Sie sind auf Seite 1von 10

Journal of Colloid and Interface Science 302 (2006) 537546 www.elsevier.

com/locate/jcis

Soft water-soluble microgel dispersions: Structure and rheology


A. Omari a, , R. Tabary b , D. Rousseau b , F. Leal Calderon a , J. Monteil a , G. Chauveteau b
a Laboratoire des Milieux Disperss Alimentaires, ISTAB, Avenue des Facults, 33405 Talence cedex, France b Institut Franais du Ptrole, 1-3 avenue de bois Prau, 92506 Rueil Malmaison cedex, France

Received 24 March 2006; accepted 6 July 2006 Available online 11 July 2006

Abstract The size and structural characteristics of polyacrylamide-based water-soluble microgel dispersions were investigated by optical and rheological methods. Microgel hydrodynamic radii Rh were measured by light scattering and derived from intrinsic shear viscosity []0 . The variations of 3 and [] with the crosslink density N , follow the scaling law R 3 N with close to 0.63, in good agreement with the simple structural Rh x 0 h= x model proposed in this paper showing how the exact value of depends on inner structural details of the microgel. The plateau viscosity versus particle apparent volume fraction shows a monotonous change from hard sphere dispersions (high crosslink density of microgels) to exible linear polymer solutions. Measurements of the rst normal stress difference N1 show that increasing the microgel crosslink density affects the system viscosity more than its elasticity. Under oscillatory shear ow, loss and storage moduli undergo both qualitative and quantitative changes with crosslink density. At moderate concentrations, the elastic modulus is the most affected and its slope in low frequency regime decreases from two to less than one as Nx increases. We discuss the experimental results within the frame of knowledge on linear, branched polymer solutions and soft microgel suspensions. 2006 Elsevier Inc. All rights reserved.
Keywords: Microgel; Crosslink; Shear viscosity; Loss modulus; Storage modulus; First normal stress difference

1. Introduction Micrometric and sub-micrometric particles are widely used on an industrial scale and are usually designed to have suitable properties depending on the target application [1]. For controlled drug delivery and, more recently, in Enhanced Oil Recovery (EOR) processes, new soft particles, which we refer to as microgels, are very attractive [2]. For medical diagnosis, surface grafting of special molecules on rigid magnetic particles is aimed at [3]. Different techniques have been developed to manufacture microgel particles. One of them consists in crosslinking a primary linear water-soluble polymer solution under controlled shear [4]. The emulsion polymerization technique is more attractive since it leads very easily to well-calibrated particles, which softness is mainly monitored through monomerto-crosslinker concentration ratio [57]. In the past, signicant efforts were focused on investigating the swelling/de-swelling
* Corresponding author. Fax: +33 5 56 37 03 36.

E-mail address: a.omari@istab.u-bordeaux1.fr (A. Omari). 0021-9797/$ see front matter 2006 Elsevier Inc. All rights reserved. doi:10.1016/j.jcis.2006.07.006

mechanism that particles undergo by changing crosslink density [8,9], temperature [10], particle charge density and/or solvent parameters such as pH [11] or salt content [9,12]. For that purpose, photon correlation spectroscopy (PCS) and smallangle neutron scattering (SANS) [1] are more suitable in dilute regime and turbidimetry in higher concentrations [13]. The particle swelling magnitude is then evaluated through the measurement of Q = (Rh /Rh0 )3 where Rh and Rh0 stand for the hydrodynamic radius of particles in swelled and collapsed states, respectively. As expected, Rh is controlled by the particle crosslink density Nx . In a recent paper dealing with nano [9] where colloidal particles, Q was found to vary as Nx depends on monomer solvent quality. When used for Enhanced Oil Recovery purposes, for instance in water shutoff and conformance control operations, microgels adsorb onto the pore surface oil-bearing layers, reduce effective pore size and thus water permeability. To be efcient, microgels must be calibrated, highly water-soluble and soft enough for adsorbed layer thickness to be easily reduced by capillary pressure so that oil permeability is not affected. In other terms, both particle size

538

A. Omari et al. / Journal of Colloid and Interface Science 302 (2006) 537546

and their intrinsic mechanical characteristics must be specically designed for the reservoir to be treated. In most applications, these microgels are submitted to hydrodynamic forces and their rheological behaviour should be well understood and monitored. The main key parameters that govern such rheology are: particle surface charge, particle size and distribution, microgel phase volume fraction and particle internal structure. Under shear ow, both the plateau viscosity and the shear-thinning behaviour deviate appreciably from that of hard spheres [8]. Generally speaking, these dispersions are expected to exhibit rheological behaviours ranging from that of hard sphere suspensions to that of linear polymer solutions with a determining role of crosslink density. So far, most published investigations dealt with highly crosslinked particles and were restricted to linear properties under steady and oscillatory shear ow. Moreover, little is known about the rheological properties of this type of suspensions in the non-linear regime. In this paper, we propose to investigate the effects of change of crosslink density on the structural and rheological properties of well-characterized microgel dispersions. A wide range of crosslink density was explored, from linear polymer to highly crosslinked microgels. Furthermore, the concentration range is such that both dilute and concentrated regimes were explored. In steady shear experiments, viscosity and rst normal stress difference were measured in both linear and non-linear regimes. Dispersions were also submitted to oscillatory shear ow, and measurements of loss and storage moduli were performed in the linear regime, within a frequency range accessible to the equipment used. 2. Experimental The particles used in this study were acrylamide-based microgels kindly supplied by SEPPIC Company (France). They were chemically similar to those previously used by Senff and Richtering [10]. They were obtained by using emulsion polymerization technique, which generally leads to a narrow size distribution. The monomer concentration in emulsion droplets was kept constant, and the particle internal crosslink density was monitored by varying the amount of added crosslinker. As the emphasis here is on general features, the detailed chemistry of particles is not relevant and is not described. A linear polymer with similar chemical composition was also used. The solvent was brine containing 20 g l1 NaCl and 0.4 g l1 NaN3 (used as a bactericide and as a stabilizer to prevent free radical degradation), both dissolved in distilled water. The pH was adjusted to 7 for all samples. Microgel sizes were measured by the dynamic light scattering technique (Photon Correlation Spectroscopy), which allows the determination of the hydrodynamic radius Rh of dispersed particles through their translational diffusion coefcient D0 : Rh = kB T /6s D0 , where kB is the Boltzmanns constant, T is the absolute temperature and s is the solvent viscosity. Experiments were done at sufciently low angles in order to meet the condition qRh < 1, q being the scattering vector [q = (4n/) sin(/2)]

where n is the solvent refractive index and the wavelength of the laser source ( = 632.8 nm). The dispersions used were sufciently diluted (<200 ppm) to prevent multiple scattering and interaction effects. The experiments were performed using the N4 Plus 3000 apparatus from Beckman Coulter at 3 different angles (15.7 , 23 and 30.1 ). Further characterizations were then carried out by measuring the viscosity for very dilute dispersions using a low shear viscometer (LS 30 from Contraves) equipped with a Couette geometry (0.5 mm gap) at shear rates ranging from 0.1 to 70 s1 . In this way, it was possible to accurately determine the relative viscosity r0 , dened as the plateau viscosity ratio between the dispersion and the solvent. For higher concentrations, a controlled-stress rheometer (MCR 300 from Anton-Paar Physica) and a controlled-strain rheometer (ARES from Rheometric Scientic) were used, both equipped with coaxial cylinders or cone and plate geometry. These instruments enabled us to measure the steady shear viscosity in linear (if it does exist) and in non linear regimes. The rst normal stress difference, N1 , and the dynamic moduli versus frequency in oscillatory shear linear regime were also measured. All measurements were performed at 20 C unless otherwise specied. 3. Results and discussion 3.1. Microgel size determination Microgels with a high crosslink density were rst examined with an optical microscope, which qualitatively reveals that particles are spherical and closely monodisperse. The crosslink density Nx is dened as the ratio of crosslinker-to-monomer concentrations, assuming that both of them are homogenously distributed. The value of Nx ranges from 0 to 0.3%; zero corresponding to linear polymer. In Fig. 1, particle radii measured by dynamic light scattering were plotted versus Nx . Although the results are scattered, this gure shows that microgel radius decreases with crosslink density according to a scaling law:
3 Rh = Nx

(1)

with close to 0.63. For microgels obtained by emulsion polymerization of methacrylic acid and ethyl acrylate cross-linked with diallyl phthalate, Tan et al. [9] found a larger power law exponent ( 1.5). In the model we propose, microgel size and structure derive from well known scaling laws [14] as follows. If we consider a microgel as made of nb subunits linked by the covalent bonds formed by the crosslinker, which we will refer to as blobs, characterized by a mass m and a distance between two consecutive crosslinks, the microgel radius of gyration Rg is given by:
3 Rg = nb 3 .

(2)

As a rst approximation, and m should be related by a Mark Houwink-type relationship: = m , (3)

A. Omari et al. / Journal of Colloid and Interface Science 302 (2006) 537546

539

Fig. 1. Plot of hydrodynamic radius (Q) and particle intrinsic viscosity (F) of microgel suspensions as a function of crosslink density. ( ) []0 = (Nx )0.63 ; 3 0 . 63 ( - -) Rh = (Nx ) .

where the exponent is the excluded volume parameter, the values of which depend on chain stiffness and solvent quality. For exible chains, we expect values to be 0.5 and 0.6, for monomers in and good solvent respectively, while for fully stiff chains = 1. The microgel mass M is related to m by: M = mNx . Combining Eqs. (2) to (4) gives:
3 (13) Rg . = M 3 Nx

3.2. Shear viscosity measurement Relative shear viscosity measurements were performed on each sample at variable microgel concentrations, C . At low shear rates and low concentrations, the reduced specic viscosity spr is expressed by: spr = (r0 1)/C, where spr is a linear function of C , r0 being the relative plateau viscosity. From the spr (C) plot, the microgel intrinsic viscosity []0 is determined by extrapolation at C = 0, and the Huggins constant KH from the slope. []0 and KH , respectively, represent the hydrodynamic volume occupied by an isolated particle and the particleparticle interactions. Table 1 summarizes the results obtained for most microgel samples used, and for the linear polymer. On the one hand, we observe an increase in the Huggins constant with crosslink density from 0.2 for linear polymer to 1 for the highest Nx value. Even for highly crosslinked microgel dispersions, KH = 1 as expected for repulsive hard spheres without exceeding this value, thus showing the absence of attractive interactions between microgels. On the other hand, microgel intrinsic viscosity (which is proportional to particle volume) decreases when Nx increases. Intrinsic viscosity values are also reported in Fig. 1, showing that both []0 3 exhibit the same scaling law versus N . For linear polyand Rh x mers of molecular weight M w, []0 is linked to the radius of 3 /M w, where g gyration Rg through the relationship: []0 = R is a universal constant. The absolute values of intrinsic vis cosities are signicantly larger (nearly one order of magnitude larger) than that previously studied in literature. The microgels used in this study can thus be considered to be soft particles. As it is customary to do when dealing with microgels, experimental results are more conveniently expressed in terms of apparent particle volume fraction , rather than in mass concentration [8,15,16]. This is achieved in dilute regime by matching the linear terms in Huggins and Einstein equations: r0 = 1 + C []0 + KH C 2 []2 0, (7)

(4)

(5)

3 is related For monodisperse microgels, M is constant so that Rg to Nx by a scaling law: 3 Rg = Nx

(6)

with = 3 1, so we can predict = 0.5, 0.8 and 2, for exible chains in solvent, exible chains in good solvent and for fully stiff chains, respectively. This simple model has predictive capability, as soon as the chain chemistry and solvent quality are known. In the case of our experiments, with the chemistry of microgels used and the experimental conditions (temperature and water salinity), our model predicts an value between 0.5 and 0.8 in good agreement with experimental results ( = 0.63), since the mean monomer number between successive crosslinks is small enough for volume exclusion effects to be limited. In the case of the experimental results reported in Ref. [9] where was found to be equal to 1.5, the above model suggests that the polymer chain is rigid, in agreement with polymer chemistry and water salinity. Indeed, such exponent was observed for free salt medium and for neutralisation degree of polymer acidic groups exceeding 0.4 (signicantly charged microgels). In the concentrated regime, the radius of soft microgels is expected to be reduced because of its compressibility. However, more sophisticated techniques are needed to determine the particle radius in that regime.

540

A. Omari et al. / Journal of Colloid and Interface Science 302 (2006) 537546

Table 1 Microgel characteristics Sample number 1 2 3 4 5 6 Crosslink density, Nx (%) 0 0.05 0.10 0.15 0.20 0.30 []0 (cm3 g1 ) 2840 284 1335 134 884 88 670 67 559 56 428 43 KH 0.26 0.03 0.21 0.02 0.40 0.04 1.0 0.1 Rh (m) 0.20 0.02 1.60 0.16 1.40 0.14 1.33 0.13 1.40 0.14 1.15 0.12

crosslink density. At low Nx values, interpenetration is possible over large scales, and dominates both structure and rheology in semi-dilute regimes, so that the behaviour is close to that of linear polymers. By contrast, at large Nx values, microgels are more and more expected to behave like hard spheres. When particle inter-penetration can be neglected (high crosslink density), one can dene an effective volume fraction, eff , by the Batchelor equation [10,13]:
2 r0 = 1 + 2.5eff + 5.9eff .

r0 = 1 + 2.5.

(8)

Thus giving: = 2C []0 /5. For high concentrations (i.e., high volume fractions), we measured shear viscosities versus shear rate. Fig. 2 shows the variation of the low-shear relative viscosity as a function of for the six samples. For some of them, no plateau viscosity was observed at high volume fractions, and the data between brackets corresponds to the lowest accessible shear rate. For estimation of microgels softness (deviation from hard sphere suspensions), it is convenient to use the empirical KriegerDougherty equation: r0 = (1 /m )2.5m . The curve corresponding to hard spheres is plotted on the same gure for comparison. In this relation, the maximum randompacking fraction m is taken as being equal to 0.64 (monodisperse spheres). It is to be noted that, for general use, the quantity 2.5m should be replaced by []0 m . As pointed out earlier, microgel dispersions exhibit lower dependence of r0 on than hard sphere dispersions [8], but signicantly higher than linear polymer [2]. Such dependence becomes weaker as crosslink density decreases, owing to partial particle inter-penetrability. The range of such inter-penetrability is expected to be of the order of the blob mesh size of the internal network. The actual conformation of the microgels in concentrated regime is the result of two opposite factors: particle inter-penetrability, given by shell properties, and compressibility, governed by core

The ratio of the apparent volume fraction to the effective volume fraction eff is a useful parameter, since it is representative of particle compressibility. Nevertheless, since only rst order and quadratic terms are involved, the data does not t at high concentrations. However, such an approach is not pertinent in the present study, for at least two reasons: our microgels are very soft, and can interpenetrate substantially, and concentrations used are too high. Nevertheless, one should keep in mind that imbedding a particle in an imposed shear ow causes an extra shear rate by modifying the ow pattern [17]. Such shear rate excess is, of course, a decreasing function of the particles inner permeability. In another picture, the shared zone of neighbouring particles (increasing with shell thickness) decreases as crosslink density increases. As a consequence, the local shear rate becomes higher, and the local shear stress rises. Conversely, decreasing crosslink density induces signicant particle interpenetration, thus weakening the dependence of r0 on . Roovers [18] has reported similar behaviour for concentrated star polymer solutions when the number of branches per molecule increases. For linear polymer, as indicated by the dashed line in Fig. 2, the relative viscosity closely follows the power law, r0 = 3.4 , which is known to t experimental data correctly in the high concentration regime [19]. As expected, the rate of viscosity rising for microgel suspensions increases monotonically with Nx . In the high shear rate regime, as shown in Fig. 3, shearthinning index n increases signicantly with . Such shearthinning index n is the exponent in the pseudo-plastic regime,

Fig. 2. Variation of relative plateau shear viscosity with apparent volume fraction for different crosslink densities: (!) 0%; (E) 0.05%; (P) 0.10%; (1) 0.15%; (+) 0.20%; ( |) 0.30%; ( ) KriegerDougherty relation; ( ) correspond to the power law: r0 = 3.4 .

A. Omari et al. / Journal of Colloid and Interface Science 302 (2006) 537546

541

Fig. 3. Shear-thinning index versus apparent volume fraction for various Nx . (") 0% (linear polymer); (F) 0.05%; (Q) 0.10%; (+) 0.20%; ( |) 0.30%.

crease in concentration is located in the inter-particles shared zone, where viscous dissipation is predominant. Therefore the observed shear thinning is more pronounced. 3.3. Measurement of rst normal stress differences First normal stress difference N1 was measured as a function of the macroscopic applied shear rate using a cone and plate geometry. Most measurements were carried out in the high apparent volume fraction range, because data at low was too scattered. They were also restricted to shear rates high enough to ensure acceptable measurement accuracy, but nevertheless not too high, in order to prevent ow instabilities [20,21]. In polymers, normal stress difference in shear ow results from the anisotropic alignment of molecular segments. It depends therefore on the tension along streamlines. Fig. 5a gives the variation of N1 with for each sample at apparent volume fraction = 5.75. It shows that suspensions of more dense microgels clearly exhibit higher elastic characteristics, when compared at the same imposed shear rate. To explain such behaviour, we emphasize again that, instead of the imposed shear rate, one should use the effective shear rate that segments actually experience in the overlap region. In that region, tightening of ow streamlines and consequently, the increase of local shear rate is more important as the inner density of microgels is increased. Moreover, in order to have an overall view of the incidence of structural change on viscoelasticity under permanent shear ow, it is more convenient to represent N1 versus shear stress [22]. By doing so, the hierarchy is inverted (Fig. 5b) and N1 is a decreasing function of Nx at a xed shear stress, meaning that the elastic-to-viscous force ratio is a decreasing function of microgel crosslink density. In other words, increasing the internal density of microgels affects viscosity more than elasticity for the same . This is more clearly evidenced when the microgel volume fraction is increased further, as shown in Fig. 5c. At that volume fraction, it was unfortunately not possible to obtain a homogeneous preparation for microgels with the highest density.

Fig. 4. Normalized relative shear viscosity versus shear rate at = 0.34 for linear polymer (Nx = 0%) and a microgel suspension (Nx = 0.3%), temperature is 30 C. The continuous lines are data ts using the Carreau model.

where viscosity decreases according to the power law: r = n . For linear polymer, n rises smoothly and should later reach a limiting value close to 0.8, which characterizes concentrated high-molecular-weight linear polymers [19]. For microgel dispersions, the asymptotic value obviously increases with crosslink density, due to the increase of local connement of peripheral microgels shell. Furthermore, an isolated particle with low crosslink density should deform easily under shear ow, for volume fractions far below m . Thus, increasing particle internal density is expected to reduce shear-thinning behaviour. The results shown in Fig. 4 conrm that, in dilute regime ( < m ), shear thinning is due to individual particle deformation under shear forces, even if a larger shear rate interval should be scanned to conrm this nding. By contrast, for > m , the shear-thinning index increases with crosslink density. Moreover, the shape of the n vs curves is noticeably modied as Nx varies, but the asymptotic limit is somewhat similar. Such behaviour is indebted to the intrinsic structure of the system. When particles are highly penetrable (low Nx ), increasing the system concentration results in an homogeneous increase of concentration. For partially penetrable particles (higher Nx ), the resulting in-

542

A. Omari et al. / Journal of Colloid and Interface Science 302 (2006) 537546

(a)

(b)

(c) Fig. 5. First normal stress difference for = 5.75: (a) versus shear rate and (b) versus shear stress. (c) First normal stress difference versus shear stress for = 7.

3.4. Oscillatory shear measurement In this section, we report measurements of both loss and storage modulus (G and G ) as a function of frequency, under small amplitude oscillatory shear. All samples were tested

but only some results are presented for clarity. For the linear polymer (Fig. 6a), we observe the expected behaviour. Polymer solutions are essentially viscous at low frequencies, tending to t the scaling laws: G = 2 and G = . At high frequencies, elasticity dominates (G > G ). This corresponds to

A. Omari et al. / Journal of Colloid and Interface Science 302 (2006) 537546

543

(a)

(b)

(c) Fig. 6. Loss modulus (solid symbols) and storage modulus (hallow symbols) versus frequency at indicated apparent volume for linear polymer (a) and for microgel samples: (b) Nx = 0.10%; (c) Nx = 0.30%.

a Maxwell-type behaviour with a single relaxation time that may be determined from the crossover point and, as expected, this relaxation time increases with concentration. Therefore microgel dispersions display a similar liquid-like behaviour at low volume fractions but deviate from this behaviour at high values. Moreover, for identical apparent volume fractions,

the deviation widens out when particles internal density increases. For instance, at intermediate crosslink densities and low volume fractions, the behaviour is similar to that of linear polymer, except in the low frequency regime, where the exponent corresponding to the storage modulus is severely reduced to less than one. In the low frequency regime, G and G

544

A. Omari et al. / Journal of Colloid and Interface Science 302 (2006) 537546

(a)

(b)

(c) Fig. 7. Plot of tan() versus frequency at indicated apparent volume fraction: (a) linear polymer; (b) Nx = 0.10%; (c) Nx = 0.30%.

change in the same way, with a subsequent change in the G slope below the crossover point (Fig. 6b). In the picture proposed above, soft microgels are described as deformable (soft) core-shell particles. Their elasticity at high is attributed to the shell constituted by the dangling polymer branches located at the periphery, while elasticity at low reects the slow dynamics of cores. Similar behaviour was reported earlier by

Pham et al. [23] for latex dispersions containing associative copolymers, with either attractive or repulsive particleparticle interactions. These authors also proposed a simple model to extract contributions from hard spheres and polymer solutions. Unfortunately, because of microgel compressibility and interpenetrability, such a method cannot be used for quantitative treatment of our data.

A. Omari et al. / Journal of Colloid and Interface Science 302 (2006) 537546

545

Fig. 8. Plot of G values at frequency of 1 rad s1 , as a function of apparent volume fraction for various Nx . (") Linear polymer; (F) 0.05%; (Q) 0.10%; (2) 0.15%; (+) 0.20%; ( |) 0.30%. (The dashed lines are only guides for eyes.)

For higher crosslink densities and high volume fractions, microgel dispersions exhibit a power-law dependence of the viscoelastic moduli, with at intermediate frequencies (see Fig. 6c for example) where we have G = kG = p . This behaviour reveals the existence of ordered structures in the dispersions and the exact value of the exponent p is dependent on structural details. In that frequency range, their viscoelasticity is therefore governed by a relaxation time spectrum of the form: H () = H0 (/0 )p , where H0 (Pa) and 0 (s) are scaling factors [24]. This is more clearly shown in Figs. 7a7c, where the viscous-to-elastic modulus ratio (or the dissipated to stored energy ratio): k = tan() = tan(p/2) is plotted versus frequency for some samples at various . For the linear polymer, the behaviour is usual inasmuch as k is a monotonically decreasing function of frequency and is lowered by increasing polymer concentration. The corresponding p value is greater than 0.5 at intermediate frequencies, since polymer solutions are more viscous than elastic; G > G . For all microgel samples, deviation from linear polymer behaviour is both qualitative and quantitative. First, no asymptotic divergence in the low frequency regime is observed, and secondly tan() is quasi-insensitive to frequency changes, especially for high volume fractions and high crosslink densities. For low crosslink density (Fig. 7b), the viscous-to-elastic modulus ratio is greater than 1 (giving 0.5 < p < 1 since p = 2/ tan1 (G /G )) in the intermediate frequency range. This indicates that such dispersions behave as very soft gels or semi-dilute solutions of branched polymers. The exponent is similar to that obtained by crosslinking exible linear polymers below the gel point with a lack of crosslinking agent [22]. As internal crosslink density increases, particles steric connement increases and suspensions take a gel-like structure with p < 0.5 at signicant values (see Fig. 7c). Indeed, for a solution-like system, the viscous part dominates, giving p > 0.5. For gelled systems however, the elastic part dominates, and p is then less than 0.5. At the transition from solution-like state to gel-like state (the gel point), mechanical and structural properties of the system are described by scaling laws which may be derived using percolation theory. In that regime, both G and G scales

as p where the exponent p is constant over a large frequency interval and which exact value is dependent on the involved system. In other respects, since for microgel dispersions elasticity arises from particleparticle contacts, a critical volume fraction c can be dened, below which G should drop to very low values. For > c , G rises sharply before reaching a pseudo-plateau. Moreover, as the distance between neighbouring particles scales as 1/3 , G may be written in the form: G = Gp 1 (/c )1/3 , where Gp represents the intrinsic elasticity of microgels and c is usually taken equal to m . Some authors have already used this relation in an attempt to plot data on a master curve experimental data obtained for various crosslink densities [25]. For that purpose, Gp was taken as equal to the plateau modulus of the bulk gel at the considered crosslink concentration. The data did not t with a master curve, probably due to particle compressibility, and adjustment parameters were necessary to make them collapse. Therefore, and since in the present study crosslink densities are an order of magnitude lower, we did not perform any scaling of our data and only qualitative arguments are put forward to explain the observed behaviour. Fig. 8 shows the evolution of the elastic shear modulus G ( = 1 rad s1 ) as a function of . For low crosslink densities, G is a growing function of , as for linear polymers, since particles may easily entangle without signicant compression. Further increase in crosslink density obviously reduces the critical volume fraction, and levels off the elasticity at high volume fractions, approaching the behaviour of dense microgel suspensions. 4. Conclusion Microgels structure and rheological properties of microgel suspensions were studied with a particular emphasis on the inuence of internal crosslink density. Crosslink densities were in the low range, allowing particles to partially interpenetrate. We proposed a simple model that was able to predict the dependence of hydrodynamic radius of microgel particles on their inner crosslink density, at least qualitatively. Rheological behav-

546

A. Omari et al. / Journal of Colloid and Interface Science 302 (2006) 537546

iour of studied Microgel suspensions is shown to lie between that of linear polymer solutions and that of hard sphere suspensions with a smooth change as crosslink density increases. In this work, we only propose a phenomenological interpretation of experimental evidences and more efforts are to be undertaken to establish quantitative modelling of the rheology of such systems. Acknowledgments The authors wish to thank IFP for nancial support and SEPPIC Company for supplying microgels. References
[1] B.R. Saunders, B. Vincent, Adv. Colloid Interface Sci. 80 (1999) 125. [2] D. Rousseau, G. Chauveteau, M. Renard, R. Tabary, A. Zaitoun, P. Mallo, O. Braun, A. Omari, SPE 93254, Presented at the International Symposium on Oileld Chemistry, Houston, TX, 24 February 2005. [3] J. Baudry, E. Bertrand, N. Lequeux, J. Bibette, J. Phys. Condens. Matter 16 (2004) R469R480. [4] A. Omari, G. Chauveteau, R. Tabary, Colloids Surf. A 225 (2003) 3748. [5] D.N. Bikiaris, G.P. Karayannidis, Polym. Int. 52 (2003) 12301239. [6] L.-W. Chen, B.-Z. Yang, M.-L. Wu, Prog. Org. Coat. 31 (1997) 393399. [7] Q. Ye, Z. Zhang, H. Jia, W. He, X. Ge, J. Colloid Interface Sci. 253 (2002) 279284.

[8] M.S. Wolfe, C. Scopazzi, J. Colloid Interface Sci. 133 (1) (1989) 265277. [9] B.H. Tan, K.C. Tam, Y.C. Lam, C.B. Tan, Langmuir 20 (2004) 11380 11386. [10] H. Senff, W. Richtering, J. Chem. Phys. 111 (4) (1999) 17051711. [11] B.E. Rodriguez, M.S. Wolfe, M. Fryd, Macromolecules 27 (1994) 6642 6647. [12] B.H. Tan, K.C. Tam, Y.C. Lam, C.B. Tan, J. Rheol. 48 (4) (2004) 915926. [13] B.H. Tan, K.C. Tam, Y.C. Lam, C.B. Tan, Langmuir 21 (10) (2005) 4283 4290. [14] P.G. de Gennes, Scaling Concepts in Polymer Physics, Cornell Univ. Press, 1979. [15] S.E. Paulin, B.J. Ackerson, M.S. Wolfe, J. Colloid Interface Sci. 178 (1996) 251262. [16] G.L. Flickinger, I.S. Dairanieh, C.F. Zukoski, J. Non-Newtonian Fluid Mech. 87 (1999) 283305. [17] G.K. Batchelor, An Introduction to Fluid Dynamics, Cambridge Univ. Press, Cambridge, 2000. [18] J. Roovers, Macromolecules 27 (1994) 53595364. [19] J.D. Ferry, Viscoelastic Properties of Polymers, second ed., Wiley, New York, 1970. [20] J.J. Magda, C.S. Lee, S.J. Muller, R.G. Larson, Macromolecules 26 (1993) 16961706. [21] V.R. Mhetar, L.A. Archer, J. Rheol. 40 (4) (1996) 549571. [22] L. Pellens, R.G. Corrales, J. Mewis, J. Rheol. 48 (2) (2004) 379393. [23] Q.T. Pham, W.B. Russel, J.C. Thibeault, W. Lau, J. Rheol. 43 (6) (1999) 15991615. [24] M.E. De Rosa, H.H. Winter, Rheol. Acta 33 (1994) 220237. [25] S. Adams, W.J. Frith, J.R. Stokes, J. Rheol. 48 (6) (2004) 11951213.

Das könnte Ihnen auch gefallen