You are on page 1of 13

micromachines

Article
Simulation of the Slip Velocity Effect in an AC
Electrothermal Micropump
Fraj Echouchene 1 , Thamraa Al-shahrani 2 and Hafedh Belmabrouk 1,3, *
1 Electronic and Microelectronics Laboratory, Department of Physics, Faculty of Science of Monastir,
University of Monastir, Monastir 5000, Tunisia; frchouchene@yahoo.fr
2 Department of Physics, College of Science, Princess Nourah Bint Abdulrahman University,
Riyadh 11671, Saudi Arabia; thmalshahrani@pnu.edu.sa
3 Department of Physics, College of Science at Zulfi, Majmaah University, Majmaah 11952, Saudi Arabia
* Correspondence: ha.belmabrouk@mu.edu.sa

Received: 27 July 2020; Accepted: 28 August 2020; Published: 31 August 2020 

Abstract: The principal aim of this study was to analyze the effect of slip velocity at the microchannel
wall on an alternating current electrothermal (ACET) flow micropump fitted with several pairs of
electrodes. Using the finite element method (FEM), the coupled momentum, energy, and Poisson
equations with and without slip boundary conditions have been solved to compute the velocity,
temperature, and electrical field in the microchannel. The effects of the frequency and the voltage,
and the electrical and thermal conductivities, respectively, of the electrolyte solution and the substrate
material, have been minutely analyzed in the presence and absence of slip velocity. The slip velocity
was simulated along the microchannel walls at different values of slip length. The results revealed
that the slip velocity at the wall channel has a significant impact on the flow field. The existence of
slip velocity at the wall increases the shear stress and therefore enhances the pumping efficiency.
It was observed that higher average pumping velocity was achieved for larger slip length. When a
glass substrate was used, the effect of the presence of the slip velocity was more manifest. This study
shows also that the effect of slip velocity on the flow field is very important and must be taken into
consideration in an ACET micropump.

Keywords: slip velocity; AC electrothermal; micropump; microfluidics

1. Introduction
Alternating current electrokinetics (ACE) such as AC electrothermal (ACET), AC electro-osmosis
(ACEO), and dielectrophoresis (DEP) are a promising technology for development in lab-on-a-chip
systems and Bio-MEMS [1–7]. Dielectrophoresis (DEP) has been extensively studied in order to
manipulate particles, capture DNA, and perform cell separation [3]. AC electro-osmosis (ACEO)
has been used for driving, mixing, or pumping in several kinds of biochips or micro-fluidic chips
under low frequencies [2,7]. AC electrothermal flow has been tremendously well-used for biological
applications in low voltages and at higher frequencies, i.e., improvement in quality and performance
of heterogeneous immuno-sensors in microfluidic systems [5], DNA hybridization [1], and on-chip
mixing of fluids [8]. In previous studies, we used the ACET flow combined with obstacles inside the
microchannel in order to improve the binding efficiency of an immunoassay for a biosensor [9,10].
The effect of the slip velocity was ignored. Selmi and Belmabrouk [11] investigated the repercussions
of the AC electroosmosis force on a microfluidic biosensor. The main purpose was to prove how the
AC electroosmosis could modify the flow topology and improve the binding reaction inherent to
the biosensor.

Micromachines 2020, 11, 825; doi:10.3390/mi11090825 www.mdpi.com/journal/micromachines


Micromachines 2020, 11, 825 2 of 13

Pumping of electrolytes using the ACET has important relevance for microfluidic devices [12–15].
The pumping using the AC electrokinetics is characterized by low power consumption and the
electrodes can be inserted inside the microfluidic devices. ACET electrothermal flow is commanded by
an electrostatic force created by the local gradient of the temperature and the electrical conductivity
(σ) and permittivity (ε) of the fluid, while taking into account the interaction induced by the electric
field [16]. The gradients of the physical parameters of the fluid (such as σ, ε, dynamic viscosity µ,
and density ρ) are actually caused by the non-uniform distribution of the temperature which is itself
due to Joule heating [17].
In recent years, ACET flow has garnered increasing interest worldwide regarding fundamental
and applied aspects. ACET flow was initiated by Green et al. [18] in the wake of the experimental
investigation of ACET flow in microfluidic chips containing microelectrode pairs. Hong et al. [19]
proposed a two-dimensional study of ACET flow. The microchannel contains asymmetrical planar
electrode pairs. This configuration is similar to an ACEO pump. In a complementary study,
Hong et al. [20] realized a numerical study of an ACET micropump. The thermal, physical, and electrical
parameters of the fluids are considered variable versus the temperature. Two ACET pump designs
with biobuffers such as lysogeny broth medium at 0.754 S/m and minimal salts medium at 0.145 S/m
were optimized by Wu et al. [21]. High-efficiency pumping was achieved for T array. An improvement
of AC electrothermal micropumping was studied by Du and Manoochehri [22,23]. The capacity of
micro-grooved surface channels has been proved at low voltage. Zhang et al. [24] investigated a
two-phase ACET micropump fitted with a coplanar asymmetric electrode array. Using this device,
the fluid flow rates can be 50% faster than those obtained with a single-phase structure.
Gao and Li [25] studied an ACET micropump integrating an asymmetric spiral microelectrode
pair in a cylindrical microchannel in order to drive rapidly and to mix efficiently, high-conductivity
fluids. Salari et al. [26] proposed an original AC multiple-array electrothermal micropump. In a recent
study, Salari and Dalton [27] introduced a double-array ACET equipment involving two opposite
microelectrode arrangements. This device can be used for simultaneous pumping and mixing. Ren [28]
has investigated the microfluidic pumping of blood flow. The blood is considered a non-Newtonian
fluid. The ACET force is taken into account. The immersed boundary-lattice Boltzmann method
(IB-LBM) and the hybrid boundary element method (BEM) have been used.
The slip velocity phenomenon at the liquid/solid interface, observed in polymer solutions [29,30],
colloidal particles [31], and considerably in red blood cell suspensions [31], plays an important role in
electro-osmotic flow through micro-channels.
Goswami and Chakraborty [32] theoretically studied the electro-osmotic flow in a micro-channel
using a slip velocity condition. Zhao and Yang [33] have investigated the combined effects of
streaming potential and hydrodynamic Navier slip on pressure-driven flows in a micro-channel.
Rivero and Cuevas [34] analyzed fluid/wall slippage in magnetohydrodynamic (MHD) micropumps for
low-Hartmann-number flow. The influence of the slip length on the flow rate was tested. Shit et al. [35]
have studied the effects of slip velocity on rotating electro-osmotic flow in a non-uniform micro-channel.
It has been observed that the no-slip condition is not valid whatever the configuration—
in particular, in many microfluidic and nanofluidic applications whose hydraulic diameter ranges
from 10 µm to 200 µm [36]. The interfacial slip is characterized by a Navier-slip condition and the slip
velocity is written against the normal gradient of the tangential velocity component and the tangential
temperature gradient at the interface.
In the current paper, two-dimensional models available in the literature for describing the AC
electrothermal flow in micropump are revisited. The numerical study considers slip and no-slip
boundary conditions. These models have been used to investigate an ACET flow micropump having
asymmetric planar electrode pairs and using slip and no-slip boundary conditions. The effects
of frequency, voltage, concentration, and substrate material under slip velocity at the wall of the
microchannel were also discussed.
2. Physical Configuration
Figure 1a displays the three-dimensional configuration of an ACEO micropump having
asymmetric pairs of interdigitated electrodes fabricated on the substrate. The consecutive electrodes
are supposed to be in antiphase. Therefore, there is no need to use complex notation and only the real
parts of AC2020,
Micromachines quantities
11, 825 are computed. We suppose that the channel height H is negligible compared 3 of 13
to the length l of the electrode. Considering the periodicity of the asymmetric electrodes, we can,
therefore, reduce the 3-D geometry to a 2D geometry (x, y) [19,20]. Only one pair of electrodes is
2. PhysicalasConfiguration
involved, indicated in Figure 1b.
Figure 1a displaysofthe
The dimensions the computational domain
three-dimensional of the micropump
configuration were taken from
of an ACEO micropump having the literature:
asymmetric
a microchannel
pairs heightelectrodes
of interdigitated H = 50 μm and length
fabricated on Lthe
= 60 μm. TheThe
substrate. substrate (glasselectrodes
consecutive or silica) isare
ofsupposed
thickness
H = 500 μm and the cover (PDMS) is of thickness H = 100 μm. A pair of electrodes
to be in antiphase. Therefore, there is no need to use complex notation and only the real parts
s c of small thickness
of AC
is on the bottom of the microchannel. L and L are the smaller electrode’s
quantities are computed. We suppose that the channel height H is negligible compared to the length
e g width and gap,l
respectively.
of the electrode.LE Considering
and LG are the
the periodicity
larger electrode’s width andelectrodes,
of the asymmetric gap. An we ACcan,
electric field reduce
therefore, is appliedthe
through the electrodes. The nanofluid used is the KCl electrolyte solution. The physical
3-D geometry to a 2D geometry (x, y) [19,20]. Only one pair of electrodes is involved, as indicated in properties of
KCl, the
Figure 1b.cover, and the substrate are summarized in Table 1.

(a)
Hc

Cover

H
Electrodes
L

y
l Hs

x
Substrate
z

(b) y

Cover Hc

L H

LG/2 Le Lg LE LG/2

x
−Vrms Electrodes +Vrms
Hs
Substrate

Figure 1. Geometry of an alternating current electrothermal (ACET) micropump involving the


Figure 1. Geometry of an alternating current electrothermal (ACET) micropump involving the
electrodes. (a) 3D view, (b) 2D view of the reduced computational domain.
electrodes. (a) 3D view, (b) 2D view of the reduced computational domain.
The dimensions of the computational domain of the micropump were taken from the literature:
a microchannel height H = 50 µm and length L = 60 µm. The substrate (glass or silica) is of thickness
Hs = 500 µm and the cover (PDMS) is of thickness Hc = 100 µm. A pair of electrodes of small thickness
is on the bottom of the microchannel. Le and Lg are the smaller electrode’s width and gap, respectively.
LE and LG are the larger electrode’s width and gap. An AC electric field is applied through the
Micromachines 2020, 11, 825 4 of 13

electrodes. The nanofluid used is the KCl electrolyte solution. The physical properties of KCl, the cover,
and the substrate are summarized in Table 1.

Table 1. Physical properties of the fluid, cover, and substrate.

Substrate
Property Fluid (KCl) Cover (PDMS)
Glass Silica
Thermal conductivity λ
0.61 0.18 1.38 150
(W/K·m)
Relative permittivity εr 80 - - -
Electrical conductivity σ (S/m) 0.001–0.1 - - -
Dynamic viscosity µ (Pa·s) 10−3 - - -

3. Physical Model
In order to analyze the electrothermal mechanism, the mathematical formulation requires solving
the Navier–Stokes, heat, and Poisson equations in order to find the velocity, the temperature, and the
electrical field. As mentioned above (Figure 1b), a 2D configuration is considered in this work.
The Poisson equation that governs the electrical potential is solved in Cartesian coordinates:

∂2 ϕ ∂2 ϕ
+ 2 =0 (1)
∂x2 ∂y

where ϕ indicates the root mean square (rsm) voltage, and x, y denotes the Cartesian coordinates.
Since the relative permittivity of the fluid is considered to be independent of the temperature,
the electrical field E is easily obtained from the electrical potential. The boundary conditions related to
the above equation are:

• At the left and right walls (x = 0 and x = L), a periodic condition is applied.
• At the anode and the cathode, the potential is respectively equal to −Vrsm and Vrms where Vrms is
an adjustable value.
• At the other walls (y = H and the remaining part of y = 0), an insulation condition is applied.

On the other hand, by neglecting the heat convection, the energy equation in all domains can be
expressed as:   
∂2 T ∂2 T
λ + + hσE2 i = 0 in the f luid region


∂x 2 ∂y 2

(2)

 ∂ T2 + ∂ T2 = 0 in the cover and the substrate

 2 2

∂x ∂y

where σ and λ are respectively the electrical and thermal conductivities of the fluid. Those quantities
are assumed to be constant. The term σ × E2 is the volumetric source term due to Joule heating,
and the symbol hi represents the time-averaged value. The boundary conditions related to the heat
equation are:

• At the left and right walls (x = 0 and x = L), the inlet and the outlet walls are assumed to be
adiabatic. The heat flux is equal to zero at these boundaries.
• At the walls (y = 0 and y = H), the temperature is assumed to be continuous.
• At the external walls of the cover and the substrate (y = −Hs and y = H + Hc ), an isothermal
condition is applied: Ts = Tc = 298 K.

With regard to the flow velocity field in the microfluidic channel, it is governed by the continuity
and Navier–Stokes equations. The continuity equation is given by:

∂u ∂v
+ =0 (3)
∂x ∂y
Micromachines 2020, 11, 825 5 of 13

To write the Navier–Stokes equations involving the electrothermal force, we assume that the fluid
is incompressible, that the flow is steady, and that the flow occurs at low Reynolds numbers. It is
straightforward to obtain:
 ∂p  
 − ∂x + µ ∂ u2 + ∂ u2 + < Fex > = 0
 2 2

 ∂x ∂y 

(4)

 ∂p

 − + µ 2 + ∂ v2 + < Fey > = 0

 2v 2
∂y ∂x ∂y

Here, p is the pressure, and u and v are the velocity components. The fluid is Newtonian and
its viscosity is µ = 10−3 Pa·s. The temperature rise is less than 2 K. Therefore, the viscosity may be
considered as constant during the simulations. No effects of heat convection or buoyancy take place.
The two electrodes produce an AC electric field which creates the electrothermal force per unit

volume hFe i [9,10,37]. Its expression is derived using a linear perturbation method:

1  (α − β)
  → → → 
→ 1 →
2
= ε gradT. E . E − α|E| gradT

hFe i (5)
2 1 + (ωτ)2 2

α and β represent the expansion coefficients of the fluid permittivity ε and the electrical conductivity
σ, respectively. They are given by:

1 ∂ε 1 ∂σ
α= and β =
ε ∂T σ ∂T

For water, the values of these parameters are α = −0.004 K−1 and β = 0.02 K−1 [19]. ω represents
the electric field angular frequency. τ = ε/σ is the relaxation time of the liquid.
The boundary conditions related to the Navier–Stokes equations are:

• At the left and right walls (x = 0 and x = L), the normal stress is equal to zero at the inlet and
outlet walls.
• At the walls (y = 0 and y = H), a slip or no-slip velocity condition is applied.

It should be mentioned that the Poisson and Navier–Stokes equations are solved only inside the
microchannel, while the heat equation is solved throughout the structure.
In order to analyze the effect of slip velocity on 2D ACE electrothermal flow, we use the following
expressions [38]:
σT µ ∂T
!
(2 − αv ) ∂u ∂v

uslip = u − uw = Λ + + (6)
αv ∂y ∂x w ρT ∂x w

The first term in Equation (6) is due to the velocity gradient normal to the surface, whereas the
second term results from the temperature gradient along the surface. In this study, the tangential
temperature gradient is very low; the second term can be ignored. uw is the solid wall velocity
assumed to be equal to zero, Λ is the molecular mean free path, αv is the tangential momentum
accommodation coefficient (TMAC) varying from 0.0 to 1.0, and σT = 0.75 [39] is the thermal slip
coefficient. Ls = Λ (2 − αv )/αv is the slip length which experiments have shown that can take values
ranging from nanometers to microns [40,41].
The computations were carried out using the finite element method. Applying the Galerkin
weighted residual method, Equations (2)–(4) can be transformed into a matrix form as

KX = Q (7)

where X is the unknown vector; K and Q are the stiffness matrix and load vector, respectively.
First, the electric potential is solved using Equation (1). Then the resulting thermal field is
calculated from Equation (2), and finally the induced flow velocity is solved from the Navier–Stokes
equation, Equations (3) and (4). A triangular mesh containing 15,988 triangular elements was used for
Micromachines 2020, 11, 825 6 of 13

all our simulations. The mesh refinement was especially achieved close to the electrodes. Figure 2
shows this mesh
Micromachines 2020,where
11, x the region near the wall of the microchannel is refined for a better resolution.
6 of 13

Micromachines 2020, 11, x 6 of 13

Figure
Figure 2. 2. 2D
2D unstructured
unstructured mesh
mesh with
withtriangular
triangularelements; thethe
elements; mesh is refined
mesh near near
is refined the wall of
the wall
microchannel.
of microchannel.
Figure 2. 2D unstructured mesh with triangular elements; the mesh is refined near the wall of
4. Results and
4. Results andDiscussion
Discussion
microchannel.

4.1.4.1.
Effects of of
4.Effects
ResultsSlip
SlipVelocity
and VelocityononCharacteristics
Discussion Characteristicsof
of Electrothermal Flow
Electrothermal Flow
Numerical results for
forthe velocity fieldstaking
takingintointo account the electrothermal force and with
4.1.Numerical results
Effects of Slip Velocity theCharacteristics
on velocity fields of Electrothermal account
Flow the electrothermal force and with or
or without
without slipslip velocity
velocityare arepresented
presentedininFigure Figure3;3;computations
computationswere wereperformed
performed forfor a microchannel
a microchannel
containing Numerical
containing a glass results
a glasssubstrate
substrateforwith
the
withvelocity
LLee ==LLggfields
== taking
5 µm
5 μm and
and into
L EL
account
= EL ==L25
G = the
μm. 25electrothermal
µm.
The The force and
electrical
electrical withσor
conductivity
conductivity was σ
G
without slip velocity are presented in Figure 3; computations were performed for a microchannel
equal to 0.01 S/m and the frequency f was 100 kHz, whereas
was equal to 0.01 S/m and the frequency f was 100 kHz, whereas the RMS voltage V rms was 1 Thethe RMS voltage V rms was 1 or 3 V. or 3 V.
containing a glass substrate with Le = Lg = 5 μm and LE = LG = 25 μm. The electrical conductivity σ was
Theslip velocity
slip velocity boundary
boundary condition
condition was wasapplied
applied on on
thethe
toptop andand bottom walls
bottom of the
walls of microchannel
the microchannel for afor
equal to 0.01 S/m and the frequency f was 100 kHz, whereas the RMS voltage Vrms was 1 or 3 V. The
a slip length LLss == 11 μm.
sliplength µm. In Figure
Figure3,3,
Incondition the smaller
theapplied
smaller and larger electrodes areare indicated in in blue and red,
slip velocity boundary was onand larger
the top and electrodes
bottom walls ofindicated
the microchannel blue and
for a red,
respectively.
respectively. TheThe electrothermal flow was generated at the narrow electrode. Then it was oriented to to
slip length Ls =electrothermal
1 μm. In Figure flow
3, thewas generated
smaller at theelectrodes
and larger narrow electrode.
are indicatedThen it was
in blue and oriented
red,
the
the widewide electrodes.
electrodes.
respectively.
Furthermore,
The Furthermore,
electrothermal the
the velocity
flowvelocity
reached
reached
was generated
its
at its
maximal
themaximal
value at the
value atThen
narrow electrode.
electrode
the electrode borders
borders
it was oriented
and
to and
thethe velocity
velocity
the decayedaway
widedecayed
electrodes. away fromthe
from
Furthermore, the bottom
bottom
the velocity wall
wall because
because
reached of
of the
its maximaltheviscous
value ateffects.
viscous effects.
the TheThe
electrode influence
influence
borders of of
and thethe
slipslip velocity
the onon
velocity
velocity the
decayed
the fluid
fluid flowis
away
flow ispresented
from presented
the bottom onFigure
on Figure
wall 3c,dfor
because
3c,d for
of VVrmsviscous
the = 1 V and
and 3 V,
rms = 1 V effects.
respectively.
3 V,The influenceItofis
respectively. clear
Itthe
is clear
that the
slip slip velocity
velocity on the has an
fluid floweffect
is on the distribution
presented on Figure of fluid
3c,d for V velocity.
= 1 V We3 V,
and note an extensive
respectively. It isincrease
clear
that the slip velocity has an effect on the distribution of fluid velocity. We note an extensive increase in
rms
in that
the velocity
the field field due
slip velocity to theeffect
increasethe of the shear stress.
the velocity due to has the an increaseon of thedistribution
shear stress. of fluid velocity. We note an extensive increase
in the(a) velocity
50 field due to the increase of the shear
µm/s stress.
(b) 50 µm/s
0.18 14
(a) 50 µm/s (b) 50 1 µm/s
0.18
0.16 1 14
0.02
1 2 12
0.02
0.02 0.16 1
0.14 2 2 3 12
0.02 0.04
10
No slip velocity

0.14 2 3 4
0.04 0.12
3 10
No slip velocity

0.04 0.06 4 5
0.12
y (µm)

8
y (µm)

0.04 0.1 3 5 6
0.06
0.08 4
y (µm)

8
y (µm)

0.1 6 7
0.06 0.08 0.08 4 5 6
7
0.06 0.1 0.08 8
0.08 5 6 6
0.1 0.06
0.1 8 8
0.1 0.08 68 7 7 4
0.12 0.06 10
0.1 0.08 8 9 8 7 7 4
0.04 9 11 6
0.12
0.14 0.08 7 10 10
0.10.12 0.06 0.04 9 6 5
0.08 0.16
76 10 11 12
13 13 9 12 11 4 2
0.14
0.16 0.14 0.12 5 10 5
0.060.08 0.12 0.16 0.04
0.06 0.02 64 13 12 3
0.04 0.14 0.18
0.16
0.12 523 11
1213 14 10
9 4 2 2
0.06 0.02
0.04 0.02 4 3 1
0.02
0.04 0.18 0.1
0.08
0.06
0.04
0.02 0231 1 141 8 7 165
924
3 2
0 0.02 1 0
0.02 0.1
0.08
0.06
0.04
0.02 0 01 1 1 8 7162
5
4
3 600
0 0 60
0 60
0 x (µm) 60 x (µm)
x (µm) x (µm)

Figure 3. Cont.
Micromachines 2020, 11, x 7 of 13
Micromachines 2020, 11, 825 7 of 13
Micromachines
(c) 50 2020, 11, x (d) 50 7 of 13
µm/s µm/s
0.02 22
0.25 2
(c) 50 (d) 50 20
0.04 µm/s µm/s
0.02 0.02 4 22
0.25 2 2 18
0.06 4
0.04 0.2 20
0.06 0.02 4 6 16
Slip velocity

0.08 2 18
0.04 0.06 4
0.2 4 14
0.06 6 6 16

y )(µm)
y (µm)
Slip velocity

0.08 0.06 0.08 0.15 8


0.04 0.1 12
4 6 14
0.08 6
y (µm)

0.15

y (µm
0.1 0.08 0.06
0.1 0.1 8 8 8
12 10
6
0.12 0.08 10
0.1 0.12 0.1 8
0.12 0.1 8 10 8
0.12 0.12
0.12 0.1 10 10
0.16 0.12 0.12 8 6
0.12 10
10 12 12
0.14 0.16 0.14 0.16 0.05 14 6
10
0.1 0.120.16
0.14
0.22 0.18 8 12 18
12
4
0.2 0.2 0.14 0.140.16 0.05 141618 16 14
0.08 0.1 0.180.16 0.22 0.18 86 12 4
0.06 0.24
0.2 0.2 0.12 0.10.080.060.04 4 141618 18 2016 8 6 4
2
0.04 0.08 0.18
0.26
0.14
0.02
6
20 22 12
10 8 6 2 2
0 0.02 0.06 0.24
0.12
0.10.08
0.060.04 0 04 2 4 2
0.04 0.26 0.02 22 10
0 0 0.02 60 0 0 2
0 60
0 60 0 60
x (µm) x (µm)
x (µm) x (µm)
= 1=V1 V
VrmsVrms Vrms
Vrms = 3=V3 V
Figure 3. Velocity
Figure contours,
3. Velocity without
contours, without slip
slipvelocity for
velocityforforVV rms == 11VV(a)
Vrms (a)and
andVrms
Vrms
= 3=V3 (b)
V (b)
andandwithwith
slip slip
Figure 3. Velocity contours, without slip velocity rms = 1 V (a) and V rms = 3 V (b) and with slip
velocity for Vfor
velocity =rms
rms V 1V= 1(c)
V and Vrms
(c) and = 3= V
Vrms 3 V(d).
(d).The
Theslip
slip length
length andandthethefrequency
frequency areare taken
taken as equal
as equal to Ls to Ls
velocity for V rms = 1 V (c) and V rms = 3 V (d). The slip length and the frequency are taken as equal
= 1μm = 1μmfand f = kHz,
100 kHz, respectively. Theslip
slipvelocity
velocity has
has an
aneffect ononthe distribution of fluid velocity.
to Ls =and
1µm =and100 f = 100respectively. The
kHz, respectively. The slip velocity effect
has an the distribution
effect of fluid
on the distribution velocity.
of fluid
An An extensive
extensive increase
increase in in
the the velocity
velocity field
field due
due to
to the
the increase
increase of
ofthe
theshear
shearstress is
stress noted.
is noted.
velocity. An extensive increase in the velocity field due to the increase of the shear stress is noted.
From Figure 3c,d, it appears that the effects of slip velocity on the flow field depend mainly on
From Figure
From Figure3c,d,
3c,d,ititappears
appearsthatthatthe theeffects
effectsofof slip
slip velocity
velocity ononthethe
flowflow
fieldfield depend
depend mainly
mainly on
on the
the applied voltage. The increase of the velocity field is more and more important when Vrms is large.
the applied voltage. The increase of the velocity field is more and more important when Vrms is large.
applied voltage.
In orderThe increase
to evaluate theofimpact
the velocity field
of the slip is more
condition atand
themore
wall onimportant
the velocity whenfield,Vwe
rms is large.
plotted
In
In order
theorder to evaluate
to evaluate
slip velocity at both the
the impact
impact
upper andofof the slip
the slip
down condition
condition
boundary at
wallsat the
ofthe wall
the wall on the velocity
on the velocity
microchannel field,
with andfield, we
we the
without plotted
plotted
the slip
the slip velocity
slipvelocity
condition at
atforboth
both upper
Vrmsupper and down
andLsdown
= 3 V and boundary
= 1 μmboundary
(Figure 4). Thewalls
walls of the microchannel
slipofvelocity
the microchannel
under the no-slip with and
withconditionwithout
and without was thethe
slip condition
slip condition for
for V
equal to zero Vrms =3V
rms = 3
because and
the
V wall
andLsshear
= s1=μm
L (Figure
1stress
µm 4). The
was set
(Figure to slipFigure
4).zero.
The velocity
slip under
4 also
velocity shows thethat
under no-slip
thethe condition
biggest
no-slip slip was
condition
equal to zero
velocity because
took place the
at wall
the shear
bottom stress
wall was
and set
more to zero.
preciselyFigure
near 4 also
the
was equal to zero because the wall shear stress was set to zero. Figure 4 also shows that the biggest shows
small that
region the biggest
between the slip
slip
electrodes,
velocity took where
place atthe shear
the stress
bottom is
wall important.
and In
more addition,
precisely the slip
near velocity
the
velocity took place at the bottom wall and more precisely near the small region between the electrodes, at
small the top
region wall is
betweenalso the
very
electrodes, small due
where to the
the is viscous
shear effects.
stress isInimportant. In addition, theatslip
where the shear stress important. addition, the slip velocity thevelocity
top wallat is the
alsotopverywall
smallis also
due
very small due to the
to the viscous effects. 20 viscous effects. 0.5

UTop with slip velocity


20 18 UBottomwithout slip velocity 0.5
UTop w
U without slip
ith slip velocity
velocity
Top
16
18 U
UB
with slip velocity
Bottom without slip velocity
ottom
UTop without slip velocity
14
16 UBottomwith slip velocity

12
UBottom (μm/s)
UTop (μm/s)

14

10
12
UBottom (μm/s)
UTop (μm/s)

10
6

8
4

6
2

4 0 0
0 10 20 30 40 50 60
x (μm)
2

Figure 4. Slip velocity profile on the top and bottom walls for Vrms = 3 V and Ls = 1 μm compared to
0 0
the no-slip boundary
0 condition.
10 20 30 40 50 60
x (μm)
Figure 5a shows the transversal velocity distribution at x = 20 μm for Vrms = 1, 2, or 3 V and in the
Figure 4. Slip velocity profile
profile on
on the
thetop
topand
andbottom
bottomwalls
wallsfor rms =
forVVrms = 3 V and Ls == 11 μm
µm compared to
presence or not of a slip velocity. The slip length was fixed to 0.5 μm. It appears that the maximum
the no-slip boundary condition.
velocities occurred at y ≈ 3 μm. This was valid for the three voltages and for both kinematic conditions
(with or without slip). With the increase in voltage, the slip velocity and the maximum velocity
Figure
Figure 5a5a shows
showsthe thetransversal
transversalvelocity
velocitydistribution
distributionatat = μm
x =x 20 20 µm
for for 1, 2,=or
VrmsV=rms 1, 32,Vor 3 Vinand
and the
in the presence or not of a slip velocity. The slip length was fixed to 0.5 µm. It appears
presence or not of a slip velocity. The slip length was fixed to 0.5 μm. It appears that the maximum that the
maximum velocitiesatoccurred
velocities occurred y ≈ was
y ≈ 3 μm.atThis 3 µm. Thisfor
valid was
thevalid
threefor the three
voltages voltages
and for bothand for bothconditions
kinematic kinematic
conditions (with or
(with or without without
slip). Withslip). With theinincrease
the increase voltage,inthe
voltage, the slip and
slip velocity velocity and the maximum
the maximum velocity
Micromachines 2020, 11, 825 8 of 13
Micromachines 2020, 11, x 8 of 13

velocity
increase increase
rapidly.rapidly.
From this From this figure,
figure, we canwe can conclude
conclude that
that the the maximum
maximum velocity
velocity for slipfor
andslipno-slip
and
no-slip and slip conditions is proportional to V 4 (Figure 5b) which is in accordance with the existing
and slip conditions is proportional to 𝑉
Micromachines 2020, 11, x rms
(Figure 8 of 13
5b) which is in accordance with the existing results
results in the
in theincrease
literatureliterature
[19,20]. [19,20].
The The discrepancy
discrepancy in maximum velocity velocity
in maximum betweenbetween theand
the no-slip no-slip and slip
slip conditions
conditions rapidly.when
increases From V this figure, we can conclude that the maximum velocity for slip and no-slip
increases.
increases when Vrms increases. rms
and slip conditions is proportional to 𝑉 (Figure 5b) which is in accordance with the existing results
20in the literature [19,20]. The discrepancy in maximum velocity
20 between the no-slip and slip conditions
10

increases when V rms increases.


umax without slip velocity
18 No-slip V =1V 18 rms
umax with slip velocity
16 20
Slip Velocity Vrms=1V 2016 10

No-slip Vrms uswithout


umax slip velocity
slip velocity
14 18 No-slip V=2V
=1Vrms 18
14
SlipSlip
Velocity Vrms =2V umax with slip velocity
16
Velocity Vrms =1V 16
12

Umax (μUm/s) (μ m/s)


12 us slip velocity
u (µ m /s )

No-slip Vrms
No-slip =3V
Vrms =2V

Us (μ m/s)
14 14
10
SlipSlip Velocity
Velocity Vrms
Vrms =3V=2V 10 5
12

max
12
u (µ m /s )

No-slip Vrms=3V

Us (μ m/s)
8 8
10
6 Slip Velocity Vrms=3V 10 5

8 6
8
4
6 4
6
2 4
42
0 2
20 0
0 10 20 30 40 50 60 70 80
0 0 5 10 15 20 25 30 35 40 45 50
y20(µm)25 0
0 10 20 30 40
V4rms50 60 70
0
80
0 5 10 15 30 35 40 45 50
y (µm) V4rms
(a) (b)
(a) (b)
Figure
Figure 5. 5.(a)(a) Transversal
Transversal velocity
velocity profile
profile x =x 20
at at = 20
µm μm
forfor Vrms==1,1,2,2,and
V rms and3 3V;V;(b)
(b)maximum
maximumvelocity
velocity
Figure 5. (a) Transversal velocity profile4at x = 20 μm for Vrms = 1, 2, and 3 V; (b) maximum velocity
and
andslip velocity
slip velocityprofiles asas
profiles a function ofof
a function 𝑉 with
Vrms withand without
and withoutslip
slipvelocity.
velocity.
and slip velocity profiles as a function of 𝑉 with and without slip velocity.
InInorder
order totograsp
grasp the
theimpact
impact of the applied voltage V rms on the slip velocity at the wall, the slip
In order to grasp the impactofofthe
theapplied voltage
applied voltage VV rms on the slip velocity at the wall, the slip
rms on the slip velocity at the wall, the slip
velocities
velocitieson the
on the top and
toptop
andbottom
bottom walls have been plotted for V
forVrms = 22and 3 V,V,respectively (Figure 6).
velocities on the and bottomwalls
wallshave
have been plottedfor
been plotted Vrms
rms = 2=and
and
3 V,3 respectively
respectively (Figure
(Figure 6). 6).
The
Theprofiles
profiles arearesomewhat
The profiles somewhat
are somewhat similar.
similar.However,
similar.However,
However, when
whenthe
when thevoltage
the voltage
voltage increases,
increases,
increases, the
the themaximum
maximum
maximum slipslip
slipvelocity
velocity
velocity
exhibits a sharp
exhibits
exhibits increase.
a sharp
a sharp The same
increase.The
increase. profiles
Thesame were
same profiles observed
profiles were with
were observed a
observedwith significant
with increase
a significant in the
increase
a significant maximum
in thein the
increase
slip velocity
maximum
maximum when
slip slip the applied
velocity
velocity whenwhenvoltage increased.
theapplied
the applied voltage increased.
voltage increased.
20 0.45
20 0.45
Vrms=2V Vrms=2V
18 Vrms=2V 0.4 Vrms=2V
18 Vrms=3V 0.4 Vrms=3V
Vrms=3V Vrms=3V
16
0.35
16
0.35
14
0.3
(µm/s

14
uTop(µm/s

12 0.3
(µm/s
uTop(µm/s

0.25
12
10 0.25
uBottom

0.2
10
uBottom

8
0.2
0.15
8 6
0.15
0.1
6 4

2 0.1
0.05
4
0 0
0.05
2 0 20 40 60 0 20 40 60

0
x(µm) 0
x(µm)
0 20 40 60 0 20 40 60
Figure Figure
6. Slip6.velocity
Slip velocity profile
profile on the
on the toptop
andand bottom
bottom wallsfor
walls forVVrms ==2 2oror
3V3 and Ls =L1 μm.
V and = 1 µm.
x(µm) x(µm)
rms s

4.2. Slip
4.2.Velocity’s EffectEffect
Slip Velocity’s on theon Average Pumping
the Average Velocity
Pumping Velocity
Figure 6. Slip velocity profile on the top and bottom walls for Vrms = 2 or 3 V and Ls = 1 μm.
The average
The average pumping
pumping velocity
velocity uaveuave
cancanbe
bedefined
defined according
according to:
to:
4.2. Slip Velocity’s Effect on the Average Pumping Velocity
Z H
1
The average pumping velocity uaveucan = definedudy
ave be according to: (8)
H 0
Micromachines 2020, 11, x 9 of 13

1
u = 𝑢𝑑𝑦 (8)
𝐻
Micromachines 2020, 11, 825to
It is worthwhile 9 of 13
discuss how the slip velocity and the applied voltage can affect the average
velocity uave. Figure 7 shows the effects of these two parameters on the average velocity uave. It can
Article
also be seen from Figure 7b that the uave increases almost linearly with the slip length. The maximum
Simulation of the Slip Velocity Effect in an AC
It is worthwhile to discuss how the slip velocity and the applied voltage can affect the average
average velocity was achieved for larger Ls and larger Vrms.
velocity uave . Figure 7 shows the effects of these two parameters on the average velocity uave . It can
Electrothermal Micropump
also be seen from Figure 7b that the uave increases almost linearly with the slip length. The maximum
average
(a)
7
velocity was achieved for larger Ls and larger V rms .(b)
No slip 8
6
Slip velocity Ls=0.5 祄 (b)
Slip velocity Ls=1 祄 6
5

uave (µm/s)
8

4
uave(µm/s)

4
6

uave (µm/s)
3 2
4

2 0
2 2
1.5 3
1
1 2.5
0 2
2 Ls ( 0.5 1.5
0 µm)
1.0 1.5 2.0 2.5 3.0 1.5 0 1 (V) 3
Vrms 2.5
1
Vrms (V) 2
Ls ( 0.5 1.5
µm)
0 1 (V)
Vrms

Average pumping
Figure7.7.(a)(a)Average velocity
pumpingvelocity versusVV
versus
velocityversus rms, , with
Vrms with and
and without slip velocity, for Ls == 0.5
without slip 0.5 μm
µm
Figure pumping rms, with and without slip velocity, for Ls = 0.5 µm
and 1 μm.
µm. (b) (b) Isovalues
Isovalues of uave asaafunction
aveas ofLLs sand
functionof andVV rms .
and 1 µm. (b) Isovalues of uave as a function of Ls and Vrms.rms.
4.3. Effects of Slip Velocity and Electrical Conductivity on Average Velocity
4.3. Effects of Slip Velocity and Electrical Conductivity on Average Velocity
(a) 7Figure 8 shows the influence of electrolyte conductivity on the average velocity u obtained
Figure 8 shows the influence of electrolyte conductivity on the average velocity uave ave obtained
with and without slip velocity for Ls = 0.5 and 1 µm. In all the studied cases, the average velocity
No slip:Glass
with6and without slip velocity for Ls = 0.5 and 1 μm. In all the studied cases, the average velocity uave
Slip :Glass
uave exhibited practically a linear increase against the electrical conductivity. For a fixed value of
No slip:Si
exhibited practically a linear increase against the electrical conductivity. For a fixed value of the
5 Slip :Sithe discrepancies between velocity profile with and without slip velocity become
the conductivity,
conductivity, the discrepancies between velocity profile with and without slip velocity become more
more4 and more important when Ls increases. More precisely, for the same value of the conductivity,
uave(µm/s)

and more important when Ls increases.Glass More precisely, for the same value of the conductivity, the
the discrepancy on the average velocities obtained with and without slip velocity is about unity when
discrepancy on the average velocities obtained with and without slip velocity is about unity when Ls
L = 30.5 µm and increases to 2 for L = 1 µm. These curves prove that the average velocity depends
= s0.5 μm and increases to 2 for Ls = s1 μm. These curves prove that the average velocity depends on
on many parameters. Until now, the effects of the presence or absence of slip velocity, the applied
many2 parameters. Until now, the effects of the presence or absence of slip velocity, the applied
voltage, and the electrolyte conductivitySihave been revealed. The effects of other parameters will be
voltage, and the electrolyte conductivity have been revealed. The effects of other parameters will be
discussed
1 hereafter.
discussed hereafter.
0 (a)
1.0 1.5 2.0 2.5 3.0

V -1 (V)
1 σ=10rms
S/m
1.0x10
-2
Figure0 9. (a) Variation of σ=10average
the S/m velocity versus the rsm voltage for two kinds of substrates. (b)
1.0x10
Axial profile of the bottom wall velocity for Ls = 1 µm and for two kinds of substrates.
uave(µm/s)

-1
1.0x10
-3
σ=10 S/m No slip
-2
1.0x10 Slip
No slip
Slip
-3
1.0x10 No slip
Slip
-4
1.0x10
1.0 1.5 2.0 2.5 3.0
Vrms(V)

Figure 8. The effect of electrolyte conductivity on average velocity without slip velocity and with slip
velocity for (a) Ls = 0.5 µm and (b) Ls = 1 µm.

Micromachines 2020, 11, x; doi: www.mdpi.com/journal/micromachines


Article

Simulation of the Slip Velocity Effect in an AC


Electrothermal Micropump
Micromachines 2020, 11, 825 10 of 13

(b)
4.4. Effects of Slip Velocity and Thermal Conductivity on Average Pumping Velocity
8
To investigate the effect of the thermal conductivity on the average velocity, two substrates were
used. One was made of glass, whereas the second was made 6 of silica. Figure 9a exhibits the average

uave (µm/s)
velocity versus rms voltage for the two substrates when the slip condition is or is not taken into account.
For the high thermal conductivity of the substrate (case of4 silicon λSi = 150 W/K·m), more heat will be
dissipated through this substrate, causing a lower temperature
2
gradient. On the other hand, when λ is
low (case of silicon λGlass = 1.38 W/K·m), the heat flux will be accumulated in the microchannel leading
0
to an increase of temperature gradient. 2
Figure 9b depicts the axial profiles of the bottom velocities
1.5 for the two substrates. When the glass3
1 2.5
is used as a substrate, we notice: 2
Ls ( 0.5 1.5
µm)
1 0 (V)
• A significant increase in average velocity. Vrms

• The effect of the slip velocity is very significant with a rate of increase of 33%. However, this change
Figure
is only7.3%
(a) for
Average pumping velocity versus Vrms, with and without slip velocity, for Ls = 0.5 µm
silicon.
and 1 µm. (b) Isovalues of uave as a function of Ls and Vrms.
• A significant increase of the slip velocity at the bottom wall of the microchannel.

(a) 7

No slip:Glass
6
Slip :Glass
No slip:Si
5 Slip :Si
uave(µm/s)

4 Glass

2
Si

0
1.0 1.5 2.0 2.5 3.0

Vrms(V)

Figure 9. (a) Variation of the average velocity versus the rsm voltage for two kinds of substrates.
Figure 9. (a) Variation of the average velocity versus the rsm voltage for two kinds of substrates. (b)
(b) Axial profile of the bottom wall velocity for Ls = 1 µm and for two kinds of substrates.
Axial profile of the bottom wall velocity for Ls = 1 µm and for two kinds of substrates.
4.5. Effects of Slip Velocity and Frequency on Average Pumping Velocity
Figure 10a shows the average pumping velocity uave at different AC frequencies for V rms = 2 V
and 3 V, with and without slip velocity for the case of Ls = 1 µm, and σ = 0.01 S/m. In this part,
glass was used as a substrate. The diagrams in Figure 10a can be divided into three regions. The first
region is related to low frequencies. The second one is related to intermediate frequencies and called
the transition region. The third part is related to high frequencies. It appears noticeably that uave (f )
takes high values in the low-frequency region while in the high-frequency range it takes much lower
values. On the other hand, uave (f ) takes a constant value independent of the frequency, whether in
the low-frequency range or the high-frequency range. It varies clearly on the frequency only in the
transition zone. For low-frequency region I, the Coulomb force is dominant, while for high-frequency
region III, where the velocity becomes almost zero, the dielectric force is dominant.
The Coulomb force FC , the first term in Equation (5), is due to
Micromachines 2020, 11, x; doi:
the net charge induced by the
www.mdpi.com/journal/micromachines
electrical conductivity gradient in an electrical field which is related to the AC frequency. For low
frequency, the free electric charge is almost saturated, and as a result the dominant Coulomb force is
almost constant and the electrothermal flow velocity is relatively large. On the other hand, when the
frequency increases significantly, the free electric charge density increases, and subsequently the effect
of the Coulomb force decreases. In this high frequency domain, the electrothermal flow velocity is
relatively small.
Micromachines 2020, 11,
Micromachines 2020, 11, 825
x 11 2ofof132

Figure 10. (a) Average pumping velocity versus the AC frequency and (b) slip velocity along the
Figure 10. (a) Average pumping velocity versus the AC frequency and (b) slip velocity along the
bottom wall of microchannel for several AC frequencies.
bottom wall of microchannel for several AC frequencies.
The effects of slip velocity and applied voltage are much larger at low frequencies. For the
low-frequency region I and V rms = 2 V, the average pumping velocity is approximately 1.24 µm/s
when a slip velocity condition is used, which is 33% greater compared to the flow without slip
velocity, which is approximately 0.93 µm/s. In the same low-frequency region, for V rms = 3 V, uave is
approximately 6.28 µm/s when a slip velocity condition is used, which is 34% greater compared to the
flow without slip velocity, which is approximately 4.70 µm/s. The rate of increase depends obviously
on the slip length Ls .
The effect of AC frequency on the slip velocity along the bottom wall is depicted in Figure 10b.
There is a significant reduction of slip velocity when the AC frequency increases. For the low frequency
(f = 10 kHz), the peak of slip velocity at the bottom wall is great where the shear stress is large. For high
frequencies, we notice a significant decrease in the slip velocity.

5. Conclusions
This work presented a physical model to describe the AC electrothermal flow in a microchannel.
The existence or not of a slip velocity at the microchannel walls was considered. The governing
equations have been solved numerically with the finite element method. In addition to the presence or
not of slip velocity, the effects of the electrical conductivity of the liquid, substrate material, voltage,
and frequency have been investigated. The main findings concerned are listed hereafter.

• Shear stress increases with increasing slip length and applied voltage.
• Slip velocity in the region between the two electrodes is very higher than the slip velocity at the
top wall.
• The effect of the slip velocity is very significant with a rate of increase of 33% of the average
pumping velocity when a glass substrate is used. However, for a silicon substrate, this rate is
only 3%.
• The electrothermal flow does not depend on frequency but depends on the slip velocity and the
applied voltage for the frequency range below 100 kHz.

This investigation may be helpful for further studies on ACET micropumps. It is interesting to
note that the slip velocity condition has a significant influence on pumping efficiency.
In future work, the findings obtained by FEM simulations shall be confirmed by experimental data.
Micropump design considerations, based on these findings, could also be more explicitly provided.
Micromachines 2020, 11, 825 12 of 13

Author Contributions: Conceptualization, F.E. and T.A.-s.; methodology, F.E. and H.B.; software, F.E.; validation,
F.E., T.A.-s. and H.B.; formal analysis, T.A.-s.; investigation, F.E.; resources, F.E.; data curation, H.B.;
writing—original draft preparation, F.E., T.A.-s. and H.B.; writing—review and editing, H.B.; visualization, T.A.-s.;
supervision, H.B.; project administration, T.A.-s.; funding acquisition, T.A.-s. All authors have read and agreed to
the published version of the manuscript.
Funding: This research was funded by the Deanship of Scientific Research at Princess Nourah bint Abdulrahman
University through the Fast-track Research Funding Program.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Bottausci, F. DNA hybridization enhancement in microarrays using AC-electrothermal flow. In Proceedings
of the ASME 2008 Fluids Engineering Division Summer Meeting Collocated with the Heat Transfer, Energy
Sustainability, and 3rd Energy Nanotechnology Conferences, Jacksonville, FL, USA, 10–14 August 2008;
pp. 629–636.
2. Cao, J.; Cheng, P.; Hong, F. Applications of electrohydrodynamics and Joule heating effects in microfluidic
chips: A review. Sci. China Ser. E Technol. Sci. 2009, 52, 3477–3490. [CrossRef]
3. Doh, I.; Cho, Y.-H. A continuous cell separation chip using hydrodynamic dielectrophoresis (DEP) process.
Sens. Actuators A Phys. 2005, 121, 59–65. [CrossRef]
4. Lu, Y. AC electrokinetics of physiological fluids for biomedical applications. J. Lab. Autom. 2015, 20, 611–620.
[CrossRef] [PubMed]
5. Sigurdson, M.; Wang, D.; Meinhart, C.D. Electrothermal stirring for heterogeneous immunoassays. Lab Chip
2005, 5, 1366–1373. [CrossRef]
6. Sin, M.L. Advances and challenges in biosensor-based diagnosis of infectious diseases. Expert Rev. Mol. Diagn.
2014, 14, 225–244. [CrossRef] [PubMed]
7. Urbanski, J.P. Fast ac electro-osmotic micropumps with nonplanar electrodes. Appl. Phys. Lett. 2006,
89, 143508. [CrossRef]
8. Ghandchi, M.; Vafaie, R.H. AC electrothermal actuation mechanism for on-chip mixing of high ionic strength
fluids. Microsyst. Technol. 2017, 23, 1495–1507. [CrossRef]
9. Selmi, M. Electrothermal effect on the immunoassay in a microchannel of a biosensor with asymmetrical
interdigitated electrodes. Appl. Therm. Eng. 2016, 105, 77–84. [CrossRef]
10. Selmi, M. Enhancement of the analyte mass transport in a microfluidic biosensor by deformation of fluid
flow and electrothermal force. J. Manuf. Sci. Eng. 2016, 138, 081011. [CrossRef]
11. Selmi, M.; Belmabrouk, H. AC electroosmosis effect on microfluidic heterogeneous immunoassay efficiency.
Micromachines 2020, 11, 342. [CrossRef]
12. Gao, X.; Li, Y. Biofluid pumping and mixing by an AC electrothermal micropump embedded with a spiral
microelectrode pair in a cylindrical microchannel. Electrophoresis 2018, 39, 3156–3170. [CrossRef] [PubMed]
13. Kim, B.J. Development of a microfluidic device for simultaneous mixing and pumping. Exp. Fluids 2009, 46,
85–95. [CrossRef]
14. Lastochkin, D. Electrokinetic micropump and micromixer design based on ac faradaic polarization.
J. Appl. Phys. 2004, 96, 1730–1733. [CrossRef]
15. Ramos, A. Pumping of liquids with ac voltages applied to asymmetric pairs of microelectrodes. Phys. Rev. E
2003, 67, 056302. [CrossRef] [PubMed]
16. Lu, Y. Long-range electrothermal fluid motion in microfluidic systems. Int. J. Heat Mass Transf. 2016, 98,
341–349. [CrossRef] [PubMed]
17. Ramos, A. AC electrokinetics: A review of forces in microelectrode structures. J. Phys. D Appl. Phys. 1998,
31, 2338. [CrossRef]
18. Green, N.G. Fluid flow induced by nonuniform ac electric fields in electrolytes on microelectrodes.
I. Experimental measurements. Phys. Rev. E 2000, 61, 4011. [CrossRef]
19. Hong, F.; Cao, J.; Cheng, P. A parametric study of AC electrothermal flow in microchannels with asymmetrical
interdigitated electrodes. Int. J. Heat Mass Transf. 2011, 38, 275–279. [CrossRef]
20. Hong, F.; Bai, F.; Cheng, P. Numerical simulation of AC electrothermal micropump using a fully coupled
model. Microfluid. Nanofluidics 2012, 13, 411–420. [CrossRef]
Micromachines 2020, 11, 825 13 of 13

21. Wu, J.; Lian, M.; Yang, K. Micropumping of biofluids by alternating current electrothermal effects.
Appl. Phys. Lett. 2007, 90, 234103. [CrossRef]
22. Du, E.; Manoochehri, S. Enhanced ac electrothermal fluidic pumping in microgrooved channels. J. Appl. Phys.
2008, 104, 064902. [CrossRef]
23. Du, E.; Manoochehri, S. Microfluidic pumping optimization in microgrooved channels with ac electrothermal
actuations. Appl. Phys. Lett. 2010, 96, 034102. [CrossRef]
24. Zhang, R.; Dalton, C.; Jullien, G.A. Two-phase AC electrothermal fluidic pumping in a coplanar asymmetric
electrode array. Microfluid. Nanofluidics 2011, 10, 521–529. [CrossRef]
25. Gao, X.; Li, Y. Simultaneous microfluidic pumping and mixing using an array of asymmetric 3D ring electrode
pairs in a cylindrical microchannel by the AC electroosmosis effect. Eur. J. Mech B Fluids 2019, 75, 361–371.
[CrossRef]
26. Salari, A.; Navi, M.; Dalton, C. A novel alternating current multiple array electrothermal micropump for
lab-on-a-chip applications. Biomicrofluidics 2015, 9, 014113. [CrossRef]
27. Salari, A.; Dalton, C. Simultaneous pumping and mixing of biological fluids in a double-array electrothermal
microfluidic device. Micromachines 2019, 10, 92. [CrossRef]
28. Ren, Q. Investigation of pumping mechanism for non-Newtonian blood flow with AC electrothermal forces
in a microchannel by hybrid boundary element method and immersed boundary-lattice Boltzmann method.
Electrophoresis 2018, 39, 1329–1338. [CrossRef]
29. Degré, G. Rheology of complex fluids by particle image velocimetry in microchannels. Appl. Phys. Lett. 2006,
89, 024104. [CrossRef]
30. Tuinier, R.; Taniguchi, T. Polymer depletion-induced slip near an interface. J. Phys. Condens. Matter 2004,
17, L9. [CrossRef]
31. Kok, P.H. Effects of particle size on near-wall depletion in mono-dispersed colloidal suspensions. J. Colloid
Interface Sci. 2004, 280, 511–517.
32. Chakraborty, S. Dynamics of capillary flow of blood into a microfluidic channel. Lab Chip 2005, 5, 421–430.
[CrossRef] [PubMed]
33. Zhao, C.; Yang, C. On the competition between streaming potential effect and hydrodynamic slip effect in
pressure-driven microchannel flows. Colloids Surf. A Physicochem. Eng. Asp. 2011, 386, 191–194. [CrossRef]
34. Rivero, M.; Cuevas, S. Analysis of the slip condition in magnetohydrodynamic (MHD) micropumps.
Sens. Actuators B Chem. 2012, 166, 884–892. [CrossRef]
35. Shit, G. Effects of slip velocity on rotating electro-osmotic flow in a slowly varying micro-channel. Colloids Surf.
A Physicochem. Eng. Asp. 2016, 489, 249–255. [CrossRef]
36. Karniadakis, G.; Beskok, A.; Aluru, N. Microflows and Nanoflows: Fundamentals and Simulation; Springer Science
& Business Media: Berlin/Heidelberg, Germany, 2005.
37. Selmi, M.; Gazzah, M.H.; Belmabrouk, H. Numerical study of the electrothermal effect on the kinetic reaction
of immunoassays for a microfluidic biosensor. Langmuir 2016, 32, 13305–13312. [CrossRef] [PubMed]
38. Lockerby, D.A. Velocity boundary condition at solid walls in rarefied gas calculations. Phys. Rev. E 2004,
70, 017303. [CrossRef]
39. Sharipov, F. Data on the velocity slip and temperature jump coefficients [gas mass, heat and momentum
transfer]. In Proceedings of the 5th International Conference on Thermal and Mechanical Simulation and
Experiments in Microelectronics and Microsystems, EuroSimE 2004, Brussels, Belgium, 10–12 May 2004.
40. Priezjev, N.V.; Darhuber, A.A.; Troian, S.M. Slip behavior in liquid films on surfaces of patterned wettability:
Comparison between continuum and molecular dynamics simulations. Phys. Rev. E 2005, 71, 041608.
[CrossRef]
41. Tabeling, P. Introduction to Microfluidics; Oxford University Press: Oxford, UK, 2005.

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like