Enhancement of Mass Transport and Separation of Species by Oscillatory Electroosmotic Flows
Enhancement of Mass Transport and Separation of Species by Oscillatory Electroosmotic Flows
Enhancement of Mass Transport and Separation of Species by Oscillatory Electroosmotic Flows
1. Introduction
Biochips are the essential elements of bioengineering designs and are often
applied in drug diagnosis, tumour cell analyses, DNA sequencing systems and
biological/environmental-monitoring sensors. Because of this important role in
its own discipline, the technical integration and the design of ‘Lab-on-a-Chip’
devices become a major research interest both academically and industrially.
Micro-fluidic components such as micro-channels, micro-mixers, micro-pumps
and micro-reaction-chambers are commonly implemented in biochip designs. In
order to build biochips with high efficiency and quality, the physical factors and
phenomena that will affect the performance of the micro-fluidic components are
needed to be studied carefully and thoroughly. For example, electrokinetic (EK)
effects, originated from the presence of the electrical double layer (EDL) at the
contact interface of the working liquid and the solid substrate, are long discussed
in literature (Hunter 1981; Probstein 1994; Karniadakis & Beskok 2002), and are
* Author for correspondence (cllai@ntu.edu.tw).
Table 1. Nomenclature.
the presence of the EDL. This phenomenon is called ‘electroosmosis,’ which was
first observed by Reuss who performed a series of experiments on EK effects in
1809 (Probstein 1994). The resulting flow is called the electroosmotic flow (EOF)
with a shear layer formed between the motionless Stern layer and the electric-
active diffusive region. Many novel studies about EOF have been carried out
recently. For instance, laminar oscillatory EOFs were described and discussed in
detail by Dutta & Beskok (2001b) and Erickson & Li (2003); unstable oscillatory
EOFs and their application to the design of high efficiency micro-mixers were
presented by Oddy et al. (2001); in later studies performed by Lin et al. (2004)
and Suresh & Homsy (2004), the physical mechanisms of electrokinetic
instability (EKI) were investigated systematically and thoroughly.
Laminar oscillatory flow fields are remarkable because of the dramatic mass
dispersion enhancement when concentration gradients of miscible mass species
are introduced into the flow field. The basic mechanism that contributes to the
above process is the non-uniform velocity distribution between the wall
boundaries, which require no-slip of the fluid, and the bulk flow region; radial
or lateral concentration gradients across the flow region are therefore generated.
This is the basic idea of ‘Taylor dispersion’ first proposed by Taylor (1953). Aris
(1956, 1960) then extended the idea into more general flow situations by the
method of moments, and broadened the adequate parametric ranges and usage of
the dispersion concept. Recently, Probstein (1994) presented an inductive scaling
illustration of the dispersion enhancement under various conditions when flow
convection is introduced in a circular tube.
In addition to the enhancement of mass dispersion (or transfer), the separation
of different species and cross-over phenomenon in which the slow diffuser
eventually travels faster than the fast diffuser in an oscillatory flow under specific
frequencies may exist and have been shown experimentally by Kurzweg & Jaeger
(1987). Such an effect has been applied to the air revitalization process in the
space station life-supporting systems as described by Thomas (2003) and
Thomas & Narayanan (2002b). Thomas and Narayanan (2001, 2002a) also
presented results due, respectively, to the pressure-driven and boundary-driven
flow oscillations in their studies. Three dimensionless parameters, namely, the
Womersley number W (the ratio of viscous diffusion time-scale to the oscillation
time-scale), the Schmidt number Sc (the ratio of species diffusion time-scale to
the viscous diffusion time-scale), and length ratio of the oscillation amplitude to
the channel-width, were identified by them to be essential in understanding the
separation process among species. This separation process was also studied
analytically by Kurzweg (1988) for oscillatory liquid flows in a bundle of
cylindrical capillaries. Therein, he suggested that despite the considerably low
diffusivities of mass species in liquid solutions, the method using flow oscillation
for species separation was still better than the traditionally used chromatography
method because of the relatively low-operation pressure gradient needed in the
oscillatory system. Nevertheless, little effort has been devoted to understand the
fundamental physics of the enhancement of mass transport and separation of
species in an oscillatory EOF.
In order to understand the fundamental physics as well as provide useful
information and criteria for designing micro-fluidic devices, the present study is,
thus aimed at the theoretical investigation of the transport and separation
phenomena of mass species in a periodically oscillatory EOF in two-dimensional
A/C
Exsin(w t)
C1
ux h
A C
CL
y
x C2
Figure 1. Model of the system considered with the associated coordinates and dimensions.
(a ) Electrical field
Due to the symmetry of the system, the EDL potentials extending from the
upper and lower walls are symmetrical with respect to the centreline (depicted
by CL in figure 1) of the channel and are assumed not to overlap with each other.
Such a non-overlapping condition is valid when the Debye length of EDL is much
smaller compared with the width of the channel as suggested by Probstein
(1994), e.g. l! h h=lD O 10. In the above expression, lD is the Debye length
defined as
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3RT
lD Z ; ð2:1Þ
2F 2 z 2 c 0
with F being the Faraday constant, z the valence of the co- and counter-ions in
the carrier liquid (the solvent is assumed to be a 1 : 1 symmetric electrolyte), 3
the permittivity of the carrier liquid, R the universal gas constant, T the absolute
temperature in Kelvin and c 0 the averaged molar concentration of the counter-
ions.
To further simplify the calculation of the EDL potential, other assumptions
are needed. Firstly, the excessive charge distribution is assumed to vary
significantly merely in the y-direction and hence the electrical potential of the
EDL is viewed as a function of y only. The Boltzmann distribution of the charge
density then applies, which gives
" #
zFj
re ZK2Fzc0 sinh ; ð2:2Þ
RT
where re represents the charge density, j, the EDL potential and c0, the ion
concentration far from the charged-walls. Secondly, because of the slow velocities
of EOFs (about the order of 10K4 m sK1) and the relatively small quantities of
the excessive charges, the convective effect of ions, i.e. the possible electro–
magneto interactions are assumed negligible, as suggested by Karniadakis &
Beskok (2002). The EDL potential is then described by the following Poisson–
Boltzmann equation
" #
2 d2 j re 2Fzc0 zFj
V j Z 2 ZK Z sinh : ð2:3Þ
dy 3 3 RT
Equation (2.3) can be further simplified. With the Debye–Huckel approxima-
tion, the above equation for the electrical potential distribution can then be
linearized as
d2 j 1
2
Z j: ð2:4Þ
dy lD
To non-dimensionalize the above equation, the following scheme is employed:
y!Zy/h, j!Zj/z, where the variables with a superscript ‘!’ denote the
dimensionless ones and z is the Zeta potential at the Stern layer. With such a
non-dimensionalization scheme, the governing equation for the EDL potential in
dimensionless form becomes
d2 j!
!2
Z l!2 j! : ð2:5Þ
dy
The associated boundary conditions are
at the boundaries; y ! Z 0 : j! Z 1; ð2:5aÞ
1 dj!
!
at the centreline; y Z : ! Z 0: ð2:5bÞ
2 dy
In equation (2.5a), the electrical potential at the channel wall is approximated by
the Zeta potential at the Stern layer based on the argument of immobility of ions
therein.
(b ) Velocity field
The flow considered is assumed to be continuum, isothermal, Newtonian, two-
dimensional and incompressible. The governing equations are the continuity and
momentum equations, i.e.
vux vuy
C Z 0; ð2:6Þ
vx vy
" #
vu
rf C ðu$VÞu ZKVp C mV2 u C re E: ð2:7Þ
vt
In the above equations, (ux, uy) denotes the velocity vector u in the coordinate
system (x, y); rf and m are the density and dynamic viscosity of the carrier liquid;
E is the applied electrical field with re being the charge density. The last term in
equation (2.7), i.e. reE, denotes the Lorenz force which is the main driving force
to generate the EOF.
By assuming the external electrical field be applied only in the x-direction, E
then reduces to the following form:
h i
E Z E x Z Ex sinðutÞ Z Im E fx eiut : ð2:8Þ
As proposed by Oddy et al. (2001) and Morgan & Green (2003), the magnitude
of Ex should be maintained below 100 V mmK1 to avoid possible EK instability
and significant Joule’s heating for an oscillatory EOF, respectively.
By further assuming the flow be fully developed in the x-direction and mainly
driven by the electroosmotic effect, which Dutta & Beskok (2001b) termed ‘pure
electroosmotic flow’, the y-component velocity, i.e. uy, vanishes from equation
(2.6) and the pressure effect in the momentum equation, i.e. equation (2.7),
becomes negligibly small when compared to the Lorenz effect. The momentum
equation (2.7) then reduces to
vux v2 u r h i
Z n 2x C e Im E fx eiut : ð2:9Þ
vt vy rf
To non-dimensionalize the above equation, the following scheme is applied:
g
ux! Z ux =UHS , where
fx z
3E
g
U HS hK ; ð2:10Þ
m
is a moving referenced velocity (Thomas & Narayanan 2002a) suggested by the
Helmholtz–Smoluchowski equation; t!Zt/t with tZ2p/u; and y!Zy/h. In terms
of the above non-dimensional variables and using the linearized EDL potential,
i.e.
3
re ZK 2 j; ð2:11Þ
lD
the momentum equation then assumes the following dimensionless form:
1 vu! v2 u ! !
W 2 !x Z !2x C l!2 j! ei2pt ; ð2:12Þ
2p vt vy
where
rffiffiffiffi sffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffi
u h 2 =n tn
W hh Z Z ; ð2:13Þ
n 1=u tu
is the Womersley number, a ratio of the viscous diffusion time-scale tn to the
oscillation time-scale tu.
The associated boundary conditions for ux are the no-slip condition at the
channel walls, i.e.
ux! Z 0 at y ! Z 0; 1: ð2:12aÞ
The above condition implies that finite Debye length effects (Dutta & Beskok
2001a) are taken into account in the present study so that a more general
situation of the EOF can be analysed.
(c ) Concentration field
The species to be transported through the carrier liquid is assumed neutral so
that the transport phenomenon will not be affected by any of the electrical
potentials. In addition, the species concentration is also assumed to be infinitely
dilute, so that the concentration gradients of all the species (if there are two or
more species present in the system) will not interfere with each other.
Meanwhile, for an infinitely dilute solution, the molar average velocity is the
same as the mass average velocity which is simply the flow velocity u of the
carrier liquid. Assuming Fick’s law apply, the mass transport equation can then
be written as " 2 #
vc vc v c v2 c
C ux ZD C ; ð2:14Þ
vt vx vx 2 vy 2
where D is the diffusivity coefficient and uyZ0 from the discussion of §2b has
been applied.
The micro-channel in figure 1 is installed between two reservoirs where the
concentration of species are maintained at constant values, c1 for the left
reservoir and c2 for the right one, with c1Oc2. If there is no flow motion of the
carrier liquid, diffusion will be the only mechanism responsible for the transport
of mass species throughout the channel, and the steady-state concentration
distribution will assume a linear form in the x-direction and be uniform in the
y-direction. Now, with a fully developed oscillatory EOF, i.e. uxZux (y, t), as
discussed in §2b, the concentration distribution of the species will not remain
precisely uniform in any cross-section of the channel because of the non-uniform
velocity distribution (no-slip required on the channel walls) across the channel-
width. However, the linear variation of the species concentration along the flow
direction remains unchanged except near the end regions when the system
reaches steady-state. Therefore, it is reasonable to assume the following form for
the species distribution when a fully developed, steady-state oscillatory EOF in a
micro-channel connecting two reservoirs is considered, i.e.
c Kc1
cðx; y; tÞ Z c1 C 2 x C cu ðy; tÞ; ð2:15Þ
L
where cu(y, t) denotes the imposed flow oscillation effect on the concentration
distribution. The boundary conditions at the two ends of the channel, i.e.
c(0, y, t)Zc1 and c(L, y, t)Zc2, are not satisfied exactly by the above equation
3. Mathematical solutions
In this section, the analytic solutions for the electrical potential, velocity and
concentration fields are to be determined; the time and space-averaged mass
transfer rate will also be defined and calculated as an estimation of the overall
mass transport. While deriving the expressions for the averaged mass-transfer
rate, the Lagrangian displacement a and the tidal displacement b as suggested by
Kurzweg (1985, 1988) will be employed so that a reasonable comparison of the
overall mass transport among various flow situations can be achieved.
(a ) Electrical field
From equations (2.5)–(2.5b), the linearized EDL potential is solved as
'" ! # " l! # (
! ! 1KeKl y ! l! e K1 Ky! l!
j ðy Þ Z ! ! e C l! Kl! e ; ð3:1Þ
el KeKl e Ke
(b ) Velocity field
Due to the linearity of the ux-momentum equation, i.e. equation (2.12), the
expression for ux! is assumed to possess the following form:
h !
i
ux! ðy ! ; t ! Þ Z Im uex! ðy! Þei2pt ; ð3:2Þ
as suggested by the imposed electrical field which is the only forcing term of the
equation. With the substitution of the above equation and the EDL potential
(3.1) into equation (2.12), the following equation governing the space
distribution of ux! is thus obtained
'" ! # " l! # (
d2 uex! 2 e! !2 1KeKl l! y ! e K1 Kl! y!
K iW ux ZKl ! ! e C l! Kl! e : ð3:3Þ
dy!2 el KeKl e Ke
The boundary conditions become as follows:
uex! ð0Þ Z uex! ð1Þ Z 0: ð3:3aÞ
The Green’s function method is then applied to solve equations (3.3) and (3.3a),
and the corresponding Green’s function is derived as
8 $ ! pffi pffi %
> ðy K1ÞW i Kðy!K1ÞW i $
> e Ke pffi pffi %
>
> KxW i
>
>
> pffi Wpffii KWpffii exW i
Ke ; 0% x! y! ;
< 2W iðe Ke Þ
G u ðx; y! Þ Z $ ! pffi p ffi % ð3:4Þ
> y W i Ky! W i $ %
>
> e Ke pffi p ffi
>
> ðxK1ÞW i
KeKðxK1ÞW i ; y ! ! x% 1:
>
> pffi $ p ffi pffi % e
: 2W i eW i KeKW i
The distribution of uex! is thus solved and the expression of ux is obtained, i.e.
h !
i
! ! g e! ! i2pt
ux ðy ; t Þ Z Im UHS ux ðy Þe ;
with
!
pffi !
pffi ! ! ! !
uex! ðy! Þ Z 41 ey W i C 42 eKy W i C 43 ey l C 44 eKy l ; ð3:5Þ
where the coefficients 4i’s are given as
$ ! pffi !
pffi ! !
%
l!2 eKðl CW iÞ Keðl KW iÞ C el KeKl
41 Z ) !2 *+ ) pffi * ) pffi *, ; ð3:6aÞ
2 l K iW2 cosh l! C W i Kcosh l! KW i
$ ! pffi !
pffi ! !
%
l!2 eðl CW iÞ KeKðl KW iÞ C eKl Kel
42 Z ) !2 *+ ) pffi * ) pffi *, ; ð3:6bÞ
2 l K iW2 cosh l! C W i Kcosh l! KW i
$ ! pffi !
pffi pffi pffi %
l!2 eKðl KW iÞ KeKðl CW iÞ C eKW i KeW i
43 Z ) !2 *+ ) pffi * ) pffi *, ; ð3:6cÞ
2 l K iW2 cosh l! C W i Kcosh l! KW i
and
$ pffi pffi pffi pffi %
!2 ðl!KW iÞ ðl!CW iÞ W i KW i
l e Ke Ce Ke
44 Z ) !2 * + ) p ffi * ) pffi * , : ð3:6dÞ
2 l K iW2 cosh l! C W i Kcosh l! KW i
(d ) Concentration field
From the previous discussion in §2c, the convective effect on the total
concentration distribution depends only on the oscillatory flow motion. Thus, in
order to be timewise consistent with the flow field, the concentration distribution
due to the convective effect cu! ðy! ; t ! Þ is assumed to possess the following form
h i
! ! ! e! ! Ki2pt !
cu ðy ; t Þ ZKIm cu ðy Þe : ð3:13Þ
By substituting equations (3.2) and (3.13) into equation (2.17), the time and
space coordinates are decoupled, and equation (2.17) reduces to the following
form
d2 cu! gD Auex! ;
K iScW2 cu! Z Pe ð3:14Þ
dy !2
where cu! is the complex conjugate of ceu! and PegD ðuÞ is the frequency-dependent
Peclet number. With the non-penetration boundary conditions
&
dcu! &&
Z 0; ð3:14aÞ
dy ! &y!Z0;1
the Green’s function of equations (3.14) and (3.14a) is derived as
8 $ ! pffiffiffiffiffi pffiffiffiffiffi %
> ðy K1ÞW iSc Kðy! K1ÞW iSc $
> e Ce pffiffiffiffiffi pffiffiffiffiffi %
>
> KxW iSc
>
> $
pffiffiffiffiffiffiffi KWpffiffiffiffi ffi pffiffiffiffiffi % e xW iSc
Ce ; 0%x!y! ;
>
< 2W iSc e iSc Ke W iSc
G c ðx;y! ÞZ $ ! pffiffiffiffiffi p ffiffiffiffiffi %
>
> e y W iSc
Ce Ky ! W iSc $ pffiffiffiffiffi pffiffiffiffiffi %
>
> ðxK1ÞW iSc KðxK1ÞW iSc
>
>
>
$
pffiffiffiffiffiffiffi KWpffiffiffiffi ffi pffiffiffiffiffi % e Ce ; y! !x%1:
: 2W iSc e iSc KeW iSc
ð3:15Þ
After integrating
ð1
cu! ðy! Þ gD
Z APe G c ðx; y ! Þuex! ðxÞdx; ð3:16Þ
0
pffi % ð3:17Þ
Ky! Wi i l! y ! Kl! y !
Cg4 e C g5 e C g6 e ;
with PeD being the complex conjugate of Pe gD , and the coefficients gi’s being
!
functions of l , W and Sc, i.e.
$ pffi pffiffiffiffiffi % $ pffiffiffiffiffi pffi %
KWi i KWi iSc KWi iSc Wi i
41 e Ke C 42 e Ke
g1 Z p ffiffiffiffiffi ) p ffiffiffiffiffiffi
ffi*
2ðScK1ÞW2 i Sc sinh Wi iSc
$ ! pffiffiffiffiffi % $ pffiffiffiffiffi !
%
l! 43 el KeKWi iSc C l! 44 eKWi iSc KeKl
K pffiffiffiffiffiffiffi) * ) pffiffiffiffiffiffiffi* ; ð3:18aÞ
2Wi iSc l!2 C iScW2 sinh Wi iSc
$ pffi pffiffiffiffiffi % $ pffiffiffiffiffi pffi %
41 eKWi i KeWi iSc C 42 eWi iSc KeWi i
g2 Z pffiffiffiffiffi ) pffiffiffiffiffiffiffi*
2ðScK1ÞW2 i Sc sinh Wi iSc
$ ! pffiffiffiffiffi % $ pffiffiffiffiffi !
%
l! 43 el KeWi iSc C l! 44 eWi iSc KeKl
K pffiffiffiffiffiffiffi) * ) pffiffiffiffiffiffiffi* ; ð3:18bÞ
2Wi iSc l!2 C iScW2 sinh Wi iSc
42
g3 Z ; ð3:18cÞ
ðScK1ÞiW2
41
g4 Z ; ð3:18dÞ
ðScK1ÞiW2
43
g5 Z !2
; ð3:18eÞ
l C iScW2
and
44
g6 Z !2
; ð3:18f Þ
l C iScW2
where 4i ’s are the complex conjugates of 4i’s. From equations (2.15), (3.13) and
(3.17), the total concentration distribution can then be expressed as
c Kc1 + ,
cðx; y; tÞ Z c1 C 2 x KIm ceu ðyÞeKiut : ð3:19Þ
L
where cu and ux are the complex conjugates of ceu and uex , respectively. The above
relation can also be expressed in the dimensionless form as
ð1 $ %
1
!
Qx Z 1K jPeD j 2
ceu! uex! C cu! ux! dy! ; ð3:23Þ
4 0
with the parameters defined as Qx! Z ðQx LÞ=ðDðc1 Kc2 ÞÞ, ceu! Z ceu! =PeD A and
gD A. The integral in equation (3.23) is expressed in terms of 4i’s
cu! Z cu! =Pe
and gi’s, and their complex conjugates as well, i.e.
ð1 $ %
ceu! uex! C cu! ux! dy! Z 43 g6 C 44 g5 C 43 g6 C 44 g5
0
+) ! * ) ! * ,
e2l K1 ð43 g5 C 43 g5 ÞK eK2l K1 ð44 g6 C 44 g6 Þ
C
2l!
) !2 *n) ! pffi *h $ l!CWipffii % $ ! pffi % io
l KW2 i l KWi i e K1 ð43 g3 C 42 g5 ÞK eKðl CWi iÞ K1 ð44 g4 C 41 g6 Þ
C
l!4 C W4
) !2 *n) ! pffi *h $ l!KWipffii % $ pffi % io
K1 ð43 g4 C 41 g5 ÞK eKðl KWi iÞ K1 ð44 g3 C 42 g6 Þ
!
l KW2 i l C Wi i e
C
l!4 C W4
) !2 *n) ! pffi *h $ l!CWpffii % $ pffi % io
K1 ð41 g5 C 43 g4 ÞK eKðl CW iÞ K1 ð42 g6 C 44 g3 Þ
!
l C W2 i l KW i e
C
l!4 C W4
) !2 *n) ! pffi *h $ l!KWpffii % $ pffi % io
K1 ð42 g5 C 43 g3 ÞK eKðl KW iÞ K1 ð41 g6 C 44 g4 Þ
!
l C W2 i l CW i e
C
l!4 C W4
) pffi pffi *h $ WpffiiCWipffii % $ pffi pffi % i
W i KWi i e K1 ð41 g3 C 42 g4 ÞK eKðW iCWi iÞ K1 ð42 g4 C 41 g3 Þ
C
2W2 i
) pffi pffi *h $ WpffiiKWipffii % $ pffi pffi % i
W i C Wi i e K1 ð41 g4 C 41 g4 ÞK eKðW iKWi iÞ K1 ð42 g3 C 42 g3 Þ
C
2W2 i
) pffi pffiffiffiffiffiffiffi*h $ pffi pffiffiffiffiffi % $ pffi pffiffiffiffiffi %i
W i KWi iSc 41 g1 eW iCWi iSc K1 K42 g2 eKðW iCWi iScÞ K1
C
ð1 C ScÞW2 i
) pffi pffiffiffiffiffiffiffi*h $ pffi pffiffiffiffiffi % $ pffi pffiffiffiffiffi %i
W i C Wi iSc 41 g2 eW iKWi iSc K1 K42 g1 eKðW iKWi iScÞ K1
C
ð1 C ScÞW2 i
) pffi pffiffiffiffiffiffiffi*h $ pffi pffiffiffiffiffi % $ pffi pffiffiffiffiffi %i
Wi i KW iSc 42 g2 eWi iCW iSc K1 K41 g1 eKðWi iCW iScÞ K1
K
ð1 C ScÞW2 i
) pffi pffiffiffiffiffiffiffi*h $ pffi pffiffiffiffiffi % $ pffi pffiffiffiffiffi %i
Wi i C W iSc 42 g1 eWi iKW iSc K1 K41 g2 eKðWi iKW iScÞ K1
K
ð1 C ScÞW2 i
) !2 *n) ! pffiffiffiffiffiffiffi*h $ ! pffiffiffiffiffi % $ pffiffiffiffiffi %io
l KWi iSc 43 g1 el CWi iSc K1 K44 g2 eKðl CWi iScÞ K1
!
l KW2 iSc
C
l!4 C W4 Sc2
Table 3. Mass transport properties and the corresponding Schmidt number for four example mass
species.
gradually near the solid wall with a smaller value of l!, which indicates a larger
Debye length. As the Womersley number increases, the velocity profile tends to
be non-uniformly distributed across the channel-width as shown in figure 2b and
c, indicating phase lags between the velocity distribution and the imposed
electrical field. Since the counter-ions spread farther into the bulk liquid with a
larger Debye length, i.e. a smaller value of l!, more fluid particles are dragged by
the counter-ions driven by the electrical force. Consequently, the velocity
boundary layer becomes thicker with the peak value of the non-dimensional
velocity distribution being smaller.
(a) 1.00
0.98
0.96
u*x
0.94
0.92
(b)
0.8
0.6
u*x 0.4
0.2
(c)
0.8 l∗=700
l∗=70
0.6 l∗=20
u*x 0.4
0.2
0
0 0.2 0.4 0.6 0.8 1.0
y*
Figure 2. The dimensionless velocity distributions ux! ðy ! Þ across the channel-width for various
values of Womersley number W and the dimensionless Debye length l! at tuZ(2nC0.5)p (a)
WZ0.5, l!Z20, 70 and 700. (b) WZ5, l!Z20, 70 and 700.(c) WZ15, l!Z20, 70 and 700.
–2
W= 0.1
–4 W= 0.3
W= 0.6
–6
0 0.2 0.4 0.6 0.8 1.0
y*
Figure 3. The dimensionless convective concentration distributions cu! across the channel-width for
WZ0.1, 0.3 and 0.6 at tuZ(2nC0.5)p, with ScZ1000, l!Z70 and Dx 0 /hZ1.0.
and 2.0 are qualitatively similar to those shown in figure 3 deferring only slightly
in the numerical values and hence, not included.
1.35
1.30
1.25 ∆ x'/ h = 2.0
Figure 4. The dimensionless averaged mass transport rate Qx! as a function of Womersley number
W with l!Z70, ScZ1000 and Dx 0 /hZ1.0, 1.5 and 2.0.
(a)
1.035
1.030
1.025
1.020
Q*x
1.015
1.010
1.005
1
0 0.1 0.2 0.3 0.4 0.5
W
(b)
2
0
–2
c*u × 10 5
–4 l*= 20
–6 l* = 70
l* = 700
–8
– 10
0 0.2 0.4 0.6 0.8 1.0
y*
Figure 5. The dimensionless averaged mass transport rates Qx! and the associated convective
concentration distributions cu! for l!Z20, 70 and 700 with Dx/hZ1.0. (a) Qx! versus W. (b) cu!
distribution across the channel-width.
nw10K6 m2 sK1); figure 6a is for l!Z20, figure 6b is for l!Z70, and figure 6c is
for l!Z700, all with Dx 0 /hZ1.0. The results indicate that an oscillatory EOF is
always more conducive to the enhancement of mass transport rate of the species
with a larger Sc number, or a smaller mass diffusivity and this phenomenon
becomes more and more obvious with a smaller value of l!. Thus, with a
sufficiently low value of l! and at a sufficiently high frequency, the cross-over
(a) 1.7
1.6 Sc = 2000
1.5 Sc =1500
Q*x 1.4 Sc =1000
1.3 Sc =500
1.2 Sc =100
1.1
1.0
(b)
1.030
1.025
1.020
Q*x 1.015
1.010
1.005
1.000
(c) 1.0004
1.0003
Q*x 1.0002
1.0001
1.0000
0 0.1 0.2 0.3 0.4 0.5
W
Figure 6. The variations of Qx! versus W with Dx 0 /hZ1.0 and ScZ100, 500, 1000, 1500 and 2000.
(a) l!Z20. (b) l!Z70. (c) l!Z700.
(a) 15.0
12.5 sucrose
10.0 glucose
Qx ×10 8
7.5
5.0 2nd cross-over
2.5
1st cross-over
0
(b) 15.0
12.5 sucrose
Qx ×10 8
10.0 ethanol 2nd cross-over
7.5
5.0
2.5
1st cross-over
0 0.05 0.10 0.15 0.20 0.25 0.30
W
Figure 7. The cross-over phenomenon with l!Z20 and Dx 0 /hZ10.0. (a) Glucose and sucrose. (b)
Ethanol and sucrose. The material properties are listed in table 3.
5. Concluding remarks
(v) With proper choices of the Debye length, oscillation frequency and tidal
displacement, a first-step separation of species by means of an oscillatory
EOF becomes achievable.
Future work may include the removal of the linear electrical potential
assumption; the transient effect on the velocity and concentration distributions
may also need to be considered for the flow situations with WR1. Numerical
results will then become inevitable.
The authors gratefully acknowledge the financial support from the National Science Council of
Taiwan, ROC, through grant no. NSC93-2212-E-002-036.
References
Aris, R. 1956 On the dispersion of a solute in a fluid flowing through a tube. Proc. R. Soc. A 235,
67–77.
Aris, R. 1960 On the dispersion of a solute in pulsating flow trough a tube. Proc. R. Soc. A 259,
370–376.
Çengel, Y. A. 2003 Heat transfer: a practical approach, 2nd edn, p. 726. New York: McGraw-Hill.
Dutta, P. & Beskok, A. 2001a Analytical solution of combined electroosmotic/pressure-driven
flows in two-dimensional straight channels: finite Debye layer effects. Anal. Chem. 73,
1979–1986. (doi:10.1021/ac001182i)
Dutta, P. & Beskok, A. 2001b Analytical solution of time periodic electroosmotic flows: analogies
to Stokes’ second problem. Anal. Chem. 73, 5097–5102. (doi:10.1021/ac015546y)
Erickson, D. & Li, D. 2003 Analysis of alternating current electroosmotic flows in a rectangular
microchannel. Langmuir 19, 5421–5430. (doi:10.1021/la027035s)
Green, N. G., Ramos, A., Gonzalez, A., Morgan, H. & Castellanos, A. 2000 Fluid flow induced by
nonuniform ac electric fields in electrolytes on microelectrodes. I. Experimental measurements.
Phys. Rev. E 61, 4011–4018. (doi:10.1103/PhysRevE.61.4011)
Hunter, R. J. 1981 Zeta potential in colloid science, pp. 359–359, 1–121, London: Academic Press.
Karniadakis, G. E. & Beskok, A. 2002 Micro flows: fundamentals and simulation, pp. 193–210. New
York: Springer.
Kurzweg, U. H. 1985 Enhanced heat-conduction in oscillating viscous flows within parallel-plate
channels. J. Fluid Mech. 156, 291–300.
Kurzweg, U. H. 1988 Enhanced diffusional separation in liquids by sinusoidal oscillations. Sep. Sci.
Tech. 23, 105–117.
Kurzweg, U. H. & Jaeger, M. J. 1987 Diffusional separation of gases by sinusoidal oscillations.
Phys. Fluids 30, 1023–1025. (doi:10.1063/1.866300)
Lin, H., Storey, B. D., Oddy, M. H., Chen, C. H. & Santiago, J. G. 2004 Instability of electrokinetic
microchannel flows with conductivity gradients. Phys. Fluids 16, 1922–1935. (doi:10.1063/1.
1710898)
Morgan, H. & Green, N. G. 2003 AC electrokinetics: colloids and nanoparticles, pp. 81–158.
Oxford: Research Studies Press.
Oddy, M. H., Santiago, J. G. & Mikkelsen, J. C. 2001 Electrokinetic instability micromixing. Anal.
Chem. 73, 5822–5832. (doi:10.1021/ac0155411)
Probstein, R. F. 1994 Physicochemical hydrodynamics: an introduction, pp. 21–29, 37–50, 82–96,
190–202, 2nd edn. New York: Wiley
Suresh, V. & Homsy, G. M. 2004 Stability of time-modulated electroosmotic flow. Phys. Fluids 16,
2349–2356. (doi:10.1063/1.1736677)
Taylor, G. I. 1953 Dispersion of soluble matter in solvent flowing slowly through a tube. Proc. R.
Soc. A 219, 186–203.
Thomas, A. M. 2003 Unusual effects of oscillating flows in an annulus on mass transfer and
separation. Adv. Space Res. 32, 279–285. (doi:10.1016/S0273-1177(03)90263-8)
Thomas, A. M. & Narayanan, R. 2001 Physics of oscillatory flow and its effect on the mass transfer
and separation of species. Phys. Fluids 13, 859–866. (doi:10.1063/1.1351549)
Thomas, A. M. & Narayanan, R. 2002a A comparison between the enhanced mass transfer in
boundary and pressure driven oscillatory flow. Int. J. Heat Mass Transfer 45, 4057–4062.
(doi:10.1016/S0017-9310(02)00111-4)
Thomas, A. M. & Narayanan, R. 2002b The use of pulsatile flow to separate species. Ann. NY
Acad. Sci. 974, 42–56.