolecules, Vol. 10, No. 1, 2009 Houga et al.

4πn θ
q)
λ
sin ()
2
(2)

and λ is the wavelength of the incident laser beam (632.8 nm), θ is the
scattering angle, and n is the refractive index of the media. Single
nanoparticle diffusion coefficient were determined by extrapolation to
zero concentration and hydrodynamic radius (RH) was calculated from
the Stokes-Einstein relation

kBT 2 kBT
RH ) q ) (3)
6πηΓ 6πηDreal
where kB is the Boltzmann constant, Γ is the relaxation frequency, T is
the temperature, and η is the viscosity of the medium.
The size dispersity of the particles was then calculated by the 〈∆Γ2〉/
Γ2 ratio, in which 〈∆Γ2〉 was determined by analysis of the first-order
correlation function by cumulant analysis.

1 1 Figure 1. Autocorrelation function at 90 ° and time distribution function


g1(t)H ) exp[-〈Γ〉t + 〈∆Γ2 〉t2 - 〈∆Γ3 〉t3 + ...] (4) by CONTIN analysis for dex40-b-PS5 at 0.05 mg/mL in water. Inset:
2 6
Relaxation frequency vs q2.
The reduced elastic scattering I(q)/kC, with K ) 4π2n0(dn/dc)2(I090°/
R90°)/λ04NA, was measured in steps from 40 to 120° scattering angle,
where n0 is the refractive index of the standard (toluene), I090° and R90° sulfuric acid and water). The detached replicas were then rinsed with
are the intensity and the Rayleigh ratio of the standard at θ ) 90°, water and cleaned from copolymer with DMSO/THF mixture, before
respectively, dn/dc is the increment of the refractive index, C is the being collected on the 200 mesh copper grid. Observations by trans-
concentration, and I(q) is the intensity scattered by the sample. Elastic mission electron microscopy were made with a FEI Tecnai 12
(static) intensity was calculated according to standard procedures using Microscope working at 120 keV.
toluene as the standard with known absolute scattering intensity. A Some nanoparticules were directly observed by TEM after evapora-
curvature of the angular dependence in a Zimm plot is often observed tion of the solvent. For that purpose, one drop of solution (∼20 µL) at
in our self-assembled nanoparticles which present hydrodynamic radii 0.1 mg/mL was deposited on copper EM grid coated with Formvar
larger than about 100 nm. This curvature becomes largely linearized film. The solvent was dried at atmospheric pressure at room temperature.
by the Berry representation.36 The radius of gyration (Rg) and the second
virial coefficient (A2) were obtained using the equation
Results and Discussion

  (
Kc 1 1
R(q, c)
)
Mw 6 )
1 + Rg2q2 (1+A2Mwc) (5) I. Self-Assembly of a 7% w/w PS Fraction dex-b-PS
Copolymer. The solution properties of dex40-b-PS5 with a PS
(Kc/∆R(θ))1/2 was plotted against (sin 2(θ/2) + kC), with k being an
weight fraction of 7% were first investigated. Dilute solutions
adjustable constant. Extrapolation of the experimental data to zero were analyzed by dynamic light scattering (DLS) at different
concentration and zero angle gave the micellar parameters of MW, Rg, angles at 25 °C. Figure 1 illustrates the autocorrelation function
and A2. We only considered the calculated radius of gyration in our obtained at a scattering angle of 90° with the corresponding
work. CONTIN analysis. Three main time distributions were observed:
Small Angle Neutrons Scattering (SANS). SANS experiments were the main population exhibits a hydrodynamic radius RH ) 28
performed at the Léon Brillouin Laboratory (Orphée reactor, Saclay) nm that was calculated from the diffusion coefficient obtained
on the PACE spectrometer. Two spectrometer configurations have been by plotting the characteristic relaxation frequency versus q2
used to cover a q range from 5 × 10-3 Å-1 to 0.15 Å-1. The main (inset in Figure 1). The other populations were not observed at
parameters used to calculate the corresponding resolution functions37,38 all the scattering angles and were thus considered as physically
are listed in Supporting Information (SI; Table S1). meaningless. Moreover, the few characteristic times that could
The sample (a solution of dex40-b-PS270 at 10 mg/mL in a mixture be extracted were not representative of a diffusive mode, as
DMSO-d/THF-d (40/60)) was introduced into a 5 mm thick rectangular illustrated in Figure S1, SI. The hydrodynamic radius (28 nm)
quartz cell. The blank sample was pure DMSO-d/THF-d (40/60) v/v. of the main population is characteristic of micelles formed by
Data treatment was done with the PAsidur software (LLB). Absolute self-assembly. These micelles were found to be stable for weeks.
values of the scattering intensity (I(q) in cm-1) were obtained from Given the block copolymer composition (∼7% w/w PS),
the direct determination of the number of neutrons in the incident beam core-shell micellar structure with a dextran-based corona
and the detector cell solid angle.39 Contribution due to incoherent oriented toward the continuous aqueous phase is the expected
scattering of the dex40-b-PS270 solute was determined by plotting q4I(q)
morphology and should be at thermodynamic equilibrium.
versus q4 of the sample signal. At large q values, this plot is linear and
Indeed, the size and dispersity of these micelles remained
its slope gives the incoherent contribution of the solute. The magnitude
unchanged even after prolonged heating time. The formation
of this slope was around 0.0065 cm-1 and was subtracted from the
data.
of spherical objects was confirmed by AFM experiments carried
Freeze-Fracture Transmission Electron Microscopy (FFTEM) out in a tapping mode (Figure 2): the measured diameter (∼40
Technique. A drop of the water solution of dex-b-PS (0.1 mg/mL) nm) being in reasonable agreement with DLS findings.
was placed between two copper planchettes of a sandwich holder and A comparison of the size of the micelles (RH ∼ 28 nm in
freezed by plunging into liquid propane. Sample was then fractured at solution) with the characteristic length of the block copolymer
-150 °C and pressure of the order of 10-6 mBar in a BAF 300 Balzers analyzed has been made to check the compatibility with a
apparatus. The fractured surfaces were replicated with platinum core-shell structure. Assuming a fully extended conformation
evaporated at a 45° angle, followed by carbon deposition normal to and average monomer length of 0.25 nm, the contour length
the fracture surface to increase mechanical strength. The copper for the PS block would be 1.2 nm. A solvated dextran of the
planchettes were dissolved in chromerge (a mixture of chromic acid, same molar mass than in our block copolymers, presents a
Micelles and Polymersomes Biomacromolecules, Vol. 10, No. 1, 2009 35

Figure 4. Autocorrelation function at 90° and time distribution function


after CONTIN analysis for a dex40-b-PS270 in THF region (92% THF).
Figure 2. AFM image obtained from a drop of a solution in water at Frequency relaxation vs q2 are represented in inset.
5 mg/L of dex40-b-PS5 evaporated on a mica surface.

mixture. Insets in Figure 3 show a schematic representation of


the obtained morphologies in the following three distinct
domains: DMSO region, THF region, and water, as discussed
in the following sections.
In all cases, self-assembly occurred irrespective of the solvent
composition, giving self-assembled structures with hydrody-
namic radii ranging from 70 to 145 nm. For dex40-b-PS270, RH
values of about 140 nm were measured at low DMSO volume
fraction. Upon increasing the DMSO content, no scattered signal
(dotted curve in Figure 3) could be detected from ∼50 to ∼75%
of DMSO. Further addition of DMSO (>80%) brought about
the formation of smaller nanoparticles (RH ) 120 nm). For the
dex40-b-PS775 sample the behavior was similar with hydrody-
namic radii ranging from ∼115 to ∼65 nm, when increasing
the DMSO content. The domain with no scattering signal was
Figure 3. Evolution of hydrodynamic radius vs DMSO volumic fraction
for different block-copolymers: (O) dex40-b-PS270, (4) dex40-b-PS775. estimated from ∼30 to 45% DMSO. It has to be noted that the
Lines are guides for the eyes. Dotted curve indicates the area where same hydrodynamic radii are obtained, whatever the evolution
no scattered signal could be detected. Schematic morphologies are of this solvent composition (from DMSO to the THF region or
inserted in the graph. Red color illustrates dextran and blue color from THF to the DMSO region), attesting to the reversibility
illustrates polystyrene. of the self-assembly process.
Given the sizes of the nanoparticles obtained for dex40-b-
hydrodynamic radius in the range of 2 nm. Therefore, it is PS775 and dex40-b-PS270 samples and the fact that the RH of dex40-
expected that the hydrophilic chains in core-shell micelles adopt b-PS775 is smaller than that of dex40-b-PS270, the formation of
an extended rather than a coil conformation in water.39 Assum- spherical core-shell micelles is very unlikely. Thus, further
ing a fully extended or “rod-like” conformation of dextran and investigations were carried out to identify the type of nano-
considering a characteristic size of 0.5 nm for each sugar unit, structures formed in THF and DMSO regions.
a value of 20 nm (for DP ) 40) for its contour length is II-1. BehaVior in THF Region. In a first approach, both DLS
obtained. This hypothesis is in good agreement with our and SLS experiments served to assess the Rg/RH value. This
experimental observations, confirming the formation of micellar ratio is well-known to be indicative of the morphology
core shell structure, with dextran chains in their extended con- obtained,36,40 (Rg/RH ∼ 0.77 for a hard sphere, Rg/RH ) 1 for a
formation. vesicle, Rg/RH ) 1.7 for a coil), provided the nanoparticles
II. Self-Assembly of Larger PS Fraction dex-b-PS (>81%) present a relatively narrow size distribution. The different
Block Copolymers. In the first instance, each solution was structures were also characterized by AFM and FFTEM. A
analyzed by DLS at different angles, and the resulting RH were relatively narrow distribution of relaxation time was detected
plotted as a function of the solvent composition (Figure 3). As by DLS in all cases. It is worth mentioning that the slow mode
the procedure is based on the addition of THF or DMSO to observed in Figure 4 (∼2000 nm) was not reproducible and
change the solvent composition, this of course induces dilution not representative of a diffusive mode. This can be attributed
of the solution. We have first checked that dilution had no to the existence of few and nonreproducible aggregates and was
influence on the morphology of the nanostructures. This was therefore considered as physically meaningless. Autocorrelation
realized by diluting up to four times a given solution in which function and relaxation frequency versus q2 are illustrated in
self-assembled structures were obtained, without changing its Figure 4 for dex40-b-PS270. Similar results are available as ESI
solvent composition and by checking the invariance of hydro- for dex40-b-PS775 (Figure S2, SI).
dynamic radius. We can thus conclude that variations observed Values of the hydrodynamic radii extracted from these
in Figure 3 are exclusively due to the modification of solvent experiments are summarized in Table 2. Static light scattering
36 Biomacromolecules, Vol. 10, No. 1, 2009 Houga et al.

Table 2. Radius or Gyration, Hydrodynamic Radii, and Polydispersity Indexa


solvent system Rg (nm) RH (nm) Rg/RH expected morphology dispersity index 〈∆Γ2〉/Γ2
THF region dex40-b-PS270 (92% THF) 200 145 1.38 polydisperse vesicles 0.30
dex40-b-PS775 (95%THF) 115 110 1.04 vesicle 0.20
DMSO region dex40-b-PS270 (95%DMSO) 250 115 2.17 elongated nanoparticle 0.11
dex40-b-PS775 (90%DMSO) 77 77 1 vesicle 0.10
water dex40-b-PS270 60 64 0.94 vesicle 0.074
dex40-b-PS775 70 77 0.91 vesicle 0.10
a
Given at 90° (cumulant analysis) obtained by dynamic light scattering.

Figure 6. SANS intensity as a function of q for dex40-b-PS270 in


DMSO-d6/THF-d5 (40/60) mixture (10 mg/mL).

with the inner membrane of dextran playing the role of a


plasticizer and giving the vesicles the aspect of circular plates
on the mica surface. Very similar AFM results have been
Figure 5. (Top) AFM height micrographs obtained in the THF region obtained on vesicles resulting from the self-assembly of discostic
(92%) for (a) dex40-b-PS270 and (b) dex40-b-PS775 (95% THF). (Bottom)
Section profile of (b).
liquid crystalline molecules42 and diblock codendrimers based
on poly(benzylether) and poly(methallyl dichloride), above the
Tg of both blocks.43 In both cases, results were also interpreted
experiments were also performed and a Berry plot was used to in terms of deformation of the spherical vesicle following
extract the Rg values (SI, Figures S4,5). Size dispersity of the adsorption on the surface, this deformation being possibly
objects formed, estimated by the cumulant analysis, is also enhanced by the force of the AFM probe. Based on these
indicated in Table 2. Considering the light scattering results, elements, it can be concluded that dex40-b-PS775 and dex40-b-
dex40-b-PS775 would afford vesicular morphology (Rg/RH ∼ 1), PS270 self-assemble into vesicles when dissolved as unimers in
the case of dex40-b-PS270 being less clear. Indeed, the ratio Rg/ a mixture of DMSO and THF and then further diluted with THF
RH of 1.4 in that case does not correspond to any known (92% for dex40-b-PS270 and 95% for dex40-b-PS775). It has to be
morphology in literature. This can be explained by the relatively noted that a few very small objects appearing as non hollow
high polydispersity (0.30) of the structures obtained. structures on AFM images could be due to the existence of
To get more insight into the morphologies obtained, AFM spherical micelles which could result from a “residual” kinetic
experiments were performed on dex40-b-PS270 and dex40-b-PS775. influence on the micellar structure formation. Due to their size
As illustrated in Figure 5, for both block copolymers, a donut- and their low number compared to the vesicular morphologies,
type morphology is observed. In the case of dex40-b-PS270 these objects are not detected in light scattering experiments.
(Figure 5a, top), it appears that large number of objects are in The solvent composition (DMSO/THF 40/60) was also
the range of ∼100-400 nm in reasonable agreement with the investigated in the case of the dex40-b-PS270 sample, to gain
distribution of RH observed in Figure 4. However, it is difficult more information on the morphology generated before dilution
to evaluate the polydispersity of these objects because of their to a high THF content (>92%). The Rg/RH ratio was found
insufficient number on the picture. In the case of dex40-b-PS775 roughly the same as in the case of the THF-rich solutions (∼1.4),
(Figure 5b, top), less objects are observed, with a characteristic which is not relevant to draw a clear conclusion about the type
size between 300 nm and 1 µm, in relative good agreement of self-assembled nanostructures. SANS experiments were thus
with distribution of RH obtained by DLS (Figure S2, SI). Again, performed on a solution that was prepared in this case with a
this technique confirms the formation of vesicles (polymer- deuterated mixture of solvents (DMSO-d6/THF-d5 40/60 v/v).
somes) that appear as being adsorbed on the surface. The force The corresponding scattering curve is shown in Figure 6. Due
induced by the AFM probe causes deformation of the vesicles to the large diameter and high polydispersity of the particles,
as verified by analysis of cross section profiles (Figure 5b, the scattered intensity did not present any characteristic oscil-
bottom). Such a profile was drawn using Monte Carlo simula- lation from the form factor that could easily help the data
tions on adsorbed lipid vesicles.41 This deformation is presum- interpretation. However, this curve presented a clear q-2 slope
ably due to the presence of residual DMSO, which interacts of the scattered intensity as a function of q, attesting to the
Micelles and Polymersomes Biomacromolecules, Vol. 10, No. 1, 2009 37

Figure 8. TEM image obtained for the dex40-b-PS270 in water (0.1


mg/mL).

Figure 7. Autocorrelation function at 90° and time distribution function


after CONTIN analysis for dex40-b-PS270 in DMSO region (95%
DMSO). Frequency relaxation vs q2 is presented in inset.

existence of a flat interface, which is characteristic of vesicle


membranes. We thus conclude that the dex40-b-PS270 copolymer
self-assemble in the DMSO/THF 40/60 solution into vesicles
with its PS chains oriented toward the solvent. The size and
morphology of these vesicles remained unaffected upon addition
of THF.
II-2. BehaVior in the DMSO Region. Each block copolymer
solutions exhibited a transition region where no scattering signal
was detected “entering” either the THF- or the DMSO-rich
regions. This means that morphologies obtained in THF dis-
assemble into free copolymer chains (unimers) when the solvent
Figure 9. Freeze-fractured TEM image of dex40-b-PS775 (down) and
mixture solvates efficiently both blocks. These free chains dex40-b-PS270 (up; from water solution, 0.1 mg/mL). Inset: schematic
subsequently self-assemble into reverse nanoparticles when representation of the fracture and platinum shadowing (arrows).
DMSO content increases. In this case, the nanoparticles consist
of dextran blocks oriented toward the solvent, whereas PS chains
are shielded from the continuous DMSO medium in the inner within the core of the nanoparticles, water was added dropwise
part of the nanoparticles. Each solution of the block copolymers until a total volume fraction of 50% was reached. The
has been studied in the DMSO region by DLS and SLS. hydrodynamic radii measured under these conditions remained
Autocorrelation functions and CONTIN analysis are shown in unchanged upon addition of water onto dex40-b-PS775. In
Figure 7 for dex40-b-PS270 as an example, the other results being contrast, for dex40-b-PS270, the RH value decreased from 115 to
provided in SI (Figure S3). 64 nm (Figure 7 and Figure S9, SI). This could be ascribed to
All the block copolymers exhibit a single relaxation time with a morphology change of the self-assemblies and will be
a relatively narrow size distribution, as indicated by cumulant discussed further. Dialysis was performed against distilled water
analysis (Table 2). The Berry plot are available in SI (Figures to remove all traces of DMSO and THF. Results of SLS and
S6, S7). One can note that the characteristic sizes of the DLS experiments are given in Table 2, and those drawn from
nanoparticles are lower compared to the ones measured in the imaging techniques (TEM and FFTEM) are shown in Figures
THF region. When diluting the THF-rich solution with DMSO, 8 and 9. After dialysis, no significant change of the hydrody-
the morphology of the dex40-b-PS270 is modified as the Rg/RH namic radius of the nanoparticles could be noted for both block
ratio stands above 2, a value that is characteristic of an copolymers.
anisotropic structure. It has to be noted that no rotational In water, the Rg/RH ratio is close to one, irrespective of the
diffusion mode in depolarized DLS experiments was detected,44 block copolymer composition, suggesting a vesicular morphol-
meaning that the anisotropy of the nanoparticles formed is ogy. This confirms the occurrence of a morphological change
relatively weak. In the case of dex40-b-PS775, a vesicular structure from an anisotropic structure to spherical vesicles for dex40-b-
is expected in DMSO according to its Rg/RH ratio. Unfortunately, PS270 from its initial DMSO-rich solution, where the Rg/RH was
no reliable results could be obtained by an imaging technique higher than 2. Hollow spheres with low dispersity are clearly
in the DMSO-rich region to confirm these results. Indeed, visible on TEM image (Figure 8), their characteristic sizes being
evaporating DMSO under vacuum caused the disruption of the in good agreement with those obtained by DLS. The replica of
nanoparticles formed. In order to further elucidate the exact freeze fractured samples of dex40-b-PS775 and dex40-b-PS270 were
structure, we then tried to transfer the nanoparticles from DMSO further analyzed by TEM (Figure 9). In the case of dex40-b-
to water. PS775, the holes inside the spheres are clearly evidenced by the
II-3. BehaVior in Water. Starting from solution containing presence of a shadow. The fracture seems to occur as schema-
95% of DMSO and 5% THF, which means that the dextran tized in Figure 9. The shadowing by platinum images the
blocks are oriented toward the solvent and PS chains are shielded vesicular structure of the sample. For the dex40-b-PS270 nano-
38 Biomacromolecules, Vol. 10, No. 1, 2009 Houga et al.

Figure 10. AFM image obtained for cross-linked dex40-b-PS270 nanoparticles from solution DMSO/THF 95/5.

particles, one can notice a black core surrounded by a thick


structured ring. We believe that the particles were just cut off
in the plane as shown in the inset, which leads the inside of the
vesicle to appear black. The images obtained are compatible
with the formation of nanoparticles, exhibiting a vesicular
structure, and thus confirm the results obtained from scattering
experiments.
III. Cross-Linking of the Nanoparticles. To better elucidate
the morphologies obtained in the DMSO-rich region, we
attempted to cross-link them directly from the organic solution.
Indeed, cross-linking is an appropriate method to fix the self-
assembled morphologies and modify their properties such as
their permeability or strength. This can be realized by different
approaches that have been recently reviewed,45 with radical
cross-linking polymerization and chemical reaction between two
functional groups being mainly used. Cross-linking can be
performed within the core 46-50 or in the shell of the micellar Figure 11. AFM image obtained for cross-linked dex40-b-PS270
structures.51-60 If the cross-linking of core shell micelles has nanoparticles after dialysis in water.
been widely investigated, vesicles made from the self-assembly
of block copolymer have also been the subject of such treatment.
61-64
In this work, DVS was used to cross-link the dextran In the case of dex40-b-PS270, we anticipated the formation of
chains of the self-assembled nanostructures in DMSO/THF an anisotropic (elongated) self-assembled structure in DMSO.
solution. This cross-linker has been successfully used in This hypothesis could not be confirmed by imaging techniques.
polyacrylate derivatives54,65 or to cross-link vesicles made from Indeed, as indicated above, the evaporation of DMSO under
the self-assembly of amphiphilic comb-like PEG.66 In DMSO/ vacuum led to a disruption of the nanoparticles. After cross-
THF medium, nanoparticles with a core constituted of dextran linking, this could be done without destroying the nanoparticles.
or PS can be obtained reversibly, depending on the composition AFM imaging was then performed as shown in Figure 10. Ovoïd
of the solvent mixture. The cross-linking of dextran was nanoparticles are clearly observed with characteristic dimensions
successfully accomplished irrespective of the location of dextran of ∼250 nm along the main axis and ∼100 nm perpendicular
in the nanoparticle (in the inner core or as the shell). To this to it, confirming the anisotropy and the size measured by
end, a solution of dex40-b-PS270 in DMSO/THF (40/60) was scattering techniques. Detailed information about these aniso-
prepared to obtain a vesicle with dextran oriented in the inner tropic objects, whether they are filled or hollow structures, have
membrane before adding DVS. The evolution of the nanopar- not yet been obtained. However, as vesicles were formed from
ticles was then followed by DLS. A slight increase of RH (from noncross-linked samples upon addition of water and dialysis
∼140 to ∼155 nm) was observed, which could be explained (see Figure 9 and Table 2), the formation of hollow anisotropic
by an increase of the rigidity of the dextran membrane, resulting objects can be contemplated here. Upon addition of water (up
in a slight decrease of its curvature. Dilution with THF of this to 50% v/v) and dialysis, these cross-linked anisotropic nano-
solution left RH almost unchanged, confirming the stability of particles became spherical as shown in Figure 11. The charac-
the nanoparticles with their PS chains located at the surface teristic size is in reasonable agreement with the one obtained
of the vesicle. In contrast, a progressive addition of DMSO by dynamic and static light scattering on the cross-linked system
induced the irreversible precipitation of the cross-linked nano- in water (RH ) 145 nm; Rg ) 210 nm; Rg/RH ) 1.44, see Figure
particles. Indeed, the structure cannot disassemble and form S11). The Rg/RH ratio is different from what is expected for a
stable reverse vesicles with dextran oriented toward the continu- vesicular structure. At this point, it is difficult to conclude on
ous medium because of intermolecular cross-linking of dextran. the morphology obtained in water from cross-linked samples.
The same approach was used for nanoparticles presenting We think that vesicular structure with relatively pronounced size
dextran at the surface. A solution of dex40-b-PS270 was prepared dispersity could be possible. Finally, these results suggest that
in 90/10 DMSO/THF, thereafter, DVS was added. A marked in spite of cross-linking, the mobility of some dextran chains
increase of the hydrodynamic radius from 115 to 140 nm is increases when DMSO is replaced by water, the constraints and
again observed, and the subsequent dilution of this solution with the stresses existing in the anisotropic structure being likely
THF brought about the particles precipitation, attesting to the relaxed by a morphology change. For noncross-linked mor-
efficiency of the cross-linking process. phologies, the phenomenon is supposed to be the same but with
Micelles and Polymersomes Biomacromolecules, Vol. 10, No. 1, 2009 39

greater amplitude, as a decrease of hydrodynamic radius from the transfer of the morphologies from organic solvent to water,
115 to 65 nm is observed when water is added to a solution of as well as dynamic and static light scattering results. This
dex40-b-PS270 in DMSO. Obviously, it is difficult to describe material is available free of charge via the Internet at http://
precisely what really happens at the molecular level during the pubs.acs.org.
transfer from DMSO to water as the morphology in DMSO is
not yet fully elucidated (full or hollow ovoïds). References and Notes
(1) Rodriguez-Hernandez, J.; Checot, F.; Gnanou, Y.; Lecommandoux,
Conclusions S. Prog. Polym. Sci. 2005, 30 (7), 691–724.
The self-assembly process of dex-b-PS diblock copolymers (2) Zhang, L. F.; Eisenberg, A. Polym. AdV. Technol. 1998, 9 (10-11),
677–699.
in organic and aqueous media is strongly influenced by their
(3) Riess, G. Prog. Polym. Sci. 2003, 28 (7), 1107–1170.
overall composition. The reversibility of the self-assembly (4) Gohy, J. F. AdV. Polym. Sci. 2005, 190, 65–136.
process (from DMSO to THF and vice versa) demonstrates that (5) Soo, P. L.; Eisenberg, A. J. Polym. Sci., Part B: Polym. Phys. 2004,
some thermodynamic equilibrium can be reached. Core-shell 42 (6), 923–938.
micelles are generated in water for a content of dextran equal (6) Zhang, W. Q.; Shi, L. Q.; An, Y. L.; Gao, L. C.; Wu, K.; Ma, R. J.
Macromolecules 2004, 37 (7), 2551–2555.
to 93% (w/w). In all other conditions, block copolymers self- (7) De Gennes, P. J.; In Solid State Physics, Academic Press: New York,
assemble into vesicular structures. These morphologies were 1978; Vol. 14.
not expected and do not correspond to the prediction and (8) Shusharina, N. P.; Nyrkova, I. A.; Khokhlov, A. R. Macromolecules
classification proposed by Discher and Eisenberg11 (e.g., for 1996, 29 (9), 3167–3174.
(9) Wu, C.; Gao, J. Macromolecules 2000, 33 (2), 645–646.
hydrophilic fraction of 19 and 8%) for coil-coil type block
(10) Zhulina, Y. B.; Birshtein, T. M. Polym. Sci. U.S.S.R. 1985, 27 (3),
copolymers, because crew-cut micelles would have been 570–578.
expected. Although dextran is considered to exhibit a coil (11) Discher, D. E.; Eisenberg, A. Science 2002, 297 (5583), 967–973.
conformation, it can adopt a rod-like conformation for molar (12) Ahmed, F.; Discher, D. E. J. Controlled Release 2004, 96, 37–53.
masses below 2000 g/mol.32 Because the average molar mass (13) Ahmed, F.; Pakunlu, R. I.; Brannan, A.; Bates, F.; Minko, T.; Discher,
D. E. J. Controlled Release 2006, 11, 150–158.
of the dextran in this study is equal to 6600 g/mol and its (14) Liu, D. Z.; Hsieh, J. H.; Fan, X. C.; Yang, J. D.; Chung, T. W.
polydispersity index is around 1.6, it is obvious that some of Carbohydr. Polym. 2007, 68 (3), 544–554.
the dextran chains are below the 2000 g/mol range and (15) Elhasi, S.; Astaneh, R.; Lavasanifar, A. Eur. J. Pharm. Biopharm.
accordingly adopt a rod conformation. It is well-known in the 2007, 65 (3), 406–413.
(16) Geng, Y.; Discher, D. E. Polymer 2006, 47 (7), 2519–2525.
self-assembly of rod-coil block copolymer that their stiffness
(17) Discher, D. E. Abstr. Pap.sAm. Chem. Soc. 2003, 225, U679–U679.
asymmetry promotes self-assembly into flat membranes and in (18) Ziegast, G.; Pfannemuller, B. Makromol. Chem., Rapid Commun. 1984,
diluted state, into vesicles.67-70 Therefore, the slight stiffness 5 (7), 373–379.
asymmetry induced by the dextran chains that adopt a rod (19) Bosker, W. T. E.; Agoston, K.; Cohen Stuart, M. A.; Norde, W.;
conformation could play a crucial role in the formation of a Timmermans, J. W.; Slaghek, T. M. Macromolecules 2003, 36, 1982–
1987.
planar bilayer membrane, resulting in the formation of vesicles (20) Liu, J.-Y.; Zhang, L.-M. Carbohydr. Polym. 2007, 69 (1), 196–
either in water, THF, or DMSO. The study of block copolymers 201.
with longer dextran chains and narrow size distribution could (21) Loos, K.; Stadler, R. Macromolecules 1997, 30 (24), 7641–7643.
bring interesting information about this aspect. (22) Hernandez, O. S.; Soliman, G. M.; Winnik, F. M. Polymer 2007, 48
(4), 921–930.
This study revealed some specificities of dextran-based block
(23) Yang, Y.; Kataoka, K.; Winnik, F. M. Macromolecules 2005, 38 (6),
copolymers (unexpected vesicles, ovoïds,...) and that further 2043–2046.
studies are needed to get a better insight, especially on the (24) Akiyoshi, K.; Kohara, M.; Ito, K.; Kitamura, S.; Sunamoto, J.
influence of the dextran conformation to the self-assembly Macromol. Rapid Commun. 1999, 20, 112–115.
behavior. Indeed, if vesicular structure could be explained by a (25) Loos, K.; Müller, A. H. E. Biomacromolecules 2002, 3 (2), 368–373.
(26) Haddleton, D. M.; Ohno, K. Biomacromolecules 2000, 1 (2), 152–
rod conformation of short dextran chains, the origin of the 156.
formation of the ovoı¨d structures is not yet understood. In (27) Kakuchi, T.; Narumi, A.; Miura, Y.; Matsuya, S.; Sugimoto, N.; Satoh,
addition, the exact role of the dextran chains in the morphologi- T.; Kaga, H. Macromolecules 2003, 36 (11), 3909–3913.
cal transition observed during transfer from DMSO to water (28) Houga, C.; Le Meins, J.-F.; Borsali, R.; Taton, D.; Gnanou, Y. Chem.
Commun. 2007, 3063–3065.
has to be further elucidated.
(29) Akiyoshi, K.; Maruichi, N.; Kohara, M.; Kitamura, S. Biomacromol-
Morphologies obtained in DMSO could be transferred in ecules 2002, 3 (2), 280–283.
water paving the way of potential applications as carriers or (30) Narumi, A.; Miura, Y.; Otsuka, I.; Yamane, S.; Kitajy, Y.; Satoh, T.;
vehicles. The vesicles (or polymersomes) obtained are particu- Hirao, A.; Kaneko, N.; Kaga, H.; Kakuchi, T. J. Polym. Sci., Part A:
larly interesting because of their ability to encapsulate both Polym. Chem. 2006, 44 (16), 4864–4879.
(31) Loos, K.; Böker, A.; Zettl, H.; Zhang, M.; Krausch, G.; Müller,
hydrophobic and hydrophilic molecules. The morphologies can A. H. E. Macromolecules 2005, 38 (3), 873–879.
be efficiently cross-linked via the hydroxyl groups of the (32) Gekko, K. ACS Symp. Ser. 1981, 150, 415–437.
dextran, which can be a way to tune the membrane permeability (33) Yu, K.; Eisenberg, A. Macromolecules 1998, 31 (11), 3509–3518.
of these vesicles. Moreover, low concentration regime can be (34) Shen, H.; eisenberg, A. J. Phys. Chem. B 1999, 103, 9473–9487.
(35) Yu, K.; Bartels, C.; Eisenberg, A. Langmuir 1999, 15 (21), 7157–
reached without disruption of the nanoparticles, which could
7167.
be an advantage in many applications. (36) Burchard, W. In Polysaccharides: Structural DiVersity and Functional
Versatility, 2nd edition; Dimitriu, S., Ed.; Marcel Dekker: New York,
Acknowledgment. The authors would like to thank Prof. 2005; pp 189-236.
Olivier Mondain Monval (CRPP, Bordeaux) for helpful com- (37) Mildner, D. F. R.; Carpenter, J. M. J. Appl. Crystallogr. 1984, 17,
ments on freeze-fractured TEM images, Michel Schappacher 249–256.
(38) Pedersen, J. S.; Posselt, D.; Mortensen, K. J. Appl. Crystallogr. 1990,
for AFM measurements, Annie Brûlet (LLB, CEA Saclay) for
23, 321.
SANS measurements, and referees for their helpful comments. (39) Daoud, M.; Cotton, J. P. J. Physiol. (Paris) 1982, 43 (3), 531–538.
(40) Burchard, W. AdV. Polym. Sci. 1983, 48, 1–124.
Supporting Information Available. Experimental details (41) Dimitrievski, K.; Zäch, M.; Zhdanov, V. P.; Kasemo, B. Colloids Surf.,
about light scattering measurements, experimental details about B 2006, 47, 115–125.
40 Biomacromolecules, Vol. 10, No. 1, 2009 Houga et al.

(42) Seo, S. H.; Chang Young, J.; Tew, G. N. Angew. Chem., Int. Ed. 2006, (56) Thurmond Ii, K. B.; Huang, H.; Clark, C. G., Jr.; Kowalewski, T.;
45, 7526–7530. Wooley, K. L. Colloids Surf., B 1999, 16 (1-4), 45–54.
(43) Yang, M.; Wang, W.; Yuan, F.; Zhang, X.; Li, J.; Liang, F.; He, B.; (57) Thurmond Ii, K. B.; Kowalewski, T.; Wooley, K. L. J. Am. Chem.
Minch, B.; Wegner, G. J. Am. Chem. Soc. 2005, 127, 15107–15111. Soc. 1997, 119 (28), 6656–6665.
(44) Pecora, R. J. Chem. Phys. 1968, 48, 4126–4128. (58) Thurmon Ii, K. B.; Wooley, K. L. Polym. Prepr. (Am. Chem. Soc.,
(45) O’Reilly, R. K.; Hawker, C. J.; Wooley, K. L. Chem. Soc. ReV. 2006, DiV. Polym. Chem.) 1998, 303.
35 (11), 1068–1083. (59) Wooley Karen, L. J. Polym. Sci., Part A: Polym. Chem. 2000, 38 (9),
(46) Ishizu, K.; Onen, A. J. Polym. Sci., Part A: Polym. Chem. 1989, 27 1397–1407.
(11), 3721–3731. (60) Zhang, Q.; Remsen, E. E.; Wooley, K. L. J. Am. Chem. Soc. 2000,
(47) Saito, R.; Ishizu, K.; Fukutomi, T. Polymer 1990, 31 (4), 679–683. 122 (15), 3642–3651.
(61) Discher, B. M.; Bermudez, H.; Hammer, D. A.; Discher, D. E.; Won,
(48) Tian, L.; Yam, L.; Wang, J. Z.; Tat, H.; Uhrich, K. E. J. Mater. Chem. Y. Y.; Bates, F. S. J. Phys. Chem. B 2002, 106 (11), 2848–2854.
2004, 14 (14), 2317–2324. (62) Nardin, C.; Hirt, T.; Leukel, J.; Meier, W. Langmuir 2000, 16 (3),
(49) Wilson, D. J.; Riess, G. Eur. Polym. J. 1988, 24 (7), 617–621. 1035–1041.
(50) Guo, A.; Liu, G.; Tao, J. Macromolecules 1996, 29, 2487–2493. (63) Ding, J.; Liu, G. Chem. Mater. 1998, 10 (2), 537–542.
(51) Huang, H.; Kowalewski, T.; Remen, E. E.; Gertzmann, R.; Wooley, (64) Sauer, M.; Meier, W. Chem. Commun. 2001, (1), 55–56.
K. L. J. Am. Chem. Soc. 1997, 119 (48), 11653–11659. (65) Liu, S.; Ma, Y.; Armes, S. P.; Perruchot, C.; Watts, J. F. Langmuir
(52) Huang, H.; Remsen, E. E.; Kowalewski, T.; Wooley, K. L. J. Am. 2002, 18 (21), 7780–7784.
Chem. Soc. 1999, 121 (15), 3805–3806. (66) Li, X.; Ji, J.; Shen, J. Macromol. Rapid Commun. 2006, 27, 214–218.
(53) Weaver, J. V. M.; Tang, Y.; Liu, S.; Iddon, P. D.; Grigg, R.; (67) Lecommandoux, S.; Borsali, R. Polym. Int. 2006, 55 (10), 1161–1168.
Billingham, N. C.; Armes, S. P.; Hunter, R.; Rannard, S. P. Angew. (68) Halperin, A. Polym. ReV. 2006, 46 (2), 173–214.
Chem., Int. Ed. 2004, 43 (11), 1389–1392. (69) Klok, H. A.; Lecommandoux, S. AdV. Mater. 2001, 13 (16), 1217–
1229.
(54) Pilon, L. N.; Armes, S. P.; Findlay, P.; Rannard, S. P. Eur. Polym. J.
(70) Rodriguez-Hernandez, J.; Lecommandoux, S. J. Am. Chem. Soc. 2005,
2006, 42 (7), 1487–1498.
127 (7), 2026–2027.
(55) Rodriguez-Hernandez, J.; Babin, J.; Zappone, B.; Lecommandoux, S.
Biomacromolecules 2005, 6 (4), 2213–2220. BM800778N

You might also like