You are on page 1of 16

Exp Fluids (2016) 57:160

DOI 10.1007/s00348-016-2242-5

RESEARCH ARTICLE

In‑flight active wave cancelation with delayed‑x‑LMS control


algorithm in a laminar boundary layer
Bernhard Simon1 · Nicolò Fabbiane2 · Timotheus Nemitz1 · Shervin Bagheri2 ·
Dan S. Henningson2 · Sven Grundmann3

Received: 31 March 2016 / Revised: 1 July 2016 / Accepted: 21 August 2016


© Springer-Verlag Berlin Heidelberg 2016

Abstract This manuscript demonstrates the first successful laminar flow (NLF) airfoils, active methods can be used to
application of the delayed-x-LMS (dxLMS) control algo- attenuate disturbances in the boundary layer and move the
rithm for TS-wave cancelation. Active wave cancelation of transition to turbulence further downstream. Predetermined
two-dimensional broadband Tollmien–Schlichting (TS) dis- flow control techniques (Gad-el Hak 2000), e.g., steady
turbances is performed with a single DBD plasma actuator. boundary-layer suction, have been investigated intensively
The experiments are conducted in flight on the pressure side but require relatively high actuation power levels because
of a laminar flow wing glove, mounted on a manned glider. the mean boundary-layer profile is altered. Reactive flow
The stability properties of the controller are investigated in control techniques such as active wave cancelation (AWC)
detail with experimental flight data, DNS and stability anal- only act on the fluctuations in the laminar boundary layer
ysis of the boundary layer. Finally, a model-free approach and therefore have the potential to require less energy but
for dxLMS operation is introduced to operate the control- have their own complexities.
ler as a ‘black-box’ system, which automatically adjusts the Early attempts of AWC with moving wall actuators
controller settings based on a group speed measurement of (Liepmann et al. 1982; Liepmann and Nosenchuck 1982;
the disturbance wave packets. The modified dxLMS control- Milling 1981; Thomas 1983) successfully attenuated the
ler is operated without a model and is able to adapt to vary- amplitude of artificially generated TS disturbances in a
ing conditions that may occur during flight in atmosphere. laminar boundary layer, while Sturzebecher et al. (2003)
showed the cancelation of natural occurring TS-waves on a
glider wing in flight. Motivated by the previous work with
1 Introduction moving wall actuators, Grundmann and Tropea (2008) first
showed that the plasma actuator (PA) is also able to cancel
Delaying laminar-turbulent transition of a boundary layer out TS-waves in a laminar boundary layer. Further stud-
has been a major topic of fluid mechanic research during ies with PAs in AWC mode by Kurz (2013) and Kotsonis
the last decades. Besides passive control techniques that (2013, 2015) suggest that PAs can be used to delay the
are related to the favorable pressure distribution of natural onset of transition with a modulation of the high voltage.
Several in-flight measurement projects have been con-
ducted with the PA at Technische Universität Darmstadt by
* Bernhard Simon Duchmann et al. (2014), Kurz et al. (2014) and Simon et al.
simon@sla.tu‑darmstadt.de (2015). Despite the low fluid-dynamic efficiency (Kriegseis
1 et al. 2013) and the limited body force of the PA, the actua-
Institute for Fluid Mechanics and Aerodynamics, Technische
Universität Darmstadt, Flughafenstraße 19, 64347 Griesheim, tor is well suited for research applications on flow control
Germany topics. The lack of moving parts and short response times
2
Linné FLOW Centre, KTH Mechanics, 10044 Stockholm, makes PAs attractive for many applications. Recent articles
Sweden by Wang et al. (2013) and Kriegseis (2016) show compre-
3
Department of Fluid Mechanics, University of Rostock, hensive reviews of PAs as flow control devices, in particu-
Rostock, Germany lar for low speed boundary-layer flow control.

13
160 Page 2 of 16 Exp Fluids (2016) 57:160

Model-based AWC algorithms showed promising


results in DNS as reported by Semeraro et al. (2013) and 2 2

Dadfar et al. (2013), whereas experiments were mainly


1 Laminar wing glove
conducted with adaptive control algorithms as the filtered- 2 Underwing pods for
x-LMS (fxLMS) algorithm (Sturzebecher et al. 2003; Kot- 1
measurement equipment
sonis 2013, 2015). The fxLMS is an extended version of 3 Environmental data
acquisition
the adaptive ’least mean square’ (LMS) algorithm (Elliott
3
and Nelson 1993) that compensates the distance and there-
fore the transfer path from the actuator to the error sen-
sor downstream with a digital filter. Fabbiane et al. (2015) Fig. 1  Grob G109b motor glider
and Simon et al. (2015) investigated the adaptivity of the
fxLMS algorithm and its limits. The required model of the
so-called secondary path, between actuator and error sen-
sor, should not change significantly during the experiment,
otherwise the algorithm becomes unstable.
For narrowband disturbances in laminar boundary lay-
ers, the dxLMS algorithm can be used as shown later in this
Fig. 2  Shape of the MW-166-39-44-43 airfoil
manuscript. Until now, the dxLMS algorithm has only been
applied to acoustic problems (Kuo and Morgan 1995; Kim
and Park 1998; Snyder and Hansen 1990; Hansen and Sny- weight of mMTOW = 950 kg. Due to flight speed limita-
der 1996), but the savings in computational power make its tions, the angle of attack exploitable for the experiments
application promising for future 3D application with multi- ranges between αmin = −3◦ and αmax = 13◦ which is close
ple PAs in spanwise direction for the delay of natural tran- to stall speed. The flight speed is around U∞ = 41 m/s for
sition. The reduction in the complexity of the secondary α = 2.5◦, and the actual value depends on the environmen-
path model to a simple delay of N samples combined with tal conditions at flying altitude. The chord-based Reynolds
a model-free approach for the determination of the required number of the wing glove is Re ≈ 3.75 × 106.
N leads to a promising ‘black-box’ system as shown in this The right wing of the glider can be equipped with a
manuscript for the first time. natural laminar flow (NLF) airfoil wing glove. The airfoil
This work presents a detailed study on the application of (Fig. 2) was developed by Weismüller (2012) to quantify the
the dxLMS algorithm for laminar flow control. It is struc- influence of atmospheric turbulence on NLF airfoils. Reeh
tured as follows: and Tropea (2015) intensively investigated the flow around
Section 2 provides an overview of the experimental the airfoil in flight. The flat airfoil shape on the pressure side
setup, while Sect. 3 presents the applied control theory. The creates an almost linear pressure gradient, which is adjusta-
numerical tools used in this work are described in Sect. 4, ble between moderate positive and negative values by chang-
and the in-flight measurements of the base flow as well ing α. The flow was shown to be 2D in the inner region of the
as the DNS results are shown in Sect. 5. Sections 6 and 7 wing glove by Reeh (2014). All base flow considerations and
deal with the dxLMS operation in flight and a model-free resulting pressure distributions are discussed later in Sect. 5.
approach for online-delay adaption, respectively. Conclud- The wing glove itself (Fig. 3) forms a square shape with
ing remarks summarize the study in Sect. 8. a chord length of l = 1.35 m and is slipped over the wing.
The pressure distribution is measured with 64 pressure taps
distributed over the suction and pressure sides of the wing
2 Experimental setup glove and acquired with 10 Hz by a Pressure Systems ESP-
64HD pressure transducer with a range of ±2490.82 Pa.
The in-flight experiments are conducted with a manned The reference atmospheric pressure p∞ is obtained using a
Grob G109b motorized glider, which combines the advan- boom protruding upstream into the flow.
tages of vibrationless gliding flight with a 96 kW engine An exchangeable plexiglass insert is arranged flush
for take off and altitude gain. The three-view drawing in mounted to the surface (see Fig. 3) and accommodates the
Fig. 1 shows the glider with a wing span of 17.4 m and a sensors and actuators. The disturbance source d is located
wing aspect ratio of 15.9. The sweep angle is close to zero, at xd /l = 0.18 and consists of 12 miniature loudspeakers
which leads to a negligible spanwise pressure gradient and installed underneath the measurement insert in spanwise
an almost two-dimensional flow along the chord. A tem- direction with a spacing of 20 mm. A circular array of six
porary ’permit to fly’ allows an additional payload for the 0.2-mm holes connects each speaker to the aerodynamic
measurement equipment and leads to a maximum take-off surface.

13
Exp Fluids (2016) 57:160 Page 3 of 16 160

wing glove LEADING EDGE socket for boom signals for the disturbance source and the high voltage gen-
DBD disturbance source erator. Data acquisition and the control algorithm (Sect. 3)
wing
actuator run at a sampling frequency of fS = 20 kHz.
hot-wire
sensors An environmental data acquisition setup mounted on the
left glider wing completes the in-flight testing setup (see
OUTBOARD

INBOARD
Fig. 1). The angle of attack α and the sideslip angle β are
storage
IR camera pod
measured with a Dornier Flight Log, which is a wind vane
plexi- heated insert
mounted on a second boom. In addition, the atmospheric
glass
insert pressure p∞, dynamic pressure q and temperature T∞ are
DBD controller measured upstream of the left wing. All ambient flow con-
pressure
taps
(covered) ditions are acquired with a sampling frequency of 1 kHz by
a NI6221 USB A/D converter, synchronized with the other
measurement data with a digital trigger.
Fig. 3  Sketch of the wing glove pressure side
All measurement equipment is operated by an onboard
24 V DC lead acid battery with a capacity of 16 Ah, which
allows one a system operation time of about 1 h. More
Three surface hot-wires probes p, r and e are located at details about the measurement procedure are given in
xp /l = 0.26, xr /l = 0.29 and xe /l = 0.38. The in-house Sect. 5.1.
manufactured hot-wire sensors are flush mounted to the sur-
face and consist of two needles molded in a plastic case with
a 1.25-mm-long and 5-μm-thin gold-plated tungsten wire 3 Control theory
welded on top. All hot-wires are operated by Dantec Min-
iCTA constant temperature anemometers. A signal condi- This section briefly explains the implemented control algo-
tioner filters the analog signals (first-order bandpass 50 Hz to rithms and provides an insight into the application of these
3.8 kHz) and amplifies the signal fluctuations by a factor of algorithms for active flow control in a laminar boundary
400. Further information about in-house made sensors can be layer. Further detailed information on the control theory
found in Simon et al. (2015). background can be found in several text books (Elliott and
The PA c is positioned flush mounted in a groove at Nelson 1993; Kuo and Morgan 1995).
xc /l = 0.33. It consists of a 10-mm-wide grounded lower
copper electrode of 35 µm thickness and a 5-mm-wide 3.1 Filtered‑x‑LMS control algorithm
upper electrode divided by five layers of polyimide tape
with a total thickness of 0.3 mm. The driving high voltage The flow control system sketched in Fig. 4 is a single-input-
signal is generated at a fixed frequency of fPA = 9.8 kHz single-output (SISO) system as it consists of one reference
by a GBS Minipuls 2.1, while the high voltage amplitude sensor r, an error sensor e and the actuator c. An adaptive
can be modulated with an analog input signal. The modu- feed-forward least mean squares (LMS) control algorithm
lation of the driving frequency enables the generation of which is applied adapts a digital finite impulse response
a variable volume force in time (Kurz 2013), which can (FIR) filter in order to minimize the signal at the error sen-
be divided into a stabilizing steady ’force offset’ and an sor. The filter output y(n) of a FIR filter w = [w0 ... wM−1 ]T
unsteady part, which is used for active wave cancelation. and an input signal x is given by
In spanwise direction, the PA is 230 mm long and a 2D
M−1
behavior is assumed in the following. The following geo- 
metrical distances between sensors and actuator have been y(n) = wi x(n − i). (1)
i=0
measured at the wing glove:
pr = 35 mm The variable n indicates the discrete-time step, and the
filter order is M = 256 for most examples presented in this
rc = 47 mm
manuscript. The filter w is adapted by the LMS control
ce = 76 mm. algorithm as indicated in Fig. 4 by the arrow. The adapta-
The dominant wave length of the TS-waves  is in the order tion is based on
of  = 50 mm.
A dSPACE digital signal processor, consisting of a w(n + 1) = w(n) + α r(n) e(n), (2)
DS1007 processor board, a DS2004 A/D board and a where α is the step size, which is set to α = 10−3,
and
DS2102 D/A board mounted in a dSPACE AutoBox r = [r(n) ... r(n − M + 1)]T a vector of the last M values
acquires the hot-wire signals and generates the input of the reference sensor signal.

13
160 Page 4 of 16 Exp Fluids (2016) 57:160

0.5

amplitude
wing surface
Hec
d 0
r c e
s δ
n cω
U∞

-0.5
r(n) 0 50 100 150 200 250

c(n) FIR coefficient


w(n)
adaptive
controller e(n) Fig. 5  Secondary path model Ĥec (dashed line) for fxLMS controller
r'(n) operation and corresponding delay of NP = 124 samples (solid line)
Ĥec(z) LMS for the dxLMS algorithm

Fig. 4  2D wing setup and extended fxLMS control algorithm


sketched below. The red triangles represent reference sensor r and behavior as long as the phase-angle error between the phys-
error sensor e. Disturbance source d and plasma actuator c are indi- ical secondary path Hec and the secondary path model Ĥec
cated as blue triangles is between ±90◦. This is a well-known boundary for the
controller stability (Elliott and Nelson 1993; Hansen and
Snyder 1996).
The transfer function from x to y is defined as Hyx
(to actuator/sensor y, from actuator/sensor x); therefore, 3.2 Delayed‑x‑LMS control algorithm
Her denotes the primary path (to e, from r), while Hec is
the secondary path (to e, from c). Thus, the secondary For a narrowband disturbance problem, such as TS-waves
path describes the transfer function between the plasma in a laminar boundary layer, a delay z−N with N sam-
actuator c and the error sensor e, while the primary path ples might be good enough to approximate a model for
describes the transmission behavior between reference the convective behavior of the secondary path. Figure 5
sensor r and error sensor e. The feedback path Hrc (to shows such a model for the secondary path Ĥec as well as
r, from c) is negligible in the presented flow control the approximation of Hec with a delay of NP = 124 sam-
system as the measured Tollmien–Schlichting distur- ples. The index P indicates a delay corresponding to the
bances in the laminar boundary layer only propagate global maximum (peak) of Ĥec, see Fig. 5. Li and Gaster
downstream. (2006) proposed a related approach for the simplification
The LMS control algorithm adapts the filter w , which of a transfer function for active flow control by simpli-
describes the behavior of the control path Hcr. The LMS fying the already calculated control kernel with a delay.
algorithm requires the same position for the actuator c and With the delay z−N , the LMS algorithm (Eq. 2) can be
the error sensor e, otherwise the phase-angle shift due to written as follows and is now called delayed-x-LMS
the secondary path Hec would lead to an unstable control- (dxLMS):
ler behavior (Elliott and Nelson 1993). The experimental
setup does not allow one to place the sensors very close to w(n + 1) = w(n) + α r(n) z−N e(n)
the plasma actuator due to the high voltage and geometrical (4)
= w(n) + α r(n − N) e(n).
concerns. This is often the case also for acoustic or struc-
tural vibration problems. Therefore, the physical second- Snyder and Hansen (1990) as well as Kim and Park
ary path Hec is modeled with another FIR filter Ĥec to filter (1998) applied the dxLMS algorithm for acoustic problems
the reference signal r(n) and compensate the phase shift. with narrowband noise. The experiments presented in this
An example for a filter Ĥec is presented in Fig. 5, which manuscript show its application for laminar boundary-layer
shows the convective behavior of the transmission path. flow control for the first time.
The filtered signal r ′ (n) is now fed into the LMS algorithm Figure 6 shows the dxLMS controller sketched below
as shown in Fig. 4, and the algorithm is now called filtered- the 2D wing setup. Besides the dxLMS algorithm, a second
x-LMS or fxLMS: upstream sensor p is introduced, which serves as an input
signal p(n) for the block C together with the reference sen-
w(n + 1) = w(n) + α r′ (n) e(n). (3) sor signal r(n). Block C is responsible for the identification
Former investigations by the authors (Fabbiane et al. of the delay N in Eq. 4. The delay N can be set manually
2015; Simon et al. 2015) showed a robust controller or calculated in real time by online measurements. In the

13
Exp Fluids (2016) 57:160 Page 5 of 16 160

wing surface where fS is the sampling frequency.


Hec
d p As the dSPACE digital signal processor executes the
r c e
s δ
n
algorithm at a rather low frequency of fS = 20 kHz , the
U∞ cω
phase-angle resolution δΦ(f ) = ffS 360◦ has to be taken into
account for the accuracy of the signal shift measurement
p(n) r(n) L. If the assumption of an almost constant average group
c(n) speed is correct, the signal shift L in samples can be nor-
w(n) malized by the distance between the sensors p and r ( pr )
adaptive
controller and a specific time delay γ is introduced as follows:
e(n)
r'(n)
z-N LMS L N fS
C N γ = = = .
pr ce cg (10)

Fig. 6  2D wing setup and dxLMS control algorithm sketched below. This is valid if the group speed is constant between the
The red square represents upstream sensor p; reference sensor r and sensors p and r: This assumption is verified later in the
error sensor e are shown as red triangles. Disturbance source d and
manuscript in Sect. 5.2.
plasma actuator c are indicated as blue triangles
Equation (10) indicates that the required delay N for the
dxLMS operation can then directly be calculated from a
following, the procedure for calculating the time delay N is
measurement of the signal shift L by
described.
N = γ × ce, (11)
3.2.1 Delay identification
where N is rounded to the closest integer. This link leads
to a ‘black-box’ system, which could operate the controller
The downstream propagating disturbance a is a superposi-
without any knowledge of the laminar boundary-layer flow
tion of traveling waves:
around the wing glove. Section 7 discusses the application
a(x, t) = A ei(ωt−kx) = A eiΦ , (5) of the dxLMS in real flight application and the advantages
in comparison with the well-known fxLMS approach.
where ω is the angular frequency, k the streamwise wave
The key to this reduction in the complexity of the sec-
number and Φ the phase angle of the wave. The growth in
ondary path is the similarity between the phase response
amplitude A is neglected, since it is not relevant for the fol-
given by the time delay and the actual phase response by a
lowing analysis. The phase speed cω is defined by cω = ωk ,
TS-wave. The phase response associated with a time delay
while group speed cg is defined as the derivative of ω with
reads:
respect to the modulus of the wave number k:
 
∂ω ◦N
�Φ = − 360 f. (12)
cg = . (6) fS
∂k
As mentioned before, a similar expression can be derived
Assuming a group speed cg, the cross-correlation func-
for the disturbances in the TS-wave band. A downstream
tion of two sensor signals at location x1 and x2 gives the
propagating wave in (Eq. 5) causes a phase-angle shift �Φ
time shift between both sensor signals τ :
 x2 of the sensor signals between the two different streamwise
τ (ω) =
1 positions. It can be shown that the derivative of the phase-
dx. (7)
x1 cg (ω) angle shift �Φ with respect to the angular frequency ω is
equal to minus the previously computed time delay τ:
If a parallel flow or a slowly varying boundary layer is
considered, the group speed is constant between the two ∂ ∂
(�Φ) = (−k �x)
locations (x2 − x1 = x). If a constant group speed is ∂ω ∂ω
assumed in the range of the amplified frequencies, Eq. 7 ∂k �x (13)
=− �x = − = −τ .
reduces to ∂ω cg
x2 − x1 �x Hence, by integrating ∂ω ∂
(�Φ) in this region, we obtain
τ (ω) ≈ τ = = . (8)
cg cg that the phase-angle shift (in deg) is given by:
This translates in a time-discrete delay or lag L between
 
◦ �x
�Φ(f ) = − 360 f + �Φ0 , (14)
the sensors p and r cg
�x where f = 2π ω is the frequency and �Φ0 ∈ [0◦ , 360◦ ) is
L = fS τ = f S , (9)
cg the zero-cross phase. By comparing Eqs. 12 and 14, the

13
160 Page 6 of 16 Exp Fluids (2016) 57:160

possible values of �Φ0 reduce to 0◦, if a positive gain is e(n)


used in front of the time delay, or 180◦, if a negative gain is
considered. Hence, once the time delay τ is identified via a r(n) w(n) c(n) Ĥec(z) -+ e‘(n)
measurement of the group speed cg, the only free parameter
for the dxLMS design is the positive/negative gain associ- Ĥec(z) r'(n) LMS
ated with the time delay.

3.3 Offline controller simulation (a) fxLMS.

The knowledge of all transmission paths of the flow control e(n)


problem makes an offline investigation of the fxLMS and
dxLMS controller behavior possible. Either the sensor sig- r(n) w(n) c(n) Ĥec(z) -+ e‘(n)
nals of a flight test or the DNS solution of a flow field can
describe the required paths between sensors and actuators. z-N r'(n) LMS
p(n) C N
This offline controller simulation in MATLAB/Simulink
saves measurement time and allows a detailed preparation
for the flight tests, where time is valuable. In the following, (b) dxLMS.
the controller simulation based on flight data and DNS is
discussed. Fig. 7  Block diagram for controller simulation with recorded flight
data
For the identification of the secondary path model Ĥec
during an experiment, the Minipuls high voltage amplifier
is operated with a pseudo-random binary signal (PRBS). Ĥed(z) e(n)
This results in a modulated high voltage signal at the PA,
d(n) Ĥrd(z) r(n) w(n) c(n) Ĥec(z) -+ e‘(n)
which creates broadband TS-waves traveling downstream.
The transmission behavior of Hec can be either estimated
z-N r'(n) LMS
C N
online with a LMS or calculated offline with the power Ĥpd(z) p(n)

spectral density Scc (Ω) and the cross-spectral density


Sec (Ω) (Elliott and Nelson 1993): Fig. 8  Block diagram for controller simulation with transfer paths
obtained with DNS
1  ∗ 
Scc (Ω) = lim √ CT (Ω)CT (Ω) (15)
T →∞ 2T

simulated error sensor signal e′ (n) is the second input for


1  ∗
the LMS adaptation algorithm. It is obvious that this pro-

Sec (Ω) = lim √ ET (Ω)CT (Ω) . (16)
T →∞ 2T
cedure does not completely describe the flight conditions
The transmission behavior Ĥec (Ω) is defined as follows: because the estimated paths Ĥec in the upper and lower row
in Fig. 7a are exactly the same. In a real case, the physical
Sec
Ĥec (Ω) = . (17) transmission path Hec changes due to varying inflow condi-
Scc
tions and can even cause an unstable controller behavior as
The magnitude and phase information, which is con- discussed later in Sect. 7 and shown by Simon et al. (2015).
tained in Ĥec (Ω) , can then be converted to a digital FIR Figure 7b shows the block diagram for the dxLMS
filter Ĥec (z), e.g., with the invfreqz command in MAT- offline controller simulation with recorded flight data. The
LAB. For offline controller simulation Ĥec (z), a record of only difference to the fxLMS approach in Fig. 7a is the fil-
a flight case with only the disturbance source in operation tering of r(n) and the online adaptation of the delay N as
is required (uncontrolled case). The discrete reference and described earlier in this section.
error sensor signals r(n) and e(n) are the input signals for Direct numerical simulations (DNS) allow one to extract
the fxLMS offline controller simulation as indicated in detailed information of the boundary-layer transmission
Fig. 7a. The incoming reference sensor signal r(n) is there- behavior. For the given problem, the FIR filter coefficients
fore the same as if the controller would be operated dur- of the transfer path models Ĥpd, Ĥrd, Ĥed and Ĥec are avail-
ing flight and the filtering of r(n) with Ĥec is exactly the able and implemented in the controller simulation as shown
same as in flight. Because the controller actuation is only in Fig. 8. Analogous to the in-flight data controller simula-
performed offline, the physical transmission path Hec has tion, the reference sensor filtering can be changed for the
to be modeled and c(n) is filtered by Ĥec. The result is then fxLMS controller simulation, see Fig. 7. The incoming
subtracted from the measured ’desired’ signal e(n), and the disturbances are not given by in-flight sensor data, a white

13
Exp Fluids (2016) 57:160 Page 7 of 16 160

0.2
wall free-stream outflows
4.1 Linear simulations
0.1
Linear simulations are performed with respect to the steady
y/c (-)

0 solution. The boundary condition for the perturbation


-0.1 velocity is similar to the baseflow boundary conditions:
no-slip on the surface, homogeneous Dirichlet on the free-
-0.2 stream boundary and homogeneous outflows (i.e., pa = 0).
0 0.2 0.4 0.6 0.8 1
x/l (-) In addition, sponges are placed in front of each outflow
in order to avoid reflections: Both upper and lower sponges
Fig. 9  DNS mesh have the thickness �x/l = 0.05.
Input and output devices used in the experimental setup
are modeled in the numerical simulations.
noise signal d(n) is the only input signal required for the Surface hot wires are represented by a weighted average
DNS data controller simulation. of the wall-shear stress: A Gaussian function with variance
10−3 l is used as a weight function.
The disturbance introduced by the loudspeakers is mod-
4 Numerical tools eled by a train of synthetic vortices introduced at the same
location as the loudspeakers in the experimental setup,
A spectral-element method (SEM) code—Nek5000—is Bagheri et al. (2009).
used to perform two-dimensional (2D) direct numerical The plasma actuator, instead, is represented by a vol-
simulations (DNS) of the incompressible flow close to the ume forcing based on the experimental results by Kriegseis
wing (Fischer et al. 2008). The code is based on a discre- (2013) at the PA’s location (Fabbiane et al. 2015).
tization of the computational domain in spectral elements
(Patera 1984), and in each element the flow is approxi-
mated by 2D Legendre–Gauss–Lobatto polynomials up to 5 Base flow
degree N.
The Reynolds number Re = U∞ l
ν for the simulations is All test flights are conducted in gliding flight without
3.75 × 10 . The computational domain extends along the
6 engine, which requires a special in-flight testing procedure,
airfoil from x/l = 0.1 of its upper surface to x/l = 0.65 as explained in Sect. 5.1. Based on these experiments, the
on the lower side, see Fig. 9. The domain is discretized in base flow around the wing is characterized and the bound-
20000 2D spectral elements of order N = 12 on both direc- ary-layer properties are investigated in Sect. 5.2.
tions: Four hundred elements are distributed along the air-
foil surface and 50 along the wall normal direction with 5.1 In‑Flight testing procedure
cosine distribution to form a curvilinear grid. The wall-nor-
mal and wall-wise size of the elements goes, respectively, The in-flight tests presented in this paper enable boundary-
from a minimum of 4 × 10−5 and 7 × 10−4 close to the layer measurements in a realistic flight environment. It
wall to a maximum of 1 × 10−2 and 7 × 10−3 in the free does not, however, determine the performance parameters
stream. of the aircraft as flight testing is usually understood. None-
No-slip boundary condition is enforced on the airfoil theless, the measurements are related to flight mechanics,
surface. On the free-stream boundary, a Dirichlet bound- environmental conditions and flight planning—challenges
ary condition enforces the solution to the potential solution which only exist for measurements under flight conditions
of the flow, corrected for the presence of a boundary layer. (Joslin 1998). Despite the difficulties of in-flight measure-
The two outflow boundaries are treated according to Bryn- ments compared to wind tunnel tests, there are reasons for
jell-Rahkola (2015): On the boundary pressure and velocity experiments on a airplane wing in flight. The investiga-
are linked by the relation tions of Weismüller (2012) on the atmospheric turbulence
indicate the need for in-flight experiments as the turbu-
1 ∂un
− p = −pa , (18) lence intensity is very low. More importantly, isentropic
Re ∂n turbulence is achievable for high Reynolds numbers. In
where n is the boundary-normal direction and un the com- addition, acoustic disturbances make the investigation on
ponent of the velocity normal to the boundary. The pressure natural transition in the wind tunnel difficult as the ambi-
is indicated by p and the boundary pressure pa is computed ent noise can trigger laminar-turbulent transition. Last but
from the superposition of the potential flow and Falkner– not least, a flow control technology in flight is very close
Skan boundary-layer solution. to application.

13
160 Page 8 of 16 Exp Fluids (2016) 57:160

The measurement time, preparation, maintenance and


costs exceed the effort for wind tunnel experiments by far.
For the current setup, one test flight of 1.5 h includes only
20 min of measurement time, the rest of the time is spent
on cruise flight and altitude gain. During a measurement
flight, the glider climbs up to an altitude of 10,000 ft and
the pilot switches off the engine. As soon as the propeller
is turned away from the wind, the pilot starts to adjust the
angle of attack α according to the flight display in front of
him; an online visualization of the low-pass-filtered values
for α and the sideslip angle β assists the pilot in setting and
Fig. 10  Measured and calculated pressure distribution at α= 2.5◦.
maintaining the desired flight state. Flight mechanics cou-
The error bars indicate the standard deviation of the measured pres-
ple flight speed U∞ and α depending on the weight, required sure distribution, calculated over 20 s
lift, density of the air and other parameters and therefore
do not enable the same boundary conditions for each flight,
even during one single gliding flight. Numerical studies Due to the high Reynolds number, small displacement
based on linear stability analysis concerning these issues thickness and moderate α potential flow theory provides a
have been conducted by Duchmann (2012). They revealed good estimate of the pressure around the wing glove. The
a stronger sensitivity of the streamwise position of the natu- potential flow solution, calculated with a Morino method
rally occurring transition to changes of the angle of attack (Morino and Kuot 1974), is drawn as a dashed line in
than to a slowly drifting Reynolds number. Therefore, the Fig. 10. All experiments are conducted on the pressure side
active flow control experiments in gliding flight were con- of the wing glove. Therefore, the angle of attack is iterated
ducted for a constant angle of attack conditions instead of for the potential flow solution in order to match the pres-
constant Reynolds number condition. Nevertheless, only sure distribution and, more important, the pressure gradient
small Reynolds number fluctuations are observed in the pre- on the lower side.
sented data. The first measurement run can be recorded in In the following discussions, only the value for α meas-
gliding flight at 9000 ft, and the measurements have to be ured during the experiment in flight is given.
stopped at an altitude of about 3000 ft to turn back on the The potential flow solution serves as an initial condition
engine. The measurement time of one gliding flight is about for DNS calculations which aim at two points;
10 min, while typically 15 measurement runs of 20–40 s
are recorded. All experiments have been conducted early in 1. comparison between flight and numerical simulations;
the morning under calm conditions to avoid changes in α. 2. parameter studies under controlled conditions.
Unavoidable limitations to in-flight experiments are given
by traffic, inversion layers and battery power supply, which A comparison between numerical and experimental base-
operates the whole measurement equipment. flow is presented in Fig. 10. It shows a good agreement on
the pressure side.
5.2 Base flow characterization Active wave cancelation in the laminar boundary layer
requires a sufficiently low amplitude of the disturbances at
As discussed above, the angle of attack measurement with the actuator position such that the transition process is in
the Dornier Flight Log only gives a rough estimate of the the linear regime. For the in-flight experiments, this leads
in-flow conditions as the systematical error due to position- to a range of 1.5◦ < α < 3.5◦ in which the experiments
ing of the probe is not taken into account. Therefore, the with artificially introduced 2D disturbances are reasonable.
pressure distribution around the wing glove as well as the Lower values for α would lead to interaction of the artificial
dynamic pressure q measured upstream of the wing are the 2D waves with natural 3D disturbances. Higher values for
reference values for comparison with numerical investiga- α would lead to more stable laminar boundary layer at the
tions. Figure 10 shows the averaged pressure distribution disturbance source position and no amplification of the arti-
for α = 2.5◦ measured in flight. The error bars indicate the ficially introduced disturbances. The pressure distributions
standard deviation caused by inflow fluctuations. The fluc- for the considered range are shown in Fig. 11. It is obvi-
tuations are small because the flights are conducted early ous that the pressure gradient in this region can be altered
in the morning. The boundary layer is tripped artificially from slightly negative for α = 3.5◦ to slightly positive for
on the suction side of the wing glove by surface roughness, α = 1.5◦.
which can be seen in the pressure distribution as a rising The boundary-layer profiles extracted from DNS calcu-
pressure at x/l = 0.3. lations enable investigations on the linear stability of the

13
Exp Fluids (2016) 57:160 Page 9 of 16 160

10
α = 1.5° (exp.)
-1 1400
α = 1.5° (DNS)
8
α = 2.5° (exp.) 1200
-0.5 α = 2.5° (DNS)

f (Hz)
1000 6
α = 3.5° (exp.)

N (-)
α = 3.5° (DNS)
cp

0 800 4

600 2
0.5
400
0
0.1 0.2 0.3 0.4 0.5 0.6
1
0 0.2 0.4 0.6 0.8 1 x/l (-)
x/l
(a) α = 2◦ .

Fig. 11  Measured and calculated pressure distribution for different 1400


10

angles of attack 8
1200

f (Hz)
1000 6

N (-)
800 4
boundary layer in the considered range of flight states.
600
Three neutral stability curves and N-factor contour plots, 2
400
calculated with PSE (Juniper et al. 2014), are shown in 0
0.1 0.2 0.3 0.4 0.5 0.6
Fig. 12. The sensor positions are indicated as triangles x/l (-)
(see Fig. 6) and the disturbance source d with a . Because
(b) α = 2.5◦ .
the experiments are conducted on the pressure side of the
10
wing glove, the pressure gradient is decreasing for higher 1400
α and therefore stabilizes the laminar boundary layer. This 1200
8

results in a downstream-shifted neutral stability curve


f (Hz)

1000 6

N (-)
(αi = 0) and a slightly shifted amplified frequency band. 800 4
The flat shape of the middle region of the airfoil (Fig. 2) 600
2
leads to rather low amplification transition scenarios but 400
also a significant movement of the transition region for a 0.1 0.2 0.3 0.4 0.5 0.6
0

change of α which was shown by Reeh (2014) for the same x/l (-)
airfoil in flight. For all three presented cases, the actua- (c) α = 3◦ .
tion with the PA ( is active in a low amplified region
(

where N ≈ 4. The low amplification rates are necessary to


Fig. 12  Neutral stability curve (αi = 0, black solid line) and N-factor
introduce artificial 2D waves which dominate the natural contours for different α and Re = 3.75 × 106. The dashed black line
disturbances. Future experiments on wave cancelation of shows the most amplified frequency (αi = min), while the dashed
natural disturbances with distributed actuators would not white line indicates the local N-factor maximum Nmax. The triangles
require such a low amplification because no artificially below indicate the sensor and actuator positions (see Fig. 6)
created 2D wave front is present.
Based on the PSE results, the propagation behavior of
the disturbances can be investigated more in detail. Fig-
0.46
ure 13 shows the non-dimensionalized phase speed cω and α = 1.5° : phase-speed

group speed cg for the local maximum N-factor Nmax and 0.44 α = 1.5° : group-speed
cg /U∞ (-)

1.5◦ ≤ α ≤ 3◦ . The PSE results are the basis for the calcu- 0.42 α = 2.0° : phase-speed

lation of the phase speed cω. 0.4


α = 2.0° : group-speed

α = 2.5° : phase-speed
As indicated in Fig. 13, the phase speed cω and the group 0.38
α = 2.5° : group-speed
speed cg slow down about 5 % in the region between the
cω /U∞ (-)

0.36 α = 3.0° : phase-speed


sensors p ( and e ( . The consequences of a changing
( (
0.34 α = 3.0° : group-speed
group speed cg for the specific time delay γ Eq. 10 are dis- α = 3.5° : phase-speed
0.32
cussed later in Sect. 7. α = 3.5° : group-speed
The flow control setup is a SISO system as mentioned 0.3
0.2 0.3 0.4 0.5 0.6
before in Sect. 3. Since only one actuator is mounted in x/l (-)
spanwise direction, a 2D wave front is required for the
experiment. Natural TS-waves occur in wave packets with Fig. 13  Non-dimensionalized phase speed cω /U∞ and group speed
a modulation in spanwise direction (Peltzer et al. 2009; cg /U∞ of the disturbances at the frequency for Nmax. The triangles
Saric et al. 2002). Therefore, artificial disturbances with below indicate the sensor and actuator positions (see Fig. 6)

13
160 Page 10 of 16 Exp Fluids (2016) 57:160

-20 -30
α = 2.5° , natural disturbances
uncontrolled
-30 α = 2.5° , disturbance source on -35

amplitude (dB)
fxLMS
amplitude (dB)

α = 3° , natural disturbances
-40 α = 3° , disturbance source on -40
-50
-45
-60
-50
-70
-55
-80
200 400 600 800 1000 1200 1400 1600 1800 2000 -60
400 500 600 700 800 900 1000 1100 1200
f (Hz)
f (Hz)

Fig. 14  Power density spectra of the error sensor signal e(t) for natu- (a) fxLMS.
ral and artificially induced TS-wave disturbances
-30
uncontrolled
-35

amplitude (dB)
dxLMS
an even wave front are created by the disturbance source. -40
Figure 14 shows the power spectra of the reference sensor
-45
signal r(t) for the natural and artificial case (disturbance
-50
source on). The artificially created 2D waves show ampli-
tudes, which are 12–15 dB higher compared to the natural -55

case, depending on the angle of attack α. This emphasizes -60


400 500 600 700 800 900 1000 1100 1200
that the naturally occurring 3D waves can be neglected f (Hz)
and a 2D problem is present for the experiments. Limited
space for sensors and measurement time causes the lack (b) dxLMS.
of measurements to proof the 2D character of the wave
front as it was conducted by the authors in wind tunnel Fig. 15  Control success of the active wave cancelation for fxLMS
and dxLMS (NP = 128) algorithm at α = 3◦. The spectra show the
experiments (Fabbiane et al. 2015). The speakers are posi- signal reduction at the error sensor e
tioned directly beneath the surface, and the resulting small
dead volume leads to the assumption that the speakers
create a 2D wave. The steady part of the PA force alters -30 uncontrolled
the boundary-layer profile and therefore also the stability -35 fxLMS
amplitude (dB)

dxLMS (N=118)
properties of the boundary layer (Duchmann et al. 2014). -40 dxLMS (N=115)

For the present case, the high flight speed in combination -45
with the rather low steady forcing of the PA leads to the -50
assumption of a negligible effect of the boundary-layer
-55
stabilization. Changing flow conditions during flight do
-60
not allow one to extract this small effect for the low actua- 400 500 600 700 800 900 1000 1100 1200
tor control authority from the recorded data, see Sect 5.1. f (Hz)

Fig. 16  Controller simulation (offline) of fxLMS and dxLMS for the


6 Controller operation in flight same test case. The transmission path is recorded at α = 2.5◦ and the
dxLMS algorithm operated with delays of NP = 118 and N = 115

Several experiments with successful TS-wave cancelation


of broadband disturbances with adaptive fxLMS algorithms digital bandpass filter (400–1100 Hz) is implemented in the
have been reported in the literature (Fabbiane et al. 2015; Simulink model for all sensor signals. The filter is neces-
Simon et al. 2015; Sturzebecher et al. 2003). The successful sary because the dxLMS controller adaptation can become
application of the fxLMS controller for U∞= 40.3 m/s and unstable for low-frequency disturbances due to a phase-
α = 3◦ with a broadband disturbance is shown in Fig. 15a. angle error, which is discussed later in this section.
The disturbances at the error sensor e, which range from For direct comparison of both controller concepts,
about 500–1100 Hz, are almost completely damped. the boundary conditions have to be the same. Wind tun-
The dxLMS control algorithm works as good as the nel experiments can fulfill this requirement but for the
fxLMS algorithm even with this very simplified model of flight test case an offline simulation of the controller
the transmission behavior of the secondary path Hec. Fig- behavior is necessary, see Sect. 3.3. The physical sec-
ure 15b shows a amplitude reduction in 12–15 dB, if the ondary path Hec is modeled with a FIR filter Ĥec. Fig-
controller is operated. For dxLMS controller operation, a ure 16 shows the offline simulated controller behavior for

13
Exp Fluids (2016) 57:160 Page 11 of 16 160

exactly the same boundary conditions. The direct com- -500


phase response filter
parison shows that both controllers perform equally well phase response filter +-90deg

while small deviations are caused by an adaptation of the -1000


phase response delay (N=118)
phase response delay (N=115)
LMS algorithm. It should be noted here that the dxLMS frequency response filter

Φ (deg)
controller shows almost exactly the same signal reduc- -1500
tion for NP = 118 samples compared to a shorter delay of
N = 115 samples. The delay N = 115 samples is obtained -2000
with the specific time delay approach (11) and discussed
more in detail later in this manuscript in Sect. 7.
-2500
The LMS adaptation algorithm estimates the control 400 500 600 700 800 900 1000 1100 1200
path Hcr with a FIR filter w, which describes the transmis- f (Hz)
sion behavior from the reference sensor r to the PA c. The
filter w is also known as ‘kernel’ and is shown in Fig. 17 Fig. 18  Phase response of the secondary path model filter Ĥec and a
for the offline simulated cases presented in Fig. 16. The delay of NP = 118 for dxLMS operation for the test case shown in
Fig. 16
FIR filters differ only for the first twenty coefficients, but
the most characteristic peaks, which are responsible for a
successful TS-wave damping (Fabbiane et al. 2015), match indicated in Fig. 18. The dxLMS controller requires an
perfectly. additional bandpass filter to avoid an unstable control-
The control success with the dxLMS algorithm can ler adaptation caused by disturbances below the TS-wave
be explained by looking at the phase response of the frequency band, which are not amplified by the boundary
transmission paths. As shown by Simon et al. (2015), layer but can exceed the ±90◦ boundary.
the fxLMS controller performance is almost constant for The in-flight measurements are limited to the measure-
the stable controller parameters but decreases abruptly ment with un-calibrated surface hot-wires. A measurement
if ±90◦ phase-angle error of the secondary path model of the control success in terms of disturbance amplitude
is exceeded. The solid blue line in Fig. 18 shows the growth along the streamwise direction is not possible. The
phase response of the secondary path model filter Ĥec , DNS results allow one to evaluate the control success from
while the dotted lines indicate the ±90◦ boundary. The the flow field. Figure 19a shows the error sensor spectra
phase response of a delay is a straight line with the slope for the uncontrolled case as well as for dxLMS and fxLMS
∂Φ 360◦ N
∂f = − fS and for the presented case with NP = 118 operation. The fxLMS controller reaches about 15 dB reduc-
the phase response lies in-between the ±90◦ bound- tion in the amplified region, while the dxLMS control algo-
ary for the amplified region of the boundary layer. The rithm performs equally well for f < 900 Hz but slightly
frequency response of Ĥec is plotted as a black solid better for f > 900 Hz, which is consistent with the experi-
line and marks the important region below the TS-wave mental results in Fig. 16. The converged kernels w for both
‘hump’ where the disturbances in the boundary layer are control approaches (Fig. 19b) match well but the dxLMS
amplified. For the optimal delay of NP = 118, the phase kernel differs slightly, similar to the experimental case shown
responses of Ĥec and the delay meet at the most amplified in Fig. 17.
frequency f ≈ 720 Hz , whereas the curve for N = 115 An evaluation of the whole flow control approach is not
intersect above the amplified band at f ≈ 1030 Hz. Both possible only by analyzing the error sensor signals because
lie in-between the ±90◦ boundary for the whole band as it includes only the performance at one point. The flow
field obtained by DNS is the basis for the prediction of the
perturbation energy evolution for both control approaches,
0.015 dxLMS and fxLMS. An integral value for the perturbation
fxLMS
0.01 dxLMS (N=118) amplitude at a certain streamwise location is defined as
dxLMS (N=115)
follows:
amplitude

0.005

0
 ∞
A(x) = �(u′ (x, y))2 �dy. (19)
-0.005 y=0
-0.01
Figure 19c shows the streamwise evolution of A. The
-0.015
amplitude drops down at the actuator position PA ( and
(
0 50 100 150 200 250
FIR coefficient grows again. The DNS results demonstrate that e(t) is a
good measure for the control success and the wave cancela-
Fig. 17  Kernel w of the dxLMS and fxLMS controller for the test tion has a sustained effect on the disturbance amplitude
case shown in Fig. 16 downstream.

13
160 Page 12 of 16 Exp Fluids (2016) 57:160

uncontrolled 0.6
-40 α = 1.5° U∞ = 43.1 m/s
fxLMS 0.4
dxLMS α = 2° U∞ = 42.8 m/s
0.2
amplitude (dB)

-50

amplitude
α = 2.5° U∞ = 42 m/s
0 α = 3° U∞ = 41.6 m/s
-60 α = 3.5° U∞ = 40.5 m/s
-0.2
α = 2.5° U∞ = 37.2 m/s
-70 -0.4
-0.6
-80 60 80 100 120 140 160 180 200 220
400 500 600 700 800 900 1000 1100 1200 FIR coefficient
f (Hz)

(a) Power density spectra of the error sensor signal. Fig. 20  Secondary path model Ĥec for different angles of attack and
flight altitudes
fxLMS
dxLMS
5
online adaptation during the experiment. The black curve
amplitude

(α = 2.5◦ , U∞ = 37.2 ms) in Fig. 20 shows that such a pre-


0
calibration of the system is not possible. Because of flight
mechanics, the flow around the wing changes during flight.
-5 All transfer functions but the black curve are recorded at a
flight altitude of about 8000 ft, while the black curve is
0 2 4 6 8 10
t (ms)
recorded at 3000 ft during the same gliding flight. Even if the
mass of the glider (wing surface area S) and required lift FL
(b) Kernel w.
remain the same, the air density ρ changes with altitude, and
therefore, the flight speed U∞ changes for a constant α (or lift
coefficient CL):
ρ 2
FL = CL × U S. (20)
2 ∞
The increased density ρ at the lower altitude leads to a lower
flight speed U∞ and therefore to a lower group speed cg of
the Tollmien–Schlichting waves, which is visible in a shift to
the right in Fig. 20. The amplification of the disturbances is
mainly influenced by the pressure gradient or α, respectively.
(c) Perturbation energy development.
This is why the amplitude of both measurements for α = 2.5◦
remains constant. A pre-calibrated system would lead to an
Fig. 19  Comparison between fxLMS and dxLMS by linear DNS for
α = 2.5◦ and Re = 3.75 × 106. Data are normalized with respect to unstable controller behavior, if the phase-angle shift because
the introduced disturbance of the decreased group speed is higher than 90◦.
Due to the changing boundary conditions, the flight
data are not ideal for a parametric study of α and U∞. The
6.1 Variable flow conditions in flight DNS calculations shown in Fig. 21 enable to investigate
the influence of each parameter independently. A variation
The stability and robustness of the fxLMS and dxLMS con- of α in Fig. 21a is associated with an increasing amplifi-
trol algorithm are mainly dependent on the secondary path cation for lower α , but also the shape of the filter Ĥec is
model Ĥec and the delay N. The corresponding FIR filters shifted to the left because of the increased group speed
are shown in Fig. 20 for different angles of attack α and cg, see Fig. 13. The flow speed U∞ is varied by ±10 % at
flight speeds U∞. It is obvious that the curves of Ĥec are α = 2.5◦, and the resulting secondary path model Ĥec is
stretched for higher angles of attack α and the amplitude is shown in Fig. 21b. The shift of the curves is more signifi-
higher for lower α. cant for the change of U∞ compared to α , but the ampli-
Simon et al. (2015) introduced scaling and stretching fac- fication is almost not influenced by U∞. Compared to the
tors for a model adaptation and extended the stable operation experimental values, the DNS results do not match exactly
range in a wind tunnel experiment for the fxLMS control- because α and U∞ are coupled in-flight, but the observa-
ler. This was done by a system identification and controller tions of the parameter study can also be made with the
calibration for a certain range of wind tunnel speeds and an experimental results in Fig. 20.

13
Exp Fluids (2016) 57:160 Page 13 of 16 160

0.04 2
L
p(t)
1.5 r(t)
α = 1.5°
amplitude (-)

0.02 α = 2° 1

amplitude (V)
α = 2.5°
α = 3° 0.5
0
α = 3.5°
0
-0.02 -0.5
-1
-0.04 -1.5
60 80 100 120 140 160 180 200 220 0 0.005 0.01 0.015 0.02 0.025
FIRcoefficient t (s)
(a) Variation of α.
Fig. 22  Upstream sensor signals p(t) and r(t). The lag sensor signal
0.04
p(t) is amplified by a factor of 2 for better comparison
U = 0.9
amplitude (-)

0.02 U = 0.95
U = 1.00
U = 1.05
0 U = 1.10 L
50
-0.02

amplitude
-0.04 0
60 80 100 120 140 160 180 200 220
FIR coefficient
-50
(b) Variation of the normalized flow speed U = U∞
U∞,ref
.
-250 -200 -150 -100 -50 0 50 100 150 200 250
samples
Fig. 21  Parameter study of the secondary path model Ĥec for param-
eters α and U∞, computed with DNS data
Fig. 23  Cross-correlation of two upstream sensor signals p(t) and
r(t), presented in Fig. 22

7 dxLMS stability and a model‑free approach


for online‑delay adaptation Compared to the optimal delay NP = 118samples , the
delay determined by cross-correlation is slightly underes-
The discussion of the varying flow conditions during flight timated. The change of the group speed cg leads to an error
in the last paragraphs shows the need for a model adapta- of N = −3samples , but it can be compensated by the
tion during operation of the LMS algorithm. In Sect. 3, the adaptive character of the LMS algorithm as shown earlier
calculation of the required delay N with an online-meas- in Fig. 16.
ured specific time delay γ was introduced (Eq. 11). The phase-angle resolution δΦ(f ) of a disturbance with
Figure 22 shows two upstream sensor signals p(t) and the frequency f is critical for the stability of the controller
r(t) extracted from the test case shown in Fig. 16. The char- because a phase-angle error for the secondary path model
acteristics of the signals are almost the same, while r(t) is Ĥec of less than ±90◦ is required as explained in Sect. 3.
slightly shifted to the right due to the sensor distance pr A sample rate of fS = 20 kHz leads to a phase-angle res-
and the group speed cg of the downstream traveling waves. olution δΦ = 14.4◦ for a disturbance with a frequency of
The shift between the signals, L, is extracted from the f = 800 Hz. Flow control at higher flow speeds requires
cross-correlation of both signals, presented in Fig. 23. A higher sampling rates. The model-free approach can only
number of 2000 samples were found to give reliable results work, if the phase-angle shift �Φ can be resolved. In addi-
for the online calculation on the dSPACE system, which tion, the ratio of the sensor distance pr compared to the dis-
implies an update of L every 0.1 s. The determined value tance ce influences the resolvable signal shift. For the given
of L = 53 samples can now be used to determine γ . For the example, the required delay N can only be determined with
presented example, this leads to a resolution of
L 53 samples ce
γ = = = 1.514 samples/mm. (21) δN = 1 sample = 2.17 samples. (23)
pr 35 mm pr
The required delay N can now be calculated based on γ : In conclusion, the sample rate and the upstream sen-
sor distance are most important for the success of the pre-
N = γ × ce = 115 samples. (22) sented model-free control approach. The positioning of the

13
160 Page 14 of 16 Exp Fluids (2016) 57:160

0.02
110
0.01

amplitude
100
N (-)

0
90 group speed impulse response
peak delay: group-speed
-0.01
cross-correlation delay: peak
80 envelope delay: cross-correlation
-0.02 delay: envelope
1.5 2 2.5 3 3.5
0 50 100 150 200 250
α (deg)
FIR coefficient

Fig. 24  Delay N for dxLMS operation, computed using group speed,


the (negative) peak of Hec and the cross-correlation or the envelope Fig. 25  Secondary paths for fxLMS operation and different delays
of the cross-correlation from p(t) and r(t), respectively. The data are for the dxLMS approach. The data are obtained via DNS simulations
obtained from the DNS results at α = 2.5◦

lag sensor p between r and c could provide a possibility to 6


×10-7
minimize the influence of a changing group speed in the cross-correlation
4 envelope
streamwise direction.

amplitude
2
The numerical simulation results allow one the compu-
tation of the time delay N based on different methods; 0

-2
–– group speed cg, based on the stability properties; -4
–– based on a peak in the secondary path model Ĥec; -6
–– correlation techniques, based on the sensor signals p(t) -200 -150 -100 -50 0 50 100 150 200
samples
and r(t)
Fig. 26  Cross-correlation of two upstream sensor signals p(t) and r(t)
Figure 24 shows a comparison between the listed meth- computed from DNS data at α = 2.5◦
ods for determining the required delay N for the dxLMS
algorithm. Based on the stability of the boundary layer, the
average group speed cg between PA (c) and e can be cal- 100
culated (see Fig. 13). Equations 8 and 9 then translate the
50
time shift τ with the sample rate fS to a delay N. The result-
(deg)

ing N increases with α because cg slows down due to the 0


delay: group speed
decreasing flight speed U∞. delay: peak
-50 delay: cross-correlation
Plotting the position of the delay in a secondary path delay: envelope
FIR filter curve (Fig. 25) shows an interesting fact: The -100 stability limit (±90° )

group speed is represented by the global minimum of the 400 500 600 700 800 900 1000 1100 1200
impulse response curve. A reason for this behavior is that f (Hz)
the peak (global minimum) of the impulse response is a
very good measurement of the center of the wave packet. Fig. 27  Phase error �Φe with respect to the secondary path Hec. The
A model-free dxLMS operation based on the cross-corre- data are obtained via DNS simulations at α = 2.5◦. A negative gain is
used in the time-delay approximations, i.e., �φ0 = π= 180◦, for the
lation of the upstream sensor signals p(t) and r(t) has been time delays N computed via group speed, (negative) peak of Hec and
introduced and works well for the data recorded in flight, the cross-correlation or the envelope of the cross-correlation from p(t)
see Fig. 23. However, for the cases calculated by DNS, the and r(t), respectively
dxLMS controller is unstable because the second positive
peak is lower in amplitude, compared to the experimen-
tally obtained curves in Fig. 20 which leads to a slightly speed cg and therefore the minimum peak in Fig. 25. As
changed phase response. already discussed earlier cg is decreasing with x and the
Another correlation technique is a time shift measure- envelope technique underestimates the delay slightly.
ment based on the envelope of the cross-correlation, see The most critical part which can be well investigated
Fig. 26. The maximum of the envelope is the lag L between with the DNS results is the phase-angle error �Φe between
the sensor signals p(t) and r(t). Due to the envelope tech- the secondary path model Ĥec and the actual secondary
nique, the corresponding delay N is now close to the group path Hec. Figure 27 shows �Φe for all four different delay

13
Exp Fluids (2016) 57:160 Page 15 of 16 160

approaches. The two methods based on the group speed cg The introduced model-free ‘black-box’ system suc-
and the (negative) peak lie on top of each other due to the cessfully works without any previous information about
same delay N. The curves are shifted by a zero-cross phase the environmental conditions and successfully cancels out
of �Φ0 = 180◦ because of the negative gain, described in TS-wave disturbances with the presented SISO system.
Eq. 14. In comparison with the group speed method, the An advanced method for dxLMS controller operation and
cross-correlation method fulfills the stability criteria of the determination of N, based on the group speed meas-
the LMS (±90◦ limit) only in a short band and does not urement, has been introduced and theoretically derived. In
match the slope of Hec well but crosses the phase response, addition, the dxLMS algorithm requires less computational
as already seen in the experimental results in Fig. 18. The power because a FIR filter is replaced by a simple delay
group speed and peak-based delay match the slope much N. With regard to future MIMO systems for active wave
better, which leads to a phase-angle error �Φe close to cancelation of naturally occurring 3D wave packets, the
zero for a wide range of the amplified frequency band, see dxLMS approach is a key development for less complex
Fig. 12b. A model-free dxLMS operation is more robust in controllers and lower requirements on the computational
terms of LMS controller stability with the envelope method power.
introduced above; the DNS results in Fig. 19 are obtained
by using this method. Acknowledgments This work was supported by the German research
foundation (DFG) under the Grant No.GR3524/4-1. Simulations have
For dxLMS controller operation, the zero-cross phase of been performed at the National Supercomputer Centre (NSC) with
�Φ0 = 180◦ can be realized by adjusting the control law computer time granted by the Swedish National Infrastructure for
(Eq. 4) with a negative gain (−1) as follows: Computing (SNIC). We also wish to thank Jens Rohlfing from Fraun-
hofer LBF (Darmstadt) for the fruitful discussions. Finally we appre-
w(n + 1) = w(n) + (−1)αr(n)z−N e(n) (24) ciate the support of our student and flight test pilot Tobias Hofmann.

The discussion above illustrates that particular care has


to be taken in identifying the center of the wave-packet
trace. A cross-correlation method may lead to an error in
References
the time delay or phase angle, respectively. Moreover, the Bagheri S, Brandt L, Henningson DS (2009) Input-output analysis,
error of the correlation is also due to the variation in the model reduction and control of the flat-plate boundary layer. J
group speed between p-r and c-e location: In this setup, Fluid Mech 620(1):263–298
this error does not compromise the stability of the control Brynjell-Rahkola M (2015) Global stability analysis of three-dimen-
sional boundary-layer flows. Lic. thesis, KTH, Stockholm
algorithm and the error reduces by increasing the angle of Dadfar R, Semeraro O, Hanifi A, Henningson DS (2013) Output feed-
attack, see Fig. 24. However, a small systematic error will back control of flow on a flat plate past a leading edge using
always be present. plasma actuators. AIAA J 51(9):2192–2207
Duchmann A (2012) Boundary-layer stabilization with dielectric bar-
rier discharge plasmas for free-flight application. Ph.D. thesis,
TU Darmstadt
8 Conclusions Duchmann A, Simon B, Tropea C, Grundmann S (2014) Dielectric
barrier discharge plasma actuators for in-flight transition delay.
The main focus of the work is the controller stability paired AIAA J 52(2):358–367
Elliott S, Nelson P (1993) Active noise control. Signal Process Mag
with a detailed investigation of the boundary-layer trans- IEEE 10(4):12–35
mission behavior, supported by DNS. The performance of Fabbiane N, Simon B, Fischer F, Grundmann S, Bagheri S, Henning-
the reliable fxLMS control algorithm and the newly intro- son DS (2015) On the role of adaptivity for Robust Laminar-flow
duced modified dxLMS control algorithm has been inves- control. J Fluid Mech 767:R1
Fischer PF, Lottes JS, Kerkemeier SG (2008) Nek5000 - open source
tigated for active wave cancelation in flight under realistic spectral element cfd solver. http://nek5000.mcs.anl.gov
atmospheric conditions. Grundmann S, Tropea C (2008) Active cancellation of artificially
Performance-wise the dxLMS and fxLMS control algo- introduced Tollmien–Schlichting waves using plasma actuators.
rithms work equally well, if the delay N is chosen in such Exp Fluid 44(5):795–806
Gad-el Hak M (2000) Flow control: passive, active, and reactive flow
a way that the phase-angle error �Φ ≤ ± 90◦. In-flight management. Cambridge University Press, Cambridge
measurements and DNS simulations showed that the group Hansen C, Snyder S (1996) Active control of noise and vibration.
speed of the TS-wave disturbances changes significantly CRC Press, Boca Raton
dependent on the environmental conditions and therefore Joslin R (1998) Aircraft laminar flow control 1. Ann Rev Fluid Mech
30(1):1–29
also, e.g., on the altitude. The resulting phase-angle error Juniper MP, Hanifi A, Theofilis V (2014) Modal stability theorylec-
�Φ therefore leads to an unstable controller behavior, if a ture notes from the FLOW-NORDITA summer school on
previously identified model of the boundary-layer transmis- advanced instability methods for complex flows, Stockholm,
sion is not valid any more. Sweden, 2013. Appl Mech Rev 66(2):024804

13
160 Page 16 of 16 Exp Fluids (2016) 57:160

Kim HS, Park Y (1998) Delayed-x lms algorithm: an efficient ANC Patera AT (1984) A spectral element method for fluid dynamics: lami-
algorithm utilizing robustness of cancellation path model. J nar flow in a channel expansion. J Comp Phy 54(3):468–488
Sound Vib 212(5):875–887 Peltzer I, Pätzold A, Nitsche W (2009) In-flight experiments for
Kotsonis M, Giepman R, Hulshoff S, Veldhuis L (2013) Numerical delaying laminar—turbulent transition on a laminar wing glove.
study of the control of Tollmien-Schlichting waves using plasma Proc Inst Mech Eng Part G J Aerosp Eng 223(6):619–626
actuators. AIAA J 51(10):2353–2364 Reeh AD (2014) Natural laminar flow airfoil behavior in cruise flight
Kotsonis M, Shukla RK, Pröbsting S (2015) Control of natural Tollm- through atmospheric turbulence. Ph.D. thesis, TU Darmstadt
ien-Schlichting waves using dielectric barrier discharge plasma Reeh AD, Tropea C (2015) Behaviour of a natural laminar flow aer-
actuators. Int J Flow Control 7(1–2):37–54 ofoil in flight through atmospheric turbulence. J Fluid Mech
Kriegseis J, Duchmann A, Tropea C, Grundmann S (2013) On the 767:394–429
classification of dielectric barrier discharge plasma actuators: Saric WS, Reed HL, Kerschen EJ (2002) Boundary-layer receptivity
a comprehensive performance evaluation study. J Appl Phy to freestream disturbances. Ann Rev Fluid Mech 34(1):291–319
114(5):053,301 Semeraro O, Bagheri S, Brandt L, Henningson DS (2013) Transition
Kriegseis J, Schwarz C, Tropea C, Grundmann S (2013) Velocity- delay in a boundary layer flow using active control. J Fluid Mech
information-based force-term estimation of dielectric-barrier dis- 731(9):288–311
charge plasma actuators. J Phy D Appl Phy 46(5):055,202 Simon B, Nemitz T, Rohlfing J, Fischer F, Mayer D, Grundmann S
Kriegseis J, Simon B, Grundmann S (2016) Towards in- fight applica- (2015) Active flow control of laminar boundary layers for vari-
tions? A review on dielectric barrier discharge-based boundary- able flow conditions. Int J Heat Fluid Flow 56:344–354
layer control. Appl Mech Rev 68(2) Simon B, Schnabel P, Grundmann S (2016) IR measurements for
Kuo SM, Morgan D (1995) Active noise control systems: algorithms quantification of laminar boundary layer stabilization with dbd
and DSP implementations. Wiley, Hoboken plasma actuators. In: New results in numerical and experimental
Kurz A, Goldin N, King R, Tropea C, Grundmann S (2013) Hybrid fluid mechanics X. Springer
transition control approach for plasma actuators. Exp Fluid Snyder S, Hansen C (1990) The influence of transducer transfer func-
54(11):1–4 tions and acoustic time delays on the implementation of the
Kurz A, Simon B, Tropea C, Grundmann S (2014) Active wave LMS algorithm in active noise control systems. J Sound Vib
cancelation using plasma actuators in flight. In: 52nd Aerospace 141(3):409–424
Sciences Meeting, AIAA SciTech. American Institute of Aero- Sturzebecher D, Nitsche W (2003) Active cancellation of Tollmien–
nautics and Astronautics Schlichting instabilities on a wing using multi-channel sensor
Li Y, Gaster M (2006) Active control of boundary-layer instabilities. J actuator systems. Int J Heat Fluid Flow 24:572–583
Fluid Mech 550:185–206 Thomas AS (1983) The control of boundary-layer transition using a
Liepmann H, Brown G, Nosenchuck D (1982) Control of lami- wave-superposition principle. J Fluid Mech 137:233–250
nar-instability waves using a new technique. J Fluid Mech Wang JJ, Choi KS, Feng LH, Jukes TN, Whalley RD (2013) Recent
118:187–200 developments in dbd plasma flow control. Prog Aerosp Sci
Liepmann H, Nosenchuck D (1982) Active control of laminar-turbu- 62:52–78
lent transition. J Fluid Mech 118:201–204 Weismüller (2012) A new approach to aerodynamic performance of
Milling RW (1981) Tollmien–schlichting wave cancellation. Phy aircraft under turbulent atmospheric conditions. Ph.D. Thesis,
Fluid (1958–1988) 24(5):979–981 TU Darmstadt
Morino L, Kuot CC (1974) Subsonic potential aerodynamics for com-
plex configurations: a general theory. AIAA J 12(2):191–197

13

You might also like