Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015...

12
Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal, David Castañeda, Nathan Webb, and Mo Samimy The Ohio State University, Columbus, Ohio 43235 DOI: 10.2514/1.J056071 Dynamic stall is observed in numerous applications, including sharply maneuvering fixed-wing aircraft, biomimetics, wind turbines, and most notably, rotorcraft. The associated unsteady loading can lead to aerodynamic flutter and mechanical failure in the system. The present work explores the ability of nanosecond pulse-driven dielectric barrier discharge plasma actuators to control dynamic stall over a NACA 0015 airfoil. The Reynolds number, reduced frequency, and excitation Strouhal number were varied over large ranges: Re 167;000500;000, k 0.0250.075, and St e 010, respectively. Surface pressure measurements were taken for each combination of Reynolds number, reduced frequency, and excitation Strouhal number. Phase- locked particle image velocimetry measurements were acquired for select cases. It was observed that the trends of effect of St e were similar for all combinations of Reynolds number and reduced frequency, and three major conclusions were drawn. First, it was observed that low Strouhal number excitation (St e < 0.5) results in oscillatory aerodynamic loading in the stalled stage of dynamic stall. This oscillatory behavior was gradually reduced as St e increased and was not observed beyond St e > 2. Second, all excitation resulted in earlier flow reattachment. Last, it was shown that excitation, especially at high St e , resulted in reduced aerodynamic hysteresis and dynamic stall vortex strength. The decrease in the strength of the dynamic stall vortex is achieved by the formation of large-scale structures induced by the excitation that bleed the leading-edge vorticity before the ejection of the dynamic stall vortex. At sufficiently high excitation Strouhal numbers (St e 10), the dynamic stall vortex was completely suppressed. Nomenclature C L = lift coefficient, −∫ 1 1 C p sinθ ds C M = moment coefficient, −∫ 1 1 C p × 14 sin θ x sin θ y cos θ ds C p = pressure coefficient, p p q c, L = airfoil chord, 203, mm f = frequency, Hz k = reduced frequency, πfcu p = static pressure, Pa p o = stagnation pressure, Pa p = freestream static pressure, Pa q = dynamic pressure, Pa, 1.05 × p o p , where 1.05 is the tunnel calibration constant Re = Reynolds number, u cν St e = excitation Strouhal number, fcu U = velocity vector, u; v, m/s U = nondimensional velocity vector, Uu u = freestream velocity, m/s x = streamwise position, mm y = vertical position, mm α = angle of attack, deg β = moment stall magnitude Δα = motion amplitude, deg θ = surface normal angle λ = swirling strength, Hz λ = nondimensional swirling strength, λcu Ξ = damping coefficient, 1πΔα 2 H C M dα ϕ = phase, deg I. Introduction T HE primary focus of this work is exploring flow physics and control of dynamic stall using nanosecond pulse-driven dielectric barrier discharge (NS-DBD) plasma actuators. Dynamic stall occurs in a variety of applications, including wind turbines and sharply maneuvering fixed-wing aircraft, and occurs in rotorcraft primarily due to the change in rotor angle of attack α required to maintain lift symmetry [1,2]. Dynamic stall occurs when the rate of change of α is fast enough to maintain attached flow past the static stall angle of attack. As the airfoil motion slows and begins to pitch down, a vortex forms at the leading edge, convected over the airfoil, and shed. This dynamic stall vortex momentarily increases the lift production. Once it has convected past the airfoil and shed, the flow is fully stalled. This process results in large oscillatory loads that can significantly reduce the life of the blades [3]. This problem has attracted much attention in the rotorcraft community because it is desirable to eliminate the oscillatory loads. Previous research has focused on the use of flow control devices to mitigate the dynamic stall vortex. These devices consist of both passive (such as geometric modifications to the airfoil) and active (such as momentum injection) techniques [48]. For practical application, it is pertinent to ensure such a device can be incorporated into the blade of a rotorcraft. Adding geometric modifications, such as vortex generators, would result in heavier rotor blades and may not easily be incorporated into an existing rotorcraft fleet. Furthermore, geometric modifications tend to work well at design conditions, but become ineffective or detrimental at off-design conditions. Active techniques aim to provide efficacy over a broad operating range. Active techniques typically involve momentum injection [4]. Although potentially effective, it is not a desirable solution due to the increased mechanical complexity, weight, and cost. Recently, dielectric barrier discharge (DBD) plasma actuators have been explored as a potential alternative to provide momentum injection. These devices consist of two copper electrodes separated by a dielectric barrier and use the electrohydrodynamic effect to generate a secondary flow near the surface to energize the boundary layer, Presented as Paper 2017-1687 at the 55th AIAA Aerospace Sciences Meeting, Grapevine, TX, 913 January 2017; received 14 February 2017; revision received 7 July 2017; accepted for publication 16 August 2017; published online 19 September 2017. Copyright © 2017 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved. All requests for copying and permission to reprint should be submitted to CCC at www. copyright.com; employ the ISSN 0001-1452 (print) or 1533-385X (online) to initiate your request. See also AIAA Rights and Permissions www.aiaa.org/ randp. *Graduate Student, Department of Mechanical and Aerospace Engineering. Student Member AIAA. Research Engineer, Aerospace Research Center. Member AIAA. John B. Nordholt Professor of Mechanical and Aerospace Engineering, Director of Aerospace Research Center; [email protected]. Fellow AIAA (Corresponding Author). 78 AIAA JOURNAL Vol. 56, No. 1, January 2018 Downloaded by OHIO STATE UNIVERSITY LIBRARIES on March 9, 2018 | http://arc.aiaa.org | DOI: 10.2514/1.J056071

Transcript of Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015...

Page 1: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

Control of Dynamic Stall over a NACA 0015Airfoil Using Plasma Actuators

Achal Singhal,∗ David Castañeda,∗ Nathan Webb,† and Mo Samimy‡

The Ohio State University, Columbus, Ohio 43235

DOI: 10.2514/1.J056071

Dynamic stall is observed in numerous applications, including sharply maneuvering fixed-wing aircraft,

biomimetics, wind turbines, and most notably, rotorcraft. The associated unsteady loading can lead to

aerodynamic flutter and mechanical failure in the system. The present work explores the ability of nanosecond

pulse-driven dielectric barrier discharge plasma actuators to control dynamic stall over a NACA 0015 airfoil.

The Reynolds number, reduced frequency, and excitation Strouhal number were varied over large ranges:

Re � 167;000–500;000, k � 0.025–0.075, and Ste � 0–10, respectively. Surface pressure measurements were

taken for each combination of Reynolds number, reduced frequency, and excitation Strouhal number. Phase-

locked particle image velocimetry measurements were acquired for select cases. It was observed that the trends of

effect of Ste were similar for all combinations of Reynolds number and reduced frequency, and three major

conclusions were drawn. First, it was observed that low Strouhal number excitation (Ste < 0.5) results in

oscillatory aerodynamic loading in the stalled stage of dynamic stall. This oscillatory behavior was gradually

reduced as Ste increased and was not observed beyond Ste > 2. Second, all excitation resulted in earlier flow

reattachment. Last, it was shown that excitation, especially at high Ste, resulted in reduced aerodynamic

hysteresis and dynamic stall vortex strength. The decrease in the strength of the dynamic stall vortex is achieved by

the formation of large-scale structures induced by the excitation that bleed the leading-edge vorticity before the

ejection of the dynamic stall vortex. At sufficiently high excitation Strouhal numbers (Ste ≈ 10), the dynamic stall

vortex was completely suppressed.

Nomenclature

CL = lift coefficient, −∫ 1−1Cp sin�θ� ds

CM = moment coefficient, −∫ 1−1Cp × ��1∕4� sin θ − x sin θ −

y cos θ� dsCp = pressure coefficient, �p − p∞�∕q∞c,L = airfoil chord, 203, mmf = frequency, Hzk = reduced frequency, πfc∕u∞p = static pressure, Papo = stagnation pressure, Pap∞ = freestream static pressure, Paq = dynamic pressure, Pa, 1.05 × �po − p∞�, where 1.05 is

the tunnel calibration constantRe = Reynolds number, u∞c∕νSte = excitation Strouhal number, fc∕u∞U = velocity vector, �u; v�, m/sU� = nondimensional velocity vector, U∕u∞u∞ = freestream velocity, m/sx = streamwise position, mmy = vertical position, mmα = angle of attack, degβ = moment stall magnitudeΔα = motion amplitude, degθ = surface normal angleλ = swirling strength, Hz

λ� = nondimensional swirling strength, λc∕u∞Ξ = damping coefficient, −�1∕πΔα2� H CM dαϕ = phase, deg

I. Introduction

T HE primary focus of this work is exploring flow physics andcontrol of dynamic stall using nanosecond pulse-driven

dielectric barrier discharge (NS-DBD) plasma actuators. Dynamicstall occurs in a variety of applications, including wind turbines andsharply maneuvering fixed-wing aircraft, and occurs in rotorcraftprimarily due to the change in rotor angle of attack α required tomaintain lift symmetry [1,2]. Dynamic stall occurs when the rate ofchange of α is fast enough to maintain attached flow past the staticstall angle of attack. As the airfoil motion slows and begins to pitchdown, a vortex forms at the leading edge, convected over the airfoil,and shed. This dynamic stall vortex momentarily increases the liftproduction. Once it has convected past the airfoil and shed, the flow isfully stalled. This process results in large oscillatory loads that cansignificantly reduce the life of the blades [3]. This problem hasattracted much attention in the rotorcraft community because it isdesirable to eliminate the oscillatory loads.Previous research has focused on the use of flow control devices to

mitigate the dynamic stall vortex. These devices consist of bothpassive (such as geometric modifications to the airfoil) and active(such as momentum injection) techniques [4–8]. For practicalapplication, it is pertinent to ensure such a device can be incorporatedinto the blade of a rotorcraft. Adding geometric modifications, suchas vortex generators, would result in heavier rotor blades andmay noteasily be incorporated into an existing rotorcraft fleet. Furthermore,geometric modifications tend to work well at design conditions, butbecome ineffective or detrimental at off-design conditions. Activetechniques aim to provide efficacy over a broad operating range.Active techniques typically involve momentum injection [4].Although potentially effective, it is not a desirable solution due to theincreased mechanical complexity, weight, and cost. Recently,dielectric barrier discharge (DBD) plasma actuators have beenexplored as a potential alternative to provide momentum injection.These devices consist of two copper electrodes separated by adielectric barrier and use the electrohydrodynamic effect to generate asecondary flow near the surface to energize the boundary layer,

Presented as Paper 2017-1687 at the 55th AIAA Aerospace SciencesMeeting, Grapevine, TX, 9–13 January 2017; received 14 February 2017;revision received 7 July 2017; accepted for publication 16 August 2017;published online 19 September 2017. Copyright © 2017 by the AmericanInstitute ofAeronautics andAstronautics, Inc.All rights reserved.All requestsfor copying and permission to reprint should be submitted to CCC at www.copyright.com; employ the ISSN 0001-1452 (print) or 1533-385X (online) toinitiate your request. See also AIAA Rights and Permissions www.aiaa.org/randp.

*Graduate Student, Department of Mechanical and Aerospace Engineering.Student Member AIAA.

†Research Engineer, Aerospace Research Center. Member AIAA.‡John B. Nordholt Professor of Mechanical and Aerospace Engineering,

Director of Aerospace Research Center; [email protected]. Fellow AIAA(Corresponding Author).

78

AIAA JOURNALVol. 56, No. 1, January 2018

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 2: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

thereby maintaining minimally complex construction. These devicesare generally referred to as AC-DBD actuators, because they operateusing an alternating current.The primary disadvantage of momentum injection techniques is

that, at high speeds,moremomentum is required tomaintain efficacy.Meeting this increasing demand for high Reynolds numberapplication flows is often impractical. The focus of this paper will beon the use of NS-DBD plasma actuators. These actuators areconstructed identically to AC-DBD plasma actuators. However,instead of an AC signal, high-voltage, short-duration (approximately100 nanoseconds) DC pulses are used to create localized rapidheating, which results in the formation of a compression wave at theDC pulse frequency [9–13]. If introduced in the proper locationand at the proper frequency, the localized heating can be used toexcite natural flow instabilities, which then grow and roll up intolarge flow structures. This is the flow control mechanism for theNS-DBD plasma actuators, and because it is based on excitation ofinstabilities, relatively low power input is required and theactuation remains effective at high speeds. These actuators excitethe flow instabilities in a manner similar to another class of plasmaactuators called localized arc filament plasma actuators, whichhave been used for excitation of instabilities in several flows withinstabilities [14,15].NS-DBD plasma actuators have been shown to reattach flow in

stalled airfoils for a variety of airfoil shapes and flow velocities in astatic configuration [9–12]. The static stall control work includesflow control of an airfoil in fully reversed flow [11]. Thiswork aims toextend the existing body of work on NS-DBD flow control tounsteady flows.

II. Experimental Arrangement

Experiments were performed in the recirculating wind tunnellocated at the Gas Dynamics and Turbulence Laboratory, withinthe Aerospace Research Center at The Ohio State University [9–12].The tunnel has an optically clear acrylic test section measuring61 × 61 cm in cross section and 122 cm in length. A NACA 0015airfoil with a 203 mm chord was used in the experimentation.Although not a typical cross section for rotorcraft blades, it has beenwell characterized in static flow, and for this first effort it was deemedappropriate. The experimental arrangement used in this work isbriefly recalled here. More details are found in Singhal [13].Two coordinate systems are used throughout this paper. The first

coordinate system has its origin at the aerodynamic leading edge, asshown in Fig. 1. The axis of the first system lies along the chord lineand the coordinate is normalized by the chord length, denoted x∕c.A positive coordinate indicates the aerodynamic suction side and anegative coordinate indicates the aerodynamic pressure side. Thissystem is in the airfoil reference frame and is used to specify thelocation of onboard instrumentation, actuators, and the flowseparation line. The second coordinate system is a two-dimensionalgrid aligned with the test section. These coordinates are alsonormalized by the chord, denoted x∕L and y∕L (L, instead of c, isused to distinguish between the two normalized x coordinates). Thissystem is in the test section reference frame and is used to define thelocation and direction of velocity data. The origin of this coordinatesystem is at the point of rotation, or x∕c � 0.25. A moment M

is defined to be positive when it results in the pitch-down motion of

the nose, as defined in Fig. 1.The airfoil oscillation mechanism was designed to maximize

optical access andmodeled after that ofGreenblatt andWygnanski [16].

It consists of two acrylic disks secured in aluminum rings. The ends ofthe airfoil are mounted to these disks. Thin ball bearings are used to

mount the airfoil support rings to the tunnel sidewalls. The aluminumrings are driven by a servo via timing belts, as shown in Fig. 2.The motion of the airfoil a�t� in the remainder of the paper is

described by three parameters: themean angle αm, amplitudeΔα, andreduced frequency (k � πfc∕u∞, where f is the physical oscilla-

tion frequency of the airfoil). The angle of attack is defined as

a�t� � am − Δa ⋅ cos�2ku∞ t∕L�, where t is the time in seconds.During experimentation, αm � 10 deg and Δα � 10 deg.A single actuator, powered by a custom in-house manufactured

pulse generator, was placed on the airfoil to explore the efficacyof excitation. The actuator (shown in Fig. 3) is constructed of two

0.09-mm-thick copper tape electrodes; the exposed high-voltage(HV) electrode is 6.35 mmwide and the covered ground electrode is

12.70 mm wide. The dielectric layer is composed of three layers of

Kapton tape, each 0.09 mm thick with a dielectric strength of 10 kV.The total thickness of the entire actuator is 0.45mm.The actuatorwas

placed on the suction side of the airfoil with the electrode junctionat x∕c � 0.01 (see Fig. 1) and covers the entire span of the airfoil.

As previously mentioned, the actuators are driven by a nanosecond

pulse to produce perturbations. This placement was motivated by thelocation of the receptivity region of the shear layer instabilities, which

is the location of maximum sensitivity to perturbations.Static pressure measurements on the airfoil surface were acquired

using three Scanivalve digital pressure sensor arrays (DSA-3217).

A total of 35 taps are located on the surface of the airfoil. The pressurecoefficient was phase averaged over 4 sets of 16,384 samples

acquired at 400 Hz near the airfoil centerline. The sectional lift and

moment coefficients CL andCM were calculated from these data andused to determine the effect of excitation on the dynamic stall

process. Figures containing pressure data, or results calculated frompressure data, include error bars, unless their presence would make

the figure difficult to read. These bars are measures of statistical

uncertainty and represent the 95% confidence interval.Particle image velocimetry (PIV) was the primary diagnostic

technique. Two-component planar PIV data were acquired for both

the flow surrounding the suction side of the airfoil and behind thetrailing edge. Olive oil seed particles were injected upstream of the

test section. An approximately 2-mm-thick laser sheet illuminatedthe measurement plane at 60.4% of the airfoil span. Two com-

mercially available LaVision cameras (Imager Pros) and software

(DaVis) were used to acquire 5 sets of 100 images pairs for eachexcited case. For the baseline cases, a single set of 500 images was

acquired. Acquisition was locked to the airfoil motion to providephase-averaged flowfield information. Swirling strength was used as

the vortex identification technique, as detailed by Adrian et al. [17],

and was nondimensionalized using the freestream velocity and thechord length. In the presentation of the PIV results, the airfoil is

plotted in black along with an arc. The arc is a manifestation of anoptical obtrusion and is not indicative of any flow features.PIV acquisition was phase locked to the motion of the airfoil

and is described by the phase ϕ. It is easier to consider the camerasto be acquiring snapshots of the flowfield at the same rate as

the oscillation frequency of the airfoil. Then, the equation of the

airfoil motion (relative to the cameras) becomes a�t� � am−Δα ⋅ cos�2ku∞t∕L� ϕ�. Ideally, phase and angle of attack are

injective; however, due to small errors in the motion, this may not bethe case. As such, the phase will always be provided in discussions

involving PIV results. For convenience, the angle of attack will also

be provided.It is also important to note that, for a given excitation Strouhal

number, any given excitation pulse will always occur at the same

phase for any cycle of the airfoil motion. As such, in the excited PIVresults, PIVacquisition is phase locked to both the airfoil motion and

the actuation.Fig. 1 Schematic of experimental coordinate systems showing theorigins.

SINGHAL ETAL. 79

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 3: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

III. Error Analysis

A detailed analysis of the error is provided by Singhal [13].

For brevity, the main conclusions of the error analysis in terms of

the motion profile, pressure measurements, and PIVare provided in

this section. In particular, the error will be described by 95%

confidence intervals (abbreviated as “ci” or “CI” in the figures) for

Re � 300;000 and k � 0.05.For a majority of the data presented herein, the desired motion is

described by k � 0.05, am � 10 deg, andΔα � 10 deg. The actualmotion is described by k � 0.05� 0.00, αm � 10.20� 0.05 deg,and Δα � 10.03� 0.02 deg. As such, the cycle-to-cycle repeat-

ability was approximately 0.07 deg.Pressure measurements used pressure taps connected to a pressure

transducer via tubing. Empirical correlationswere used to account for

the gain and phase delay introduced by the tubing length. The effect

of these empirical correlations, as well as the repeatability of the

motion profile, can be characterized by the confidence interval,

which is shown in Fig. 4, for each excitation Strouhal number. In this

figure, the mean confidence interval is normalized by the baseline

(Ste � 0) lift peak or themagnitude of themoment stall. As indicated

in the figure, the confidence intervals are nearly three orders of

magnitude lower than the lift peak or the magnitude of the moment

stall, which indicate high repeatability.When analyzing the phase-locked PIV data, it is important to

understand the uncertainty of the data. Shown in Fig. 5 is the root

mean square of a particular phase where the flow is completely

separated. In this figure, it is observed that the root mean square is

on the order of the freestream velocity in the separation region.

Three distinct processes could cause this: velocity fluctuations

in the separation region, repeatability error in the airfoil motion,

and error inherent in the PIV process. The latter two can be

characterized by the 95% confidence interval of the data, which is

shown in Fig. 6. The confidence interval is an order of magnitude

smaller than the root mean square (Fig. 5), indicating that the

majority of the fluctuations are indeed flow related and not an

artifact of the experimental setup.

Fig. 3 NS-DBD plasma actuator schematic. Vertical dimension exaggerated for clarity.

Fig. 2 Photographs of the oscillating mechanism.

Fig. 5 Root mean square of the streamwise velocity at Re � 300;000and k � 0.05.

Fig. 4 Normalized confidence interval forRe � 300;000 and k � 0.05.

80 SINGHAL ETAL.

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 4: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

IV. Results and Discussion

Before examining the effect of excitation, the baseline flow results

were documented. Although the actuators are thin (0.0022c), thephysical presence of the actuator near the leading edge does impact

the flow. Shown in Fig. 7 are lift andmoment curves for two differentcases at Re � 167;000. The red curves indicate that the actuator wasmounted on the airfoil (but not active). The blue curves are the cases

without an actuator present. In this figure, the motion is described by

αm � 10 deg, Δα � 10 deg, and k � 0.050. As demonstrated by

the plots, the lift peak andmoment stall due to the dynamic stall vortex

are slightly delayed. Furthermore, themoment stall isweakened by the

presence of the actuator. Thus, excited datawill be comparedwith data

acquired with the actuator mounted to the airfoil but turned off

(referred to as baseline or Ste � 0) throughout this paper.Flows were investigated for various Reynolds numbers (Re �

167;000; 300;000; 500,000) and reduced frequencies (k � 0.025;0.05; 0.075). Twenty excitation Strouhal numbers were tested for

each combination of Reynolds number and reduced frequency. The

only motion profile tested was a sinusoidal waveform with α m �10 deg and Δα � 10 deg. These parameters place the motion

within the regime of deep stall. The results presented herein are at

Re � 300;000 and k � 0.05. The trends for each combination of

Reynolds number and reduced frequency are similar with respect to

the excitation Strouhal number, and no significant information is lostby the omission of the other combinations of Reynolds number and

reduced frequency.

PIV results were acquired at one combination of Reynolds numberand reduced frequency. This case was Re � 300;000 and k � 0.05.PIV data were obtained for only two excitation Strouhal numberscases (Ste � 0.3 and 9.9), in addition to the baseline case. Theseconditionswere chosen as representative cases after an analysis of thepressure data. Phase-locked (to the airfoil motion) PIV data werecollected at 17 phases for the baseline case and 9 phases for theexcited cases.

A. Baseline Results

Figure 8 shows the baseline phase-averaged lift and momentcoefficients. In this figure, the five stages of dynamic stall, as outlinedby Corke and Thomas [4], are numbered. The process of dynamicstall begins as the airfoil starts pitching up. In this stage, the flow isattached and the lift increases steadily with α. When α exceeds thestatic stall α, the lift continues to increase; this is the second stage ofdynamic stall. In literature, the delayed separation is attributed to theairfoil motion and/or the formation of a closed separation bubble nearthe leading edge of the airfoil [4].In the second stage, the vorticity at the leading edge accumulates

and the dynamic stall vortex begins to form. This accumulation ofvorticity is better visualized by the PIV results, shown in Fig. 9. In thisfigure, the phase of motion ϕ and angle of attack are shown in thebottom left of each image.As α increases, the airfoil undergoes the third stage of dynamic

stall [4]: the ejection of this vorticity in the form of the dynamic stallvortex. The ejection of this vortex is attributed to vortex-inducedseparation [18]. As observed by Mulleners and Raffel [18], thedynamic stall vortex formation is accompanied by counter-rotatingvortices on the airfoil’s surface. These vortices travel upstream,pushed by the dynamic stall vortex, resulting in the detachment of thedynamic stall vortex. This phenomenon is termed vortex-inducedseparation [19]. Vorticity data clearly indicate the presence ofcounter-rotating vorticity near the surface (Fig. 10). Swirling strengthresults, shown in Fig. 11, also show the two sets of vortices. Figure 11also shows the dynamic stall vortex convecting over the airfoil atϕ � 73.5 deg (α � 19.6 deg). However, the vortex does not appearto be well defined, due to cycle-to-cycle variation in the phase of thevortex.The shape of the dynamic stall vortex in the PIV results can be

explained by the results in Fig. 12. The figure illustrates thesuction side pressure coefficient as it varies chordwise (horizontalaxis) and in time (vertical axis). Vortices produce pressure spikesthat move along the chord and can be identified [20] by red streaksin the surface pressure. The dynamic stall vortex is shown by thegreen dashed line on the left plot. On the right plot, the phase

Fig. 6 Confidence interval of the streamwise velocity at Re � 300;000and k � 0.05.

Fig. 7 Lift and moment curves at Re � 167;000 and k � 0.05 without an actuator (blue) and with a plasma actuator mounted on the airfoil but notturned on (red). The darker and lighter colors represent pitch-up and pitch-down motion, respectively.

SINGHAL ETAL. 81

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 5: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

ϕ � 73.5 deg (α � 19.6 deg) is shown by the blue dashed line,

which does not exactly follow the red streak. This supports the

conclusion that the dynamic stall vortex is convecting over the

airfoil as it is pitching. However, the width of the red streak, not

being narrow and sharp, indicates that the phase at which the

dynamic stall vortex convects is not consistent from cycle-to-

cycle. This is attributed to the stochastic nature of dynamic stall

process.The convection of the dynamic stall vortex results in an additional

increase in lift, as shown by the lift peak near α � 19 deg in Fig. 8.The vortex shedding also results in a sharp decrease in pitching

moment, because the vortex moves the airfoil center of pressure

downstream as it convects [21]. It is known that moment stall

precedes lift stall [4]; however, due to the limited temporal and spatial

resolution of the pressure taps, they appear to occur nearly

simultaneously. After the dynamic stall vortex has convected, theairfoil is fully stalled. This marks the fourth stage of the dynamic stall

process [4] and corresponds to the sharp drop in the lift observed in

Fig. 8. The final stage of dynamic stall is flow reattachment. The flowreattachment is a stochastic process, in which random features of the

shear layer result in large cycle-to-cycle variation in aerodynamicloads during recovery [4,22,23]. The time of reattachment can be

determined from the pressure data.

B. Excited Results

After characterizing the baseline flow, a sweep of excitation

Strouhal numberswas used to perform an assessment of theNS-DBDactuator’s potential as a dynamic stall flow control device. Figure 13

displays the phase-averaged lift coefficient for the baseline and

Fig. 9 Close up of baseline phase-averaged vorticity maps at Re � 300;000 and k � 0.050.

Fig. 10 Baseline phase-averaged normalized vorticity at Re � 300;000 and k � 0.050.

Fig. 8 Baseline phase-averaged lift and moment coefficients at Re � 300;000 and k � 0.050.

82 SINGHAL ETAL.

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 6: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

excitation results for Re � 300;000 and k � 0.05. Although a total

of 20 different excitation Strouhal numbers were tested only four

(Ste � 0; 0.3; 0.96, and 9.9), which are representative, will be shownto keep the figures uncluttered.The trend of the effects of excitation on the overall progression of

dynamic stall appears to be similar for different combinations of

Reynolds number and reduced frequencies. Each combination

follows three primary trends:1) Low excitation Strouhal numbers (Ste < 0.5) result in

oscillatory lift and drag following the dynamic stall vortex shedding.This behavior smooths out as the excitation Strouhal numberincreases.2) All excited cases result in earlier flow reattachment than the

baseline.3) Excited cases (in general) result in reduced lift and moment

hysteresis and decreased dynamic stall vortex strength.These trends will be further discussed in the next three sections.

1. Effects of Excitation in the Stalled Regime

The first trend, that excitation introduces oscillatory behavior intothe stalled regime of dynamic stall, is indicative of the formation oflarge coherent flow structures in the shear layer over the separatedregion. As shown by previous work with the actuators [9,11,12], lowStrouhal number excitation leads to the formation of coherentstructures at poststall α. If the excitation pulse timing is plotted withthe moment and lift curves for the low excitation Strouhal numbercase, it is easy to observe that every pulse results in the formation of astructure (Fig. 14). The static stall research previously conducted inour laboratory has demonstrated that the flow is receptive toexcitation over a range of Strouhal numbers [13]. This is alsoobserved in the unsteady flow results, particularly at low excitationStrouhal numbers,where the large structureswere also detected usingPIV. The swirling strength for the stalled phases at a low (left) andhigh (right) Strouhal number excitation is shown in Fig. 15. At highexcitation Strouhal numbers, this oscillatory behavior is not

Fig. 11 Baseline phase-averaged normalized swirling strength at Re � 300;000 and k � 0.050.

Fig. 12 Baseline suction side pressure coefficient atRe � 300;000 and k � 0.050, showing the (left) dynamic stall vortex convection time and (right) PIVphase acquisition time (right).

SINGHAL ETAL. 83

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 7: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

observed. This is expected given that higher excitation Strouhalnumbers result in the formation of smaller structures. Thesestructures also form further upstream on the airfoil and quicklydevelop, disintegrate, and dissipate [9,11,12]. This too is wellcaptured by PIV data, shown in Fig. 15. In the low excitation Strouhal

number case, one can observe a large structure near 40% chord overthe airfoil. In the high excitation Strouhal number case, there areseveral smaller structures convecting over the airfoil. It also bearsmentioning that, due to the limited spatial and temporal resolution ofpressure measurements, high-frequency events would be filtered out.

Fig. 13 Phase-averaged lift and moment coefficients for various excitation Strouhal numbers at Re � 300;000 and k � 0.050.

Fig. 14 Phase-averaged lift and moment coefficients for Ste � 0.3 at Re � 300;000 and k � 0.050.

Fig. 15 Phase-averaged normalized swirling strength at Re � 300;000 and k � 0.050 for (left) Ste � 0.35 and (right) Ste � 9.9.

84 SINGHAL ETAL.

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 8: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

Thus, although similar signature oscillations, but with much lower

amplitude, may exist for the high Strouhal number cases, they would

not manifest in these data.

2. Effect of Excitation on Reattachment

The second trend, that excitation results in earlier reattachment, is

easily observed in Fig. 13. It can be seen by the increased lift on the

pitch-downmotion of the airfoil, the lift on the pitch-downmotion for

excited cases exceeds the lift on the pitch-up motion sooner than the

baseline case. This phenomenon is related to the first trend. The

formation of coherent structures entrains high-momentum flow from

the freestream and results in momentary reattachment [9,11,12]. The

formation of small structures (such as those produced by high

excitation Strouhal numbers) only results in partial reattachment

[9,11,12]. Earlier flow reattachment reduces the aerodynamic

hysteresis.

In Fig. 16, the reattachment α (which is defined as α such that the

coefficient of lift is equal on the down- and upstroke) is plotted as a

function of the excitation Strouhal number. At low excitation

Strouhal numbers (Ste < 2), there is a stronger dependence betweenthe excitation Strouhal number and the reattachment α. This is

attributed to the timing of the actuator pulse relative to the motion of

the airfoil. If vortex-induced flow attachment does not occur before

the airfoilα is less than the criticalα, then reattachment does not occur

until the next pulse. Thus, timing is especially critical at low

excitation Strouhal numbers, because the period between pulses is

relatively large. However, at high excitation Strouhal numbers

(Ste > 6), the flow remains attached, and thus there is less

dependence between the reattachment α and the Strouhal numbers.

Flow attachment is best observed by the suction side pressure

coefficient. Shown in Fig. 17 is the suction side pressure coefficient

for the baseline (left) and Ste � 9.9 (right). In the baseline case, from20 deg ↑ to 15 deg ↓ (the stalled stage of dynamic stall), the pressure

coefficient is relatively constant with respect to the chordwise

position.When excitation is applied (right figure), we can observe the

emergence of a pressure peak (indicated by the dark colors near

x∕c � 0) in the stalled stage of dynamic stall, indicating that the flow

is attached.

3. Effect of Excitation on Aerodynamic Hysteresis

The final primary trend is that excitation results in decreased lift

and moment hysteresis and decreased dynamic stall vortex strength.

The first part of this trend is related to the generation of coherent

structures in the stalled stage of dynamic stall and earlier flow

reattachment: The excitation perturbations grow and roll up into

large-scale structures, which entrain high-momentum freestream

fluid into the separated region, energizing and reattaching the flow.

This results in increased lift production and thus reduces the lift

hysteresis for all excitation Strouhal numbers. At high excitationFig. 16 Reattachment α versus the excitation Strouhal number atRe �300;000 and k � 0.050.

Fig. 17 Suction side pressure distribution at Re � 300;000 and k � 0.050 for (left) Ste � 0 �baseline� and (right) Ste � 9.9.

SINGHAL ETAL. 85

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 9: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

Strouhal numbers, when the vortices are small and the inducedunsteady loading due to the convection and shedding of vorticalstructures is small, the moment hysteresis is significantly reduced, asshown in Fig. 13. Note that, at low excitation Strouhal numbers, themoment hysteresis is larger than at high excitation Strouhal numbers,due to the formation of large structures.An important measure of the aerodynamic hysteresis is the

damping coefficient. In particular, regions of negative damping (inthe stalled stage of dynamic stall) can lead to aerodynamic flutter [4].Negative damping is associated with clockwise loops in the momentcoefficient curve [4], as shown in Fig. 18. The damping coefficient isdefined as the closed-loop integral of the moment coefficient curve.However, to understand the effect of the actuation on the negativedamping, the integration is only performed over the closed loop of theclockwise trajectory of the moment coefficient. This quantity isreferred to as the negative damping coefficient Ξ−.Shown in Fig. 19 is the negative damping coefficient versus the

excitation Strouhal number. It is observed that low excitationStrouhal numbers result in aworse negative damping coefficient thanthe baseline. As indicated by Fig. 13, low Strouhal number excitationresults in the formation of large structures that cause large-amplitudeunsteadiness during the pitch down of the airfoil motion. Thisunsteadiness increases the magnitude of the negative dampingcoefficient. However, as the excitation Strouhal number increasesbeyond around Strouhal number of 0.3 (the natural sheddingStrouhal number), the magnitude of the negative damping

coefficient is reduced, particularly at much higher excitation

Strouhal numbers. This is attributed to the constant partial reattach-ment seen in high Strouhal number excitation, as well as the

decreased strength of the dynamic stall vortex. As the strength ofthe dynamic stall vortex decreases, the magnitude of the moment

peak also decreases [6], further reducing the magnitude of thenegative damping coefficient.Phase-locked PIVmeasurements were used to understand why the

strength of the dynamic stall vortex was decreased by excitation.

Figure 20a shows the normalized swirling strength for various phases(at angle of attack near the peak lift) at Re � 300;000 and k � 0.05for the pitch-up motion. The images are ordered by the excitationStrouhal number columnwise, that is, the left, middle, and right

columns contain the data for Ste � 0, Ste � 0.35, and Ste � 9.9,respectively. Baseline lift curves shown in Fig. 20b provide a

reference for when the PIV data were acquired with respect to the liftand moment coefficients.As discussed previously, the ejection of the dynamic stall vortex

is preceded by the accumulation of vorticity at the leading edge. This

is shown by the high swirling strength in Fig. 20 at Ste � 0 andϕ � 44.5; 53.0; 64.2 deg (α � 17–19 deg). At ϕ � 73.5 deg(α � 19.6 deg), it is seen that the dynamic stall vortex has beenejected and is convecting downstream. Because of the unsteady

nature of dynamic stall, the vortex is notwell defined.However,whencoupled with pressure results, it is apparent that the vortex is

convecting over the airfoil at this angle of attack. Compared with thebaseline case, the accumulation of vorticity is much smaller in the

excited cases, particularly for the high excitation Strouhal number.At Ste � 9.9, this accumulation of vorticity is replaced by a stream of

small, coherent vortices, which appear to remove the accumulatingvorticity near the leading edge. Thus, it follows that, as the number of

vortices increases (i.e., as Ste increases), more vorticity is removedfrom the leading edge and the dynamic stall vortex should be weaker

(as observed in the results).To an extent, this is also observed in Fig. 20 at Ste � 0.35 and

ϕ � 73.5 deg (α � 19.6 deg), where the swirling strength of thedynamic stall vortex is considerably weaker than its baseline

counterpart. However, at the highest excitation Strouhal number(Ste � 9.9), the formation and ejection of the dynamic stall vortex

is no longer observed. Although one might think that the dynamicstall vortex could have shed between the captured phases, suction

side pressure coefficient data (Fig. 21) suggests that this is not thecase. In Fig. 21, vortex shedding is marked in the figure by green

lines for the dynamic stall vortex and purple lines for the vorticesresulting from excitation. At the high excitation Strouhal number

case, the pressure spikes associated with vortex convection are notobserved. Thus, the high Strouhal number excitation seems to have

depleted the leading-edge vorticity, completely suppressing thedynamic stall vortex.Closer examination of pressure results (Fig. 13) indicates that there

is a fairly monotonic relationship between the excitation Strouhal

number and the lift peak, at least in the limited cases shown in thefigure. This is consistent with the prior discussion of the formation of

vortices, namely, that the formation of structures due to excitationremoves accumulated vorticity from the leading edge, thereby

weakening the dynamic stall vortex and reducing the magnitude ofthe lift and moment peaks.Because of the poor temporal resolution in the phase-locked PIV

data, it is difficult to directly determine the strength of the dynamic

stall vortex. As a simple alternative, the reduction in themagnitude ofmoment stall due to excitation is shown in Fig. 22. The reduction in

the moment coefficient is normalized by the baseline value. Thefigure indicates that, in general, as the excitation Strouhal number

increases, the magnitude of the peak moment coefficient decreases.This continues until the formation of the dynamic stall vortex is

suppressed, at which point the moment coefficient value seems toplateau with respect to the excitation Strouhal number, as the limited

points show in the figure. However, a more direct metric should beobtained with time-resolved PIV, in which the dynamic stall vortex

and its boundaries could be better defined.

Fig. 18 Negative damping (red) and positive damping (green) shown onmoment coefficient curve for Re � 300;000, k � 0.05, and Ste � 0.

Fig. 19 Negative damping coefficientΞ− versus the excitation Strouhalnumber Ste for Re � 300;000 and k � 0.05.

86 SINGHAL ETAL.

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 10: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

Fig. 20 Phase-averaged normalized swirling strength at Re � 300;000 and k � 0.050.

Fig. 21 Suction side pressure coefficient results for Re � 300;000, k � 0.05, and various excitation Strouhal numbers.

SINGHAL ETAL. 87

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 11: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

V. Conclusions

Dynamic stall is present inmany applications, including rotorcraft,sharply maneuvering fixed-wing aircraft, and wind turbines.Associated unsteadiness and the potential for aerodynamic flutter hasmade it a problem of interest to the aerodynamics and flow controlcommunities. Nanosecond pulse-driven dielectric barrier discharge(NS-DBD) plasma actuators have previously demonstrated theability to reattach separated flow in a variety of static stall conditions.This work, which is a first step, has sought to explore the use of theseactuators to control dynamic stall in an airfoil and to provide a betterunderstanding of dynamic stall flow physics.A single NS-DBD plasma actuator was mounted to the airfoil

near the leading edge, and baseline pressure for αm � 10 deg andΔα � 10 deg was taken for each combination of three Reynoldsnumbers and three reduced frequencies. Pressure data were acquiredfor all nine combinations of Reynolds number and reduced frequency.Phase-locked particle image velocimetry (PIV) measurements wereacquired at Re � 300;000 and k � 0.05. The baseline pressure andPIV results compare well with the stages of dynamic stall in theliterature: attached flow, the accumulation of vorticity at the leadingedge, the ejection of the accumulated vorticity in the form of thedynamic stall vortex, stalled flow, and reattachment. PIV resultsindicated that the ejection of the dynamic stall vortex was due tovortex-induced separation, a counterclockwise rotating stream ofvorticity generated over the airfoil, moving upstream during the pitch-up motion and resulting in the ejection of the dynamic stall vortex.Over all the combinations of Reynolds numbers and reduced

Strouhal numbers, the observed trends were remarkably similar.From these results, three major trends were observed. The first wasthat excitation at low Strouhal number results in the formation oflarge-scale flow structures in the stalled regime, which result inunsteadiness in the lift andmoment curves.As the excitation Strouhalnumber increases, this unsteadiness decreases and eventuallydisappears. As observed in static stall flow control experiments withNS-DBDplasma actuators, low Strouhal number excitation results inthe formation of large coherent structures, whereas high Strouhalnumber excitation results in the formation of small structures thatquickly develop, break down, and result in only partial reattachmentof the flow. Thus, high Strouhal number excitation results insmoother lift and moment coefficient curves.In a second trend, excitation also results in earlier reattachment due

to the formation of coherent structures that entrain the high-momentum freestream flow and reattach the flow. Once α of theairfoil decreases past the static stall angle, the flow remains attached.Last, excitation results in the reduction of the lift and momenthysteresis. This is reflected in the increased damping coefficient. Thisis partially due to the reattachment of the flow in the stalled regime.The other major factor is the decreased dynamic stall vortex strength,which subsequently results in a decreased peak lift and momentcoefficients. The decreased strength of the dynamic stall vortex is

attributed to the decreased vorticity accumulation at the leading edgeduring the pitch-up motion. Excitation results in the formation ofvortices before the ejection of the dynamic stall vortex. Thesevorticesremove some of the accumulated vorticity. At high excitationStrouhal numbers, this results in the suppression of the dynamic stallvortex.Although the experimental results clearly demonstrate the efficacy

of the actuators, reducing the sensor noise and uncertainty in therepeatability of the airfoil motion requires that a new mechanismshould be designed and built for improving the results. Higher phaseresolution in the phase-locked PIV will provide higher data fidelityand may reveal a well-defined relationship between the excitationStrouhal number and the dynamic stall vortex strength. Phase-lockedPIV data collected at other combinations of Reynolds numbers andreduced frequencies will better support the generalization of trendspresented in this paper.

Acknowledgments

The support of this research by the U.S. Air Force ResearchLaboratory Collaborative Center for Aerospace Sciences at the OhioState Aerospace Research Center with Miguel Visbal is greatlyappreciated. The fellowship supports for Achal Singhal from theNASA/Ohio Space Grant Consortium and for David Castañeda fromthe Fulbright-Colciencias Convocatoria from Colombia are alsogreatly appreciated.

References

[1] Leishman, J. G., Principles of Helicopter Aerodynamics, 2nd ed.,CambridgeUniv. Press, Cambridge, England, U.K., 2006, pp. 423–464.

[2] McCroskey, W. J., “The Phenomenon of Dynamic Stall,” NASA TM81264, Moffett Field, CA, 1981.

[3] Ham,N.D., “AerodynamicLoading on aTwo-DimensionalAirfoil DuringDynamic Stall,” AIAA Journal, Vol. 6, No. 10, 1968, pp. 1927–1934.doi:10.2514/3.4902

[4] Corke, T. C., and Thomas, F. O., “Dynamic Stall in Pitching Airfoils:AerodynamicDamping andCompressibility Effects,”Annual Review of

Fluid Mechanics, Vol. 47, No. 1, 2015, pp. 479–505.doi:10.1146/annurev-fluid-010814-013632

[5] Joo, W., Lee, B., Yee, K., and Lee, D., “Combining Passive ControlMethod for Dynamic Stall Control,” Journal of Aircraft, Vol. 43, No. 4,2006, pp. 1120–1128.doi:10.2514/1.17957

[6] Heine, B., Mulleners, K., Jourbert, G., and Raffel, M., “Dynamic StallControl by Passive Disturbance Generators,” AIAA Journal, Vol. 51,No. 9, 2013, pp. 2086–2097.doi:10.2514/1.J051525

[7] Weaver, D., McAlister, K. W., and Tso, J., “Suppression of DynamicStall by Steady and Pulsed Upper-Surface Blowing,” NASATP 3600,Moffett Field, CA, 1996.

[8] Post, M. L., and Corke, T. C., “Separation Control Using PlasmaActuators: Dynamic Stall Vortex Control on Oscillating Airfoil,” AIAAJournal, Vol. 44, No. 12, 2006, pp. 3125–3135.doi:10.2514/1.22716

[9] Little, J., Takashima, K., Nishihara, M., Adamovich, I., and Samimy,M., “Separation Control with Nanosecond-Pulse-Driven DielectricBarrier Discharge Plasma Actuators,” AIAA Journal, Vol. 50, No. 2,2012, pp. 350–365.doi:10.2514/1.J051114

[10] Rethmel, C., Little, J., Takashima, K., Sinha, A., Adamovich, I., andSamimy, M., “Flow Separation Control Using Nanosecond PulseDrivenDBDPlasmaActuators,” International Journal of FlowControl,Vol. 3, No. 4, 2011, pp. 213–232.doi:10.1260/1756-8250.3.4.213

[11] Clifford, C., Singhal,A., andSamimy,M., “FlowControl over anAirfoilin Fully Reversed Condition Using Plasma Actuators,” AIAA Journal,Vol. 54, No. 1, 2016, pp. 141–149.doi:10.2514/1.J054157

[12] Esfahani, A. G., Singhal, A., Clifford, C., and Samimy, M., “FlowSeparation Control over a Boeing Vertol VR-7 Using NS-DBD PlasmaActuators,” 54th AIAA Aerospace Sciences Meeting, AIAA Paper2016-0843, 2016.

[13] Singhal, A., “Unsteady Flow Separation Control over a NACA 0015Using NS-DBD Plasma Actuators,” M.S. Thesis, Dept. of MechanicalandAerospace Engineering, TheOhio State Univ., Columbus, OH, 2016.

Fig. 22 Reduction in moment coefficient peak versus excitationStrouhal number at Re � 167;000 and k � 0.05.

88 SINGHAL ETAL.

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071

Page 12: Control of Dynamic Stall over a NACA 0015 Airfoil …...Control of Dynamic Stall over a NACA 0015 Airfoil Using Plasma Actuators Achal Singhal,∗ David Castañeda,∗ Nathan Webb,†

[14] Samimy, M., Kim, J. H., Kastner, J., Adamovich, I., and Utkin, Y.,“Active Control of High-Speed and High-Reynolds-Number Jets UsingPlasma Actuators,” Journal of Fluid Mechanics, Vol. 578, May 2007,pp. 305–330.doi:10.1017/S0022112007004867

[15] Samimy, M., Webb, N., and Crawley, M., “Excitation of Free Shear-Layer Instabilities in High-Speed Flows,” AIAA Journal (in review).

[16] Greenblatt, D., and Wygnanski, I., “Dynamic Stall Control by PeriodicExcitation: Part 1: NACA 0015 Parametric Study,” Journal of Aircraft,Vol. 38, No. 3, 2001, pp. 430–438.doi:10.2514/2.2810

[17] Adrian, R. J., Christensen, K. T., and Liu, Z. C., “Analysis andInterpretation of Instantaneous Turbulent Velocity Fields,” Experimentsin Fluids, Vol. 29, No. 3, 2000, pp. 275–290.doi:10.1007/s003489900087

[18] Mulleners, K., and Raffel, M., “The Onset of Dynamic Stall Revisited,”Experiments in Fluids, Vol. 52, No. 3, 2012, pp. 779–793.doi:10.1007/s00348-011-1118-y

[19] Obabko, A., and Cassel, K., “Detachment of the Dynamic Stall VortexAbove a Moving Surface,” AIAA Journal, Vol. 40, No. 9, 2002,pp. 1811–1822.doi:10.2514/2.1858

[20] Frankhouser, M.W., “Nanosecond Dielectric Barrier Discharge PlasmaActuator Flow Control of Compressible Dynamic Stall,” M.S. Thesis,Dept. of Mechanical and Aerospace Engineering, The Ohio State Univ.,Columbus, OH, 2015.

[21] Harris, F. D. J., Tarzanin, F. J. J., and Fisher, R. K., “Rotor High SpeedPerformance, Theory Versus Test,” Journal of the American HelicopterSociety, Vol. 15, No. 3, 1970, pp. 35–44.doi:10.4050/JAHS.15.35

[22] Shih, C., Lourenco, L. M., van Dommelen, L. L., and Korthapalli, A.,“Unsteady Flow past an Airfoil Pitching at a Constant Rate,” AIAA

Journal, Vol. 30, No. 5, 1992, pp. 1153–1161.doi:10.2514/3.11045

[23] Wernert, P., Geissler, W., Raffel, M., and Kompenhans, J.,“Experimental and Numerical Investigations of Dynamic Stall on aPitching Airfoil,” AIAA Journal, Vol. 34, No. 5, 1996, pp. 982–989.doi:10.2514/3.13177

A. NaguibAssociate Editor

SINGHAL ETAL. 89

Dow

nloa

ded

by O

HIO

ST

AT

E U

NIV

ER

SIT

Y L

IBR

AR

IES

on M

arch

9, 2

018

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

6071