Вы находитесь на странице: 1из 33

Home

Search

Collections

Journals

About

Contact us

My IOPscience

Airflow control by non-thermal plasma actuators

This article has been downloaded from IOPscience. Please scroll down to see the full text article.
2007 J. Phys. D: Appl. Phys. 40 605
(http://iopscience.iop.org/0022-3727/40/3/S01)
View the table of contents for this issue, or go to the journal homepage for more

Download details:
IP Address: 128.173.127.127
The article was downloaded on 12/02/2012 at 01:34

Please note that terms and conditions apply.

INSTITUTE OF PHYSICS PUBLISHING

JOURNAL OF PHYSICS D: APPLIED PHYSICS

J. Phys. D: Appl. Phys. 40 (2007) 605636

doi:10.1088/0022-3727/40/3/S01

Airflow control by non-thermal plasma


actuators
Eric Moreau
Laboratoire dEtudes Aerodynamiques, Groupe Electrofluidodynamique, CNRS 6609,
University of Poitiers, Bd Curie, 86962 Futuroscope, France
E-mail: eric.moreau@lea.univ-poitiers.fr

Received 24 July 2006, in final form 23 October 2006


Published 19 January 2007
Online at stacks.iop.org/JPhysD/40/605
Abstract
Active flow control is a topic in full expansion due to associated industrial
applications of huge importance, particularly for aeronautics. Among all
flow control methods, such as the use of mechanical flaps, wall synthetic jets
or MEMS, plasma-based devices are very promising. The main advantages
of such systems are their robustness, simplicity, low power consumption and
ability for real-time control at high frequency. This paper is a review of the
worldwide works on this topic, from its origin to the present. It is divided
into two main parts. The first one is dedicated to the recent knowledge
concerning the electric wind induced by surface non-thermal plasma
actuators, acting in air at atmospheric pressure. Typically, it can reach
8 m s1 at a distance of 0.5 mm from the wall. In the second part, works
concerning active airflow control by these plasma actuators are presented.
Very efficient results have been obtained for low-velocity subsonic airflows
(typically U  30 m s1 and Reynolds number of a few 105 ), and
promising results at higher velocities indicate that plasma actuators could be
used in aeronautics.
(Some figures in this article are in colour only in the electronic version)

1. Introduction
Active airflow control consists of manipulating a flow to affect
a desired change [1]. For example, efficient flow control
systems could modify the laminarturbulent transition inside
the boundary layer, to prevent or to induce separation, to reduce
the drag and to enhance the lift of airfoils, to stabilize or to
mix airflow in order to avoid unsteadiness which generates
unwanted vibrations, noise and energy losses. This is of large
technological importance for industries where internal and
external airflows occur, and more specifically for aeronautics.
In order to manipulate a free airflow, three main
phenomena may be modified: the laminar-to-turbulent
transition, the separation and the turbulence. Delaying
laminar-to-turbulent transition of a boundary layer has a lot
of advantages. For instance, skin-friction drag of a laminar
boundary layer may be in certain conditions one order of
magnitude lower than turbulent drag. For an aircraft, reduced
drag means reduced fuel cost, longer range and higher speed
[1]. To delay this transition, wall suction devices and MEMS
may be effective. Flow separation may be illustrated by the
example of a plane wing [2]. The maximum lift and stall
0022-3727/07/030605+32$30.00

2007 IOP Publishing Ltd

characteristics of a wing affect many performance aspects of


aircraft, including take-off and landing distance for instance.
The lift at a given angle of attack can be increased by increasing
camber, but the maximum achievable lift is limited by the
ability of a flow to follow the curvature of the airfoil. When it
cannot, the flow separates. One solution to prevent separation
is to use a leading edge slat, trailing edge flaps or wall synthetic
jets. The last aerodynamic phenomenon is turbulence. An
increase in turbulence can lead to better flow mixing, in the
case of mixing layers for example. A decrease in turbulence
can play a fundamental role in aerodynamic noise reduction.
For a few years, the topic of active flow control has been
growing constantly. For instance, in 2002 the well-known
American Institute of Aeronautics and Astronautics (AIAA)
held a conference especially dedicated to this subject called
Flow Control Conference. The third and last one was held
at San Francisco in June 2006. The book of Gad-El-Hak [1],
intended for engineers and researchers, is a very clear review
of flow control.
Among all the active methods, a new and original
technology using non-thermal surface plasmas is in full
expansion. Indeed, although mechanical devices may be

Printed in the UK

605

E Moreau

Figure 1. 2D visualization of a manipulated airflow along a flat


plate.

effective, they have some drawbacks. In particular, they are


complicated, add weight, have volume and are sources of noise
and vibration. Moreover, they are composed of mechanical
parts that wear away and that may break down. Consequently,
plasma actuators are now in full expansion because they do not
exhibit all these drawbacks. These plasma actuators consist of
using the discharge-induced electric wind within the boundary
layer to modify its properties and then to actively manipulate
the airflow. In most of the cases, the actuator is composed of
at least two electrodes flush-mounted at the wall of the profile
between which a high voltage is applied, resulting in a cold
plasma sheet. In the case of a dc discharge for example, the
actuator is extremely simple: it consists of two low-diameter
wires flush-mounted on the surface between which a dc high
voltage is applied. In ambient air, a corona is formed around
the lowest diameter electrode (usually the positive one) and an
electric wind is created, tangentially to the wall. Under certain
conditions, the plasma may extend to the second electrode.
The aim of using electric wind is in most cases to accelerate
the airflow tangentially and very close to the wall in order
to modify the airflow profile within the boundary layer. As
an example, figure 1 shows a manipulated airflow along a flat
plate. In the absence of discharge, the smoke wire is horizontal,
as the streamlines. When the discharge is established, the
electric wind induces a depression at the anode, resulting in
a deviation of the smoke wire and the airflow is accelerated
inside the plasma area.
The main advantage of this process is that it directly
converts electric energy into kinetic energy without involving
moving mechanical parts. Thus it may be considered to
be a very simple MEMS. Secondly, its response time is
very short and enables a real-time control at high frequency.
Its disadvantage is the low efficiency of energy conversion.
Surface discharges may also modify the gas properties at the
wall, such as density or viscosity. However it seems that this
effect is negligible in the case of velocities below 30 m s1 .
Indeed, all the authors consider only the effect of the electric
wind, excepted for high velocities. This point will be discussed
at the end of this paper.
Before 2000, very few works were published on this
subject. In fact, preliminary research was started in the 1950s
and a few patents were developed, in Europe [3] and the USA
[4, 5]. However, the first scientific papers were not published
before the 1968 by Velkoff and Ketchman [6] and by Yabe
et al in 1978 [7]. In the 1980s, a few papers were published
( [8, 9] for instance) but the topic really emerged in the middle
of the 1990s. During that period, several research groups
worked on airflow control by dc surface corona discharges,
experimentally [10, 11] and theoretically [1215].
606

At the end of the 1990s, two groups started to work more


continuously on this subject. The first group included people
from the University of Poitiers (France) and the University of
Buenos Aires (Argentina). At this period, this group focused
on the study of dc coronas for airflow applications. They
published a dozen papers between 1999 and 2002 ([1628]
for instance). The second group was the group directed by
J R Roth, from the University of Tennessee, who worked with
the NASA Langley Research Center, USA. In 1992, J R Roths
group perfected and developed a new kind of surface plasma,
primarily dedicated to decontamination. Realizing that this
discharge could induce a secondary airflow of several m s1 ,
they proposed to use it for airflow control by 1994. In 1998 they
published their first results [29, 30]. This new discharge was a
TM
surface dielectric barrier discharge (DBD), called OAUGDP
and was protected by a US patent since 1995 [31]. It is clear
that this surface plasma has considerably influenced research
on airflow control by plasmas because the simplicity of its
use allowed many researchers in aerodynamics to work on
this subject, without necessarily being a specialist in plasma
generation.
Therefore, one can say that airflow control by a plasma
actuator was really born in 2000. Several papers were written
for the general audience, such as in Air and Cosmos in
France [32,33]. As an example, if one considers only the works
presented at the AIAA meetings, three papers were published
in 2000 while about fifteen papers were published in 2003 and
as many in 2004. At that time, several major groups were
known for their works on this topic: University of Tennessee
with NASA Langley (USA), University of Poitiers (France)
with University of Buenos Aires (Argentina), University of
Notre-Dame with US Air Force Academy (USA), University
of Moscow (Russia). On the one hand, since 2005, the subject
has been growing considerably, and about 30 groups now work
on plasmas for aerodynamic applications in Europe and the
USA, and a few others worldwide. As a result, about 100150
papers have been published to date. On the other hand, there
is one particularity: in most cases US groups use actuators
based on surface DBDs, following Roths technique, whereas
others works sometimes use the DBD actuator or sometimes
the corona one.
In this paper, the goal is to present a review covering the
worldwide development of plasma actuators, from their origin
to the present. It is divided into two main parts.
The first part deals with the electrical and mechanical
characterization of plasma actuators in the absence of a free air
stream. Indeed, although many researchers study the effects
of plasma actuators on airflow, only a few study the dischargeinduced electric wind in the absence of free airflow. However,
that part seems crucial to be able to optimize and to well know
the mechanical effects of such actuators in order to use them
efficiently for flow control. Consequently, in this part, the goal
is to describe briefly the physics and electrical parameters of
atmospheric plasmas, and then to focus more particularly on
the mechanical effects of the discharge, such as electric wind,
induced body force and induced kinetic power.
The second part deals with the application of plasma
actuators for airflow control. A review of the worldwide works
is presented. Regarding the large number of publications,
plasma actuators have shown their good efficiency for airflow

Airflow control by non-thermal plasma actuators

control at velocities up to 30 m s1 , and some significant results


have been obtained up to 110 m s1 and more. A great number
of well-known aerodynamic applications have been studied.
One can mention the case of airflow around cylinders, flat
plates and airfoils, and the case of free jets and mixing layers.
However, it is not possible to report precisely all these works
in the present paper. Most papers will be referenced, but one
will focus more precisely on typical examples, covering the
most-known aerodynamic cases and industrial applications.

HV point

Ionization zone

Ion drift zone

2. Plasma actuators
As previously indicated, this part deals with the secondary
airflow, usually called electric wind or ionic wind, induced
by atmospheric pressure non-thermal plasma discharges
established in air. Indeed, although it seems crucial to know
well the electro-mechanical effects of plasma actuators in order
to use them efficiently in flow control, a few researchers did
this work. Here, one is going to focus on these works.
This part is divided into four sections. First, a brief
introduction concerning atmospheric cold plasmas deals with
electric wind. The next two sections concern the two mostused discharge actuators, i.e. the surface corona discharge
actuator and the single dielectric barrier discharge actuator,
respectively. Although the DBD-based devices are now the
most used, the corona-based devices are presented first because
they were historically the first ones. Then the last section
deals with other types of plasma actuators. Indeed, a few
authors have proposed other non-thermal discharge actuators,
by either modifying originally the electrode geometry or the
HV excitation, or by using another type of discharge.
2.1. Atmospheric cold plasmas
2.1.1. Discharge mechanisms. For a few decades, nonthermal atmospheric pressure plasmas have been studied for
numerous industrial applications such as ozone generation,
pollutant removal and surface treatment. Several papers and
books give a complete review of this type of discharge [3438].
These non-thermal plasmas may be produced by a variety of
electrical discharges and are very low energy cost because they
have the particularity that the majority of the electrical energy
primarily goes into the production of energetic electrons,
instead of heating the surrounding gas.
Briefly, the formation of such a discharge is based
on the Townsend mechanism, or electron avalanche which
corresponds to the multiplication of some primary electrons in
cascade ionization. Let us consider the simple case of a dc high
voltage applied between two plane electrodes in atmospheric
pressure air. In the gap, electrons are usually formed by
photo ionization. Under the electric field, these electrons are
accelerated towards the anode and ionize the gas by collisions
with neutral molecules such as A+e A+ +2e where A is a
neutral particle and A+ a positive ion. An avalanche develops
because the multiplication of electrons proceeds along their
drift from the cathode to the anode. A discharge current is
then created. Different current behaviours may be obtained
when the high voltage increases. The voltagecurrent curve
usually allows one to determine the discharge regime. More
details may be found in [3538].

Figure 2. Schematic of a point-to-plate corona.

The discharge plasmas used for airflow control are usually


atmospheric pressure corona discharges and dielectric barrier
discharges. Their characteristics are typically as follows: high
voltage of a few kV to several tens of kV with dc or ac
excitation, with a frequency from 50 Hz to 50 kHz, electrical
current from a few A to a few mA. In these conditions, the
density of charged species is between 109 and 1013 cm3 and
the electron temperature is a few eV. More accurate values will
be given in the next two parts.
2.1.2. Dc corona discharge. Corona discharge is a weakly
luminous discharge, which usually appears at atmospheric
pressure near sharp points or thin wires, where the electric field
is sufficiently large. Corona may be considered as a Townsend
discharge or a negative glow discharge, depending on field and
potential distribution [36].
Figure 2 shows a schematic of the point-to-plane corona,
inducing a strong electric field, ionization and luminosity
around the point where the high voltage is applied. Several
authors gave an expression of this electric field ( [39, 40]
for instance). The mechanism of sustaining the continuous
ionization near the high voltage electrode depends on its
polarity. If the high electric field is located at the cathode,
it is a negative corona. If the strong electric field is near the
anode, it is a positive corona.
In the case of a positive corona, it is assumed that electrons
are produced by photo ionization in the electrode gap. These
electrons are accelerated towards the anode, resulting in an
avalanche. An ionization zone is created around the point,
where the number of positive ions and electrons is equal. This
corresponds to a plasma zone of a few tenths of a millimetre.
Then ions are repulsed towards the cathode by Coulombian
forces, and constitute the drift region, where one can assume
that there is no recombination because the electric field is
not sufficiently large. In the case of the negative corona,
Goldman and Sigmond [41] assume that positive ions created
by detachment in the ionization zone go rapidly towards the
cathode. Then negative ions created by attachment drift
towards the grounded plane electrode.
In the case of positive corona, one can assume that
the discharge current is composed of two components: an
alternative streamer current due to the streamer propagation,
and a continuous unipolar current due to the ion drift that
are collected by the grounded plane. These streamers extend
between the electrodes. They have a typical frequency of
607

E Moreau

10 kHz, a diameter of 100 m and a displacement velocity


of a few 105 m s1 . This value exceeds by a factor of 10 the
typical electron drift velocity in an avalanche. In the plasma
near the point, the density of charged species rapidly decreases
with distance from about 1013 to 109 cm3 [41]. In the case of
a negative corona, there are Trichel pulses at a frequency that
depends on the applied high voltage. The electrical power of
the continuous coronas is very low because when the voltage
and current increase above a given threshold, there is a coronato-arc transition.
When the applied high voltage is not dc (ac excitation),
the discharge mechanisms are similar when the frequency is
small and when the residual charges have time to be collected
between two successive half-cycles. Then the current is
composed of three components: a capacitive current due to
the gas in the electrode gap, a synchrone current that is in
phase with the applied voltage and a pulsed current due to
the streamer pulses during the positive half period and Trichel
pulses during the negative one. An increase in corona voltage
and power without a spark becomes possible by using a pulse
HV.
The corona discharge generates lots of charged species.
More details may be found in [36] and the large number of
papers referenced in this paper.
2.1.3. Electric wind. As explained by Robinson in his
famous paper [42], the phenomenon variously known as the
electric wind, corona wind and electric aura refers to the
movement of gas induced by the repulsion of ions from the
vicinity of a high voltage electrode. This phenomenon was
reported for the first time in 1709 by Hauksbee and the first
explanation was given by Faraday in 1838. A history of the
electric wind is given in [43]. Many current works are still
dealing with electric wind.
Indeed, the electric wind is due to the collisions between
the ions that drift and the neutral particles in the electrode gap
region. However, although the electron velocity is much higher
than the ion one, one can neglect the role of electrons because
their mass is very low compared with the ion one.
The first expression of the electric wind velocity was given
in 1961 by Robinson [42]:

i
,
(1)
vG = k

where k is a constant that depends mainly on the electrode


geometry, i the time-averaged discharge current, the gas
density and the ion mobility. This shows that the electric
wind velocity is proportional to the square root of the current.
One must distinguish the gas velocity and the ion velocity that
may reach a few thousands of m s1 and which is given by:
vi = E,

(2)

where E is the electric field. From 1970 to 2000, the group


of Goldman (Supelec, Paris, France) worked especially on
coronas, and the induced electric wind. In 1993, the expression
of Robinson was completed by Sigmond and Lagstadt [44]:

id
(3)
vG =
AG
608

with d the electrode gap and AG the discharge cross-section.


In practice, the plasma researchers do not agree on the
phenomena that induce the electric wind. A part of them
considers that it is only due to ion drift while the other part
assumes that it is due to the streamer propagation. In fact, and
this point have been recently confirmed, it seems that electric
wind is due to both phenomena [45].
Several studies showed that the maximum achievable
velocity is about 10 m s1 in the case of positive coronas ([46
48] for instance).
2.2. Surface corona discharge actuator
In this section, we are going to focus on plasma actuators based
on corona discharge. In the next one, we will see the dielectric
barrier discharge-based actuator.
2.2.1. History. The first works on airflow control by electrical
discharges dealt with dc coronas. For example, Velkof et al
[6, 49] demonstrated that the transition point on a flat plate
could be affected by the application of an electric field. In
this application, the plasma actuator consisted in four HV
wire electrodes placed above the wall surface of a flat plate
(figure 3(a)). There was no grounded electrode. Bushnel
[8] and Malik et al [9] reported an electric wind of several
m s1 contributing to drag reduction. In 1992, Soetomo [10]
experimentally observed a drag reduction effect induced by ac
and dc corona discharges along a flat plate in the case of flow
velocities up to 2 m s1 . In this case, the corona discharge was
established between two razor blades flush-mounted on the
wall of a glass flat plate (figure 3(e)). In 1997, Noger et al [11]
used a point-to-plane configuration in order to manipulate an
airflow around a cylinder. More recently, new geometries
have been perfected. All these geometrical configurations are
summarized in figure 3. However, only one group has been
working continuously (from 1998 until now) on the surfacecorona-based-actuator, and characterizing its electrical and
mechanical properties in the absence of a free air stream. That
group is composed of people from the University of Poitiers
(France) and the University of Buenos Aires (Argentina). In
this section, we will focus on their works.
2.2.2. Electrical properties. Here, we are going to consider
the geometry presented in figure 3(g). The plasma actuator
consists of two wire electrodes placed inside a groove at the
wall surface. The electrode diameter is nearly 1 mm. In order
to induce a stronger electric field at the anode compared with
the cathode, the anode diameter is lower than the cathode one.
One can then assume that the anode plays the role of the point
and the cathode corresponds to the plane, in a point-to-plane
configuration. For instance, the anode diameter may be equal
to 0.6 mm when the cathode one is 2 mm. The electrode gap
here is equal to 40 mm. To establish the discharge, a negative
HV of 10 kV is applied at the cathode instead of ground, and
the anode HV is progressively increased.
In these conditions, above the onset voltage, the discharge
current increases with the applied potential difference.
Figure 4 shows a typical example of the time-averaged
discharge current per unit length I (current i divided by the
electrode length in mA m1 , and called current density) as a

Airflow control by non-thermal plasma actuators


HV

HV

(a)

(b)

HV

HV

(d)

(c)

HV

HV

(f)

(e)

Y
X

HV
(g)

Figure 3. Different electrode geometrical configurations found in the literature as dc plasma actuators: (a) volume multiple wire [6], (b)
volume wire-to-plate in plane configuration [7], (c) volume wire-to-plate in cylindrical configuration [50], (d) surface wire-to-plate or
point-to-plate in cylindrical configuration [11, 21], (e) surface plate-to-plate [10], (f) surface wire-to-plate [19], (g) surface wire-to-wire [20].

Current density I (mA/m)

2.5
2.0

Experiments
Fitting

1.5
1.0
0.5
0.0

Electric field E (kV/cm)


Figure 4. Current density I versus reduced electric field E in the
case of wire-to-wire dc surface discharge (the gap is 40 mm and the
dielectric is Plexiglas).

function of the reduced electric field E (potential difference


V divided by the electrode gap in kV cm1 ). Usually, in
point-to-plane configuration, i.e. for volume coronas, the V i
characteristic may be fitted by this expression:
i = CV (V V0 ),

(4)

where i is the time-averaged discharge current, V the potential


difference, V0 the onset voltage and C a constant depending
on the electrode gap (typically between 0.1 and 1 A (kV)1
in air [47]). In figure 4, the thick line corresponds to the best
fit obtained with equation (4), when the symbols represent the

behaviour of the surface corona experimental measurements.


This shows that the surface corona and the volume one are
strongly different, because from 7 kV cm1 , the surface corona
current increases suddenly. This important difference is
certainly due to the gassolid interface, where the corona acts.
Indeed, in this case, the discharge is located just above the
dielectric wall, where the surrounding conditions are particular.
In the case of surface coronas, five corona discharge
regimes may be observed when the voltage between both
electrodes is increased. Briefly, above the corona-starting
voltage V0 , the first regime is the spot regime. The discharge is
concentrated within some visible spots on the smaller diameter
wire. The current density I is smaller than 0.2 mA m1 and the
electric wind is negligible. As the electric field is increased, a
thin sheet of blue ionized air between both electrodes may be
observed. This mode is called the streamer discharge. In this
regime, the current density values vary from 0.2 to 0.5 mA m1
and the electric power consumption Pelec is about 50 mW per
square centimetre of plasma sheet. For higher voltage/gap
ratios (current density from 0.5 to several mA m1 and 50 <
Pelec < 250 mW cm2 ), the glow discharge is obtained. In
this regime, there is no thin sheet of blue ionized air between
both electrodes but only a set of adjacent luminescent spots
around both electrodes. This is a typical corona. Compared
with the streamer discharge, the two main advantages of the
glow discharge are its stability and the fact that high current
values may be reached. If the voltage/gap ratio increases more,
the entire current is concentrated within a few filaments; this
609

E Moreau

1000

Current (A)

800
600
400
200
0
0.000

0.025

(a)

0.050

0.075

0.100

0.075

0.100

Time (ms)

Figure 5. Photography of the dc surface discharge, (a) in streamer


regime with one filament and (b) in glow regime.

Current (A)

1000
800
600
400
200

the discharge is more stable (fewer glow-to-arc transitions)


when the anode diameter is very small compared with the
cathode one [51];
temperature has no effect on the discharge up to 60 C,
whereas the pressure plays a fundamental role [52];
the discharge is less stable when the relative air humidity
increases [27];
in the case of a discharge established in the presence of
a free air stream, the airflow modifies considerably the
physical mechanisms, and limits significantly the glowto-arc transition [53];
610

0
0.000

0.025

(b)

0.050

Time (ms)

Figure 6. Discharge current versus time, for a time-averaged


current of 0.45 (a) and 1.5 mA m1 (b).

E = 7.5 kV/cm
E = 8.25 kV/cm

1.6

Current (mA/m)

is the filamentary regime. Thereafter, sparks may appear and


the discharge may become very difficult to control. Figure 5
shows an example of the surface discharge in streamer regime
and in glow regime.
For airflow control applications, the typical electrical
values are as follows: E 8 kV cm1 , I 1 mA m1 and
Pelec 80 mW cm2 of surface discharge.
Figure 6 presents the discharge current versus time, in
the case of the streamer regime, for 150 and 500 A. The
electrode length here is equal to 33 cm and the current densities
are, respectively, 0.45 and 1.5 mA m1 . This shows that
the current is composed of two components: a continuous
component due to ion drift and an alternative component due
to streamer propagation. It is the reason for this regime being
called the streamer regime. In the glow regime, the alternative
component is negligible: this means that there is no streamer in
this regime, and that the discharge behaves like a low-pressure
glow.
In several works [27, 5155], the electrical properties of
this surface discharge as a function of different atmospheric
parameters (such as ambient air humidity, pressure, different
atmospheric parameters (such as ambient air humidity,
pressure, temperature, etc) and geometric parameters (such
as the electrode position and gap, their shape, etc) have been
studied. They demonstrate that the surface discharge is very
sensitive to these parameters. More details may be found
in [56]. Briefly, some tendencies may be given as follows:

1.2

0.8

Glow

0.4
0

10

Streamer

15

20

25

30

Velocity U0 (m/s)
Figure 7. Current versus airstream velocity for 2 voltage/gap ratio
values.

the discharge electrical properties depend highly on the


dielectric wall surface [54].
As an example, figure 7 presents the current density versus
free air stream velocity, in the case of a surface discharge
established along a flat plate. In the absence of a free air stream,
the current is equal to about 0.4 mA m1 and 0.9 mA m1 , for
E = 7.5 kV cm1 and E = 8.25 kV cm1 , respectively. When
the airflow velocity increases (here in the same direction as the
electric wind), the current density increases too, nearly linearly
with U . Furthermore, from about 15 m s1 , the discharge
regime changes. The velocity favours the streamer regime in
spite of the glow one.
2.2.3. Induced electric wind. In this section we focus on
the mechanical properties of such discharges. First, we are

Airflow control by non-thermal plasma actuators

VGmax (m/s)

4
3
2
1
0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4

Current (mA/m)

going to present the electric wind velocity profile. A velocity


profile is built by measuring the discharge-induced velocity
from the profile wall (y = 0) to y = 20 mm, at a given
x location (for the coordinate system, see figure 3(g)). The
velocity is measured with the help of a glass capillary (as a Pitot
tube) connected to a micro-manometer. The glass capillary,
which is displaced along the x axis with a step-by-step motor,
measures the total pressure. The static pressure is detected
with a typical Pitot tube located above the flat plate. Then the
electric wind velocity is deduced from the dynamic pressure,
corresponding to the total pressure minus the static pressure.
Here, the capillary internal radius is 0.2 mm and the electrodes
are 0.6 mm and 2 mm in diameter for the anode and the cathode,
respectively [55]. Figure 8 presents an example of velocity
profiles measured at x = 10 mm (10 mm upstream from the
cathode), for several discharge current values. The two main
features of the velocity distribution are as follows:
the electric wind velocity increases with the current. For
a time-averaged current of 0.3 mA m1 , the maximum
velocity is about 2.1 m s1 whereas it reaches almost
3.5 m s1 for a current equal to 1.2 mA m1 .
the maximum velocity is usually obtained at a distance of
about 1 mm from the wall (y = 1 mm).
Other authors have measured the electric wind induced
by a dc corona surface discharge and have found the same
behaviour. Practically, the maximum velocity ever measured to
date was 5 m s1 . For example, figure 9 presents the maximum
velocity vGmax versus discharge current density I : it shows that
vGmax is nearly proportional to I 1/2 . Moreover, measurements
along the x axis demonstrated that the discharge-inducedmomentum does not come from the region located upstream
of the anode but from the region placed above the anode,
at x = 40 mm. This means that the discharge induces a
depression at the anode.
Figure 10 presents the instantaneous velocity versus time
(during 2 s) at x = 10 mm and y = 1 mm, for two
current values (0.3 and 1.5 mA m1 ). This shows that the
instantaneous velocity is less stable when the current increases.
At 0.3 mA m1 , the standard deviation (sd) value of the
velocity is 0.07 m s1 whereas it is equal to 0.32 m s1 for a
current of 1.5 mA m1 . However, this effect may be attenuated
by increasing the electrode dissymmetry.

Figure 9. Maximum velocity versus time-averaged discharge


current (dc surface corona), at x = 10 mm.
0.3 mA/m, Umean = 1.3 m/s, sd = 0.07
1.5 mA/m, Umean = 3.0 m/s, sd = 0.32

4.0
3.5

Velocity (m/s)

Figure 8. Velocity profiles at x = 10 mm for time-averaged


current values between 0.3 and 1.2 mA m1 .

3..0
2.5
2.0
1.5
1.0
0.5
0.0
0.0

0.5

1.0

1.5

2.0

Time (s)
Figure 10. Instantaneous velocity versus time at x = 10 mm and
y = 1 mm, for two current values (0.3 and 1.5 mA m1 ).

From the velocity profiles, the induced kinetic power and


then the electro-mechanical efficiency of such actuators may
be computed. In mechanical engineering, the power induced
by blade fans or the power of an airflow inside a duct is given
by
1
(5)
Pmec = VG3 ,
2
where is the air density and VG the airflow velocity. For
corona-induced kinetic power, Robinson [42] assumed that the
kinetic power corresponds to the energy derivative


d 1
1
1 dM 2
(6)
MVG2 =
V = AG VG3 ,
Pmec =
dt 2
2 dt G
2
with M the air mass in motion. Sigmond and Lagstad [44]
established a theoretical model where they considered that
Pmec = SG VG2 /2,

(7)

with SG the averaged gas flow rate through the discharge crosssection AG . More, assuming that the electrical power Pelec
might be expressed as follows:

Pelec = V I = EI dl vi I d/ = vi SG VG
(8)
with vi the ion drift velocity; they demonstrated that the electromechanical efficiency of the corona discharges is given by
=

Pmec
= VG /2vi  1
Pelec

(9)
611

E Moreau
Pmec

0.20

0.6

0.18
0.16

0.4

0.14

0.2

Efficiency (%)

Pmec (mW/cm)

++++++

0.22

Efficiency

0.8

0.12

0.0

+ HV
(a)
++ +

- --

0.10

0.0

0.5

1.0

1.5

2.0

Current (mA/m)

+ HV

Figure 11. Dc surface discharge-induced kinetic power and


electro-mechanical efficiency versus time-averaged current.

(b)

and is then very low. Indeed, a large part of the electrical power
goes into gas heating and a per cent into direct gas motion.
The main problem of both models is that they do not
consider the velocity variations along the y axis. We
then introduced in [57, 58] a method based on the energy
conservation equation, allowing us to take into account the
velocity profile. Under the assumption of a stationary flow,
the mechanical power corresponds to the kinetic energy density
flow rate, which may be expressed by

1
VG3 (y) dy
(10)
Pmec = L
2
0
with L the electrode length. To determine the mechanical
power of surface discharges per unit of electrode length (here
in mW cm1 ) this value is divided by L. The electrical power
per unit length Pelec (in mW/cm) is also divided by L:
Pelec = V i/L = V I.

(11)

Then the electro-mechanical efficiency of the plasma actuator


is the absence of a free air stream and is given by
=

Pmec
.
Pelec

(12)

Figure 11 presents the surface discharge-induced kinetic power


and efficiency versus discharge current. It shows that Pmec I
for I up to 1.5 mA cm1 . Above 1.5 mA cm1 , the discharge
becomes unstable and the electric wind velocity does not
increase any more. However, efficiency decreases nearly
linearly when I increases and is rather low (a few tenths of
a percent). One can notice that these values are smaller than
those obtained with volume dc coronas [45, 58]. On the one
hand, it is not surprising because in the case of a surface
discharge, the skin-friction at the wall reduces significantly the
electric wind velocity. Indeed, it is a heresy to want to create
velocity at the wall surface, where velocity is equal to zero by
definition. However, previous works that will be reported in
the part concerning airflow control, showed that this efficiency
reaches several per cent in the presence of free airflow [53].
However, a few remarks may be made about the dc surface
wire-to-wire discharge. Indeed, to look as much as possible
like the positive point-to-plate corona discharge and taking
into account that positive ions are produced more easily than
612

Figure 12. Schematic of the dc surface electric wind (a) with a thin
anode, (b) with a thin cathode.

negative ions in air, one uses usually a thin anode and a wider
cathode. Then the space charge between both electrodes is
mainly positive, and the electric wind is due to the positive
ion motion (figure 12(a)). If anode and cathode are reversed
(figure 12(b)), then the electric wind is composed of two
opposite components: a positive electric wind, due to the
motion of a positive space charge from the anode to the cathode,
but that is limited because the anode is not thin enough, and a
negative electric wind, favoured because the cathode is thinner
than the anode. In this case, the electric wind is globally
unstable, and directed towards the cathode near the anode, and
vice-versa (figure 12(b)). This point is discussed in [27, 51].
More, the most influential electrical parameter on the
discharge properties is not the potential of each electrode, but
the potential difference between them. For instance, for a 4 cm
electrode gap, a HV of about 32 kV must be applied between
both electrodes, but the discharge is rather similar if the cathode
is grounded and 32 kV are applied at the anode, or if 16 kV
are applied to the cathode and 16 kV at the anode.
2.2.4. Ac high voltage. Until now, the case of the timeaveraged electric wind velocity with a dc high voltage has
been presented. However several studies used ac excitation
[51, 55] and the instantaneous electric wind velocity has been
measured [59]. In the range considered in these papers, there
was no advantage in using ac excitation instead of dc, except
to produce a pulsed electric wind. However, it seems that the
maximum frequency of this pulse electric wind is limited to
about 100 Hz [59]. This point must be clarified because the
authors do not know if this is due to a physical mechanism of
the surface discharge, or if it is due to the power supply used
in this study.
2.2.5. Numerical works. Although a large number of
works deal with modelling volume coronas, and the associated
electric wind, the mechanisms occurring in a surface corona
discharge actuator are so complex that only a few researchers
try to model it.
The first known publication concerning the modelling of a
discharge used for airflow control is certainly the paper of Yabe

Airflow control by non-thermal plasma actuators

2.2.6. Conclusion. Although this dc actuator showed its


possibility to induce an electric wind of several m s1 close to
a wall surface, and it is very easy to produce because it needs
a very simple power supply, one of its main disadvantages is
the glow-to-arc transition that limits the maximum velocity.
Thus its geometrical configuration must be optimized, as has
been done for the DBD actuator. Indeed, it is easily possible
to reach electric wind velocity up to 10 m s1 in a point-to-grid
without arcing, so it might be feasible in the case of a surface
discharge, by replacing, for instance, the wire anode by a set
of very sharp pins.

10
case 1
case 2

9
8

y (mm)

7
6
5
4
3
2

2.3. Surface dielectric barrier discharge actuator

1
0
0.0

0.5

1.0

1.5
2.0
2.5
-1
velocity (m.s )

3.0

3.5

4.0

Figure 13. Velocity profile without airflow 10 mm upstream of the


cathode, for I = 0.24 mA m1 [60].

et al in 1978 [7]. However, it was not a surface discharge but a


volume discharge established in the geometrical configuration
presented in figure 3(b). Consequently, the first publication
concerning the surface discharge modelling is [13] in 1997,
followed by [14] two years later. In these papers, coronainduced drag reduction was numerically computed over a finite
region. The model simulated a corona discharge along a
surface from two parallel wire electrodes of infinite length,
immersed flush on the surface, and oriented perpendicular
to the flow. The analysis was limited to positive ions,
which was considered a space charge. Five coupled partial
differential equations govern the numerical model including
continuity, momentum gas phase conservation of charge,
Poissons equation and conservation of charge at the solid
interface. The effects obtained from this model were more
efficient than experimental measurements. In the same way,
Vilela Mendes and Dente developed a numerical model in
1998 [15]. However, in both cases, the numerical tool did not
allow modelling the electric wind in the absence of discharge,
but the electric force added in the NavierStokes equation.
Using a model rather similar to the one of Colver and
El-Khabiry, Baudoin et al [60] published a numerical work
in 2005. The originality of this paper was that it proposed
a confrontation of different results, issued from modelling
and experimental measurements. Firstly, 2D modelling of
the electric wind was performed. Then the influence of the
electrode geometry was studied. Secondly, the model was
adapted to the case of a dc corona discharge in the presence of
an airflow. For instance, figure 13 presents velocity profiles,
where the electrodes are placed at the surface wall (case 2) and
embedded inside a groove (case 1). The modelled electric wind
velocity profiles were in good agreement with experimental
measurements (see figure 8) but the velocity is slightly overestimated. Moreover, it shows that the maximum velocity is
slightly closer to the wall in case 2.
In addition one can mention the work of Mateo-Velez
et al [61] who worked on an electric wind model using a
kinetic scheme. In their last works, they simulated the unsteady
electric wind versus time and the associated electric forces.

In this section, we are going to focus on plasma actuators based


on dielectric barrier discharges, usually called DBD.
2.3.1. Volume DBD. As indicated before, the corona-tospark transition may be prevented by employing a very short
pulsed high voltage (a few ns to s). Another approach is
based on the use of a dielectric barrier in the discharge gap,
which stops the electric current and prevents spark formation.
Because of the dielectric barrier, dc HV cannot be used. Then
DBDs are usually excited by ac or pulsed HV, with frequencies
between 50 Hz and 500 kHz [37]. DBDs have a large number
of industrial applications because they can operate in air at
atmospheric pressure, and do not need sophisticated pulsed
power supplies; a DBD may be excited by a sine HV, delivered
by a simple transformer.
Although DBDs that operate at atmospheric pressure have
a long history (middle of the 19th century), it seems that the
first DBD at atmospheric pressure was established by von
Engle et al in 1933 [62]. They reported the operation of a dc
normal glow discharge in air at one atmosphere. However their
procedure required ignition of the discharge under vacuum,
followed by a gradual increase in pressure to one atmosphere.
It also required cooling of the cathode in order to remove the
glow-to-arc transition. More, this discharge was not so stable.
Since the middle of the 1980s, many works concerning the
physics of DBD have been published. One can mention the
fundamental and experimental works of Yokoyama et al [63],
Massines et al [64, 65] and the reviews of Wagner et al [66],
Fridman et al [37], Kogelschatz [67] and Schutze et al [36].
Briefly, the plasma in the electrode gap is generated
through a succession of microdischarges, randomly distributed
in time and space. These streamers have a width of about
100 m and a duration of a few ns (streamer velocity in the
range 107 108 m s1 ). The charged species density and the
electron temperature are rather near the corona ones.
2.3.2. Surface DBD. In the middle of the 1990s, Roths group
perfected and developed a new atmospheric pressure DBD. It
was a surface DBD, established in air between at least two electrodes placed asymmetrically on each side of a dielectric. They
protected it by a patent in 1994 and called it One Atmosphere
TM
Uniform Glow Discharge Plasma (OAUGDP ). Figure 14
shows the first three configurations tested by Roth et al and
presented for the first time for airflow control in 1998 [29, 30]
at the AIAA conference and 25th ICOPS, and published in a
journal in 2000 [68]. This first paper is very interesting. The
613

E Moreau

Figure 14. Plasma panel design concepts. (a) Symmetric staggered,


(b) asymmetric staggered, (c) symmetric planar lower electrodes.
(From Roth et al [29]; reprinted with permission of the American
Institute of Aeronautics and Astronautics, Inc.)

Figure 16. Schematic side view of a single DBD actuator, (a) area
of plasma, (b) electric wind direction, and (c) photograph (top view)
of a surface DBD established on a glass plate.

y
x
t

wA
Figure 15. OAUGDP-induced velocity in still air. (From Roth
et al [29]; reprinted with permission of the American Institute of
Aeronautics and Astronautics, Inc.)

authors show that the discharge induces a secondary airflow of


several m s1 tangentially to the wall, such that the resulting
force increases with the applied voltage, up to 11 mN, and that
this secondary airflow can modify the free air stream, resulting in a drag modification. For example, figure 15 presents
velocity profiles, measured above the surface discharge.
Then, a large number of geometrical configurations have
been proposed to produce surface DBD actuators. They will be
presented in section 2.4. Here we will only focus on the mostused DBD actuator, which consists of a single actuator (see
figure 16). Typically, this plasma actuator is composed of two
plane electrodes flush-mounted on both sides of a dielectric
plate. One electrode is excited by an ac high voltage (usually a
sine waveform) and the other one is grounded (figure 16(a)). In
these conditions, and above the ignition voltage V0 , a plasma
sheet appears on both sides of the dielectric, resulting in an
electric wind on both sides of the dielectric, as illustrated in
figure 16(b). Figure 16(c) shows a picture of a surface DBD.
The electrical and mechanical characteristics of such plasma
depend strongly on different parameters, such as electrode
width (wA and wB ), electrode gap (g), dielectric thickness (t)
and the nature of the dielectric (figure 17). This point will be
discussed later in this paper.
However, in most cases, these parameters are as follows:
electrode width of a few mm, electrode gap equal to zero or a
few mm and a dielectric in Teflon, kapton, glass, ceramics or
Plexiglas. Its thickness is usually between 0.1 and a few mm.
614

wB

Figure 17. Geometrical electrode configuration of a single DBD


actuator.

To ignite the surface plasma, an ac HV is applied.


Magnitude, frequency and waveform vary with users, but the
most frequent excitations are: magnitude from a few kV to
30 kV and frequency f between 100 Hz and a few tens of kHz.
2.3.3. Electrical properties. Figure 18 presents a typical
behaviour of the discharge current versus time, in the case of
a sine HV [69, 70]. Here, the actuator is constituted of two
aluminium foils flush-mounted on each side of a 5 mm-thick
glass plate. The electrode gap is g = 5 mm, the electrode
length is equal to 20 cm, V = 20 kV and f = 300 Hz.
The frequency is voluntarily low, in order to separate well
the positive and the negative half-cycles. This figure shows
that the discharge current, corresponding to the plasma ignited
on both sides of the dielectric, is composed of short pulses
appearing at the beginning of each inversion of polarity. Each
pulse corresponds to a microdischarge or streamers. Here,
these pulses are limited to about 20 mA because of the power
supply. But these peaks may reach several amperes in certain
cases [71]. Moreover, there is a capacitive component, mainly
due to the dielectric between both electrodes, and the air
gap. Concerning the current pulses due to streamers, they are
positive during the positive voltage half-period and negative
during the negative one, contrary to recent results that show
negative and positive pulses simultaneously [72, 73].
If one wants to prevent plasma formation below the
dielectric, the grounded electrode must be encapsulated in a
dielectric, as illustrated in figure 19. Then there is plasma

Airflow control by non-thermal plasma actuators

20
20

10

10

5
0

-5
-10

-10
-15
-20

Voltage (kV)

Current (mA)

15

-20

Time (ms)
Figure 18. Current versus time, in the case of surface DBD.

Figure 19. Schematic side view of the DBD actuator, when the
grounded electrode is encapsulated.

Figure 21. Schematic illustration of the charge build-up on the


dielectric surface. (From Enloe et al [76]; reprinted with permission
of the American Institute of Aeronautics and Astronautics, Inc.)

20

3000

2
0

0
-2

-10

-4

2000
1000
0
-1000
-2000
-3000

-6
-8

Q (nC)

10

Voltage (kV)

Current (mA)

30 kV
20 kV
10 kV

-20

-30 -20 -10

10

20

30

V (kV)

Time (ms)
Figure 20. Current versus time when the grounded electrode is
encapsulated.

only on the upper side of the dielectric, near the air-exposed


electrode. Figure 20 presents the current that is obtained in
this case. It shows that the plasma is composed of a set of
microdischarges during the positive half-period when it seems
that the plasma is more homogeneous during the negative
one. This demonstrates that the plasma is different during
the negative and the positive half-cycles [70].
In [7476], Enloe et al seem to confirm this result. In
these papers, they did optical measurements that revealed that
the structure of the plasma is different in both space and
time on the negative-going versus the positive-going applied
voltage. More precisely, they concluded that the plasma is
not as it appears to the unaided eye, that is to say a uniform
plasma at the dielectric surface but rather consists of a series
of individual filaments or microdischarges occurring in rapid
sequence. During the forward stroke, microdischarges happen
in rapid succession with a relatively low charge transferred per
event whereas during the backstroke, there is a relatively small
number of more intense microdischarges.

Figure 22. Typical voltagecharge V Q characteristics, for


voltages of 10, 20 and 30 kV.

Moreover, they give a simple physical explanation of


the surface DBD, summarized in figure 21. Figure 21(a)
illustrates the half-cycle for which the exposed electrode
is negative, and emits electrons. Because the discharge
terminates on the dielectric surface, the build-up of surface
charges opposes the applied voltage, and the discharge shuts
itself off unless the magnitude of the applied voltage is
continually increased. When the voltage reverses, they think
that the charge transferred through the plasma is limited to that
deposited on the dielectric surface (figure 21(b)).
It is clear that the dielectric and the solidgas interface
play a fundamental role in the plasma mechanisms. In fact,
when one increases the time-averaged transferred charge
by increasing voltage and/or frequencythe charge buildup is so high that the discharge becomes unstable because
filaments appear between different points on the surface where
space charge has not been relaxed. Moreover, if voltage and
frequency are increased too much, filaments or arcs may occur
across the dielectric, resulting in its destruction. This point has
been underscored by two recent papers [77, 78].
615

40
35
30
25
20
15
10
5
0

Pel = 0.0065 (V- V0 )

Power (W)

Pel (W)

E Moreau

10

15

20

25

30

V (kV)
Figure 23. Example of time-averaged electrical power consumption
versus applied voltage, where V0 is the minimum voltage to ignite
the discharge.

With regard to the electrical power consumption, a few


studies have been made. In using the method described
in [66], Pons et al [70] computed the electrical power as
a function of several parameters. This method consists in
placing a capacitor between the grounded electrode and earth,
and in plotting the V Q curve, as illustrated in figure 22.
The area of this curve corresponds to the energy per period;
then the electrical power is obtained by multiplying this value
with the waveform frequency. Energies here are between 1
and 10 mJ and the transferred charge per period may reach
a few C. For instance, figure 23 presents the electrical
power consumption versus applied voltage, for the actuator
presented at the beginning of this section (5 mm-thick glass
dielectric, electrode length 20 cm, f = 300 Hz). It shows
that the electrical power increases according to a parabolic
function (Pel = K (V V0 )2 ) in these geometrical and
electrical conditions. This has been confirmed by [78] in other
geometrical and electrical configurations. Other authors found
slightly different behaviours when the dielectric is thinner
[75, 76]. As a result, one can assume that the electrical power
consumption of the surface DBD is in the same order of
magnitude as the surface corona. However, the instantaneous
power versus time is completely different. In fact, the electrical
power consumption is nearly constant and continuous in the
case of a dc surface corona, because voltage and current
are continuous. In the case of a DBD, the power versus
time presents a large number of peaks. Figure 24 presents
the instantaneous electrical power, for a time-averaged power
equal to 16 W. It shows that power peaks up to 350 W appear,
whereas the current peaks were strongly limited by the power
supply. Other authors found power peaks attaining 40 kW
[71, 81].
Roth et al [78, 79] studied which part of this electrical
power is dissipated by heating the dielectric. They showed
that this depends highly on the dielectric, the applied voltage
and the waveform frequency.
Concerning the geometrical properties of the surface
DBD, Enloe et al [75] showed that its expansion speed from the
active electrode (air-exposed electrode in figure 19) toward the
dielectric surface, is near 100 m s1 and increased lightly with
the applied voltage. More, its extension always stops at the
downstream edge of the encapsulated electrode. In practice,
616

400
350
300
250
200
150
100
50
0
-50

Time (ms)
Figure 24. Typical instantaneous electrical power consumption
versus time, for a surface DBD.

this extension is limited to about 2 cm when the thickness of


the dielectric is constant [77].
2.3.4. Mechanical effects. Three parameters may illustrate
the mechanical effect induced by such an actuator: electric
wind velocity, mechanical power and electric force. These
three points will be treated in this section.
Concerning the DBD-induced velocity, the group from the
University of Tennessee measured it for the first time in 1998
(see figure 15). Since that date, two groups have been studying,
rather accurately, the mechanical effects of surface DBD:
the group composed of the US Air Force and the University
of Notre-Dame (USA) and the group from the University
of Poitiers (France). At Poitiers, experimental works have
dealt with stationary and non-stationary measurements of
the electric wind induced by such discharges ( [59, 77] for
example). The group from the US Air Force focused mainly
on electric force induced by the plasma. A few other groups
have presented electric wind velocity profile measurements
[71, 81, 83].
In [70], Pons et al measured velocity profiles in the
absence of a free air stream, with the help of a glass Pitot tube,
along the x and y axes (see figure 17), by modifying several
parameters such as the applied voltage, frequency, electrode
gap, dielectric thickness and dielectric nature. They showed
that:
the velocity induced by the grounded electrode (when it
is not encapsulated) is slightly smaller than the velocity
induced by the HV electrode;
as for the surface corona, the momentum comes from the
region placed above the air exposed electrode, meaning
that the discharge induces a depression towards the wall;
the maximum velocity is always reached at the limit of the
plasma extension, usually placed at the downstream edge
of the grounded electrode;
at low frequency (from 300 to 700 Hz), and up to 20 kV,
the maximum velocity which is usually at about 0.5 mm
from the wall, increases linearly with the applied voltage.
With regard to this last point, others authors found that the
maximum velocity was located between 1 and 2 mm above the
wall, certainly because their dielectrics were thinner [78, 84].
For example, figure 25 presents velocity profiles,
measured above an actuator consisting of a 4 mm-thick glass

Airflow control by non-thermal plasma actuators

5.5

Umax (m/s)

Y (mm)

6.0

20kV
18kV
16kV
14kV
12kV
10kV

4.5
4.0
3.5
3.0

0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5

Velocity (m/s)
Figure 25. Typical velocity profiles measured with a glass Pitot
tube, for different voltages.
4.5
4.0

Umax (m/s)

5.0

3.5
3.0
2.5
2.0
1.5
-5

10

15

Electrode Gap (mm)


Figure 26. Evolution of the maximum induced velocity versus
electrode gap.

plate, with an electrode gap and electrode width equal to 5 mm.


Frequency is still 300 Hz here.
More recently, this experimental study has been completed
by Forte et al [77]. The ultimate goal of this work was to
perform a parametric study in order to increase the induced
velocity. It allowed inducing velocities up to 8 m s1 . For
example, figures 26 and 27 present the maximum velocity
(almost always at y = 0.5 mm) as a function of the electrode
gap g and the encapsulated electrode width wB (see figure 17),
respectively. In figure 26, a 2 mm-thick Plexiglas plate is used,
with two 5 mm-wide electrodes. V = 20 kV and f = 700 Hz.
One can remark that the optimum electrode gap is 5 mm,
because when g is bigger, the electric field may fall down and
the space charge cannot continue anymore to move towards the
downstream electrode. For the next measurements, the same
actuator is used, but g = 0 and the encapsulated electrode
width is increased little by little. In the first part of the curve,
the velocity increases with the electrode width, and reaches
a plateau. On the one hand, the wider the electrode is, the
farther the plasma can expand; thus ions are accelerated for a
longer distance, resulting in a velocity increase. On the other
hand, the phenomenon which makes the plasma self-sustaining
is dissipative, and the plasma cannot expand more than about
20 mm. In conclusion, the best configuration is g = 0 and
wB = 20 mm, or g = 5 mm and wB = 15 mm. A similar
tendency had been found previously by Enloe et al [76], but

10

20

30

Grounded Electrode Width (mm)


Figure 27. Evolution of the maximum induced velocity versus
grounded electrode width.

with a very different configuration of electrodes and different


voltages and frequency.
Moreover, three studies demonstrate that the induced
velocity increases with both voltage and frequency, but the
behaviour of the curves are completely different [7678]. On
the one hand, figure 28(a) [77] shows that the maximum
velocity increases asymptotically with the applied voltage. The
same behaviour is observed with an increase in frequency.
In fact, the discharge loses a part of its mechanical effect
when the transferred space charge is too high because the
discharge becomes unstable and filamentary. This has been
confirmed by [78]. On the other hand, the results obtained
by [76] show an exponential behaviour where the maximum
velocity Umax = 0.000166 V 7/2 (figure 28(b)). This
may be explained by the fact that the actuator geometries
and the power supplies used to apply the HV are different.
Moreover, measuring methods are different. In [76] the
velocity is deduced from particle image velocimetry (PIV)
measurements and the actuator is as follows: the electrode
is in 0.025 mm-thick copper foils with 0.15 mm-thick kapton
polyimide as the dielectric. However the maximum cannot
increase indefinitively, as shown figure 28(b).
In order to measure the instantaneous velocity versus
time, Forte et al [77] did laser doppler velocimetry (LDV)
measurements. For example, figure 29(a) presents the nonstationary velocity induced by a surface DBD (V = 18 kV,
f = 700 Hz, 1.5 mm-thick Teflon plate, electrode width
w = 30 mm, and electrode gap g = 0), when the discharge
is ignited at t = 0.5 s. Before t = 0.5 s, the measured
velocity corresponds to a free air stream in the wind tunnel.
This airflow allows dragging away the smoke needed for
LDV measurements. One can remark that this velocity is not
perfectly constant because the wind tunnel is not optimized
to work at such a low velocity. At t = 0.5 s, there is a fluid
acceleration (transient phenomenon), up to about 3.5 m s1 , in
10 ms. This value corresponds to the actuator rise time at this
space point (y = 1 mm, x = 5 mm). If one zooms the velocity
when it reaches the permanent regime, one can see that there
are 35 velocity peaks during 50 ms (figure 29(b)). This means
that the induced velocity follows the HV waveform frequency.
In the next study [77], Forte et al performed the same kind
of measurements, but they were now able to synchronize
the records of the voltage, the current and the velocity (xcomponent and y-component). A typical measurement is
617

E Moreau

7
6

Umax (m/s)

5
4
3
2
1
0
6

(a)

10

12

14

16

18

20

22

24

26

Voltage (kV)
(a)

(b)

Figure 28. Evolution of the maximum induced velocity versus


applied voltage. ((a) From Forte et al [77]; reprinted with
permission of the American Institute of Aeronautics and
Astronautics, Inc. (b) From Enloe et al [76]; reprinted with
permission of the American Institute of Aeronautics and
Astronautics, Inc.)

given in figure 30. One can notice that the discharge does
not behave similarly during the positive and the negative halfcycles. One can see clearly that the negative half-cycle induces
more horizontal velocity (3.6 m s1 ) than the positive one
(2.4 m s1 ). This may be linked to the fact that the negative
half-cycle produces a more uniform discharge than the positive
one, which is composed of current peaks, i.e. microdischarges.
Would this tend to prove that this surface DBD is composed of
two successive discharges: a positive corona and a negative
one? One thing is sure: the mechanical effects of such
discharges are lower when it is a streamer discharge, instead
a more homogeneous discharge. Moreover, the y-component
curve shows negative values during the positive half cycle.
This means that the depression above the actuator is formed
during the positive half-cycle.
From the velocity profiles measured with a Pitot
tube (figure 25 for instance), the mechanical power and
then the electro-mechanical efficiency may be computed
(equations (10) and (12)). It has been demonstrated that the
mechanical power, which is maximum at the right extremity
of the plasma extension, increases linearly with the discharge
current (in this range, because for higher currents, there is an
618

(b)

Figure 29. Velocity versus time. (a) The discharge is established at


t = 0.5 s, (b) zoom of the velocity after t = 0.5 s.

asymptotic behaviour). However, if one compares figure 31


with figure 11, one can notice that DBD efficiency is lower
than the corona one [58,85], even when the grounded electrode
is encapsulated.
Concerning the electric force due to the space charge
displacement inside the plasma, a few researchers measured it.
As indicated previously, Roth [30] was the first to do it. More
recently, Van Dyken et al [86] did a study concerning the net
force produced by a single actuator. In this experimental work,
the influence of several parameters (signal waveform, signal
frequency, electrode geometry, electrical power consumption)
was investigated in order to increase the induced net force.
They conclude that:
the waveform is of great importance in the net force. In
fact, the positive saw tooth waveform produces the greatest
net force for a given input power;
the optimum frequency is 5 kHz (under their conditions);
at 5 kHz, the maximum net force is about 0.9 g for an
electrical power of 55 W, i.e. about 16 mg W1 ;
thicker dielectric configurations are generally able to
produce higher maximum net forces because the dielectric
is able to withstand higher currents.

Airflow control by non-thermal plasma actuators

Voltage
Current

5.0

u
v

4.5

10

20

16

12

Velocity (m/s)

3.0
2.5

2.0
0

1.5

4
0
-4

1.0

-2

0.5

-4

-8

-6

-12

-8

-16

0.0

Voltage (kV)

3.5

Current (mA)

4.0

-0.5
-1.0
-1.5
252.5

253.0

253.5

254.0

254.5

255.0

255.5

-10
256.0

-20

Time (ms)
Figure 30. Synchronized records of voltage, current, horizontal velocity (u) and vertical velocity (v) of the surface DBD-induced velocity,
versus time. (Current is here filtered by a 10 kHz low-pass filter.)
Pmec

0.5

0.05

Efficiency

0.04
-4

Pmec = 4 x 10 x I
0.3

0.03

0.2

0.02

0.1

0.01

0.0

Efficiency (%)

Pmec (mW/cm)

0.4

0.00

0.0

0.5

1.0

1.5

2.0

2.5

Current (mA/m)
Figure 31. DBD surface discharge-induced kinetic power and
electro-mechanical efficiency versus rms discharge current.

More recently, other measurements have been performed


[84, 87]. They showed that
the net force is proportional to the frequency;
the efficiency is typically between 0.15 and 0.20 mN per
electrical watt;
the force induced by a positive pulse is greater than the
one induced by a negative pulse. This last result does not
agree with figure 30. This point will have to be studied
more accurately in the future.
2.3.5. Numerical works There are more and more numerical
works in order to model surface DBD mechanisms, in the case
of the single actuator configuration. However, in the present
paper, one will focus on works concerning the mechanical
effects of surface DBDs.
The first known work is certainly the paper of Shyy
et al [88]. This objective of this work is to estimate the
body force term acted upon by the plasma on the fluid. The
electric field lines are then simplified geometrically, and the

authors consider that the force acts only during one half-cycle,
because they consider that the plasma is active only during the
positive half-period. This means that they neglect the negative
space charge. It is known now that this hypothesis is not
correct. However, this approached electric force is added into
the NavierStokes equation. As a result, in this preliminary
study, the electric force is very overestimated compared with
experimental results.
In [89], surface discharge has been modelled using the
chemistry of formation of different ion and neutral species
of nitrogen and oxygen. The continuity equation governing
densities of electrons, ions and neutral species is solved with
Poissons equation to obtain spatial and temporal profiles of
densities and voltage. The electric body force per unit volume
has been predicted. This showed that the time-averaged
force, predominantly downstream with a transverse component
towards the wall, acts on the plasma in the forward direction.
Recently, Boeuf et al [90, 91] published two papers on
the subject. A fluid model is used to describe the space and
time evolution of the plasma and the force acting on neutral
molecules. The results show that for a ramp or a sine voltage,
the discharge current consists of large amplitude short current
pulses during which a filamentary plasma spreads along the
surface, separated in time by ion drift duration and low current
discharge phases of corona types. The main information is that
they assume that the contribution of the low current phases to
the total force may be predominant because the force acts over a
much larger volume. Figures 32 and 33 present the computed
current during 250 s and the corresponding forces, parallel
and perpendicular to the wall.
Last year, several other papers were presented at the AIAA
conferences at Reno in January 2006 and at San Francisco in
June 2006.
2.3.6. Conclusion. Compared with the dc surface corona
discharge, the main advantage of the ac surface barrier
discharge is that the dielectric between both electrodes prevents
the glow-to-arc transition, resulting in a more stable discharge
619

Current

E Moreau

Figure 32. DBD current. (From Boeuf et al [91]; reprinted with


permission of the American Institute of Aeronautics and
Astronautics, Inc.)

Figure 33. DBD-induced forces. (From Boeuf et al [91]; reprinted


with permission of the American Institute of Aeronautics and
Astronautics, Inc.)

and in a faster electric wind. Its main disadvantage is that the


electrical input power is composed of high peaks, up to 40 kW
under certain conditions. However, this may be significantly
reduced by using inductive filters between the power supply
and the actuator.
2.4. Other discharge actuators
In the previous two sections, we focused on the two most-used
actuator types. In the literature, a few authors have proposed
other non-thermal discharge actuators, by either modifying
originally the electrode geometry or the HV excitation or by
using another type of discharge. We are going to present them
briefly in this section.
2.4.1. DBD-based actuators In 1998, Roth et al proposed
the well-known single DBD actuator. In the same period, they
presented several other geometrical configurations, with various types of HV excitation. For instance, these works may be
found in [35] and they are summarized in [92]. In [92], Roth
classified DBD actuators in two kinds: paraelectric and peristaltic. We have to notice that this classification is very much
debated and both these terms are used only by Roths group.
The paraelectric forces would be due to the plasma
acceleration towards increasing field gradients, while dragging
the neutral gas with it. It is the case for example for the single
DBD actuator and other geometrical configurations given in
[92] (from figures 34(a)(d)). The problem is that the electric
620

wind induced by paraelectric EHD force is limited to about


10 m s1 .
The second EHD body force would be due to peristaltic
flow acceleration, produced by a travelling electrostatic wave
TM
in an AOUGDP surface plasma. Then one uses a polyphase
power supply to excite the plasma at progressive voltage phase
angles, on successive linear electrode strips, as illustrated
by figure 34(f). The main advantage of this method is that
Roth assumes that the induced neutral gas drift velocity in
atmospheric air may reach 100 m s1 . To be honest, it is
necessary to recall that this is a theoretical assumption based
on a certainly too simple model and that this velocity has never
been reached yet. Experimentally, Roth perfected a panel
that was energized with an 8-phase polyphase power supply,
capable of supplying voltages up to 8 kV and frequencies up to
8 kHz. The panel was set up for three complete periods of 8phase excitation, with successive strip electrodes energized at
a phase angle of 45 leading or lagging behind the previous
electrode. He showed that velocities up to 6 m s1 were
achievable. Similar velocity has been measured with a single
actuator. Then this does not demonstrate the best efficiency of
travelling electrostatic waves. However, the main advantage
of this actuator is that by reversing the phase angle of the
electrostatic wave, the induced flow can be reversed too. This
effect is not possible with paraelectric effects alone. To
illustrate this, Roth prepared a well-known video where he
shows the effect of the induced velocity on smoke emitted
from a vertical tube 2.5 cm above the panel (figure 35).
Other configurations similar to the ones presented by
Roths group have been studied in [9395] for example.
Concerning DBD-based actuators, a last and original
configuration may be presented. The plasma actuator design
consists of two annular electrodes, one air-exposed and the
other one embedded, in order to generate a plasma jet actuator.
Indeed, the goal here is not to induce velocity tangentially to the
wall but perpendicularly [96, 97]. A schematic top view of the
annular actuator and cross-stream streamlines of the induced
vortex are given in figures 36(a) and (b), respectively.
2.4.2. Other plasma actuators In this section, three actuators
are going to be presented: from the cooler to the hotter.
The first actuator is based on a sliding discharge [98, 99].
Initially, this type of discharge was developed for laser pumping applications [100102] and was excited by nanosecondwidth pulse HV. In 2004, Louste et al [98] perfected a similar design for the generation of atmospheric pressure plasmas
(figure 37). This design is based on a three-electrode geometry: two air-exposed electrodes (#1 and #2) flush mounted on
the wall surface of a dielectric material and another planar electrode (#3) on the other side. The electrode placed below the
dielectric may be encapsulated in a dielectric. Electrodes #2
and #3 are connected together, and usually grounded, whereas
electrode #1 is excited. On one hand, if electrode #1 is excited
by an ac sine HV, without a dc component, then there is a typical DBD between electrodes #1 and #3 (figure 38(a)). On the
other hand, if one applies simultaneously a sufficient ac component to establish a DBD, plus a dc component sufficient to
establish a typical corona between electrodes #1 and #2, then
a sliding discharge is produced (figure 38(b)). In fact, it seems
that the DBD plays the role of ionizer around electrode #1

Airflow control by non-thermal plasma actuators

Figure 35. Video samples showing the effect of the induced


peristaltic velocity on smoke emitted 2.5 cm above the panel, (a)
with a electrostatic wave from left to right, (b) from right to left.
(Reused with permission from [92] Roth J R 2003 Phys. Plasmas 10
2117, Copyright 2003, American Institute of Physics.)

Figure 34. Electrode geometries used to generate a flat surface layer


TM
of OAUGDP , (a) with electrode embedded in dielectric material,
(b) symmetric electrode of opposite polarity on each side of a
dielectric, (c) active electrode on one side of the panel, with a sheet
electrode on the other side, (d) asymmetric electrode configuration
for paraelectric acceleration, (e) peristaltic flow acceleration by
travelling wave, (f) configuration for combined paraelectric and
peristaltic flow acceleration. (Reused with permission from [92]
Roth J R 2003 Phys. Plasmas 10 2117, Copyright 2003, American
Institute of Physics.)

(with the help of electrode #3), and the dc component induces


a sliding corona discharge between electrodes #1 and #3. The
two main advantages of this discharge are that large plasma
sheets may be produced (one needs about 7 kV per centimetre
between #1 and #2) and that it is very stable with no glowto-arc transition, excepted if the dc component is very high.
The induced electric wind has been measured in [99], but the
mechanical effect was rather similar to a typical DBD. However, best results should be reached when the sliding discharge
actuator will have been geometrically optimized.
The second actuator presented here is the localized arc
filament plasma actuator [103105]. The main objective of
these researches is to produce strong localized perturbations in
high-pressure transonic and supersonic flows. This is achieved
by a very rapid localized heating of the flow by high-current
filaments formed by constriction of an electric discharge at

Figure 36. Annular jet actuator, (a) schematic top view, (b)
cross-stream streamlines of the induced vortex and jet (from [96]).

relatively high pressures. Figure 39 presents the actuator


arrangement in the axisymmetric nozzle extension. In fact,
filaments are produced between two successive electrodes, one
excited and the other one grounded. Typically, the distance
between two adjacent electrodes is a few millimetres. Each pair
of electrodes constitutes a localized plasma actuator powered
by a variable frequency (2100 kHz) high voltage power
supply, which produces an arc filament discharge between
the electrodes. The time-averaged power dissipated in such
discharge is quite low, in the order of 50100 W.
621

E Moreau

Figure 37. Schematic side view of the sliding discharge actuator.

Figure 40. Schematic view of the sparkjet actuator. (From Cybik


et al [106]; reprinted with permission of the American Institute of
Aeronautics and Astronautics, Inc.)

Figure 38. Photograph (top view) of a typical DBD (a), and a


sliding discharge (b). The gap between #1 and #3 is 4 cm.

Figure 39. Schematic view of the localized arc filament plasma


actuator. (From Samimy et al [104]; reprinted with permission of
the American Institute of Aeronautics and Astronautics, Inc.)

The last actuator is a sparkjet actuator [106]. The sparkjet


is a completely solid-state device that consists of a small
chamber with embedded electrodes and a discharge orifice.
High pressure in the chamber is generated by rapidly heating
the gas inside it, using an electrical discharge. The pressure
is relieved by exhausting the heated air through the orifice
(figure 40). A single cycle of sparkjet consists of three distinct
stages: energy deposition, discharge and recovery. Cycles
are repeated to produce a sustainable synthetic jet. Up to
now, no experimental measurements have been performed,
but simulations showed that the chamber temperatures are
typically 3001000 K, resulting in jet velocities from 100 to
400 m s1 . This device should enable us to control high-speed
flows.
622

3. Airflow control
This part is dedicated to a few typical examples of subsonic
airflow control by non-thermal plasma actuators. The aspect
concerning the control of high-speed airflows by thermal
plasmas will not be presented here. A review may be found
in [107] for instance.
Then, this part is divided into four sections. The first
one deals with the manipulation of a flow along a flat plate and
phenomena occurring in boundary layers are more particularly
studied. The second section presents a few results concerning
the control of airflows around cylinders. This part is linked
to turbulence and wake vortex shedding. In the third part,
one focuses on the most-studied aerodynamic profile type, i.e.
airfoil, and more specifically on the flow separation associated
with high angles of attack. Then, the last section deals with
other applications, such as jets, mixing layers and 3D vehicles.
In most cases, the airflow will be considered a 2D airflow,
excepted in the last section concerning 3D vehicles.
Note that only experimental works will be presented
here. Until 2006, there were only a few numerical works ( [13,
14, 60] for dc actuators and [108, 109] for DBD actuators for
instance). However, a large number of papers were published
last year at the AIAA conferences ([110, 111] for instance).
3.1. Flat plate
Figure 41 illustrates the action of electric wind in the case
of a corona discharge created by application of a dc high
voltage between a wire anode and an aluminium foil cathode.
When the air stream flows along the flat plate in the absence
of discharge (figure 41(a)), it separates from the wall at the
leading edge, forming a significant wake. When the actuator
is on (figure 41(b)), the airflow is reattached to the wall.
In 1968, Velkoff and Ketchman published the first paper
concerning airflow control by electrostatic process and they
demonstrated that the transition point on a flat plate could be

Airflow control by non-thermal plasma actuators

Figure 41. 2D visualization of the airflow at 0.4 m s1 , (a) in the


absence, and (b) in the presence of discharge.

affected by the application of an electric field [6]. The plasma


actuator was the one presented in figure 3(a). It was placed
32 cm downstream of the leading edge and 1.4 cm above the
plate wall. The applied electric field was composed of a continuous component of 9 kV, plus an alternative component of
2 kV. The time-averaged current was 1 mA. At 60 m s1 , the use
of this first plasma actuator induced a transition delay of 7 cm.
In 1992, Soetomo [10] experimentally observed a drag
reduction effect induced by ac and dc corona discharges along
a flat plate in the case of flow velocities up to 2 m s1 .
In 1998, Roth et al [29] published a complete paper
concerning flow manipulation along a flat plate boundary layer
TM
by means of their new plasma actuator, the OAUGDP .
The goal of this study was to demonstrate that EHD
forces could be sufficient to alter wall turbulence and drag.
Indeed, they performed flow visualizations, force and drag
measurements and discharge-induced velocity profiles. They
then investigated the influence of the body force direction
with the direction of the free air stream, in using span wise
and stream wise configurations. These works will be more
precisely detailed in section 3.1.2.
3.1.1. Momentum addition When the direction of the electric
wind and the free air stream are similar, one considers that the
goal of the manipulation is to increase the velocity inside the
boundary layer, i.e. to bring positive momentum. This may be
achieved by dc coronas and ac DBDs.
Several works were published by the University of Poitiers
and the University of Buenos Aires on airflow control by dc
surface coronas along flat plates [19, 20, 22, 23, 2628, 53]. I
am going to briefly summarize them. At the beginning, a large
number of electrode configurations were tested to limit the

corona-to-arc transition. They finally used the two configurations presented in figures 3(f) and (g). They then studied
the effect of such surface discharges on several aerodynamic
parameters. For instance, in [28], a dc-energized wire-toplane actuator placed at the leading edge of a flat plate parallel to the free air stream allowed them to increase significantly the velocity inside the boundary layer, up to 10 m s1
at U = 17.5 m s1 . The actuator acted inside the leadingedge recirculation bubble. In [23, 53], they presented works
based on velocity profile measurements, by means of a glass
Pitot tube, for velocities up to 25 m s1 (chord-based Reynolds
number Re of 375 000). In these experiments, the profile consists of a 50 cm wide flat plate of PMMA, with a chord of
21 cm. Its thickness was equal to 25 mm (figure 42). The
anode consisted of a 0.7 mm diameter copper wire electrode
placed inside a 0.7 mm deep groove located 10 cm downstream
of the leading edge. The cathode was a 2 mm diameter wire
placed inside a 2 mm deep groove, located 4 cm downstream
of the anode. Figure 43 shows velocity profiles for different
current values at x = 1 cm (1 cm downstream of the cathode),
at U = 5 m s1 . These profiles indicate that the discharge
induces an increase in airflow velocity close to the wall and
that the momentum addition increases with the discharge current (this is in agreement with figure 8). Figure 44 shows
velocity profiles at x = 1 cm, with discharge on and off, for
U values of 5, 10 and 17 m s1 . One can see that the velocity added by the plasma actuator decreases when the free air
stream velocity increases. This is certainly due to the fact that
the power added by the actuator is constant (here equal to 10 W
i.e. 830 W m2 ) whereas the kinetic power of the free airflow
increases with velocity. From their measurements, the authors
computed several aerodynamic parameters, such as drag, and a
few electro-mechanical parameters such as efficiency and electrohydrodynamic number NEHD . This dimensionless number,
which is a measure of the ratio of corona discharge-induced
body force to inertial force on the gas, is given by
NEHD =

i
,
lu2

(13)

where i is the time-averaged discharge current, the fluid density, l the electrode length, the ion mobility and u the local air
flow velocity. When NEHD 0, the airflow is hardly affected
by the EHD electric wind. Conversely, when NEHD +,
the electric wind velocity is so high compared with the free air
stream one that it considerably modifies the air flow. Then the
authors demonstrated that the electric wind has to act as close
to the wall as possible, for example at 0.1 mm from the wall for
velocities up to 50 m s1 . Moreover, this study showed that the
efficiency of such actuators (rate of electrical power converted
into kinetic power) is rather low (a few per cent) and that it
decreases with the discharge current. However, it is higher
in the presence than in the absence of the free air stream (see
figure 11).
A few other authors published papers where they describe
the effect of electric wind in the same direction as the free
air stream on a flat plate boundary layer. One can refer
to the recent works of Seraudie et al [71] who studied the
effect of dc coronas and DBDs on the laminar-to-turbulent
transition for velocities up to 50 m s1 and those of Borghi
et al [81,112,113] who used single and multiple surface barrier
discharge actuators.
623

E Moreau

Airflow

40mm

102.3 mm

70mm

X
ANODE 0.7mm

25mm

ANODE 0.7mm
CATHODE 2mm

CATHODE 2mm

Figure 42. Schematic of the flat plate used in [53].

i=0
i = 0.9 mA/m
i = 1.5 mA/m
i = 2.1 mA/m

14
12

Y (mm)

10
8
6
4
2
0
2.5

3.0

3.5

4.0

4.5

5.0

5.5

Velocity (m/s)
Figure 43. Velocity profiles in the boundary layer of a 5 m s1 free
airflow along a flat plate.

Several researchers worked on the separation along flat


profiles, linked to an inclination of the profile or depression
above it. For example, Hultgren and Ashpis [114] studied
the control of boundary layer separation using a single DBD
actuator, for 50 000 < Re < 300 000 and various free
stream turbulence intensities (from 0.2% to 7%). The results
suggested that the actuator works by promoting early transition
in the shear layer above the separation bubble thus leading to
rapid reattachment. Rivir et al [81, 115, 116] recently did nonstationary velocity measurements of a low-velocity airflow that
separates from the wall [116]. The data showed a phasedependence in the U -component of velocity (parallel to the
wall) when no similar phase-dependence was observed for the
V -component. Labergue et al [117] used a low-frequency
square wave corona discharge to reattach a naturally detached
airflow along a 17 inclined wall. They showed that each
reattachment induced a vortex downstream of the separation
point, for frequencies up to 20 Hz.
In all the above presented cases, the EHD body forces and
the free air stream were in the same direction. It is not the case
in the following presented works.
3.1.2. Boundary layer manipulation. Figure 45 presents
velocity profiles at U = 5 m s1 , when a dc actuator acts
in both directions: co-flow and counter-flow. It shows that the
electric wind may increase or decrease the velocity.
In [29], several EHD body force directions have been
TM
investigated. The actuator consisted of a surface OAUGDP ,
with span wise electrodes (co-flow and counter-flow) and
stream wise electrodes (in this case, the electric wind is
perpendicular to the free air stream). For instance, figure 46
624

presents the plate drag, in the absence of discharge and with


co- and counter-flow electric winds, for free velocities up to
25 m s1 . It shows that the drag is reduced when the electric
wind and the free air stream are in the same direction, whereas
it increases in the other case. However, the most interesting
point in this study is the effect of span wise body forces on
the boundary layer. In figure 47, one can see that these forces
modify significantly a 4 m s1 airflow at the wall, resulting
in a flow separation. As a result, figure 48 shows velocity
profiles measured between two adjacent electrodes, for the
case of a laminar (a), transitional (b) and turbulent (c) flow
at the panel leading edge. It shows a large acceleration of the
fluid near the wall and a retardation farther out. These results
are very surprising because they indicate that this actuator
configuration induces in the same time a flow acceleration
very close to the wall (2 mm, figure 48), and a flow separation
usually followed by a velocity decrease in the boundary layer
(see figure 47). Moreover, in this paper, it is demonstrated that
the span wise electric wind adds more velocity close to the wall
than the stream wise one, particularly at higher free air stream
velocities.
Jukes et al [95, 118] tried to reduce skin-friction drag
by applying a span wise oscillation in the near-wall region
of a turbulent boundary layer, using a surface DBD in a
geometrical electrode configuration that looks like the one
presented in figure 34(b). In such a configuration, one can
assume that the plasma induces a wall-jet flow, similar to
the one presented in figure 36, but in a 2D linear geometry.
Experiments are performed at low velocity (1.8 m s1 ) and
consist of visualizations and velocity measurements. Figure 49
presents a typical visualization of the wall flow induced by
the oscillatory plasma actuators in initially static air, at two
different times. It is clearly observed that a series of co-rotating
vortices are created by the plasma, depending on time. The
distance between two successive three-electrode actuators and
the voltage frequency may be adjusted to control the vortex
generation. Wilkinson [94] studied this actuator too.
In [96], the authors focus more specifically on the annular
wall-jet actuator that they perfected (see figure 36). For
instance, figure 50 presents time-averaged airflow streamlines
obtained from PIV measurements, for an increasing velocity.
At such low velocities (<1.75 m s1 ), a clear separation can be
seen, but the penetration of the jet decreases with increasing
cross flow velocity, as expected. In addition, they characterized
the non-stationary airflow streamlines obtained from phaselocked PIV measurements (figure 51). This shows that
the actuation is strongly non-stationary, because it modifies
dramatically the natural airflow at certain times (figure 51(a))
and has no effect at other times (figure 51(b)).

Airflow control by non-thermal plasma actuators

Y (mm)

Figure 44. Velocity profiles in the boundary layer at 5, 10 and 17 m s1 , with and without corona discharge.

20
18
16
14
12
10
8
6
4
2
0

Off
Co-flow
Counter-flow

0.2

0.4

0.6

0.8

1.0

Velocity (m/s)
Figure 45. Velocity profiles in the boundary layer of a 5 m s1 free
airflow along a flat plate, with co- and counter-flows.

Figure 46. Drag on a panel: baseline and plasma on with co- and
counter-flows. (From Roth et al [29]; reprinted with permission of
the American Institute of Aeronautics and Astronautics, Inc.)

3.2. Cylinder
In the previous section concerning airflow control along a
flat plate, the goal was mainly to manipulate the transition
position and the skin-friction drag. In the present section,
the case is aerodynamically very different: the goal of airflow
control around a cylinder is linked to the manipulation of
wake by preventing or provoking boundary layer separation.
The pressure distribution may be modified, resulting in vortex
shedding manipulation and reduction or enhancement of the
pressure drag. Here, the goal is not to modify skin friction.
Figure 52 presents an example of bluff body flow control
by a surface DBD at a few m s1 [119]. The plasma actuators

Figure 47. Top view visualizations of a manipulated airflow by a


stream wise electrode actuator, (a) wide view, (b) zoomed view of a
single actuation. (From Roth et al [29]; reprinted with permission of
the American Institute of Aeronautics and Astronautics, Inc.)

are located at 90 and 135 with respect to the approach


flow direction, inducing a tangential electric wind of several
m s1 (figure 52(a)). With the actuators off, the flow undergoes
subcritical separation leading to a large-scale separated
flow region, accompanied by unsteady vortex shedding
(figure 52(b)). With the actuators turned on (figure 52(c)),
there is a substantial reduction of the separated flow region,
resulting in an elimination of the associated vortex shedding.
The first study of flow control around cylinders was published in 1997 [11] and completed in [21,120]. It dealt with dc
surface coronas established at the cylinder surface, between a
0.9 mm diameter wire anode and a thin aluminium foil (50 m
thick) located at 0 and 135 with respect to the forward
stagnation point, respectively (figure 53(a)). By applying dc
625

E Moreau

Figure 49. Visualization of the vortices induced by successive


three-electrode DBD actuators. (From Jukes et al [118]; reprinted
with permission of the American Institute of Aeronautics and
Astronautics, Inc.)

counter-rotating vortices, well known in this Re range. The


contour of these vortices defines the limits of the mean recirculation bubble. The presence of the discharge induces a slight
narrowing and shortening of this zone. Figure 54 presents the
cylinder pressure coefficient CP with discharge off and on, for
several discharge currents. CP is given by
CP =

Figure 48. Non-dimensional velocity profiles measured between


two adjacent electrodes, for the case of a laminar (a), transitional (b)
and turbulent (c) flow at the panel leading edge. (From Roth
et al [29]; reprinted with permission of the American Institute of
Aeronautics and Astronautics, Inc.)

high voltage, a streamer plasma sheet covered the cylinder


surface, inducing an electric wind from the wire anode to the
plane cathode. In order to investigate the discharge effects
for velocities up to 27 m s1 (Re = 58 000), the authors performed flow visualizations and PIV measurements to analyse
the near-wake manipulation and wall pressure measurements
with the help of 20 small orifices located around the cylinder, every 9 and connected to pressure sensors (figure 53(b)).
With plasma off, the PIV time-averaged flow fields show two
626

P P0
,
2 /2
U

(14)

where P is the measured pressure, P0 the static pressure and


the air density. Figure 54 shows that the curves have a minimum, followed by a plateau usually associated with the boundary layer separation. On one hand, it seems that the actuation
produces no significant delay in the separation. On the other
hand, the plateau pressure is significantly reduced, whatever
the current value. This means that the suction increases in the
recirculation bubble.
Hyun and Chun [50] investigated the wake flow control
behind a circular cylinder using dc corona-induced electric
wind. In these works, the electrode geometry was the one
presented in figure 3(c), resulting in a volume corona instead
of a wall surface one. The goal was not to modify the boundary
layer but rather the wake characteristics. They performed
visualizations and pressure measurements, at very low
Reynolds number (Re < 8000). They showed that the pressure
drag may be dramatically reduced by the plasma actuator.
For comparable Reynolds numbers (Re < 7400),
McLaughlin et al [121] used a single DBD to manipulate the
flow separation and wake behaviour downstream of a circular
cylinder. Hot wire measurements showed that vortex shedding
frequency can be driven to the actuator forcing frequency. In
the same way, Thomas et al [119] studied the bluff body flow
control, motivated by the need to reduce landing gear noise
for commercial transport aircraft via an effective streamlining
created by the surface DBD presented in figure 52(a). Note

Airflow control by non-thermal plasma actuators

(a)

(a)

(b)

(b)

Figure 51. Phase-locked PIV results showing the manipulated


airflow streamlines at different times at 0.85 m s1 . (From
Santhanakrishnan and Jacob [97]; reprinted with permission of the
American Institute of Aeronautics and Astronautics, Inc.)

(c)

Figure 50. Time-averaged PIV results showing the manipulated


airflow streamlines at 0.85, 1.25 and 1.75 m s1 . (From
Santhanakrishnan and Jacob [97]; reprinted with permission of the
American Institute of Aeronautics and Astronautics, Inc.)

that Re < 33 500 in this study, corresponding to a maximum


velocity equal to 5 m s1 . The electric wind velocity is
then in the same order of magnitude as the free air stream
velocity. However, the results are fundamentally interesting.
Steady manipulation has been shown to drastically reduce
flow separation and the associated Karman vortex shedding is
eliminated. An additional result is that peak turbulence levels
downstream of the cylinder are reduced by 50%. Moreover, the
authors used an unsteady actuation, as illustrated in figure 55.
On one hand, in steady actuation, the waveform frequency
is sufficiently high in relation to any relevant frequency of
the baseline flow such that the associated body force may be
considered effectively steady. On the other hand, for unsteady
actuation, the signal sent to the actuator has a characteristic
frequency f = 1/T1 , which is much lower than that of the
sinusoid and is comparable to some relevant frequencies of the
flow around the cylinder. Duty is given by 1/T1 divided by the
DBD sine waveform frequency. By using unsteady actuation,
with a frequency corresponding to a Strouhal number St = 1,
the effects were optimum. Peak turbulence level was reduced
by 66% with only 25% of the input power required for steady
actuation. Consequently, the near-sound pressure levels were
reduced by 13.3 dB in the wake.
Sung et al [122] showed by PIV measurements and
visualizations that the discharge alters the location of the

flow separation for Re < 40 000. In [123], an experimental


study deals with the use of plasma actuators in order to phase
synchronize vortex shedding from two side-to-side circular
cylinders, in the Reynolds number range Re = 16 70076 000
(U  30 m s1 ). Plasma actuators were placed at 90
from the forward stagnation point of each cylinder, along
the full span. The vortex shedding phase between the two
cylinders was determined from unsteady velocities measured
by two hot wires, located behind the cylinders. This showed
the effectiveness of the plasma actuators in synchronizing the
vortex shedding.
Recently, McLaughlin et al [124] maintained their low
velocity study presented in [121] but at higher velocities
(up to 110 m s1 , Re = 300 000). They showed that the
plasma actuator was capable of triggering vortex shedding at
its switching rate. Indeed, very promising results with a nonstationary actuation have been obtained, suggesting that the
plasma actuator is more efficient when it is used in a noncontinuous operation.
3.3. Airfoil
Aerodynamic performance of a plane wing is of great industrial
interest. It is the reason why this profile is the most-studied
one. Two main aspects are investigated. The first one is the
reduction of skin-friction drag, which would lead to fuel cost
decreases, longer ranges and higher speeds because it can be
used during cruise conditions. The second one deals with flow
separation control, resulting in lift control without flaps for
instance. This aspect is more important during phases of takeoff and landing.
The first study was published by Shcherbakov et al in
2000 [125]. In this paper, drag reduction caused by a surface
DBD along a wing-like profile was studied, for velocities
up to 53 m s1 . The profile had a chord of 11 cm and the
627

E Moreau

discharge device, presented in figure 14(c), was placed on


both sides of the profiles, along the chord, from about 30
100 mm downstream of the leading edge. The airflow was
naturally laminar up to 100 mm downstream of the leading
edge. To make it turbulent, a trip-wire (turbulisater) consisting
of a 0.5 mm diameter wire might be placed 15 mm behind
the leading edge. In these conditions, Re  4 105 .
As a result, relative drag reduction up to 5.3% at 35 m s1
has been found for turbulentairflow whereas the maximum

Figure 52. Schematic of the DBD plasma actuators mounted on the


cylinder profile in (a) where (2) are the air-exposed electrodes, (3)
the encapsulated electrodes and (5) the plasma formations and
smoke visualizations around a cylinder (b) with plasma off, (c) with
plasma. (From Thomas et al [119]; reprinted with permission of the
American Institute of Aeronautics and Astronautics, Inc.)

reduction was 0.8% for an initially laminar flow. The authors


consider that the effect seems to relate to suppression of the
turbulent pulsations originating from the transient boundary
layer between the inner viscous boundary sub-layer and the
outer turbulent boundary layer. One can think that the
effects are not due to the laminar-to-turbulent transition, or
the separation, but certainly to a skin-friction modification.
This was a rare case where the skin friction was
investigated, because in most cases, the authors are interested
in separation control. For example, Roth [92] studied the effect
of a single DBD and an 8-phase travelling electrostatic wave on
the separation at high angles of attack by flow visualizations, at
low velocities. For instance, figure 56 presents visualizations
of a naturally detached airflow along a NACA 0015 airfoil
inclined at = 12 . Its chord is 12.7 cm and the free air stream
velocity is 2.85 m s1 (Re 104 ). Although it seems that these
visualizations are slightly perturbed by the smoke injector
(indeed, the baseline airflow separates from 4 in this study
whereas the separation starts usually at about 12 , displacing
from the trailing edge to the leading edge with increasing angle
of attack), he shows clearly the plasma actuator effect: the
electric wind reattaches the airflow along the wall or at least
delays the separation.
The group of Corke at the University of Notre-Dame have
been working on separation control over airfoils for about 5
years [2 ,126131]. Their first paper dealing with this subject
was published in 2002 [2]. In these preliminary experiments
using a single DBD actuator, measurable lift enhancement
for a full range of angles of attack had been demonstrated,
accompanied by an increase in drag. The plasma effect was
compared with an increase in camber. One year later, they
published a paper in which two methods of control were
compared: a passive method with mechanical vortex generators
and a plasma-based active method [126]. The airfoil was a 663 018 with a chord c = 12.7 cm. Two actuators were used: the
first one was placed at the leading edge (x/c = 0) and the
second one at x/c = 0.5 (50% of the chord). The airfoil was
instrumented for surface pressure measurements in order to
calculate lift coefficients. Mean velocity profiles downstream
of the airfoil were used to determine drag coefficients.
Measurements were performed over a range of free air stream
velocities from 10 to 20 m s1 , that is to say 79 000 < Re <
158 000. The main results showed that the actuators were
found to lead to reattachment for angles of attack that were
8 past the stall angle. It was accompanied by a full pressure
recovery and up to a 400% increase in the lift-to-drag ratio.
Moreover, the authors demonstrated that the most effective
actuator location to reattach the leading-edge flow separation
was at the exact leading edge. This last point is important

Figure 53. Schematic side view of the cylinder, (a) with location of both electrodes and (b) location of the 20 pressure measurement orifices.

628

Airflow control by non-thermal plasma actuators

Figure 54. Pressure distribution on the cylinder surface for several


current values, at U = 27 m s1 .

Figure 56. Visualisations of the airflow streamlines along a 12


inclined NACA 0015 airfoil at 2.85 m s1 , (a) without actuation,
TM
(b) manipulated by the OAUGDP . (Reused with permission
from [92] Roth J R 2003 Phys. Plasmas 10 2117, Copyright 2003,
American Institute of Physics.)

Figure 55. Steady and unsteady actuation. (From Thomas


et al [119]; reprinted with permission of the American Institute of
Aeronautics and Astronautics, Inc.)

but leads to a question: is this effect only due to the velocity


addition at the wall of the leading edge or might it be due to
the laminar-to-turbulent transition induced by the pulse electric
wind of such a surface DBD actuator (see figures 29 and 30).
Since 2003, Corke et al have worked with a NACA
0015 airfoil and performed a large number of experiments
concerning separation control at high angles of attack, in
stationary and periodic oscillating aerodynamic conditions,
with steady and unsteady actuations.
As previously,
measurements have mainly consisted of flow visualizations,
wall pressure distribution records for lift computation and
downstream velocity profile acquisitions for drag estimation.
For instance, figures 5759 present a typical example of
their measurements in the case of steady actuation with an
immovable airfoil at 20 m s1 : wall pressure distribution CP
at = 16 (see equation (14)), lift coefficient CL versus and
a velocity profile downstream of the airfoil, with and without
plasma actuation. Figure 57 shows that the actuation induces
a pressure recovery near the leading edge. By integrating
the CP distribution, the coefficient of lift may be obtained,
as illustrated in figure 58. With the plasma actuator on, CL
is consistently higher compared with the original post-stall
condition (  15 ). Mean velocity profiles in the wake of the
airfoil for the same conditions are presented in figure 59. With
plasma on, an increase in the minimum peak deficit velocity
and a substantial reduction in the wake width are observed.
This induces drag reduction for high angles of attack.

Figure 57. Coefficient of pressure distribution with plasma actuator


on and off, for = 16 , U = 20 m s1 and Re = 158 000. (From
Post and Corke [127]; reprinted with permission of the American
Institute of Aeronautics and Astronautics, Inc.)

In [129], Corke et al used an airfoil mounted on a lift-drag


force balance and investigated more accurately the effect of
unsteady actuation (see figure 55) instead of a steady actuation
and the influence of the position of the actuator: one is located
at the leading edge while the other one is placed at the trailing
edge. The goal here was to mimic the effects of wing leading
edge slat and trailing edge flaps. The airfoil is still a NACA
0015 profile, with a chord of 12.7 cm. Flow control tests have
been conducted at 21 and 30 m s1 , corresponding to a maximum Re = 257 000 (307 000 if one takes into account the correction due to the blockage). The results showed that unsteady
actuation was more efficient. Indeed, steady actuation was able
629

E Moreau

Figure 58. Comparison of computed lift coefficient with plasma on


and off, at U = 20 m s1 and Re = 158 000. (From Post and
Corke [127]; reprinted with permission of the American Institute of
Aeronautics and Astronautics, Inc.)

Figure 59. Wake mean velocity profiles with plasma actuator on


and off, for = 16 , U = 20 m s1 and Re = 158 000. (From
Post and Corke [127]; reprinted with permission of the American
Institute of Aeronautics and Astronautics, Inc.)

to reattach the flow for angles of attack up to 19 , which was


4 past the normal stall angle while unsteady actuation was
able to reattach up to 9 past the normal stall angle. More, the
most efficient results were obtained for an unsteady actuation
frequency corresponding to a Strouhal number equal to one,
and the duty of the unsteady actuation was equal to 10%. This
means that the unsteady power consumption was 2 W instead
of 20 W in the steady case. For instance, figure 60 confirms
this result: unsteady actuation is more efficient than the steady
one on the lift coefficient. Concerning flow control over an
oscillating airfoil, all the results may be found in [127,128,130]
which consist of a summary of previous works. Briefly, the
main result indicates that the best actuation frequency appeared
to be approximately eight times the airfoil pitching frequency.
During the same period as Corke et al in 2002, Sosa et al
conducted the same kind of experiments, but the actuator was
a surface corona-based one. These works were published two
years later, in [132] and completed in [133, 134]. In [131], the
630

Figure 60. Comparison of measured lift coefficient with plasma off,


steady actuation and unsteady actuation, at U = 30 m s1 and
Re = 257 000. (From Corke et al [129]; reprinted with permission
of the American Institute of Aeronautics and Astronautics, Inc.)

airfoil is a 20 cm chord NACA 0015 profile, with two electrodes


flush-mounted at the leading-edge wall (from x/c = 00.18)
energized by dc and ac high voltages, for frequencies up to
2 kHz. The velocity range was 1025 m s1 , corresponding
to a maximum Re = 375 000. The authors conducted wall
pressure measurements by means of 20 holes on each side of
the airfoil, characterized by velocity fields around the airfoil
from PIV and computed drag and lift. For example, figure 61
presents a typical result of PIV measurements, at 25 m s1
(Re = 375 000) and = 19.8 . Without actuation, it is
clearly observed that the airflow is fully detached from the
leading edge (figure 61(a)). In the presence of the discharge,
the actuator effect depends dramatically on the discharge
frequency and the input electrical power. In figure 61(b),
frequency and input power are, respectively, 30 Hz and 5 W; the
airflow is not fully reattached. By increasing the input power
up to 10 W, the separation is displaced downstream but the
airflow is still partially detached. By adjusting the discharge
frequency at 50 Hz, even with an input power of only 5 W
(figure 61(d)), the airflow can be fully reattached, resulting in
an increase in the lift-to-drag ratio.
All these studies are very interesting to understand
the physical phenomena involved in such actuations and to
characterize the capability of plasma actuators. However, the
concerned Reynolds number is still widely smaller than that in
cruise aircraft conditions. Recently, a Russian group published
a paper dealing with pressure distribution of the suction side of
a 9 cm chord NACA 0015 airfoil, for velocities up to 110 m s1
[135137]. The actuator is composed of three successive
single DBDs, from the leading edge to the trailing edge.
These three very similar publications are difficult to understand
because the experimental setup and the measurements are not
precisely explained. For example, it seems that the electric
wind and the free air stream are in the same direction, but
this is never clearly indicated. However, the results are very
interesting. In fact, the main result at 110 m s1 is as follows:
up to 12 , the plasma actuator has no effect because there is no
separation. Between 12 and 20 , the actuation is very effective,

Airflow control by non-thermal plasma actuators

leading to a strong increase in the pressure distribution along


the suction side of the airfoil. This means that the airflow is
reattached. Then, above 20 , the effects are negligible. The
authors assume that they can significantly change the flow only
at the separation point, where the velocity is relatively low.
In 2006, they maintained the same measurements [138]
but with a stream wise electrode actuator, resulting in an
electrical wind direction perpendicular to the free air stream.
They obtained similar results, indicating that the airflow could
be reattached with this device. These results are important,
because, as was suspected, this means that the discharge effect
which induces the reattachment is not a flow acceleration at
the wall. The effect may be explained by the formation of 3D
vortexes or by the laminar-to-turbulent transition inside the
boundary layer.
Concerning the location where the actuator must be placed
to be the most efficient, Jolibois et al [139] published recently a
paper where the authors mounted seven independent actuators
along the suction side of a 1 m chord NACA 0015 airfoil, from
x/c = 0.30.8. This study had two advantages compared with
previous ones. Firstly, the boundary layer was transitioned
at the leading edge by a trip-surface (turbulator). Then
one can be sure that the actuator effects are not due to the
laminar-to-turbulent transition but only due to velocity addition
close to the wall. Secondly, flow visualizations and PIV
measurements enabled one to understand well the physical
effect of each actuator, as a function of the flow separation
point which depends on the angle of attack. The authors
investigated the ability of a surface DBD actuator to displace
(upstream and downstream) the position of the separation
point, by reattaching naturally detached airflows or, inversely,
by detaching naturally attached airflows. The main result
was that, whatever the angle of attack and the location of the
separation point, the actuator must act at the separation point
to be the most efficient with a minimum input power.
3.4. Other applications
Airflows along flat plates, cylinders and airfoils are widely
studied because they cover a large number of aerodynamic
applications. However, there are a few other fundamental
aerodynamic cases, such as jets and mixing layers and a few
typical aerodynamic cases usually associated with industrial
applications, such as turbine blade, 3D airfoils and 3D vehicles.
These cases are briefly presented in this section.
In [140], Van Dycken et al focus on a plasma actuator
installed on a blade of a gas turbine blade cascade. These highcamber angle blades are used for transportation and stationary
applications. At partial load, i.e. at low flow velocities, they
exhibit flow separation on the suction side. Then a plasma
actuator is placed on the suction side, near the trailing edge,
such as on an airfoil. Separation is detected via surface pressure
measurements and loss of stagnation pressure via Pitot pressure
measurements. The results indicate a reduction of stagnation
pressure losses by 50 % at 3 m s1 (Re = 30 000) for an
input power of 45 W. With a similar profile, Morris et al [141]
obtained qualitative changes in the structures of downstream
wake profiles, and the results were strongly dependent on the
unsteady actuation frequency.
Concerning the case of mixing layers, it seems that only
two papers have been published [142, 143]. The objective

here was to manipulate the mixing layer downstream of a thin


flat plate, by detachment and reattachment of both airflows
by means of a co- or counter-electric wind induced by a
surface corona actuator (dc and ac). The results showed that
the actuation modifies significantly the mixing layer thickness
for velocities up to 16 m s1 . This study has been followed
by several experiments dealing with jet manipulation for
velocities up to 30 m s1 [144, 145, 146] and is summarized
in a paper published in the present special issue [147].
Two other works investigated the ability of plasma
actuators for jet control. The first one deals with the use of an
original travelling wave plasma device applied to the control
of the exciting helical modes in a 30 m s1 axisymmetric
jet [93]. The plasma device looks like the one presented
in figure 34(f) but the originality is due to the fact that the
air-exposed electrodes are grounded, and square wave phase
shifted high voltages are applied to the embedded electrodes.
This device is installed along the inner side of the jet diffuser,
in order to induce an electric wind perpendicularly to the
jet. Measurements consist of hot-wire velocity records. The
main result is that the time-averaged jet velocity is slightly
affected by the actuation, whereas the rms fluctuations are
strongly modified, according to the azimuthal mode number.
The second work investigated the ability of the localized arc
filament plasma actuator (see figure 39) for high-speed and
high Reynolds number jet control [103105]. Typically, the
jet, which can be axisymmetric or rectangular, is created by a
converging nozzle with a design Mach number of 1.3. Its size
is a few centimetres, then Re is equal to a few 105 . The authors
have demonstrated the potential of this actuator in generating
stream wise vortices, and they assume that it could be used to
excite the jet instabilities. In fact, plasma actuation is explained
by a local pressure rise due to a local heating up to high
temperature inside the filaments. The authors assume that this
actuator could be effective in heated jets, such as exhaust jets
of typical commercial aircrafts where the temperature is 700
800 K, because the filament temperature considerably exceeds
1250 K.
Finally, one can indicate the last three studies concerning
airflow control around 3D vehicles: the works of Vorobiev et al
[148] who investigated lift enhancement and roll control over a
3D NACA 0009, the works of Patel et al [149] who used plasma
actuators for hingeless aerodynamic control of an unmanned
air vehicle and the works of Goksel and Rechenberg [150]
dealing with airflow control on a plane-like flying wing.
Note that the use of non-thermal surface plasmas for
supersonic airflow control has not been discussed since few
works have been published ([152] for instance) and that a
review dealing with the use of thermal plasmas for high-speed
airflow control has been published recently [107].
3.5. Conclusion
The manipulation of a flat plate boundary layer has shown
significant effects on velocity profiles, resulting in transition
delay or drag reduction for instance. However, this airflow
case is the less advantageous for manipulation because it is
aerodynamically very stable. And although several authors
used electric wind in the same direction as free air stream, it
seems that efficient results may come by using counter-flow
631

E Moreau

(a)

(b)

(c)

(d)

Figure 61. Velocity fields and associated streamlines obtained from PIV measurements of an airflow over an inclined NACA 0015 airfoil,
for = 19.8 and U = 25 m s1 (Re = 375 000); (a) without actuation, (b) with actuation where the discharge frequency f = 30 Hz and
the electrical input power Pelec = 5 W, (c) with actuation, f = 30 Hz, Pelec = 10 W, (d) with actuation, f = 50 Hz, Pelec = 5 W. (From
Corke et al [131]; reprinted with permission of the American Institute of Aeronautics and Astronautics, Inc.)

electric winds, span wise-directed electric winds or vortex


generators induced by plasma-induced wall jets. Concerning
separation control, results are very promising because they
show that this phenomenon can be manipulated with very little
injected input power. The effectiveness is not only due to
the electric wind velocity, but also to how the momentum
is brought about, spatially and temporally. Indeed, to be
efficient, the plasma must act at the right location, at the
right time. To do that, one needs a better knowledge of the
near-wall aerodynamic and more specifically of the turbulence
characteristics.

4. Conclusion
The first part of this paper was dedicated to the electrical and
mechanical characterization of plasma actuators. It has been
clearly shown that both corona-based and dielectric barrierbased discharges could be established at the wall of a dielectric,
at atmospheric pressure in air. Typically, the mechanical effect
of such actuators is due to the induced-electric wind. Its timeaveraged velocity is typically a few m s1 , and it can reach
8 m s1 at about 0.5 mm distance from the wall. The body force
632

created by the ion motion is globally about 0.15 mN per watt


of input electrical power, and the electro-mechanical efficiency
is a few tenths of a percent in the absence of free air stream.
Each kind of discharge has its advantages and drawbacks. On
one hand, the main advantage of the surface DBD compared
with the surface corona is its stability, because the glow-toarc transition is limited by the barrier. On the other hand, the
electro-mechanical efficiency of coronas is better than that of
DBDs, and the electrical devices are simpler in the case of
coronas. Consequently, researchers in electrical engineering
and plasma physics must continue to work on both discharges
to optimize them and to imagine other configurations, such as
that presented at the end of the first section. Moreover, these
actuators must be tested in a hostile environment, in terms of
temperature, pressure and relative humidity, for instance, to be
used industrially. As an example, the environmental conditions
around a cruise aircraft wing are a temperature near 40 C,
a pressure of about 250 kPa and a variable humidity, while at
the turbine exit or inside the motor, the temperature may reach
a few thousand kelvin with over or under-pressures.
The experimental works presented in the second part have
shown the ability of plasma actuators to manipulate airflows in

Airflow control by non-thermal plasma actuators

various aerodynamic conditions. Very efficient results have


been obtained for low-velocity subsonic airflows (typically
U  30 m s1 and Reynolds number of a few 105 ), and some
promising results at higher velocities make one think that the
plasma actuators could be used in aeronautics. However, note
that the general aviation Reynolds number is typically around
107 108 . Consequently, efforts are required in the near future
to reach velocities of aeronautics and to conduct experiments
with large-scale profiles. Otherwise, other applications at low
velocities will have to be found, such as the case of internal
flows.
Moreover, it seems that the induced-electric wind is not
the only property of a surface discharge to modify the airflow
at the wall. A recent study showed a slight modification of
the density inside the plasma. In future, density and viscosity
of these surface discharges must be accurately estimated, to
know if their potential decrease can play a role in wall airflow
control.

Acknowledgments
The author greatly acknowledges all the researchers of his
laboratory who have efficiently contributed to the works
performed at the University of Poitiers and more especially
his PhD students who have always had a passion for this
exciting topic, Pr Gerard Touchard for sharing his experience
on EHD phenomena and Jean-Paul Bonnet for his help. He
also acknowledges the researchers of other laboratories with
whom he worked and more especially Pr Guillermo Artana
and Frederic Thivet. Moreover, he acknowledges the French
Ministry of Research, Airbus Industries (EADS) and Snecma
Moteurs (Safran) for their financial support.

References
[1] Gad-El-Hak M 2000 Flow Control (Cambridge: Cambridge
University Press)
[2] Corke T C, Jumper E J, Post M L, Orlov D and
McLaughlin T E 2002 Applications of weakly-ionized
plasmas as wing flow-control devices AIAA Meeting (Reno,
USA, January 2002) paper #20020350
[3] Turck J 1951 French Patent # 1.031.925
[4] Bahnson A G 1959 French Patent # 1.266.476
[5] Hill G A 1963 US Patent # 3.095.163
[6] Velkoff H and Ketchman J 1968 Effect of an electrostatic field
on boundary layer transition AIAA J. 16 13813
[7] Yabe A, Mori Y and Hijikata K 1978 AIAA J. 16 3405
[8] Bushnell D 1983 AIAA Meeting (Reno, USA, January 1983)
paper #1983-0227
[9] Malik M R, Weinstein L and Hussaini M 1983 AIAA Meeting
(Reno, USA, January 1983) AIAA paper #1983-0231
[10] Soetomo, F 1992 The influence of high voltage discharge on
flat plate drag at low Reynolds number air flow MS Thesis
Iowa State University
[11] Noger C, Touchard G and Chang J S 1997 Proc. ISNTP-2
(Salvador, Brazil) pp 136141
[12] Soldati A and Banerjee S 1998 Turbulence modification by
large scale organized electrohydrodynamic flows Phys.
Fluids 10 174256
[13] El-Khabiry S and Colver G M 1997 Drag reduction by a DC
corona discharge along an electrically conductive flat plate
for small Reynolds number flow Phys. Fluids 9 58799
[14] Colver G and El-Khabiry S 1999 Modeling of DC corona
discharge along an electrically conductive flat plate with gas
flow IEEE Trans. Indust. Appl. 35 38794

[15] Vilela Mendes R and Dente J A 1998 Boundary layer control


by electric fields J. Fluid Eng. 120 6269
[16] Leger L, Moreau E, Artana G and Touchard G 2000
Modification de lecoulement dair autour dune plaque
plane par une decharge couronne Proc. Colloque de la
Societe Francaise dElectrostatique (Montpellier, France,
July 2000) pp 97101
[17] Artana G, DAdamo J, Desimone and Diprimio G 2000 Air
flow control with electrohydrodynamic actuators 2nd Int.
Workshop on Conduction Convection and Breakdown in
Fluid (Grenoble, France) pp 1904
[18] Artana G, Desimone G and Touchard G 1999 Study of the
changes in the flow around a cylinder caused by
electroconvection Electrostatics 99 (IOP Publishing Ltd,
BristolPhiladelphia) pp 14752
[19] Leger L, Moreau E, Artana G and Touchard G 2001 Influence
of a DC corona discharge on the airflow along an inclined
flat plate J. Electrostat. 5051 3006
[20] Leger L, Moreau E and Touchard G 2001 Control of low
velocity airflow along a flat plate with a DC electrical
discharge Proc. IEEE-IAS World Conf. on Industrial
Applications of Electrical Energy (Chicago, USA, 30
September4 October)
[21] Artana G, DiPrimio G, Desimone G, Moreau E and Touchard
G 2001 Electrohydrodynamic actuators on a subsonic
airflow around a circular cylinder Proc. 4th AIAA Weakly
Ionized Gases International Conference (Anaheim, USA,
June 2001) paper #2001-3056
[22] Artana G, DAdamo J, Leger L, Moreau E and Touchard G
2001 Flow control with electrohydrodynamic actuators
Proc. 39th AIAA Meeting (Reno, USA, January 2001)
paper #2001-0351
[23] Leger L, Moreau E and Touchard G 2002
Electrohydrodynamic airflow control along a flat plate by a
DC surface corona dischargevelocity profile and wall
pressure measurements Proc. 1st AIAA Flow Control Conf.
(St Louis, June 2002) paper #2002-2833
[24] Moreau E, Leger L and Touchard G 2002 Three-phase
traveling wave surface discharge along an insulating flat
plate in air: application to electrohydrodynamically airflow
control IEEE-CEIDP Annual Report 272-278 (Cancun,
Mexique, 2024 October 2002)
[25] Louste C, Moreau E and Touchard G 2002 DC corona surface
discharge along an insulating flat plate in air: experimental
results IEEE-CEIDP Annual Report 822-826 (Cancun,
Mexique, 2024 October 2002)
[26] DAdamo J, Artana G, Moreau E and Touchard G 2002
Control of the airflow close to a flat plate with
electrohydrodynamic actuators Proc. ASME-FEDSM 2002
(Canada, 1418 July 2002)
[27] Leger L, Moreau E and Touchard G 2002 Effect of a DC
corona electrical discharge on the airflow along a flat plate
IEEE Trans. Indust. Appl. 38 147885
[28] Artana G, DAdamo J, Leger L, Moreau E and Touchard G
2002 Flow control with electrohydrodynamic actuators
AIAA J. 40 17739
[29] Roth J R, Sherman D M and Wilkinson S P 1998 Boundary
layer flow control with a one atmosphere uniform glow
discharge surface plasma AIAA Meeting (Reno, USA,
January 1998) paper #98-0328
[30] Roth J R 1998 Electrohydrodynamically induced airflow in a
one atmosphere uniform glow discharge surface plasma
25th IEEE Int. Conf. Plasma Science (Raleigh, USA)
[31] Roth J R, Tsai P P and Liu C 1995 US Patent #5387842
[32] Szames A D 2000 Lelectroaerodynamique, secret de la
furtivite du B-2 Air Cosmos 1757 223
[33] Szames A D 2004 Les enjeux de lelectroaerodynamique Air
Cosmos 1918 2430
[34] Roth J R 1995 Industrial Plasma Engineering vol 1 (Bristol:
Institute of Physics Publishing)
[35] Roth J R 2001 Industrial Plasma Engineering vol 2 (Bristol:
Institute of Physics Publishing)

633

E Moreau

[36] Schutze A et al 1998 The atmospheric-pressure plasma jet: a


review and comparison to other plasma sources IEEE Trans.
Plasma Sci. 26 168594
[37] Fridman A, Chirokov A and Gutsol A 2005 Non-thermal
atmospheric pressure discharges J. Phys. D: Appl. Phys.
38 R124
[38] Loeb L B 1965 Electrical Coronas, Their Basic Physical
Mechanisms (Berkeley: University of California Press)
[39] Lacoste A, Pai D and Laux C 2004 Ion wind effect in a positive
DC corona discharge in atmospheric pressure air AIAA
Meeting (Reno, USA, January 2004) paper #2004-0354
[40] Hartmann G 1977 Spectrographie de la decharge couronne:
e tude des mecanismes de collisions dans le dard PhD Thesis
University of Paris XI
[41] Goldman M and Sigmond R S 1982 Corona insulation IEEE
Trans. Electr. Insul. 90105
[42] Robinson M 1961 Movement of air in the electric wind of the
corona discharge AIEE Trans. 80 14350
[43] Robinson M 1962 A history of the electric wind Am. J. Phys.
30 36672
[44] Sigmond R S and Lagstadt I H 1993 Mass and species transport
in corona discharges High Temp. Chem. Proces. 2 2219
[45] Moreau E, Afande Y and Touchard G Electric wind in
coronasapplication to the perfecting of a wall injection jet
plasma actuator ISHED 2006 (Buenos Aires, Argentina,
December 2006)
[46] Loiseau J F, Batina J, Noel F and Peyrous R 2002
Hydrodynamical simulation of the electric wind generated
by successive streamers in a point-to-plane reactor J. Phys.
D: Appl. Phys. 35 102031
[47] Zouzou N, Moreau E and Touchard G 2006 Precipitation
e lectrostatique dans une configuration pointe-plaque J.
Electrostat. 64 53742
[48] Rickard M, Dunn-Rankin D, Weinberg F and Carleton F 2006
Maximizing ion-driven gas flows J. Electrostat.
64 36876
[49] Velkoff H, Godfrey R 1979 Low velocity heat transfer to a
plate in the presence of a corona discharge in air J. Heat
Transfer 101 15763
[50] Hyun K T and Chun C H 2003 The wake flow control behind a
circular cylinder ion wind Exp. Fluids 35 54152
[51] Labergue A, Moreau E and Touchard G 2005 Proc. CEIDP
(Nashville, USA, October 2005) pp 46973
[52] Moreau E, Artana G and Touchard G 2004 Surface corona
discharge along an insulating flat plate in air applied to
electrohydrodynamical airflow control: electrical properties
Electrostatics 2003 (Institute of Physics Conference Series
vol 178) (Bristol: Institute of Physics Publishing) pp 28590
[53] Moreau E, Leger L and Touchard G 2006 Effect of a DC
surface non-thermal plasma on a flat plate boundary layer
for airflow velocity up to 25 m s1 J. Electrostat.
64 21525
[54] Louste C, Moreau E and Touchard G 2004 Influence of an
insulating flat plate on a DC surface corona discharge at
various air relative humidity Electrostatics 2003 (Institute
of Physics Conference Series vol 178) (Bristol: Institute of
Physics Publishing) pp 2738
[55] Moreau E, Labergue A and Touchard G 2005 DC and pulse
surface corona discharge along a PMMA flat plate in
air: electrical properties and discharge-induced ionic wind
J. Adv. Oxydation 8 2417
[56] Moreau E 2004 Application des plasma non thermiques au
controle e lectrofluidodynamique des e coulement
Habilitation Thesis Universite de Poitiers, France
[57] Leger L 2003 PhD Thesis (University of Poitiers, France)
[58] Moreau E, Labergue A and Touchard G 2005 About the
kinetic power induced by AC and DC discharges Proc.
IEEE-CEIDP (Nashville, USA, October 2005) pp 4904
[59] Forte M, Leger L, Pons J, Moreau E and Touchard G 2005 DC
and pulse plasma actuator for airflow control:
measurements of the instationary ionic wind velocity J.
Electrostat. 63 92936

634

[60] Baudoin F, Moreau E and Touchard G 2005 Proc. CAPPSA


2005 (Bruges, Belgium, 30 AugustSeptember 2) pp 1969
[61] Mateo-Velez J C, Thivet F, Rogier F, Quinio G and Degond P
2005 Numerical modelling of corona discharges and their
interaction with aerodynamics Proc. EUCASS 2005
(Moscow, Russia, July 2005)
[62] Von Engle A, Seeliger R and Steenback M 1933 On the glow
discharge at high pressure Z. fur Physik 85 14460
[63] Yokoyama S, Kogoma M, Morikawi T and Okazaki S 1990
The mechanisms of the stabilized glow plasma at
atmospheric pressure J. Phys. D: Appl. Phys. 23 11258
[64] Massines F, Rabehi A, Decomps P, Gadri R B, Segur P and
Mayoux C 1998 Experimental and theoretical study of a
glow discharge at atmospheric pressure controlled by
dielectric barrier J. Appl. Phys. 83 29507
[65] Massines F, Segur P, Gherardi N, Khamphan C and Ricard A
2003 Surf. Coat. Technol. 8 1745
[66] Wagner H E, Brandenburg R, Kozlov K V, Sonnenfeld A,
Michel P and Behnke J F 2003 The barrier discharge: basic
properties and applications to surface treatment Vacuum
71 41736
[67] Kogelschatz U 2002 Filamentary, patterned and diffuse barrier
discharges IEEE Trans. Plasma Sci. 30 14008
[68] Roth J R, Sherman D M and Wilkinson S P 2000
Electrohydrodynamic flow control with a glow discharge
surface plasma AIAA J. 38 11729
[69] Pons J, Moreau E and Touchard G 2004 Electrical and
aerodynamic characteristics of atmospheric pressure barrier
discharges in ambient air Proc. ISNTPT-2004 (Florida,
USA, May 2004) pp 30710
[70] Pons J, Moreau E and Touchard G 2005 Asymmetric surface
barrier discharge in air at atmospheric pressure: electric
properties and induced airflow characteristics J. Phys. D :
Appl. Phys. 38 363542
[71] Seraudie A, Aubert E, Naude N and Cambronne J P 2006
Effect of plasma actuators on a flat plate laminar boundary
layer in subsonic conditions AIAA Meeting (San Francisco,
USA, June 2006) paper #2006-3350
[72] Enloe C L, McLaughlin T E, Font G I and Baughn J W
Frequency effect on the efficiency of the aerodynamic
plasma actuator AIAA Meeting (Reno, USA, January 2006)
paper #2006-166
[73] Orlov D M, Corke C and Patel M P Electric circuit model for
aerodynamic plasma actuator AIAA Meeting (Reno, USA,
January 2006) paper #2006-1206
[74] Enloe C L, McLaughlin T E, VanDyken R D and Fischer J C
2004 Plasma structure in the aerodynamic plasma actuator
AIAA Meeting (Reno, USA, January 2004) paper
#2004-0844
[75] Enloe C, McLaughlin T E, VanDyken R D, Kachner K D,
Jumper E J and Corke T C 2004 Mechanisms and responses
of a single dielectric barrier plasma actuator: plasma
morphology AIAA J. 42 58994
[76] Enloe C L, McLaughlin T E, VanDyken R D, Kachner K D,
Jumper E J, Corke T C, Post M and Haddad O 2004
Mechanisms and responses of a single dielectric
barrier plasma actuator: geometric effect AIAA J. 42
595604
[77] Forte M, Jolibois J, Moreau E, Touchard G and Cazalens M
2006 Optimization of a dielectric barrier discharge actuator
by stationary and instationary measurements of the induced
flow velocity, application to airflow control AIAA
Meeting (San Francisco, USA, June 2006) paper
#2006-2863
[78] Roth J R and Dai X 2006 Optimization of the aerodynamic
plasma actuator as an EHD electrical device AIAA Meeting
(Reno, USA, January 2006) paper #2006-1203
[79] Roth J R, Dai X, Rahel J and Shermann M 2005 The physics
and phenomenology of paraelectric one atmosphere glow
discharge plasma actuators for aerodynamic flow
control AIAA Meeting (Reno, USA, January 2005) paper
#2005-781

Airflow control by non-thermal plasma actuators

[80] Roth J R, Dai X, Rahel J and Shermann M 2005 The physics


and phenomenology of one atmosphere glow discharge
plasma reactors for surface treatment applications J. Phys.
D: Appl. Phys. 38 55567
[81] Rivir R, White A, Carter C and Ganguly B 2004 AC and
pulsed plasma flow control AIAA Meeting (Reno, USA,
January 2004) paper #2004-0847
[82] Borghi C A, Carraro M R and Cristofolini A 2005
Experimental investigations on the EHD interaction on a
flat panel barrier discharge Proc. 15th ICMHD 186-191
(Moscow, Russia; May 2005)
[83] Hong D, Dong B, Bauchire J M and Pouvesle J M 2006
Experimental study of a dielectric barrier discharge
dedicated to airflow control Proc. 5th ISNTPT (Ile
dOleron, France, June 2006)
[84] Baughn J W, Porter C O, Peterson B L, McLaughlin T E,
Enloe C L, Font G I and Baird C 2006 Momentum transfer
for an aerodynamic plasma actuator with an imposed
boundary layer AIAA Meeting (Reno, USA, January 2006)
paper #2006-168
[85] Pons J, Moreau E and Touchard G 2004 Surface DC corona
discharge and AC barrier discharge in ambient air at
atmospheric pressure: measurements of the induced ionic
wind velocity 15th Conf. on Gas discharge and their
Applications (Toulouse, France, September 2004)
[86] Van Dyken R, McLaughlin T M and Enloe C L 2004
Parametric investigations of a single dielectric barrier
plasma actuator AIAA Meeting (Reno, USA, January 2004)
paper #2004-0846
[87] Porter C O, Baughn J W, McLaughlin T E, Enloe C L and
Font G I 2006 Temporal measurements on an aerodynamic
plasma actuator AIAA Meeting (Reno, USA, January 2006)
paper #2006-104
[88] Shyy W, Jayaraman B and Andersson A 2002 Modeling of
glow discharge-induced fluid dynamics J. Appl. Phys.
92 643443
[89] Singh K P, Roy S and Gaitonde D V 2006 Modeling of
dielectric barrier discharge plasma actuator with
atmospheric air chemistry AIAA Meeting (San Francisco,
USA, June 2006) paper #2006-3381
[90] Boeuf J P and Pitchford L C 2005 Electrohydrodynamic force
and aerodynamic flow acceleration in surface dielectric
barrier discharge J. Appl. Phys. 97 103307
[91] Boeuf J P, Lagmich Y, Callegari T and Pitchford L C 2006
Electrohydrodynamic force and acceleration in surface
discharges AIAA Meeting (San Francisco, USA, June 2006)
paper #2006-3374
[92] Roth J R 2003 Phys. Plasmas 10 211726
[93] Corke T C and Matlis E 2000 Phased plasma arrays for
unsteady flow control AIAA Meeting (Denver, USA, June
2000) paper #2000-2323
[94] Wilkinson S P 2003 Investigation of an oscillating surface
plasma for turbulent drag reduction AIAA Meeting (Reno,
USA, January 2003) paper #2003-1023
[95] Jukes T N, Choi K S, Jonhson G A and Scott S J 2004
Turbulent boundary layer control for drag reduction using
surface plasma AIAA Meeting (Portland, USA, June 2004)
paper #2004-2216
[96] Santhanakrishnan A and Jacob J D 2006 On plasma synthetic
jet actuators AIAA Meeting (Reno, USA, January 2006)
paper #2006-0317
[97] Santhanakrishnan A and Jacob J D 2006 Flow control using
plasma actuators and linear/annular synthetic jet actuators
AIAA Meeting (San Francisco, USA, June 2006) paper
#2006-3033
[98] Louste C, Artana G, Moreau E and Touchard G 2005 Sliding
discharge in air at atmospheric pressure: electrical
properties J. Electrost. 63 615620
[99] Louste C, Moreau E and Touchard G 2006 Sliding discharge
in air at atmospheric pressure: mechanical properties
Proc. ESA-IAS-IEJ Joint Symp. (Berkeley, USA, June
2006)

[100] Arad B, Gazit Y and Ludmirsky A 1987 A sliding discharge


device for producing cylindrical shock waves J. Phys. D:
Appl. Phys. 20 3607
[101] Laqua H, Bluhm H, Buth L and Hoppe P 1995 Properties of
the nonequilibrium plasma from a pulsed sliding discharge
in a hydrogen gas layer desorbed from a metal hybrid film J.
Appl. Phys. 77 554552
[102] Tsikrikas G N and Serafetinides A A 1996 The effect of
voltage pulse polarity on the performance of a sliding
discharge pumped HF laser J. Phys. D: Appl. Phys.
29 280610
[103] Samimy M, Adamovich I, Webb B, Kastner J, Hileman J,
Keshav S and Palm P 2004 AIAA Meeting (Reno, USA,
January 2004) paper #2004-0184
[104] Samimy M, Adamovich I, Kim J H, Webb B, Keshav S and
Utkin Y 2004 AIAA Meeting (Portland, USA, June 2004)
paper #2004-2130
[105] Samimy M, Adamovich I, Webb B, Kastner J, Hileman J,
Keshav S and Palm P 2004 Development and
characterization of plasma actuators for high-speed jet
control Exp. Fluids 37 57788
[106] Cybik B Z, Wilkerson J T and Grossman K R 2004
Performance characteristics of the sparkjet flow control
actuator AIAA Meeting (Portland, USA, June 2004) paper
#2004-2131
[107] Bletzinger P, Ganguly B N, Van Wie D and Garscadden
2005 Plasmas in high speed aerodynamics J. Phys. D
38 R3357
[108] Font G I 2006 Boundary layer control with atmospheric
plasma discharges AIAA J. 44 15728
[109] Hall K D, Jumper E J, Corke T C and McLaughlin T E 2005
Potential flow model of a plasma actuator as a fifth
enhancement device AIAA Meeting (Reno, USA, January
2005) paper #2005-0783
[110] Van Ness II D K, Corke T C and Morris S C 2006 Turbine tip
clearance flow control using plasmas actuators AIAA
Meeting (Reno, USA, January 2006) paper #2006-0021
[111] Visbal M R, Gaitonde D V and Roy S 2006 Control of
transitional and turbulent flows using plasma-based
actuators AIAA Meeting (San Francisco, USA, June 2006)
paper #2006-3230
[112] Borghi C A, Carraro M and Cristofolini A 2005 Plasma
characterization of OAUBD AIAA Meeting (Reno, USA,
January 2005) paper #2005-1179
[113] Borghi C A, Cristofolini A, Carraro M and Neretti G
2006 An analysis of a three phase flat panel uniform
barrier discharge at atmospheric pressure AIAA
Meeting (San Francisco, USA, June 2006) paper
#2006-3380
[114] Hultgren L S and Ashpis D E 2003 Demonstration of
separation delay with glow-discharge plasma actuators
AIAA Meeting (Reno, USA, January 2003) paper
#2003-1025
[115] Jacob J, Rivir C, Carter C and Estevadeordal J 2004
Boundary layer flow control using AC discharge plasma
actuators AIAA Meeting (Portland, USA, June 2004) paper
#2004-2128
[116] Boxx I G, Rivir R B, Newcamp J M and Woods N M 2006
Reattachment of a separated boundary layer on a flat plate
in high adverse pressure gradient using a plasma actuator
AIAA Meeting (San Francisco, USA, June 2006) paper
#2006-3023
[117] Labergue A, Leger L, Moreau E, Touchard G and Bonnet J P
2004 Detachment and reattachment of a low velocity
airflow along an inclined wall actively controlled by a low
frequency square wave corona discharge Electrostatics
2003 (Institute of Physics Conference Series vol 178)
(Bristol: Institute of Physics Publishing) pp 27984
[118] Jukes T N, Choi K S, Jonhson G A and Scott S J 2006
Turbulent drag reduction by surface plasma through
spanwise flow oscillation AIAA Meeting (San Francisco,
USA, June 2006) paper #2006-3693

635

E Moreau

[119] Thomas F O, Kozlov A and Corke T C 2006 Plasma


actuators for bluff body flow control AIAA Meeting (San
Francisco, USA, June 2006) paper #2006-2845
[120] Artana G, Sosa R, Moreau E and Touchard G 2003 Control of
the near-wake flow around a circular cylinder with
electrohydrodynamic actuators Exp. Fluids 35 5808
[121] McLaughlin T E, Munska M D, Vaeth J P, Dauwalter T E,
Goode J R and Siegel S G 2004 Plasma-based actuators for
cylinder wake vortex control AIAA Meeting (Portland, USA,
June 2004)
[122] Sung Y, Kim W, Mungal M G and Cappeli M A 2006
Aerodynamic modification of flow bluff objects by plasma
actuation Exp. Fluids 41 47986
[123] Asghar A and Jumper E J 2004 Phase synchronized of vortex
shedding from two side-by-side circular cylinders using
plasma actuators AIAA Meeting (Portland, USA, June 2004)
[124] McLaughlin T E, Felker B, Avery J C and Enloe C L 2006
Further experiments in cylinder wake modification with
dielectric barrier discharge forcing AIAA Meeting (Reno,
USA, January 2006) paper #2006-1409
[125] Shcherbakov Y V, Isanov N S, Baryshev N D, Frolovskij V S
and Syssoev V S 2000 AIAA Meeting (Reno, USA, January
2000) paper #2000-2670
[126] Post M L and Corke T C 2003 Separation control on high
angle of attack airfoil using plasma actuators AIAA Meeting
(Reno, USA, January 2003) paper #2003-1024
[127] Post M L and Corke T C 2004 Separation control using
plasma actuatorsstationary and oscillating airfoils AIAA
Meeting (Reno, USA, January 2004) paper #2004-0841
[128] Post M L and Corke T C 2004 Separation control using
plasma actuatorsdynamic stall control on an oscillating
airfoil AIAA Meeting (Portland, USA, June 2004) paper
#2004-2517
[129] Corke T C, He C and Patel M P 2004 Plasma flaps and slats:
an application of weakly-ionized plasma actuators AIAA
Meeting (Portland, USA, June 2004) paper #2004-2127
[130] Post M L and Corke T C 2005 Overview of plasma flow
control: concepts, optimization and applications AIAA
Meeting (Reno, USA, January 2005) paper #2005-0563
[131] Corke T C, Mertz B and Patel M P 2006 Plasma flow control
optimized airfoil AIAA Meeting (Reno, USA, January 2006)
paper #2006-1208
[132] Sosa R, Moreau E, Touchard G and Artana G 2004 Stall
control at high angle of attack with periodically excited
EHD actuators AIAA Meeting (Portland, USA, June 2004)
paper #2004-2738
[133] Sosa R and Artana G 2006 Steady control of laminar
separation over airfoils with plasma sheet actuators J.
Electrost. 64 60410
[134] Sosa R, Moreau E, Touchard G and Artana G 2006 Stall
control at high angle of attack with plasma sheet actuators
Exp. Fluids 42 14367
[135] Opaits D F, Roupassov D V, Starikovskaia A Y, Zavialov I N
and Saddoughi S G 2005 Plasma control of boundary layer
using low-temperature non equilibrium plasma of gas
discharge AIAA Meeting (Reno, USA, January 2005) paper
#2005-1180
[136] Zavialov I N, Opaits D F, Roupassov D V, Starikovskaia A Y
and Saddoughi S G 2005 Boundary layer control for
naca-0015 airfoil in subsonic regime Proc. 15th ICMHD
(Moscow, Russia, May 2005) pp 18691

636

[137] Zavyalov I N, Roupassov D V, Starikovskii A Y and


Saddoughi S G 2005 Boundary layer control by gas
discharge plasma Proc. EUCASS 2005 (Moscow, Russia,
July 2005)
[138] Roupassov D V, Zavyalov I N and Starikovskii A Y 2006
Boundary layer separation plasma control using
low-temperature non-equilibrium plasma of gas discharge
AIAA Meeting (Reno, USA, January 2006) paper
#2006-373
[139] Jolibois J, Forte M and Moreau E IUTAM Symp. Flow
Control and Mems (London, England, September
2006)
[140] Van Dycken R, Perez-Blanco H, Byerley A and McLaughlin
T 2004 Plasma actuator for wake flow control of high
camber blades during part load operation Proc. ASME
GT2004-53227 (Vienna, Austria, June 2004)
[141] Morris S C, Corke T C, VanNess D, Stephens J and Douville
T 2005 Tip clearance control using plasma actuators AIAA
Meeting (Reno, USA, January 2005) paper #2005-0782
[142] Labergue A, Leger L, Moreau E, Touchard G and Bonnet J P
2003 Effect of a corona discharge on a plane mixing layer
Proc. IEEE-IAS World Conf. on Industrial Applications of
Electrical Energy (Little Rock, USA, June 2003)
pp 17286
[143] Labergue A, Leger L, Moreau E, Touchard G and Bonnet J P
2004 Experimental study of the detachment and the
reattachment of an airflow along an inclined wall by a
surface corona dischargeApplication to a plane turbulent
mixing layer IEEE Trans. Indust. Appl. 40 120514
[144] Labergue A, Moreau E, Younis G and Touchard G 2005
Plasma applications in aerospace. Recent advances in
airflow control Proc. Workshop on Plasma Technology and
Applications (Cebu City, Philippines, December 2005)
pp 416
[145] Labergue A, Moreau E and Touchard G 2006 DC corona
discharge and AC barrier discharge for separation control in
the case of a turbulent jet diffuser Proc. ESA-IEEE-IEJ
(Berkeley, USA, June 2006)
[146] Jolibois J, Labergue A, Moreau E and Touchard G 2006 Proc.
ISNTPT-2006 (Ile dOleron, France, June 2006)
[147] Labergue A, Moreau E, Zouzou N and Touchard G 2007
J. Phys. D: Appl. Phys. (Special issue)
[148] Vorobiev A N, Rennie R M, Jumper E J, McLaughlin T E
2006 An experimental investigation of lift enhancement and
roll control using plasma actuators AIAA Meeting (San
Francisco, USA, June 2006) paper #2006-3383
[149] Patel M P, Ng T T and Vasudevan S 2006 Plasma actuators
for hingeless aerodynamic control of unmanned air vehicle
AIAA Meeting (San Francisco, USA, June 2006) paper
#2006-3495
[150] Goeksel B and Rechenberg I 2004 Active separation flow
control experiments in weakly-ionized gas Proc. 10th
European Turbulence Conf. (Barcelona, Spain)
[151] Sosa R and Artana G 2004 Steady control of laminar
separation over airfoils with plasma sheet actuators Proc.
SFE 2004 (Poitiers, France, September 2004)
[152] Menier E, Lago V, Depussay E, Leger L and Martin J P 2005
Emission spectroscopy in a mach 2 air flow with an
electrical discharge Proc. 6th Workshop on Frontiers in Low
Temperature Plasma Diagnostics (Les Houches, France,
April 2005)

Вам также может понравиться