Skip to main content

Advertisement

Log in

Energy conversion by surface-tension-driven charge separation

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

Abstract

In this work, the shear-induced electrokinetic streaming potential present in free-surface electrolytic flows subjected to a gradient in surface tension is assessed. Firstly, for a Couette flow with fully resolved electric double layer (EDL), the streaming potential per surface stress as a function of the Debye parameter and ζ-potential is analyzed. By contrast to the Smoluchowski limit in pressure-driven channel flow, the shear-induced streaming potential vanishes for increasing Debye parameter (infinitely thin EDL), unless the free surface contains (induced) surface charge or the flow at the charged, solid wall is permitted to slip. Secondly, a technical realization of surface-tension-induced streaming is proposed, with surface stress acting on the free (slipping) surfaces of a micro-structured, superhydrophobic wall. The streaming potential is analyzed with respect to the slip parameter and surface charge. Finally, the surface tension is assumed to vary with temperature (thermocapillarity) or with surfactant concentration (destillocapillarity). The maximal thermal efficiency is derived and compared to the Carnot efficiency. For large thermal Marangoni number, the efficiency is severely limited by the large heat capacity of aqueous solvents. By contrast, destillocapillary flows may reach conversion efficiencies similar to pressure-driven flow.

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

Similar content being viewed by others

Notes

  1. Since only half of the channel height is considered, the corresponding total external resistance for the full channel, i.e., two half-systems in parallel, is \(R/2\).

References

  • Bell LE (2008) Cooling, heating, generating power, and recovering waste heat with thermoelectric systems. Science 321:1457–1461

    Article  Google Scholar 

  • Biswas K, He J, Blum ID, Wu CI, Hogan TP, Seidman DN, Dravis VP, Kanatzidis MG (2012) High-performance bulk thermoelectrics with all-scale hierarchical architectures. Nature 489:414–418

    Article  Google Scholar 

  • Chen H, Goswami DY, Stefanakos EK (2010) A review of thermodynamic cycles and working fluids for the conversion of low-grade heat. Renew Sustain Energy Rev 14:3059–3067

    Article  Google Scholar 

  • de Surgy GN, Chabrerie JP, Denoux O, Wesfreid JE (1993) Linear growth of instabilities on a liquid metal under normal electric field. J Phys II France 3:1201–1225

    Article  Google Scholar 

  • Dietzel M, Hardt S (2012) Streaming potential of an electrolyte in a microchannel with a lateral temperature gradient, In: Proceedings of 3rd conference on microfluidics, Heidelberg 39 (\(\mu\)Flu12-108), pp 1–9

  • Dietzel M, Hardt S (2015) Flow and streaming potential of an electrolyte in a channel with an axial temperature gradient. J Fluid Mech (in preparation)

  • Drelich J, Chibowksi E, Meng DD, Terpilowski K (2011) Hydrophilic and superhydrophilic surfaces and materials. Soft Matter 7:9804

    Article  Google Scholar 

  • Duffin AM, Saykally RJ (2008) Electrokinetic power generation from liquid water microjets. J Phys Chem C 112:17018–17022

    Article  Google Scholar 

  • Gao Y, Wong TN, Yang C, Ooi KT (2005) Transient two-liquid electroosmotic flow with electric charges at the interface. Colloids Surf A 266(1):117–128

    Article  Google Scholar 

  • Gradshteyn IS, Ryzhik IM (2007) Table of integrals, series, and products, 7th edn. Academic Press, New York

    MATH  Google Scholar 

  • Grosu FP, Bologa MK (2010) Thermoelectrohydrodynamic methods of energy conversion. Surf Eng Appl Electrochem 46(6):582–588

    Article  Google Scholar 

  • Joo SW (2008) A new hydrodynamic instability in ultra-thin film flows induced by electro-osmosis. J Mech Sci Technol 22:382–386

    Article  Google Scholar 

  • Karakashev SI, Teskov R (2011) Electro-Marangoni effect in thin liquid films. Langmuir 27:2265–2270

    Article  Google Scholar 

  • Lawrentjew MA, Schabat BW (1967) Methoden der komplexen Funktionentheorie. VEB Deutscher Verlag der Wissenschaften, Berlin

    MATH  Google Scholar 

  • Lee JSH, Barbulovic-Nad I, Wu Z, Xuan X, Li D (2006) Electrokinetic flow in a free surface-guided microchannel. J Appl Phys 99:054905

    Article  Google Scholar 

  • Marín AG, van Hoeve W, García-Sánchez P, Shui L, Xie Y, Fontelos MA, Eijkel JCT, van den Berg A, Lohse D (2013) The microfluidic Kelvin water dropper. Lab Chip 13(23):4489–4682

    Article  Google Scholar 

  • Mayur M, Amiroudine S, Lasseux D (2012) Free-surface instability in electro-osmotic flows of ultrathin liquid films. Phys Rev E 85:046301

    Article  Google Scholar 

  • Melcher JR, Taylor GI (1969) Electrohydrodynamics: a review of the role of interfacial stresses. Annu Rev Fluid Mech 1:111–146

    Article  Google Scholar 

  • Mugele F, Baret JC (2005) Electrowetting: from basics to applications. J Phys Condens Mater 17:R705–R774

    Article  Google Scholar 

  • Oelkers EH, Helgeson HC (1989) Calculation of the transport properties of aqueous species at pressures to 5 kb and temperatures to 1000 \(^\circ \text{ C }\). J Sol Chem 18(7):601–640

    Article  Google Scholar 

  • Oh JM, Manukyan G, den Ende DV, Mugele F (2011) Electric-field-driven instabilities on superhydrophobic surfaces. Europhys Lett 93(5):56001

    Article  Google Scholar 

  • Philip JR (1972) Flows satisfying mixed no-slip and no-shear conditions. Z Angew Math Phys 23:353–372

    Article  MathSciNet  MATH  Google Scholar 

  • Philip JR (1972) Integral properties of flows satisfying mixed no-slip and no-shear conditions. Z Angew Math Phys 23:960–968

    Article  MathSciNet  MATH  Google Scholar 

  • Qian S, Joo SW, Jiang Y, Cheney MA (2009) Free-surface problems in electrokinetic micro- and nanofluidics. Mech Res Commun 36:82–91

    Article  MATH  Google Scholar 

  • Riffat SB, Ma X (2003) Thermoelectrics: a review of present and potential applications. Appl Therm Eng 23:913–935

    Article  Google Scholar 

  • Russel WB, Saville DA, Schowalter WR (1989) In: Batchelor GK (ed) Colloidal dispersion. Cambridge University Press, Cambridge

  • Saha BB, Koyama S, Kashiwagi T, Akisawa A, Ng KC, Chua HT (2003) Waste heat driven dual-mode, multi-stage, multi-bed regenerative adsorption system. Int J Refrig 26:749–757

    Article  Google Scholar 

  • Salata OV (2005) Tools of nanotechnology: electrospray. Curr Nanosci 1:25–33

    Article  Google Scholar 

  • Saqr KM, Musa MN (2009) Critical review of thermoelectrics in modern power generation applications. Therm Sci 13(3):165–174

    Article  Google Scholar 

  • Schönecker C, Baier T, Hardt S (2014) Influence of the enclosed fluid on the flow over a microstructured surface in the Cassie state. J Fluid Mech 740:168–195

    Article  MathSciNet  Google Scholar 

  • Seshadri G, Baier T (2013) Effect of electro-osmotic flow on energy conversion on superhydrophobic surfaces. Phys Fluids 25:042002

    Article  Google Scholar 

  • Shakouri A (2011) Recent developments in semiconductor thermoelectric physics and materials. Ann Rev Mater Res 41:399–431

    Article  Google Scholar 

  • Squires TM (2008) Electrokinetic flows over inhomogeneously slipping surfaces. Phys Fluids 20(9):092105

    Article  Google Scholar 

  • Steffes C, Baier T, Hardt S (2011) Enabling the enhancement of electroosmotic flow over superhydrophobic surfaces by induced charges. Colloids Surf A 376(1):85–88

    Article  Google Scholar 

  • Taylor GI, McEwan AD (1965) The stability of a horizontal fluid interface in a vertical electric field. J Fluid Mech 22:1–15

    Article  MATH  Google Scholar 

  • Tchanche BF, Lambrinos G, Frangoudakis A, Papadakis G (2011) Low-grade heat conversion into power using organic Rankine cycles—a review of various applications. Renew Sustain Energy Rev 15:3963–3979

    Article  Google Scholar 

  • Thomson W (1867) On a self-acting apparatus for multiplying and maintaining electric charges, with applications to illustrate the voltaic theory. Proc R Soc Lond 16(67):67–72

    Article  Google Scholar 

  • Tsekov R, Ivanova DS, Slavchov R, Radoev B, Manev ED, Nguyen AV, Karakashev SI (2010) Streaming potential effect on the drainage of thin liquid films stabilized by ionic surfactants. Langmuir 27(7):4703–4708

    Article  Google Scholar 

  • van der Heyden FHJ, Stein D, Dekker C (2005) Streaming currents in a single nanofluidic channel. Phys Rev Lett 95:116104

    Article  Google Scholar 

  • van der Heyden HJ, Bonthuis DJ, Stein D, Meyer C, Dekker C (2006) Electrokinetic energy convertion in nanofluidic channels. Nano Lett 6(10):2232–2237

    Article  Google Scholar 

  • Watanabe T, Nakajima A, Wang R, Minabe M, Koizumi S, Fujishima A, Hashimoto K (1999) Photocatalytic activity and photoinduced hydrophilicity of titanium dioxide coated glass. Thin Solid Films 351:260–263

    Article  Google Scholar 

  • Xuan X, Li D (2005) Thermodynamic analysis of electrokinetic energy conversion. J Power Sources 156:677–684

    Article  Google Scholar 

  • Yang J, Lu F, Kostiuk LW, Kwok DY (2003) Electrokinetic microchannel battery by means of electrokinetic and microfluidic phenomena. J Micromech Microeng 13:963–970

    Article  Google Scholar 

  • Zhao H (2011) Streaming potential generated by a pressure-driven flow over superhydrophobic stripes. Phys Fluids 23:022003

    Article  Google Scholar 

Download references

Acknowledgments

This work was in part supported by the German Research Foundation (DFG) through Cluster of Excellence 259, ‘Center of Smart Interfaces.’ Steffen Hardt is acknowledged for fruitful discussion.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mathias Dietzel.

Appendix: Solution details of the Laplace equation

Appendix: Solution details of the Laplace equation

The velocity field between parallel plates containing periodic patches of no-slip and constant shear regions was estimated using the result by Philip, Eq. (30), for such a flow in an infinite half-plane, \(y\ge 0\). No analytical result is known for a finite plate separation. However, the range of validity of this approximation can be assessed numerically. For this, the Laplace equation, \(\nabla ^2 w =0\), was discretized using the finite element method as implemented in the commercial code Comsol Multiphysics. By symmetry, the computational domain can be restricted to a unit cell indicated by the gray area in Fig. 4. The boundary conditions are \(w(x,0) = 0\) for \(0<x<(1-a)W\) (at the solid wall) and \(\eta \partial _y w(x,0) = -\tau\) for \((1-a)W<x<W\) (at the constant shear surface). On all other boundaries symmetry conditions apply, i.e., \((\mathbf {n} \cdot \mathbf {\nabla }) w = 0\) with \(\mathbf {n}\) being the outward normal at the boundary. The average velocity, \((HW)^{-1}\int _0^H\!\!\int _0^W\!\! w\,{\mathrm{d}}x\,{\mathrm{d}}y\), normalized with the analytic value corresponding to Philip’s solution, \(\beta _\Vert W \tau /\eta\), is tabulated in Table 1 for different values of the free-surface fraction, \(a=B/W\), and aspect ratio, \(H/W\). It is evident from the table that for \(H/W=1\) the numerically obtained results deviate by only \(\mathcal{O}(10^{-3})\) from the corresponding analytical result and even for \(H/W=0.75\) the agreement is \(\mathcal{O}(10^{-2})\).

Table 1 Numerical values for the normalized average velocity, \((\beta _\Vert HW^2 \tau /\eta )^{-1}\int _0^H\!\!\int _0^W\!\! w\,{\mathrm{d}}x\,{\mathrm{d}}y\), for different values of the free-surface fraction, \(a=B/W\), and aspect ratio, \(H/W\), as obtained using a finite element discretization
Fig. 6
figure 6

Sketch of the integration contour, \(\gamma\). Due to periodicity, the integration along sections (2) and (4) cancel

For the analysis presented in the main text, flow rates and line averages of the velocity field are needed. Normalized with the length or area of the integration region, these averages turn out to be identical, attesting the relevance of Table 1. In fact, even for finite H, one can show that

$$\begin{aligned} W^{-1}\int _0^W\! w(x,y_0)dx=W^{-1}\int _0^W\! w(x,y_1){\mathrm{d}}x \end{aligned}$$
(68)

for any \(0\le y_0 < y_1 \le H\); thus, line averages and the flow rate are inherently linked. A sketch of a proof of this relation is as follows: Since \(w\) is harmonic, \(\nabla ^2 w=0\), it is the imaginary part of a holomorphic function \(f(x+iy)=v(x,y)+iw(x,y)\) with \(v,w:\mathbb {R}^2\rightarrow \mathbb {R}\), (Lawrentjew and Schabat 1967). By Cauchy’s integral theorem, \({\text {Im}}[\oint _\gamma f({\xi })d\xi ]=0\) for any closed path \(\gamma\). Choose \(\gamma\) as the rectangle with vertices \(\xi =iy_0\), \(2W+iy_0\), \(2W+iy_1\), \(iy_1\) as shown in Fig. 6. Due to the Cauchy–Riemann conditions, \(v(x,H)\) is constant since \(\partial _x v|_{y=H} =-\partial _y w|_{y=H}=0\); similarly, \(v(0,y)\) and \(v(2W,y)\) are constant since \(\partial _x w|_{x=0}=0=\partial _x w|_{x=2W}\). Thus, not only \(w(x,y)=w(x+2W,y)\) is periodic in \(x\), but so is \(v\) and thus \(f\). The line integrals on the legs with constant x thus cancel, which completes the proof.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pini, C., Baier, T. & Dietzel, M. Energy conversion by surface-tension-driven charge separation. Microfluid Nanofluid 19, 721–735 (2015). https://doi.org/10.1007/s10404-015-1597-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10404-015-1597-x

Keywords

Navigation