Hostname: page-component-848d4c4894-tn8tq Total loading time: 0 Render date: 2024-06-19T21:50:51.406Z Has data issue: false hasContentIssue false

Thermocapillary flow between grooved superhydrophobic surfaces: transverse temperature gradients

Published online by Cambridge University Press:  24 May 2019

Ehud Yariv*
Affiliation:
Department of Mathematics, Technion – Israel Institute of Technology, Haifa 32000, Israel
Darren Crowdy
Affiliation:
Department of Mathematics, Imperial College London, 180 Queen’s Gate, London SW7 2AZ, UK
*
Email address for correspondence: udi@technion.ac.il

Abstract

We consider the thermocapillary motion of a liquid layer which is bounded between two superhydrophobic surfaces, each made up of a periodic array of highly conducting solid slats, with flat bubbles trapped in the grooves between them. Following the recent analysis of the longitudinal problem (Yariv, J. Fluid Mech., vol. 855, 2018, pp. 574–594), we address here the transverse problem, where the macroscopic temperature gradient that drives the flow is applied perpendicular to the grooves, with the goal of calculating the volumetric flux between the two surfaces. We focus upon the situation where the slats separating the grooves are long relative to the groove-array period, for which case the temperature in the solid portions of the superhydrophobic plane is piecewise uniform. This scenario, which was investigated numerically by Baier et al. (Phys. Rev. E, vol. 82 (3), 2010, 037301), allows for a surprising analogy between the harmonic conjugate of the temperature field in the present problem and the unidirectional velocity in a comparable longitudinal pressure-driven flow problem over an interchanged boundary. The main body of the paper is concerned with the limit of deep channels, where the problem reduces to the calculation of the heat transport and flow about a single surface and the associated ‘slip’ velocity at large distance from that surface. Making use of Lorentz’s reciprocity, we obtain that velocity as a simple quadrature, providing the analogue to the expression obtained by Baier et al. (2010) in the comparable longitudinal problem. The rest of the paper is devoted to the diametric limit of shallow channels, which is analysed using a Hele-Shaw approximation, and the singular limit of small solid fractions, where we find a logarithmic scaling of the flux with the solid fraction. The latter two limits do not commute.

Type
JFM Papers
Copyright
© 2019 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

Baier, T., Steffes, C. & Hardt, S. 2010 Thermocapillary flow on superhydrophobic surfaces. Phys. Rev. E 82 (3), 037301.Google Scholar
Balasubramaniam, R. & Subramaniam, R. S. 1996 Thermocapillary bubble migration – thermal boundary layers for large Marangoni numbers. Intl J. Multiphase Flow 22 (3), 593612.Google Scholar
Balasubramaniam, R. & Subramanian, R. S. 2000 The migration of a drop in a uniform temperature gradient at large Marangoni numbers. Phys. Fluids 12, 733743.Google Scholar
Brown, J. W. & Churchill, R. V. 2003 Complex Variables and Applications. McGraw-Hill.Google Scholar
Choi, C.-H. & Kim, C.-J. 2006 Large slip of aqueous liquid flow over a nanoengineered superhydrophobic surface. Phys. Rev. Lett. 96 (6), 066001.Google Scholar
Crowdy, D. 2011 Frictional slip lengths and blockage coefficients. Phys. Fluids 23 (9), 091703.Google Scholar
Crowdy, D. 2017a Effect of shear thinning on superhydrophobic slip: perturbative corrections to the effective slip length. Phys. Rev. Fluids 2 (12), 124201.Google Scholar
Crowdy, D. G. 2017b Perturbation analysis of subphase gas and meniscus curvature effects for longitudinal flows over superhydrophobic surfaces. J. Fluid Mech. 822, 307326.Google Scholar
Feuillebois, F., Bazant, M. Z. & Vinogradova, O. I. 2009 Effective slip over superhydrophobic surfaces in thin channels. Phys. Rev. Lett. 102 (2), 026001.Google Scholar
Happel, J. & Brenner, H. 1965 Low Reynolds Number Hydrodynamics. Prentice-Hall.Google Scholar
Hasimoto, H. 1958 On the flow of a viscous fluid past a thin screen at small Reynolds numbers. J. Phys. Soc. Japan 13 (6), 633639.Google Scholar
Lauga, E. & Stone, H. A. 2003 Effective slip in pressure-driven Stokes flow. J. Fluid Mech. 489, 5577.Google Scholar
Lee, C., Choi, C.-H. & Kim, C.-J. 2008 Structured surfaces for a giant liquid slip. Phys. Rev. Lett. 101 (6), 064501.Google Scholar
Philip, J. R. 1972a Flows satisfying mixed no-slip and no-shear conditions. Z. Angew. Math. Phys. 23 (3), 353372.Google Scholar
Philip, J. R. 1972b Integral properties of flows satisfying mixed no-slip and no-shear conditions. Z. Angew. Math. Phys. 23 (6), 960968.Google Scholar
Pozrikidis, C. 1992 Boundary Integral and Singularity Methods for Linearized Viscous Flow. Cambridge University Press.Google Scholar
Quéré, D. 2005 Non-sticking drops. Rep. Prog. Phys. 68 (11), 24952532.Google Scholar
Quéré, D. 2008 Wetting and roughness. Annu. Rev. Mater. Res. 38 (1), 7199.Google Scholar
Schnitzer, O. & Yariv, E. 2017 Longitudinal pressure-driven flows between superhydrophobic grooved surfaces: large effective slip in the narrow-channel limit. Phys. Rev. Fluids 2 (7), 072101.Google Scholar
Subramanian, R. S. 1981 Slow migration of a gas bubble in a thermal gradient. AIChE J. 27, 646654.Google Scholar
Teo, C. J. & Khoo, B. C. 2009 Analysis of Stokes flow in microchannels with superhydrophobic surfaces containing a periodic array of micro-grooves. Microfluid. Nanofluid. 7 (3), 353382.Google Scholar
Yariv, E. 2017 Velocity amplification in pressure-driven flows between superhydrophobic gratings of small solid fraction. Soft Matt. 13, 62876292.Google Scholar
Yariv, E. 2018 Thermocapillary flow between longitudinally grooved superhydrophobic surfaces. J. Fluid Mech. 855, 574594.Google Scholar
Young, N. O., Goldstein, J. S. & Block, M. J. 1959 The motion of bubbles in a vertical temperature gradient. J. Fluid Mech. 6 (3), 350356.Google Scholar