arxiv:cond-mat/ v1 [cond-mat.soft] 9 Aug 1997

Similar documents
Liquid crystal phase transitions in dispersions of rod-like colloidal particles

Generalized depletion potentials

Depletion interactions in binary mixtures of repulsive colloids

Interfacial forces and friction on the nanometer scale: A tutorial

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

INTERMOLECULAR AND SURFACE FORCES

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

Important practical questions:

Entropic wetting and the fluid fluid interface of a model colloid polymer mixture

Applied Surfactants: Principles and Applications

Colloidal Suspension Rheology Chapter 1 Study Questions

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

Fluid-solid transitions on walls in binary hard-sphere mixtures

Phase behaviour of very asymmetric binary mixtures

arxiv:cond-mat/ v1 2 Feb 94

Complete and precise descriptions based on quantum mechanics exist for the Coulombic/Electrostatic force. These are used to describe materials.

Measuring particle aggregation rates by light scattering

Ken-ichi Amano, Kota Hashimoto, and Ryosuke Sawazumi. Department of Energy and Hydrocarbon Chemistry, Graduate School of Engineering,

A theory for calculating the number density distribution of. small particles on a flat wall from pressure between the two

Supporting Information

Proton and electron mass derived as the vacuum energy displaced by a Casimir cavity. By: Ray Fleming

arxiv:cond-mat/ v2 [cond-mat.dis-nn] 3 Oct 2002

COLLOIDAL SELF ASSEMBLY I: INTERACTIONS & PACMEN

SOLUTIONS TO CHAPTER 5: COLLOIDS AND FINE PARTICLES

Particle Characterization Laboratories, Inc.

A phenomenological model for shear-thickening in wormlike micelle solutions

Supplementary Information: Triggered self-assembly of magnetic nanoparticles

Confinement of polymer chains and gels

Experimental Soft Matter (M. Durand, G. Foffi)

arxiv: v1 [cond-mat.soft] 3 Feb 2011

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

Generalizations for the Potential of Mean Force between Two Isolated Colloidal Particles from Monte Carlo Simulations

Electrostatic Double Layer Force: Part III

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

Fluid structure in colloid-polymer mixtures: the competition between electrostatics and depletion

Gas-liquid phase separation in oppositely charged colloids: stability and interfacial tension

Influence of solvent quality on polymer solutions: A Monte Carlo study of bulk and interfacial properties

Micromechanics of Colloidal Suspensions: Dynamics of shear-induced aggregation

Isotropic-nematic phase transition in suspensions of filamentous virus and the neutral polymer Dextran

Measuring colloidal interactions with confocal microscopy

6 Hydrophobic interactions

Phase separation in suspensions of colloids, polymers and nanoparticles: Role of solvent quality, physical mesh, and nonlocal entropic repulsion

Randomly Triangulated Surfaces as Models for Fluid and Crystalline Membranes. G. Gompper Institut für Festkörperforschung, Forschungszentrum Jülich

Self-Assembly of Colloidal Particles

Depletion forces induced by spherical depletion agents

Intermolecular and Surface Forces

Thermodynamically Stable Emulsions Using Janus Dumbbells as Colloid Surfactants

Cavity QED induced colloidal attraction?

Biochemistry,530:,, Introduc5on,to,Structural,Biology, Autumn,Quarter,2015,

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

Colloidal stabilization via nanoparticle halo formation

Polymers in a slabs and slits with attracting walls

Steric stabilization. Dispersions in liquids: suspensions, emulsions, and foams ACS National Meeting April 9 10, 2008 New Orleans

2 Structure. 2.1 Coulomb interactions

arxiv: v1 [cond-mat.soft] 22 Oct 2007

arxiv:cond-mat/ v1 17 Mar 1993

arxiv:cond-mat/ v1 [cond-mat.other] 4 Aug 2004

ON ALGORITHMS FOR BROWNIAN DYNAMICS COMPUTER SIMULATIONS

Forces Acting on Particle

Origins of Mechanical and Rheological Properties of Polymer Nanocomposites. Venkat Ganesan

Lines of Renormalization Group Fixed Points for Fluid and Crystalline Membranes.

Depletion interactions in colloid-polymer mixtures

Scanning Force Microscopy

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 5 Jul 2005

Chitin nanocrystal dispersions: rheological and microstructural properties

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

Dispersion systems. Dispersion system = dispersed phase in a continuum phase (medium) s/l, l/l,... According to the size of the dispersed phase:

Critical Depletion Force

Swelling and Collapse of Single Polymer Molecules and Gels.

Contents. Preface XIII

VERIFYING COULOMB S LAW

Chapter 2 Controlled Synthesis: Nucleation and Growth in Solution

Overview of DLVO Theory

Supporting information

arxiv:cond-mat/ v3 [cond-mat.soft] 12 Sep 2001

Jean-François Dufrêche

Steric stabilization i the role of polymers

Stability of colloidal systems

arxiv:cond-mat/ v1 [cond-mat.soft] 19 Mar 2001

Proteins polymer molecules, folded in complex structures. Konstantin Popov Department of Biochemistry and Biophysics

IPLS Retreat Bath, Oct 17, 2017 Novel Physics arising from phase transitions in biology

Contents. Preface XIII. 1 General Introduction 1 References 6

On the potential of mean force of a sterically stabilized dispersion

Colloids as nucleons

Devitrification of colloidal glasses in real space

Theory of Interfacial Tension of Partially Miscible Liquids

A General Equation for Fitting Contact Area and Friction vs Load Measurements

Interfacial Tension of Electrolyte Solutions

arxiv:cond-mat/ v1 2 Jul 2000

Capillary-gravity waves: The effect of viscosity on the wave resistance

26.542: COLLOIDAL NANOSCIENCE & NANOSCALE ENGINEERING Fall 2013

Long-range Casimir interactions between impurities in nematic liquid crystals and the collapse of polymer chains in such solvents

Lecture 5: Macromolecules, polymers and DNA

CHEMISTRY PHYSICAL. of FOODS INTRODUCTION TO THE. CRC Press. Translated by Jonathan Rhoades. Taylor & Francis Croup

Field Ionization of C 2 H 5 I in Supercritical Argon near the Critical Point

arxiv:cond-mat/ v2 [cond-mat.soft] 10 Dec 2002

Proteins in solution: charge-tuning, cluster formation, liquid-liquid phase separation, and crystallization

M1.(a) (i) giant lattice allow each carbon atom is joined to three others 1

we1 = j+dq + = &/ET, (2)

Transcription:

Depletion forces between two spheres in a rod solution. K. Yaman, C. Jeppesen, C. M. Marques arxiv:cond-mat/9708069v1 [cond-mat.soft] 9 Aug 1997 Department of Physics, U.C.S.B., CA 93106 9530, U.S.A. Materials Research Laboratory, U.C.S.B., CA 93106, U.S.A. C.N.R.S.-Rhône-Poulenc, Complex Fluids Laboratory UMR166 Cranbury, NJ 08512 7500, USA (October 5, 2018) Abstract We study the depletion interaction between spherical particles of radius R immersed in a dilute solution of rigid rods of length L. The computed interaction potential is, within numerical accuracy, exact for any value of L/R. In particular we find that for L R, the depth of the depletion well is smaller than the prediction of the Derjaguin approximation. Our results bring new light into the discussion on the lack of phase separation in colloidal mixtures of spheres and rods. PACS: 82.70.Dd; 64.75.+g Typeset using REVTEX 1

Mixtures of colloids are abundant in industry as paints, glues, lubricants and other materials [1]. They are also present in the preparation of foods and drugs, and in the biological realm: many living organisms or components of living organisms are colloidal suspensions of a variety of sizes and shapes. A pervasive issue in the colloidal domain is the stability of colloidal suspensions. Stability is often necessary for practical purposes as in paint formulation for instance, but certain applications like water treatment or mineral recovery might instead require aggregation or flocculation to occur. The determination of the stability criteria or the study of the flocculation kinetics can be achieved with reasonably good accuracy once the interparticle interaction potentials are known [2]. In a system of a pure colloidal species the two main interactions are the van-der-waals attraction and the hard-core repulsion. Such a system is intrinsically unstable, the van-der-waals attractive component of the potential always leading to flocculation. In practice the stabilization of the suspension is enforced by using the screened electrostatic repulsion in aqueous solutions or steric polymer layers in organic solvents. To a good approximation the stabilized suspension can then be regarded as hard-core particles with no attractive potential component. One way of treating the stability issue in a colloidal mixture of two components is by considering the effective potential that the second species induces between two particles of the first species. A well studied case [3,4] is the attractive potential that a solution of hard-core spheres of radius r 0 induce between two spheres of radius R, (see fig. 1): U s (H) = k B T 3 8 φ R ( 2 H ) 2 ( 1+ 2 r 0 r 0 r 0 3R + 1 ) H ; for H 2r 0 (1) 6 R where H is the distance between the two large spheres and φ the volume fraction of the small spheres. Of course, U s is zero if H 2r 0 ; by definition of the depletion potential the depletion at H = 2r 0 is the zero-point. We will only give the potential in the regions where it is non-zero from now on. The depletion attraction has been known since the pioneering work [5] of Asakura and Oosawa who calculated the interaction energy between two flat plates immersed in a hard-sphere solution: 2

u f (H) = k ( BT 3 r0 2 4π φ 2 H ) r 0 (2) The depletion potential has a purely entropic origin: the exclusion of the small particles from the gap creates a pressure deficit and hence an effective attraction between the plates. When theexactanalyticalformoftheinteractionisonlyknownforflatplates, onecanstillcompute the interaction potential between spheres much larger than the range of the potential by employing the so-called Derjaguin approximation[6]. For instance, the asymptotic behaviour (as R/(2r 0 ) 1) of eq. (1) can be obtained from eq. (2): U Der. (H) = πr dh u f (H ) = k B T 3 H 8 φ R (2 H ) 2 (3) r 0 r 0 Note that the Derjaguin approximation underestimates the attraction in this case. Indirect experimental evidence for the depletion effect were since long ago provided by the flocculation observed in colloids and emulsions, but more direct observations of the attraction have been only recently performed by force measurements [7] with surface force apparatus or microscopy-techniques [8]. A much less-studied case is the depletion induced by hard-core rods, a surprising fact considering that rod-shaped particles in the colloidal range are present in a large variety of mineral and organic systems [9]; they also exist in the biological realm where examples range from TMV-like virus to fibrils of amyloid β-protein, the molecular agent at the origin of the Alzheimer disease. The depletion interaction between two flat surfaces immersed in a dilute rod solution was studied in [5]: u f (H) = k BT LD 4.2 φ ( 1 1 H 2 (4) φ 2 L) with D the diameter of the rod, and φ the Onsager volume fraction, below which the solution is isotropic. For spheres in a rod solution the potential can be written in general as: U s (H) = k B T 4.2 φ φ R D K 1(H/L) (5) 3

but no explicit form of K 1 is available except in the Derjaguin approximation: K Der. 1 (x) = (π/6) (1 x) 3, which is expected to apply only for L/R 1. Since this potential has quite a large contact value compared to k B T, one would expect on this basis to observe flocculation also in rod/sphere mixtures. Surprisingly, such a phase separation has not yet been clearly identified experimentally. A possible explanation for this lack of experimental evidence has recently been proposed [10], based on repulsive contributions from the surface modification of the rod-rod excluded volume interactions. However such contributions are second order in the rod-concentration, and thus small below the Onsager concentration [11], ρ b = 4.2/(L2 D). The authors of reference [10] also note that most systems of interest involve rods that are of comparable size to the spherical particles; in particular the experiments [12] cited in [10] have 1.16 < L/R < 4.34, a range where the use of the Derjaguin approximation to compute the depletion potential in questionable. In this letter we present exact results for the depletion potential to first order in rod density. Our results are numerical in nature, but exact, i.e. the inaccuracy of our calculation is only limited by the numerical nature of the integration performed. It is for our case reducible to levels at which the error bars on the figures given below are not visible. The method of our calculation is based on the work on surface tension of objects immersed in rigid rod solutions that we published elsewhere [13,14]. We now briefly discuss this method and present our results. We consider two spheres of radius R in a solution of rods of length L, and thickness D, with L/D 1. We work to first order in the density of the rods, hence we really consider one rod and two spheres. A priori this is only valid when the rod-rod interactions can be ignored: for rod concentrations below the threshold ρ b. The interaction potential is given [14] by the differences in the grand potentials of the spheres at infinite separation and inside the range of the potential, L. These grand potentials are integrals over the phase space available to a rigid rod: If the rod s center of mass is at a perpendicular distance z greater than L/2 to the surface of a sphere, then the rod is free to rotate, otherwise its rotational degrees of freedom are limited. The depletion region of 4

the rod is shown in fig. 1. This region can be further broken into two: In the first one (I) the rod interacts only with one sphere, whereas in the second one (II) interactions with both spheres are possible. Our previous work [13] provides exact expressions for the contribution of region I to the free energy for all values of (L/R). Region II needs to be further broken down to three sub-regions where the functional form of the phase space allowed to the rod is different. The integration over the angular degrees of freedom can still be performed analytically, but integration over the possible center of mass positions is complicated by the lack of analytical expressions for the boundaries between the sub-regions where different functional forms apply. We therefore use the following procedure: once the position of the center of mass is given, we compute two angles, θ a and θ b, (see fig. 2), from its coordinates, and then numerically decide what functional form to use. Since we have exact expressions for the allowed phase spaces, this procedure reduces the problem to a two-dimensional numerical integration from a four-dimensional one, which can be performed rapidly on a personal computer. One can, of course, do the calculation using Monte-Carlo integration, and we have checked our results using this technique. When region II is not too small as compared to the total depletion region, Monte-Carlo yields reasonable accuracy, moderately fast. The results for these small H values are in perfect agreement, (see fig. 3), with the exact method, for values of (L/R) anywhere from 0.1 to 15.0. When H gets larger, though, Monte-Carlo results fluctuate considerably; longer machine time is needed to reduce these fluctuations. Representative results for the function K 1 (x) defined in eq. (5) are shown in fig. 4. The Derjaguin function, K Der. 1 (x) is also shown for comparison. As can be seen from this plot, Derjaguin is indeed an excellent approximation to the depletion potential when (L/R) 1. What is noteworthy is the large and systematic deviation from this approximation when (L/R) grows. This deviation is about 50% when (L/R) = 1, and becomes more pronounced as (L/R) increases. This can be understood in the light of general results we presented [14] recently. These results relate the second derivative of the free energy of a rod system to the measure of configurations where the rod can wedge between two points on the surface. 5

Clearly, in our case for a fixed H, this measure goes to zero as the rod gets longer. This implies that the free energy becomes a linear function of (L/R), and the function K 1, being related to the free energy by a factor of (L/R) 2, vanishes as R/L. Figure 5 shows the behaviour of the contact value, K 1 (0), as a function of the rod/sphere ratio. We fitted the data of fifteen points to a function with three parameters. The resulting fit is: K 1 (0) = π 6 1+0.8762(L/R) 1+1.33198(L/R)+0.98225(L/R) 2 (6) as can be seen in fig. 5 the fit is very good indeed for the range 0.05 < L/R < 13. The lack of phase separation in the experiments [12] is less surprising in the light of our results. Previous expectations based on Derjaguin approximation estimate the value of the attraction minimum at 10 20k B T. Our results show that the real value is much lower. This reduction amounts to a factor of three to five for these experiments and reduces the depth of the depletion potential to a few k B T. Note also that other factors like chain flexibility are likely to reduce the strength of the attraction further. We thank Phil Pincus for useful discussions. KY acknowledges support from the National Science Foundation(NSF) awards DMR96-32716, and DMR8-442490-22213-3, CJ from NSF Grant CDA96-01954. 6

REFERENCES [1] Russel W. B., Saville D. A. and Schowalter W.R., Colloidal Dispersions,(Cambridge University Press, Cambridge, England) 1989 [2] Proceedings of the First Meeting of the European Network on Colloid Physics, Physica A, 235 (1997) [3] Asakura S. and Oosawa F., Journal of Polymer Science, 33 (1958) 183 [4] Vrij A. and Overbeek J. Th. G., J. Colloid Interface Sci., 17 (1962) 570 [5] Asakura S. and Oosawa F., J. Chem. Phys., 22 (1954) 1255 [6] Israelachvili J. N., Intermolecular and Surface Forces, (Academic Press, London) 1985 [7] Richetti P. and Kekicheff P., Phys. Rev. Lett., 68 (1992) 1951 [8] Dinsmore A. D., Yodh A. G. and Pine D. J., Nature, 383 (1996) 239 [9] Buining P. A., J. Phys. Chem., 97 (1993) 11510 [10] Mao Y., Cates M. E. and Lekkerkerker H. N. W., Phys. Rev. Lett., 75 (1995) 4548 [11] Onsager L., Ann. N.Y. Acad. Sci., 51 (1949) 627 [12] Tracy M. A., Garcia J. L. and Pecora R., Macromolecules, 26 (1993) 1862 [13] Yaman K., Pincus P. and Marques C.M., Phys. Rev. Lett., 78 (1997) 4514 [14] Yaman K., Jeng M., Jeppesen C., Pincus P. and Marques C.M., to appear in Physica A, e-print: http://xxx.lanl.gov/abs/cond-mat/9706005 7

FIGURES a) I I L/2 R II H b) I I L/2 R II H FIG. 1. The depletion region: a) for H > L/2, b) for H < L/2. 8

θ a θ b FIG. 2. The basic construction of the numerical integration method. 9

0-0.1-0.2 Derjaguin MC (425,000,000) MC (25,000,000) integral K1(x) -0.3-0.4-0.5-0.6 0 0.2 0.4 0.6 0.8 1 x = H/L FIG. 3. Monte-Carlo compared to the numerical integral for L/R = 0.1. 10

0-0.1-0.2 Derjaguin L/R = 0.05 L/R = 0.20 L/R = 1.00 L/R = 2.00 L/R = 3.50 L/R = 5.00 L/R = 15.0 L/R = 30.0 K1(x) -0.3-0.4-0.5-0.6 0 0.2 0.4 0.6 0.8 1 x = H/L FIG. 4. The function K 1 (x) for several values of L/R. 11

-0.5-1 -1.5 ln(k1(0)) -2-2.5-3 -3.5-4 -3-2 -1 0 1 2 3 x = ln(l/r) FIG. 5. Contact values.the function plotted here is the one given in the text. Notice the drastic change in the functional form of the data when L crosses 2R. 12