Skip to main content
Log in

Mobility tensor of a sphere moving on a superhydrophobic wall: application to particle separation

  • Research Paper
  • Published:
Microfluidics and Nanofluidics Aims and scope Submit manuscript

Abstract

The paper addresses the hydrodynamic behavior of a sphere close to a micropatterned superhydrophobic surface described in terms of alternated no-slip and perfect-slip stripes. Physically, the perfect-slip stripes model the parallel grooves where a large gas cushion forms between fluid and solid wall, giving rise to slippage at the gas–liquid interface. The potential of the boundary element method in dealing with mixed no-slip/perfect-slip boundary conditions is exploited to systematically calculate the mobility tensor for different particle-to-wall relative positions and for different particle radii. The particle hydrodynamics is characterized by a nontrivial mobility field which presents a distinct near-wall behavior where the wall patterning directly affects the particle motion. In the far field, the effects of the wall pattern can be accurately represented via an effective description in terms of a homogeneous wall with a suitably defined apparent slippage. The trajectory of the sphere under the action of an external force is also described in some detail. A “resonant” regime is found when the frequency of the transversal component of the force matches a characteristic crossing frequency imposed by the wall pattern. It is found that under resonance, the particle undergoes a mean transversal drift. Since the resonance condition depends on the particle radius, the effect can in principle be used to conceive devices for particle sorting based on superhydrophobic surfaces.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

Notes

  1. The library is available under the terms of the GNU Lesser General Public License at http://dehesa.freeshell.org/BEMLIB/.

References

  • Asmolov ES, Belyaev AV, Vinogradova OI (2011) Drag force on a sphere moving toward an anisotropic superhydrophobic plane. Phys Rev E 84(2):026330

    Article  Google Scholar 

  • Batchelor G (1976) Brownian diffusion of particles with hydrodynamic interaction. J Fluid Mech 74(01):1

    Article  MATH  MathSciNet  Google Scholar 

  • Belyaev AV, Vinogradova OI (2010) Effective slip in pressure-driven flow past super-hydrophobic stripes. J Fluid Mech 652:489

    Article  MATH  Google Scholar 

  • Benzi R, Biferale L, Sbragaglia M, Succi S, Toschi F (2006) Mesoscopic two-phase model for describing apparent slip in micro-channel flows. EPL (Europhys Lett) 74:651

    Article  MathSciNet  Google Scholar 

  • Blake J (1971) A note on the image system for a Stokeslet in a no-slip boundary. In: Proc. Camb. Phil. Soc, vol 70. Cambridge University Press, Cambridge, pp 303–310

  • Bolognesi G, Cottin-Bizonne C, Guene EM, Teisseire J,  Pirat C (2013) A novel technique for simultaneous velocity and interface profile measurements on micro-structured surfaces. Soft Matter 9(7):2239–2244

    Google Scholar 

  • Bottiglione F, Carbone G (2012) Role of statistical properties of randomly rough surfaces in controlling superhydrophobicity. Langmuir 29(2):599

    Article  Google Scholar 

  • Boymelgreen AM, Miloh T (2011) A theoretical study of induced-charge dipolophoresis of ideally polarizable asymmetrically slipping Janus particles. Phys Fluids 23:072007

    Article  Google Scholar 

  • Brady JF, Bossis G (1988) Stokesian dynamics. Annu Rev Fluid Mech 20:111

    Article  Google Scholar 

  • Chinappi M, Casciola CM (2010) Intrinsic slip on hydrophobic self-assembled monolayer coatings. Phys Fluids 22:042003

    Article  Google Scholar 

  • Chinappi M, Melchionna S, Casciola CM, Succi S (2008) Mass flux through asymmetric nanopores: microscopic versus hydrodynamic motion. J Chem Phys 129:124717

    Article  Google Scholar 

  • Chinappi M, Gala F, Zollo G, Casciola CM (2011) Tilting angle and water slippage over hydrophobic coatings. Philos Trans R Soc A Math Phys Eng Sci 369(1945):2537

    Article  Google Scholar 

  • Cottin-Bizonne C, Barentin C, Charlaix É, Bocquet L, Barrat J (2004) Dynamics of simple liquids at heterogeneous surfaces: molecular-dynamics simulations and hydrodynamic description. Eur Phys J E: Soft Matter Biol Phys 15(4):427

    Article  Google Scholar 

  • Cottin-Bizonne C, Steinberger A, Cross B, Raccurt O, Charlaix E (2008) Nanohydrodynamics: The intrinsic flow boundary condition on smooth surfaces. Langmuir 24(4):1165

    Article  Google Scholar 

  • Gentili D, Chinappi M, Bolognesi G, Giacomello A, Casciola CM (2013) Water slippage on hydrophobic nanostructured surfaces: molecular dynamics results for different filling levels. Meccanica. doi:10.1007/s11012-013-9717-8

  • Giacomello A, Meloni S, Chinappi M, Casciola CM (2012) Cassie–Baxter and Wenzel states on a nanostructured surface: phase diagram, metastabilities, and transition mechanism by atomistic free energy calculations. Langmuir 28(29):10764

    Article  Google Scholar 

  • Giacomello A, Chinappi M, Meloni S, Casciola CM (2012) Metastable wetting on superhydrophobic surfaces: continuum and atomistic views of the Cassie–Baxter–Wenzel Transition. Phys Rev Lett 109(22):226102

    Article  Google Scholar 

  • Goldman A, Cox RG, Brenner H (1967) Slow viscous motion of a sphere parallel to a plane wall-I motion through a quiescent fluid. Chem Eng Sci 22(4):637

    Article  Google Scholar 

  • Happel JR, Brenner H (1965) Low Reynolds number hydrodynamics: with special applications to particulate media, vol 1. Springer, Berlin

    Google Scholar 

  • Huang D, Sendner C, Horinek D, Netz R, Bocquet L (2008) Water slippage versus contact angle: a quasiuniversal relationship. Phys Rev Lett 101(22):226101

    Article  Google Scholar 

  • Jeffrey D, Onishi Y (1984) Calculation of the resistance and mobility functions for two unequal rigid spheres in low-Reynolds-number flow. J Fluid Mech 139(1):261

    MATH  Google Scholar 

  • Kamrin K, Bazant MZ, Stone HA (2010) Effective slip boundary conditions for arbitrary periodic surfaces: the surface mobility tensor. J Fluid Mech 658:409

    Article  MATH  MathSciNet  Google Scholar 

  • Kim S, Karrila S (2005) Microhydrodynamics: principles and selected applications. Dover Publications, Mineola

    Google Scholar 

  • Landau LD (1987) Fluid mechanics: volume 6 (course of theoretical physics) Author: LD Landau, EM Lifshitz, Publisher: Bu (Butterworth-Heinemann)

  • Lauga E, Brenner M, Stone H (2007) Anisotropic flow in striped superhydrophobic channels. Handbook of experimental fluid mechanics

  • Lee C, Kim C (2011) Influence of surface hierarchy of superhydrophobic surfaces on liquid slip.. Langmuir: ACS J Surf Colloids 27(7):4243

    Article  Google Scholar 

  • Li Z (2009) Critical particle size where the Stokes–Einstein relation breaks down. Phys Rev E 80(6):061204

    Article  Google Scholar 

  • Ng C, Wang C (2010) Apparent slip arising from Stokes shear flow over a bidimensional patterned surface. Microfluid Nanofluidics 8(3):361

    Article  Google Scholar 

  • Nosonovsky M, Bhushan B (2009) Superhydrophobic surfaces and emerging applications: non-adhesion, energy, green engineering. Curr Opin Colloid Interface Sci 14(4):270

    Article  Google Scholar 

  • Pan Y, Bhushan B (2012) Role of surface charge on boundary slip in fluid flow. J Colloid Interface Sci

  • Philip J (1972) Flows satisfying mixed no-slip and no-shear conditions. Zeitschrift für Angewandte Mathematik und Physik (ZAMP) 23(3):353

    Article  MATH  MathSciNet  Google Scholar 

  • Pozrikidis C (1992) Boundary integral and singularity methods for linearized viscous flow, vol 8. Cambridge University Press, Cambridge

  • Pozrikidis C (2002) A practical guide to boundary element methods with the software library BEMLIB. CRC, Boca Raton, FL

    Book  MATH  Google Scholar 

  • Sbragaglia M, Prosperetti A (2007) A note on the effective slip properties for microchannel flows with ultrahydrophobic surfaces. Phys Fluids 19:043603

    Article  Google Scholar 

  • Shum H, Gaffney E (2012) The effects of flagellar hook compliance on motility of monotrichous bacteria: a modeling study. Phys Fluids 24:061901

    Article  Google Scholar 

  • Steinberger A, Cottin-Bizonne C, Kleimann P, Charlaix E (2007) High friction on a bubble mattress. Nat Mater 6(9):665

    Article  Google Scholar 

  • Teo C, Khoo B (2010) Flow past superhydrophobic surfaces containing longitudinal grooves: effects of interface curvature. Microfluid Nanofluidics 9(2):499

    Article  Google Scholar 

  • Vinogradova OI (1995) Drainage of a thin liquid film confined between hydrophobic surfaces. Langmuir 11(6):2213

    Article  Google Scholar 

  • Vinogradova OI, Belyaev AV (2011) Wetting, roughness and flow boundary conditions. J Phys: Condens Matter 23(18):184104

    Google Scholar 

  • Ybert C, Barentin C, Cottin-Bizonne C, Joseph P, Bocquet L (2007) Achieving large slip with superhydrophobic surfaces: Scaling laws for generic geometries. Phys Fluids 19:123601

    Article  Google Scholar 

  • Zhang R, Koplik J (2012) Separation of nanoparticles by flow past a patterned substrate. Phys Rev E 85(2):026314

    Article  Google Scholar 

  • Zhang H, Zhang Z, Ye H (2012) Molecular dynamics-based prediction of boundary slip of fluids in nanochannels. Microfluid Nanofluidics 1–9

  • Zhou J, Belyaev AV, Schmid F, Vinogradova OI (2012) Anisotropic flow in striped superhydrophobic channels. J Chem Phys 136(19):194706. doi:10.1063/1.4718834. http://link.aip.org/link/?JCP/136/194706/1

    Google Scholar 

  • Zhu L, Neto C, Attard P (2012) Reconciling slip measurements in symmetric and asymmetric systems. Langmuir 28(20):7768

    Article  Google Scholar 

  • Zhu L, Lauga E, Brandt L (2013) Low-Reynolds-number swimming in a capillary tube. J Fluid Mech 726

Download references

Acknowledgements

Computer resources were made available by CASPUR under HPC grant 2012.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to C. M. Casciola.

Appendix: Boundary integral formulation

Appendix: Boundary integral formulation

In this section, the application of the boundary element method (BEM) to the present geometrical configuration is shortly described (see, e.g., Pozrikidis (1992); Kim and Karrila (2005) for further details).

Due to linearity, the constant coefficient Stokes problem, Eq. (2), can be recast into a boundary integral representation formula

$$\begin{aligned} E({\boldsymbol \xi}) u_j({\boldsymbol \xi}) &= \frac{1}{8 \pi} \int\nolimits_{\partial B} t_i({\bf x}) G_{ij}({\bf x},{\boldsymbol \xi}) {\hbox{d}}S_{B} \\ &\quad-\frac{1}{8 \pi } \int\limits_{W_{NS}} t_i({\bf x}) G_{ij}({\bf x},{\boldsymbol \xi}) dS_{NS} \\ &\quad-\frac{1}{8 \pi } \int\limits_{ W_{PS}} t_i({\bf x}) G_{ij}({\bf x},{\boldsymbol \xi}) {\hbox{d}}S_{PS} \\ &\quad- \frac{1}{8\pi} \int\limits_{\partial B} u_i({\bf x}) {\mathcal T}_{ijk} ({\bf x},{\boldsymbol \xi}) n_k({\bf x}) {\hbox{d}}S_{B} \\ &\quad- \frac{1}{8\pi} \int\limits_{W_{NS}} u_i({\bf x}) {\mathcal T}_{ijk} ({\bf x},{\boldsymbol \xi}) n_k({\bf x}) {\hbox{d}}S_{NS} \\ &\quad- \frac{1}{8\pi} \int\limits_{W_{PS}} u_i({\bf x}) {\mathcal T}_{ijk} ({\bf x},{\boldsymbol \xi}) n_k({\bf x}) {\hbox{d}}S_{PS}. \end{aligned}$$
(26)

In Eq. (3) the boundary \(\partial \varOmega\) of the fluid domain is explicitly decomposed into three parts: the particle surface ∂B, the no-slip stripes on the patterned wall, collectively denoted W NS , and the complementary part of the wall with the perfect-slip stripes W PS . The contributions arising from the portion of the boundary at infinity (not included in the present formulation where the fluid is assumed to be at rest) can be easily incorporated under suitable assumption on the asymptotic behavior of the field. The effects of body forces such as gravity and electric fields can be accounted for by a convolution integral extended to the fluid domain between the force and the free-space Green’s tensor (see below). In representation (26), t i and \(u_i, i=1,\ldots,3\), are the Cartesian components of the surface stress and velocity, respectively. The free-space Green’s function (the so-called steady Stokeslet) is defined as

$$G_{ij}({\bf r})=\left(\frac{\delta_{ij}}{r} + \frac{r_i r_j}{r^3}\right) ,$$
(27)

and the associated stress tensor is

$${\mathcal T}_{ijk}({\bf r}) =-6 \frac{r_i r_j r_k }{r^5}.$$
(28)

In the above expression, G ij (r) provides the contribution to the jth velocity component at \({\boldsymbol \xi}\) due to a concentrated force acting in the ith direction at x. The associated Green’s stress tensor, as always, should be contracted with the outward unit normal n k (x) to the boundary \(\partial \varOmega\) in order to provide the effect on the jth velocity component at \({\boldsymbol \xi}\) of the ith boundary velocity at x. The vector r is defined as \({\bf r}={\bf x}-{\boldsymbol \xi}\) with \(r=\sqrt{r_k r_k}\) its modulus. In Eq. (3) \(E({\boldsymbol \xi})=1\) when \({\boldsymbol \xi}\in\varOmega\) and \(E({\boldsymbol \xi})=1/2\) for \({\boldsymbol \xi}\in\partial\varOmega\) (the existence of a regular tangent plane is assumed throughout). When \({\boldsymbol \xi}\in\partial\varOmega\), representation (3) becomes a boundary integral equation where the unknowns can either be the three stress vector components t i , the three velocity components u i , or a combination thereof, depending on the boundary conditions assigned on the specific portion of boundary. This approach allows us to discretize only the boundary surfaces of the flow domain instead of considering the entire volume. This results in: (1) a substantial reduction in the number of unknowns; (2) the possibility to easily specify different kinds of boundary condition on different surface patches; and (3) the simple update of the geometry when dealing with time-dependent configurations. Once the boundary \(\partial\varOmega\) is discretized into panels, Eq. (3) is recast into an algebraic linear system whose solution can be achieved by standard linear algebra packages. In simple cases, part of the boundary can be accounted for by symmetry, as it happens for a flat homogeneous wall.

In this paper, given the generality of the boundary condition to be used at the wall (either patterned perfect/no-slip stripes or effective slip Navier-like boundary conditions), the complete formulation of the boundary integral problem based on the free-space Green’s function has been retained, with the use of the wall Green’s function demanded of providing reference results for accuracy tests.

Finally, a few more words may be useful concerning the specific boundary conditions used in the paper. On the perfect-slip boundary patches, W PS , the normal velocity vanishes u  = 0 due to impermeability, while the tangential velocity u (two Cartesian components) is unknown. Moreover, the tangential stress t vanishes by perfect slip such that the stress is aligned to the normal, \(t_i=-\varPhi({\bf x}) n_i\), with \(\varPhi\) representing a further scalar unknown. On the no-slip surfaces, W NS and ∂B, velocities are completely assigned while stresses are unknown. Concerning the partial-slip condition used as an effective model of the stripe pattern, zero normal velocity u  = 0 at the wall is implied, while tangential velocities and stresses are coupled by the Navier condition (Lauga et al. 2007),

$${\bf u}_{\parallel} = {\boldsymbol \ell}_{\bf s} {\bf n} \cdot ( {\boldsymbol \nabla}{{\bf u}} + ({\boldsymbol \nabla}{{\bf u}})^T) \cdot ({\bf 1}-{\bf n} \otimes {\bf n}) .$$
(29)

Here \({\boldsymbol \ell}_{\bf s}\) is a 2 × 2 symmetric (Kamrin et al. 2010) tensor describing the directionally dependent slip length. In the present case, this tensor is diagonalized when expressed in stripe-parallel and stripe-normal Cartesian coordinates. In this case the two diagonal entries are different as a consequence of the orientation of the stripe pattern. For a flat wall, Eq. (29) is rewritten as

$${\bf u}_{\parallel}={\boldsymbol \ell}_{\bf s} \cdot {\bf t_{\parallel}} ,$$
(30)

that is the vectorial form of Eq. (7), the two nonzero components of \({\boldsymbol \ell}_{\bf s}\) being reported in Eq. 8 (Philip 1972; Ng and Wang 2010). The boundary integral Eq. (3) supplemented with Eq. (30) provides a closed system that after inversion determines all the unknowns involved in the problem.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Pimponi, D., Chinappi, M., Gualtieri, P. et al. Mobility tensor of a sphere moving on a superhydrophobic wall: application to particle separation. Microfluid Nanofluid 16, 571–585 (2014). https://doi.org/10.1007/s10404-013-1243-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10404-013-1243-4

Keywords

Navigation