Sunteți pe pagina 1din 49

CHAPTER 1

THERMO-ACOUSTIC SYSTEM MODELLING


AND STABILITY ANALYSIS
CONVENTIONAL APPROACHES
1.1 INTRODUCTION
In science and technology, the term instability characterizes situations where small
(input) perturbations of initial or boundary conditions result in a violent re-
sponse, such that system (output) variables change drastically or grow to very
large amplitudes. After a transition period, a new system state is established, which
is usually quite dierent from the initial state. Indeed, a complete breakdown of
the initial state is often observed. Many instabilities involve large-amplitude spatio-
temporal oscillations of the system variables; in this case a so-called limit cycle may
be established as a nal state.
A wide variety of instabilities are known in mechanics, hydrodynamics, plasma
physics, geophysics, control theory, aeronautics, or even economics. Well-known
examples are acoustic feedback in public address systems (electro-mechanical feed-
back), the collapse of the Tacoma Narrows bridge (aeroelastic utter), the oscilla-
tions of the London Millennium bridge (an instability involving positive biofeed-
back), the formation of large scale vortices in shear layers (Kelvin-Helmholtz in-
stability; Karman vortex street), blade utter or surge and stall in the compressor
of gas turbines (aeroelastic utter or pressure rise / mass ow instability), etc.
Thermo-acoustic instabilities arise primarily from an interaction of acoustic waves
and unsteady heat release. In general thermo-acoustic instabilities are undesirable,
because they can produce excessive noise, limit operational exibility and can even
result in structural damage. This is particularly true for combustion instabilities,
Nonnormality in Fluid Systems. By XXX
ISBN A-BBB-CCCCC-D c 2009 John Wiley & Sons, Inc.
1
Invited lecture presented at workshop
''Advanced Instability Methods''
IIT Madras, Chennai, India, January 2009
WOLFGANG POLIFKE
2 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
which are a cause for concern in applications as diverse as small household burners,
process heaters, gas turbines, or rocket engines.
Unlike other ow instabilities, thermo-acoustic instabilities are not a local phe-
nomenon in the sense that stability properties are not determined solely by the
dynamics of the heat source and the ow eld in its immediate surrounding. In-
stead, acoustic waves travel back and forth across the extent of the system, and
as a consequence acoustic boundary conditions far away from the heat source at
the top of the chimney, say may strongly inuence stability properties. In (pre-
mixed) ames, convective transport of fuel inhomogeneities from the fuel injector
to the ame, or of entropy inhomogeneities (hot spots) from the ame to the
exhaust, are other non-local phenomena which inuence system stability. All these
eects are linked together in feedback loops, as illustrated in Fig. 1.1. for a premix
combustor. It should be obvious that the complete and accurate description of a
thermo-acoustic instability can be a daunting task, since the system under consid-
eration may involve a variety of uid-dynamic and physico-chemical phenomena,
covering a wide range of space and time scales.
Whether the interaction of the uctuations of heat release rate, pressure, velocity,
fuel concentration, etc., actually gives rise to an instability depends in an essential
manner on the relative phases, and in particular on the phase between uctuations
of heat release and pressure at the heat source. This has been known since Rayleigh
(1878); the stability criterion named after him is discussed in the next section, where
the basic physics of thermo-acoustic instabilities is outlined. It is also shown
why the Rayleigh criterion represents a necessary, but not sucient criterion for
instability.
Heat Release
Fluctuations
Turbulence
Flame
Combustion
Chamber
Acoustic
Waves
Air / Fuel
Supply
Equivalence Ratio
Fluctuations
Coherent
Structures
Burner
Pressure Loss
Volume
Flow Velocity
Pressure Entropy
waves
Figure 1.1. Interactions between ow, acoustics and heat release in a combustor as an
example of a thermo-acoustic system which may exhibit instability.
In Section 3, an overview of various methods for stability analysis is given. Since
researchers in thermo-acoustics often come from dierent backgrounds, it should
not come as a surprise that a wide variety of methods have been proposed. An
introduction to possible alternatives for modelling the system comes next, ranging
from use of full-scale computational uid dynamics (CFD) to simple, low-order
network models as they are popular in duct acoustics. These models are somewhat
limited in geometrical exibility and may seem simplistic nevertheless they often
THE PHYSICS OF THERMO-ACOUSTIC INSTABILITIES 3
produce non-trivial results and are most helpful to develop insight into instability
mechanisms.
More details on network models are given in Section 4. It is shown how a math-
ematical description of network model elements may be derived from fundamental
principles for simple geometries. A few example computations obtained with sim-
ple network models are presented and discussed. In the last section CFD-based
approaches are discussed in more detail.
In these notes, it has not been attempted to give reference to, let alone discuss
in adequate detail all publications which have contributed to the development of
the eld. Suggestions for further reading are given in the short review sections at
the end of some chapters but the selection is admittedly not free from personal
bias and features the work of the combustion dynamics groups at ABB Corporate
Research and TU M unchen more prominently than would be adequate for a proper
review of the subject.
1.2 THE PHYSICS OF THERMO-ACOUSTIC INSTABILITIES
Consider ow of a gas past a heat source, e.g. a premix or diusion ame, or a
surface with convective heat transfer to the gas (the so-called stack in a thermo-
acoustic machine, or the wire in a Rijke tube, see below.). Mass conservation in
steady state requires that
(u)
c
= (u)
h
. (1.1)
With the density
h
on the downstream side (index h for hot) being lower than
the density of the upstream side (index c for cold), the velocity u, i.e. the volume
ux, must increase across the heat source. Now, if the rate of heat released by the
source uctuates, the volume produced will also uctuate, thereby generating
sound
1
just like a loudspeaker box with its oscillating membrane.
The heat release rate in a ame, say, may be perturbed by turbulent uctuations
of the velocity eld upstream of the ame front. This gives rise to combustion noise,
e.g. a camping burner or a blow torch which hisses or roars. Combustion noise
often exhibits a broad band frequency distribution, which derives from the size
distribution of the turbulent eddies perturbing the ame. Combustion noise may
also be generated by fairly large scale, vortical coherent structures, originating from
hydrodynamic instability of the base ow (e.g. a shear layer or swirling ow). In
any case, if one speaks of combustion noise, it is usually implied that there is no
signicant feedback from the sound emitted back to the ow uctuations which
perturbed the heat release in the rst place.
However, if the heat source is enclosed in a chamber that acts as an acoustic
resonator, sound will in general be reected back to the ame such that a feedback
loop is established (see again Fig. 1.1.). If the phase between the sound eld estab-
lished in the chamber and the uctuations of heat release is just right, a self-excited
feedback instability may occur. Then small (innitesimal) perturbations are am-
plied ever more, until eventually some kind of saturation mechanism kicks in and
a limit cycle is established. For saturated thermo-acoustic combustion instabilities,
limit cycle velocity uctuations often exceed the mean ow velocities, amplitudes of
1
Note that sound production by heat release with its monopole character is a very strong sound
source and if present usually dominates other source processes.
4 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
p
v
2'
3'
4'
2"
4"
3"
t
t
t
p , u
Q
Q
2'
3'
4'
1
1
1
2"
3"
4"
Figure 1.2. Thermodynamic interpretation of Rayleighs stability criterion. Left: p-
v diagram of isentropic compression / expansion ( ), with heat addition in-phase
with pressure uctuations (1-2-3-4) and out-of phase (1-2-3-4). The dashed gray lines
represent lines of constant entropy. Right-top: Fluctuations of pressure ( ) and velocity
( - - - ) in a standing wave, Right-middle/bottom: in-phase / out-of-phase uctuations of
heat release rate.
pressure uctuations can reach more than 120 dB in atmospheric ames, and sev-
eral MPa in rocket engines. Damage to the combustion equipment can then result
very quickly due to excessive mechanical or heat loads. If such thermo-acoustic
instability occurs, the frequency spectrum of the resulting pressure and velocity
oscillations typically exhibit one or several distinct peaks, with frequencies often
(but not necessarily) close to the acoustic eigenfrequencies of the enclosure (or the
complete combustion system) without unsteady heat release.
Rayleigh (1878) has proposed a criterion which tells us when the phases between
the sound eld and the heat release uctuations are just right: For instability to
occur, heat must be released at the moment of greatest compression. A more general
formulation of the Rayleigh criterion states that a positive correlation between
the uctuations of heat release and pressure, respectively, is required for thermo-
acoustic instability to be possible:

dt > 0. (1.2)
Here the integration runs over one period of the oscillation, p

denotes uctuations
of pressure at the heat source,

Q

the uctuations of the heat transfer (or release)


rate. There is also a straightforward generalization of this Rayleigh integral for
spatially distributed heat release rate, which includes a spatial integration

. . . dV
over the volume where heat is released.
The Rayleigh criterion (1.2) may be made plausible by analogy with a thermo-
dynamic cycle. Consider a small volume of gas, which is compressed and expanded
by a standing acoustic wave. Sound waves are isentropic, so in a p-v diagram the
volume moves back and forth on an isentrope (see the line in the diagram on the
left side of Fig. 1.2.). What happens if heat is added and extracted periodically to
the gas? Well, addition of heat will result in a comparative increase of the specic
volume v of the gas, and if the heat addition is in-phase with pressure uctuations,
the state of the gas volume moves clockwise around a thermodynamic cycle (curve
1-2-3-4 in the diagram). So this is a thermo-acoustic heat engine, which feeds
mechanical energy into the sound wave and a self-excited instability may occur,
THE PHYSICS OF THERMO-ACOUSTIC INSTABILITIES 5
as Rayleigh suggested. If uctuations of heat addition are not perfectly in phase
with pressure uctuations, then the area of the cycle 1-2-3-4 will be smaller, and
the eciency of the engine will decrease. In the extreme case where uctuations of
heat release rate

Q

are opposite in phase with p

, the system moves counterclock-


wise through the cycle 1-2-3-4 and mechanical energy is extracted from the
acoustic wave.
The mechanical work performed by the thermodynamic cycle resulting from
acoustic perturbations with heat release illustrated in Fig. 1.2. may be formulated
as follows:

dv

=
v
p

dp

dv
(Q)
= 0 +

dv
(Q)
dt
dt

dt. (1.3)
Here the changes v

in specic volume have been split into an isentropic part, for


which v

= vp

/p (where is the ratio of specic heats), and a part v


(Q)
which
is due to heat addition (or removal). The rate of change of this term with time is
proportional to uctuations of the rate of heat addition. One may conclude that
the work done by the thermo-acoustic engine is positive (energy is added to the
acoustic eld), if the integral of p

over one period of oscillation is positive just


as Rayleigh has argued.
Rayleighs criterion is necessary, but not sucient for instability to occur
2
.
This means that a system is guaranteed to be thermo-acoustically stable, if the
stability is not fullled but if the criterion holds, it is not guaranteed that an
instability will indeed develop. The reason for this limitation is the following: if
the criterion is fullled, Rayleighs thermo-acoustic engine feeds mechanical energy
into the acoustic eld. However, oscillation amplitudes will only grow, if losses of
acoustic energy, which may occur elsewhere in the system, do not exceed the rate
of energy generated by the uctuating heat source. More complete stability criteria
have been developed, they are discussed in Section 1.3.
Lets conclude this section with an important observation: Rayleighs criterion is
stated in terms of uctuations of heat release rate and pressure at the heat source.
Nevertheless, acoustic boundary conditions far away from the heat source can and
in general will determine stability properties in a decisive manner. Why is that?
To answer this question, one must consider that convective heat transfer rates as
well as the heat release rate in a premix ame usually respond much stronger
to changes in velocity u

than in pressure p

. The phase between uctuations of


pressure p

and u

at the ame is determined by the acoustics of the system, e.g. the


acoustic impedance at the boundaries, and the wave pattern perhaps with partial
reections and transmissions inside the enclosure. As a very simple example,
consider the fundamental /2-mode in a straight duct of length L with two open
ends and a compact heat source at position x = L/4 as shown in Fig. 1.3.. Such a
system has been known to exhibit thermo-acoustic instability for a long time and
is known as aRijke tube Rayleigh (1878); McManus et al. (1993); Heckl (1988).
Acoustic uctuations of velocity in the left half of the tube are opposite to those in
the right half of the tube (velocity u

is ahead of pressure p

by /2 in the left half


and lags behind by /2 in the right half). If we assume that a heat source placed
in the tube responds to changes in velocity, then clearly the sign of the Rayleigh
2
Other stability criteria found in the literature, which are based in a similar manner only on phase
relationships, suer also from this limitation.
6 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
Q
.
p =0 p =0
i c h x
Figure 1.3. Rijke tube of length L with two open end boundary conditions (inlet i,
exit x) and a compact heat source

Q at x
c
= L/4 with x
h
x
c
L. Distribution of
uctuating velocity and pressure / density at one particular instant during the oscillation
are indicated by arrows and shading, respectively.
TECHNISCHE UNIVERSITT MNCHEN
Growth rate of instabilities
Q = 385 7 W; v
mean
= 0.0218 0.0002 m/s
( )
1
ft
gr +
Figure 1.4. Time trace of pressure in an electrically heated Rijke tube. After a limit
cycle is established, a plug is pulled from the tube, such that the instability collapses.
Integral

p

dt will depend on the position of the heat source: a heat source,


which gives rise to self-excited instability when placed in the left half of the tube,
will not go unstable when placed in the right half, and vice versa. This is explained
in more detail in the following.
It follows that it is always the combination of heat release dynamics and system
acoustics which controls stability. System acoustics determines largely both the
impedance at the heat source (the phase between p

and u

) as well as the total losses


of acoustic energy (dissipation inside the system, also radiation to the environment
at the boundaries). For this reason stability analysis without a model for the
acoustics of the enclosure a system model is in general not possible for thermo-
acoustic systems.
STABILITY ANALYSIS 7
1.3 STABILITY ANALYSIS
A system (or a system state) may be called stable against a perturbation, if some
time after the perturbation is imposed, the initial system state is re-established.
It follows quite naturally from this denition that in order to investigate the sta-
bility properties of a system, one must determine the response of the system to
perturbations.
Often it is fairly easy to compute the response to perturbations with innites-
imally small amplitudes. In this case the governing equations may be simplied
by neglecting the so-called non-linear terms, i.e. terms which are higher order in
perturbation amplitudes. A system is accordingly called linearly stable, if it returns
to its initial state after a slight perturbation. Note that it is not specied what
kind of perturbation is imposed, except that the amplitude be small in some sense.
In linearly unstable systems, a dominant or most unstable mode i.e. a distinct
pattern of vibration with a particular frequency develops in the very early stages
of the instability. This most unstable mode has as the term implies a growth rate
that is larger than that of any other unstable mode. After a while, the most unstable
mode dominates the behaviour of the system such that the whole system oscillates
at the frequency of the mode. In this stage, the frequency spectrum typically
shows a single peak at the frequency . The oscillation amplitudes keep growing
exponentially, until saturation of amplitude sets in and eventually a limit cycle,
i.e. oscillation with large, but nite amplitudes, is established. Signicant growth
of amplitudes is associated with linear instability, correspondingly the departure
from exponential growth is due to non-linear terms, i.e. those terms of higher order
in oscillation amplitudes, which were neglected in the linearized analysis. During
the non-linear phase of the evolution of the instability, the frequency spectrum
typically shows more than one peak often a dominant peak at the frequency
and lower-amplitude, but still rather distinct peaks at integer multiples n of the
fundamental frequency (harmonics).
This scenario of linear instability growth of dominant mode non-linear
saturation limit cycle is a well established paradigm in instability theory (Drazin
2002; Keller 1995; Dowling 1995). Experimental results for a Rijke tube obtained
by Lumens (2006) are shown in Fig. 1.4.. Nevertheless, there is evidence that this
scenario does not always apply. For example, some linearly stable thermo-acoustic
systems operate stably for a long time, but exhibit strong instability after suering a
strong external perturbation. This kind of non-linear instability behaviour is called
triggering. It has been observed in thermo-acoustic systems ranging from Rijke
tubes to rocket engines (Culick 1989, 1994; Flandro et al. 2007). More recently, the
relevance of non-normal eects for thermo-acoustic instabilities has been brought
to attention by Balasubramanian and Sujith (2007a, b).
Furthermore, many combustion systems, while not exhibiting a strong thermo-
acoustic instability, show signicant uctuation levels (of pressure, velocity, heat
release rate) over a band of frequencies (often in the vicinities of combustor eigen-
frequencies). Such a state should perhaps not be described as the limit cycle of an
instability, but rather as resonant amplication of combustion noise by the com-
bustion chamber.
8 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
1.3.1 Unsteady simulation
Probably the most straightforward method to assess the stability of a thermo-
acoustic system makes use of unsteady simulation in the presence of small (initial)
perturbations: If the perturbations grow in amplitude with time, the system is
understood to be unstable, and stable otherwise. Provided that non-linear eects
are taken into account properly by the simulation model which is in principle
possible for many models in use a limit cycle is established once the simulation
is continued for a suciently long time,
As an example of this strategy, Fig. 1.5. shows the temporal development of the
velocity just upstream of the heat source in a generalized Rijke tube, i.e. a resonator
with closed-open boundary and heat source with time lag placed in the center. The
computational model used to generate the data shown in Fig. 1.5. was based on
simplied 1-D conservation equations for mass, energy and momentum in laminar
compressible ow. Lumped parameter source terms for energy and momentum
were used to model the heat source, see Polifke et al. (2001b). Obviously, the
simulation exhibits instability: the oscillation amplitudes grow until a limit cycle
with negative velocities, i.e. reverse ow past the heat source during part of the
oscillation cycle, is established
3
.
The use of unsteady simulation for stability analysis is conceptually a straight-
forward approach, and has been used by a number of authors (Janus and Richards
1996; Dowling 1997; Veynante and Poinsot 1997; Peracchio and Proscia 1998; Po-
lifke et al. 2001b; Pankiewitz and Sattelmayer 2003b; Sung et al. 2000). This does
not imply, however, that this approach is easy to implement or generally superior
to alternative strategies. The following drawbacks may be identied:
Unsteady simulation is a brute force approach which can be computation-
ally very expensive. This is in particular the case if a computational uid
dynamics simulation model for (turbulent, reacting), incompressible ow is
used, i.e. the unsteady Reynolds-averaged Navier-Stokes (URANS) or Large
Eddy Simulation (LES).
In general, only the most unstable mode is identied. Stable modes, or modes
that are unstable, but have smaller growth rates than the most unstable mode,
cannot be observed. In this sense, the stability assessment based on unsteady
simulation is not comprehensive.
In some instances, it is not easy to discern a spurious numerical instability
from a genuine physical instability. A related problem is that it can be di-
cult to numerically generate the initial state, from which the instabilities are
supposed to develop.
Results can depend on initial conditions, i.e. details of the initial perturbation
(typically a random perturbation of very small amplitude) can inuence which
mode develops. For example Evesque et al. (2003), a Finite-Element based
model of linear acoustics, showed that for non-plane acoustic waves in an
annular combustor, the initial conditions can determine whether a standing or
3
It is straightforward to interpret this result: if hot gases are swept back to the heat source,
the temperature dierence and therefore also the heat transfer between the hot wire and the gas
decreases, thereby reducing with increasing uctuation amplitude the gain of the transfer function
and the Rayleigh integral until the instability saturates.
STABILITY ANALYSIS 9
30
20
10
0
-10
v
c

[
m
/
s
]
0.7 0.6 0.5 0.4 0.3 0.2 0.1 0.0
t [s]
Figure 1.5. Temporal development of velocity v
c
upstream of the heat source in a
generalized Rijke tube. Time lag of the heat source was adjusted to fulll the Rayleigh
criterion.
a rotating mode develops. Further analysis showed that rotating modes with
clockwise and counterclockwise sense of rotation are degenerate eigenmodes
for this problem. This explains the sensitivity against initial conditions in
this particular case.
In general, the acoustic impedance at the boundary of the computational
domain is a complex-valued function Z() which depends on frequency .
The implementation of appropriate acoustic boundary conditions in the time
domain is a non-trivial problem. See Schuermans et al. (2005); Huber et al.
(2008) for recent advances on this problem in the context of CFD-based mod-
els of thermo-acoustics in the presence of mean ow.
1.3.2 Determination of eigenmodes and eigenfrequencies
The stability of a system can be determined by identifying the eigenmodes and in
particular the eigenfrequencies of the system. This classical method, known also as
dynamic stability analysis, is very popular and can be applied to a wide variety of
problems.
What is an eigenmode and an eigenfrequency? The German prex Eigen can
be translated as own, peculiar to, characteristic or individual. An eigenfrequency
is then a frequency that is easily excited in a system. Once excited, the system,
left to itself, will continue to oscillate for some time at that frequency. The corre-
sponding pattern of vibration, e.g. the spatial distribution of nodes and anti-nodes,
is identied as the shape of the eigenmode. Typically, a dynamic system has more
than one even innitely many eigenmodes and corresponding eigenfrequencies.
Trying for a bit more mathematical rigor, the eigenvectors x
m
of an operator A
are those vectors which are left unchanged under the operator up to factors
m
,
the eigenvalues of the operator:
Ax
m
=
m
x
m
. (1.4)
This should be familiar from linear algebra. The notions of eigenvector (or eigen-
function or eigenmode) and operator may be interpreted in a rather general
sense. For example, consider acoustic oscillations of pressure p

in a duct of length
10 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
L with two open end boundary conditions p

(0) = p1(L) = 0, obeying the 1D


Helmholtz-equation,
1
c
2

2
p

t
2


2
p

x
2
= 0. (1.5)
The eigenmodes are the well-known standing waves
p

(x, t) exp(i
m
t) sin(k
m
x).
The wave numbers k
m
, which are restrained to the values k
m
= n/L in order
to fulll the boundary conditions, determine the shape of each eigenmode. The
Helmholtz-equation then is re-written as

2
p

t
2
=
2
m
p

, (1.6)
for the eigenfrequencies
m
= k
m
c. When formulated in this way, the analogy with
(1.4) should be obvious: The eigenmodes correspond to particular oscillations of
the gas column in the duct, such that the second derivative of pressure with respect
to time is simply the shape of the pressure distribution scaled by a factor. This
factor, the eigenvalue, turns out to be
2
m
, where
m
is the angular frequency of
the oscillation. The fact that the time derivative operator is Hermitian
4
leads to
several important properties, such as that the standing wave patterns are orthogonal
functions.
For ideal systems without dissipative eects, an eigenmode oscillates without
decay, corresponding to a purely real eigenfrequency R. If damping (due to
viscous losses or radiation losses at the system boundary) is non-negligible, oscil-
lation amplitudes decrease in time. If amplitude levels are small, non-linear terms
may be neglected, then an exponential decrease of amplitudes is typical. Assuming
harmonic time dependence exp(it) an eigenfrequency
m
with positive imagi-
nary part (
m
) > 0 describes this situation
5
. Unstable modes in systems which
can exhibit self-excited instability this is of course the case most interesting in
the present context are characterized by eigenfrequencies with negative imaginary
part (
m
) < 0.
A growth rate or cycle increment of an eigenmode m can be derived from its
eigenfrequency:
exp

2
(
m
)
(
m
)

1. (1.7)
The growth rate indicates by which amplitude ratio a mode increases or decreases
per cycle. For example, a growth rate of = 0.2 indicates an increment of the
amplitude of 20% per cycle.
To illustrate this method of stability analysis, it is shown in the next subsec-
tion how the eigenfrequencies of a Rijke tube can be determined from a low-order
thermo-acoustic model.
4
In mathematics, on a nite-dimensional inner product space, a self-adjoint operator is one that
is its own adjoint, or, equivalently, one whose matrix is Hermitian, where a Hermitian matrix is
one which is equal to its own conjugate transpose. By the nite-dimensional spectral theorem
such operators have an orthonormal basis in which the operator can be represented as a diagonal
matrix with entries in the real numbers (from Wikipedia).
More details on Sturm-Liouville theory (the theory of linear dierential equations, dierential
operators and their eigenvalues / eigenfunctions) are found in many mathematics text books.
5
Note that often the opposite convention exp(it) for the time dependence is chosen. In this
case, a negative imaginary part indicates a decaying eigenmode.
STABILITY ANALYSIS 11
1.3.2.1 Eigenfrequencies of a Rijke tube The Rijke tube is a simple, yet interesting
example of a system that can exhibit self-excited thermo-acoustic instability. It has
been discussed previously by many authors for its pedagogical value (McManus
et al. 1993; Dowling 1995; Polifke et al. 2001b; Polifke 2004). Here it is shown how
the eigenfrequencies and thereby the stability characteristics of a Rijke tube can
be determined from the characteristic equation (also dispersion relation), which in
turn is derived from an approximate low-order model of the Rijke tube.
We consider a system as shown in Fig. 1.3., i.e. a heat source placed in a duct.
It is assumed in the following that there is an acoustically closed end (u

= 0) at
the left boundary. This little trick simplies the analysis a bit. Also, the position of
the heat source is not at position x
c
= L/4, as it is for a real Rijke tube. Instead,
the length of the duct to the left of the heat source is denoted as l
c
, while to the
right we have l
h
= L l
c
.
It is assumed that the reader is familiar with basic acoustics, so instead of starting
from scratch, we just introduce some notation and go into medias res quickly:
Riemann Invariants f and g (also called p
+
and p

by some authors) are related


to the primitive acoustic variables as follows,
f =
1
2

c
+u

, g =
1
2

c
u

. (1.8)
One may think of these Riemann invariants (also called characteristic wave
amplitudes) simply as acoustic waves propagating in the down- and upstream
direction, respectively
6
.
Now consider the length of duct to the left of the heat source, i.e. the cold
part of the Rijke tube. A downstream traveling wave f undergoes a change in phase
exp{il
c
/c
c
} as it travels with the speed of sound c across the distance l
c
= x
c
x
i
from the inlet to the cold side of the heat source. Similar for the wave g traveling
in the upstream direction. Along the length of the duct, there is no interaction
between the waves f and g. From these coupling relations the transfer matrix of
the duct is readily obtained:

f
c
g
c

e
ik
c
l
c
0
0 e
ik
c
l
c

f
i
g
i

. (1.9)
with a wave number k
c
/c
c
for plane waves without mean ow and dissipative
eects. For the hot duct to the right of the heat source, the transfer matrix is of
the same form, but of course c =

RT and l may be dierent, so the coecients


of the transfer matrix may be dierent, too.
For the heat source (which is a wire mesh or gauze in a Rijke tube), one
observes that the pressure drop across the heat source is negligible for suciently
small mean-ow Mach number, p

c
= p

h
, while for the velocity we assume that
u

h
(t) = u

c
(t) +nu

c
(t ). (1.10)
This is the famous n model for a heat source with time lag. This simple
model has played an important role in the development of the theory of combustion
6
A note on nomenclature: we make in these notes no explicit distinction between acoustic variables
in the time domain and in the frequency domain. Most of the time we are in frequency space,
and the fs and gs are to be understood as complex-valued Fourier transforms of time series.
12 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
instabilities in rocket engines, is used frequently as a pedagogical example (as in
these notes), and is even relevant as a building block for models of real-world
turbulent premix ames. One can show (see e.g. Polifke (2004); Poinsot and
Veynante (2005)) that the second term on the r.h.s. appears if the heat release of
a ame (or some other source of heat) responds to a change in velocity u with a
certain time lag . The important thing about the time-lag is that it may allow some
phase-alignment between uctuations of pressure and heat release, respectively,
even with standing acoustic waves (where the phase between p

and u

is /2).
So, for given impedance Z = p

/u

at the heat source, the time lag in combination


with the frequency controls the sign of the Rayleigh integral and therefore system
stability. For the interaction index n, one nds under certain assumptions which
hold true for the gauze of a Rijke tube, but not necessarily for a premix ame
that n = T
h
/T
c
1, i.e. n is determined by the increase in mean temperature T
across the heat source.
In terms of the Riemann invariants, the coupling relations across the heat source
are

h
c
h

c
c
c
(f
h
+g
h
) = (f
c
+g
c
), (1.11)
f
h
g
h
=

1 +n e
i

(f
c
g
c
). (1.12)
At the left boundary u

= 0 and therefore f
i
g
i
= 0, while at the right boundary,
p

= 0 and correspondingly f
x
+g
x
= 0. This completes the construction of a low-
order model for the Rijke tube: there are eight equations (2 boundary conditions,
22 equations from the ducts on both sides of the heat source, and 2 equations from
the heat source) for the eight unknowns f
i
, g
i
, . . . g
x
. In matrix & vector notation,
this reads

Matrix
of
Coecients

f
i
.
.
.
g
x

0
.
.
.
0

. (1.13)
As indicated, the system is homogeneous (the right hand side is a vector with all
0s), so a non-trivial solution exists only if the determinant of the system matrix S
vanishes, i.e. if
Det(S) = 0. (1.14)
How can this condition be satised? Inspection of the coupling relations (1.9)
- (1.12) shows, that the coecients of the system matrix depend on geometrical
and physical parameters (length l, speed of sound c, density in the ducts plus
the interaction index n and the time lag ), which are xed for a given system.
However, the frequency , which appears in several of the coecients, is not xed
a priori, so we may hope to nd one or indeed several eigenfrequencies
m
of the
system such that Det(S)|
=
m
= 0. In this sense Eq. (1.14) is the equivalent to
the characteristic equation Det(AI) = 0 in linear algebra.
For the Rijke tube with
h
c
h
/
c
c
c
= c
c
/c
h
,
Det(S) = 4{cos(k
c
l
c
) cos(k
h
l
h
) sin(k
c
l
c
) sin(k
h
l
h
)

1 +n e
i

}, (1.15)
see McManus et al. (1993) and Appendix 1.7 for the derivation.
The characteristic equation Det(S) = 0 is a transcendental equation, i.e. it
cannot be solved explicitly for eigenfunctions
m
. In general one has to resort
STABILITY ANALYSIS 13
1 2
m=0
m=1
m=2
m=3
3

0
Figure 1.6. Stability map of the rst four modes in a Rijke tube with one closed end
and weak coupling n 0. Gray regions indicate stability, i.e. positive imaginary part of
the eigenfrequency
m
.
to numerical root nding to determine the eigenfrequencies (see the Appendix for
a simple implementation in Matlab). An approximate analytical solution for the
special case l
c
= l
h
,
h
=
c
and c
c
= c
h
= c has been derived by McManus et al.
(1993). Introducing non-dimensional variables with L and L/c as characteristic
lenght- and time-scales, respectively, the characteristic equation (1.15) reduces for
this conguration to
cos() n e
i
sin
2
(/2) = 0. (1.16)
For n = 0, i.e. in the absence of thermo-acoustic coupling between velocity and
heat release at the gauze, the solutions to this equation are the familiar quarter-wave
duct eigenmodes with wave lengths = L/4, 3L/4, . . . and frequencies

(0)
m
=

2
,
3
2
,
5
2
, . . . = (2m+ 1)

2
; m = 0, 1, 2, . . . . (1.17)
These eigenfrequencies are real-valued, so there is no amplication or damping of
the eigenmodes.
For non-zero, but small interaction index n 1, the eigenfrequencies can be
determined approximately as
m

(0)
m
+

m
, where the deviations

m
from the
eigenfrequencies
(0)
m
are assumed to be small,

m

(0)
m
. Then, to rst order in

m
,
cos(
(0)
m
+

m
) (1)
m

m
,
sin
2
(
(0)
m
+

m
)
1
2
+
(1)
m
2

m
.
Inserting this in Eq. (1.16) and retaining only terms that are 1st order in n or

m
one obtains

m
(1)
m+1
n
2
e
i
m

or (1.18)

m

(0)
m
+ (1)
m+1
n
2
e
i
(0)
m

, (1.19)
where
(0)
m
= (2m + 1)/2, see Eq. (1.17). Obviously, coupling between acoustics
and heat release (non-zero interaction index n) results in complex-valued eigenfre-
quencies, i.e. and eigenmodes that asymptotically grow or decay in time. Adopting
the convention exp (it) for the time dependence, an eigenmode is unstable if
14 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
0.5 1 1.5 2 2.5 3

1.45
1.5
1.55
1.6
1.65
1.7
1.75
Rew
0.5 1 1.5 2 2.5 3

-0.15
-0.1
-0.05
0.05
0.1
0.15
0.2
Rew
Figure 1.7. Comparison of eigenfrequencies determined with numerical () and
approximate analytical solutions (dashed lines) of the dispersion relation (1.16) with
interaction index n = 0.3 for the fundamental mode m = 0. With increasing n, m the
dierences between numerical and analytical solutions will increase (not shown).
the imaginary part of its eigenfrequency is negative,
(
m
) (1)
m
n
2
sin(
m
) < 0, (1.20)
because in this case the amplitude grows with time as exp ((
m
)). The stability
limits for the time lag resulting from this criterion for eigenfrequencies according
to Eq (1.19) are shown in Fig. 1.6. for modes 0 to 3. For almost any value of
the time lag , one or several of the rst four eigenmodes is unstable. In a more
realistic model, viscous losses and especially radiation of acoustic energy to the
environment at the open end should be taken into account. Both eects tend to
increase stability, and they both tend to increase with frequency, so that the width
of the stable regions in Fig. 1.6. would increase, and especially so for the higher
modes (Indeed, in a real Rijke tube, one usually observes only the fundamental
mode m = 0). Note that the Rayleigh criterion, which cannot take into account
the benecial eect of losses on system stability, would then be overly pessimistic.
The real part of

m
shows that the frequency of an unstable mode will in general
dier slightly from the eigenfrequency of a corresponding stable mode (with n = 0):
(

m
) (1)
m+1
nc
2L
cos
m
, (1.21)
This is due to the phase change introduced by the time-lagged response of the heat
source.
It is not dicult to solve the dispersion relation numerically for larger values of
n (and indeed for arbitrary values of the parameters l, c, n, as well as arbitrary
reection coecients at the duct ends). A simple Mathematica script is given in
the Appendix, results are shown in Fig. 1.7..
1.3.2.2 Further remarks on dynamic stability analysis
What has been done here by hand for the simplistic model of a Rijke tube
can be formalized and cast into a numerical algorithm a network model.
With this approach, the thermo-acoustic system is modelled as a collection
of network elements. Each network element is represented mathematically
by its transfer matrix (or scattering matrix), derived from coupling relations
akin to Eq. (1.9). With given network structure and a library of network
STABILITY ANALYSIS 15
Burner Flame Combustor Air Supply
Fuel Supply
Figure 1.8. Network model of a combustion system.
elements, the system of equations (1.13) (or the system matrix) representing
the complete network may then be generated automatically by the computer
without any paper and pencil work. This can be done for networks of
arbitrary topology, representing combustion systems with fuel and air supply
ducts, cooling channels, etc. as suggested in Fig. 1.8.. Of course, the search
for eigenfrequencies is then also performed numerically.
The computation and inspection of eigenfrequencies as presented above for the
Rijke tube is compatible with Rayleighs criterion: If one determined from a
network model the relative phases between uctuations of pressure p

and heat
release

Q

at the heat source and evaluated the Rayleigh integral (1.2), one
would obtain results in complete agreement with Eq. (1.20). However, this is
only so because we neglected losses and assumed ideal boundary conditions
with acoustic reection coecients |r| = 1. Because loss mechanisms can
easily be taken into account with a low-order model, the dynamic stability
criterion is more powerful than Rayleighs criterion.
The dynamic stability criterion is in principle applicable to any homogeneous
linear system. However, for complicated systems, the roots of the charac-
teristic equation (1.14), i.e. the eigenfrequencies
m
, can in general not be
determined analytically, because the system of equations (1.13) is simply too
large for paper-and-pencil work. Instead, iterative numerical algorithms for
root-nding are required. The iterative search for the roots gives rise to a
number of problems: In order to assure that a system is stable with respect
to self-excited thermo-acoustic oscillations, it is necessary (and sucient) to
verify that all eigenfrequencies
m
are located in the upper half of the com-
plex plane. However, numerical root-nding always starts from an initial
guess and moves iteratively by trial and error towards a root. The roots
may be located anywhere in the complex plane; the basins of attraction of
the various roots may be quite dierent in size, and some roots may indeed
be very hard to nd. So one can never be sure that one has really found all
eigenfrequencies but one unstable root that goes unnoticed may render an
otherwise stable system unstable!
Sometimes the coecients of the system matrix are known only for real-
valued frequencies (e.g. if transfer matrices are determined by experiment or
from computational uid dynamics). In this situation, one cannot solve for
complex-valued roots of the characteristic equation, because the determinant
of the system matrix away from the real axis is not known.
In the next section, an alternative approach making use of Nyquist diagrams as
they are familiar from control theory is presented, which does not suer from some
of the shortcomings mentioned. In particular, iterative searches for eigenfrequencies
in the complex plane are not required with that approach.
16 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
G(s)
1
x
1
x
0
Figure 1.9. Open loop system with unit negative feedback.
1.3.3 The Nyquist Criterion
The Nyquist stability criterion is a well-known tool in control theory. It allows to
test for stability of a closed-loop system by inspection of the Nyquist plot of the
open-loop transfer function (OLTF) (Jacobs 1993). In this section, a brief review of
the method is given, then the application to thermo-acoustic systems is discussed.
1.3.3.1 Nyquist plots in control theory Consider a system as shown in Fig. 1.9.
with open loop transfer function G(s) and unit negative feedback H(s) = 1, such
that
x
1
= G(s) (x
0
x
1
), or x
1
=
G(s)
G(s) + 1
x
0
. (1.22)
The characteristic equation, from which stability can be deduced, is then
G(s) + 1 = 0. (1.23)
The system is stable, if all the roots s
n
of this equation are located in the left
half of the complex plane. In that case, all perturbations of the system will decay
exponentially e
st
with time.
To locate all the roots can be tedious, therefore alternative methods to assess the
stability have been developed. For example, in control theory the transfer function
is usually a fraction of polynomials,
G(s) =
P
N
(s)
P
D
(s)
=
a
n
s
n
+. . . +a
0
b
m
s
m
+. . . +b
0
. (1.24)
The Routh-Hurwitz stability criterion exploits this fact and allows to determine
system stability from the polynomial coecients (Jacobs 1993).
Alternatively, one can deduce stability from the Nyquist plot, i.e. a polar plot
of the imaginary axis
7
mapped through the open loop transfer function. The asso-
ciated Nyquist stability criterion is based on Cauchys argument principle, which
states the following: Consider an analytical function f(z) with a number Z of zeros
f(z) = 0 and a number P of poles f(z) within a simple closed contour C in
the complex plane. Then the winding number of the image curve of the contour C
mapped f(z) around the origin 0 +i0 is equal to Z P.
7
Strictly speaking, the Nyquist plot is the image of the Nyquist contour, which is a closed contour
encompassing the right half of the complex plane. Starting from i one moves along the
imaginary towards +iand then in a clockwise circular arc with radius r back to the point
i.
STABILITY ANALYSIS 17
For the characteristic equation (1.23) and the Nyquist plot, Cauchys argument
principle implies that the number Z of zeros of the OLTF G(s) in the right half
of the complex plane is related to the number N of positive encirclements (i.e. in
clockwise direction) of the critical point (1 + 0i) and the number of poles P of
G(s) as follows:
N = Z P. (1.25)
For a stable system, no roots of the characteristic equation (1.23) should be on the
right side of the s-plane, i.e. Z = 0. Nyquists criterion follows with Eq. (1.25):
for stability, the number N of anticlockwise encirclements about the critical point
(1 +0i) must be equal to P, the number of open loop poles in the right half plane.
1.3.3.2 Application of the Nyquist criterion to thermo-acoustic systems It is by no
means obvious how the Nyquist criterion can be employed to analyze the stability
of thermo-acoustic systems. A network-model-based approach for generation of a
Nyquist plot has been proposed by Polifke et al. (Polifke et al. 1997; Sattelmayer
and Polifke 2003a): To establish an analogy to control systems and to dene the
equivalent to the OLTF, the network model must be cut open. This can be done
for networks of arbitrary topology by introducing a diagnostic dummy element
in the network.
As indicated in Fig. 1.10., two of the four variables of this network element
are simply shortened out, e.g. g
d
= g
u
, while the second pair of variables is
not connected at all. Instead, a xed value is assigned to one of those ports, e.g.
f
d
= 1. Note that by inserting the diagnostic dummy, the homogeneous system
of equations (1.13) changes into an inhomogeneous system S

= b with r.h.s.
b = (0, . . . , 1, . . . , 0). Now one can dene the OLTF of the thermo-acoustic system
with diagnostic dummy as
G() =

f
u

f
d
. (1.26)
The minus sign is introduced to maintain close analogy with negative feedback
control systems.
In general, i.e. for arbitrary frequency , the solution x

() of the inhomoge-
neous system with diagnostic dummy will not be equal to any of the eigenmodes of
the original system Eq. (1.13). In this case the values of the unconnected variables
across the diagnostic dummy will not be equal, i.e.

f
d
=

f
u
. However, for every
eigenfrequency
n
, the solutions x

(
n
) of the system with diagnostic dummy will
be identical to the corresponding eigenvector x
n
of the original system (up to an
arbitrary scaling factor), and the acoustic variables will match across the cut,

f
d
=

f
u
. It follows that Eq. (1.26) denes a mapping, which maps every eigen-
mode of the homogeneous system Eq. (1.13) to the critical point (-1 + 0i), i.e.
G(
n
) = 1 for every eigenfrequency
n
. This important property of the OLTF
will be exploited in the following.
Although the equivalent to the open loop transfer function is now dened, the
classical Nyquist criterion, as it was formulated in the previous section, is not di-
rectly applicable to a thermo-acoustic system: As already mentioned, low-order
models for thermo-acoustic stability analysis are commonly formulated with har-
monic time dependence e
it
, and the imaginary part of the angular frequency
determines stability. For a stable system, no eigenfrequencies
n
must be located in
18 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
f
d

g
d

g
u

f
u
1
Figure 1.10. Diagnostic dummy
the lower half of the complex plane. This implies for the Nyquist plot, that the real
axis of the -plane is mapped by the open loop transfer function to the G()-plane.
Once these modications are taken into account, one could in principle proceed
with application of the criterion in complete analogy to control theory. However,
transfer functions in thermo-acoustics are in general not polynomials in or frac-
tions thereof, but involve harmonic or exponential functions
8
. The identication of
poles is in this case no easier than the determination of eigenfrequencies by iterative
numerical solution of the characteristic equation (1.23). Moreover, the coecients
of acoustic transfer matrices the building blocks of network models are not
always given in analytical form. In this case, the OLTF cannot be evaluated for
complex-valued frequencies with imaginary part () = 0. These diculties are
also discussed by Polifke et al. (1997); Sattelmayer and Polifke (2003a).
Therefore, a modied rule for the interpretation of Nyquist plots has been put
forward, which is more suitable for application to thermo-acoustic systems (Polifke
et al. 1997; Sattelmayer and Polifke 2003a). The proposed rule is based on a
property of analytical functions: An analytic function is conformal, i.e. it preserves
local angles or handedness, at any point where it has a nonzero derivative (Janich
1983). Consider now the open loop transfer function G() as a conformal mapping
from the -plane onto the G-plane, see Fig. 1.11.. The real axis () = 0 in the
-plane (left graph) and its image in the G-plane (right graph) are indicated by the
thick dashed line with arrow head. According to Eq. (1.26), the eigenfrequencies
(roots of the characteristic equation (1.23))
n
are mapped to the critical point
(-1 + 0i). Because the conformal mapping G() preserves handedness, the
critical point will lie to the left (right) of the image curve of the real axis if the
corresponding root
n
lies in the upper (lower) half of the complex -plane. The
situation shown in the Fig. 1.11. corresponds to an unstable mode.
These deliberations suggest the following modied Nyquist criterion: Consider
the image curve of the positive half of the real axis = 0 under the OLTF
mapping in the G()-plane, as shown in Fig. 1.11.. As one moves along the image
curve in the direction of increasing frequency , an eigenmode with eigenfrequency

n
is encountered each time the image curve passes the critical point (1 +0i) (In
Fig. 1.11., only one sweep past an eigenfrequency is shown). If the critical point
lies to the right of the image curve, the eigenmode is unstable, because then its
frequency
n
is located below the real axis in the -plane. On the other hand, if
the critical point is located to the left, the eigenmode is stable. If the image curve
passes through the critical point, the mode is neutrally stable.
In comparison to the classical Nyquist stability criterion, it is perhaps easier
to foster an intuitive understanding of the modied Nyquist rule. However, it is
admitted that no strict mathematical proof for the modied criterion is known. It
8
Incidentally, for this reason application of the Routh-Hurwitz criterion to thermo-acoustic sys-
tems is not possible.
STABILITY ANALYSIS 19

n
Figure 1.11. Conformal mapping G() of the real axis under the OLTF for an
unstable eigenfrequency.
has been validated successfully for a number of cases, where eigenfrequencies and
growth rates can be determined by solution of the characteristic equation (Polifke
et al. 1997; Sattelmayer and Polifke 2003a). Nevertheless, one must concede that
erroneous predictions may be obtained if the derivative of the OLTF vanishes for
some real-valued frequency R (the mapping is then not conformal). Further-
more, if an eigenfrequency has a large imaginary part such that the OLTF curve
passes the critical point at a large distance, the proposed criterion may fail because
conformality is a local property, i.e. it holds only in a nite-size neighborhood of
the point considered. Fortunately, very large growth rates with () 0 are not
observed in realistic network models, while very large damping rates with () 0
correspond to strongly damped modes, which are of no concern for the overall sta-
bility of a combustion system.
1.3.3.3 Identication of eigenfrequencies and growth rates from a Nyquist plot Con-
formality, i.e. the local preservation of angles under a mapping f : z f(z), implies
that an orthogonal grid of lines with constant real or imaginary part, respectively,
is mapped to an orthogonal grid of lines in the image plane (with the exception
of points where the derivative of f is zero). In other words, under a conformal
mapping, the neighborhood of any point is rotated and stretched or shrunk, as
illustrated in Fig. 1.11. (Janich 1983).
This interpretation of conformality implies that it is possible to estimate the
frequency of an eigenmode as well as its rate of growth or decay from the image
curve of the OLTF:
1. the real part of the eigenfrequency is approximately equal to the frequency
for which the distance between the image curve and the critical point attains
a local minimum.
2. the imaginary part (
n
) is approximately equal to the minimum distance
from the critical point to the OLTF image curve divided by the scaling factor
of the mapping.
The scaling factor can be roughly determined by evaluating how an interval ((
n
)
; (
n
)+), is mapped to a segment G((
n
)) G((
n
)+) of the
OLTF image curve. The scaling factor is then estimated as the arc length divided
by 2.
20 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
A more precise way of deducing the growth rate from the image curve has been
proposed by Sattelmayer and Polifke, based on the identity theorem of functional
analysis (Sattelmayer and Polifke 2003b). The theorem assures that a polynomial t
for the OLTF, which approximates the transfer function G() with good accuracy
for a range of purely real frequencies
1

2
, with ,
1
,
2
R will locally
approximate the OLTF also for complex-valued frequencies C.
This consideration suggests to determine the growth rate of an eigenmode as
follows: generate a polynomial t P
G
() = g
m

m
+. . . +g
0
which approximates the
OLTF curve close to the critical point, i.e. in the vicinity of an eigenmode
n
, see
Fig. 1.11.. Then determine with a numerical root nding algorithm the frequency

for which the approximating polynomial P


G
(

) = 1. If the eigenmode
n
is
not too far away from the real axis, then
n

.
It is remarkable that in this way complex-valued eigenfrequencies of an acoustical
system can be determined from the OLTF image curve, which is computed for
purely real frequencies R. This is very convenient when analytical expressions
for transfer matrix coecients are not known, which is usually the case for transfer
matrices or response functions determined from experiment.
In concluding this section we remark that with the proposed rule for interpre-
tation of Nyquist plots, it is obvious that the frequency at which the OLTF curve
crosses the real axis should not be identied with an eigenfrequency. Indeed, it has
been shown by example that this popular, but incorrect heuristic version of the
Nyquist criterion does lead to erroneous predictions (Lundberg 2002; Sattelmayer
and Polifke 2003a).
1.3.4 Acoustic energy budget
The physical interpretation of the Rayleigh criterion implies that correlated uctu-
ations of heat release and pressure generate acoustic energy, such that self-excited
thermo-acoustic instabilities may develop. On the other hand, loss of acoustic
energy through viscous dissipation, through transfer of acoustic energy to hydro-
dynamic or evanescent modes, or through radiation losses at the system boundary
tends to stabilize a thermo-acoustic system.
Stability considerations based on a balance between generation and loss of acous-
tic energy may appear to be an almost self-evident generalization of Rayleighs
criterion
9
. However, quantitative methods for stability analysis based on this idea
have been proposed only recently (Ibrahim et al. 2006, 2007; Nicoud and Poinsot
2005; Giauque et al. 2006, 2007).
Ibrahim et al. (2006, 2007) base their analysis of thermo-acoustic instability in a
generic premix combustor on an expression for the instantaneous density of acoustic
energy,
e =

2

2
+u

i
u

. (1.27)
The rst term is the potential acoustic energy, the second term (with summation
over repeated indices implied) the kinetic acoustic energy. Morfey (1971) suggests
that in the presence of mean ow, there should be additional terms of the form
9
Recall that the Rayleigh criterion is necessary, but not sucient for instability to occur,
because loss mechanisms are not taken into account.
STABILITY ANALYSIS 21
p

u
i
u

i
/c
2
a kind of convective acoustic energy density but these terms have
not been considered by Ibrahim et al.
The conservation of acoustic energy is then formulated as a transport equation
for the energy density e,
e
t
+

x
i
(p

i
) +

x
i
(eu
i
) = . (1.28)
The second and third term on the l.h.s. represent the ux of acoustic energy, with
convection by the mean ow u retained. The source term comprises thermo-
acoustic and viscous contributions,
=
1
c
2
p

i
x
j

ij
, (1.29)
where the former is obviously related to the Rayleigh integral.
Introducing an average over the oscillation period T,
. . . =
1
T

T
0
. . . dt,
integrating over the domain considered and employing the divergence theorem, the
energy balance can be written as follows

V
edV =

A
p

i
dA
i

A
eu
i
dA
i
+

V
dV. (1.30)
From the balance equation (1.30), an amplication coecient is then dened,

A
p

i
dA
i

A
eu
i
dA
i
+

V
dV.
2

V
edV
. (1.31)
The rst term in the numerator represents boundary work, the second term rep-
resents advection of acoustic energy across the boundary by the mean ow, the
third term is the net production (the thermo-acoustic source vs. the viscous sink).
Some of these contributions to the overall amplication increase the acoustic en-
ergy and are denoted
amp
, while the magnitude of contributions that represent a
loss of energy are denoted
damp
, such that for the overall amplication coecient
=

amp

damp
. Separation of destabilizing and damping eects in this
way allows to introduce a stabilizing factor
S() =

amp

damp
. (1.32)
Ibrahim et al. (2006, 2007) then argue that if an acoustic eigenfrequency
m
of the
combustor with S(
m
) > 1 exists, that eigenmode will be unstable.
The appealing feature of this method is that it should allow the combustor
designer to investigate quickly and inexpensively a wide variety of potential design
congurations and operating conditions. Validation studies carried out by Ibrahim
et al. (2006, 2007) met moderate success. It was possible to rationalize experimental
results to some extent. However, it was also found that results are very sensitive
against the model for the heat release employed.
22 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
Ibrahim et al. (2006, 2007) used in their validation studies rather simplistic as-
sumptions concerning the acoustic wave eld in the combustor, i.e. the acoustic
impedance at the location of heat release. Furthermore, some of the models pro-
posed for the dynamics of the heat release in a premix ame are quite unreasonable
in our opinion (although it is of course interesting to study the sensitivity of pre-
dictions against model details). In any case, acoustic impedance at the heat source
and the dynamics of the heat source are, as explained above, decisive for thermo-
acoustic stability. More dependable predictions with the energy-budget approach
proposed by Ibrahim et al. require more accurate models for the dynamics of the
heat source (i.e. its transfer function) as well as the combustor acoustics.
A conceptually related approach has been proposed by Nicoud and Poinsot
(2005); Giauque et al. (2006, 2007). It is suggested to derive the acoustic energy
budget from simulation data generated with large eddy simulation (LES). Indeed,
with computational uid dynamics a complete closure of the energy budget should
be possible, and all terms appearing in the budget may be analyzed in great de-
tail. However, Nicoud and Poinsot (2005) realized that this hitherto not available
level of detail leads to unexpected diculties: it is by no means obvious what the
appropriate denition of acoustic energy or uctuation energy should be. Giauque
et al. (2006) have identied additional and signicant energy density, ux, and
source terms, thereby generalizing the recent results of Nicoud and Poinsot (2005)
to include non-zero mean ow quantities, large amplitude disturbances, and varying
specic heats. The associated stability criterion is therefore signicantly dierent
from the Rayleigh criterion in several ways. The closure of the exact equation is
performed on an oscillating 2-D laminar ame. Results show that in this case the
general equation can be substantially simplied by considering only entropy, heat
release, and heat ux terms. The rst one behaves as source term whereas the
latter two dissipate the disturbance energy. Moreover, terms associated with the
non-zero baseline ow are found to be important for the global closure of the bal-
ance even though the mean Mach number is small. (Quotation from the abstract
of the paper).
The most recent results (Giauque et al. 2007) conrm the importance of the
entropy uctuations and the non-linear terms. However, it is also pointed out that
to account for these terms is analytically tedious and numerically challenging. It
remains to be seen whether stability analysis for systems of practical interest can
prot from these developments.
1.3.5 Suggestions for further reading
Stability criteria which are like the Rayleigh criterion based on relative phases
between uctuations of velocity, pressure, fuel concentration, heat release rate, en-
tropy, etc., have been developed and used by Richards and Janus (1998); Lawn
(2000); Polifke et al. (2001a). Such criteria are certainly helpful for qualitative
analysis of relevant feedback mechanisms, but one must be aware of their limita-
tions! As discussed above, phase-based criteria can be overly pessimistic regarding
the stability of a combustion system, because they cannot include the stabilizing ef-
fects of loss of acoustic energy. Furthermore, Polifke et al. (2001a) have shown that
it is not always possible to determine the relative phases involved with sucient
accuracy without a full network model of the combustion system.
SYSTEM MODELLING 23
In Fig. 1.6. a simple stability map is shown, i.e. the stability limits vs. the value
of the time lag . Similar maps can be constructed in many ways to illustrate how
the stability properties of a system depend on one or more system parameters. Of
course, one can do more than merely indicate parameter regions of stable/unstable
behavior. For example, one may plot frequency of limit cycle oscillation, instability
growth rates, mode switching, etc. as a function of time lags, spread of time lags,
combustor residence times, or Helmholtz resonator location, to name but a few
possible arrangements (Richards and Janus 1998; Lieuwen and Zinn 1998; Polifke
et al. 2001a; Flohr et al. 2001; Stow and Dowling 2001, 2003).
A comparison of the dynamic and diagrammatic approaches for system sta-
bility analysis has been carried out by Sattelmayer and Polifke (2003a). It is found
that the approach proposed by Polifke et al. (1997), which has been presented
above, produces the same results as the search for eigenmodes and eigenfrequen-
cies, while the earlier rules for interpretation of the Bode diagram (Deuker 1995;
Kr uger et al. 2000) are often in error.
1.4 SYSTEM MODELLING
Miscellaneous methods of stability analysis for thermo-acoustic systems have been
discussed in the previous section. It should have become apparent that stability
analysis in general requires an underlying system model, which represents the rel-
evant thermo-uid- dynamic and acoustic phenomena. A wide variety of system
models have been proposed for thermo-acoustic stability analysis, ranging from
simple estimates for phase lags based on convective time scales to CFD simulation
notably Large Eddy Simulation (LES). In this section, it is attempted to give an
overview of the most important types of system models, see also Fig. 1.12..
Some system models notably CFD-based approaches include sub-models for
all relevant processes as they may occur in (turbulent, reacting) compressible ow
in complicated geometries. However, due to the high numerical eort associated
with this brute force approach, a strategy of divide and conquer is often pre-
ferred, where the propagation (and dissipation, and reection) of acoustic waves is
modelled explicitly, while information on the dynamics of the heat source is sup-
plied as input to the model. This input data is obtained from experiment or CFD
models, and supplied to the system model in terms of a ame transfer function, or
a burner matrix, say.
1.4.1 Finite-Volume / Finite-Element models
1.4.1.1 Computational uid dynamics Modelling of unsteady (turbulent, reacting)
compressible ow is a powerful tool, particularly attractive for thermo-acoustic in-
stabilities because advanced CFD tools can capture in principle all relevant pro-
cesses. An introduction to unsteady ow modelling is given in companion lectures
of this lecture series by Nicoud (2007), the book by Poinsot and Veynante (2005)
is recommended for further reading.
An obvious drawback of this approach is the high computational cost, which is
certainly one of the reasons why it has been used only rarely until now (Hantschk
and Vortmeyer 1999; Murota and Ohtsuka 1999; Wall 2005; Schmitt et al. 2007).
24 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
Time Domain Frequency Domain
CFD
nonlinear PDEs -
Navier-Stokes
Computational Acoustics
linearized PDEs -
often extended Helmholtz
Galerkin Methods
ODE
Network Models
algebraic equations
Nonlinear Linearized Equations
F
i
n
i
t
e

V
o
l
u
m
e
,
F
i
n
i
t
e

E
l
e
m
e
n
t
M
o
d
e
-
B
a
s
e
d
Figure 1.12. Overview of models for thermo-acoustic systems.
If CFD is used for thermo-acoustic stability analysis, a signicant technical dif-
culty is the implementation of boundary conditions which enforce not only the
required mean and turbulent ow conditions, but also the appropriate acoustic
boundary conditions (impedance or reection coecient). See Selle et al. (2004);
Polifke et al. (2006); Schuermans et al. (2005); Huber et al. (2008) for recent pub-
lications on this problem.
A third shortcoming of the CFD approach is a point that is not always appre-
ciated by practitioners of that art: interpretation of CFD results in the thermo-
acoustic context is usually not straightforward! Insight into a thermo-acoustic in-
stability mechanism is often gained only after sophisticated post-processing of CFD
data, usually supplemented with acoustic analysis. See the publication of Martin
et al. (2006) for an example along these lines.
Note that brute force simulation of transient system behaviour in the presence
of small perturbations is not the only possible way of applying CFD to the study
of thermo-acoustic instabilities. CFD can also be used as an element in a divide
and conquer strategy. In such a framework, ow simulation is often used to deter-
mine the ame transfer function or a transfer matrix (Sklyarov and Furletov 1975;
Deuker 1995; Kr uger et al. 1998; Bohn et al. 1998; Polifke et al. 2001b; Polifke and
Gentemann 2004; Gentemann et al. 2004; Armitage et al. 2004; Gentemann and
Polifke 2007; Zhu et al. 2005; Giauque et al. 2008; Huber and Polifke 2008). It is
also possible to use ow simulation to compute the OLTF of a (sub-)system, which
is then used to generate a Nyquist plot (see section 1.3.3). Explicit determination
of a ame transfer function or burner transfer matrix is not required, which can
be an advantage for some system congurations (Kopitz and Polifke 2005, 2008;
Neunert et al. 2007).
SYSTEM MODELLING 25
1.4.1.2 Computational acoustics In this section, system models for thermo-acoustic
analysis that make use of a nite-element- or nite-volume-formulation of linearized
perturbation equations are discussed.
From the equations for conservation of mass and momentum in ow of an ideal
gas, the linearized Euler equations for small uctuations of velocity u

, pressure p

,
and density

can be derived

t
+u
i

x
i
+u

x
i
+
u

i
x
i
+

u
i
x
i
= 0, (1.33)
u

i
t
+u
j
u

i
x
j
+u

j
u
i
x
j
+
1

x
i

2
p
x
i
= 0, (1.34)
p

t
+u
i
p

x
i
+u

i
p
x
i
+ p
u

i
x
i
+ p

u
i
x
i
= ( 1) q

. (1.35)
Here q denotes a volumetric rate of heat release, is the ratio of specic heats.
The variables without apostrophe

represent a mean ow state, and may depend
on position x, but not on time.
One can show see e.g. Chu and Kovasznay (1958); Pierce (1981) that every
perturbation of the mean ow state described by this system of equations can be
interpreted as a superposition of three dierent modes: an acoustic mode, a vorticity
mode, and an entropy mode. If the mean ow is homogeneous and heat sources
are absent, then the three modes propagate independently from each other. In that
case, the acoustic mode represents propagation of pressure waves (with the speed
of sound), the vorticity mode represents production and transport of uctuations
of vorticity, the entropy modes describes production and transport of hot spots
(regions of temperature and density dierent from the mean value). The latter two
modes are transported convectively.
If the mean elds are not spatially homogeneous, coupling between the three
modes is to be expected. This can give rise, e.g., to low-frequency oscillations in
combustors, where so-called entropy waves are an important element of the thermo-
acoustic feedback loop (Keller et al. 1985; Polifke et al. 2001a; Eckstein 2004).
The linearized Euler equations Eqs. (1.33) (1.35) represent a very general
description of the generation and propagation or transport of small disturbances in
compressible ow. This set of coupled, partial dierential equations is although
linearized notoriously dicult to handle numerically. Therefore, the linearized
Euler equations are not often used for computational acoustics in the form shown
above .
Ewert and Schroder (2003) have shown how a set of acoustic perturbations equa-
tions (APEs) can be derived from from Eqs. (1.33) (1.35), where after suitable
variable decomposition, each equation represents one of the perturbation modes.
This formulation is a good basis to introduce simplications and isolate the phe-
nomena of interest in a specic problem setup.
In FE/FV-based system models for thermo-acoustic analysis, even simpler for-
mulations are typically employed. Often the vorticity mode is neglected completely,
while entropy waves are considered only in a rudimentary manner (if at all). For
example, Pankiewitz and Sattelmayer (2003a, b) have developed a nite-element-
based model of acoustic wave propagation in 3D geometries in the presence of heat
release with non-constant speed of sound and density. The model is based on a
26 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
0.8 1 1.2 1.4 1.6 1.8 2 2.2 2.4
0.2
0.3
0.4
0.5
0.6
0.7
0.8

f
(1,0,0) (1,1,0) (1,0,0)
Figure 5. FREQUENCIES AND CORRESPONDING MODE TYPES
FOR DIFFERENT DELAY TIMES. (,) LOW ORDER MODEL, (,)
TIME DOMAIN SIMULATION. (,) UNSTABLE MODE, (,) STABLE
MODE.
domain simulation, from the time evolution of the acoustic quan-
tities the frequency and the cycle increment (and thus the growth
rate) can be determined.
For a comparison of the two techniques we carried out cal-
culations for a model combustor with a geometry similar to g. 1
with twelve burners. We employed a simple heat release model
adopted from Dowling [8]. At one burner j this model reads
1

0
d q

j
dt
+ q

j
=u

B, j
(t ), j = 1, . . . , 12, (7)
where =

q/ u
B
is the ratio of the mean heat release rate to the
mean burner exit velocity and
0
is the cut-off frequency of the
low-pass lter. For the calculations with the low order model
eq. (7) is transfered to the frequency domain.
Figure 4 shows the frequencies (given in a non-dimensional
form) as a function of the delay time for a given value of
0
.
Additionally the type of the (unstable and least stable, resp.)
mode is given. Either the pure axial mode (1,0,0) or the rst
combined axial-circumferential mode (1,1,0) (see g. 4) is ob-
served in distinct regions for the value of .
Both the low-order model (model A) and the time domain
simulation predict the same type of mode for the same values
of . The corresponding frequencies are similar in both models,
though in the time domain simulation the values are somewhat
lower. However, in the low-order model the calculated level of
instability is generally higher. Therefore some modes which are
stable in the time domain simulation, are predicted to be unstable
in the low-order model.
The differences in frequency and stability can be attributed
to the sudden change of temperature from the plenum/burner
section to the combustion chamber. This discontinuity at the mo-
ment is treated differently in the two approaches. So already
the purely acoustical eigenfrequencies (not shown here) differ,
which affects the stability analysis. We are condent that a more
elaborate treatment of this problem will remove the discrepan-
cies and lead to a not only qualitatively but quantitatively good
agreement.
CONCLUSIONS
We have presented two novel techniques for the thermoa-
coustic analysis of annular combustors. The low-order model is
especially attractive for its speed of execution. The time-domain
simulation has the advantage that it can readily be applied to ar-
bitrary geometries and does not rely on any assumptions about
the propagation and coupling of the acoustic modes. First results
indicate that both approaches are suitable for a stability analysis
and have the potential to be developed to design tools.
REFERENCES
[1] S. Hubbard and A. P. Dowling. Acoustic Instabilities in Pre-
mix Burners. AIAA Paper AIAA-98-2272, 1998.
[2] T. Sattelmayer. Inuence of the Combustor Aerodynamics
on Combustion Instabilities from Equivalence Ratio Fluctu-
ations. ASME Paper 2000-GT-0082, 2000.
[3] F. E. C. Culick. Some Recent Results for Nonlinear Acous-
tics in Combustion Chambers. AIAA Journal, 32(1), pp.
146169, 1994.
[4] T. Murota and M. Ohtsuka. Large-Eddy Simulation of Com-
bustion Oxcillation in the Premixed Comubustor. ASME Pa-
per 99-GT-274, 1999.
[5] U. Kr uger, J. H uren, S. Hoffmann, W. Krebs, P. Flohr, and
D. Bohn. Prediction and Measurement of Thermoacous-
tic Improvements in Gas Turbines With Annular Combus-
tion Systems. Journal of Engineering for Gas Turbines and
Power, 123, pp. 557566, 2001.
[6] S. R. Stow and A. P. Dowling. Thermoacoustic Oscilla-
tions in an Annular Combustor. ASME Paper 2001-GT-0037,
2001.
[7] W. Krebs, G. Walz, and S. Hoffmann. Thermoacoustic Anal-
ysis of Annular Combustor. AIAA Paper AIAA 99-1971,
1999.
[8] A. P. Dowling. Nonlinear self-excited oscillations of a
ducted ame. Journal of Fluid Mechanics, 346, pp. 271
290, 1997.
4
Figure 1.13. Eigenfrequency of dominant modes in an annular combustor computed
with an FE (triangles) and a low-order model (circles), respectively. Filled symbols indicated
unstable modes. As the frequency increases, the non-plane (1,1,0) mode becomes dominant.
generalized wave equation for pressure uctuations
1
c
2
D
2
p

Dt
2


x
i

x
i

=
1
c
2
D q

Dt
. (1.36)
In low-Mach-number ow, the substantial derivative operator D. . . /Dt may be
replaced by the standard partial derivative w.r.t. time . . . /t. Then it is apparent
that Eq. (1.36) is a generalized Helmholtz equation, with spatial dependence of
speed of sound c and density taken into account, and a source term due to
uctuations of the heat release rate q

.
For the heat source, models based on time lags between the acoustic velocity
u

b
(computed from the gradient of pressure p

) at a reference position and the


momentary heat release rate in a certain region of the combustion chamber are
employed. In the simplest case, a formulation equivalent to the famous n model
is used
q

(x, t)
q(x)
= n
u

b
(t (x))
u
b
. (1.37)
More advanced formulations are discussed by Pankiewitz (2004). The interaction
index n which is a measure of the strength of the thermo-acoustic coupling and
the distribution of time lags (x) cannot by computed by the FE-model, but must
be determined by other means (experiment, CFD, . . . ) In this sense, the approach
of Pankiewitz and Sattelmayer (2003a, b) is one example of the divide and conquer
strategy mentioned above.
With a time-domain, nite-element based solver for Eq. (1.36), the investigation
of system stability is now in principle a straightforward task, if one follows the
strategy outlined in Section 1.3.1: rst a reference solution for Eq. (1.36) without
the source term is computed, then the response to small initial perturbations is
simulated. Results obtained for the eigenfrequency of the dominant mode in an an-
nular combustor are shown in Fig. 1.13.. This gure is taken from Pankiewitz et al.
(2001). It is seen that with varying time lag of the heat release model, dierent
plane (mode (1,0,0)) and non-plane (mode (1,1,0)) eigenmodes are dominant. As
expected, the stability of the modes also depends on the time lag.
SYSTEM MODELLING 27
the swirler. Since circumferential modes of the combustor are not purely azimuthal and comprise not only
a longitudinal component in the combustor but also in the plenum. Thus, through this component these
modes are inuenced by the introduction of the swirler as it can be seen on Fig. 2.6 and Fig. 2.7. For this
harmonics, the role of the swirler can be sgnicant particulary when acoustic ame is taken into account: in
the conguration without swirler, the peak of pressure is located in the ame zone so the Rayleigh criterion
[11] is potentially high whereas in the case with swirler, the amplitude of the mode is far less important in
this zone.
Z
X
Y
Figure 2.6: 1st circumferential mode of the combustor in the conguration without swirler, f 836hz, acoustic
pressure modulus.
Z
X
Y
Figure 2.7: 1st circumferential mode of the combustor in the conguration with swirler, f 753hz, acoustic
pressure modulus.
14
Figure 1.14. Mixed axial/azimuthal acoustic mode in an annular combustor. No
unsteady heat release (Courtesy of L. Benoit).
Note that local nonlinearities in the response of the heat source to large ampli-
tude oscillations can be implemented in the formulation of Pankiewitz et al.. If
dissipative eects are also included e.g. by imposing non-ideal acoustic bound-
ary conditions a limit cycle can then be observed as an asymptotic state of the
simulation. Of course, the formulation of physically reasonable, let alone accurate
models for non-linear eects and loss mechanism remains a formidable challenge.
Pankiewitz (2004) has also implemented a frequency-domain solver for Eq. (1.36)
in the presence of a harmonic forcing term. This allows to compute the response of
a thermo-acoustic resonator to an external excitation. From pressure distributions
obtained in this way, the transfer matrix of geometrically non-trivial elements can
be computed.
A frequency-domain formulation that determines eigenfrequencies
m
to assess
the stability see Section 1.3.2 of a thermo-acoustic system has been proposed by
Benoit et al. (Benoit 2005; Benoit and Nicoud 2005; Martin et al. 2004; Selle et al.
2006). A wave equation for small pressure perturbations (Poinsot and Veynante
2005) is considered,

2
p

t
2


x
i

c
2
p

x
i

= ( 1)
q

t
, (1.38)
which is for small Mach numbers and homogeneous mean pressure p equivalent
to Eq. (1.36). In the simplest case, the ame is modelled as a purely acoustic
element. Neglecting the eects of local turbulence, chemistry or heat loss, the heat
release rate responds to uctuations of velocity at a reference position, see again Eq.
(1.37). For the boundary conditions, the acoustic impedance must be expressed in
the form
1
Z
=

1

+
2
+
3
with C. (1.39)
Using a classical Galerkin nite element method to discretize Eq. (1.38) on a
nite element mesh with m nodes, one obtains an algebraic system of equations
[A][P] + [B][P] +
2
[P] = [D()][P]. (1.40)
28 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
Here [P] is the vector of unknowns p

, the matrix [A] represents the spatial-


derivative operator on the r.h.s. of Eq. (1.38), [B] represents the boundary terms.
If the source term due to unsteady heat release is neglected, [D] = 0, a suitable
variable transformation yields a problem of dimension 2m that is linear in the fre-
quency . This problem can be solved by a direct method, e.g. QR-based, or an
Arnoldi method. The latter is preferable, because only the rst few modes are
usually of interest.
If the source term on the r.h.s. of Eq. (1.38) is included, the resulting problem
cannot be solved by standard methods of linear algebra. Benoit and co-workers
propose two methods to nd eigenfrequencies
1. an expansion for weak thermo-acoustic interaction with an expansion pa-
rameter

1
V

V
n(x) dV.
Eigenmodes (
m
, p

m
) are sought as rst-order expansions around the modes
(
(0)
m
, p
(0)
m
) without heat release uctuations,

m
=
(0)
m
+
(1)
m
+ O(
2
), (1.41)
p

m
= p
(0)
m
+ p
(1)
m
+ O(
2
). (1.42)
After some algebra, an explicit expression for
(1)
in terms of the pressure
eld p
(0)
m
(x), the heat release uctuations q

(x) and the boundary condi-


tions is obtained. The imaginary part of
(1)
then indicates whether the
thermo-acoustic coupling destabilizes the system, see Benoit (2005); Benoit
and Nicoud (2005) for details.
A validation of this approach against a resonator with local, time-delayed heat
source again a generalized Rijke tube gives excellent agreement for very
weak thermo-acoustic coupling (interaction index n = 0.01 and expansion
parameter = 0.003). However, with coupling strength representative of
combustion, i.e. interaction index n = 5 and expansion parameter = 1.6
signicant deviations in the computed growth rates are observed.
2. The non-linear eigenvalue problem Eq. (1.40) can be solved iteratively for a
sequence of eigenfrequencies
(k)
m
, k = 1, 2, . . . from the quadratic problem

[A] [D(
(k1)
)]

[P] +
(k)
[B][P] + (
(k)
)
2
[P] = 0. (1.43)
Martin et al. (2004) report that usually less than 5 iterations are enough to
achieve convergence (starting from the homogeneous problem [D] = 0 for the
rst step of the iteration. Using this system model in combination with LES
results, Martin et al. (2004) discuss acoustic energy and modes in a turbulent
swirled combustor.
The computational acoustics models based on linearized PDEs for perturbations
in compressible ow with heat release presented in this section have not yet been
validated in a comprehensive, quantitative manner against experimental data. A
signicant diculty is the implementation of realistic models for the uctuating
heat release q

(x).
SYSTEM MODELLING 29
1.4.2 Galerkin method
In the remainder of this section, a system modelling approach that relies on a modal
representation of the acoustic eld is introduced: the Galerkin method. Another
noteworthy approach that makes use of a modal expansion is the state-space ap-
proach developed by Schuermans and co-workers (Schuermans et al. 2002, 2003;
Schuermans 2003). Although that approach appears to be quite exible, powerful
and ecient, it is not discussed further in these introductory notes.
The Galerkin method is a classical method, that can be interpreted as a par-
ticular variant of a weighted residual method (Fletcher 1991). The application of
Galerkin methods is by no means restricted to thermo-acoustic problems on the
contrary the method is applicable to a wide variety of dierential equations. The
presentation here follows the discussion of Culick (1989) and is restricted to a one-
dimensional situation for ease of presentation. Note, however, that the Galerkin
approach may very well be applied to geometries of applied interest, e.g. annular
gas turbine combustion chambers (Krebs et al. 1999). Other recent applications are
those of Deo and Culick (2005); Tyagi et al. (2007); Huang and Baumann (2007).
Consider wave propagation in the absence of a uctuating heat source, repre-
sented by the Helmholtz equation (like Eq. (1.36), but with zero right-hand side).
In the one-dimensional case with 0 < x < L and with constant speed of sound,
eigenmodes
m
(x) of the pressure perturbations satisfy the equation

m
x
2
+k
2
m

m
= 0, (1.44)
with wave number k
m
=
m
/c = m/L for a chamber open at both ends, p

= 0 (a
chamber closed at both ends, u

= 0, has the same set of possible wave numbers).


The solutions to this homogeneous problem are the well-known orthogonal modes,
i.e.

m
(x) = sin(k
m
x), (1.45)
where orthogonality implies

L
0

n
dx = E
2
m

mn
. (1.46)
The modes are orthogonal, because simply speaking the Laplace operator in the
Helmholtz-equation is Hermitian. More mathematical details on Sturm-Liouville
theory (the theory of linear dierential equations, dierential operators and their
eigenvalues / eigenfunctions) are found in many text books.
The essential feature of the Galerkin method is that the acoustic eld in the
presence of the heat source is expressed as superposition of these homogeneous
eigenmodes
m
,
p

(x, t) =

m
(t)
m
(x), (1.47)
u

(x, t) =

m

m
(t)
k
2
m
d
dx

m
(x). (1.48)
Substituting this expansion into the governing equations (1.36), multiplying with a
basis function
n
, integrating over the domain

L
0
, and exploiting the orthogonality
30 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
property (1.46), a system of ordinary dierential equations for the amplitudes
m
(t)
is obtained:
d
2

m
dt
2
+
2
m

m
= F
m
, (1.49)
with a forcing term
F
m
=
1
E
2
m

L
0
q

(x)
m
(x) dx. (1.50)
Note that there may be an additional term that stems from the boundary con-
ditions, see (Culick 1989). Also, non-linear terms, which may arise from local
non-linearities in the response of the heat source to large amplitude ow perturba-
tions, or from non-linear acoustics, can be integrated in the formulation. In that
case nonlinear terms, which may take the form

l
a
ml

l
or similar, appear in
Eq. (1.49).
From the Galerkin expansion, a system of ordinary dierential equations for
the amplitudes
m
(t) is obtained. Naturally then, stability analysis proceeds by
unsteady simulation as outlined in Section 1.3.1. Note that the shape and the
frequency of the dominant mode, which develops in the course of the simulation,
need not be equal (not even similar) to one of the eigenmodes
m
of the homoge-
neous problem. In general, an eigenmode of the inhomogeneous problem (with heat
source) is a collective phenomenon, to which many normal modes
m
contribute.
Especially in the case of nonlinear interactions, very complex behaviour with trans-
fer of energy between normal modes during the time evolution of the ODE (1.49)
is to be expected.
When complex geometries of technical interest an annular gas turbine combus-
tor are to be considered, the normal modes
m
can no longer be determined an-
alytically. Instead, they are determined by computational acoustics, usually based
on an FE formulation (and of course without the heat source term). The projection
of the partial dierential equation to generate the ODEs for the amplitudes (t),
which requires integration over the computational domain with the acoustic vari-
ables and the
m
s as integrands, is then also performed numerically (Krebs et al.
1999). One must make certain that from such a procedure indeed a normal set of
base functions
m
is produced. This property is often taken for granted but as
Benoit (2005) has shown, non-trivial boundary conditions are enough to lead to a
non-normal set of eigenmodes for the homogeneous problem!
1.5 NETWORK MODELS OF ACOUSTIC SYSTEMS
In section 1.3.2.1, a low order model of the Rijke tube has been constructed. This
model represents the Rijke tube as an assembly of three elements the upstream
duct, the gauze, the downstream duct plus the boundary conditions. Assuming
linear acoustics and harmonic time dependence exp{it}, a mathematical model
of the relevant thermo-acoustic processes is obtained in terms of a linear system of
equations, see Eq. (1.13). The unknowns are the (Fourier-transformed) acoustic
variables
10
of pressure p

and velocity u

at the junctions the ports of the


elements.
10
Instead of the primitive acoustic variables p

and u

, the so-called Riemann invariants f, g are


often used, but this is a matter of convention and convenience.
NETWORK MODELS 31
Many thermo-acoustic systems can be represented in an approximate manner as
such a network of acoustic elements, with each element corresponding to a particular
component of the system, e.g. a duct, a nozzle or orice, a burner, a ame, some
kind of acoustic termination, etc., see Fig. 1.8.. The coupling relations for the
unknowns across an element are combined into the transfer or scattering matrix of
the element. The transfer matrix coecients of all network elements are combined
into the system matrix S of the network. What has been done by hand for the
Rijke tube, can also be done by software, which is particularly useful if an network
comprises many elements. The numerical tools that build up a representation of
the system as an assembly of interconnected elements, construct the corresponding
system matrix from this network, and then solve the algebraic problem, are often
called network models or low-order network tools in the acoustics community.
In the EU Research and Training Network AETHER, TU M unchen is responsible
for System Modelling and Stability Analysis. The network tool taX will be
made available to network members. A structured training event is planned in
March 2008, where AETHER researchers will have the opportunity to learn how to
use the software. In the following, previous work on network models is reviewed and
suggestions for further reading are given. Then, a few simple examples of network
elements and results obtained with network models are discussed.
1.5.1 Review of previous work
An abundant literature exists on low-order models of acoustic wave propagation in
ducts, good introductions are found in Munjal (1986) and Poinsot and Veynante
(2005) (the latter with consideration also of combustion instabilities).
The use of low order models for thermo-acoustic stability analysis has been pio-
neered by Merk (1956). The low-order method has been applied to study combus-
tion instabilities in afterburners (Bloxsidge et al. 1988), Rijke tubes (Heckl 1988)
and lean premixed stationary gas turbines (Keller 1995). Validations of 1D low-
order models against experimental data from a single burner test rig have been
published by Bloxsidge et al. (1988) and Schuermans et al. (2000).
For the particular case of annular gas turbine combustors, the low-order method
has been extended to quasi 2D or 3D geometries. Under the guidance of J. J.
Keller, Curlier (1996) developed a numerical tool to study the propagation of acous-
tic waves in annular geometries in the presence of a mean ow. The underlying
approach to represent the acoustic wave eld goes back to Tyler and Sofrin (1962).
Kr uger et al. (1999) used a network of four-pole acoustic elements to investigate the
stability limits of azimuthal modes in a heavy-duty annular gas turbine combus-
tor. Their particular approach, however, seems to be limited inasmuch as it is not
capable of handling mixed modes see Fig. 1.16. below in a reasonable manner.
The interaction of entropy waves with acoustics in an annular combustor with
a choked exit has been investigated by Polifke et al. (1999), details of the method
and additional results have been published by Polifke et al. (2001a); Polifke (2004).
An extension of the method, capable of describing the eect of burner-to-burner
variability (broken symmetry) and mode coupling has been proposed and vali-
dated by Evesque and Polifke (2002). This is important, because Berenbrink and
Homann (2000) showed experimentally that judiciously introduced dierences be-
tween burners could be exploited as a means of passive control. The dierence
32 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
x
u
x
d
A
u
A
d
l
Figure 1.15. Acoustically compact contraction.
and signicance of spinning vs. standing modes in annular combustors has been
discussed by Evesque et al. (2003).
Bohn and Deuker (1993) were perhaps the rst to propose the implementation
of low-order methods as a network tool, where a library of acoustic elements is gen-
erated, each of them represented by its respective transfer matrix, and combined
in a exible and to some extent automated manner to represent the dynamic prop-
erties of complicated systems. Since then, network tools have been implemented
by the research groups at RWTH Aachen for Siemens Power Generation (Kr uger
et al. 1999), at ABB Corporate Research in Dattwil (Polifke et al. 1997, 2001a)
and more recently also at the University of Cambridge (Stow and Dowling 2001).
Stow and Dowling (2001, 2003) also implemented a low-order model for annular
combustors. Particular attention was paid to the boundary conditions appropriate
for a gas turbine combustor, otherwise the formulation is in essence identical to
the method proposed by Keller and co-workers above Keller (1995); Curlier (1996);
Polifke et al. (2001a).
A new, qualitatively dierent and seemingly quite powerful variant of the network
approach has been proposed recently by Schuermans et al. (2003); Schuermans
(2003), which makes use of a state space approach.
1.5.2 Compact element with inertia and loss
Consider the propagation of acoustic perturbations through a contraction as shown
in Fig. 1.15. with length l = (x
d
x
u
) such that kl 1 (acoustically compact).
A transfer matrix for such an element may be derived in an approximate manner
from momentum and mass conservation.
Conservation of momentum is exploited to derive a coupling relation between
acoustic perturbations p

, u

on both sides of the contraction by integrating the


unsteady Bernoulli equation (Hirschberg 2001):

x
i

t
+
u
2
2
+

1
p

= 0. (1.51)
along a streamline of the mean ow from x
u
to x
d
. First consider the term with
the velocity potential :

x
d
x
u

x
i
dx
i
=

t

x
d
x
u
u
i
dx
i


t
u
u

x
d
x
u
A
u
A(x)
dx. (1.52)
The approximation u(x) u
c
A(x) is valid because it was assumed that the el-
ement is acoustically compact then the ow through the element is eectively
incompressible.
NETWORK MODELS 33
Introducing an extended length
l
ext

x
d
x
u
A
u
A(x)
dx, (1.53)
and assuming harmonic time-dependence, one obtains:

x
d
x
u

x
i
dx
i
= i l
ext
u

u
. (1.54)
The remaining two terms in (1.51) are simply evaluated at x
u
and x
d
. Linearization
and dividing by c yields
i k l
ext
u

u
+

M u

+
p

d
u
= O(kl)
2
, (1.55)
with [v]
d
u
v
d
v
u
for any variable v. This result is exact to rst order in the
Helmholtz number k l, because the time derivative in (1.52) introduces a factor k l,
so that any deviations of u(x) from u
c
A(x) appear as second order terms in (1.55).
Neglecting terms of rst order in Mach number, this result may be simplied
p

d
=
p

u
i k l
ext
u

u
+ O(kl)
2
+ O(M). (1.56)
Note that l
ext
may be signicantly larger than the geometrical length l of the
element if the area A(x) < A
u
, A
d
between the up- and downstream terminations.
Mass conservation between upstream (u) and downstream (d) demands that
d
dt

dV +

u
i
dA
i
= 0. (1.57)
For the geometry considered (see Fig. 1.15.) and assuming quasi-one-dimensional
ow, this can be simplied as follows
d
dt

x
d
x
u
(x) A(x) dx + [uA]
d
u
= 0. (1.58)
For the density,

(x) =

u
+O(kl) and with the assumed harmonic time dependence
i

x
d
x
u
A(x) dx + [(

u + u

) A]
d
u
= O(kl)
2
. (1.59)
Diving by the mean density and dening a reduced length
l
red

x
d
x
u
A(x)
A
d
dx (1.60)
one obtains:
i k l
red
A
d
p

u
+

+M
p

d
u
0. (1.61)
Neglecting terms of rst order in M, this simplies further to
u

d
A
d
= u

u
A
u
i k l
red
A
d
p

u
+ O(kl)
2
+ O(M). (1.62)
34 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
Finally, the transfer matrix for a compact acoustic element with loss is approxi-
mated as follows:

c
u

d
=

1 i k l
e
M
i k l
red

c
u

u
(1.63)
This is a combination of (1.56) and (1.62), with an area ratio A
u
/A
d
and a
pressure loss term M. The loss term M has been introduced to account in an
approximate manner for loss of acoustic energy due to, e.g., ow coupling. With
this term included, a change in volume ow rate u

A through the element results


in a change in pressure drop p

u
p

d
across the element.
The loss coecient for the acoustic pressure that appears in Eq. (1.63)
is in general not equal to the hydrodynamic loss coecient
d
for the total
head. It is fortunately not dicult to establish a relation between the two
loss coecients, which may be expected to hold with reasonable accuracy a)
in the quasi-stationary limit and b) if the relevant loss mechanisms ow and
acoustics are the same.
Consider Bernoullis equation from position u upstream to position d
downstream with a loss term:
p
d
+

2
u
2
d
= p
u
+

2
u
2
u

d
u
2
d
. (1.64)
By convention, the loss is expressed in terms of the downstream velocity u
d
.
Continuity implies that u
d
= u
u
with = A
u
/A
d
, and the downstream
pressure p
d
can be expressed in terms of upstream pressure p
u
and velocity
u
u
as follows:
p
d
= p
u
+

2
u
2
u
(1
2
(1 +
d
)). (1.65)
Linearization and division by c yields
p

d
c
=
p

u
c
[
2
(1 +
d
) 1] M
u
u

d
. (1.66)
The term in angular[. . .] brackets appears as the acoustic loss coecient in
Eq. (1.63).
For a sudden change in ow cross-sectional area, conservation of momentum
implies for the loss coecient
d
= (1 1/)
2
(neglecting the vena contracta
eect for contractions). Then = 2( 1), which is small and negative for
0 < < 1 and increases rapidly for area ratios > 1 (contractions).
The eective length introduced in (1.63) combines the extended length see
the denition (1.53) with end corrections l
ec
, which may appear at sharp
corners,
l
e
l
ext
+l
ec,u
+l
ec,d
(1.67)
For an element with signicant intermediate contraction, A(x) < A
u
, A
d
over some length, the following inequalities should hold:
l
red
< l < l
ext
< l
e
.
NETWORK MODELS 35
(a) Axial mode (1,0) (b) Azimuthal mode (0,1) (c) Mixed mode (1,1)
Figure 1.16. Plane and higher order eigenmodes in a thin annular duct.
Physically, the eective length accounts for the inertia eects, i.e. a change in
pressure dierence p

u
p

d
will accelerate the uid between the two reference
positions u and d, resulting in a gradual change of velocity u

through
the element, which leads to a phase dierence between p

u
p

d
and u

.
An expression similar to (1.63) has been introduced by Paschereit and Polifke
(1998) to approximate the transfer matrix of a premix burner (without ame). It
turns out that this form is indeed quite general and applicable to all sorts of compact
elements, e.g. sudden changes in cross sectional area, an orice, etc. (Gentemann
et al. 2003). Computational uid dynamics may be used to determine the values
of the coecients appearing in (1.63), see e.g. Flohr et al. (2003)
1.5.3 Ducts
In this section, we present transfer matrices for a few non-trivial duct elements.
1.5.3.1 Thin annular duct The equation (1.9) given above for wave propagation
in a duct with wave number k = /c is valid only for plane waves without mean
ow. In general, non-plane waves (higher-order modes) may propagate in a duct
at suciently high frequency (above the cut-o frequency). Of particular interest
for gas turbine combustion are thin annular ducts, with the height of the annulus
signicantly smaller than the radius R or the length L, such that the amplitude of
acoustic waves depends on the axial coordinate x and the circumferential position
, but not on the radius r,
p

exp (it ik
x
x ik

R) , (1.68)
and similarly for velocity u

or the Riemann invariants (but see the discussion of


standing vs. rotating modes by Evesque et al. (2003)). Dierent mode shapes in
such a thin annulus are shown in Fig. 1.16.. For the azimuthal or perpendicular
component of the wave vector, k

m/R with mode index m, because the wave


eld must be periodic in the azimuthal direction. If one plugs the expression (1.68)
into the 2-D form of the convective wave equation,
(

t
+u )
2
p

c
2

2
p

= 0, (1.69)
36 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
one obtains the following condition for the axial component of the wave vector
k
x

=
/c
1 M
2

/c

2
(1 M
2
)

. (1.70)
Here M U/c is the mean ow Mach number for uniform mean ow in the axial
direction with velocity U. The transfer matrix for non-plane wave propagation in
a thin annular duct with mean ow is similar in form to the previous result (1.9):

f
d
g
d

e
ik
x+
l
0
0 e
ik
x
l

f
u
g
u

. (1.71)
Note, however, that the relation between Riemann invariants and the primitive
acoustic variables is a bit more complicated,
p

c
= f +g, u

=
+
f +

g. (1.72)
with a coecient


k
x

k Mk
x

. (1.73)
For zero Mach number and plane waves m = M = 0, this reduces to the previous
results (1.9) with k
x

= /c.
With this formalism, it is possible to describe eciently the acoustic wave eld
in a modern gas turbine with an annular combustor from the compressor exit to
the turbine inlet, say.
1.5.3.2 Ducts with varying cross-sectional area A duct with varying cross sectional
area also an annular duct with curved walls, like a gas turbine combustion chamber
can be eectively modelled by stair-stepping the contour of the duct, i.e. as
an alternating succession of short straight duct elements and small discontinuous
changes in cross sectional area. The transfer matrices of these building blocks
have been introduced above (a change in cross-sectional area can be described as a
compact element as in 1.5.2), the transfer matrix of the complete duct is obtained
as the matrix-product of the individual transfer matrices. Note that in general for
each sub-element, the cross-sectional area A and the radius R may be dierent,
therefore the wave vector (k
x
, k

), the Mach number and the area ratio will not be


constant. Experience shows that geometries of applied interest can be represented
with sucient accuracy with about one dozen of sub-elements.
1.5.3.3 Joints and forks In networks with non-trivial topology, c.f. Fig. 1.8.,
elements are needed to describe a situation where two ow paths are joined into
one, or one ow path splits into two. In principle, this requires a 3 3 transfer
matrix to interconnect the six Riemann invariants at the three ports of such an
element. In the simplest case, one may assume equal pressures, i.e. p

i
= p

j
= p

k
and conservation of mass, i.e. u

i
A
i
+ u

j
A
j
+ u

k
A
k
= 0 to construct the transfer
matrix. Of course, if end corrections, loss coecients, etc. must be considered, the
coupling relations are more complicated, and one has to take into account mean
ow direction but there is no essential diculty in constructing transfer matrices
for such elements.
NETWORK MODELS 37
1.5.4 Transfer matrix of a compact ame
In this section a closure formulation for the transfer matrix of a compact ame (more
generally speaking: a compact source of heat) is presented. Compact means
that the heat release is concentrated in a region much smaller than acoustic wave
lengths. In this approximation, a premix ame front is considered as a discontinuity
of negligible thickness, which adds a certain amount of heat q per unit mass (units
J/kg = m
2
/s
2
) to the ow of an ideal gas.
For steady ow of gas, Chu (1953) has shown that conservation of energy, mass
and momentum across a thin source of heat leads to the following Rankine-Hugoniot
relations:
c
2
h
= c
2
c
+ ( 1) q + O(M
2
), (1.74)
u
h
= u
c
+ ( 1) M
c
q
c
c
+ O(M
2
), (1.75)

p
c

h
=

p
c

c
( 1) M
c
q
c
c
+ O(M
2
), (1.76)
with
h
c
h
/
c
c
c
. Here the index c stands for the cold (upstream) side of the
discontinuity, h stands for the hot (downstream) side A detailed derivation of
these relations (using the present notation) is given in (Polifke et al. 2001a). A
useful result follows from Eq. (1.74):
T
h
T
c
1 = ( 1)
q
c
2
c
. (1.77)
The linearization of the Rankine-Hugoniot relations (1.74) - (1.76) in the presence
of (small) acoustic uctuations shows how uctuations of pressure and velocity up-
and downstream of the ame are related to uctuations of the heat release rate:

h
=

T
h
T
c
1

u
c
M
c

c
u
c
+

Q

, (1.78)
u

h
= u

c
+

T
h
T
c
1

u
c

c
p
c

. (1.79)
where

Q u q is dened as rate of heat addition per unit area, and terms of order
M
2
or higher are neglected. Again, a detailed derivation of these results is found
in (Polifke et al. 2001a).
Eqns. (1.78) and (1.79) do not yet allow to write down the transfer matrix of the
ame, which would relate acoustic perturbations on both sides of the ame sheet to
each other, because it is not yet clear how acoustic waves inuence the uctuations

of the heat release rate we have a closure problem! In general, the heat release
rate

Q

may respond to perturbations of pressure or velocity at the ame front, or


at the burner mouth, or at the location of fuel injection, as discussed above. A
quantitative model for the ame response,

=

Q

(u

h
, u

c
, u

i
, . . . , p

i
, . . . ,

; L, M, , S
L
, . . .) , (1.80)
is obviously needed to achieve closure, i.e. formulate a consistent network model.
As indicated, the heat release rate will in general not only depend on velocity
38 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
Sheet1
Page 1
refin_duct_areachange_duct_refout, A2/A1-1/3, M-0.1, loss_fact-0
refin_duct_areachange_duct_refout, A2/A1-3, M-0.1, loss_fact-0
inMachCorr_duct_areachange_duct_outMachCorr, A2/A1-3, M-0.1, loss_fact-(1-A2/A1)
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
n-0
n-1
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
0.2
0.4
0.6
0.8
1
n-0
n-1
x/L
G
r
o
w
t
h

r
a
t
e
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
1
1.2
1.4
1.6
1.8
2
x/L
G
r
o
w
t
h

r
a
t
e
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.75
0.8
0.85
0.9
0.95
1
x/L
G
r
o
w
t
h

r
a
t
e
Sheet1
Page 1
refin_duct_areachange_duct_refout, A2/A1-1/3, M-0.1, loss_fact-0
refin_duct_areachange_duct_refout, A2/A1-3, M-0.1, loss_fact-0
inMachCorr_duct_areachange_duct_outMachCorr, A2/A1-3, M-0.1, loss_fact-(1-A2/A1)
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
n-0
n-1
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
0.2
0.4
0.6
0.8
1
n-0
n-1
x/L
G
r
o
w
t
h

r
a
t
e
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
1
1.2
1.4
1.6
1.8
2
x/L
G
r
o
w
t
h

r
a
t
e
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.75
0.8
0.85
0.9
0.95
1
x/L
G
r
o
w
t
h

r
a
t
e
Figure 1.17. Frequencies (left) and growth factors (right) of rst two eigenmodes in a
duct with area change vs. position of the area change. Loss coecient = 0, boundary
reection coecients r = 1 at both ends.
uctuations u

c
, u

h
just up- and downstream of the ame, but may also involve
perturbations of velocity, or pressure, or equivalence ratio elsewhere in the system.
As indicated the ame response also depends on length scales (of the burner / the
ame), Mach numbers, equivalence ratios, ame speeds, etc.
In the simplest case, the ame reponse may be described in terms of a transfer
function that depends on the upstream (cold) velocity u

c
,
F()

/

Q
u

c
/u
c
. (1.81)
Then the corresponding transfer matrix for a compact ame follows from the lin-
earized Rankine-Hugoniot relations Eqns. (1.78) and (1.79);
T =

c
c
c

h
c
h

T
h
T
c
1

M
h
(1 +F())

T
h
T
c
1

M
c
1 +

T
h
T
c
1

F()

(1.82)
1.5.5 Example network calculations
To conclude, we present here a result of a very simple network, which nevertheless
produces non-trivial results. The conguration considered consists of the following
elements:
open end duct area change duct open end
For the ducts, a transfer matrix as given in Eq. (1.9) or Eq. (1.71) is used. A Mach
number M = 0.1 is assumed, mean ow eects on wave propagation are taken into
account by setting the wave numbers k
x
= /c(1 M) (this follows from Eq.
(1.70) for plane waves with k

= 0.) ). The area change is an expansion with


area ratio = A
u
/A
d
= 1/3, it is described as a compact element according to
Eq. (1.63). For simplicity the eective lengths are set to zero for the example
considered.
Fig. 1.17. shows frequencies and growth factors (computed as exp(i ()) of
the rst two eigenmodes with a loss coecient = 0. It should come as a surprise
that growth factors larger than unity are predicted. Clearly, there is no thermo-
acoustic engine in this conguration so what is the source of acoustic energy
that is responsible for the predicted growth of amplitudes?
ACKNOWLEDGMENTS 39
The answer is that an overly simplistic model for the boundary conditions was
used: If p

= 0 is assumed at the open end boundaries, the amplitudes of the


reected characteristic wave amplitude g at the downstream end, say, is equal to
the amplitude of the outgoing wave amplitude f. However, this boundary condi-
tion does not conserve acoustic energy, because the convective transport of energy,
represented by the factors (1 M) in the denition of the wave number, is not
considered. So, a boundary condition with g = f actually loses acoustic energy.
In the present conguration with an expansion placed between two open ends, the
convective loss of acoustic energy at the downstream end (where the Mach number
is smaller) is smaller than the inow of acoustic energy at the upstream end (where
the Mach number is larger). So in sum energy is added to the system.
A boundary condition that conserves acoustic energy can be derived from Bernoullis
equation,
p +
1
2
u
2
= p

. (1.83)
Here the l.h.s. stands for the conditions at the open end, the r.h.s. for the
surrounding. Assuming that there are no uctuations of the surrounding pressure
p

, linearization of Eq. (1.83) yields to rst order in Mach number


p

c
+Mu

= 0 or (1.84)
f(1 +M) +g(1 M) = 0. (1.85)
If this boundary condition is used, growth factors equal to unity (constant oscilla-
tion amplitude) are computed if the loss coecient is also set to zero (not shown).
With a non-zero loss coecient, the growth factorss are always less than unity, i.e.
the eigenmodes are damped, see Fig. 1.18.. The results indicate furthermore that
the losses vanish whenever the area change, which is the only element where losses
may occur in the present conguration, is located at a velocity anti-node with
u

= 0. This is close to the center of the tube at x/L 0.5 for the fundamental
mode, and at x/L 0.25 and x/L 0.75 for the second mode.
It is also interesting to discuss the eect that the position of the area change
has on the value of the frequency of the eigenmodes, see the right graph in Figs.
1.17. and 1.18.. The frequencies in these plots are normalized with the frequency
f
0
of the fundamental mode in a duct of length L with constant cross-sectional
area and two open ends. For the fundamental mode, the area change reduces the
eigenfrequency of the ducts with area change for x/L < 0.5, while it increases the
eigenfrequency for x/L > 0.5. This can be explained by considering ignoring for
the sake of the argument the loss term how the acoustic variables u

, p

on both
sides are related to each other, and what this implies for the characteristic wave
amplitudes f, g. Details are left to the reader it is useful to think of the variables
u

, p

, f, g as pointers in the complex plane!


1.6 ACKNOWLEDGMENTS
Thanks to doctoral students Jan Kopitz, Robert Leandro, and Andreas Huber.
Discussions with Raman Sujith on non-normal modes in thermo-acoustics are ap-
preciated. My apologies to Bruno Schuermans for not discussing his state space
approach time and space did not permit this.
40 THERMO-ACOUSTIC SYSTEM MODELLING AND STABILITY ANALYSIS CONVENTIONAL APPROACHES
Sheet1
Page 1
refin_duct_areachange_duct_refout, A2/A1-1/3, M-0.1, loss_fact-0
refin_duct_areachange_duct_refout, A2/A1-3, M-0.1, loss_fact-0
inMachCorr_duct_areachange_duct_outMachCorr, A2/A1-3, M-0.1, loss_fact-(1-A2/A1)
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
n-0
n-1
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
0.2
0.4
0.6
0.8
1
n-0
n-1
x/L
G
r
o
w
t
h

r
a
t
e
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
1
1.2
1.4
1.6
1.8
2
x/L
G
r
o
w
t
h

r
a
t
e
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.75
0.8
0.85
0.9
0.95
1
x/L
G
r
o
w
t
h

r
a
t
e
Sheet1
Page 1
refin_duct_areachange_duct_refout, A2/A1-1/3, M-0.1, loss_fact-0
refin_duct_areachange_duct_refout, A2/A1-3, M-0.1, loss_fact-0
inMachCorr_duct_areachange_duct_outMachCorr, A2/A1-3, M-0.1, loss_fact-(1-A2/A1)
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
n-0
n-1
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0
0.2
0.4
0.6
0.8
1
n-0
n-1
x/L
G
r
o
w
t
h

r
a
t
e
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.6
0.8
1
1.2
1.4
1.6
1.8
2
2.2
x/L
E
i
g
e
n
f
r
e
q
u
e
n
c
i
e
s

f
/
f
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
1
1.2
1.4
1.6
1.8
2
x/L
G
r
o
w
t
h

r
a
t
e
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
0.75
0.8
0.85
0.9
0.95
1
x/L
G
r
o
w
t
h

r
a
t
e
Figure 1.18. Frequencies (left) and growth factors (right) of rst two eigenmodes in a
duct with area change vs. position of the area change. Loss coecient = 4, boundary
reection coecient r = (1 M)/(1 M).
1.7 APPENDIX: CHARACTERISTIC EQUATION OF THE RIJKE TUBE
The following lines of Mathematica code determine the determinant of the Rijke
Tube network model discussed in this lecture:
ClearAll[A,nEt,deta,"@"];
A = {
{-1,1,0,0,0,0,0,0}, (* left b.c. *)
{0,0,1,1,-xi,-xi,0,0},(* pressure coupling across heat source *)
{0,0,(1+nEt),-(1+nEt),-1,1,0,0}, (* velocity coupling across heat source *)
{0,0,0,0,0,0,-1,-1}, (* right b.c. *)
{e1m,0,-1,0,0,0,0,0},(* next four lines: ducts *)
{0,e1p,0,-1,0,0,0,0},
{0,0,0,0,e2m,0,-1,0},
{0,0,0,0,0,e2p,0,-1}} ;
MatrixForm[A] (* this should output the 8x8 system matrix *)
deta = Det[A]//Simplify (* compute the matrix, simplify *)
(* insert phase propagators for ducts:
*)
e1p = Exp[I k1]; e1m = Exp[-I k1 ];
e2p = Exp[I k2]; e2m = Exp[-I k2 ];
deta = ComplexExpand[deta]//Simplify (* voila *)
These commands should produce rst a nice 8 8 system matrix and then the
following output statement:
Out[...]= e1m (-e2p (-1 + xi + nEt xi) + e2m (1 + xi + nEt xi)) + ...
+ e1p (- e2m (-1 + xi + nEt xi) + e2p (1 + xi + nEt xi))
Out[...]= 4 Cos[k1] Cos[k2] - 4 (1+nEt) xi Sin[k1] Sin[k2]
The latter is obviously the determinant of the system matrix (up to an irrelevant
factor of 4).
1.8 FINDING ROOTS
The following bit of Mathematica code searches for the eigenfrequency of the fun-
damental mode m = 0 of the Rijke tube. First values are assigned to lengths and
physical parameters (in arbitray units), then the coecients k1, k2, nEt so far left
undened in the variable deta are assigned.
lc = 0.5; lh = 0.5;
cc = 1.; ch = 1.0;
xi = 1./ch; n = (ch/cc)^2-1;
n=0.1;
tau = 1.5 Pi; (* most unstable for ground mode m = 0*)
k1 = Pi/2. omega lc / cc;
k2 = Pi/2. omega lh / ch;
nEt := n Exp [ - I omega tau ];
FindRoot[deta==0.,{omega,1}]
Out[...[] = {omega -> 0.998921 - 0.0278638 i}
Bibliography
Armitage, C. A., Riley, A. J., Cant, R. S., Dowling, A. P., and Stow, S. R. (2004).
Flame transfer function for swirled LPP combustion from experiments and
CFD. In Intl Gas Turbine and Aeroengine Congress & Exposition, number
ASME GT2004-53831, Vienna, Austria.
41
42 BIBLIOGRAPHY
Balasubramanian, K. and Sujith, R. I. (2007a). The role of nonnormality in com-
bustion interactions in diusion ames. In 45th Aerospace Sciences Meeting
and Exhibit, number AIAA-2007-0567, Reno, Nevada, USA. (accepted for
publication in Journal of Fluid Mechanics).
Balasubramanian, K. and Sujith, R. I. (2007b). Thermoacoustic instability in a Ri-
jke tube - nonnormality and nonlinearity. In 13th AIAA/CEAS Aeroacoustics
Conference, number AIAA 2007-3428, Rome, Italy.
Benoit, L. (2005). Prediction Des Instabilites Thermoacoustiques Dans les Tur-
bines a Gaz. Dissertation, Universite Montpellier II, Sciences et Techniques du
Languedoc.
Benoit, L. and Nicoud, F. (2005). Numerical assessment of thermo-acoustic insta-
bilities in gas turbines. Int. J. of Numerical Methods in Fluids, 47:849855.
Berenbrink, P. and Homann, S. (2000). Suppression of Dynamic Combustion In-
stabilities by Passive and Active Means. In Intl Gas Turbine and Aeroengine
Congress & Exposition, ASME paper, number 2000-GT-0079, Munich, Ger-
many.
Bloxsidge, G., Dowling, A., Hooper, N., and Langhorne, P. (1988). Active control
of reheat buzz. AIAA J., (Bloxsidge88):783790, volume 26.
Bohn, D. and Deuker, E. (1993). An acoustical model to predict combustion driven
oscillations. In 20th Intl Congress on Combustion Engines, number G20 in ,
London, UK. CIMAC.
Bohn, D., Deutsch, G., and Kr uger, U. (1998). Numerical Prediction of the Dynamic
Behaviour of Turbulent Diusion Flames. J. Eng. for Gas Turbines and Power,
(120):713720.
Chu, B. T. (1953). On the Generation of Pressure Waves at a Plane Flame Front.
In 4th Symposium (International) on Combustion, pages 603612.
Chu, B. T. and Kovasznay, L. S. G. (1958). Non-linear interactions in a viscous
heat-conducting compressible gas. J. Fluid Mech., (3):495514.
Culick, F. E. C. (1989). Combustion Instabilities in Propulsion Systems. In Hersh,
A. S., Catton, I., and Keltie, R. F., editors, Combustion Instabilities Driven by
Thermo-Chemical Acoustic Sources, number NCA-Vol. 4, HTD-Vol.128, pages
3352. ASME, ASME.
Culick, F. E. C. (1994). Some recent results for nonlinear acoustics in combustion
chambers. AIAA Journal, 32(1):146169.
Curlier, J. (1996). Acoustic feedback program. Masters thesis, Univ. of Washing-
ton, Dept. of Aeronautics & Astronautics.
Deo, N. A. S. and Culick, F. E. C. (2005). Reduced-order modeling and dynamics
of nonlinear acoustic waves in a combustion chamber. Combust. Sci. Tech.,
177(2):221 248.
BIBLIOGRAPHY 43
Deuker, E. (1995). Ein Beitrag zur Vorausberechnung des akustischen Sta-
bilitatsverhaltens von Gasturbinen-Brennkammern mittels theoretischer und
experimenteller Analyse von Brennkammerschwingungen. PhD thesis, RWTH
Aachen.
Dowling, A. P. (1995). The calculation of thermoacoustic oscillation. J. of Sound
and Vibration, 180:557581.
Dowling, A. P. (1997). Nonlinear Self-Excited Oscillations of a Ducted Flame. J.
Fluid Mech., 346:271290.
Drazin, P. G. (2002). Introduction to Hydrodynamic Stability. Cambridge University
Press.
Eckstein, J. (2004). On the Mechanisms of Combustion Driven Low-Frequency
Oscillations in Aero-Engines. PhD thesis, TU M unchen.
Evesque, S. and Polifke, W. (2002). Low-Order Acoustic Modelling for Annular
Combustors: Validation and Inclusion of Modal Coupling. In Intl Gas Tur-
bine and Aeroengine Congress & Exposition, number ASME GT-2002-30064,
Amsterdam, NL.
Evesque, S., Polifke, W., and Pankiewitz, C. (2003). Spinning and azimuthally
standing acoustic modes in annular combustors. In 9th AIAA/CEAS Aeroa-
coustics Conference, number AIAA 2003-3182, Hilton Head, S.C., U.S.A.
Ewert, R. and Schroder, W. (2003). Acoustic perturbation equations based on
ow decomposition via source ltering. Journal of Computational Physics,
188(2):365398.
Flandro, G. A., Fischbach, S. R., and Majdalani, J. (2007). Nonlinear rocket
motor stability prediction: Limit amplitude, triggering, and mean pressure
shift. Phys. Fluids, 19(094101).
Fletcher, C. (1991). Computational Techniques for Fluid Dynamics, volume I.
Springer-Verlag.
Flohr, P., Paschereit, C. O., and Bellucci, V. (2003). Steady CFD analysis for gas
turbine burner transfer functions. In 41st AIAA Aerospace Sciences Meeting
& Exhibit, Reno, NV. AIAA.
Flohr, P., Paschereit, C. O., van Roon, B., and Schuermans, B. B. H. (New Orleans,
2001). Using CFD for time-delay modeling of premix ames. Intl Gas Turbine
and Aeroengine Congress & Exposition, ASME 2001-GT-376.
Gentemann, A. and Polifke, W. (2007). Scattering and generation of acoustic energy
by a premix swirl burner. In Intl Gas Turbine and Aeroengine Congress &
Exposition, number ASME GT2007-27238, Montreal, Quebec, Canada.
Gentemann, A. M. G., Fischer, A., Evesque, S., and Polifke, W. (2003). Acoustic
transfer matrix reconstruction and analysis for ducts with sudden change of
area. In 9th AIAA/CEAS Aeroacoustics Conference, number AIAA 2003-3142,
Hilton Head, S.C., U.S.A. AIAA.
44 BIBLIOGRAPHY
Gentemann, A. M. G., Hirsch, C., Kunze, K., Kiesewetter, F., Sattelmayer, T.,
and Polifke, W. (2004). Validation of ame transfer function reconstruction for
perfectly premixed swirl ames. In Intl Gas Turbine and Aeroengine Congress
& Exposition, number ASME GT-2004-53776, Vienna, Austria.
Giauque, A., Nicoud, F., and Brear, M. (2007). Numerical assessment of stability
criteria from disturbance energies in gaseous combustion. In 13th AIAA/CEAS
AeroAcoustics Conference, number AIAA-2007-3425, Rome, Italy.
Giauque, A., Poinsot, T., Brear, M., and Nicoud, F. (2006). Budget of disturbance
energy in gaseous reacting ows. In Proc of the Summer Program, Stanford,
CA, USA. Center for Turbulence Research, NASA Ames/Stanford Univ.,.
Giauque, A., Poinsot, T., and Nicoud, F. (2008). Validation of a Flame Tranfer
Function Reconstruction Method for Complex Turbulent Congurations. In
14th AIAA/CEAS Aeroacoustics Conference (29th AIAA Aeroacoustics Con-
ference), number AIAA-2008-3046, Vancouver, Canada. AIAA/CEAS.
Hantschk, C. and Vortmeyer, D. (1999). Numerical simulation of self-excited ther-
moacoustic instabilities in a rijke tube. J. of Sound and Vibration, 227(3):511
522.
Heckl, M. (1988). Active Control of the Noise from a Rijke tube. J. of Sound and
Vibration, 124:117133.
Hirschberg, M. (2001). Introduction to aero-acoustics of internal ow. In Advances
in Aeroacoustics, Brussels, BE. Von Karman Institute.
Huang, X. and Baumann, W. T. (2007). Reduced-order modeling of dynamic
heat release for thermoacoustic instability prediction. Combustion Science and
Technology, 179(3):617636.
Huber, A. and Polifke, W. (2008). Impact of fuel supply impedance on combustion
stability of gas turbines. In Intl Gas Turbine and Aeroengine Congress &
Exposition, number ASME GT2008-51193, Berlin.
Huber, A., Romann, P., and Polifke, W. (2008). Filter-based time-domain
impedance boundary conditions for CFD applications. In Intl Gas Turbine and
Aeroengine Congress & Exposition, number ASME GT2008-51195, Berlin.
Ibrahim, Z. M., Williams, F. A., Buckley, S. G., and Lee, J. C. Y. (2006). An acous-
tic energy approach to modeling combustion oscillations. In Intl Gas Turbine
and Aeroengine Congress & Exposition, number ASME GT2006-90956.
Ibrahim, Z. M., Williams, F. A., Buckley, S. G., and Twardochleb, C. Z. (2007).
A method for estimating the onset of acoustic instabilities in premixed gas-
turbine combustion. In Intl Gas Turbine and Aeroengine Congress & Exposi-
tion, number ASME GT2007-27037, Montreal, Canada.
Jacobs, O. L. R. (1993). Introduction to control theory. Oxford University Press.
Janich, K. (1983). Analysis f ur Physiker und Ingenieure. Springer-Verlag, Berlin
Heidelberg New York.
BIBLIOGRAPHY 45
Janus, M. C. and Richards, G. A. (1996). Results of a model for premixed com-
bustion oscillations. In American Flame Research Committee International
Symposium, Baltimore, MD.
Keller, J. J. (1995). Thermoacoustic Oscillations in Combustion Chambers of Gas
Turbines. AIAA Journal, 33(12):22802287.
Keller, J. J., Egli, W., and Hellat, J. (1985). Thermally induced low-frequency
oscillations. Zeitschrift f ur angewandte Mathematik und Physik, 3(2):250274.
Kopitz, J. and Polifke, W. (2005). Stability analysis of thermoacoustic systems
by determination of the open-loop-gain. In 12th Int. Congress on Sound and
Vibration (ICSV12), Lisbon, Portugal. IIAV.
Kopitz, J. and Polifke, W. (2008). CFD-based application of the Nyquist criterion
to thermo-acoustic instabilities. J. Comp. Phys, 227:67546778.
Krebs, W., Walz, G., and Homann, S. (1999). Thermoacoustic analysis of annular
combustor. In 5th AIAA Aeroacoustics Conference, number AIAA 99-1971,
Seattle, WA.
Kr uger, U., Homann, S., Krebs, W., Judith, H., Bohn, D., and Matouschek,
G. (1998). Inuence of Turbulence on the Dynamic Behaviour of Premixed
Flames. In Intl Gas Turbine and Aeroengine Congress & Exposition, volume
ASME 98-GT-323, Stockholm, Sweden.
Kr uger, U., H uren, J., Homann, S., Krebs, W., and Bohn, D. (2000). Predic-
tion and measurement of thermoacoustic improvements in gas turbines with
annular combustion systems. In Intl Gas Turbine and Aeroengine Congress &
Exposition, number ASME 2000-GT-95, Munich.
Kr uger, U., H uren, J., Homann, S., Krebs, W., and Bohn, D. (Indianapolis, USA
1999). Prediction of Thermoacoustic Instabilities with Focus on the Dynamic
Flame Behavior for the 3A-Series Gas Turbine of Siemens KWU. Intl Gas
Turbine and Aeroengine Congress & Exposition, ASME 99-GT-111.
Lawn, C. J. (2000). Thermo-acoustic frequency selection by swirled premixed
ames. In 28th Symp. (International) on Combustion, pages 823830. The
Combustion Institute.
Lieuwen, T. and Zinn, B. T. (1998). The role of equivalence ratio oscillations in
driving combustion instabilities in low NO
x
gas turbines. In Proc. Comb. Inst.,
pages 18091816, Pittsburg, PA. The Combustion Institute.
Lumens, P. G. E. (2006). The rijke tube: a validating journey. Technical Report
R-1702-S, Tu Eindhoven / TU M unchen.
Lundberg, K. H. (2002). Barkhausen stability criterion.
http://web.mit.edu/klund/www/weblatex/node4.html.
Martin, C., L. Benoit, L., Nicoud, F., and Poinsot, T. (2004). Analysis of acoustic
energy and modes in a turbulent swirled combustor by. In Proceedings of the
Summer Programm 2004.
46 BIBLIOGRAPHY
Martin, C. E., Benoit, L., Sommerer, Y., Nicoud, F., and Poinsot, T. (2006). Large-
eddy simulation and acoustic analysis of a swirled staged turbulent combustor.
AIAA Journal, 44(4).
McManus, K. R., Poinsot, T., and Candel, S. M. (1993). A review of active control
of combustion instabilities. Prog. Energy Combust. Sci., 19:129.
Merk, H. J. (1956). Analysis of heat-driven oscillations of gas ows. i. general
considerations. Appl. Sci. Research, pages 317335.
Morfey, C. L. (1971). Acoustic energy in non-uniform ow. J. of Sound and Vibra-
tion, 14(2):159170.
Munjal, M. L. (1986). Acoustics of Ducts and Muers. John Wiley & Sons.
Murota, T. and Ohtsuka, M. (Indianapolis, 1999). Large-eddy simulation of com-
bustion oscillation in premixed combustor. Intl Gas Turbine and Aeroengine
Congress & Exposition, ASME 99-GT-274.
Neunert, U., Kopitz, J., Sattelmayer, T., and Polifke, W. (2007). Numerical
eigen-mode analysis of an acoustic-duct system by CFD-OLG and comparison
against experiment. In 6th International Congress on Industrial and Applied
Mathematics, number GA/CTS4837/12-4550, Z urich, Switzerland.
Nicoud, F. (2007). Unsteady ow modelling. In Basics of Aero-Acoustics and
Thermo-Acoustics, VKI LS 2007-02, Brussels, BE. Von Karman Institute.
Nicoud, F. and Poinsot, T. (2005). Thermoacoustic instabilities: Should the
rayleigh criterion be extended to include entropy changes? Combust. and
Flame, 142(1-2):153159.
Pankiewitz, C. (2004). Hybrides Berechnungsverfahren f ur thermoakustische Insta-
bilitaten von Mehrbrennersystemen. PhD thesis, TU-M unchen.
Pankiewitz, C., Evesque, S., Polifke, W., and Sattelmayer, T. (2001). Stability
Analysis of Annular Gas Turbine Combustors. In LECT Workshop on Insta-
bilities in Aero-Engine Combustors, Stuttgart.
Pankiewitz, C. and Sattelmayer, T. (2003a). Hybrid methods for modelling com-
bustion instabilities. In 10th Int. Conf. on Sound and Vibration, Stockholm,
Sweden. IIAV.
Pankiewitz, C. and Sattelmayer, T. (2003b). Time domain simulation of combustion
instabilities in annular combustors. Transactions of the ASME, Journal of
Engineering for Gas Turbines and Power, (125):677685.
Paschereit, C. O. and Polifke, W. (1998). Investigation of the Thermo-Acoustic
Characteristics of a Lean Premixed Gas Turbine Burner. In Intl Gas Turbine
and Aeroengine Congress & Exposition, number ASME 98-GT-582, Stock-
holm, Sweden.
Peracchio, A. A. and Proscia, W. M. (1998). Non-linear heat-release / acoustic
model for thermoacoustic instability in lean premixed combustors. In Intl Gas
BIBLIOGRAPHY 47
Turbine and Aeroengine Congress & Exposition, volume ASME 98-GT-269,
Stockholm, Sweden.
Pierce, A. D. (1981). Acoustics. McGraw-Hill.
Poinsot, T. and Veynante, D. (2005). Theoretical and Numerical Combustion. Ed-
wards, R. T. Incorporated, 2 edition.
Polifke, W. (2004). Combustion instabilities. In Advances in Aeroacoustics and
Applications, VKI LS 2004-05, Brussels, BE. Von Karman Institute.
Polifke, W. and Gentemann, A. M. G. (2004). Order and realizability of impulse
response lters for accurate identication of acoustic multi-ports from transient
CFD. Int. J. of Acoustics and Vibration, 9(3):139148.
Polifke, W., Paschereit, C. O., and Dobbeling, K. (1999). Suppression of combustion
instabilities through destructive interference of acoustic and entropy waves. In
6th. Int. Conf. on Sound and Vibration, Copenhagen, Denmark.
Polifke, W., Paschereit, C. O., and Dobbeling, K. (2001a). Constructive and De-
structive Interference of Acoustic and Entropy Waves in a Premixed Combustor
with a Choked Exit. Int. J. of Acoustics and Vibration, 6(3):135146.
Polifke, W., Paschereit, C. O., and Sattelmayer, T. (1997). A Universally Ap-
plicable Stability Criterion for Complex Thermoacoustic Systems. In 18.
Deutsch-Niederlandischer Flammentag, VDI Bericht, number 1313, pages 455
460, Delft, NL.
Polifke, W., Poncet, A., Paschereit, C. O., and Dobbeling, K. (2001b). Reconstruc-
tion of acoustic transfer matrices by instationary computational uid dynamics.
J. of Sound and Vibration, 245(3):483510.
Polifke, W., Wall, C., and Moin, P. (2006). Partially reecting and non-reecting
boundary conditions for simulation of compressible viscous ow. J. of Comp.
Phys., 213(1):437449.
Rayleigh, J. W. S. (1878). The explanation of certain acoustical phenomena. Nature,
18:319321.
Richards, G. A. and Janus, M. C. (1998). Characterization of Oscillations During
Premix Gas Turbine Combustion. J. Eng. for Gas Turbines and Power, 120:294
302.
Sattelmayer, T. and Polifke, W. (2003a). Assessment of methods for the com-
putation of the linear stability of combustors. Combust. Sci. and Techn.,
175(3):453476.
Sattelmayer, T. and Polifke, W. (2003b). A novel method for the computation of
the linear stability of combustors. Combust. Sci. and Techn., 175(3):477498.
Schmitt, P., Poinsot, T., Schuermans, B., and Geigle, K. P. (2007). Large-eddy
simulation and experimental study of heat transfer, nitric oxide emissions and
combustion instability in a swirled turbulent high-pressure burner. J. Fluid
Mech., 570:1746.
48 BIBLIOGRAPHY
Schuermans, B. (2003). Modeling and Control of Thermoacoustic Instabilities. PhD
thesis,

Ecole Polytechnique Federale de Lausanne.
Schuermans, B., Bellucci, V., Nowak, D., and Paschereit, C. O. (2002). Modelling
of complex thermoacoustic systems: A state-space approach. In Ninth Int.
Congress on Sound and Vibration, ICSV9, Orlando, FL, U.S.A. IIAV.
Schuermans, B., Bellucci, V., and Paschereit, C. O. (2003). Thermoacoustic model-
ing and control of multi-burner combustion systems. In Intl Gas Turbine and
Aeroengine Congress & Exposition, number ASME 2003-GT-38688, Atlanta,
GA, U.S.A.
Schuermans, B., Luebcke, H., Bajusz, D., and Flohr, P. (2005). Thermoacoustic
analysis of gas turbine combustion systems using unsteady CFD. In Intl Gas
Turbine and Aeroengine Congress & Exposition, number GT 2005-68393,
Reno, Nevada. ASME.
Schuermans, B. B. H., Polifke, W., and Paschereit, C. O. (2000). Prediction of
Acoustic Pressure Spectra in Gas Turbines based on Measured Transfer Ma-
trices. In Intl Gas Turbine and Aeroengine Congress & Exposition, number
ASME 2000-GT-105, Munich, Germany.
Selle, L., Benoit, L., Poinsot, T., and Krebs, W. (2006). Joint use of compressible
LES and Helmholtz solvers for analysis of rotating modes in an industrial
swirled burner. Combust. and Flame, 145:194205.
Selle, L., Nicoud, F., and Poinsot, T. (2004). Actual impedance of nonreect-
ing boundary conditions: Implications for computation of resonators. AIAA
Journal, 42(5):958964.
Sklyarov, V. A. and Furletov, V. I. (1975). Frequency characteristics of a laminar
ame. Plenum Publishing, (UDC 536.46:533.6,):6977. Translated from Zhur-
nal Prikladnoi Mekhaniki i Tekhnicheskoi Fiziki, No. 1, pp. 84-94, Jan. -Feb.
1974.
Stow, S. R. and Dowling, A. P. (2003). Modelling of circumferential modal coupling
due to helmholtz resonators. In Intl Gas Turbine and Aeroengine Congress &
Exposition, number ASME GT2003-38168, Atlanta, GA, U.S.A.
Stow, S. R. and Dowling, A. P. (New Orleans, LA, 2001). Thermoacoustic oscilla-
tions in an annular combustor. Intl Gas Turbine and Aeroengine Congress &
Exposition, ASME 2001-GT-37.
Sung, H. G., Hsieh, S., and Yang, V. (2000). Combustion Dynamics of a Lean-
Premixed Swirl-Stabilized Injector. In Proc. Comb. Inst. 28, Edinburgh, Scot-
tland.
Tyagi, M., Chakravarthy, S. R., and Sujith, R. I. (2007). Unsteady combustion
response of a ducted nonpremixed ame and acoustic coupling. Combust.
Theory Modelling, 11(2):205 226.
Tyler, J. M. and Sofrin, T. G. (1962). Axial ow compressor studies. SAE Trans-
actions, 70:309332.
BIBLIOGRAPHY 49
Veynante, D. and Poinsot, T. (1997). Large Eddy Simulation of Combustion In-
stabilities in Turbulent Premixed Burners. Technical report, Center for Tur-
bulence Research, Stanford, CA.
Wall, C. (2005). Numerical Methods for Large Eddy Simulation of Acoustic Com-
bustion Instabilities. PhD thesis, Stanford University.
Zhu, M., Dowling, A. P., and Bray, K. N. C. (2005). Transfer function calcula-
tions for aeroengine combustion oscillations. Journal of Engineering for Gas
Turbines and Power, 127(1):1826.

S-ar putea să vă placă și