Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-23T21:52:23.876Z Has data issue: false hasContentIssue false

High-frequency effective viscosity of a dilute suspension of particles in Poiseuille flow between parallel walls

Published online by Cambridge University Press:  30 June 2016

François Feuillebois*
Affiliation:
LIMSI-CNRS, UPR 3251, BP 133, 91403 Orsay CEDEX, France
Maria L. Ekiel-Jeżewska
Affiliation:
Institute of Fundamental Technological Research, Polish Academy of Sciences, Pawińskiego 5b, 02-106 Warsaw, Poland
Eligiusz Wajnryb
Affiliation:
Institute of Fundamental Technological Research, Polish Academy of Sciences, Pawińskiego 5b, 02-106 Warsaw, Poland
Antoine Sellier
Affiliation:
LadHyX, École Polytechnique, 91128 Palaiseau CEDEX, France
Jerzy Bławzdziewicz
Affiliation:
Department of Mechanical Engineering, Texas Tech University, Lubbock, TX 79409, USA
*
Email address for correspondence: Francois.Feuillebois@limsi.fr

Abstract

It is shown that the formal expression for the effective viscosity of a dilute suspension of arbitrary-shaped particles in Poiseuille flow contains a novel quadrupole term, besides the expected stresslet. This term becomes important for a very confined geometry. For a high-frequency flow field (in the sense used in Feuillebois et al. (J. Fluid Mech., vol. 764, 2015, pp. 133–147), the suspension rheology is Newtonian at first order in volume fraction. The effective viscosity is calculated for suspensions of $N$-bead rods and of prolate spheroids with the same length, volume and aspect ratio (up to 6), entrained by the Poiseuille flow between two infinite parallel flat hard walls. The numerical computations, based on solving the Stokes equations, indicate that the quadrupole term gives a significant positive contribution to the intrinsic viscosity $[{\it\mu}]$ if the distance between the walls is less than ten times the particle width, or less. It is found that the intrinsic viscosity in bounded Poiseuille flow is generally smaller than the corresponding value in unbounded flow, except for extremely narrow gaps when it becomes larger because of lubrication effects. The intrinsic viscosity is at a minimum for a gap between walls of the order of 1.5–2 particle width. For spheroids, the intrinsic viscosity is generally smaller than for chains of beads with the same aspect ratio, but when normalized by its value in the bulk, the results are qualitatively the same. Therefore, a rigid chain of beads can serve as a simple model of an orthotropic particle with a more complicated shape. The important conclusion is that the intrinsic viscosity in shear flow is larger than in the Poiseuille flow between two walls, and the difference is significant even for relatively wide channels, e.g. three times wider than the particle length. For such confined geometries, the hydrodynamic interactions with the walls are significant and should be taken into account.

Type
Papers
Copyright
© 2016 Cambridge University Press 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Batchelor, G. K. 1970 The stress system in a suspension of force-free particles. J. Fluid Mech. 41, 545570.CrossRefGoogle Scholar
Bhattacharya, S. 2008 Cooperative motion of spheres arranged in periodic grids between two parallel walls. J. Chem. Phys. 128, 074709.Google Scholar
Bhattacharya, S. & Bławzdziewicz, J. 2008 Effect of small particles on the near-wall dynamics of a large particle in a highly bidisperse colloidal solution. J. Chem. Phys. 128, 214704.Google Scholar
Bhattacharya, S., Bławzdziewicz, J. & Wajnryb, E. 2006 Far-field approximation for hydrodynamic interactions in parallel-wall geometry. J. Comput. Phys. 212, 718738.CrossRefGoogle Scholar
Bławzdziewicz, J. & Wajnryb, E. 2008 An analysis of the far-field response to external forcing of a suspension in Stokes flow in a parallel-wall channel. Phys. Fluids. 20, 093303.CrossRefGoogle Scholar
Brenner, H. 1970 Pressure drop due to the motion of neutrally buoyant particles in duct flow. J. Fluid Mech. 43, 641660.Google Scholar
Chwang, A. T. & Wu, T. Y. 1975 Hydromechanics of low-Reynolds-number flow. Part 2. Singularity method for Stokes flow. J. Fluid Mech. 67, 787815.CrossRefGoogle Scholar
Cox, R. G. & Mason, S. G. 1971 Suspended particles in fluid flow through tubes. Annu. Rev. Fluid Mech. 3, 300321.CrossRefGoogle Scholar
Einstein, A. 1906 Eine neue Bestimmung der Molekuldimensionen. Ann. Phys. 19, 289306.CrossRefGoogle Scholar
Ekiel-Jeżewska, M. L. & Wajnryb, E. 2009 Precise multipole method for calculating hydrodynamic interactions between spherical particles in the Stokes flow. In Theoretical Methods for Micro Scale Viscous Flows (ed. Feuillebois, F. & Sellier, A.), pp. 127172. Transworld Research Network, ISBN: 978-81-7895-400-4.Google Scholar
Ezhilan, B. & Saintillan, D. 2015 Transport of a dilute active suspension in pressure-driven channel flow. J. Fluid Mech. 777, 482522.Google Scholar
Feuillebois, F., Ekiel-Jeżewska, M. L., Wajnryb, E., Sellier, A. & Bławzdziewicz, J. 2015 High-frequency viscosity of a dilute suspension of elongated particles in a linear shear flow between two walls. J. Fluid Mech. 764, 133147.Google Scholar
Feuillebois, F., Ghalya, N., Sellier, A. & Elasmi, L. 2012 Influence of wall slip in dilute suspensions. J. Phys.: Conf. Ser. 392, 012012, 1–19.Google Scholar
Jeffery, G. B. 1922 The motion of ellipsoidal particles immersed in a viscous fluid. Proc. R. Soc. Lond. 102, 161179.Google Scholar
Navardi, S. & Bhattacharya, S. 2010 Effect of confining conduit on effective viscosity of dilute colloidal suspension. J. Chem. Phys. 132 (11), 114114.Google Scholar
Park, J., Bricker, J. M. & Butler, J. E. 2007 Cross-stream migration in dilute solutions of rigid polymers undergoing rectilinear flow near a wall. Phys. Rev. E 76, 040801.Google Scholar
Park, J. & Butler, J. E. 2009 Inhomogeneous distribution of a rigid fibre undergoing rectilinear flow between parallel walls at high Péclet numbers. J. Fluid Mech. 630, 267298.Google Scholar
Pasol, L.2003 Interactions hydrodynamiques dans les suspensions. Effets collectifs. PhD thesis, Université Pierre et Marie Curie, Paris VI.Google Scholar
Pasol, L., Martin, M., Ekiel-Jeżewska, M. L., Wajnryb, E., Bławzdziewicz, J. & Feuillebois, F. 2011 Motion of a sphere parallel to plane walls in a Poiseuille flow. Application to field-flow fractionation and hydrodynamic chromatography. Chem. Engng Sci. 66, 40784089; Corrigendum: ibid. 90 (2013) 51–52.Google Scholar
Sangani, A. S., Acrivos, A. & Peyla, P. 2011 Roles of particle-wall and particle-particle interactions in highly confined suspensions of spherical particles being sheared at low Reynolds numbers. Phys. Fluids 23, 083302.Google Scholar
Sheraga, H. A. 1955 Non-Newtonian viscosity of solutions of ellipsoidal particles. J. Chem. Phys. 23 (8), 15261532.Google Scholar
Zurita-Gotor, M., Bławzdziewicz, J. & Wajnryb, E. 2007 Motion of a rod-like particle between parallel walls with application to suspension rheology. J. Rheol. 51, 7197.CrossRefGoogle Scholar