Hostname: page-component-77c89778f8-m42fx Total loading time: 0 Render date: 2024-07-21T11:23:11.190Z Has data issue: false hasContentIssue false

Non-continuum lubrication flows between particles colliding in a gas

Published online by Cambridge University Press:  26 April 2006

R. R. Sundararajakumar
Affiliation:
School of Chemical Engineering, Cornell University, Ithaca, NY 14853, USA
Donald L. Koch
Affiliation:
School of Chemical Engineering, Cornell University, Ithaca, NY 14853, USA

Abstract

Solid-body collisions between smooth particles in a gas would not occur if the lubrication force for a continuum incompressible fluid were to hold at all particle separations. When the gap between the particles is of the order of the mean free path λ0 of the gas, the discrete molecular nature of the gas becomes important. For particles of radii a smaller than about 50 μm colliding in air at a relative velocity comparable to their terminal velocity, the effects of compressibility of the gas in the gap are not important.

The nature of the flow in the gap depends on the relative magnitudes of the minimum gap thickness h0aε, the mean-free path λ0, and the distance aε1/2 over which the effects of curvature become important. The slip-flow regime, a[Gt ]λ0, was analysed by Hocking (1973) using the Maxwell slip boundary condition at the particle surface. To find the lubrication force in the transition regime (aε ∼ O(λ0)), we use the results of Cercignani & Daneri (1963) for the flux as a function of the pressure gradient in a Poiseuille channel flow. When aε[Lt ]λ0[Lt ]aε1/2, one might expect the local flow in the gap to be governed by Knudsen diffusion. However, an attempt to calculate the Knudsen diffusivity between parallel plates leads to a logarithmic divergence, which is cut off by intermolecular collisions, and the flux is therefore proportional to h0c log(λ0/h0), where c is the mean molecular speed. The non-continuum lubrication force is shown to have a weak, log - log divergence as the particle separation goes to zero. As a result, the energy dissipated in the collision is finite. In the limit of large particle inertia, the energy dissipated is 6πμU0a2(log h00 – 1.28), where 2U0 is the relative velocity of the particles.

When λ0[Gt ]aε1/2, we have a free molecular flow in the gap. In this case, owing to the curvature of the particles, the flux versus pressure gradient relation is non-local. We analyse the free molecular flow between two cylinders and obtain scalings for the lubrication force.

Type
Research Article
Copyright
© 1996 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

Abramowitz, M. & Stegun, I. A. 1993 Handbook of Mathematical Functions. Dover
Barnocky, G. & Davis, R. H. 1989 The influence of pressure dependent density and viscosity on the elastohydrodynamic collision and rebound of two spheres. J. Fluid Mech. 209, 501.Google Scholar
Bender, C. L. & Orszag, S. A. 1978 Advanced Mathematical Methods for Scientists and Engineers. McGraw-Hill
Bohnet, M. 1983 In Handbook of Fluids in Motion (ed. N. P. Cheremisinoff & R. Gupta). Ann Arbor.
Cercignani, C. 1963 Plane poiseuille flow and Knudsen minimum effect. In Rarefied gas dynamics ed. (J. A. Laurmann), vol. 2, p. 92. Academic.
Cercignani, C. 1975 Theory and Application of the Boltzmann Equation. Elsevier
Cercignani, C. & Daneri, A. 1963 Flow of a rarefied gas between two parallel plates. J. Appl. Phys. 34, 3509.Google Scholar
Chapman, S. & Cowling, T. G. 1990 The Mathematical Theory of Non-uniform Gases. Cambridge University Press
Davis, R. H. 1984 The rate of coagulation of a dilute polydisperse system of sedimenting spheres. J. Fluid Mech. 145, 179.Google Scholar
Davis, R. H., Serraysol, J. & Hinch, E. J. 1986 The elastohydrodynamic collisions of two spheres. J. Fluid Mech. 163, 479.Google Scholar
Hickey, K. A. & Loyalka, S. K. 1990 Plane poiseuille flow: Rigid sphere gas. J. Vac. Sci. Technol. A 8, 957.Google Scholar
Hocking, L. M. 1973 The effect of slip on the motion of a sphere close to a wall and of two adjacent spheres. J. Engng Maths 7, 207.Google Scholar
Hocking, L. M. & Jonas, P. R. 1970 The collision efficiency of small drops. Q. J. R. Met. Soc. 96, 722.Google Scholar
Jeffrey, D. J. & Onishi, Y. 1984 Calculation of the resistance and mobility functions for two unequal sized spheres in low-Reynolds-number flow. J. Fluid Mech. 139, 261.Google Scholar
Jenkins, J. T. & Savage, S. B. 1983 A theory for the rapid flow of identical, smooth, nearly elastic spherical particles. J. Fluid Mech. 130, 187.Google Scholar
Koch, D. L. 1990 Kinetic theory for a monodisperse gas-solid suspension. Phys. Fluids A 2, 1711.Google Scholar
Kumaran, V. & Koch, D. L. 1993 Properties of a bidisperse particle-gas suspension. Part 1. Collision time small compared with viscous relaxation time. J. Fluid Mech. 247, 623.Google Scholar
Kytömaa, H. K. & Schmid, P. J. 1992 On the collision of rigid spheres in a weakly compressible fluid. Phys. Fluids A 4, 2683.Google Scholar
Loyalka, S. K. & Ferziger, J. H. 1967 Model dependence of the slip coefficient. Phys. Fluids 10, 1833.Google Scholar
Ochs, H. T. & Beard, K. V. 1985 Effects of coalescence efficiencies on the formation of precipitation. J. Atmos. Sci. 42, 1451.Google Scholar
Ying, R. & Peters, M. H. 1989 Hydrodynamic interaction of two unequal-sized spheres in a slightly rarefied gas: resistance and mobility functions. J. Fluid Mech. 207, 353.Google Scholar