You are on page 1of 9

Eur. Phys. J. E (2011) 34: 4 DOI 10.

1140/epje/i2011-11004-1

THE EUROPEAN PHYSICAL JOURNAL E

Regular Article

Rheological behaviour of polyoxometalate-doped lyotropic lamellar phases


J.P. de Silva1,a , A.S. Poulos1 , B. Pansu1 , P. Davidson1 , B. Kasmi1 , D. Petermann1 , S. Asnacios2 , F. Meneau3 , and M. Impror1 e
1 2 3

Laboratoire de Physique des Solides-UMR 8502-Universit Paris-Sud, F-91405 Orsay, France e Laboratoire Mati`re et Syst`mes Complexes-10 rue Alice Domon et Lonie Duquet, 75205 Paris, France e e SOLEIL Synchrotron-LOrme des Merisiers Saint-Aubin, BP 48 91192 Gif-sur-Yvette, France Received 29 March 2010 and Received in nal form 23 August 2010 Published online: 10 January 2011 c EDP Sciences / Societ` Italiana di Fisica / Springer-Verlag 2011 a Abstract. We study the inuence of nanoparticle doping on the lyotropic liquid crystalline phase of the industrial surfactant Brij R 30 (C12 E4 ) and water, doped with spherical polyoxometalate nanoparticles smaller than the characteristic dimensions of the host lamellar phase. We present viscometry and in situ rheology coupled with small-angle X-ray scattering data that show that, with increasing doping concentration, the nanoparticles act to decrease the shear viscosity of the lamellar phase, and that a shear-induced transition to a multilamellar vesicle onion phase is pushed to higher shear rates, and in some cases completely suppressed. X-ray data reveal that the nanoparticles remain encapsulated within the membranes of the vesicles, thus indicating a viable method for the fabrication of nanoparticle incorporating organic vesicles.

1 Introduction
The rheological properties of complex uids is an active eld of research: areas of particular interest concern if and how such uids ow, as well as any structural changes or transitions associated with an imposed shear rate or stress. Fine control over the ow of structured uids is highly desirable for a wide range of applications, while the rheological properties of non-Newtonian, structured uids is also an important subject, as such uids often exhibit strongly nonlinear behaviour under shear [1]. Here we consider how the rheological properties of a complex uid, in this case a lamellar phase of water and nonionic surfactant, are modied by the addition of spherical nanoparticles (1.1 nanometers in diameter) that are smaller than the characteristic dimensions of the host lamellar phase. If one considers the rheological properties of dilute surfactant/water lyotropic lamellar phases, observations of shear-induced alignment, thinning, and formation of multilamellar vesicles (MLVs) of monodisperse dimensions initiated by buckling instabilities, have been reported by Diat and coworkers using a variety of microscopy and scattering techniques [2,3]. In such systems one may observe three regimes [4]: an initial shear thinning regime corresponding to shearing of the lamellar phase; a shear thickening regime corresponding to the
a Current address: Dpartement de Physique-Universit libre e e de Bruxelles, U.L.B CP223, B-1050 Brussels, Belgium. e-mail: j.desilva@physics.org, imperor@lps.u-psud.fr

transition from lamellae to MLVs; a nal shear thinning regime at high shear rates corresponding to shearing of the MLV state. It is suggested that lamellar phases ow under shear through the transportation of defects (such as dislocations) through the bulk phase [5,6]. In this paper we discuss the rheological properties of a lyotropic, liquid crystalline lamellar phase of Brij R 30/water doped with polyoxometalate (POM) nanoparticles derived from phosphotungstic acid [7]. One interesting property of the POM doped lamellar phases that is immediately visible is that the shear viscosity of the phase is signicantly reduced with increased doping concentration, as illustrated in g. 1, where the dierence in ow under the inuence of gravity between doped and undoped phases can be visualised simply by tilting the samples. The extended phase diagram and properties of this ternary system have been previously characterised in detail [8,9]. Brij R 30 surfactant is industrial grade C12 E4 that contains a small amount of homologue impurities, but is signicantly cheaper than pure C12 E4 and is therefore more suitable for wide scale application. It is observed in the ternary system that the POMs are preferentially located close to the EO groups of the surfactant membrane rather than remaining dispersed within water between the bilayers, while the interlamellar spacing remains unmodied by the presence of the particles [8]. Previous studies of C12 E4 /water [10] and C10 E3 /water [11] lamellar phases under shear have characterised the shear viscosity and lamellar-MLV transition in these pure systems. Recent

Page 2 of 9

The European Physical Journal E

Fig. 1. A simple visual illustration of the signicant dierence in ow properties for 40/0 and 40/4 doped lamellar phases under the inuence of gravity: (a) initial state and (b) after 10 seconds inclined at approximately 70 degrees. Note also the higher turbidity of the 40/0 sample.

bilayer spacing of between 7 and 11 nm depending on the surfactant/water ratio, but importantly invariant with respect to POM doping concentration [8]. The samples used for this study comprise either 40 or 50% C12 E4 volume fraction with up to 4% POMs by volume in water; these surfactant/water/POM ratios are chosen to keep the system well inside the L lamellar region of the phase diagram, and the upper limit on POM doping is imposed by a phase separation that is observed at POM concentrations greater than this. Hereafter the samples will be labelled using the nomenclature of surfactant volume fraction of total sample/POM volume fraction of water; thus 40/1 refers to a sample composed of 40% surfactant by volume and 1% POMs by volume in water. 2.2 Optical microscopy Optical observations of the sample texture between crossed polarisers is achieved using Olympus microscopes with digital image capture via CCD camera. Optical grade, at glass capillaries (2 0.2 mm, Vitrocom U.S.A.) are lled by slowly drawing the lamellar phase into the capillary, while a temperature stage permits the samples to be accurately thermostated. 2.3 Laboratory rheology and rheo-microscopy Laboratory rheology measurements are performed using Anton-Paar Physica MCR 301 and 501 series rheometers in cone-plate geometry. One and two degree truncated cones (52 m truncation) of metal and 0.5 degree cones of glass, both of radius 50 mm, were available. Rheomicroscopy experiments are performed using a Rheometrics DSR500 rheometer retrotted with a cross-polarised optical capture system, using an inverted microscope and CCD or digital camera allowing in situ capture of the sheared texture [5]. A truncated glass cone (40 mm in diameter) and plate are roughened using precision sandblasting, while two small polished windows are retained in the cone and plate in order to provide clear optical image capture. This procedure is performed in order to reduce slip at the polished surfaces of this geometry. Rheology measurements are performed with both imposed shear rate and stress, and anti-evaporation traps are used where available; measurements are performed at room temperature including all laboratory and subsequent synchrotron RheoSAXS experiments. 2.4 RheoSALS Rheological small-angle light scattering experiments were performed at the laboratories of Anton-Paar France using an AP 501 rheometer tted with a solid-state laser (658 nm wavelength) illuminating the sample vertically down through a glass plate-plate geometry (0.5 mm gap) between crossed polarisers. Scattering patterns are captured using a CCD camera and screen. A value of n =

studies of the eect of temperature on the MLV transition in nonionic surfactant lamellar phases have been made using coupled rheology and scattering techniques [12]. Preceding studies of the rheological properties of doped lamellar phases have considered dopants such as water soluble polymers [1315], colloidal particles [1618] or clays [19] that are of the same order as or of dimensions much larger than the interlamellar spacing. Doping with nanoparticles such as POMs, where the nanoparticles are smaller than the characteristic dimensions of the lamellar phase, has not been investigated in detail; more precisely, the 1.1 nm diameter POMs considered here are less than the bilayer thickness (3.4 nm) and about one quarter of the water channel thickness (4.6 nm) for a lamellar period of 8 nm [8].

2 Experimental
2.1 Sample preparation The nonionic surfactant polyethylene glycol dodecyl ether, C12 E4 , forms a lyotropic liquid crystalline phase when mixed with water. Lamellar phases are prepared using surfactant, water and POMs at ratios that place the resultant doped and undoped systems in the L lamellar region of the phase diagram, as previously deduced for the same experimental system [8,20]. Here we use the commercial surfactant Brij R 30 sourced from Sigma-Aldrich in lieu of pure C12 E4 ; Brij R 30 contains an unspecied small amount of impurities by way of organic homologues. Nanoparticles are derived from Sigma-Aldrich sourced 12phosphotungstic acid, 3H+ [PW12 O40 ]3 ; the POM particles are in the form of the Keggin structure [7] with a central phosphor homoatom and a diameter of 1.1 nm. The surfactant is mixed with water or an aqueous POM solution of various concentrations in a test tube; samples are centrifuged in normal and inverted congurations in order to achieve good homogeneity and evacuate trapped air bubbles that will adversely aect rheological measurements. The resulting birefringent lamellar phases have a

J.P. de Silva et al.: Rheological behaviour of polyoxometalate-doped lyotropic lamellar phases

Page 3 of 9

1.38 is taken for the refractive index of Brij R 30/water samples [10] where the expression of the scattering vector is given by q = 4n sin(/2)/ and is the scattering angle. 2.5 RheoSAXS RheoSAXS experiments were performed at the SWING beamline of the Soleil Synchrotron, using a custom-built concentric cylinder Couette shear cell constructed of polycarbonate. A synchrotron light source provides sucient beam intensity for kinetic studies, where the time for a single acquisition is minimised to at most a few seconds in order to acquire a clear snapshot of a dynamic process such as this. The polycarbonate cell is 20 mm in diameter with a gap of 0.5 mm, height of 17 mm and wall thickness of 0.3 mm. As the radius of the interior cylinder is much greater than that of the gap, a linear approximation to the velocity gradient across the gap is valid: thus we derive the shear viscosity as a function of the mean 2 2 2 2 shear rate = (R2 + R1 )/(R2 R1 ), and shear stress 2 2 2 2 = M (R2 + R1 )/(4HR1 R2 ), where R1 and R2 are the interior and exterior cylinder radii, respectively, H is the cylinder height, M is the measured torque and is the angular velocity. The shear viscosity simply relates the mean shear rate and stress: = . The polycarbonate Couette shear cell is utilised by two dierent experimental set-ups: a custom shear rheometer that provides a torque measurement is installed onto the beamline, where the shear rate is imposed by controlling the speed of rotation of the external cylinder; an identical polycarbonate shear cell has also been fabricated by Anton-Paar to permit the Physica 501 controlled stress rheometer to be installed onto the beamline, giving access to both imposed shear rate and stress experiments. Contrary to the custom shear rheometer, the interior cylinder is rotated by the Anton-Paar rheometer and thus the system is susceptible to Taylor-Couette instabilites at higher shear rates. Further details of these experimental apparatus are available [21]. During shear, scattering data at both radial and tangential positions of the shear cell (with respect to the ow direction these correspond to a perpendicular and parallel alignment of the incident beam) is recorded by horizontally displacing the rheometer with respect to the incident X-ray beam (horizontal and vertical spot size approximately 400 200 m) via a motorised table. The X-ray transmission at both radial and tangential positions of the cell is accurately modelled taking into account the width of the beam, and data are treated using software provided by the SWING beamline. The typical A accessible q-range is 4104 0.55 1 at an X-ray energy of 11 keV, which minimises absorption by the polycarbonate cell. RheoSAXS measurements are limited to a maximum of approximately 2 h as the interlamellar spacing will be rather sensitive to drying of the sample. Laboratory X-ray experiments indicate that the interlamellar spacing remains constant under static conditions within the bulk of the Couette cell for at least 10 hours without any form of anti-evaporation measures.

Fig. 2. (a) Measured shear viscosity against shear rate for 40/0, 40/1, 40/2 and 40/4 samples for a sequential series of constant shear rates in cone-plate geometry, after 2 h at each shear rate step. In situ optical textures between crossed polarisers are shown for 40/2 sample corresponding to viscosity data before shear, and after the corresponding 10 and 20 s1 shear steps. (Scale bar is 50 m.) (b) Corresponding SALS patterns for the 40/2 sample at the indicated shear rates.

3 Results and discussion


3.1 Macroscopic observations A quantitative global overview of the eect of POM doping is illustrated by g. 2(a) for 40% surfactant lamellar phases at 0, 1, 2, 3 and 4% POM doping concentrations. This gure represents a comparison of the measured shear viscosity of each particular phase during a single imposed shear rate rheology experiment consisting of a continuous series of constant shear rate steps, each for a duration of 2 h in the case of these data. The data presented are the value of the shear viscosity at the end of each shear rate step. A pre-shear protocol at a constant shear rate of 0.1 s1 for 15 minutes is implemented in order to attain a reproducible starting state for each repeat measurement without large deformation [22,23]. From g. 2 one can see that the trend is for the lamellar phase shear viscosity to systematically decrease with increasing POM doping concentration, where the basic form of the ow curve is oset towards higher shear rate and lower viscosity. This is in contrast to the inuence of adding a small amount of ionic surfactant such as sodium dodecylsulfate (SDS) to

Page 4 of 9

The European Physical Journal E

a C12 E4 lamellar phase for example, which results in an increase in shear viscosity [24]. Between 0 and 4% POM doping the ow curve is displaced by approximately an order of magnitude in both viscosity and shear rate. The general form of each individual ow curve is in good agreement u with that observed for pure C12 E4 by Mller et al. [10]: an initial shear thinning, followed by a shear thickening region corresponding to a lamellar-MLV transition, and a nal shear thinning regime of the MLV or onion phase. As g. 2(a) illustrates the lamellar-MLV transition region is also systematically shifted to higher shear rates with increasing POM doping concentration, and the origin of this eect will be discussed in detail later. Optical micrographs between crossed polarisers of the lamellar and MLV phase textures are in good agreement with those observed by Mller et al. [10]: the optical imu ages presented in g. 2(a) for the 40/2 sample before and after shearing at = 10 and 20 s1 closely resemble those of the C12 E4 lamellar phase and MLV phase, respectively, where the texture of the MLV phase becomes ner with increasing shear rate as the close-packed MLVs decrease in diameter. This behaviour is also supported by the typical four-lobe SALS patterns observed between crossed polarisers (see g. 2(b)) for the same 40/2 sample at shear rates from 10 to 40 s1 . These SALS patterns are in good agreement with those observed for a pure C12 E4 system indicating the presence of densely packed MLVs of tens of microns in diameter [25]. Here we conrm for the studied system that an MLV phase can be formed in both doped and undoped samples; from these patterns, an average size of a few tens of microns for the MLV diameter is derived, and this size decreases with increasing shear rate from about 30-40 microns to 20 microns for the 40/2 sample from 10 to 40 s1 , for example, again consistent with previous studies of undoped lamellar phases [26]. The important conclusion to draw from these data is that the classical lamellar-MLV transition is observed in both doped and undoped samples even though it is shifted to higher shear rates with increased POM doping concentration. The optical texture of the doped lamellar phases prior to any shear exhibits a much ner texture than the classical mosaic texture of a nonoriented lamellar phase [10]. In fact the undoped lamellar phases are consistently visually more turbid than the doped samples, as seen in g. 1 for instance, which indicates that POM doping has some eect on the basic macroscopic texture of the prepared lamellar phase prior to shear.

Fig. 3. (a) Normalised measured steady-state shear viscosity and t as a function of POM area fraction Fpom for 40 and 50% surfactant volume fraction, for imposed shear rate = 0.1 s1 . (b) Measured yield stress y and linear t as a function of Fpom . (c) Corresponding instantaneous elastic modulus G0 and linear t as a function of Fpom .

and A are the volume and cross-section of the POM particle, taken as 0.685 nm3 and 0.93 nm2 , respectively [27]. In g. 3(a) we present the normalised steady-state shear viscosity of the lamellar phase for 40 and 50% surfactant concentration as a function of Fpom , measured for an imposed shear rate = 0.1 s1 . Here we propose a suitable model, where Fpom is the relevant parameter used to describe the decrease in shear viscosity with increasing doping concentration: Fpom (Fpom ) = 1 A 1 exp 0 B
n

(1)

3.2 Bulk properties of doped lamellar phases This viscosity decrease with increasing POM doping can be quantied by examining the eect of POM doping on the shear viscosity as a function of the POM area fraction of the surfactant membrane, Fpom , for dierent surfactant concentrations. The POM area fraction is given by Fpom = [A(1 s )pom ]/2V0 s , where s and pom are the surfactant and POM volume fractions, respectively, is the surfactant bilayer spacing, taken as 3.4 nm, and V0

where 0 is the zero-shear viscosity of the undoped phase, (1 A) is the plateau reduced viscosity at high doping concentrations and B is the critical value of Fpom . Here the data for the two dierent surfactant concentrations collapse onto the same curve, where the data for both 40 and 50% samples are well tted by n = 3, A = 0.85 and B = 0.092. This result strongly indicates that the eect of the POMs in reducing the shear viscosity depends purely on the areal density of the POMs adsorbed onto the surfactant membrane, regardless of the surfactant/water ratio. It can be seen from g. 3(a) that there is no further reduction in viscosity after Fpom 0.15, which probably corresponds to a physical limit on the density of nanoparticles that can be adsorbed onto the surfactant membrane [8]. Also presented in g. 3(b) are yield stress (y ) measurements for the same range of surfactant concentration

J.P. de Silva et al.: Rheological behaviour of polyoxometalate-doped lyotropic lamellar phases

Page 5 of 9

and POM areal fraction; yield stress data are determined from creep experiments, where a series of increasing shear stress steps are applied to the sample until a detectable deformation is recorded. A linear t to the data shows that the trend is for the yield stress to decrease with increasing POM areal fraction, falling from around 8 2 Pa for undoped phases to around 2 1 Pa at Fpom = 0.15. The values of y measured here for the undoped phases are also in good agreement with the values measured by Paasch et al. [28] for similar undoped lamellar phases. (It should be noted that the measured value of a yield stress for a lamellar phase, or indeed even the existence of a yield stress, has been shown to often depend on the measuring technique itself, for example [29].) Finally, the instantaneous elastic modulus G0 is presented as a function of Fpom in g. 3(c) for the two surfactant concentrations, determined from creep-recovery experiments, where a constant stress 0 < y is applied to the sample for a time T and then removed, while the strain is measured during the stress application and subsequent relaxation period. The creep data are well described by the Kelvin-Voigt model derived by Panizza et al. for viscoelastic phases [30], where typical values from the ts for the 40/4 samples, for example, are of the order of J0 , Jr , Jn 102 Pa1 , v 500 s, rel 500 s and m 105 Pa s. The introduction of POMs tends to slightly reduce the elastic modulus, as shown in g. 3(c), where the instantaneous elastic modulus decreases from about 300 to 100 Pa by Fpom 0.15. The correlated decrease in viscosity, yield stress and elastic modulus with increasing POM doping suggests that the POMs act to uidify the lamellar phase. If one considers that the yield stress and elastic response of a lamellar phase are due to defects within the lamellar phase that create internal stress [31], we can postulate that the POMs act to reduce the defect density resulting in the experimentally observed reduction in y and G0 . Such an effect would also explain the reduction in turbidity observed with increasing POM doping concentration (see g. 1 for example), corresponding to the measurements made by Horn et al. relating turbidity to the density of defects within the phase [31]. 3.3 Shear rate dependence of lamellar and MLV phase viscosity It is now pertinent to examine the shear rate dependence of the viscosity for these doped phases. The shear thinning behaviour of both a lamellar and MLV phase can be mod elled using the Sisko equation [32]: () = + K n , where is the innite shear viscosity, K the consistency and n the power law t parameter. We present in g. 4 shear viscosity against shear rate data for various POM doping concentrations for the lamellar phase, i.e. strictly prior to the onset of the lamellar-MLV transition. By contrast, in g. 5 the equivalent data for the MLV phase is presented for the various POM doping concentrations. The inset in g. 5 shows a typical complete rheological curve for an undoped 40/0 sample, where the classical hysteresis of

Fig. 4. Measured shear viscosity against shear rate and ts for lamellar phases before the onset of an MLV transition for various POM doping concentrations. Inset: tted consistency K as a function of POM volume fraction.

Fig. 5. Measured shear viscosity against shear rate and t for onion phases at various POM doping concentrations. Inset: measured shear viscosity against time for a 40/0 sample at various imposed shear rates for cone plate (c-p) and Couette cylinder (cyl) geometries for ascending and descending shear rate ramps. Dashed line is a guide to the eye.

a lamellar phase that has undergone a lamellar-MLV transition when subjected to an ascending and descending series of constant shear rate steps is observed, cf. Oliviero et al. [11], for example. This gure also illustrates that data derived from a polycarbonate concentric cylinder shear cell (cyl) and metal cone-plate geometries (c-p) are in good agreement. The data in g. 4 and 5 are obtained by multiple experiments; again with imposed shear rate steps over narrow shear rate ranges. It should be noted that, because the shear history of the sample is dierent, in contrast to the data presented in g. 2 that are from a single experiment over a wide shear rate range, the exact point of the lamellar-MLV transition may appear to be slightly shifted

Page 6 of 9

The European Physical Journal E

between the two data sets for samples of the same doping concentration. Firstly, upon examination of the shear rate dependence of the viscosity for the lamellar phases, data presented in g. 4, it can be seen that the shear thinning behaviour of the doped and undoped lamellar phases when tted by the Sisko equation are all well described by a 0.5 dependence that is typical of a lamellar phase that ows by the displacement of defects through the bulk phase [5, 11]. However, it is clear that the viscosity is systematically decreased by the addition of POMs without aecting this fundamental behaviour, while the consistency K falls from around 25 to 3 Pa sn with 0 to 4% POM doping fraction, as presented in g. 4 inset. This behaviour is consistent with our supposition that the POMs act to uidify the lamellar phase via a reduction of defect density. In g. 5 we present similar data and analysis for the viscosity of the MLV phase for various POM doping concentrations. Here we see a rather dierent behaviour from that of the lamellar phase: all the data fall on the same curve described by a power law index of n = 1 0.1 from a t to the Sisko equation with 2 Pa s. This result indicates that POM doping has a negligible eect on the viscosity of the corresponding MLV phase (note that the viscosity of the MLV phase is typically higher than that of the lamellar phases, K 200 Pa sn compared to the 253 Pa sn range for the lamellar phases), a rather surprising result considering the profound eect on the viscosity of the lamellar phase seen previously. This 1 depen dence is of a comparable order to that of 0.8 observed by Bergenholtz et al. [33] for an AOT/brine system and 0.89 observed by Nettesheim et al. for a Laponite clay particle doped C12 E4 /water system [26], but greater than the 0.53 dependence observed for a pure C12 E4 /water system by Mller et al. [10]. The calculation of such expou nents depends on the range of shear rates studied and the uniformity of the sample at high shear rates, where voids and solvent lubrication layers forming on the cell walls can result in a measured viscosity lower than the true viscosity. Here we may also see an eect of the small amounts of Brij R 30 impurities versus pure C12 E4 , as the undoped phase is also aected by the same discrepancy. 3.4 Critical shear stress for MLV formation Presented in g. 6 is the mean shear stress, m , against the shear rate during which the lamellar-MLV transition occurs for a range of POM doping concentrations. m is the average stress recorded during the lamellar-MLV transition over the total time of the imposed shear rate step, where the error bar gives an indication of the total stress range during the shear step. Typical measured shear stress against shear rate data are shown in the inset of g. 6, where the curves for 40/0 and 40/4 samples exhibit a similar form that is oset in shear rate by an order of magnitude. The form of these data is consistent between doped and undoped samples and similar to previous studies such as those of cationic lamellar phases for example [34]. A zero-order t to the data indicates that

Fig. 6. Mean critical shear stress against shear rate and t for 40% surfactant samples for a range of doping concentrations. Inset: measured shear stress against shear rate data for a series of shear rate steps for 40/0 and 40/4 samples.

there is an invariant (with respect to POM doping concentration) critical stress, c = 13020 Pa for 40% surfactant phases, and it is this invariant critical stress that lies at the root of the displacement of the lamellar-MLV transition to higher shear rates with increased POM doping (see g. 2): the POMs act to systematically reduce the viscosity of the lamellar phase but c remains constant, thus the doped system must be sheared at a correspondingly higher shear rate in order to attain an average stress of c across the sample and induce the lamellar-MLV transition.

3.5 Microscopic observations SAXS experiments can be employed to shift the focus of the study towards the nanoscopic scale structure corresponding to the various observed bulk states and transitions. Presented in g. 7 are rheological, SAXS and optical data for a 50% surfactant system at two dierent doping concentrations for a range of sequentially increasing shear rate steps, obtained through a combination of RheoSAXS and rheo-optical experiments. In g. 7 we present data for 50/2 and 50/4 samples sheared at 1, 4, 8, 10, and 15 s1 for 10 minutes per step in the Couette cylinder geometry of the RheoSAXS instruments. For the 50/4 sample comparative data from glass cone-plate geometry is presented in order to give an indication of the consistency between data from the two dierent measurement systems [35]. Also presented are optical micrographs between cross-polarisers of the sample texture and corresponding X-ray scattering patterns at selected shear rates, in this case at the tangential position for the 50/2 sample and the radial position for the 50/4 sample. The most striking feature of this data is that the 50/2 samples undergo a lamellar-MLV transition at approximately 8 s1 while the 50/4 sample continues to exhibit the shear thinning behaviour of a lamellar phase up to shear rates of 15 s1 . If one estimates the critical stress

J.P. de Silva et al.: Rheological behaviour of polyoxometalate-doped lyotropic lamellar phases

Page 7 of 9

Fig. 8. Radially averaged SAXS data for 40/0 and 40/2 samples at shear rates pre- and post-transformation from a lamellar to MLV phase (radial cell position).

Fig. 7. Measured shear viscosity against time for 50/2 and 50/4 samples at shear rates of 1, 4, 8, 10 and 15 s1 at 10 minutes per shear step (15 min pre-shear data not shown). Corresponding cross-polarised optical textures and comparative rheological data derived from glass cone-plate geometries under identical shear conditions. (Scale bar is 100 microns, arrow indicates direction of shear.) Corresponding SAXS spectra are acquired at the indicated cell position.

for the 50% surfactant systems to be about 160200 Pa, then at a shear rate of 20 s1 only approximately 20 Pa of stress is generated across the sample; thus a vastly increased shear rate, probably well above that practically possible before the sample is ejected from the shear cell, would be required to reach the transition threshold. Such suppression of MLV formation is similar to that observed by Berghausen et al. upon the addition of 0.5% poly(Nisopropyl acrylamide) (PNIPAm) to an SDS/water lamellar phase [15]. Optical micrographs for the 50/4 sample at 1 s1 are representative of a disoriented lamellar phase, and the corresponding SAXS spectra exhibit typical isotropic, narrow Bragg rings corresponding to the interlamellar repeat distance and indicating no preferential alignment in the sample; here the rst two Bragg orders are clearly visible by eye. The broad halo signal towards the extremities of the pattern is due to strong scattering from the POM particles. As the shear rate increases to 10 s1 the 50/2 sample has already formed an MLV phase, where a characteristic uniform texture is seen in the optical micrograph and the corresponding SAXS pattern remains isotropic in both tangential (pictured) and radial cell positions, as the MLV phase is isotropic. It is interesting to note that the Bragg rings observed for the 50/2 sample MLV phase are slightly broadened and displaced from those of the corresponding lamellar phase prior to any lamellar-MLV transition, and the origin of this eect will be discussed in detail below.

Up to a shear rate of 15 s1 the 50/4 sample continues to shear align, showing ne graining of the defects in the lamellar phase into streaks parallel to the ow direction, similar in structure to that observed in shearaligned C12 E5 /water lamellar phases [36] or AOT/brine and SDS/oil/water lamellar phases [37]. This is also conrmed by SAXS data that shows an anisotropic enhancement attributed to the scattering by the lamellae inside defects oriented under the ow. This enhancement must be attributed to the alignment of defects, as in the radial cell position the lamellae oriented by the ow (with their normal along the X-ray beam) do not contribute to the scattering.

3.6 Vesicle formation in doped phases One can more closely study the kinetics of the lamellarMLV structural transformation using in situ rheoSAXS data. Radially averaged SAXS data for 40/0 and 40/2 samples at shear rates of 1 and 10 s1 are presented in g. 8, where the data are representative of the lamellar phase and MLV phases respectively; the peaks correspond to the rst and second Bragg orders and the strong scattering prole from the POM particles is once again visible in the 40/2 sample data. Future detailed line shape analysis of such data should provide some insight into the effect of adsorbed POMs on the bilayer bending moduli [38, 39]; here we focus only on the Bragg peaks in the region 0.06 < q < 0.2 1 because the polycarbonate cell scatA ters signicantly at small q. The characteristic features of X-ray data from doped and undoped phases have been previously described in detail [8], where one nds that the lamellar period is invariant with POM doping and the POMs preferentially adsorb onto the surface of the surfactant bilayers, as indicated by the change in relative intensities of the rst- and second-order Bragg peaks with POM doping.

Page 8 of 9

The European Physical Journal E

With regards to the lamellar-MLV transition the Xray data reveal two important points: upon forming an MLV state there is a broadening and shift to higher q of the Bragg peaks from that of the pre-transition lamellar phase, consistently around six percent for the rst and second Bragg orders. Pre-transition rst and second order Bragg peaks that are relatively sharp can be seen for both the 40/0 and 40/2 samples, giving a well-dened interlamellar spacing of around 8 nm for both the 40/0 and 40/2 samples at = 1 s1 . In the MLV state the peak cen tre shifts towards larger q, indicating a reduced lamellar spacing, possibly a result of the compression or collapse of the structure within the vesicle or the expulsion of solvent [25,40,41]. It is important to note that this is not the eect of drying in the sample over the duration of the experiment, as SAXS data from lamellar phases sheared for many hours at a constant shear rate below the lamellarMLV shear transition show no change in interlayer spacing from that of the lamellar phase. We derive a coherence length from the FWHM of the broadened Bragg peak in the MLV state (correcting for the experimental resolution) of around 30 lamellae, or around 250 nm, which shows that the X-rays do not probe the total size of the MLV (2040 microns or more than 3000 lamellae) but rather some smaller coherent length scale with a reduced lamellar spacing, perhaps the result of imperfect stacking in the multilayers where several coherent domains are separated by less ordered regions. The second important point derived from data in g. 8 is that the nanoparticles remain encapsulated within the surfactant layers of the MLVs, as the relative intensities of the peaks are preserved pre- and post-transformation between the doped and undoped samples, indicating that the nanoparticles remain adsorbed to the surfactant membranes after the lamellar-MLV phase transition. This result implies a viable a method of encapsulating nanoparticles within inexpensive and relatively easily formed surfactant vesicles, which may have a wide variety of applications including biological payload delivery. In g. 9 we present the interlamellar spacing, d, as a function of shear rate for a range of POM doping fractions, as derived from radially averaged rheoSAXS data. At low shear rates in a lamellar state, d is well dened for all POM doping concentrations at 8.05 nm. At the onset of the lamellar-MLV transition d begins to decrease in all cases, reaching a steady 7.4 nm for the undoped sample but increasingly smaller values for the doped phases, down to around 6.5 nm for the 4% POM doped sample. The origin for such a change in compression with POM doping is unknown; surfactant vesicles have already been shown to deform into polyhedral shapes that ll available space without excess space-lling solvent [42]. Such an effect may also be responsible for the reduced length scales derived from X-ray experiments, as discussed previously. Moreover, it was observed that after cessation of shear the interlamellar spacing may increase again at rest, indicating the reabsorption of solvent and the swelling or rupture of the vesicle phase, consistent with freeze-fracture electron microscopy measurements on AOT/brine phases [42]. Structural dierences in doped and undoped vesicle phases

Fig. 9. Interlamellar spacing d derived from rheoSAXS data as a function of shear rate and POM doping concentration.

could be revealed by electron microscopy experiments, for example [42,34].

4 Conclusions
We have studied the inuence of [PW12 O40 ]3 polyoxometalate nanoparticle doping on the rheological properties of a Brij R 30/water lyotropic lamellar phase. We show that the shear viscosity, yield stress and instantaneous elastic modulus of the lamellar phase are all systematically decreased with increased POM doping fraction. A shear-induced transition to a multilamellar vesicle phase can be oset to higher shear rates, or even completely suppressed, with increasing POM doping fraction. We measure a critical stress of around 130 Pa for the lamellar-MLV transition in the 40% surfactant systems, which is invariant with respect to POM doping fraction; the constant nature of this critical stress causes a shift of the lamellar-MLV transition to higher shear rates with increasing POM doping. We show that the areal density of POMs adsorbed to the surfactant bilayer is the key parameter required to quantitatively describe the viscosity variation with POM doping. From the decrease in turbidity, elastic modulus and yield stress, we postulate that POMs act to suppress defects within the lamellar phase. X-ray data shows that the nanoparticles remain encapsulated within the MLV membranes, thus indicating a rather inexpensive methodology for the encapsulation of metallic nanoparticles within surfactant vesicles.

JPdS thanks Triangle de la Physique for a postdoctoral fellowship (under project RheoSAXS), C. Baravian, I. Bihannic, L. Michot, E. Paineau and D. Constantin for help with RheoSAXS experiments and useful discussions. ASP beneted from a Marie Curie fellowship under contract no. MEST-CT2004-514307. We thank D. Langevin for the use of an AP MCR 301 rheometer, Anton-Paar France for RheoSALS experiments and the impromptu loan of an air-compressor, Laboratoire

J.P. de Silva et al.: Rheological behaviour of polyoxometalate-doped lyotropic lamellar phases FAST for the use of a glass rheometer measurement system and SOLEIL for beamtime allocation. We thank the referees for pertinent comments that improved the manuscript. 23.

Page 9 of 9

References
1. R.G. Larson, The Structure and Rheology of Complex Fluids (Oxford University Press, New York, 1999) pp. 3-4. 2. O. Diat, D. Roux, J. Phys. II 3, 9 (1993). 3. O. Diat, D. Roux, F. Nallet, J. Phys. II 3, 1427 (1993). 4. D. Roux, F. Nallet, O. Diat, Europhys. Lett. 1, 53 (1993). 5. C. Meyer, S. Asnacios, M. Klman, Eur. Phys. J. E 6, 245 e (2001). 6. P. Oswald, M. Klman, J. Phys. (Paris) Lett. 12, L411 e (1982). 7. J.F. Keggin, Proc. R. Soc. London, Ser. A 144, 75 (1934). 8. A.S. Poulos, D. Constantin, P. Davidson, M. Impror, B. e Pansu, P. Panine, L. Nicole, C. Sanchez, Langmuir 24, 6285 (2008). 9. A.S. Poulos, D. Constantin, P. Davidson, M. Impror, P. e Judeinstein, B. Pansu, J. Phys. Chem. B 114, 220 (2010). 10. S. Mller, C. Brschig, W. Gronski, C. Schmidt, D. Roux, u o Langmuir 15, 7558 (1999). 11. C. Oliviero, L. Coppola, R. Gianferri, I. Nicotera, U. Olsson, Colloids Surf. A 228, 85 (2003). 12. Y. Kosaka, M. Ito, Y. Kawabata, T. Kato, Langmuir 26, 3835 (2010). 13. H.E. Warriner, S.H. Idziak, N.L. Slack, P. Davidson, C.R. Sanya, Science 271, 969 (1996). 14. S.L. Keller, H.E. Warriner, C.R. Sanya, J.A. Zasadzinski, Phys. Rev. Lett. 78, 4781 (1997). 15. J. Berghausen, J. Zipfel, P. Lindner, W. Richtering, J. Phys. Chem. B 105, 11081 (2001). 16. V. Ponsinet, P. Fabre, J. Phys. II 6, 955 (1996). 17. J. Arrault, C. Grand, W.C.K. Poon, M.E. Cates, Europhys. Lett. 38, 625 (1997). 18. Y. Suganuma, M. Imai, K. Nakaya, J. Appl. Crystallogr. 40, s303 (2007). 19. F. Nettesheim, I. Grillo, P. Lindner, W. Richtering, Langmuir 20, 3947 (2004). 20. D.J. Mitchell, G.J.T. Tiddy, L. Waring, T. Bostock, M.P. McDonald, Faraday Trans. I 79, 975 (1983). 21. J.P. de Silva, D. Petermann, B. Kasmi, M. Imperor, P. Davidson, B. Pansu, F. Meneau, J. Perez, E. Paineau, I. Bihannic, L. Michot, C. Baravian, J. Phys. Conf. Ser. 247, 012052 (2010). 22. We use a pre-shear treatment rather than a quench into the low-temperature isotropic region of the phase diagram [10], as at 50% surfactant concentration there is no low-temperature lamellar-isotropic phase transition, as

24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35.

36. 37. 38. 39. 40. 41. 42. 43.

seen in the C12 E4 phase diagram [20] and conrmed by observations of cross-polarised optical texture after a quench to 0 C. B. Medronho, M. Miguel, U. Olsson, Langmuir 23, 5270 (2007). F. Nettesheim, J. Zipfel, P. Lindner, W. Richtering, Colloids Surf. A 183-185, 563 (2001). O. Diat, D. Roux, F. Nallet, Phys. Rev. E 51, 3296 (1995). F. Nettesheim, J. Zipfel, U. Olsson, F. Renth, P. Lindner, W. Richtering, Langmuir 19, 3603 (2003). A.S. Poulos, PhD thesis (2009). S. Paasch, F. Schambil, M.J. Schwuger, Langmuir 5, 1344 (1989). J. Munoz, C. Gallegos, V. Flores, J. Disp. Sci. Technol. 7, 453 (1986). P. Panizza, D. Roux, V. Vuillaume, C.-Y.D. Lu, M.E. Cates, Langmuir 12, 248 (1996). R.G. Horn, M. Kleman, Ann. Phys. (Paris) 3, 229 (1978). P. Versluis, J.C. van de Pas, J. Mellema, Langmuir 13, 5732 (1997). J. Bergenholtz, N.J. Wagner, Langmuir 12, 3122 (1996). P. Partal, A.J. Kowalski, D. Machin, N. Kiratzis, M.G. Berni, C.J. Lawrence, Langmuir 17, 1331 (2001). Shear viscosities measured here for the pure undoped lamellar phase at 40% surfactant concentration are around an order of magnitude higher than those given by Mller u et al. for a pure 40% C12 E4 lamellar phase at equivalent shear rates [10]. We can identify two reasons for this: previous experiments demonstrate that there is a clear eect of the nature of the cell material and surface roughness [43]; the second is the possible inuence of impurities, however this is less likely to explain the observed discrepancy. Our measurements indicate that this discrepancy is due to the use of smooth glass shear cells rather than metal or polycarbonate; a 40% Brij R 30 sample sheared with a smooth glass geometry very closely reproduces the viscosities measured by M ller et al. u C.-Y.D. Lu, P. Chen, Y. Ishii, S. Komura, T. Kato, Eur. Phys. J. E 25, 91 (2008). L. Courbin, P. Panizza, Phys. Rev. E 69, 021504 (2004). J.T. Brooks, C.M. Marques, M.E. Cates, J. Phys. II 1, 673 (1991). M.-F. Ficheux, A.-M. Bellocq, F. Nallet, Eur. Phys. J. E 4, 315 (2001). A. Lutti, P.T. Callaghan, Eur. Phys. J. E 24, 129 (2007). A. Leon, D. Bonn, J. Meunier, J. Phys.: Condens. Matter 14, 4785 (2002). T. Gulik-Krzywicki, J.C. Dedieu, D. Roux, C. Degert, R. Laversanne, Langmuir 12, 4668 (1996). N. Jager-Lzer, J.F. Tranchant, V. Alard, J. Doucet, J.L. e Groissiord, J. Rheol. 42, 417 (1998).

You might also like