Dynamic electrophoretic mobility of concentrated suspensions Comparison between experimental data and theoretical predictions

Similar documents
Sedimentation velocity and potential in a concentrated colloidal suspension Effect of a dynamic Stern layer

Dielectric Dispersion of Colloidal Suspensions in the Presence of Stern Layer Conductance: Particle Size Effects

An experimental study on the influence of a dynamic Stern-layer on the primary electroviscous effect

Stern-layer parameters of alumina suspensions

Analysis of the Dielectric Permittivity of Suspensions by Means of the Logarithmic Derivative of Its Real Part

The CMP Slurry Monitor - Background

Electrophoretic Light Scattering Overview

CHARACTERIZING CEMENT DISPERSIONS USING ACOUSTICS

ISO Colloidal systems Methods for zetapotential. Part 1: Electroacoustic and electrokinetic phenomena

Primary Electroviscous Effect with a Dynamic Stern Layer: Low κa Results

Influence of cell-model boundary conditions on the conductivity and electrophoretic mobility of concentrated suspensions

Stability of Dispersions of Colloidal Hematite/Yttrium Oxide Core-Shell Particles

MEGASONIC CLEANING OF WAFERS IN ELECTROLYTE SOLUTIONS: POSSIBLE ROLE OF ELECTRO-ACOUSTIC AND CAVITATION EFFECTS. The University of Arizona, Tucson

Correlation between rheological parameters and some colloidal properties of anatase dispersions

Electroacoustic and Dielectric Dispersion of Concentrated Colloidal Suspensions

Review: ISO Colloidal systems Methods for zeta potential determination

Introduction to the calculators in the Zetasizer software

The effect of surface dipoles and of the field generated by a polarization gradient on the repulsive force

The dielectric properties of charged nanoparticle colloids at radio and microwave frequencies

Initial position, x p (0)/L

A comparison of theoretical and experimental aggregation stability of colloidal silica

South pacific Journal of Technology and Science

ENV/JM/MONO(2015)17/PART1/ANN2

Particle Characterization Laboratories, Inc.

Electrostatic Forces & The Electrical Double Layer

An electrokinetic LB based model for ion transport and macromolecular electrophoresis

Electrostatic Double Layer Force: Part III

Conductivity dependence of the polarization impedance spectra of platinum black electrodes in contact with aqueous NaCl electrolyte solutions

Electrophoresis and electroosmosis as determined on the level of a single isolated colloid by use of optical tweezers

Modeling Viscosity of Multicomponent Electrolyte Solutions 1

Contents. Preface XIII

Specific ion effects on the interaction of. hydrophobic and hydrophilic self assembled

Zeta potential - An introduction in 30 minutes

Electrophoretic Deposition. - process in which particles, suspended in a liquid medium, migrate in an electric field and deposit on an electrode

Effect of charged boundary on electrophoresis: Sphere in spherical cavity at arbitrary potential and double-layer thickness

The Utility of Zeta Potential Measurements in the Characterization of CMP Slurries David Fairhurst, Ph.D. Colloid Consultants, Ltd

Methods for charge and size characterization colloidal systems


Viscosity of magmas containing highly deformable bubbles

An Overview of the Concept, Measurement, Use and Application of Zeta Potential. David Fairhurst, Ph.D. Colloid Consultants, Ltd

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

Part II: Self Potential Method and Induced Polarization (IP)

Optical bistability in metal/dielectric composite with interfacial layer


ELECTRIC CHARGING AND COLLOID STABILITY OF FABRICATED NEEDLE-LIKE TiO 2 NANOPARTICLES

Suspension Stability; Why Particle Size, Zeta Potential and Rheology are Important

CHAPTER TWO: EXPERIMENTAL AND INSTRUMENTATION TECHNIQUES

Colloidal Materials: Part III

PARTICLE SIZE ANALYSIS OF GOLD NANOPARTICLES

Where do ions solvate?

A NEW WAY TO CHARACTERIZE STABILITY AND PERFORMANCE OF COSMETIC EMULSIONS AND SUSPENSIONS

Colloidal Suspension Rheology Chapter 1 Study Questions

Electro-osmotic Flow Through a Rotating Microchannel

MAGNETIC PROPERTIES OF FERROFLUID EMULSIONS: THE EFFECT OF DROPLET ELONGATION

Dielectric Behavior of Suspensions of Polystyrene#Zinc Oxide Composite Microspheres

Porous Media Induced Aggregation of Protein- Stabilized Gold Nanoparticles

Foundations of. Colloid Science SECOND EDITION. Robert J. Hunter. School of Chemistry University of Sydney OXPORD UNIVERSITY PRESS

CEINT/NIST PROTOCOL REPORTING GUIDELINES FOR THE PREPARATION OF AQUEOUS NANOPARTICLE DISPERSIONS FROM DRY MATERIALS. Ver. 2.0

Revisit of wall-induced lateral migration in particle electrophoresis through a straight rectangular microchannel: Effects of particle zeta potential

Electrophoretic mobility of silica particles in a mixture of toluene and ethanol at different particle concentrations

Particle charge and Rheology

Effective charges of colloidal particles obtained from collective diffusion experiments

Ajaya Bhattarai * and Bijan Das. Department of Chemistry, North Bengal University, Darjeeling, , India.

Induced-charge Electrophoresis of Metallo-dielectric Particles

Zeta Potential Analysis using Z-NTA

Interaction of Gold Nanoparticle with Proteins

Studies on dielectric properties of a conducting polymer nanocomposite system

Highly stable and AC electric field-activated electrorheological fluid. based on mesoporous silica-coated graphene nanosheets

Viscoelastic Effects on Dispersion due to Electroosmotic Flow with Variable Zeta Potential

n v molecules will pass per unit time through the area from left to

Cracking in Drying Colloidal Films of Flocculated Dispersions

Nanoparticle Analyzer

arxiv:physics/ v2 [physics.chem-ph] 8 Dec 2004

Viscosity and aggregation structure of nanocolloidal dispersions

Rate of change of velocity. a=dv/dt. Acceleration is a vector quantity.

Charged Interfaces & electrokinetic

Preparation and Sedimentation Behavior in Magnetic Fields of Magnetite-Covered Clay Particles

The e!ects of the concentration of a polymer dispersant on apparent viscosity and sedimentation behavior of dense slurries

Stability of colloidal systems


Electroosmotic Flow Mixing in a Microchannel

Sample measurements to demonstrate Zetasizer specifications (Zetasizer Nano, Zetasizer APS, Zetasizer μv)

Colloid Chemistry. La chimica moderna e la sua comunicazione Silvia Gross.

Effect of Particle Size on Thermal Conductivity and Viscosity of Magnetite Nanofluids

Modified Maxwell-Garnett equation for the effective transport coefficients in inhomogeneous media

Demystifying Transmission Lines: What are They? Why are They Useful?

Multivariate Statistical Process Control for On-line Monitoring Size of Milling Process

SOLUTIONS TO CHAPTER 5: COLLOIDS AND FINE PARTICLES

Introduction to Zeta Potential Measurement with the SZ-100

Electro-hydrodynamic propulsion of counter-rotating Pickering drops. P. Dommersnes*, A. Mikkelsen, J.O. Fossum

Ultrasound propagation in concentrated suspensions: shear-mediated contributions to multiple scattering

Influence of ph and type of acid anion on the linear viscoelastic properties of Egg Yolk

Understanding the colloidal stability of protein therapeutics using dynamic light scattering

Observable Electric Potential and Electrostatic Potential in Electrochemical Systems

Measuring particle aggregation rates by light scattering

Protein Synthetic Lipid Interactions

PLUS. Zeta/Nano Particle Analyzer

Surfactant-free exfoliation of graphite in aqueous solutions

Ali Eftekhari-Bafrooei and Eric Borguet. Department of Chemistry, Temple University, Philadelphia PA Supporting information

Multiple Light Scattering by Spherical Particle Systems and Its Dependence on Concentration: A T-Matrix Study

Transcription:

Colloids and Surfaces A: Physicochem. Eng. Aspects 267 (2005) 95 102 Dynamic electrophoretic mobility of concentrated suspensions Comparison between experimental data and theoretical predictions A.V. Delgado a,, S. Ahualli a, F.J. Arroyo b, F. Carrique c a Department of Applied Physics, University of Granada, 18071 Granada, Spain b Department of Physics, University of Jaén, 23071 Jaén, Spain c Department of Applied Physics I, University of Málaga, 29071 Málaga, Spain Available online 1 August 2005 Abstract In this paper, we first report dynamic mobility (u d ) data obtained with an electroacoustic technique (electrokinetic sonic amplitude (ESA)) on suspensions of alumina with varying volume fractions and ionic strengths. The frequency interval was 1 18 MHz for these measurements. We found the expected reduction of the modulus of u d with volume fraction associated to the hindering of the particle motion by the interaction with its neighbors. More illustrative is the behavior of the phase angle, normally negative (the mobility lags behind the applied field) but sometimes positive in the frequency region of the Maxwell Wagner relaxation. This behavior is related to the phase angle of the induced dipole moment. From phase data, we also concluded that the effect of the dipole mentioned is weakened by partial neutralization of the dipoles induced on neighbor particles. Our results indicate that dynamic mobility data are little affected by the possible presence of stagnant-layer conductivity, which adds to the advantage of being suitable for concentrated systems. Furthermore, a good sensitivity to changes in particle size by, for instance, aggregation was also detected. The main features of the experimental mobility spectra were compared with the predictions of two models, one based on direct estimation of particle particle interactions and the other based on the cell concept. We found that the predictions of the cell model agree best with data obtained at the lowest frequencies (close to the Maxwell Wagner relaxation), whereas for higher frequencies it appears as if the interaction model would better capture the effects of inertia. 2005 Elsevier B.V. All rights reserved. Keywords: Dynamic mobility; Electrophoresis; Concentrated suspensions; Alumina 1. Introduction Electroacoustic techniques are recently gaining interest from the colloid science community as powerful tools to obtain information on the electrical state of many interfaces, particularly in a number of situations where traditional electrokinetic methods are of limited use. One of such situations is that of concentrated suspensions and slurries, in which we can get no information on the behavior of individual particles by performing, for instance, microelectrophoresis experiments. Take off of electroacoustic methods has also been favored by the commercial availability of sophisticated, highly automated devices which, in turn, Corresponding author. Tel.: +34 958 243209; fax: +34 958 243214. E-mail address: adelgado@ugr.es (A.V. Delgado). have driven theoretical treatments dealing with very different problems. There are two such techniques. One involves the generation of a pressure wave when an ac electric field is applied to the suspension: the amplitude of the sound wave, A ESA, is known as electrokinetic sonic amplitude, and so we speak of the ESA effect. The second method, reciprocal of ESA, is based on the determination of the electric potential induced by the passage of a sound wave through the system. It is called colloid vibration potential (CVP) or colloid vibration current (CVI) depending on the quantity measured. After the very early works in the subject, particularly those by Debye, and Booth and Enderby [1 4], O Brien [5,6] carried out an investigation on the physical foundations of electroacoustic techniques, based on the concept of dynamic electrophoretic mobility, u d, a complex quantity that is in 0927-7757/$ see front matter 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.colsurfa.2005.06.047

96 A.V. Delgado et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 267 (2005) 95 102 fact proportional to the ESA signal, from which it can be derived [6]: A ESA φ ρ p ρ m u d ρ, (1) m where φ is the volume fraction of solids and ρ p (ρ m )isthe density of the particles (the dispersion medium), and the dynamic mobility relates the electrophoretic velocity of the particle to the electric field, if this is alternating. In the case of the CVI method, its interpretation in terms of dynamic mobility is also possible, and a formally identical expression can be given for colloid vibration current, I CV, [7]: I CV φ ρ p ρ m u d ρ. (2) m Current technology allows to obtain the dynamic mobility from either A ESA or I CV data, and the problem can thus be reduced to evaluate u d and relate it to the zeta potential, the particle shape and dimensions, the ionic composition of the medium, and, very important, the volume fraction of solids. Two independent approaches have been followed to solve specifically the problem of volume fraction dependence in the case of concentrated dispersions. O Brien et al. use a first-principles approach, considering the ac electrophoresis of pairs of interacting particles [8] or the more realistic situation of a particle interacting with its nearest neighbors [9 12]. Another evaluation, first carried out by Dukhin et al. [13,14], is based on a so-called cell model, that has also been successfully applied to the analysis of a variety of electrokinetic phenomena in concentrated suspensions, including electrophoresis, sedimentation potential or dielectric dispersion [15 19]. In this sort of calculation, the hydrodynamic and electrical interactions between particles are accounted for in an averaged form by solving the electrokinetic equations for a single spherical particle of radius a, enclosed by a spherical shell of solution of radius b, concentric with the particle. The radius of the sphere is chosen in such a way that the volume fraction of solids is the same in the cell and in the suspension: ( a ) 3 φ = (3) b suspension in an unknown dispersion medium. SEM micrographs showed that the particles were crystalline and moderately polydisperse. Their average hydrodynamic radius was 165 ± 30 nm, as measured by photon correlation spectroscopy in a Malvern 4700c PCS device (Malvern Instruments, UK). The original suspensions were repeatedly washed with KCl solutions of concentration ranging between 5 10 5 and 10 3 mol/l. The process was carried out by centrifugation at 8500 g in a Kontron T-124 centrifuge, and redispersion in the desired KCl solution. The volume fraction of solids, φ, in the suspensions prepared was determined from the dry weight of known volumes using ρ p = 2.42 g/cm 3 as the density of alumina. 2.2. Methods The classical or dc electrophoretic mobility of the particles, u e, was measured in dilute suspensions (φ 10 4 )in the desired KCl concentration at ph 6, using a Malvern Zetasizer 2000 (Malvern Instruments, UK). These measurements provided an alternative or complementary way to evaluate the electrokinetic or zeta potential of the particles. The basic quantity of interest in this work was the dynamic or ac mobility of the particles, u e, a complex quantity measured for a frequency range between 1 and 18 MHz in an Acoustosizer II (Colloidal Dynamics, USA). Like other electroacoustic techniques, the method is particularly suitable for the evaluation of the electrokinetic properties of concentrated systems [1,7]. 3. Results and discussion 3.1. dc mobility and zeta potential Fig. 1 shows the electrophoretic mobility of alumina particles as a function of KCl concentration. Previous In this paper, our aim is to report experimental dynamic mobility data on suspensions of alumina of different volume fractions and ionic strengths, analyzing the features presented by the modulus and phase of u d and checking these results against predictions of the two types of models described. 2. Experimental 2.1. Materials The hydrated alumina used for the preparation of the suspensions was provided by Dispersion Technology Inc. (USA), and was given to us in the form of a concentrated Fig. 1. Dc electrophoretic mobility ( ) and zeta potential ( ) of alumina particles as a function of KCl molar concentration.

A.V. Delgado et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 267 (2005) 95 102 97 Fig. 2. Modulus (left) and phase angle (right) of the dynamic mobility u e of alumina particles as a function of frequency and volume fraction, in 5 10 5 mol/l KCl solutions. determination of the isoelectric point (ph iep or ph of zero zeta potential) yielded a value of ph 4, hence the negative mobility at the ph of the measurements. A significant feature of Fig. 1 is the increase in u e with ionic strength, using a complete theory of electrophoresis, like O Brien and White s [20], the zeta potential, ζ, was also found to increase in absolute value with the concentration of KCl. Such an increase is against the predictions of classical models of electrokinetic phenomena, the compression of the electrical double layer upon adding indifferent ions should lead to a reduction in ζ. The anomalous behavior of ζ has been reported by many other authors [21,22], and there seems to be a general agreement that it is originated by the existence of a finite stagnant-layer conductivity (SLC), and that a theory assuming that SLC 0 leads to correct ζ-concentration dependences [23,24]. We will not pursue this topic at this stage, but will come back to it the following sections. 3.2. Dynamic mobility of alumina suspensions: experimental data We measured the dynamic mobility of alumina particles in suspensions with varying volume fractions (φ 2.5 20%) and KCl concentrations. The ESA instrument provides experimental data of the modulus and phase of the complex dynamic mobility u e. Because this quantity has been found to depend strongly on both the ionic strength and the volume fraction, we must show all data in order to discuss such effects. This is done in Figs. 2 5. The following features of these figures must be mentioned: (i) In whatever conditions, increasing the volume fraction provokes a reduction in the mobility. This was to be expected, as it is predicted by all theoretical models [7,11,12] and justified because of the hydrodynamic and Fig. 3. Modulus (left) and phase angle (right) of the dynamic mobility u e of alumina particles as a function of frequency and volume fraction, in 2 10 4 mol/l KCl solutions.

98 A.V. Delgado et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 267 (2005) 95 102 Fig. 4. Modulus (left) and phase angle (right) of the dynamic mobility u e of alumina particles as a function of frequency and volume fraction, in 5 10 4 mol/l KCl solutions. electrostatic interactions (at non-zero frequencies of the applied field) of a given particle with its neighbors. (ii) In addition, the phase angle is in general negative, i.e. the velocity lags behind the applied ac fields. This must be due to two phenomena that, for certain frequency ranges, act against the electrophoretic motion. One is the particle and fluid inertia that becomes increasingly important for frequencies above f I η/ρ m a 2 ( 5 MHz for our particles), where η is the viscosity of the liquid medium, this is the manifestation of the time needed for the particles to acquire their steady velocity after a variation of the applied force. Because of this inertia, the phase lag of the mobility will become increasingly negative with frequency for f > f I. Furthermore, the action of the field provokes the polarization of the electrical double layer inducing a frequency, and zeta potential, dependent dipole moment. In order to understand in a qualitative way the influence of that dipole on the dynamic mobility, we have used the Delacey and White s theory [25], valid for dilute suspensions, to calculate the induced dipole coefficient C0, related to the dipole moment d* as follows: d = 4πε m a 3 C0 E, where ε m is the electric permitivity of the dispersion medium. Recall that in this theory no contribution of stagnant-layer conductivity is assumed, but this is not essential for our qualitative arguments. It is known [26] that, for large κa, the dc electrophoretic mobility is very approximately determined by the induced dipole coefficient by means of the relation: u e = ε m η ζ(1 C 0), (4) Fig. 5. Modulus (left) and phase angle (right) of the dynamic mobility u e of alumina particles as a function of frequency and volume fraction, in 1 10 3 mol/l KCl solutions.

A.V. Delgado et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 267 (2005) 95 102 99 Fig. 6. Modulus (a) and phase angle (b) of(1 C 0 ) for 150 nm particles with the zeta potentials indicated, in 2 10 4 mol/l KCl solutions. The suspensions are assumed to have low volume fractions of solids. which O Brien [5] has shown to be also valid for alternating fields. We can wonder whether a similar dependence between u e and C 0 exists for ac fields. Such dependence would fail for frequencies above f I, since inertia effects play obviously no role in dc fields, so we will focus ourselves on the frequency range of interest from the electrokinetic point of view. From the ac version of Eq. (4) we can conclude that the phase angle of u e is given by Arg(1 C 0 ), and u e will depend on 1 C 0. Fig. 6 shows 1 C 0 and Arg(1 C0 ) as a function of frequency for low and moderate zeta potential. Both the low-frequency (or α, with frequency f α ) and the Maxwell Wagner (MW, with characteristic frequency f MW ) relaxations are observed. Note how, for sufficiently high ζ, u e should increase around f MW, as experimentally found, because of the raise in 1 C0 (Fig. 6a). For similar reasons, the phase angle of u e can increase, and even become positive, as that of (1 C0 ) displays a maximum around the Maxwell Wagner relaxation frequency (Fig. 6b). Summarizing in the f MW region the dynamic mobility may show a phase lead associated to the behavior of C0, and a phase delay because of inertia. Since f MW K m /ε m (K m is the conductivity of the medium), increasing the ionic strength will shift f MW to higher frequencies, thus hiding the raise in the phase by its inertial decrease. This occurs in Fig. 5 (10 3 mol/l KCl), whereas in Fig. 4 and, mainly, Fig. 3, the phase increase, and even its positive values are observed. Note also how the phase difference between the applied field and the particle motion decreases when the content of solids in the suspension increases; in the frequency region controlled by inertia, this phenomenon may be due to the fact that the volume of fluid associated with each particle decreases and so does the inertia of the particle-fluid ensemble. Below f MW, the effect of the induced dipole must be weakened by the fields from the dipoles of the surrounding particles [11]. (iii) Concerning the modulus of u e, note how at the lowest electrolyte concentration, it is found to increase with frequency due to the Maxwell Wagner relaxation around 0.5 MHz (Fig. 2). When the electrolyte concentration is increased, the mobility maximum shifts to higher frequencies, as explained above, in such conditions the fast inertial decrease dominates and almost fully buries the MW increase. 3.3. Comparison with models Two kinds of theoretical models discussed in Section 1 were used to compare our experimental results to the predictions of the theory. These are the cell model [7,12,14] and the particle interaction model [11]. In all cases, the particle radius was considered fixed (165 nm), and only the zeta potential was used as an adjustable parameter minimizing the differences between calculated and measured mobilities. Fig. 7 shows the comparison between experimental u e and calculated data of the modulus of the mobility in 5 10 5 mol/l KCl solutions. Note that in general both models are capable of properly fitting the data, although cell model calculations sometimes fail to reach the high frequency side because of convergence limitations. In contrast, this model can properly reproduce the maxima in u e at frequencies close to the experimental ones. Similarly, a good agreement is found in Fig. 8 for a concentration of 2 10 4 mol/l KCl. Nevertheless, at higher ionic strengths (Figs. 9 and 10), both models find characteristic inertia frequencies, f I, higher than those experimentally detected. These differences can be explained on the simple basis that, for the particle concentration involved, the aggregation between particles is more likely for these ionic strengths. The dynamic mobility may be sufficiently sensitive as to sense such processes; in fact, if we take an equiva-

100 A.V. Delgado et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 267 (2005) 95 102 Fig. 7. Comparison between the modulus of the dynamic mobility (symbols) lines) and interaction (dashed lines) model calculations, for the zeta potentials indicated in Table 1. The dc mobility measured in dilute suspensions is shown as a short dotted line labelled u e. Electrolyte concentration 5 10 5 mol/l. lent particle size approximately equal to twice the measured one (a = 300 nm), the agreement is much better, as Fig. 11 demonstrates for the case of the cell model. The fact that collective electrokinetic phenomena are very sensitive to particle size variations or particle aggregation has also been demonstrated for dielectric dispersion phenomena [27]. There are other aspects where the models do not follow the data. Thus, the calculated maxima observed in u e f plots shifts to higher frequencies when the volume fraction is increased, a fact not found experimentally, and that will require a more detailed investigation. Similarly, the calculated data disagree from measured mobilities at the highest volume fractions investigated (φ 20%), when the mobility Fig. 9. Comparison between the modulus of the dynamic mobility (symbols) lines) and interaction (dashed lines) model calculations, for the zeta potentials indicated in Table 1. The dc mobility measured in dilute suspensions is shown as a short dotted line labelled u e. Electrolyte concentration of 5 10 4 mol/l. curve is shifted to larger frequencies. This may be due to the fact that f I is, as mentioned, inversely proportional to the square of a characteristic length, which can be considered equal to the radius for low solids concentration, but becomes volume fraction dependent for high φ. According to the cell model, for a 20% particle concentration the distance from the particle surface to the cell boundary is 120 nm, i.e. smaller than the radius itself, suggesting a critical frequency for inertia effects that is higher than for more dilute suspensions. This does not mean that the model does not catch the main features of the hydrodynamic interactions; it is crucial, how- Fig. 8. Comparison between the modulus of the dynamic mobility (symbols) lines) and interaction (dashed lines) model calculations, for the zeta potentials indicated in Table 1. The dc mobility measured in dilute suspensions is shown as a short dotted line labelled u e. KCl concentration = 2 10 4 mol/l. Fig. 10. Comparison between the modulus of the dynamic mobility (symbols) lines) and interaction (dashed lines) model calculations, for the zeta potentials indicated in Table 1. The dc mobility measured in dilute suspensions is shown as a short dotted line labelled u e. Electrolyte concentration of 1 10 3 mol/l.

A.V. Delgado et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 267 (2005) 95 102 101 Fig. 11. Comparison between the modulus of the dynamic mobility (symbols) and the theoretical prediction obtained from cell model calculations (solid lines), for the particle size equal to 300 nm and the zeta potential equal to 70 mv. Electrolyte concentration 5 10 4 mol/l. ever, to properly choose the parameters of the system. Note, for instance, how having set the particle size equal to 300 nm in Fig. 11 leads to a very reasonable agreement between the slopes of the decrease in u e with f in the high frequency (inertia) regime. It is interesting to mention that the phase angle is also very useful in evaluating the validity of the models. Figs. 12 14 show the comparison between theory and experiment (the 5% volume fraction has not been included in these figures to avoid confusion). Although no experimental data are available below 1 MHz, we have performed the calculations down to 100 khz. Contrary to the interaction model, the cell one can properly describe the effect of volume fraction on the phase angle, in particular the phase lead in the MW fre- Fig. 13. Comparison between the phase angle of the dynamic mobility (symbols) lines) and interaction (dashed lines) model calculations. Electrolyte concentration of 2 10 4 mol/l. quency. This is more significant at low volume fractions, and in some cases it gives rise to crossings (also found in the series of experimental data) of the phase curves at frequencies depending on the ionic strength (which determines the Maxwell Wagner relaxation frequency). Let us finally refer to the zeta potentials needed to fit the data. They are summarized in Table 1, and the following features are worth to be mentioned: (i) ζ decreases with ionic strength, contrary to the trends found in dc electrophoresis (Fig. 1). Although this deserves further investigation, it suggests that at the relatively high frequencies used in ESA experiments stagnant-layer conductivity, if present, plays a rather Fig. 12. Comparison between the phase angle of the dynamic mobility (symbols) lines) and interaction (dashed lines) model calculations. Electrolyte concentration 5 10 5 mol/l. Fig. 14. Comparison between the phase angle of the dynamic mobility (symbols) lines) and interaction (dashed lines) model calculations. Electrolyte concentration of 5 10 4 mol/l.

102 A.V. Delgado et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 267 (2005) 95 102 Table 1 Zeta potential (mv) best fitting the dynamic mobility spectra of alumina suspensions in the indicated KCl solutions, for different volume fractions of solids, φ KCl (mol/l) φ = 2.5% φ =5% φ = 10% φ = 20% I II I II I II I II 5 10 5 100 80 100 80 100 80 120 90 2 10 4 90 80 90 80 95 80 100 90 5 10 4 70 60 70 60 70 60 80 70 Column I corresponds to cell model calculations and column II refers to particle interaction calculations. little significant role in determining the electrokinetic behavior of the system. (ii) A slightly higher ζ value seems to be needed, whatever the model, to fit the results at highest volume fractions. This probably reflects the limitations of the models to properly account for the collective behavior of the suspensions. (iii) Although the trends of ζ are similar for the two models, it appears that a higher zeta potential results from the cell calculations. However, the differences are only significant at the lowest ionic strengths. We have no clear explanation for this, and more cases should be studied prior to deciding which approximation to the mobility is better, if any. (iv) We have calculated the best-fit zeta potential for the most dilute suspensions (φ = 2.5%) using a singleparticle dynamic electrophoresis model [5]. The results are 75, 80 and 60 mv for KCl concentrations 5 10 5,2 10 4 and 5 10 4 mol/l, respectively. Note that they are close to the data in Table 1, and agree with the results obtained using concentrated-suspension approaches for the evaluation of the dynamic mobility. 4. Conclusions Dynamic mobility data obtained from the ESA signal are coherent with conventional dc electrophoresis measurements performed on dilute suspensions. It appears as if the high frequency data would have the additional advantage of being little affected by the existence of finite stagnant-layer conductivity. Our results also indicate that electroacoustic measurements are extremely sensitive to variations in particle size associated to, for instance, particle growth or aggregation in case of high ionic strength or concentrated suspensions. These advantages of dynamic mobility will be even more appreciated as the theoretical models describing the phenomenon in the case of concentrated systems continue to be improved. In this work, we have compared the predictions of two models, describing the effects of neighbor particles on a given one either through the evaluation of particle particle interactions or by the use of cell models. We have found that the predictions of the cell model are closer to the experimental frequency dependences at the lowest frequencies (in the vicinity of Maxwell Wagner relaxation), whereas for high frequency (in the region of inertia effects), the situation is reversed, especially for the most concentrated suspensions. Acknowledgements Financial support from CICYT, Spain, under Project MAT2004-00866, and from FEDER funds (EU) is gratefully acknowledged. References [1] R.J. Hunter, Colloids Surf. A 141 (1988) 37. [2] R.J. Hunter, R.W. O Brien, in: A.T. Hubbard (Ed.), Encyclopedia of Surface and Colloid Science, Marcel Dekker Inc., USA, 2002, p. 1722. [3] J.A. Enderby, Proc. R. Soc. Lond., Ser. A 207 (1951) 329. [4] F. Booth, J.A. Enderby, Proc. Phys. Soc. 65 (1952) 321. [5] R.W. O Brien, J. Fluid Mech. 190 (1988) 71. [6] R.W. O Brien, J. Fluid Mech. 212 (1990) 81. [7] A.S. Dukhin, V.N. Shilov, H. Ohshima, P.J. Goetz, in: A.V. Delgado (Ed.), Interfacial Electrokinetics and Electrophoresis, Surfactant Science Series, Marcel Dekker Inc., New York, 2002 (Chapter 17). [8] P. Rider, R.W. O Brien, J. Fluid Mech. 257 (1993) 607. [9] R.W. O Brien, T.A. Wade, M.L. Carasso, R.J. Hunter, W.N. Rowlands, J.K. Beattie, Am. Chem. Soc. Symp. Ser. 693 (1998) 311. [10] R.J. Hunter, Colloids Surf. A 195 (2001) 205. [11] R.W. O Brien, A. Jones, W.N. Rowlands, Colloids Surf. A 218 (2003) 89. [12] F.J. Arroyo, F. Carrique, S. Ahualli, A.V. Delgado, Phys. Chem. Chem. Phys. 6 (2004) 1446. [13] A.S. Dukhin, H. Ohshima, V.N. Shilov, P.J. Goetz, Langmuir 15 (1999) 3445. [14] A.S. Dukhin, V.N. Shilov, H. Oshima, P.J. Goetz, Langmuir 15 (1999) 6692. [15] S. Kuwabara, J. Phys. Soc. Jpn. 14 (1959) 527. [16] H. Ohshima, J. Colloid Interface Sci. 208 (1998) 295. [17] F. Carrique, F.J. Arroyo, A.V. Delgado, Colloids Surf. A 195 (2001) 157. [18] F. Carrique, F.J. Arroyo, A.V. Delgado, J. Colloid Interface Sci. 243 (2001) 351. [19] F. Carrique, F.J. Arroyo, A.V. Delgado, J. Colloid Interface Sci. 252 (2002) 126. [20] R.W. O Brien, L.R. White, J. Chem. Soc., Faraday Trans. 2 74 (1978) 1607. [21] R.R. Midmore, R.J. Hunter, J. Colloid Interface Sci. 122 (1988) 521. [22] C.F. Zukoski, D.A. Saville, J. Colloid Interface Sci. 114 (1986) 32. [23] C.S. Mangeldorf, L.R. White, J. Chem. Soc., Faraday Trans. 86 (1990) 2859. [24] A.S. Dukhin, V.N. Shilov, in: A.V. Delgado (Ed.), Interfacial Electrokinetics and Electrophoresis, Surfactant Science Series, Marcel Dekker Inc., New York, 2002 (Chapter 2). [25] E.H.B. DeLacey, L.R. White, J. Chem. Soc., Faraday Trans. 2 (77) (1981) 2007. [26] V.N. Shilov, A.V. Delgado, F. González-Caballero, J. Horno, J.J. López-García, C. Grosse, J. Colloid Interface Sci. 232 (2000) 141. [27] F. Carrique, F.J. Arroyo, M.L. Jiménez, A.V. Delgado, J. Chem. Phys. 118 (2003) 1945.