You are on page 1of 241

High-Lift Airfoil Separation Control with Dielectric Barrier Discharge Plasma Actuators

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy
in the Graduate School of The Ohio State University

By

Jesse C. Little

Graduate Program in Mechanical Engineering

The Ohio State University

2010

Dissertation Committee:

Igor Adamovich

Michael Dunn

James Gregory

Mo Samimy, Advisor

Andrea Serrani
Copyright by

Jesse C. Little

2010
Abstract

This work examines the performance of dielectric barrier discharge (DBD)

plasma actuators for controlling separation from the leading edge and trailing edge flap

shoulder of a supercritical high-lift airfoil. DBD plasma actuators driven by both typical

AC voltages (AC-DBD) and more developmental nanosecond duration pulses (NS-DBD)

are investigated. Characterization of the two actuators shows that very different behavior

is created when exciting the plasma discharge using these two waveforms. The AC-DBD

plasma actuator functions through electrohydrodynamic effects that introduce zero net

mass, but nonzero net momentum into the flow. Conversely, the electrohydrodynamic

effects of the NS-DBD are quite weak suggesting thermal effects from rapid localized

heating by the plasma are responsible for control authority. The performance of both

devices as separation control actuators is tested on a high-lift airfoil system.

The AC-DBD is effective for controlling turbulent boundary layer separation

from a deflected trailing edge flap between Reynolds numbers of 240,000 and 750,000.

Momentum coefficients for the AC-DBD plasma actuator are generally an order of

magnitude lower than those usually employed for such studies yet control authority is

still realized through amplification of natural vortex shedding from the flap shoulder. The

corresponding lift enhancement is primarily due to upstream effects from increased

circulation around the entire model rather than full separated flow reattachment to the

ii
deflected flap surface. Lift enhancement via instability amplification is found to be

relatively insensitive to changes in angle of attack provided that the separation location

and underlying dynamics do not change. Control authority decreases with increasing

Reynolds number and flap deflection highlighting the necessity for further optimization

of AC-DBD plasma actuators for use in realistic takeoff and landing transport aircraft

applications. As a whole, these findings compare favorably to studies on a similar high-

lift platform using piezoelectric driven zero net mass flux actuation.

The NS-DBD plasma actuator is ineffective for controlling separation from the

deflected trailing edge flap. However, the device is found to be superior to the tested AC-

DBD plasma actuators for controlling leading edge separation and rivals the performance

of a passive droop by extending the stall angle by six degrees in the Reynolds number

range 750,000-1,000,000. Detailed flow diagnostics show the NS-DBD plasma actuator

functions as an active trip for pre-stall incidence angles and generates coherent spanwise

vortices that entrain freestream momentum into the separated region at post-stall angles.

These structures are generated across all surveyed frequencies, but optimal dimensionless

frequencies for controlling separation are in the range four to six depending on the

incidence angle. The contrasting performance of the NS-DBD plasma actuator at the

leading and trailing edge in comparison to the AC-DBD is discussed and

recommendations for future work are provided.

iii
Acknowledgments

The world class facilities, technical guidance, interesting projects and professional

development opportunities provided to me by Prof. Mo Samimy have been invaluable. I

will be forever grateful for his tutelage. I wish to acknowledge Prof. Igor Adamovich for

being the driving force behind much of the NS-DBD plasma work. Thanks to my

dissertation committee for providing comments that have improved the quality of this

manuscript and challenged me to explore additional related avenues of study. I have had

the opportunity to work with incredible people during my time at the OSU GDTL. A list

of these names would require its own appendix. The stimulating discussions and unique

ideas proposed by these individuals have substantially influenced my growth as a

researcher. I will always be thankful for these interactions and the friendships that have

subsequently developed. I wish to specifically acknowledge Dr. Munetake Nishihara and

Dr. Keisuke Takashima for acquiring the electrical measurements and schlieren images

used in the actuator characterization as well as their work on the NS-DBD power supply

development. Thanks to Kristine McElligott for assisting in the wind tunnel experiments.

A very special recognition is reserved for my family and especially my wife, Misty.

Thank you for everything. Aspects of this work have been supported by AFRL, DAGSI,

The Boeing Company and the Howard D. Winbigler Professorship at OSU.

iv
Vita

October 24, 1980 ………………………………………………. Born Athens, OH, USA

2002-2004……………………………………………Undergraduate Research Associate


Department of Mechanical Engineering
The Ohio State University

2003-2004…………………………………………………………………Research Intern
Aerosol and Process Technologies
Battelle Memorial Institute
Columbus, OH, USA

2004…………………………………………………………B.S. Mechanical Engineering


The Ohio State University

2005…………………………………………………………M.S. Mechanical Engineering


The Ohio State University

2009…………………………………………………………Graduate Teaching Associate


Department of Mechanical Engineering
The Ohio State University

2004-present…………………………………………………Graduate Research Associate


Department of Mechanical Engineering
The Ohio State University

Publications

Book Chapters
1. Samimy, M., Debiasi, M., Caraballo, E., Serrani, A., Yuan, X., Little, J., and Myatt,
J., “Reduced-order Model-based Feedback Control of Subsonic Cavity Flows,” In
Active Flow Control, R. King, editor, Volume 95 of Notes on Numerical Fluid
Mechanics and Multidisciplinary Design (NNFM), pages 211-229. Springer Verlag,
2007. ISBN: 978-3-540-71438-5.

v
Refereed Journal Articles
1. Little, J., and Samimy, M., “Control of Separation from the Flap of a High-Lift
Airfoil using DBD Plasma Actuation,” submitted to AIAA Journal, January 2010.
2. Little, J., Nishihara, M., Adamovich, I. and Samimy, M., “High-Lift Airfoil Trailing
Edge Separation Control Using a Single Dielectric Barrier Discharge Plasma
Actuator,” Experiments in Fluids, DOI 10.1007/s00348-009-0755-x, 2009.
3. Yuan, X., Caraballo, E., Little, J., Debiasi, M., Serrani, A., Ozbay, H., Myatt, J. and
Samimy, M., “Feedback Control Design for Subsonic Cavity Flows,” Applied and
Computational Mathematics, Vol. 8, No. 1, 2009, pp. 70-91.
4. Malone, J., Debiasi, M., Little, J. and Samimy, M., “Analysis of the Spectral
Relationships of Cavity Tones in Subsonic Resonant Cavity Flows,” Physics of
Fluids, Vol. 21, No. 055103, 2009.
5. Caraballo, E., Little, J., Debiasi, M., and Samimy, M., “Development and
Implementation of an Experimental Based Reduced-order Model for Feedback
Control of Subsonic Cavity Flows,” Journal of Fluids Engineering, Vol. 129, No. 7,
pp. 813-824, 2007.
6. Little, J., Debiasi, M., Caraballo, E., and Samimy, M., “Effects of Open-loop and
Closed-loop Control on Subsonic Cavity Flows,” Physics of Fluids, Vol. 19, No. 6,
065104, 2007.
7. Samimy, M., Debiasi, M., Caraballo, E., Serrani, A., Yuan, X., Little, J., and Myatt,
J. H., “Feedback Control of Subsonic Cavity Flows Using Reduced-order Models,”
Journal of Fluid Mechanics, Vol. 579, pp. 315-346, 2007.
8. Yan. P., Debiasi, M., Yuan, X., Little, J., Özbay, H., and Samimy, M., “Experimental
Study of Linear Closed-Loop Control of Subsonic Cavity Flow,” AIAA Journal, Vol.
44, No. 5, pp. 929-938, 2006.

Fields of Study

Major Field: Mechanical Engineering

Studies in: Aerodynamics, Experimental Techniques, Flow Control, Fluid Mechanics,


Optical Diagnostics, Turbulence

vi
Table of Contents

Abstract ............................................................................................................................... ii

Acknowledgments.............................................................................................................. iv

Vita...................................................................................................................................... v

Publications ......................................................................................................................... v

Fields of Study ................................................................................................................... vi

Table of Contents .............................................................................................................. vii

List of Tables ...................................................................................................................... x

List of Figures .................................................................................................................... xi

Nomenclature ................................................................................................................... xxi

Chapter 1: Introduction ....................................................................................................... 1

Chapter 2: Background ....................................................................................................... 5

2.1 Separation Control..................................................................................................... 5

2.1.1 LE Separation Control ........................................................................................ 6

2.1.2 TE Separation Control ........................................................................................ 8

2.2 Plasma Actuators ..................................................................................................... 10

vii
2.2.1 AC-DBD Plasma Actuators .............................................................................. 10

2.2.2 NS-DBD Plasma Actuators .............................................................................. 17

2.3 Separation Control with Dielectric Barrier Discharge Plasma Actuators ............... 20

Chapter 3: Experimental Facilities and Measurement Techniques................................... 24

3.1 Wind Tunnel............................................................................................................ 24

3.2 Airfoil Model........................................................................................................... 36

3.3 Plasma Actuator Hardware...................................................................................... 40

3.4 Diagnostics .............................................................................................................. 43

3.4.1 Static Pressure................................................................................................... 43

3.4.2 Fluctuating pressure .......................................................................................... 43

3.4.3 PIV .................................................................................................................... 49

3.4.4 Accuracy ........................................................................................................... 57

3.5 Test Conditions ....................................................................................................... 59

Chapter 4: Dielectric Barrier Discharge Plasma Actuators .............................................. 60

4.1 AC DBD Plasma Actuator Design .......................................................................... 60

4.2 AC-DBD Plasma Actuator Characterization........................................................... 66

4.3 NS DBD Plasma Actuator Design .......................................................................... 87

4.4 NS-DBD Plasma Actuator Characterization ........................................................... 87

Chapter 5: Baseline Verification and Characterization .................................................. 103

viii
Chapter 6: Trailing Edge Separation Control ................................................................. 119

6.1 Single AC-DBD Plasma Actuator Control Results ............................................... 119

6.2 Perspective on OSU Single AC-DBD Control Results ......................................... 147

6.3 Attempts to Extend AC-DBD Plasma Control Authority ..................................... 148

6.4 Single NS-DBD Separation Control Results......................................................... 153

Chapter 7: Leading Edge Separation Control ................................................................. 156

7.1 Single DBD Plasma Actuator Control Results: Downstream Orientation ............ 157

7.2 Single DBD Plasma Actuator Control Results: Upstream Orientation ................. 174

7.3 Comparison of Single NS-DBD Plasma Actuators to Passive Flow Control ....... 196

Chapter 8: Summary and Conclusions............................................................................ 199

Chapter 9: Discussion of Future Work ........................................................................... 205

References ....................................................................................................................... 208

ix
List of Tables

Table 3.1: PIV Uncertainty ............................................................................................... 58

Table 4.1: Properties of various dielectric materials (Roth and Dai 2006). ..................... 64

Table 4.2: Momentum and power characteristics for 3 kHz actuation at 20 kVpp............ 80

Table 4.3: Momentum and power characteristics for 3 kHz actuation at 20 kVpp with

various burst frequencies at dc=50%. ............................................................................... 81

Table 4.4: Electrical properties of NS-DBD plasma on a 46 cm long flat plate actuator. 90

Table 4.5: Electrical properties of AC-DBD plasma on a 46 cm long flat plate actuator. 91

Table 5.1: Boundary layer properties at x/c=0.70 measured using PIV. ........................ 118

x
List of Figures

Figure 1.1: Simplified high-lift airfoil system with deflected LE droop and TE simple

flap. ..................................................................................................................................... 2

Figure 1.2: Typical high-lift airfoil system with deflected multi-element LE slat and TE

flap (Lin and Dominik 1997). ............................................................................................. 2

Figure 1.3: Further simplified high-lift airfoil system with deflected TE flap. .................. 3

Figure 2.1: Typical asymmetric DBD plasma actuator geometry (Corke et al. 2010). .... 11

Figure 2.2: Simultaneous traces of voltage and current for a typical AC-DBD plasma

actuator. ............................................................................................................................. 12

Figure 2.3: High-speed photography of the forward stroke (a) and reverse/back stroke (b)

of a typical AC-DBD plasma actuator (Enloe et al. 2008a). ............................................ 14

Figure 2.4: Schematic of negatively charged species movement for the forward (a) and

reverse/back stroke (b) in a typical AC-DBD plasma actuator (Enloe et al. 2008a). ....... 14

Figure 2.5: Visual appearance of a typical ~10 cm long AC-DBD plasma actuator

operating in the normal glow regime (a) and at maximum thrust (b) (Thomas et al. 2009)

........................................................................................................................................... 15

Figure 3.1: Axial fan used in the subsonic recirculating wind tunnel. ............................. 25

Figure 3.2: Operator keypad for the subsonic recirculating wind tunnel. ........................ 26

Figure 3.3: Subsonic recirculating wind tunnel contraction section. ................................ 26

xi
Figure 3.4: Flow conditioning screens at the subsonic recirculating wind tunnel

contraction entrance. ......................................................................................................... 27

Figure 3.5: Safety catch screens downstream of the subsonic recirculating wind tunnel

test section......................................................................................................................... 27

Figure 3.6: Turning cascades used in the elbows of the subsonic recirculating wind

tunnel................................................................................................................................. 28

Figure 3.7: Gravity filter, heat exchanger, modulating valve, flow meter and drain line on

the subsonic recirculating wind tunnel. ............................................................................ 29

Figure 3.8: Test section arrangement with wall plugs on both infield and outfield sides. 31

Figure 3.9: Modified pitot-static probe assembly. ............................................................ 32

Figure 3.10: Dimensionless dynamic pressure profiles measured 0, 30.5, 61 and 85 cm (0,

12, 24 and 33.5 in) (left to right) from the test section inlet. ............................................ 34

Figure 3.11: Near-wall dimensionless dynamic pressure profiles measured at 0, 30.5, 61

and 85 cm (0, 12, 24 and 33.5 in) (left to right) from the test section inlet. ..................... 34

Figure 3.12: Modular instrument panel for the subsonic recirculating wind tunnel. ........ 36

Figure 3.13: 2D profile of the airfoil in cruise configuration showing the approximate

location of static pressure taps and high bandwidth pressure transducers near the airfoil

centerline. .......................................................................................................................... 38

Figure 3.14: OSU version of the simplified high-lift EET airfoil with TE flap deflected.39

Figure 3.15: Model bending produced by loading at high Re and α with DBD plasma at

the LE. ............................................................................................................................... 40

Figure 3.16: Setup used for distinguishing flow phenomena from EMI. ......................... 47

xii
Figure 3.17: Example of pressure spectra used for distinguishing flow phenomena from

EMI. .................................................................................................................................. 47

Figure 3.18: Example of randomly occurring voltage spikes in pressure traces due to

EMI. .................................................................................................................................. 48

Figure 3.19: PIV experimental setup for LE airfoil measurements. ................................. 50

Figure 3.20: Example of laser reflections from the model surface and their effect on PIV

data. ................................................................................................................................... 52

Figure 3.21: Sample raw PIV data images for LE (a) and TE (b) experiments. ............... 52

Figure 3.22: Experimental setup for flat plate PIV measurements. .................................. 54

Figure 4.1: Airfoil damage due to arc formation. ............................................................. 63

Figure 4.2: Degradation of the dielectric surface after substantial plasma run-time. ....... 66

Figure 4.3: Simultaneous current and voltage traces for a typical AC-DBD plasma

actuator. ............................................................................................................................. 67

Figure 4.4: Example of Q-V data for AC-DBD plasma. .................................................. 69

Figure 4.5: Dissipated power per unit length as a function of applied voltage. ............... 70

Figure 4.6: AC-DBD time-averaged induced velocity magnitude, W , in quiescent air

(m/sec)............................................................................................................................... 72

Figure 4.7: Modulation waveforms for AC-DBD plasma actuation................................. 73

Figure 4.8: Phase-averaged AC-DBD plasma induced velocity fields, V , (left, m/sec) and

 , (right sec-1) in quiescent air for modulation using AM (top) and various BM
vorticity, Ω

(2nd from top to bottom) waveforms. ................................................................................ 75

xiii
Figure 4.9: Mean (a) and rms (b) velocity profiles, U, at x=20 mm for AC-DBD plasma

operating in quiescent air for various BM frequencies at 50% dc. ................................... 78

Figure 4.10: Exponential profiles used to interpolate momentum values for various AC-

DBD plasma modulation frequencies. .............................................................................. 80

Figure 4.11: Mean and oscillatory momentum values for the AC-DBD plasma induced

flow in quiescent air for modulation using AM and various BM waveforms. ................. 82

Figure 4.12: Q-V diagram for AC-DBD plasma actuation using AM at 100 Hz. ............ 84

Figure 4.13: AC-DBD plasma integrated energies for various modulation waveforms... 86

Figure 4.14: Typical voltage, current (a) and power, energy (b) traces for an NS-DBD

plasma actuator. ................................................................................................................ 89

Figure 4.15: Startup vortex (b) and dielectric charging effect (c) created by NS-DBD

plasma actuator. ................................................................................................................ 93

Figure 4.16: Mean velocity profiles, U, for NS and AC-DBD plasma operating in

quiescent air with 2 kHz carrier frequency using 100% (a) and 50% (dc) at 90 Hz

measured 20 mm downstream. ......................................................................................... 95

Figure 4.17: Schlieren imaging of compression waves generated by NS-DBD plasma

actuator viewed along the major axis of the actuator. ...................................................... 97

Figure 4.18: Schlieren photography of vertical density gradients for NS-DBD plasma

generated compression waves viewed along the minor axis of the actuator. ................... 98

Figure 4.19: Schlieren photography of horizontal density gradient for NS-DBD plasma

generated compression waves viewed along the minor axis of the discharge. ................. 98

Figure 5.1: Simplified high-lift airfoil with LE droop and TE flap. ............................... 103

xiv
Figure 5.2: Baseline CL vs. α for sample Re and δf. ........................................................ 105

Figure 5.3: Baseline CP behavior for Re=410k and δf=20o at various α. ........................ 105

Figure 5.4: Baseline PSD of fluctuating pressure, cp, measured at x/c=0.95 for Re=410k

and α=0o at various δf. ..................................................................................................... 106

Figure 5.5: Baseline PSD of fluctuating pressure, cp, measured at x/c=0.95 for Re=410k

and δf=20o at various α. ................................................................................................... 107

Figure 5.6: Baseline PSD of fluctuating pressure, cp, measured at x/c=0.90 for α=0o and

δf=30o at various Re. ....................................................................................................... 107

Figure 5.7: Baseline PSD of fluctuating pressure, cp, measured at x/c=0.40 for Re=750k,

δf=0 at various α. ............................................................................................................. 108

Figure 5.8: Baseline comparison of OSU and NASA (Melton et al. 2006) results for

tripped and untripped behavior of CL vs. α for Re=750k and various δf. ........................ 109

Figure 5.9: OSU (a) and NASA (b) CP curves at Re=750k, α=0 and δf=0 (Melton et al.

2003). .............................................................................................................................. 109

Figure 5.10: OSU (a) and NASA (b) stall characteristics at Re=410k and δf =12o (Melton

2006). .............................................................................................................................. 110

Figure 5.11: Re and δf effects on OSU airfoil CP distributions....................................... 113

Figure 5.12: OSU (a) and NASA (b) PSD of fluctuating pressure, cp, at Re=240k, α =0, δf

=45, δs =-25 in (Melton et al. 2006). ............................................................................... 114

Figure 5.13: 3D behavior of CP for Re=410k, α=0 and δf=20 (a), δf =30 (b) and δf =40 (c).

......................................................................................................................................... 115

xv
Figure 5.14: PIV-measured boundary layer velocity profile with power law fit for

Re=240k, α=0, δf=30. ...................................................................................................... 117

Figure 6.1: Example of a DBD plasma actuator applied at the flap shoulder (x/c=0.77) of

the simplified NASA EET airfoil, δf=30o. ...................................................................... 120

Figure 6.2: Effects of passive (power off) DBD plasma actuator on CP (a) and PSD of

fluctuating pressure, cp, at x/c=0.90 (b). ......................................................................... 121

Figure 6.3: Effect of dimensionless actuation frequencies FL,m+ (a) and F* (b) on ΔCL at

α=0o for various Re and δf. .............................................................................................. 123

Figure 6.4: Baseline and controlled CP behavior for Re=240k, α=0o and δf =20. .......... 125

Figure 6.5: Baseline and controlled (dc=50%) time-averaged streamlines for Re=240k,

α=0o and δf =20o. ............................................................................................................. 127

*
Figure 6.6: Baseline and controlled (dc=50%) time-averaged normalized vorticity, Ω ,

for Re=240k, α=0o and δf =20o. ....................................................................................... 129

*
Figure 6.7: Baseline and controlled time-averaged vorticity magnitude, Ω at x/c=1.1 for

Re=240k, α=0o and δf =20o. ............................................................................................ 130

 * , for Re=240k, α=0o


Figure 6.8: Phase-averaged normalized vorticity fields (ΔΦ=π), Ω

and δf =30o forced at FL+=8.5, FL,m+=0.4 (F*=1). .......................................................... 132

Figure 6.9: Baseline and controlled PSD of fluctuating pressure, cp, measured at x/c=0.90

Re=240k, α=0o and δf=30o. ............................................................................................. 133

Figure 6.10: Spatial correlation of the v component of velocity, Rvv, at y/c=yT for

Re=240k, α=0o and δf=30o for baseline and controlled (dc=50%) flows. ....................... 133

xvi
Figure 6.11: First four normalized baseline and controlled (AM) POD modes of the v

component of velocity, ϕ1− 4 ( y ) , for Re=240k, α=0o and δf=30o. ................................. 136

Figure 6.12: Modal energy (a) and cumulative energy (b) for baseline and controlled

(AM) POD modes for Re=240k, α=0o and δf=30o. ......................................................... 137

Figure 6.13: Effect of modulation waveform on ΔCL for Re=240k, α=0o and variable δf

when forcing at FL+=8.5, F*=1. ...................................................................................... 139

Figure 6.14: ΔCL and <J/ρ> as a function of BM dc for Re=240k, α=0, δf=30, FL+=8.5,

FL,m+=0.4, F*=1. ............................................................................................................. 140

Figure 6.15: Effect of Re, α, and δf on ΔCL for forcing at F*=1 using dc=50%. ............ 143

Figure 6.16: Mean U velocity profiles (a) and Reynolds stresses uu (b) and vv (c) used

in CD calculations for Re=240k, α=0 and δf=20 at x/c=1.1, 1.2, 1.3, 1.4 and 1.5 (left to

right). ............................................................................................................................... 146

Figure 6.17: DBD plasma actuators straddling the flap shoulder. .................................. 149

*
Figure 6.18: Normalized time-averaged vorticity magnitude profiles, Ω at x/c=1.1 for

Re=240k, α=0, δf=20 for baseline and AC-DBD actuators operating in phase at FL+ =12.7

and F*=1. ........................................................................................................................ 149

Figure 6.19: Effect of reversing AC-DBD momentum direction on CP. ........................ 151

Figure 6.20: First four normalized baseline and controlled (AM) POD modes of the v

component of velocity, ϕ1− 4 ( y ) , for Re=240k, α=0o and δf=30o with reverse actuator at

x/c=0.77. .......................................................................................................................... 152

Figure 6.21: NS-DBD effect on CP for Re=240k, α=0 and δf=30................................... 154

xvii
Figure 6.22: NS-DBD effect on CP for Re=240k, α=0 and δf=30 with increased energy

input. ............................................................................................................................... 155

Figure 7.1: DBD plasma actuator mounted at the LE of the OSU simplified NASA EET

airfoil. .............................................................................................................................. 158

Figure 7.2: CL curves for the OSU airfoil with and without passive LE actuator and Re

overshoot compared to NASA (Melton et al. 2006). ...................................................... 159

Figure 7.3: Various examples of baseline (plasma off) behavior observed when an

actuator is applied to the LE. .......................................................................................... 161

Figure 7.4: Effect of DBD plasma at x/c=0 on CP for Re=750k (a) and Re=1000k (b) at

α=10o. .............................................................................................................................. 163

Figure 7.5: Effect of DBD plasma at x/c=0 on CP for Re=750k (a) and Re=1000k (b) at

α=12o. .............................................................................................................................. 164

Figure 7.6: Effect of DBD plasma at x/c=0 on CP for Re=750k (a) and Re=1000k (b) at

α=14o. .............................................................................................................................. 166

Figure 7.7: Effect of DBD plasma at x/c=0 on CP for Re=750k (a) and Re=1000k (b) at

α=16o. .............................................................................................................................. 168

Figure 7.8: Effect of NS-DBD plasma actuation at x/c=0 on CP for Re=750k at α=16o

showing preference for higher frequency forcing. .......................................................... 170

Figure 7.9: Comparison of the effects of NS (a) and AC (b) DBD plasma on CP for

Re=750k. ......................................................................................................................... 172

Figure 7.10: Effect of Fc+ of AC and NS-DBD plasma at x/c=0 on suction side CP at

x/c=0.05. .......................................................................................................................... 174

xviii
Figure 7.11: Photograph of the DBD plasma actuator mounted near the airfoil LE in

reverse arrangement ds=6 mm from x/c=0. .................................................................... 176

Figure 7.12: Effect of AC and NS-DBD plasma for reverse actuator arrangement ds=6

mm from x/c=0................................................................................................................ 177

Figure 7.13: Effect of Fc+ of NS-DBD plasma for reverse actuator arrangement ds=6 mm

from x/c=0 on suction side CP at x/c=0.05 at Re=750k. ................................................. 178

Figure 7.14: Effect of Fc+ of NS-DBD plasma at for reverse actuator arrangement ds=6

mm from x/c=0 on suction side CP at x/c=0.05 at Re=750k and Re=1000k. .................. 179

*
Figure 7.15: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline

(b) and NS-DBD forcing at Fc+=4 (c) at Re=750k, α=12o. ............................................. 181

Figure 7.16: Phase-averaged normalized velocity fluctuations, v* , for NS-DBD plasma

forcing at Fc+=4 (c) at Re=750k, α=12o. ......................................................................... 182

*
Figure 7.17: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline

(b) and NS-DBD forcing at Fc+=4 (c) at Re=750k, α=14o. ............................................. 183

Figure 7.18: Phase-averaged normalized velocity fluctuations, v* (a), swirling strength,

 * , (c) for NS-DBD forcing at Fc+=4 for Re=750k, α=14o. ...... 186
λci* , (b) and vorticity, Ω

*
Figure 7.19: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline

(b) and NS-DBD forcing at Fc+=0.6 (c) at Re=750k, α=16o. .......................................... 188

Figure 7.20: Phase-averaged normalized velocity fluctuations, v* (a) and swirling

strength, λci* , (b) for NS-DBD forcing at Fc+=0.6 for Re=750k, α=16o. ........................ 189

xix
*
Figure 7.21: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline

(b) and NS-DBD forcing at Fc+=5.6 (c) at Re=750k, α=16o. .......................................... 191

Figure 7.22: Phase-averaged normalized velocity fluctuations, v* (a) and swirling

strength, λci* , (b) for NS-DBD forcing at Fc+=5.6 for Re=750k, α=16o. ........................ 192

*
Figure 7.23: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline

(b) and NS-DBD forcing at Fc+=11.3 (c) at Re=750k, α=16o. ........................................ 194

Figure 7.24: Phase-averaged normalized velocity fluctuations, v* (a) and swirling

strength, λci* , (b) for NS-DBD forcing at Fc+=11.3 for Re=750k, α=16o. ...................... 195

Figure 7.25: Effect of AFC with NS-DBD plasma actuation on CL for the OSU airfoil.197

Figure 7.26: Effect of PFC with using LE droop on CL for the simplified NASA EET

airfoil (Melton et al. 2005) .............................................................................................. 198

xx
Nomenclature

a = POD modal amplitude


c = model chord, 25.4 cm
CD = sectional drag coefficient
Cdp = sectional pressure drag coefficient
CE = power coefficient, P/q∞U∞c
CL = sectional lift coefficient
CP = pressure coefficient
cP = fluctuating pressure coefficient
cˆP = Fourier transform of cP denoted by ^
Cv = Specific heat at constant volume
Cµ = momentum coefficient, J/q∞c
<Cµ> = oscillatory momentum coefficient, <J>/q∞c
Cµ,tot = total momentum coefficient, Cµ + <Cµ>
Cµ,Δp = modified momentum coefficient based on pressure difference, Δpz/q∞c
ds = arc length measured along airfoil suction surface from x/c=0
E = energy
f = frequency
fc = plasma carrier frequency
fm = modulation frequency
fTE = characteristic frequency measured on the flap
Fc+ = reduced frequency based on model chord, fc/U∞
FL+ = reduced frequency based on flap length, fL/U∞
FL,m+ = reduced modulation frequency based on flap length, fmL/U∞
F* = normalized modulation frequency, fm/fTE
h = dielectric thickness
J = time-averaged actuator induced momentum
<J> = oscillatory actuator induced momentum
k = wind tunnel calibration constant
l = actuator length
L = flap length, 6.35 cm
m = mass
N = sample size
P = time-averaged power consumed by plasma
p = static pressure
po = stagnation pressure
xxi
Δp = pressure difference across compression wave
Q = charge
q = dynamic pressure, ρU2/2
q∞ = freestream dynamic pressure, ρU∞2/2
Rvv = Normalized spatial correlation of v
Re = chord based Reynolds number, U∞c/ν
ΔT = temperature rise due to plasma heating
t = time
u = fluctuating streamwise velocity
U = streamwise velocity
U = time-averaged velocity denoted by _
U = phase-averaged velocity denoted by ~
U* = normalized velocity, U/ U∞
Urms = rms of streamwise velocity
U∞ = freestream velocity
Vac = AC voltage
v = fluctuating vertical velocity
V = instantaneous vertical velocity

W = 2D vector field [U,V]
xsp = length of separation region
x = two dimensional space (x,y)
x/c = normalized streamwise coordinate
y/c = normalized vertical coordinate
yT = y/c coordinate of the tail of the deflected trailing edge flap
z = width of fluid heated by plasma
α = angle of attack in degrees
δf = flap deflection angle in degrees
ε = dielectric constant
λci = swirling strength
λci* = normalized swirling strength, λcic/U∞
ρ = density
ν = kinematic viscosity
δ* = boundary layer displacement thickness
θ = boundary layer momentum thickness
σ = sample standard deviation
H = shape factor, δ*/ θ
Ω = spanwise component of vorticity
Ω* = normalized spanwise vorticity, Ωc/U∞
Φ = POD mode
Φ* = Normalized POD mode, Φ/U∞
CI = 95% confidence interval
PSD = power spectral density
TE = trailing edge

xxii
LE = leading edge
PFC = passive flow control
AFC = active flow control
AM = amplitude modulation with sine wave
BM = burst modulation with square wave
DBD = dielectric barrier discharge
AC = alternating current
DC = direct current
NS = nanosecond pulse
LAFPA = localized arc filament plasma actuator
EHD = electrohydrodynamic
EMI = electromagnetic interference
PIV = particle image Velocimetry
ZNMF = zero net mass flux
HV = high voltage
dc = duty cycle

xxiii
Chapter 1: Introduction

High-lift airfoils typically employ leading edge (LE) droops or slats and trailing

edge (TE) flaps that can be deflected during take-off or landing and stowed during cruise

(Figure 1.1). The former acts to extend the stall angle while the latter acts to increase the

airfoil camber by creating greater CL at a given α at the expense of decreasing the stall

angle (Hoener and Borst 1975). Consequently, the two devices are often used in tandem.

Simple flaps can impose a penalty due to flow separation that occurs when the

momentum of fluid in the boundary layer is not sufficient to overcome wall friction and

the adverse pressure gradient encountered as it travels over the deflected surface.

Traditional methods of eliminating flow separation on high-lift airfoils utilize multi-

element slats and flaps that allow mixing of fluid between the pressure (bottom) and

suction (top) sides like those shown in Figure 1.2. These systems are effective for

augmenting lift, but increase mechanical complexity, manufacturing cost, weight and

parasitic drag even when stowed during cruise. A system study indicates that significant

decreases in manufacturing cost, weight and drag could be realized if the complex multi-

element high-lift system is simplified like Figure 1.1 while maintaining CL,max (McLean

et al. 1999). To obtain similar performance, passive flow control (PFC) previously in the

1
form of slots must be replaced by some active flow control (AFC) device that would not

reintroduce additional detrimental factors.

Figure 1.1: Simplified high-lift airfoil system with deflected LE droop and TE simple
flap.

Figure 1.2: Typical high-lift airfoil system with deflected multi-element LE slat and TE
flap (Lin and Dominik 1997).

This works explores the use of two types of dielectric barrier discharge (DBD)

plasma actuators for controlling LE and TE separation on a further simplified high-lift

airfoil (Figure 1.3). The most common of these is the DBD plasma actuator driven by an

AC voltage waveform (AC-DBD). This device has become very popular in recent years

2
due to its simple construction, lack of moving parts, fast time response and low power.

AC-DBD plasma is well-established as a flow control actuator at airfoil LEs for

freestream velocities up to 30 m/sec (Moreau 2007). Its use on the more challenging

problem of controlling flow separation at the TE flap shoulder has not been fully

explored and is one of the primary focuses of this work. The control of flow separation

over a deflected TE flap using more common piezoelectric type actuators has been

documented and serves as a basis for judging the AC-DBD plasma on a similar scaled

version of the same airfoil model (Pack et al. 2002; Melton et al. 2003; Melton et al.

2004; Melton et al. 2005; Melton et al. 2006; Melton et al. 2007).

Figure 1.3: Further simplified high-lift airfoil system with deflected TE flap.

DBD plasma actuators driven by repetitive nanosecond pulses (NS-DBD) are less

established, but appear quite promising for flow control applications at higher speeds.

The construction of this actuator is identical to the AC version, but it has shown airfoil

LE separation control authority in isolated tests up to Mach 0.74 which is well beyond

any published results of AC-DBD plasma control authority (Roupassov et al. 2009). The

3
efficacy of this unique device is primarily explored at the LE of a supercritical airfoil and

compared to the more established AC-DBD as well as published results for a common

PFC device on a similar scaled version of the same model.

A characterization of both types of DBD plasma actuators is performed in

quiescent air to highlight the substantially different behavior created by producing the

discharge with these radically different input waveforms. In addition to the obvious

metric of CL, detailed flow diagnostics reveal the physics behind separation control and

lift enhancement using both types of DBD plasma actuators. This work is intended to

provide a benchmark comparison of AC-DBD plasmas to piezoelectric actuators for

controlling flow separation over a deflected TE flap while clearly documenting the

potential of NS-DBD plasma actuators for LE airfoil separation control.

4
Chapter 2: Background

Separation control is a broad and widely studied topic and thorough reviews on its

various applications have been published (Gad-el-Hak and Bushnell 1991; Greenblatt and

Wygnanski 2000). Passive separation control techniques which generally constitute

geometric changes such as slotted LE slats and TE flaps are employed on many

operational aircraft. The slotted portions of the PFC devices are necessary to energize the

boundary layer on the suction surface via mixing with high momentum fluid from the

pressure surface allowing it to follow the curvature of the deflected system. These control

elements are effective for reducing or eliminating separation during takeoff and landing

when aircraft speeds are low and high-lift is required. However, they are expensive to

manufacture, heavy, mechanically complex and introduce parasitic drag effects even

when stowed during cruise. Despite these drawbacks, the benefits of PFC outweigh the

incurred cost created by their application to the aerodynamic surface.

2.1 Separation Control

Active separation control has gained popularity in recent years due to its potential

for maintaining or enhancing the benefits of passive control techniques without the

penalty associated with many of the detrimental factors listed above. The main difference

between PFC and AFC strategies is that the latter can be turned on and off by command

5
at time scales consistent with relevant flow dynamics. This also gives that potential for

implementation in a feedback control system that coupled with adequate sensors and

controller could create even greater benefits in flight efficiency and maneuverability. A

complete review of active separation control is a subject in itself. Rather, the following

background information focuses on separation control studies that examine the effect of

nominally two-dimensional actuation on two dimensional airfoil models. It should be

noted that governing parameters and their optimal values such as frequency scale

favorably from 2D airfoil to 3D wing configurations at modest sweep angles (Seifert and

Tillman 2009).

2.1.1 LE Separation Control

Technological advances over the last few decades have allowed researchers to

more fully explore the wide parameter space associated with this research topic.

Accordingly, significant advances in the understanding of separated flow phenomena in

response to actuation have followed. Among the most widely accepted is that unsteady

actuation via pulsed blowing, pulsed suction or both (zero net mass flux (ZNMF)) is

more effective than steady forcing (Seifert et al. 1996). The range of effective

dimensionless frequencies associated with separation control is on the order of unity. The

dimensionless frequency often termed reduced frequency, F+, or more commonly

Strouhal number is defined as:

fxsp
F+ =
U∞ 2.1

6
where f is the forcing frequency, xsp is the length of the separated region and U∞ is the

freestream velocity (Darabi and Wygnanski 2004; Glezer et al. 2005). This parameter

underscores the importance of the characteristic length scale of separated flow

phenomena, xsp, which is generally assumed to be the length of the separation zone over

the body in question (Seifert et al. 1996). This variable also depends on the response of

the flow to actuation (Wygnanski 2004). Physically, the reduced frequency of unity

requires that a perturbation must be introduced during the time that the freestream flow

propagates over the separated region. The importance of actuator location is closely

related to this expression since the shear layer created between the freestream and the low

speed separated region by nature selectively amplifies small perturbations if they are

introduced near its receptivity region. The optimum choice of this location for unsteady

actuation is generally at or slightly upstream of the separation line. This ensures that the

shear layer is excited by the control perturbations near its receptivity point. Successful

introduction of such forcing creates large spanwise vortices that develop via the Kelvin-

Helmholtz instability. These vortices encourage momentum transport between the

freestream and separated region thus reattaching the flow (Darabi and Wygnanski 2004;

Melton et al. 2005). Separation control by forcing at higher frequencies (F+ > 10) has

been classified as a different phenomenon characterized by enhanced dissipation. The

more desirable aspect of this forcing is a reduction in flow unsteadiness and subsequent

dynamic loads on the aerodynamic surfaces under the separated region (Amitay and

Glezer 2002).

7
2.1.2 TE Separation Control

Studies of specifically TE airfoil separation control are less prevalent, but two

major efforts in recent years have resulted in additional understanding of this system. The

ADVINT program on the Boeing Tilt-Wing SSTOL transport (Kiedaisch et al. 2006;

Nagib et al. 2006; Smith et al. 2006; Kiedaisch et al. 2007; Nagib et al. 2007; DeSalvo et

al. 2010) and a parallel work by NASA (Pack et al. 2002; Melton et al. 2003; Melton et

al. 2004; Melton et al. 2005; Melton et al. 2006; Melton et al. 2007) have produced

significant developments using AFC via both nonzero and ZNMF periodic excitation

(Seifert and Tillman 2009). Perhaps the most significant of these is that more momentum

input is required in comparison with LE control (Melton et al. 2006). This is

commensurate with the existence of a thicker, turbulent boundary layer that develops

along the main element of the airfoil. Consequently, simply tripping the boundary layer

approaching the flap shoulder, which can be effective for LE separation control at low α,

is not sufficient for controlling separation at high Re and flap deflection, δf. These results

also support LE separation control findings that show greater centripetal acceleration

created by airfoil surface curvature requires additional momentum for realizing similar

control authority (Greenblatt and Wygnanski 2003).

Because of these challenges, it is difficult to fully attach the flow over a simple

TE flap. Instead, separation control and circulation control are inherently linked in these

systems (Cerchie et al. 2006). Evidence of this effect is observed when increases in CL

are attributed to upstream effects rather than complete or partial reattachment to the flap

surface (Kiedaisch et al. 2006; Melton et al. 2006). This is especially true for low

8
frequency forcing (F+~0.3-1.5) if separation is not radically delayed (Melton et al. 2004).

Such behavior does not occur in more canonical studies of separation control for wall-

mounted humps (Seifert and Pack 2002). Additionally, the control authority for a given δf

is highly sensitive to actuation location in regions of high surface curvature encountered

near the flap shoulder (Melton et al. 2006). This is especially true for typical ZNMF type

actuation issuing from small often 2D slots on the model surface. Such results can be

traced back to the importance of separation location on the actuator placement in that

control introduced at or slightly upstream of the separation point generally produces the

greatest effect (Greenblatt and Wygnanski 2000).

For purely increasing CL, high frequency excitation (F+ > 10) is less efficient in

terms of both momentum and power than low frequency modulated versions of the same

signal at F+~0.3-1.5 (Seifert et al. 1996; Melton et al. 2006). The latter mechanism relies

on the existence of natural instabilities in the flow field which can selectively amplify

small disturbances if introduced near a region of strong receptivity (Darabi and

Wygnanski 2004). In either the low or high frequency forcing cases above, amplifying the

disturbance generally serves to increase the control authority (Melton et al. 2004;

Kiedaisch et al. 2006; Melton et al. 2006), but the scaling laws governing these effects

are not well-defined (Seifert and Tillman 2009). Additional open questions on the

directional effects of the momentum addition as well as the size and shape of the slot or

hole for which the momentum is introduced remain. It also appears that flows with large

separated regions such as those over deflected TE flaps may benefit substantially from

3D forcing and the subsequent production of streamwise vorticity (DeSalvo et al. 2010).

9
2.2 Plasma Actuators

The recent interest in plasma actuators for aerodynamic flow control is motivated

by their simple construction, lack of moving parts, fast time response and low power. The

two types of plasma actuators used in this work are believed to function through

electrohydrodynamic (EHD) and thermal mechanisms. The former relies on momentum

transfer from charged species to neutral air molecules while the latter works through

Joule heating effects. These two very different mechanisms are produced using a DBD

arrangement. The construction of the device is identical in both cases, but the input

waveform is substantially different. DBD plasmas are relatively new to the aerodynamic

community, but have long been used in a variety of industrial applications such as ozone

generation (Kogelschatz 2003).

2.2.1 AC-DBD Plasma Actuators

The DBD plasma actuator for aerodynamic flow control is usually composed of

two electrodes separated by a dielectric material arranged in an asymmetric fashion

shown in Figure 2.1. Application of a sufficiently high voltage (≥5 kV) AC (1-10 kHz)

signal between the electrodes weakly ionizes the air over the dielectric covering the

encapsulated electrode. The dielectric barrier allows the generation of a large volume of

plasma by preventing the discharge from collapsing into an arc. The DBD plasma

actuator is a self limiting device in that the accumulation of charged particles onto the

dielectric surface opposes the electric field requiring consistently higher voltages to

sustain the discharge. This is circumvented using an AC waveform which, because of a

change in polarity, creates movement of charged species back and forth between the
10
exposed electrode and the dielectric surface at the AC driving frequency. The movement

of these charged particles transfers momentum to the flow via ion-neutral collisions. In

quiescent conditions, the asymmetric plasma actuator creates suction above the exposed

electrode and a pseudo wall jet over and downstream of the covered electrode that has

been shown to follow a skewed Gaussian profile (Hoskinson et al. 2008). Maximum near

wall streamwise velocities generated by a single actuator in air are less than 10 m/sec

measured a few millimeters from the wall just downstream of the plasma extent (Moreau

2007; Corke et al. 2009).

Figure 2.1: Typical asymmetric DBD plasma actuator geometry (Corke et al. 2010).

The induced flow is predominantly directed away from the exposed electrode due

to the asymmetry of the actuator geometry and behavior of the discharge over the two

waveform half cycles (Enloe et al. 2004a; Enloe et al. 2008b). In air, this force

production is dominated by interactions from negatively charged species, most notably

11
the negative oxygen ion (Enloe et al. 2006; Kim et al. 2007; Font et al. 2010). Each cycle

of the AC waveform produces a dominant velocity pulse due to movement of these

species. Time-resolved force measurements show that the dominant momentum

producing phase of the waveform is associated with the negative going half of the

negative half cycle and that this can produce up to 97% of the generated thrust (Enloe et

al. 2008b). Referring to Figure 2.2, this region is seen as the portion of the AC waveform

with negative polarity and negative slope.

Figure 2.2: Simultaneous traces of voltage and current for a typical AC-DBD plasma
actuator.

12
This portion of the discharge cycle is characterized by more uniform behavior

where negative species are deposited on to the dielectric surface. This is often referred to

as the forward stroke with respect to electron movement (Corke et al. 2010). The more

diffuse behavior can be seen in light emission, current measurements and high speed

photography which all show more uniform low intensity microdischarges in comparison

to the positive half cycle. The reverse or back discharge is substantially more filamentary

in nature with concentrated streamers at distinct, but unpredictable locations.

High speed photography shown in Figure 2.3 has revealed this behavior in detail.

The current traces in Figure 2.2 correlate with this behavior as the current spikes in the

reverse/back discharge case are less uniform and reach peaks of >200 mA while the

forward stroke is more uniform with lower intensity peaks. It has been postulated that this

asymmetry is created by the inability of the dielectric surface to freely give electrons to

the bare electrode similar to the schematic in Figure 2.4 (Corke et al. 2010). This

asymmetric behavior has motivated the use of a negative sawtooth high voltage signal

which acts to extend the lifetime of the momentum generating portion of the waveform

thereby increasing the generated thrust (Enloe et al. 2004a; Balcon et al. 2009).

13
a)

b)

Figure 2.3: High-speed photography of the forward stroke (a) and reverse/back stroke (b)
of a typical AC-DBD plasma actuator (Enloe et al. 2008a).

Figure 2.4: Schematic of negatively charged species movement for the forward (a) and
reverse/back stroke (b) in a typical AC-DBD plasma actuator (Enloe et al. 2008a).

14
It is generally accepted that AC-DBD plasma body forces are voltage driven

phenomenon as both the thrust and dissipated power are proportional to Vac7/2 when the

device is operating in the normal glow regime similar to Figure 2.5a (Corke et al. 2009).

Beyond this region, the dissipated power tends to increase while the induced flow

saturates at some maximum value determined by the excitation waveform and actuator

construction (Forte et al. 2007). The power increase seen beyond the Vac7/2 regime is

believed to be dissipated through heating in randomly occurring thick filaments that are

constant in space and are easily visible with the naked eye (Figure 2.5b). These should

not be confused with the microdischarges in Figure 2.3. Note that the localized heating in

these thick filaments eventually leads to a breakdown of the dielectric material and the

formation of an arc filament which eliminates velocity production.

Figure 2.5: Visual appearance of a typical ~10 cm long AC-DBD plasma actuator
operating in the normal glow regime (a) and at maximum thrust (b) (Thomas et al. 2009)

15
There is an optimal carrier frequency, fc, for DBD plasma induced thrust that is

dependent on both the ambient gas and the bulk capacitance of the dielectric, ε/h, where ε

and h are the dielectric constant and thickness respectively. In addition, for a given

dielectric operating at its optimized frequency, the thrust created from the actuator is

dependent on the bulk capacitance of the dielectric (Corke et al. 2009). For a given

voltage input, dielectrics with larger bulk capacitance tend to produce greater thrust when

operated in the Vac7/2 region due to an enhanced electric field. Consequently, an idealized

dielectric will have a large dielectric constant and a small thickness to create a larger bulk

capacitance. The caveat here is that dielectrics with smaller thickness generally have

lower dielectric strength and cannot withstand high voltages without entering the arc

regime. The use of relatively thick materials increases dielectric strength at the expense

of requiring higher voltages to initially ionize the gas and sustain the discharge in the

Vac7/2 region. This results in high values of thrust, but an increased actuator thickness and

voltage requirement (Corke et al. 2009). Using this knowledge, the generated thrust from

an AC-DBD plasma actuator can be increased by an order of magnitude using thicker

dielectric materials with lower dielectric constants and higher AC voltages (Thomas et al.

2009).

As previously discussed, the force production of AC driven DBD plasma

actuators is strongly dependent on the oxygen content of the environment (Enloe et al.

2006; Kim et al. 2007) However, ambient conditions such as pressure and humidity also

play a role. The dependence of thrust generation on these parameters has not been

exhaustively studied, but early reports suggest the actuator may still be viable under these

16
variable conditions (Abe et al. 2008; Benard et al. 2009a). Still, considerably more work

must be done to fully characterize their performance across a wide range of flight

conditions. Other aspects such as the effect of charge accumulation on the dielectric

surface appear to be crucial, but conflicting reports exist on the wheter this has a positive

or negative effect on the induced flow (Opaits et al. 2009b; Font et al. 2010). Despite

these many open questions, strides continue to be made for implementation as AC-DBD

plasma actuators have been used with varying degrees of success in feedback control and

flight testing (Patel et al. 2007; Sidorenko et al. 2008).

Numerous models have been developed for DBD plasma actuators that include

varying degrees of complexity in simulating plasma chemistry (Golubovskii et al. 2002;

Singh and Roy 2008; Likhanskii et al. 2009). These approaches are certainly useful, but

depending on the level of complexity can be computationally expensive. In many

practical flow control applications an empirical space-time lumped element circuit model

appears sufficient for predicting body forces that can be used in the Navier Stokes

equations for design purposes (Jayaraman and Shyy 2008; Rizzetta and Visbal 2009;

Corke et al. 2010).

2.2.2 NS-DBD Plasma Actuators

The AC-DBD plasma actuator is currently undergoing rapid development, but the

most important aspect of this remains the low momentum of the induced flow of the

device which limits control authority at high speed. This continues to be explored through

optimization of a single actuator, the use of multiple actuators and more novel

arrangements like sliding discharges that rely on additional DC bias voltages (Forte et al.
17
2007; Thomas et al. 2009; Corke et al. 2010). The common goal in all these cases is to

increase the velocity generated by the device. This is certainly a useful endeavor, but all

flow control actuators that rely on ZNMF momentum introduction eventually lose

effectiveness as requirements increase at progressively higher flow speeds.

This has motivated the study of actuation techniques based on other mechanisms,

namely thermal effects. Flow control using this method of actuation can be developed via

bulk or localized heating and was originally intended for use in supersonic flows to

weaken or manipulate shock structure (Palm et al. 2003; Adelgren et al. 2005). The

former requires the production of a very large volume of diffuse plasma which is

expensive from a power standpoint and difficult if not impossible to sustain at

atmospheric pressure. The latter can be produced using high power lasers or microwave

beams, but a more practical method is through an arc discharge between pin electrodes.

This has been extensively explored with localized arc filament plasma actuators

(LAFPAs) in high speed jets (Samimy et al. 2007). The control mechanism in this case is

believed to be a thermal effect that generates compression waves. The flow amplifies

these LAFPA generated perturbations when introduced at the correct frequency and

mode. LAFPAs also have the advantage of high bandwidth which allows them to be used

to excite and manipulate a variety of flow instabilities (Utkin et al. 2007).

These devices appear nearly ideal for exciting various instabilities in free jets.

However, they generate a considerable amount of heat requiring they be housed in high

temperature non-conducting materials such as ceramics. Additionally, multiple arc

filaments are required for control authority in most systems. For, example eight actuators

18
distributed around the circumference of a one inch diameter jet have considerable control

authority. For transport aircraft separation control applications, placing multiple pin

electrodes along an aircraft wing may not be practical.

The use of AC-DBD plasmas for flow control via thermal mechanisms was

originally considered, but this has since been dispelled by multiple studies (Enloe et al.

2004a; Jukes et al. 2006; Sung et al. 2006). Instead, a different high voltage waveform in

the form of very short pulses is employed. The construction of this device is identical to

the AC-DBD plasma actuator, but the input waveform now has pulse width of a few tens

of nanoseconds with pulse voltages in the range 10-50 kV. The NS-DBD plasma actuator

relies on a very short rise time to generate rapid localized heating by the plasma. The

dielectric barrier along with the short pulse width serves to prevent the discharge from

collapsing into an arc. This allows the thermal effect to become distributed along the span

of the device. The discharge can now be sustained on simple plastic dielectric surfaces

rather than the more extreme ceramics required in the arc filament case.

The physics of NS-DBD plasma actuators as flow control devices have not been

extensively studied. This stems from the developmental nature of this type of discharge

as well as the difficulty in procuring power supplies capable of producing such

waveforms. EHD effects produced by these devices appear quite weak and compression

waves generated by NS-DBD plasma actuators have been detected in both modeling and

experiments in still air (Roupassov et al. 2009). Details of these effects as well as

quantitative measurements of heating and compression wave strength remain open for

investigation, but estimates of near surface temperature rises of ~200 K corresponding to

19
compression waves at Mach 1.13 have been suggested (Roupassov et al. 2009). There

have also been attempts to combine both AC and NS-DBD waveforms in hopes of

increasing EHD effects, but these results are still very preliminary (Opaits et al. 2009b).

2.3 Separation Control with Dielectric Barrier Discharge Plasma Actuators

Early examples of AC-DBD plasma actuators as flow control devices

demonstrated their potential for boundary layer and LE airfoil separation control

applications (Roth et al. 2000; Post and Corke 2004). More recently, experimental studies

using such devices have broadened to include jet mixing, cavity tone attenuation, noise

control and aero-optics (Benard et al. 2007; Chan et al. 2007; Freeman and Catrakis

2008; Thomas et al. 2008). While these new applications have gained popularity in recent

years, the majority of DBD plasma work is still applied to separation control. These

actuators are particularly appealing for this application due to the nature of their induced

flow when arranged asymmetrically. In this configuration, the induced flow produces a

wall jet with maximum velocity a few mm away from the surface which is often

amenable for influencing boundary layers.

The mechanism responsible for separation control by AC-DBD plasma is most

often associated with the wall jet generation described above, but whether this results in

boundary layer tripping, energizing or amplification of instabilities depends on the flow

system under consideration. For separation control explicitly, the state of the boundary

layer (laminar or turbulent) just upstream of the actuator will also play a role.

20
Like other methods of periodic excitation (Greenblatt and Wygnanski 2000), AC

driven DBD plasma actuators are often most effective for airfoil separation control and

lift enhancement when excitation is created with F+ on the order of unity (Huang et al.

2006; Greenblatt et al. 2008; Patel et al. 2008). To operate in this fashion, the actuator

must be powered with a sufficiently high fc to produce the plasma (1-10 kHz) and

modulated at a lower frequency to excite the longer wavelength instabilities associated

with most separated flow dynamics. This behavior is analogous to ZNMF type actuation

created by piezoelectric diaphragms which produce the highest intensity fluctuations

when excited near the resonant frequency of the disc and/or cavity which is often on the

order of a few kHz. Studies of separation control with AC-DBD plasma actuators often

assume the flow does not respond to perturbations created by the high frequency carrier

signal, but instead responds as if exposed to steady wall jet (Thomas et al. 2008). For

many low-speed applications, this is true because the instabilities involved are not

receptive to high frequency perturbations and instead feel their effect as a quasi-steady

phenomenon. However, it has been confirmed that in air, the movement of charged

species in the plasma creates a dominant velocity pulse at the fc of plasma generation and

thus suggests the possibility of using AC-DBD plasma actuators for high frequency

forcing applications (Glezer et al. 2005) if sufficient amplitude can be produced

(Takeuchi et al. 2007; Boucinha et al. 2008).

In terms of frequency, the excitation signals driving piezoelectric and AC-DBD

plasma of actuators are quite similar, but the characteristics of the momentum production

and delivery are fundamentally different. In a typical ZNMF type actuator, periodic

21
blowing and suction is 180o out of phase and occurs through a slot or orifice mounted at

an angle relative to the freestream or surface. AC-DBD plasma actuators are completely

surface mounted and the exact location at which the plasma actuator accomplishes

control is not immediately obvious although actuators placed at or slightly upstream of

the separation location give favorable results (Huang et al. 2006; Sosa et al. 2007;

Jolibois et al. 2008). This appears consistent with modeling results that show the highest

force density associated with such devices is near the edge of the exposed electrode

(Enloe et al. 2004b; Corke et al. 2007). As described above, there are periodic blowing

and suction effects for a AC-DBD plasma actuator, but these predominantly occur during

the same phase of the input waveform, do not emanate from a common location (i.e. slot)

and are far from equal in magnitude. This behavior causes AC-DBD plasma to perform

more similarly to a pulsed blowing device, while still retaining ZNMF properties.

The use of AC-DBD plasma actuators for airfoil separation control at locations

other than the LE has been limited. Studies have reported that actuators placed near the

TE of airfoils can produce effects similar to simple flaps with deflections of a few

degrees (Vorobiev et al. 2008). This results in a uniform increase in CL across all α and a

slight reduction in minimum drag coefficient, CD, at Re on the order of 105 corresponding

to velocities of a few tens of meters per second (He et al. 2009).

While the potential of AC-DBD plasma actuation for separation control is

apparent, they also possess some drawbacks. They have primarily been limited to

relatively low speed (U∞ < 30 m/s) low Re (~105) applications such as those associated

with micro air vehicles due to the weak flows induced by the plasma discharge

22
(Greenblatt et al. 2008). More recently, claims of control authority for freestream

velocities as high as 60 m/sec with Re=106 have been presented in the literature (Patel et

al. 2008). To date, DBD plasma actuators have not produced sufficient momentum to

eliminate separation for flows over simple deflected flaps at Re > 105 unless the

freestream velocity is quite low (Mabe et al. 2009), although novel arrangements of these

devices show promise (Poggie et al. 2010). Recent publications suggesting momentum

production can be increased by an order of magnitude using thicker dielectric materials

with lower dielectric constants and higher AC voltages imply that the technology is

continuing to mature and may allow use in transport aircraft applications (Thomas et al.

2009). These devices also present the possibility of placing multiple actuators in various

locations and orientations on the airfoil to allow spatially distributed forcing over 180

degrees referenced to the model surface which can produce streamwise vorticity and

further freestream momentum entrainment (Porter et al. 2009).

NS-DBD plasma actuator testing in flow control applications has been limited to

one isolated publication (Roupassov et al. 2009). This work demonstrated the efficacy of

NS-DBD plasma actuators for controlling flow separation on an airfoil LE up to Mach

0.74. This flow speed is substantially greater than any associated with AC-DBD control

authority in existing literature for airfoils. Unfortunately, attempts at replicating this

success have been unsuccessful. As such, it remains an exciting, but not well-documented

actuation technique.

23
Chapter 3: Experimental Facilities and Measurement Techniques

3.1 Wind Tunnel

A Gottingen-type, closed, recirculating wind tunnel (SN: 55706) manufactured by

Engineering Laboratory Design, Inc. (ELD) is used as the test bed for this work. Viewed

from above, the standard version of this tunnel creates counterclockwise flow. At the

request of the GDTL, the tunnel was modified by the manufacturer to produce clockwise

flow. This places the large cross section elements upstream of the test section in a more

convenient location maximizing practical laboratory work space. The various wind tunnel

components were shipped in a dedicated trailer and assembled on site. Overall

dimensions of the assembled tunnel are 9.8 x 2.2 x 4.1 m3 (32.2 x 7.2 x 13.5 ft3) with a

test section centerline height of 1.4 m (4.6 ft).

Air flow in the tunnel is created by an axial fan powered with a 200 hp variable

speed AC induction motor (Figure 3.1). The fan speed is set using an operator keypad on

the tunnel outfield just downstream of the test section (Figure 3.2). The contraction

before the test section has a symmetrical cross section and an area ratio of 6.25:1 (Figure

3.3). The primary diffuser downstream of the test section expands with a total angle of

6°. This assembly allows continuously variable air velocity in the tunnel from 3-90 m/s

24
(10-300 ft/s). Flow conditioning upstream of the test section includes a hexagonal cell

aluminum honeycomb (Figure 3.4). High porosity stainless steel screens are mounted

downstream of the test section as a safety catch (Figure 3.5). Four high efficiency turning

cascades fabricated of galvanized steel are installed in each of the four tunnel elbows

(Figure 3.6). This assembly results in freestream turbulence specifications on the order of

0.25% with +/-1% variation in mean freestream velocity across the tunnel span measured

15 cm (6 in) from the test section inlet. These measurements were performed by factory

personnel at ELD (ELD 2006).

Figure 3.1: Axial fan used in the subsonic recirculating wind tunnel.

25
Figure 3.2: Operator keypad for the subsonic recirculating wind tunnel.

Figure 3.3: Subsonic recirculating wind tunnel contraction section.

26
Figure 3.4: Flow conditioning screens at the subsonic recirculating wind tunnel
contraction entrance.

Figure 3.5: Safety catch screens downstream of the subsonic recirculating wind tunnel
test section.
27
Figure 3.6: Turning cascades used in the elbows of the subsonic recirculating wind
tunnel.

The tunnel is also equipped with a commercial aluminum fin/copper tube, double

row heat exchanger with electronic modulating valve and set point controller located on

the wind tunnel operator keypad. The heat exchanger is positioned upstream of last elbow

before the nozzle contraction (Figure 3.7). This arrangement allows the tunnel freestream

operating temperature to be maintained at +/- 1 °C from the ambient when supplied a

sufficient source of cooling water (max 189 lpm (50 gpm)). Cooling water is taken from

the lines normally supplying the compressors at the Aeronautical and Astronautical

Research Laboratory (AARL) at OSU by opening a valve on the east wall of room 129. A

significant amount of slag tends to build in these lines necessitating a gravity and course

mesh filter upstream of the heat exchanger to prevent fin clogging. No detrimental effects

28
on cooling either the wind tunnel or the compressors have been observed when operating

both simultaneously.

Figure 3.7: Gravity filter, heat exchanger, modulating valve, flow meter and drain line on
the subsonic recirculating wind tunnel.

During the course of the plasma actuator experiments, the modulating valve

associated with cooling the facility was found to be highly sensitive to electromagnetic

interference (EMI). The effect of EMI is to fully open the valve which can drop the air

temperature in the tunnel below the desired set point. To prevent a change in tunnel

29
temperature during plasma tests, the modulating valve is always forced to 100% by

setting the temperature to an unrealistically low value (i.e. 50 F). The temperature of the

tunnel is measured by a thermocouple placed just before the first elbow downstream of

the test section.

The 61 x 61 x 122 cm3 (2 x 2 x 4 ft3) facility test section is constructed of

optically accessible super abrasion resistant acrylic walls that are 25.4 mm (1 in) thick.

The infield side wall of the standard test section is fitted with a blank 30.5 cm (12 in)

diameter port that is located 30.5 cm (12 in) from the test section floor and 61 cm (24 in)

downstream of the test section entrance. The infield plug is equipped with a protractor for

setting model angles of rotation and attack using manual thumb screws. The standard

outfield wall has no such port. To facilitate use with 2D full-span airfoil models, the

outfield wall panel has been replaced with a mirror image of the infield wall complete

with an additional wall plug port shown in Figure 3.8.

30
Figure 3.8: Test section arrangement with wall plugs on both infield and outfield sides.

A two axis traversing assembly (Velmex, Inc. Unislide) driven by a DC stepper

motor and controller is mounted on top of the test section allowing introduction of

instrumentation such as pitot-static and hot wire probes through a 109 cm (42.8 in) long

high density nylon brush seal along the tunnel ceiling centerline. The use of such

instrumentation requires an aluminum support to prevent deflection of the probes at high

tunnel speeds. It is also necessary to modify the downstream side of the support with

some device to prevent tuning of the probe assembly to the freestream velocity. Without

this addition, the probe can vibrate significantly similar to a high amplitude tuning fork.

The final arrangement for the 91.44 cm (36 inch) pitot-static probe is shown in Figure

3.9. In this case, the probe is mounted on the front of the support and a small piece of

31
round stock is used on the downstream side. Both are secured to the support using

aluminum tape.

Figure 3.9: Modified pitot-static probe assembly.

Two piezometer rings consisting of 4 static pressure taps located on the centerline

of each panel are installed 50.8 mm (2 in) from both the contraction entrance and exit.

Pressures at these locations are used to set the tunnel operating conditions. Vertical

profiles of the dynamic pressure, q, measured along the tunnel centerline have been

acquired using a pitot-static probe at streamwise locations of 0, 30.5, 61 and 85 cm (0,

12, 24 and 33.5 in) relative to the test section entrance for velocities in the range 10-90

32
m/s with a 10 m/s increment (Figure 3.10). Maximum boundary layer thickness on the

tunnel floor is less than 20 mm (0.8 in) for all velocities and locations surveyed (Figure

3.11). The profiles have been offset for clarity by a dimensionless dynamic pressure,

q/q∞, of 0.5 for each 12 inches of streamwise location. Average freestream measurements

of q across the tunnel length surveyed compared to outputs from the piezometric rings

upstream and downstream of the contraction show that the tunnel has a calibration

constant, k, of 1.05 calculated from

q∞
k=
po − p 3.1
where q∞ is the dynamic pressure measured with the pitot-static probe and po and p are

the static pressures measured at the contraction entrance and exit respectively (Rae and

Pope 1984). This factor has been employed in plotting the profiles in Figure 3.10 and

Figure 3.11 such that the q/q∞ is essentially unity in the freestream. Note that the profile

furthest downstream (x=33.5 cm) is shifted less than other cases to represent the actual

streamwise development of the boundary layer. While tunnel specifications state low

speed velocities such as 3 m/sec are obtainable, it is not advisable to acquire data below

10 m/sec due to significant flow unsteadiness. It is also clear that the boundary layer is

transitional on the tunnel floor for freestream velocities of 30 m/sec and below, however

this does not affect the freestream velocity profile.

33
Figure 3.10: Dimensionless dynamic pressure profiles measured 0, 30.5, 61 and 85 cm (0,
12, 24 and 33.5 in) (left to right) from the test section inlet.

Figure 3.11: Near-wall dimensionless dynamic pressure profiles measured at 0, 30.5, 61


and 85 cm (0, 12, 24 and 33.5 in) (left to right) from the test section inlet.

34
Components of the wind tunnel not in use during airfoil and traverse experiments

include a blank acrylic top panel, blank acrylic outfield wall panel and blank wall plug.

Two additional minor modifications to the tunnel include threaded ports in the test

section floor for mounting various assemblies and a port upstream of the tunnel

contraction for introducing particle seed used in optical diagnostics. Additional details on

the wind tunnel including electrical information, installation, operation and maintenance

instructions can be found in the ELD Inc. supplied manual (ELD 2006).

A modular instrument panel near the tunnel operation station houses additional

hardware (Figure 3.12). The panel includes two sets of differential static pressure

transducers (Omega Engineering, Inc. PX655-25DI and PX655-5DI) used for setting the

tunnel test conditions and three process meters (Omega Engineering, Inc. DP-25-E-A) for

displaying the output. The Velmex controller for the 2D traverse is also housed on the

instrument panel and can be controlled via computer or manually. The anemometer for

use in hot wire experiments (TSI 1750) is mounted on the instrument panel as well as

Scanivalve pressure sensors (DSA-3217/16px 5 psid), quick connectors for use with 1.6

mm (0.063) inch inner diameter polyurethane static pressure tubing and an five port

Netgear ethernet switch used for routing information to a dedicated wind tunnel

computer. Standard manometers serve as a means of verifying the tunnel operating

conditions.

35
Figure 3.12: Modular instrument panel for the subsonic recirculating wind tunnel.

3.2 Airfoil Model

A simplified high-lift version of the NASA Energy Efficient Transport (EET)

airfoil has been chosen as the test model based on recommendations from the initial

sponsor (AFRL). The 2D EET airfoil was thoroughly examined (Lin and Dominik 1997),

but more recently significant studies on active separation control with synthetic jets have

been completed for a similar simplified version (Pack et al. 2002; Melton et al. 2003;

Melton et al. 2004; Melton et al. 2005; Melton et al. 2006; Melton et al. 2007). The OSU

version has a chord of 25.4 cm (10 in) and fully spans the 61 cm (24 in) test section in a

36
horizontal configuration. The model is equipped with a deflectable (0-60°) TE flap that is

25% of the airfoil chord, but for simplicity lacks the leading edge droop used by NASA.

It is constructed of a nylon compound (Duraform GF) and has been fabricated using

selective laser sintering (SLS) technology.

The model is assembled in three elements consisting of the suction surface,

pressure surface and flap. The model elements are fabricated by General Pattern, Inc. and

assembled and instrumented by ELD according to specifications provided by OSU. There

are 45 staggered static pressure taps located near the test section centerline and 15 static

pressure taps at ¼ and ¾ spans used for checking flow three dimensionality. The urethane

tubing described above connects the pressure taps from the infield side of the tunnel to

the static pressure sensors mounted on the instrument panel. Care is taken to ensure no

kinks or leaks exist in this tubing before testing.

The model is also instrumented with 7 high bandwidth 2.5 mm (~0.1 in) diameter

Kulite pressure transducers (XCQ-080-25A and LQ-062-25A) flush mounted near the

centerline. Figure 3.13 shows the airfoil profile and the location of static pressure taps

and transducers near the centerline. There were significant issues during the pressure

transducer installation as many were broken by ELD, Inc. due to tight space constraints in

the model. This was especially difficult for the miniature sensors (LQ-062-25A) in the

TE flap. Details on model construction and disassembly can be found in the ELD

supplied manual (ELD 2007).

37
Figure 3.13: 2D profile of the airfoil in cruise configuration showing the approximate
location of static pressure taps and high bandwidth pressure transducers near the airfoil
centerline.

Independent settings for the incidence and flap deflection angles are set manually

using separate wall plugs with protractors on the tunnel infield side. The zero angle of

attack position is scribed on the infield wall plug that is attached to airfoil model. This

scribe aligned with 20 degrees on the uncalibrated protractor sets the zero angle of attack

position. The flap protractor is calibrated separately and documented on the flap wall

plug. A digital photograph showing the airfoil with trailing edge flap deflected and flap

wall plug is shown in Figure 3.14.

38
Figure 3.14: OSU version of the simplified high-lift EET airfoil with TE flap deflected.

The model has been extensively tested at multiple incidence and flap angles up to

Re=750k and has been run less extensively at Re=1000k at α as high at 18o. Attempts to

further increase Re at high α are not advisable based on visual observations of model

bending and flutter (Figure 3.15). Operation at higher Re and low α may be possible, but

this has not been attempted. The model is not designed for use in dynamic α studies.

39
Figure 3.15: Model bending produced by loading at high Re and α with DBD plasma at
the LE.

3.3 Plasma Actuator Hardware

Input signals for the AC DBD plasma actuators are generated using a dSpace DSP

1103 board and hardware system. The dSpace system was originally purchased for use in

feedback control, but serves only as a function generator for this work. Outputs of the

dSpace D/A board include both the full modulated excitation signal supplied to the high

voltage power supply as well as the modulation signal alone for use in phase-locked data

acquisition timing.

Signals generated by dSpace are used as inputs to a Powertron Model 1500S AC

power supply (1500 W) and step-up high voltage transformer. Two identical power

supply models have been used, but the majority of single actuator AC DBD plasma tests

have been performed using amplifier 1 as labeled. Amplified signals from the power

40
supply are sent to a low power (200 W) high voltage (0-20 kVrms) transformer designed

to operate in the frequency range of 1-5 kHz. Ballast resistors were originally employed

in the plasma circuit, but were found to be unnecessary and subsequently removed. The

output of the transformer is connected directly to the high voltage electrode. No

impedance matching or EMI hardening circuit elements are employed. The high voltage

15 kVDC rated lead wires must be insulated from the laboratory floor and ground wire as

a weak discharge can ensue even through the insulation. This is accomplished using

sheets of cardboard. A safety circuit consisting of various fuses was investigated to

protect the airfoil from damage in the event of arc formation, but close monitoring of the

signal during operation using a Tektronix P6015A high voltage probe and a Tektronix

TDS 220 oscilloscope was found to be sufficient. The AC power supply and hardware is

located behind the wind tunnel operator station to facilitate one person operation.

High voltage nanosecond pulses are generated using an in-house constructed

pulser. The pulser is a magnetic compression type currently capable of repetition rates up

to 10 kHz. Peak voltage developed on an open load can reach ~18 kV with maximum

pulse energy of ~100 mJ. Pulses can be produced with either positive or negative

polarity, but only the positive pulse is examined in this work. The device is relatively

inexpensive in comparison to commercial pulsers of similar specification and can be

repaired using mostly off-the-shelf components. Dimensions of the pulser are

approximately 40 x 40 x 40 cm3. A 650 V DC Power Supply (Sorenson DCR 600-4.5B)

is used to power the system and a Tektronix AF6310 function generator provides input

signals. The pulser is located in the wind tunnel infield during operation due to the

41
requirements that lead wires be of equal length and as short as possible to minimize pulse

propagation effects and the possibility of reduced peak voltages (Opaits et al. 2009a).

This requires two operators during testing.

AC voltage measurements are acquired and monitored at the secondary side of the

high voltage transformer with a Tektronix P6015A high voltage probe. A LeCroy current

probe (CP031) is also employed. The power dissipated by the AC-DBD is calculated with

the Lissajous figure using charge-voltage measurements (Falkenstein and Coogan 1997).

A 47 nF capacitor is connected in series with the covered ground electrode in this case.

The voltage across the capacitor is measured using a Tektronix P6111B voltage probe.

The corresponding signals are monitored on a LeCroy Waverunner 6050A oscilloscope,

but the actual power calculation is performed offline. The AC-DBD plasma actuator is

operated using a 1-3 kHz sinusoidal carrier frequency with voltage levels up to 30 kVpp

and various modulation waveforms and frequencies. Electrical measurements for the NS-

DBD are acquired at the high voltage electrode in order to most accurately resolve the

actual voltage developed on the load. EMI is relatively minor when all external

equipment is properly grounded. The only devices that show considerable EMI influence

are the modulating valve for the heat exchanger and the differential pressure sensors used

for monitoring the tunnel air speed. Neither of these is essential once test conditions are

established.

42
3.4 Diagnostics

3.4.1 Static Pressure

Measurements of static pressure from ~1 mm (0.040 inch) inner diameter taps on

the airfoil model surface are acquired using three Scanivalve digital pressure sensor

arrays. Values of dimensionless pressure (CP) are averaged over 50 samples acquired at

10 Hz near the model centerline and used to calculate sectional lift coefficient (CL) using

numerical integration according to:

p − p∞
CP = 3.2
q∞
 x
1
CL ≈ ∫ CP , pressure − CP , suction d  
0 side side c 3.3

where p is the static pressure on the surface, p∞ is the freestream static pressure, q∞ is the

freestream dynamic pressure and x/c is the chordwise position.

3.4.2 Fluctuating pressure

Kulite pressure transducers installed in the model are powered using an in-house

constructed signal conditioner that amplifies each sensor output by 1000. The signal

conditioner has a variable low pass filter, but noise levels are found to be substantially

reduced when a separate Kemo filtering unit is employed (Benchmaster 21M). The Kemo

unit has variable low/high pass filter settings, but the majority of tests have been

performed using a low pass filters at 10 kHz. Attempts to filter out the AC plasma carrier

frequency during low frequency modulating signal testing were met with limited success.

43
The resulting pressure traces are sampled simultaneously at 50 kHz using a National

Instruments PCI-6143 data acquisition board. Average spectra are calculated from 32

blocks of 8192 samples which results in a frequency resolution of approximately 6 Hz. A

Hanning window function is applied to each block.

The power spectral density (PSD) for each block of dimensionless fluctuating

surface pressure, cP, is calculated using

c=
P (t j ) CP (t j ) − CP 3.4

( ) = ∑ w( j)c (t ) e
N
+ −2π i ( j −1)( k −1)
cˆP F xsp , k P j ,
j =1
3.5
k ∈ [1, (N+1)/2] and Fx+sp ,k =
(k − 1) Fx+sp , s / N

( ) (
PSD Fx+sp ,k = 2 cˆP ∗ Fx+sp ,k cˆP Fx+sp ,k
 ) ( )(
CNFx+sp , s  ,
 )
1 N
for k ∈ [1, N / 2] with C = ∑ ( w( j )) 2
N j =1
3.6

where, CP is the time-averaged surface pressure, w(j) is the Hanning window function, tj

is the time index of the jth sample, Fx,s + is the dimensionless sampling frequency, Fx,k + is

the measured frequency and N is the number of samples acquired. The dimensionless

units of PSD are then expressed as cP2/ Fx+ in the corresponding figures where xsp is the
sp

length scale used for expressing the dimensionless frequency which in this work is the
44
chord, c, or flap length, L. The corresponding blocks are then averaged to obtain the

properly scaled result.

Pressure transducer measurements near plasma actuators are quite challenging

due to EMI and arc formation. During the course of this work, transducers near the TE

flap shoulder and LE have been damaged due to direct arcing from the high voltage

electrode for both AC and NS-DBD plasma. The resulting current spike can also

propagate through the transducer cabling causing damage to the individual channels of

the signal conditioner. At this time, only three transducers (x/c=0.40, 0.95 and 0.40

(pressure side)) remain operational. Less catastrophic effects have also been observed in

that noise levels for individual sensors increase with EMI exposure time such that after

months of use the transducer signal becomes excessively noisy. It is also suspected that

transducer installation during which shielding wire was pulled away from the sensor

element has caused a less robust shield effect and resulted in higher noise levels in some

sensors. Strides must continue to be made to improve pressure and velocity diagnostics

near plasma actuation.

Since significant uncertainty exists around measurements of pressure spectra in

the presence of plasma, a study was performed in an attempt to distinguish real effects

from electrical noise. Figure 3.16 shows the experimental setup in which an actuator is

placed upstream of the TE flap shoulder and an additional pressure transducer is taped to

the test section wall outfield near the high voltage cable. Figure 3.17 shows the resulting

pressure spectra calculated using this arrangement for transducer 5 (x/c=0.90). In this

case, properly magnitude scaling has not been employed as only the relative effects are of

45
interest. The baseline flow shows an oscillation near 90 Hz with a high frequency noise

floor near -15 dB beginning around 1 kHz. Actuation at 2 kHz with sinusoidal

modulation of 90 Hz increases the magnitude across all frequencies and shows a very

strong peak at the carrier frequency (2 kHz). An additional large amplitude peak is also

visible at the modulation frequency (90 Hz). With this data only, it is unclear if the 90 Hz

peak and the broadband increase are actual flow changes or merely artifacts of electrical

noise. The spectrum of the transducer outside the wind tunnel gives a representation of

EMI propagation through the air from the HV line. The 2 kHz carrier waveform is quite

apparent, but there is no peak associated with the 90 Hz modulation signal. The

broadband signal also matches the controlled spectra above ~500 Hz suggesting this is

not a real flow artifact. Finally, the plasma actuator is operated in still air (tunnel off)

using the same excitation inputs as the control case. The spectrum from transducer 5

(x/c=0.90) is also shown in the figure. It is obvious that as in the outside sensor case, the

2 kHz carrier frequency is quite strong and the broadband levels above ~200 Hz are likely

an EMI artifact only. However, there is no evidence of a strong peak at the modulation

frequency and the controlled flow is found to have a broadband level above the tunnel off

case below ~200 Hz. From this information, it can then be seen that flow behavior cannot

be distinguished from electrical noise above ~200 Hz in this case. However, both the

coherent and incoherent dynamics below 200 Hz appear quite physical. Note that these

frequency values are not constant across the parameter space surveyed and similar

analysis should be performed for each forcing case. The use of pressure spectra is only

employed briefly in Section 6.1 for a case in which this verification has been used.

46
Figure 3.16: Setup used for distinguishing flow phenomena from EMI.

Figure 3.17: Example of pressure spectra used for distinguishing flow phenomena from
EMI.
47
One additional comment on the analysis of pressure traces near plasma is

warranted. The broadband increases shown in Figure 3.17 are partially attributed to an

impulse like spike in the pressure trace which, in the frequency domain, is composed of

all frequencies. Consequently, amplification across the entire frequency range is visible.

A sample of pressure data for this behavior is shown in Figure 3.18. Since only a finite

number of these impulses appear in the pressure trace, blocks of data which contain these

events can simply be neglected in the analysis since they represent unphysical flow

behavior. Note again that this is only one example of this behavior and that other cases

can contain more or less of these spikes. The source of this behavior is not known.

Figure 3.18: Example of randomly occurring voltage spikes in pressure traces due to
EMI.

48
3.4.3 PIV

Two-component particle image velocimetry (PIV) is used to obtain quantitative

velocity measurements for the various flow fields discussed. Nominally submicron olive

oil seed particles are introduced upstream of the test section contraction using a 6-jet

atomizer (TSI 9306A). The beam from a dual-head Spectra Physics PIV-400 Nd:YAG is

directed over a distance of ~6 m (~20 ft) then focused using a 1 m focal length spherical

lens and two -25 mm focal length cylindrical lenses. The additional cylindrical lens is

used to increase the spreading rate of the beam in an effort to provide a uniform light

distribution over the entire field of view which for the airfoil cases is approximately 300

x 200 mm2 (12 x 8 in2). Only ~50% of the maximum laser energy is necessary for ample

signal to noise ration in this flow. The beam forming hardware is mounted on top of the

wind tunnel and can be located in two possible positions depending on the airfoil case

tested (LE or TE) by removing the various 1/4 -20 screws connecting it to the tunnel

ceiling. In the LE case, the optical assembly is positioned as far upstream as possible

while the TE case requires the assembly in the more downstream position so that the

wake can be visualized. The spanwise location of the beam can be adjusted using an

optical rail and slider while the streamwise position can be fine-tuned using thumbscrews

on the final turning optic. No serious vibrations have been encountered for the beam

forming optics on the tunnel ceiling. However, the turning optics on the structural support

near the shear layer/SWBLI facility have been found to vibrate during operation of the

AARL transonic wind tunnel. Beam alignment is ensured by first using burn paper to

match the tails of the sheets and the fine tuned by visual inspection of the particle

49
movement between two frames of the raw data. The experimental setup for airfoil LE

PIV measurements is shown in Figure 3.19. The two axis traverse assembly is also visible

in this image.

Figure 3.19: PIV experimental setup for LE airfoil measurements.

Significant laser reflections exist in the test section primarily due to reflections

from the airfoil model surface. These contaminate not only data near the airfoil surface,

but also reflect from the test section walls and contaminate other regions of the flow field.

An example of this behavior is shown in Figure 3.20. These effects are minimized by
50
adhering blue masking tape to the infield interior wall of the wind tunnel. The reflections

are further reduced by covering the tape with flat black paint (see Figure 3.15).

Reflections from the model surface can be reduced using fluorescent paint created by

solving rhodamine powder in ethanol and mixing with water soluble acrylic paint.

Rhodamine absorbs the green wavelengths (532 nm) and fluoresces red (~600 nm), but

its efficiency is only ~30%. By using a narrow band pass optical filter (CVI F10-532.0-4-

2.00) over the camera lens, a portion of the original green intensity is eliminated.

However, when the beam is directly incident on the surface, the paint loses effectiveness

due to burnoff after ~100 high power laser shots. Thus, the rhodamine paint is more

effective when the laser sheet is not directly incident on the surface. Note that thin pieces

of Kapton tape perform similarly by diffusing the laser reflections, but these can only be

used in large separated regions such that the boundary layer is not disturbed by their

addition to the surface. Additional localized regions of reflection such as white high

voltage wires can minimized using a simple black marker. By using rhodamine, Kapton

tape, paint, marker and finely tuning the beam location and intensity, these reflections are

minimized and high quality raw data is produced. Sample raw data from both the LE and

TE experiments is shown in Figure 3.21.

51
Figure 3.20: Example of laser reflections from the model surface and their effect on PIV
data.

a)

b)

Figure 3.21: Sample raw PIV data images for LE (a) and TE (b) experiments.

52
Images are acquired and processed using a LaVision PIV system operating

software version DaVis 7.2. Two images corresponding to the pulses from each laser

head are acquired by a LaVision 14 bit 2048 by 2048 pixel Imager Pro-X CCD camera

equipped with a Nikon Nikkor 50 mm f/1.2 lens. The camera is mounted on an optical

table on the wind tunnel outfield using an optical rail and support. PIV measurements

over the chord of the airfoil in the LE case are acquired by viewing the light sheet

orthogonally through the main wall plug. Measurements over the TE flap and in the

airfoil wake must be acquired from a downstream angle of approximately 14° due to the

wall plug arrangement. An image correction algorithm provided in the commercial

software is applied to the data in this case. The boundary layer and flat plate PIV data are

acquired using an additional Vivitar 2x teleconvertor to increase spatial resolution at the

expense of field of view. In that latter, the flat plate substrate is mounted vertically in the

test section to reduce reflections from the surface (Figure 3.22). Finely machined

Lavision calibration plates (Type 31 and 10) are used for the airfoil and flat plate/BL

cases respectively.

53
Figure 3.22: Experimental setup for flat plate PIV measurements.

A narrow band pass optical filter (CVI F10-532.0-4-2.00) is often used when

acquiring PIV data. The effect of this component is to eliminate contamination from

ambient lighting as well as rhodamine painted surfaces. The optical filter has an

additional effect of slightly blurring the particle size on the CCD. In airfoil experiments,

this effect can be beneficial for increasing the particle size such that peak locking effects

that can occur for small particle sizes (< 1 pixel) on the CCD are minimized (Stanislas et

al. 2008). In cases where the spatial resolution is increased (boundary layer and flat

plate), the particle size can become too large due to the addition of the optical filter. For

these tests, no optical filter is used necessitating the lights in the laboratory must be
54
substantially dimmed. Additional ambient light is eliminated by draping a black felt cloth

over the infield side of the wind tunnel. For measurements of the actuator in still air, the

flow is seeded and the tunnel is energized to uniformly distribute the particles. Once

adequate seeding is ensured, the tunnel is shut off and particles remain suspended

Regions of interest are adjusted depending on the flow field in an effort to reduce

file size and processing time while maximizing sampling rate (max 10 Hz). Note that

cropping the raw image in x does not improve the sampling rate for this hardware system.

Also, the camera cropping in x is forced to be symmetric while cropping y is more

flexible. Regions outside the area of interest are excluded from the analysis using a

manually drawn data mask. For each image pair, subregions are cross-correlated using

decreasing window size (642-322 pixel2) multi-pass processing with 50% overlap in high

accuracy Whittaker reconstruction mode (Lavision 2007). The time separation between

laser pulses used for particle scattering is tuned according to the flow velocity, camera

magnification and correlation window size using the known freestream velocity and the

criteria which requires particle movement to be less than ¼ of the correlation window

size. The particle size on the CCD is in the 1-4 pixel range and the seeding density is

tuned to ensure no less than 6 particles are present in a given correlation window in all

cases. The resulting velocity fields are post-processed to remove any remaining spurious

vectors using a correlation peak ratio criteria, allowable vector range and median filter

that uses contributions from the eight closest neighboring vectors. Removed vectors are

replaced using an eight vector standard interpolation scheme and a 3x3 Gaussian

smoothing filter is also applied to the calculated velocity fields. The PIV data for used in

55
creating ensemble averages are sampled at 10 Hz. Statistics for the airfoil, boundary layer

and actuator characterization are calculated from 1000, 500 and 500 instantaneous

velocity fields respectively. Details on the Lavision algorithms are documented in the

corresponding manual (Lavision 2007).

Conditional sampling of PIV data (phase-locking) is accomplished using the

programmable timing unit of the LaVision system. In this case, the acquisition is synced

with the modulation frequency of the actuation signal. The baseline pressure signal is not

sufficiently periodic to allow this acquisition in the airfoil case. Velocity fields at various

phases of the actuator modulation frequency are investigated by stepping through the

actuator period using time delays. Phase locked data sets are averaged over at least 50

images for each phase which is found to be sufficient for resolving the primary flow

features (velocity and vorticity). Conditionally sampled (phase-locked) PIV data is

acquired at 5 Hz. The spatial resolutions of PIV data for the flat plate, boundary layer and

airfoil data sets are approximately 0.2, 0.4 and 2.4 mm, respectively which corresponds to

magnification values of ~44 and ~7 pixels/mm. Near surface measurements for the flat

plate and boundary layer data sets are obtained within 0.3 and 0.4 mm of the substrate

respectively.

The timing between the laser pulse and actuator signal must be checked for proper

synchronization. In the phase-locked case, this is obvious since the actuator and laser

signal must be have a constant delay at a given phase to ensure phase averages are

accurate. The opposite effect must occur for a true ensemble average in that the PIV

acquisition should sample over many phases of the actuation signal such that time-

56
averaged results are not biased toward any phase. Under certain circumstances the PIV

system and function generator running the actuator can run on similar clocks. What

appears to be a random acquisition can actually be inadvertently phase-locked. This

effect is known to occur when using Lavision PTU version 9 and the Tektronix AF6310

function generator used for NS-DBD plasma actuation. A simple solution to this problem

is to slightly increase or decrease the PIV acquisition rate to some non integer multiple of

the actuation signal (9.99 or 10.01 Hz). This effect does not occur for actuation used

dSpace for input signals, but for good practice one should employ this non-integer

acquisition rate in all cases.

3.4.4 Accuracy

Full scale accuracy for measurements of instantaneous velocity and the

uncertainty on mean velocity calculations are listed in Table 3.1. The former is calculated

by assuming negligible laser timing errors and a correlation peak estimation error of 0.1

pixels. This is the generally accepted error estimation for modern PIV systems (Stanislas

et al. 2008). The smaller error in the boundary layer and airfoil cases results from better

choices of the time separation between laser pulses due to the a priori knowledge of

maximum velocities in the freestream flow. The uncertainties in measurements of mean

velocity are presented using 95% confidence intervals, CI:

σ 2 
CI= x ± 1.96   3.7
 N 
where x is the sample mean, σ, is the sample standard deviation and N is the sample size

(Benedict and Gould 1996).


57
The values in Table 3.1 are in terms of relative error. The actuator and boundary

layer cases are estimated using the average standard deviation and average velocity in the

near wall profiles which gives a more representative estimate than the maximum or

freestream velocity. The uncertainty in the airfoil cases is estimated using the maximum

standard deviation in the turbulent wake and the freestream velocity. A range of values

that encompasses all the flow fields surveyed is listed in the interest of brevity. The

actuator and airfoil measurements have higher values of uncertainty due to the unsteady

nature these flow fields. The V component of velocity has not been considered for the

actuator and boundary layer measurements since it is not relevant to calculations of

integral parameters.

Mean Instantaneous

U V U, V

Actuator Characterization 4-6% - 1.7%

Boundary Layer 0.5-0.7% - 0.9%

Airfoil 2-3% 2-3% 0.9%

Table 3.1: PIV Uncertainty

58
3.5 Test Conditions

Experiments on the controlling separation from the TE flap have been conducted

for Re between 240k (15 m/s) and 750k (45 m/s) for δf of 20-40o at atmospheric

temperature and pressure. These test conditions correspond to a sample published results

of NASA work (Melton et al. 2004; Melton et al. 2006), but do not contain studies on the

effect of LE droop which limits our analysis to lower α. LE separation control studies

have been performed at primarily Re=750k (45 m/sec) and 1000k (62 m/sec) with no flap

deflection. For all airfoil measurements, data is acquired by establishing a separated flow

baseline condition then energizing the actuator. For repeated samples of different forcing

cases the baseline separated condition is reestablished between consecutive control cases

to eliminate confusion on the results due to hysteresis effects.

59
Chapter 4: Dielectric Barrier Discharge Plasma Actuators

4.1 AC DBD Plasma Actuator Design

As described in Section 2.2.1, AC-DBD plasma actuators which integrate

relatively thick dielectrics (on the order of a few mm) into the model geometry create the

greatest induced flows due to their ability to withstand higher voltage inputs without

entering the Corona or streamer mode (Corke et al. 2009). These types of designs are

preferred for second generation type studies where the actuator is cast into the model

design. For this initial work it is essential to maintain both the flexibility and modular

nature of the actuators due to the variable separation location and number of parameters

(actuator location, geometry, orientation and number) which are intended for

investigation. More importantly, we wish to be able to move these devices without

modifying the airfoil. For these reasons, DBD plasma actuators whose construction is

based on thin flexible adhesive materials (i.e. tapes) are chosen. Such devices are

inexpensive, readily available and easily removable which allows the actuator location

and orientation to be varied. The thin profile and ability to conform to surface curvature

are also appealing since this allows the application of the device to the surface with

minimal alteration of the basic features of the flow field.

60
It is widely believed that the electrode material is much less important than the

dielectric when inducing flows based on DBD plasma discharges (Hoskinson et al. 2008).

Accordingly, the most common electrode used in the literature, copper, is selected for this

study and no attempt is made to optimize this choice. It has a total thickness of 0.09 mm

(0.0035 in) and is bonded with an acrylic adhesive that is 0.05 mm (0.0021 in) thick. The

exposed and covered electrodes have widths of 6.35 mm (0.25 in) and 12.7 mm (0.50 in),

respectively. A ground electrode width of 12.7 mm (0.50 in) allows the use of standard

25.4 mm (1 in) wide dielectric tapes. The ground electrode width must be large enough

so the plasma formation is not limited in the streamwise extent since this has a strong

effect on the induced flow. For the input waveforms in this work, a covered electrode

width of 12.7 mm is sufficient.

Various studies recommend the use of a slight gap (1-5 mm (0.08-0.20 in))

between exposed and covered electrodes (Roth and Dai 2006; Forte et al. 2007). The

latter work showed a modest velocity increase (50 cm/sec) from the 0 to 5 mm (0.20 in)

gap case. Because of the difficulty repeating this exact gap size for multiple iterations,

some prefer the use of no gap or slight overlap between exposed and covered electrodes

(Corke et al. 2007). The zero gap recommendation has been employed in this work.

The self-imposed mandate that the actuator materials (both electrode and

dielectric) be composed of thin adhesive tapes limits the dielectric material selection to a

few basic choices. At the onset of this study, significant work was performed to

determine the best dielectric material and even combinations of various materials for

sustaining the glow discharge and inducing flow. At this time (2006), plasma actuator

61
studies were becoming very popular, but archived literature on optimizing the device

design was not readily available and various researchers had their own preference for

materials. Limiting these choices to readily available adhesives left primarily Kapton,

Teflon and/or ultra high molecular weight (UHMW) tape as possibilities. A significant

amount of time was spent to optimize the dielectric choice based on a balance of adhesive

properties, dielectric strength and induced flow capability.

Initial designs were composed of very thick layers of dielectric with an additional

layer of protective material under the ground electrode such as mica or thermally

conductive Kapton tape in order to prevent damage to the airfoil. With experience this

became less of a concern. The voltage trace could easily be monitored in real time and

when dielectric breakdown occurred, immediately shutting down the power supply was

sufficient to prevent damage to the airfoil substrate for normal operating parameters. For

more catastrophic events with significant arc formation, a small pit a few mm in diameter

is burned in the model similar to Figure 4.1. Even in this case, the damage can be

repaired by employing simple body filler. In the final design, the ground electrode is

directly applied to the model surface.

62
Figure 4.1: Airfoil damage due to arc formation.

The induced flow created by the device was only mildly affected by the dielectric

material. Rather the adhesive properties and ability to withstand arcing were deemed

most important in this application. Teflon and UHMW adhesive tapes were found

substantially inferior to Kapton in this regard when used in thin layers. This selection can

also be supported based on material properties. Referring to Table 4.1, which summarizes

various properties of dielectric materials, it can be seen that Kapton has medium to low

dielectric constant (ε~3.5) compared to other dielectrics surveyed, but its dielectric

strength is approximately an order of magnitude greater (154 kV/mm). Assuming that the

dielectric strength is an indicator of the ability for a dielectric to maintain operation in the

normal glow regime makes this a superior choice, especially for the purposes of this work

in which very thin adhesive materials are necessary. As mentioned earlier, the adhesive

and layering ability of Kapton tape has been found to be superior to others explored. This

layering decreases the bulk capacitance, but the increased dielectric strength allows
63
application of higher voltages without entering the arc regime. The Kapton tape in this

study has thickness of 0.09 mm (0.0035 in) and dielectric strength of 10 kV. Each tape

has a silicon adhesive layer that is 0.04 (0.0015 in) thick. The effect of the silicon

adhesive on the dielectric performance is not examined here. The total thicknesses of the

dielectric and the device as a whole are 0.44 mm (.0175 in) and 0.62 mm (.025 in),

respectively, unless otherwise noted. This actuator is stable for AC-DBD plasma studies

at frequencies of 1-3 kHz at voltages of ~20 kVpp. Voltages up to 30 kVpp have been

tested, but these higher potentials are substantially more prone to arc formation and

sparking to the model surface.

Table 4.1: Properties of various dielectric materials (Roth and Dai 2006).

64
The recommended carrier frequency for the Kapton dielectric in air is 5 kHz

(Corke et al. 2009). This has not been used due to transformer power saturation issues

and limited run time at this high frequency due to higher thermal loads on the dielectric.

Results suggest that this should have little effect on the maximum body force generated

due to the relatively broad peaks associated with thrust optimized excitation frequency

(Corke et al. 2009). A typical actuator like this lasts roughly one hour of non-continuous

actual run time. Note that in practical flight applications, more robust dielectrics that are

embedded in the substrate as well as close monitoring of the dielectric fidelity will be

required. Cursory investigations of induced flow created by negative going sawtooth

waveforms showed little difference compared to sinusoidal version of the same signal.

Consequently, sinusoidal carrier frequencies are used for simplicity.

Even though Kapton is preferred in this work, it is still found to degrade with run

time. Figure 4.2 shows the dielectric surface after a significant amount of run time (~1

hr). Note the white regions in the dielectric which appear almost like traces of plasma

filaments. This changes the dielectric properties and a quantitative monitoring system for

such material changes is required for the practical application of DBD plasma actuators.

Some researchers have suggested using combinations of Kapton and Teflon to increase

plasma induced flow and allow longer run time (Poggie et al. 2010). Visual inspection of

the latter seemed to show promise, but dielectric breakdown was not found to be

substantially delayed in practice nor was the induced flow radically changed. Thus, the

extra layers of Teflon are not employed in the final design. Note that rhodamine paint

(pink region) is also visible above and below the device.

65
Figure 4.2: Degradation of the dielectric surface after substantial plasma run-time.

The application of the actuator to the airfoil is carefully done by applying the

device in the selected location using pressure transducers or static pressure taps as datums

such that each layer of tape is adhered at the proper streamwise location. The electrodes

overlap for at least 3/4 span of the airfoil (~46 cm). To prevent arcing, the high voltage

cable is connected on the tunnel outfield side while the ground is connected on the tunnel

infield side. Care is taken to use sufficient insulation at the soldered joints of the leads

especially on the high voltage side. Structural joints and steel tabulations that form the

pressure taps in the airfoil must also be insulated using Kapton tape to prevent discharge

to these regions.

4.2 AC-DBD Plasma Actuator Characterization

Figure 4.3 shows simultaneously sampled traces of current and voltage for a

typical the AC-DBD plasma actuator test. As previously discussed, the majority of the

momentum is generated by the negative going portion of the negative voltage half-cycle

when the amplitude of microdischarges is lower and the discharge is more diffuse. The

large current spikes during the positive going half-cycle of the positive waveform
66
represent the more localized filamentary discharge. This portion of the cycle has been

experimentally determined to be less efficient for generating flow in air due to intense,

but more localized streamer discharges. As such, it functions more like a reset switch for

the momentum generating portion of the waveform.

Figure 4.3: Simultaneous current and voltage traces for a typical AC-DBD plasma
actuator.

The power dissipated in the discharge is most simply determined by integrating

the product of voltage and current over a given cycle with data like that shown in Figure

4.3. While the calculation is quite simple, simultaneous measurements of the voltage and

current is difficult due to the radically different time scales involved. The voltage
67
waveform has a period on the order of a few hundred microseconds depending on the

carrier frequency, but the current spikes associated with microdischarges occur on time

scales of a few tens of nanoseconds (Falkenstein and Coogan 1997). Thus, one must

either sacrifice resolution of current spikes in order to get a statistically significant

number of voltage periods or limit the number of voltage periods acquired in an effort to

resolve all the microdischarges in the current trace.

A more accepted approach is to use charge-voltage methods (Falkenstein and

Coogan 1997). In this case, a capacitor is connected in series with the ground electrode

and charge on the capacitor is measured simultaneously with the high voltage applied to

the exposed electrode. Since the high voltage waveform determines the charge

distribution on the capacitor, these two signals are on similar time scales. A plot of these

two simultaneously sampled parameters forms a parallelogram as shown in Figure 4.4

where each high voltage cycle represents one revolution around the figure. The area

enclosed by this curve is the energy dissipated per high voltage cycle. The corresponding

power, P, can then be calculated by simply multiplying by the frequency of the applied

voltage:

P= f ∫V
cycle
ac dQ
4.1

where f is the AC frequency, Vac is the voltage applied to the HV electrode and Q is the

charge on the capacitor.

68
[×103] 4

2
Charge [nC]
0

-2
41.2W

-4
-10 0 10
Voltage [kV]
Figure 4.4: Example of Q-V data for AC-DBD plasma.

Figure 4.5 gives an example of the typical power per unit length dissipated by the

actuator and calculated using Q-V measurements as a function of the applied voltage. In

this case, the discharge is generated by a 2 kHz carrier frequency and the dielectric is

composed of 5 layers of Kapton tape. The dissipated power is found to increase

proportionally with voltage to the power 3.51, consistent with existing literature (Corke et

al. 2010), where the constant of proportionality, α, is approximately 2x10-5. The data has

been plotted on a log-log scale to emphasize the validity of the fit at higher voltages

while highlighting considerable scatter of the data below 5 kVpp. This scatter occurs over

a voltage region where no discharge exists and the Vac7/2 power law is not valid. This

region is often referred to as being characterized by dielectric heating (Roth and Dai

69
2006). The electrical power per unit length, P/l, dissipated by the discharge is 0.74 W/cm

at 20 kVpp. In later experiments, the plasma carrier frequency is increased to 3 kHz and

the dielectric thickness is decreased to 3 layers (0.27 mm (0.01 in)). This changes the

power dissipation at 20 kVpp to ~1.7 W/cm. A plot of dissipated power as a function of

applied voltage for this updated arrangement would still follow the Vac7/2 power law, but

have a different constant of proportionality.

Figure 4.5: Dissipated power per unit length as a function of applied voltage.

70
The velocity induced by a single DBD plasma actuator in still air has been

characterized using PIV. The characterization is performed by mounting actuators on a

flat plate substrate made from the same material as the airfoil model (Duraform GF

resin). The generated plasma spans approximately 16 cm (6 in) of the flat plate substrate.

The flat plate is mounted parallel to the laser sheet (vertically) to minimize reflections

from the surface (Figure 3.22). The data is acquired near the mid-plane of the device.

Figure 4.6 shows a sample time-averaged total velocity field generated by the actuator

operating with a 2 kHz sinusoidal carrier frequency at a voltage of 20 kVpp with no

modulation. The actuator schematic is to scale in the streamwise direction, but expanded

vertically to allow visualization. For example, the dielectric is composed of 5 layers of

Kapton tape which has an actual thickness of ~0.5 mm (0.018 in). Figure 4.6 is presented

to show the global characteristics of DBD plasma induced flow in still air, but subsequent

tests using thinner dielectrics have been found to produce more induced flow at the same

input voltage and frequency. As documented in the literature, a quasi-steady suction

develops above the exposed electrode, but the flow field is dominated by the near wall jet

downstream of the actuator. The actuator in Figure 4.6 has been mounted near the edge of

the flat plate (white region) hence the jet spreads on both the positive and negative sides

downstream beginning near x~50 mm since there is no shear force from the surface in

this region. Time-resolved velocity and thrust measurements in the literature show that

the AC-DBD plasma actuator in air produces a dominant velocity pulse at the high

frequency driving signal corresponding to the negative going portion of the negative half

cycle as mentioned earlier (Enloe et al. 2008b). Because of the asymmetry of both the

71
electrode geometry and force production over the two waveform half-cycles, the AC-

DBD is more similar to a pulsed wall jet than a traditional ZNMF device in that a

nonzero mean flow is established.


Figure 4.6: AC-DBD time-averaged induced velocity magnitude, W , in quiescent air
(m/sec).

The usefulness of AC-DBD plasma for energizing boundary layers is certainly

evident from the near wall jet effect, however its efficacy as a more general flow control

actuator is dependent on the ability to influence natural flow instabilities. The majority of

experiments for separation control on airfoils deal with flow instabilities that are

approximately an order of magnitude less than frequencies associated with AC-DBD

plasma generation (1-10 kHz). To create excitation of this nature, the high frequency

carrier waveform is modulated using a low frequency signal in the range associated with

separated flow dynamics. Two examples of these modulating waveforms are shown in

Figure 4.7. The voltage traces have been acquired at the secondary (HV) side of the

transformer. The carrier frequency of the waveforms is 2 kHz and the modulation
72
frequency is 100 Hz. In Figure 4.7a, the high frequency carrier has been modulated using

a sine wave envelope, termed amplitude modulation (AM). Modulation using a variable

duty cycle as shown in Figure 4.7b allows more control over the input power and actuator

on/off time. The duty cycle shown is 50%. Modulation using this waveform is hereafter

termed burst modulation (BM).

a) AM b) BM

Figure 4.7: Modulation waveforms for AC-DBD plasma actuation.

The ability of these waveforms to produce velocity fluctuations in quiescent

conditions is examined using phase-averaged velocity and vorticity data for AM and BM

of 10%, 30%, 50%, 70% and 90%. The vorticity is calculated according to:

∂V ∂U
Ω= −
∂x ∂y
z 4.2

with a 2nd order accurate central difference scheme. For brevity, only one phase of the

oscillation is shown in Figure 4.8, but animations of four phases of the modulation period

73
confirm that the structures generated by plasma convect in still air. As in Figure 4.7, the

plasma is created using a carrier frequency of 2 kHz at 20 kVpp with modulation

frequency of 100 Hz. The velocity fields are averaged over 50 instantaneous phase-

locked samples which is sufficient for resolving the primary features of the pulsed flow.

Recall that AC-DBD plasma induced flows are dominated by the near wall jet (U

component). However, the vertical component of velocity (V) gives a better emphasis of

the pulsating nature of the device when excited using these modulated waveforms. Note

that the dominant feature shown is the pulsed suction initialized near the electrode

interface that is released over the covered electrode before creating a vortex train further

downstream as confirmed by vorticity data for the same phase. The sinusoidal

modulation (AM) displays a well organized vortex train commensurate with the

modulation frequency that is sustained for approximately 40 mm before dissipating. At

the lowest duty cycle (10%), the pulsed suction is quite weak and no vortex is visible

because not enough carrier cycles are produced to create significant plasma induced

flows. The 30% duty cycle case shows greater suction and some sustained vortex

behavior. As the duty cycle is further increased, pulsed suction near the electrode

interface is stronger and covers a much larger region while the generated vortices persist

further downstream. In the 70% duty cycle case, the suction near the electrode interface

also generates a secondary flow in the form of a stationary clockwise rotating vortex that

is visible just upstream of the exposed electrode. At 90% the pulsing nature of the

actuation is essentially lost with little primary vortex generation and no recognizable

structure content in the vorticity.

74
Figure 4.8: Phase-averaged AC-DBD plasma induced velocity fields, V , (left, m/sec) and
 , (right sec-1) in quiescent air for modulation using AM (top) and various BM
vorticity, Ω
(2nd from top to bottom) waveforms.

While the behavior of actuators in quiescent flows is not necessarily indicative of

their behavior in a flow control environment, these results suggest that greater control

authority can be achieved by changing the modulation waveform. This will be examined

in Section 6.1. Other methods of creating low frequency perturbations include providing

separate excitation signals to each electrode (Post 2004) or more simply using a very low

carrier frequency that is not optimal for plasma generation. The majority of the remaining
75
results are for BM using 50% duty cycle which is a reasonable choice for generating

control authority across the test cases surveyed.

Despite the usefulness of the wall normal velocity component for observing the

pulsing actuator behavior, the U component of the velocity is the dominant momentum

source. The momentum introduced by AC-DBD plasma is quantified using the PIV data

in Figure 4.9. In this case, the input frequency and dielectric thickness have been

optimized for practical application, run-time and flow generation. The dielectric is

composed of 3 layers of Kapton tape with total thickness of 0.27 mm (0.0105 in) and the

AC driving frequency is now 3 kHz. The plasma induced flow for this optimized actuator

has been measured for a range of forcing cases that encompass relevant excitation

frequencies for airfoil TE separation control. Profiles of U and Urms created by the

actuator 20 mm downstream of the electrode interface are shown in Figure 4.9. The

profiles have been forced to obey a no-slip condition at the wall. Plasma generation at a

frequency of 3 kHz creates a wall jet with maximum mean velocity of ~3.5 m/sec

measured ~1 mm from the surface (Figure 4.9a). Operating the actuator in burst mode

using square wave modulation at 50% duty cycle lowers the maximum induced velocity

but creates profiles that become substantially fuller as the burst frequency is decreased.

The fluctuating behavior of burst mode actuation follows a similar trend in profile

fullness, but the maximum value of the fluctuations now decreases with increasing

frequency (Figure 4.9b). Note that this behavior is also dependent on the carrier

frequency and duty cycle. The profiles in Figure 4.9 have been used to calculate both

76
mean, J/ρ and oscillatory <J/ρ> components of momentum with a constant density, ρ,

assumption (Enloe et al. 2006) using:


J
= ∫ U 2 dy
ρ 0
4.3


J
< >= ∫ U rms
2
dy
ρ 0
4.4

where U and Urms are the time-averaged and rms velocity profiles. The x=20 mm

location is chosen since it is near the discharge, but just downstream of the region where

visible radiation from the discharge can contaminate the PIV data. It should be noted that

the downstream location at which the momentum should be calculated is not well defined

for AC-DBD plasma (Pons et al. 2005; Porter et al. 2007; Greenblatt et al. 2008;

Hoskinson et al. 2008). Thus, reported values should be taken as order of magnitude

approximation.

77
a)

b)

Figure 4.9: Mean (a) and rms (b) velocity profiles, U, at x=20 mm for AC-DBD plasma
operating in quiescent air for various BM frequencies at 50% dc.

78
The momentum values are cast in terms of Cµ, <Cµ> and Cµ,tot for the various Re

surveyed in Section 6.1 to allow comparison to more common flow control actuators

using:

J
Cµ = 4.5
q∞ c
<J>
< Cµ >= 4.6
q∞ c
Cµ ,tot= Cµ + < Cµ > 4.7

where q∞ is the freestream dynamic pressure and c is the airfoil chord. It should be noted

that each δf has an optimal control frequency of burst mode excitation ranging from 70-

400 Hz and these are not all encompassed by the data in Figure 4.9. Consequently, values

of J/ρ and <J/ρ> for a given burst frequency are interpolated using an exponential fit to

the three burst mode cases as shown in Figure 4.10. This interpolated value is then used

for expressing momentum coefficients in Table 4.3 for the various optimal forcing

frequencies. No such interpolation is necessary for the unmodulated data in Table 4.2.

79
Figure 4.10: Exponential profiles used to interpolate momentum values for various AC-
DBD plasma modulation frequencies.

Re/1000 J/ρ (N/mm) <J/ ρ> (N/mm) Cμ % <Cμ> % Cμ,tot% CE%

240 11.8 1.3 0.041 0.005 0.046 15

410 11.8 1.3 0.015 0.002 0.017 3.3

750 11.8 1.3 0.005 0.001 0.006 0.57

Table 4.2: Momentum and power characteristics for 3 kHz actuation at 20 kVpp.

80
Re/1000 fm (Hz) J/ρ (N/mm) <J/ρ> (N/mm) Cμ % <Cμ> % Cμ,tot% CE%

δf=20o

240 120 11.9 2.3 0.042 0.008 0.050 7.5

410 210 9.3 1.3 0.012 0.002 0.014 1.7

750 400 5.6 0.7 0.002 <0.001 0.002 0.29

δf=30o

240 90 12.9 2.6 0.045 0.009 0.054 7.5

410 160 10.7 1.9 0.014 0.002 0.016 1.7

750 270 8.1 1.2 0.003 0.001 0.004 0.29

δf=40o

240 70 13.5 3.0 0.047 0.011 0.058 7.5

410 120 11.9 2.3 0.015 0.003 0.018 1.7

750 230 8.9 1.4 0.004 0.001 0.005 0.29

Table 4.3: Momentum and power characteristics for 3 kHz actuation at 20 kVpp with
various burst frequencies at dc=50%.

Similar momentum calculations for the conditions of Figure 4.8 have also been

performed in an effort to quantify the dependence of momentum generation on the

modulation signal. An examination of the integral values in Figure 4.11 shows some

weak correlation with the results observed in Figure 4.8. For example the oscillatory

momentum coefficient peaks at 70% dc which is consistent with observations of the

81
phase-averaged V component in Figure 4.8, especially in the oscillatory case. The drop

off above 70% dc is also captured. The mean momentum coefficient values are shown to

saturate at 50% dc. The displays one of the limitations of using such integral parameters

as a comparison of the 50% and 100% dc profiles in Figure 4.9 reveal very different

behavior. The use of momentum coefficient will be revisited in the context of control

results in Section 6.1.

Figure 4.11: Mean and oscillatory momentum values for the AC-DBD plasma induced
flow in quiescent air for modulation using AM and various BM waveforms.

82
Before discussing control results, some comments on the Cμ calculations are

warranted. ZNMF actuation is generally characterized using only oscillatory momentum

coefficient <Cμ> due to the periodic blowing suction behavior. The DBD plasma actuator

is certainly a ZNMF device, but the analogy with more traditional ZNMF devices such as

those used by NASA is not warranted since the plasma also produces a nonzero net mean

flow (Greenblatt et al. 2008). Thus, the characterization of such a device is more similar

to nonzero net mass flux devices. When the wall normal plasma induced flow is

neglected as is the case here, the actuator is essentially a pulsed jet and thus its

characterization requires the inclusion of both mean and oscillatory components of

momentum (Greenblatt and Wygnanski 2000). Despite this additional mean component,

Cμ,tot is still approximately an order of magnitude less than used in the majority NASA

experiments for a scaled version of the same airfoil. It is also important to note that

values of Cμ for periodic excitation produced by piezoelectric devices in the presence of

cross-flow have been obtained by correlating the pressure in the actuator cavity with the

velocity at the slot exit in bench top experiments. The pressure in the actuator cavity can

then be monitored during cross flow experiments giving a more accurate representation

of the actual momentum production (Melton et al. 2006). Calibrations for DBD plasma

actuators performed using nonzero freestream have resulted in even lower estimates of

momentum coefficient, thus the values reported in Table 4.2 and 4.3 are most likely an

overestimate further emphasizing the low momentum nature of the device (Goksel et al.

2006). The power dissipated by the discharge is also provided in Table 4.2 and 4.3 in

dimensionless form as power coefficient, CE, using:

83
P
CE = 4.8
q∞U ∞ c
where P is the time-averaged power dissipated by the plasma. Q-V power measurements

for modulated signals are more complex than the pure sine waves discussed previously.

Figure 4.12 gives an example of the Q-V curves for a 2 kHz carrier waveform modulated

using 100 Hz AM. The parallelogram is formed due to oscillations of the carrier signal,

which when modulated using the sine wave, expands and shrinks according to the peak

voltages. This complicates the Q-V area calculation and necessitates that the modulation

frequency must be taken into account.

Figure 4.12: Q-V diagram for AC-DBD plasma actuation using AM at 100 Hz.

84
Figure 4.12 shows the integrated energy dissipated on the actuator for

approximately one period of various 100 Hz modulation waveforms. The unmodulated

signal is also included for reference. The slopes for each condition are nearly identical

and the only obvious difference is the continuity of energy addition which changes based

on the modulation signal. Note that the curves are offset on the time axis for clarity. The

energy dissipated in the modulation period is found by noting the value associated with

the maximum energy (i.e. when the curve flat lines) and the corresponding power is

found by dividing by the modulation period. From this figure, it is apparent that power

and energy calculations for modulated waveforms can be simplified by multiplying the

power of 100% duty cycle case by the duty cycle of the modulation waveform in

question. For example, the 50% duty cycle case corresponds to approximately half the

energy of the 100% duty cycle case. The AM case is a more complex since the integrated

energy tends to have rounded corners associated with the more gradual sinusoidal

envelope rather than the abrupt on/off seen in the duty cycle case. Note that slightly more

than one modulation period has been captured in this case as another cycle begins near

t=5 ms. Despite these complexities, one can see that the energy associated with AM

saturates near ~40% of the maximum power (i.e 100% duty cycle).

85
Energy [mJ]
200 AM=100 Hz
Duty Cycle 100%
Duty Cycle 75%
150 Duty Cycle 50%
Duty Cycle 25%
Duty Cycle 10%
100

50

-5 0 5 10
Time [ms]
Figure 4.13: AC-DBD plasma integrated energies for various modulation waveforms.

The simplification outlined above for calculating energy and power of modulated

signals is sufficient in this work, but for cases in which very few high frequency carrier

signals are developed (i.e. very low duty cycle), the phase of the carrier frequency at

which modulation is turned on/off becomes important due to the current behavior of the

AC-DBD actuator (see Figure 4.3). This low duty cycle complexity will also be relevant

to momentum calculations. While this is an interesting thought exercise, optimizing

power requirements to this degree is not crucial until these devices become more

established in practical applications.

86
4.3 NS DBD Plasma Actuator Design

The NS-DBD plasma actuator is constructed exactly like the AC-DBD plasma

actuator. The only additional consideration now is that the cables connecting the high

voltage and ground electrodes must be as short as possible and of similar length as

previously described. To facilitate this, both leads are connected on the tunnel infield

side. Note that a typical NS-DBD plasma actuator lasts longer roughly twice as long as an

AC-DBD actuator due to short time scales involved with the NS duration pulse which

limits the time the dielectric is actually exposed to the discharge.

4.4 NS-DBD Plasma Actuator Characterization

Characterization of nanosecond pulse (NS) DBD plasma actuators highlights the

fundamentally different nature of the device when driven by these short high voltage

waveforms. Typical traces of simultaneously sampled voltage and current for one pulse

are shown in Figure 4.14. These traces have been acquired on a 30 cm long DBD load.

The construction of the actuator in the NS-DBD case is identical to the AC-DBD. It is

composed of copper tape electrodes as described previously and 3 layers of Kapton tape

dielectric (0.27 mm (0.0105 in) thick). The voltage pulse width (<100 ns) is at least three

orders of magnitude shorter than a typical AC driven DBD plasma period. The pulse

voltage is ~15 kV for this particular load (~30 cm). For longer actuators such as those on

the airfoil (46 cm) the peak voltage is slightly reduced to 12 kV due to the additional

length of the actuator. In either case, the zero to peak voltage amplitude is not

substantially different than the AC-DBD plasma actuator, but the short pulse width and

rise time has a drastic effect on the current developed. The peak current in the discharge
87
is 50 A in the NS-DBD case compared to ~200 mA for the AC-DBD. The effect of this

high current is seen in the energy and power developed in the pulse. In the latter case,

peak powers of up to 600 kW are created. While this power is exceptionally high due to

the short pulse width and high current, the average power for the NS-DBD is slightly less

than the AC-DBD plasma actuator operating at the same frequency and similar voltage

levels.

88
20 60

a)
40
Voltage [kV]
10

Current [A]
20
0
0

-10 -20
-200 -100 0 100 200 300 400
Time [ns]
[×105] 8 20

6 b) 15

Energy [mJ]
Power [W]

4
10
2
5
0

-2 0
-200 -100 0 100 200 300 400
Time [ns]

Figure 4.14: Typical voltage, current (a) and power, energy (b) traces for an NS-DBD
plasma actuator.

89
Tables 4.4 and 4.5 compare various electrical parameters for the two discharges.

The NS-DBD pulse energy is found to vary slightly with frequency and run time. The

latter has the form of a first order response to a step input with time constant of 400-600

seconds that is likely due to thermal effects in the power supply. Given that control

experiments were run for periods of less than 600 sec, the results in Table 4.4 are likely

an upper bound. Comparison of the two tables shows that for the 46 cm long actuator, the

NS-DBD consumes slightly less power than the AC-DBD plasma actuator operating at

the same frequency. Note that none of the issues encountered with simultaneous AC-

DBD measurements are encountered for the NS-DBD plasma actuator since the current

and voltage time scales are quite similar. In all the control cases explored in Chapter 7

which include tests in the range 60-2750 Hz, CE is less than 0.6%.

Frequency (Hz) Voltage (kV) Energy (mJ/pulse) Power (W)

100 ~12 < 11 < 1.1

500 ~12 < 14 <7

1000 ~12 < 17 < 17

Table 4.4: Electrical properties of NS-DBD plasma on a 46 cm long flat plate actuator.

90
Frequency (Hz) Voltage (kVpp) Energy (mJ/cycle) Power (W)

1000 18 26 26

2000 17 21 42

3000 18 27 81

Table 4.5: Electrical properties of AC-DBD plasma on a 46 cm long flat plate actuator.

The ability of NS-DBD plasma to induce flow in quiescent air is investigated

using PIV experiments similar to the AC-DBD case. As before, the length of the actuator

on the flat plate is ~16 cm and this smaller load allows the generation of higher NS-DBD

peak voltages. Previous literature suggests the NS-DBD produces very weak velocity in

the neutral species (Roupassov et al. 2009) although some authors have studied the NS-

DBD induced vortex as a thrust generating mechanism (Opaits et al. 2008). The latter is

imaged using PIV raw data in Figure 4.15. The flat plate is arranged vertically and the

plate surface is visible by the laser reflection near x=-10 mm. When the NS-DBD plasma

is off (Figure 4.15a), there are some very weak ambient air motions seen by the vertical

striations near the flat plate surface, but these do not influence the discharge induced

flow. Initiation of the plasma discharge, seen in Figure 4.15b as the bright region on the

left side of the image creates a very large counterclockwise rotating vortex that is

qualitatively similar to published schlieren images (Opaits et al. 2007). This vortex is

quite slow (~4 cm/sec) and only occurs during the initial plasma generation at high

frequencies (≤~1 k suggesting the ambient fluid requires multiple plasma bursts to
91
produce this response. Once it propagates out of the frame of view, no other structures

are observed in the PIV raw data. Consequently, these effects do not influence the PIV

data. The rapid heating and expulsion of fluid near the surface by the initial plasma

propagation is believed to create this behavior and this is not observed in ensemble

averages. When the plasma is turned off (Figure 4.15c), charging the dielectric and its

repulsion of the olive oil seed particles is very apparent. This is displayed as a void in

particles near the dielectric and flat plate surfaces. The only region in the image that still

has particles near the surface is the area near the exposed electrode which does not suffer

from a similar surface charging effect. This surface charge buildup has been postulated as

an important phenomenon for AC-DBD plasma induced flow since it changes the electric

field distribution, positively or negatively affecting flow generation (Opaits et al. 2009b;

Font et al. 2010), but behavior to the extent observed in Figure 4.15 is not observed for

AC-DBD PIV raw data in this work.

92
a) b) c)

10 mm

Figure 4.15: Startup vortex (b) and dielectric charging effect (c) created by NS-DBD
plasma actuator.

The weak induced velocity of the NS-DBD is confirmed by the data in Figure

4.16 which shows the mean velocity created by NS-DBD operating at 2 kHz for 100%

and BM at 90 Hz with 50% duty cycle measured 20 mm downstream of the electrode

interface. The weak magnitude of the NS-DBD induced flow is highlighted by plotting

the velocity generated by an AC-DBD operating at the same frequency at 20 kVpp. In

both cases, the AC-DBD is found to generate at least five times more peak velocity than

the NS-DBD. The limitations of the AC-DBD plasma actuator due to its low momentum

nature are well-established. By comparison, the NS-DBD is even weaker such that no

control authority is likely manifested by NS-DBD generated momentum at the Re

surveyed in this work. Consequently, values for momentum and corresponding

momentum coefficient for the NS-DBD are irrelevant. The weak velocity created by the

93
NS-DBD is thought to be a consequence of the extremely short time scales associated

with the plasma discharge which are not sufficiently long to allow substantial fluid

response.

94
a)

b)

Figure 4.16: Mean velocity profiles, U, for NS and AC-DBD plasma operating in
quiescent air with 2 kHz carrier frequency using 100% (a) and 50% (dc) at 90 Hz
measured 20 mm downstream.

95
In Section 7.1 the NS-DBD is shown to generate superior control authority over

AC-DBD plasma actuators for LE airfoil separation control at high speeds. The preceding

velocity measurements suggest that the momentum addition cannot be a control

mechanism for the Re surveyed here. Instead, it is postulated that a thermal mechanism is

responsible. This scenario is well established for LAFPAs that have shown control

authority in subsonic and supersonic cold and hot jet experiments (Samimy et al. 2007)

and have also recently been supported by computations (Gaitonde 2009). One of the

characteristics of this control mechanism is the generation of compression waves due to

rapid local heating by the plasma. In the case of LAFPAs, these waves propagate

spherically as if generated by point source (Samimy et al. 2010). Similar, but

substantially more complex behavior has been confirmed for NS-DBD plasma using

phase-averaged schlieren imaging.

Figure 4.17 shows the compression waves imaged by viewing the discharge along

the major axis of an actuator. In this case, the plasma forms from right to left and can be

seen as the bright region near y=1.25 mm. The behavior imaged in Figure 4.17 is

relatively complex as the wavefront is composed of both a quasi-planar and cylindrical

compression wave, but appears relatively uniform along the viewing axis. This behavior

is consistent with numerical results of (Roupassov et al. 2009). However, the schlieren

technique by nature performs a spatial average normal to the field of view. The three

dimensionality of the discharge is observed by performing schlieren measurements

viewing the plasma from behind the high voltage electrode along its minor axis (Figure

4.18 and Figure 4.19). It is now apparent that what appeared to be relatively two

96
dimensional behavior is substantially more complex in that the cylindrical compression

wave along the span of the device is actually the summation of multiple spherical

compression waves created by localized regions of the discharge. Figure 4.18 shows

measurements taken with the schlieren knife edge vertically. This is done to highlight

density gradients along the y-coordinate and the quasi-planar shock behavior is well

imaged using this technique. Rotating the knife edge allows visualization of density

gradients along the x coordinate. In this case (Figure 4.19), the quasi-planar compression

wave is no longer visible. Note that the spherical compression waves are visible for the

knife edge in both orientations.

Figure 4.17: Schlieren imaging of compression waves generated by NS-DBD plasma


actuator viewed along the major axis of the actuator.

97
Δt=20 μs Δt=130 μs Δt=240 μs

Figure 4.18: Schlieren photography of vertical density gradients for NS-DBD plasma
generated compression waves viewed along the minor axis of the actuator.

Δt=49 μs Δt=159 μs Δt=269 μs

Figure 4.19: Schlieren photography of horizontal density gradient for NS-DBD plasma
generated compression waves viewed along the minor axis of the discharge.

98
While the compression wave behavior away from the surface is visually

compelling, the efficacy of the device for flow control is more likely dependent on the

compression wave strength in the boundary layer of the flow field to be controlled. For

example, in the airfoil results that follow the maximum boundary layer thickness (~7

mm) is encountered on the TE and the LE should be considerably thinner. Thus it is more

appropriate to concentrate on the NS-DBD behavior near the surface. At the onset of the

NS pulse, both the planar and localized compression waves are visible, but the latter

appears to be the stronger of the two as it propagates slightly faster as seen from the

curved region protruding out of the quasi-planar wavefront in Figure 4.18. As the time

delay increases, the spherical and planar waves merge such that the only evidence of the

spherical pattern is the various circular arcs beneath planar wave. Recall that these are

phase-averaged data which mean the local “hot spots” in the discharge that produce these

stronger spherical wavefronts must be somewhat spatially stationary. At this time, it is

not possible to predict the location of these strong filaments, however it is postulated that

these intense regions could be manipulated by shaping the electrodes, perhaps in a

sawtooth fashion, to create stronger regions of electric field.

To truly understand this device, quantitative near surface high spatial resolution

measurements of temperature rise and compression wave strength are necessary.

Measurements of the latter a few mm from the surface using existing data show

approximately sonic compression wave speeds (Mach 1) which are consistent with

LAFPA generated compression waves. However, existing literature and limited near

surface data suggest that the compression wave strength here may be substantially

99
greater. A rough estimate of this effect can be made based on the energy deposited by a

pulse, ~15 mJ from Table 4.4. Let us assume that 100% of this pulse energy is deposited

into a two dimensional near surface gas layer as heat. The temperature rise associated

with this heating can be estimated simply by:

E
∆T = 4.9
mCv
where ΔT is the temperature rise, E is the energy provided by the pulse, m is the mass of

the heated air and Cv is specific heat of air at constant volume which is used based on the

assumption that the fluid does not have sufficient time to expand during the heating

process (~50 ns). The mass of the air, m, or more importantly the volume of air used to

calculate this mass requires another assumption. The compression wave behavior near the

surface appears to show a wavefront that originates near the electrode interface rather

than over the entire extent of the 12.7 mm ground electrode. This heated region has been

estimated at 0.4 x 0.4 mm2 in the literature based on ICCD images of the NS-DBD

discharge (Roupassov et al. 2009). With this and the span of the device for the airfoil

cases in Chapters 6 and 7, (46 cm) the temperature rise near the surface is estimated to be

~235 K consistent with the literature (Roupassov et al. 2009). In reality, this is likely an

upper bound since the heating efficiency of the pulse energy is certainly not 100%. Also,

the two dimensional plasma assumption and the simplified calculation of heated air mass

are clearly not valid based on the schlieren images of Figures 4.18 and 4.19.

Using a similar temperature rise of ~200 K, compression wave strength near the

surface has been estimated at Mach 1.13 using one dimensional theory (Roupassov et al.

2009). This is again probably an upper bound since it is unlikely that a Mach 1.13
100
compression wave dissipates to approximately Mach 1 over one mm of propagation as

calculations from schlieren images like those in Figure 4.17 suggest. Still this may

provide some insight for quantifying a momentum or better amplitude coefficient for

these thermal perturbations. If we assume a Mach 1.1 compression wave that behaves

like a normal shock in atmospheric pressure (~100 kPa), the pressure difference across

the wave Δp is ~25 kPa using simple one dimensional normal shock relations. This can

be cast as a momentum coefficient analogous to Equation 4.6:

∆pz
Cµ ,∆p = 4.10
q∞ c
where z is the width of the heating region, assumed to be 0.4 mm and the supscript, Δp,

denotes the coefficient is defined based on a pressure difference rather than a

conventional momentum value, J. Using the pressure difference across a Mach 1.1

compression wave at atmospheric pressure, 25 kPa, and Re=750k, consistent with test

cases from Chapter 7, Cµ ,∆p =~4%. This is a reasonable number compared to more

standard ZNMF devices (i.e. synthetic jets) with substantial control authority (Cμ~1%).

This very simplified analysis may hint at the robustness of thermal effects for controlling

high speed flows. For example, even a very weak shock (Mach 1.02) for the Re=750k

airfoil case results in Cµ ,∆p =~0.6%.

An open question is whether the strength of the near surface compression wave

can be increased using bursts of these short duration pulses similar to the AC-DBD

signals of Figure 4.7. For this to occur, localized heating the near surface gas layer should

occur on a time scales shorter than the acoustic time scale. Again using an estimate of a

101
0.4 mm gas layer for air at 300 K, this heating should occur in less than ~1.2 µs which

means that if bursts of the NS-DBD pulses are used, they should probably be provided at

frequencies greater than 800 kHz.

It should be noted that similar compression wave behavior does not exist in the

literature for AC-DBD plasma actuators and thermal effects from these devices have been

ruled out as a control mechanism by multiple studies (Enloe et al. 2004b; Jukes et al.

2006; Sung et al. 2006). The compression wave behavior described above is also quite

repeatable for bursts of multiple pulses suggesting the dielectric charging has little

influence on the thermal mechanism.

102
Chapter 5: Baseline Verification and Characterization

Supercritical airfoils like the simplified NASA EET model are prone to stall at

relatively low α due to a small radius of curvature at the LE (Abbot and vanDoenhoff

1959). In practice, this is circumvented using a deflectable LE droop or slotted slat which

allows extension of the stall angle to higher α by promoting boundary layer transition and

weakening the adverse pressure gradient near the LE (Figure 5.1). This PFC device has a

significant effect on the CP distribution around the model as suction peaks are now

developed at the LE and over slat/droop shoulder (Melton et al. 2005). Such PFC devices

are also quite common on airfoils with deflectable TE flaps since these systems alone

serve to increase the effective camber of the model by increasing CL at a given α, but also

introduce stall at lower α.

Figure 5.1: Simplified high-lift airfoil with LE droop and TE flap.

103
The OSU version of the NASA EET is designed without a LE droop for

simplicity and the consequences of this decision are visible in Figure 5.2, which shows

CL versus α for selected Re and δf. The OSU airfoil reaches CL,max at relatively low α (6-

10o). Deflecting the flap increases CL over all α, but introduces stall at lower α as

expected from a change in effective camber. In addition, some Re effects are apparent

especially between the low (240k) and high Re (750k) test cases in that the generated CL

increases with Re at a given α. Sample CP curves in Figure 5.3 give more detail of the LE

stall behavior. The flow abruptly separates from the LE rather than gradually moving

upstream from the flap shoulder as α is increased, beginning with a reduced and

broadened LE suction peak at α=6o. This is also accompanied by a closed separation

bubble slightly downstream of the LE near x/c=0.1 in the α=6o and α=8o cases before

deep stall, recognized by the near-zero pressure gradient on the suction surface is

established at α=10o.

104
Figure 5.2: Baseline CL vs. α for sample Re and δf.

Figure 5.3: Baseline CP behavior for Re=410k and δf=20o at various α.


105
The flow field over the deflected flap is characterized by vortex shedding at Fc+

between 1-3 depending on δf where larger flap deflections create lower frequency

shedding (Figure 5.4). The dynamic signature of vortices shed from the flap shoulder is

independent of α until the separation point moves upstream after which a turbulent

spectrum with no recognizable frequency peak is established (Figure 5.5). The frequency

of vortex shedding obeys F+ scaling as expected (Figure 5.6). The flat spectrum

beginning around Fc+=10 in Figure 5.6 is created by an overly noisy measurement system

and is unphysical. The behavior of flow separating from the deflected flap shoulder as

described in Figures 5.2-5.6 is consistent across the parameter space surveyed. Separation

from the LE is much more chaotic, with no recognizable dynamic signature (Figure 5.7).

Figure 5.4: Baseline PSD of fluctuating pressure, cp, measured at x/c=0.95 for Re=410k
and α=0o at various δf.
106
Figure 5.5: Baseline PSD of fluctuating pressure, cp, measured at x/c=0.95 for Re=410k
and δf=20o at various α.

Figure 5.6: Baseline PSD of fluctuating pressure, cp, measured at x/c=0.90 for α=0o and
δf=30o at various Re.

107
Figure 5.7: Baseline PSD of fluctuating pressure, cp, measured at x/c=0.40 for Re=750k,
δf=0 at various α.

A direct comparison of OSU and NASA results without LE droop in Figure 5.8

shows three significant differences: 1) The pre-stall dCL/dα for the two airfoils is

considerably different when δf=0; 2) the difference in stall angle between the two airfoils

ranges from ~2o for δf=0 to ~4o for δf =30o and 3) even when the pre-stall dCL/dα are

similar, the OSU results are shifted by CL~-0.1. The latter discrepancy is most easily

explained. A comparison of the CP curves for the two models at α=8o with δf=0 shows

good LE suction peak agreement, however NASA EET creates a slightly fuller profile

over the main element (Figure 5.9). This is partially due to the joint between the flap and

airfoil main element in the OSU case. This CP agreement persists for moderate flap

deflections (< 20o).

108
Figure 5.8: Baseline comparison of OSU and NASA (Melton et al. 2006) results for
tripped and untripped behavior of CL vs. α for Re=750k and various δf.

a) b)

Figure 5.9: OSU (a) and NASA (b) CP curves at Re=750k, α=0 and δf=0 (Melton et al.
2003).

109
Figure 5.10 compares the stall characteristics of the OSU and NASA airfoils at

Re=410k with δf=12o. The only obvious discrepancy between the two data sets occurs at

α=8o where the NASA model appears to enter a deeper stall than OSU which retains a

broad, but visible suction peak at the LE with secondary peak downstream near x/c=0.15.

This dual peak signifies a closed separation bubble on the LE and is prevalent at many of

the post-stall conditions. This is the cause for the OSU flat post-stall CL behavior in

Figure 5.8 compared to the drastic CL decrease seen in the NASA curves. Note also that

the unpublished NASA data from Figure 5.10b shows that some Re effects were

prevalent at the low Re cases since the NASA CL curves at α=8o show no sign of stall at

Re=750k even for δf=30o.

-6

P=3166, cL=1.21, a=2.0, ds=0.1, df=12.2


P=3167, cL=1.33, a=4.0, ds=0.0, df=12.2
P=3168, cL=1.47, a=6.0, ds=-0.3, df=12.3
P=3169, cL=1.59, a=8.0, ds=-0.3, df=12.2
-4

a) b)
Cp

-2

0 0.2 0.4 0.6 0.8 1


x/c

Figure 5.10: OSU (a) and NASA (b) stall characteristics at Re=410k and δf =12o (Melton
2006).

110
The first two NASA/OSU discrepancies mentioned above can be addressed

together, but first note that the OSU version is 5/8 scale of NASA and thus has a more

limited array of static pressure taps due to size constraints. This is especially true near the

flap shoulder as the NASA model is instrumented with taps that expose upon flap rotation

whereas OSU model does not. Nonzero δf, alone can rectify the dCL/dα discrepancy at

OSU pre-stall (Figure 5.8). The CP curves in Figure 5.11 highlight this behavior. At

δf=10o, Re=240k shows separated flow on the flap (Figure 5.11). Increasing the

freestream Re to 410k is sufficient transition the boundary layer to turbulent and prevent

separation at the flap shoulder indicated by the large suction peak now visible. This effect

can be seen over the entire suction surface as an increase in circulation is apparent. In this

case, sufficient resolution is available to capture this TE suction peak, but this is not the

case for δf=20o in Figure 5.11. Despite this lack of resolution, it is clear that δf=20o

rectifies the Re discrepancy and shows CP agreement at all cases surveyed. This suggests

that the increased circulation caused by the flap deflection encourages turbulent transition

farther upstream on the airfoil. However, this alone is not sufficient to cause immediate

transition at the LE since the stall angle is still quite different.

To further explore this behavior in cruise configuration (δf=0o), various tripping

mechanisms in form of grit roughness, low profile vortex generators and

backward/forward facing steps created by spanwise tape were applied to the airfoil LE.

Tripping devices are common on many airfoils and are used to accelerate the laminar to

turbulent transition in an effort to delay separation in regions of high adverse pressure

gradient. A sample of these results is also shown in Figure 5.8. Note that the trip strip

111
increases the stall angle by approximately two degrees in both cases and the δf=0 dCL/dα

now agrees well with NASA results. The presence of a passive plasma actuator at the

airfoil LE can also serve as a trip. This effect will be considered in section 7.1 which

describes studies of LE airfoil separation control. Additional tripping experiments are

able to more closely match NASA results, but this effect is limited to Re=750k and

similar improvements in performance are not repeatable at lower Re. Consequently, no

LE trip has been used in the TE separation control results that follow. Because of these

issues, results are limited to α≤6o for δf =20-40o due to boundary layer separation

upstream of the flap shoulder. Smaller values of δf are not analyzed because separation

occurs much further downstream of the flap shoulder (Figure 5.11). As a final check of

the validity of comparison of OSU baseline results to NASA, the pressure spectrum at

x/c=0.95 is shown in Figure 5.12. Despite the deflected droop in the NASA work, the

dynamics of the TE flow field remain quite similar.

112
a)

b)

Figure 5.11: Re and δf effects on OSU airfoil CP distributions.

113
a) b)

Figure 5.12: OSU (a) and NASA (b) PSD of fluctuating pressure, cp, at Re=240k, α =0, δf
=45, δs =-25 in (Melton et al. 2006).

Measurements depicting the three-dimensionality of the baseline flow are shown

in Figure 5.13 for the three primary flap deflections surveyed in this work at Re=410k

and α=0. Measurements are shown at ¼, ½ and ¾ span. The agreement is quite good even

for cases in which large separated regions which are inherently 3D exist on the flap.

Since baseline agreement between OSU and NASA has been verified, no blockage or

wind tunnel corrections for integral parameters have been employed. These corrections

will change values of integral parameters (CL, CD), but the effects of control are the

primary focus of this work and are believed to be conservative irrespective of whether

such corrections are employed.

114
a)

b)

c)

Figure 5.13: 3D behavior of CP for Re=410k, α=0 and δf=20 (a), δf =30 (b) and δf =40 (c).

115
The state of the boundary layer is an important parameter for separation control.

Since transition occurs somewhere on the main element downstream of the LE, it is

important to verify the boundary layer approaching the flap is turbulent so that results are

viable for practical large-scale applications. The boundary layer velocity profile has been

measured upstream of the flap shoulder (x/c=0.7) using PIV. A sample result is shown in

Figure 5.14. The results of these measurements are summarized in Table 5.1. In all cases,

the boundary layer follows a turbulent profile according to:

1/ n
U  y
= 
U∞  δ  5.1

where U is the average velocity profile, U∞ is the freestream velocity, δ is the 99%

boundary layer thickness. The exponent, n, varies from 4 to 6 and δ decreases in

thicknesses with increasing Re as expected. The shape factor, H= δ*/θ, where δ* is the

displacement thickness and θ is the momentum thickness remains relatively constant

around 1.4 ensuring the boundary layer approaching the flap is similar in all cases. The

baseline results show that even though LE stall characteristics between the NASA and

OSU models are different, the results are consistent at OSU pre-stall α as the separating

boundary layer is turbulent and has similar dynamic features upon separation.

116
Figure 5.14: PIV-measured boundary layer velocity profile with power law fit for
Re=240k, α=0, δf=30.

117
Re/1000 δf (deg) n δ (mm) δ* (mm) θ (mm) H

240 20 5.32 5.14 0.81 0.59 1.38

410 20 5.51 4.77 0.73 0.54 1.36

750 20 5.9 4.41 0.64 0.48 1.34

240 30 5 6.24 1.04 0.74 1.40

410 30 4.86 5.14 0.88 0.62 1.41

750 30 5.76 4.41 0.65 0.48 1.35

240 40 4.89 6.98 1.18 0.84 1.41

410 40 5.14 5.87 0.96 0.69 1.39

750 40 4.93 4.77 0.81 0.57 1.41

Table 5.1: Boundary layer properties at x/c=0.70 measured using PIV.

118
Chapter 6: Trailing Edge Separation Control

6.1 Single AC-DBD Plasma Actuator Control Results

A single DBD plasma actuator is applied to the flap shoulder of the airfoil as

shown in Figure 6.1. This is done by first deflecting the flap and then applying the

various layers of adhesive tape according to the description in Section 4.1. The resulting

actuator location is x/c~0.77 independent of δf. This location is an acceptable estimate of

the separation point where actuators are generally found to be effective for controlling

flow separation. Note that this is in contrast to the fixed actuator locations used by

NASA. Thus, the most relevant comparison of forthcoming results are made in reference

to NASA actuation locations at and between FWD and slot 3 which are located at

x/c=0.725 and 0.757 respectively in reference to the cruise configuration (δf=0) (Melton

et al. 2006) Despite the documented sensitivity of TE control authority to actuator

location, the following results are quite repeatable for DBD plasma actuators manually

applied to the airfoil surface.

119
Exposed HV electrode

Dielectric extent

Figure 6.1: Example of a DBD plasma actuator applied at the flap shoulder (x/c=0.77) of
the simplified NASA EET airfoil, δf=30o.

It should be noted that each time an actuator is placed on the model the baseline

flow changes slightly due to the presence of the actuator alone, most notably from the

surface discontinuity created at the actuator leading and trailing edge. Figure 6.2a shows

the effect of the actuator on the CP distribution around the model. There is some change

most notably near the LE, but overall this effect is marginal. For practical applications,

this is not a concern since the actuator should be embedded in the airfoil substrate nearly

flush mounted to the surface. A more important result in Figure 6.2b shows that the

dynamics of the flow field are relatively unaffected by the passive presence of an actuator

on surface. These results are consistent for the conditions surveyed. For the remainder of

the work, any reference to the baseline configuration pertains to the case where an

actuator is present on the model surface without plasma (power off).


120
a)

b)

Figure 6.2: Effects of passive (power off) DBD plasma actuator on CP (a) and PSD of
fluctuating pressure, cp, at x/c=0.90 (b).

121
Results of frequency sweeps for various Re, δf, actuator locations, input voltages,

modulation waveforms and carrier and modulation frequencies are presented in Figure

6.3a. The data are plotted against FL,m+. Only results outside of the ±0.02

uncertainty/repeatability band are relevant. The actuation cases that do not employ low

frequency modulation, referred to as quasi-steady actuation hereafter, have been plotted

at FL+=0 to reduce the abscissa range for clarity. In reality, FL+ is in the range 2.8-8.5 for

the cases shown. Only a relatively narrow band of frequencies, FL,m+=0.3-0.5, shows

clear increases in CL. In reality, this effective band of frequencies is even narrower than it

appears. Portions of the data in Figure 6.3a are plotted using a different scale in Figure

6.3b. The modulation frequencies are normalized by fTE in this case. A discrete frequency

preference is now obvious in that excitation at F*=1 generates the most lift enhancement.

In some cases, the subharmonic of this frequency (F*=1/2) also has an effect, but there is

no experimental evidence to support that vortex shedding from the flap is shifted to this

lower frequency. Rather, the increased CL value seems to occur through a realizable, but

less efficient amplification of the natural shedding. In general, the control authority is

found to decrease with increasing Re and δf. The frequency preference in Figure 6.3 is

similar to presented NASA results of lift enhancement for a slightly different version of

the same airfoil using piezoelectric driven ZNMF actuators. However, the DBD plasma

actuator control authority for low frequency excitation above FL,m+=1 and quasi-steady

forcing is significantly reduced in comparison. This is due to the low momentum nature

of DBD plasma compared to the NASA actuation which limits control authority to very

narrow frequency bands corresponding to natural flow instabilities.

122
a)

b)

Figure 6.3: Effect of dimensionless actuation frequencies FL,m+ (a) and F* (b) on ΔCL at
α=0o for various Re and δf.

123
A sample set of CP curves for the results of Figure 6.3 are presented in Figure 6.4.

The difference between the baseline and both actuation cases is clearly visible on the

flap. The quasi-steady actuation case at FL+=12.7 generates less suction on the flap than

the baseline case. Forcing at F*=1 increases suction on the flap and more importantly

enhances circulation around the entire model as the area enclosed by the CP curve

expands especially on the suction surface. This circulation increase and corresponding

upstream effect create the majority of the lift enhancement. This behavior is typical of TE

airfoil separation control in that a significant portion of the lift enhancement is due to

upstream effects rather than full reattachment of flow to the flap (Kiedaisch et al. 2006;

Melton et al. 2006). It should also be noted that the model scale and obstruction of static

pressure taps by the actuator do not allow the resolution of any suction peak near the flap

shoulder (x/c=0.75) as exhibited in NASA results (Melton et al. 2006). However, a slight

favorable pressure gradient is visible around x/c=0.7 for both forcing cases suggesting it

does exist. The lack of CP resolution here at least partially accounts for the increased ΔCL

values in the NASA case over OSU.

124
Figure 6.4: Baseline and controlled CP behavior for Re=240k, α=0o and δf =20.

Figure 6.5 provides a more illustrative example of the behavior exhibited in the

CP curves. The streamlines calculated from two-component PIV are shown for the three

cases of Figure 6.4. The effect of actuation is readily apparent on the time-averaged flow

field. Quasi-steady forcing (Figure 6.5b) serves to energize the boundary layer and move

the separation point downstream. This delayed separation has the effect of lengthening

and flattening the recirculation region and confirms the suggestion that a slightly stronger

suction peak should exist in the CP curve near the flap shoulder in comparison to the

baseline. The persistence of the recirculation region on the flap explains the diminished

circulation increase in comparison to forcing at F*=1 (Figure 6.5c) which has a drastic

effect of the time-averaged streamlines. In this case, the separation location has again

125
moved downstream, but now the time-averaged streamlines resemble almost potential

flow with only a slight recirculation region visible. This also confirms the existence of a

stronger suction peak near the flap shoulder in comparison to both the baseline and quasi-

steady forcing cases.

126
a) Baseline

b) FL+=12.7

c) FL+=12.7
FL,m+=0.5
F*=1

Figure 6.5: Baseline and controlled (dc=50%) time-averaged streamlines for Re=240k,
α=0o and δf =20o.

127
*
The time-averaged spanwise vorticity fields, Ω =Ωc / U ∞ , corresponding to the

streamlines of Figure 6.5 are shown in Figure 6.6. As expected, strong regions of

vorticity exist above and below the flap corresponding to the shear layer between the

high-speed freestream and recirculation region. Strong vorticity on the flap persists for all

cases emphasizing that even when time-averaged streamlines more closely follow the

flap surface as in

Figure 6.5c, there is still considerable interaction here from flow structures

excited by the plasma (Figure 6.6c). This is further discussed in the context of phase-

averaged measurements that follow. The vorticity fields additionally show the extent of

recirculation region and its effect on the time-averaged wake. Quasi-steady forcing serves

to delay separation, but lengthens the extent of the recirculation region in comparison to

both the baseline and F*=1 forcing cases. The latter case is found to reduce the size of the

recirculation region and pull the region of strong vorticity closer to the flap element.

These effects are visible in the vorticity on both the pressure and suction sides of the

airfoil downstream of the flap and are most evident from profiles at x/c=1.1 shown in

Figure 6.7.

128
a) Baseline

b) FL+=12.7

c) FL+=12.7
FL,m+=0.5
F*=1

*
Figure 6.6: Baseline and controlled (dc=50%) time-averaged normalized vorticity, Ω ,
for Re=240k, α=0o and δf =20o.

129
*
Figure 6.7: Baseline and controlled time-averaged vorticity magnitude, Ω at x/c=1.1 for
Re=240k, α=0o and δf =20o.

The time-averaged results of Figures 6.5-6.7 do not truly represent the physical

behavior of the flow. Recall that the most striking benefits of control are realized by

amplifying the natural instability of the trailing edge flow field. In a temporal sense, this

slightly delays boundary layer separation, but a more obvious result is the amplification

and organization of the natural vortex shedding downstream. This increases entrainment

of freestream momentum into the shear layer and recirculating region even beyond the

point of separation from the flap surface. While this does not serve to fully reattach the

flow, the reduced size of the recirculating region has a profound effect on the circulation

around the model and subsequently on the sectional CL as seen in Figure 6.3.

130
This organized nature of shedding structures is most apparent from measurements

of phase-locked vorticity (Figure 6.8) and pressure fluctuations on the TE flap where both

the broadband and the coherent portions become amplified considerably (Figure 6.9). The

latter case has been verified using the techniques from Section 3.4.2. This highlights the

detrimental aspect of forcing at F*=1 in that the dynamic loads on the flap due to vortex

shedding are substantially increased. Unfortunately, transducer and hot wire

measurements in the presence of plasma actuation are quite difficult due to EMI and

reliable results have not been obtained for the quasi-steady forcing cases. Instead, the

spatial correlation of PIV data is used to investigate the organization of the flow field in

the wake in Figure 6.10. The normalized spatial correlation of the fluctuating vertical

component of velocity, Rvv, is computed along the y/c coordinate associated with the

trailing edge of the deflected flap (y/c=yT) using

∫ v ( x, y ) v ( x + ∆x, y )dx
T T

Rvv ( ∆x, yT ) =∞ −∞
6.1
∫ v ( x, y )
2
T dx
−∞

This location is a reasonable estimate of the symmetry plane associated with the

spatial dynamics of the vertical fluctuating velocity component in the turbulent wake

(Figure 6.11). A correlation function is computed for each instantaneous fluctuating

vertical velocity field and the average correlation waveform is computed using the 1000

image sample set. The baseline and F*=1 forcing cases behave as expected from the

131
results in Figure 6.9. Both exhibit a similar wavelength with the latter being amplified

considerably due to forcing near the natural vortex shedding frequency. The quasi-steady

forcing case has similar correlation levels as the baseline, but with a slightly shorter

wavelength consistent with oscillations at a higher frequency. While the shift in the

correlation peak toward higher frequencies is weak, the trend is in general agreement

with pressure spectra of NASA using piezoelectric ZNMF actuation at high frequency

(Melton et al. 2006).

 , for Re=240k, α=0o


Figure 6.8: Phase-averaged normalized vorticity fields (ΔΦ=π), Ω
*

and δf =30 forced at FL =8.5, FL,m =0.4 (F*=1).


o + +

132
Figure 6.9: Baseline and controlled PSD of fluctuating pressure, cp, measured at x/c=0.90
Re=240k, α=0o and δf=30o.

Figure 6.10: Spatial correlation of the v component of velocity, Rvv, at y/c=yT for
Re=240k, α=0o and δf=30o for baseline and controlled (dc=50%) flows.
133
The POD method is used to gain additional insight into the dynamics of the flow

field (Holmes et al. 1996). The method decomposes an uncorrelated set of velocity field

snapshots into a set of spatial orthonormal modes or POD modes, ϕ i ( x ) , and modal

amplitude coefficients for these modes, ai(t). The decomposition is proper in the sense

that the modes are organized in order of decreasing energy content to produce an optimal

set of basis functions that capture the dominant characteristics of the fluctuating velocity

field. The POD expansion (6.2) expresses the fluctuations of the two-dimensional vector

field w ( x, t ) = u ( x, t ) , v ( x, t )  in terms of the modes that contain the major portion of

the kinetic energy in the flow.

N
w ( x, t ) ≅ ∑ ai ( t ) ϕ i ( x ) 6.2
i =1

where x in this case represents two-dimensional space (x,y). Each snapshot contains u and

v components of instantaneous velocity on the x-y plane near the tunnel centerline. For

the purposes of this work 1000 snapshots are used to derive the POD modes and

statistical quantities.

The first four normalized POD modes for the v component of velocity,

ϕ1− 4 ( y ) =
ϕ1− 4 ( y ) / U ∞ , for the baseline and controlled flows using quasi-steady

(F+L=8.5), F*=1 and F*=2 forcing at Re=240k, α=0 and δf=30 are shown in Figure 6.11.

In all cases, a recognizable oscillation exists in the trailing edge wake. This is expected

for the baseline and F*=1 forcing case and further supports the suggestion that

134
oscillations persist near this low frequency even for quasi-steady forcing. Forcing at

F*=1 shows more amplified spatial modes compared to the other cases and a

recognizable signature of structures emanating from the flap shoulder near the actuator

location. At F*=2, a clear oscillation exists in modes 3 and 4 at a wavelength consistent

with higher frequency forcing. These shorter wavelength cases have little effect on the

overall flow field and most importantly CL.

Figure 6.12 presents corresponding energy results for the data of Figure 6.11 and

further emphasizes the low dimensional organized nature of the low frequency forced

flow field (F*=1) as nearly 60% of the total energy is contained in modes 1 and 2. The

quasi-steady forcing removes energy from the first 2 modes and redistributes it to mode 4

in particular. The F*=2 case has lower energy content in modes 1 and 2, but slightly

higher values for modes 3-7. The low energy levels of modes 3 and 4 explain why the

interesting high frequency dynamic content in Figure 6.11 at F*=2 has little effect on the

overall flow field. Figure 6.12b shows cumulative energy of the data of Figure 6.12a

where it is seen that the increased energy in modes 1 and 2 for the F*=1 forcing case

accelerates convergence, while the quasi-steady does the opposite. The F*=2 case also

has weaker modes 1 and 2, but the higher energy levels in modes 3-7 are sufficient to the

make summation of this energy quite similar to the baseline. The data in Figures 6.11-

6.12 are representative examples for the parameter space surveyed. In general, the trends

outlined above tend to increase and decrease with increasing and decreasing control

authority. For example, at higher Re such as 410k, the quasi-steady forcing has little

effect on the POD energy content.

135
Mode 1 Mode 2 Mode 3 Mode 4

Baseline

FL+=12.7

FL+=12.7, F*=1

FL+=12.7, F*=2

Figure 6.11: First four normalized baseline and controlled (AM) POD modes of the v component of velocity, ϕ1− 4 ( y ) , for

Re=240k, α=0o and δf=30o.

136
a)

b)

Figure 6.12: Modal energy (a) and cumulative energy (b) for baseline and controlled
(AM) POD modes for Re=240k, α=0o and δf=30o.

137
NASA results showed modulating the high frequency waveform using a square

wave (BM) rather than a sine wave (AM) could increase the efficiency of actuation since

the duty cycle can now be lowered thereby decreasing power consumption (Melton et al.

2004). A similar study for three δf at Re=240k is shown in Figure 6.13 using DBD

plasma. Recall from Section 4.2 that the power consumption of AM is approximately

equal to BM at 40% duty cycle. Duty cycles of at least 20% create a significant

improvement over the lower cases and each flap configuration has a preferred range of

duty cycle percentage for control that increases with δf. These results show that BM can

be both more efficient and more effective for enhancing lift. This is due to the response

of the natural instability to on/off nature of the BM perturbation in comparison to the

more gradual sinusoidal envelope associated with the AM case although quantification of

this effect requires future study. While this conclusion generally agrees with NASA’s

results, the functional relationship between ΔCL and duty cycle is quite different (Melton

et al. 2004). In their work, ΔCL was shown to saturate at 45% duty cycle for a δf=20o at

α=6o for excitation near the flap shoulder. Further increases in duty cycle required more

power input, but showed none of the detrimental effects on CL exhibited here. This

discrepancy is likely due to the pulsed blowing nature of the DBD plasma compared to

pulsed blowing/suction in the NASA case although this has not been show in the

literature to the author’s knowledge. Note that studies of LE separation control with

plasma actuators report significant authority and even complete reattachment for duty

cycles as low as 6% (Benard et al. 2009b). The results of Figure 6.13 highlight the more

138
challenging problem of separation control over a TE flap which requires greater periodic

momentum input in comparison (Melton et al. 2006).

Figure 6.13: Effect of modulation waveform on ΔCL for Re=240k, α=0o and variable δf
when forcing at FL+=8.5, F*=1.

An attempt to capture the pulsing effect of the AC-DBD plasma actuator in still

air is shown in Figure 6.14 using the momentum definitions outlined in Section 4.2.

These measurements have been acquired for similar forcing conditions as those shown in

the δf=30 case. The intent of this calculation was to show some correlation between the

139
oscillatory momentum coefficient and ΔCL. There does appear to be some relationship

here as both functions peak around 70% dc, but the momentum curve does not drop off as

significantly at higher dc. This highlights the difficulty in comparing actuator still air

measurements to behavior in a flow control environment and shows that considerably

more work is required to quantify these amplitude effects (Seifert and Tillman 2009).

Figure 6.14: ΔCL and <J/ρ> as a function of BM dc for Re=240k, α=0, δf=30, FL+=8.5,
FL,m+=0.4, F*=1.

With this knowledge, the effect of modulation and carrier frequency can be

further discussed. It is well known that AC-DBD plasmas create most of the momentum

transfer in the forward stroke half cycle of the carrier frequency (Forte et al. 2007; Enloe
140
et al. 2008b). For example, 100 Hz modulation of a 2 kHz carrier frequency with 50%

duty cycle corresponds to 10 high frequency cycles per modulation cycle. A 4 kHz carrier

frequency with the same modulation frequency (100 Hz) and duty cycle (50%) would

contain 20 high frequency cycles per modulation cycle. This implies that for a given

carrier and modulation frequency, there may exist an optimum number of high frequency

cycles and subsequent relaxation time for creating the strongest perturbations. This

optimum would obviously be governed by the system in question, the time response of

the induced flow to the pulsed signal and how receptive a particular flow field is to

excitation. This also implies that for a given dielectric with optimum carrier frequency,

there exists an upper limit for low frequency modulation. As before, this is governed by

both a minimum number of high frequency cycles to affect the flow and necessary

relaxation time to create the perturbation effect.

These findings indicate that dielectrics with high optimized frequencies give the

most flexibility for exciting high bandwidth low frequency modulations. In practical

applications, this point may be moot since scaling by reduced frequency (F+) dictates that

as length scales increase frequency scales decrease thus easing requirements for high

frequency excitation. However, for the sake of argument, a future AC DBD plasma

actuator design criteria could center around selecting a suitably robust dielectric that has

a dimensionless optimum carrier frequency on the order of F+~10 to allow the high

frequency forcing benefits (Glezer et al. 2005) while retaining significant bandwidth in

the low frequency regime more traditionally investigated. The success of such a device

141
would hinge on the selection or design of dielectric materials specifically optimized for

DBD plasma actuation.

The efficacy of DBD plasma actuators for TE separation control at nonzero α is

shown in Figure 6.15. Characterization of the baseline flow indicates that the dynamics of

the TE flow field remain relatively consistent as long as the separation location does not

change (α<6o) (Figure 5.5). This is also apparent in control results as forcing at F*=1 for

a given δf creates a near constant increase in CL for α<6o. The diminished control

authority at α=6o is due to the movement of the separation point upstream of the actuation

location. To retain the benefits of AFC at α≥6o, some PFC or AFC actuator is required at

the LE to maintain attached flow along the main element up to the flap shoulder. As

expected, the control authority decreases with increasing δf and Re, respectively. In the

former, the progressively more severe adverse pressure gradient and centripetal

acceleration imposed along flap shoulder requires greater momentum input to delay

separation and enhance circulation around the model. The reduction in actuator control

authority as Re is increased is well-established in the literature for both DBD plasma and

more traditional ZNMF devices operating at a fixed amplitude (Melton et al. 2006;

Greenblatt et al. 2008).

142
Re=240k

Re=410k

Re=750k

Figure 6.15: Effect of Re, α, and δf on ΔCL for forcing at F*=1 using dc=50%.

143
The effect of control on CD has yet to be considered. NASA results show that

forcing can have a significant effect on Cdp that is strongly dependent on both Cμ and

actuation location and does not necessarily correlate with CL (Melton et al. 2004; Melton

et al. 2006). Due to the inaccuracy of Cdp for quantifying total drag values and the

obstruction of static pressure taps by the plasma actuator, these effects are not considered.

Instead, calculations of CD from PIV wake surveys that include contributions of turbulent

stresses are employed (Lu and Bragg 2003).

U  U   y  vv − uu   y 
.1 .1
CD = 2 ∫ 1 − d   + ∫  d   6.3
   c
2
−0.3
U ∞  U ∞  c −0.3 
U ∞

where U and U∞ are the time-averaged local and freestream dynamic pressure

respectively. The turbulent stresses must be included since measurements of static

pressure in highly turbulent regions are difficult to obtain.

The effect of control is apparent in the near wake velocity profiles, but this

phenomena disappears downstream (Figure 6.16a). Note that the profiles are offset for

clarity and the spanwise Re normal stress is assumed negligible. The uu stress behaves

similarly to the mean velocity in that the effects of control are seen in the near wake, but

quickly dissipate downstream (Figure 6.16b). Only the vv stress shows an effect of

control significantly downstream (Figure 6.16c). This profile with equation 6.3 suggests

that CD should increase for F*=1 forcing. While this effect seems apparent in the wake

profiles, the contribution of this term to the CD calculation is quite small, less than 10%

144
on average across the parameter space, such that overall calculations of CD at x/c=1.5

show no significant change between the baseline and control cases at α=0o due to slight

variations in the U and uu profiles. The uncertainty surrounding CD calculations from

wake surveys for these highly separated flows is well established (Zaman and Culley

2008). Thus, significantly more work is required to quantify the effects of DBD plasma

on CD in this flow field, but at present the effect of increasing CL via instability

amplification does not appear to be prohibitive from this perspective.

145
a)

b)

c)

Figure 6.16: Mean U velocity profiles (a) and Reynolds stresses uu (b) and vv (c) used
in CD calculations for Re=240k, α=0 and δf=20 at x/c=1.1, 1.2, 1.3, 1.4 and 1.5 (left to
right).

146
6.2 Perspective on OSU Single AC-DBD Control Results

As the NASA and ADVINT programs have shown, increasing the amplitude of

forcing generally serves to improve control authority until some saturation level is

reached. The scaling laws for this behavior are not well defined and an open question is

whether to cast this in terms of Cμ, velocity ratio, etc (Seifert and Tillman 2009). In this

work, the amplitude effect has been cast as Cμ and values are generally one order of

magnitude less than those used by NASA. Yet control results at low Re are comparable

when forcing at F*=1 seemingly rendering DBD plasma actuators quite efficient from a

Cμ standpoint when used primarily to amplify the natural instability. This analysis is

complicated by the vague definition of Cμ calculations for DBD plasmas as well as the

uncertainty on both the effects of direction and surface area/volume over which control is

applied. It should be noted that a complete optimization for DBD plasma induced flow

has not been employed in this work and significant improvements have been

demonstrated in the literature. Recent studies that show plasma induced thrust can be

increased by an order of magnitude suggest that application of these devices at more

practical takeoff and landing velocities speeds is realizable (Thomas et al. 2009). The

obvious question is whether additional increases in Cμ from DBD plasma follow similar

trends as those shown for Cμ increases in the ZNMF NASA work (Melton et al. 2006).

An investigation of this possibility requires more robust dielectrics and higher voltage

inputs which are beyond our capabilities at this time.

147
6.3 Attempts to Extend AC-DBD Plasma Control Authority

Literature has shown that the simultaneous use of multiple actuators distributed

along the airfoil chord has been more successful than the contribution from each actuator

alone (Greenblatt 2007). NASA studies of specifically TE separation control showed that

additional CL gains could be realized using actuators both upstream and downstream of

the flap shoulder. The relative phase between the two modulation waveforms was found

to be an important parameter. Operating the two devices in phase created additional lift

gains, but operating the two with a pi phase shift had a detrimental effect (Melton et al.

2004). This was studied using two AC-DBD plasma actuators straddling the flap shoulder

operating as a parallel circuit (Figure 6.17). Figure 6.18 shows that using both actuators

can create additional increases in control authority, but this has not yet been superior to a

single actuator placed in an optimized location. Future work is required to examine the

effect actuator phase on this system, but these studies are limited by the size of the airfoil

model and actuators. The latter cannot be significantly reduced in size without reducing

the induced flow generated by the device. Thus, there is a limitation on spacing these

devices around the airfoil. Quantifying the effect of two AC-DBD plasma actuators for

TE separation control will likely require a larger airfoil to allow more options for spacing

the devices.

148
Figure 6.17: DBD plasma actuators straddling the flap shoulder.

*
Figure 6.18: Normalized time-averaged vorticity magnitude profiles, Ω at x/c=1.1 for
Re=240k, α=0, δf=20 for baseline and AC-DBD actuators operating in phase at FL+ =12.7
and F*=1.

149
The effect of forcing direction on control authority of AC-DBD plasma for TE

separation control has also been investigated. The majority of published studies use

devices that produce force dominant body forces in the streamwise direction. Figure 6.19

shows the result of producing the body force upstream. In this case, the actuator has been

rotated about the electrode interface which is believed to be the region where force

density created by the plasma should be near maximum (Enloe et al. 2004b; Corke et al.

2007). The flow is forced using both quasi-steady and F*=1 forcing. The quasi-steady

forcing has a slightly detrimental effect on the CP distribution. Forcing at F*=1 still

generates some lift enhancement although the effect is decreased in comparison to the

normal orientation. This may seem like a surprising result, but given the control

mechanism, it is explainable. Recall that the effect of F*=1 forcing is to amplify the

natural instability rather than simply energize the boundary layer. Thus, the introduction

of any perturbation should be able to partially accomplish this. This is clear from Figure

6.19 although it is apparent that the downstream orientation is more effective. This

reverse orientation even creates similar dynamic features in POD results (Figure 6.20).

While in this case, the downstream orientation is most effective, the influence of the

direction of DBD induced forcing is an open question. Arrangements that generate

momentum over 180 degrees can easily be constructed allowing detailed studies of this

effect not easily obtainable with typical ZNMF actuators (Porter et al. 2009).

150
Figure 6.19: Effect of reversing AC-DBD momentum direction on CP.

151
Mode 1 Mode 2 Mode 3 Mode 4

Baseline

FL+=12.7

FL+=12.7, F*=1

FL+=12.7, F*=2

Figure 6.20: First four normalized baseline and controlled (AM) POD modes of the v component of velocity, ϕ1− 4 ( y ) , for

Re=240k, α=0o and δf=30o with reverse actuator at x/c=0.77.

152
6.4 Single NS-DBD Separation Control Results

In an attempt to extend control authority to higher Re, the effect of DBD plasma

actuators driven by repetitive nanosecond pulses is investigated. These devices have

demonstrated control authority up to Mach 0.74 in isolated testing for airfoil LE

separation control (Roupassov et al. 2009). An in-house developed pulser is used to test

the efficacy of this type of plasma actuator for controlling TE airfoil separation. In this

work, the construction of this device is identical to AC driven plasma actuators which

offer a good benchmark comparison especially in this flow field where effective control

parameters like frequency and actuator location have already been established.

Amplification of natural flow instabilities over a deflected flap has been identified

as an effective control strategy for the airfoil TE system. Control using NS-DBD plasma

has been investigated similarly. The characterization results shown previously suggest

that velocities generated by the device are too weak to have an effect especially when

compared to AC-DBD plasma. Instead, manipulation of the natural instability is proposed

to be accomplished using thermal perturbations. Results of multiple attempts at TE

separation control using this device are shown in Figure 6.21. The plasma is generated

using a 2 kHz carrier similar to the AC-DBD case and burst modulated using dc=50% at

F*=1. Pulses at fTE alone (FL+=0.4) are also tested. This is a representative example of

multiple experiments, none of which have yet to demonstrate that the NS-DBD plasma

possesses control authority at the TE. This is an unexpected result given the efficacy of

the NS-DBD plasma for LE separation control at high speeds in the literature (Roupassov

et al. 2009) and in-house (Chapter 7). Increasing the burst frequency in an effort to
153
increase the amount of energy deposited into the flow at the optimal forcing frequency is

also not successful (Figure 6.22). This is likely due to the reasons outlined in Section 4.4

where it is postulated that bursts of pulses should be applied at frequencies greater than

800 kHz to have a cumulative effect on the thermal mechanism. These results will be

further discussed in comparison to contrasting results for LE separation control in

Chapter 9.

Figure 6.21: NS-DBD effect on CP for Re=240k, α=0 and δf=30.

154
Figure 6.22: NS-DBD effect on CP for Re=240k, α=0 and δf=30 with increased energy
input.

155
Chapter 7: Leading Edge Separation Control

ZNMF periodic excitation is widely established as an effective forcing

mechanism for separation control in many aerodynamic systems. This has been

extensively explored for LE stall control on airfoils due to drastic lift and drag

performance gains realized by reattaching these massively separated flows. Initial studies

of AC-DBD plasma actuators also focused on this system (Roth et al. 2000; Post and

Corke 2004) and it has remained widely popular for demonstrating AC-DBD plasma

control efficacy. The authority of AC-DBD plasma for controlling LE separation has

been mostly limited to freestream velocities below ~30 m/sec although some claims of

control at higher speeds have been published and more are certainly to follow given the

new optimization techniques being suggested (Thomas et al. 2009).

The control of LE stall with AC-DBD plasma at relatively low speed is a mature

research area and further work is not compelling unless control authority is extended to

higher speeds. There is nothing radically new or different about the AC-DBD plasma

used in this work and similar performance as seen in the literature is expected when

applied to the LE of the OSU version of the simplified NASA EET airfoil. Consequently,

a detailed study of this type of actuation is not employed on the LE. Rather, the AC-DBD

plasma actuator is used as a benchmark for the performance of the NS-DBD plasma

actuator. The latter has shown control authority on an airfoil LE up to Mach 0.74 in
156
isolated testing (Roupassov et al. 2009), but both flow physics and control using this

actuator are not well-understood.

7.1 Single DBD Plasma Actuator Control Results: Downstream Orientation

The NS-DBD plasma actuator as described in section 4.3 is mounted at the LE of

the airfoil near x/c=0 similar to Figure 7.1. The OSU airfoil requires a boundary layer trip

to perform similarly to the simplified NASA EET. When the NS-DBD plasma actuator is

mounted at the LE near x/c=0, the passive (plasma off) presence of the device can

function in this fashion. An example of this behavior is shown in Figure 7.2. The NASA

airfoil reaches CLmax at 12o with a sharp drop-off at higher α indicating stall. The OSU

airfoil with no LE passive actuator has a different pre-stall dCL/dα. The stall

characteristics of the OSU airfoil are also quite different with a gradual saturation of CL

and no indication of a sharp drop like the NASA case. The passive plasma actuator

curves are acquired by creating holes in the actuator materials such that CP and

subsequent CL measurements can be made. When these holes exist, the actuator cannot be

used for plasma generation because arcs form between the high voltage and ground

electrodes. The addition of the actuator to the LE rectifies the dCL/dα discrepancy and a

more uniform CL offset of the OSU model is now apparent. The additional curve in this

case shows the result of overshooting the target Re then gradually reducing the freestream

velocity back to the target value to take advantage of hysteresis effects often encountered

in separation control (Greenblatt and Wygnanski 2000). This overshoot allows extension

of the stall angle to more closely match NASA results by forcing the boundary layer at

the LE to undergo turbulent transition farther upstream. In doing so, the flow is now able
157
to negotiate the adverse pressure gradient encountered over the model surface. Note that

the post-stall characteristics of the OSU airfoil differ from NASA until α=16o where all

cases now agree. The majority of testing with actuators mounted near x/c=0 is performed

here.

Exposed HV electrode

Dielectric extent

Figure 7.1: DBD plasma actuator mounted at the LE of the OSU simplified NASA EET
airfoil.

158
Figure 7.2: CL curves for the OSU airfoil with and without passive LE actuator and Re
overshoot compared to NASA (Melton et al. 2006).

The stall characteristics of the OSU version of the NASA EET are very sensitive

to LE geometry due to a small radius of curvature in this region. Slight changes to the LE

boundary conditions caused by the application of actuators can change the stall

characteristics considerably. This is particularly frustrating when using DBD plasma

actuators composed of adhesive tape that must be applied to the airfoil surface by hand.

The actuator inevitably fails by arcing between the HV and ground electrodes requiring

replacement. Application of a new actuator to the LE can result in a different post-stall

CP distribution. These changes in the CP distribution can also occur within a given run

due to slight variations in the freestream velocity, turbulence intensity and LE boundary

159
conditions. Figure 7.3 gives an example of the various types of baseline behavior at post

stall with an actuator at the LE for Re=750k and α=16o. As in Chapter 6, baseline refers

to the condition when an actuator is applied to the LE without plasma (power off). Note

that static taps at x/c<0.05 are covered by the actuator and preclude measurements at

these points. In some cases, the post-stall baseline can exhibit a flat, zero pressure

gradient on the suction surface, while other cases can exhibit closed separation bubbles of

various size at various locations in the range x/c=0 to x/c=0.4. The time scale of this

changing baseline state can be as long as five minutes. The imprecise nature of applying

adhesive layers of tape to the leading edge by hand makes this a very time consuming

problem. In addition, hysteresis and Re effects are common, most notably in the α=10-12o

range.

160
Figure 7.3: Various examples of baseline (plasma off) behavior observed when an
actuator is applied to the LE.

The difficulty in reproducing baseline data has a strong effect on control

authority. When post-stall CP characteristics exhibit a near-zero pressure gradient on the

suction surface, NS-DBD plasma control authority is quite effective. However, when

closed separation bubbles of various size and location exist, the flow is much more

difficult to affect. This is due to variations of the separation characteristics near the LE.

The zero pressure gradient case indicates stall with separation near the LE which is near-

optimal for control with this actuator location. The nonzero pressure gradient on the

suction surface is created by post-stall closed separation bubbles and variable separation

location. This introduces a complicating factor in the analysis. To remove this variable,

the focus of analysis for actuators located near x/c=0 is on post-stall baselines that exhibit
161
a near zero pressure gradient along the suction surface. For completeness, the lower α

cases are discussed briefly.

At pre-stall α (10-12o), both AC and NS-DBD plasma actuation at 1 kHz can

substantially improve the CP distribution at Re=750k and 1000k as seen in Figures 7.4

and 7.5. The Re=750k cases also include CP profiles with (Passive Actuator) and without

(No Actuator) a passive actuator with holes for comparison (Figure 7.4a and 7.5a). The

No Actuator cases do not exhibit a zero pressure gradient on the suction surface. The

passive actuator alone can substantially improve the CP distribution in the α=12o case as

suction increases and the closed separation bubble near x/c=0.1 is weakened

considerably. In all cases, the DBD plasma actuator seems to function as an active trip

that is a substantial improvement over passive presence of the device alone. In these

cases, the flow remains attached after turning off both plasma discharges (not shown).

This indicates incipient separation at these nominally pre-stall angles of attack where a

well-designed boundary layer trip can and should result in attached flow for this airfoil

shape.

162
a)

b)

Figure 7.4: Effect of DBD plasma at x/c=0 on CP for Re=750k (a) and Re=1000k (b) at
α=10o.

163
a)

b)

Figure 7.5: Effect of DBD plasma at x/c=0 on CP for Re=750k (a) and Re=1000k (b) at
α=12o.

164
At α=14o in Figure 7.6, the NS-DBD plasma actuator consistently outperforms the

AC-DBD discharge by attaching the normally separated flow at both Re although some

isolated cases do not follow this trend. The CP curves in Figure 7.6 contain data from

various tests using different actuators applied to the same location on the airfoil LE.

Thus, there are repeated data series in some cases. This is done to highlight the variable

performance and complication that arises from applying the DBD very near the LE. For

example, Figure 7.6a shows two data series for NS-DBD forcing at 2 kHz which show

considerably different levels of control authority. Similarly, in Figure 7.6b, AC-DBD

forcing at 2 kHz has variable performance. This phenomenon will be discussed

throughout this section. However, the data in Figure 7.6 begin to show more reliable

performance for the NS-DBD plasma actuator.

165
a)

b)

Figure 7.6: Effect of DBD plasma at x/c=0 on CP for Re=750k (a) and Re=1000k (b) at
α=14o.

166
Figure 7.7 shows additional examples of the variable performance encountered

for both the NS and AC-DBD plasma at the airfoil LE at α=16o for Re=750k and 1000k.

As in Figure 7.6, the plots are created from multiple data sets with different actuators

placed at x/c=0. In the low Re case, both AC and NS-DBD plasma can reattach the

normally separated flow as evidenced by the suction peak created near the LE, but some

isolated cases do not follow this trend. In the Re=750k case, the 2 kHz NS-DBD performs

poorly during one test and the AC-DBD at 1 kHz performs well. At Re=1000k, the NS-

DBD at 1 kHz has some variable performance and little control authority is demonstrated

by the AC-DBD plasma. Excitation using 2 kHz AC-DBD plasma is not effective for

Re=1000k in any of the surveyed cases. This result is not presented in Figure 7.7. The

lack of control authority for AC-DBD plasma at higher Re and freestream velocity is

expected.

167
a)

b)

Figure 7.7: Effect of DBD plasma at x/c=0 on CP for Re=750k (a) and Re=1000k (b) at
α=16o.

168
The preceding results have been shown in the interest of full disclosure of the

variable performance associated with using both AC and NS-DBD plasma actuators near

x/c=0. Because of the sensitivity of the developing boundary layer to LE boundary

conditions, only results that have been repeated at least twice are used for analysis.

Figure 7.8 shows CP profiles developed on the airfoil for forcing at various

frequencies. In this case, higher frequencies appear to generate the greatest effect with

control authority progressively increasing up to 1 kHz where saturation ensues. This

result is consistent with measurements of NS-DBD applied to a NACA 0015 airfoil

(Roupassov et al. 2009), but is somewhat contrary to the majority of literature that shows

a preferred frequency should exist for controlling airfoil LE separation at post-stall

(Seifert and Tillman 2009).

169
Figure 7.8: Effect of NS-DBD plasma actuation at x/c=0 on CP for Re=750k at α=16o
showing preference for higher frequency forcing.

Figure 7.9 shows another frequency investigation of both the AC and NS-DBD

plasma at the airfoil LE. The frequency sweep is actually more detailed than shown, but

has been down-sampled for clarity. In this case, a repeatable baseline condition with zero

pressure gradient stall on the suction surface is established and detailed frequency sweeps

are performed using both AC and NS-DBD plasma by merely changing the excitation

input to the actuator. As such, this figure represents a direct comparison of the two types

of discharges for controlling flow separation. Low frequency AC signals are created

using a 2 kHz carrier frequency and modulated by the low frequency indicated in the

figure legend using 25% dc. Both the NS and AC DBD plasma produce the greatest

effect at frequencies in the range 600-800 Hz. Note that 2 kHz forcing does not produce a
170
significant change in the Cp distribution in these results. This is in contrast to Figure 7.9

and again highlights the sensitivity of the system to LE boundary conditions for actuators

placed at x/c=0.

171
a)

b)

Figure 7.9: Comparison of the effects of NS (a) and AC (b) DBD plasma on CP for
Re=750k.

172
Clearly, the NS-DBD outperforms the AC-DBD in this case as a significantly

stronger suction peak is visible Figure 7.9. The results of Figure 7.9 are more succinctly

summarized in Figure 7.10 where the full frequency data set is plotted. The ordinate is

created by plotting the change in CP value on the suction side closest to the LE

(x/c=0.05). The lack of CP resolution at the LE prevents the accurate calculation of CL,

but the minimum CP value should correlate quite well. Thus, these results can be

interpreted as a scaled measure of CL. The results are reported in dimensionless form

(Fc+) based on the airfoil chord (25.4 cm (10 in)). The results clearly show the NS-DBD

outperforms the AC discharge in this case and both have a preferred frequency for control

around Fc+=3-4. Also note that the performance of this particular actuator is not as

impressive as some previously shown as CP nearest the LE is 2.5 compared to what

would be near 3.0 if a similar metric is employed in Figure 7.8. Despite this, the NS-DBD

appears superior to the AC-DBD for actuators operating at similar energy and power

levels as the discharge power for the peak minimum CP values is created using between

10-12 W in each case corresponding to CE=~0.08%.

173
Figure 7.10: Effect of Fc+ of AC and NS-DBD plasma at x/c=0 on suction side CP at
x/c=0.05.

7.2 Single DBD Plasma Actuator Control Results: Upstream Orientation

The previous results have shown the capability and at times superiority of NS

over AC-DBD plasma for controlling LE separation with actuators mounted near x/c=0.

However, the variable CP response with respect to actuation frequency is troubling and

also exists in archived literature (Roupassov et al. 2009). To understand this behavior

more fully, a baseline repeatability study was performed. In the course of this work,

results showed that placement of a surface discontinuity due to tape layers near x/c=0

created variable baseline and controlled flow behavior. The sensitivity of the flow to this

discontinuity can be explained by considering the boundary layer developing on the

174
airfoil LE. The boundary layer begins to develop at the stagnation point which for airfoils

at high angle of attack is on the pressure side of the model indicated by CP=1. The

boundary layer always begins in a laminar state and over some distance transitions to

turbulent depending on Re, surface roughness, freestream turbulence and the pressure

gradient developed by flow over the model. When this developing boundary layer is in a

laminar or transition state, it is very sensitive to small changes such as those encountered

by the discontinuity created on both sides of the HV electrode. To make the LE boundary

layer less sensitive to these effects, various tripping attempts both upstream and on the

HV electrode were attempted. If the transition to turbulence could be accelerated, it is

suspected that more repeatable performance could be established. These attempts were

met with little success due to the small scale of the model. Instead, a study on HV

electrode location was employed based on the assumption that discontinuities created by

this element caused the majority of repeatability issues. After significant effort,

repeatable baseline and control results were established by moving the interface of the

HV and ground electrodes downstream a distance of 6 mm in arc length, ds, and

reorienting the discharge such that plasma is formed on the upstream side of the high

voltage electrode. To further remove additional surface discontinuities near x/c=0, wide

sections of Kapton tape (5 cm (2 in)) were used as the dielectric and wrapped further

around the model LE as shown in Figure 7.11. Note that (10 cm (4 in)) wide dielectrics

were also tested, but these performed similarly to the more narrow variety. The reverse

actuator orientation is used to remove discontinuities, but also to keep the actuator

interface region which is believed to be most important for separation control as near as

175
possible to the LE in an effort to remain near or slightly upstream of the separation line.

This is another advantage of the NS-DBD. It is postulated that, unlike momentum

producing devices, the direction of actuator layout is not important for control authority

using thermal effects. This reverse arrangement has been previously explored (Roupassov

et al. 2009). Note that in this arrangement, a direct comparison of NS and AC-DBD

plasma is not as appropriate since the AC induced flow counters the freestream velocity

and has not been as widely examined in the literature. However, limited testing (Figure

7.12) shows, to some degree, similar AC-DBD performance as in the normal arrangement

(Figure 7.9b).

Exposed HV electrode

Dielectric extent

Figure 7.11: Photograph of the DBD plasma actuator mounted near the airfoil LE in
reverse arrangement ds=6 mm from x/c=0.

176
Figure 7.12: Effect of AC and NS-DBD plasma for reverse actuator arrangement ds=6
mm from x/c=0.

With the newly developed baseline and control repeatability, detailed flow

diagnostics can now be employed. Minimum CP values developed on the model as a

function of Fc+ for various α are shown in Figure 7.13. At nominally pre-stall α (10-12o),

no frequency preference is visible and once a threshold level is reached, control authority

remains relatively constant. In the post-stall region beginning at α=14o, some optimal

frequencies become apparent around Fc+=4. As α is further increased, the preferred

frequency increases up to a maximum of around 5.5 at α =16o while the CP value tends to

increase. At α =18o, control authority is quite weak. From this information, it is apparent

that control authority at α>15o has been reduced by using the repeatable actuator

177
arrangement. This is an expected result since the actuator interface is now further

downstream (ds=6 mm). From this data, it appears this is near the optimal actuator

location for controlling flow over the airfoil at α < 14o. As α is progressively increased,

the separation line moves upstream and the actuator is now downstream of the optimal

location for producing control authority.

Figure 7.13: Effect of Fc+ of NS-DBD plasma for reverse actuator arrangement ds=6 mm
from x/c=0 on suction side CP at x/c=0.05 at Re=750k.

Despite this reduced control authority at high α, some significant flow physics are

revealed. Figure 7.14 shows a sample of the results of Figure 7.13, but now adds similar

curves for Re=1000k (62 m/sec). It is clear that control authority does not suffer due to
178
the increased freestream velocity and in the two lower α cases, it is actually augmented

considerably. This is even more promising when considering that q∞, which may be a

more reliable scaling parameter than velocity only, increases in proportion to U∞2. Note

also that Fc+ scaling holds quite well as the frequency preference for all three α is quite

similar. It is also clear that there is some effect on the flow for all Fc+ as the CP value has

been reduced by at least 0.5 in all cases. With this in mind, forcing in the various

frequency regions is explored using PIV.

Figure 7.14: Effect of Fc+ of NS-DBD plasma at for reverse actuator arrangement ds=6
mm from x/c=0 on suction side CP at x/c=0.05 at Re=750k and Re=1000k.

179
Figure 7.15a shows the NS-DBD plasma actuator effect on CP for α=12o. This

coincides with CL,max for the NASA airfoil and for the OSU model when a properly

designed and placed trip can effectively attach the flow (Figure 7.2). The forcing

frequency shown in Figure 7.15 corresponds to a condition where the threshold frequency

has been reached and further increases to do not result in significant CP improvement.

The baseline flow is characterized by a near zero pressure gradient along the suction

surface indicating separated flow. This is apparent in the time-averaged vorticity field in

Figure 7.15b as a strong region of vorticity signifies the boundary between the freestream

and separated region over the model. Control at Fc+=4 reattaches the normally separated

flow creating a strong suction peak at the LE. The time-averaged vorticity field of Figure

7.15c shows stronger levels of vorticity near the LE indicating a significant effect of the

actuator on the flow in this region. It can also be seen that the flow then separates from

the TE near x/c=0.8. Figure 7.16 shows phase-locked normalized velocity fluctuations,

v* = v / U ∞ , over the airfoil calculated according to:

v= V − V 7.1
where V and V denote phase and time-averaged quantities respectively.

The figure in question is one representative sample of four phases of the actuation

period. None of the phase-averaged images show any recognizable dynamic content.

Thus, reattachment of the nominally separated flow does not appear to occur through

excitation of large coherent structures. Instead, the flow has become reattached through

the use of the actuator as an active tripping device. This is supported by both NASA and

OSU results passive control results.


180
a)

b)

c)

*
Figure 7.15: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline
(b) and NS-DBD forcing at Fc+=4 (c) at Re=750k, α=12o.

181
Figure 7.16: Phase-averaged normalized velocity fluctuations, v* , for NS-DBD plasma
forcing at Fc+=4 (c) at Re=750k, α=12o.

Figure 7.17a shows the baseline and controlled CP distributions for α=14o. In this

case, the baseline is characterized by an even more zero pressure gradient stall along the

suction surface with control being quite effective for reattaching the separated flow along

nearly the entire suction surface. The time-averaged normalized vorticity field clearly

shows this effect as the region of strong vorticity bounding the freestream and separated

region is significantly more intense than the same figure for the α=12o case (Figure

7.15b). Control moves the region of strong vorticity near the LE further upstream. This

strong vorticity extends downstream to approximately x/c=0.2 after which a more diffuse

region is observed (Figure 7.17c). The controlled vorticity in this case differs

substantially from the α=12o case.

182
a)

b)

c)

*
Figure 7.17: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline
(b) and NS-DBD forcing at Fc+=4 (c) at Re=750k, α=14o.

183
The nature of flow structures in of Figure 7.18a gives the most striking

description of the controlled flow field. Clearly, coherent structures dominate the flow

and the number of these structures is consistent with forcing at Fc+=4. The effectiveness

of such dynamic structures for entraining freestream momentum and reattaching the

nominally separated flow of the airfoil is well established (Darabi and Wygnanski 2004).

Only one phase of the actuation signal is shown, but remaining phases confirm the

propagation of these organized regions along the airfoil chord. Further downstream, the

organized nature of these structures is broken. A pair of these positive and negative

regions corresponds to the outer edges of a clockwise rotating structure. However, from

Figure 7.18a one cannot discern which pair of these regions correspond to a vortex.

Normalized swirling strength, λci*=λcic/U∞, which separates regions of rotation

from pure shear is used to further identify this behavior. This technique is based on

critical point analysis of the local velocity gradient tensor and its eigenvalues (Adrian et

al. 2000). In two-dimensional form, the velocity gradient tensor is:

 ∂U ∂U
  
  ∂x ∂y 
∇W =
 ∂V ∂V 
7.2
 
 ∂x ∂y 


where in this case the gradient has been applied to the phase-averaged vector field, W .

The parameter of interest is the imaginary component of the eigenvalues of Equation 7.2

which is nonzero only if:

184
 ∂V 1   ∂U
 ∂V 1 ∂U
∂U   2  ∂V  2 
− +   +  <0
∂y ∂x 2 ∂x ∂y 4   ∂x   ∂y   7.3
 
The imaginary component of the eigenvalues of Equation 7.2, λci, is calculated using a 2nd

order accurate central difference scheme and the normalized results are shown in Figure

7.18b. It is now apparent that the regions on either side of x/c=0.15 form a coherent

structure. The same can be said for x/c=0.35 and 0.6. The more commonly used vorticity

field is shown in Figure 7.18c. By definition this term also includes components of pure

shear, which in this flow masks the location of vortex cores. Vorticity is well-suited for

identifying vortex shedding from the airfoil trailing edge since vorticity of different signs

is alternatively shed form the flap similar to the Von Karmen vortex sheet observed for

cylinders in cross flow (Figure 6.8). Clearly, it is not as useful for identifying coherent

structures over the airfoil as all vorticity is of the same sign.

185
a)

b)

c)

Figure 7.18: Phase-averaged normalized velocity fluctuations, v* (a), swirling strength,
 * , (c) for NS-DBD forcing at Fc+=4 for Re=750k, α=14o.
λ* , (b) and vorticity, Ω
ci

186
The frequency preference indicated in Figure 7.14 for α=16o shows three

relatively distinct forcing regimes. The flow response will be analyzed in order of

increasing frequency. Low frequency forcing (Fc+=0.6) has an effect on the naturally

stalled flow as seen in the CP curves, but in this case the minimum value is not observed

nearest LE (Figure 7.19a). Rather it appears further downstream near x/c=0.15. The time-

averaged vorticity certainly shows this change in behavior as the very apparent shear

layer is diffused considerably (Figure 7.19b,c). As before, the dynamic content of the

flow is most compelling (Figure 7.20). In this low frequency case, the phase averaged

velocity fluctuations are massively amplified as they move over the model surface. The

vortex cores (x/c=0.2 and 0.8) are again identified using swirling strength. Note that these

structures do not follow the airfoil surface as closely as those shown in Figure 7.18 and

instead are swept downstream by freestream momentum rather quickly once they move

away from the LE. Also, the magnitude of the phase averaged velocity fluctuations is

approximately twice as large are those seen in Figure 7.18a as noted by the change in

color scale. These massive coherent structures have been identified for all the low

frequency forcing cases in Figure 7.14. The vortex shedding associated with this low

frequency forcing has a rather strong audible signature that is quite obvious during wind

tunnel operation.

187
a)

b)

c)

*
Figure 7.19: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline
(b) and NS-DBD forcing at Fc+=0.6 (c) at Re=750k, α=16o.

188
a)

b)

Figure 7.20: Phase-averaged normalized velocity fluctuations, v* (a) and swirling
strength, λci* , (b) for NS-DBD forcing at Fc+=0.6 for Re=750k, α=16o.

Optimal forcing (Fc+=5.6) at α=16o produces reattachment at the LE and over

approximately half the model chord (Figure 7.21). The control authority in this case is

diminished compared to optimal forcing in the α=14o case (Figure 7.17a) as the suction

peak only reaches CP=-3. This is likely due to the movement of the separation line

upstream resulting in a less than optimized position for the plasma actuator which was
189
knowingly traded for repeatability. Forcing still has a significant effect on the time-

averaged vorticity field as the intense shear layer over the entire model is eliminated

(Figure 7.21b,c). As usual, notable effects of control can be seen in the dynamic features

of the separated flow (Figure 7.22). In this case, the coherent structures do not propagate

as far downstream as those at α=14o before dissipating. However, their organization near

the LE is quite apparent in both the phase averaged velocity fluctuations and the swirling

strength that shows a clear vortex core at x/c=0.15.

190
a)

b)

c)

*
Figure 7.21: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline
(b) and NS-DBD forcing at Fc+=5.6 (c) at Re=750k, α=16o.

191
a)

b)

Figure 7.22: Phase-averaged normalized velocity fluctuations, v* (a) and swirling
strength, λci* , (b) for NS-DBD forcing at Fc+=5.6 for Re=750k, α=16o.

Lastly the effect of high frequency forcing (Fc+=11.3) is investigated. Although

the LE CP value is similar to the low frequency forcing case (Figure 7.19), the location of

this minimum is further upstream signifying a very different flow field (Figure 7.23). As

in all cases, the time-averaged vorticity shows a significant effect, but merely breaking up

the intense layer of shear over the airfoil is not sufficient to substantially reduce
192
separation. The dynamic features are less apparent, but still exist near the LE (Figure

7.24). In this case, the color scale has been reduced considerably in an effort to bring out

the short wavelength low amplitude, but measureable phase averaged velocity

fluctuations generated by the high frequency forcing. This reduced color scale also

emphasizes the chaotic features of the dissipated region downstream. The robustness of

the swirling strength parameter is exhibited in Figure 7.24b. Despite the weak oscillation

observed in Figure 7.24a, as very distinct vortex core is identified at x/c=0.1.

193
a)

b)

c)

*
Figure 7.23: Time-averaged behavior of CP (a) and normalized vorticity, Ω , for baseline
(b) and NS-DBD forcing at Fc+=11.3 (c) at Re=750k, α=16o.

194
a)

b)

Figure 7.24: Phase-averaged normalized velocity fluctuations, v* (a) and swirling
strength, λci* , (b) for NS-DBD forcing at Fc+=11.3 for Re=750k, α=16o.

The preceding results for Re=750k have also been acquired at Re=1000k. As the

Fc+ in Figure 7.14 predicts, the flow fields are nearly identical in terms of control

authority, time-averaged and phase-averaged content. This is an encouraging result given

that this higher Re flow is in the upper range at which standard AC-DBD plasma

actuators lose control authority as documented in Figure 7.10. The efficiency at which the
195
NS-DBD plasma actuator excites coherent structures is particularly appealing. Although

optimal frequencies exist for control, it seems that the NS-DBD possesses sufficient high

amplitude characteristics to create structures at any frequencies in the range surveyed.

7.3 Comparison of Single NS-DBD Plasma Actuators to Passive Flow Control

The NS-DBD plasma controlled CL curve is shown in Figure 7.25 as a final

metric for demonstrating its capability. Recall that CP values are not measurable here due

to the actuator application. The CL values are estimated using the measured CP values and

linearly extrapolating the NS-DBD controlled data at the LE based on pre-stall values for

passive actuator with holes for CP measurements. This methodology is similar to other

simplified CL estimates (Janiszewska and Lee 2005), but also employs the measured data

away from the airfoil LE. The error bars on the figure are quite large due to the

uncertainty introduced by this extrapolation. As suggested by previous data, the NS-DBD

is capable of extending the stall angle to 16o and this effect is not diminished at

Re=1000k shown by the single data point. This figure is compared to a similar

experiment from NASA using a well-established PFC device in the form of a LE droop

(Figure 7.26) similar to the schematic in Figure 5.1. In terms of stall angle, the

performance rivals that of the NASA droop in these conditions. The actual CL values

between the two cases are considerably different as CL for NASA peaks near 2.0 for the

maximum droop angle (-30o) while the NS-DBD controlled CL is approximately 1.7.

These discrepancies can be explained. Baseline characterization showed that the OSU lift

curve is shifted by approximately -0.1 in comparison to NASA. Additionally, the

drastically different CP distribution created by the droop in the NASA case changes CL
196
considerably. For example, the NASA CL value at α=12o increases by approximately 0.2

when the maximum droop angle (-30o) is used. Thus, within the CL variation described

above the extrapolated values are comparable.

Figure 7.25: Effect of AFC with NS-DBD plasma actuation on CL for the OSU airfoil.

197
Figure 7.26: Effect of PFC with using LE droop on CL for the simplified NASA EET
airfoil (Melton et al. 2005)

198
Chapter 8: Summary and Conclusions

Control of flow separation from a high-lift airfoil has been examined

experimentally using dielectric barrier discharge (DBD) plasma actuators driven by high

voltage AC and nanosecond pulse (NS) waveforms. Actuators are composed of copper

tape electrodes and Kapton tape dielectric arranged in an asymmetric fashion. Actuators

are adhered to the leading edge (LE) and trailing edge (TE) flap shoulder of a simplified

supercritical high-lift airfoil with chord length of 25.4 cm in an effort to reduce or

eliminate flow separation from these regions. Surveyed flow conditions include Reynolds

numbers in the range 240k-1000k (15-62 m/sec) for incidence angles between 0-18

degrees and flap deflections between 0-40 degrees.

Actuator characterization in quiescent air shows that very different behavior is

generated when creating plasma discharges with these two radically different waveforms.

The AC-DBD plasma actuator generates near wall jets with maximum mean velocities of

~3.5 m/sec approximately 1 mm from the surface when exciting using a 3 kHz waveform

at 20 kVpp. Modulating this high frequency carrier signal with a low frequency signal

(50-400 Hz) results in fuller mean velocity profiles with lower maximum values. The

unsteady character of the modulated signals creates a near surface vortex train in still air

whose strength varies depending on the character of the modulating waveform. The

electrohydrodynamic effects of the AC-DBD plasma actuator are quantified by


199
integrating the near-wall profiles and casting the result as a dimensionless momentum

coefficient commonly used to describe flow control actuator amplitude. This metric is

found to be only weakly correlated with the control authority of the device when operated

with different modulation waveforms highlighting the limitations of quiescent air

measurements for quantifying actuator performance.

The NS-DBD plasma actuator induced mean velocity is approximately five times

lower than the AC-DBD. The low momentum production of the AC-DBD and its limited

control authority coupled with this result suggest that velocity production of the NS-DBD

plasma actuator is not a control mechanism for the flow conditions surveyed in this work.

schlieren imaging of the discharge shows that a very complex compression wave

structure is generated by the NS-DBD thermal effects in quiescent conditions. A single

planar and multiple localized spherical shock waves are generated for each high voltage

pulse. The speed of these waves a few mm from the discharge is approximately sonic

(Mach 1), but compression wave strength near the surface may be substantially greater.

Simplified estimates predict near surface temperature rises of ~200 K and compression

wave strength of approximately Mach 1.1. A modified momentum coefficient is

suggested for thermal effects based on pressure difference across the near surface

compression wave. This new coefficient produces values similar to those expected for

more traditional fluidic actuation techniques that demonstrate similar airfoil LE control

authority. High spatial resolution near surface measurements of this behavior are required

to fully characterize the device as a flow control actuator.

200
The time-averaged power dissipated by the two DBD plasma actuators is

relatively small. The AC-DBD plasma actuator consumes ~26 W when operated at 1 kHz

and 18 kVpp on a 46 cm long DBD load while the NS-DBD consumes <17 W operating

at the same frequency and 12 kV. The instantaneous power of the NS-DBD can be as

high as 600 kW due to the very short pulse duration (~100 ns) which deposits ~10 mJ

into the load generating the strong thermal effect seen in the schlieren photography.

Control of turbulent boundary layer separation over a deflected TE flap is

demonstrated using a single AC-DBD plasma actuator placed near the flap shoulder.

Quasi-steady forcing at 3 kHz slightly delays separation and increases static pressure on

the flap due to a lengthening and flattening of the separated region, but does not have a

drastic effect on the measured lift. Amplifying the natural instability associated with

vortex shedding from the flap shoulder by forcing with low frequency modulated

versions of the high frequency waveform is most effective for increasing lift and reducing

the time-averaged recirculation region on the flap. This comes at the expense of increased

dynamic loading and more organized turbulent fluctuations in the wake. Amplification of

this natural instability creates more negative static pressure values on the flap, but the

majority of the lift increase is due to upstream effects from enhanced circulation around

the entire model. These results persist even when significant reattachment downstream of

the flap shoulder is not achieved. The lift increase due to amplification of the natural

instability is relatively independent of incidence angle as long as the separation location

and the underlying dynamics do not change. Calculations of sectional total drag from

wake surveys that include turbulent stresses suggest lift increase via instability

201
amplification is not prohibitive. The control authority of the actuator generally decreases

with increasing Reynolds number and flap deflection consistent with the limited nature of

AC-DBD plasma momentum addition.

Both the effectiveness and efficiency of the actuator for forcing at low frequencies

can be further enhanced using a square wave modulation signal (BM) instead of a

sinusoidal signal (AM). This is due to the on/off nature of the BM perturbation in

comparison to the more gradual sinusoidal envelope associated with the AM case. Each

flap deflection angle is found to have a preferred duty cycle range for BM that increases

with increasing deflection angle. Beyond this preferred duty cycle, control authority is

found to decrease considerably. The falloff at high dc is believed to be due to the zero net

mass flux (ZNMF) pulsed blowing nature of the AC-DBD plasma actuator and is in

contrast to existing data for a similar airfoil using more traditional piezoelectric ZNMF

excitation (Melton et al. 2004). The NS-DBD plasma actuator has shown no control

authority at the TE flap shoulder.

The findings, aside from the dc and NS-DBD results, are generally consistent with

a similar study by NASA using piezoelectric driven ZNMF actuators on larger scale

version of the same airfoil (Melton et al. 2004; Melton et al. 2006). However, actuator

characterization shows that values of momentum coefficient for DBD plasma on the

airfoil are generally an order of magnitude lower than those used at NASA yet control

authority, especially at low Reynolds number, is comparable. Two open questions then

remain: do further increases in momentum coefficient generate a similar trend as shown

by NASA and is DBD plasma more efficient from a momentum perspective in this flow

202
field as momentum coefficient appears to show? Definitive answers require future work,

but such studies are certainly warranted given recent advances in DBD plasma thrust

generation that appear to make their use in transport aircraft feasible (Thomas et al.

2009).

Control of separation from the LE of the same airfoil is investigated primarily

with NS-DBD plasma actuators. The AC-DBD plasma actuator is well-established for

controlling LE separation up to 30 m/sec and is used as a metric for evaluating the NS-

DBD performance (Moreau 2007). The actuator is mounted across the span of the airfoil

to produce nominally two-dimensional perturbations. Locating the actuator at the LE near

x/c=0 facing downstream is very effective for controlling separation, however

repeatability of the both baseline and control conditions is difficult due to the sensitivity

of the flow field to small changes in the LE boundary conditions caused by the passive

presence of the device. Despite this complexity, repeated tests indicate the NS-DBD

outperforms the AC-DBD plasma actuator used in this work for controlling LE separation

and demonstrates similar scaling with dimensionless frequency while increasing the stall

angle by six degrees at Reynolds numbers up to 1000k (62 m/sec). This rivals the

performance of a passive LE droop used on a similar airfoil by NASA.

To improve repeatability, the actuator is moved downstream by 6 mm in arc

length along the suction surface and reversed such that plasma is formed in the upstream

direction using the assumption that control via thermal effects is not strongly dependent

on forcing direction. The slight movement of the surface discontinuity created by the tape

layers is sufficient to produce both repeatable baseline and control results. The control

203
authority at high incidence (α=16o) in this configuration is slightly reduced in comparison

to the x/c=0 case. This is likely due to a less optimal location of the actuator. Detailed

diagnostics in this configuration reveal the NS-DBD plasma actuator can act as an active

trip for nominally pre-stall incidence angles for this airfoil (≤1 2o). At high incidence, the

NS-DBD is found to reattach the normally separated flow through excitation of coherent

spanwise vortices that entrain freestream momentum into the separated region. The

device is found to generate these structures across the frequency range surveyed, but the

optimal dimensionless frequency for controlling separation is in the range four to six

depending on the incidence angle. The control authority in this configuration is

independent of the Reynolds numbers surveyed and obeys dimensionless frequency

scaling.

The contrasting performance of the NS-DBD plasma actuator at the LE and TE in

comparison to the AC-DBD is the most compelling question left unanswered in this

work. It is postulated that the near surface behavior of DBD plasma actuators coupled

with the varying state of the boundary layer at LE and TE is a possible cause for this

inconsistency. High resolution near surface DBD measurements that include the effects

of cross-flow are required to make a substantive conclusion on this topic.

204
Chapter 9: Discussion of Future Work

The previous sections have demonstrated some contrasting effects that deserve

discussion. The AC-DBD plasma actuator is found to be superior to NS-DBD plasma for

controlling separation from a deflected TE flap in the Re range 240k-750k (15-45 m/s). In

direct contrast, the NS-DBD plasma actuator is found to be superior to AC-DBD plasma

for controlling separation from the airfoil LE in the Re range 750k-1000k (45-62 m/s).

The exact cause of this discrepancy is unknown at this time, but some possibilities can be

considered.

Let us first examine the systems under consideration. The boundary layer

approaching the deflected flap in TE separation control studies has been characterized as

turbulent with thickness of 4.5-7 mm depending on Re. Separation control at the LE is a

substantially easier problem in terms of momentum requirements due to the state of the

boundary layer here (Melton et al. 2006). For airfoils at high α, the boundary layer

develops from the stagnation point on the pressure side around the suction surface of the

airfoil. At high α when flow separates abruptly from the LE, it is difficult to define the

state of this boundary layer (laminar/transitional/turbulent) especially in small scale

laboratory applications. However, one can assume that it is substantially thinner than the

boundary layer measured in the TE case. Despite the different boundary layer

characteristics at the LE and TE flap shoulder, literature shows that similar dimensionless
205
frequency scaling applies to both if the separation length scales are properly identified

(Seifert et al. 1996).

Actuator characterization showed two fundamentally different mechanisms are

involved when generating plasma with these two radically different waveforms. The NS-

DBD showed relatively weak compression waves propagating away from the surface. It

is suspected that the compression wave strength is substantially greater very near the

surface but quickly weakens as it propagates away. This is complicated by the highly

three dimensional nature of the compression waves generated by the device. To quantify

these effects, high spatial resolution near surface measurements must be performed.

However, if this assumption is correct it may explain the superior performance of the NS-

DBD for controlling LE airfoil separation where a very thin boundary layer is present. If

this is the case, it would also suggest that the AC-DBD plasma, while not as strong near

the surface, may produce a slightly stronger effect than the NS-DBD away from the

surface. A more fundamental question is whether there is any difference in instability

response to thermal versus EHD generated perturbations.

Another possibility for the discrepancy in control authority between these two

devices is the streamwise distribution of the forcing effect. For example, it is assumed

that the asymmetric electrode interface is the most important region for control authority

in both cases. This assumption is based on modeling results for the AC-DBD plasma

actuator (Corke et al. 2007), but similar studies have not been employed for the NS-DBD.

The effects of freestream flow on the discharge characteristics have also not been fully

206
investigated. Thus, it is possible that the optimized actuator location may not be

equivalent for the two discharges.

Lastly, an obvious missing piece to this work is the combination of NS-DBD

forcing on the LE and AC-DBD forcing on the TE to even more fully enhance lift. While

time constraints did not permit the experiments, both existing literature and TE results

suggest that if flow can be attached up to the flap shoulder using NS-DBD, the additional

CL lift gains from TE should be conserved (Melton et al. 2006).

207
References

Abbot I, vanDoenhoff A (1959) Theory of Wing Sections. New York, NY: Dover

Abe T, Takizawa Y, Sato S, Kimura N (2008) Experimental study for momentum transfer
in a dielectric barrier discharge plasma actuator. AIAA Journal 46[9]:2248-2256

Adelgren R, Yan H, Elliott G, Knight D, Zheltovodov A (2005) Control of Edney IV


Interaction by Pulsed Laser Energy Deposition. AIAA Journal 43[2]:256-269

Adrian R, Christensen K, Liu Z (2000) Analysis and interpretation of instantaneous


turbulent velocity fields. Experiments in Fluids 29:275-290

Amitay M, Glezer A (2002) Role of Actuation Frequency in Controlled Flow


Reattachment over a Stalled Airfoil. AIAA Journal 40[2]:209-216

Balcon N, Benard N, Lagmich Y, Boeuf J, Touchard G, Moreau E (2009) Postive and


Negative Sawtooth Signals Applied to a DBD Plasma Actuator - Influence on the
Electric Wind. Journal of Electrostatics 67:140-145

Benard N, Balcon N, Moreau E (2009a) Electric Wind Produced by a Surface Dielectric


Barrier Discharge Operating Over a Wide Range of Relative Humidity. AIAA
Paper 2009-0488

Benard N, Jolibois J, Forte M, Touchard G, Moreau E (2007) Control of an


Axisymmetric Subsonic Air Jet by Plasma Actuator. Experiments in Fluids
43[4]:603-616

Benard N, Jolibois J, Moreau E (2009b) Lift and Drag Performances of an Axisymmetric


Airfoil Controlled by Plasma Actuator. Journal of Electrostatics 67:133-139

208
Benedict L, Gould R (1996) Towards better uncertainty estimates for tublence statistics.
Experiments in Fluids 22:129-136

Boucinha V, Magnier P, Leroy-Chesneau A, Weber R, Joussot R, Dong B, Hong D


(2008) Characterization of the Ionic Wind Induced by a Sine DBD Actuator used
for Laminar-to-Turbulent Transition Delay. AIAA Paper 2008-4210

Cerchie D, Halfron E, Hammerich A, Han G, Taubert L, Trouve L, Varghese P,


Wygnanski I (2006) Some Circulation and Separation Control Experiments. In:
Applications of Circulation Control Technology. eds Joslin, R and Jones, G. Vol.
214, pp. 113-165, Reston, VA: AIAA Inc.

Chan S, Zhang X, Gabriel S (2007) Attenuation of Low-Speed Flow-Induced Cavity


Tones Using Plasma Actuators. AIAA Journal 45[7]:1525-1538

Corke T, Enloe C, Wilkinson S (2010) Dielectric Barrier Discharge Actuators for Flow
Control. Annual Review of Fluid Mechanics 42:505-529

Corke T, Post M, Orlov D (2007) SDBD Plasma Enhanced Aerodynamics: Concepts,


Optimization and Applications. Progress in Aerospace Sciences 43:193-217

Corke T, Post M, Orlov D (2009) Single Dielectric Barrier Discharge Plasma Enhanced
Aerodynamics: Physics, Modeling and Applications. Experiments in Fluids 46:1-
26

Darabi A, Wygnanski I (2004) Active Management of Naturally Separated Flow over a


Solid Surface. Part 1. The Forced Reattachment Process. Journal of Fluid
Mechanics 510:105-129

DeSalvo M, Whalen E, Glezer A (2010) High-Lift Enhancement using Fluidic Actuation.


AIAA Paper 2010-0863

ELD (2006) 24 inch High Speed Recirculating Wind Tunnel: Installation, Operation and
Maintenance Instructions. Lake City, MN

ELD (2007) 10 inch Chord Two-dimensional Airfoil Model with Adjustable Flap:
Installation, Operation and Maintenance Instructions. Lake City, MN
209
Enloe C, Font G, McLaughlin T, Orlov D (2008a) Surface Potential and Longitudinal
Electric Field Measurements in the Aerodynamic Plasma Actuator. AIAA Journal
46[11]:2730-2740

Enloe C, McHarg M, McLaughlin T (2008b) Time-Correlated Force Production


Measurements of the Dielectric Barrier Discharge Plasma Aerodynamic Actuator.
Journal of Applied Physics 103[073302]:1-7

Enloe C, McLaughlin T, Font G, Baughn J (2006) Parameterization of Temporal


Structure in the Single-Dielectric-Barrier Aerodynamic Plasma Actuator. AIAA
Journal 44[6]:1127-1136

Enloe C, McLaughlin T, VanDyken R, Kachner K, Jumper E, Corke T (2004a)


Mechanisms and Responses of a Single Dielectric Barrier Plasma Actuator:
Plasma Morphology. AIAA Journal 42[3]:589-594

Enloe C, McLaughlin T, VanDyken R, Kachner K, Jumper E, Corke T, Post M, Haddad


O (2004b) Mechanisms and Responses of a Single Dielectric Barrier Plasma
Actuator: Geometric Effects. AIAA Journal 42[3]:595-604

Falkenstein Z, Coogan J (1997) Microdischarge Behaviour in the Silent Discharge of


Nitrogen Oxygen and Water Air Mixtures. Journal of Physics D: Applied Physics
30:817-825

Font G, Enloe C, Newcomb J, Teague A, Vasso A, McLaughlin T (2010) Effects of


Oxygen Content on the Behavior of the Dielectric Barrier Discharge
Aerodynamic Plasma Actuator. AIAA Paper 2010-0545

Forte M, Jolibois J, Pons J, Moreau E, Touchard G, Cazalens M (2007) Optimization of a


Dielectric Barrier Discharge Actuator by Stationary and Non-Stationary
Measurements of the Induced Flow Velocity: Application to Airflow Control.
Experiments in Fluids 43:917-928

Freeman A, Catrakis H (2008) Direct Reduction of Aero-Optical Aberrations by Large


Structure Suppression Control in Turbulence. AIAA Journal 46[10]:2582-2590

210
Gad-el-Hak M, Bushnell D (1991) Separation Control: Review. Journal of Fluids
Engineering 113:5-29

Gaitonde D (2009) Simulation of Supersonic Nozzle Flows with Plasma-based Control.


AIAA Paper 2009-4187

Glezer A, Amitay M, Honohan A (2005) Aspects of Low- and High-Frequency Actuation


for Aerodynamic Flow Control. AIAA Journal 43[7]:1501-1511

Goksel B, Greenblatt D, Rechenberg I, Nayeri C, Paschereit C (2006) Steady and


Unsteady Plasma Wall Jets for Separation and Circulation Control. AIAA Paper
2006-3686

Golubovskii Y, Maiorov V, Behnke J, Behnke J (2002) Influence of Interaction between


Charged Particles and Dielectric Surface over a Homogeneous Barrier Discharge
in Nitrogen. Journal of Physcis D: Applied Physics 35:751-761

Greenblatt D (2007) Dual Location Separation Control on a Semispan Wing. AIAA


Journal 45[8]:1848-1860

Greenblatt D, Goksel B, Rechenberg I, Schule C, Romann D, Paschereit C (2008)


Dielectric Barrier Discharge Flow Control at Very Low Flight Reynolds
Numbers. AIAA Journal 46[6]:1528-1541

Greenblatt D, Wygnanski I (2000) The Control of Flow Separation by Periodic


Excitation. Progress in Aerospace Sciences 36:487-545

Greenblatt D, Wygnanski I (2003) Effect of Leading-Edge Curvature on Airfoil


Separation Control. Journal of Aircraft 40[3]:473-481

He C, Corke T, Patel M (2009) Plasma Flaps and Slats: An Application of Weakly


Ionized Plasma Actuators. Journal of Aircraft 46[3]:864-873

Hoener S, Borst H (1975) Fluid-Dynamic Lift. Brick Town, NJ: Hoerner Fluid Dynamics

211
Holmes P, Lumley J, Berkooz G (1996) Turbulence, Coherent Structures, Dynamical
Systems and Symmetry. Cambridge, UK: Cambridge University Press

Hoskinson A, Hershkowitz N, Ashpis D (2008) Force measurements of single and double


barrier DBD plasma actuators in quiescent air. Journal of Physics D: Applied
Physics 41[245209]:1-9

Huang J, Corke T, Thomas F (2006) Plasma Actuators for Separation Control of Low-
Pressure Turbine Blades. AIAA Journal 44[1]:51-57

Janiszewska J, Lee J (2005) A Simple Method for Determining Lift and Drag on a Wing.
AIAA Paper 2005-1061

Jayaraman B, Shyy W (2008) Modeling of Dielectric Barrier Discharge-Induced Fluid


Dynamics and Heat Transfer. Progress in Aerospace Sciences 44:139-191

Jolibois J, Forte M, Moreau E (2008) Application of an AC Barrier Discharge Actuator to


Control Airflow Separation above a NACA 0015 Airfoil: Optimization of the
Actuation Location Along the Chord. Journal of Electrostatics 66:496-503

Jukes T, Choi K, Johnson G, Scott S (2006) Characterization of Surface Plasma-Induced


Wall Flows Through Velocity and Temperature Measurements. AIAA Journal
44[4]:764-771

Kiedaisch J, Demanett B, Reinhard P, Nagib H (2007) Active Flow Control for High Lift
Airfoils: Dynamic Flap Actuation. AIAA Paper 2007-1120

Kiedaisch J, Nagib H, Demanett B (2006) Active Flow Control Applied to High-Lift


Airfoils Utilizing Simple Flaps. AIAA Paper 2006-2856

Kim W, Do H, Mungal G, Cappelli M (2007) On the Role of Oxygen in Dielectric


Barrier Discharge Actuation of Aerodynamic Flows. Applied Physics Letters
91[18]:181501

Kogelschatz U (2003) Dielectric-barrier Discharges: Their History, Discharge Physics,


and Industrial Applications. Plasma Chemistry and Plasma Processing 23[1]:1-46

212
Lavision (2007) FlowMaster: Product Manual.

Likhanskii A, Semak V, Schneider M, Opaits D, Miles R, Macheret S (2009) The Role of


the Photoionization in the Numerical Modeling of the DBD Plasma Actuator.
AIAA Paper 2009-0841

Lin J, Dominik C (1997) Parametric Investigation of a High-Lift Airfoil at High


Reynolds Numbers. Journal of Aircraft 34[4]:485-491

Lu B, Bragg M (2003) Airfoil Drag Measurement with Simulated Leading-Edge Ice


using the Wake Survey Method. AIAA Paper 2003-1094

Mabe J, Calkins F, Wesley B, Woszidlo R, Taubert L, Wygnanski I (2009) Single


Dielectric Barrier Discharge Plasma Actuators for Improved Airfoil Performance.
Journal of Aircraft 46[3]:847-855

McLean J, Crouch J, Stoner R, Sakurai S, Seidel G, Feifel W, Rush H (1999) Study of


the Application of Separation Control by Unsteady Excitation to Civil Transport
Aircraft. NASA CR 209338

Melton L (2006) Private Communication

Melton L, Schaeffler N, Lin J (2007) High-Lift System for a Supercritical Airfoil:


Simplified by Active Flow Control. AIAA Paper 2007-0707

Melton L, Schaeffler N, Yao C-S, Seifert A (2005) Active Control of Flow Separation
from Supercritical Airfoil Leading-Edge Flap Shoulder. Journal of Aircraft
42[5]:1142-1149

Melton L, Yao C-S, Seifert A (2003) Active Control of Separation from the Flap of a
Supercritical Airfoil. AIAA Paper 2003-4005

Melton L, Yao C-S, Seifert A (2004) Application of Excitation from Multiple Locations
on a Simplified High-Lift System. AIAA Paper 2004-2324

213
Melton L, Yao C-S, Seifert A (2006) Active Control of Separation from the Flap of a
Supercritical Airfoil. AIAA Journal 44[1]:34-41

Moreau E (2007) Airflow Control by Non-Thermal Plasma Actuators. Journal of Physics


D: Applied Physics 40:605-636

Nagib H, Kiedaisch J, Reinhard P, Demanett B (2006) Control Techniques for Flows


with Large Separated Regions: A New Look at Scaling Parameters. AIAA Paper
2006-2857

Nagib H, Kiedaisch J, Reinhard P, Demanett B (2007) Active Flow Control for High Lift
Airfoils: Separation versus Circulation Control. AIAA Paper 2007-1119

Opaits D, Neretti G, Likhanskii A, Zaidi S, Shneider M, Miles R, Macheret S (2007)


Experimental Investigation of DBD Plasma Actuators Driven by Repetitive High
Voltage Nanosecond Pulses with DC or Low-Frequency Sinusoidal Bias. AIAA
Paper 2007-4532

Opaits D, Neretti G, Zaidi S, Shneider M, Miles R (2008) DBD Plasma Actuators Driven
by a Combination of Low Frequency Bias Voltage and Nanosecond Pulses. AIAA
Paper 2008-1372

Opaits D, Shneider M, Miles R (2009a) Electrodynamic Effects in Nanosecond-Pulse-


Sustained Long Dielectric-Barrier-Discharge Plasma Actuators. Applied Physics
Letters 94[061503]:3

Opaits D, Zaidi S, Shneider M, Miles R (2009b) Improving Thrust by Suppressing


Charge Build-up in Pulsed DBD Plasma Actuators. AIAA Paper 2009-0487

Pack LG, Schaeffler NW, Yao. CS, Seifert A (2002) Active Control of Separation from
the Slat Shoulder of a Supercritical Airfoil. AIAA Paper 2002-3156

Palm P, Meyer R, Plonjes E, Rich J, Adamovich I (2003) Nonequilibrium Radio


Frequency Discharge Plasma Effect on Conical Shock Wave: M=2.5 Flow. AIAA
Journal 41[3]:465-469

214
Patel M, Ng T, Vasudevan S, Corke T, Post M, McLaughlin T, Suchomel C (2008)
Scaling Effects of an Aerodynamic Plasma Actuator. Journal of Aircraft
45[1]:223-236

Patel M, Sowle Z, Corke T, He C (2007) Autonomous Sensing and Control of Wing Stall
Using a Smart Plasma Slat. Journal of Aircraft 44[2]:516-527

Poggie J, Tillman C, Flick P, Silkey J, Osborne B, Ervin G, Maric D, Mangalam S,


Mangalam A (2010) Closed-Loop Stall Control on a Morphing Airfoil using Hot-
Film Sensors and DBD Actuators. AIAA Paper 2010-0547

Pons J, Moreau E, Touchard G (2005) Asymmetric surface dielectric barrier discharge in


air at atmospheric pressure: electrical properties and induced airflow
characteristics. Journal of Physics D: Applied Physics 38:3635-3642

Porter C, Abbas A, Cohen K, McLaughlin T, Enloe CL (2009) Spatially Distributed


Forcing and Jet Vectoring with a Plasma Actuator. AIAA Journal 47[6]:1368-
1378

Porter C, Baughn J, McLaughlin T, Enloe C, Font G (2007) Plasma Actuator Force


Measurements. AIAA Journal 45[7]:1562-1570

Post M (2004) Plasma Actuators for Separation Control on Stationary and Oscillating
Airfoils. Dissertation, University of Notre Dame

Post M, Corke T (2004) Separation Control on High Angle of Attack Airfoil Using
Plasma Actuators. AIAA Journal 42[11]:2177-2184

Rae W, Pope A (1984) Low Speed Wind Tunnel Testing. 2nd edn New York: John Wiley
& Sons, Inc.

Rizzetta D, Visbal M (2009) Large Eddy Simulation of Plasma-Based Control Strategies


for Bluff Body Flow. AIAA Journal 47[3]:717-729

Roth J, Dai X (2006) Optimization of the Aerodynamic Plasma Actuator as an


Electrohydrodynamic (EHD) Electrical Device. AIAA Paper 2006-1203

215
Roth J, Sherman D, Wilkinson S (2000) Electrohydrodynamic Flow Control with a
Glow-Discharge Surface Plasma. AIAA Journal 38[7]:1166-1172

Roupassov D, Nikipelov A, Nudnova M, Starikovskii A (2009) Flow Separation Control


by Plasma Actuator with Nanosecond Pulsed-Periodic Discharge. AIAA Journal
47[1]:168-185

Samimy M, Kim J-H, Kastner J, Adamovich I, Utkin Y (2007) Active Control of High-
Speed and High-Reynolds-Number Jets using Plasma Actuators. Journal of Fluid
Mechanics 578:305-330

Samimy M, Kim J, Kearney-Fischer M, Sinha A (2010) Acoustic and Flow Fields of an


Excited High Reynolds Number Axisymmetric Perfectly-Expanded Mach 1.3 Jet.
Journal of Fluid Mechanics accepted for publication

Seifert A, Darabi A, Wygnanski I (1996) Delay of Airfoil Stall by Periodic Excitation.


Journal of Aircraft 33[4]:691-698

Seifert A, Pack L (2002) Active Flow Separation Control on Wall-Mounted Hump at


High Reynolds Numbers. AIAA Journal 40[7]:1363-1372

Seifert A, Tillman C (2009) Fixed Wing Airfoil Applications. In: Fundamentals and
Applications of Modern Flow Control. eds Joslin, R and Miller, D. Vol. 231, pp.
231-257, Reston, VA: AIAA Inc.

Sidorenko A, Budovsky A, Pushkarev A, Maslov A (2008) Flight Testing of DBD


Plasma Separation Control System. AIAA Paper 2008-0373

Singh K, Roy S (2008) Force approximation for a plasma actuator operating in


atmospheric air. Journal of Applied Physics 103[013305]:1-6

Smith D, Dickey E, Klein TV (2006) The ADVINT Program. AIAA Paper 2006-2854

Sosa R, Artana G, Moreau E, Touchard G (2007) Stall control at high angle of attack
with plasma sheet actuators. Experiments in Fluids 42:143-167

216
Stanislas M, Okamoto K, Kähler C, Westerweel J, Scarano F (2008) Main results of the
third international PIV Challenge. Experiments in Fluids 45[1]:27-71

Sung Y, Kim W, Mungal M, Cappelli M (2006) Aerodynamic Modification of Flow over


Bluff Objects by Plasma Actuation. Experiments in Fluids Vol. 41:479-486

Takeuchi N, Yasuoka K, Ishii S (2007) Inducing Mechanisms of Electrohydrodynamic


Flow by Surface Barrier Discharge. IEEE Transactions on Plasma Science
35[6]:1704-1709

Thomas F, Corke T, Igbal M, Kozlov A, Schatzman D (2009) Optimization of Dielectric


Barrier Discharge Plasma Actuators for Active Aerodynamic Flow Control.
AIAA Journal 47[9]:2169-2178

Thomas F, Kozlov A, Corke T (2008) Plasma Actuators for Cylinder Flow Control and
Noise Reduction. AIAA Journal 46[9]:1921-1931

Utkin YG, Keshav S, Kim J-H, Kastner J, Adamovich IV, Samimy M (2007)
Development and Use of Localized Arc Filament Plasma Actuators for High-
Speed Flow Control. Journal of Physics D: Applied Physics 40[3]:685-694

Vorobiev A, Rennie RM, Jumper E, McLaughlin T (2008) Experimental Investigation of


Lift Enhancement and Roll Control using Plasma Actuators. Journal of Aircraft
45[4]:1315-1321

Wygnanski I (2004) The Variables Affecting the Control of Separation by Periodic


Excitation. AIAA Paper 2004-2505

Zaman K, Culley D (2008) Flow Separation Control over an Airfoil: Implication of Wake
Velocity-Deficit Reduction. AIAA Paper 2008-3768

217

You might also like