Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

13th AIAA/CEAS Aeroacoustics Conference (28th AIAA Aeroacoustics Conference) AIAA 2007-3540

An Active Control Scheme for Tuning


Acoustic Impedances

Jonas P. Moeck∗, Mirko R. Bothien∗ and Christian Oliver Paschereit†


Institute of Fluid Dynamics and Engineering Acoustics
Technical University Berlin, 10623 Berlin, Germany

Acoustic boundary conditions are a crucial component in various technical sys-


tems. Experimental investigations in laboratory test rigs often do not implement
the same acoustic characteristics present in the actual application, in particular if
only individual system components are studied. In this work, a method is proposed
which uses an active control scheme to mimic a certain prescribed acoustic reflec-
tion coefficient. The control concept is explained and results from simulations and
experiments are presented to demonstrate feasibility of the method. It is shown
that generic boundary conditions like fully reflecting (pressure and velocity node)
and fully absorbing can be realized. Moreover, it is possible to impose a virtual ad-
ditional length by manipulating the phase of the reflection coefficient. In this way,
the impact of modified acoustic resonance frequencies on the system can be studied.

Nomenclature

A B state space matrices Symbols


C D
c speed of sound α reflectivity
D wave decomposer ∆x microphone spacing
e actuator command λ wavelength
f, g Riemann invariants ω angular frequency
G actuator transfer function ρ fluid density
K controller transfer function τ time-delay

k wave number i −1
M Mach number
Subscripts
p scaled acoustic pressure
R reflection coefficient d downstream
S scattering matrix u upstream
u axial particle velocity cl closed-loop
u0 mean flow velocity exp experiment
Z impedance ol open-loop
here Superscripts
more
+ in mean flow direction
need
− against mean flow direction
some ˙
(·) derivative with respect to time
space ˆ
(·) Fourier transform
∗ PhD student.
† Professor, Experimental Fluid Dynamics.

1 of 14

Copyright © 2007 by the authors. Published by the AmericanAmerican Institute of


Institute of Aeronautics Aeronautics
and Astronautics, and
Inc., Astronautics
with permission.
I. Introduction
A. Background
lane wave acoustic fields in ducted geometries are an important component in various technical appli-
P cations. Thermoacoustic instabilities arise in combustion systems due to the presence of a low frequency
1
acoustic field that couples with the unsteady heat release. The acoustic boundary conditions are of eager
importance in this phenomenon since the reflection of acoustic waves at the boundaries represents a crucial
component in the thermoacoustic feedback loop. In muffler design, it is of great importance to model the
plane wave acoustic field accurately to have a proper prediction of the transmission loss.2 In general, when
considering damping characteristics of acoustic liners, the amount of dissipated sound power depends on the
incident sound field.
In the applications mentioned above, experimental work is of high practical value since analytical and
numerical work is often not capable to fully take into account the complete acoustic-flow and acoustic-
flow-combustion interaction. Yet, the experimental characterization of a (thermo-) acoustic system only
corresponds to a certain geometrical system set-up of the full device. As a matter of fact, laboratory
tests with the aim to assess acoustic characteristics of technical components do not use the full actual
geometrical set-up. For example, in gas turbine burner development, single burner test rigs are used to
investigate emissions and stability characteristics. The combustion chamber geometry in the full scale engine
is completely different, often being of annular type. It is apparent that the acoustic characteristics, as, e.g.,
the resonance frequencies or the damping can be essentially different in the two cases. From thermoacoustic
system analysis it is known, however, that combustion instabilities preferably occur at frequencies close to
the acoustic modes.1, 3 Also, the degree of reflectivity of acoustic waves at the combustor outlet has a strong
impact on thermoacoustic stability and the associated oscillation amplitudes.
When assessing the efficiency of acoustic dampers or liners, the amount of dissipated sound power typically
depends on the incident sound field. Although the minimum and maximum amount of dissipated sound power
can be determined from the scattering matrix of the acoustic element (see Aurégan and Starobinski4 ), the
actual amount for a given set-up will depend on the sound field (i.e., on the acoustic boundary conditions).
One factor typically causing different acoustic fields in the test bed and in the actual device is the
geometry. To investigate the impact of a change in geometry on the system characteristics, time-consuming
and expensive design changes have to be made. Physically, the impact of a geometry change on the plane wave
acoustic field can be represented by a change of the impedance at a certain position (see Fig. 1). Therefore,
instead of changing the actual geometry, the same effect could be achieved by changing the impedance at
some reference position.

B. Control Objective
The aim of the control scheme to be presented in this work is to expose a system to acoustic boundary
conditions which are different than the actual ones. In the following, the set-up depicted in Fig. 1 is
considered. The system is connected to a duct, in which only plane waves propagate, terminated with an
acoustic boundary condition Z1 . Here, Z denotes the acoustic impedance, a complex valued function of
frequency. The acoustic impedance is defined as

Z = p̂/û, (1)

where p̂ and û denote Fourier transforms of acoustic pressure (which is considered to be scaled by the
characteristic impedance ρc, ρ being fluid density and c the speed of sound) and particle velocity normal to

SYSTEM Z1 SYSTEM Z2 SYSTEM Z3

Figure 1. System with different acoustic boundary conditions induced by a change in geometry or by imple-
mentation of a liner

2 of 14

American Institute of Aeronautics and Astronautics


ˆ denotes the Fourier transform of a variable.
the boundary, respectively. Here, as in the following, (·)
The system considered in Fig. 1 has certain characteristics depending on the acoustic field associated
with Z1 . For other acoustic boundary conditions, e.g., induced by a change in geometry (middle in Fig. 1) or
by some damping device mounted in the duct (right in Fig. 1), the system may exhibit essentially different
characteristics. This can be due to higher or lower resonance frequencies or increased damping, for example.
The aim of the control scheme to be presented is to mimic certain prescribed impedances. In this way,
the impact of different acoustic boundary conditions on the system characteristics can be studied, without
actually implementing the devices that alter the boundary impedance.
Guicking & Karcher5 introduced the concept of manipulating the impedance through active control of
the acoustic boundary condition in 1984. They tuned a duct end with a speaker mounted to fully absorbing.
However, they only considered fixed frequency harmonic signals and set the control parameters empirically.
Furthermore, they only considered a duct without mean flow. Application of this concept was also presented
by Li et al.6 They used a similar method to reduce the resonator length of a thermoacoustic cooler.
The active control scheme that is presented here is able to mimic a broad class of boundary conditions
for plane acoustic waves over a wide range of frequencies in ducts with and without flow. The control scheme
consists of a sensor array, a linear controller and an acoustic actuator. In contrast to Guicking & Karcher’s
approach,5 there is no need for manual parameter tuning and the controller acts over a range of frequencies.
In particular, the scheme presented here will not only work for time-harmonic but also for transient signals.
As will be shown in the next sections, the sensor signals have to be processed in a suitable way so that the
acoustic field is decomposed into downstream and upstream propagating components. This is conveniently
expressed by using the Riemann invariants f and g. At a fixed axial location, they are related to the scaled
acoustic pressure p and the axial acoustic particle velocity u by

p = f + g, (2a)
u = f − g. (2b)

Note here that the linear relations expressed in Eqs. (2) hold in frequency as well as in time domain.
In terms of the Riemann invariants, the acoustic boundary condition is expressed as the reflection coef-
ficient, which represents the ratio of the reflected and the incident wave. The reflection coefficient is related
to the impedance by
Z −1
R= . (3)
Z +1
Therefore, equivalently to imposing a certain desired impedance, the control scheme will be designed so as
to mimic a certain reflection coefficient.

II. Control Design


To develop the control scheme, the set-up in Fig. 2 is considered. The duct, whose end-impedance is
to be manipulated, is equipped with a sensor array, consisting of several microphones at different axial
positions, and an actuator. To achieve the control objective, the sensor signals are fed to a control scheme
which generates the command signal and acts on the acoustic field via the actuator. It is more instructive
to consider the downstream propagating wave f as the actual control input. For this reason, the control

sensor array actuator


p f e
D K

Rcl f
SYSTEM
Rol g

Figure 2. System with controlled boundary conditions. The wave decomposer D extracts the downstream
propagating wave f from multiple pressure measurements p. Based on f , the controller K generates the
actuation command e. The uncontrolled reflection coefficient Rol is manipulated by the control scheme and
the actuator so that the system is exposed to a reflection coefficient given by Rcl

3 of 14

American Institute of Aeronautics and Astronautics


scheme depicted in Fig. 2 has been split in two parts: i) a wave decomposer, labeled D, that extracts the
downstream propagating wave from the pressure signals p; and ii) the actual control law K. Accordingly,
finding a wave decomposition scheme D that works accurately over a range of frequencies is necessary. A
method for time domain wave separation is presented in Sec. III. For the remainder of this part, the f -wave
is considered to be known so that it can serve as the input to the controller.
The problem now is to find a suitable K that, given the downstream propagating wave f , is able to drive
the actuator in such a way that the closed-loop reflection coefficient Rcl is close to the desired one. The
control law K can be built if the uncontrolled (or “open-loop”) reflection coefficient Rol and the actuator
transfer function (in the following denoted as G) are known, as will be shown below.
The actuated end element is modeled as a linear system with two inputs and one output. The inputs
are given by the incident wave f and the actuator command e and the output is the reflected wave g.
It is assumed here that the upstream traveling wave can be obtained from a linear superposition of the
uncontrolled passive reflection of the incident wave and the wave generation due to the actuator. Validity
of this assumption for the test rig used in the experiments was shown in.7, 8 The frequency domain model
for the actuated end element then reads
ĝ = Rol fˆ + Gê, (4)
where Rol is the uncontrolled reflection coefficient. Rol and G can both be obtained experimentally from
frequency response measurements in conjunction with the Multi-Microphone-Method (MMM,9, 10 see also
Sec. IV).
In the closed-loop case, in which the fˆ wave is fed back to the actuator via the control law K, the end
element’s ĝ response (see Eq.(4)) becomes

ĝ = Rol fˆ + GK fˆ. (5)

Consequently, the closed-loop reflection coefficient takes the form

Rcl = Rol + GK. (6)

If R and G are assumed to be known from the frequency response measurements, the control law K can be
formally calculated from Eq. (6) if a desired closed-loop reflection coefficient Rcl is prescribed:
Rcl − Rol
K= . (7)
G
In order to implement the control scheme on a digital control board, a model for the control law K has to
be found. In the present work, this is done by applying a frequency domain system identification algorithm
(see Gustavsen and Semlyen11 ) to the discrete frequency response data, which is obtained from Eq. (7).

III. Online Wave Decomposition


As the f wave is the control input, a suitable algorithm is required that separates the up- and downstream
traveling waves from the pressure sensor signals. In frequency domain, this can be accomplished by using the
well-known MMM.9, 10 However, in contrast to the works cited, the wave decomposition cannot be performed
here as a postprocessing step. The f wave has to be identified online and its calculation from the sound
pressure sensors has to be accomplished in time domain. With reference to Fig. 3 (left), the pressure at two
axial locations in the duct can be written as

p1 (t) = f (t) + g(t − τ − ), (8)


+
p2 (t) = f (t − τ ) + g(t), (9)

where τ ± = ∆x/(c ± u0 ). In these equations, g can be eliminated by evaluating Eq. (9) at t − τ − and using
the result in Eq. (8). This gives

p1 (t) = f (t) + p2 (t − τ − ) − f (t − τ − − τ + ), (10)

from which it is concluded that the f wave can be identified in real time by using the scheme shown in Fig. 3
(right).

4 of 14

American Institute of Aeronautics and Astronautics


p1 p2
p1
f f
+
− +
U0 p2
g t − τ−
Δx t − τ+ − τ−

Figure 3. Two-microphone scheme for online wave identification

One drawback of this approach is that the feedback loop in Fig. 3 (right) is only marginally stable. To
see this, Eq. (10) is solved for fˆ in frequency domain:

p̂1 − p̂2 e−iωτ
fˆ = . (11)
1 − eiω(τ − +τ + )
The purely real eigenvalues are located at
2πn
ωn = , n = 0, ±1, ±2... (12)
τ+ + τ−
Note that this relation can be recast in the form

kn ∆x = πn(1 − M 2 ), (13)

which was given, e.g., by Åbom and Bodén12 in the analysis of the Two-Microphone-Method (TMM) in
frequency domain. It is important to note that, as the above authors pointed out, not only are the frequencies
given in Eq. (12) excluded from data processing, but the error in the TMM increases dramatically when
approaching theses frequencies. Also, decreasing the microphone spacing to move ω1 to higher frequencies
is not an option since, in this case, the accuracy of the TMM significantly deteriorates at low frequencies.12
In practice, this means that the identification algorithm cannot be used close to frequencies for which
Eq. (12) holds. Moreover, oscillations can be excited that would have to be suppressed with suitable notch
filters.
p1 p2 pn-1 pn

f

U0
g
x1 x2 … xn-1 xn

Figure 4. Set-up for online wave identification with multiple microphones

In the frequency domain, this issue is resolved by using the Multi-Microphone-Method.9, 10 Here, several
microphones at different axial locations, corresponding to unequal propagation delays, are used. This resolves
the issue of singularities at frequencies given by Eq. (12). The wave decomposition is then obtained from a
least squares fit to the measured complex pressure amplitudes. Another advantage of the MMM is that by
averaging over several microphones, flow noise is, at least partially, suppressed. This is even more important
in a real-time application where noise rejection techniques, such as cross-correlations with the excitation
signal, cannot be used. However, the MMM is applied in a post-processing step, where pressure records for
a certain time interval are available. For the impedance tuning control scheme, the decomposition has to be
accomplished in real-time. Therefore, application of the traditional MMM is not straightforward. However, a
time domain scheme, having similar advantages like the MMM in frequency domain (i.e., using more pressure

5 of 14

American Institute of Aeronautics and Astronautics


signals to reject flow noise and removal of singular frequencies) can be constructed as follows. With respect
to the set-up shown in Fig. 4, n equations for the pressure signals are obtained as

p1 (t) = f (t) + g(t), (14a)


+ −
p2 (t) = f (t − τ1,2 ) + g(t + τ1,2 ), (14b)
..
.
+ −
pn (t) = f (t − τ1,n ) + g(t + τ1,n ), (14c)
±
where τn,m = (xm − xn )/(c ± u0 ). The aim is now to find a causal relation for f (t) that accounts for all n

pressure recordings. The pressure at location xk is written at time t − τ1,k , viz.,

p1 (t) = f (t) + g(t), (15a)


− + −
p2 (t − τ1,2 ) = f (t − τ1,2 − τ1,2 ) + g(t), (15b)
..
.
− + −
pn (t − τ1,n ) = f (t − τ1,n − τ1,n ) + g(t). (15c)

Equations (15) are now considered as an overdetermined linear system for f (t) and g(t), whose least squares
solution takes the form
 
p1 (t)
− + − 
" # " #
f (t) 1 −1/(n − 1) . . . −1/(n − 1)  p (t − τ1,2 ) − f (t − τ1,2 − τ1,2 )
=  2 .. . (16)
g(t) 0 −1/(n − 1) . . . −1/(n − 1)  
.


− + −
pn (t − τ1,n ) − f (t − τ1,n − τ1,n )

A feedback scheme similar to the one shown in Fig. 3 can be constructed, mapping the n pressures to
the downstream traveling wave. It is, however, more instructive to consider the frequency domain solution
for fˆ, which is obtained from (16) as
 − −

1
p̂1 − n−1 p̂2 e−iωτ1,2 + . . . + p̂n e−iωτ1,n
fˆ =  + − + −
. (17)
1
1 − n−1 e−iω(τ1,2 +τ1,2 ) + . . . + e−iω(τ1,n +τ1,n )

+ −
The denominator in Eq. (17) has real zeros only where ω(τ1,k +τ1,k ) mod 2π = 0 for all k. If the propagation
delays associated with the microphone spacings are incommensurate, this will happen only at large frequen-
cies (except at ω = 0). Apart from this, all poles lie in the upper half of the ω-plane. A plot of the first few
poles of Eqs. (11) and (17) for the case n = 3 is shown in Fig. 5. Accordingly, a state space realization of

1.2

1
=(ω)(τ + + τ −)

0.8

0.6
0.4
0.2
0

−0.2
0 π 2π 3π 4π 5π 6π 7π 8π 9π
<(ω)(τ + + τ −)

Figure 5. Poles of the two-microphone identification scheme (Eq. (11), squares) and a three-microphone method
+ − + − √
(Eq. (17), circles) with (τ1,1 + τ1,1 )/(τ1,2 + τ1,2 ) = 2

6 of 14

American Institute of Aeronautics and Astronautics


Eq. (17) is asymptotically stable and can be used to serve as the wave decomposer D, mapping n pressures
to the downstream propagating wave, as shown in Fig. 2. Note, however, that when implementing (17) on
a digital control board, Padé approximations for the time-delays should be used. Otherwise, rounding the
time-delays to sample time will introduce inaccuracies which might have a significant effect on the quality
of the wave decomposition – even if the control board is run at several kilohertz.

IV. Application of the Control Scheme in Simulation and Experiment


In this section, results from the application of the impedance tuning control scheme in model simulations
and in an experimental facility are presented. The configuration considered is a straight duct with flow,
terminated at the upstream boundary with a low-reflecting condition and an open end downstream. The
simulations were designed to match the acoustic characteristics of the experiment. Therefore, boundary con-
ditions and actuator transfer functions were identified from measured data and the models were incorporated
into the simulation.

A. Experimental Set-up
The experimental set-up considered is the acoustic test facility depicted schematically in Fig. 6. Seven 1/4”
microphones were installed at different axial positions. All microphones were used for frequency domain
model identification but only three of them for time domain wave decomposition in the control scheme.
The duct is made of aluminum with a wall thickness of 10 mm and has a diameter of 140 mm. Accordingly,
the cut-on frequency for the first non-planar mode is at f10 = 1435 Hz at ambient conditions.2 The upstream
end was equipped with an anechoic termination, the downstream end was open and unflanged. Speakers are
laterally mounted to the duct, close to the up- and downstream ends. For some experiments, an additional
low-reflecting end was mounted at the downstream end.

p e
DSP
anechoic end

open end
G
Rcl
flow Rol

reference plane

Figure 6. Experimental set-up used for the application of the control scheme

B. Model identification
As shown in Sec. II, the uncontrolled reflection coefficient Rol and the actuator transfer function G have to
be known to calculate the control law. Both can be determined from frequency response measurements of
the plane wave field. This procedure is explained in the following.
To identify the end element’s ĝ response according to Eq. (4), two independent acoustic states are
required. The first is generated by upstream acoustic forcing. In this case, ê = 0 and Rol can be obtained by
calculating the complex wave ratio ĝ/fˆ. A second state can now be realized by either only downstream or
simultaneous up- and downstream excitation. Since Rol is already known from the first state, the actuator
transfer function can be computed from Eq. (4) as

ĝ − Rol fˆ
G= . (18)

The Riemann invariants fˆ and ĝ cannot be measured directly. However, as mentioned in Sec. III, they
can be estimated from frequency response measurements at multiple axial locations along a uniform duct.
For the test facility considered in this work, all 7 pressure measurements were used for frequency domain
wave decomposition.

7 of 14

American Institute of Aeronautics and Astronautics


The uncontrolled reflection coefficient and the actuator transfer function for mean flow Mach numbers
M = (0, 0.05) are shown in Fig. 7. The time delays apparent in both phase responses are due to the fact
that the reference plane is located 0.4 m upstream of the actual duct outlet (see Fig. 6). Although the duct
is terminated by an unflanged open end, the magnitude of the reflection coefficient does not fully agree with
the theoretical result from Levine & Schwinger13 (Fig. 7) for the zero mean flow case. This is because the
actuator is mounted to the main duct via a side channel, whose λ/4-frequency is about 380 Hz. Around this
frequency, the impact of mean flow on the actuator transfer function is also largest (Fig. 7, right). Also note
that the reflection coefficient has been scaled with (1 + M )/(1 − M ).

1 1

0.8 0.8
|R|

0.6 0.6

|G|
1−M
1+M

0.4 M =0 0.4
M = 0.05 M =0
0.2 0.2 M = 0.05
Levine & Schwinger
0 0
-π -π
π π
-2 -2

arg G
arg R

0 0
π π
-2 -2

-π -π
200 300 400 500 600 Hz 800 200 300 400 500 600 Hz 800
frequency frequency

Figure 7. Reflection coefficient (left) and actuator transfer function (right) with and without mean flow

It may be desirable to model the open-loop reflection coefficient and/or the actuator transfer function
analytically (see, e.g., Venugopal & Bernstein14 or Pota and Kelkar15 ) and, in this way, be able to design the
controller completely a priori. However, to achieve a reasonable result, Rol and K have to be known quite
accurately, and it seems doubtful that a completely theoretical model will suffice, except for the simplest
configurations.

C. Impedance control simulation


The impedance tuning approach developed in the preceding sections was validated in simulations before it
was applied in the experimental facility. This was done in order to test the control scheme and to ensure a
stable closed-loop system.

Model set-up
The system considered in this work (see Fig. 6) is modeled as a network of acoustic elements. The structure
of the network model is shown in Fig. 8. Individual elements are described by 2 × 2 linear time-invariant
input-output relations (so called acoustic two-ports). Plane wave modeling is sufficient since only frequencies
are considered for which all higher modes are cut-off.
An interior element is sufficiently described by a 2×2 mapping of acoustic variables. There are various
representations of this mapping but only the scattering form will be considered here. The scattering matrix
of an acoustic element relates the scattered and reflected waves to the incident waves in frequency domain
" # " #" #
fˆd S11 S12 fˆu
= , (19)
ĝu S21 S22 ĝd

where subscripts u and d denote up- and downstream positions, respectively. An equivalent time domain

8 of 14

American Institute of Aeronautics and Astronautics


f
D K
e p1 p3 e
p2

duct
duct
duct
G duct
G
R R

Figure 8. Set-up of the network model

representation has the form of a general state space system with two inputs and two outputs
" #
f
ẋ = Ax + B u ,
gd
" # " # (20)
fd f
= Cx + D u .
gu gd

For a duct of constant cross-section, the scattering matrix simply reads


" ikL #
e− 1+M 0
Sduct = ikL , (21)
0 e− 1−M

where L is the duct length. To use this relation in a time domain simulation, a finite-dimensional truncation
is obtained via Padé approximants.
The up- and downstream ends are represented as actuated end elements, as described in Sec. II. In time
domain, Eq. (4) can be written as
" # " #" #
AR 0 BR 0 f
ẋ = x+ , (22)
0 AG 0 BG e
" #
h i h i f
g = CR CG x + DR DG , (23)
e

where matrices with subscripts R and G correspond to state-space realizations of the uncontrolled reflection
coefficient and the actuator transfer function, respectively. Note that R and G for the up- and downstream
ends are different due to the anechoic end mounted at the upstream side.
The pressure at a certain axial location is obtained by inserting a static gain element of the form
   
fd 1 0 " #
 fu
gu  = 0 1 (24)
  
gd
p 1 1

between two ducts with respective lengths.


A model for the complete system is now set up by connecting all respective subsystems. In this way,
an input-output model is derived, where the actuator command acts as the input and the pressure at some
desired location represents the output. Once the complete model is set up, balanced truncation is used to
reduce the state dimension. The model obtained from this procedure is accurate in frequency as well as in
time domain, as was shown in.7, 8 The modeled transfer function from the upstream excitation command e
to microphone p2 in the uncontrolled case is shown in Fig. 9. For comparison, the measured transfer function
is also added. Based on this result, the model is found to be accurate in the frequency range considered. To
test the impedance tuning approach, the wave decomposition scheme D and the control law K are added to
the model, as shown in Fig. 8.

9 of 14

American Institute of Aeronautics and Astronautics


1

measured
0.8 model
π
-2

arg(p̂2 /ê)
0.6
|p̂2 /ê| 0
0.4
π
-2
0.2


0
200 300 400 500 600 Hz 800 200 300 400 500 600 Hz 800
frequency frequency

Figure 9. Transfer function from upstream speaker command to pressure in the duct; results from network
model and experiment

Simulation results
The acoustic system was simulated in time domain with reflection coefficients and actuator transfer functions
identified from the frequency response measurements at the experimental facility. Simulations with control
schemes imposing several different acoustic boundary conditions were run.
Figure 10 (left) shows the measured reflection coefficient without control and the results of the simulation
for different prescribed closed-loop reflection coefficients Rcl . Without actuation (black curve) the reflection
coefficient approximately corresponds to that of an open end, i.e., Rol = −1 for low frequencies. The phase of
R corresponds to the length of the duct plus the end-correction of an open end.13 With the control scheme, a
non-reflecting end could be established. For Rcl = 0, the reflection coefficient calculated from the simulation
takes values below 0.05 (red curve), according to a quasi non-reflecting end. (Technically realizable non-
reflecting terminations have reflection coefficients with magnitudes between 0.1 and 0.2.) Discrete K data
and the identified model for the case of tuning the impedance to an anechoic end are shown in the right
frame of Fig. 10. A distinct increase of the magnitude of K for decreasing frequencies can be observed. This
is due to two reasons: i) the uncontrolled reflection coefficient increases for lower frequencies and ii) the
control authority of the speaker that was used deteriorated for smaller frequencies. This could have been
partially overcome by using a woofer.

1.2 10
1 8
0.8 discrete data
6 identified model
|K|
|R|

0.6
uncontrolled
4
0.4 Rcl = 0
Rcl = −1 2
0.2 Rcl = 1, ∆l = 1.3 m
0 0
-π -π
π π
-2 -2
arg K
arg R

0 0
π π
-2 -2

-π -π
250 300 350 400 450 500 550 600 650 Hz 750 200 300 400 500 600 Hz 800
frequency frequency

Figure 10. Right: Reflection coefficients (uncontrolled, no reflection, sound soft, sound hard with additional
virtual length). Left: Controller transfer function to mimic an anechoic end computed from Eq. (7) with
experimental open-loop reflection coefficient and actuator transfer function; discrete data and identified model
are plotted

10 of 14

American Institute of Aeronautics and Astronautics


The green curve (left frame, Fig. 10) presents results for the simulation of a sound hard boundary condition
(Rcl = +1, velocity node) with an additional phase lag of 2k∆l, where ∆l = 1.3 m. In this way, a virtual
additional length is added. All values of |R| lie between 0.9 and 1.1. Compared to the case without control,
the slope is steeper, due to the additional length. Hence, the controller simulates the acoustic properties of
a test rig, which is longer than the actual one and exhibits different resonance frequencies, accordingly. If
this length is taken into account, the phase has the same slope as in the uncontrolled case and differs by π.
As a last example, the simulation was run with Rcl = −1. Here, the results are even better (blue curve).
Almost in the complete frequency range considered, |R| takes values between 0.98 and 1.02. This is because
the uncontrolled open end has a reflection coefficient of R = −1 (pressure node) for low frequencies and,
therefore, the desired boundary condition is much closer to the uncontrolled case. This is also apparent in
the phase plot; both are almost identical. Consequently, the controller has to work less than in the case of
establishing a velocity node.

D. Impedance control in the experiment


Experiments were conducted without mean flow and with a Mach number of 0.05. Wave decomposition
and control scheme ran on a DS 1103 PPC (dSPACE) digital control board. Some of the desired closed-loop
reflection coefficients were the same as in the simulations in the previous section.

Anechoic end
Results for the case of an imposed anechoic end with Rcl = 0 are shown in Fig. 11. The experimentally
obtained reflection coefficient magnitudes are slightly higher than the one from the simulation. In the case
of no mean flow, the reflection coefficient magnitude remains below 0.2 in the frequency range considered.
This can be regarded as a non-reflecting end. The reflection coefficient obtained with mean flow (blue curve)
is lower than that without flow, except for frequencies below 350 Hz. Here, the controller was actually only
required to work above 300 Hz. The increase for low frequencies stems from the fact that the loudspeaker’s
response deteriorates (see Fig. 7, right). Also, the controller transfer function has a high gain (Eq. (7)) and,
therefore, the absolute error of the model grows quickly. Better results could certainly be obtained by using
a woofer with a higher response at low frequencies. However, as a result, it can be stated that the impedance
control scheme successfully managed to tune an almost fully reflecting boundary with a pressure node to an
anechoic end.

0.8

0.6
|R|

uncontrolled
0.4 Rcl = 0, M = 0
Rcl = 0, M = 0.05

0.2

0
250 300 350 400 450 500 550 600 650 Hz 750
frequency

Figure 11. Control results for a desired anechoic end. Magnitudes of the uncontrolled and the closed-loop
reflection coefficients with and without mean flow

Fully reflecting boundary and virtual length


In a next step, the control scheme was used to generate fully reflecting terminations. Figure 12 presents
results for imposed pressure and velocity nodes. Here, the nodes were assumed to be located at the axial
position of the open end – not at the reference plane (see Fig. 6). When Rcl = ∓1 is mentioned in the
following, this refers to a pressure/velocity node at the duct end. The phase of Rcl was then required to
have a phase lag corresponding to the propagation time from the reference plane to the duct end and back.

11 of 14

American Institute of Aeronautics and Astronautics


1.2 1.2
1 1
0.8 0.8
|R|

|R|
0.6 0.6
uncontrolled uncontrolled
0.4 0.4
Rcl = −1, M = 0 Rcl = 1, ∆l = 1.3 m, M = 0
0.2 Rcl = −1, M = 0.05 0.2 Rcl = 1, ∆l = 1.3 m, M = 0.05
0 0
-π -π
π π
-2 -2
arg R

arg R
0 0

π π
-2 -2

-π -π
200 300 400 500 600 Hz 800 200 300 400 500 600 Hz 800
frequency frequency

Figure 12. Boundary condition tuned to fully reflecting, Rcl = −1 (left) and Rcl = +1 (right). In case of the
sound hard end, the phase of the reflection coefficient was additionally tuned so as to mimic an additional
length of 1.3 m

In the left frame of Fig. 12, closed-loop reflection coefficients for an imposed pressure node are presented.
For the case of no mean flow (red curve), the control objective is almost exactly achieved. Over the whole
frequency range considered, the reflection coefficient magnitude deviates less than 5 % from the desired value,
except close to 800 Hz. With mean flow, the results are almost as good as in the no flow case. Overall, the
control scheme manages to establish a pressure node, however, around 260 and 380 Hz, |R| reaches values of
up to 1.2. Less successful results at 380 Hz are probably due to resonance effects in the speaker side channel.
Results for an imposed sound hard end are shown in the right frame of Fig. 12. In addition to a velocity
node, the phase of Rcl was required to mimic a duct elongated by 1.3 m, as in the simulation (see Fig. 10). The
reflection coefficient magnitudes achieved are not as good as in the case of a pressure node. This is reasonable
since the uncontrolled boundary condition more closely resembles a pressure node. Consequently, the control
scheme and the actuator have to perform less when mimicking a sound soft termination. However, good
results are obtained at all frequencies except for the resonance frequency of the laterally mounted speaker
ducts. Also, the phase with the additional delay for an elongated end could be well established. This is
a particularly important case that can be used to generate resonance frequencies which are different from
those imposed by the actual geometry.

0
10

−1
|p̂|

10

uncontrolled
Rcl = −1
−2
Rcl = +1, ∆l = 1.3 m
10
200 300 400 500 600 Hz 800
frequency

Figure 13. Pressure spectra in the duct for the uncontrolled case, an imposed pressure node, and an imposed
velocity node with virtual length of 1.3 m

12 of 14

American Institute of Aeronautics and Astronautics


Pressure spectra in the duct corresponding to the uncontrolled case, a simulated pressure node, and
a velocity node with additional virtual length of 1.3 m are presented in Fig. 13. As expected, the results
for the uncontrolled end and a closed-loop reflection coefficient of −1 are quite similar. It can be noted,
however, that the nodes are more pronounced in case of the simulated fully reflecting boundary. For the
imposed elongated duct end, the number of nodes and lobes increases due to an increased total length of the
system. The lobes represent the duct’s resonance frequencies. These do not take the form of sharp peaks,
as the upstream end is still low reflecting. Bothien et al.16 used this technique to shift the frequency of
thermoacoustic oscillations in a laboratory premixed combustor.

Different levels of reflection


To show that different levels of reflection can be realized with the control scheme, a second low-reflecting
end was mounted at the downstream boundary. The reflection coefficient magnitude for this case (without
control) corresponds to the lowest curve (purple) in the left frame of Fig. 14. Different levels of reflection
can now be established by choosing a control law
Rcl − Rol
K=α , (25)
G
where Rcl corresponds to a pressure node at the duct exit, and α takes values between 0 and 1. Results for
α = 0, 0.2, ..., 1 are presented in the left frame of Fig. 14. By raising α, the reflection coefficient magnitude
gradually increases. |R| slightly varies with frequency because the uncontrolled reflection coefficient is not
strictly zero. However, the level of reflection can be set as desired. Pressure spectra in the duct corresponding
to the different values of α are shown in the right frame of Fig. 14. For strong levels of reflection, distinct
pressure nodes and frequencies with large response can be observed. For smaller values of α, i.e., less
reflection, nodes and lobes vanish, and the pressure spectrum becomes more flat, corresponding to a purely
propagating wave. Note that even for α = 1, no sharp resonance peaks can be observed since the upstream
end is still low-reflecting.
Imposing different levels of reflection at the outlet of a combustion chamber is a useful method to
study thermoacoustic instabilities in premixed combustors. Since stability and oscillation amplitude strongly
depend on the degree of reflectivity, a controlled transition from stable to unstable conditions can be invoked
with the method presented.17

0
1.2 10

0.8
|R|

α↑ −1
|p̂|

0.6 10

α = 1.0
0.4 α = 0.8
α = 0.6
0.2 α = 0.4
α = 0.2
−2
α = 0.0
0 10
200 300 400 500 600 Hz 800 200 300 400 500 600 Hz 800
frequency frequency

Figure 14. Increasing levels of reflection, realized with a low-reflecting end mounted. Left: Reflection coefficient
magnitude. Right: Sound pressure spectrum in the duct

V. Conclusions & Outlook


It was shown how a feedback control scheme can be used to change an acoustic boundary condition to a
desired reflection coefficient. Multiple pressure measurements were used to decompose the wave field online
to create the control input. The signal was manipulated by a suitable control law and was fed back via
an acoustic actuator mounted to the lateral boundary. In this way, it was possible to adjust the reflection
coefficient to fully reflecting (sound hard and sound soft) and non-reflecting. In addition to that, a virtual

13 of 14

American Institute of Aeronautics and Astronautics


length could also be established to simulate different acoustic resonance frequencies at the same test rig.
Simulations and experiments showed similarly good results.
The presented control scheme can be used to expose a combustion system to different resonance frequen-
cies. This could be a crucial step in burner design phases since combustion test rigs usually do not exhibit
the same acoustic behavior as the full scale engine. A further application of the proposed approach is the
active control of combustion instabilities. Since the reflection of acoustic waves at the combustor outlet is a
crucial component for the thermoacoustic feedback loop, imposing an absorbing boundary condition has a
highly stabilizing effect.

Acknowledgements
We enjoyed interesting discussions on online wave identification with Gregor Gelbert. Financial support
from the German Science Foundation through the Collaborative Research Center 557 “Control of Complex
Turbulent Shear Flows” as well as from the Bundesministerium für Wirtschaft und Technologie in the
framework of AG TURBO COOREFF-T is gratefully acknowledged.

References
1 Lieuwen, T. C. and Yang, V., editors, Combustion Instabilities in Gas Turbine Engines, Vol. 210 of Progress in Aero-
nautics and Astronautics, AIAA, Inc., 2005.
2 Munjal, M. L., Acoustics of Ducts and Mufflers, Wiley & Sons, Inc., 1987.
3 Poinsot, T. J., Trouve, A. C., Veynante, D. P., Candel, S. M., and Esposito, E. J., “Vortex-driven acoustically coupled

combustion instabilities,” Journal of Fluid Mechanics, Vol. 177, 1987, pp. 265–292.
4 Aurégan, Y. and Starobinski, R., “Determination of Acoustical Energy Dissipation/Production Potentiality from Acous-

tical Transfer Functions of a Multiport,” Acustica, Vol. 85, 1999, pp. 788–792.
5 Guicking, D. and Karcher, K., “Active Impedance Control for One-Dimensional Sound,” Journal of Vibration, Acoustics,

Stress, and Reliability in Design, Vol. 106, 1984, pp. 393–396.


6 Li, Y., Chiu, G. T.-C., and Mongeau, L. G., “Dual-Driver Standing Wave Tube: Acoustic Impedance Matching with

Robust Repetitive Control,” Proceedings of the American Control Conference, Anchorage, AK, 2002.
7 Bothien, M. R., Moeck, J. P., and Paschereit, C. O., “A Modular Approach for Time Domain Modelling of Complex

(Thermo-)Acoustic Systems,” Proceedings of the 13th International Conference on Modelling Fluid Flow, Budapest, Hungary,
2006.
8 Paschereit, C. O., Moeck, J. P., and Bothien, M. R., “State-space Modeling of Thermoacoustic Systems for Stability

Analysis and Time Domain Simulation,” Proceedings of the 13th International Congress on Sound and Vibration, Vienna,
Austria, 2006.
9 Schuermans, B., Bellucci, V., Guethe, F., Meili, F., Flohr, P., and Paschereit, C. O., “A Detailed Analysis of Thermoa-

coustic Interaction Mechanisms in a Turbulent Premixed Flame,” ASME Paper 2004-GT-53831.


10 Paschereit, C. O., Schuermans, B., Polifke, W., and Mattson, O., “Measurement of Transfer Matrices and Source Terms

of Premixed Flames,” Journal of Engineering for Gas Turbines and Power , Vol. 124, No. 2, 2002, pp. 239–272.
11 Gustavsen, B. and Semlyen, A., “Rational Approximation of Frequency Domain Responses by Vector Fitting,” IEEE

Transactions on Power Delivery, Vol. 14, No. 3, 1999, pp. 1052–1061.


12 Åbom, M. and Bodén, H., “Error analysis of two-microphone measurements in ducts with flow,” Journal of the Acoustical

Society of America, Vol. 83, No. 6, 1988, pp. 2429–2438.


13 Levine, H. and Schwinger, J., “On the Radiation of Sound from an Unflanged Circular Pipe,” Physical Review , Vol. 73,

No. 4, 1948, pp. 383–406.


14 Venugopal, R. and Bernstein, D. S., “State Space Modeling of an Acoustic Duct with an End-Mounted Speaker,” Pro-

ceedings of the 1999 International Conference on Control Applications, MI, USA, 1999.
15 Pota, H. R. and Kelkar, A. G., “Modeling and Control of Acoustic Ducts,” Journal of Vibration and Acoustics, Vol. 123,

2001, pp. 2–10.


16 Bothien, M. R., Moeck, J. P., and Paschereit, C. O., “Impedance Tuning of a Premixed Combustor Using Active Control,”

2007, ASME Paper GT2007-27796.


17 Bothien, M. R., Moeck, J. P., and Paschereit, C. O., “Experimental Validation of Linear Stability Analysis in Premixed

Combustors Supported by Active Control,” Proceedings of the 14th International Congress on Sound and Vibration, Cairns,
Australia, 2007.

14 of 14

American Institute of Aeronautics and Astronautics

You might also like