+ All Categories
Home > Documents > Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators

Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators

Date post: 08-Dec-2016
Category:
Upload: sven
View: 216 times
Download: 2 times
Share this document with a friend
7
Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators Jochen Kriegseis, Daniel Schröter, Katrin Barckmann, Alexander Duchmann, § Cameron Tropea, and Sven Grundmann ** Technische Universität Darmstadt, D-64347 Griesheim, Germany DOI: 10.2514/1.J052159 The counteraction of changing environmental conditions (i.e., changes of pressure level and airflow speed) on the resulting plasma-actuator performance is demonstrated in the present work. The impact of these changing (fluctuating and/or transient) airflow conditions on the performance of dielectric-barrier-discharge plasma actuators is suppressed using a novel closed-loop performance-control procedure. The goal of controlling a preset plasma- actuator performance online and in situ is achieved and successfully demonstrated. This novel approach represents the first step toward optimal-discharge-based flow control because, beyond the common purpose of favorably manipulating the airflow, any advanced dielectric-barrier-discharge-based flow-control system will necessarily require an appropriate closed-loop performance control of the discharge device. Nomenclature C eff = effective discharge capacitance, pF C p = probe capacitance, nF C 0 = passive-component (cold) capacitance, pF d = dielectric thickness, mm F = actuator thrust production, N f = operating frequency, kHz G b = background value of Gx G p = peak value of Gx Gx = streamwise gray-value distribution h = relative humidity, % J = evaluation interval j = index (time step) K = number of recorded discharge cycles K d = control parameter K p = control parameter k = index (discharge cycle), k K M = Mach number P A = actuator power consumption, W P A = preset power consumption, W P min A = minimum-recorded power consumption, W p = static pressure, bar p max = maximum pressure, bar p min = minimum pressure, bar p t = total pressure, bar p 0 = ambient pressure, bar Q = charge, nC T = temperature, K T i = control parameter t = time, s t 1 , t 2 , t 3 = characteristic times, s V = operating voltage, kV V p = probe voltage, V w 1 = upper electrode width, mm w 2 = lower electrode width, mm x = streamwise direction, mm x max = downstream boundary of plasma domain, mm x min = upstream boundary of plasma domain, mm γ = isentropic exponent ΔP j A = power-consumption deviation, W Δt = time interval, s Δx = plasma length, mm Q = relative performance ρ = density, kg=m 3 σ P A = relative standard deviation ϕ = general variable Ω = control signal, V I. Introduction I N RECENT years, plasma actuators have proven to be an attractive and promising alternative to conventional flow-control devices [1], and therefore enjoy increasing importance for various flow-control scenarios in the community [2,3]. The variety of investigated flow-control applications ranges from velocities of a few meters per second (e.g., [4]) to supersonic conditions (e.g., [5]). However, it is commonly assumed in the flow-control community that the actuator discharge manipulates the airflow, but not vice versa. The dielectric-barrier-discharge plasma-actuator performance is influenced by numerous thermodynamic and kinematic environ- mental parameters. The effect of changing ambient conditions (humidity [68], temperature [8,9], and ambient pressure [915]) on the discharge intensity has been documented in several publications. Changes of these background conditions act differently on the plasma-actuator performance. An increase in relative humidity, for instance, monotonously reduces the power consumption, but first leads to a flow-rate augmentation of the resulting wall jet at moderate humidity levels before an overall decrease occurs for both quantities. Similar nontrivial interrelationships are observed when the ambient pressure changes. The close interrelation between breakdown voltage and ambient pressure is well known from the Paschen curves [16]. In consequence, above the Stoletow point [17], any reduction of pressure increases the discharge intensity. In contrast, the transferred momentum does not change monotonously with decreasing pressure, Presented as Paper 2012-2803 at the 6th AIAA Flow Control Conference, New Orleans, Louisiana, 2528 June 2012; received 3 July 2012; revision received 26 October 2012; accepted for publication 4 November 2012; published online 27 February 2013. Copyright © 2012 by Jochen Kriegseis. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission. Copies of this paper may be made for personal or internal use, on condition that the copier pay the $10.00 per-copy fee to the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; include the code 1533-385X/13 and $10.00 in correspondence with the CCC. *Postdoctoral Associate, Flughafenstr. 19; currently Mechanical and Manufacturing Engineering, University of Calgary, Calgary, AB, Canada; [email protected]. Member AIAA (Corresponding Author). Masters Student, Center of Smart Interfaces, Flughafenstr. 19. Research Assistant, Center of Smart Interfaces, Flughafenstr. 19; [email protected]. Member AIAA. § Research Assistant, Center of Smart Interfaces, Flughafenstr. 19; [email protected]. Member AIAA. Professor, Institute of Fluid Mechanics and Aerodynamics, Center of Smart Interfaces, Flughafenstr. 19; [email protected]. Member AIAA. **Research Coordinator, Center of Smart Interfaces, Flughafenstr. 19; [email protected]. Member AIAA. 961 AIAA JOURNAL Vol. 51, No. 4, April 2013 Downloaded by UNIVERSITY OF MICHIGAN on April 3, 2013 | http://arc.aiaa.org | DOI: 10.2514/1.J052159
Transcript
Page 1: Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators

Closed-Loop Performance Control of Dielectric-Barrier-DischargePlasma Actuators

Jochen Kriegseis,∗ Daniel Schröter,† Katrin Barckmann,‡ Alexander Duchmann,§

Cameron Tropea,¶ and Sven Grundmann**

Technische Universität Darmstadt, D-64347 Griesheim, Germany

DOI: 10.2514/1.J052159

The counteraction of changing environmental conditions (i.e., changes of pressure level and airflow speed) on the

resulting plasma-actuator performance is demonstrated in the present work. The impact of these changing

(fluctuating and/or transient) airflow conditions on the performance of dielectric-barrier-discharge plasma actuators

is suppressed using a novel closed-loop performance-control procedure. The goal of controlling a preset plasma-

actuator performance online and in situ is achieved and successfully demonstrated. This novel approach represents

the first step toward optimal-discharge-based flow control because, beyond the common purpose of favorably

manipulating the airflow, any advanced dielectric-barrier-discharge-based flow-control system will necessarily

require an appropriate closed-loop performance control of the discharge device.

Nomenclature

Ceff = effective discharge capacitance, pFCp = probe capacitance, nFC0 = passive-component (cold) capacitance, pFd = dielectric thickness, mmF = actuator thrust production, Nf = operating frequency, kHzGb = background value of G�x�Gp = peak value of G�x�G�x� = streamwise gray-value distributionh = relative humidity, %J = evaluation intervalj = index (time step)K = number of recorded discharge cyclesKd = control parameterKp = control parameterk = index (discharge cycle), k ∈ KM = Mach numberPA = actuator power consumption, WP�A = preset power consumption, WPminA = minimum-recorded power consumption, Wp = static pressure, barpmax = maximum pressure, barpmin = minimum pressure, barpt = total pressure, barp0 = ambient pressure, bar

Q = charge, nCT = temperature, KTi = control parametert = time, st1, t2, t3 = characteristic times, sV = operating voltage, kVVp = probe voltage, Vw1 = upper electrode width, mmw2 = lower electrode width, mmx = streamwise direction, mmxmax = downstream boundary of plasma domain, mmxmin = upstream boundary of plasma domain, mmγ = isentropic exponentΔPjA = power-consumption deviation, WΔt = time interval, sΔx = plasma length, mmQ

= relative performanceρ = density, kg=m3

σPA = relative standard deviationϕ = general variableΩ = control signal, V

I. Introduction

I N RECENT years, plasma actuators have proven to be anattractive and promising alternative to conventional flow-control

devices [1], and therefore enjoy increasing importance for variousflow-control scenarios in the community [2,3]. The variety ofinvestigated flow-control applications ranges from velocities of a fewmeters per second (e.g., [4]) to supersonic conditions (e.g., [5]).However, it is commonly assumed in the flow-control communitythat the actuator dischargemanipulates the airflow, but not vice versa.The dielectric-barrier-discharge plasma-actuator performance isinfluenced by numerous thermodynamic and kinematic environ-mental parameters. The effect of changing ambient conditions(humidity [6–8], temperature [8,9], and ambient pressure [9–15]) onthe discharge intensity has been documented in several publications.Changes of these background conditions act differently on the

plasma-actuator performance. An increase in relative humidity, forinstance, monotonously reduces the power consumption, but firstleads to a flow-rate augmentation of the resulting wall jet at moderatehumidity levels before an overall decrease occurs for both quantities.Similar nontrivial interrelationships are observed when the ambientpressure changes. The close interrelation between breakdownvoltageand ambient pressure is well known from the Paschen curves [16]. Inconsequence, above the Stoletow point [17], any reduction ofpressure increases the discharge intensity. In contrast, the transferredmomentumdoes not changemonotonouslywith decreasing pressure,

Presented as Paper 2012-2803 at the 6th AIAA Flow Control Conference,New Orleans, Louisiana, 25–28 June 2012; received 3 July 2012; revisionreceived 26 October 2012; accepted for publication 4 November 2012;published online 27 February 2013. Copyright © 2012 by Jochen Kriegseis.Published by the American Institute of Aeronautics and Astronautics, Inc.,with permission. Copies of this paper may be made for personal or internaluse, on condition that the copier pay the $10.00 per-copy fee to the CopyrightClearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923; includethe code 1533-385X/13 and $10.00 in correspondence with the CCC.

*Postdoctoral Associate, Flughafenstr. 19; currently Mechanical andManufacturing Engineering, University of Calgary, Calgary, AB, Canada;[email protected]. Member AIAA (Corresponding Author).

†Masters Student, Center of Smart Interfaces, Flughafenstr. 19.‡Research Assistant, Center of Smart Interfaces, Flughafenstr. 19;

[email protected]. Member AIAA.§Research Assistant, Center of Smart Interfaces, Flughafenstr. 19;

[email protected]. Member AIAA.¶Professor, Institute of Fluid Mechanics and Aerodynamics, Center of

Smart Interfaces, Flughafenstr. 19; [email protected]. MemberAIAA.

**Research Coordinator, Center of Smart Interfaces, Flughafenstr. 19;[email protected]. Member AIAA.

961

AIAA JOURNALVol. 51, No. 4, April 2013

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

ICH

IGA

N o

n A

pril

3, 2

013

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

2159

Page 2: Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators

but rather a (local [15]) maximum is observed for the induced thrustbelow ambient condition before it decreases under further pressurereductions. Changes in altitude for any aerial vehicle imply simulta-neous changes of all state variables, in which decreasing pressure andtemperature act competitively on the discharge performance. Inaddition, the aerodynamic influence of the freestream velocity on thegas-discharge intensity has recently been identified to significantlyreduce the actuator performance [18–20]. Obviously, the variety ofpossible combinations of promoting and diminishing effects on thedischarge intensity limits the predictability of environmental influ-ences on the plasma-actuator performance.Kriegseis et al. [21,22] introduced a performance-characterization

procedure, which is based on an improved interpretation of voltage–charge cyclograms (Lissajous figures) and a light-emission analysisof the discharge domain.With these new diagnosis approaches, it hasbeen possible to quantify the performance drop of plasma actuatorsfor different airflow speeds (up to 30% [19,20]). Assuming aninitially optimized plasma-actuator setup, precisely adjusted powerlevel, and perfectly matched impedance of the electrical circuit, theseenvironmental influences on the discharge result in a detuned setup ofthe plasma actuator, which consequently operates at unnecessarilyhigh energy requirements. The more important consequence is asignificantly reduced control authority of the plasma actuator as aflow-control device. To ensure a constant plasma-actuator perfor-mance during operation, a real-time and in situ evaluation of theperformance is desirable.The present study addresses this requirement with a real-time-

characterization approach, which allows a closed-loop performancecontrol of the electrical plasma-actuator setup. Because of the closed-loop character of the proposed control concept, a compensation forany performance changes can be successfully achieved without adetailed knowledge about the complex interrelationship between theinvolved thermodynamic and kinematic environmental parameters.In the first step, the relevant performance-identifying approaches

[21,22] are modified such that the required data are continuouslyacquired and processed to provide an online evaluation basis asproposed by Kriegseis et al. [23]. Then, in the second step, anyidentified performance changes can be counteracted in a closed-loopcircuit, based on the known interrelations of the involved quantities.The novel tool is tested and optimized by applying it to an artificialflow situation of changing pressure andMach number in a blowdownwind tunnel. However, the main objective of the present study was tooutline the potential of the presented closed-loop-control concept fordischarge-based aerodynamic flow control in general, rather than anoptimization of one particular closed-loop-controller design for anartificial test case in a blowdown wind tunnel.

II. Experimental Procedure

The experiments were conducted in the blowdown wind tunnel ofthe Technische Universität Darmstadt. This wind tunnel is designedto operate at Mach numbers in the range of 0.4 < M < 4. Thedimensions of the test section are 150 × 150 mm2. The plasmaactuator consisting of two copper electrodes (w1 � 2.5 mm,w1 � 10 mm) and a Kapton dielectric (d � 0.4 mm) was mountedon the side window of the test section, as shown in Fig. 1. Toretroactively verify the online evaluation of the performance, similarto previous investigations [19–21], a complementary metal-oxidesemiconducter camera (Phantom V12.1, 512 × 512 pixels, 24 fps;Nikon 105 mm, AF Micro-Nikkor f/2.8D) was used to record thespatiotemporal light emission of the discharge during the power-consumption analysis.The electrical control circuit, as shown in Fig. 1b, was built up

using a computer-based digital oscilloscope (Picotech PicoScope4424, four channels, 2500 points per channel, 10 MS/s) to record theoperating voltage V (TESTEC HVP-15HF, 1000:1) and the voltageVp across the charge probe capacitorCp � 22 nF (LeCroy PP006A,10:1). The operating voltage V was generated by a high-voltagetransformer circuit board (GBS Elektronik, Minipuls 2 [24]), whichwas driven by an external laboratory power supply (Voltcraft VSP2410, variable output 0–24 V dc) and a function generator [GW

Instek, SFG-2004, fixed frequency (sine): f � 12.0 kHz]. For moredetails about the chosen high-voltage transformer-circuit-boardworking principle, the reader is referred to the manufacturer’s webpage [24].The governing equation to calculate the power consumption PA

from Lissajous figures (see Fig. 2a) is

PA � fIQ dV � f

K

XKk�1

IkCpVp dV (1)

in which CpVp � Q is the charge crossing the actuator [22].Equation (1) represents the average ofK discharge cycles. It has beendemonstrated by Kriegseis et al. [21] that changes of the plasmalength and corresponding discharge light-emission distribution onthe dielectric are a robust means to quantify changes of the actuatorperformance (i.e., power consumption PA and thrust production F).Therefore, the simultaneously recorded light-emission data are usedsubsequently to retroactively validate the online diagnostic tool bymeans of the temporal plasma-length evolutionΔx�t� as discussed inSec. III.A. The gray-value analysis of the light-emission images leadsto the (chordwise) plasma length:

Δx � xmax�G�x� > Gb� − xmin�G�x� > Gb� (2)

in whichGb andGp are the gray values of the background and light-emission peak, respectively. A typical result of the gray-valueanalysis is shown qualitatively in Fig. 2b.Reproducible airflow conditions were achieved, operating the

blowdown wind tunnel at a constant valve position, thus generating ablowdown from initiallyM � 0.85 (cruise flight) to quiescent air atchanging pressure levels (p � 0.9–1.5 bar). The resulting timetraces of the state variables static and total pressurep,pt are shown inFig. 3, alongside the Mach numberM.In Fig. 3, three characteristic times are highlighted for p�t1� �

pmax, p�t2� � p0, p�t3� � pmin with the purpose of distinguishingpressure and air-speed effects on the discharge performance. For thesake of completeness, the interdependency of pressure and Machnumber is stated, that is,

p

pt��γ − 1

2M2 � 1

� γ1−γ

(3)

a) Wind-tunnel test section b) Detail: closed-loop circuit

Fig. 1 Sketch of experimental setup: a) wind-tunnel test section and

overhead camera (CAM); and b) detailed view of electrical plasma-actuator setup and closed-loop circuit comprising a function generator(FG), power supply (PS), high-voltage (HV) transformer, notebookcomputer (NB), and plasma actuator.

a) Power consumption b) Plasma length

Fig. 2 Qualitative example results of a) electric measurements for theLissajous figure and b) gray-value analysis of the light-emission images(according to Kriegseis et al. [21,22]).

962 KRIEGSEIS ETAL.

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

ICH

IGA

N o

n A

pril

3, 2

013

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

2159

Page 3: Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators

in which γ is the isentropic exponent of gas (γ � 1.4 fordiatomic gases).

III. Online Diagnostic and Control Approach

The recorded data of operating voltage V and probe voltage Vpwere continually acquired by the notebook computer from the digitaloscilloscope. The consumed power PA for monitoring or controllingpurposes was computed from voltage–charge cyclograms. Based onEq. (1), the evaluation algorithm calculates the power consumptionfor every time step tj according to

PjA � fItj

tj−1Q dV with Q � CpVp (4)

When using the diagnostic tool in the closed-loop-control mode, thealgorithm furthermore comparesPjAwith a preset power levelP

�A, and

calculates the deviation

ΔPjA � PjA − P�A (5)

as the relevant control error. The control signal for the next time steptj�i is estimated based on a proportional–integral derivative (PID)control algorithm:

Ωj�1 � Kp�ΔPjA� �1

Ti

Ztj

0

�ΔPjA� dt� Kdd�ΔPjA�

dt(6)

This type of controller has been chosen for the ease of adjusting thecontroller gains for robust and stable controller characteristics.After the output Ωj�1 has been calculated, it is fed into the controlinput of the power supply, where it is transformed directly into thecorresponding voltage for the high-voltage transformer (see Fig. 1b).For all experiments, the quiescent-air operating conditions of

f � 12 kHz, V � 10 kV, and p0 � 1.028 bar resulted in an initialpower consumption PA � P�A � 7.1 W for t < 0. The experimentsand corresponding basic parameter settings are listed in Table 1.

A. Proof of Concept

In a preliminary proof of concept by Kriegseis et al. [23], twoexperiments (ExpA, ExpB) were conducted to demonstrate thecapabilities and the two operating modes of the online diagnostictool, and to verify the validity of the close correlation of powerconsumption and resulting plasma length across the entire range ofrapidly changing environmental conditions during the experimentalruns. This set of experiments is briefly reviewed in the following toexplain the control concept. The advanced control concept and theachieved results are discussed in Secs. III.B and IV, respectively.Notethat the wind-tunnel blowdown for the preliminary experimentsExpA and ExpB was slightly faster than for the main experimentsExp1–Exp5, which does not affect the qualitative comparability ofthe discussed results for the characteristic times t1 − t3.In the first experiment (ExpA — monitoring mode), the

performance changes for a preset power consumption of P�A �7.1 Wweremonitored online according to Eq. (4) in time intervals ofΔt � tj − tj−1 ≈ 0.5 s. Because the hardware/software interfaceswere preimplemented in LabVIEW, but the source code of theLissajous-figure analysis was implemented in MATLAB, for thepreliminary proof of concept, every time step required the time-consuming communication between both software packages. Inthe second experiment (ExpB — controlling mode), a simpleproportional-derivative (PD) control algorithm according to Eq. (6)remote controlled the laboratory power supply of the electricalplasma-actuator setup, thus closing the control loop. The results areshown in Fig. 3. Because of the large time steps of the controller(Δt ≈ 0.5 s), the integral term of Eq. (6) had to be set to zero in thepreliminary control concept.The light-emission data of ExpA and ExpB are used to

retroactively validate the online diagnostic tool by means of thetemporal plasma-length evolution Δx�t� based on the closecorrelation of Δx and PA according to Kriegseis et al. [21].In Fig. 4a, the results of the monitoring experiment (ExpA) are

depicted. They clearly reveal the impact of the transient flowconditions on the actuator power PA. Immediately after the wind-tunnel valve was opened at t � 0, a power peak occurred due to aninitial expansion wave passing the test section, which was followedby a significant performance drop (PA � 4.8 W) once the blowdownscenario was fully developed at t1 under adverse pressure conditionsat maximum airflow speed (see Table 2). For decreasing Machnumber and pressure at t2, a constantly reduced performance(PA � 4.9 W) was observed, solely due to the impact of high-speedairflow (M � 0.75) at ambient-pressure conditions p0, which agreeswith the reports of Barckmann et al. [20] andKriegseis et al. [19]. Theinfluence of theminimum pressure at t3 exceeded the adverse airflowimpact (M � 0.69), which resulted in an increased performancePA � 8.7 W as compared to the preset initial value P�A � 7.1 W.This augmentation is in good agreement with the reports of Abe et al.[14] and Bénard et al. [12,25]. Thereafter, for t > t3, all quantitiesasymptotically returned to their initial values again. The plasmalength Δx of the simultaneously recorded light emissions,retroactively determined according to Kriegseis et al. [21], furtherconfirms the correctness of the online characterization of the powerconsumption.

Fig. 3 Transient flowconditions during experimentation.Time traces of

static and total pressure p, pt, Mach numberM; characteristic times arelabeled t1, t2, t3.

Table 1 List of conducted experiments and corresponding basic parameter settingsa

Experiment number Controller concept Time step Δt, ms Characteristictimes, s

Mach numberM Maximum/minimumstatic

pressure, bar

t1 t2 t3 pmax pmin

ExpA — 500 2.4 7.8 14.5 0.84 1.45 0.90ExpB PDb 500 2.4 7.8 14.5 0.84 1.45 0.90Exp1 — 17.3 3.1 13.8 22.3 0.86 1.52 0.89Exp2–5 PIc 17.3 3.1 13.8 22.3 0.86 1.52 0.89

aConstant settings: f � 12 kHz, V � 10 kV, P�A � 7.1 W, and p0 � 1.028 bar.bProportional derivative.cProportional–integral.

KRIEGSEIS ETAL. 963

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

ICH

IGA

N o

n A

pril

3, 2

013

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

2159

Page 4: Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators

For identical initial and airflow conditions, the results of theclosed-loop-control experiment (ExpB) are shown in Fig. 4b. Thecontrol algorithm failed for the very strong initial power oscillation ofthe passing expansion wave. Thereafter, the performance drop wasidentified at t1 and the algorithm counteracted this drop, as shown bythe slope of the control signal Ω. At t � 5.5 s, the control algorithmcollapsed for a single time step and caused a power overshoot due toan erratic signal. Apart from this peak, the algorithm successfullyconducted a closed-loop control of the power variations, which againis confirmed by the results for the plasma length Δx.

B. Advanced Control Circuit

To develop a more effective control algorithm, it is necessary toreduce the size of the control loop (i.e., minimize Δt). To eliminatethe time-consuming communication between both softwarepackages, the source code for the Lissajous-figure analysis wastranslated from MATLAB to LabVIEW. In addition, a programparallelizing multiple tasks has been developed to optimize thecomputational time for each time step. This program uses pipeliningtechnology and is especially powerful on multicore processors.The data acquisition of the digital oscilloscope samples the two

channels (V and Vp) in block mode with 250 data points at a samplerate of 1 MS∕s. These data points are filtered using a Savitzky–Golayfilter [26], and then separated cyclewise (K � 1) into individualLissajous figures with 83 data points each. For each block, the powerconsumption is calculated from Eq. (1), based on which the newremote signal of the closed-loop control circuit is estimated accordingto Eq. (6). The basic tasks processed for each iteration aresummarized in Table 3. Furthermore, the table shows the average run

time of the tasks for a standard dual-core laptop computer (2.2 GHz,4 GB RAM).For an optimized pipelining program, all parallelized tasks have to

be distributed over the available CPU cores to minimize theoverall run time.Obviously, the data acquisition (task 1) takesmost ofthe time. Therefore, at the same time as the current data block isacquired using CPU#1, the program performs all other computations(tasks 2–5) with the previous data block using CPU#2 as sketched inFig. 5. In total, the program needs about Δt � 17.3 ms to perform acomplete cycle of data acquisition and computations for 250data points per channel at a sampling rate of 1 MS∕s, which isan improvement of one order of magnitude with respect toconsumed time.As the repetition rate of the controller was increased, the system

dynamics also changed. Practice showed that the PD controller usedbefore was not suitable for the given noise level, as the Δt reductionresulted in higher noise gradients. Consequently, a proportional–integral (PI) controller based on Eq. (6) was chosen.To determine the controller gains, the open-loop step responsewas

analyzed using the Ziegler–Nichols step-response method [28]. Inthe literature, there are different ways shown for determining thecontroller gainsKp and Ti. Åström and Hägglund [29] determine thesystem latency with the help of a line of best fit through the point ofthe highest slope of the process variable, whereas the PID ControlToolkit user manual from National Instruments [30] uses a referencepoint of the process variable, which has 63.2%of themaximumvalueof the process variable.Additionally, the time gapΔt between twomeasurements is still in

the same order of magnitude as the system latency (see Fig. 6), suchthat an accurate determination of the system latency beyond thetemporal resolution of the control circuit is impossible. For the stepresponse shown in Fig. 6, this led to a proportional gain Kp rangingfrom 0.2 to 1, and an integral time Ti ranging from 0.05 to 0.1.Therefore, these values served as a first starting point for tuning thecontroller gains manually. Based on this semimanual procedure, itwas possible to implement a PI controller with reduced overshoot,minimized steady-state error, and suppressed oscillations.

IV. Results

Five experiments were conducted with the advanced controlalgorithm based on the PI-control concept. Exp1 was operated inmonitoring mode to provide a reference for the discharge-performance variations. Subsequently, four experiments in thecontrol mode (Exp2–Exp5) were conducted (see Table 1).

a) Monitoring mode (ExpA) b) Controlling mode (ExpB)

Fig. 4 Proof-of-concept results of consumed power PA, operating voltage V, plasma length Δx and remote-control signal Ω for a) monitoring andb) controlling modes of operation (cp. Table 2).

Table 2 Measured data at characteristic timesa

p, bar M PA, W V, kV Δx, mm Ω, Vt < 0 1.03 0 7.2m;c 10.0m;c 2.9m;c 0.89c

t1 � 2.4 s 1.45 0.84 4.8m 5.7c 108m 10.1c 1.9m 1.6c 0.97c

t2 � 7.8 s 1.03 0.75 4.9m 7.5c 9.4m 10.2c 2.0m 2.8c 0.95c

t3 � 14.5 s 0.89 0.69 8.7m 7.2 10.3m 9.8c 3.3m 2.8c 0.78c

aThe superscript m means monitoring mode (ExpA), whereas cmeans controlling mode (ExpB); (cp. time traces in Figs. 3and 4).

Table 3 List of tasks to be processed duringeach performance-control iteration and

corresponding (average) run time

Task Description Average run time, ms

1 Data acquisition 15.92 Filtering 0.83 Data separation 0.34 Power computation 0.15 Control computations 0.1

Fig. 5 Timing diagram for the pipelined performance-control circuitrunning on two CPU cores (according to NI Developer Zone [27]).

964 KRIEGSEIS ETAL.

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

ICH

IGA

N o

n A

pril

3, 2

013

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

2159

Page 5: Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators

A. Monitoring/Controlling

Figure 7a shows the result of the new monitoring experiment(Exp1). The improvement of the diagnostic tool due to the increasedsampling rate is clearly visible even in the monitoring mode.At the first instant of time after the wind-tunnel valve is opened

at t � 0, the typical actuator power peak can be recognized inthe diagram due to the passing expansion wave. Thereafter, thesignificant power drop occurred at t1 (PA � 3.7 W) under adversepressure and airflow conditions. At t2, the performance reductionsolely due to airflow influence is less pronounced (PA � 6.1 W),before the favorable pressure conditions lead to an increasedperformance of PA � 9.6 W at t3. The wind-tunnel valve is closedat t � 43 s.The results of Exp2 (controlling mode) are shown in Fig. 7b as a

direct comparison to Exp1. Both initial performance peak andsubsequent drop are reduced significantly as the controlled mimicsthe respective slopes with opposite signs. At t ≈ 7 s, the variations ofenvironmental conditions are entirely compensated, that is, theremaining control error ΔPjA is minimized.Qualitatively, the slope of the control signal Ω of the controlled

case directly mirrors the slope of the power PA of theuncontrolled case.

B. Parameter Optimization

Although considered an artificially strong change of theenvironmental conditions, the extreme performance gradients inthe time span between t � 0 and t1 emerged as an excellent test-casescenario for further control-circuit-optimization efforts. In particular,an interval J was defined, which was used to determine the relativestandard deviation:

σPA �1

P�A

������������������������������������������1

J − 1

XJj�1�PjA − P�A�

2

vuut (7)

of the controlled cases as a measure of control success. Differentcombinations of control parameters Kp and Ti were implemented in

the control algorithm (6), as listed in Table 4 together with therespective values of σPA . The results of all experiments are shownin Fig. 8.The reduction of the control error is reflected directly in the

standard deviation, which is 24.22% for the open-loop (monitoring)experiment Exp1, and has been reduced to 1.7% for the optimizedclosed-loop controller in experiment Exp5. However, the success ofcontrol is highly dependent on the choice of the control parameters. Ahigher control gain Kp increases the speed of the controller, but

Fig. 6 Open-loop step response of the control circuit; control signal Ω � process variable, power consumption PA � control variable.

a) Monitoring mode (Exp1) b) Controlling mode (Exp2)Fig. 7 Results of consumed power PA and control signal Ω for a) monitoring and b) controlling modes of operation (advanced PI-control concept).

Table 4 List of control parametersKp andTi as used forequation (6)

Kp Ti σPA , %a

Exp1 Monitoring 24.22Exp2 0.2 0.05 5.99Exp3 0.2 0.03 5.73Exp4 0.06 0.005 9.07Exp5 0.1 0.005 1.70

aRelative standard deviation σPA for the chosen intervalJ of tmax − tmin � 10 s (cp. Fig. 8).

Fig. 8 Comparison of the actuator power consumptionPA for differentcombinations of control parameters Kp and Ti as used for Eq. (6);considered evaluation interval J is indicated by arrows (cp. Table 4).

KRIEGSEIS ETAL. 965

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

ICH

IGA

N o

n A

pril

3, 2

013

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

2159

Page 6: Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators

comes at the cost of an increasing actuating variable, which isnaturally limited. The influence of the controller gain can be seen inthe beginning of the experiment at the difference of Exp4 incomparison to Exp5, in which the first performance drop is reducedfrom Pmin

A � 4.43 W to PminA � 5.57 W by an increase of Kp.

Figure 8 also yields that a reduction of Ti enhances the controlperformance by decreasing the steady-state error, but this alsodecreases the stability margin of the controller. For the value of Tichosen to low the controller shows heavy oscillations, whichtherefore leads to instability.

V. Conclusions

The present work successfully demonstrates a counteraction ofchanging environmental conditions (i.e., changes of pressure leveland airflow speed) on the resulting plasma-actuator performance.Continuing in the direction of previous studies, a novel online

characterization approach is introduced, allowing an in situ conditionmonitoring of the discharge performance. Moreover, this new toolfeatures a closed-loop control circuit, which assures constant plasma-actuator performance independently of the environmental con-ditions. A proportional–integral control algorithm is successfullyimplemented, closing the loop between discharge measurements andpower supply of the electric setup. Based on the relative performanceΠ as evaluation measure, the slope of the control signal Ω of thecontrolled casewas identified to directlymirror the slope of the powerPA of the uncontrolled case.From this result, the conclusion can be drawn that by means of

steadily processed real-time performance data, it is possible toachieve a constant plasma-actuator performance during operationunder fluctuating and transient flow conditions. This is an importantinsight implying significant consequences, because beyond thecommon purpose of favorably manipulating the airflow, anyadvanced dielectric-barrier-discharge-based flow-control systemwill necessarily require an appropriate closed-loop performancecontrol of the discharge device.The first successful incorporation of the closed-loop control

circuit for in-flight experiments on a full-sized motorized gliderwas demonstrated recently by the authors. Application of the onlinecontrol tool assures constant control performance of the actuatordespite variations of the environmental conditions (relative humidity,temperature, density) during transition control experiments. Furtherenhancements of the closed-loop control system for flight applicationare projected.

Acknowledgments

The authors gratefully acknowledge the financial support by theGerman Research Foundation (Deutsche Forschungsgemeinschaft)under grant EXC 259.

References

[1] Cattafesta, L. N., and Sheplak,M., “Actuators for Active FlowControl,”Annual Review of Fluid Mechanics, Vol. 43, No. 1, 2011, pp. 247–272.doi:10.1146/annurev-fluid-122109-160634

[2] Moreau, E., “Airflow Control by Non-Thermal Plasma Actuators,”Journal of Physics D: Applied Physics, Vol. 40, No. 3, 2007,pp. 605–636.doi:10.1088/0022-3727/40/3/S01

[3] Corke, T. C., Enloe, C. L., and Wilkinson, S. P., “Dielectric BarrierDischarge Plasma Actuators for Flow Control,” Annual Review of Fluid

Mechanics, Vol. 42, No. 1, 2010, pp. 505–529.doi:10.1146/annurev-fluid-121108-145550

[4] Grundmann, S., Frey, M., and Tropea, C., “Unmanned Aerial Vehicle(UAV) with Plasma Actuators for Separation Control,” AIAA Paper2009-0698, Jan. 2009.

[5] Leonov, S. B., andYarantsev, D. A., “Near-Surface Electrical Dischargein Supersonic Airflow: Properties and Flow Control,” Journal of

Propulsion and Power, Vol. 24, No. 6, 2008, pp. 1168–1181.doi:10.2514/1.24585

[6] Anderson, R., and Roy, S., “Preliminary Experiments of BarrierDischarge Plasma Actuators Using Dry and Humid Air,” AIAA Paper2006-0369, Jan. 2006.

[7] Bénard, N., Balcon, N., and Moreau, E., “Electric Wind Producedby a Surface Dielectric Barrier Discharge Operating Over aWide Range of Relative Humidity,” AIAA Paper 2009-0488,Jan. 2009.

[8] Opaits, D. F., Neretti, G., Zaidi, S. H., Shneider, M. N., Miles, R. B.,Likhanskii, A. V., andMacheret, S. O., “DBD Plasma Actuators Drivenby a Combination of Low Frequency Bias Voltage and NanosecondPulses,” AIAA Paper 2008-1372, Jan. 2008.

[9] Versailles, P., Gingras-Gosselin, V., andVo, H., “Impact of Pressure andTemperature on the Performance of Plasma Actuators,” AIAA Journal,Vol. 48, No. 4, 2010, pp. 859–863.doi:10.2514/1.43852

[10] Gregory, J. W., Enloe, C. L., Font, G. I., and McLaughlin, T. E., “ForceProduction Mechanisms of a Dielectric-Barrier Discharge PlasmaActuator,” AIAA Paper 2007-0185, Jan. 2007.

[11] Wu, Y., Li, Y., Jia,M., Song, H., Guo, Z., Zhu, X., and Pu, Y., “Influenceof Operating Pressure on Surface Dielectric Barrier Discharge PlasmaAerodynamic Actuation Characteristics,” Applied Physics Letters,Vol. 93, No. 3, 2008, p. 031503.doi:10.1063/1.2964193

[12] Bénard, N., Balcon, N., and Moreau, E., “Electric Wind Produced by aSingle Dielectric Barrier Discharge Actuator Operating in AtmosphericFlight Conditions—Pressure Outcome,” AIAA Paper 2008-3792,June 2008.

[13] Takagaki, M., Isono, S., Nagai, H., and Asai, K., “Evaluation of PlasmaActuator Performance in Martian Atmosphere for Applications to MarsAirplanes,” AIAA Paper 2008-3762, June 2008.

[14] Abe,T., Takizawa,Y., andSato, S., “Experimental Study forMomentumTransfer in a Dielectric Barrier Discharge Plasma Actuator,” AIAA

Journal, Vol. 46, No. 9, 2008, pp. 2248–2256.doi:10.2514/1.30985

[15] Valerioti, J. A., and Corke, T. C., “Pressure Dependence of DielectricBarrier Discharge Plasma Flow Actuators,” AIAA Journal, Vol. 50,No. 7, 2012, pp. 1490–1502.doi:10.2514/1.J051194

[16] Paschen, F., “Ueber die zum Funkenübergang in Luft, Wasserstoff undKohlensäure bei Verschiedenen Drucken Erforderliche Potentialdiffer-enz,” Annalen der Physik, Vol. 273, No. 5, 1889, pp. 69–96.doi:10.1002/andp.18892730505

[17] Raju, G. G., Dielectrics in Electric Fields, Marcel Dekker, New York,NY, 2003, pp. 416–425.

[18] Pavon, S., Dorier, J.-L., Hollenstein, C., Ott, P., andLeyland, P., “Effectsof High-Speed Airflows on a Surface Dielectric BarrierDischarge,” Journal of Physics D: Applied Physics, Vol. 40, No. 6,2007, pp. 1733–1741.doi:10.1088/0022-3727/40/6/021

[19] Kriegseis, J., Grundmann, S., and Tropea, C., “Airflow Influence on theDischarge Performance of Dielectric Barrier Discharge PlasmaActuators,” Physics of Plasmas, Vol. 19, No. 7, 2012, p. 073509.doi:10.1063/1.4736995

[20] Barckmann, K., Kriegseis, J., Grundmann, S., and Tropea, C.,“Dielectric-Barrier Discharge Plasmas for FlowControl at HigherMachNumbers,” AIAA Paper 2010-4258, June–July 2010.

[21] Kriegseis, J., Grundmann, S., and Tropea, C., “Power Consumption,Discharge Capacitance and Light Emission as Measures for ThrustProduction of Dielectric Barrier Discharge Plasma Actuators,” Journalof Applied Physics, Vol. 110, No. 1, 2011, p. 013305.doi:10.1063/1.3603030

[22] Kriegseis, J., Möller, B., Grundmann, S., and Tropea, C., “Capacitanceand Power Consumption Quantification of Dielectric Barrier Discharge(DBD) Plasma Actuators,” Journal of Electrostatics, Vol. 69, No. 4,2011, pp. 302–312.doi:10.1016/j.elstat.2011.04.007.

[23] Kriegseis, J., Schröter, D., Grundmann, S., and Tropea, C., “Online-Characterization of Dielectric Barrier Discharge Plasma Actuators forOptimized Efficiency of Aerodynamical Flow Control Applications,”Journal of Physics: Conference Series, Vol. 301, No. 1, 2011,p. 012020.doi:10.1088/1742-6596/301/1/012020

[24] GBS ELEKTRONIK, “Special Pulse Generator Solutions,” http://www.gbs-elektronik.de/en/services/hv-pulse-technology/special-pulse-generator-solutions [retrieved 21 Jan. 2013].

[25] Bénard, N., Balcon, N., and Moreau, E., “Electric Wind Produced by aSurface Dielectric Barrier Discharge Operating in Air at DifferentPressures: Aeronautical Control Insights,” Journal of Physics D:

Applied Physics, Vol. 41, No. 4, 2008, p. 042002.doi:10.1088/0022-3727/41/4/042002.

[26] Savitzky, A., and Golay, M. J. E., “Smoothing and Differentiation ofData by Simplified Least Squares Procedures,” Analytical Chemistry,

966 KRIEGSEIS ETAL.

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

ICH

IGA

N o

n A

pril

3, 2

013

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

2159

Page 7: Closed-Loop Performance Control of Dielectric-Barrier-Discharge Plasma Actuators

Vol. 36, No. 8, 1964, pp. 1627–1639.doi:10.1021/ac60214a047

[27] National Instruments, “Programming Strategies for MulticoreProcessing: Pipelining,” Multicore Programming FundamentalsWhitepaper Series, http://zone.ni.com/devzone/cda/tut/p/id/6425 [re-trieved 21 Jan. 2013].

[28] Ziegler, J., and Nichols, N., “Optimum Settings for AutomaticControllers,” Transactions of the ASME, Vol. 64, No. 8, 1942,pp. 759–768.

[29] Åström, K. J., and Hägglund, T., PID Controllers: Theory, Design and

Tuning, Instrumentation, Systems, and Automation Society, ResearchTriangle Park, NC, 1995, pp. 11–24.

[30] National Instruments, “PID Control Toolkit User Manual,” http://www.ni.com/pdf/manuals/372192a.pdf [retrieved 21 Jan. 2013], pp. 3–6.

L. CattafestaAssociate Editor

KRIEGSEIS ETAL. 967

Dow

nloa

ded

by U

NIV

ER

SIT

Y O

F M

ICH

IGA

N o

n A

pril

3, 2

013

| http

://ar

c.ai

aa.o

rg |

DO

I: 1

0.25

14/1

.J05

2159


Recommended