Curvature-induced secondary microflow motion in steady electro...

10
Curvature-induced secondary microflow motion in steady electro-osmotic transport with hydrodynamic slippage effect Jin-Myoung Lim and Myung-Suk Chun a) Complex Fluids Research Laboratory, National Agenda Research Division, Korea Institute of Science and Technology (KIST), Seongbuk-gu, Seoul 136-791, South Korea (Received 17 February 2011; accepted 9 September 2011; published online 19 October 2011) In order to exactly understand the curvature-induced secondary flow motion, the steady electro- osmotic flow (EOF) is investigated by applying the full Poisson-Boltzmann/Navier-Stokes equa- tions in a whole domain of the rectangular microchannel. The momentum equation is solved with the continuity equation as the pressure-velocity coupling achieves convergence by employing the advanced algorithm, and generalized Navier’s slip boundary conditions are applied at the hydro- phobic curved surface. Two kinds of channels widely used for lab-on-chips are explored with the glass channel and the heterogeneous channel consisting of glass and hydrophobic polydimethylsi- loxane, spanning thin to thick electric double layer (EDL) problem. According to a sufficiently low Dean number, an inward skewness in the streamwise velocity profile is observed at the turn. With increasing EDL thickness, the electrokinetic effect gets higher contribution in the velocity profile. Simulation results regarding the variations of streamwise velocity depending on the electrokinetic parameters and hydrodynamic fluid slippage are qualitatively consistent with the predictions docu- mented in the literature. Secondary flows arise due to a mismatch of streamline velocity between fluid in the channel center and near-wall regions. Strengthened secondary flow results from increas- ing the EDL thickness and the contribution of fluid inertia (i.e., electric field and channel curva- ture), providing a scaling relation with the same slope. Comparing with and between the cases enables us to identify the optimum selection in applications of curved channel for enhanced EOF and stronger secondary motion relevant to the mixing effect. V C 2011 American Institute of Physics. [doi:10.1063/1.3650911] I. INTRODUCTION The electro-osmotic flow (EOF) is a fundamental electroki- netic phenomenon, which plays an important role for improve- ments and optimum design of the devices related to lab-on-chips (LOCs). 1 Exemplary advantages of EOFs over pressure-driven flows are ease of fabrication, flowing control, and precise particle manipulations. A uniform external electric field E applied in the channel induces the movement of an electrolyte solution (with dielectric constant e and viscosity l) relative to a stationary sur- face (with zeta potential f). A relationship between EOF velocity and E is given by the convenient Helmholtz-Smoluchowski (H-S) formula, v EOF ¼ef E=l. 13 In the case of very thin electric double layer (EDL), v EOF for Newtonian fluids appears to slip at the wall represented by this formula, and then one needs to consider the velocity field outside the EDL. It is noted that a plug-like velocity profile in EOF provides reduced dispersion of sample species, realizing the electropho- resis become one of the most effective technologies for chem- ical and biomedical analyses. A design of microfluidic devices requires the effort to consider the flow condition and flow pattern depending on the channel geometry, where a curved rectangular channel can commonly be dealt with. Curved channels become an in- dispensable tool in the chip-based electrophoresis system 46 and the controllable reaction residence to be required for synthetic applications. 7,8 Chaotic advection generating pas- sive mixing at low Reynolds number (Re) is achieved by introducing turn geometry, acting to enhance stretching and breaking of the flow. 9 Although lots of theoretical studies have contributed to elucidating the EOF, those studies were almost confined to an analysis with applying straight channel with uniform cross sections. Simulations of the EOF were performed in various geometries, demonstrating that the linearized Poisson- Boltzmann (P-B) referred to as Debye-Hu ¨ckel (D-H) ansatz electric field can still produce plug velocity predictions for large zeta potentials. 10 Note that a few previous studies have dealt with compensations for turn-induced broadening and how channels can be designed to minimize such effects. Grif- fiths and Nilson 11 derived the analytical method of a closed form expression for the increased band variance induced by a turn. Paegel et al. 12 proposed narrowing the channel width upstream of a turn followed by widening the channel once the turn is completed. Molho et al. 13 tried to come up with opti- mal shapes for low-dispersion geometries using computer simulations. Yang and co-workers explored the transient EOFs in curved channels applying an AC field with a single frequency 14 and suddenly induced DC and AC fields of various frequencies. 15 Later, Woo et al. 16 treated the problem of electro-osmotic flow in a U-turn microchannel to determine the optimal zeta potential distributions for minimal dispersion. In Fig. 1, the external electric field is tangential to the wall and same potential drop occurs over a shorter distance a) Author to whom correspondence should be addressed. Electronic mail: [email protected]. Fax: þ82-2-958-5205. 1070-6631/2011/23(10)/102004/10/$30.00 V C 2011 American Institute of Physics 23, 102004-1 PHYSICS OF FLUIDS 23, 102004 (2011) Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

Transcript of Curvature-induced secondary microflow motion in steady electro...

  • Curvature-induced secondary microflow motion in steady electro-osmotictransport with hydrodynamic slippage effect

    Jin-Myoung Lim and Myung-Suk Chuna)

    Complex Fluids Research Laboratory, National Agenda Research Division, Korea Institute of Science andTechnology (KIST), Seongbuk-gu, Seoul 136-791, South Korea

    (Received 17 February 2011; accepted 9 September 2011; published online 19 October 2011)

    In order to exactly understand the curvature-induced secondary flow motion, the steady electro-

    osmotic flow (EOF) is investigated by applying the full Poisson-Boltzmann/Navier-Stokes equa-

    tions in a whole domain of the rectangular microchannel. The momentum equation is solved with

    the continuity equation as the pressure-velocity coupling achieves convergence by employing the

    advanced algorithm, and generalized Navier’s slip boundary conditions are applied at the hydro-

    phobic curved surface. Two kinds of channels widely used for lab-on-chips are explored with the

    glass channel and the heterogeneous channel consisting of glass and hydrophobic polydimethylsi-

    loxane, spanning thin to thick electric double layer (EDL) problem. According to a sufficiently low

    Dean number, an inward skewness in the streamwise velocity profile is observed at the turn. With

    increasing EDL thickness, the electrokinetic effect gets higher contribution in the velocity profile.

    Simulation results regarding the variations of streamwise velocity depending on the electrokinetic

    parameters and hydrodynamic fluid slippage are qualitatively consistent with the predictions docu-

    mented in the literature. Secondary flows arise due to a mismatch of streamline velocity between

    fluid in the channel center and near-wall regions. Strengthened secondary flow results from increas-

    ing the EDL thickness and the contribution of fluid inertia (i.e., electric field and channel curva-

    ture), providing a scaling relation with the same slope. Comparing with and between the cases

    enables us to identify the optimum selection in applications of curved channel for enhanced EOF

    and stronger secondary motion relevant to the mixing effect. VC 2011 American Institute of Physics.[doi:10.1063/1.3650911]

    I. INTRODUCTION

    The electro-osmotic flow (EOF) is a fundamental electroki-

    netic phenomenon, which plays an important role for improve-

    ments and optimum design of the devices related to lab-on-chips

    (LOCs).1 Exemplary advantages of EOFs over pressure-driven

    flows are ease of fabrication, flowing control, and precise particle

    manipulations. A uniform external electric field E applied in thechannel induces the movement of an electrolyte solution (with

    dielectric constant e and viscosity l) relative to a stationary sur-face (with zeta potential f). A relationship between EOF velocityand E is given by the convenient Helmholtz-Smoluchowski(H-S) formula, vEOF ¼ �e fE=l.1–3 In the case of very thinelectric double layer (EDL), vEOF for Newtonian fluidsappears to slip at the wall represented by this formula, and

    then one needs to consider the velocity field outside the EDL.

    It is noted that a plug-like velocity profile in EOF provides

    reduced dispersion of sample species, realizing the electropho-

    resis become one of the most effective technologies for chem-

    ical and biomedical analyses.

    A design of microfluidic devices requires the effort to

    consider the flow condition and flow pattern depending on

    the channel geometry, where a curved rectangular channel

    can commonly be dealt with. Curved channels become an in-

    dispensable tool in the chip-based electrophoresis system4–6

    and the controllable reaction residence to be required for

    synthetic applications.7,8 Chaotic advection generating pas-

    sive mixing at low Reynolds number (Re) is achieved by

    introducing turn geometry, acting to enhance stretching and

    breaking of the flow.9

    Although lots of theoretical studies have contributed to

    elucidating the EOF, those studies were almost confined to an

    analysis with applying straight channel with uniform cross

    sections. Simulations of the EOF were performed in various

    geometries, demonstrating that the linearized Poisson-

    Boltzmann (P-B) referred to as Debye-Hückel (D-H) ansatz

    electric field can still produce plug velocity predictions for

    large zeta potentials.10 Note that a few previous studies have

    dealt with compensations for turn-induced broadening and

    how channels can be designed to minimize such effects. Grif-

    fiths and Nilson11 derived the analytical method of a closed

    form expression for the increased band variance induced by a

    turn. Paegel et al.12 proposed narrowing the channel widthupstream of a turn followed by widening the channel once the

    turn is completed. Molho et al.13 tried to come up with opti-mal shapes for low-dispersion geometries using computer

    simulations. Yang and co-workers explored the transient

    EOFs in curved channels applying an AC field with a single

    frequency14 and suddenly induced DC and AC fields of

    various frequencies.15 Later, Woo et al.16 treated the problemof electro-osmotic flow in a U-turn microchannel to determine

    the optimal zeta potential distributions for minimal dispersion.

    In Fig. 1, the external electric field is tangential to the

    wall and same potential drop occurs over a shorter distance

    a)Author to whom correspondence should be addressed. Electronic mail:

    [email protected]. Fax: þ82-2-958-5205.

    1070-6631/2011/23(10)/102004/10/$30.00 VC 2011 American Institute of Physics23, 102004-1

    PHYSICS OF FLUIDS 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

    http://dx.doi.org/10.1063/1.3650911http://dx.doi.org/10.1063/1.3650911http://dx.doi.org/10.1063/1.3650911

  • of arc AB on the inner side than on the outer side of arc

    A0B0. Thus, a stronger field is created on the inside edge ofthe channel wall, driving a faster EOF. Furthermore, the

    pressure distribution along the spanwise direction results

    from a gap of the flow path between the inner and outer walls

    of a curved channel. By extending previous work on the

    explicit model for electrokinetic pressure-driven flow in

    straight microchannels,17 Chun and co-workers18 investi-

    gated recently the problem of steady flow in a curved rectan-

    gular channel to figure out the skewed pattern in a

    streamwise velocity with stressing the geometry effect. They

    found the inward skewness in the streamwise axial velocity

    due to a dominant effect of the spanwise pressure gradient

    over the fluid inertia by centrifugal force. Re and nondimen-

    sional curvature (W/RC) are combined into the Dean number(De¼Re

    ffiffiffiffiffiffiffiffiffiffiffiffiffiW=RC

    p) that quantifies the inertial force in the

    curvature.19 As De gets higher enough, the axial velocity

    profile becomes skewed into the outer wall as can be

    observed in the general narrow-bore channels.

    One can presume that the spanwise velocity, pressure

    gradient, and skewed streamwise velocity are formed by

    channel curvature, although those subjects were not quantita-

    tively emphasized in the previous studies. This paper

    addresses the numerical investigation of EOFs in curved rec-

    tangular microchannels of uniform width that provides quan-

    titative information useful in designing either channels or

    devices and their fabrications. In pursuit of more effective

    approach for the pressure-velocity coupling, an alternating

    direction implicit (ADI) method was implemented in the

    SIMPLE (semi-implicit method for pressure-linked equa-

    tions) algorithm.20 Numerical results of the streamwise veloc-

    ity and vorticity change are reported to examine the unique

    behavior of secondary microflow in shallow and deep chan-

    nels with either thick or thin EDL thickness, providing origi-

    nal insights. Unlike the thin EDL situation, the H-S slip

    velocity at the edge of the EDL cannot be applied in the thick

    EDL case which makes the analysis complicated. Our frame-

    work enables to conduct the dissimilar configurations of the

    channel wall taking unequal surface charge and fluid slippage

    with adopting hydrophilic glass and hydrophobic polydime-

    thylsiloxane (PDMS), which has not been fully attempted.

    II. MODEL FORMULATION

    A. The geometry and Navier’s fluid slip at curved wall

    Let us consider a situation for steady-state EOF through

    a curved rectangular microchannel with a uniform curvature.

    The geometry of the channel is schematically presented in

    Fig. 1. Here, a fluid flows through the channel of width Wand height (or, referred to as the depth) H that makes a leftturn. Coordinate transformation is available between the

    global toroidal system and the Cartesian system, where r(¼RCþ x) and y correspond to spanwise and longitudinaldistances. The curvature radius RC ¼ dz=dh is measuredfrom the axis of curvature, and z is the streamwise distancealong the axis of a microchannel.

    While no-slip boundary conditions (BCs) are applied to

    the wall of hydrophilic surface, the hydrodynamic BC at the

    hydrophobic surface undergoes the fluid slippage that is a

    function of the wettability indicated by the contact angle.

    Fluid slippage at the stationary interface can be interpreted

    as the continuity of shear stress r for the tangential velocity

    component, vt¼m � r/fs, with the interfacial friction coeffi-cient fs and the unit normal m directed into the fluid. Com-bining the constitutive equation for the Newtonian fluid

    allows a slip length (sometimes, Navier length) b at a uni-form hydrophobic rigid boundary to linearly relate vt at thewall and the wall shear strain rate, so-called the Navier’s

    fluid slip condition21–23

    vt ¼ b m � rvþ rvð ÞTh i

    : (1)

    Here, b (¼ l/fs) can be regarded as the characteristic lengthnormal to the wall, representing the degree of absence of the

    shearing force. Since the normal velocity component vn¼ v � mð Þ is governed by conservation of mass, vn ¼ 0 for a

    stationary and impermeable solid referred to as the no-

    penetration condition.

    The slip length b is a material parameter, and it is thelocal equivalent distance below the solid surface at which

    the no-slip BC (b¼ 0) would be satisfied if the flow fieldswere extended linearly outside of the physical domain. It is

    possible to derive the streamwise velocity by applying Eq.

    (1) to the curved region in Fig. 1, expressed as

    vh ¼ b1

    r

    @vr@hþ r @

    @r

    vhr

    � �� �: (2)

    Since the normal velocity vr vanishes on the surface, weobtain a generalized slip condition at a surface with RC as

    vh ¼1

    bþ 1

    RC

    � ��1 @vh@r¼ bRC

    bþ RCð Þ@vh@r

    : (3)

    Then, bRC/(bþRC) corresponds to the slip length bC on acurved surface.

    B. Electrokinetic governing equations

    We recall here briefly the background of electrokinetic

    transport. Once the non-conductive and non-neutral

    (charged) surface is in contact with an electrolyte in solution,

    the electrostatic charge would influence the distribution of

    nearby ions and the electric field is established.3,24 For a uni-

    formly charged rectangular channel, the nonlinear P-B equa-

    tion governing the electric potential w field is given as

    FIG. 1. The curved channel with relevant coordinates system and each

    boundary condition.

    102004-2 J.-M. Lim and M.-S. Chun Phys. Fluids 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

  • r2W ¼ j2 sinh W: (4)

    Here, the dimensionless electric potential W denotes Kew/kTand the EDL thickness (or Debye thickness) j�1 is defined

    by j ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2nbK

    2i e

    2=ekTq

    with the valence of i ions Ki. Apply-

    ing symmetric (K-K) electrolytes provides the same valencesof cations and anions. The electrolyte ionic concentration in

    the bulk solution at the electroneutral state nb (1/m3) equals

    to the product of the Avogadro’s number NA (1/mol)and bulk electrolyte concentration (mM). The dielectric

    constant e is given as a product of the dielectric permittivityof a vacuum (¼ 8.854� 10�12 C2/J�m) and the relativepermittivity for aqueous fluid (¼ 78.9), e the elementarycharge, and kT the Boltzmann thermal energy at roomtemperature.

    The equilibrium Boltzmann distribution of the ionic

    concentration of type i (i.e., ni¼ nb exp(�Kiew/kT)) pro-vides a local charge density Kieni. One can determine the netcharge density qe (: Ri Kieni¼Ke(nþ � n� )), as follows:

    qe ¼ Kenb exp �Wð Þ � exp Wð Þ½ � ¼ � 2Kenb sinh W: (5)

    The Dirichlet BCs of constant-potential surface are imposed

    on each side of the wall displayed in Fig. 1: W¼Ws,L atx¼�W/2, W¼Ws,R at x¼W/2, W¼Ws,B at y¼ 0, andW¼Ws,T at y¼H. The subscripts T, B, R, and L correspondto top, bottom, right, and left-sided channel wall.

    Regarding the velocity field for an incompressible New-

    tonian fluid at the steady state, the continuity equation and

    Navier-Stokes (N-S) equation are provided as

    r � v ¼ 0; (6)

    q v � rvð Þ ¼ �rpþ lr2vþ F; (7)

    where q is the density of the fluid. If the flow field exists byaxially applied pressure-driven, the velocity profile in

    straight channel with uncharged (i.e., inert) wall is described

    by the Stokes flow (l52v¼5p). In the case of EOF withoutthe end effects (i.e., @/@h¼ 0), the fluid velocity andthe pressure due to curvature effect are expressed as

    v ¼ vr r; yð Þ; vy r; yð Þ; vh r; yð Þ

    and p¼ p(r,y), respectively.Gravitational forces and the effect of Joule heating are

    neglected. The body force F per unit volume ubiquitouslycaused by the action of electric field E on the net charge den-sity qe(r,y) can be written as F¼qeE.

    The electric field can be given by the linear superposi-

    tion of the external electric potential / imposed by end-channel electrodes and the inherent electric potential w,providing that E¼�5(/þw).3 Originally, / satisfies herethe 2-dimensional Laplace equation (i.e., 52/(r,h)¼ 0) inthe outside region of the EDL. As the BCs, the Dirichlet con-

    ditions can normally be given at the inlet and outlet of the

    channel, and the Neumann conditions are assigned at the

    inside and outside walls of impermeable solid channel

    (@/wall/@r¼ 0). It is acceptable that the spanwise depend-ence of / is much smaller than its streamwise axial depend-ence, whereas the streamwise dependence of w is negligible.Then, streamwise electric field Eh is defined by the externalpotential / as Eh¼� d/(h)/dh.

    With above conditions and pertinent approximations,

    the N-S equation readily reduces to

    r : q vr@vr@rþ vy

    @vr@y� vh

    2

    r

    � �¼�@p

    @rþl 1

    r

    @

    @rr@vr@r

    � ��

    �vrr2þ@

    2vr@y2

    ��qe

    @w@r;

    (8)

    y : q vr@vy@rþ vy

    @vy@y

    � �¼� @p

    @yþ l @

    2vy@r2þ 1

    r

    @vy@r

    þ@2vy@y2

    �� qe

    @w@y

    ;

    (9)

    h : q vr@vh@rþ vy

    @vh@yþvrvh

    r

    � �¼�qe

    r

    @/@hþl 1

    r

    @

    @rr@vh@r

    � ��

    �vhr2þ@

    2vh@y2

    �: (10)

    Utilization of the Cartesian coordinates leads to the follow-

    ing equations:

    x : q vx@vx@xþ vy

    @vx@y� vz

    2

    xþ RC

    � �

    ¼ � @p@xþ l 1

    xþ RC@

    @xxþ RCð Þ

    @vx@x

    � �� vx

    xþ RCð Þ2

    "

    þ @2vx@y2

    �� qe

    @w@x

    ; (11)

    y : q vx@vy@xþ vy

    @vy@y

    � �¼ � @p

    @yþ l @

    2vy@x2þ 1

    xþ RC@vy@x

    þ @2vy@y2

    �� qe

    @w@y

    ; (12)

    z : q vx@vz@xþ vy

    @vz@yþ vxvz

    xþ RC

    � �¼ � RC

    xþ RCqe@/@z

    � �þ

    l1

    xþ RC@

    @xxþ RCð Þ

    @vz@x

    � �� vz

    xþ RCð Þ2þ @

    2vz@y2

    " #: (13)

    For hydrophobic curved surfaces in Fig. 1, each BC is

    applied as

    vxjy¼0 ¼ bB @vx=@yð Þy¼0; vxjy¼H ¼ �bT @vx=@yð Þy¼H; (14)

    vyjx¼�W=2 ¼ bL @vy=@x� �

    x¼�W=2;

    vyjx¼W=2 ¼ �bR @vy=@x� �

    x¼W=2;(15)

    vzjy¼0 ¼ bB @vz=@yð Þy¼0; vzjy¼H ¼ �bT @vz=@yð Þy¼H; (16)

    vzjx¼�W=2 ¼ bCL @vz=@xð Þx¼�W=2;vzjx¼W=2 ¼ �bCR @vz=@xð Þx¼W=2:

    (17)

    Here, bCL and bCR are displayed in Fig. 2 for inner and outer

    walls, respectively, expressed as

    bCL ¼ bL RC �W=2ð Þ= RC �W=2ð Þ � bL½ �; (18)

    bCR ¼ bR RC þW=2ð Þ= RC þW=2ð Þ � bR½ �: (19)

    102004-3 Curvature-induced secondary microflow motion Phys. Fluids 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

  • The straight channel (W/RC! 0) takes a change in Eq. (17),leading to vzjx¼�W=2 ¼ bL @vz=@xð Þx¼�W=2 and vzjx¼W=2¼ �bR @vz=@xð Þx¼W=2. Since the velocity becomes a maxi-mum at the center of the channel, the sign of the velocity

    gradient changes into negative along the axis.

    We substitute Eq. (5) into Eq. (13), and the velocity

    fields are normalized by the characteristic velocity l=qdh soas to elucidate the secondary flow effect along with each ve-

    locity component: Vx ¼ vx qdh=lð Þ, Vy ¼ vy qdh=lð Þ, andVz ¼ vz qdh=lð Þ. Other nondimensionalized parameters areemployed as R0 ¼ RC=dh, X ¼ x=dh, Y ¼ y=dh, Z ¼ z=dh,P ¼ p qdh2=l2

    � �, Cw ¼ 2kTnb sinhW qdh2=l2

    � �, C/ ¼ 2Kienb

    sinhW qdh2wo=l

    2� �

    , and dH=dZ¼ d/=dzð Þ dh=woð Þ. Here, dhis the hydraulic diameter, the external potential difference

    D/ equals EzL with channel length L, and wo is the referenceelectric potential. Then, one obtains the nondimensionalized

    equations of motion, such that

    Vx@Vx@Xþ Vy

    @Vx@Y� V

    2z

    X þ R0

    ¼ � @P@Xþ 1

    X þ R0@

    @XX þ R0ð Þ @Vx

    @X

    � �� Vx

    X þ R0ð Þ2

    þ @2Vx@Y2þ Cw

    @W@X

    ;

    (20)

    Vx@Vy@Xþ Vy

    @Vy@Y¼� @P

    @Yþ @

    2Vy@X2

    þ 1X þ R0

    @Vy@X

    þ @2Vy@Y2þ Cw

    @W@Y

    ;

    (21)

    Vx@Vz@Xþ Vy

    @Vz@Yþ Vx Vz

    X þ R0

    ¼ R0

    X þ R0 C/@H@Z

    � �þ 1

    X þ R0@

    @XX þ R0ð Þ @Vz

    @X

    � �

    � VzX þ R0ð Þ2

    þ @2Vz@Y2

    : (22)

    III. COMPUTATIONAL METHODS AND PARAMETERS

    A precedent requirement is the electric potential profile

    obtained by solving Eq. (4) with the finite difference method

    (FDM), of which basic procedures are analogous to the pre-

    vious work.18 The five-point central difference is taken on

    the left-hand side of Eq. (4) and we adopt the successive

    under-relaxation iterative scheme. The sinh W on the

    right-hand side can be linearized with the Taylor formula. For

    the external electric potential, 52H(X,Z)¼ 0 is treated as aspecial case of Eq. (4) subject to the relevant BCs. Unlike W,the solution for H shows negligibly small gradient in the span-wise direction. The fluid velocity profile can be computed by

    solving the aforementioned N-S equation of Eqs. (11)–(13).

    The well-established SIMPLE algorithm represented by the fi-

    nite volume method (FVM) was employed in this study to

    solve the unknown pressure terms p(x,y) of the momentumequation by using the continuity equation as the pressure-

    velocity coupling. Detailed discussion of the SIMPLE is avail-

    able in the literature.20 The velocity and pressure corrections

    are accomplished by repeating the sequence of operations

    until the convergence is guaranteed. Note that the ADI

    method and staggered grid are adopted when the procedure

    comes to treat x and y components of discretized N-S equationand the continuity equation. As mentioned in the previous

    work,18 the staggered grid is half-node shifted from the origi-

    nal grid, which prevents the failure of pressure estimations

    and assures calculations that are more accurate.

    Computations are performed with the 1-1 type electro-

    lyte of KCl fluid (q¼ 103 kg/m3, l¼ 10�3 Pa�s), where itsEDL thickness j�1 in nanometers is expressed as [fluid elec-trolyte concentration (M)]�1/2/3.278. We consider two kinds

    of configurations with setting each side of channel walls as

    (I) glass (i.e., glass in all sides) and (II) glass in bottom and

    PDMS in the other sides, which are the most widely prepared

    and used in microchip electrophoresis. As the prototype of

    channel cross section, the dimension is set 25 lm wide and 5lm deep for shallow geometry (H/W¼ 0.2, low-aspect-ratio)and 5 lm wide and 25 lm deep for deep geometry (H/W¼ 5,high-aspect-ratio), with the hydraulic radius Rh (¼HW/(HþW)) of 4.17 lm.

    After the mesh refinement with grid convergence test to

    yield satisfactory accuracy level, the computational domain

    is discretized into 151� 151 equally spaced grid points intransverse (x and y) directions. A very fine grid with largernumber of points is employed for the case of thin EDL (i.e.,

    j�1¼ 9.7 nm), because of steep potential profile near thechannel wall. Convergence criteria are specified with the rel-

    ative variation between two successive iterations to be

    smaller than the pre-assigned accuracy order of O(10�8).

    The convergence is fast even for small j�1, and 500 itera-tions were enough for the present channel dimension. The

    under-relaxation coefficients were chosen in the range of

    0.3–0.7. The velocity prediction matches the analytical prob-

    lem in a straight channel, validating our codes. Axial veloc-

    ity profiles at the horizontal cut (i.e., H/2) along thestreamwise direction maintain a fully developed state in the

    entire region of the curved channel except the entrance. In

    order to ensure its axial independency (@vz/@z¼ 0), the ve-locity and pressure fields are analyzed in the transverse sec-

    tion at the middle of the turn.

    To provide an insight for realistic predictions, rigorous

    values of material parameters are required. A number of ex-

    perimental studies conducted to determine the slip length

    b reveal a fact that it depends on the experimental condi-tions, such as solution environment, shear rate, and surface

    roughness.25–27 Here, b of 100 nm is taken for partially non-

    FIG. 2. The fluid slippage at hydrophobic curved surfaces.

    102004-4 J.-M. Lim and M.-S. Chun Phys. Fluids 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

  • wetting PDMS surfaces (contact angle of ca. 98�). Althoughwe need pursue its accurate value, a more detailed discussion

    of this issue is beyond the scope of this study. According to

    Eqs. (18) and (19), the slip lengths for each BC at left-side

    and right-side walls have rightly been corrected, and result-

    ant changes compared to uncorrected slip lengths are esti-

    mated in Fig. 3. It is emphasized that the relative differences

    increase with increasing slip length, presenting scaling rela-

    tions as �ba with the consistent slope a. A higher differencecan be obtained at either inner wall (i.e., left-side wall) or

    deep channel for the specified conditions, where the increas-

    ing ratio of bCL to bCR gets much larger in response to an

    increase of W/RC.The magnitude of dimensionless surface potential Ws

    (¼ws e/kT) at medium pH 7 can be taken by the zeta poten-tial data. According to a great deal of previous reports,28–30

    the zeta potential for 1� 10�4mM 1-1 type electrolyte isfound to vary from� 20 to� 100 mV for PDMS and� 35to� 200 mV for glass surfaces, depending on the electrolyteconcentration and the condition of surface treatment. We

    take here Ws of PDMS as� 0.78 (ws¼� 20 mV),� 1.56(ws¼� 40 mV),� 2.33 (ws¼� 60 mV), and Ws of glassas� 1.36 (ws¼� 35 mV),� 3.11 (ws¼� 80 mV),� 4.67(ws¼� 120 mV) for 1.0, 10�2, and 10�4mM KCl electro-lyte, respectively.

    IV. SIMULATION RESULTS

    A. Spanwise and streamwise velocities in EOF

    Figure 4 presents the spanwise velocity and pressure

    gradient observed in the glass channel with shallow geometry

    for the thick EDL thickness j�1¼ 965 nm. It provides a

    strong electrokinetic interaction with low concentration limit

    of 10�4mM KCl electrolyte that is similar to a state of

    the deionized and distilled water. Note that jRh> 3 is validfor applying the P-B equation with acceptable accuracy.

    The spanwise velocity becomes minimum in top and bottom

    regions, where the negative spanwise velocity means

    FIG. 3. Dependence on the slip length for left-side and right-side walls of

    shallow and deep channels, where b¼bL for left-side and b¼bR for right-side.

    FIG. 4. (Color online) The dimension-

    less profiles of spanwise velocity and

    pressure field for the case of glass with

    H¼ 5 lm, W¼ 25 lm, W/RC¼ 0.5,E¼ 100 V/cm, j�1¼ 965 nm (10�4 mMelectrolyte solution), and jRh¼ 4.3.Here, vx,max¼ 1.92 nm/s and pmax¼ 8.4� 10�10 bar.

    102004-5 Curvature-induced secondary microflow motion Phys. Fluids 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

  • directing to inner wall. The higher pressure in the outer

    region results in the spanwise pressure gradient directing to

    inner wall from outer one. As noted before, this spanwise

    pressure distribution results from a gap of the flow path

    between the inner and the outer walls of a curved channel.

    The fluid in the outer region passes relatively slow, due to a

    longer distance of arc A0B0. Since the PDMS/glass channeldoes not show a qualitative difference, its figures are not pre-

    sented here. However, we can observe that the negatively

    minimum created in the bottom region of the glass surface

    has vanished.

    Figure 5 demonstrates how the streamwise velocity pro-

    file, corresponding to the EOF velocity, changes for each

    case of channels. The magnitude of the streamwise velocity

    was found to be greater in the order of about 104 than that of

    the spanwise velocity, supporting the aforementioned

    assumption of purely streamwise dependence of Ez (i.e., @//@z in Eq. (13)). The De number of this study is estimated as

    FIG. 5. (Color online) The profiles and contour plots of streamwise velocity of glass and PDMS/glass channels for (a) thick and (b) thin EDL thickness, where

    other conditions are same with Fig. 4.

    102004-6 J.-M. Lim and M.-S. Chun Phys. Fluids 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

  • very low values in the order of 10�2� 10�3, from which thebehavior of inward skewness is explained by a dominant

    effect of the spanwise pressure gradient over the fluid inertia

    by centrifugal force, similarly as reported in the pressure-

    driven flow.18,31 Further computations provide that the mag-

    nitudes of spanwise and streamwise velocities as well as

    their ratio increase with increasing either dimension of chan-

    nel cross-section or external electric field. Hence, the ability

    to utilize the role of curvature-induced secondary flow by

    control both the streamwise and spanwise electric fields can

    be linked with the rational design and operation for the axial

    dispersion and the enhanced mixing performance. De also

    increases in response to an increase of Rh, implying the con-tribution of fluid inertia. In the case of thick EDL, the

    streamwise velocity becomes decreased in the PDMS/glass

    channel due to a lower magnitude of surface potential ws atthe other sides of PDMS wall, nonetheless the fluid slippage

    occurs. Against, PDMS/glass channel has higher streamwise

    velocity in the case of thin EDL, except at the bottom of

    glass wall. It allows us to figure out that the electrokinetic

    effect is dominant in the lower jb (thick EDL), while thefluid slippage gets higher contribution in the higher jb(thin EDL).

    In order to justify the contribution of surface potential in

    the thick EDL case (i.e., lower jb), streamwise velocity pro-files obtained at the horizontal cut are compared for each

    case, as shown in Fig. 6. One needs remark again on the fact

    that, regardless of the existence of slip, the higher ws resultsin the higher EOF velocity. Thickening the EDL results in an

    increase of the electrokinetic strength originated from long-

    range electrostatic repulsion. In Fig. 7, we have changed the

    EDL thicknesses j�1 as 9.7 nm (1.0 mM), 96.5 nm (10�2

    mM), and 965 nm (10�4mM), which correspond to jb % 10,1, and 0.1. The fluid slippage in the PDMS/glass channel

    leads to an arched shape at the side walls around the horizon-

    tal cut of the profile, and the linear profile is found in the

    glass channel due to the H-S slip alone.

    The increasing behavior of streamwise velocity in the

    purely glass channel with increasing magnitude of ws is con-sistent with theoretical predictions according to the H-S

    equation for the EOF velocity. The enhancement of stream-

    wise velocity by ws is weaker in the PDMS/glass channel,because the individual effects exerted by PDMS and glass

    surfaces, occupying a larger area of top and bottom in shal-

    low geometry, are almost equivalent. If the purely PDMS

    channel is considered, one can find a qualitative agreement

    with the theoretical analysis on non-wetting charged surfaces

    reported in the literature,32 given by

    vEOF¼�efEz

    l1þjeff b� � ws

    f

    � �¼�ews Ez

    l1þjeff b� �

    : (23)

    Here, jeffb¼ (f�ws)/ws is the ratio of the slip length to theEDL characteristic thickness jeff

    �1 (¼�ws/5wjs), repre-senting contributions due to fluid slip. jeff

    � 1 is given as

    j�1 2Ws=sinh 2Wsð Þð Þ and reduces to j�1 in the case of D-Happroximation.

    B. Vorticity pattern and magnitude of circulation

    It is known that curvature introduces a secondary rota-

    tional flow-field perpendicular to the flow direction (namely

    Dean flow). From the spanwise and longitudinal velocities,

    the stream function S is defined as vr¼ 1=rð Þ @S=@yð Þ andvy¼� 1=rð Þ @S=@rð Þ. Taking the vorticity vector x(¼5� v) yields its h� and z�components, given by

    xh ¼ @vy=@r � @vr=@y; xz ¼ @vy=@x� @vx=@y: (24)

    The vorticity provides information on the flow structure

    inside the channel. For local rectangular coordinates, it is

    obtained after some rearrangement of terms that

    FIG. 6. The variations of streamwise velocity profile at the position of H/2with different values of Ws for glass and PDMS/glass channels, where otherconditions are same with Fig. 4. Dotted curves correspond to virtual cases.

    FIG. 7. The variations of streamwise velocity profile at the position of H/2with different values of EDL thickness for glass and PDMS/glass channels,

    where the values of WPDMS and Wglass depend on j�1. Other conditions are

    same with Fig. 4.

    102004-7 Curvature-induced secondary microflow motion Phys. Fluids 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

  • @2S

    @x2� 1

    xþ RC@S

    @xþ @

    2S

    @y2¼ � xþ RCð Þxz ¼

    xþ RCdh

    � �r2S:

    (25)

    Figure 8 provides the vortices results in two channels

    with shallow and deep geometry, where the magnitude of dif-

    ference between each contour line is all equal as 2� 10�7s�1. The rotational direction is determined by a sign of the

    magnitude of vortex cell, with plus for anticlock-wise and

    minus for clock-wise. Evidently, the positions of maximum in

    the vorticity profile are located with shifting close to inner

    wall. A pair of counter-rotating vortices placed above and

    below the plane of symmetry of the channel is commonly

    found in the glass channel. On the contrary, PDMS/glass

    channel does not show this trend, but the close pattern appears

    to be developed near the bottom wall of the glass surface. In

    light of the EDL effect displayed in Fig. 7, we obtained the

    variations of vorticity pattern in the PDMS/glass channel with

    shallow geometry. In Fig. 9, as j�1 increases, one can see thehigher magnitude of vortex cell as well as the extended vortic-

    ity layer, representing the strengthened secondary flow. In the

    vorticity profile, the splitting pattern into above and below the

    symmetric plane becomes more obvious.

    Estimating the average magnitude of total vorticity

    xzj jh i at a specified cross section of the channel allows oneto predict the circulation strength in the secondary flow.

    Such a flow motion provides useful information in the micro-

    fluidic mixing and the manipulation of particles or cells.33 In

    Fig. 10, xzj jh i increases with increasing either D//L orchannel curvature W/RC, evidencing the contribution of fluidinertia. The magnitude of vorticity depends on the slip

    length, by noting that actual changes gain even in several

    percentages of the relative difference between the corrected

    slip length and uncorrected one. The log-log plot provides a

    result of scaling relations as xzj jh i!(D//L)a with the slopea. Each exponent was obtained as almost the same valuefor the same channel, regardless of the cross sectional geom-

    etry and channel curvature W/RC. The effect of W/RC repre-sents complicated behavior. The shallow geometry keeps

    consistency of the higher xzj jh i value in both channels,except in the glass channel at W/RC¼ 0.1. Moreover, ahigher W/RC leads to the enhancement of axial velocity

    FIG. 8. (Color online) The contour plots of vorticity xzfor (a) glass and (b) PDMS/glass channels with cross

    sections of shallow (lying figures: H¼ 5 lm, W¼ 25lm) and deep (standing figures: H¼ 25 lm, W¼ 5 lm)geometry, where W/RC¼ 1.0, and other conditions aresame with Fig. 4.

    102004-8 J.-M. Lim and M.-S. Chun Phys. Fluids 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

  • relative to one in the straight channel. Figure 10 points out

    that, besides the fluid inertia in terms of D//L and W/RC,both the enhancement of EOF velocity and the secondary

    flow are strictly influenced by aspect ratio and surface prop-

    erties of the curved channel.

    V. CONCLUSIONS

    The curved channel is frequently encountered in the LOC

    system because it provides a convenient way for increasing

    the effective channel length per unit chip area. The behavior

    of steady EOF with secondary motion in curved microchan-

    nels was explored based on a model with P-B/N-S and the nu-

    merical framework developed by advanced scheme. Correct

    BCs of Navier’s slip derived at curved surfaces were applied

    at each hydrophobic wall. Both the glass and the PDMS/glass

    channels were considered in the cases of shallow and deep ge-

    ometry with taking the coupled effect of the EDL and the fluid

    slip for jb % 0.1 – 10.Each channel revealed different features in the velocity

    profile, vorticity pattern, and its magnitude, according to var-

    ious conditions. The streamwise velocity profile was domi-

    nated by the electrokinetic effect in the thick EDL, while the

    fluid slippage gets higher contribution in the thin EDL case.

    Our simulation results regarding the variations of streamwise

    velocity according to j�1, ws, and b were qualitativelymatched with the analytical predictions reported in the litera-

    ture. Strengthened secondary flow results from increasing

    the EDL thickness and the contribution of fluid inertia (D//Land W/RC), with scaling relations as xzj jh i!(D//L)a. It isfound that, in a view of the curvature-induced secondary

    flow, a channel design of the shallow geometry can take

    advantages in either channel if one needs to obtain more

    enhanced mixing. The framework established in this study

    has an impact on applications to the situation in which the

    channel may have arbitrary aspect ratios of cross section as

    well as any kinds of wall condition.

    ACKNOWLEDGMENTS

    This work was supported by the Basic Research Fund

    (No. 20100021979) from the National Research Foundation

    of Korea.

    1T. M. Squires and S. R. Quake, “Microfluidics: Fluid physics at the nanoli-

    ter scale,” Rev. Mod. Phys. 77, 977 (2005).2S. Ghosal, “Electrokinetic flow and dispersion in capillary electro-

    phoresis,” Annu. Rev. Fluid Mech. 38, 309 (2006).3J. H. Masliyah and S. Bhattacharjee, Electrokinetic and Colloid TransportPhenomena (Wiley, Hoboken, 2006).

    4D. J. Harrison, K. Fluri, K. Seiler, Z. Fan, C. S. Effenhauser, and A. Manz,

    “Micromachining a miniaturized capillary electrophoresis-based chemical

    analysis system on a chip,” Science 261, 895 (1993).5M. Ueda, T. Hayama, Y. Takamura, Y. Horiike, T. Dotera, and Y. Baba,

    “Electrophoresis of long deoxyribonucleic acid in curved channels: The

    effect of channel width on migration dynamics,” J. Appl. Phys. 96, 2937(2004).

    6J. Zhu and X. Xuan, “Particle electrophoresis and dielectrophoresis in

    curved microchannels,” J. Colloid Interface Sci. 340, 285 (2009).7W.-H. Tan and S. Takeuchi, “A trap-and-release integrated microfluidic

    system for dynamic microarray applications,” Proc. Natl. Acad. Sci.

    U.S.A. 104, 1146 (2007).

    FIG. 9. The variations of vorticity profile with different values of EDL

    thickness for glass shallow channel, where conditions are same with Fig. 8.

    Vorticity magnitudes given as numbers on each contour line have an order

    of 10�8 s�1.

    FIG. 10. The evolution of average magnitude of total vorticity for shallow

    and deep geometry at different W/RC with variations of external electric field(D//L), where j�1¼ 965 nm.

    102004-9 Curvature-induced secondary microflow motion Phys. Fluids 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

    http://dx.doi.org/10.1103/RevModPhys.77.977http://dx.doi.org/10.1146/annurev.fluid.38.050304.092053http://dx.doi.org/10.1126/science.261.5123.895http://dx.doi.org/10.1063/1.1776625http://dx.doi.org/10.1016/j.jcis.2009.08.031http://dx.doi.org/10.1073/pnas.0606625104http://dx.doi.org/10.1073/pnas.0606625104

  • 8L. Gervais and E. Delamarche, “Toward on-step point-of-care immuno-

    diagnostics using capillary-driven microfluidics and PDMS substrates,”

    Lab Chip 9, 3330 (2009).9H. Chen and J. C. Meiners, “Topological mixing on a microfluidic chip,”

    Appl. Phys. Lett. 84, 2193 (2004).10M. J. Mitchell, R. Qiao, and N. R. Aluru, “Meshless analysis of

    steady-state electro-osmotic transport,” J. Microelectromech. Syst. 9,435 (2000).

    11S. K. Griffiths and R. H. Nilson, “Band spreading in two-dimensional

    microchannel turns for electrokinetic species transport,” Anal. Chem. 72,5473 (2000).

    12B. M. Paegel, L. D. Hutt, P. C. Simpson, and R. A. Mathies, “Turn geome-

    try for minimizing band broadening in microfabricated capillary electro-

    phoresis channels,” Anal. Chem. 72, 3030 (2000).13J. I. Molho, A. E. Herr, B. P. Mosier, J. G. Santiago, T. W. Kenny,

    R. A. Brennen, G. B. Gordon, and B. Mohammadi, “Optimization of

    turn geometries for microchip electrophoresis,” Anal. Chem. 73, 1350(2001).

    14W.-J. Luo, Y.-J. Pan, and R.-J. Yang, “Transient analysis of electro-

    osmotic secondary flow induced by dc or ac electric field in a curved rec-

    tangular microchannel,” J. Micromech. Microeng. 15, 463 (2005).15J.-K. Chen, W.-J. Luo, and R.-J. Yang, “Electro-osmotic flow driven by

    DC and AC electric fields in curved microchannels,” Jpn. J. Appl. Phys.

    45, 7983 (2006).16H. S. Woo, B. J. Yoon, and I. S. Kang, “Optimization of zeta potential dis-

    tributions for minimal dispersion in an electroosmotic microchannel,” Int.

    J. Heat Mass Transfer 51, 4551 (2008).17M.-S. Chun, T. S. Lee, and N. W. Choi, “Microflow analysis of electroki-

    netic streaming potential induced by microflows of monovalent electrolyte

    solution,” J. Micromech. Microeng. 15, 710 (2005).18J. H. Yun, M.-S. Chun, and H. W. Jung, “The geometry effect on steady

    electrokinetic flows in curved rectangular microchannels,” Phys. Fluids

    22, 052004/1-10 (2010).19W. R. Dean, “Note on the motion of fluid in a curved pipe,” Philos. Magn.

    4, 208 (1927).20S. V. Patankar, Numerical Heat Transfer and Fluid Flow (McGraw-Hill,

    New York, 1980).

    21C. L. M. H. Navier, “Mémoire sur les lois du mouvement des fluids,”

    Mem. Acad. Sci. Inst. Fr. 6, 389 (1823).22E. Lauga and H. A. Stone, “Effective slip in pressure-driven Stokes flow,”

    J. Fluid Mech. 489, 55 (2003).23L. Bocquet and J.-L. Barrat, “Flow boundary conditions from nano- to

    micro-scales,” Soft Matter 3, 685 (2007).24D. P. J. Barz and P. Ehrhard, “Model and verification of electrokinetic

    flow and transport in a micro-electrophoresis device,” Lab Chip 5, 949(2005).

    25D. C. Tretheway and C. D. Meinhart, “Apparent fluid slip at hydrophobic

    microchannel walls,” Phys. Fluids 14, L9 (2002).26P. Joseph and P. Tabeling, “Direct measurement of the apparent slip

    length,” Phys. Rev. E 71, 035303R (2005).27P. Huang, J. S. Guasto, and K. S. Breuer, “Direct measurement of slip

    velocities using three-dimensional total internal reflection velocimetry,”

    J. Fluid Mech. 566, 447 (2006).28M. Mala, D. Li, C. Werner, H. Jacobasch, and Y. Ning, “Flow characteris-

    tics of water through a microchannel between two parallel plates with

    electrokinetic effects,” Int. J. Heat Fluid Flow 18, 489 (1997).29M.-S. Chun and S. Lee, “Flow imaging of dilute colloidal suspension in

    PDMS-based microfluidic chip using fluorescence microscopy,” Colloids

    Surf. A 267, 86 (2005).30J. Song, J. F. L. Duval, M. A. C. Stuart, H. Hillborg, U. Gunst, H. F.

    Arlinghaus, and G. J. Vancso, “Surface ionization state and nanoscale

    chemical composition of UV-irradiated poly(dimethylsiloxane) probed by

    chemical force microscopy, force titration, and electrokinetic meas-

    urements,” Langmuir 23, 5430 (2007).31Y. Yamaguchi, F. Takagi, K. Yamashita, H. Nakamura, H. Maeda, K.

    Sotowa, K. Kusakabe, Y. Yamasaki, and S. Morooka, “3-D simulation and

    visualization of laminar flow in a microchannel with hair-pin curves,”

    AIChE J. 50, 1530 (2004).32L. Joly, C. Ybert, E. Trizac, and L. Bocquet, “Liquid friction on charged

    surfaces: From hydrodynamic slippage to electrokinetics,” J. Chem. Phys.

    125, 204715/1-14 (2006).33H. Song, Z. Cai, H. Noh, and D. J. Bennett, “Chaotic mixing in microchan-

    nels via low frequency switching transverse electro-osmotic flow gener-

    ated on integrated microelectrodes,” Lab Chip 10, 734 (2010).

    102004-10 J.-M. Lim and M.-S. Chun Phys. Fluids 23, 102004 (2011)

    Author complimentary copy. Redistribution subject to AIP license or copyright, see http://phf.aip.org/phf/copyright.jsp

    http://dx.doi.org/10.1039/b906523ghttp://dx.doi.org/10.1063/1.1686895http://dx.doi.org/10.1109/84.896764http://dx.doi.org/10.1021/ac000595ghttp://dx.doi.org/10.1021/ac000054rhttp://dx.doi.org/10.1021/ac001127+http://dx.doi.org/10.1088/0960-1317/15/3/006http://dx.doi.org/10.1143/JJAP.45.7983http://dx.doi.org/10.1016/j.ijheatmasstransfer.2008.02.008http://dx.doi.org/10.1016/j.ijheatmasstransfer.2008.02.008http://dx.doi.org/10.1088/0960-1317/15/4/007http://dx.doi.org/10.1063/1.3427572http://dx.doi.org/10.1080/14786440708564324http://dx.doi.org/10.1017/S0022112003004695http://dx.doi.org/10.1039/b616490khttp://dx.doi.org/10.1039/b503696hhttp://dx.doi.org/10.1063/1.1432696http://dx.doi.org/10.1103/PhysRevE.71.035303http://dx.doi.org/10.1017/S0022112006002229http://dx.doi.org/10.1016/S0142-727X(97)00032-5http://dx.doi.org/10.1016/j.colsurfa.2005.06.046http://dx.doi.org/10.1016/j.colsurfa.2005.06.046http://dx.doi.org/10.1021/la063168shttp://dx.doi.org/10.1002/aic.v50:7http://dx.doi.org/10.1063/1.2397677http://dx.doi.org/10.1039/b918213f

    s1cor1s2s2AE1E2E3s2BF1E4E5E6E7E8E9E10E11E12s2BE13E14E15E16E17E18E19E20E21E22s3F2s4s4AF3F4F5E23s4BE24F6F7E25F8s5B1B2B3B4B5B6B7F9F10B8B9B10B11B12B13B14B15B16B17B18B19B20B21B22B23B24B25B26B27B28B29B30B31B32B33