On the Origin of Stationary Electroosmotic Flow Driven by AC Fields Around Insulators

V´ıctorCalero,1, ∗ Ra´ulFern´andez-Mateo,1, ∗ Hywel Morgan,1 Pablo Garc´ıa-S´anchez,2 and Antonio Ramos2, † 1School of Electronics and Computer Science, University of Southampton, United Kingdom. 2Depto. Electr´onica y Electromagnetismo. Facultad de F´ısica. Universidad de Sevilla. Avda. Reina Mercedes s/n, 41012. Sevilla (Spain). (Dated: March 2, 2021) Electric fields are commonly used for manipulating particles and liquids in microfluidic systems. In this work, we report stationary electroosmotic flow vortices around micro-pillars induced by AC electric fields in electrolytes. The flow characteristics are theoretically predicted based on the well-known phenomena of surface conductance and concentration polarization around a charged object. The stationary flows arise from two distinct contributions working together: an oscillating non-uniform zeta-potential induced around the pillar and a rectified electric field induced by the ion concentration gradients. We present experimental data in support of the theoretical predictions. The magnitude and frequency dependence of the electroosmotic velocity are in agreement with the theoretical estimates and are significantly different from predictions based on the standard theory for induced-charge electroosmosis, which has previously been postulated as the origin of the stationary flow around dielectric objects. In addition to furthering our understanding of the influence of AC fields on fluid flows, we anticipate that this work will also expand the use of AC fields for flow control in microfluidic systems. Keywords: AC Electrokinetics, Microfluidics, Electroosmosis, Surface Conductance, Concentration Polariza- tion.

I. INTRODUCTION However, recent experimental reports have shown a non-zero time-average slip velocity of electrolytes around Solid surfaces in contact with aqueous electrolytes usu- dielectric pillars in an AC field [5]. Similar stationary ally carry a net surface charge arising from the different (or rectified) flows have been observed around dielectric affinities of cations and anions [1]. This surface charge is corners [6–8] and have been attributed to induced-charge screened by a diffuse ionic layer on the electrolyte side of electroosmosis (ICEO) [9], i.e. electroosmosis generated the interface, and liquid streaming can occur when an ex- by the action of an electric field on the charges in the ternal electric field acts on these charges. This fluid flow diffuse layer induced by the same field [10], see figure is known as electroosmosis (EO) [2] and is a common 1(a). In this work we show that ICEO is not the origin way of driving liquids within capillaries and microfluidic of these flows and demonstrate that a rectified fluid structures [3]. The thickness of the diffuse layer (Debye length) for typical aqueous electrolytes is around tens of (a) (b) Insulating Cylinder Uncharged Metal Cylinder

nanometers or smaller [1]. Thus, at length scales of mi- with Surface Charge

-

- -

-

- -

-

- ++ crometers or larger, the fluid motion that occurs within +

+

+ + - - + - -

+ +

+ + + + - the diffuse layer can be modelled via an effective slip ve- - + + locity tangential to the solid wall, uslip. The Helmholtz- E E Smoluchowski formula relates the slip velocity with the + - - + +- - + + - - + +- - + - - - applied electric field (E) and the (ζ) of the +--+ - + + + interface [1]: + εζ uslip = E, (1) − η FIG. 1. (a) An applied electric field induces charges on a metal (conducting) cylinder that are screened by ions in the where ε and η are, respectively, the electrolyte permi- liquid phase – a diffuse layer is induced around the metal ob- tivitty and . The zeta potential is commonly ject. ICEO occurs due to the action of the applied field on defined as the electrical potential at the slip plane with these diffuse-layer charges. (b) A dielectric cylinder has an the bulk solution [4]. intrinsic surface charge. Upon application of an electric field, surface conductance creates both a non-homogeneous elec- In the case of AC electric fields, eq. (1) predicts an trolyte concentration and zeta potential around the cylinder. oscillating slip velocity with a zero time-average value. In (b), the distance of the positive ions (counterions) to the particle surface changes with position - this indicates that the Debye length varies due to concentration polarization. For both (a) and (b), and using equation (1), the induced zeta ∗ Both authors contributed equally to this work. potentials generate stationary quadrupolar flows around the † Corresponding author: [email protected] cylinder. 2

(a) (b) (c) - - + - + + - + - + - + - + - + + + + + + + + + + + ------

FIG. 2. (a) Negative charges on the dielectric surface attract (positive) counterions. The mean ion concentration increases near the surface leading to a surface current density Js. Charge conservation implies that variations in surface current must be balanced by a counterion flux (F+). (b) Color map showing the perturbation in electrolyte concentration (δc) around a charged cylinder. Darker blue indicates a higher value of c. As a consequence of the concentration polarization, the screening thickness (Debye length) depends on position around the cylinder. The black arrows indicate the direction of the stationary electroosmotic flow. (c) A consequence of concentration polarization is induced free charge that is in addition to the concentration gradients. The rectified electric field due to these charges acts on the diffuse-layer charges and generates an electroosmotic flow.

flow arises from the polarization of modified electrolyte insulating pillar. We show that two distinct mechanisms concentration (i.e. concentration polarization) that acting together give rise to stationary flows: an oscil- results from the surface conductance around a dielectric lating non-uniform zeta-potential and a rectified elec- pillar [2]. tric field induced by the concentration gradients. We also present experimental data of these stationary flows. Figure 1(b) shows a diagram of how concentra- Both the magnitude and frequency dependence of the tion polarization (CP) leads to a non-homogeneous electroosmotic velocity are in agreement with the theo- zeta potential. We have assumed that the pillar car- retical estimates and differ significantly from predictions ries a negative intrinsic surface charge density (qs) based on the theory for induced-charge electroosmosis. which is linked to ζ via the Gouy-Chapman equation qs = 2√2cεeφther sinh(ζ/2φther), with c the electrolyte concentration and φther = kBT/e the thermal voltage (kB is Boltzmann’s constant, T absolute temperature and e the proton charge; φther 25 mV at 20ºC) [1]. A II. THEORY consequence of the surface charge≈ is a local increase in counterion concentration within the diffuse layer and an accompanying enhancement in electrical conductivity Our analysis for an insulating cylinder follows the work that manifests itself as a surface conductance that is of Schnitzer and Yariv [14, 15] for the of in addition to the bulk electrolyte conductivity [2], see charged particles immersed in a symmetrical electrolyte. Figure 2(a). When an external electric field acts on the We extend their analysis to the case of AC signals (see interface, the additional current near the wall leads to appendix A for details). We perform a linear expansion of depletion of electrolyte on one side of the pillar, and a the governing equations for small Dukhin number (Du); corresponding enhancement on the opposite side. Figure the ratio of surface to bulk conductance [2]. The lin- 2(b) depicts the variation in electrolyte concentration earization of the electrokinetic equations for ac voltages that appears near a charged cylinder subjected to an ex- gives rise to a steady velocity field that scales linearly ternal electric field. Thus, the electrolyte concentration with Du but is quadratic with the amplitude of the elec- near the wall is not homogeneous and since qs is fixed, tric field. In the approximation, the electrical potential the Gouy-Chapman relation implies that ζ also varies is written as φ = φ0 + δφ, where φ0 is the potential over the solid wall. The effects of CP on colloids have around a dielectric cylinder for Du = 0, and δφ is the been extensively studied [11–13]; it is responsible for perturbation as a consequence of surface conductance. the well-known low-frequency dispersion of a colloidal φ0 satisfies Laplace equation with zero normal derivative suspension - the so called α-relaxation. as the boundary condition on the cylinder surface. For an applied AC field of magnitude E0 and angular fre- ˜ In the following sections we focus on developing a the- quency ω, φ0 can be written as φ0(t) = e[φ0 exp(iωt)], R oretical model for the effect of CP on the electroosmotic where e[ ] means the real part of the function be- R ··· slip velocity induced by an AC electric field around an tween the brackets and φ˜0 is the potential phasor, which 3 in cylindrical coordinates is written as: for the co-field electrorotation observed in polystyrene microspheres at low frequencies (below 100 Hz).  a2  φ˜0 = E0 r + cos θ. (2) − r Another effect of the concentration polarization is the induction of a net electrical charge arising from the con- Likewise, the salt concentration can be written as c = centration gradients [18, 19]. In fact, current conserva- c + δc, where c is the bulk concentration and δc is the 0 0 tion for a symmetrical electrolyte leads to the following perturbation due to the applied field. For a relatively equation for the perturbation of the electrical potential, small surface conductance, and neglecting advection, δc 2δφ = φ δc/c (see Appendix A). Since δc and is given by the solution of the diffusion equation: ∇ −∇ 0 · ∇ 0 φ0 are oscillating functions with angular frequency ω, δφ 2 D δc = ∂tδc, (3) has a non-zero time-averaged component satisfying: ∇ 2 δφ = (1/2) e[ φ˜ δc˜∗/c ]. (7) where D is the diffusion coefficient of the ions in the ∇ h i − R ∇ 0 · ∇ 0 electrolyte. Figure 2(c) shows a schematic representation of the induced charges associated with the rectified potential For thin diffuse layers, the ion flux balance shown in (δρ = ε 2 δφ ) for a cylinder with negative surface Figure 2(a) can be written as an effective boundary con- charge.− The∇ rectifiedh i electric field corresponding to these dition that incorporates the surface conductance as a pa- induced charges acts on the intrinsic charges of the diffuse rameter. Assuming that negative ions (co-ions) are ex- layer generating an electroosmotic slip velocity with non- pelled from the diffuse layer, the divergence of the surface zero time average given by: current must be balanced by the normal flux of positive ions. The boundary condition for δc on the insulating uslip B = (ε/η)ζ0 s δφ . (8) surface is written as [15]: h i ∇ h i To the best of our knowledge, this contribution to the n δc/c = Du a 2(φ /φ ), (4) slip velocity has not been previously considered. − · ∇ 0 ∇s 0 ther where n is a unit vector normal to the wall, a is the Evaluation of eq. 8 requires a solution for δφ . To 2 cylinder radius and φ0 is the electrical potential. s is this end, we write equation (7) as: h i the Laplacian operator tangential to the wall surface.∇ As mentioned above, Du is the ratio of surface to 2 2 bulk conductance (Du = Ks/aσ, with Ks the surface δφ /φther = Du(E0a/φther) e[f(r) + g(r) cos(2θ)], conductance and σ the electrolyte conductivity). ∇ h i R (9) where δc is also an oscillating function with angular frequency   1 K2(kr) ω and with a frequency-dependent phasor given by: f(r) = 0 K0(kr) 2 , (10) 2K1(ka) − (r/a)   E0a K1(kr) 1 K (kr) δc/c˜ = 2Du cos θ, (5) 0 0 0 g(r) = 0 K2(kr) 2 . (11) − φther kaK1(ka) 2K1(ka) − (r/a) p where k = iω/D, Kn(x) is the modified Bessel func- From this the solution is δφ /φther = 2 h i 0 Du(E0a/φther) e[F (r) + G(r) cos(2θ)]. G(r) is tion of the second kind of order n and Kn(x) = dKn/dx. R The relaxation angular frequency for δc is the reciprocal the only function that contributes to the slip velocity of the typical time in the diffusion equation, τ = a2/D. (the derivation of G(r) is outlined in the appendices). Figure 2(b) shows the solution for δc around a cylinder Thus, the combination of the contributions given by eqs. for ω = 0, i.e. a DC field. (6) and (8) provides the rectified slip velocity for the cylinder. This is of the form u = U sin(2θ), with U h slipi According to the Gouy-Chapman relation, a change in the frequency dependent, maximum slip velocity: local concentration δc implies a perturbation in zeta po- U tential given by δζ/φ = δc tanh(ζ /2φ )/c = (1/2) ζ0 f1 + 2 tanh( ζ0 /2)f2, (12) ther r=a 0 ther 0 (εaE2/2η)Du | | | | and thus, ζ can be written as− shown| in Figure 1(b). For 0 the case of an AC excitation, the inhomogeneous part where we define the functions f1 and f2 as of ζ is an oscillating function with angular frequency ω. f (ωa2/D) = 4 e[G(a)], (13) Using the Helmholtz-Smoluchowski equation, it can be 1 R 2 0 readily shown that a non-zero time-averaged electroos- f2(ωa /D) = e[K1(ka)/(kaK1(ka))]. (14) motic velocity appears, given by: −R The functions f1(x) and f2(x) are plotted in figure 3. For ∗ uslip A = (ε/2η) e[δζ˜ sφ˜ ], (6) zero frequency (k = 0), f = f = 1 and h i R ∇ 0 1 2 where ∗ indicates complex conjugate. Grosse and Shilov U 2 = ζ0 /2 + tanh( ζ0 /2), (15) [16, 17] proposed a similar mechanism as the explanation (εaE0 /2η)Du | | | | 4

while for high frequencies f f and 1  2 U 1 tanh( ζ0 /2). (16) 2 p 2 (εaE0 /2η)Du ∼ 2ωa /D | |

EAAAB8XicbVDLSgMxFL3js9ZX1aWbYBFclZkq6LIogssK9oFtKZn0ThuayQxJRihD/8KNC0Xc+jfu/Bsz7Sy09UDgcM695Nzjx4Jr47rfzsrq2vrGZmGruL2zu7dfOjhs6ihRDBssEpFq+1Sj4BIbhhuB7VghDX2BLX98k/mtJ1SaR/LBTGLshXQoecAZNVZ67IbUjPwgvZ32S2W34s5AlomXkzLkqPdLX91BxJIQpWGCat3x3Nj0UqoMZwKnxW6iMaZsTIfYsVTSEHUvnSWeklOrDEgQKfukITP190ZKQ60noW8ns4R60cvE/7xOYoKrXsplnBiUbP5RkAhiIpKdTwZcITNiYgllitushI2ooszYkoq2BG/x5GXSrFa880r1/qJcu87rKMAxnMAZeHAJNbiDOjSAgYRneIU3RzsvzrvzMR9dcfKdI/gD5/MHsQiQ7Q==

1

5AAAB9XicbVDLSgMxFL1TX7W+qi7dBIvgqsxURZdFNy4r2Ad0xpJJ0zY0yQxJRilD/8ONC0Xc+i/u/Bsz7Sy09UDgcM693JMTxpxp47rfTmFldW19o7hZ2tre2d0r7x+0dJQoQpsk4pHqhFhTziRtGmY47cSKYhFy2g7HN5nffqRKs0jem0lMA4GHkg0YwcZKDxe+SHyBzUiJVEx75YpbdWdAy8TLSQVyNHrlL78fkURQaQjHWnc9NzZBipVhhNNpyU80jTEZ4yHtWiqxoDpIZ6mn6MQqfTSIlH3SoJn6eyPFQuuJCO1kllAvepn4n9dNzOAqSJmME0MlmR8aJByZCGUVoD5TlBg+sQQTxWxWREZYYWJsUSVbgrf45WXSqlW9s2rt7rxSv87rKMIRHMMpeHAJdbiFBjSBgIJneIU358l5cd6dj/lowcl3DuEPnM8f56CSxw== µm

0.8 (a) (b) AAAB83icbVBNSwMxFHxbv2r9qnr0EixCvZTdKuix6MVjBVsL3aVk02wbmmSXJCuUpX/DiwdFvPpnvPlvzLZ70NaBwDDzHm8yYcKZNq777ZTW1jc2t8rblZ3dvf2D6uFRV8epIrRDYh6rXog15UzSjmGG016iKBYhp4/h5Db3H5+o0iyWD2aa0EDgkWQRI9hYya/7ApuxEhmenQ+qNbfhzoFWiVeQGhRoD6pf/jAmqaDSEI617ntuYoIMK8MIp7OKn2qaYDLBI9q3VGJBdZDNM8/QmVWGKIqVfdKgufp7I8NC66kI7WQeUS97ufif109NdB1kTCapoZIsDkUpRyZGeQFoyBQlhk8twUQxmxWRMVaYGFtTxZbgLX95lXSbDe+i0by/rLVuijrKcAKnUAcPrqAFd9CGDhBI4Ble4c1JnRfn3flYjJacYucY/sD5/AHJfJGF (c) AAAB83icbVBNSwMxFHxbv2r9qnr0EixCvZTdKuix6MVjBVsL3aVk02wbmmSXJCuUpX/DiwdFvPpnvPlvzLZ70NaBwDDzHm8yYcKZNq777ZTW1jc2t8rblZ3dvf2D6uFRV8epIrRDYh6rXog15UzSjmGG016iKBYhp4/h5Db3H5+o0iyWD2aa0EDgkWQRI9hYya/7ApuxElk4Ox9Ua27DnQOtEq8gNSjQHlS//GFMUkGlIRxr3ffcxAQZVoYRTmcVP9U0wWSCR7RvqcSC6iCbZ56hM6sMURQr+6RBc/X3RoaF1lMR2sk8ol72cvE/r5+a6DrImExSQyVZHIpSjkyM8gLQkClKDJ9agoliNisiY6wwMbamii3BW/7yKuk2G95Fo3l/WWvdFHWU4QROoQ4eXEEL7qANHSCQwDO8wpuTOi/Ou/OxGC05xc4x/IHz+QPLApGG

AAAB83icbVBNSwMxFHxbv2r9qnr0EixCvZTdKuix6MVjBVsL3aVk02wbmmSXJCuUpX/DiwdFvPpnvPlvzLZ70NaBwDDzHm8yYcKZNq777ZTW1jc2t8rblZ3dvf2D6uFRV8epIrRDYh6rXog15UzSjmGG016iKBYhp4/h5Db3H5+o0iyWD2aa0EDgkWQRI9hYya/7ApuxEhmZnQ+qNbfhzoFWiVeQGhRoD6pf/jAmqaDSEI617ntuYoIMK8MIp7OKn2qaYDLBI9q3VGJBdZDNM8/QmVWGKIqVfdKgufp7I8NC66kI7WQeUS97ufif109NdB1kTCapoZIsDkUpRyZGeQFoyBQlhk8twUQxmxWRMVaYGFtTxZbgLX95lXSbDe+i0by/rLVuijrKcAKnUAcPrqAFd9CGDhBI4Ble4c1JnRfn3flYjJacYucY/sD5/AHMiJGH f1 f2 1.75 mS/m 0.6

2 5.01 mS/m f

1.75 mS/m

, 11.23 mS/m m/s) 5.01 mS/m 1 f 11.23 mS/m 0.4 m/s)

101 0.2 101

0 0.01 0.1 1 10

1/2

AAAB63icbVBNSwMxEJ2tX7V+VT16CRbBi3W3FfRY9OKxgv2AdinZNNuGJtklyQpl6V/w4kERr/4hb/4bs+0etPXBwOO9GWbmBTFn2rjut1NYW9/Y3Cpul3Z29/YPyodHbR0litAWiXikugHWlDNJW4YZTruxolgEnHaCyV3md56o0iySj2YaU1/gkWQhI9hk0oV3WRuUK27VnQOtEi8nFcjRHJS/+sOIJIJKQzjWuue5sfFTrAwjnM5K/UTTGJMJHtGepRILqv10fusMnVlliMJI2ZIGzdXfEykWWk9FYDsFNmO97GXif14vMeGNnzIZJ4ZKslgUJhyZCGWPoyFTlBg+tQQTxeytiIyxwsTYeEo2BG/55VXSrlW9erX2cFVp3OZxFOEETuEcPLiGBtxDE1pAYAzP8ApvjnBenHfnY9FacPKZY/gD5/MHwvyNZQ== Mean Velocity Magnitude ( 2

⁄ Mean Velocity Magnitude (

휔푎 퐷 2 0 FIG. 3. Plot of the functions f1 and f2 versus ωa /D. The 10 102 103 frequency dependence of the slip velocity is contained within 100 2 f (Hz) 3 these functions (eq. (12)). 10 AAAB9XicbVBNT8JAEN3iF+IX6tHLRmKCF9KiiR6JXjhiIh8JrWS7bGHDdtvsTlVs+B9ePGiMV/+LN/+NC/Sg4EsmeXlvJjPz/FhwDbb9beVWVtfWN/Kbha3tnd294v5BS0eJoqxJIxGpjk80E1yyJnAQrBMrRkJfsLY/up767XumNI/kLYxj5oVkIHnAKQEj3QUudoE9QlquP51OesWSXbFnwMvEyUgJZWj0il9uP6JJyCRQQbTuOnYMXkoUcCrYpOAmmsWEjsiAdQ2VJGTaS2dXT/CJUfo4iJQpCXim/p5ISaj1OPRNZ0hgqBe9qfif100guPRSLuMEmKTzRUEiMER4GgHuc8UoiLEhhCpubsV0SBShYIIqmBCcxZeXSatacc4q1ZvzUu0qiyOPjtAxKiMHXaAaqqMGaiKKFHpGr+jNerBerHfrY96as7KZQ/QH1ucP3ESSGA== f (Hz) 10 f (Hz) The rectified slip velocity gives rise to a quadrupolar FIG. 4. (a) Experimental streamlines around an insulating flow of the liquid around the cylinder with a velocity field pillar of 20 µm diameter. The electrolyte conductivity was given by: 1.75 mS/m and the voltage amplitude and frequency were 1600 Vpp and 190 Hz, respectively (see video in supplemen- 1 r2 1  tary material [23]). (b) Numerically calculated streamlines ˆ for the rectified electroosmotic flow around a dielectric cylin- u = U −3 cos(2θ)ˆr + 3 sin(2θ)θ . (17) h i r r der. The color map represents the magnitude of the fluid velocity field. (c) Experimental data for the average fluid ve- This expression gives rise to a velocity amplitude that locity magnitude as a function of signal frequency for three p 2 approximately decays as D/ωa at frequencies larger electrolyte conductivities (KCl in water). The amplitude of than D/a2. ICEO flows around a cylinder are also the applied voltage√ is 1600 Vpp. The velocity approximately quadrupolar (they satisfy eq. (17)) and scale with decays as 1/ f. 2 E0 [20]. However, assuming that the permittivity of the cylinder is much smaller than that of water (true for all experimental cases), the ICEO theory predicts CPEO 2 vslip = (εaE0 /2η)Du( ζ0 /φther + 2 tanh( ζ0 /2φther)), a velocity with a frequency dependence consisting of where we introduce the| acronym| CPEO for| concentra-| a plateau followed by a decay around frequencies of tion polarization electroosmosis. Using values of the the order of the reciprocal of the charge relaxation parameters obtained from our experiments, and typical time of the electrolyte (σ/(2πε) 0.3 3 MHz for our values for the permittivity of PDMS (εd = 2.8ε0), experimental conditions. This frequency≈ − is several orders λD 30 nm and Du [0.01, 0.1], the ratio between the of magnitude greater than the typical frequencies for ≤ ∈ICEO CPEO two slip velocities is vslip /vslip [0.004, 0.04], which the flows observed in our experiments [9, 21]). It is also demonstrates that CPEO stationary∈ flows completely enlightening to compare the slip velocities predicted by dominate over ICEO for charged dielectric obstacles. both theories. According to ICEO theory for dielectric objects [20, 22], the maximum induced zeta potential for a DC field with amplitude E0 is δζICEO = 2(εd/ε)E0λD, where εd is the permittivity of the dielectric object III. EXPERIMENTAL RESULTS AND and λD the thickness of the diffuse layer, i.e. the Debye length. Thus, from eq. (1), the maximum COMPARISON WITH THEORY time-averaged slip velocity for an AC field of amplitude ICEO 2 E0 is vslip = (εd/η)λDE0 . On the other hand, The CPEO flows were experimentally validated using the maximum slip velocity for the stationary flows is simple microfluidic devices made using standard soft 5 lithography. Channels (1 cm long, 50 µm tall, 200 µm (8). We also imposed periodicity for the velocity and wide) containing a periodic square array of cylindrical pressure fields on the boundaries of the unit cell. micropillars (20 µm diameter) were made from PDMS. The separation between the centers of neighboring pillars Figure 4(b) shows the streamlines obtained for the was 40 µm. The channel was filled with KCl electrolyte rectified electroosmotic flow around a post in the with conductivities σ = 1.75, 5.01, 11.23 mS/m. For periodic array. As expected, the time-averaged flow { } flow visualization, fluorescent nanoparticles (500 nm pattern shows four recirculating vortices as found diameter) were dispersed in the electrolyte and imaged experimentally. Fig. 5 shows a comparison between with a fluorescence microscope. Before experiments, the the experimental data from figure 4(c) and numerical PDMS channels were primed for at least 30 minutes results with the following parameters: ζ = 110 mV 0 − with a solution of 0.1% Pluronics F-127 – a non-ionic (for the lowest conductivity) [24] and E0 = 96.25 kV/m. surfactanct that adsorbs onto the PDMS walls to For the other conductivities, ζ0 was calculated from the minimise sticking of the tracer particles. Metal needles Gouy-Chapman relation for fixed surface charge; these were inserted at the inlet and outlet of the channel and values are in agreement with the measurements in [25]. AC voltages applied, with amplitude up to 2000 Vpp The experimental data in fig. 5 are scaled with u0Du, and frequencies up to 1 kHz. Videos of the fluorescent 2 where u0 = εaE0 /2η and Du is calculated by assuming particles were analyzed with Particle Image Velocimetry a fixed surface conductance of 1 nS – independent of (PIV) (described in section S3 of supplementary material electrolyte conductivity. This is a typical value obtained [23]). from experimental data for the electrokinetic properties of submicrometre latex particles [26] and it is larger Figure 4(a) shows results of the superposition of that the estimation obtained for Ks when using the experimental images showing the trajectories of the theory of the diffuse layer. The difference is attributed fluorescent particles in an electrolyte. Four symmetrical to the contribution to surface conductance arising from flow rolls are seen, as predicted from the theory for the a layer of mobile ions adsorbed on the wall [12] – the rectified electroosmotic flow field around a cylinder. so-called Stern layer. Experimental trends are correctly Figure 4(c) shows the mean value of the velocity mag- described by the theoretical model: the rectified fluid nitude as a function of the applied AC frequency for three electrolyte conductivities. Following convention, the frequency in experiments f is related to the angular 100 frequency by ω = 2πf. The data were obtained by 1.75 mS/m averaging the fluid velocity magnitude within a unit 5.01 mS/m cell of the periodic array of cylinders. Error bars 11.23 mS/m 1.75 mS/m (numerical) correspond to the dispersion in the measurements within 10-1 5.01 mS/m (numerical) six different unit cells. Importantly, a strong decrease 11.23 mS/m (numerical) 0 of the stationary velocity is observed for frequencies Du) 0 u/u

around and above tens of Hertz, in accordance with u 2 ( a /D 0.1 s, i.e. the timescale introduced by the -2 AAAB+XicbVDLSgNBEJz1GeNr1aOXwSDES9yNgh6DevAYwTwgWZbZyWwyZPbBTE8wLPkTLx4U8eqfePNvnCR70MSChqKqm+6uIBVcgeN8Wyura+sbm4Wt4vbO7t6+fXDYVImWlDVoIhLZDohigsesARwEa6eSkSgQrBUMb6d+a8Sk4kn8COOUeRHpxzzklICRfNvW52XtO11gT5Dd6cmZb5ecijMDXiZuTkooR923v7q9hOqIxUAFUarjOil4GZHAqWCTYlcrlhI6JH3WMTQmEVNeNrt8gk+N0sNhIk3FgGfq74mMREqNo8B0RgQGatGbiv95HQ3htZfxONXAYjpfFGqBIcHTGHCPS0ZBjA0hVHJzK6YDIgkFE1bRhOAuvrxMmtWKe1GpPlyWajd5HAV0jE5QGbnoCtXQPaqjBqJohJ7RK3qzMuvFerc+5q0rVj5zhP7A+vwBxbKTGg== diffusion≈ equation, eq. (3). Additionally, the velocity for u/ 10 2 1/2 a given conductivity approximately decays as 1/√f, also u = "aE /2⌘ AAAB63icbVBNSwMxEJ2tX7V+VT16CRbBi3W3FfRY9OKxgv2AdinZNNuGJtklyQpl6V/w4kERr/4hb/4bs+0etPXBwOO9GWbmBTFn2rjut1NYW9/Y3Cpul3Z29/YPyodHbR0litAWiXikugHWlDNJW4YZTruxolgEnHaCyV3md56o0iySj2YaU1/gkWQhI9hk0oV3WRuUK27VnQOtEi8nFcjRHJS/+sOIJIJKQzjWuue5sfFTrAwjnM5K/UTTGJMJHtGepRILqv10fusMnVlliMJI2ZIGzdXfEykWWk9FYDsFNmO97GXif14vMeGNnzIZJ4ZKslgUJhyZCGWPoyFTlBg+tQQTxeytiIyxwsTYeEo2BG/55VXSrlW9erX2cFVp3OZxFOEETuEcPLiGBtxDE1pAYAzP8ApvjnBenHfnY9FacPKZY/gD5/MHwvyNZQ== AAACCHicbVDLSgNBEJz1GeMr6tGDg0HwFHejoBchKILHCOYBSVx6J51kyOzsMjMbCCFHL/6KFw+KePUTvPk3Th4HTSxoKKq66e4KYsG1cd1vZ2FxaXllNbWWXt/Y3NrO7OyWdZQohiUWiUhVA9AouMSS4UZgNVYIYSCwEnSvR36lh0rzSN6bfoyNENqStzgDYyU/c5D4Lr2k9R4ojDUXkaRAb3z3IX+Sr6MBP5N1c+4YdJ54U5IlUxT9zFe9GbEkRGmYAK1rnhubxgCU4UzgMF1PNMbAutDGmqUSQtSNwfiRIT2ySpO2ImVLGjpWf08MINS6Hwa2MwTT0bPeSPzPqyWmddEYcBknBiWbLGolgpqIjlKhTa6QGdG3BJji9lbKOqCAGZtd2obgzb48T8r5nHeay9+dZQtX0zhSZJ8ckmPikXNSILekSEqEkUfyTF7Jm/PkvDjvzsekdcGZzuyRP3A+fwDtu5gB 0 0 in agreement with theoretical predictions. We have also f = D/(2⇡a2) performed experiments using a single post (rather than AAAB+3icbVDLSsNAFL3xWesr1qWbwSLUTU2ioBuhqAuXFewD2hgm00k7dPJgZiKW0F9x40IRt/6IO//GaZuFth64cDjnXu69x084k8qyvo2l5ZXVtfXCRnFza3tn19wrNWWcCkIbJOaxaPtYUs4i2lBMcdpOBMWhz2nLH15P/NYjFZLF0b0aJdQNcT9iASNYackzS4FnoUt0c1JxuglD+ME59syyVbWmQIvEzkkZctQ986vbi0ka0kgRjqXs2Fai3AwLxQin42I3lTTBZIj7tKNphEMq3Wx6+xgdaaWHgljoihSaqr8nMhxKOQp93RliNZDz3kT8z+ukKrhwMxYlqaIRmS0KUo5UjCZBoB4TlCg+0gQTwfStiAywwETpuIo6BHv+5UXSdKr2adW5OyvXrvI4CnAAh1ABG86hBrdQhwYQeIJneIU3Y2y8GO/Gx6x1ychn9uEPjM8fXZaSEQ== 0 10-3 an array). The observed flow is completely analogous to 1 2 the flow observed with the array of posts. An example 10 f/f 10 AAAB7HicbVBNS8NAEJ3Ur1q/qh69LBbBU02qoMeiF48VTFtoQ9lsN+3SzSbsToRS+hu8eFDEqz/Im//GbZuDtj4YeLw3w8y8MJXCoOt+O4W19Y3NreJ2aWd3b/+gfHjUNEmmGfdZIhPdDqnhUijuo0DJ26nmNA4lb4Wju5nfeuLaiEQ94jjlQUwHSkSCUbSSH11EPbdXrrhVdw6ySrycVCBHo1f+6vYTlsVcIZPUmI7nphhMqEbBJJ+WupnhKWUjOuAdSxWNuQkm82On5MwqfRIl2pZCMld/T0xobMw4Dm1nTHFolr2Z+J/XyTC6CSZCpRlyxRaLokwSTMjsc9IXmjOUY0so08LeStiQasrQ5lOyIXjLL6+SZq3qXVZrD1eV+m0eRxFO4BTOwYNrqMM9NMAHBgKe4RXeHOW8OO/Ox6K14OQzx/AHzucPHeiOOg== f/f 0 of streamlines is shown in section S1 of supplementary 0 material [23]. FIG. 5. Comparison between experimental data in figure 4(c) and numerical calculations for a periodic array of di- The stationary electroosmotic velocity for a periodic electric cylinders. The solid lines correspond to the aver- array of dielectric cylinders as used experimentally was age velocity magnitude determined from simulations with E = 96.25 kV/m and ζ = −110 mV for the lowest con- calculated using the commercial finite element solver 0 0 ductivities. For the other conductivities, ζ0 was calculated COMSOL Multiphysics. The governing equations for from the Gouy-Chapman relation for fixed surface charge. φ (Laplace equation), δc (eq.(3)) and δφ (eq. (7)) 2 0 h i Frequencies are nondimensionalized with f0 = D/(2πa ). were solved in a 2D domain corresponding to a unit Velocities are scaled with the product of a typical velocity 2 cell centered on a cylinder, see Figure 4(b). Boundary (u0 = εaE0 /2η) and the Dukhin number. Note that the nu- conditions on the cylinder surface were n φ = 0, merical curves do not exactly have the same frequency de- · ∇ 0 eq.(4) and n δφ0 = 0. Periodicity was imposed on pendence and, therefore, they do not collapse onto a master the boundaries· ∇h of thei unit cell. The velocity field within curve. The reason is that the relative contribution of the two the unit cell satisfies Stokes equations with slip velocity mechanisms to rectified electroosmosis varies with ζ0, which on the cylinder wall given by the sum of eqs. (6) and in turn decreases with electrolyte conductivity. 6 velocity decreases with electrolyte conductivity and, (ωε/σ 1, with ε and σ, the liquid permittivity and significantly, its frequency dependence is close to 1/√f. conductivity, respectively). In the following we use a The magnitude of the velocity is clearly overestimated dimensionless formulation [14]: length is nondimension- by the numerical simulations by a factor of around 4.5. alised with a typical distance a (radius of the pillar in It is important to note that the channels were primed our experiments), potential with the thermal voltage with the surfactant Pluronics before each experiment φther = kBT/e (kB is Boltzmann’s constant, T absolute which is known to significantly reduce electroosmotic temperature and e elementary charge), time with 2 2 velocities [27]. η/εEther where Ether = φther/a, pressure with εEther, and concentrations with typical salt concentration c0. Thus, diffusion constants are nondimensionalised with 2 2 2 εa Ether/η, velocities with εEthera/η, and the typical 2 2 2 IV. CONCLUSIONS is Re = ρmεEthera /η , with ρm the liquid mass density. The surface charge on the dielectric We have presented a mathematical model that pre- is nondimensionalised with εφther/λD, where λD is the dicts stationary fluid flow of electrolytes induced by Debye length. AC electric fields around charged dielectric objects. We show experimental data for electrolyte flow around The nondimensional equations for the conservation of micropillars in a microfluidic channel and demonstrate positive and negative ions in the bulk electrolyte (outside that the experimental trends are in agreement with the electrical double layer) are, respectively, numerical calculations for typical values of surface ∂c conductance and ζ-potential for PDMS. ( c φ c) + α u c + α = 0, (A1) ∇ · − ∇ − ∇ + · ∇ + ∂t The magnitude of the fluid velocity is overestimated by the numerical calculations which can be attributed to the ∂c (c φ c) + α−u c + α− = 0, (A2) fact that adsorption of Pluronics to PDMS significantly ∇ · ∇ − ∇ · ∇ ∂t reduces electroosmosis. The theoretical model correctly describes the amplitude and frequency dependence of where electro-neutrality has been taken into account so the rectified flow, in contrast with previous work that at- that the concentrations of positive and negative ions are tributes the flow around dielectric structures to classical equal c+ = c− = c. The nondimensional parameters α+ ICEO in an AC field [9, 28]. Beyond the fundamental and α− are the reciprocals of the nondimensional dif- interest in these stationary flows, we anticipate that this fusion constants D+ and D− of positive and negative model will expand the understanding of the behavior of ions, respectively. Adding and substracting equations systems that employ AC electric fields for micro- and (A1) and (A2), we get nanoparticle manipulation [29, 30] and lead to ways ∂c of locally controlling fluid flow using dielectric structures. D 2c = u c + , (A3) ∇ · ∇ ∂t

∂c  ACKNOWLEDGMENTS (c φ) = γ + u c , (A4) ∇ · ∇ ∂t · ∇

PGS and AR acknowledge financial support by ERDF where D = 2/(α+ +α−) and γ = (α+ α−)/2. Equation and Spanish Research Agency MCI under contract (A3) is the diffusion equation for the− salt concentration, PGC2018-099217-B-I00. with D a nondimensional ambipolar diffusion constant. Equation (A4) can be read as the equation for the electrical potential where γ is a parameter that controls the ion mobility asymmetry. Appendix A: Theory for Surface Conductance and Electroosmosis The boundary conditions on the surface of the charged dielectric object are the following [14]. Zero normal flux We consider a negatively charged dielectric object of co-ions (anions in our case) immersed in a binary electrolyte subjected to an ∂φ ∂c AC electric field and follow the theory developed by c = 0. (A5) Schnitzer and Yariv [14], but extending it to the AC ∂n − ∂n case. This theory is a thin-double-layer analysis of the The normal flux of counter-ions (cations in our case) electrokinetic equations for charged dielectric solids equate the surface divergence of EDL cation fluxes that considers surface conduction effects. We assume that the frequency ω of the applied electric field is low ∂φ ∂c c = 2Du 2(φ + ln c), (A6) enough to consider the EDL is in quasi-equilibrium − ∂n − ∂n ∇s 7 where Du is the Dukhin number defined as cylinder is given in equation (2)); and d) the velocity iωt Du = (1 + 2α+) qs λD, with qs and λD the nondi- field and pressure are, respectively, u0(t) = e[u˜0e ] | | iωt R mensional intrinsic surface charge and Debye length, and p0(t) = e[˜p0e ], where u˜0 andp ˜0 satisfy respectively. Here the normal derivative is from the R 2 dielectric to the electrolyte. The Dukhin number is iωReu˜0 = p˜0 + u˜0, u˜0 = 0, (A10) −∇ ∇ ∇ · defined in the literature [4] as Du = K /(σa), where s ˜ Ks is the surface conductance. This expression is fully with boundary condition u˜0 = ζ0 φ0 at the dielectric surface. The analytical solution∇ for the velocity and equivalent to the definition Du = (1 + 2α+) qs λD for symmetrical electrolyte with equal diffusivities| | and pressure generated around a cylinder is shown below large zeta potential (the case we deal with in this (Appendix D). work). The supplementary material contains a detailed derivation of the equivalence of the two definitions [23]. The concentration δc satisfies In previous equations, we assumed that the particle is 2 ∂δc nonpolarizable, i.e. the charge induced in the EDL by D δc = u0 δc + . (A11) ∇ · ∇ ∂t the external electric field is negligible. This condition will be examined later. Here we neglect the advection term in order to obtain an analytical solution. For this to be valid, the Peclet (P e) Liquid velocity and pressure satisfy the Navier- must be negligibly small (P e = U0L/D¯, where equation for negligible Reynolds number: U0 is a typical velocity, L is the characteristic length for the concentration gradients and D¯ the dimensional dif- ∂u Re = p + 2u + 2φ φ, u = 0, (A7) fusivity of the ions). For DC, the characteristic length ∂t −∇ ∇ ∇ ∇ ∇ · is a, the pillar radius. For AC fields, the characteris- where the Coulomb term is present because gradients tic length is the smaller of a or the diffusion penetration p ¯ ¯ of concentration can lead to induced charge in the depth (the penetration depth is D/ω¯, with D andω ¯ bulk, through equation (A4), and the time derivative of dimensional quantities). P e is negligible when either U0 velocity is present because it may not be negligible for or L is very small. In our experimental system, this im- high frequency. plies that either U0 is much smaller than 200 µm/s or ω 2π10 rad/s. The boundary condition for equation  The boundary conditions on the charged dielectric sur- (A11) at the dielectric surface is face are (i) the wall is impermeable, u n = 0, and (ii) · ∂δc there is a slip velocity generated at the EDL [31] = Du 2(φ ). (A12) − ∂n ∇s 0 us = ζ sφ 4 ln (cosh(ζ/4)) sc, (A8) ∇ − ∇ Since φ0 is an oscillating function in time with angular frequency ω and we neglect the advection term, we find where the zeta potential ζ is related to the intrinsic iωt charge by [1] a solution of δc that is of the form δc = e[δce˜ ]. The complex function δc˜ satisfies R qs = 2√c sinh(ζ/2). (A9) D 2δc˜ = iωδc.˜ (A13) ∇ Here we see that perturbations of salt concentra- tion lead to perturbations of zeta potential, i.e. with boundary condition δζ = δc tanh(ζ0/2)/c0. ∂δc˜ − = Du 2(φ˜ ). (A14) ∂n − ∇s 0 Following Schnitzer and Yariv [15], we now perform a linear expansion in the parameter Du (which is small We further assume that positive and negative ions have in our case): φ = φ0 + δφ, c = c0 + δc, ζ = ζ0 + δζ, the same mobility so that γ = 0 in equation (A4). The u = u0 + δu, with δφ, δc, δζ, δu of the order of Du. In potential δφ satisfies then addition, the equations will be simplified in order to 2 obtain an analytical solution. δφ + φ0 δc = 0, (A15) ∇ ∇ · ∇ Consider an applied AC electric field of nondimen- with boundary condition on the dielectric surface sional angular frequency ω. The solution at order zero ∂δφ ∂δc (Du = 0) is: a) the salt concentration is unperturbed, = . (A16) ∂n ∂n c0 = 1; b) the zeta potential in the absence of concentra- tion polarization is uniform and given by the relation Equation (A15) and b.c. (A16) implies that the general qs = 2 sinh(ζ0/2); c) the potential in the liquid bulk solution for δφ has three Fourier independent frequency iωt  iωt 2iωt is φ0(t) = e[φ˜0e ], where φ˜0 is the potential phasor components: 1, e , e . We are interested in the that satisfiesR Laplace’s equation with boundary condition time-independent component, because this contributes ∂φ˜0/∂n = 0 at the dielectric surface (the solution for a to the rectified velocity, as shown below. Thus we solve 8

2 δφ + φ0 δc = 0. δc given by eq. (5) in nondimensional form and evalu- ∇ h i h∇ · ∇ i ated on the cylinder surface, the maximum induced zeta The time-averaged velocity and pressure at the first potential due to the concentration perturbation is order satisfy ζCP = 2DuE0 tanh( ζ0 /2), (B2) 2δu + 2δφ φ = δp, δu = 0. (A17) | | ∇ h∇ ∇ 0i ∇ ∇ · We see that for thin EDL (λ 1) and a non- The Coulomb term has zero time average, D negligible Dukhin number (i.e. Du> λ ), ζ /ζ 2δφ φ = 0. Effectively, the Laplacian of δφ D IC CP 0 (ε /ε )(λ /Du) 1 for common . Thus, it∼ is h∇is equal∇ toi φ δc because of equation (A15). It d w D 0 justifiable to neglect ζ compared with ζ . has Fourier−∇ components· ∇ 1, e2iωt and when multiplied IC CP by φ0 the Coulomb term has Fourier components eiωt∇, e3iωt . Appendix C: Functions for analytical solution The boundary conditions for the time-averaged veloc- ity on the dielectric surface at the first order are δu n = 0 The functions F (r) and G(r) are obtained from the and the time-averaged slip velocity · equations:

δus = ζ0 s δφ + δζ sφ0 4 ln (cosh(ζ0/4)) s δc = 1 ∂  ∂F  ∇ h i h ∇ i − ∇ h i r = f, (C1) = ζ s δφ + δζ sφ , (A18) r ∂r ∂r ··· 0∇ h i h ∇ 0i where the last equality comes from δc = 0. Here h i δζ = δc tanh(ζ0/2). 1 ∂  ∂G 4G − r = g, (C2) r ∂r ∂r − r2 Summary of approximations in the model: 1. Thin EDL and highly charged surface, as required with boundary conditions of ∂F/∂r = ∂G/∂r = 0, both for the Schnitzer-Yariv model; for r = 1 and r . Only the function G(r) contributes to the rectified→ slip ∞ velocity (equation 12). We solve for 2. Frequencies much smaller than the electrolyte G using Green’s function that satisfies equation (C2) but charge relaxation frequency, ensuring that the EDL with source equal to the Dirac delta function δ(r r0) is in quasi-equilibrium; and with boundary conditions of zero radial derivatives− at r = 1 and r . The solution for G is then 3. High frequencies or weak electric fields so that ad- → ∞ 0 04 0 vection of ions can be neglected; −2 R r g(r )(r +1) 0 2 −2 R ∞ g(r ) 0 G(r) = r 0 dr (r + r ) 0 dr . − 1 4r − r 4r 4. Small Du to justify the use of a linear expansion; (C3) According to equation (12), we require the value of G 5. Equal ion diffusivities so that the charge induced evaluated on the cylinder surface G(1): in the bulk due to different diffusivities can be ne- 0  0 0  glected; R ∞ g(r ) 0 1 R ∞ K0(kr ) K2(kr ) 0 G(1) = r0 dr = K0 k r03 r0 dr . − 1 2 4 1( ) 1 − 6. Negligible induced charge in the EDL (valid for (C4) common dielectrics). The integral (C4) is evaluated using Mathematica.

Appendix B: Negligible induced charge Appendix D: Oscillating velocity and pressure around a cylinder At this point we examine the charge induced by the applied field on a dielectric cylinder. First we compare This appendix derives the oscillating velocity at zero the maximum induced zeta potential from the induced order. This does not affect the stationary electroosmotic charge on a dielectric and also from the change in con- velocity since we have neglected the advection term in centration due to surface conduction. Both induced zeta equation (5). The solution to equation (A10) for the potentials are maximum at zero frequency. According to case of a cylinder can be found by writing [32] [20] the maximum induced zeta potential on a dielectric cylinder due to induced charge is p˜0 u˜0 = ∇ + (ψzˆ), (D1) εd −iωRe ∇ × ζIC = 2 E0λD, (B1) εw with equations forp ˜0 and ψ given by, respectively, where εd and εw are the dielectric constants of the solid and water, respectively. From δζ = δc tanh(ζ /2) with 2p˜ = 0, 2ψ = iωReψ (D2) − 0 ∇ 0 ∇ 9

Imposing boundary conditions of zero velocity at infinity leads to and slip velocity on the cylinder surface u˜0 = 2ζE0 sin θθˆ K1(α) cos θ p˜0 = iωRe2ζE0 0 , (D3) K1(α) + αK1(α) r

K1(αr) ψ = 2ζE0 0 sin θ, (D4) − K1(α) + αK1(α)

where α = √iωRe.

[1] R.J. Hunter, Introduction to Modern Colloid Science (2010). (Oxford University Press, 1993). [14] Ory Schnitzer and Ehud Yariv, “Macroscale description [2] J. Lyklema, Fundamentals of Interface and Colloid Sci- of electrokinetic flows at large zeta potentials: nonlin- ence (Academic Press Limited, 1995). ear surface conduction,” Physical Review E 86, 021503 [3] H.A. Stone, A.D. Stroock, and A. Ajdari, “Engineering (2012). flows in small devices: Microfluidics toward a lab-on-a- [15] Ory Schnitzer and Ehud Yariv, “Nonlinear electrophore- chip,” Annu. Rev. Fluid Mech. 36, 381411 (2004). sis at arbitrary field strengths: small-Dukhin-number [4] Angel V. Delgado, Fernando Gonzalez-Caballero, R.J. analysis,” Physics of Fluids 26, 122002 (2014). Hunter, L.K. Koopal, and J. Lyklema, “Measurement [16] Constantino Grosse and Vladimir Nikolaievich Shilov, and interpretation of electrokinetic phenomena (IUPAC “Theory of the low-frequency electrorotation of technical report),” Pure and Applied Chemistry 77, polystyrene particles in electrolyte solution,” Jour- 1753–1802 (2005). nal of Physical Chemistry 100, 1771–1778 (1996). [5] Victor Calero, Pablo Garcia-Sanchez, Antonio Ramos, [17] Constantino Grosse and Vladimir Nikolaievich Shilov, and Hywel Morgan, “Electrokinetic biased deterministic “Theory of the low frequency electrorotation of disperse lateral displacement: Scaling analysis and simulations,” particles in electrolyte solution,” Colloids and Surfaces A: Journal of Chromatography A , 461151 (2020). Physicochemical and Engineering Aspects 140, 199–207 [6] Yuval Eckstein, Gilad Yossifon, Avraham Seifert, and (1998). Touvia Miloh, “Nonlinear electrokinetic phenomena [18] Veniamin G. Levich, Physicochemical Hydrodynamics around nearly insulated sharp tips in microflows,” Jour- (Prentice-Hall, 1962). nal of colloid and interface science 338, 243–249 (2009). [19] John S. Newman and Karen E. Thomas-Alyea, Electro- [7] Sunil Kumar Thamida and Hsueh-Chia Chang, “Nonlin- chemical systems (Wiley-IEEE, 2004). ear electrokinetic ejection and entrainment due to polar- [20] Todd M Squires and Martin Z Bazant, “Induced-charge ization at nearly insulated wedges,” Physics of Fluids 14, electro-osmosis,” Journal of Fluid Mechanics 509, 217– 4315–4328 (2002). 252 (2004). [8] Paul Takhistov, Ksenia Duginova, and Hsueh-Chia [21] G Yossifon, I Frankel, and T Miloh, “Macro-scale de- Chang, “Electrokinetic mixing vortices due to electrolyte scription of transient electro-kinetic phenomena over po- depletion at microchannel junctions,” Journal of colloid larizable dielectric solids,” Journal of fluid mechanics and interface science 263, 133–143 (2003). 620, 241 (2009). [9] Matan Zehavi, Alicia Boymelgreen, and Gilad Yossifon, [22] Cunlu Zhao and Chun Yang, “Analysis of induced-charge “Competition between induced-charge electro-osmosis electro-osmotic flow in a microchannel embedded with and electrothermal effects at low frequencies around a polarizable dielectric blocks,” Physical Review E 80, weakly polarizable microchannel corner,” Physical Re- 046312 (2009). view Applied 5, 044013 (2016). [23] See Supplemental Material at [URL will be inserted by [10] M. Z. Bazant and T. M. Squires, “Induced-charge elec- publisher]. trokinetic phenomena: theory and microfluidic applica- [24] Alice Sze, David Erickson, Liqing Ren, and Dongqing tions,” Physical Review Letters 92, 066101 (2004). Li, “Zeta-potential measurement using the smoluchowski [11] S.S Dukhin and Vladimir Nikolaievich Shilov, Dielectric equation and the slope of the current–time relationship phenomena and the double layer in disperse systems and in electroosmotic flow,” Journal of colloid and interface polyelectrolytes (John Wiley and Sons, 1974). science 261, 402–410 (2003). [12] VN Shilov, AV Delgado, F Gonzalez-Caballero, and [25] Mario A Saucedo-Espinosa and Blanca H Lapizco- C Grosse, “Thin double layer theory of the wide- Encinas, “Refinement of current monitoring methodol- frequency range dielectric dispersion of suspensions of ogy for electroosmotic flow assessment under low ionic non-conducting spherical particles including surface con- strength conditions,” Biomicrofluidics 10, 033104 (2016). ductivity of the stagnant layer,” Colloids and Surfaces A: [26] Irina Ermolina and Hywel Morgan, “The electrokinetic Physicochemical and Engineering Aspects 192, 253–265 properties of latex particles: comparison of electrophore- (2001). sis and dielectrophoresis,” Journal of colloid and interface [13] Nataliya A Mishchuk, “Concentration polarization of in- science 285, 419–428 (2005). terface and non-linear electrokinetic phenomena,” Ad- [27] Martina Viefhues, S Manchanda, T-C Chao, Dario Ansel- vances in Colloid and Interface Science 160, 16–39 metti, Jan Regtmeier, and Alexandra Ros, “Physisorbed 10

surface coatings for poly (dimethylsiloxane) and quartz opments, applications and guidelines,” Lab on a Chip 9, microfluidic devices,” Analytical and bioanalytical chem- 2551–2567 (2009). istry 401, 2113 (2011). [34] M. Raffel, C. E. Willert, F. Scarano, C. J. K¨ahler,S. T. [28] Qianru Wang, Naga Neehar Dingari, and Cullen R Buie, Wereley, and J. Kompenhans, Particle Image Velocime- “Nonlinear electrokinetic effects in insulator-based di- try: A Practical Guide, 3rd ed. (Springer International electrophoretic systems,” Electrophoresis 38, 2576–2586 Publishing AG, 2018). (2017). [35] C. E. Willert and M. Gharib, “Digital particle image ve- [29] Blanca H Lapizco-Encinas, Blake A Simmons, Eric B locimetry,” Experiments in Fluids 10, 181–193 (1991). Cummings, and Yolanda Fintschenko, “Insulator-based [36] W. Thielicke, “PIVlab - particle image velocimetry (PIV) dielectrophoresis for the selective concentration and sep- tool, matlab central file exchange.” . aration of live bacteria in water,” Electrophoresis 25, [37] W. Thielicke and E. J. Stamhuis, “PIVlab towards user- 1695–1704 (2004). friendly, affordable and accurate digital particle image ve- [30] Victor Calero, Pablo Garcia-Sanchez, Carlos Honrado, locimetry in MATLAB,” Journal of Open Research Soft- Antonio Ramos, and Hywel Morgan, “Ac electrokinetic ware 2 (2014), 10.5334/jors.bl. biased deterministic lateral displacement for tunable par- [38] C. D. Meinhart, S. T. Wereley, and J. G. Santiago, ticle separation,” Lab on a Chip 19, 1386–1396 (2019). “A PIV algorithm for estimating time-averaged veloc- [31] DC Prieve, JL Anderson, JP Ebel, and ME Lowell, “Mo- ity fields,” Journal of Fluids Engineering 122, 285–289 tion of a particle generated by chemical gradients. part 2. (2000). electrolytes,” Journal of Fluid Mechanics 148, 247–269 [39] M. G. Olsen and R. J. Adrian, “Out-of-focus effects on (1984). particle image visibility and correlation in microscopic [32] J Milton Andres and Uno Ingard, “Acoustic streaming particle image velocimetry,” Experiments in Fluids 29, at low reynolds numbers,” The Journal of the Acoustical S166–S174 (2000). Society of America 25, 932–938 (1953). [33] R. Lindken, M. Rossi, S. Große, and J. Westerweel, “Micro-particle image velocimetry (PIV): Recent devel-