Correlation between rheological parameters and some colloidal properties of anatase dispersions

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

Rights statement Post print of work supplied. Link to Publisher's website supplied in Alternative Location.

Stern-layer parameters of alumina suspensions

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

SURFACE FORCES ARISING FROM POLYMERS AND SMALL IONIC ADDITIVES YIELD STRESS AND ZETA POTENTIAL RELATIONSHIP CHARLES ONG BAN CHOON

Interfacial forces and friction on the nanometer scale: A tutorial

Contents. Preface XIII

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

Zeta potential - An introduction in 30 minutes

Porous Media Induced Aggregation of Protein- Stabilized Gold Nanoparticles

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

Rheology of strongly sedimenting magnetite suspensions

Rheological study on the aging process in a polymeric fumed silica suspension

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

SOLUTIONS TO CHAPTER 5: COLLOIDS AND FINE PARTICLES

CHAPTER TWO: EXPERIMENTAL AND INSTRUMENTATION TECHNIQUES

Stability of colloidal systems

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

Applied Surfactants: Principles and Applications

Aqueous Colloidal Processing and green sheet properties of. Lead Zirconate Titanate (PZT) ceramics made by Tape. Casting.

Module 8: "Stability of Colloids" Lecture 37: "" The Lecture Contains: DLVO Theory. Effect of Concentration. Objectives_template

*blood and bones contain colloids. *milk is a good example of a colloidal dispersion.

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

Contents. Preface XIII. 1 General Introduction 1 References 6

István Bányai, University of Debrecen Dept of Colloid and Environmental Chemistry

Electrophoretic Light Scattering Overview

Overview of DLVO Theory

The CMP Slurry Monitor - Background

Electrostatic Forces & The Electrical Double Layer

NANO-RDX ELECTROSTATIC STABILIZATION MECHANISM INVESTIGATION USING DERJAGUIN-LANDAU AND VERWEY-OVERBEEK (DLVO) THEORY

South pacific Journal of Technology and Science

Electrostatic Double Layer Force: Part III

Effect of sodium dodecyl sulfate on flow and electrokinetic properties of Na-activated bentonite dispersions

Particle Characterization Laboratories, Inc.

Methods for charge and size characterization colloidal systems

Attraction or repulsion between charged colloids? A connection with Debye Hückel theory

Zeta Potential Analysis using Z-NTA

Kinetics of gold nanoparticle aggregation: Experiments and modeling

Supporting Information for. Concentration dependent effects of bovine serum albumin on graphene

Colloid stability. Lyophobic sols. Stabilization of colloids.

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

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

Modeling of colloidal gels

Review: ISO Colloidal systems Methods for zeta potential determination

Factors governing the tendency of colloidal particles to flocculate

Rheology of cementitious suspensions containing weighting agents.

Pharmaceutics I صيدالنيات 1. Unit 6

Colloidal Suspension Rheology Chapter 1 Study Questions

Multimedia : Boundary Lubrication Podcast, Briscoe, et al. Nature , ( )

Contents XVII. Preface

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

Single action pressing (from top)

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

Introduction to Zeta Potential Measurement with the SZ-100

Effect of aggregation on viscosity of colloidal suspension

A comparison of theoretical and experimental aggregation stability of colloidal silica

Thermodynamic properties of the S/L interfacial layer: Stabilization of the colloidal system in binary liquids

Contents. Preface XI Symbols and Abbreviations XIII. 1 Introduction 1

Effect of physicochemical conditions on crossflow microfiltration of mineral dispersions using ceramic

CH5716 Processing of Materials

Viscosity and aggregation structure of nanocolloidal dispersions

Supporting Information

Aggregate Structures and Solid-Liquid Separation Processes

Determination and Assessment of the Rheological Properties of Pastes for Screen Printing Ceramics

Heteroaggregation, repeptization and stability in mixtures of oppositely charged colloids

Anomalous ph Dependent Stability Behavior of Surfactant-Free Nonpolar Oil Drops in Aqueous Electrolyte Solutions

Surface interactions part 1: Van der Waals Forces

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

Aadil Zayid. Submitted in partial fulfilment of the requirements for the degree of Master of Applied Science

Rheological performance of cementitious materials used in well cementing

Module 8: "Stability of Colloids" Lecture 38: "" The Lecture Contains: Calculation for CCC (n c )

Aging in laponite water suspensions. P. K. Bhattacharyya Institute for Soldier Nanotechnologies Massachusetts Institute of Technology

Forces across thin liquid film

Probing the Hydrophobic Interaction between Air Bubbles and Partially. Hydrophobic Surfaces Using Atomic Force Microscopy

Engineering Nanomedical Systems. Zeta Potential

Ceramic Processing Research

Module 8: "Stability of Colloids" Lecture 40: "" The Lecture Contains: Slow Coagulation. Other Factors affecting Kinetics of Coagulation

Relative Viscosity of Non-Newtonian Concentrated Emulsions of Noncolloidal Droplets

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

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

Effect of solution physical chemistry on the rheological properties of activated sludge

Sedimentation acceleration of remanent iron oxide by magnetic flocculation

Rheology of Dispersions

Supplementary Information: Triggered self-assembly of magnetic nanoparticles

Deposition of Titania Nanoparticles on Spherical Silica

Interaction forces between oil water particle interfaces Non-DLVO forces

Stability of concentrated suspensions of Al 2 O 3 SiO 2 measured by multiple light scattering

Electrical double layer interactions between dissimilar oxide surfaces with charge regulation and Stern Grahame layers

Measuring particle aggregation rates by light scattering

Anindya Aparajita, Ashok K. Satapathy* 1.

! Importance of Particle Adhesion! History of Particle Adhesion! Method of measurement of Adhesion! Adhesion Induced Deformation

COLLOIDAL processing offers the potential to reliably produce

EMPIRICAL ESTIMATION OF DOUBLE-LAYER REPULSIVE FORCE BETWEEN TWO INCLINED CLAY PARTICLES OF FINITE LENGTH

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

Analysis on the birefringence property of lyotropic liquid crystals below Krafft temperature

Influence of Enterococcal Surface Protein (esp) on the Transport of Enterococcus faecium within Saturated Quartz Sands

THE RHEOLOGICAL PROPERTIES OF DISPERSIONS OF LAPONITE, A SYNTHETIC HECTORITE-LIKE CLAY, IN ELECTROLYTE SOLUTIONS

Surfactant-free exfoliation of graphite in aqueous solutions

Research Article A Modified Method to Calculate Critical Coagulation Concentration Based on DLVO Theory

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

Transcription:

ANNUAL TRANSACTIONS OF THE NORDIC RHEOLOGY SOCIETY, VOL. 15, 7 Correlation between rheological parameters and some colloidal properties of anatase dispersions A.I. Gómez-Merino, F. J. Rubio-Hernández, J.F. Velázquez-Navarro, F.J. Galindo-Rosales, P. Fortes-Quesada Rheology and Electrokinetics Group, University of Málaga, Spain ABSTRACT The linear relationship of the yield stress with the square zeta potential may be used to determine the Hamaker constant in solid dispersions. In this work we have obtained the Hamaker constant for the attractive force between anatase aqueous suspensions using this method and compared with those obtained by other techniques. INTRODUCTION The van der Waals (attractive) and electrostatic (repulsive) forces play a critical role in determining the stability of dispersions. They form the basis of the well-known DLVO theory for colloid stability 1,. Depending on the stability level, differences in properties such as viscosity, flow and sedimentation behaviour in colloidal dispersions can be observed. For example, the viscosity of a flocculated dispersion can be several orders of magnitude larger than that for a dispersed dispersion at the same solid concentration 3. Considering the role played in the stability of dispersions, it is suggested the necessity to characterize the van der Waals interaction between colloidal particles. The value of the van der Waals force for pair interaction is greatly determined by the Hamaker constant. This parameter is material dependent and cannot be directly measured. Its theoretical value can be calculated from Lifshitz s theory 4,5. This calculation requires that some dielectric or optical properties of the material be known at a wide frequency spectrum. As these full spectra data are difficult to obtain for most materials, a number of approximate models, derived from Lifshitz s theory, that use a limited frequency data were developed 6,7. The theoretical value of the Hamaker constant corresponding to some materials agrees with the experimental, however, in other cases, the disagreement was as high as a factor of seven. Surface force apparatus (SFA) 8,9 and atomic force microscope (AFM) 1 are sophisticated experimental techniques that can be used to determine the Hamaker constant from DLVO theory. However, the results are very sensitive to the presence of impurities. Alternatively, the yield stress-zeta potential technique here used works with concentrated dispersions and, consequently, the results obtained are less sensitive to the relative smaller amount of impurities introduced from external sources. When flocculated, the particles in the colloidal dispersion are linked together by a net attractive force. These form a 3-D network that occupies the whole volume of the dispersion. The magnitude of the net attractive force determines the strength of the structure. The yield stress is a direct measure of the strength of this structure 3,11,1 and gets its maximum value at the isoelectric point (IEP). The IEP is the ph value corresponding to zero repulsive interaction, therefore only the van der Waals force contributes to strength of the structure 3,11,1. If the ph is different from the IEP, the yield stress should decrease as the particles develop a larger repulsive potential.

Eventually, at a certain ph, the yield stress becomes zero. The surface potential (zeta potential) on the particles at this ph-value, characterizes the transition from a flocculated to a wholly dispersed state 13,14. This critical zeta potential characterises the electrostatic repulsive potential that exactly counters the van der Waals attractive potential. Therefore, the critical zeta potential can be used to calculate the Hamaker constant. In this paper we will measure the flow behaviour of anatase concentrated dispersions to determine the Hamaker constant of TiO particles in water. For this purpose, a combined yield stress-zeta potential technique has been used. THEORETICAL BACKGROUND As it is well established in Electrokinetics, the zeta potential, measured at a shear plane near the surface particle, can be used as a good approximation to the surface potential, mainly when no specific adsorption takes place on the particles. The zeta potential is greatly dependent of the ionic strength at the liquid phase. At high ionic strength, the potential decreases much more sharply over the distance from the surface to the shear plane. This means a smaller zeta potential. It is therefore important to maintain a relatively constant ionic strength while characterizing the zeta potential of dispersion as a function of ph. A linear relationship between yield stress and the square zeta potential has been validated by many aqueous dispersions 13,14,15,16. By assuming that the yield stress is proportional to the particle pair DLVO interaction potential (or force), Hunter and Hunter et al. obtained an equation that predicts such a linear relationship. The yield stress of a particle network structure in a flocculated dispersion is proportional to the number of particle-particle links that cross a unit area of the sample and the strength of the links between particles. The relationship between yield stress ( τ ) ( ς ) is given by 17, y and zeta potential φ A C τ y = (1) a D D ( ζ ) 1 where A is the Hamaker constant of the particle in water, C = πε ln 1 (1 e κ D ), D o is the minimum surface separation distance between interacting particles in the flocculated state, є is the permittivity of water, κ is the inverse of the double-layer thickness which surrounds the colloidal particle, a is the particle size and φ is the solid volume fraction of the dispersion. At the flocculateddispersed state transition, the yield stress is zero. Essentially, at this state, the electrostatic repulsive potential is equal to the van der Waals attractive potential. The critical zeta potential ζ crit characterizing this transition is therefore given by, A ζ crit = () 1DC As can be seen, for a constant ionic strength, the critical zeta potential is proportional to the square root of Hamaker constant. According with Eq., the critical zeta potential should not be dependent upon the particle size, size particle distribution, and solid concentration. Zhou et al. 18 confirmed this fact with a study on α-al O 3 dispersions. MATERIALS AND METHODS The titanium oxide (TiO ), supplied by Sigma Aldrich, has a purity of 99.9 %, and a density of 3.9 g/cm 3, a BET area of 1 m /g, and an average primary particle size (TEM) of 1 nm. The water was double distilled and deionised with a column of mixed bed ions (Millipore). All chemicals were of analytical grade. All measurements were performed at 5.±.1 ºC.

concentration and can be characterised by the Debye length, κ -1, which characterises the range of the electrostatic repulsion between particles. 4 Zeta Potential (mv) - Fig. 1. Particle morphology of the TiO s nanoparticles used in the study. The dispersions were prepared by sonication the powder in water that was previously acidified with HCl or made alkaline with NaOH. Concentrated HCl or NaOH solutions 1 M were used to change de dispersions ph so as to minimize dilution. The zeta potential was calculated from electrophoretic mobility measurements, which were made with a Zetasizer (Malvern Instruments). A fixed solid concentration of 4 1-4 g/ml was used in all electrophoresis measurements. At least six measurements were taken at the stationary level in a rectangular cell. The phmeter used was calibrated with buffer solutions of ph 4. and 7.. The rheological measurements were made using a strain-controlled Bohling-Vor rheometer with a cone-plate geometry (1º and 4 mm diameter). Dispersions containing 1 or 5 vol. % solids were used. The contact angle was measured using a CAM Optical Angle Meter. RESULTS AND DISCUSSION For the metal oxide dispersions, the magnitude of electrical double layer repulsive force is dependent on the surface charge density and the thickness of the double layer. The former is determined by the suspension ph. The latter is determined by the electrolyte -4 3 4 5 6 7 8 9 1 Fig.. The effect of ph on the zeta potential of anatase TiO dispersions, for 1-5 M KCl. The zeta potential for TiO suspensions was measured as a function of ph at various levels of ionic strength and the results are shown in Fig.. Since adjustment of ph with acid or base will lead to the change in the ionic concentration of each suspension, the ionic strength was controlled by adding an appropriate amount of salt (KCl) after each Stress (Pa) 7 6 5 4 3 1 1 3 4 5 6 7 8 9 1 11 ph Shear rate (1/s) ph=3 ph=6 ph=7 ph=3.6 ph=9 ph=6.6 ph=4.4 Figure 3. Flow curves for anatase TiO dispersion as a function of ph, φ =.5, for 1-5 M KCl. adjustment of ph to ensure the total ionic concentration coming from the salt and acid or base is nearly constant. The value for the IEP, ph=4.7, occurs at a lower ph than those given by other authors 14,18. This can be due to the trace quantities of rutile found in this sample.

At ph near IEP, the effect of conductivity on zeta potential was not pronounced probably because the change in zeta potential with ph is quite sharp in this region. At ph further away from IEP, the effect of ionic strength is quite discernable for TiO. The zeta potential is usually slightly higher for the lower ionic strength dispersion. But above IEP, the magnitude of the zeta potential is higher at lower ionic strength. The zeta potential attained a limiting value at ph well away from IEP. For TiO, the limiting value at low ph is about 4 mv and, at high ph, it was about -48 mv at ionic strength between 1-3 M and 1-5 M of indifferent electrolyte. Figures 3 and 4 show the flow curves for TiO suspensions of φ =.5 and φ =.1 at different ph-values but a constant ionic strength. In both cases, the dispersions exhibited a shear-thinning flow behaviour over the shear rate range examined, which revealed that the particle aggregates in the suspensions were broken down into smaller flow units by the applied forces, so that the resistance to flow was reduced, leading to the lower viscosity as the shear rate increased. The Stress (Pa) 4 3 1 5 5 75 1 15 15 175 Shear rate (1/s) ph=3.6 ph=7. ph=4.8 ph=6.4 ph=3. ph=9- Figure 4. Flow curves for anatase TiO dispersion as a function of ph, φ =.1, for 1-5 M KCl. suspension began to flow so long as the applied stress was higher than the yield stress (τ y ). Various empirical models 19 were used to estimate τ y : Bingham plastic model: τ = τ yb, + η & sγ (3) 1/ 1/ 1/ Casson model: τ = τ, + ( η & γ) (4) yc n Herschel-Bulkley model: τ = τ yh, + K & γ (5) where τ y,b, τ y,c and τ y,h are the yield-stress parameters determined from Bingham, Casson and Herschel-Bulkley models, respectively; η s is the plastic viscosity, and K and n are structure-dependent parameters that can be determined experimentally. As can be expected, the yield stress, τ y increases withφ, regardless of the model used. The Bingham plastic model supplies the highest yield stress comparing with the other models examined. This is presumably due to the linear τ-γ& dependence in Eq. 3. The yield stress for each ph was calculated from Eqs. 3, 4 and 5. The criterion followed to choose the yield stress value was established on the basis of the most accurate adjustment obtained according to the fit parameters. The yield stress vs. ph data for TiO dispersions is shown in figure 5. As can be seen, the maximum shear yield stress for each volume fraction concentration matches well with the isoelectric point. As the ph moves away from the IEP for both volume fractions, the shear yield stress decreases as the net attractive force decreases due to the increase in the electric double layer force as a result of the increase of surface charge density. When this repulsive force exceeds the van der Waals attraction at the ph far away from the IEP, dispersed suspensions are generated and the yield stress disappears. The dispersion at Φ=.5 is prepared in a dispersed state at a native ph=7.. This dispersion displays a yield stress of.14 Pa. The ph of the dispersion is lowered in a stepwise manner. With a further reduction in the ph the yield stress increases and reaches a maximum value of 1.5 Pa at ph=4, however, a reduction in ph causes a decrease in the yield stress which eventually disappears at s

ph=3. A similar behaviour was observed when the ph increases from the IEP. Yield stress (Pa) 1 1 8 6 4 3 4 5 6 7 8 9 1 ph 1 vol.% 5 vol.% Figure 5. The yield stress of anatase TiO dispersion as a function of ph. Yield stress (Pa) 14 1 1 8 6 4 1 vol.% 5 vol.% 4 6 8 1 1 Zeta Potential (mv) Figure 6. Yield stress vs. square of zeta potential for anatase TiO dispersion at two volume fractions. Thus, at ph=6, the value obtained for the yield stress is 4,8 Pa, and at ph=9, when the experiment is terminated, a small change in the yield stress is observed from the value found at the native ph. The same tendency described above is also observed in the case of Φ=.1. For this volume fraction the maximum yield stress attained is 8. Pa at ph=4.4. Further reduction of ph causes a decrease in the yield stress until the very small value. Pa at ph=3. Far away from the IEP, at ph=9 the yield stress almost disappears, as could be expected. Not surprisingly, the maximum yield stress τ y,max exhibited a concentration dependence. At the IEP, the yield stress is maximum because the electrostatic repulsive potential is absent. The linear relationship between yield stress and square of zeta potential ζ for TiO dispersions is shown in figure 6. The critical zeta potential is determined from the intercept on the ζ axis. The critical zeta potential was 33 mv for Φ=.5 and 31 mv for Φ=.1. The difference in the critical zeta potential of only mv may not be experimentally significant. Although the ionic strength was not monitored during the yield stress measurement, 1-5 M KCl is added to the dispersions as background electrolyte. Other authors 18, have recently reported yield stress vs. zeta potential data for oxides suspensions at different concentrations and with different particles sizes. However, the ionic strength in these dispersions was not monitored during the yield stress measurement, but.1 M KNO 3 was added to the dispersions as background electrolyte. The critical zeta potential obtained from these data was found to be not dependent upon solids concentration and particle size. Leong et al. 14, have also reported data of yield stress vs square of zeta potential for anatase TiO and γ- Al O 3 dispersions, and the critical zeta potential determined using this method was not significantly affected by the solid concentration. From Eq. the critical zeta potential is proportional to the square root of the Hamaker constant. The mean value obtained for anatase suspensions according with the data shown in figure 5 was 44x1-1 J. We have also estimated the Hamaker constant through the contact angle measurements, and the value obtained is 34.85x1-1 J. Both values are in good agreement with those reported by other authors 1. The large discrepancy in the Hamaker constant data in the literature clearly reflects the imprecise and inaccurate nature of the current techniques or methods for determining the Hamaker constant of solids.

With this new zeta potential yield stress technique, the precision for Hamaker constant determination may be improved. CONCLUSIONS The relationship between yield stress and the square of zeta potential is linear with a negative slope for TiO dispersions, indicating that the suspension behaviour obeys the DLVO theory. A critical zeta potential of 3 mv has been obtained for TiO suspensions, at an ionic strength of 1-5 M KCl. From the critical zeta potential, the estimated Hamaker constant is 37x1-1 J, which is in good agreement with the experimental value obtained by the contact angle technique of 34.85 x1-1 J. The yield stress-zeta potential technique can be used as an indirect measure of the Hamaker constant, and it is capable of characterising dispersions with solid concentration as high as that used in the yield stress measurement, and the results obtained are usually insensitive to the relatively small amount of impurities. This technique can also be used as an indirect measure of the strength of the attractive interaction between particles in dispersions in the flocculated state. ACKNOWLEDGMENTS A.I.G.M. would like to acknowledge the Group of Rheology of the Twente University (The Netherlands), in which part of this work was carried out, and the University of Málaga for the financial support. REFERENCES 1. Dejarguin B.V., Landau, L. (1941) Acta Phys. Chim. URSS, 14, 633.. Verwey, E.J.W., Overbeek, J.Th.G. (1948), Theory of the Stability of Lyophobic Colloid, Elsevier, Amsterdam. 3. Leong, Y.K., Boger, D.V., Parris, D. (1991), J. Rheol., 35, 149. 4. Lifshitz, E.M. (1956), Sov. Phys. JETP, 73. 5. Dzyaloshinskii, I:E:, Lifshitz, E.M., Pitaevskii, L.P. (1961), Ad. Phys., 1, 165. 6. Tabor, D., Winterton, R.H.S. (1969), Proc. R. Soc. London., A 31, 435. 7. Hough, D.B., White, L.R. (198), Ad. Colloid Interface Science, 14, 41. 8. Israelachvili, J.N., Adams, G.E. (1978), J. Chem. Soc., Faraday Trans., 74, 975. 9. Luckham, P.F. (1989), Powder Tecnology, 58, 75. 1. Larson, I., Drummond, C.J., Chan, D.Y.C., Grieser, F. (1993), J. Am. Chem. Soc., 115, 11885. 11. Leong, Y.K., Scales, P.J., Healy, T.W., Boger, D.V. (1995), J. Am. Chem. Soc., 78, 9. 1. Leong, Y.K., Scales, P.J., Healy, T.W., Boger, D.V., Buscall, R. (1993), J. Chem. Soc. Faraday Trans., 89, 473. 13. Leong, Y.K. (), Chemeca, Perth, CD Proceedings, 314-319, ISBN -646-3991-1. 14. Leong, Y.K., Ong, B.C. (3), Powder Technology, 134, 49. 15. Hunter, R.J., Matarese, R., Napper, D.H. (1983), Colloids Surf.,7, 1. 16. Avramidis, K.S., Turian, R.M. (1991), J. Colloid Interface Sci., 143, 54. 17. Larson, R.G. (1999), The Structure and Rheology of Complex Fluids, Oxford Univ. Press, New York. 18. Zhou, Z.W., Scales, P.J., Boger, D.V. (1), Chem. Eng. Sci., 56, 91. 19. Reed, J.S. (1995), Principles of Ceramics Processing. Wiley, New York, USA.. Scales, P.J., Kapur,P.C., Johnson, S.B., Healy, T.W. (1998), A.I.Ch.E. Journal, 44, 538. 1. Ackler, H.D., French, R.H., Chaing, Y.M.(1996), J. Colloid Interface Sci., 179, 46.