Sei sulla pagina 1di 11

424

M. R. Pearson and J. Seyed-Yagoobi: Advances in Electrohydrodynamic Conduction Pumping

Advances in Electrohydrodynamic Conduction Pumping


Matthew R. Pearson and Jamal Seyed-Yagoobi
Two-phase Flow and Heat Transfer Enhancement Laboratory
Mechanical, Materials, and Aerospace Engineering Department
Illinois Institute of Technology
10 W. 32nd Street, Chicago, IL 60616, USA

ABSTRACT
A detailed review of progress-to-date in the field of electrohydrodynamic conduction
pumping is presented. Analytical, numerical, and experimental research studies that
have been published in this field are summarized and a detailed scientific overview of
the conduction pumping model is provided. Conduction pumping has already been
investigated successfully to resolve several commonplace issues relating to pumping,
flow-distribution, and heat transfer and these applications are also presented.
Index Terms Electrohydrodynamics, conduction phenomenon, dielectric fluids,
non-mechanical pumping.

NOMENCLATURE
b
C0
C
c
cp
D
d

charge mobility coefficient


dimensionless parameter, defined by equation (11)
same as C0 but based on length scale instead of d.
concentration of neutral species
specific heat
charge diffusion coefficient
characteristic length (typically electrode spacing distance or liquid film
thickness)
e
elementary charge of an electron
E
electric field intensity
E
electric vector field
F() see equation (19)
electric body force
fe
G
dimensionless parameter, defined by equation (13)
g
gravitational acceleration vector
modified Bessel function of first kind
I1
j
current density
Boltzmann constant
kB
constant dissociation rate
kD
field enhanced dissociation rate, see equation (19)
kd
constant recombination rate
kR
mobility ratio, defined by equation (16)
M0
m
number of electrode pairs
n
negative charge density
n
unit normal vector to surface
P
pressure
p
positive charge density
Re
Reynolds number
ReEHD EHD Reynolds number, defined by equation (17)
T
temperature
t
time
ionic transit time
tT
u
velocity
V
applied voltage

dimensionless parameter, defined by equation (12)

electric permittivity

efficiency

heterocharge layer thickness


Debye layer thickness
D

dynamic viscosity

mass density
net charge density
e
electric conductivity
e
Manuscript received on 31 May 2008, in final form 15 September 2008.

potential field
space charge relaxation time (/e)
dimensionless dissociation rate parameter, defined by equation (19)

SUBSCRIPTS
+
positively-charged ions

negatively-charged ions
eq
equilibrium
ohm ohmic
SCL space charge limited
SUPERSCRIPTS
*
dimensionless

INTRODUCTION

All electrohydrodynamic (EHD) pumping methods rely on


the interaction between electric fields and the flow fields of a
dielectric fluid. The electric body force that is produced by
this interaction is [1],
1
1
f e = e E E 2 + E 2

2
2 T

(1)

The first term represents the Coulomb force that acts on free
charges in an electric field. The second and third terms, titled
dielectrophoretic and electrostriction forces respectively,
represent the polarization force acting on polarized charges.
The dielectrophoretic term requires a gradient in permittivity
that does not exist in an isothermal liquid, while the
electrostriction term is relevant only for compressible fluids.
EHD pumping has shown extensive potential due to its
simple, lightweight, non-mechanical design, low power
consumption, low acoustical noise, and the ease with which
pumping can be controlled by adjusting the applied voltage.
A variety of EHD pumping mechanisms are based on the
Coulomb force: conduction pumping, induction pumping, and
ion-drag pumping. These methods differ in the manner of
generating the non-zero net charge density that is required for a
Coulomb force to exist. Ion-drag pumping relies on the injection
of ions into the liquid from sharp liquid/solid interfaces, while
induction pumping uses an ac travelling wave to attract and repel

1070-9878/09/$25.00 2009 IEEE

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 16, No. 2; April 2009

charges induced in the liquid due to gradients or discontinuities


of electric conductivity, often due to temperature gradients [2].
Conduction pumping produces a non-zero charge density in
heterocharge layers near the electrodes through bulk electric
conduction through the liquid. As a result, the problems of iondrag pumping (specifically, degradation of the electrical
properties of the working fluid and potentially hazardous
operation) and induction pumping (specifically, the need for a
gradient in electric conductivity) can be avoided while still
maintaining the ability to pump using the Coulomb force [2].
For a neutral species, AB, and its positive and negative ions, A+
and B there is reversible process of dissociation and
recombination:

AB A+ + B

(2)

Under conditions of low electric field intensity, the dissociation


and recombination rates are in dynamic equilibrium [3]. Any
ions that are present in the fluid, for instance due to impurities,
will be drawn along lines of constant electric field, attracted to
the charges of opposite polarity on the electrode surface. This
rearrangement of charges causes an electric double layer, with
charges of one polarity on the solid electrode surface and
charges of opposite polarity on the liquid side of the surface.
This layer is further divisible into two sub-regions. The first
region is a compact layer, typically about 0.5 nm thick, where a
layer of ions is strongly attracted to the surface and
immobilized as a result. The second region is the diffuse layer,
where the ionic concentration gradually decreases from that of
the compact layer to that of the liquid bulk and where ions
remain mobile.
When the electric field exceeds a certain threshold value (on
the order of 1 kV/cm, depending on the working fluid [4]), the
rate of dissociation begins to exceed the rate of recombination,
with the difference in rates further increasing with increasing
electric field intensity. Therefore, an additional layer forms,
known as the heterocharge layer, where the dissociation and
recombination processes are not in equilibrium. Atten and
Seyed-Yagoobi [3] describe these heterocharge layers as
characteristic of dielectric liquids having sufficiently low
conductivity; generally the case for non-polar or mildly-polar
fluids. The charges in the heterocharge layer are attracted to the
nearby electrode, thus causing a propensity for bulk fluid flow.
Figure 1, first presented by Jeong et al [5], provides a very clear
graphical representation of the way in which a non-uniform
electric field, coupled with attractive forces from the
heterocharge layers to the electrodes, can produce a net fluid
flow when the electrodes are suitably designed. The need for
low conductivity fluids relates to the Debye length, D, which
can be viewed as the typical length of diffusion of the ions

during the relaxation time of the space charge, = /e. If the


conductivity is too high, the heterocharge layers have a
thickness on the order of D and the electric field intensity in
the layers greatly exceeds that in the bulk [3].
It should be noted that understanding of the conduction
pumping phenomenon occurred recently compared to the other
principal EHD pumping techniques (ion-drag pumping,
induction pumping, and pumping/extraction based on the use of
polarization forces). Some early studies of EHD incorrectly
attributed the observed EHD pumping motion to polarization
forces [68] because it was not well-understood that the
electrostriction term in equation (1) vanishes for the case of an
incompressible fluid. It is now accepted that conduction
pumping was the primary pumping mechanism in these studies.
This work summarizes the progress that has been made todate in the field of EHD conduction pumping since the
scientific theory was first formulated and presented by Atten
and Seyed-Yagoobi [9]. Section 2 presents the theoretical
model that has been developed and refined to explain the
conduction pumping phenomenon. Section 3 summarizes the
initial numerical and experimental studies that focused on the
pumping of a single-phase (liquid) fluid. These studies lay the
foundation for understanding of the conduction pumping
phenomenon. More recent studies, which have investigated
two-phase pumping of liquid films, are summarized in Section
4. Finally, Section 5 summarizes the applications in which
conduction pumping has been investigated and where the
technology has shown great potential and applicability in
several problems relating to pumping, flow-distribution, and
heat transfer. Throughout this review paper, some nomenclature
has been adjusted from the original sources so that a consistent
usage can be presented.

THEORETICAL MODEL

A simple theoretical model of the conduction pumping


phenomenon was first presented by Atten and Seyed-Yagoobi
[3]. The model was based on the reversible dissociationrecombination process described by equation (2), and at
thermodynamic equilibrium, it follows that
k D c = k R peq neq = k R peq2 = k R neq2

(3)

where electroneutrality requires that peq = neq. As a result, the


value of neq is given as
neq = k D c k R

(4)

The rate of dissociation was assumed to be very small, such


that neq/c << 1, and the mobility and diffusion coefficients of
positive and negative ions were assumed to be equal, such
that b+ = b = b and D+ = D = D. Subsequently, the constant
recombination rate was assumed to be kR = 2b/ based on the
upper bound given by Langevin [10]. Length variables were
nondimensionalized by the inter-electrode spacing d, and the
other dimensionless variables were defined as

p* =
Figure 1. Schematic of heterocharge layers, the attractive force toward the
electrode, and the use of suitable electrode geometries to generate pumping
using this force [4].

425

E
n
, n* =
, E* =
, * = ,
V d
V
neq
neq

P* =

d
2

, t* =

t
tT

(5)

M. R. Pearson and J. Seyed-Yagoobi: Advances in Electrohydrodynamic Conduction Pumping

426

The conservation equations for the positive and negative


species are
p*
+ * ( p*E* * p* ) = 2C0 (1 p* n* )
t *
n*
+ * ( n*E* * n* ) = 2C0 (1 p* n* )
t *

(7)

* E* = C0 ( p* n* )

(8)

E* = * *

(9)

By assuming that the fluid is static, the momentum equation is


simply a balance of pressure gradient and electric body force:
* P* = G ( p* n* ) E*

(10)

The dimensionless parameters in equations (6)(10) are


neq d 2

(11)

D k BT
=
bV
eV
neq d 2V
G=
2

(12)
(13)

Noting that is very small for the large applied voltages of


interest, the diffusion terms have been neglected in some
studies. In this case, the ionic species and electric field depend
solely on the parameter C0, with the parameter G needed only
for computing the pressure field, which is decoupled from the
electrical equations. It was further noted that C0 is a
dimensionless form of neq and could be viewed as: 1) the ratio
of ionic transit and relaxation times: tT/ ~ 2C0, and 2) the
ratio of ohmic and space charge limited currents:
johm/jSCL = 16C0/9 [3].
The most significant evolution of the model was presented
by Jeong and Seyed-Yagoobi [11] by incorporating the effect
of fluid flow (i.e. u* 0) As a result, the potential, electric,
ionic, and flow fields became coupled due to the presence of
convective terms in the charge conservation equations. The
dimensionless continuity equation, along with the full
momentum equation containing fluid convection and
diffusion terms, were
* u* = 0

(u

) u =
*

1
Re EHD

(14)

P + Re1EHD u + M C0 ( p n ) E*
*

*2

2
0

(15)
where the mobility ratio M0 and the EHD Reynolds number,
ReEHD, were defined as
M0 =
Re EHD

b2
bV
=

u* =

(6)

The electric field with space charges and uniform permittivity


is given by Gauss Law as

C0 =

and the velocity field was nondimensionalized using

(16)
(17)

u
bV d

(18)

Note that the parameter G in the early model is related to the


parameters in the later model by G = C0(ReEHDM0)2.
Jeong and Seyed-Yagoobi [11] further improved the ionic
conservation equations (6) and (7) by incorporating Onsagers
field-enhanced dissociation model which replaces kD in
equation (3) with kd defined as
kd = k D F ( ) = k D

I1 ( 2 )

,=

e3 E
4 k B2T 2

(19)

With this modification and the incorporation of convective


terms, the dimensonless charge conservation equations (6)
and (7) become
p*
+ * p* ( u* + E* ) * p* = 2C0 ( F ( ) p* n* ) (20)
t *
n*
+ * n* ( u* E* ) * n* = 2C0 ( F ( ) p* n* ) (21)
t *

Note that F() = 1 when the dissociation rate coefficient is


independent of electric field, i.e. = 0. When the dissociation
rate is dependent on the electric field, > 0 and F() > 1.
2.1 BOUNDARY CONDITIONS
For the purpose of numerical or analytical solutions to the
governing equations, the following electrical boundary
conditions have generally been adopted [12]. Some earlier
studies (e.g. [5]) considered the potential field distribution in
both the liquid and the solid phases, with p* = n* = 0 assumed
in the solid phase. At the high voltage and ground electrodes,
the potential is fixed, ion-injection is assumed to not be
present, and the flux of positive and negative charges on the
ground and high-voltage electrodes, respectively, is assumed
to be small. Therefore, the high voltage and ground electrode
boundary conditions are, respectively,

* = 1 , p* = 0 , n * n* = 0
* = 0 , n * p* = 0 , n* = 0

(22)

(23)

At any surface not electrically connected to a high voltage


power supply or to ground, the zero-flux condition requires
that
2

p*n * * , n * n* = n*n * * (24)

When the liquid conductivity is low and the length scales are
sufficiently large (on the order of millimeters), the effect of
electric double layer and the induced zeta potential are
negligible and the conditions simplify to
n * * = 0 , n * p* = 0 , n * n* = 0
(25)

n * p* =

For two-phase studies that feature a liquid-vapor interface, the


boundary conditions for an insulating surface, equation (25),
have been used, since the vapor electrical conductivity of
most fluids is several orders of magnitude smaller than the
liquid electrical conductivity.

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 16, No. 2; April 2009

SINGLE-PHASE STUDIES

3.1 THEORETICAL/NUMERICAL
Atten and Seyed-Yagoobi [3] applied their steady-state
model to three electrode configurations: parallel plane
electrodes, point-plane electrodes, and blade-plane electrodes.
Of particular interest was the point-plane configuration, since
it presented significant field asymmetry that could be
harnessed for pumping head generation. The thickness of the
heterocharge layer was approximated in two asymptotic limits
and the net force exerted on the liquid bulk was estimated, by
a series of approximations and simplifications, to vary
linearly with and V2.
Jeong et al [5] were the first to take the model developed
by Atten and Seyed-Yagoobi [3] and obtain a full numerical
solution to the governing equations (6)(10) to numerically
investigate the prospects of an electrode configuration
consisting of a ground ring and a high-voltage hollow tube
centered in the flow channel (refer to Table 1 for a sketch of
a hollow-tube electrode). The study neglected fluid motion
and focused on static pressure-head generation. The
computation considered the potential field in both the liquid
and the solid insulating material that surrounds the liquid
channel, with appropriate boundary conditions at the solidliquid interface. The numerical solution provided valuable
insight by allowing the easy visualization of contours of
potential field intensity, electric field intensity, and
positive/negative/net charge density. From the results, the
heterocharge layers around the electrodes were clearly
visible. The numerical results suggested that a pressure
generation of 1 kPa was plausible by combining 15 electrode
pairs in series, with sufficient spacing to minimize
interaction between pairs.
Jeong and Seyed-Yagoobi [11] used the addition of the
continuity and momentum equations to the theoretical model
in order to numerically solve the fluid motion inside a twodimensional enclosure when subject to an electric field
imposed by the electrodes shown in Figure 2. Numerical
results showed two counter-rotating vortices to each side (left
and right) of the high voltage electrode. To study the effect
that the charge convection term and the field-enhanced
dissociation term, F(), could have on the charge distribution,
three distinct cases were considered and compared: 1) zero
charge convection (p*u* = n*u* = 0) and field-independent
dissociation (F() = 1), 2) charge convection and fieldindependent dissociation, and 3) both charge convection and
field-enhanced dissociation. Charge convection was shown to
deform the heterocharge layers in certain regions, while fieldenhanced dissociation served to cause heterocharges with
high-density values to exist in larger areas around the
electrodes, implying an increase in the resultant net electric
body force compared to the field-independent dissociation
case. This higher body force resulted in stronger flow
circulation within the enclosure.
Feng and Seyed-Yagoobi [13] investigated equations (6)
(8) with the same simplifications as Atten and Seyed-Yagoobi
[3] (i.e. steady-state, spatial variations limited to x-direction,
diffusion terms neglected). However, the characteristic length
scale used to nondimensionalize was the characteristic

427

heterocharge layer thickness, . Consequently, the parameter


C0 was replaced with C in equations (6) and (7), defined as
C =

neq d

(26)
V
The characteristic heterocharge layer thickness, , was
defined as the point x = at which the non-dimensional
positive ion density, p*, is equal to 0.99, corresponding to the
location where the charge concentration has almost reached
the equilibrium level. Furthermore, the exact analytical
solutions to equations (6)(8) were used, and from these
solutions the parameter C was shown to assume a constant
value of 1.8, meaning that the parameter is independent of the
properties of the dielectric fluid and the applied electric field
provided that << d. The theoretical solution was shown to
have an error of just 0.0001% even for = d/3, suggesting
that the developed model is widely applicable. The analysis
predicted the thickness of the heterocharge layer, as a function
of the fluid properties and the electric field intensity, as

V
= 1.8
, neq e
(27)
2b

neq d

For the refrigerant HCFC-123, the heterocharge layer


thickness was calculated as 0.03 m and 0.3 mm at applied
electric field intensities of 100 V/m and 1 MV/m,
respectively.
Feng and Seyed-Yagoobi [13] next considered a grounded
ring electrode and a perforated high voltage electrode (refer to
Table 1 for a sketch of a perforated electrode). A numerical
analysis of the two-dimensional electric field revealed that the
electric field decreased significantly in the holes of the
perforated electrode, signifying that the hetercharge layer
would be mostly confined to the non-punctured areas of the
electrode. To provide simple theoretical approximations, the
electric field in the vicinity of the perforated electrode was onedimensionally treated as the electric field between parallel
electrodes. Based on this simplification, the pressure head
generation due the axial electric body force was estimated as
P = 0.85m

V2
d2

(28)

where m denotes the number of electrode pairs. To account


for the fact that the heterocharge layer may not extend into the
punctured sections of the perforated disk, the authors correctly
predicted that multiplying the right-hand-side of equation (28)
by the percentage of the cross-sectional area that is not

Figure 2. Schematic of circulation enclosure studied numerically by Jeong


and Seyed-Yagoobi [11].

M. R. Pearson and J. Seyed-Yagoobi: Advances in Electrohydrodynamic Conduction Pumping

428

perforated would yield a much better estimate of the pressure


generation.
In a second paper, Feng and Seyed-Yagoobi [14]
considered the same electrode configuration and used the
same one-dimensional approximate formulation of the electric
field and flow field. This treatment amounted to the field
between parallel plane electrodes that are permeable, with an
arbitrary fluid velocity in the direction perpendicular to the
electrode plane, as shown in Figure 3. For much of the study,
only the left half of this domain, near the permeable anode,
was considered. The study focused on providing asymptotic
theoretical analyses of the EHD pumping performance in the
presence of liquid flow.
Results considering a perfectly permeable parallel plane
electrode showed that when the pumping velocity reaches the
mobility limit (u* = 1), the EHD-induced pressure generation
decreases to zero because the convective cross-flow serves to
blow the heterocharge layer off the permeable electrode.
Since the actual electrode is not perfectly permeable, with
significant solid area on which the flow stagnates, an inviscid
two-dimensional flow simplification was used to predict the
convective flow around the actual, imperfectly-permeable
electrode. This modification demonstrated that even at the
critical velocity of u* = 1, non-zero pressure generation
remained. The authors noted that while the inviscid velocity
profiles differ significantly from the actual profiles near the
perforated edges of the electrode, the influence of these
regions on the pressure generation and electric current
consumption was expected to be small since these phenomena
primarily depend on the solid electrode surface normal to the
electric field, where the heterocharge layer is present.
By again normalizing all lengths by the heterocharge layer
thickness, the theoretical model depended only on the flow
velocity, applied electric field, and the properties of the
working fluid. The analysis revealed that the convection in the
fluid had a significant effect on the charge distribution and the
pumping pressure head, and that the electric conductivity of
the working fluid does not affect pressure generation provided
that the conductivity is low enough to allow the electric field
to be established. Furthermore, an approximate expression for
the pumping efficiency was given as

= 3.6

V b
f ( u* )
e d d

(29)

3.2 EXPERIMENTAL
A number of electrode designs have been considered for the
pumping of single-phase liquid dielectrics using conduction
pumping. Not-to-scale sketches that illustrate the most
commonly utilized electrodes are shown in Table 1. In all
single-phase studies, the ground electrodes have been ring
electrodes that are flushed to the circumference of the tube.
For additional details about electrodes and dimensions,
readers should refer to the references. Additional electrode
designs that have been considered for the pumping of singlephase and two-phase fluids can be found in [15, 16].
For the experimental portion of their study, Atten and
Seyed-Yagoobi [3] considered two different electrode
configurationssingle-tube and multi-needle. The minimum
axial distance between the electrode pair (high-voltage and
ground) was 4.5 mm (single-tube) and 4.2 mm (multi-needle).
Both electrodes were designed to create a strong electric field
in the axial direction of the tube in order to produce
appreciable liquid flow rates. The pressure generation of the
conduction pump was measured by using a manometer.
Results were presented for two different working fluids: nHexane and low-pressure refrigerant HCFC-123.
For HCFC-123 with the single-tube electrode configuration,
the pressure head reached approximately 500 Pa at an applied
voltage of 20 kV and a power level of 1.4 W, as shown in
Figure 4. In the absence of viscous pressure losses, this
pumping head was shown to correspond to a flow velocity of
approximately 1 m/s, demonstrating the potential of EHD
conduction pumping for practical use. The needle electrode
configuration was not as effective, producing only 100 Pa of
pressure head at an applied voltage of 15 kV. Further increases
in the applied voltage caused the pressure generation to cease
(at 17.5 kV) and then reverse (at 20 kV). This flow reversal was
attributed to the effect of direct ion-injection by the sharp
needle electrodes overcoming the effect of the heterocharge
layers. For both electrode configurations, the pressure
generation was a quadratic function of the applied voltage;
however, departure from this quadratic behavior began around
10 kV for the needle electrode configuration, due to the onset of
direct ion injection. Pressure head generation using n-Hexane
as the working fluid was shown to be approximately one-sixth
of the pressure head generation using HCFC-123.

suggesting that efficiency improvements could be gained by


using a more intense electric field, by reducing the distance
between the electrodes, or by choosing working fluids with a
longer charge relaxation time (/e). However, it must be noted
that an upper limit to efficiency exists because the electric field
intensity (V/d) is limited by the onset of ion injection.

Figure 3. Schematic of permeable parallel plane electrode configuration


studied in a theoretical context by Feng and Seyed-Yagoobi [14].

Figure 4. Pressure generation and current consumption for the hollow-tube


electrode configuration considered by Atten and Seyed-Yagoobi [3]

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 16, No. 2; April 2009

429

Table 1. A selection of single-phase electrode designs that have been


considered.
Electrode Geometry

Name and Refs.


Hollow tube
Refs. [3,4]
Four needle
Refs. [3]
Three needle
Refs. [4]
Perforated/Porous
Refs. [13,17]
Multi tube
Refs. [4,23,24]

Jeong and Seyed-Yagoobi [4] performed a similar


experimental study, but using three different electrode
designsthree-needle, hollow-tube, and multi-tube. The axial
spacing between the high voltage and ground electrode was
5 mm for the three-needle design and 2.64 mm for the other
two designs. Additionally, two electrode pairs were arranged
in series to increase the pumping capacity. As shown in
Figure 5, the multi-tube electrode configuration was found to
produce the highest pressure head due to the higher available
projected area and because of a shorter absolute distance
between the high voltage and ground electrodes. This
electrode design was also found to operate with higher
efficiency. At the maximum applied voltage of 20 kV, the
generated pressure head reached almost 600 Pa compared to
250 Pa with the hollow-tube design and 75 Pa with the threeneedle design. The power consumption for the multi-tube
electrode at 20 kV was 0.27 W, decaying to a final steadystate level of 0.06 W. The ion-drag phenomenon was not
observed even at the maximum applied voltage. The pressure
head generation was found to have a quadratic behavior with
respect to applied voltage, agreeing with the theoretical study
by Atten and Seyed-Yagoobi [3].
For the hollow tube and three-needle designs, the pressure
generation resulting from multiple pairs of electrodes acting
in series was also studied. One-, three-, and five-pair
configurations were considered, with results showing an
approximately linear increase in pumping performance with
respect to the number of electrode pairs. However, the head
generation of some m electrode pairs was found to be less
than m-times the head generation of one electrode pair. This
was attributed to negative interactions between the electric
fields of each electrode pair; just as a strong electric field
directed in one direction existed between a high-voltage
electrode and its associated ground electrode, a small electric
field directed in the opposite direction would exist between a
high-voltage electrode and the ground electrode of the
adjacent pair, even though the spacing distance between pairs
was significantly larger than the spacing distance between the
two electrodes of a given pair.

Figure 5. Pressure generation and current consumption for three-needle,


hollow-tube, and multi-tube electrodes as presented by Jeong and SeyedYagoobi [4].

Figure 6. Pressure generation and current consumption for a number of


perforated/porous disk electrodes as presented by Jeong and Seyed-Yagoobi
[17]. Design1: 1.14 mm perforated, Design2: 1.59 mm perforated, Design3:
0.2 m porous, Design4: 40.0 m porous.

Noting that electrodes with higher projected area performed


better in their previous study [4], Jeong and Seyed-Yagoobi
[17] further improved the pressure generation performance by
considering a variety of perforated and porous disk electrodes.
The perforated disks had 1.14 and 1.59 mm diameter holes
punched through the disk, while the porous disks had 0.2 and
40 m pores. The projected areas were 64%, 59%, 78%, and
50%, respectively. As shown in Figure 6, the porous
electrodes generated the highest pressures, although the
corresponding electrical currents were moderately higher too.
However, it was noted that while pressure generation is large,
the pressure loss due to flow through the electrodes would
also be large (net flow was absent in these static pressure
generation studies). Porous electrodes could also present
operational issues due to the blockage of pores by impurities.
Therefore, porous electrodes are appropriate only for very
clean working fluids with very low flow rates and high
pressure generation requirements.

M. R. Pearson and J. Seyed-Yagoobi: Advances in Electrohydrodynamic Conduction Pumping

430

Endurance tests were conducted by Jeong and SeyedYagoobi [4, 17] to investigate the long-term performance of
EHD conduction pumps, with multiple electrode pairs
operating continuously in series for 8 hours. The pressure
generation and current levels were found to slowly decrease
over time to some asymptotic value. These decays were
attributed to a purification process caused by gradual
precipitation of fine particles, which is a phenomenon that is
generally seen in electrical conduction measurements of a
fluid.
Feng and Seyed-Yagoobi [13] used the 1.59 mm perforated
high voltage electrode for the experimental portion of their
study. Experiments were conducted using a single-phase loop,
fitted with a flow-meter, valve, and a pressure-transducer to
measure the pumping head generated by the EHD pump.
HCFC-123 was chosen as the working fluid. Due to friction
losses within the pump itself and the external load placed on
the pump due to the setup of the single-phase loop, the
maximum flow velocity was 8.9 cm/s at an applied voltage of
15 kV and a generated pressure head of 425 Pa. The
corresponding mass flux was found to be 130 kg/m2s. The
pressure head generation and current level were shown to
agree with the theoretically predicted values in the absence of
net flow. To explain the discrepancies caused by the presence
of net flow, a magnitude analysis of the current level was
completed, which showed that charge mobility and charge
convection could significantly affect the current level, while
charge diffusion had a negligible effect at high voltages [13].

TWO-PHASE STUDIES

4.1 NUMERICAL
The application of the model described in Section 2 to
examine the pumping of a liquid-film (stratified two-phase
flow) was first conducted by Yazdani and Seyed-Yagoobi
[12] by considering electrodes on the base of a channel that
were designed to remain submerged in the liquid film
(without penetrating into the vapor phase). The design of the
electrodes was motivated by two simple and counter-acting
concepts: 1) that greater penetration of the electrodes into the
liquid film, away from the surface, would generate a greater
electric body force in the fluid and provide greater pumping

Figure 7. Comparison between numerical results [12] and experimental


results [18] for the pumping of a 6 mm liquid film using flushed, asymetrical
electrodes.

potential, and 2) that greater penetration of the electrodes


would also inhibit the flow by acting as an obstruction to
liquid motion. From these considerations, four designs were
explored, all having a flushed ground electrode with a width
equal to the film thickness. The liquid film thickness was
taken as the normalizing length scale, d, for the purposes of
the model. By matching the flow parameters and electrode
dimensions to an experimental study by Siddiqui and SeyedYagoobi [18], and by imposing the experimentally-measured
pressure load on the EHD conduction pump, the numerical
simulation was able to predict the experimental results for 4,
6, and 8 mm films with good accuracy, as shown in Figure 7
for the case of a 6 mm film.
For the simple case of a flushed, high voltage electrode
having the same width as the ground electrode, no net flow
resulted and two symmetrical, counter-rotating vortices
formed above each electrode. However, by halving the width
of the flushed, high-voltage electrode, a resultant flow was
generated, directed from the narrower electrode to the wider
one, caused by a net body force in the region between the
electrodes of a given pair that acts towards the wider
electrode. By keeping the width of the high-voltage electrode
as half that of the ground electrode, but by increasing the
height from zero so that the electrode would partially
penetrate into the liquid film, then the flow eventually
reversed and became directed from the wide, flushed, ground
electrode to the narrow, raised, high-voltage electrode. The
flow reversal was shown to occur at approximately 5%
penetration into the channel. When the electrode penetrated
through 50% of the liquid film thickness, the obstruction
caused by the raised electrode caused flow separation and a
large wake, but the resultant net flow was still greater than for
the flushed, asymmetrical electrodes for the same operating
parameters.
The numerical study of the conduction pumping of liquid
films was further extended by Yazdani and Seyed-Yagoobi
[19] to account for a non-rigid liquid-vapor interface and the
presence of heat transfer and phase-change. These
developments required the incorporation of the energy
equation for the control volume and the Young-Laplace
equation as a boundary condition at the liquid-vapor interface,
along with an interface tracking method for the numerical
scheme. For the numerical domain, the study considered an
entrance region, a single pair of asymmetrical, flushed
electrodes, an evaporator region where heat flux was applied
to the liquid-solid interface, and finally an adiabatic exit
region. The results showed bumps in the liquid-vapor
interface due to strong, local circulations in the vicinity of and
downstream from the electrode pair. These circulations also
caused an increase in the overall heat transfer coefficient for
the evaporator region, when compared to a conceptual
pressure-driven flow with the same volumetric flow rate at the
inlet (such a flow being theoretical in nature since it is hard to
mechanically introduce a pressure gradient solely into a
horizontal liquid film).
4.2 EXPERIMENTAL
Siddiqui and Seyed-Yagoobi [18] performed an
experimental study of the flushed electrodes geometry that
was studied by Yazdani and Seyed-Yagoobi [12]. Six

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 16, No. 2; April 2009

electrode pairs were used, having dimensions shown in Figure


8. A perforated electrode design was also considered where
the wide electrodes in Figure 8 were replaced with a thin,
perforated electrode penetrating 2.5 mm vertically into the
liquid film. The liquid film was pumped around a racetrackstyle channel with a rectangular cross section and the flowrate was estimated based on the motion of visible particles
within the fluid and a detailed video recording of the flow
channel. The pressure drop along the length of the racetrack
was also measured. Liquid film thicknesses ranging from 2 to
8 mm were considered. The perforated electrode design
performed better in thicker films, approaching an average
velocity of 100 mm/s for the 8 mm film. The flushed
electrode performed better with thinner films by reducing the
drag on the flow, achieving approximately 25 mm/s with a
2 mm film and 70 mm/s with a 5 mm film. Voltages of up to
15 kV were applied. The reduction in flow velocity at thinner
film thicknesses was attributed to the high shear stresses
along the floor of the flow channel.

DEMONSTRATED APPLICATIONS
5.1

LIQUID AND TWO-PHASE FLOW


DISTRIBUTION
Active control of liquid or two-phase flow distribution has
been one very successful demonstration of the simplicity and
practical applicability of EHD conduction pumping
technology. Considering a heat exchanger with multiple,
parallel evaporators or branch lines, optimal operation
requires a uniform flow distribution in all branches.
Maldistribution can cause local hot areas and regions of dryout, but this maldistribution can be easily caused by uneven
heat loads, non-uniform pressure drops, or gravity [20]. To
control the flow distribution, an EHD conduction pumping
mechanism was proposed by Feng and Seyed-Yagoobi [20].
The EHD pump was installed on one branch of a two-branch
single-phase (liquid) system. A mechanical pump was used as
the principal method of flow generation and the EHD pump
acted as an auxiliary pump on one of the two branches. The
experimental setup ensured that, without EHD pumping
assistance, the flow through each branch tube was
approximately equal. The EHD pump was then powered with
various applied voltages and the resulting flow distribution
was monitored.
For an HVAC&R relevant total mass flux of 100 kg/m2s
and an applied voltage of 10 kV, the mass flux difference
between the two branch lines reached 30 kg/m2s (see Figure
9). Furthermore, an applied voltage of 15 kV produced such a
high pressure head generation that essentially the entire flow
was completely directed through branch tube 1. Increasing the
total mass flux to 200 kg/m2s, the difference in mass fluxes
bus line for ground
electrodes

431
2

between the branch tubes was approximately 20 kg/m s and


30 kg/m2s for applied voltages of 10 kV and 15 kV,
respectively. Under all test conditions, the power
consumption ranged from 0.56 W to 1.17 W, which is
negligible compared to typical heat transfer rates found in
HVAC&R equipment (on the order of kilowatts).
The use of an EHD conduction pump as an active flow
control device was extended to two-phase flow scenarios by
Feng and Seyed-Yagoobi [21]. Experiments were conducted
at mass fluxes of 50 and 100 kg/m2s with vapor quality levels
from 4% to 26%. The distribution control was less effective
because the EHD conduction pump can only act in the liquid
phase of the two-phase flow, but nonetheless under certain
conditions of mass flux, applied voltage, and vapor quality the
flow distribution remained significant. Furthermore, the
authors proposed that improved electrode designs, optimized
for the two-phase nature of the flow, together with the
possibility of adding additional electrode pairs, might serve to
better alter the flow distribution under higher mass flux
conditions.
5.2 HEAT PIPES OR TWO-PHASE LOOPS
EHD conduction pumps have been demonstrated as the
pumping mechanism for two-phase loops and heat pipes. Heat
pipes, capillary pumped loops, and loop heat pipes
traditionally have used capillary force generation provided by
a wicking structure to generate flow. However, a number of
limitations exist including limitations to pumping capacity,
start-up difficulties, and poor recovery from evaporator dryout. For this reason, it has been repeatedly proposed that an
EHD conduction pump could be applied to resolve many of
these issues and act as the pumping mechanism in lieu of or in
addition to a wick.

The first such demonstration came from Bryan and SeyedYagoobi [8], but at that time the conduction pumping
phenomenon was not understood and the flow generation was
erroneously credited to polarization forces. Once the
conduction pumping phenomenon had been clarified [3, 9],
the EHD conduction driven heat pipe was revisited by Jeong
and Seyed-Yagoobi [22] using electrodes better optimized for
conduction pumping (the same porous electrodes as studied

bus line for high voltage


electrodes

50.80

pressure port
3.17

17.46

1.59
9.53

ground
electrode

high voltage
electrode

Figure 8. Flushed, asymmetrical electrodes studied experimentally by


Siddiqui and Seyed-Yagoobi [18].

Figure 9. Experimental results by Feng and Seyed-Yagoobi [20] for flow


distribution control using an EHD conduction pump with an applied voltage
of 10 kV. Flow is initially distributed evenly; activation of the EHD pump in
one branch causes a large change in flow distribution.

M. R. Pearson and J. Seyed-Yagoobi: Advances in Electrohydrodynamic Conduction Pumping

432

Table 2. Heat transport capacity of an EHD conduction-driven heat pipe [22].


Tilt
No EHD
10 kV

3 mm
(Favorable)
520 W
920 W

0 mm
(Horizontal)
400 W
840 W

3 mm
(Adverse)
200 W
800 W

by Jeong and Seyed-Yagoobi [4]). The conduction pump was


installed in the liquid line in the adiabatic section of the heat
pipe (between the condenser and the evaporator). Experiments
were conducted in the horizontal position and also with 3 mm
of adverse and favorable tilt (the adverse case meaning that
the evaporator is raised 3 mm above the condenser, vice versa
for the favorable case). As shown in Table 2, substantial
increases in heat transport capacity of the heat pipe were
demonstrated, particularly when an adverse tilt was imposed.
Fast recovery from dryout was also demonstrated.
In a recent study, Jeong and Didion [23] experimentally
investigated the feasibility of using an EHD conduction pump
for thermal control of a two-phase loop with a high heat flux
source, demonstrating that the pump could be used as a standalone system for high heat flux thermal control. The EHD
pump consisted of six electrode pairs, with a ring ground
electrode flushed to the tube wall and a high-voltage six-tube
electrode with 58.3% open area. The two-phase loop
contained a pre-heater, evaporator, condenser, and the EHD
pump installed on the liquid line from the condenser to the
evaporator. The total length of the loop was approximately
10 m and the working fluid was the refrigerant HFC-134a.
The results showed that by varying the temperature of the
heat sink (removing heat from the condenser) from 20 C to
20 C, pumping performance was shown to have a fairly
strong dependence on operating temperature. For example, at
an applied voltage of 15 kV and at sink temperatures of 20
C, 0 C, and 20 C, the generated pressure heads were
approximately 9.6 kPa, 11.2 kPa, and 13.2 kPa, respectively.
This phenomenon was explained by the dependence of
relaxation time (/e) on temperature, and the corresponding
dependence that heterocharge layer thickness has on
relaxation time (cf. equation (27)). For the experimental
conditions that gave the highest pressure head generation
(13.2 kPa), the heat pipe was shown to provide thermal
control capacity for 35.8 W/cm2 of heat flux and 1500 Wm of
heat transport. Furthermore, it was noted that this capacity
could be enhanced simply by increasing the number of
electrode pairs or enhancing the surface of the evaporator.
A second study by Jeong and Didion [24] sought to
characterize and quantify the performance of the conduction
pump in two different two-phase loop setups. The geometric
dimensions of the conduction pumping electrodes were
identical to those of the earlier study [23] but fabrication
improvements yielded a more consistent separation between
electrodes and a smoother electrode surface, resulting in
reduced ion-injection. As a result, the voltage at which
electrical breakdown began increased from 16 to 22 kV, and
significantly higher pressures were generated with the same
power consumption. The effect of sink temperature was
reconfirmed and a number of other effects were studied. In
particular, deaeration of the working fluid HFC-134a

(removal of noncondensables from the liquid phase) was


shown to yield up to a 23.3% increase in pressure head
generation at 20 kV of applied voltage and the coldest sink
temperature. The following reasons were given to explain the
performance reduction caused by air: 1) air, having an
electrical permittivity less than liquid HFC-134a, is drawn by
dielectrophoretic force to the region of weakest electric field,
which opposes the direction of liquid flow towards the region
of stronger field; 2) the effect is greater at colder temperatures
because the permittivity of air is approximately constant while
the permittivity of HFC-134a liquid increases with decreasing
temperaturetherefore, the dielectrophoretic force, which is a
function of permittivity difference, also increases with
decreasing temperature; 3) gas bubbles can collect due to
surface tension and dielectrophoretic force, causing a
significant flow obstruction in the EHD pump.
In a very recent study, Pearson and Seyed-Yagoobi [25]
investigated the prospects of a heat pipe that is driven by
EHD conduction pumping of liquid films. Flushed electrodes,
very similar to those studied by Yazdani and Seyed-Yagoobi
[12] and Siddiqui and Seyed-Yagoobi [18], were placed along
the floor of the channel in the adiabatic section. These
electrodes were then used to pump the liquid film from the
condenser section to the evaporator section. This study has
demonstrated, as a proof-of-concept, that a heat pipe can be
operated by replacing the traditional wicking structure with an
EHD conduction pump designed for the pumping of liquid
films.
5.3 JET FORMATION
Yabe and Maki [6] and Sato et al [7] developed an EHDdriven nozzle using a ring electrode and a plate electrode.
When a dc voltage was applied to the ring electrode, there
was flow generation through the ring electrode, directed away
from the grounded plate electrode, and flow velocities of
1 m/s were achieved. The conduction pumping mechanism
was not well understood at the time and the flow generation
was erroneously attributed to the electrostriction term of
equation (1). The electrode configuration has since been
revisited by Hanaoka et al [26, 27], who demonstrated that the
actual cause of flow generation was conduction pumping.
Both studies [26, 27] considered the same high voltage ring
electrode, while the later study [27] varied the geometry of the
ground electrode.
Hanaoka et al [26] considered four different working fluids:
di-n-butyl dodecanedioate, 2-(2-butoxyethoxy)ethyl acetate,
JIS-C2320 transformer oil, and fluorinated silicon oil. These
fluids are referred to as DBDN, BCRA, Tr-Oil and FS-Oil,
respectively. In the later study, Hanaoka et al [27] considered
two additional refrigerant fluids: HCFC-123 and HFC-43-10.
Using BCRA as the working fluid, a jet velocity of just over
1.5 m/s was achieved at an applied voltage of 23 kV [26].
Similar velocities were achieved at slightly lower voltages for
HFC-43-10. For HCFC-123, a peak velocity of approximately
0.8 m/s was obtained with an applied voltage of 16 kV [27].
Hanaoka et al [27] also demonstrated that the use of a large
plate electrode as the ground electrode produced higher
velocities than a smaller plate electrode or a sphere electrode,
due to greater electric field asymmetry.

IEEE Transactions on Dielectrics and Electrical Insulation

Vol. 16, No. 2; April 2009

For BCRA and Tr-Oil, the flow was always directed


through the ring electrode and away from the plate electrode,
regardless of whether a positive or negative voltage was
applied. For DBDN, FS-Oil, HCFC-123, and HFC-43-10, the
flow direction was dependent upon the sign of the applied
voltageDBDN flowed away from the plate electrode when
a negative voltage was applied to the ring electrode and
flowed in the opposite direction when a positive voltage was
applied. The opposite dependence was exhibited by the other
three fluids. To explain this dependence, the authors
postulated that in addition to the dissociation/recombination
of impurities, which should be independent of applied
polarity, the fluid molecules themselves were also undergoing
dissociation and recombination, with the potential for positive
and negative ions to have different mobility coefficients. In
particular, it was noted that liquids containing the strongly
electronegative fluorine atom all exhibited flow towards the
positively charged ring electrode. This suggests that the
electro-chemical behavior of the fluid can greatly influence
the heterocharge layers and cause a directionality of the flow
[26, 27]. Additional understanding of the electrical behavior
of the working fluids is necessary because a differing mobility
of positive and negative ions could greatly affect the design of
electrodes. For example, if the positive and negative ions have
significantly different mobility coefficients then pumping may
be generated even with symmetrical electrodes.
A different jet design that also harnesses EHD conduction
pumping has been demonstrated numerically by Yazdani and
Seyed-Yagoobi [28]. This design uses pairs of asymmetrical
ring electrodes that are flushed within the inner wall of a
cylindrical or conical jet nozzle. Local heat transfer
coefficients of up to approximately 2500 W/m2K at the
impingement point have been predicted for the working fluid
HCFC-123.
5.4 STORAGE TANK AND FLOW MIXING
Yazdani and Seyed-Yagoobi [29] have demonstrated,
through numerical simulation that conduction pumping can
effectively produce circulations that can non-mechanically
mix a single-phase dielectric fluid stored within a spherical
storage tank. Two electrode designs were considered: 1)
flushed, asymmetrical electrodes (like those studied by
Yazdani and Seyed-Yagoobi [12] and Siddiqui and SeyedYagoobi [18] for the pumping of liquid films) arranged along
latitudinal lines on the inner surface of the spherical tank, and
2) concentric ring electrodes arranged in pairs, just above and
just below a plane passing through the center of the tank. The
study investigated the resulting steady-state flow field for
each electrode design. The first electrode design produced a
net down-flow near the wall, directed from the narrow
electrode to the wide electrode of each pair, and there was
resultant up-flow in the center of the tank, thus creating a
donut-shaped circulation pattern that effectively mixed the
fluid in the tank. As with previous studies of the
asymmetrical, flushed electrode design [12, 18, 19],
additional local circulations also existed near the electrodes.
The second electrode design was designed to produce only
local circulations without a net flow generation and this
behavior was verified numerically. The motivation behind the
second electrode design was to cause EHD-induced

433

circulation within the sphere, not just at the surface. However,


because of the absence of net flow generation, the resulting
mixing was inferior to that of the first electrode design, even
though the first design could produce an electric body force
only near the tank surface.
The mixing concept was further demonstrated by Yazdani
and Seyed-Yagoobi [30] by considering a single-phase fluid
in a spherical tank that initially has a jump in temperature
along a horizontal plane through the center of the sphere.
Only the flushed, asymmetrical electrode configuration was
considered. The unsteady numerical simulation then
demonstrated the evolution of temperature distribution with
time, requiring incorporation of the energy equation and the
Boussinesq approximation of buoyancy-driven natural
convection effects. Joule heating due to the applied electric
field was also included. Increasing the size of the tank, which
corresponds to increases in C0 and Gr = |g|d/cp, caused the
effect of the conduction pumping to diminish in comparison
to natural convection. When the initial temperature
distribution favored natural convection (i.e. the hotter fluid
below the colder fluid), the mixing remained efficient, but
when the initial distribution opposed natural convection the
time to reach homogeneity increased rapidly with the size of
the tank.
As a generalization of storage tank mixing, it has been
recognized that the high amount of circulation produced by
flushed electrodes can be used to produce secondary
velocities within a laminar pipe or channel flow, with an
associated increase in convective heat transfer. This concept
was demonstrated by Yazdani and Seyed-Yagoobi [31] with
asymmetric, flushed electrodes embedded on the inside wall
of a circular channel. The EHD pumping mechanism was
secondarythe primary flow production mechanism was an
externally-applied pressure head that created a Pouseille flow
in the channel. When the Reynolds number of the pressure
driven flow was low, significant enhancements in Nusselt
number caused by large secondary velocities were achieved
by the application of an electric field. These effects
diminished quickly, and at Re = 1500 the enhancements were
almost zero due to the suppression of EHD-induced secondary
flow by the larger pressure-driven flow.

CONCLUSIONS

When the conduction pumping phenomenon was originally


observed, it was incorrectly attributed to electrostriction
effects. Several years on, conduction pumping is now a well
understood part of EHD that has both theoretical advantages
(such as no need for temperature gradients, no degradation of
the working fluid) and proven applications. The theoretical
model behind conduction pumping has evolved such that
numerical results can predict actual, experimental data with
good accuracy. Conduction pumping has been successfully
applied to resolve common problems such as liquid-film
pumping, nonmechanical mixing, and active flow control,
while retaining its simplicity.
Many researchers have proposed that EHD conduction
pumping might be well-suited to electronic-cooling
applications, but this particular application remains largely
unexplored. The dielectric fluids required for conduction

M. R. Pearson and J. Seyed-Yagoobi: Advances in Electrohydrodynamic Conduction Pumping

434

pumping are generally very compatible with electronic


components, and the strong electric field in conduction
pumping is very localized and unlikely to cause interference,
although this has not yet been confirmed experimentally.
Furthermore, as the scale of the device is reduced, the voltage
required to maintain the necessary electric field intensity is
also reduced proportionally. Future studies at the micro-scale
must explore the combined effect of the heterocharge layer
and the electric double layer on EHD-driven flow, as the latter
layer becomes increasingly important with decreasing size.

REFERENCES
[1]
[2]
[3]
[4]
[5]
[6]
[7]

[8]
[9]

[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]

J. R. Melcher, Continuum Electromechanics. Cambridge, MA: MIT


Press, 1981.
J. Seyed-Yagoobi, "Electrohydrodynamic pumping of dielectric
liquids", J. Electrostatics, Vol. 63, pp. 861869, 2005.
P. Atten and J. Seyed-Yagoobi, "Electrohydrodynamically induced
dielectric liquid flow through pure conduction in point/plane geometry",
IEEE Trans. Dielectr. Electr. Insul., Vol. 10, pp. 2736, 2003.
S.-I. Jeong and J. Seyed-Yagoobi, "Experimental study of
electrohydrodynamic pumping through conduction phenomenon," J.
Electrostatics, Vol. 56, pp. 123133, 2002.
S.-I. Jeong, J. Seyed-Yagoobi, and P. Atten, "Theoretical/numerical
study of electrohydrodynamic pumping through conduction
phenomenon", IEEE Trans. Ind. Appl., Vol. 39, pp. 355361, 2003.
A. Yabe and H. Maki, "Augmentation of convective and boiling heat
transfer by applying an electro-hydrodynamical liquid jet", Int. J. Heat
Mass Transfer, Vol. 31, pp. 407417, 1988.
M. Sato, A. Yabe, and T. Taketani, "Heat transfer enhancement by
applying an electro-hydrodynamical pump utilizing dielectrophoretic
force", ASME-JSME Thermal Eng. Joint Conf., Reno, Nevada,
USA,Vol. 3, pp. 38, 1991.
J. E. Bryan and J. Seyed-Yagoobi, "Heat transport enhancement of
monogroove heat pipe with electrohydrodynamic pumping", J.
Thermophysics Heat Transfer, Vol. 11, pp. 454460, 1997.
P. Atten and J. Seyed-Yagoobi, "Electrohydrodynamically induced
dielectric liquid flow through pure conduction in point/plane
geometryTheory", IEEE 13th Int. Conf. on Dielectric Liquids, Nara,
Japan, pp. 231234, 1999.
P. Langevin, "Recombinaison et mobilites des ions dans les gaz," Ann.
Chim. Phys., Vol. 28, p. 433, 1903.
S.-I. Jeong and J. Seyed-Yagoobi, "Fluid circulation in an enclosure
generated by electrohydrodynamic conduction phenomenon", IEEE
Trans. Dielectr. Electr. Insul., Vol. 11, pp. 899910, 2004.
M. Yazdani and J. Seyed-Yagoobi, "Electrically induced dielectric
liquid film flow based on conduction phenomenon", Phys. Fluids,
submitted for publication.
Y. Feng and J. Seyed-Yagoobi, "Understanding of electrohydrodynamic
conduction pumping phenomenon", Phys. Fluids, Vol. 16, pp. 2432
2441, 2004.
Y. Feng and J. Seyed-Yagoobi, "Electrical charge transport and energy
conversion with fluid flow during electrohydrodynamic conduction
pumping", Phys. Fluids, Vol. 19, Paper No. 057102, 2007.
J. Seyed-Yagoobi and J. E. Bryan, "Electrohydrodynamic conduction
pump", U.S. Patent 6 932 580, 2005.
J. Seyed-Yagoobi, "Electrode design for electrohydrodynamic
conduction pumping", U.S. Patent 7 261 521, 2007.
S.-I. Jeong and J. Seyed-Yagoobi, "Innovative electrode designs for
electrohydrodynamic conduction pumping", IEEE Trans. Ind. Appl.,
Vol. 40, pp. 900904, 2004.
M. A. W. Siddiqui and J. Seyed-Yagoobi, "Experimental study of
pumping of liquid film with electric conduction phenomenon", IEEE
Trans. Ind. Appl., submitted for publication.
M. Yazdani and J. Seyed-Yagoobi, "Numerical investigation of
electrohydrodynamic-conduction pumping of liquid film in the presence
of evaporation", J. Heat Transfer, Vol. 131, Paper No. 011602, 2009.
Y. Feng and J. Seyed-Yagoobi, "Control of liquid flow distribution
utilizing EHD conduction pumping mechanism", IEEE Trans. Ind.
Appl., Vol. 42, pp. 369377, 2006.

[21] Y. Feng and J. Seyed-Yagoobi, "Control of adiabatic two-phase


dielectric fluid flow distribution with EHD conduction pumping", J.
Electrostatics, Vol. 64, pp. 621627, 2006.
[22] S.-I. Jeong and J. Seyed-Yagoobi, "Performance enhancement of a
monogroove heat pipe with electrohydrodynamic conduction pumping",
ASME Int. Mech. Eng. Congr. Expo., New Orleans, LA, USA, Paper
No. 33531, 2002.
[23] S.-I. Jeong and J. Didion, "Thermal control utilizing an
electrohydrodynamic conduction pump in a two-phase loop with high
heat flux source", J. Heat Transfer, Vol. 129, pp. 15761583, 2007.
[24] S.-I. Jeong and J. Didion, "Performance characteristics of
electrohydrodynamic conduction pump in two-phase loops", J.
Thermophysics Heat Transfer, Vol. 22, pp. 9097, 2008.
[25] M. R. Pearson and J. Seyed-Yagoobi, "Experimental study of a twophase heat transport device driven by electrohydrodynamic conduction
pumping", ASME Int. Mech. Eng. Congr. Expo., Boston, MA, Paper
No. 68478, 2008.
[26] R. Hanaoka, S. Takata, M. Murakumo, and H. Anzai, "Properties of
liquid jet induced by electrohydrodynamic pumping in dielectric
liquids", Elect. Eng. Japan, Vol. 138, pp. 19, 2002, [Denki Gakkai
Ronbunshi, Vol. 121-A, pp. 224230, 2001].
[27] R. Hanaoka, H. Nakamichi, S. Takata, and T. Fukami, "Distinctive flow
properties of liquid jet generated by EHD pump and conical nozzle",
Elect. Eng. Japan, Vol. 154, pp. 919, 2006, [Denki Gakkai Ronbunshi,
Vol. 124-A, pp. 399406, 2004].
[28] M. Yazdani and J. Seyed-Yagoobi, "An electrically driven impinging
liquid jet for direct cooling of heated surfaces", IEEE Trans. Ind. Appl.,
submitted for publication.
[29] M. Yazdani and J. Seyed-Yagoobi, "Fluid circulation within a spherical
reservoir with EHD conduction pumping", IEEE Trans. Ind. Appl.,
submitted for publication.
[30] M. Yazdani and J. Seyed-Yagoobi, "Thermal homogenization in
spherical reservoir by EHD conduction phenomenon", J. Heat Transfer,
submitted for publication.
[31] M. Yazdani and J. Seyed-Yagoobi, "Heat transfer enhancement of a
Poiseuille Flow by means of electric conduction phenomenon", IEEE
16th Int. Conf. Dielectric Liquids, Poitiers, France, pp. 6265, 2008.

Matthew R. Pearson received the B.S. and M.S.


degrees in mechanical engineering from the Illinois
Institute of Technology, Chicago, IL, in 2004 and
2006, respectively. He remains at the Illinois Institute
of Technology as a student pursuing a Ph.D. degree
in mechanical engineering. He is a recipient of a
three-year Graduate Research Fellowship from the
National Science Foundation that began in September
2006. His research is focused on the enhancement of
heat transfer utilizing electrohydrodynamics in twophase flows. Mr. Pearson is a Student Member of the American Society of
Mechanical Engineers, the Sigma Xi Scientific Research Society, and the Tau
Beta Pi Engineering Honor Society.
Jamal Seyed-Yagoobi (M89-SM02) received the
Ph.D. degree from the University of Illinois, UrbanaChampaign, in 1984. He was a Research Engineer for
three years with Westvaco Corporation. Following
his industrial career, he was with the Department of
Mechanical Engineering, Texas A&M University,
College Station, from 1987 to 2001. He is currently a
Professor and Chair of the department of Mechanical,
Materials, and Aerospace Engineering, Illinois
Institute of Technology, Chicago. His research
specialties are enhancement of heat transfer and mass transport with
electrohydrodynamics, heat and mass transfer in porous media (drying), and
innovative impinging jets. Prof. Seyed-Yagoobi is a Fellow of the American
Society of Mechanical Engineers and a Member of the American Society of
Engineering Education, American Society of Heating, Refrigeration and Air
Conditioning Engineers, and Technical Association of the Pulp and Paper
Industry.

Potrebbero piacerti anche