Hostname: page-component-7c8c6479df-r7xzm Total loading time: 0 Render date: 2024-03-28T04:04:43.680Z Has data issue: false hasContentIssue false

Geometric phase methods with Stokes theorem for a general viscous swimmer

Published online by Cambridge University Press:  06 April 2021

Lyndon Koens*
Affiliation:
Department of Mathematics and Statistics, Macquarie University, 12 Wally's Walk, Sydney, New South Wales2109, Australia
Eric Lauga
Affiliation:
Department of Applied Mathematics and Theoretical Physics, University of Cambridge, Wilberforce Road, CambridgeCB3 0WA, UK
*
Email address for correspondence: lmk42@cam.ac.uk

Abstract

The geometric phase techniques for swimming in viscous flows express the net displacement of a swimmer as a path integral of a field in configuration space. This representation can be transformed into an area integral for simple swimmers using the Stokes theorem. Since this transformation applies for any loop, the integrand of this area integral can be used to help design these swimmers. However, the extension of this Stokes theorem technique to more complicated swimmers is hampered by problems with variables that do not commute and by how to visualise and understand the higher-dimensional spaces. In this paper, we develop a treatment for each of these problems, thereby allowing the displacement of general swimmers in any environment to be designed and understood similarly to simple swimmers. The net displacement arising from non-commuting variables is tackled by embedding the integral into a higher-dimensional space, which can then be visualised through a suitability constructed surface. These methods are developed for general swimmers and demonstrated on three benchmark examples: Purcell's two-hinged swimmer, an axisymmetric squirmer in free space and an axisymmetric squirmer approaching a free interface. We show in particular that, for swimmers with more than two modes of deformation, there exists an infinite set of strokes that generate each net displacement. Hence, in the absence of additional restrictions, general microscopic swimmers do not have a single stroke that maximises their displacement.

Type
JFM Papers
Copyright
© The Author(s), 2021. 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

REFERENCES

Alapan, Y., Yigit, B., Beker, O., Demirörs, A.F. & Sitti, M. 2019 Shape-encoded dynamic assembly of mobile micromachines. Nat. Mater. 18, 12441251.CrossRefGoogle ScholarPubMed
Avron, J.E. & Raz, O. 2008 A geometric theory of swimming: Purcell's swimmer and its symmetrized cousin. New J. Phys. 10, 063016.CrossRefGoogle Scholar
Barta, E. & Liron, N. 1988 Slender body interactions for low Reynolds numbers–Part I: Body-wall interactions. SIAM J. Appl. Maths 48, 9921008.CrossRefGoogle Scholar
Becker, L.E., Koehler, S.A. & Stone, H.A. 2003 On self-propulsion of micro-machines at low Reynolds number: Purcell's three-link swimmer. J. Fluid Mech. 490, 1535.CrossRefGoogle Scholar
Bianchi, S., Saglimbeni, F. & Di Leonardo, R. 2017 Holographic imaging reveals the mechanism of wall entrapment in swimming bacteria. Phys. Rev. X 7, 011010.Google Scholar
Bittner, B., Hatton, R.L. & Revzen, S. 2018 Geometrically optimal gaits: a data-driven approach. Nonlinear Dyn. 94, 19331948.CrossRefGoogle Scholar
Blake, J.R. 1971 A spherical envelope approach to ciliary propulsion. J. Fluid Mech. 46, 199208.CrossRefGoogle Scholar
Borker, N.S. & Koch, D.L. 2019 Slender body theory for particles with non-circular cross-sections with application to particle dynamics in shear flows. J. Fluid Mech. 877, 10981133.CrossRefGoogle Scholar
Burton, L.J., Hatton, R.L., Choset, H. & Hosoi, A.E. 2010 Two-link swimming using buoyant orientation. Phys. Fluids 22, 091703.CrossRefGoogle Scholar
Chakrabarti, B., Liu, Y., LaGrone, J., Cortez, R., Fauci, L., du Roure, O., Saintillan, D. & Lindner, A. 2020 Flexible filaments buckle into helicoidal shapes in strong compressional flows. Nat. Phys. 16, 689694.CrossRefGoogle Scholar
Chakrabarti, B. & Saintillan, D. 2019 a Hydrodynamic synchronization of spontaneously beating filaments. Phys. Rev. Lett. 123, 208101.CrossRefGoogle ScholarPubMed
Chakrabarti, B. & Saintillan, D. 2019 b Spontaneous oscillations, beating patterns, and hydrodynamics of active microfilaments. Phys. Rev. Fluids 4, 043102.CrossRefGoogle Scholar
Cicconofri, G. & DeSimone, A. 2016 Motion planning and motility maps for flagellar microswimmers. Eur. Phys. J. E 39, 72.CrossRefGoogle ScholarPubMed
Colin, R., Drescher, K. & Sourjik, V. 2019 Chemotactic behaviour of Escherichia coli at high cell density. Nat. Commun. 10, 5329.CrossRefGoogle ScholarPubMed
Cortez, R. 2001 The method of regularized stokeslets. SIAM J. Sci. Comput. 23, 12041225.CrossRefGoogle Scholar
Cortez, R., Fauci, L. & Medovikov, A. 2005 The method of regularized Stokeslets in three dimensions: analysis, validation, and application to helical swimming. Phys. Fluids 17, 031504.CrossRefGoogle Scholar
Crane, K., de Goes, F., Desbrun, M. & SchrŁoder, P. 2013 Digital geometry processing with discrete exterior calculus. In ACM SIGGRAPH 2013 courses, SIGGRAPH ‘13. ACM.Google Scholar
Das, D. & Lauga, E. 2018 Computing the motor torque of Escherichia coli. Soft Matter 14, 59555967.CrossRefGoogle ScholarPubMed
Desimone, A., Heltai, L., Alouges, F. & Lefebvre-Lepot, A. 2012 Computing Optimal Strokes for Low Reynolds Number Swimmers, pp. 177184. Springer.Google Scholar
DeSimone, A. & Tatone, A. 2012 Crawling motility through the analysis of model locomotors: two case studies. Eur. Phys. J. E 35, 85.CrossRefGoogle ScholarPubMed
Drescher, K., Goldstein, R.E. & Tuval, I. 2010 Fidelity of adaptive phototaxis. Proc. Natl Acad. Sci. USA 107, 1117111176.CrossRefGoogle ScholarPubMed
Eastham, P.S. & Shoele, K. 2019 Axisymmetric squirmers in Stokes fluid with nonuniform viscosity. arXiv:1904.01946.CrossRefGoogle Scholar
Felderhof, B.U. & Jones, R.B. 2016 Stokesian swimming of a sphere at low Reynolds number by helical surface distortion. Phys. Fluids 28, 073601.CrossRefGoogle Scholar
Gaffney, E.A., Gadêlha, H., Smith, D.J., Blake, J.R. & Kirkman-Brown, J.C. 2011 Mammalian sperm motility: observation and theory. Annu. Rev. Fluid Mech. 43, 501528.CrossRefGoogle Scholar
Goldstein, R.E. 2015 Green algae as model organisms for biological fluid dynamics. Annu. Rev. Fluid Mech. 47, 343375.CrossRefGoogle ScholarPubMed
Golestanian, R. & Ajdari, A. 2008 Analytic results for the three-sphere swimmer at low Reynolds number. Phys. Rev. E 77, 036308.CrossRefGoogle ScholarPubMed
Gray, J. & Hancock, G.J. 1955 The propulsion of sea-urchin spermatozoa. J. Expl Biol. 32, 802814.Google Scholar
Gueron, S. & Liron, N. 1992 Ciliary motion modeling, and dynamic multicilia interactions. Biophys. J. 63, 10451058.CrossRefGoogle ScholarPubMed
Gutman, E. & Or, Y. 2016 Symmetries and gaits for Purcell's three-link microswimmer model. IEEE Trans. Robot. 32, 5369.CrossRefGoogle Scholar
Hatton, R.L. 2020 Osu-lram/geometricsystemplotter. Available at: https://github.com/OSU-LRAM/GeometricSystemPlotter/.Google Scholar
Hatton, R.L. & Choset, H. 2011 Geometric motion planning: the local connection, Stokes’ theorem, and the importance of coordinate choice. Intl J. Rob. Res. 30, 9881014.CrossRefGoogle Scholar
Hatton, R.L. & Choset, H. 2013 Geometric swimming at low and high Reynolds numbers. IEEE Trans. Robot. 29, 615624.CrossRefGoogle Scholar
Hatton, R.L. & Choset, H. 2015 Nonconservativity and noncommutativity in locomotion. Eur. Phys. J 224, 31413174.Google Scholar
Hatton, R.L., Dear, T. & Choset, H. 2017 Kinematic cartography and the efficiency of viscous swimming. IEEE T. Robot. 33, 523535.CrossRefGoogle Scholar
Hayashi, R. & Takagi, D. 2020 Metachronal swimming with rigid arms near boundaries. Fluids 5, 24.CrossRefGoogle Scholar
Hewitt, D.R. & Balmforth, N.J. 2018 Viscoplastic slender-body theory. J. Fluid Mech. 856, 870897.CrossRefGoogle Scholar
Higdon, J.J.L. 1979 A hydrodynamic analysis of flagellar propulsion. J. Fluid Mech. 90, 685711.CrossRefGoogle Scholar
Hu, J., Yang, M., Gompper, G. & Winkler, R.G. 2015 Modelling the mechanics and hydrodynamics of swimming E. coli. Soft Matter 11, 78677876.CrossRefGoogle ScholarPubMed
Huang, H.-W., Uslu, F.E., Katsamba, P., Lauga, E., Sakar, M.S. & Nelson, B.J. 2019 Adaptive locomotion of artificial microswimmers. Sci. Adv. 5, eaau1532.CrossRefGoogle ScholarPubMed
Ishikawa, T., Simmonds, M.P. & Pedley, T.J. 2006 Hydrodynamic interaction of two swimming model micro-organisms. J. Fluid Mech. 568, 119160.CrossRefGoogle Scholar
Ishimoto, K. & Gaffney, E.A. 2013 Squirmer dynamics near a boundary. Phys. Rev. E 88, 062702.CrossRefGoogle Scholar
Ishimoto, K. & Gaffney, E.A. 2018 An elastohydrodynamical simulation study of filament and spermatozoan swimming driven by internal couples. IMA J. Appl. Math. 83, 655679.CrossRefGoogle Scholar
Johnson, R.E. 1979 An improved slender-body theory for Stokes flow. J. Fluid Mech. 99, 411431.CrossRefGoogle Scholar
Karani, H., Pradillo, G.E. & Vlahovska, P.M. 2019 Tuning the random walk of active colloids: from individual run-and-tumble to dynamic clustering. Phys. Rev. Lett. 123, 208002.CrossRefGoogle ScholarPubMed
Keaveny, E.E, Walker, S.W. & Shelley, M.J. 2013 Optimization of chiral structures for microscale propulsion. Nano Lett. 13, 531537.CrossRefGoogle ScholarPubMed
Keller, J.B. & Rubinow, S.I. 1976 Slender-body theory for slow viscous flow. J. Fluid Mech. 75, 705714.CrossRefGoogle Scholar
Kim, M. & Powers, T.R. 2005 Deformation of a helical filament by flow and electric or magnetic fields. Phys. Rev. E 71, 021914.CrossRefGoogle ScholarPubMed
Kim, S. & Karrila, S.J. 2005 Microhydrodynamics: Principles and Selected Applications. Courier Corporation.Google Scholar
Koens, L. & Lauga, E. 2016 a Rotation of slender swimmers in isotropic-drag media. Phys. Rev. E 93, 043125.CrossRefGoogle ScholarPubMed
Koens, L. & Lauga, E. 2016 b Slender-ribbon theory. Phys. Fluids 28, 013101.CrossRefGoogle Scholar
Koens, L. & Lauga, E. 2018 The boundary integral formulation of Stokes flows includes slender-body theory. J. Fluid Mech. 850, R1.CrossRefGoogle Scholar
Koens, L., Wang, W., Sitti, M. & Lauga, E. 2019 The near and far of a pair of magnetic capillary disks. Soft Matter 15, 14971507.CrossRefGoogle ScholarPubMed
Koens, L., Zhang, H., Moeller, M., Mourran, A. & Lauga, E. 2018 The swimming of a deforming helix. Eur. Phys. J. E 41, 119.CrossRefGoogle ScholarPubMed
Lauga, E. 2016 Bacterial hydrodynamics. Annu. Rev. Fluid Mech. 48, 105130.CrossRefGoogle Scholar
Lauga, E. & Powers, T.R. 2009 The hydrodynamics of swimming microorganisms. Rep. Prog. Phys. 72, 096601.CrossRefGoogle Scholar
Lighthill, J. 1976 Flagellar hydrodynamics: The John von Neumann Lecture, 1975. SIAM Rev. 18, 161230.CrossRefGoogle Scholar
Lighthill, M.J. 1952 On the squirming motion of nearly spherical deformable bodies through liquids at very small reynolds numbers. Commun. Pure Appl. Maths 5, 109118.CrossRefGoogle Scholar
Maggi, C., Saglimbeni, F., Dipalo, M., De Angelis, F. & Di Leonardo, R. 2015 Micromotors with asymmetric shape that efficiently convert light into work by thermocapillary effects. Nature Commun. 6, 7855.CrossRefGoogle ScholarPubMed
Man, Y., Koens, L. & Lauga, E. 2016 Hydrodynamic interactions between nearby slender filaments. Eur. Phys. Lett. 116, 24002.CrossRefGoogle Scholar
Man, Y., Ling, F. & Kanso, E. 2020 Cilia oscillations. Philos Trans. Roy. Soc. B 375, 20190157.CrossRefGoogle ScholarPubMed
Martinez, V.A., et al. 2020 A combined rheometry and imaging study of viscosity reduction in bacterial suspensions. Proc. Natl Acad. Sci. USA 117, 23262331.CrossRefGoogle ScholarPubMed
Omori, T. & Ishikawa, T. 2019 Swimming of spermatozoa in a Maxwell fluid. Micromachines 10, 78.CrossRefGoogle Scholar
Pedley, T.J., Brumley, D.R. & Goldstein, R.E. 2016 Squirmers with swirl: a model for Volvox swimming. J. Fluid Mech. 798, 165186.CrossRefGoogle Scholar
Perez Ipiña, E., Otte, S., Pontier-Bres, R., Czerucka, D. & Peruani, F. 2019 Bacteria display optimal transport near surfaces. Nature Phys. 15, 610615.CrossRefGoogle Scholar
Pozrikidis, C. 1992 Boundary Integral and Singularity Methods for Linearized Viscous Flow. Cambridge University Press.CrossRefGoogle Scholar
Purcell, E.M. 1977 Life at low reynolds number. Am. J. Phys. 45, 3.CrossRefGoogle Scholar
Quispe, J.E., Oulmas, A. & Regnier, S. 2019 Geometry optimization of helical swimming at low Reynolds number. In 2019 International Conference on Manipulation, Automation and Robotics at Small Scales, pp. 1–6. IEEE.CrossRefGoogle Scholar
Radford, J. & Burdick, J.W. 1998 Local motion planning for nonholonomic control systems evolving on principal bundles. In Proc. Math. Theory of Networks and Systems.Google Scholar
Ramasamy, S. & Hatton, R.L. 2016 Soap-bubble optimization of gaits. In 2016 IEEE 55th Conference on Decision and Control, pp. 1056–1062. IEEE.CrossRefGoogle Scholar
Ramasamy, S. & Hatton, R.L. 2017 Geometric gait optimization beyond two dimensions. In 2017 American Control Conference, pp. 642–648. IEEE.CrossRefGoogle Scholar
Ramasamy, S. & Hatton, R.L. 2019 The geometry of optimal gaits for drag-dominated kinematic systems. IEEE Trans. Robot. 35, 10141033.CrossRefGoogle Scholar
du Roure, O., Lindner, A., Nazockdast, E.N. & Shelley, M.J. 2019 Dynamics of flexible fibers in viscous flows and fluids. Annu. Rev. Fluid Mech. 51, 539572.CrossRefGoogle Scholar
Shapere, A. & Wilczek, F. 1987 Self-propulsion at low Reynolds number. Phys. Rev. Lett. 58, 2051.CrossRefGoogle ScholarPubMed
Shapere, A. & Wilczek, F. 1989 a Efficiencies of self-propulsion at low Reynolds number. J. Fluid Mech. 198, 587599.CrossRefGoogle Scholar
Shapere, A. & Wilczek, F. 1989 b Geometry of self-propulsion at low Reynolds number. J. Fluid Mech. 198, 557685.Google Scholar
Smith, D.J., Gaffney, E.A., Blake, J.R. & Kirkman-Brown, J.C. 2009 Human sperm accumulation near surfaces: a simulation study. J. Fluid Mech. 621, 289320.CrossRefGoogle Scholar
Taylor, G. 1951 Analysis of the swimming of microscopic organisms. Proc. R. Soc. A 209, 447461.Google Scholar
Turner, L., Ryu, W. & Berg, H. 2000 Real-time imaging of fluorescent flagellar filaments. J. Bacteriol. 182, 27932801.CrossRefGoogle ScholarPubMed
Valdés, R., Angeles, V., de la Calleja, E. & Zenit, R. 2019 Self-propulsion of a helical swimmer in granular matter. Phys. Rev. Fluids 4, 084302.CrossRefGoogle Scholar
Vizsnyiczai, G., Frangipane, G., Maggi, C., Saglimbeni, F., Bianchi, S. & Di Leonardo, R. 2017 Light controlled 3D micromotors powered by bacteria. Nature Commun. 8, 15974.CrossRefGoogle ScholarPubMed
Walker, B.J., Wheeler, R.J., Ishimoto, K. & Gaffney, E.A. 2019 Boundary behaviours of Leishmania Mexicana: a hydrodynamic simulation study. J. Theor. Biol. 462, 311320.CrossRefGoogle ScholarPubMed
Wiezel, O., Giraldi, L., DeSimone, A., Or, Y. & Alouges, F. 2018 Energy-optimal small-amplitude strokes for multi-link microswimmers: Purcell's loops and Taylor's waves reconciled. arXiv:1801.04687.Google Scholar
Wróbel, J.K., Lynch, S., Barrett, A., Fauci, L. & Cortez, R. 2016 Enhanced flagellar swimming through a compliant viscoelastic network in Stokes flow. J. Fluid Mech. 792, 775797.CrossRefGoogle Scholar
Zhang, L., Peyer, K.E. & Nelson, B.J. 2010 Artificial bacterial flagella for micromanipulation. Lab on a Chip 10, 2203–15.CrossRefGoogle ScholarPubMed