You are on page 1of 12

Microfluid Nanofluid (2013) 15:859–870

DOI 10.1007/s10404-013-1196-7

RESEARCH PAPER

High field asymmetric waveform for ultra-enhanced


electroosmotic pumping of porous anodic alumina membranes
D. Piwowar • M. E. Tawfik • F. J. Diez

Received: 9 December 2012 / Accepted: 19 April 2013 / Published online: 7 May 2013
Ó Springer-Verlag Berlin Heidelberg 2013

Abstract An electroosmotic EO process is presented for 1 Introduction


nanoporous membranes capable of generating EO flow
rates over thirty times higher than previously possible with An EO pump generates a flow rate or pressure change due
the same membrane and solution. In generating high EO to the ion migration in the electric double layer toward the
flows, a limiting factor is faradaic reactions which appear electrode which causes viscous drag creating a net flow.
at high electric fields. A process is presented capable of These pumps can be fabricated using a variety of methods
limiting and even canceling these reactions allowing from capillary columns packed with particles (Zeng et al.
electric field between one and two orders of magnitude 2001; Wang et al. 2006; Chen 2005) to porous membranes
higher. This is achieved by applying an asymmetric bipolar (Yao et al. 2003; Chen et al. 2005; Yao et al. 2006; Cao
rectangular voltage waveform. The results show the et al. 2012; Kwon et al. 2012; Kim et al. 2013; Garg et al.
enhanced EO pumping capabilities of membranes under a 2012). EO pumps have several advantages compared
high electric field asymmetric waveform which prevents to typical mechanical pumps. These include having no
gas generation at high voltages. A baseline is established moving or mechanical parts, being compatible with lab-on-
by measuring the EO pump performance when a constant a-chip components and being used for accurate and
voltage is applied to SiO2-coated nanoporous anodic alu- adjustable flow rate control. EO pumps are used in a
minum oxide membranes. The analysis compares the effect variety of applications including power electronic cooling
of the applied voltage type on the maximum flow rate, (Berrouche et al. 2009), fuel cells (Buie et al. 2006),
power consumption, and maximum pressure. Results show actuation (Prakash et al. 2006), drug delivery (Pikal 1992),
that large gas generation prevents membrane operation chromatography (Chen et al. 2004), and lab-on-a-chip
when direct current DC voltages above 50 V are applied. systems (Kim et al. 2013; Li and Harrison 1997).
On the other hand, it operates normally under an asym- High flow rates are of interest in many applications, but they
metric voltage ?1,800/-900 V applied, with negligible require high electric fields. A well-known issue at those con-
gas generation. This results in a thirty-time flow rate ditions has been faradaic reactions. These limit the maximum
increase. Larger flow rates/voltages are possible but were electric field that can be applied and the flow rate that can be
not considered due to hardware limitations. obtained. Faradaic reactions can result in pH changes, elec-
trode degradation, and electrolysis. Gas evolution due to
electrolysis occurs at the electrodes and can decrease the
efficiency, the flow rate performance and increase the possi-
Electronic supplementary material The online version of this bility of channel blockage. To reduce pH changes, large res-
article (doi:10.1007/s10404-013-1196-7) contains supplementary
material, which is available to authorized users. ervoirs can be used, but are not practical for on chip devices. To
prevent electrode erosion, many EO pumps utilize platinum
D. Piwowar  M. E. Tawfik  F. J. Diez (&) electrodes due to their high inert properties. Nevertheless, they
Department of Mechanical and Aerospace Engineering,
still produce bubbles from the reaction of hydrogen and oxy-
Rutgers, The State University of New Jersey,
98 Brett Road, Piscataway, NJ 08854, USA gen at the cathode and anode, respectively. One method of
e-mail: diez@rutgers.edu reducing this reaction is to stay below the voltage limit of 2 V

123
860 Microfluid Nanofluid (2013) 15:859–870

of the reaction (Ai et al. 2010). This method is impractical for pumping characteristics of membranes are presented. These
high flow rate EO pumps where higher voltages are needed. show that under an asymmetric applied voltage, flow rates
Several other methods have been proposed including the use of over an order of magnitude higher than under an applied DC
ion exchange membranes and meshes among others to block voltage can be obtained. Also, applied voltages over an
the gas bubbles from entering the pores (Yao et al. 2003; order higher are possible without faradaic reactions.
Berrouche et al. 2009; Burke et al. 2010). In some cases, this
has been effective but leaves the device with an overload of
gases or the need for platinum catalysts for recombination. Lin 2 Theoretical model
et al. (2007) used Nafion tubing to funnel the gases out of the
electroosmotic pump into a recombiner to create water. Some 2.1 Bipolar rectangular voltage
EO pumps have replaced the inert electrodes with disinte-
grating electrodes such as in Heuck et al. (2011). In that case, A process for canceling or limiting faradaic reactions in EO
the electrodes erode with the reactions to prevent bubble flow will be presented by applying an asymmetric bipolar
production but these pumps have a short life usage before the rectangular voltage waveform. As will be shown, this method
electrodes need to be replaced. The use of palladium electrodes not only cancels the gas generation during EO flow, but it can
has also been proposed (Brask et al. 2006) due to their also significantly enhance the flow rate capabilities of
hydrogen absorbing properties, but this is only effective for membranes used as EO pumps. The type of applied voltage
small quantities of hydrogen gas such as that typically pro- proposed is shown schematically in Fig. 1 and given by,
duced from de-ionized water EO flow. It has also been sug- (
Vþ if tþ þt 1
t\ tþtþt
þ

gested to include additives in the liquid solution to reduce the V¼ 1





ð1Þ
hydrogen and oxygen reaction at the electrodes (Kohlheyer V if tþ þt t [ tþ þt
et al. 2008). But, it is limited or impractical for high molarity
with a high voltage value V? and a low value V-, and time
solutions and high electric fields applied or in cases where the
periods t? and t-, respectively. Also, the average voltage
solution cannot be contaminated.
can be obtained by integrating over one wave period to give,
Recently, the idea of pulsing the voltage or current has 0 1
been proposed for electroosmotic-driven flows. Selvagana- Ztþ þ þt
tZ

pathy et al. (2002) applied a periodic zero net current pulse to ¼ 1 B C tþ Vþ þ t V


V @ Vþ dt þ V dtA ¼ :
generate a flow rate while preventing electrolysis. They tþ þ t tþ þ t
0 tþ
report that this is possible at low frequencies where the
ð2Þ
current–voltage response of the system is nonlinear and can
generate an EO flow. Xu et al. (2011) used positive voltage For the case when the area per pulse are equal to each
pulsing with different duty cycles to digitally control the other,
flow. They found it to be an effective way to stabilize the Areaþ Vþ tþ
¼ ¼ 1; ð3Þ
flow rate (no decay of the flow rate with time and with Area V t
minimal fluctuation). It also helped the EO pump to behave
the net current in the system is zero, and faradaic reactions
linearly with increasing electric field. There has also been
are canceled. McBride (1999) has suggested that for this
work done on using arrays of electrodes where an alternating
equal area case, there is a net fluid flow due to the pressure
current AC electric field is used to generate an EO flow with
developed by ion drag pumping. Although this is important
no electrolysis (Chen et al. 2009). This flow is driven by the
for electro-hydro-dynamics EHD applications such as in
electric double-layer EDL around the electrodes and the
corona discharges (Stuetzer 1959), ion drag pumping as
vortices generated between adjacent electrodes in the array.
The main application is for micromixers, but the low voltages
used, and the need for assembly of electrode arrays in micro-
channels is not suitable for large EO pumping in membranes.
The present work is motivated by the lack of EO pro-
cesses capable of preventing faradaic reactions that will
allow high electric fields and flow rates. The work starts
with a brief analytical description of the proposed asym-
metric voltage waveform and the EO theory for membranes.
A brief description of the experimental setup follows. Next,
baseline measurements for DC voltage EO pumping for
Fig. 1 A sketch of the bipolar rectangular voltage waveform used for
different membranes are obtained. Also, measurements the electroosmotic pump. The positive and negative areas do not need
showing the effect of asymmetric voltage in the EO to be equal to each other

123
Microfluid Nanofluid (2013) 15:859–870 861

defined by (McBride 1999; Stuetzer 1959), is negligible for Za


pa4 efpa2 f
EO flow applications. Therefore, maintaining the ratio in Q ¼ 2p ruðrÞdr ¼ Pz  Ez ð5Þ
Eq. (3) equal to 1 will result in a negligible flow rate for EO 8g g
0
pumps as we will show. But more importantly, the exper-
iments will prove that by relaxing the condition imposed by where
Eq. (3) to values as low as 0.1, high flow rates are possible Za  
wðrÞ 2r
with negligible faradaic reactions. f ¼ 1 dr: ð6Þ
The response of the electrokinetic flow in a typical nano- f a2
0
pore of radius a to an applied bipolar rectangular pulse wave
described by Eq. (1) can be approximated by a simple f tends to unity when the channel radius a is much larger
hydrodynamic model and classical steady EO flow theory. than the electric double-layer EDL thickness j-1. Briefly, an
Quasi-steady EO flow conditions can be expected by con- EDL is a thin region of nonzero net charge density near the
sidering that the motion induced in the double layer by either a wall that forms when a solid surface (such as the walls of a
symmetric or an asymmetric applied electric field is equiva- channel) enter in contact with an aqueous solution resulting
lent to a variation of Stokes’ problem for the flow induced by an interfacial charge that causes a rearrangement of the free
viscous diffusion above a moving wall. A bipolar rectangular ions in the solution. The EDL thickness can be estimated by
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
time-varying applied field produces a nonsymmetric oscil- j1 ¼ ekT=2e2 z2 MNA , where T is the temperature of the
lating motion in the thin double layer directly adjacent to the fluid, k = 1.381 9 1023 l/mol is the Boltzmann constant,
wall which induces a motion in the bulk electrolyte outside the e = 1.602 9 1023 C is the elementary charge, z = 1 is the
double layer. This motion is confined to a ‘Stokes layer’ with charge number, M is the molarity (i.e., 0.1, 1, 10 mM), and
pffiffiffiffi
thickness d  mt, where v is the kinematic viscosity. For the NA = 6.022 9 1023 l/mol is the Avogadro constant. For
present work in AAO membranes with a * 75 nm and instance, for a molarity M of 0.1, 1, and 10 mM, the cal-
applied frequency of 1,000 Hz, it can be shown that culated j-1 are 30, 9, and 3 nm, respectively. Normalizing
d  a. For instance, defining d as the thickness where these values by the radius a = 75 nm, we obtain the elec-
ud = 0.9uw (uw is the wall velocity) results in d = 1.88 lm. trokinetic radius ja with values 2.5, 8.3, 25 for M of 0.1, 1,
For the membranes in the present work, the upper limit fre- and 10 mM, respectively. For large ja values, f tends to
quency is 353 kHz. This ensures quasi-steady flow conditions unity, but for smaller values such as the one in this work
during each voltage pulsed applied so that a total flow rate (i.e., ja = 2.5 and 8.3), Eq. (6) was solved.
Q = Q? - Q- results from such an asymmetric-induced From Eq. (5), it can be shown that the maximum flow rate
motion during each volumetric displacement ‘‘stroke’’ of the occurs when the back pressure is zero, and the maximum
actuator and classic EO flow theory can be applied. pressure occurs when the flow rate equals zero yielding,
efpa2 f
2.2 Electroosmotic theory Qmax ¼  Ez ð7Þ
g

An analytical model is briefly presented that will be com- 8eff


Pz;max ¼  Ez ð8Þ
pared with experimental results and help to better understand a2
the behavior of the EO pump. Using the electroosmotic Thus, both Qmax and Pz,max are linearly related to the
theory for flow in a cylindrical pore (Levine et al. 1975; Rice applied electric field. Also, the current can be obtained by
and Whitehead 1965), the maximum flow rates, maximum integrating the total current density over the channel to give
pressure, and current consumption can be obtained. The
velocity profile in a nano-pore in presence of both an axial Za
pressure field Pz and axial electric field Ez assuming steady, I ¼ 2p ðqðrÞuðrÞ þ KEz Þrdr
low Reynolds number flow and a 1:1 electrolyte is 0

1  2  e Za   Za  2
uðrÞ ¼ a  r 2 Pz  ðf  wðrÞÞEz ð4Þ pa2 efPz w 2r pa2 e2 Ez dw 2r
4g g ¼ 1 dr þ dr
g f a2 g dr a2
0 0
where g is the viscosity, f potential is the effective surface Za  
electric potential, and e the permittivity of the liquid. The 2r zew
þ pa2 KEz cosh dr; ð9Þ
induced potential w(r) in a pore can be described using the a2 kT
0
solution to the Poisson–Boltzmann (PB) equation from
Levine et al. (1975). Then, the volumetric flow rate can be where q(r) is the charge density at a distance r from the
obtained by integrating the velocity over the cross- axis, and K is the electrolyte conductivity. The first term in
sectional area of the channel to give the integration gives the convection current, and it is due to

123
862 Microfluid Nanofluid (2013) 15:859–870

the transport of ions through the bulk fluid by its velocity.


The second term is the conduction current, and it is due to
the movement of ions relative to the bulk fluid. This is
derived from Ohm’s Law and simplified in terms of the
solution’s conductivity when the surface conductance is
sufficiently small as suggested by Rice and Whitehead
(1965). The ratio of the flow rate to the current can be
written in terms of a nondimensional parameter g (Yao and
Santiago 2003), given by
,8 Z a   Za  
9
< e2 dw 2r2
zew 2r =
g¼f dr þ cosh dr :
:gK dr a 2 kT a2 ;
0 0
ð10Þ
The functions f and g are evaluated using the solution to
the PB equation from Levine et al. (1975). In most cases of
interest, ja is large so that f and g tend to unity. For the
present study with 150-nm-pore size membranes and a
typical 1 mM phosphate buffer solution, the calculated ja Fig. 2 A 3D model of the EO pump housing showing the anodic
is 8, and f and g are 0.766 and 0.212, respectively, aluminum oxide membrane in the middle, with platinum wire mesh
electrodes and reservoirs in each side
showing that these correction factors cannot just be
approximated by 1.
To obtain the membrane performance, the single chan- degradation. These are fabricated from 0.25-mm-diameter
nel theory needs to be modified to include the multiple platinum wire (99.9 % pure) and platinum 52 mesh (99.9 %
channels in a membrane. This is done by comparing the pure) from Sigma Aldrich. A Keithley 6517B electrometer or
open area of the membrane to the single pore area which a Trek model 5/80 amplifier with a Tektronix AFG 3021B
gives the number of pores in the membrane as function generator is used to supply power to the EO pump.
An SEM photograph of the anodic aluminum oxide
HAmem (AAO) membranes obtained from Synkera Technologies,
N¼ ; ð11Þ
pa2 Inc is shown in Fig. 4. These have a 150-nm pore size and a
where H is the porosity, and Amem is the membrane’s area. 50-lm thickness. Their characteristics include symmetric
Flow rate and current calculations for single channel theory cylindrical pore radius throughout the thickness of the
can now be multiplied by N to obtain the membrane values. membrane, porosity of 30–32 %, and tortuosity of 1. These
membranes are heat-treated up to 1,000 °C by the manu-
facturer, so they can be operated in the pH range of 4.7–9.
3 Experimental setup Vajander et al. (2007) coated their AAOs with a 5-nm layer
of SiO2 and reported a performance increase. Thus, we had
To study the performance of the membranes, they are Synkera coat the membranes with a custom coating of 10 nm
mounted in an enclosure machined in acrylic sketched in of SiO2 which is added to the heat-treated AAOs by atomic
Fig. 2. This housing for the EO pump has a reservoir in each layer deposition. This resulted in a measured pore size
side of the membrane. The inner reservoirs are connected to opening of 135 ± 10 nm. Miao et al. (2007) coated their
the large diameter outer reservoirs using 00 tubing for flow AAOs with a layer of Pt on both faces to act as electrodes,
rate measurements. Optical access to the inner reservoirs in whereby maximizing the effective voltage. To evaluate this
each side of the membrane allowed monitoring gas genera- effect, a 100-nm Pt-coating is also added to a batch of SiO2-
tion. A scale (Ohaus Scout Pro ± 0.001 g) monitored the coated AAOs. This resulted in a membrane porosity of only
output flow as seen in Fig. 3. For maximum pressure mea- 13 % and a measured pore size opening of 99 nm. This
surements, the tubing is replaced with a connector to a coating is only added to a small batch due the high cost which
pressure transducer (Omegadyne PX309) as seen in Fig. 3. resulted in limited testing. During EO testing, the average
Silicone rubber gaskets are used to seal the pump. Mem- exposed area for all AAOs was 0.400 cm2.
branes are epoxied to PVC sheets of 0.8 mm thickness to The primary solutions used are deionized ultra-filtered
hold the membranes inside the housing. Platinum wire and water and phosphate buffer made with Sodium Phosphate
mesh is used as electrodes for the electroosmotic pump. Monobasic (C99 % pure) and Sodium Phosphate Dibasic
Using inert platinum electrodes minimizes electrode (C99 % pure) from Sigma Aldrich. The corresponding

123
Microfluid Nanofluid (2013) 15:859–870 863

Fig. 3 Sketch showing the


experimental setup used to
study the EO pump
performance. For the pressure
measurements, a pressure
transducer is used, and for the
flow rate measurements, this is
replaced with a reservoir and
precision scale

Table 1 Measured Phosphate buffer conductivity for the different prevented gas generation. A baseline is established first by
molarities used measuring the EO pump performance for a constant
Molarity mM pH Conductivity applied voltage. This analysis will allow us to compare and
(lS/cm) show the effect of the type of voltage applied on the
maximum flow rate, current consumption, and maximum
0.1 5.7 6.2
pressure of the membranes.
0.5 6.1 45
1 6.2 90
4.1 Constant voltage EO pump performance
2 6.3 168
10 6.3 861
The EO pumping capabilities of membranes are evaluated
using the setup described in Fig. 3. Flow rate measurements
are conducted in two different types of membranes. The first
is a SiO2-coated AAO membrane where the electrode is a
thin Pt film coated on both sides of the membrane surface.
The second is also a SiO2-coated AAO membrane, but the
electrode is formed by a Pt wire, and mesh not in contact
with the membrane but rather positioned 1–2 mm from the
membrane. The flow rate measurements are normalized by
the applied voltage and open area of the membrane in order
to compare these results to published work. This normalized
flow rate can now be plotted against the applied electric
field in Fig. 5, and it allows for the comparison of mem-
branes with different open areas, thickness, and applied
voltage. For instance, the SiO2-coated AAOs with the Pt
wire mesh achieved for at 30 V and a 1 mM buffer solution,
a maximum flow rate of 0.561 ml/min which when nor-
malized gives 0.102 ml/min/V/cm2. This is comparable to
the flow rate measured for SiO2-coated AAO membranes in
the literature (Vajandar et al. 2007) and also for porous
silicon membranes (Yao et al. 2006) as shown in Fig. 5. The
best performing EO pumps were the SiO2-coated AAO with
Fig. 4 SEM image of SiO2-coated AAO with 150-nm pore size Pt film-coated electrodes. The Pt coating minimizes any
voltage loss from the electrodes to the membrane and thus,
phosphate buffer (pH of 6.2) conductivities are listed in allows the applied voltage to be approximately equal to the
Table 1 and the measured deionized water conductivity is effective voltage. The flow rate from these AAOs improved
1.51 lS/cm. Conductivity and pH measurements are con- by an order of magnitude compared with the non-Pt-coated
ducted using an Oakton 510 series meter. AAOs and outperformed the SiO2- and Pt-coated AAOs
from Miao et al. (2007). They generated a flow rate of
2.10 ml/min at 10 V which equates to a normalized value of
4 Results 4.69 ml/min/V/cm2 shown in Fig. 5. To our knowledge,
this normalized flow rate is the highest obtained to date with
The objective is to evaluate the enhanced EO pumping a membrane.
capabilities of membranes when a bipolar rectangular In order to compare the experimentally measured flow
voltage waveform is applied which limited and even rate results to those predicted by the theory, the effective

123
864 Microfluid Nanofluid (2013) 15:859–870

10 The 10 mM case could not be run at 30 V due to the


presence of electrolysis while the 0.1 mM case could be
run at 50 V (not shown in the figure) with a maximum flow
rate of 0.099 ml/min before gas generation is observed. A
Q/V/Am (mL/min/V/cm2)

more detailed effect of the molarity on the flow rate is


1
shown in Fig. 6b. It compares the flow rates generated by
varying the molarity of the phosphate buffer using the
SiO2-coated AAOs. It also shows that molarities greater
than 2 mM do not contribute to a significant increase in
0.1 flow rate. This is comparable to the result obtained by
Vajandar et al. (2007) where a 2.5 mM solution generated
their greatest normalized flow rate. As discussed by
Vajandar et al. (2007), in the regime of 10 C ja C 1, the
flow rate decreases as the EDL overlaps. Present results
0.01 have the same tendency with 1 mM phosphate buffer
0.01 0.1 1
(ja = 8) having a greater flow rate than a 0.1 mM phos-
V/L (V/µm)
phate buffer (ja = 2). On the other hand, the increase in
Fig. 5 Flow rate measurements of SiO2-coated AAOs normalized by molarity has the undesirable effect of increasing the current
the applied voltage and open area of the membrane. Results are as shown in Fig. 6c and by extension the power con-
compared to published work for various types of membranes. The
obtained value of 4.69 ml/min/V/cm2 for the platinum and SiO2-
sumption. Similar to the flow rate results, the increase in
coated AAOs represents the highest normalized flow rate measured to current consumption is linear with the applied voltage as
date for an EO pump membrane predicted by the theory. The deviation between the theory
and the experiments are due to the errors in estimating Veff
voltage Veff across the membrane needs to be calculated. needed for the theoretical calculations. There, the largest
This is the actual voltage that generates the EO flow. This error comes from measuring the electrodes spacing (i.e.,
Veff is lower than the voltage V applied by the power supply 1.0 ± 0.3 mm) which varies between experiments and
as a result of voltage losses between the electrodes which showed the largest deviation in the theoretical predictions
can be substantial even for millimeter-spaced electrodes. of the flow rate in Fig. 6a.
Miao et al. (2007) reduced these losses by sputtering the Another method for evaluating the performance of EO
platinum electrodes on both surfaces of the AAO. To solve pumps is measuring the maximum pressure achieved. The
for Veff, a simple potential model by Yao et al. (2003) is measurements are conducted using the setup described in
used Fig. 3 with the pressure transducer. A 1 mM phosphate
Veff ¼ V  Vdec  2IRel ð12Þ buffer is used due to the higher flow rates obtained previ-
ously. Results are compared to published data in Fig. 7. It
where Vdec is the electrode decomposition voltage and shows that the SiO2-coated AAOs generated a slightly
combines the voltage needed for the reactions at the greater maximum pressure than the APS (NH2–(CH2)3–
cathode and the anode and the over-potential voltage Si(OCH3)3)-coated AAOs from Chen et al. (2010). This is
needed to start the process. It tends to be small (i.e., 4 V for most likely due to the pore size difference of 150–200 nm.
1 mM borate buffer Yao et al. 2003), and it varies with the As noted by Chen et al. (2007) and shown by Eq. (8), smaller
molarity, the reactants, the electrode material, and current pore sizes tend to produce a greater maximum pressure.
supplied. In the other hand, the voltage drop occurring The f potential is an important parameter in character-
between the electrodes and the membrane can be much izing membranes’ EOP capabilities. It can be calculated
larger and is given by IRel where the resistance of the fluid from Pz;max ; Qmax ; and Imax measurements using the fol-
between the electrode and membrane is Rel = D/AelK lowing relations (Yao et al. 2003; Reichmuth et al. 2003)
where D is the distance between the electrodes, and Ael is
the electrode area. It can also be approximated experi- Pz;max 8ef
¼ 2 ð13Þ
mentally by removing the membrane, calculating the I–V Veff a
curve and measuring the slope (Cao et al. 2012). Qmax ef
The theoretical flow rate Eq. (5) can now be calculated ¼ g ð14Þ
Imax gK
by using the Veff for each applied V. As predicted by the
theory, the measured flow rates of the SiO2-coated AAOs Theoretically, the f potentials for the 10-nm SiO2-coated
increased linearly with voltage as shown in Fig. 6a. It AAOs and for borosilicate glass capillaries should be
shows that the flow rate is higher at higher molarities. comparable due to similar chemical composition. Our

123
Microfluid Nanofluid (2013) 15:859–870 865

a
0.6
30
0.5

0.4
20
Q (mL/min)

Pz (kPa)
0.3

0.2 10

0.1

0
0.0 0 10 20 30
0 10 20 30
V (V) V (V)
b
0.6 Fig. 7 Pressure measurements for SiO2-coated AAOs using 1 mM
phosphate buffer with theoretical predictions (solid line) along with
0.5 data from Arulanandam and Li (2000) that used an APS-coated AAO
with 200-nm pores and deionized water

0.4
Q (mL/min)

measurements for the f potential in borosilicate capillaries


0.3 using current monitoring (Arulanandam and Li 2000) gives
-97 mV for 0.1 mM phosphate buffer (-90 mV in Kirby
0.2
and Hasselbrink 2004) and -120 mV for 1 mM phosphate
buffer (-100 mV in Kirby and Hasselbrink 2004). These f
potential values are much larger than those obtained for
0.1
SiO2-coated AAO membranes using our measurements in
combination with Eq. (14) and listed in Table 2. Briefly,
0.0
0.01 0.1 1 10 100 when comparing the measured f potential values to those in
Molarity (mM) Xu et al. (2011) and Miao et al. (2007) in Table 2, their
c values are higher. It is speculated that this is due to the
30
fabrication process and membrane treatment applied which
varied significantly from the vendor used (Synkera Inc) to
25
those in Xu et al. (2011) and Miao et al. (2007) which must
affect the properties of the SiO2 coating. Following the
20 discussion by Cao et al. (2012) where it has been reported
significant differences in f potential depending on the
I (mA)

15 method used, this is also calculated using Eq. (13) to give


-11 mV. This similarity among values calculated from
10 two different measurements supports that these membranes
have a f potential between -11 and -13 mV.
5 The EO efficiency of the AAO can be calculated from
the ratio of the hydraulic output power to the electrical
input power. It can also be written in terms of the measured
0
0 10 20 30 variables, Qmax and Pmax (Chen et al. 2007), to yield
V (V)
QP Qmax Pmax
v¼ ¼ : ð15Þ
Fig. 6 Effect of a applied voltage and b solution molarity on flow IV 4IV
rates and effect of c molarity on current for EO pumps. For a and c,
measurements are compared to theoretical predictions with similar
Using Eq. (15), the calculated experimental efficiencies
trends and linear behavior. In b measurements show that molarities for the SiO2-coated AAOs with and without the Pt-coated
greater than 2 mM do not significantly contribute to flow rate increase surfaces were 0.161 and 0.028, respectively. These values

123
866 Microfluid Nanofluid (2013) 15:859–870

Table 2 Measured f potential for SiO2-coated AAOs and published f potential values for different types of membranes
Source Membrane Molarity (mM) Solution pH Zeta potential (mV)

Present AAO w/SiO2 0.5, 1, 2, 10 Phosphate 6.5 -8, -13, -13, -21
Miao et al. (2007) AAO N/A DI water 7.0 -19.3
Miao et al. (2007) AAO w/H2SO4 N/A DI water 7.0 -30
Xu et al. (2011), Miao et al. (2007) AAO w/SiO2 N/A DI water 5.8, 7 -50, -42
Prakash et al. (2006) Alumina 0.2 NaCl Phosphate 10.3 -9, -30
Prakash et al. (2006) Silica 0.2 NaCl Phosphate 10.3 -63,-75, -6.3
Cao et al. (2012) MCP N/A DI water N/A -15
Wang et al. (2012) PC N/A DI water N/A -23

contrast with the higher values in the literature due to their 100
ff = =
1 (1
t+ /+(t−t+) + t− )
higher zeta potentials. For instance, an efficiency of 0.900
is obtained for SiO2- and Pt-coated AAOs by Miao et al.
(2007) although conductivity and pump resistance were 10
similar. Also, an v = 0.100 is obtained for APS-coated
AAOs by Chen et al. (2010) or an v = 0.430 is obtained

Q (mL/min)
for bare AAOs by Chen et al. (2007). These higher values
can be attributed to the lower current consumption. 1

Although we would prefer membranes with a higher


value for improved performance and efficiency values, this
lower values do not affect the scope of the method 0.1
proposed for enhancing the flow rate by applying an
asymmetric bipolar rectangular voltage waveform for EO
pumping.
0.01
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
4.2 Bipolar rectangular voltage waveform EO pump V+t+/V-t-
performance
Fig. 8 Measured flow rate for an applied asymmetric bipolar
Having established a baseline for EO pump performance rectangular voltage to SiO2-coated AAO with 1 mM (closed symbols)
and 0.1 mM (open symbols) phosphate buffer. Results are evaluated
when a constant voltage is applied, we now consider the in terms of the area ratio V?t?/V-t-. An asymmetric voltage as high
EO pump capabilities when an asymmetric bipolar rect- as ?1,800/-900 V is applied generating a flow rate of 2.95 ml/min
angular voltage is applied. The same setup used for the DC with no gas generation
voltage tests described in Fig. 3 is used for these flow
measurements. Figure 8 shows the measured flow rates A similar behavior is observed for all the other V?/V-
when a bipolar rectangular pulse wave is applied to SiO2- cases considered in Fig. 8. Also, 1 mM phosphate buffer
coated AAO membranes. Results are plotted as a function (solid symbols) generated a greater flow rate than 0.1 mM
of the ratio V?t?/V-t- where V?t? and V-t- correspond to phosphate buffer (open symbols) similar to the case for
the area underneath the positive and negative part of the constant voltage.
waveform in Fig. 1. By varying these positive and negative What is remarkable about these results is not so much
areas, it can be shown that the flow rate is driven by the that a flow rate can be generated by an asymmetric bipolar
greater area and that for equal areas (V?t?/V-t- = 1) the rectangular pulse wave, but rather the unexpected high
flow rate is negligible. For instance, let’s consider an voltages that can be applied with negligible gas generation.
example where V?/V- = ? 80/-20 V shown by the solid For instance, when a constant voltage is applied, gas gen-
circles in Fig. 8. For this case, when the areas are the same eration is observed between 30 and 50 V for the buffers
(i.e., t?/t- = 1/4 which corresponds to t? = 0.2 ms, t- = and concentrations used. In the other hand, when an
0.8 ms for a frequency of 1,000 Hz), the flow rate is about asymmetric pulse is applied, there was no gas generation
0.04 ml/min. In the other hand, when the areas are not the for voltage as high as 1,800 V. Furthermore, higher volt-
same (i.e., t?/t- [ 1/4 or t?/t- \ 1/4) the flow rate ages might be possible but could not be tested due to
increases significantly until reaching a plateau of *0.8 and testing hardware limitations. The effect of the frequency,
*0.6 ml/min for V?t?/V-t- of 0.2 and 1.7, respectively. 1/(t? ? t-), on gas generation is also considered. The

123
Microfluid Nanofluid (2013) 15:859–870 867

results show that changing the frequency from 100 to a


0.7
1,000 Hz made no difference on the flow rate generated.
Also, negligible faradaic reactions are present except on
0.6
the limits when V?t?/V-t- *0.2 or *1.7 where a small
amount is observed but not to the extent seen for constant
0.5
voltages. Higher frequencies could not be tested due also to

Q (mL/min)
amplifier limitations. In the other hand, frequencies below 0.4
50 Hz resulted in gas generation which shows that this is a
lower bound frequency value where the proposed technique 0.3
cannot be applied. This is also an indicator that the time
scale for gas generation H2(g) and/or O2(g) is probably of 0.2
this order at least for the cases studied here. The highest
flow rate measured is 2.95 ml/min at ?1,800/-900 V for 0.1
0.1 mM phosphate buffer and 1.69 ml/min at ?400/
-200 V for 1 mM phosphate buffer as shown in Fig. 8. 0.0
As with all EO pumps, power consumption is an 0.001 0.01 0.1 1 10
Pin (W)
important characteristic that needs to be considered. The
b
goal is to achieve higher flow rates Q with the lower power 2.5
consumption Pin possible. Figure 9 shows the EO mea-
sured flow rates and corresponding power consumption
2.0
when a bipolar rectangular pulse wave is applied to SiO2-
coated AAO membranes in a 1 mM phosphate buffer
Q/Pin (mL/min/W)

solution. The plot in Fig. 9a shows the effect that the 1.5
variation of the area ratio V?t?/V-t- has on Q and Pin
when the bipolar rectangular waveform is applied. It shows
that the asymmetric voltage pulsing follows the same trend 1.0
as the constant voltage but it is not as efficient. For these
measurements, the most efficient area ratio is 0.44 and the
least efficient ratio or the worst performer in terms of 0.5
power to flow rate is the equal area pulsing which has
already been shown in Fig. 8. Also, the effect that the
0.0
variation of the voltage ratio has on Q/Pin when the bipolar 0.0 0.5 1.0 1.5 2.0 2.5
rectangular waveform is applied is considered in Fig. 9b. A V+ t+ / V− t−
voltage ratio of 2:1 provided the best flow rate per supplied
power, and as the ratio is increased to 4:1, 7:1 and lastly Fig. 9 Measured flow rates and power consumption when a bipolar
9:1 higher flow rates are obtained but the power con- rectangular voltage waveform is applied to AAO membranes in a
1 mM phosphate buffer. Results in a show the effect that the variation
sumption increases at a faster rate resulting in a less of area ratio V?t?/V-t- has on Q and Pin. Results in b show the effect
effective system. that the variation of voltage ratio has on Q/Pin
Similar to the constant voltage analysis, the performance
of the EO pump can be evaluated by measuring the max-
imum pressure achieved by the SiO2-coated AAOs in These results are comparable to the linear increase in
1 mM phosphate buffer. The measurements are conducted pressure with voltage observed for the constant voltage
using the setup described in Fig. 3 with the pressure analysis in Fig. 7. There is nevertheless a loss in efficiency
transducer. The obtained pressures are plotted as a function estimated to be 0.005 using Eq. (15) which is about
of the ratio V?t?/V-t- in Fig. 10. As was the case for flow *20 % the efficiency obtained in the constant voltage
rate measurements, pressure was maximized for the greater analysis. The results also show that for an equal area ratio,
negative area to positive area ratio. For a fixed area ratio the system was able to generate a small pressure of about
V?t?/V-t- and voltage ratio V?/V- such as 0.45 and 4, 2.1 kPa for ?200/-50 V which is responsible for the small
respectively, the results show an increase in pressure with flow rate observed.
voltage and power with a maximum pressure of 36 kPa for Last, an overall comparison of flow rates obtained in
?200/-50 V. Higher pressures are possible although it terms of the average voltage applied V  from Eq. (2) is
would require modifying the setup to properly support the shown in Fig. 11. By using the average voltage applied, a
membrane which kept breaking at the higher pressures. more direct way of comparing flow rate measurements

123
868 Microfluid Nanofluid (2013) 15:859–870

40

30
Pz (kPa)

20

10

0
0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
V+ t+ / V− t−

Fig. 10 Pressure measurements for SiO2-coated AAOs using 1 mM


phosphate buffer for an applied asymmetric bipolar rectangular
voltage waveform

Fig. 12 Close-up images show a partial frontal view of the platinum


mesh electrode and membrane inside a transparent acrylic housing
during pump operation. a For an applied ?600/-300 V bipolar
rectangular voltage, only a few small bubbles are formed without
affecting the pump operation. b For an applied 50 V DC voltage,
1 many small but also large bubbles form in front and behind the
electrode such as the smaller ones throughout that seem to defocus the
image or the large one in the lower part of the image which prevented
Q (mL/min)

pump operation

by nearly a factor of 30 with a peak voltage increase of


0.1 about 36 times.

4.3 Additional comments on gas generation reduction

When a potential is applied to an electrochemical cell, both


0.01 hydrogen and oxygen ions, and their respective radicals are
1 10 100 1000
(V+t++V-t-)/(t++t- ) formed. The H2 and O2 gas formed dissolves in the elec-
trolyte until it reaches supersaturation. Beyond this
Fig. 11 Flow rates measured in terms of the average voltage applied threshold and in the presence of a nucleation site, crack or
for 1 mM (closed symbols) and 0.1 mM (open symbol) phosphate cavity in the electrode surface, bubbles are initiated on the
buffer. This allowed comparing flow rates obtained by applied DC
voltage with those obtained by applied bipolar rectangular voltage electrode surface (Donose et al. 2012). Part of the dissolved
waveforms gas generated is transformed into the gaseous phase of the
bubbles adhering to the electrode, while the other part
between the DC voltage testing and asymmetric bipolar transferred to the liquid bulk (Vogt 2011). For a bipolar
rectangular voltage testing is possible. Results show that voltage, during the positive pulse, H2 is generated at the
when a DC voltage is applied, faradaic reactions are cathode, and O2 is generated at the anode. During the
important above 30 and 50 V for 1 and 0.1 mM borate negative pulse, the electrode polarity is reversed with H2
buffer solutions, respectively, with maximum delivered forming at the electrode where O2 previously formed and
flow rates of 0.56 ml/min for 1 mM and 0.099 ml/min for vice versa. For a high frequency asymmetric bipolar rect-
0.1 mM. In the other hand, the proposed asymmetric angular voltage, the hydrogen ions and hydroxide ions
voltage pulsing is capable of delivering a flow rate of generated at the electrode compensate each other, and no
2.95 ml/min for a 0.1 mM solution at V   400 V corre- gas is generated. However, in the case of voltage pulsing
sponding to ?1,800/-900 V. This demonstrates that for with different duty cycle, the excess amount of ions gen-
the current case, the proposed method raised the flow rate erated at each electrode react and form gas molecules.

123
Microfluid Nanofluid (2013) 15:859–870 869

These dissolve in the liquid until the gas concentration and 1.7, respectively. These results are independent of the
reaches supersaturation. Considering that the fluid is being frequency of the applied waveform as long as it is greater
convected by this EO pump in an open circuit, the than *50 Hz to avoid faradaic reactions. Furthermore, the
incoming fluid is not supersaturated, and the threshold voltage ratio V?/V- of 2:1 provided the best flow rate per
needs to be reached in each cycle before bubbles can be supplied power. The highest flow rate observed were
generated. It is speculated that this reduces bubble forma- 2.95 ml/min at ?1,800/-900 V for 0.1 mM phosphate
tion. More importantly, the dynamics of bubble formation buffer and 1.69 ml/min at ?400/-200 V for 1 mM phos-
and growth is a complex nonlinear process which is phate buffer.
affected by this voltage pulsing. This process is not well It is remarkable that when an asymmetric pulse is
understood, but it is speculated that the bubbles generated applied, there was negligible gas generation for applied
between cycles do not have sufficient time to grow much voltages as high as 1,800 V, but for DC, voltage gas gen-
larger than the nanochannel’s diameter. This is supported eration is observed between 30 and 50 V for the solutions
by some observations during high field asymmetric bipolar used. Furthermore, the results showed that the proposed
EO pumping of only small bubbles being carried down- method could increase the flow rate output compared to DC
stream with the flow and minimal small bubbles found in EO pumping by one to two orders of magnitude with
the electrodes as shown in Fig. 12a (see also attached similar increased in applied voltages.
supplemental video material). On the other hand, for
[30–50 V DC EO pumping, the bubbles grow quickly and Acknowledgments The authors gratefully acknowledge the finan-
cial support provided for this study by the Office of Naval Research
in just seconds bubbles with millimeter size are present (ONR), Grant No. N00014-11-1-0019, with Dr. Thomas F. Swean of
which continue expanding as shown in Fig. 12b and the Ocean Engineering and Marine Systems Program serving as
eventually prevent pump operation. The presence of bub- Program Manager. The authors are thankful to Mr. Thomas E. Hansen
bles for these applied DC voltages ([50 V) are reported by for all the support obtaining the EO pump video and images.
Vajandar et al. (2007) limiting their tests to values below it.
Furthermore, they reported the current being unstable due
to the bubbles. Thus, the high frequency asymmetric
References
bipolar rectangular voltage permits overcoming this DC
voltage limit, and it results in much higher flow rates than Ai Y, Yalcin SE, Gu D, Baysal O, Baumgart H, Qian S, Beskok A
previously possible. (2010) A low-voltage nano-porous electroosmotic pump. J Col-
loid Interface Sci 350(2):465–470
Arulanandam S, Li D (2000) Determining f potential and surface
conductance by monitoring the current in electro-osmotic flow.
5 Conclusion J Colloid Interface Sci 225(2):421–428
Berrouche Y, Avenas Y, Schaeffer C, Hsueh-Chia Chang, Ping Wang
The experimental results have shown that the use of an (2009) Design of a porous electroosmotic pump used in power
asymmetric bipolar rectangular waveform for EO pumps electronic cooling. IEEE Trans Ind Appl 45(6):2073–2079
Brask A, Snakenborg D, Kutter JP, Bruus H (2006) AC electroos-
allows applying higher voltages than those possible in DC motic pump with bubble-free palladium electrodes and rectifying
EO pumping while generating much higher flow rates and polymer membrane valves. Lab Chip 6(2):280
preventing gas generation. Results are compared to a Buie CR, Posner JD, Fabian T, Cha S-W, Kim D, Prinz FB, Eaton JK,
baseline measurement for DC voltage EO pumping in Santiago JG (2006) Water management in proton exchange
membrane fuel cells using integrated electroosmotic pumping.
SiO2-coated AAO membranes. From the DC EO pumping J Power Sources 161(1):191–202
results, faradaic reactions are observed at about 30–50 V Burke JM, Smith CD, Ivory CF (2010) Development of a membrane-
depending on the buffer concentration. For instance, for 1 less dynamic field gradient focusing device for the separation of
and 0.1 mM phosphate buffers, gas generation is observed low-molecular-weight molecules. Electrophoresis 31(5):902–909
Cao Z, Yuan L, Liu Y-F, Yao S, Yobas L (2012) Microchannel plate
at 30 and 50 V, respectively, while the flow rates measured electro-osmotic pump. Microfluid Nanofluid 13(2):279–288
are 0.561 and 0.099 ml/min, respectively. These values and Chen L (2005) Fabrication and characterization of a multi-stage
their normalized results are comparable to theory and to electroosmotic pump for liquid delivery. Sensors Actuators B
flow rate measurements for AAO membranes in the liter- Chem 104(1):117–123
Chen L, Ma J, Guan Y (2004) Study of an electroosmotic pump for
ature (Vajandar et al. 2007). liquid delivery and its application in capillary column liquid
Results from the bipolar rectangular-wave EO pumping chromatography. J Chromatogr A 1028(2):219–226
measurements are evaluated in terms of the area ratio Chen W, Yuan JH, Xia XH (2005) Characterization and manipulation
V?t?/V-t-. For instance, for V?/V- = ? 80/-20 V of the electroosmotic flow in porous anodic alumina membranes.
Anal Chem 77(24):8102–8108
when the area ratio is 1, then flow rate is negligible Chen Y-F, Li M-C, Hu Y-H, Chang W-J, Wang C-C (2007) Low-
(0.04 ml/min), but it increased significantly until reaching voltage electroosmotic pumping using porous anodic alumina
a plateau of *0.8 and *0.6 ml/min for area ratios of 0.2 membranes. Microfluid Nanofluid 5(2):235–244

123
870 Microfluid Nanofluid (2013) 15:859–870

Chen J-K, Weng C-N, Yang R-J (2009) Assessment of three AC Pikal MJ (1992) The role of electroosmotic flow in transdermal
electroosmotic flow protocols for mixing in microfluidic chan- iontophoresis. Adv Drug Deliv Rev 9(2):201–237
nel. Lab Chip 9(9):1267 Prakash P, Grissom MD, Rahn CD, Zydney AL (2006) Development
Chen Y-F, Hu Y-H, Chou Y-I, Lai S-M, Wang C-C (2010) Surface of an electroosmotic pump for high performance actuation.
modification of nano-porous anodic alumina membranes and its J Membr Sci 286(1–2):153–160
use in electroosmotic flow. Sensors Actuators B Chem Reichmuth DS, Chirica GS, Kirby BJ (2003) Increasing the perfor-
145(1):575–582 mance of high-pressure, high-efficiency electrokinetic micro-
Donose BC, Harnisch F, Taran E (2012) Electrochemically produced pumps using zwitterionic solute additives. Sensors Actuators B
hydrogen bubble probes for gas evolution kinetics and force Chem 92(1–2):37–43
spectroscopy. Electrochem Commun 24:21–24 Rice CL, Whitehead R (1965) Electrokinetic flow in a narrow
Garg R, Kumar V, Kumar D, Chakarvarti SK (2012) Electrical cylindrical capillary. J Phys Chem 69(11):4017–4024
transport through micro porous track etch membranes of same Selvaganapathy P, Yit-shun Leung Ki, Renaud P, Mastrangelo CH
porosity. Mod Phys Lett B 26(31):1250209 (2002) Bubble-free electrokinetic pumping. J Microelectromech
Heuck F, Van der Ploeg P, Staufer U (2011) Deposition and Syst 11(5):448–453
structuring of Ag/AgCl electrodes inside a closed polymeric Stuetzer OM (1959) Ion drag pressure generation. J Appl Phys
microfluidic system for electroosmotic pumping. Microelectron 30(7):984–994
Eng 88(8):1887–1890 Vajandar SK, Xu D, Markov DA, Wikswo JP, Hofmeister W, Li D
Kim Y, Cha M, Choi Y, Joo H, Lee J (2013) Electrokinetic separation (2007) SiO2-coated porous anodic alumina membranes for high
of biomolecules through multiple nano-pores on membrane. flow rate electroosmotic pumping. Nanotechnology 18(27):
Chem Phys Lett 561–562:63–67 275705
Kirby BJ, Hasselbrink EF (2004) Zeta potential of microfluidic Vogt H (2011) On the gas-evolution efficiency of electrodes
substrates: 1. Theory, experimental techniques, and effects on I-Theoretical. Electrochim Acta 56(3):1409–1416
separations. Electrophoresis 25(2):187–202 Wang P, Chen Z, Chang H-C (2006) A new electro-osmotic pump
Kohlheyer D, Eijkel JCT, Schlautmann S, Van den Berg A, based on silica monoliths. Sensors Actuators B Chem 113(1):
Schasfoort RBM (2008) Bubble-Free Operation of a Microflu- 500–509
idic Free-Flow Electrophoresis Chip with Integrated Pt Elec- Wang C, Wang L, Zhu X, Wang Y, Xue J (2012) Low-voltage
trodes. Anal Chem 80(11):4111–4118 electroosmotic pumps fabricated from track-etched polymer
Kwon K, Park C-W, Kim D (2012) High-flowrate, compact electro- membranes. Lab Chip 12(9):1710
osmotic pumps with porous polymer track-etch membranes. Sens Xu Z, Miao J, Wang N, Wen W, Sheng P (2011) Digital flow control
Actuators, A 175:108–115 of electroosmotic pump: Onsager coefficients and interfacial
Levine S, Marriott JR, Robinson K (1975) Theory of electrokinetic parameters determination. Solid State Commun 151(6):440–445
flow in a narrow parallel-plate channel. J Chem Soc, Faraday Yao S, Santiago JG (2003) Porous glass electroosmotic pumps:
Trans 2(71):1–11 theory. J Colloid Interface Sci 268(1):133–142
Li PCH, Harrison DJ (1997) Transport, manipulation, and reaction of Yao S, Hertzog DE, Zeng S, Mikkelsen JC, Santiago JG (2003)
biological cells on-chip using electrokinetic effects. Anal Chem Porous glass electroosmotic pumps: design and experiments.
69(8):1564–1568 J Colloid Interface Sci 268(1):143–153
Lin C-W, Yao S, Posner JD, Myers AM, Santiago JG (2007) Toward Yao S, Myers AM, Posner JD, Rose KA, Santiago JG (2006)
orientation-independent design for gas recombination in closed- Electroosmotic pumps fabricated from porous silicon mem-
loop electroosmotic pumps. Sensors Actuators B Chem 128(1): branes. J Microelectromech Syst 15(3):717–728
334–339 Zeng S, Chen CH, Mikkelsen JC, Santiago JG et al (2001) Fabrication
McBride SE (1999) Balanced asymmetric electronic pulse patterns and characterization of electroosmotic micropumps. Sensors
for operating electrode-based pumps. US Patent 5,964,977, filed Actuators B Chem 79(2–3):107–114
21 Mar 1997 and issued 12 Oct 1999
Miao J-Y, Xu Z-L, Zhang X-Y, Wang N, Yang Z-Y, Sheng P (2007)
Micropumps based on the enhanced electroosmotic effect of
aluminum oxide membranes. Adv Mater 19(23):4234–4237

123

You might also like