You are on page 1of 14

ARTICLE IN PRESS

Journal of Electrostatics 63 (2005) 337–350

EHD flow in air produced by electric corona


discharge in pin–plate configuration
L. Zhao, K. Adamiak*
Department of Electrical and Computer Engineering, University of Western Ontario,
London, Ontario, Canada N6A 5B9
Received 16 January 2004; received in revised form 22 April 2004; accepted 10 June 2004
Available online 8 July 2004

Abstract

The electrohydrodynamic flow in air produced by the electric corona discharge in the pin–plate
and pin–grid configurations has been investigated numerically and experimentally in this paper.
The numerical algorithm based on the boundary and finite element methods and the method of
characteristics is employed to obtain the distributions of the electric field, the electric potential and
the space charge density. After the volume force has been calculated, the commercial FLUENT
software is used to solve the airflow. Experimental axial velocity profiles outside the ground grid
obtained with a hot-wire anemometer reasonably validate the simulation results.
r 2004 Elsevier B.V. All rights reserved.

Keywords: EHD flow; Corona discharge; Electric field; Numerical simulation

1. Introduction

The electric corona discharge usually occurs when a high voltage is applied between
two electrodes with substantially different radii of curvature [1]. The high electric field
in the vicinity of the corona electrode causes gas ionization and its partial breakdown.
While the whole process is rather complicated, the net effect is that ions, of the same
polarity as that of the corona electrode, are drifting to the other electrode. A space
charge is formed and an electric current flows between both electrodes [2].
The movement of ions is obstructed by very high-frequency collisions with
electrically neutral air molecules. The complete momentum transfer from the ionic

*Corresponding author. Tel.: +1-519-661-2111X88358; fax: +1-519-850-2436.


E-mail address: kadamiak@eng.uwo.ca (K. Adamiak).

0304-3886/$ - see front matter r 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.elstat.2004.06.003
ARTICLE IN PRESS
338 L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350

space charge to the air bulk can be assumed to take place. Therefore, the Coulomb
force acting on the ions becomes an electric body force on the air molecules. This
force gives rise to the electrohydrodynamic (EHD) flow known as the ‘‘electric
wind’’, or ‘‘ionic wind’’, and was firstly reported by Hauksbee in 1709 [3].
In many industry and research applications, such as electrostatic precipitation [4]
and powder coating [5], corona discharge is often used to charge small particles and
droplets in order to electrically control their motion toward a specified target.
However, the air drag due to the EHD flow modifies the particles’ trajectories
making them more complicated. Therefore, a full analysis of such devices would be
incomplete without considering this effect. The EHD flow is also expected to have
applications in industries including ionic loudspeakers [6], heat transfer enhancement
[7] and household ionic breeze air cleaners [8].
During the last several decades, many efforts have been made to have a better
understanding of the electric corona discharge and many theoretical, experimental
and numerical results have been published [1,2,9–13]. In most cases, the interest in
the EHD flows results from its effect on the fine particle collecting efficiency of the
electrostatic precipitators [14–22]. Some authors tried to experimentally visualize the
flow patterns, as well as to measure the velocities of the EHD flows, using laser-
Doppler anemometers (LDV), particle image velocimetry (PIV) systems and hot wire
anemometers [14–18]. The leading conclusion is that the recirculating EHD flows
with the velocity of a few meters per second are generated by the electric corona
discharge. Other authors focused their studies on the theoretical and numerical
investigation of the EHD flows [14,15,19–22]. Some simulation results in the wire-to-
plane configuration of the electrostatic precipitators and in the needle-to-plane
configuration inside a closed tube have been published, but they have not yet been
experimentally verified. It is clear that a reliable numerical understanding of the
electric wind has not yet been established.
This paper presents a numerical algorithm which can be used to simulate the EHD
flow produced by the electric corona discharge in the pin–plane and pin–mesh
configurations—often used in industrial applications. All essential parameters of the
process, including the electric field, the space charge density, the airflow velocity and
the relative pressure, are predicted. A numerical technique based on the boundary
and finite elements methods (BEM and FEM) and the method of characteristics
(MOC) has been employed to obtain the distributions of the electric field, the
potential and the space charge density [12]. The airflow distribution has been
simulated by a commercial software package FLUENT. Finally, a hot-wire
anemometer was utilized to measure the gas velocities for the simulation model
verification. A close examination of computed and measured velocity profiles of the
EHD flows validate the used numerical algorithm.

2. Mathematical model

The configuration under investigation in this paper consists of the ground


electrode, which is an infinitely large metal plate or a metal grid, and the corona
ARTICLE IN PRESS
L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350 339

Fig. 1. Pin–plane corona model and the boundary conditions for airflow.

electrode in the form a sharp pin of length L supplied with a high positive DC
voltage and perpendicular to the ground electrode at a distance D. Two different
kinds of pins have been studied: a hyperboloidal needle with the tip radius of
curvature R and a cylindrical needle with a conical ending and a hemispherical tip of
radius R. The ambient gas is air at room temperature and atmospheric pressure. By
taking into account the axial symmetry of the configuration, a 2D computational
model in the cylindrical coordinate can be assumed, as shown in Fig. 1(a).

2.1. Electrical model

The volume force that drives the airflow is the Coulomb force acting on the space
charge, therefore, calculation of the electric field and space charge density is a crucial
step in the EHD flow simulation. When an electric corona discharge takes place, the
electric charges are injected from the ionization layer and form a space charge in the
air gap between the two electrodes. The electric field is thus governed by Poisson’s
equation
q
r2 F ¼  ; ð1Þ
e
where F is the scalar electric potential, q the space charge density and e the
permittivity of the ambient gas.
The ionic charges are accelerated by the Coulomb force and move towards the
ground plate. The charge drifting creates an electric current with a density defined as
, , ,
j ¼ qðKE þ V Þ  Drq; ð2Þ
,
where j is the current density, K the mobility of ions, D the ions diffusion coefficient
,
and V the gas velocity. Under steady-state conditions, the current density must
satisfy the charge conservation equation
,
r  j ¼ 0: ð3Þ
Since the drift velocity of ions is usually about two orders of magnitude bigger than
the typical velocity of the gas flow [14], the convective component in the ionic current
, ,
density can be neglected, so j ¼ KqE  Drq: Under such an assumption, and after
ARTICLE IN PRESS
340 L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350

neglecting the ion diffusion, the charge conservation equation (3) leads to [12]

rq  rF ¼ q2 =e: ð4Þ

Therefore, the electric problem of the corona discharge is governed by a set of two
partial differential equations with two unknown distributions: the linear one (1) with
the unknown potential F and the nonlinear one (4) with the unknown space charge
density q. The boundary conditions for the potential are very straightforward: a
given DC potential F0 at the corona electrode and zero at the ground plane.
However, formulation of proper boundary conditions for the space charge density is
not so easy. The Katpzov hypothesis is adopted, which suggests that the electric field
increases proportionally to the voltage below the corona onset, but will preserve its
value after the corona is initiated. Peek’s formula is used to determine the threshold
strength of electric field for the corona onset at the hyperboloidal corona electrode
[13]. This approach provides an indirect boundary condition for space charge
density. Its distribution on the corona electrode surface is iterated until the corona
electrode electric field is sufficiently close to Peek’s value [23]. Therefore, the entire
algorithm consists of two iterative loops, the inner loop for the electric field and the
space charge density, and the outer loop for the charge density on the corona
electrode surface [12].

2.2. Airflow model

Under the assumption that the ambient air is incompressible, has constant density
and viscosity, and the flow is steady and laminar, the airflow has to satisfy the
continuity equation
,
r  V ¼ 0; ð5Þ

and the Navier–Stokes equation


,, ,
r  ðrV V Þ ¼ rP þ r  ðtÞ þ F ; ð6Þ

where r is the gas density, P the static pressure, t the stress tensor and F the external
body force, in this case equal to the Coulomb force qrF:
In 2D cylindrical geometry, the continuity equation becomes
qVx qVr Vr
þ þ ¼ 0: ð7Þ
qx qr r
The momentum equation in the axial direction x can be written as
1 q 1q
ðrrVx Vx Þ þ ðrrVr Vx Þ
r qx r qr    
qP 1 q qVx 2 , 1q qVx qVr qF
¼ þ rZ  3ðr  V Þ þ rZ þ þq : ð8Þ
qx r qx qx r qr qr qx qx
ARTICLE IN PRESS
L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350 341

Similarly, for the radial direction r


1 q 1q
ðrrVx Vr Þ þ ðrrVr Vr Þ
r qx r qr     
qP 1 q qVr 2 , 1q qVx qVr
¼ þ rZ 2  3ðr  V Þ þ rZ þ
qr r qr qr r qr qr qx
Vr 2 Z , qF
 2Z 2 þ ðr  V Þ þ q ; ð9Þ
r 3r qr
where Z is the air viscosity.
The boundary conditions for the airflow are straightforward, as shown in
Fig. 1(b). The two electrodes act as stationary walls where both axial and radial
components of the velocity vector vanish. The sidewall was defined as pressure-
outlet, wherein the relative pressure is zero and the direction of the flow is normal to
the boundary, since the computational domain is open in this area and air is free to
flow in both directions.

3. Numerical simulation

3.1. Electric field

The numerical algorithm for the corona discharge simulation is based on three
different numerical techniques: BEM, FEM and MOC. The BEM is used to obtain
the solution of the Laplacian electric field where space charge is assumed to be zero
and only the applied voltage between the two electrodes is considered. The electric
field gradient is very high in the vicinity of the corona electrode tip and the BEM can
provide very accurate and smooth solution in this area without creating too large an
algebraic system. The BEM solution is also used to generate the FEM mesh. In the
next step, the conventional FEM procedure is employed to obtain the Poissonian
component of the electric potential where the space charge is considered and both
electrodes are at ground potential. The final solution for F is a superposition of the
BEM and the FEM components. The electric field can be easily calculated by
differentiating the potential distribution. The last step is to solve the space charge
density by the MOC [12].
The algorithm is arranged in a double iterative loop. The inner loop starts from
some initial guess of the space charge density, and Eqs. (1) and (4) are solved
iteratively for F and q. The outer loop is added to update the charge density on the
corona electrode surface until the electric field values there are sufficiently close to
the Peek’s value [23]. All local parameters of the electric corona discharge process,
including the space charge density, the current density, the electric field and the
potential for every point in the simulation domain, as well as the global parameters,
such as I2V characteristics, can be obtained.
The I2V characteristics for R ¼ 100 mm and D ¼ 1 cm is shown in Fig. 2. The
numerically calculated corona onset voltage is 4.625 kV, and the corona current
rises approximately as a square function of the applied voltage above the onset level.
ARTICLE IN PRESS
342 L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350

25

20

Current in uA
15

10

0
2.5 5 7.5 10 12.5
Applied voltage in kV
Fig. 2. Numerically calculated I2V characteristics for 100 mm hyperboloidal needle 1 cm above the
ground plate.

Fig. 3. (a) Space charge density distribution and (b) the equipotential lines for R ¼ 100 mm needle
D ¼ 1 cm above the plate at 7 kV.

Fig. 3 shows the space charge density contours and the equipotential lines for the pin
voltage equal to 7 kV. The maximum space charge density is equal to 0.068 C/m2 at
the point located next to the tip of the needle. The space charge density decreases
quickly towards the ground plane, as well as while going away from the tip along the
corona electrode surface. While there is still some charge density on the ground
plate, starting from certain points on the corona electrode surface, the space charge
decreases to zero. The space charge forms a conical cloud, and there is no space
charge outside this area. On the equipotential lines graph, there are 20 equipotential
lines with a 0.35 kV voltage difference between every two adjacent lines. The gradient
of the voltage near the needle tip is much higher than that near the plane, so the
electric field reaches very high values at these points.
Fig. 4 shows 3D plots of the axial and radial components of the Coulomb force
density which is numerically calculated by multiplying the space charge density and
ARTICLE IN PRESS
L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350 343

Fig. 4. (a) Axial and (b) radial Coulomb forces for 100 mm needle 1 cm above the plate at 7 kV.

the electric field for every point in the computational domain. Both components have
nonzero values in a conical region with the vertex at the tip of the pin, and zero values
everywhere else. It is obvious that the maximum force density is located at the tip of
the needle where both the space charge density and the electric field reach the
maximum values. The axial Coulomb force dominates and its maximum value is about
10 times higher than the radial one. Both components decrease drastically when
approaching the ground plane and going away from the axis. That is why the
maximum axial velocity occurs near the pin tip on the axis, as can be seen from Fig. 7.

3.2. Air flow

The FLUENT6.1.22 commercial software was used to perform the numerical


simulation of the air flow. The governing partial differential Eqs. (7)–(9) are solved
by using the finite volume method: a control-volume-based technique is used to
convert the governing equations into a set of algebraic equations that can be solved
numerically. This control-volume technique consists of integrating the governing
equations about each control volume, yielding discrete equations that conserve each
quantity on a control-volume basis. The user-defined functions (UDF) allowed us to
input the Coulomb force, calculated in the previous step, into every single volume
cell of the discretized FLUENT model.

3.2.1. R ¼ 100 mm hyperboloidal needle D ¼ 1 cm above ground plane


The path lines of the airflow in the air gap between the hyperboloidal needle and
the ground plate are displayed in Fig. 5 with the arrows showing the directions of the
flow. Near the axis of symmetry and far from the ground plate, the airflow is almost
in the axial direction only. This pattern gradually changes to more radial at points
closer to the ground plate. Finally, all the path lines are bent into the radial direction
next to the surface of the plane. This change of direction gives rise to the
recirculating EHD airflow that is shown in the right bottom corner of Fig. 5 and has
been observed by several researchers [14,15–18].
Fig. 6 presents the plots of the axial and radial velocity components of the air flow.
It can be seen that at the corona electrode surface, the velocity is equal to zero, then
ARTICLE IN PRESS
344 L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350

Fig. 5. Pathlines of the airflow for 100 mm needle 1 cm above the ground plate at 7 kV.

Fig. 6. (a) Axial and (b) radial air velocity for 100 mm needle 1 cm above a ground plate at 7 kV.

it quickly reaches the maximum axial value that occurs on the axis near the tip of the
pin, where the air is accelerated by the largest axial Coulomb force. Then, the axial
velocity decreases toward the plane as well as moves away from the axis; eventually it
drops to zero at the ground plate. Far away from the axis, the axial velocity
component would change its direction from positive to negative due to the
recirculating flow. The radial velocity, however, does not follow the same pattern as
the axial component. It changes not only in the magnitude but also in the direction:
it has zero value along the axis of symmetry, negative values at points close to the
pin, and positive values close to the plane.
ARTICLE IN PRESS
L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350 345

18

16
5kV
7kV
14 9kV

Axial velocity in m/s


11kV
12
10
8
6

4
2

0
0 0.002 0.004 0.006 0.008 0.01
Axial position in m

Fig. 7. Axial air velocity profiles on the axis of symmetry for different applied voltages.

The simulation results also predict the strong effect of the applied voltage on the
airflow velocity magnitude. Fig. 7 shows the axial velocity profiles along the axis of
symmetry for four different voltage levels. At higher voltages, both the space charge
density and the electric field increase, resulting in larger Coulomb force and higher
axial velocity. On the other hand, the axial velocity profile tends to keep in a same
pattern and the largest velocity value occurs at almost the same point.
Fig. 8 presents the numerically predicted distribution of the relative pressure in the
air gap between the two electrodes. The maximum positive pressure occurs at the
intersection of the axis and the plane where the airflow is forced to totally change
direction from axial to radial. Maximum negative pressure occurs at the tip of the
pin, where the highest acceleration and the largest axial velocity attempt to make a
vacuum around this small area.
Similarly as the velocity, the pressure is also greatly influenced by the applied
voltage. Fig. 9 shows the relative pressure profiles on the axis of the system and on
the surface of the ground plane for four different voltage levels. In both cases, the
maximum pressure drastically rises when the applied voltage increases.

3.2.2. R ¼ 95 mm needle D ¼ 2 cm above the grounded metal grid


In some practical applications, the ground plate can be replaced with a metal wire
grid. The simulation of the electric corona discharge in the needle-to-grid
configuration would be quite complex if the grid geometry is to be represented
exactly; a 3D model would be necessary, leading to a very large algebraic system.
However, the grid structure should not have a significant effect if the distance
between grid wires is much smaller than the distance between the corona electrode
and the ground grid. This conclusion can be obtained from Fig. 10 which compares
the experimental I2V characteristic of the pin–plate configuration to that of the
pin–grid configuration. Both curves are almost identical and very close to the
ARTICLE IN PRESS
346 L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350

Fig. 8. Relative pressure for 100 mm needle 1 cm above ground plate at 7 kV.

Fig. 9. Relative pressure profiles on the symmetry axis and the ground plane for different applied voltages.

simulation I2V curve with 2.2e–4 cm2/Vm as the ion mobility [24]. This part has
been done with a slightly different needle: instead of a hyperboloidal one, a 5 cm long
cylindrical needle of 1 mm in diameter with a 1.5 cm long conical ending capped with
a hemispherical tip 95 mm radius is used [12].
However, the airflow distributions in these two cases are quite different, because
the grid creates practically almost no obstruction to the moving air. Fig. 11 shows
the numerically calculated axial and radial velocity distributions for R ¼ 95 mm;
D ¼ 2 cm at 8 kV in the pin–mesh configuration. The axial velocity distribution has a
ARTICLE IN PRESS
L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350 347

20
Simulation
18 Experiment(plate)
Experiment(grid)
16
Current in uA 14
12
10
8
6
4
2
0
4 6 8 10 12 14
Applied voltage in kV
Fig. 10. Calculated and experimental I2V characteristics for metal grid and plate cathode.

Fig. 11. (a) Axial and (b) radial velocity distributions for 95 mm needle 2 cm above the ground grid.

similar pattern to that of Fig. 6, but the axial velocity in this case will not decrease to
zero at the grid. Instead, the air will flow through the grid with a speed of several
meters per second. The radial velocity near the grid is very small, in contrast to the
relatively large one for the solid plate case.
It is well known that in the pin–plate configuration, it is impossible to insert an
anemometer sensor between the two electrodes to measure the airflow velocity when
electric corona discharge occurs. This is not only because the sensor will greatly
effect the process of the corona discharge and disturb the airflow, but also because
the electric field will make the anemometer reading inaccurate and even damage the
metal sensor. The pin–grid configuration, however, allows air to flow through the
grid. Therefore, its velocity can be measured by a hot-wire anemometer placed
outside of the grid where there is no electric field. However, it should be pointed out
ARTICLE IN PRESS
348 L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350

that in the numerical model, the grid is treated as infinitely thin, while practically the
presence of the grid will slightly affect the airflow, which is the source of some
measurement errors.

4. Experimental results

The results of the numerical simulation of the air flow for the pin–grid
configuration have been verified by measurements that have been conducted on
the experimental set-up shown in Fig. 12. The steel grid has the mesh of 2  2 mm2
and the wire diameter is Øwire ¼ 0:5 mm: In the vicinity of the grid which is 2 cm
away from the needle, the air flow is non-uniform, so placing the anemometer 3.6 cm
away from the needle, instead of being adjacent to the grid, will provide more
reliable measurement data.
Fig. 13 shows the simulation axial velocity profiles versus radial distance at the
distance of 3.6 cm from the needle in comparison to the experimental ones. The
experimental and numerical curves are in very close agreement for rX1 mm at
voltage levels of both 8 and 13 kV; however, much larger differences can be observed
near the symmetry axis. The much larger difference in the region 0oro1 mm could
be due to two reasons. Firstly, the corona discharge produces a very thin air jet due
to the very sharp corona electrode, but the anemometer sensor has a relatively large
dimension, hence it is impossible to pick up the drastically changed air velocity in the
vicinity of the symmetry axis. Secondly, despite the small dimensions of the grid
wires, the grid actually tends to smooth out and disperse the very concentrated
airflow.

Fig. 12. Schematic representation of the experiment set-up used to measure the air velocity and I2V
characteristics.
ARTICLE IN PRESS
L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350 349

8
4
7 experiment

Axial velocity in m/s


Axial velocity in m/s

3.5 simulation
experiment 6
3 simulation
5
2.5
4
2
1.5 3

1 2
0.5 1
0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
(a) Radial position in mm (b) Radial position in mm

Fig. 13. Numerically calculated and experimental axial velocity profile 3.6 cm away from the tip of 95 mm
needle 2 cm above the grounded metal grid at: (a) 8 kV and (b) 13 kV.

5. Conclusions

This paper presents a numerical algorithm which can be used to predict the EHD
flow in air produced by the electric corona discharge in the positive pin–plane and
the pin–grid configurations. This algorithm is comprised of two stages: first, the
distributions of the electric field, the electric potential and the space charge density
are predicted by using a hybrid technique based on the boundary and finite element
methods and the method of characteristics. The accuracy of this numerical model
has been validated by the good agreement between the numerical and the
experimental I2V characteristics. Then, the Coulomb force is entered into
the airflow simulation by user-defined-functions of the commercial FLUENT
software. The EHD flow pattern, as well as the velocity and the pressure
distributions are predicted.
The simulation results show that the recirculating EHD flow is generated by the
electric corona discharge in the pin–plate geometry. The axial velocity reaches its
maximum at the point on the axis not far from the pin tip. Then, it decreases to zero
at the ground plane. At points close to the symmetry axis, the air flow is practically
axial, but is diverted to the radial by the plane. The airflow velocity profiles and the
pressure distribution are greatly affected by the applied voltage as well as by
the corona device configuration.
The flow characteristics in the pin–grid and the pin–plate geometries are very
similar, but air can penetrate the grid in the pin–grid configuration and no
significant radial velocity is generated. The numerical predictions have been
validated experimentally by measuring the axial velocity on the outer side of the
grid. While quite good accuracy has been obtained for points off the axis, the
difference is rather large on the axis of symmetry. A simple hot wire anemometer
was not able to accurately measure the EHD flow which has the form of a very thin
axial jet.
ARTICLE IN PRESS
350 L. Zhao, K. Adamiak / Journal of Electrostatics 63 (2005) 337–350

References

[1] L.B. Loeb, Electrical Coronas, University of California Press, London, England, 1965.
[2] J.D. Cobine, Gaseous Conductors, Dover, New York, 1958.
[3] J. Batian, F. No.el, S. Lachaud, R. Peyrous, J.F. Loiseau, Hydrodynamical simulation of the electric
wind in a cylindrical vessel with positive point-to-plane device, J. Phys. D: Appl. Phys. 34 (2001)
1510–1524.
[4] S. Oglesby, Electrostatic Precipitation, Marcel Dekker, New York, 1978.
[5] J.F. Hughes, Electrostatics Powder Coating, Wiley, New York, 1984.
[6] F. Bastien, Acoustics and gas discharges: application to loudspeakers, J. Phys. D: Appl. Phys. 20
(1987) 1547–1557.
[7] J. Darabi, M.M. Ohadi, D. Devoe, An electrohydrodynamic polarization micropump for electronic
cooling, J. Microelectromech. Syst. 10 (2001) 98–106.
[8] http://www.sharperimage.com.
[9] J.R. McDonald, W.B. Smith, H.W. Spencer III, L.E. Sparks, A mathematical model for calculating
electrical conditions in wire-duct electrostatic precipitation devices, J. Appl. Phys. 151 (1977)
2231–2243.
[10] R. Morrow, The theory of positive glow corona, J. Phys. D: Appl. Phys. 30 (1997) 3099–3114.
[11] K. Adamiak, Simulation of corona in wire-duct electrostatic precipitator by means of the boundary
element method, IEEE Trans. Indust. Appl. 30 (1994) 381–386.
[12] K. Adamiak, P. Atten, Simulation of corona discharge in point-plane configuration, Proceedings of
the ESA–IEEE Joint Meeting on Electrostatics, 2003, pp. 104–118.
[13] K. Adamiak, J. Zhang, L. Zhao, Finite element versus hybrid BEM–FEM techniques for the electric
corona simulation, Electrotech. Rev. 10 (2003) 753–756.
[14] T. Yamamoto, H.R. Velkoff, Electrohydrodynamics in an electrostatic precipitator, J. Fluid Mech.
108 (1981) 1–18.
[15] G.A. Kallio, D.E. Stock, Interaction of electrostatic and fluid dynamic fields in wire-plate
electrostatic precipitators, J. Fluid. Mech. 240 (1992) 133–166.
[16] M. Fukumoto, R. Ohyama, Image analysis of gas-phase EHD flow field for needle-plane electrode
system, 2003 Annual Report Conference on Electrical Insulation and Dielectric Phenomena,
Albuquerque, New Mexico, 2003, pp. 694–697.
[17] Ph. Bequin, K. Castor, J. Scholten, Electric wind characterisation in negative point-to-plane corona
discharges in air, Eur. Phys. J. Appl. Phys. 22 (2003) 41–49.
[18] Y.K. Stishkov, A.A. Ostapenko, M.A. Pavleyno, Velocity and power fields electrohydrodynamic
flows, IEEE Conference on Electrical Insulation and Dielectric Phenomena, Austin, TX, USA, 1999,
pp. 822–825.
[19] W.-J. Liang, T.H. Lin, The characteristics of ionic wind and its effect on electrostatic precipitators,
Aerosol Sci. Technol. 20 (1994) 330–344.
[20] J.Q. Feng, Electrohydrodynamic flow associated with unipolar charge current due to corona
discharge from a wire enclosed in a rectangular shield, J. Appl. Phys. 86 (1999) 2412–2418.
[21] T. Yamamoto, M. Okuda, M. Okubo, Three-dimensional ionic wind and electrohydrodynamics of
tuft/point corona electrostatic precipitator, IEEE Trans. Indust. Appl. 39 (2003) 1602–1607.
[22] M.H. Lean, Simulation of charging and coupled airflows in corona devices, IEEE Trans. Magn. 30
(1994) 3355–3358.
[23] A.M. Meroth, T. Gerber, C.D. Munz, P.L. Levin, A.J. Schwab, Numerical solution of nonstationary
charge coupled problems, J. Electrostat. 45 (1999) 177–198.
[24] E.W. McDaniel, E.A. Maso, The Mobility and Diffusion of Ions in Gases, Wiley, New York, 1973.

You might also like