Hostname: page-component-8448b6f56d-qsmjn Total loading time: 0 Render date: 2024-04-24T15:45:01.686Z Has data issue: false hasContentIssue false

Particle motion driven by solute gradients with application to autonomous motion: continuum and colloidal perspectives

Published online by Cambridge University Press:  03 December 2010

JOHN F. BRADY*
Affiliation:
Divisions of Chemistry & Chemical Engineering, and Engineering & Applied Science, California Institute of Technology, Pasadena, CA 91125, USA
*
Email address for correspondence: jfbrady@caltech.edu

Abstract

Diffusiophoresis, the motion of a particle in response to an externally imposed concentration gradient of a solute species, is analysed from both the traditional coarse-grained macroscopic (i.e. continuum) perspective and from a fine-grained micromechanical level in which the particle and the solute are treated on the same footing as Brownian particles dispersed in a solvent. It is shown that although the two approaches agree when the solute is much smaller in size than the phoretic particle and is present at very dilute concentrations, the micromechanical colloidal perspective relaxes these restrictions and applies to any size ratio and any concentration of solute. The different descriptions also provide different mechanical analyses of phoretic motion. At the continuum level the macroscopic hydrodynamic stress and interactive force with the solute sum to give zero total force, a condition for phoretic motion. At the colloidal level, the particle's motion is shown to have two contributions: (i) a ‘back-flow’ contribution composed of the motion of the particle due to the solute chemical potential gradient force acting on it and a compensating fluid motion driven by the long-range hydrodynamic velocity disturbance caused by the chemical potential gradient force acting on all the solute particles and (ii) an indirect contribution arising from the mutual interparticle and Brownian forces on the solute and phoretic particle, that contribution being non-zero because the distribution of solute about the phoretic particle is driven out of equilibrium by the chemical potential gradient of the solute. At the colloidal level the forces acting on the phoretic particle – both the statistical or ‘thermodynamic’ chemical potential gradient and Brownian forces and the interparticle force – are balanced by the Stokes drag of the solvent to give the net phoretic velocity.

For a particle undergoing self-phoresis or autonomous motion, as can result from chemical reactions occurring asymmetrically on a particle surface, e.g. catalytic nanomotors, there is no imposed chemical potential gradient and the back-flow contribution is absent. Only the indirect Brownian and interparticle forces contribution is responsible for the motion. The velocity of the particle resulting from this contribution can be written in terms of a mobility times the integral of the local ‘solute pressure’ – the solute concentration times the thermal energy – over the surface of contact between the particle and the solute. This was the approach taken by Córdova-Figueroa & Brady (Phys. Rev. Lett., vol. 100, 2008, 158303) in their analysis of self-propulsion. It is shown that full hydrodynamic interactions can be incorporated into their analysis by a simple scale factor.

Type
Papers
Copyright
Copyright © Cambridge University Press 2010

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

REFERENCES

Abecassis, B., Cottin-Bizonne, C., Ybert, C., Ajdari, A. & Bocquet, L. 2008 Boosting migration of large particles by solute contrasts. Nature Mater. 7, 785789.CrossRefGoogle ScholarPubMed
Anderson, J. L. 1989 Colloidal transport by interfacial forces. Annu. Rev. Fluid Mech. 21, 6199.CrossRefGoogle Scholar
Asakura, S. & Oosawa, F. 1954 On the interaction between two bodies immersed in a solution of macromolecules. J. Chem. Phys. 22, 12551256.CrossRefGoogle Scholar
Batchelor, G. K. 1970 Stress system in a suspension of force-free particles. J. Fluid Mech. 41, 545570.CrossRefGoogle Scholar
Batchelor, G. K. 1972 Sedimentation in a dilute suspension of spheres. J. Fluid Mech. 752, 245268.CrossRefGoogle Scholar
Batchelor, G. K. 1976 Brownian diffusion of particles with hydrodynamic interaction. J. Fluid Mech. 74, 129.CrossRefGoogle Scholar
Batchelor, G. K. 1982 Sedimentation in a dilute polydisperse system of interacting spheres. Part 1. General theory. J. Fluid Mech. 119, 379408.CrossRefGoogle Scholar
Batchelor, G. K. 1983 Diffusion in a dilute polydisperse system of interacting spheres. J. Fluid Mech. 131, 155175.CrossRefGoogle Scholar
Berne, B. J. & Pecora, R. 1976 Dynamic Light Scattering. Wiley.Google Scholar
Bonnecaze, R. T. & Brady, J. F. 1992 Dynamic simulation of an electrorheological fluid. J. Chem. Phys. 96, 21832202.CrossRefGoogle Scholar
Brady, J. F. 1993 Brownian motion, hydrodynamics and the osmotic pressure. J. Chem Phys. 98, 33353341.CrossRefGoogle Scholar
Brady, J. F. & Bossis, G. 1988 Stokesian dynamics. Annu. Rev. Fluid Mech. 20, 111157.CrossRefGoogle Scholar
Brady, J. F., Phillips, R. J., Lester, J. C. & Bossis, G. 1988 Dynamic simulation of hydrodynamically interacting suspensions. J. Fluid. Mech. 195, 257280.CrossRefGoogle Scholar
Córdova-Figueroa, U. M. 2008 Directed motion of colloidal particles via chemical reactions: Osmotic propulsion. PhD thesis, California Institute of Technology, Pasadena.Google Scholar
Córdova-Figueroa, U. M. & Brady, J. F. 2008 Osmotic propulsion: the osmotic motor. Phys. Rev. Lett. 100, 158303.CrossRefGoogle ScholarPubMed
Ebbens, S. J. & Howse, J. R. 2010 In pursuit of propulsion at the nanoscale. Soft Matt. 6, 726738.CrossRefGoogle Scholar
Foss, D. R. & Brady, J. F. 2000 Structure, diffusion and rheology of Brownian suspensions by Stokesian dynamics simulation. J. Fluid Mech. 407, 167200.CrossRefGoogle Scholar
Golestanian, R., Liverpool, T. B. & Ajdari, A. 2005 Propulsion of a molecular machine by asymmetric distribution of reaction products. Phys. Rev. Lett. 94, 220801.CrossRefGoogle ScholarPubMed
Golestanian, R., Liverpool, T. B. & Ajdari, A. 2007 Designing phoretic micro- and nano-swimmers. New J. Phys. 9, 126.CrossRefGoogle Scholar
Hong, Y., Blackman, N. M. K., Kopp, N. D., Sen, A. & Velegol, D. 2007 Chemotaxis of nonbiological colloidal rods. Phys. Rev. Lett. 99, 178103.CrossRefGoogle ScholarPubMed
Howse, J. R., Jones, R. A. L., Ryan, A. J., Gough, T., Vafabakhsh, R. & Golestanian, R. 2007 Self-motile colloidal particles: from directed propulsion to random walk. Phys. Rev. Lett. 99, 048102.CrossRefGoogle ScholarPubMed
Ibele, M., Mallouk, T. E. & Sen, A. 2009 Schooling behavior of light-powered autonomous micromotors in water. Angew. Chem. Intl Ed. Engl. 48, 33083312.CrossRefGoogle ScholarPubMed
Jülicher, F. & Prost, J. 2009 a Comment on ‘osmotic propulsion: the osmotic motor’. Phys. Rev. Lett. 103, 079801.CrossRefGoogle ScholarPubMed
Jülicher, F. & Prost, J. 2009 b Generic theory of colloidal transport. Eur. Phys. J. E 29, 2736.Google ScholarPubMed
Kagan, D., Calvo-Marzal, P., Balasubramanian, S., Sattayasamitsathit, S., Manesh, K. M., Flechsig, G.-U. & Wang, J. 2009 Chemical sensing based on catalytic nanomotors: motion-based detection of trace silver. J. Am. Chem. Soc. 131, 12802–12083.CrossRefGoogle ScholarPubMed
Khair, A. S. & Brady, J. F. 2006 Single particle motion in colloidal dispersions: a simple model for active and nonlinear microrheology. J. Fluid Mech. 557, 73117.CrossRefGoogle Scholar
Ladd, A. J. C. 1990 Hydrodynamic transport coefficients of random dispersions of hard spheres. J. Chem. Phys. 93, 34843494.CrossRefGoogle Scholar
Nitsche, J. M. & Batchelor, G. K. 1997 Break-up of a falling drop containing dispersed particles. J. Fluid Mech. 340, 161175.CrossRefGoogle Scholar
Paxton, W. F., Kistler, K. C., Olmeda, C. C., Sen, A., St. Angelo, S. K., Cao, Y., Mallouk, T. E., Lammert, P. E. & Crespi, V. H. 2004 Catalytic nanomotors: autonomous movement of striped nanorods. J. Am. Chem. Soc. 126, 1342413431.CrossRefGoogle ScholarPubMed
Paxton, W. F., Sundararajan, S., Mallouk, T. E. & Sen, A. 2006 Chemical locomotion. Angew. Chem. Intl Ed. Engl. 45, 54205429.CrossRefGoogle ScholarPubMed
Russel, W. B., Saville, D. A. & Schowalter, W. R. 1989 Colloidal Dispersions. Cambridge University Press.CrossRefGoogle Scholar
Squires, T. M. & Brady, J. F. 2005 A simple paradigm for active and nonlinear microrheology. Phys. Fluids 17, 073101.CrossRefGoogle Scholar
Wang, Y., Fei, S.-T., Byun, Y.-M., Lammert, P. E, Crespi, V. H., Sen, A. & Mallouk, T. E. 2009 Dynamic interactions between fast microscale rotors. J. Am. Chem. Soc. 131, 99269927.CrossRefGoogle ScholarPubMed
van der Werff, J. C. & de Kruif, C. G. 1989 Hard-sphere colloidal dispersions: the scaling of rheological properties with particle size, volume fraction and shear rate. J. Rheol. 33, 421454.CrossRefGoogle Scholar