You are on page 1of 27

ANRV400-FL42-21 ARI 13 November 2009 15:2

ANNUAL
REVIEWS Further Dielectric Barrier Discharge
Click here for quick links to
Annual Reviews content online,
including:
Plasma Actuators for
• Other articles in this volume
• Top cited articles
• Top downloaded articles
Flow Control∗
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

• Our comprehensive search


Thomas C. Corke,1 C. Lon Enloe,2
and Stephen P. Wilkinson3
1
Center for Flow Physics and Control, Aerospace and Mechanical Engineering Department,
University of Notre Dame, Notre Dame, Indiana 46556; email: tcorke@nd.edu
2
Department of Physics, U.S. Air Force Academy, Colorado Springs, Colorado 80840
3
Flow Physics and Control Branch, NASA Langley Research Center, Hampton,
Virginia 23681-2199

Annu. Rev. Fluid Mech. 2010. 42:505–29 Key Words


First published online as a Review in Advance on aerodynamic control, ionized gasses, body force
September 21, 2009

The Annual Review of Fluid Mechanics is online at Abstract


fluid.annualreviews.org
The term plasma actuator has now been a part of the fluid dynamics flow-
This article’s doi: control vernacular for more than a decade. A particular type of plasma actu-
10.1146/annurev-fluid-121108-145550
ator that has gained wide use is based on a single–dielectric barrier discharge
Copyright  c 2010 by Annual Reviews. (SDBD) mechanism that has desirable features for use in air at atmospheric
All rights reserved
pressures. For these actuators, the mechanism of flow control is through a
0066-4189/10/0115-0505$20.00 generated body-force vector field that couples with the momentum in the
∗ external flow. The body force can be derived from first principles, and the
The U.S. Government has the right to retain a
nonexclusive, royalty-free license in and to any effect of plasma actuators can be easily incorporated into flow solvers so
copyright covering this paper.
that their placement and operation can be optimized. They have been used
in a wide range of internal and external flow applications. Although ini-
tially considered useful only at low speeds, plasma actuators are effective in
a number of applications at high subsonic, transonic, and supersonic Mach
numbers, owing largely to more optimized actuator designs that were de-
veloped through better understanding and modeling of the actuator physics.
New applications continue to appear through a growing number of pro-
grams in the United States, Germany, France, England, the Netherlands,
Russia, Australia, Japan, and China. This review provides an overview of
the physics and modeling of SDBD plasma actuators. It highlights some of
the capabilities of plasma actuators through examples from experiments and
simulations.

505
ANRV400-FL42-21 ARI 13 November 2009 15:2

1. BACKGROUND
There has been increasing interest in dielectric barrier discharge (DBD) plasma actuators for flow
control in the past 10 years worldwide. The tremendous growth in research stems from their
special features, including the fact that they are fully electronic with no moving parts; a fast time
response for unsteady applications; a very low mass, which is especially important in applications
with high g-loads; the ability to apply the actuators onto surfaces without the addition of cavities or
holes; the efficient conversion of the input power without parasitic losses when properly optimized
(Enloe et al. 2004b, Orlov 2006, Roth & Dai 2006, Thomas et al. 2009); and the ability to simulate
easily their effect in numerical flow solvers (Corke et al. 2006, Orlov 2006).
The specific DBD configuration used for plasma actuators consists of two electrodes, one
uncoated and exposed to the air and the other encapsulated by a dielectric material; hence, we
refer to this configuration as single dielectric barrier discharge (SDBD). For plasma actuator
applications, the electrodes are arranged in a highly asymmetric geometry as opposed to the
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

parallel-plate arrangement common in industrial DBD uses. An example configuration is shown


in Figure 1. The electrodes are supplied with an AC voltage that, at high enough levels, causes
the air over the covered electrode to weakly ionize (typically less than 1-ppm weakly ionized gas).
In the classic description, the ionized air is a plasma, which is why these are referred to as plasma
actuators (Cavalieri 1995, Corke & Matlis 2000, Corke et al. 2001). The ionized air appears blue,
characteristic of the composition of the air as ionized components of the air recombine and de-
excite (Davidson & O’Neil 1964). The emission intensity is extremely low, requiring a darkened
space to view by eye.
In the presence of the electric field produced by the electrode geometry, the ionized air results
in a body-force vector field that acts on the ambient (nonionized, neutrally charged) air. The body
force is the mechanism for active aerodynamic control.
Langmuir (1928) introduced the term plasma into the physics literature to denote a net electri-
cally neutral region of gas discharge. This definition has been broadened since and now refers to a
system of particles whose collective behavior is characterized by long-range Coulomb interactions

a Induced flow
Plasma

Dielectric layer

Voltage AC
source

Actuator location reference

b
Covered
electrode

Exposed electrode edge

Figure 1
Schematic illustration of a single–dielectric barrier discharge plasma actuator (a) and photograph of ionized
air at 1-atm pressure that forms over an electrode covered by a dielectric layer (b).

506 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

(Kunhardt 2000). Although plasmas are often simply considered as interpenetrating fluids con-
sisting of electrons, positive ions, and neutral particles, the air discharges consist of a multiplicity
of species in numerous charge states, including negative ions. Nonetheless, the quasi-neutral ap-
proximation applies to these, as to almost all, plasmas—that is, the total density of negatively and
positively charged particles in any region of the plasma is approximately equal, with only small
deviations possible locally. Highlighting another property of plasma discharges in air, they are
often described as collisional; i.e., the electron-neutral collision frequency is of the order of, or
greater than, the plasma frequency (the characteristic frequency of electrostatic oscillations in the
plasma). This is a property of air at atmospheric pressures typical of flight, the relevant regime for
DBD plasma actuators.
A gas discharge is created when an electric field of sufficient amplitude is applied to a volume of
gas to generate electron-ion pairs through electron impact ionization of the neutral gas (Kunhardt
1980, Kunhardt & Luessen 1981, LLewellyn-Jones 1966, Raizer 1991). This requires the presence
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

of an initiating number of free electrons, which can either be present from ambient conditions or
introduced purposely (Kunhardt 1980, Kunhardt & Luessen 1981).
A traditional (industrial) arrangement for creating a self-sustained gas discharge at low pressures
of a few torr or less has involved separated facing electrodes. The electric field established by the
two electrodes can either be by direct current (DC) or alternating current (AC). The plasma is
generated by increasing the amplitude of the electric field above the breakdown electric field,
Eb , which is the value needed to sustain electron-ion pairs in the gas in the absence of space-
charge fields (Kunhardt 1980, Kunhardt & Luessen 1981, LLewellyn-Jones 1966, Raizer 1991).
The minimum breakdown electric field is a function of the driving frequency. At atmospheric
pressure, Eb is generally lower for an AC input. The optimum AC frequency depends on the static
pressure and the particular gas. Once created, the electric field needed to sustain the plasma, Es , is
lower than Eb . The difference between the breakdown and sustaining electric fields is a function
of the operating conditions (LLewellyn-Jones 1966, Meek & Craggs 1978, Nasser 1971, Raizer
1991).
As a consequence of the plasma conductivity, there is a current, I, that flows between the
electrodes. A region is formed between the plasma and the cathode electrode, whose role is to
provide current continuity at this interface. The current in the boundary region consists of two
components: the conduction current, Ic , and the displacement current, Id . For DC fields, Id is zero,
and the remaining conduction current consists of electron and ion components. For AC fields,
the contribution of the displacement current to the total current increases with frequency and can
become an important practical consideration in power-supply design.
When the operating conditions (applied field, electrode cross-sectional area, and static pressure)
are such that the current density in the boundary region near the cathode is independent of the
current flowing in the circuit, the discharge is called a normal glow discharge (Kunhardt & Luessen
1981, LLewellyn-Jones 1966, Meek & Craggs 1978, Nasser 1971, Raizer 1991, Roth 1995). For
constant current, the current density in a normal glow discharge scales with the square of the
static pressure. Therefore, the cross-sectional area of plasma decreases with increasing pressure at
constant current. As the static pressure increases at constant current, the current density increases
until the threshold for the development of instabilities leading to a transition to an arc phase is
reached. The threshold current for the development of the glow-to-arc transition depends on the
operating conditions of the discharge.
Many aerodynamic flow-control applications would require plasma actuators to operate near
atmospheric pressure. This favors AC operation over DC because of the lower breakdown voltage
requirement and lack of real currents responsible for electrode corrosion effects. In addition, when
considering any gain an application might provide in system efficiency, one needs to factor in the

www.annualreviews.org • DBD Plasma Actuators 507


ANRV400-FL42-21 ARI 13 November 2009 15:2

power required to operate the actuators. Thus there is a need to consider the actuator’s most
efficient operating conditions to maximize its effect with respect to input power.
Barrier discharges can be operated in various modes (e.g., diffuse, patterned, filamentary, mi-
crodischarge), and the terminology can often become confusing. A good review of the various
modes and terms is presented by Kogelschatz (2002). All techniques, however, use a dielectric
barrier on the surface of one or both electrodes. Okazaki and coworkers were among the first to
use this approach (Kanazawa et al. 1988, 1989, 1990; Kogoma & Okazaki 1994; Okazaki et al.
1993; Yokoyama et al. 1990). In this case, because of the dielectric layer, the electrodes must be
energized with an AC field. Barrier discharges have been operated historically in the microdis-
charge mode (Eliasson & Kogelschatz 1991, Kogelschatz et al. 1997). In this mode, the discharge
consists of a number of parallel filaments, each of which has a limited lifetime. The filaments are
essentially streamer discharges whose lifetimes are governed by the capacitance of the dielectric
barrier (Eliasson & Kogelschatz 1991). The passage of the streamer across the discharge gap
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

locally charges this capacitance, reversing the local field and thus terminating the filament.
In sinusoidally driven DBDs, the plasma forms in nonthermal equilibrium (not to be confused
with nonequilibrium plasma due to excitation time dependency) in which the electrical energy
coupled into the gas is mainly used to produce energetic electrons while the gas remains approx-
imately at ambient temperature, rising only slightly. This is because of the self-termination of
the discharge, which prevents microdischarges from degenerating into thermal arcs (Falkenstein
& Coogan 1997). Self-termination results from the accumulation of electrons at the dielectric
surface facing the cathode. The internal electric field caused by the accumulation of electrons on
the dielectric surface reduces the local field strength, and the microdischarges choke themselves
as the extinction field is reached (Falkenstein & Coogan 1997).
For an SDBD, during one half of the AC cycle, electrons leave the metal electrode and move
toward the dielectric, where they accumulate locally. In the reverse half of the cycle, electrons are
supplied by surface discharges on the dielectric and move toward the metal electrode. The timescale
of the process depends on the gas composition, excitation frequency, and other parameters. In air
at atmospheric pressure, it occurs within a few tens of nanoseconds (Falkenstein & Coogan 1997).
Kline et al. (2001) have studied time-resolved images of spatiotemporal patterns in a one-
dimensional (1D) DBD system. They obtained images of plasma filaments that revealed discharge
stages that lasted only approximately 100 ns. Several discharge stages could occur during a half-
cycle of the driving oscillation, each producing a distinct filament pattern. In some discharges,
there was a temporal structure but spatial disorder, and in others there was both temporal and
spatial disorder.
The dielectric barrier configuration also supports a uniform diffuse discharge operation, as
shown by Okazaki and coworkers (Kanazawa et al. 1988). The mechanisms that play a role in this
are well understood (e.g., Decomps et al. 1994; Roth 1995; Massines et al. 1996, 1998; Trunec
et al. 1998). The stability of the diffuse mode depends on the AC frequency, the gas type, and
the excitation power. The discharge is most stable in helium and mixtures that contain helium,
although other gases have been used including air (Decomps et al. 1994; Kanzawa et al. 1990;
Massines et al. 1996, 1998; Roth 1995; Trunec et al. 1998). The electrode separation is usually
small, of the order of a few centimeters. The electron density of the plasma generated by this mode
is of the order of 1010 cm−3 . Although interesting from a scientific standpoint, the conditions that
support this mode of the discharge are specific enough that they are unlikely to be encountered
in a device designed for aerodynamic applications.
Massines and coworkers (BenGadri et al. 1994, Massines et al. 1998, Rabehi et al. 1994) de-
veloped a 1D model for the DBD dynamics based on the numerical solution of the electron and
ion continuity and momentum transfer equations coupled to Poisson’s equation. As is typical in

508 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

Flow direction

Covered Plasma actuator


electrode (top view)
Exposed
electrode
Figure 2
Photograph of smoke tube introduced at the edge of a boundary layer that is bent toward the wall by a plasma
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

actuator oriented so that the electrodes are parallel to the streamwise direction. Figure taken from Post 2001.

high-pressure discharges, the electrons and ions were assumed to be in equilibrium with the elec-
tric field. Their model gave space and time variations in the electric field, and the electron and ion
densities. The authors accounted for the charge accumulation on the dielectric as the discharge
develops and derived the voltage boundary conditions for dielectrics by considering an equivalent
circuit of the gas gap in series with the equivalent capacitor of the dielectric.
One of the earliest low-speed demonstrations of a plasma actuator was performed by Roth
et al. (1998, 2000). They utilized an array of electrodes separated by a glass-epoxy printed-circuit
board to manipulate the boundary-layer flow over a flat plate at free-stream velocities from 4 to
26 m s−1 . They investigated a number of electrode geometries, one of which was similar to that
shown in Figure 1. With this electrode configuration, they observed that ambient air, marked by
smoke, was drawn toward the covered electrode. A similar photograph recorded by Post (2001) is
shown in Figure 2. In this case, the electrodes are aligned parallel to the mean flow direction. This
accentuates the width of the stream tube that is drawn toward the wall by the actuator. Velocity
surveys of the actuator-induced flow by Roth et al. (1998, 2000) documented a wall-normal mean
velocity profile that is similar to what might be produced by a tangential wall jet. The deflection
of the external flow toward the surface of the dielectric and the jetting of the flow in the direction
of the exposed electrode toward the covered electrode are hallmarks of this actuator design. Any
simulation models for these actuators need to produce this behavior.

2. DIELECTRIC BARRIER DISCHARGE ACTUATOR PHYSICS


2.1. Experimental Observations
In classic DBD processes, the input waveform is sinusoidal. When the AC amplitude is large
enough so that the electric field exceeds Eb , the air ionizes. The ionized air is always observed
to form over the electrode that is covered by the dielectric. To the unaided eye, the ionized air
appears to be generally uniform in color and distribution, with some structure evident that often
appears attached to a particular location on the exposed electrode. The photograph in Figure 1
is a typical example. Time-resolved images of the ionization process, however, indicate it to be
a highly dynamic, spatially evolving, nonequilibrium process with features that develop on the
timescale of the AC period (milliseconds) or less.
Enloe et al. (2004a) studied the space-time evolution of the ionized-air light emission over a
surface-mounted plasma actuator using a photomultiplier tube fitted with a double-slit aperture

www.annualreviews.org • DBD Plasma Actuators 509


ANRV400-FL42-21 ARI 13 November 2009 15:2

1.0 4000
a b
3000
normalized by maximum 0.8
2000
PMT output

0.6 1000

Voltage
0
0.4
–1000

–2000
0.2
–3000

0 –4000
–1 0 1 2 3 4 5 –1 0 1 2 3 4 5
Time (10–4 s) Time (10–4 s)
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

Figure 3
Time series of photomultiplier-tube (PMT) output (a) that is viewing ionized-air light emission at one location over an electrode
covered by a dielectric and corresponding AC input (b) to a plasma actuator. Figure taken from Orlov 2006.

to focus on a narrow 2D region of the plasma. The slit was parallel to the edge of the exposed
electrode and could be moved to different locations over the electrode covered by the dielectric. A
sample time series from Orlov (2006) of the photomultiplier-tube output that was acquired phase
locked with the AC input to the actuator is shown in Figure 3. The light emission is taken as an
indication of the plasma density, which is a good assumption based on the disparate timescales
between the recombination time (order of 10−8 s) (Vidmar & Stalder 2003) and the discharge
timescale (order of 10−3 s).
There are several fundamental features of light-emission time series. First, the air is ionized
only over part of the AC cycle. Second, when it does ionize, its character differs between the first
and second halves of the AC cycle. Finally, the light emission is made up of narrow spikes that
might indicate numerous microdischarges. Similar observations have been documented by Enloe
et al. (2004a), Massines et al. (1998), Eliasson & Kogelschatz (1991), and Kogelschatz et al. (1997),
who generally characterize this process as a DBD.
The explanation for the difference in the emission character in the two half-cycles is associated
with the source of electrons. During the negative-going half-cycle, the electrons originate from the
bare electrode, which is essentially an infinite source that readily gives them up. In the positive-
going half-cycle, the electrons originate from the dielectric surface. These apparently do not
come off as readily, or when they do, they come in the form of fewer, larger microdischarges. The
structural difference in each discharge mode is evident in high-speed images of the microdischarges
that, in fact, make up the apparently uniform actuator discharge (Figure 4). This asymmetry has
been modeled by Boeuf et al. (2007) and plays an important role in the efficiency of the momentum
coupling to the neutrals. It further suggests that some optimization can come in the selection of
the AC waveform to improve the performance of the plasma actuator.
Figure 5 shows a composite of light-emission time series similar to the one in Figure 3 but
measured at different positions over the dielectric surface. These are shown as contours of constant
light-emission intensity for one period, T, of the AC cycle.
The space-time character of the plasma light emission over the covered electrode has a number
of interesting features. For example, there is a sharp amplitude peak near the edge of the exposed
electrode at the first initiation of the plasma. As time increases, the plasma sweeps out from the
junction to cover a portion of the encapsulated electrode. This was similarly noted by Gibalov &

510 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

b
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

Figure 4
High-speed (5-ms exposure) photographs of individual microdischarges in the negative-going (a) and
positive-going (b) phases of plasma-actuator discharge, highlighting the structural asymmetry between the
phases. Figure taken from Enloe et al. 2008.

Pietsch (2000). As the plasma sweeps out away from the edge of the exposed electrode, its light
emission appears to become less intense. Estimates (Enloe et al. 2004a, Orlov 2006, Orlov et al.
2006) indicate that the intensity decreases exponentially from the junction. This led to the use of
an exponential weighting to correct the spatial dependency of the plasma actuator body force in
earlier electrostatic flow simulations (Orlov et al. 2003, Orlov & Corke 2005, Voikov et al. 2004).
Two global features of the space-time evolution of the plasma formation are the velocity at which
the plasma front moves across the dielectric and the maximum extent of the plasma during the
AC cycle. The velocity is represented by the slope, dx/dt, of the front. In Figure 5, the velocity
of the fronts is approximately the same for the two halves of the AC cycle, but the plasma extent
differs. We note that these measurements indicate the time development of the envelope of the
multiple microdischarges that compose the DBD. Although Hoskinson et al. (2008) have been
able to make some estimates of the development rate of individual microdischarges through careful
examination of fast (2–80 ns) gated images, no experimental apparatus exists that can implement
frame rates in the hundreds of megahertz that would be necessary to image the development of
an individual microdischarge event.
Orlov (2006) investigated the effects of voltage and AC frequency on the extent and propagation
velocity of the discharge. He found that the maximum extent increased linearly with increasing
AC voltage amplitude, and it was independent of the AC frequency. However, the velocity of the
plasma front increased linearly with both AC amplitude and frequency. In Orlov’s measurements,
the velocity of the discharge front ranged from 70 to 190 m s−1 .
As mentioned above, plasma actuators with the asymmetric electrode design in Figure 1 induce
a velocity field similar to that of a tangential wall jet. Enloe et al. (2004b) correlated the reaction
force (thrust) generated by the induced flow with the actuator AC amplitude. A similar experi-
ment was performed by Thomas et al. (2009) to investigate parameters in the actuator design.

www.annualreviews.org • DBD Plasma Actuators 511


ANRV400-FL42-21 ARI 13 November 2009 15:2

1.0 0.1
0.1
0.3
0.3
0.5
0.7
0.5

0.7
0.3 0.1

0.1
0.9
0.9 0.7
0.5
0.3 0.1

t/T
0.5

0.1
0.1
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

0.3
0.5
0.1

0
0 5 10 20
x (mm)
Figure 5
Space-time variation of the measured plasma light emission for a single–dielectric barrier discharge plasma
actuator corresponding to one period, T, of the input AC cycle. The x axis is the distance over the covered
electrode measured from the edge of the bare electrode at the interface of the covered electrode. Figure
taken from Orlov 2006, Orlov et al. 2006.

A schematic of their setup and sample results are shown in Figure 6. At the lower voltages, the
3.5
induced thrust was proportional to VAC . This was first observed by Enloe et al. (2004b). Thomas
et al. (2009) verified the consistency between the reaction force and the fluid momentum by in-
tegrating the velocity profiles downstream of the actuator. Post (2004) found that the maximum
3.5
induced velocity was proportional to VAC , which is consistent with conserved momentum in the
self-similar velocity-profile region near the actuator.
Post (2004) and Enloe et al. (2004b) showed that, with increasing AC amplitude, the maximum
velocity induced by the plasma actuator was limited by the area (extent for a unit spanwise width)
of the covered electrode. Thus the dielectric area needed to store charge can be too small to take
full advantage of the applied voltage. This effect can be observed in the thrust measurements in
Figure 6 at the highest voltages for the Teflon dielectric, at which the thrust no longer increases
3.5
as VAC and begins to asymptote.
Enloe et al. (2004a) computed the power dissipation in the discharge by sampling the voltage and
current waveforms across the electrodes and numerically integrating the product of the waveforms
over one period of the discharge to integrate out the reactive power and account exclusively
3.5
for the power dissipated in the plasma. The dissipated power also followed VAC , indicating a
direct proportionality with the induced momentum. They considered a model for the air above
the dielectric material covering the electrode that consisted of a capacitor and resistor in series.
Figure 7 shows this model for the SDBD plasma actuator. Before the air ionizes, the capacitor,
C1 , corresponds to the value for air. When the air ionizes, this capacitor effectively becomes

512 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

Dielectric
a b
Insulator

100

Electrodes Induced flow

Thrust (gm)
Spring scale Induced thrust
10–1
Teflon
Glass
T = C1V3.5
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org

8 10 12 14 16 18 20
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

VAC (kV–rms)

Figure 6
Schematic of experimental setup for measuring induced thrust from a single–dielectric barrier discharge
plasma actuator (a) and measured thrust versus applied AC voltage (b). Figure taken from Thomas et al.
(2009).

a short circuit. The remaining circuit elements, R1 and C2 , then form a voltage divider. The
impedance, Z2 , of the capacitance, C2 , is given as Z2 = −i/ωC2 , where ω = 2π/ f AC , and fAC is
the AC frequency. For a fixed frequency, C2 would be a constant. Then one might expect that the
electrical power dissipated by the plasma would be VAC 2
/R1 or generally be proportional to VAC 2
.
However, experiments (Enloe et al. 2004a) indicate that this is not the case.
There are several possible reasons why the power dissipated by the plasma is not proportional
to Vac2 . These include (a) that C2 is not constant but increases with the applied AC voltage and/or
(b) that R1 decreases with increasing applied AC voltage.
Considering that the sweep-out velocity of the plasma front has been documented to increase
with increasing voltage, the area over the dielectric that composes the capacitance, C2 , increases
faster with increasing voltage. Therefore, the value of C2 is not constant but increases in proportion

Plasma

Exposed electrode

Rp C1A C1B

Vac Dielectric surface

C3 C2A C2B

Insulated electrode

Figure 7
Lumped-element circuit model of a single–dielectric barrier discharge plasma actuator. Figure taken from
Enloe et al. 2004a.

www.annualreviews.org • DBD Plasma Actuators 513


ANRV400-FL42-21 ARI 13 November 2009 15:2

0.14

0.12

Efficiency (mN/W)
0.10

0.08

0.06

0.04
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

0.02
0 2 4 6 8 10 12 14 16 18 20 22
Gas composition (% O2)
Figure 8
Plasma-actuator net-force production as a function of the percentage of oxygen. Figure taken from Enloe
et al. 2006.

to VAC . With regard to R1 , numerical simulations of the circuit in Figure 7 intended to repro-
duce the output waveforms observed in experiments indicated that the resistance needed to vary
inversely with the applied voltage (Enloe et al. 2004a). Therefore, both effects are in play.
The dependency of the dissipated power in the plasma and the induced momentum on the
applied voltage is an important characteristic that any physical model needs to replicate. These
results indicate that models must include the dynamics of the plasma initiation and sweep-out
over the dielectric covered electrode that occurs twice during the AC period. Naude et al.’s (2004)
electrical model for DBD addressed this with the addition of electrical elements (Zener diodes)
that switch the current path between different passive circuit elements to control the characteristics
with voltage and frequency.
One final attribute of DBD plasmas as aerodynamic actuators (as opposed to their discharge
characteristics alone) is the importance of the species composition of the plasma, specifically the
dramatic difference that the presence of oxygen in the air makes in the actuator’s force production.
Enloe et al. (2006) showed, and other experimenters have subsequently confirmed, that removing
oxygen from the air surrounding the plasma actuator only modestly changes its discharge prop-
erties (20% in such parameters as discharge current for a given voltage). However, its removal
results in a dramatic reduction in the net force produced (by up to 80%). This is illustrated in
Figure 8. The propensity of oxygen to form negative ions (by electron attachment) adds a species
to the composition of the plasma that is usually not accounted for in the typical analysis of plasmas
that assumes that positive ions and negative electrons are formed in equal numbers.

2.2. Dielectric Barrier Discharge Body-Force Models


One of the first models for a DBD plasma actuator, developed by Massines et al. (1998), was a
1D model based on a simultaneous solution of the continuity equations for charged and excited
particles and the Poisson equation.
Paulus et al. (1999) developed a particle-in-cell simulation to study the time-dependent evo-
lution of the potential and the electrical field surrounding 2D objects during a high-voltage
pulse. The numerical procedure was based on the solution of the Poisson equation on a grid in a

514 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

domain containing an L-shaped electrode to determine the movement of the particles through the
grid. The simulation showed that the charged particles move toward the regions of high electric
potential, creating a high-electric-field strength near the electrode’s edges. In addition, it showed
that the plasma builds up on a very short, microsecond, timescale.
A model for the body force produced by the plasma on the neutral air was presented by Roth
and colleagues (Roth et al. 2000, Roth & Dai 2006). This model was based on a derivation of the
forces in gaseous dielectrics given by Landau & Lifshitz (1984). The body force is proportional
to the gradient of the squared electric field, namely
 
d 1
fb = ε0 E 2 . (1)
dx 2
This model can be problematic as it is based on a static formulation and does not account for the
presence of the charged particles, both of which have been shown to be important in experiments.
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

Boeuf & Pitchford (2005) raised the same warning in their derivation of Equation 1. Enloe et al.
(2004b) further showed that the body force given by Equation 1 is only correct in the special case
of a 1D condition where E = E x î, Ey = Ez = 0, and ∂/∂ y = ∂/∂z = 0. This special case is not
relevant to physical applications that are at least 2D.
Shyy et al. (2002) presented a model for the body force that is widely used in the literature
because of its simplicity. A basic assumption of this model is that the electric-field strength, E,
decreases linearly from the edge of the bare electrode toward the dielectric-covered electrode. This
assumption is not consistent with experiments (Enloe et al. 2004b, Orlov 2006, Orlov et al. 2006),
which show an exponential spatial decay. As a result, the model overpredicts the actuator effect.
Furthermore, it produces body-force vectors that point away from the dielectric surface, which
is again inconsistent with experiments. Finally, the body-force magnitude in the model is a linear
3.5
function of the AC voltage rather than being proportional to VAC , as observed in experiments.
Singh & Roy (2008) used the results of the body forces obtained from a first-principle simulation
along with empirical observations of actuator behavior to develop an approximation for the 2D
body-force components. This approach makes the calculation of the body force a curve-fitting
problem that is only valid for a single-actuator configuration. Similar to Shyy et al.’s model, it
does not include temporal characteristics of the body force, and the net body force does not scale
properly with voltage.
Suzen and colleagues (Suzen et al. 2005, Suzen & Huang 2006) utilized the electrostatic model
with an imposed Gaussian distribution for the spatial charge distribution to compute the plasma
body force using Enloe et al.’s (2004a,b) formulation. They proposed to split the electrostatic
equations into two parts: the first part due to the external electric field and the second part due to
the electric field created by the charged particles. Qualitatively, the net body-force vectors appear
2
to be physical. However, the scaling with AC voltage is incorrect, being proportional to VAC . Such
scaling is a closed-form solution of the electrostatic model (Orlov 2006), from which Suzen et al.’s
model originated.
Boeuf & Pitchford (2005) considered a collisional discharge in a numerical estimation of the
force acting on gas molecules in a 2D asymmetric surface DBD. For this, they considered nitrogen
at atmospheric pressure. They concluded that the asymmetry in the electrode configuration in-
duces an asymmetry in the flow, comparable with a DC force in surface corona discharges. Boeuf
et al. (2007) subsequently extended their simulation to encompass multiple microdischarge events
and not only were able to simulate the qualitative difference between positive- and negative-going
discharges that experimenters have observed, they also concluded that it is the electric-field effects
on the charged particles left after these microdischarges terminate that are the predominate cause
of force production by the actuator.

www.annualreviews.org • DBD Plasma Actuators 515


ANRV400-FL42-21 ARI 13 November 2009 15:2

There have been numerous models developed for DBDs in air that include complicated chem-
istry. These models usually include 20–30 reaction equations, each with different reaction times
and energy outputs. These equations account for electron, ion-neutral, and neutral-neutral re-
actions in different gases that are present in the air (Gibalov & Pietsch 2000, Golubovskii et al.
2002, Kozlov et al. 2001, Madani et al. 2003, Pai et al. 1996). For the most part, these models were
developed for simple 1D geometries consisting of axisymmetric facing electrodes. To simplify the
chemistry, Font and coworkers (Font 2004, Font & Morgan 2005) recently considered the plasma
discharge in a 2D asymmetric plasma actuator that included only nitrogen and oxygen reactions.
With this model, they were able to simulate the propagation of a single streamer from the bare
electrode to the dielectric surface and back.
Likhanskii et al. (2006) modeled the weakly ionized-air plasma as a four-component mixture of
neutral molecules, electrons, and positive and negative ions that included ionization and recom-
bination processes. Their simulations indicated the importance of the presence of negative ions
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

in the air. They also suggest that the charging of the dielectric surface by electrons in the cathode
phase is critical, during which they believe it acts as a “harpoon” pulling positive ions forward and
accelerating the gas in the anode phase.
Generally speaking, the charged-particle models can precisely describe all the different pro-
cesses involved in the plasma actuator. However, they are computationally time-consuming and
require significant computer resources. This is especially true if they are applied to air at near-
atmospheric pressures. Such simulations are not suitable to be a part of a design tool that would
be used in the iterative optimization of the plasma actuators and the design of flow-control appli-
cations based on plasma actuators.
Orlov and colleagues (Orlov 2006, Orlov et al. 2006) addressed the need for an efficient method
to predict the body-force field of SDBD plasma actuators by developing a space-time lumped-
element circuit model that is a variation of the one proposed by Enloe et al. (2004a) shown in
Figure 7.
A schematic of Orlov and colleagues’ (Orlov 2006, Orlov et al. 2006) model is shown in
Figure 9. The unique aspect of this model is the division of the domain over the covered electrode
into N parallel networks. The properties of each parallel network depend on its distance from the
exposed electrode. These were designated parallel network 1, which is closest to the exposed
electrode, to parallel network N, which extends the furthest distance over the covered electrode.
Each parallel network consists of an air capacitor, a dielectric capacitor, and a plasma-resistive
element, as in the earlier model (Enloe et al. 2004a, Orlov 2006). Zener diodes were added to set
a threshold voltage level at which the plasma initiates and to switch into the circuit the different
plasma-resistance values based on the current direction, which experiments had shown to be
important. The N-circuit arrangement is illustrated in Figure 9.
The values of the air capacitor and resistor in the n-th subcircuit are based on their distance
from the edge of the exposed electrode. The value of the dielectric capacitor for each subcircuit
is a property of the dielectric material. Assuming that the paths are parallel to each other, and the
length of path, In , is proportional to its position number, n, it then follows that the air capacitance
of the n-th subcircuit, Can , is proportional to 1/n, and the air resistance of the n-th subcircuit is
proportional to n. Based on this, subcircuits that are furthest from the edge of the electrodes have
the lowest air capacitance and the largest air resistance.
For a time-varying (AC) applied voltage, the voltage on the surface of the dielectric at the n-th
parallel network is given as
 
d Vn (t) d Vapp (t) Can Ipn (t)
= + kn , (2)
dt dt Can + Cdn Can + Cdn

516 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

a Parallel subcircuits

ln

Exposed electrode
1 2 3 4 N

Dielectric layer Covered electrode

AC voltage source
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

b
D1f D1b D2f D2b Dnf Dnb
C1a C2a Cna
R1f R1b R2f R2b Rnf Rnb

Exposed electrode Dielectric surface

C1d C2d Cnd

Vapp
Covered electrode

Figure 9
Space-time lumped-element circuit model for a single–dielectric barrier discharge plasma actuator that
divides the region over a covered electrode into N subregions (a) that each represents a parallel arrangement
of circuit elements (b). Figure taken from Orlov 2006, Orlov et al. 2006.

where Ipn (t) is the time-varying current through the plasma resistor, and the diodes are represented
by the variable kn . When the threshold voltage is exceeded, kn = 1. Otherwise, kn = 0. The current
through the n-th plasma resistance is given by
1  
I p n (t) = Vapp (t) − Vn (t) , (3)
Rn
where Rn = Rnf or Rn = Rnb , based on the current direction. The ratio of the two plasma resistances
used by Orlov and colleagues (Orlov 2006, Orlov et al. 2006) was Rnf /Rnb = 5, which was based
on the difference in the currents measured in experiments (Orlov 2006).
The solution of the model equations gives the voltage on the surface of the dielectric, Vn (t),
and the current, I p n (t), for each parallel circuit element. The space-time variation in the rectified
current agreed well with the experimental observations of the plasma light emission (Figures 3
and 5) for a large range of AC voltages and frequencies (Orlov 2006, Orlov et al. 2006).
The space-time dependent voltage, Vn (t), from the lumped-element model serves as the time-
dependent boundary condition for the electric potential, ϕ, found in the solution of the electrostatic
Poisson equation:
1
∇(ε∇ϕ) = ϕ. (4)
λ D2

www.annualreviews.org • DBD Plasma Actuators 517


ANRV400-FL42-21 ARI 13 November 2009 15:2

x(t)

Bare electrode Dielectric Dielectric


BC: Vapp(t) BC: Vn(t) BC: V = 0

Figure 10
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

Computational domain for calculation of unsteady plasma body force. BC refers to boundary conditions
used in solving Equation 4.

The time-dependent extent of the plasma on the surface of the dielectric, x(t), specifies the region
where charged particles are present above the covered electrode. This defines a moving boundary,
as shown in Figure 10. The boundary value of the electric potential on the bare electrode is the
applied voltage Vapp (t). At the outer boundaries at infinity, the boundary conditions are ϕ = 0.
The electric potential, ϕ(t), is determined at small time steps of the AC cycle. It is then used
to calculate the time-dependent body force produced by the plasma, given by
 
 = − ε0 ϕ(t) E(t),
fb∗ (t) = ρc E(t)  (5)
λ D2
where it is necessary to keep track of the direction of the current, which signifies the sign on the
charge, ϕ, on the dielectric, and therefore the sign convention for the body force.
This model certainly has the benefit of being computationally efficient. The question is, is it
correct? Or, more accurately, do the assumptions of the model made in the interest of efficiency
still allow it to replicate the behavior of the actuator, or is too much of the essential physics lost?
Orlov and colleagues’ model allows the plasma boundary conditions to evolve over a timescale
that is short compared with that of the AC waveform that is driving the actuator. As a direct
3.5
consequence, the net (AC cycle-averaged) body force from this model scales as VAC , which agrees
with experiments (Orlov 2006). It also predicts an asymptote in the body force at higher voltages if
the covered electrode is too small. The model also indicates that for a given plasma actuator design,
there is an optimum AC frequency that maximizes the net body force. These are considerations
that relate directly to how an actuator might be fielded in a practical system.
Nonetheless, the model, by necessity, cannot describe the evolution of the microdischarges
themselves, during which time the local nonuniformities in charge density are the most extreme.
This may not be as significant a drawback as one might think. The results of the modeling
done by Boeuf et al. (2007) indicate that although the forces on the plasma are greatest dur-
ing these short (tens of nanoseconds) periods, in fact the largest contribution to the total force
is a result of the electric field’s interaction with the charged particles remaining after the mi-
crodischarges have terminated because these particles outlive the microdischarges by orders of
magnitude (microseconds).
Orlov and colleagues’ model does not explicitly address the effect of oxygen and its propensity
to form negative as well as positive ions, but in fact it does encompass this phenomenon implicitly.
Equation 5 effectively states that the force on the neutrals is the same as the force on the plasma.
On one hand, this is an excellent assumption: The ionized particles represent such a small fraction

518 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

of the air that the probability of an ion or electron crossing the gap between the electrode and
the dielectric surface (or back) without interacting with a neutral molecule is vanishingly small.
On the other hand, the momentum-transfer cross section for electrons on neutrals is substantially
smaller than that for ions on neutrals, so the assumption that one can simply track the net charge
density, ρ c , to compute the force on the neutrals by the plasma is incorrect. If, conversely, enough
electrons attach themselves to the oxygen molecules to form large, heavy negative oxygen ions,
the electric force on the ions is the same as on the original electrons, but the momentum-transfer
cross section is increased, so that, in fact, the implicit assumption of Equation 5 is satisfied.
A frequently discussed topic is the vector direction of the plasma body force during the AC
cycle. The formulation given in Equation 5 indicates that throughout the AC cycle, the body
force is always oriented in the direction from the bare electrode toward the covered electrode.
This could be the result of the assumptions made in the formulation of the model, in particular
the quasi-steady assumption that the timescale of the electron and ion movement is much smaller
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

than the AC period.


This is certainly true for the electrons. The question is whether it is true for the ions. The
charged-particle simulations of Font and colleagues (Font 2004, Font & Morgan 2005) that in-
cluded nitrogen and oxygen reactions followed the propagation of a single streamer from the bare
electrode to the dielectric surface and back. They suggested that during the forward discharge
(when the electrons are pushed away from the bare electrode and the positive ions are pulled
back toward the bare electrode), the net momentum is not zero, but favors the ions so that there
is a net momentum toward the bare electrode. In the back discharge (when the electric field is
reversed), their simulations have a resultant force that is away from the bare electrode, toward the
dielectric [as in Orlov and colleagues’ (Orlov 2006, Orlov et al. 2006) model]. The magnitude of
body force in the back discharge was significantly larger than in the forward discharge so that the
cycle-averaged vector from the simulation was overwhelmingly toward the dielectric, as observed
in the time-averaged experiments (Thomas et al. 2009).
Font and colleagues’ scenario might be categorized as PUSH-pull, in which the upper- or low-
ercase words signify the relative magnitude. Orlov and colleagues’ model would indicate PUSH-
push. What is the experimental evidence?
Forte et al. (2006) performed time-resolved laser-Doppler-velocimetry measurements of the
flow induced by an SDBD plasma actuator similar to that shown in Figure 1, in a quiescent neutral
flow. They were able to capture the streamwise and wall-normal velocity components within a
period of the AC input. The laser-Doppler-velocimetry measurements indicated that during the
AC cycle, the u component oscillated between a large positive u and a small but positive u, and was
never negative. This result supports a PUSH-push scenario. Kim et al. (2007) arrived at the same
conclusion by observing flows with time-resolved particle-image velocimetry. Recent experiments
by Enloe and colleagues (Enloe et al. 2009, Porter et al. 2007) indicate (using an entirely different
experimental method) a PUSH-push scenario when the net effect of the actuator is concerned,
with the magnitude of the effect of the plasma alone being comparable on both the negative- and
positive-going half-cycles, bringing these experimental results into closer agreement with those
predicted by Orlov and colleagues’ (Orlov et al. 2006, Orlov 2006) model. Given that the same
effect is shown by drastically different means, it seems well established that both half-cycles of the
discharge add momentum to the flow in the same direction.

2.3. Optimization
The insight that comes from developing a better understanding of the physics behind the SDBD
plasma actuator can suggest approaches to optimize its performance. The following sections pro-
vide some examples.

www.annualreviews.org • DBD Plasma Actuators 519


ANRV400-FL42-21 ARI 13 November 2009 15:2

2.3.1. AC waveform. The observations that the ionization occurs as long as the difference be-
tween the instantaneous AC potential and the charge buildup on the dielectric exceeds a threshold
value suggest that there are AC waveforms that are optimal. For example, a square wave is least
optimum, a sine wave is better, and a triangle wave is better yet. (These are waveforms measured
at the actuator input. In practice, low-level signal inputs to high-voltage amplifiers can experience
significant filtering and shape alteration.) This can be extended further by considering a waveform
that emphasizes the time given to the PUSH and minimizes the time of the push. Such a waveform
would be a sawtooth with the greatest duty cycle possible allocated to the polarity of the dV/dt
such that electrons are emitted from the exposed electrode and deposited on the dielectric surface.
Enloe et al. (2004a) verified this experimentally.
An alternate waveform receiving some attention is very narrow (nanosecond) pulses. These
are sometimes used in combination with sinusoidal waveforms or small DC components (Opaits
et al. 2009). The addition of a DC generally leads to a so-called sliding discharge. A new plasma
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

actuator based on that effect is discussed in Section 4.1.

2.3.2. Geometry. Forte et al. (2006) experimentally examined the effect of the amount of overlap
between the bare and covered electrodes in an asymmetric arrangement similar to that shown in
Figure 1. They defined the gap spacing, g, to be positive when there was a nonoverlapping distance
between the edges. Interpreting their results, and normalizing the gap by the width of the covered
electrode Lce , which must be a factor, they found that there was little effect on the maximum
induced velocities for 0 ≤ g/Lc e ≤ 2. For larger (positive) gaps or overlap, the effectiveness of the
plasma actuator dropped off rapidly.
Because the effect of the plasma actuator on the neutral flow is through a body force, we expect
that the effect of multiple actuators is linearly additive. This was first confirmed by Post (2004)
and later by Forte et al. (2006) and Thomas et al. (2009).

2.3.3. Thick dielectrics. There are two important properties of the dielectric material: the break-
down voltage per thickness (volts per millimeter) and the dielectric coefficient, ε. The minimum
thickness of the dielectric needs to be sufficient to not break down at the applied voltage, although
for some materials, this can be accomplished with only a fraction of a millimeter of thickness (0.05
mm typical for Kapton). Recent evidence (Thomas et al. 2009) shows a benefit in using thicker
dielectric layers made of materials that have lower dielectric coefficients. The general objective
is to lower the capacitance of the actuator. The capacitance is proportional to ε/h, where h is
the thickness of the dielectric. The power loss through the dielectric is proportional to fAC ε/h.
Therefore, lowering the capacitance (ε/h) lowers the power loss through the dielectric, which is
otherwise manifest in heating, and allows higher voltages to be reached. Because the body force
3.5
is proportional to VAC , the motivation is to be able to operate at higher voltages.

2.3.4. AC frequency. Orlov and colleagues’ (Orlov 2006, Orlov et al. 2006) model indicates
that there is an optimum AC frequency to maximize the body force that depends on the actuator
capacitance. Thomas et al. (2009) investigated this using a 6.35-mm-thick glass dielectric actuator.
The results are shown in Figure 11. For this actuator design, the 8-kHz AC frequency has the
lowest maximum thrust, and the 1-kHz frequency has the highest thrust. Again at fixed power
(I × V), if the current (I) is too large, the applied voltage (V) will decrease and the body force
(thrust) will decrease. The visible indication of the increased current is the appearance of the
bright filaments in the plasma. The voltage at which these first occur varies linearly with fAC as
expected based on the relation for power loss through the dielectric.

520 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

3.0 a b 1 kHz

2.5
2 kHz

2.0
Thrust (gm)

4 kHz
1.5

1.0 8 kHz
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

0.5 8 kHz
4 kHz
2 kHz
0 1 kHz

5 10 15 20 25 30
VAC (kV–rms)

Figure 11
(a) Induced thrust from a single–dielectric barrier discharge plasma actuator for a 6.35-mm-thick glass
dielectric for different AC frequencies of the applied voltage. (b) Corresponding images of plasma for each
frequency at maximum thrust, showing the ionized air produced by the actuator. The lowest image
represents the uniform ionization that occurs within the voltage range at which the thrust is proportional to
voltage to the 3.5 power. The other four images correspond to the maximum thrust point at the respective
AC frequencies shown in the left panel. Figure taken from Thomas et al. 2009.

3. EXAMPLE APPLICATION: EXPERIMENT AND SIMULATION


A visual experimental demonstration of SDBD plasma actuators was the suppression of the von
Kármán vortex street behind a circular cylinder given by Thomas et al. (2008a,b). As an exam-
ple, Figure 12 shows two particle-image-velocimetry images taken with the plasma actuators
off and on at ReD = 33,000. Four plasma actuators were located on the downstream half of
the cylinder at the 90◦ , 135◦ , 225◦ , and 270◦ positions, as measured in the clockwise direction
from the stagnation line on the upstream side of the cylinder. The actuators were the asymmetric
electrode design shown in Figure 1. The dielectric was the 6.4-mm-thick glass wall of the cylin-
der. The plasma actuators kept the flow attached on the lee side of the cylinder, resulting in a
merged jet of fluid on the wake centerline that modified the mean flow and suppressed the vortex
shedding.
Mertz & Corke (2009) performed a simulation of Thomas et al.’s experiment and modeled the
plasma actuator in the manner of Orlov (2006), as described in Section 2.2. Figure 13a shows
the AC cycle-averaged body-force vector field computed for the simulation. This illustrates the
actuator design, which, on average, oriented the mean body-force vector in the mean flow direction
and toward the wall of the cylinder. The effect of the body force on the flow around the cylinder
was simulated using FLUENT with the body-force vector array supplied through a user-defined
input. Streamlines from the simulation with the plasma actuator effect off and on are shown in
Figure 13b. These corroborate well with the experiment.

www.annualreviews.org • DBD Plasma Actuators 521


ANRV400-FL42-21 ARI 13 November 2009 15:2

a b

Figure 12
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

Particle-image-velocimetry images of the flow behind a circular cylinder at ReD = 33,000 with plasma
actuators on the lee side of the cylinder off (a) and on (b). Figure taken from Thomas et al. 2008.

4. FUTURE EXPECTATIONS

4.1. Sliding Discharge


A relatively new design for plasma actuators is based on a sliding discharge. This concept was
first developed for laser-pumping applications (Arad et al. 1987). A number of researchers have
adapted it to atmospheric-pressure plasmas (e.g., Louste et al. 2005, Thomas et al. 2008a, Zouzou
et al. 2007). The concept is to utilize the AC DBD to weakly ionize the air, and then to superpose a
DC potential that establishes a corona discharge between spatially separated electrodes. The DC
component induces the sliding discharge. The advantages of this concept are that large plasma
sheets can be produced and the plasma is stable with no glow-to-arc transition, except when the
DC component is above the DC breakdown limit for the air.

a b
0.2
Y (m)

–0.2

–0.2 0 0.2 0.4 0.6 0.8


X (m)
0.3
0.2
0.1
Y (m)

0
–0.1
–0.2
–0.3
–0.4 –0.2 0 0.2 0.4 0.6 0.8
X (m)

Figure 13
Body-force vectors for a plasma actuator (a) and flow streamlines (b) from flow simulations that include the plasma actuator body force
off and on. Figure taken from Mertz & Corke 2009.

522 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

a Plasma b Ground Ground


DC voltage
Catode Anode
Induced air flow

7.6 cm
Dielectric AC voltage

Additional 0 kV DC –50 kV DC
Excitation insulator
electrode SDBD discharge Sliding–DBD discharge
(DC off ) (DC on)
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org

Figure 14
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

(a) Schematic of triode plasma actuator and (b) photographs of plasma for dielectric barrier discharge (DBD) operation and with sliding
discharge operation. Figure taken from Thomas et al. 2008a.

Thomas et al.’s approach is unique from the others cited above. It is referred to as a triode
plasma actuator because the electrode arrangement and the use of the dielectric material are
similar to that of a discrete triode amplifier. A schematic of the triode plasma actuator is shown
in Figure 14a. When the triode actuator is operating only with the AC input on, it functions as
an SDBD device. Figure 14b shows that the plasma generated in this case is only visible near the
edges of the two exposed electrodes. However, the addition of the DC caused the visible plasma
to completely fill the space between the electrodes.
Figure 15 shows a comparison between the thrust generated by the triode plasma actuator
when operated in DBD and sliding discharge modes. The thrust measurement was performed
in an identical manner to that shown in Figure 6. When the DC was off, the thrust followed
the power-law growth that is characteristic of SDBD plasma actuators. There was also a clear
threshold voltage below which thrust was too low to be measured. When the DC was on, there
was a thrust produced even at zero AC level. The thrust in this case then varied approximately
linearly with the AC level. Obviously, the thrust was significantly larger with the sliding discharge.
Further optimization is forthcoming, but these results suggest great potential for this approach.

4.2. Plasma Sensor


In addition to flow control, a new AC plasma sensor for velocity measurements has recently been
developed by Matlis & Corke (2005). Although it was originally intended for high–Mach number,
high-enthalpy flows, it is quite well suited for low-speed flows, or applications in which harsh
conditions make more conventional flow sensors unusable. A recent example of its use includes
the detection of traveling stall cells in a transonic compressor stage (Matlis et al. 2008). This
technology offers the unique opportunity for combined plasma actuators and sensors that could
be beneficial for closed-loop feedback control in a single element.

5. SUMMARY
There is an ever growing number of applications of SDBD plasma actuators that have appeared in
the literature. A partial list of these includes exciting boundary-layer instabilities on a sharp cone
at Mach 3.5 (Corke et al. 2001, Kosinov et al. 1990, Matlis 2004), lift augmentation on a wing

www.annualreviews.org • DBD Plasma Actuators 523


ANRV400-FL42-21 ARI 13 November 2009 15:2

0.35
DC off
DC on (–48 kV)
0.30

0.25

Thrust (N/m)
0.20

0.15

0.10
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

0.05

0
0 20 40 60
VAC (p–p kV)
Figure 15
Comparison between thrust generated by a triode plasma actuator for dielectric barrier discharge and sliding
discharge operation. Figure taken from Thomas et al. 2008a.

section (Corke et al. 2002, 2006; Goeksel & Rechenberg 2004; Goeksel et al. 2006; Nelson et al.
2006; Patel et al. 2006), low-pressure turbine blade separation control (Huang 2005; Huang et al.
2006a, 2006b; List et al. 2003; Rizzetta & Visbal 2007; Suzen et al. 2007; Wall et al. 2007), turbine
tip-clearance flow control (Douville et al. 2006, Morris et al. 2005, Van Ness et al. 2006), bluff-
body flow control (Asghar et al. 2006, Do et al. 2007, Thomas et al. 2006), turbulent boundary-
layer control (Balcer et al. 2006, Hultgren & Ashpis 2003, Porter et al. 2007, Wilkinson 2003),
unsteady vortex generation and control (Nelson et al. 2007, Visbal & Gaitonde 2006), and air-
foil leading-edge separation control (Corke et al. 2004; Post 2004; Post & Corke 2003, 2004).
New applications continue to appear as more investigators gain experience in using these flow
actuators.
Our understanding of the SDBD physics inherent to the plasma actuators has led to the
development of quantitative models that have shown remarkable agreement with experiments,
which point to improved designs and operation. The recent optimization of the actuators produced
by better choices of thick dielectric materials and AC input frequencies and waveforms has led to
order-of-magnitude improvements in their performance compared with earlier designs. Sliding
discharge approaches offer the potential for further significant improvement. All these are opening
the scope of application conditions for these flow-control devices.

DISCLOSURE STATEMENT
T.C.C. is a partial holder and C.L.E is a co-inventor of U.S. Patent No. 7,380,756, “Single
dielectric barrier aerodynamic plasma actuator,” and S.P.W. is a co-inventor of U.S. Patent
No. 6,200,539, “Paraelectric gas flow accelerator.”

524 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

LITERATURE CITED
Arad B, Gazit Y, Ludmirsky A. 1987. A sliding discharge device for producing cylindrical shock waves. J. Phys.
D 77:360–67
Asghar A, Jumper EJ, Corke TC. 2006. On the use of Reynolds number as the scaling parameter for the performance
of plasma actuator in a weakly compressible flow. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 44th, Reno,
AIAA Pap. No. 2006-21
Balcer BE, Franke ME, Rivir RB. 2006. Effects of plasma induced velocity on boundary layer flow. Presented at
AIAA Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap. No. 2006-875
BenGadri R, Rabehi A, Massines F, Segur P. 1994. Numerical modelling of atmospheric pressure low-
frequency glow discharge between insulated elecrodes. Proc. 12th ESCAMPIG, Netherlands, August 23–26,
pp. 228–29
Boeuf JP, Pitchford LC. 2005. Electrohydrodynamic force and aerodynamic flow acceleration in surface
dielectric barrier discharge. J. Appl. Phys. 97:103307
Boeuf JP, Lagmich Y, Unfer T, Callegari T, Pitchford L. 2007. Electrohydrodynamic force in dielectric
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

barrier discharge plasma actuators. J. Phys. D 40:652–62


Cavalieri D. 1995. On the experimental design for instability analysis on a cone at Mach 3.5 and 6.0 using a corona
discharge perturbation method. PhD thesis, Illinois Inst. Technol.
Corke T, Cavalieri D, Matlis E. 2001. Boundary layer instability on a sharp cone at Mach 3.5 with controlled
input. AIAA J. 40:1015–18
Corke TC, He C, Patel M. 2004. Plasma flaps and slats: an application of weakly-ionized plasma actuators. Presented
at AIAA Flow Control Conf., 2nd, Portland, AIAA Pap. No. 2004-2127
Corke TC, Jumper EJ, Post ML, Orlov D, McLaughlin TE. 2002. Application of weakly-ionized plasmas as
wing flow-control devices. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 40th, Reno, AIAA Pap. No. 2002-
0350
Corke TC, Matlis E. 2000. Phased plasma arrays for unsteady flow control. Presented at Fluids 2000 Conf. Exhibit,
Denver, AIAA Pap. No. 2000-2323
Corke TC, Mertz B, Patel MP. 2006. Plasma flow control optimized airfoil. Presented at AIAA Aerosp. Sci. Meet.
Exhibit, 44th, Reno, AIAA Pap. No. 2006-1208
Davidson G, O’Neil R. 1964. Optical radiation from nitrogen and air at high pressure excited by energetic
electrons. J. Chem. Phys. 41:3946–49
Decomps Ph, Massines F, Mayoux C. 1994. Electrical and optical diagnosis of an atmospheric pressure glow
discharge. Acta Phys. Univ. Com. 1994:47–53
Do H, Kim W, Mungal M, Capelli M. 2007. Bluff body flow control using surface dielectric barrier discharges.
Presented at AIAA Aerosp. Sci. Meet. Exhibit, 45th, Reno, AIAA Pap. No. 2007-0939
Douville T, Stephens J, Corke T, Morris S. 2006. Turbine blade tip leakage flow control by partial squealer tip
and plasma actuators. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap. No. 2006-
20
Eliasson B, Kogelschatz U. 1991. Nonequilibrium volume plasma chemical processing. IEEE Trans. Plasma
Sci. 19:1063–77
Enloe C, Font G, McLaughlin T, Orlov D. 2008. Surface potential and longitudinal electric field measurements
in the aerodynamic plasma actuator. AIAA J. 46:2730–40
Enloe C, McHarg M, Font G, McLaughlin T. 2009. Plasma-induced force and self-induced drag in the dielectric
barrier discharge aerodynamic plasma actuator. Presented at AIAA Aerosp. Sci. Meet., 47th, Orlando, AIAA
Pap. No. 2009-1622
Enloe C, McLaughlin T, Font G, Baughn J. 2006. Parameterization of temporal structure in the single-
dielectric-barrier aerodynamic plasma actuator. AIAA J. 44:1127–36
Enloe CL, McLaughlin TE, VanDyken RD, Kachner KD, Jumper EJ, Corke TC. 2004a. Mechanisms and
responses of a single-dielectric barrier plasma actuator: plasma morphology. AIAA J. 42:589–94
Enloe CL, McLaughlin TE, VanDyken RD, Kachner KD, Jumper EJ, et al. 2004b. Mechanisms and responses
of a single-dielectric barrier plasma actuator: geometric effects. AIAA J. 42:595–604
Falkenstein Z, Coogan J. 1997. Microdischarge behaviour in the silent discharge of nitrogen-oxygen and
water-air mixtures. J. Phys. D 30:817–25

www.annualreviews.org • DBD Plasma Actuators 525


ANRV400-FL42-21 ARI 13 November 2009 15:2

Font GI. 2004. Boundary layer control with atmospheric plasma discharges. Presented at AIAA/ASME/SAE/ASEE
Joint Propul. Conf. Exhibit, 40th, Fort Lauderdale, AIAA Pap. No. 2004-3574
Font GI, Morgan WL. 2005. Plasma discharges in atmospheric pressure oxygen for boundary layer separation control.
Presented at AIAA Fluid Dyn. Conf. Exhibit, 35th, Toronto, AIAA Pap. No. 2005-4632
Forte M, Jolibois J, Moreau F, Touchard G, Cazalens M. 2006. Optimization of a dielectric barrier discharge
actuator by stationary and non-stationary measurements of the induced flow velocity: application to flow control.
Presented at AIAA Flow Control Conf., 3rd, San Francisco, AIAA Pap. No. 2006-2863
Gibalov V, Pietsch G. 2000a. The development of dielectric barrier discharges in gas gaps and surfaces.
J. Phys. D 33:2618–36
Goeksel B, Rechenberg I. 2004. Active separation flow control experiments in weakly ionized air. Presented at
EUROMECH Eur. Turbul. Conf., 10th, Barcelona
Goeksel B, Rechenberg I, Greenblatt D, Paschereit C. 2006. Steady and unsteady plasma wall jets for separation and
circulation control. Presented at AIAA Flow Control Conf., 3rd, San Francisco, AIAA Pap. No. 2006-3686
Golubovskii YuB, Maiorov VA, Behnke J, Behnke JF. 2002. Influence of interaction between charged particles
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

and dielectric surface over a homogeneous barrier discharge in nitrogen. J. Phys. D 35:751–61
Hoskinson A, Oksuz L, Hershkowitz N. 2008. Microdischarge propogation and expansion in a surface dielec-
tric barrier discharge. Appl. Phys. Lett. 93:221501
Huang J. 2005. Documentation and control of flow separation on a linear cascade of Pak-B blades using plasma actuators.
PhD thesis, Univ. Notre Dame
Huang J, Corke TC, Thomas FO. 2006a. Plasma actuators for separation control of low pressure turbine
blades. AIAA J. 44:51–57
Huang J, Corke TC, Thomas FO. 2006b. Unsteady plasma actuators for separation control of low-pressure
turbine blades. AIAA J. 44:1477–87
Hultgren LS, Ashpis DE. 2003. Demonstration of separation delay with glow-discharge plasma actuators. Presented
at AIAA Aerosp. Sci. Meet. Exhibit, 41st, Reno, AIAA Pap. No. 2003-1025
Kanzawa S, Kogoma M, Kanazawa S, Moriwaki T, Okazaki SJ. 1990. The improvement of atmospheric-
pressure glow plasma method and the deposition of organic films. J. Phys. D. 23:374–77
Kanazawa S, Kogoma M, Moriwaki T, Okazaki SJ. 1988. Stable glow plasma at atmospheric pressure. J. Phys.
D 21:838–40
Kanazawa S, Kogoma M, Okazaki SJ, Moriwaki T. 1989. Glow plasma treatment at atmospheric pressure for
surface modification and film deposition. Nucl. Inst. Methods Phys. Res. B 37:842–45
Kim W, Do H, Mungal M, Cappelli M. 2007. On the role of oxygen in dielectric barrier discharge actuation
of aerodynamic flows. Appl. Phys. Lett. 91:181501
Kline M, Miller N, Walhout M. 2001. Time-resolved imaging of spatiotemporal patterns in a one-dimensional
dielectric-barrier discharge system. Phys. Rev. E 64:026402
Kogelschatz U. 2002. Filamentary, patterned and diffuse barrier discharges. IEEE Trans. Plasma Sci. 30:1400–8
Kogelschatz U, Eliasson B, Egli W. 1997. Dielectric-barrier discharges: principles and applications. J. Phys.
IV (France) 7:C4-47–66
Kogoma M, Okazaki SJ. 1994. Raising of ozone formation efficiency in a homogeneous glow discharge plasma
at atmospheric pressure. J. Phys. D 27:1985–87
Kosinov A, Maslov A, Shevelkov S. 1990. Experiments on the stability of supersonic laminar boundary layers.
J. Fluid Mech. 219:621–33
Kozlov KV, Wagner H-E, Brandenburg R, Michel P. 2001. Spatio-temporally resolved spectroscopic diag-
nostics of the barrier discharge in air at atmosperic pressure. J. Phys. D 34:3164–76
Kunhardt E. 2000. Generation of large volume atmospheric pressure nonequilibrium plasmas. IEEE Trans.
Plasma Sci. 28:189–99
Kunhardt EE. 1980. Electrical breakdown of gases: the pre-breakdown stage. IEEE Trans. Plasma Sci. PS-
8:130–38
Kunhardt EE, Luessen L. 1981. Electrical Breakdown and Discharges. New York: Plenum
Landau LD, Lifshitz EM. 1984. Electrodynamics of Continuous Media. Oxford: Pergamon
Langmuir I. 1928. Oscillations in ionized gases. Proc. Natl. Acad. Sci. USA 14:627–37

526 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

Likhanskii AV, Shneider MN, Macheret SO, Miles RB. 2006. Modeling of interaction between weakly ionized
near-surface plasmas and gas flow. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap.
No. 2006-1204
List J, Byerley AR, McLaughlin TE, VanDyken RD. 2003. Using a plasma actuator to control laminar separation
on a linear cascade turbine blade. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 41st, Reno, AIAA Pap. No.
2003-1026
LLewellyn-Jones F. 1966. The Glow Discharge and an Introduction to Plasma Physics. New York: Methuen
Louste C, Artana G, Moreau E, Touchard G. 2005. Sliding discharge in air at atmospheric pressure: electrical
properties. J. Electrost. 63:615–20
Madani M, Bogaerts A, Gijbels R, Vangeneugden D. 2003. Modelling of a dielectric barrier glow discharge at
atmospheric pressure in nitrogen. In Int. Conf. Phenom. Ionized Gases, ed. J Meichsner, D Loffhagen, HE
Wagner. Greifswald: ICPIG
Massines F, Ghadri RB, Decomps Ph, Rabehi A, Segur P, Mayoux C. 1996. Atmospheric pressure dielectric
controlled glow discharges: diagnostics and modelling. Proc. Int. Conf. Phenom. Ionized Gases, ICPIG, 22nd,
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

Hoboken, New Jersey, 363:306–15


Massines F, Rabehi A, Decomps P, Gadri RB, Segur P, Mayoux C. 1998. Experimental and theoretical study
of a glow discharge at atmospheric pressure controlled by dielectric barrier. J. Appl. Phys. 83:2950–57
Matlis E, Corke T. 2005. AC plasma anemometer for hypersonic Mach number experiments. Presented at AIAA
Aerosp. Sci. Meet. Exhibit, 43rd, Reno, AIAA Pap. No. 2005-0952
Matlis E, Corke T, Cameron J, Morris S. 2008. A.C. plasma anemometer for axial compressor stall warning.
Presented at 12th Int. Symp. Transp. Phenom. Dyn. Rotat. Mach., February 17–22, Honolulu
Matlis EH. 2004. Controlled experiments on instabilities and transition to turbulence on a sharp cone at Mach 3.5.
PhD thesis, Univ. Notre Dame
Meek JM, Craggs JD, eds. 1978. Electrical Breakdown of Gases. Chichester, UK: Wiley
Mertz B, Corke T. 2009. Time-dependent dielectric barrier discharge plasma actuator modeling. Presented at AIAA
Aerosp. Sci. Meet., 47th, Orlando, AIAA Pap. No. 2009-1083
Morris SC, Corke TC, VanNess D, Stephens J, Douville T. 2005. Tip clearance control using plasma actuators.
Presented at AIAA Aerosp. Sci. Meet. Exhibit, Reno, AIAA Pap. No. 2005-0782
Nasser E. 1971. Fundamentals of Gaseous Ionization and Plasma Electronics. New York: Wiley-Interscience
Naude N, Cambronne J-P, Gherardi N, Massines F. 2004. Electrical model of an atmospheric pressure
Townsend-like discharge (APTD). Eur. Phys. J. Appl. Phys. 29:173–80
Nelson CC, Cain AB, Patel MP, Corke TC. 2006. Simulation of plasma actuators using the wind-US code.
Presented at AIAA Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap. No. 2006-634
Nelson R, Corke T, Patel M, Ng T. 2007. Modification of the flow structure over a UAV wing for roll control.
Presented at AIAA Aerosp. Sci. Meet. Exhibit, 45th, Reno, AIAA Pap. No. 2007-0884
Okazaki SJ, Kogoma M, Uehara M, Kimura Y. 1993. Appearance of stable glow discharge in air, argon oxygen,
and nitrogen at atmospheric pressure using 50 hz source. J. Phys. D 26:889–92
Opaits D, Zaidi S, Schneider M, Miles R, Likhanskii A, Macheret S. 2009. Improving thrust by suppressing charge
build-up in pulsed DBD plasma actuators. Presented at AIAA Aerosp. Sci. Meet., 47th, Orlando, AIAA Pap.
No. 2009-487
Orlov DM. 2006. Modelling and simulation of single dielectric barrier discharge plasma actuators. PhD thesis, Univ.
Notre Dame
Orlov D, Corke T. 2005. Numerical simulation of aerodynamic plasma actuator effects. Presented at AIAA Aerosp.
Sci. Meet. Exhibit, 43rd, Reno, AIAA Paper 2005-1083
Orlov D, Corke T, Haddad O. 2003. DNS modeling of plasma array flow actuators. Am. Phys. Soc. Div. Fluid
Dyn., 48:JA.004 (Abstract)
Orlov D, Corke T, Patel M. 2006. Electric circuit model for aerodynamic plasma actuator. Presented at AIAA
Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap. No. 2006-1206
Pai ST, Guo XM, Zhou TD. 1996. Closed form analytic solution describing glow discharge plasma. Phys.
Plasmas 3:3842–52
Patel MP, Sowle ZH, Corke TC, He C. 2006. Autonomous sensing and control of wing stall using a smart plasma
slat. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap. No. 2006-1207

www.annualreviews.org • DBD Plasma Actuators 527


ANRV400-FL42-21 ARI 13 November 2009 15:2

Paulus M, Stals L, Rude U, Rauschenbach B. 1999. Two-dimensional simulation of plasma-based ion implan-
tation. J. Appl. Phys. 85:761–66
Porter C, McLaughlin T, Enloe L, Font G. 2007. Boundary layer control using DBD plasma actuator. Presented
at AIAA Aerosp. Sci. Meet. Exhibit, 45th, Reno, AIAA Pap. No. 2007-0786
Post ML. 2001. Phased plasma actuators for unsteady flow control. Master’s thesis, Univ. Notre Dame
Post ML. 2004. Plasma actuators for separation control on stationary and unstationary airfoils. PhD thesis, Univ.
Notre Dame
Post ML, Corke TC. 2003. Separation control on high angle of attack airfoil using plasma actuator. Presented at
AIAA Aerosp. Sci. Meet. Exhibit, 41st, Reno, AIAA Pap. No. 2003-1024
Post ML, Corke TC. 2004. Separation control using plasma actuators: stationary and oscillatory airfoils. Presented
at AIAA Aerosp. Sci. Meet. Exhibit, 42nd, Reno, AIAA Pap. No. 2004-0841
Rabehi A, BenGadri R, Segur P, Massines F, Decomps Ph. 1994. Numerical modelling of high pressure glow
discharges controlled by dielectric barrier. Proc. Conf. Electr. Insul. Dielectr. Phenom., Arlington, TX, October
23–26, pp. 840–45. New York: IEEE
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

Raizer YP. 1991. Gas Discharge Physics. Berlin: Springer-Verlag


Rizzetta D, Visbal M. 2007. Numerical investigation of plasma-based flow control for a transitional highly-loaded
low-pressure turbine. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 45th, Reno, AIAA Pap. No. 2007-
0938
Roth JR. 1995. Industrial Plasma Engineering. Philadelphia, PA: Inst. Phys.
Roth JR, Dai X. 2006. Optimization of the aerodynamic plasma actuator as an EHD electrical device. Presented at
AIAA Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap. No. 2006-1203
Roth JR, Sherman D, Wilkinson S. 1998. Boundary layer flow control with one atmosphere uniform glow discharge
surface plasma. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 36th, Reno, AIAA Pap. No. 1998-0328
Roth JR, Sherman DM, Wilkinson SP. 2000. Electrohydrodynamic flow control with a glow-discharge surface
plasma. AIAA J. 38:1166–72
Shyy W, Jayaraman B, Andersson A. 2002. Modeling of glow discharge-induced fluid dynamics. J. Appl. Phys.
92:6434–43
Singh K, Roy S. 2008. Force approximation for a plasma actuator operating in atmospheric air. J. App. Phys.
103:013305
Suzen Y, Huang G, Ashpis D. 2007. Numerical simulations of flow separation control in low-pressure turbines using
plasma actuators. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 45th, Reno, AIAA Pap. No. 2007-0937
Suzen YB, Huang PG. 2006. Simulation of flow separation control using plasma actuators. Presented at AIAA
Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap. No. 2006-877
Suzen YB, Huang PG, Jacob JD, Ashpis DE. 2005. Numerical simulations of plasma based flow control applications.
Presented at AIAA Fluid Dyn. Conf. Exhibit, 35th, Toronto, AIAA Pap. No. 2005-4633
Thomas F, Corke T, Iqbal M, Kozlov A, Shatzman D. 2009. Optimization of SDBD plasma actuators for
active aerodynamic flow control. AIAA J. In press
Thomas F, Corke T, Wang M. 2008a. Experimental aircraft noise control using dielectric barrier discharge
plasma actuators: benchmark experiments and LES simulations. NASA Progr. Rep. NNX07AO09A
Thomas FO, Kozlov A, Corke TC. 2006. Plasma actuators for bluff body flow control. Presented at AIAA Flow
Control Conf., 3rd, San Francisco, AIAA Pap. No. 2006-2845
Thomas F, Kozlov A, Corke T. 2008b. Plasma actuators for bluff body flow control. AIAA J. 46:1921–31
Trunec D, Brablec A, Stastny F. 1998. Experimental study of atmospheric pressure glow discharge. Contrib.
Plasma Phys. 38:435–45
Van Ness DK, Corke TC, Morris SC. 2006. Turbine tip clearance flow control using plasma actuators. Presented
at AIAA Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap. No. 2006-21
Vidmar RJ, Stalder KR. 2003. Air chemistry and power to generate and sustain plasma: plasma lifetime calculations.
Presented at AIAA Aerosp. Sci. Meet. Exhibit, 41st, Reno, AIAA Pap. No. 2003-1189
Visbal MR, Gaitonde DV. 2006. Control of vortical flows using simulated plasma actuators. Presented at AIAA
Aerosp. Sci. Meet. Exhibit, 44th, Reno, AIAA Pap. No. 2006-505
Voikov V, Corke T, Haddad O. 2004. Numerical simulation of flow control over airfoils using plasma actuators.
Am. Phys. Soc. Div. Fluid Dyn. 49:FG.008 (Abstract)

528 Corke · ·
Enloe Wilkinson
ANRV400-FL42-21 ARI 13 November 2009 15:2

Wall JD, Boxx IC, Rivir RB, Franke ME. 2007. Effects of pulsed DC discharge plasma actuators in a separated
LPT boundary layer. Presented at AIAA Aerosp. Sci. Meet. Exhibit, 45th, Reno, AIAA Pap. No. 2007-
0942
Wilkinson SP. 2003. Investigation of an oscillating surface plasma for turbulent drag reduction. Presented at AIAA
Aerosp. Sci. Meet. Exhibit, 41st, Reno, AIAA Pap. No. 2003-1023
Yokoyama T, Kogoma M, Moriwaki T, Okazaki SJ. 1990. The mechanism of the stabilization of glow plasma
at atmospheric pressure. J. Phys. D 23:1125–28
Zouzou N, Takashima K, Moreau E, Mizuno A, Touchard G. 2007. Sliding discharge study in axisymmetric
configuration. In 28th ICPIG, Prague, Czech Republic, ed. J Schmidt, M Simek, S Pekarek, V Prukner, Art.
No. 1007-10. London: Int. Union Pure Appl. Phys.
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

www.annualreviews.org • DBD Plasma Actuators 529


AR400-FM ARI 13 November 2009 15:33

Annual Review of

Contents Fluid Mechanics

Volume 42, 2010

Singular Perturbation Theory: A Viscous Flow out of Göttingen


Robert E. O’Malley Jr. p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org
by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

Dynamics of Winds and Currents Coupled to Surface Waves


Peter P. Sullivan and James C. McWilliams p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p19
Fluvial Sedimentary Patterns
G. Seminara p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p43
Shear Bands in Matter with Granularity
Peter Schall and Martin van Hecke p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p67
Slip on Superhydrophobic Surfaces
Jonathan P. Rothstein p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p89
Turbulent Dispersed Multiphase Flow
S. Balachandar and John K. Eaton p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 111
Turbidity Currents and Their Deposits
Eckart Meiburg and Ben Kneller p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 135
Measurement of the Velocity Gradient Tensor in Turbulent Flows
James M. Wallace and Petar V. Vukoslavčević p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 157
Friction Drag Reduction of External Flows with Bubble and
Gas Injection
Steven L. Ceccio p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 183
Wave–Vortex Interactions in Fluids and Superfluids
Oliver Bühler p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 205
Laminar, Transitional, and Turbulent Flows in Rotor-Stator Cavities
Brian Launder, Sébastien Poncet, and Eric Serre p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 229
Scale-Dependent Models for Atmospheric Flows
Rupert Klein p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 249
Spike-Type Compressor Stall Inception, Detection, and Control
C.S. Tan, I. Day, S. Morris, and A. Wadia p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 275

vii
AR400-FM ARI 13 November 2009 15:33

Airflow and Particle Transport in the Human Respiratory System


C. Kleinstreuer and Z. Zhang p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 301
Small-Scale Properties of Turbulent Rayleigh-Bénard Convection
Detlef Lohse and Ke-Qing Xia p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 335
Fluid Dynamics of Urban Atmospheres in Complex Terrain
H.J.S. Fernando p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 365
Turbulent Plumes in Nature
Andrew W. Woods p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 391
Fluid Mechanics of Microrheology
Annu. Rev. Fluid Mech. 2010.42:505-529. Downloaded from www.annualreviews.org

Todd M. Squires and Thomas G. Mason p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 413


by NORTH CAROLINA STATE UNIVERSITY on 09/05/12. For personal use only.

Lattice-Boltzmann Method for Complex Flows


Cyrus K. Aidun and Jonathan R. Clausen p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 439
Wavelet Methods in Computational Fluid Dynamics
Kai Schneider and Oleg V. Vasilyev p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 473
Dielectric Barrier Discharge Plasma Actuators for Flow Control
Thomas C. Corke, C. Lon Enloe, and Stephen P. Wilkinson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 505
Applications of Holography in Fluid Mechanics and Particle Dynamics
Joseph Katz and Jian Sheng p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 531
Recent Advances in Micro-Particle Image Velocimetry
Steven T. Wereley and Carl D. Meinhart p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 557

Indexes

Cumulative Index of Contributing Authors, Volumes 1–42 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 577


Cumulative Index of Chapter Titles, Volumes 1–42 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 585

Errata

An online log of corrections to Annual Review of Fluid Mechanics articles may be found
at http://fluid.annualreviews.org/errata.shtml

viii Contents

You might also like