journal of colloid and interface sciencepaginas.fe.up.pt/~fpinho/pdfs/jcis_eo-pump_2013.pdf · this...

10
Analytical solution of two-fluid electro-osmotic flows of viscoelastic fluids A.M. Afonso a,, M.A. Alves a , F.T. Pinho b a Departamento de Engenharia Quı ´mica, Centro de Estudos de Fenómenos de Transporte, Faculdade de Engenharia da Universidade do Porto, Rua Dr. Roberto Frias s/n, 4200-465 Porto, Portugal b Centro de Estudos de Fenómenos de Transporte, Faculdade de Engenharia da Universidade do Porto, Rua Dr. Roberto Frias s/n, 4200-465 Porto, Portugal article info Article history: Received 4 November 2012 Accepted 5 December 2012 Available online 22 December 2012 Keywords: Electro-osmotic flows Two-fluid pump Viscoelastic fluids abstract This paper presents an analytical model that describes a two-fluid electro-osmotic flow of stratified fluids with Newtonian or viscoelastic rheological behavior. This is the principle of operation of an electro-osmo- tic two-fluid pump as proposed by Brask et al. [Tech. Proc. Nanotech., 1, 190–193, 2003], in which an elec- trically non-conducting fluid is transported by the interfacial dragging viscous force of a conducting fluid that is driven by electro-osmosis. The electric potential in the conducting fluid and the analytical steady flow solution of the two-fluid electro-osmotic stratified flow in a planar microchannel are presented by assuming a planar interface between the two immiscible fluids with Newtonian or viscoelastic rheolog- ical behavior. The effects of fluid rheology, shear viscosity ratio, holdup and interfacial zeta potential are analyzed to show the viability of this technique, where an enhancement of the flow rate is observed as the shear-thinning effects are increased. Ó 2012 Elsevier Inc. All rights reserved. 1. Introduction Electro-osmotic flows (EOFs) in microfluidic devices have been studied extensively over the past decade [1–5], because they en- able precise liquid manipulation and are easily miniaturized to nanosized systems. The major applications of electro-osmotic (EO) pumps are in micro flow injection analysis, microfluidic liquid chromatography systems, microreactors, microenergy systems, and microelectronic cooling systems. Fluid pumps are important elements in such microchannel networks, and promising candi- dates for miniaturization are electro-hydrodynamic pumps using ion-dragging effects via the so-called electro-osmosis [6] due to the inherent simplicity in producing small-sized pumps with these techniques. A comprehensive review on electrokinetic pumps has been recently published by Wang et al. [5]. Some of the above mentioned studies focused on the transport of single-phase fluids with high electrical conductivity, for which a classical EO pump is needed. An overview of fabrication methods and working principles for such systems was presented by Zeng et al. [7]. For nonpolar fluids, such as oil, traditional EOF pumping cannot be used, due to the low fluid conductivity [2].To overcome this limitation, Brask et al. [1] proposed an alternative construction that allows the use of EOF as a driving mechanism, using an elec- trolyte fluid with high conductivity to drag the low conductivity nonpolar fluid. Their study [1] analyzed the performance of the pump by equivalent circuit theory and computational fluid dy- namic simulations. The theoretical study of electro-osmotic flows of non-Newto- nian fluids is recent and has been mainly focused in simple inelas- tic fluid models, such as the power-law, due to the inherent analytical difficulties introduced by more complex constitutive equations. Examples are the recent works of Das and Chakraborty [8] and Chakraborty [9], who presented explicit relationships for velocity, temperature, and concentration distributions in electro- osmotic microchannel flows of non-Newtonian bio-fluids de- scribed by the power-law model. Other purely viscous models were analytically investigated by Berli and Olivares [10], who considered the existence of a small wall layer depleted of additives and behaving as a Newtonian fluid (the skimming layer), under the combined action of pressure and electrical fields, thus restricting the non-Newtonian behavior to the electrically neutral region out- side the electrical double layer (EDL). More recently, these studies were extended to viscoelastic fluids by Afonso et al. [11], who pre- sented analytical solutions for channel and pipe flows of viscoelas- tic fluids under the mixed influence of electrokinetic and pressure forces, using two constitutive models: the Phan-Thien and Tanner (PTT) model [12], with linear kernel for the stress coefficient func- tion and zero second normal stress difference [13], and the Finitely Extensible Non-linear Elastic dumbbell model with a Peterlin approximation for the average spring force (FENE-P model, cf. Bird et al. [14]). An earlier investigation with the PTT model by Park and Lee [15] was concerned with EOF in a square duct for which a numerical method was used. The analysis of Afonso et al. [11] was restricted to cases with small electric double layers, where the distance between the walls of a microfluidic device is at least 0021-9797/$ - see front matter Ó 2012 Elsevier Inc. All rights reserved. http://dx.doi.org/10.1016/j.jcis.2012.12.013 Corresponding author. E-mail addresses: [email protected] (A.M. Afonso), [email protected] (M.A. Alves), [email protected] (F.T. Pinho). Journal of Colloid and Interface Science 395 (2013) 277–286 Contents lists available at SciVerse ScienceDirect Journal of Colloid and Interface Science www.elsevier.com/locate/jcis

Upload: others

Post on 25-Jul-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

Journal of Colloid and Interface Science 395 (2013) 277–286

Contents lists available at SciVerse ScienceDirect

Journal of Colloid and Interface Science

www.elsevier .com/locate / jc is

Analytical solution of two-fluid electro-osmotic flows of viscoelastic fluids

A.M. Afonso a,⇑, M.A. Alves a, F.T. Pinho b

a Departamento de Engenharia Quı́mica, Centro de Estudos de Fenómenos de Transporte, Faculdade de Engenharia da Universidade do Porto, Rua Dr. Roberto Frias s/n, 4200-465Porto, Portugalb Centro de Estudos de Fenómenos de Transporte, Faculdade de Engenharia da Universidade do Porto, Rua Dr. Roberto Frias s/n, 4200-465 Porto, Portugal

a r t i c l e i n f o a b s t r a c t

Article history:Received 4 November 2012Accepted 5 December 2012Available online 22 December 2012

Keywords:Electro-osmotic flowsTwo-fluid pumpViscoelastic fluids

0021-9797/$ - see front matter � 2012 Elsevier Inc. Ahttp://dx.doi.org/10.1016/j.jcis.2012.12.013

⇑ Corresponding author.E-mail addresses: [email protected] (A.M. Afons

Alves), [email protected] (F.T. Pinho).

This paper presents an analytical model that describes a two-fluid electro-osmotic flow of stratified fluidswith Newtonian or viscoelastic rheological behavior. This is the principle of operation of an electro-osmo-tic two-fluid pump as proposed by Brask et al. [Tech. Proc. Nanotech., 1, 190–193, 2003], in which an elec-trically non-conducting fluid is transported by the interfacial dragging viscous force of a conducting fluidthat is driven by electro-osmosis. The electric potential in the conducting fluid and the analytical steadyflow solution of the two-fluid electro-osmotic stratified flow in a planar microchannel are presented byassuming a planar interface between the two immiscible fluids with Newtonian or viscoelastic rheolog-ical behavior. The effects of fluid rheology, shear viscosity ratio, holdup and interfacial zeta potential areanalyzed to show the viability of this technique, where an enhancement of the flow rate is observed asthe shear-thinning effects are increased.

� 2012 Elsevier Inc. All rights reserved.

1. Introduction pump by equivalent circuit theory and computational fluid dy-

Electro-osmotic flows (EOFs) in microfluidic devices have beenstudied extensively over the past decade [1–5], because they en-able precise liquid manipulation and are easily miniaturized tonanosized systems. The major applications of electro-osmotic(EO) pumps are in micro flow injection analysis, microfluidic liquidchromatography systems, microreactors, microenergy systems,and microelectronic cooling systems. Fluid pumps are importantelements in such microchannel networks, and promising candi-dates for miniaturization are electro-hydrodynamic pumps usingion-dragging effects via the so-called electro-osmosis [6] due tothe inherent simplicity in producing small-sized pumps with thesetechniques. A comprehensive review on electrokinetic pumps hasbeen recently published by Wang et al. [5].

Some of the above mentioned studies focused on the transportof single-phase fluids with high electrical conductivity, for which aclassical EO pump is needed. An overview of fabrication methodsand working principles for such systems was presented by Zenget al. [7]. For nonpolar fluids, such as oil, traditional EOF pumpingcannot be used, due to the low fluid conductivity [2].To overcomethis limitation, Brask et al. [1] proposed an alternative constructionthat allows the use of EOF as a driving mechanism, using an elec-trolyte fluid with high conductivity to drag the low conductivitynonpolar fluid. Their study [1] analyzed the performance of the

ll rights reserved.

o), [email protected] (M.A.

namic simulations.The theoretical study of electro-osmotic flows of non-Newto-

nian fluids is recent and has been mainly focused in simple inelas-tic fluid models, such as the power-law, due to the inherentanalytical difficulties introduced by more complex constitutiveequations. Examples are the recent works of Das and Chakraborty[8] and Chakraborty [9], who presented explicit relationships forvelocity, temperature, and concentration distributions in electro-osmotic microchannel flows of non-Newtonian bio-fluids de-scribed by the power-law model. Other purely viscous modelswere analytically investigated by Berli and Olivares [10], whoconsidered the existence of a small wall layer depleted of additivesand behaving as a Newtonian fluid (the skimming layer), under thecombined action of pressure and electrical fields, thus restrictingthe non-Newtonian behavior to the electrically neutral region out-side the electrical double layer (EDL). More recently, these studieswere extended to viscoelastic fluids by Afonso et al. [11], who pre-sented analytical solutions for channel and pipe flows of viscoelas-tic fluids under the mixed influence of electrokinetic and pressureforces, using two constitutive models: the Phan-Thien and Tanner(PTT) model [12], with linear kernel for the stress coefficient func-tion and zero second normal stress difference [13], and the FinitelyExtensible Non-linear Elastic dumbbell model with a Peterlinapproximation for the average spring force (FENE-P model, cf. Birdet al. [14]). An earlier investigation with the PTT model by Park andLee [15] was concerned with EOF in a square duct for which anumerical method was used. The analysis of Afonso et al. [11]was restricted to cases with small electric double layers, wherethe distance between the walls of a microfluidic device is at least

Page 2: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

278 A.M. Afonso et al. / Journal of Colloid and Interface Science 395 (2013) 277–286

one order of magnitude larger than the EDL, and the fluid is wellmixed and uniformly distributed across the channel. This well-mixed approximation was also considered by Park and Lee [15].When the viscoelastic flow is induced by a combination of bothelectric and pressure potentials, in addition to the single contribu-tions from these two mechanisms, there is an extra term in thevelocity profile that simultaneously combines both forcings, whichis absent for the Newtonian fluids where the superposition princi-ple applies. This extra term can contribute significantly to the totalflow rate, depending on the value of the relative microchannel ra-tio and appears only when the rheological constitutive equation isnon-linear. Afonso et al. [16] extended this study to the flow of vis-coelastic fluids under asymmetric zeta potential forcing, Sousaet al. [17] considered the formation of a skimming layer withoutpolymer near the walls, and Dhinakaran et al. [18] analyzed the fullPTT model with non-zero second normal stress differences, butonly considering EOF without a pressure gradient. Recently, Afonsoet al. [19] derived the full analytical solution for fully-developedelectro-osmosis driven flow of polymer solutions described bythe sPTT or FENE-P models with a Newtonian solvent.

The analytical solution of the steady two-fluid electro-osmoticstratified flow in a planar microchannel is presented in this work,assuming a planar interface between the two viscoelastic immisci-ble liquids. The working principle of the two-fluid pump is de-scribed in detail in Section 2. The PTT fluid considered hereobeys the simplified model [12], with a linear kernel for the stresscoefficient function and a zero second normal stress difference inshear [13]. The PTT model also includes the limiting case of theUpper-Convected Maxwell (UCM) fluids.

The remaining of the paper starts with the flow problem defini-tion, followed by the presentation of the set of governing equationsand a discussion of the assumptions made to obtain the analyticalsolution. Using this solution, the effects of the various relevantdimensionless parameters upon the flow field characteristics andpump efficiency are discussed in detail.

2. Flow geometry and definitions

The flow under investigation is the steady, fully-developedchannel flow of two incompressible and immiscible layers of visco-elastic fluids which also have significantly different conductivities,as shown schematically in Fig. 1a. This type of flow can be found insome EOF pumps [1], where the non-conducting fluid located atthe upper part of the system is dragged by an electrically conduct-ing fluid at the bottom part, as illustrated in Fig. 1b. Although theorigin of the coordinate system is considered at the interface be-tween the two fluids, their thickness is not necessarily identical.

Fig. 1. (a) Illustration of the coordinate system and (b) schematic of the two-fluidEOF pump.

The migration of ions leading to the formation of the electric lay-ers naturally arises due to the interaction between the dielectric bot-tom wall and the conducting fluid. Concerning the wall-fluidinterface, the charged bottom wall of the channel attracts counter-ions to form a layer of charged fluid near the wall and repels theco-ions. A very thin layer of immobile counter-ions covers the bot-tom wall, known as the Stern layer, and is followed by a thicker morediffuse layer of mobile counter-ions. These two layers near the wallform the EDL. The global charge of the conducting fluid remains neu-tral, but since the EDL is thin the core of the conducting fluid is essen-tially neutral. Applying a DC potential difference between the twoelectrodes at the inlet and outlet of the bottom channel section gen-erates an external electric field that exerts a Coulombic body forceon the counter-ions of the EDL, which move along the bottom chan-nel dragging the remaining fluid above by viscous forces.

A similar situation arises at the fluid–fluid interface, wherethere is also dielectric interaction leading to the formation of a sec-ond EDL in the conducting fluid next to the interface. The conduct-ing fluid (Fluid B) moves under the forcing imposed by theCoulombic forces and drags the non-conducting fluid (Fluid A) byhydrodynamic viscous forces at the interface (cf. Fig. 1a).

The pressure difference that can be independently applied be-tween the inlets and outlets of both the upper and the lower chan-nels can act in the same or in the opposite direction of the electricfield. Alternatively, the streamwise electric potential differencemay not be imposed independently, but results from the accumu-lation of ions at the end of the channel due to the flow forced by animposed pressure difference. This particular case is known as thestreaming potential and implies a specific relationship betweenthe imposed favorable pressure gradient and the ensuing adverseexternal electric field [20], a case which will not be analyzed in thispaper for conciseness.

To analyze this system, a two-dimensional Cartesian orthonor-mal coordinate system ( x,y) is used with the origin of the y-axislocated at the fluid–fluid interface, as shown in Fig. 1a. We assumea stratified viscoelastic flow and a planar interface. The thickness ofthe conducting and non-conducting fluid is H1 and H2, respectively.The width w is assumed to be very large, such thatw� H2 þ H1 ¼ H.

The holdup of the conducting fluid (Fluid B), RB, is here definedas the ratio between the cross-section area occupied by the con-ducting fluid and the total cross-section area of the channel,

RB ¼H1

H2 þ H1¼ H1

H: ð1Þ

Similarly, the holdup of non-conducting fluid (Fluid A) is defined as

RA ¼ 1� RB ¼H2

H2 þ H1¼ H2

H: ð2Þ

The electrical double layer that forms near the bottom channelwall in contact with the conducting fluid (Fluid B) has a zeta poten-tial denoted by f1. The second EDL in Fluid B, at the interface con-tact with fluid A, has an interfacial zeta potential (fi) that dependson the properties of the two fluids and also varies with the pH va-lue, the concentration of ions in the conducting fluid and the pres-ence of ionic surfactants [2]. This interfacial zeta potentialinfluences the potential distribution in the two EDLs, hence theelectro-osmotic force distribution and consequently the flow rates.

3. Theoretical model of the two-fluid electro-osmoticviscoelastic flow

The basic field equations describing this fully-developed lami-nar flow of incompressible fluids are the continuity equation,

r � u ¼ 0; ð3Þ

Page 3: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

A.M. Afonso et al. / Journal of Colloid and Interface Science 395 (2013) 277–286 279

and the modified Cauchy equation,

qDuDt¼ �rpþr � sþ qeE; ð4Þ

where u is the velocity vector, p is the pressure, and s is the polymericextra-stress tensor. The qeE term of Eq. (4) represents a Coulombicbody force per unit volume, where E is the applied external electricfield, and qe is the net electric charge density in the fluid. This termis null for the nonpolar fluid A, but needs to be quantified for the polarfluid B. The main simplifying assumptions and considerations in thecurrent analysis are as follows: (i) the two fluids are viscoelastic (theNewtonian fluid is included in the analysis as a limiting case when therelaxation time is negligible); (ii) fluid properties are assumed to beindependent of local electric field, ion concentration, and tempera-ture (this is certainly true for dilute solutions [2], but we also considerthis assumption to hold for our case); (iii) the flow is steady and fully-developed with no-slip boundary conditions at the channel walls;(iv) the two fluids are immiscible, and there is stratification with aplanar interface between fluids where a second EDL forms; (v) a pres-sure gradient can simultaneously be imposed along the channel and(vi) the standard electrokinetic theory conditions apply [21].

3.1. PTT model constitutive equations

The polymer extra-stress s is described by an appropriate con-stitutive equation, and in this work, we consider the viscoelasticmodel of Phan-Thien and Tanner [12,13] (PTT model), derived fromnetwork theory arguments:

f ðskkÞsþ k sr¼ 2gD: ð5Þ

Here, D ¼ 5uT þ5u� �

=2 is the rate of deformation tensor, k is therelaxation time of the fluid, g is the viscosity coefficient, and s

rrep-

resents the upper-convected derivative of s, defined as

sr¼ Ds

Dt�5uT :s� s:5 u: ð6Þ

The stress coefficient function, f ðskkÞ, is given by the linear form [12]

f ðskkÞ ¼ 1þ ekg

skk; ð7Þ

where skk represents the trace of the extra-stress tensor. The extensi-bility parameter, e , bounds the steady-state elongational viscosity,which is inversely proportional to e, for small e. For e ¼ 0, the UCMmodel is recovered which has an unbounded elongational viscosityabove a critical strain rate _e ¼ 1=ð2kÞ [14]. For fully-developed flowconditions, for which u ¼ uðyÞ;0;0f g, the extra-stress field for thePTT model can be obtained from equations (5)–(7), leading to

f ðskkÞsxx ¼ 2k _csxy ð8Þ

and

f ðskkÞsxy ¼ g _c ð9Þ

where skk ¼ sxx, since syy ¼ szz ¼ 0 [22,23], and _c is the transversevelocity gradient ( _c � du=dy). Then, upon division of Eq. (8) by Eq.(9), the specific function f ðskkÞ cancels out, and a relation betweenthe normal and shear stresses is obtained,

sxx ¼ 2kgs2

xy: ð10Þ

3.2. Electric double layers in the conducting fluid (Fluid B)

The potential field within the conducting fluid B can be ex-pressed by means of a Poisson equation:

r2w ¼ �qe

�; ð11Þ

where w denotes the induced electric potential, and � is the dielec-tric constant of the fluid. In equilibrium conditions near a chargedwall, the net electric charge density, qe, can be described as

qe ¼ �2noez sinhez

kBTw

� �; ð12Þ

where no is the ion density, e is the elementary electric charge, z isthe valence of the ions, kB is the Boltzmann constant, and T is theabsolute temperature. In order to obtain the velocity field for fluidB, we first need to determine the net charge density distribution(qe). The charge density field can be calculated by combining Eqs.(11) and (12) under fully-developed flow conditions, to obtain thewell-known Poisson–Boltzmann equation

d2wdy2 ¼

2noez�

sinhez

kBTw

� �: ð13Þ

The electro-osmotic flow is primarily caused by the action of anexternally applied electric field on the charged species that existnear the bottom channel wall and in the vicinity of the interfacialsurface. The distribution of the charged species in the domain isgoverned by the potentials at the wall and at the interface and thenby the externally applied electric field. When the Debye thicknessesare small, and the charges at the wall and at the interface are notlarge, the distribution of the charged species is governed mainlyby the f1 potential at the wall and by fi at the interface, and negli-gibly affected by the applied external DC electric field (standardelectrokinetic theory). Thus, the charge distribution across fluid Bcan be determined independently of the externally applied electricfield. Indeed, the effect of fluid motion on the charge redistributioncan be neglected when the fluid velocity is small, that is, when theinertial terms in the momentum equation are not dominant (indeedthey vanish under fully-developed conditions) or when the Debyethickness is small. Additionally, for small values of w, the Debye–Hückel linearization principle (sinh x � x) can also be used, whichmeans physically that the electric potential energy is small com-pared with the thermal energy of ions, and the Poisson–Boltzmannequation can be simplified to:

d2wdy2 ¼ j2w; ð14Þ

where j2 ¼ 2noe2z2

�kBT is the Debye–Hückel parameter, related with thethickness of the Debye layer as n ¼ 1

j (normally referred to as theEDL thickness). This approximation is valid when the Debye thick-ness is small but finite, that is, for 10 K H1=nK 103.

Eq. (14) can be integrated subjected to the following boundaryconditions: zeta potential at the bottom wall wky¼�H1

¼ f1 and zetapotential at the interface wky¼0 ¼ fi. The potential field becomes

wðyÞ ¼ f1 W1ejy �W2e�jyð Þ ð15Þ

for �H1 6 y 6 0. Denoting Rf ¼ fi=f1 as the ratio of zeta potentials,

then, W1 ¼ RfejH1�12 sinh jH1ð Þ and W2 ¼ Rfe�jH1�1

2 sinh jH1ð Þ. When Rf ¼ 1 a symmetric

potential profile is observed within the conducting fluid, as ob-tained by Afonso et al. [11] for the whole channel with a single con-ducting fluid, whereas for vanishing zeta potential at the interface,Rf ¼ 0, one obtains a special case discussed by Afonso et al. [16]. Fi-nally, the net charge density distribution, Eq. (12), coupled with theDebye–Hückel linearization principle leads to

qe ¼ ��j2f1 W1ejy �W2e�jyð Þ ¼ ��j2f1X�1 ðyÞ ð16Þ

where the operator

X�p ðyÞ ¼ W1ejyð Þp � W2e�jyð Þp ð17Þ

Page 4: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

280 A.M. Afonso et al. / Journal of Colloid and Interface Science 395 (2013) 277–286

is a hyperbolic function of the transverse variable y, that depends onthe ratio of zeta potentials, Rf, and on the thickness of the Debyelayer. For the non-conducting Fluid A, there is no potential distribu-tion, that is, the induced potential is null as well as the correspond-ing electric charge density, qe.

3.3. Momentum equation of the two-fluid flow

3.3.1. Conducting fluid (Fluid B)For the conducting fluid (Fluid B), the momentum Eq. (4) re-

duces to,

dsxy;B

dy¼ p;x � qeEx ¼ p;x þ �j2f1ExX

�1 ðyÞ ð18Þ

where Ex � �d/=dx and p;x � dp=dx. The electric potential of the ap-plied external field, /, is characterized by a constant streamwisegradient. Eq. (18) is integrated to yield the following shear stressdistribution

sxy;B ¼ p;xyþ �jf1ExXþ1 ðyÞ þ sB ð19Þ

where sB is a stress integration coefficient related to the stress atthe fluid–fluid interface. It is clear that in contrast to pure Poiseuilleflow, the shear stress distribution is no longer linear on the trans-verse coordinate. Using the relationship between the normal stressand the shear stress, Eq. (10), an explicit expression for the normalstress component is also obtained,

sxx;B ¼ 2kg

p;xyþ �jf1ExXþ1 ðyÞ þ sB

� �2: ð20Þ

For simplicity, subscript B will be removed from the rheologicalparameters of Fluid B (gB ¼ g, eB ¼ e and kB ¼ k). Combining Eqs.(9), (19) and (20) allows to obtain the expression for the velocitygradient:

duB

dy¼ 1þ 2ek2 �Exf1

gjXþ1 ðyÞ þ

sB

p;xg

y� �2

" #ð21Þ

� �Exf1

gjXþ1 ðyÞ þ

sB

p;xg

y� �

:

Eq. (21) can be integrated subject to the no-slip boundary conditionat the lower wall (uBky¼�H1 ¼ 0) leading to

uB ¼sB

gyþ H1ð Þ 1þ 2ek2 sB

g

� �2 !

þ �Exf1

g

� �1þ 6

sB

g

� �2

ek2

!X�1;1ðyÞ

þ 2ek2 �Exf1

g

� �2

jsB

g6W1W2j yþ H1ð Þ þ 3

2X�2;1ðyÞ

� �

þ 2ek2 �Exf1

g

� �3

j2 13

X�3;1ðyÞ þ 3W1W2X�1;1ðyÞ

� �

þ 12

p;xg

� �y2 � H2

1

� 1þ 6ek2 sB

g

� �2

þ ek2 p;xg

� �2

y2 þ H21

� !þ 2

� sB

gek2 p;x

g

� �2

y3 þ H31

� þ 12

ek2 �Exf1g

h ip;xg

h ij

sB

gX�1;2ðyÞ �Xþ1;1ðyÞ�

þ 6ek2 �Exf1

g

h ip;xg

h i2

j2 X�1;3ðyÞ þ 2X�1;1ðyÞ � 2Xþ1;2ðyÞ�

þ 6ek2 �Exf1

g

� �2 p;xg

� �W1W2j2 y2 � H2

1

� þ 1

2X�2;2ðyÞ �

14

Xþ2;1ðyÞ� �

ð22Þ

where the operator X�p;qðyÞ, is now defined for compactness:

X�p;qðyÞ ¼ jyð Þðq�1ÞX�p ðyÞ � ð�1Þðqþ1Þ jH1ð Þðq�1ÞX�p ð�H1Þ: ð23Þ

Eq. (22) is valid for �H1 6 y < 0.It is often more convenient to work with the dimensionless

form of Eq. (22). Introducing the normalizations �y ¼ y=H1 ¼y= RBHð Þ and �j ¼ jRBH , the dimensionless velocity profile in theconducting fluid can be rewritten as

uB

ush¼ sB yþ 1ð Þ 1þ 2s2

BeDe2

jj2

!� 1þ 6s2

BeDe2

jj2

!X�1;1ðyÞ

þ 2sBeDe2

jj

6W1W2j yþ 1ð Þ þ 32

X�2;1ðyÞ� �

� 2eDe2j

13

X�3;1ðyÞ þ 3W1W2X�1;1ðyÞ� �

þ 12

C y2 � 1� �

1þ 6s2BeDe2

jj2 þ

eDe2j

j2 C2 y2 þ 1� � !

þ 2sB

� eDe2j

j2 C2 y3 þ 1� �

� 12sBeDe2

jj3 C X�1;2ðyÞ �Xþ1;1ðyÞ

þ 6eDe2

jj2 C W1W2j2 y2 � 1

� �þ 1

2X�2;2ðyÞ �

14

Xþ2;1ðyÞ� �

� 6eDe2

jj4 C2 X�1;3ðyÞ þ 2X�1;1ðyÞ � 2Xþ1;2ðyÞ

� ð24Þ

where sB ¼ sBg

RBHush

and Dej ¼ kushn ¼ kjush is the Deborah number

based on the relaxation time of the conducting fluid (Fluid B), onthe EDL thickness and on the Helmholtz–Smoluchowski electro-os-motic velocity near the bottom wall, defined as ush ¼ � �f1Ex

g . The

dimensionless parameter C ¼ � RBHð Þ2�f1

p;xEx

represents the ratio of pres-

sure to electro-osmotic driving forces. Note that for simplicity, theabove terms were based on the zeta potential at the bottom wallðwky¼�H1

¼ f1Þ, but could have been based instead on the interfacialzeta potential using the ratio of zeta potentials: ush ¼ ushi=Rf,C ¼ RfCi and Dej ¼ Deji=Rf.

The normalized flow rate of the pumping Fluid B can be deter-mined from integration of the velocity profile

Q B ¼ uBush¼R 0�H1

uBushH1

dy¼ 12sB 1þ2 eDe2

jj2 s2

B

� � 1

3C 1þ6s2B

eDe2j

j2 þ 65

eDe2j

j2 C2�

þ32sB

eDe2j

j2 C2� 1þ6 eDe2j

j2 s2B

� Xþ1;1ð0Þ

j �X�1 ð�1Þ� �

þ2 eDe2j

j sB 3W1W2jþ 34

Xþ2;1ð0Þj �2X�2 ð�1Þ

� �� �

�2eDe2j

Xþ3;1ð0Þ9j �

13X�3 ð�1Þþ3W1W2

Xþ1;1ð0Þj �X�1 ð�1Þ

� �� �

�12sBeDe2

jj4 C 2Xþ1;2ð0Þ�2X�1;1ð0Þþj2X�1 ð�1Þ

h iþ6 eDe2

jj2 C 1

2jXþ2;2ð0Þ� 23W1W2j2þ 1þ2j2

4j X�2 ð�1Þ� 14jX�2 ð0Þ

� �6 eDe2

jj5 C2 jX�1;3ð0Þ�3jXþ1;2ð0Þþ6Xþ1;1ð0Þ�6jX�1 ð�1Þ

� :

ð25Þ

3.3.2. Non-conducting fluid (Fluid A)The derivation of the analytical solution for this fluid layer fol-

lows the same steps as for the conducting fluid, with the necessaryadaptations. For the non-conducting fluid (Fluid A), the momen-tum conservation Eq. (4) reduces to

dsxy;A

dy¼ p;x; ð26Þ

since, as explained, there is no external electrical field forcing, dueto low conductivity of fluid A. Eq. (26) can be integrated to yield thefollowing shear stress distribution

Page 5: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

A.M. Afonso et al. / Journal of Colloid and Interface Science 395 (2013) 277–286 281

sxy;A ¼ p;xyþ sA ð27Þ

where sA is the shear stress at the fluid–fluid interface, to be quan-tified in the next section. Using the relationship between the nor-mal and shear stresses, Eq. (10), the following explicit expressionfor the normal stress component is obtained,

sxx;A ¼ 2kA

gAp;xyþ sA� �2

: ð28Þ

Combining Eqs. (9), (27) and (28) the velocity gradient distributionin Fluid A is given by

duA

dy¼ 1þ 2eAk

2A

p;xgA

yþ sA

gA

� �2" #

p;xgA

yþ sA

gA

� �ð29Þ

Eq. (29) can be integrated subject to the no-slip boundary conditionat the upper wall (uAky¼H2 ¼ 0), leading to

uA ¼sA

gAy� H2ð Þ 1þ 2eAk

2A

sA

gA

� �2 !

þ 2eAk2AsA

gA

p;xgA

� �2

y3 � H 32

� �þ 1

2p;xgA

� �y2 � H 2

2

� �

� 1þ 6eAk2A

sA

gA

� �2

þ eAk2A

p;xgA

� �2

y2 þ H 22

� � !ð30Þ

valid for 0 < y 6 H2. Introducing the normalizations�yA ¼ y=H2 ¼ y=RAH and �jA ¼ jRAH, the dimensionless velocity pro-file can be written as

uA

ush¼ sA �yA � 1ð Þ 1þ 2s2

AeADe2

jA

j2A

!þ 2sA

1b2

� eADe2jA

j2A

C2A y3

A � 1� �

þ 12

� 1b

CA y2A � 1

� �1þ 6s2

AeADe2

jA

j2A

þ 1b2

eADe2jA

j2A

C2A y2

A þ 1� � !

ð31Þ

where sA ¼ sAgA

RAHush

, b ¼ gA=gB is the dynamic viscosity ratio, and

DejA ¼ kAushn ¼ kAushj is the Deborah number based on the relaxation

time of fluid A, and on the EDL thickness on the bottom wall and onthe corresponding Helmholtz–Smoluchowski electro-osmotic

a1 ¼�3X�1;1ð0Þ � 3

2 C� 3 R2e

b3 jAXþ1 ð0Þ þ 3

2R2e

b3 CA

a4

a2 ¼12

j2

eDe2jþ j 6W1W2jþ 3

2 X�2;1ð0Þ�

þ C2 � 6 Cj X�1;2ð0Þ �Xþ1;1ð0Þ�

þ 12

1b

a4

þRARB

R2e

b3j2

j2AC2

A þ 3 R2e

b3RARB

j2 Xþ1 ð0Þ� �2 � 3 R2

eb3

j2

jA

RARB

CAXþ1 ð0Þ

a4

a3 ¼3C �W1W2j2 þ 1

2 X�2;2ð0Þ � 14 Xþ2;1ð0Þ

� � 1

2j2

eDe2jX�1;1ð0Þ � j2 1

3 X�3;1ð0Þ�

a4

þ14

1b

j2

eDe2jCA � 1

4j2

eDe2jCþ 1

4R2e

b3j2

j2AC3

A � 3 C2

j2 X�1;3ð0Þ þ 2X�1;1ð0Þ � 2Xþ1�

a4

þ32

R2e

b3 j2CA Xþ1 ð0Þ� �2 � 1

2jAb

j2

eDe2jXþ1 ð0Þ �

R2e

b3 j2jA Xþ1 ð0Þ� �3

a4

velocity (so, for simplicity, we use a single set of characteristicscales for normalization related to the bottom wall of the channel).

The parameter CA ¼ � RAHð Þ2�f1

p;xEx

represents the ratio of pressure to

electro-osmotic driving forces. The normalized volumetric flow rateof the pumped Fluid A in the upper part of the channel is given by

QA ¼uA

ush

� �¼R H2

0 uAdyushH2

¼ � sA

21þ 2

eADe2jA

j2A

s2A

!� 3

2sA

eADe2jA

b2j2A

C2A �

25

C3A

b3

eADe2jA

j2A

� 13

CA

b1þ 6s2

AeADe2

jA

j2A

!: ð32Þ

3.3.3. Fluid A–Fluid B interface conditionsIn deriving the shear stress profiles, Eqs. (19) and (27), and all

the subsequent quantities such as velocity and flow rates, two inte-gration coefficients appeared, sA and sB, which have to be deter-mined from the boundary conditions at the fluid–fluid interface,namely: sxy;Aky¼0 ¼ sxy;Bky¼0 and uA;ky¼0 ¼ uBky¼0.

Using the relationships between the shear stresses at the inter-face, equations (19) and (27), and those for the dimensionlessvelocity profiles, Eqs. (24) and (31), the variables sA and sB canbe determined:

sA ¼ RARB

1b sB � jA

b Xþ1 ð0Þ

sB ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi� b1

2 þffiffiffiffiffiffiffiffiffiffiffiffiffib2

14 þ a3

27

q3

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi� b1

2 �ffiffiffiffiffiffiffiffiffiffiffiffiffib2

14 þ a3

27

q3

r� a1

3

8><>: ð33Þ

where a ¼ a2 � a21=3 and b1 ¼ a3 � a1a2=3þ 2a3

1=27. The coefficients

a1, a2 and a3 are given byand a4 ¼ 1þ R2e

b3j2

j2A

RARB

� 3, where Re ¼

ffiffiffiffieAe

qkAk

is a dimensionless parameter that relates the rheological properties

of the two fluids.

4. Results and discussion

In the previous section, general equations were derived for stea-dy fully-developed two-fluid electro-osmotic stratified flow of PTTviscoelastic fluids under the mixed influence of electrokinetic andpressure gradient forcings. The different influences of the driving

RARB

j2

eDe2j

ð34Þ

þ 3W1W2X�1;1ð0Þ� 1

4 C3

;2ð0Þ� R2

eb3

j2

jAC2

AXþ1 ð0Þ

Page 6: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

u/ush

y/(R

A/BH)

-1 -0.5 0 0.5 1 1.5 2-1

-0.5

0

0.5

1

Γ=-2

Γ=2

β=1; κ=20; Rζ= 0; RΑ= 1/2

ε0.5Deκ= 0

Γ

Q

-5 -4 -3 -2 -1 0 1 2 3

0

0.5

1

1.5

2

2.5

3

Fluid B

Fluid A

β=1; κ=20; Rζ= 0; RΑ= 1/2

|

_(b)

(a)

Fig. 2. Effect of the driving forces (C ¼ �2, �1, 0, 1 and 2) on dimensionless (a)velocity profiles and (b) volumetric flow rate for the Newtonian–Newtonian flowconfiguration. Symbols represent the data from Afonso et al. [11] for (b ¼ 1 , Rf ¼ 0and C ¼ 0).

u/ush

y/(R

A/B

H)

0.50 1-1

-0.5

0

0.5

1

β= 10-2

β= 100

Γ=0; κ=20; Rζ= 0; RΑ= 1/2

ε0.5Deκ= 0

β10-3 10-2 10-1 100 101 102 103

-0.2

0

0.2

0.4

0.6

0.8

1

1.2Fluid B

Fluid A

Γ=0; κ=20; Rζ= 0; RΑ= 1/2Q|

_

(a)

(b)

Fig. 3. Effect of the viscosity ratio (b ¼ 10�2, 10�1, 0.5, 1, 2, 10 and 100) ondimensionless (a) velocity profiles and (b) volumetric flow rate for the Newtonian–Newtonian flow configuration.

282 A.M. Afonso et al. / Journal of Colloid and Interface Science 395 (2013) 277–286

forces (C), fluid rheology ðReÞ, viscosity ratio ðbÞ, fluids holdup ðRBÞ,and of the ratio of zeta potentials ðRfÞ on the velocity profile wereexplicitly incorporated in Eqs. (30), (22) and (33). In this section,we discuss in detail some limiting cases in order to understandthe system fluid dynamics.

The following set of two-fluid systems is included in the generalsolution where the second fluid is the conducting medium: (a)Newtonian–Newtonian fluid system; (b) viscoelastic–Newtonianfluid system; (c) Newtonian–viscoelastic fluid system; and (d) vis-coelastic–viscoelastic fluid system. Cases (b and d) are not dis-cussed in this work, due to space limitations, although thederived equations also include these cases. Case (a) was studiedin detail elsewhere [2], but this situation is revisited here as a start-ing point and for comparison with case (c), that is, in the followingwe analyze in detail the pumping of a Newtonian fluid by anotherNewtonian fluid, and, alternatively, by a viscoelastic fluid.

4.1. Newtonian–Newtonian EOF pump configuration

For the Newtonian–Newtonian fluid flow configuration, boththe conducting and the non-conducting fluids are NewtonianðDej ¼ DejA ¼ 0Þ. The velocity profile system equations and thedimensionless boundary condition coefficients, provided by Eq.(33), simplify to

uAush¼ sA yA � 1ð Þ þ 1

2b CA y2A � 1

� �for 0 6 yA 6 1

uBush¼ sB yþ 1ð Þ �X�1;1ðyÞ þ 1

2 C y2 � 1� �

for � 1 6 y 6 0

(ð35Þ

For small relative microchannel ratio, �j! 1, the double layerthickness is of the same order of magnitude as the Fluid B thickness,and the region of excess charge is distributed over the entire fluid.

Page 7: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

u/ush

y/(R

A/BH)

-1 0 1 2 3-1

-0.5

0

0.5

1

Rζ=0.2

Rζ=-0.2

β=1; κ=20; Γ=0; RA= 1/2

ε0.5Deκ= 0

_

Q

-0.2 0 0.2 0.4 0.6 0.8 1 1.2

0

2

4

6

8

Fluid B

Fluid A

Γ=0; κ=20; β= 1; RΑ= 1/2

|

_

(a)

(b)

Fig. 4. Effect of the ratio of zeta potentials (Rf ¼ �0:2, �0.1, 0, 0.1 and 0.2) ondimensionless (a) velocity profiles and (b) volumetric flow rate for the Newtonian-Newtonian flow configuration.

u/ush

y/(R

A/B

H)

0 0.5 1-1

-0.5

0

0.5

1

RA=2/3

RA=1/3

ε0.5Deκ= 0

β=1; κ=20; Γ=0; Rζ= 0_

RA

0 0.2 0.4 0.6 0.8 10

0.2

0.4

0.6

0.8

1Fluid B

Fluid A

Q|

Γ=0; κ=20; Rζ= 0; β= 1_

(a)

(b)

Fig. 5. Effect of the non-conducting fluid holdup on dimensionless (a) velocityprofiles (RA ¼ 1=3, 1=2 and 2=3) and (b) volumetric flow rate for the Newtonian–Newtonian flow configuration.

A.M. Afonso et al. / Journal of Colloid and Interface Science 395 (2013) 277–286 283

This situation is not compatible with this solution for which the De-bye–Hückel approximation was invoked, which requires jJ 10. Inthis work and as a typical example, we set �j ¼ 20 in all figures.

For C ¼ 0, that is, when the flow is driven only by electro-osmo-sis, the velocity profile is a function of the wall distance, of the rel-ative microchannel ratio, �j, of the ratio of zeta potentials, Rf, and ofthe viscosity ratio as shown earlier by Gao et al. [2]. Also, for a sin-gle fluid situation ðb ¼ 1Þ and in the absence of interface zeta po-tential ðRf ¼ 0Þ, the solution simplifies to a particular caseobtained by Afonso et al. [11] (no zeta potential in the upper walland no pressure gradient, symbols in Fig. 2a). The correspondingeffect of the ratio of pressure gradient to electro-osmotic drivingforces on the dimensionless flow rate is obvious (cf. Fig. 2b),increasing with favorable pressure gradients ðC < 0Þ and decreas-ing for flows with adverse pressure gradients ðC > 0Þ. Obviously,the flow rate for Fluid B is higher than for Fluid A because for iden-tical fluids height Fluid B is being forced also by electro-osmosis.

Fig. 3 shows the influence of the viscosity ratio ðb � gA=gBÞ onthe dimensionless velocity profile (a) and volumetric flow rate(b). When the viscosity ratio decreases, the dimensionless velocityincreases (cf. Fig. 3a). So, if the viscosity of the conducting fluid ismuch higher than the viscosity of the non-conducting fluid, an in-crease in the dimensionless volumetric flow rate is expected, ascan be observed in Fig. 3b. However, a higher viscosity implies alower Helmholtz–Smoluchowski electro-osmotic velocity and con-sequently the dimensional flow rate may actually decrease.

A major effect on the velocity profile is that due to non-zerointerfacial zeta potential, as presented in the profiles of Fig. 4.When fi=f1 > 0 , a favorable extra Coulombic forcing term arisesin the velocity profile at the interface of the two fluids, leading toa significant increase in the volumetric flow rate. When fi < 0,the adverse localized electrostatic force decreases the pumping ac-tion and the corresponding dimensionless flow rate (cf. Fig. 4b).

Page 8: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

u/ush

y/(R

A/BH)

0 0.5 1 1.5 2 2.5 3-1

-0.5

0

0.5

1

ε0.5Deκ= 0

β= 1; Γ=0; κ=20; Rζ= 0; RΑ= 1/2

ε0.5Deκ= 2

ε0.5Deκ= 1

ε0.5Deκ= 0.5

ε0.5Deκ= 0.25

_

ε0.5Deκ

Q

0 2 4 6 8 100

5

10

15

20

Fluid B

Fluid A

Γ=0; κ=20; Rζ= 0; β= 1

|

_

(a)

(b)

Fig. 6. Dimensionless profiles of (a) velocity and (b) volumetric flow rate asfunction of

ffiffiffiep

Dej .

u/ush

y/(R

A/BH)

0 1 2 3 4-1

-0.5

0

0.5

1

Γ=-2

Γ=2

β=1; κ=20; Rζ= 0; RΑ= 1/2

ε0.5Deκ= 2

Γ

Q

-5 -4 -3 -2 -1 0 1 20

2

4

6

8

Fluid B

Fluid A

β=1; κ=20; Rζ= 0; RΑ= 1/2

ε0.5Deκ= 0

ε0.5Deκ= 2

|

_

(a)

(b)

Fig. 7. Effect of the driving forces (C ¼ �2, �1, 0, 1 and 2 ) on dimensionless (a)velocity profiles and (b) volumetric flow rate for the Newtonian–viscoelastic flowconfiguration.

284 A.M. Afonso et al. / Journal of Colloid and Interface Science 395 (2013) 277–286

Another important effect is due to the holdup of the non-con-ducting fluid. When the height of the non-conducting fluid is largerthan the height of the conducting fluid ðRA > RBÞ, the normalizedvelocities of both fluids increase, as observed in Fig. 5a. This sug-gests that to obtain higher volumetric flow rates in Fluid A, theholdup of the conducting Fluid B should be kept small (cf.Fig. 5b). In fact, as the Helmholtz–Smoluchowski electro-osmoticvelocity is independent of the thickness of Fluid B, as RA ! 1 thefluid interface plane will tend to coincide with the regions of high-er velocities. This conclusion also suggests that a better configura-tion for an EOF pump would be a three layer fluid flow, with theconducting fluid in contact with both the upper and lower walls,and the non-conducting fluid in the middle being dragged like a so-lid body, that is, a solid lubrificated by thin layers of conductingfluid in motion.

4.2. Newtonian–viscoelastic EOF pump configuration

For the Newtonian–viscoelastic fluid flow configuration, theconducting fluid is viscoelastic dragging the non-conducting New-tonian fluid. The Deborah number of the conducting fluid is non-zero (Dej – 0 and DejA ¼ 0), and the velocity profile and the non-dimensional boundary condition coefficients are given by

uAush¼ sA yA � 1ð Þ þ 1

2b CA y2A � 1

� �for 0 6 yA 6 1

Eq:ð24Þ for � 1 6 y 6 0

(ð36Þ

Fig. 6a and b presents the dimensionless velocity and volumet-ric flow rate profiles as a function of

ffiffiffiep

Dej, respectively. We cansee that increasing the elasticity of the conducting fluid, more thandoubles the velocities due to shear-thinning effects within the EDL

Page 9: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

u/ush

y/(R

A/B

H)

0 0.5 1 1.5 2 2.5 3 3.5 4-1

-0.5

0

0.5

1

β= 10-2

β= 100

ε0.5Deκ= 2

Γ=0; κ=20; Rζ= 0; RΑ= 1/2_

β10-3 10-2 10-1 100 101 102 1030

1

2

3

4Fluid B

Fluid A

ε0.5Deκ= 0

ε0.5Deκ= 2

Q|

Γ=0; κ=20; Rζ= 0; RΑ= 1/2_

(a)

(b)

Fig. 8. Effect of the viscosity ratio (b ¼ 10�2, 10�1, 0.5, 1, 2, 10 and 100) ondimensionless (a) velocity profiles and (b) volumetric flow rate for the Newtonian–viscoelastic flow configuration.

u/ush

y/(R

A/BH)

-1 0 1 2 3 4-1

-0.5

0

0.5

1

Rζ=0.2

Rζ=-0.2

β=1; κ=20; Γ=0; RA= 1/2

ε0.5Deκ= 2

_

Q

-0.2 0 0.2 0.4 0.6 0.8 1 1.20

10

20

30

40Fluid B

Fluid A

Γ=0; κ=20; β= 1; RΑ= 1/2

ε0.5Deκ= 0

ε0.5Deκ= 2

|

_(b)

(a)

Fig. 9. Effect of the ratio of zeta potentials (Rf ¼ �0:2, �0.1, 0, 0.1 and 0.2) ondimensionless (a) velocity profiles and (b) volumetric flow rate for the Newtonian–viscoelastic flow configuration.

A.M. Afonso et al. / Journal of Colloid and Interface Science 395 (2013) 277–286 285

layer, thus raising the velocity value of the bulk transport in thecore of the channel. This also helps to increase the shear rates nearthe bottom wall and at the two fluids interface, increasing the dragforce of the non-conducting fluid by the hydrodynamic viscousforces at the interface. Consequently, there is an increase in thedimensionless volumetric flow rate (cf. Fig. 6b). Fig. 6a also showsthat in the absence of pressure gradient (and Rf ¼ 0) the EDL actslike a plate in pure Couette flow, generating a nearly constant shearstress across the channel.

Fig. 7 shows the dimensionless velocity profiles (a) and volu-metric flow rate (b) at

ffiffiffiep

Dej ¼ 2 (for comparison, the Newtonianresults of Fig. 2 are also presented) to illustrate the effect of C . Afavorable pressure gradient ðC < 0Þ leads to an increase of the flowrate and makes velocity profiles fuller. By using also pressure forc-ing, which affects directly the two fluids, the flow rate quickly in-creases. The beneficial shear-thinning effect is clear in the large

increase in the flow rate of Fig. 7b as compared to Fig. 2b, butthe impact is stronger with the fluid B, where the shear-thinningis stronger, whereas the large viscosities in Fluid A limit the flowrate enhancement. As for the Newtonian-Newtonian fluid flowconfiguration, decreasing b leads to an increase in velocity profilesand the volumetric flow rate, which is further increased by shear-thinning effects (cf. Fig. 8a and b and compare with Fig. 3). Whenusing a viscoelastic fluid as conducting fluid, it is natural to havea more viscous fluid than the Newtonian non-conducting fluid,which leads to an optimal flow situation.

The effects of the Fluid A holdup ðRAÞ and of the ratio of zetapotentials ðRfÞ are similar to what was described before, but nowthe viscoelastic flow exhibits a shear-thinning viscosity and thevelocities have increased significantly near the bottom wall (seethe higher values of u=ush) leading to the higher volumetric flowrates of Figs. 9 and 10, than in the corresponding constant viscositycase.

Page 10: Journal of Colloid and Interface Sciencepaginas.fe.up.pt/~fpinho/pdfs/JCIS_EO-pump_2013.pdf · This paper presents an analytical model that describes a two-fluid electro-osmotic

u/ush

y/(R

A/B

H)

0 1 2 3-1

-0.5

0

0.5

1

RA=2/3

RA=1/3

ε0.5Deκ= 2

β=1; κ=20; Γ=0; Rζ= 0_

RA

0 0.2 0.4 0.6 0.8 10

1

2

3

Fluid B

Fluid A

ε0.5Deκ= 0

ε0.5Deκ= 2

Q|

Γ=0; κ=20; Rζ= 0; β= 1_

(a)

(b)

Fig. 10. Effect of the non-conducting fluid holdup on dimensionless (a) velocityprofiles (RA ¼ 1=3, 1=2 and 2=3) and (b) volumetric flow rate for the Newtonian–viscoelastic flow configuration.

286 A.M. Afonso et al. / Journal of Colloid and Interface Science 395 (2013) 277–286

5. Conclusions

An analytical solution of the steady two-fluid electro-osmoticstratified flow in a planar microchannel is presented assuming a

planar interface between the two viscoelastic immiscible fluids.The PTT fluid model was used, and the effects of fluid rheology, vis-cosity ratio, fluid holdup, and interfacial zeta potential were ana-lyzed to show the viability of this pumping technique.

The flow can be induced by a combination of both electrical andpressure potentials, but in addition to the single contributionsfrom these two mechanisms, when the conducting fluid is visco-elastic, there is an extra term in the velocity profile that simulta-neously combines both effects, which is absent from conductingNewtonian fluids, in which the linear superposition principle ap-plies. This work demonstrated that higher volumetric flow ratesof a non-conducting Newtonian fluid can be achieved in EOFpumping when the conducting fluid is viscoelastic rather thanNewtonian, due to the increase of the shear-thinning effects.

Acknowledgments

The authors would like to acknowledge financial support fromFEDER and Fundação para a Ciência e a Tecnologia (FCT) throughProject PTDC/EQU-FTT/113811/2009 and Scholarship SFRH/BPD/75436/2010 (A.M. Afonso).

References

[1] A. Brask, G. Goranovic, H. Bruus, Tech. Proc. Nanotech. 1 (2003) 190–193.[2] Y. Gao, T.N. Wong, C. Yang, K.T. Ooi, J. Colloid Interf. Sci. 284 (1) (2005) 306–

314.[3] C.H. Chen, J.G. Santiago, J. Microelectromech. Syst. 11 (6) (2002) 672–683.[4] X. Xuan, D. Li, J. Micromech. Microeng. 14 (2004) 290.[5] X. Wang, C. Cheng, S. Wang, S. Liu, Microfluidics Nanofluidics 6 (2) (2009) 145–

162.[6] D. Laser, J. Santiago, J. Micromech. Microeng. 14 (2004) 290–298.[7] S. Zeng, C.H. Chen, J.C. Mikkelsen, J.G. Santiago, Sens. Actuat. B: Chem. 79 (2–3)

(2001) 107–114.[8] S. Das, S. Chakraborty, Anal. Chim. Acta 559 (2006) 15–24.[9] S. Chakraborty, Anal. Chim. Acta 605 (2007) 175–184.

[10] C.L.A. Berli, M.L. Olivares, J. Colloid Interf. Sci. 320 (2008) 582–589.[11] A.M. Afonso, M.A. Alves, F.T. Pinho, J. Non-Newtonian Fluid Mech. 159 (2009)

50–63.[12] N. Phan-Thien, R.I. Tanner, J. Non-Newtonian Fluid Mech. 2 (4) (1977) 353–

365.[13] N. Phan-Thien, J. Rheol. 22 (1978) 259–283.[14] R.B. Bird, P.J. Dotson, N.L. Johnson, J. Non-Newtonian Fluid Mech 7 (1980) 213–

235.[15] H.M. Park, W.M. Lee, Lab on a Chip 8 (7) (2008) 1163–1170.[16] A.M. Afonso, M.A. Alves, F.T. Pinho, J. Eng. Math. 71 (2011) 15–30.[17] J.J. Sousa, A.M. Afonso, F.T. Pinho, M.A. Alves, Microfluidics Nanofluidics 10

(2011) 107–122.[18] S. Dhinakaran, A.M. Afonso, M.A. Alves, F.T. Pinho, J. Colloid Interf. Sci. 344

(2010) 513–520.[19] A.M. Afonso, F.T. Pinho, M.A. Alves, J. Non-Newtonian Fluid Mech. 179–180

(2012) 55–68.[20] C. Yang, D. Li, J. Colloid Interf. Sci. 194 (1) (1997) 95–107.[21] R.F. Probstein, Physicochemical Hydrodynamics: An Introduction, second ed.,

Hoboken, Wiley Interscience, New Jersey, USA, 2003.[22] F.T. Pinho, P.J. Oliveira, J. Non-Newtonian Fluid Mech. 93 (2–3) (2000) 325–

337.[23] M.A. Alves, F.T. Pinho, P.J. Oliveira, J. Non-Newtonian Fluid Mech. 101 (2001)

55–76.