Hostname: page-component-848d4c4894-hfldf Total loading time: 0 Render date: 2024-05-21T10:14:31.263Z Has data issue: false hasContentIssue false

Isotropically active particle closely fitting in a cylindrical channel: spontaneous motion at small Péclet numbers

Published online by Cambridge University Press:  26 January 2024

Rodolfo Brandão*
Affiliation:
Department of Mechanical and Aerospace Engineering, Princeton University, Princeton, NJ 08544, USA
*
Email address for correspondence: r.brandao@princeton.edu

Abstract

Spontaneous motion due to symmetry breaking has been predicted theoretically for both active droplets and isotropically active particles in an unbounded fluid domain, provided that their intrinsic Péclet number $Pe$ exceeds a critical value. However, due to their inherently small $Pe$, this phenomenon has yet to be observed experimentally for active particles. In this paper, we demonstrate theoretically that spontaneous motion for an active spherical particle closely fitting in a cylindrical channel is possible at arbitrarily small $Pe$. Scaling arguments in the limit where the dimensionless clearance is $\epsilon \ll 1$ reveal that when $Pe=O(\epsilon ^{1/2})$, the confined particle reaches speeds comparable to those achieved in an unbounded fluid at moderate (supercritical) $Pe$ values. We use matched asymptotic expansions in that distinguished limit, where the fluid domain decomposes into several asymptotic regions: a gap region, where the lubrication approximation applies; particle-scale regions, where the concentration is uniform; and far-field regions, where solute transport is one-dimensional. We derive an asymptotic formula for the particle speed, which is a monotonically decreasing function of $\overline {Pe}=Pe/\epsilon ^{1/2}$ and approaches a finite limit as $\overline {Pe}\searrow 0$. Our results could pave the way for experimental realisations of symmetry-breaking spontaneous motion in active particles.

Type
JFM Papers
Copyright
© The Author(s), 2024. Published by 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

Anderson, J.L. 1989 Colloid transport by interfacial forces. Annu. Rev. Fluid Mech. 30, 139165.Google Scholar
de Blois, C., Bertin, V., Suda, S., Ichikawa, M., Reyssat, M. & Dauchot, O. 2021 Swimming droplets in 1D geometries: an active Bretherton problem. Soft Matt. 17 (27), 66466660.CrossRefGoogle ScholarPubMed
Bretherton, F.P. 1961 The motion of long bubbles in tubes. J. Fluid Mech. 10 (2), 166188.Google Scholar
Bungay, P.M. & Brenner, H. 1973 The motion of a closely-fitting sphere in a fluid-filled tube. Intl J. Multiphase Flow 1 (1), 2556.Google Scholar
de Buyl, P., Mikhailov, A.S. & Kapral, R. 2013 Self-propulsion through symmetry breaking. Europhys. Lett. 103 (6), 60009.CrossRefGoogle Scholar
Desai, N. & Michelin, S. 2021 Instability and self-propulsion of active droplets along a wall. Phys. Rev. Fluids 6 (11), 114103.Google Scholar
Desai, N. & Michelin, S. 2022 Steady state propulsion of isotropic active colloids along a wall. Phys. Rev. Fluids 7 (10), 100501.CrossRefGoogle Scholar
Golestanian, R., Liverpool, T.B. & Ajdari, A. 2007 Designing phoretic micro-and nano-swimmers. New J. Phys. 9, 126.Google Scholar
Happel, J. & Brenner, H. 1965 Low Reynolds Number Hydrodynamics. Prentice-Hall.Google Scholar
Hinch, E.J. 1991 Perturbation Methods. Cambridge University Press.Google Scholar
Izri, Z., Van Der Linden, M.N., Michelin, S. & Dauchot, O. 2014 Self-propulsion of pure water droplets by spontaneous marangoni-stress-driven motion. Phys. Rev. Lett. 113 (24), 248302.Google Scholar
Michelin, S. 2023 Self-propulsion of chemically active droplets. Annu. Rev. Fluid Mech. 55, 77101.Google Scholar
Michelin, S. & Lauga, E. 2014 Phoretic self-propulsion at finite Péclet numbers. J. Fluid Mech. 747, 572604.Google Scholar
Michelin, S., Lauga, E. & Bartolo, D. 2013 Spontaneous autophoretic motion of isotropic particles. Phys. Fluids 25 (6), 061701.Google Scholar
Moran, J.L. & Posner, J.D. 2017 Phoretic self-propulsion. Annu. Rev. Fluid Mech. 49, 511540.CrossRefGoogle Scholar
Paxton, W.F., Kistler, K.C., Olmeda, C.C., Sen, A., Angelo, S.K.S., Cao, Y., Mallouk, T.E., Lammert, P.E. & Crespi, V.H. 2004 Catalytic nanomotors: autonomous movement of striped nanorods. J. Am. Chem. Soc. 126 (41), 1342413431.Google Scholar
Peng, G.G. & Schnitzer, O. 2023 Weakly nonlinear dynamics of a chemically active particle near the threshold for spontaneous motion. II. History-dependent motion. Phys. Rev. Fluids 8, 033602.Google Scholar
Picella, F. & Michelin, S. 2022 Confined self-propulsion of an isotropic active colloid. J. Fluid Mech. 933, A27.Google Scholar
Saha, S., Yariv, E. & Schnitzer, O. 2021 Isotropically active colloids under uniform force fields: from forced to spontaneous motion. J. Fluid Mech. 916, A47.Google Scholar
Schnitzer, O. 2023 Weakly nonlinear dynamics of a chemically active particle near the threshold for spontaneous motion. I. Adjoint method. Phys. Rev. Fluids 8, 034201.Google Scholar
Sherwood, J.D. & Ghosal, S. 2021 Electrophoresis of tightly fitting spheres along a circular cylinder of finite length. J. Fluid Mech. 929, A45.CrossRefGoogle Scholar
Yariv, E. 2016 Wall-induced self-diffusiophoresis of active isotropic colloids. Phys. Rev. Fluids 1 (3), 032101.Google Scholar
Yariv, E. 2017 Boundary-induced autophoresis of isotropic colloids: anomalous repulsion in the lubrication limit. J. Fluid Mech. 812, 2640.Google Scholar
Yariv, E. & Kaynan, U. 2017 Phoretic drag reduction of chemically active homogeneous spheres under force fields and shear flows. Phys. Rev. Fluids 2 (1), 012201.Google Scholar