Sunteți pe pagina 1din 14

Combustion and Flame 156 (2009) 106119

Contents lists available at ScienceDirect


Combustion and Flame
www.elsevier.com/locate/combustame
Rayleigh criterion and acoustic energy balance in unconned self-sustained
oscillating ames
D. Durox

, T. Schuller, N. Noiray, A.L. Birbaud, S. Candel


Ecole Centrale Paris, EM2C Laboratory, CNRS, 92295 Chtenay-Malabry, France
a r t i c l e i n f o a b s t r a c t
Article history:
Received 30 October 2007
Received in revised form 22 April 2008
Accepted 29 July 2008
Available online 9 October 2008
Keywords:
Combustion instability
Flame dynamics
Rayleigh criterion
Instabilities of conned combustion systems are often discussed in terms of the Rayleigh criterion, which
provides a necessary condition for unstable operation and is commonly used to distinguish driving
and damping regions. The analysis is also carried out in some cases by making use of an acoustic
energy balance in which the Rayleigh term acts as a source. The case of unconned ames is less well
documented but of importance in practical systems used in heating and drying. This study is motivated
by problems of self-sustained oscillations of radiant burners for domestic or industrial processes and of
various other types of open ames. Application of the Rayleigh criterion and of the balance of acoustic
energy to oscillations arising in such unconned systems is examined. The objective is to see if the
Rayleigh condition is fullled and to show how the different perturbed variables are linked to each other
to develop an unstable oscillation. These issues are investigated by experiments in two geometries. The
rst case relates to a single V- or M-shaped ame formed by a burner behaving like a Helmholtz
resonator. The second geometry features a collection of conical ames (CCF) established by a multipoint
injector. This system is fed by a manifold that features a set of plane modes and resonates like an
organ pipe at frequencies corresponding to odd multiples of the quarter wave. The Rayleigh criterion
and a related result written in the form of an acoustic energy balance are used to dene conditions of
instability. A link is established between the pressure signal radiated by the burner and the total heat
release rate perturbation yielding the phase lag between these two variables and providing conditions
for unstable operation. Systematic experiments carried out in the two burner geometries and model
predictions are in good agreement indicating that the Rayleigh source term is positive and that the
criterion is well fullled by the waveeld component corresponding to the pressure radiated back by
the resonator. The gain and loss terms in the acoustic energy budget are estimated in one typical case,
indicating that the source term is balanced by the acoustic power radiated away from the ame region,
while dissipation of acoustic energy in the burner is shown to be essentially negligible.
2008 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
1. Introduction
Combustion instabilities commonly observed in practical sys-
tems have many undesirable features. They are a source of noise,
vibrations, and structural fatigue, they intensify heat uxes to the
chamber walls, and in extreme cases they may lead to failure of
the system. An important research effort is currently targeted at
devising solutions to these problems, mainly in relation to de-
velopment of advanced premixed combustion technologies for gas
turbines. One instability criterion that is often quoted in this eld
is due to Rayleigh [1]. Expressed in mathematical form, this cri-
terion indicates that a necessary condition for instability is that
_
V
p

dV > 0, where p

and q

, respectively, stand for pressure


*
Corresponding author.
E-mail address: durox@em2c.ecp.fr (D. Durox).
and heat release rate (per unit volume) uctuations, and V desig-
nates the ow domain. This criterion is often used to interpret ex-
perimental data and identify driving mechanisms. There are useful
extensions of the Rayleigh criterion, one of them being a balance
of acoustic energy in which the product p

appears in the source


term S. This balance, integrated over an oscillation period, can be
written in the form
E
t
+ F = S D, (1)
where E and F, respectively, designate the acoustic energy den-
sity and ux, S is a source term that under some rather general
conditions takes the form
S =
1
T
T
_
0
1

0
c
2
p

dt, (2)
0010-2180/$ see front matter 2008 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
doi:10.1016/j.combustame.2008.07.016
D. Durox et al. / Combustion and Flame 156 (2009) 106119 107
and D accounts for acoustic energy dissipation per unit volume.
Integrating the local energy balance over the volume V , one nds
that the acoustic energy will grow if
1

0
c
2
1
T
_
V
_
T
p

dV dt >
_
A
F.ndA +
_
V
DdV . (3)
In this formulation the Rayleigh integral appears as a source of
acoustic energy. This inequality indicates that perturbations will
grow if the source term on the left-hand side exceeds the losses of
energy associated with energy uxes through the domain control
surface A and dissipation integrated over the volume V . This for-
mulation is more quantitative than the Rayleigh criterion because
it indicates that when the gain associated with driving processes
exceeds losses, the acoustic energy in the system increases and the
system becomes unstable. When the system has reached a limit cy-
cle one also expects that the source term on the left-hand side will
exactly balance the losses on the right-hand side.
At this point it is worth noting that the derivation of acoustic
energy budgets in inhomogeneous ows is extensively discussed
in the literature and that this topic has received considerable at-
tention, mainly in relation to aeroacoustics (see for example [2]).
Many of these studies are concerned with the denition of acous-
tic energy and ux in cases where the mean ow velocity is nite
[35]. In the present case and in most situations of interest in
combustion, the mean ow velocity remains small and one may
use expressions that apply to the zero mean ow situation. The
next approximation would be to use the energy density and ux
introduced for example in [3] or in [6] to account for the presence
of a mean ow. It is also worth indicating that extensions or mod-
ications of the Rayleigh criterion have been proposed over the
years, one of them being reported in the work of Chu [4] and re-
cently updated in Ref. [7]. These extensions are generally less easy
to use. For example, in the source term put forward in Refs. [4,7],
the pressure uctuation is replaced by a temperature uctuation,
which is not easy to measure in ames. In addition, the forms de-
vised in these references are specically motivated by cases where
entropy waves propagate in the ow. In the present conguration
these entropy waves are convected away from the ame and do
not play a role in the instability process. For these various rea-
sons these alternative forms are not considered in what follows.
The standard Rayleigh criterion is more generally adopted in the
analysis of instabilities in conned congurations, featuring various
types of injection manifolds and combustor geometries, and it is
often used to identify processes driving instabilities [817] and has
been employed more recently to distinguish regions where ampli-
cation or damping takes place (see for example [18,19]).
At this point, it is worth summarizing some of the features ob-
served under unstable operation. It is rst noted that combustion
instabilities occur if the heat release rate can be made to uctuate.
This can be caused by the dynamical response of the ame to in-
cident perturbations. Sustained oscillations of heat release may be
highly nonlinear but their fundamental frequency coincides with
that of the pressure oscillation observed in the system. The link be-
tween these two quantities involves various types of interactions,
some of which are listed in what follows:
A pressure oscillation in the chamber induces a velocity uc-
tuation at the burner outlet. This generates a uctuation of
heat release with a time delay associated, for example, with
the time required for a hydrodynamic perturbation to reach
the ame front.
The pressure oscillation may propagate into the reactant sup-
ply manifolds and modify the injection rates, giving rise to
perturbations in equivalence ratio. These perturbations may
then be convected to the ame, inducing variations of the heat
release with a certain time lag.
Vortex structures can be generated by velocity perturbations at
the burner exhaust or at obstacles present in the ow (baes,
ame holders, etc.). These vortices, convected downstream,
impinge on the ame, inducing an accumulation of fresh re-
actants, followed by their rapid reaction and release of heat.
The phase between heat release rate and pressure is then de-
ned by a sequence of processes, which must be identied if one
wishes to predict the occurrence of instability. The list given pre-
viously serves to show that mechanisms linking uctuations of
pressure and heat release may be quite complex. In many circum-
stances multiple interactions occur simultaneously. For example,
uctuations in equivalence ratio appear in conjunction with ve-
locity oscillations; vortex shedding and roll-up arise in situations
where the ame also responds to velocity uctuations or impinges
on a boundary. Success in the control of instabilities thus requires
good knowledge of the ow and ame dynamics and of the ame
response to acoustic and convective modes. Experiments carried
out in recent years have provided a wealth of information on
these aspects, mainly in the case of turbulent combustors [20
25]. The use of numerical tools relying on large eddy simulations
has also allowed further progress in the modeling of turbulent
ame dynamics [2629]. These tools need to be validated on well-
controlled small-scale experiments, which are easier to instrument
and model. Many generic interactions can be examined in these
reduced-scale experiments, which can serve as test beds for mod-
eling and simulation. References [3037] document such experi-
ments and provide simplied descriptive model ames for laminar
ame dynamics, while [38,39] give some analytical extensions in
the case of turbulent ames.
While the case of conned combustors is suitably documented,
that of unconned ames has generally received less attention.
Burners with unconned ames also develop instabilities. This is
observed in domestic boilers for production of hot water, and in
radiant burners used for drying paper or for processing metallic
sheets. The ames are usually distributed over large areas and
their self-sustained oscillation generates important noise levels,
at a fundamental frequency and many higher order harmonics.
Regimes of strong noise radiation correspond to an acoustics
combustion coupling between the ame and the burner resonant
modes. This is documented in [4044], which focus on instabili-
ties in generic burners and describe the interaction between the
unconned ames and the system acoustics.
The coupling always involves the emission of a sound wave
generated by the time variation of the heat release rate [4548].
This wave, radiated in all space, induces external pressure pertur-
bations at the burner exhaust, which act as a source of pressure
oscillation inside this system. References [4044] already provide
conditions under which the systems become unstable but do not
consider the Rayleigh source term and its consequences in terms
of acoustic energy.
This analysis is carried out in what follows by making use of
results obtained in two different geometries: a burner featuring a
single V or M ame and operating like a Helmholtz resonator
(HRB) and a multiple ame burner behaving like an organ pipe
resonator (OPR). One diculty in the evaluation of the Rayleigh
source term is that the pressure perturbation cannot be measured
inside the ame region itself. In the conned case, pressure is usu-
ally recorded by sensors placed on the combustor side walls, and
one implicitly assumes that the pressure in the ame is equal to
that detected at the wall. In the unconned case, pressure can
only be measured on the upstream side of the burner exhaust
section near the ame-holding system. If pressure is detected at
close distance to the ame, it is possible to assume that the sig-
nal has the same phase as the pressure inside the ame region.
The amplitude of the pressure eld in the ame can be obtained
108 D. Durox et al. / Combustion and Flame 156 (2009) 106119
Fig. 1. Schematic representation of the burner acting as a Helmholtz resonator
(HRB). Outlet diameter: 22 mm. Central rod diameter: 6 mm.
by extrapolating pressure measurements in the upstream mani-
fold or by reconstruction of the pressure signal from combined
acoustic pressure and velocity data. After a brief description of
the two combustion systems (Section 2), experimental results are
used to link the main ow variables: the unsteady pressure inside
the burner, the velocity uctuation at the burner outlet, the un-
steady heat release rate in the ame, and the external pressure
perturbation at the burner exhaust induced by the ame (Sec-
tion 3). An analysis is then carried out to examine these relations
on a theoretical level (Section 4) and retrieve experimental re-
sults. The balance of acoustic energy is considered in Section 5
and the Rayleigh source term is estimated. This requires an extrap-
olation to obtain the pressure amplitude in the ame zone. It is
then possible to close the energy budget by estimating the acous-
tic energy uxes traversing a control surface including the ame
zone.
2. Experimental congurations
The rst burner (HRB) is equipped with a convergent noz-
zle and a tube with a 22-mm-inner-diameter exit (Fig. 1). Un-
der certain operating conditions, this device oscillates in the low-
frequency range like a Helmholtz resonator with a bulk pressure
oscillation inside the burner system. This oscillator may be treated
as a massspring system in which the volume acts as a spring
and the gas column inside the convergent nozzle and in the outlet
tube behaves like a mass. Details on this resonator can be found
in [4042]. A ame holder formed by a 6-mm-diameter rod is
placed on the axis of the tube. With this device, two types of
ame shapes can be obtained: ames stabilized on the central
rod and the burner lip, featuring an M shape, also designated
as fountain ames, and ames stabilized only on the inner rod,
with a V-shape, also described as inverted conical ames (ICF).
The ame conguration depends on the initial conditions. Using
a stick it is possible to transition from a V ame to an M
ame. The ow velocity of the methaneair mixture, measured
at a radial distance of 7 mm from the axis and 1.5 mm above
the burner outlet, is typically 3.5 ms
1
, but this can be adjusted
with the conguration. The equivalence ratio is set between 0.9
and 1.2, depending on the unstable conditions under investiga-
tion.
A microphone M0 is xed at the base of the burner. A second
microphone (M2) located in the exit plane, 30 cm away from the
burner axis, is used to measure the radiated noise. A photomulti-
plier, equipped with a lter for the detection of the CH* radical,
gives a signal that is nearly proportional to the heat release rate in
the case of premixed ames [45,46]. Velocity at the burner ex-
Fig. 2. Schematic representation of the multiple ame burner acting as an organ
pipe resonator (OPR). Resonant duct diameter: 70 mm. Perforated plate: 420 holes.
The 2-mm-in-diameter injection channels are arranged on a 3-mm square mesh.
The plate is 3 mm thick. L varies between 150 and 750 mm. Six pressure taps,
P1P6, are set up on an upstream duct generating line.
haust is measured with a laser Doppler velocimeter (LDV). The
ames featuring V- or M-shapes generate large variations of
heat release in response to velocity perturbations at the burner
exit, thus producing large noise levels [47,48]. A detailed descrip-
tion of the mechanisms leading to self-sustained oscillation can
be found in [41,42]. The noise radiated by the ame is periodic,
but its spectrum contains harmonics of the fundamental frequency
[4648].
The second burner (OPR), shown schematically in Fig. 2, is
equipped with a piston and behaves like an organ pipe resonator,
under operating conditions described in [43]. Four hundred twenty
small conical ames are held on a perforated plate xed at the end
of the tube. The plate porosity is 34%. Holes have a diameter of
2 mm and are located on a matrix with a square mesh of 3 mm.
Results obtained with some other grids are presented in [43,49].
The piston head can be located at different positions inside the
tube, from 150 to 750 mm away from the perforated plate. The
tube has a diameter of 70 mm. The burner is fed with premixed
air and methane through eight holes located at the top of the pis-
ton head. The equivalence ratio is typically 0.86. The ow velocity
through each of the 420 small channels is 3.5 ms
1
. Pressure taps
(P1P6) are regularly distributed on a generating line of the cylin-
drical duct. Microphones, installed on the sideline of a waveguide,
record the pressure oscillation (in Fig. 2, a microphone is xed
in P2). As in the case of the Helmholtz resonator, a photomultiplier,
equipped with a lter centered on OH* radical UV light emission,
detects heat release rate uctuations (in the case of lean premixed
ames or close to stoichiometry, the emission of OH* or CH* rad-
icals is nearly proportional to heat release rate [45]). Microphone
M2 is used to measure the noise radiated by the ame at a dis-
tance of 30 cm from the burner axis.
For certain lengths L of the upstream duct, the OPR system ex-
hibits self-sustained oscillations involving one of the longitudinal
acoustic modes of the duct. Resonances occur close to multiples
of a quarter-wave mode in these regimes. Flames oscillate collec-
D. Durox et al. / Combustion and Flame 156 (2009) 106119 109
Fig. 3. Typical views of the ames under self-sustained oscillations. On the left: V- ame. Frequency 172 Hz, V =2.7 ms
1
(measured at 7 mm of the axis),

= 0.25 ms
1
,
equivalence ratio 0.92. In the center: M-ame. Frequency 138 Hz, V = 2.26 ms
1
,

= 0.17 ms
1
, equivalence ratio 0.92. On the right: Multiple ame burner. Frequency
744 Hz. Equivalence ratio 0.86.
tively, in phase, and at one of the eigenfrequencies of the duct.
The noise radiated by the system is intense [48], with a funda-
mental frequency and some harmonics. These effects are described
in detail in [43,49], which also include a detailed account of the
instability mechanism and an analysis of passive control methods
to suppress the instability.
Typical views of the ames under self-sustained oscillations are
displayed in Fig. 3.
3. Experimental data
Measurements are carried out under strong unstable condi-
tions with the two burners described previously. The objective
is to analyze the phase relations between ow variables during
self-sustained combustion oscillations and specically examine re-
lations between heat release rate uctuations in the ame and
acoustic uctuations inside the burner.
3.1. Helmholtz resonator burner (HRB)
Two ame geometries were investigated in the HRB case shown
in Fig. 1. The rst corresponds to a V ame stabilized on the
central rod; the second case is that of an M ame stabilized
on the central rod and the burner lip. A typical example of time
signals is displayed in Fig. 4, in the case of a V ame, during
Fig. 4. V-ame. Self-sustained oscillation frequency at 206 Hz. U
0
= 3.5 ms
1
,
measured at z = 1.5 mm from the outlet and at 7 mm of the axis (see Fig. 1). The
equivalence ratio is 1.11. The rms velocity uctuation is 0.48 ms
1
.
self-sustained oscillations at 206 Hz. This value is close to the res-
onance frequency under cold conditions, which is at 198 Hz. The
pressure signal M0 measured at the base of the burner and the
velocity signal recorded at the burner exit section are almost sinu-
soidal. The photomultiplier signal is less symmetrical, indicating
nonlinear effects with the presence of harmonics of the funda-
110 D. Durox et al. / Combustion and Flame 156 (2009) 106119
Fig. 5. M-ame. Self-sustained oscillation frequency at 207 Hz. U
0
= 3.55 ms
1
,
measured at z = 1.5 mm from the outlet and at 7 mm of the axis (Fig. 1). The
equivalence ratio is 1.17. The rms velocity uctuation is 0.17 ms
1
.
mental oscillation frequency. In spite of the signal distortion, the
phase difference between the pressure signal M0 and the OH* sig-
nal emitted by the ame is of 0.2 rad and these two signals are
nearly in phase. The velocity signal is out of phase by 1.98 rad
with respect to the pressure signal. This phase shift is not very far
from the theoretical value /2 1.57 rad predicted in the analysis
carried out in the next section.
The M-ame (Fig. 5) also features an instability, but at 207 Hz,
with nearly the same operating conditions as for the V ame
case, but it is weaker. The pressure signal inside the burner is
nearly sinusoidal, indicating a harmonic oscillation, whereas the
velocity signal at the burner outlet is slightly distorted with re-
spect to a sinusoidal waveform. The heat release rate signal (pro-
portional to CH* emission) is periodic and features a slow growth,
followed by a rapid decrease. In the M-ame case, this phe-
nomenon corresponds to a fast consumption of fresh gases located
in the upper part of the ame, where neighboring reactive layers
are pinched as they interact [41]. It is interesting to mention that
this mechanism of rapid ame surface annihilation is also observed
with turbulent ames held on a central bluff-body and reattach-
ing on the outer injector lips (for example [24,50,51]), when the
ame geometry takes an M-like shape, and when the ames are
submitted to large velocity perturbations. The process of rapid de-
struction of ame area yields an asymmetrical heat release signal,
but it is nevertheless possible to determine the phase difference
between the pressure and the heat release rate by considering the
fundamental frequency of the Fourier series expansion of these
two signals. This calculation gives a phase difference of 0.05 rad,
indicating that the pressure uctuation inside the burner and the
fundamental component of the heat release rate are perfectly in
phase. The phase difference between velocity at the burner outlet
and pressure inside the burner is 2.0 rad, which is comparable to
the value observed in the V ame case and not very far from the
theoretical quadrature predicted in the analysis developed later in
this article.
3.2. Multiple ame burner (OPR)
Two duct lengths are considered in this conguration, both
equipped with a 3-mm-thick perforated plate. The rst value,
L = 183 mm, corresponds to a quarter-wave mode in the upstream
manifold; the other length, L = 417 mm, pertains to a three-
quarter-wave mode in this element. These lengths yield strong
self-sustained oscillations, and they correspond to a center point
in two instability ranges observed in this conguration [43].
Fig. 6 displays three time traces corresponding to a quarter
wave oscillation at 438 Hz: the emission signal measured by the
PM equipped with an OH* lter and two pressure signals P2 and
Fig. 6. Organ pipe burner. Cavity length L = 183 mm. Self-sustained oscillation fre-
quency at 438 Hz.
h
= 3.5 ms
1
. The equivalence ratio is 0.86. The phase dif-
ference between the photomultiplier and the pressure uctuation in P2 and P3 is
0.14 rad.
Fig. 7. Organ pipe burner resonating on a 3/4 wave mode. Upstream duct length L =
417 mm. The self-sustained oscillation is 569 Hz.
h
= 3.5 ms
1
. The equivalence
ratio is 0.86. The phase difference between the photomultiplier and the pressure
uctuation P2 is 0.61 rad. P2 and P6 are in phase opposition.
P3 recorded by microphones mounted on the burner. Pressures
uctuations P2 are detected 95 mm below the upstream side of
the plate, corresponding to a location in the middle of the burner:
x/L = 0.54, where L is the burner length. The second microphone,
P3, is located 170 mm away from the bottom of the plate, near
the piston head x/L = 0.96. The three signals P2, P3 and OH* are
perfectly in phase. The uctuation level P2 is slightly lower than
the P3 signal level, because the pressure tap is closer to the exit,
and thus closer to a pressure node. The P3 signal level exceeds
700 Pa peak to peak. The signal at the other pressure tap (P1) is
also in phase with the signals detected at P2 and P3, as expected
for a quarter-wave mode. The pressure signal in the ame zone ob-
tained by extrapolating the available pressure data is also in phase
with P2 and P3. As for the HRB cases, internal pressure uctua-
tions, extrapolated pressure in the ame region, and heat release
rate uctuations are essentially in phase. Oscillations are nearly
sinusoidal, with a limited level of harmonics. Signals for the self-
sustained oscillation at the three-quarter-wave mode are displayed
in Fig. 7. The pressure oscillation is detected at two different po-
sitions, the rst, P2, located close to the exit at x/L = 0.23, the
second, P6, located close to the piston head at x/L =0.95. The duct
length is such that the system resonance takes place on a very
strong three-quarter-wave mode at 569 Hz [43]. The signals are
slightly distorted, indicating the presence of harmonics. The peak-
to-peak amplitude of the pressure uctuation P6 reaches 700 Pa
again. The two pressure signals are in phase opposition, conrm-
ing that the oscillation involves a three-quarter-wave mode. This
time, the pressure signal in the ame region extrapolated from
D. Durox et al. / Combustion and Flame 156 (2009) 106119 111
Fig. 8. Phase difference between the pressure signal P2, measured upstream of the
perforated plate and the heat release rate uctuation. A positive value means that
the pressure is in advance with respect to heat release. The ow velocity in the
perforated plate channels is
h
=3.5 ms
1
. The equivalence ratio is 0.86. The circle
on the curve corresponds to the operating point presented in Fig. 6 and the square
on the curve to the operating point described in Fig. 7.
the pressure signals detected in the upstream manifold is in phase
with P2. The heat release rate signal is slightly out of phase with
respect to the pressure signal detected by P2 near the perforated
plate. The phase difference measured between the two signals is
0.61 rad, determined from the transfer function examined at the
system resonance frequency 569 Hz. This phase difference remains
small, indicating that external perturbations in heat release rate
and pressure oscillations in the ame region deduced from pres-
sure measurements near the perforated plate verify the Rayleigh
criterion.
It was shown that this burner exhibits wide ranges of un-
stable operation and oscillations are observed for most of the
duct lengths investigated [43,49]. It is interesting to determine
the phase difference, between the heat release rate perturbation
and the pressure close to the upstream side of the grid (P2), as
a function of the duct length. This pressure signal has the same
phase as the pressure extrapolated inside the ame region. This is
achieved by changing the piston position over a wide range of val-
ues. Results are presented in Fig. 8, which also show the specic
operating points L = 183 mm and L = 417 mm, corresponding to
Figs. 6 and 7 and, respectively, identied by a circle and a square
symbol.
In the ranges examined, the system features a quarter-wave
mode for the short lengths, and a three-quarter-wave mode for the
large values of L. For intermediate values of L, between 220 and
300 mm, the system features a stable region (the other operat-
ing conditions being kept constant). The phase difference between
external heat release rate uctuations and internal pressure uc-
tuations close to the upstream side of the perforated plate, where
ames are anchored, always lies between /4 and /4. This typ-
ies the phase difference between heat release uctuations and
pressure inside the ame region. This implies that as for self-
sustained oscillations observed in the HRB conguration, instabili-
ties observed in the OPR conguration exhibit pressure uctuations
in the ame region and heat release rate uctuations, which are
nearly in phase.
These observations, made on two burner congurations (an
organ pipe and a Helmholtz resonator) for three ame shapes
(a collection of small conical ames, a V-ame, and a M-ame)
and three different resonant oscillation modes (bulk, quarter, and
three-quarter wave modes) imply a feedback between the un-
steady external combustion process at the burner outlet and the
burner acoustics. The mechanisms linking external heat release
rate uctuations to pressure uctuations in the system and radi-
ated by the system are now examined.
4. Analysis of acousticcombustion coupling
The external pressure perturbation that drives the system into
resonance originates from combustion noise. In the unconned
congurations considered in this article, previous studies indicate
that the feedback between external and internal acoustic variables
across the burner outlet is ensured by the noise radiated by the
ame, back toward the burner outlet [4044,49]. A compact ame
behaves like a monopole and radiates noise in all space [45,52,53],
and in particular toward the burner. The multiple ames formed
on the perforated plate move synchronously and this ensemble
also acts as an acoustic monopole. The radiated pressure is directly
related to the time rate of change of the global heat release rate.
At the burner exit, this results in a pressure perturbation, which
can be expressed as
p

ext
= K(r)
_
d

Q

dt
_
t
a
, (4)
where the coecient K(r) represents the dependence with dis-
tance from the sound source constituted by the single ame or
by the collection of conical ames. This coecient is also a func-
tion of the thermal expansion ratio E = T
2
/T
1
and the laminar
burning velocity. The distance r can be identied as the distance
between the location of the main source of sound and the burner
outlet [43,47,48], typically a few cm in the HRB conguration [40
42,44] and a few millimeters in the OPR conguration [43,49]. In
expression (4),
a
is the acoustic propagation delay from the sound
source to the burner outlet. Because the ame is compact with re-
spect to the acoustic wavelength and the sound sources are close
enough to the burner outlet, this delay can be neglected. For har-
monic perturbations of the form x

= xe
it
, one nds
p
ext
= iK(r)

Q . (5)
For all frequencies, external pressure perturbations at the burner
outlet generated by the unsteady ame are in quadrature with re-
spect to heat release rate uctuations. This result may appear at
rst glance inconsistent, since at resonance, when =
0
, pres-
sure and heat release rate uctuations should be in phase, accord-
ing to the Rayleigh criterion. This conict is resolved if one remem-
bers that p
ext
does not represent the total pressure but solely the
external pressure perturbation radiated by the ame, which drives
the system into resonance.
In addition to p
ext
, the pressure eld in the vicinity of the
duct exit features a radiated pressure. In the present congura-
tion, this quantity may be evaluated by making use of the radiation
impedance. It is known that the pressure radiated by a duct, which
conveys a plane wave oscillation characterized by a velocity uc-
tuation
e
, can be written in the form [5456]
p
r
= c(
0
i
0
)
e
,
where Z
r
= c(
0
i
0
) is the radiation impedance of this duct.
In the low-frequency range corresponding to small values of ka,
where a designates the duct radius and k = /c is the wavenum-
ber, the real and imaginary parts of the radiation impedance are
given by
0
= (1/2)(ka)
2
and
0
= (8/3)ka. It is convenient to
dene a

= (8/3)a and write


0
=ka

. In the low-frequency range


the radiated eld is dominated by the imaginary part of the radia-
tion impedance:
p
r
icka


e
. (6)
This indicates that the pressure response of the system is in
quadrature with the velocity uctuation at the resonator end. This
expression will be used to estimate the pressure eld in the res-
onator near eld. The total pressure eld in the ame region is
then composed of the pressure radiated by the ame and by the
112 D. Durox et al. / Combustion and Flame 156 (2009) 106119
Fig. 9. Schematic representation of the Helmholtz resonator burner (HRB).
resonator response p
r
. It will be shown that this component of the
pressure eld has an amplitude that is of the same order as that of
the incident wave, and it will be seen that it is essentially in phase
with heat release uctuations. These features are illustrated in Ap-
pendix A, which includes a short account of an experiment where
the OPR is submitted to an external sound eld generated by a
loudspeaker. It is shown that the pressure wave radiated back by
the OPR closely matches the theoretical prediction for the radiated
pressure p
r
.
As a further demonstration of the importance of the duct re-
sponse, we also treat in Appendix B a model problem where a
source of heat release is placed in a semi-innite duct. The source
is at a position x
0
from the duct termination. It is shown that
in this case the total pressure eld formed by the pressure wave
emitted by the source and by the wave reected by the duct ter-
mination is in phase with the heat release uctuation.
4.1. Helmholtz resonator burner (HRB) case
In the Helmholtz resonator case, the burner can be represented
by a volume V followed by a small tube of length l and section S
1
(Fig. 9). When the wavelength is much larger than the size of the
system ( L, l), and when S
1
S
0
, the resonance of the system
may be calculated by separately examining the cavity of volume V
and the exhaust duct. In what follows, variables denoted x and x

,
respectively, designate average and uctuating quantities. Assum-
ing that perturbations are isentropic, a mass balance on volume
V yields the expression for the velocity uctuation at the burner
outlet

1
=
V
c
2
S
1
dp

0
dt
,
where c is the speed of sound in the cavity and p

0
represents the
bulk pressure uctuation inside the burner. Because the ame is
compact and close to the burner exhaust, the pressure in the ame
region will be close to this bulk pressure. This relation is valid
within the limit of low Mach numbers. For harmonic perturbations,
one nds

1
= i
V
c
2
S
1
p
0
. (7)
Velocity uctuations in the exhaust duct are in quadrature with
respect to the internal pressure uctuations p

0
.
Assuming that an external pressure perturbation p

1
acts at
the burner outlet, a momentum balance through the burner neck
yields
M
d

1
dt
= S
1
(p

0
p

1
) 2M

1
. (8)
In this equation, M = S
1
l
e
is the mass of the gas column set in
motion. This expression also uses the effective length l
e
= l + e,
where l is the length of the exhaust duct and e is an end cor-
rection, which corresponds to the impedance adaptation between
the ow at the exit of the burner and the atmosphere [54]. This
end correction is of the order of the exhaust duct radius. The last
term in Eq. (8) represents the dissipation in the system, which is
opposed to the motion, and is the damping coecient. For har-
monic perturbations Eq. (8) yields
p
0
p
1
= (2 i) l
e

1
. (9)
Combining Eqs. (6) and (8), one nds a relation for the evolution
of internal pressure uctuations p

0
controlled by the external pres-
sure perturbations p

1
,
d
2
p

0
dt
2
+2
dp

0
dt
+
2
0
p

0
=
2
0
p

1
, (10)
with

2
0
=
S
1
l
e
c
2
V
, (11)
where
0
is the angular frequency of the Helmholtz resonator. As-
suming harmonic forcing p

1
= p
1
e
it
, the system response inside
the cavity is given by
p
0
p
1
=
1
1 (/
0
)
2
2i/
2
0
. (12)
The interesting case is that corresponding to resonance, when =

0
. One obtains
p
0
= i

0
2
p
1
. (13)
Pressure p
0
uctuations inside the cavity are delayed by /2 with
respect to external driving uctuations p
1
, meaning that the max-
imum amplitude of the external excitation signal p

1
corresponds
to the maximum growth rate of the pressure p

0
inside the cav-
ity. Equation (12) also indicates that the phase between p
0
and p
1
lies between 0 and , but the maximum eciency for sustaining
an instability is obtained at resonance [41,42]. The internal pertur-
bation p
0
corresponds to the total pressure uctuation inside the
burner in response to an external excitation p
1
= p
ext
, which re-
sults from combustion noise (see Eq. (5)):
p
1
= iK(r)

Q . (14)
Changes in heat release rate

Q

are linked to the velocity modula-
tion

1
at the exit of the burner by the ame transfer function. In
the low-frequency range, these uctuations can be described by a
gaintime lag model similar to that devised in [57]:

Q

=n[

1
]
t
c
. (15)
In many cases, well veried in recent studies [33,35,4043], one
nds that the velocity perturbation is convected toward the ame
front at a constant velocity that is on the order of the mean ow
velocity at the burner exit. The time lag
c
thus corresponds to
the mean time required for a perturbation at the burner outlet to
reach the ame surface area. The gain n is in general not constant.
It depends on the frequency of the ame oscillation and can be
deduced from the ame transfer function [4042,58]. For harmonic
perturbations, one obtains the following relations at the resonance
frequency:

Q =ne
i
0

c

1
. (16)
The previous analysis yields a complete set of equations governing
the system at the resonance
0
(Eqs. (13), (14), (16) and using
Eqs. (7) and (11)):
D. Durox et al. / Combustion and Flame 156 (2009) 106119 113
Fig. 10. Schematic representation of the organ pipe resonator.
p
1
= iK(r)

Q ,

Q =ne
i
0

c

1
,

1
= i
1
l
e

0
p
0
,
p
0
= i

0
2
p
1
.
The pressure eld radiated by the resonator may now be ex-
pressed from Eq. (6), where the velocity
e
=
1
. Some simple
algebra yields p
r
= (a

/l
e
)(/
0
)
2
p
0
, which becomes at resonance
p
r
= (a

/l
e
) p
0
. The dispersion equation obtained by combining
these four equations provides a compatibility condition [41]:

c
=
3
2
+2k, k =1, 2, . . . . (17)
Conditions at resonance can be summarized by

Q = in
1
(quadrature), (18)
p
1
= i
0
K(r)

Q (quadrature), (19)
p
0
=

2
0
2
K(r)

Q (in phase), (20)


p
r
=
_
a

l
e
_
p
0
(in phase). (21)
At resonance, the total pressure oscillation inside the burner cavity
p
0
is in phase with the heat release rate uctuation, so that the
pressure radiated by the resonator is also in phase with that uc-
tuation, a theoretical result that is found to be in agreement with
the Rayleigh criterion. Resonance is possible because the burner
acoustics yields a /2 phase shift between the total pressure
uctuation inside the burner p
0
and the external pressure uctua-
tion p
1
(Eq. (13)) generated by combustion noise.
4.2. Case of multiple ame burner (OPR)
In the multiple ame burner case, which operates like an organ
pipe resonator, the system can be simplied as shown in Fig. 10.
There are three regions: the upstream duct of length L bounded by
the piston, the perforated plate of thickness l
h
, and the external re-
gion where the ames are anchored. Index a designates conditions
upstream of the perforated plate, while b represents conditions
prevailing downstream of this element.
One may consider that only longitudinal waves propagate when
the duct diameter is much smaller than the wavelength under con-
sideration. This condition is well fullled in the present case [43].
Assuming harmonic perturbations, it is easy to show that on the
upstream side of the perforated plate, at a distance L from the pis-
ton head, providing a perfectly closed acoustic boundary condition,
the acoustic velocity
a
and pressure p
a
inside the duct are related
by [55]

a
= i
1
c
tan
_
L
c
_
p
a
.
In the absence of a perforated plate, the resonant motion cor-
responds to odd multiples of quarter wave modes dened by
discrete eigenfrequencies, f
n
= (2n + 1)c/4L (where n is an inte-
ger), without considering end corrections.
As the plate thickness is small with respect to acoustic wave-
lengths, the unsteady ow inside the small channels of the plate
can be treated as a bulk oscillation. A relation for the pressure
jump across the plate can be obtained using a momentum balance
applied to an isolated orice. For a bulk velocity perturbation in-
side the orice

h
and a mass of uid m
h
in each small channel,
one nds
m
h
d

h
dt
= S
h
(p

a
p

b
) 2
h
m
h

h
, (22)
where the damping coecient
h
describes friction losses at the
channel wall. This coecient only weakly depends on the angular
frequency and may be considered to be constant. In this ex-
pression p

a
are pressure perturbations induced on the upstream
side of the perforated plate due to external pressure modulations
p

b
. The mass m
h
may be written as m
h
= S
h
l
eh
, where S
h
is the
surface area of the hole and l
eh
corresponds to the size of an iso-
lated channel, augmented by a small length to take into account
the mass set into motion on each side of the plate. This amounts
to about one channel diameter (l
eh
= l
h
+ d). Equation (22) then
yields
p
a
p
b
= (2
h
i) l
eh

h
. (23)
The bulk velocity oscillation
h
can be related to acoustic variables
on the upstream side of the plate,

h
=

a
P
, (24)
where P is the perforated plate porosity, i.e., the ratio of the total
area of all the holes to the duct section.
In the absence of forcing ( p
b
= 0), the previous equations may
be combined to obtain a dispersion relation. Neglecting damping
phenomena, one obtains an equation for the resonant angular fre-
quencies of the system
0
, in the absence of combustion:
1

0
l
eh
P c
tan
_

0
L
c
_
=0. (25)
The resonant frequencies
0
depend on the burner and the per-
forated plate geometrical features. It was concluded from experi-
ments that damping is, indeed, quite weak when the perforated
plate is thin [43,49], and that the frequencies of self-sustained
oscillations, in the presence of ames, are close to the values de-
termined under nonreactive conditions.
Outside the burner, on the downstream side of the plate, one
may link the bulk velocity uctuation
h
, at the perforated plate
hole outlets, to the perturbations in heat release rate

Q and then
the pressure p
b
to

Q using similar relations to those given previ-


ously in the Helmholtz resonator case (see Eqs. (14) and (16)):

Q = Nn

e
i
c

h
, (26)
p
b
= iK

(r)

Q . (27)
The main difference from the preceding analysis for the HRB is
that one now has to consider the response to perturbations of a
collection of small conical ames and the noise radiated by such
an arrangement instead of dealing with an interaction with a sin-
gle ame. The coecient N is the number of holes of the plate,
114 D. Durox et al. / Combustion and Flame 156 (2009) 106119
n

is the gain of the transfer function of a single ame, and K

(r)
accounts for the contributions of all the ames to the pressure p
b
induced on the downstream side of the plate. The coecients n

and K

(r) are not identical to N and K(r) determined for the HBR
case, because the ame geometries and dynamical responses differ
[4043,49]. By combining Eqs. (21), (23), (24) and (27), one obtains
a relation for the acoustic pressure p
a
induced on the upstream
side of the plate due to external heat release rate uctuations:
p
a
_
1
l
eh
P c
tan
_
L
c
__
+ iK

(r)

Q = i
2
h
l
eh
P c
tan
_
L
c
_
p
a
. (28)
During self-sustained oscillations, as the damping in this system is
weak, the angular frequency at resonance remains close to the
frequency
0
determined without combustion. The expression (28)
simplies using Eq. (25):
p
a
=

2
0
2
h
K

(r)

Q . (29)
Results for the OPR at resonance are synthesized by the expres-
sions
p
a
=

2
0
2
h
K

(r)

Q (in phase),

Q = i Nn


h
(quadrature), (30)
p
b
= iK

(r)

Q (quadrature), (31)
and one retrieves the dispersion relation yielding conditions for
resonance [43,49], which are similar to those of Eq. (17):

c
=
3
2
+2k.
The pressure radiated back by the resonator may again be de-
duced from Eq. (6). The velocity uctuation near the resonator
end
e
may be assimilated in this case to the velocity
a
so that
p
r
= ic(ka

)
a
. It is here convenient to express this quantity in
terms of the pressure prevailing at the base of the resonator near
the piston head, p
r
= (1)
n
(ka

) p
H
, where n designates the mode
number (n = 0 for the quarter-wave mode, n = 1 for the three-
quarter-wave mode) and p
H
is the pressure signal at the base of
the manifold near the piston. Now, p
H
and p
a
are in phase for the
quarter-wave mode and out of phase for the three-quarter-wave
mode. It is thus clear that p
r
is in phase with p
a
.
Self-sustained oscillations in this multiple injection congura-
tion with unconned ames stabilized on a perforated plate termi-
nating a cylindrical manifold are such that pressure oscillations p
a
inside the cavity below the perforated plate and external heat re-
lease rate

Q uctuations are in phase, and that as a consequence


the pressure radiated back by the resonator p
r
is also in phase
with

Q , and that this component of the pressure eld satises


the Rayleigh criterion. This phase match is once again made possi-
ble because at resonance the burner acoustics yields a phase shift
of /2 between external pressure perturbations imposed by the
ame ( p
b
= p
ext
at the burner outlet) and the pressure uctuation
inside the cavity in the vicinity of the ame ( p
a
below the perfo-
rated plate).
It would be necessary to include the damping and a more
detailed model for the convective time
c
of the velocity perturba-
tion along the ame, to establish a more rened theoretical model.
But the difference between experiments and theory is only minor
and one may conclude that uctuations of the external heat re-
lease rate and the pressure inside the cavity, just upstream of the
perforated plate, and the pressure radiated back by the resonator
verify the Rayleigh criterion.
Based on the result
0

c
= 3/2, some solutions to ght in-
stabilities can be proposed. Knowing the acoustic resonance of a
Fig. 11. Time traces recorded during instability growth. L = 397 mm. Perforated
plate thickness: 22.5 mm. The pressure tap P2 is located 95 mm upstream from
the perforated plate. The peak on the PM signal corresponds to the ignition time.
The instability frequency is 540 Hz and remains nearly constant during all the rise.
system, and the corresponding angular frequency
0
, one may se-
lect a ame shape such that the convection time
c
will differ from
that required for unstable operation. It is also possible, but per-
haps not always easy, to modify the acoustics of the system and
thus
0
, to shift away from a region of instability.
4.3. Analysis of the instability growth
It is now interesting to see if the phase difference between
pressure and heat release does not change during the onset of
the instability and before saturation of this oscillation, as observed,
for example, in Ref. [10]. This can be done by choosing operating
conditions of the OPR, which allow a precise determination of the
phase difference evolution.
The ame holding plate is now replaced by a thicker plate
(22.5 mm thick instead of 3 mm). This increases damping inside
the holes and reduces the instability bands as a function of the
cavity length [43,49]. The linear amplication rate of oscillations
is reduced and it is then easier to observe the transient growth
of oscillation. The selected position of the piston corresponds to a
strong level of ame oscillation. An example of instability growth
is given in Fig. 11, where the time reference is arbitrary. From the
ignition time, materialized by the peak in the PM signal at about
t =0.2 s, the system takes approximately 0.5 s to reach saturation.
The growth of instability is well illustrated by the two signals in
close-up views of Fig. 11 during the transient period. At t = 0.56 s
(Fig. 12, rst plot) the two signals are slightly out of phase. The
phase difference is about 0.7 rad. With the growth in amplitude,
the signals become progressively locked in phase. The phase differ-
ence is 0.53 rad at t = 0.66 s and it nearly vanishes at t = 0.75 s
and subsequently remains close to zero. Other experimental con-
ditions, not presented here, indicate that phase differences corre-
sponding to the initial oscillation or to saturation remain close to
zero and always smaller than /4.
In the sense of the Rayleigh criterion, the product p

by

Q sup-
plies energy to the system. It is then interesting to examine opera-
tion at the limit cycle and see how the energy fed by combustion is
balanced by radiation and losses and see if this can lead to satura-
tion of the perturbation amplitude. Losses induced by damping are
mainly due to friction in the perforated plate, and depend on the
geometry and velocity uctuation inside the individual channels in
the ame holding plate. The power loss associated with damping
is essentially proportional to the square of the velocity uctuation.
Acoustic power radiated away from the burner is proportional to
the square of the pressure in the fareld. This pressure is linearly
related to the time rate of change of the heat release. When the
D. Durox et al. / Combustion and Flame 156 (2009) 106119 115
Fig. 12. Time traces of pressure and OH* emission during the growth of the instability. Same conditions as in Fig. 11.
perturbation is harmonic, the pressure is proportional to the heat
release uctuations, which is in turn linearly related to the veloc-
ity uctuations in the system. It is then apparent that the source
of power, the loss by damping, and the radiated acoustic power
are all proportional to the square of the velocity perturbation in
the system. This indicates that saturation of perturbations is not a
consequence of a balance between amplication and loss terms.
Saturation is mainly caused by the nonlinearity of the ame re-
sponse. Experiments indicate that this response diminishes when
the velocity uctuation level is increased [59]. This progressively
reduces the growth of the product of p

by

Q , which nally
reaches an equilibrium level. The power radiated by the system
is merely a result of this process. The value of the radiated power
is then determined by subtracting damping losses from the source
term. This is illustrated in the next section in one specic case but
the result is probably much more general.
5. Acoustic energy balance
We now consider the OPR conguration and examine the bal-
ance of acoustic energy (1) when the instability has reached a limit
cycle. The energy density E may be evaluated in terms of rms
pressure and velocity, while the acoustic energy ux and acoustic
source may be determined by integration over a period of oscilla-
tion:
E =
1
2
p
2
rms

0
c
2
+
1
2

2
rms
, (32)
F =
1
T
T
_
0
p

1
dt, (33)
Fig. 13. Control volume used to examine the balance of acoustic energy in the OPR
burner case. A microphone M2 located at a distance of 34.5 cm from the burner
axis records the pressure level on the outer control surface A.
S =
1
T
T
_
0
1

0
c
2
p

dt. (34)
It is convenient to integrate Eq. (1) over the control volume shown
in Fig. 13. This yields

t
E +
_
A+
F ndA =
_
V
f
S dV 2
0
E, (35)
where E is the integral of E over the volume V
b
. The last term
of Eq. (35) corresponds to the integration of the dissipation D
over V
b
. This term is represented by assuming that the acous-
tic energy in the system decays exponentially at a rate
0
. The
damping factor of the system and the eigenfrequency
0
can
116 D. Durox et al. / Combustion and Flame 156 (2009) 106119
be deduced from the resonant response of the burner to an ex-
ternal excitation. The product
0
corresponds to the damping
coecient of the oscillating system. For a steady self-sustained
oscillation the integrated energy remains constant and the rst
term in Eq. (35) vanishes. The second term of this equation is re-
duced to
_
A+
F ndA A
(p
2
rms
)
M

0
c
, (36)
where the acoustic pressure is measured by the microphone M2
on the control surface A in the fareld region. In the previous ex-
pression it is implicitly assumed that radiation is nearly isotropic
and the rms pressure measured by microphone M2 is typical of
the rms pressures at other points of the control surface.
The ame is compact with respect to the acoustic wavelength
and the pressure uctuation may be considered to be uniform in
this region so that the third term of Eq. (35) can be written
_
V
f
S dV =
1

0
c
2
1
T
_
T
p


Q

dt, (37)
where

Q

represents the heat release rate uctuation integrated
over the ame region V
f
. It is convenient to introduce the relative
uctuation of heat release rate by multiplying and dividing by the
mean heat release rate

Q
0
. It is also interesting to use the ratio
between the pressure uctuation and the ambient pressure p
0
:
_
V
f
S dV =

Q
0
1

1
T
_
T
p

p
0

Q
0
dt. (38)
The energy density appearing in the fourth term of Eq. (35) can
be determined in one of the sections of the upstream manifold.
This quantity is constant in this volume and the last term is sim-
ply obtained by multiplying the value of acoustic energy density
calculated in one section (for example in Section 1, Fig. 14) by
the volume V
b
of the upstream manifold. These various operations
nally yield:
A
(p
2
rms
)
M

0
c
=

Q
0
1

1
T
_
T
p

p
0

Q
0
dt
2
0
_
1
2
(p
2
rms
)
1

0
c
2
+
1
2

0
_

2
rms
_
1
_
V
b
. (39)
The three terms of Eq. (39) can now be estimated by examining
a typical limit cycle oscillation. This is illustrated in the OPR case,
with a 3-mm-thick perforated plate, and for a tube length xed
at 326 mm. Pressure measurements were made with the ve taps
P1 to P5 (Fig. 2). The acoustic velocity
rms
is measured with a
hot wire on the upstream side of the plate, in the P1 section (see
Fig. 14). Heat release rate uctuations are measured by a pho-
tomultiplier, equipped with a bandpass lter detecting the light
emission of OH* radicals. The radiated noise is measured by a mi-
crophone, located 34.5 cm off the tube axis in the plane of the
perforated plate. The instability frequency is 750 Hz and the sys-
tem oscillates on a 3/4 wave mode. Pressure levels in the system
are plotted in rms values in Fig. 14 which shows the envelope
of pressure oscillation, corresponding to this mode. This wave-
form is obtained by tting the pressure amplitudes obtained at
the ve microphone locations along the duct. The wavelength of
0.472 m (appearing in the tting equation) is close to the value
of 0.468 m expected for a frequency of 750 Hz and a speed of
sound of 351 ms
1
(corresponding to a mixture of methane and
air at an equivalence ratio of 0.86, and at ambient temperature).
The tting curve shows that the rms pressure does not vanish in
Fig. 14. Envelope of the pressure oscillation when the system is in self-sustained os-
cillation. The pressure probes are in ports 1 to 5. The length of the tube is 0.326 m.
The grid is 3 mm thick. Triangle symbols correspond to pressure measurements un-
der self-sustained oscillation. A hot wire is placed in the same section as P1.
Table 1
Numerical values of the main variables appearing in the acoustic energy balance
Power
(kW)

Q

rms
/

Q
0
P1
(Pa)
P5
(Pa)
P
F
(Pa)
P
M
(Pa)
(
rms
)
1
(ms
1
)
2
0
(s
1
)
13.3 0.136 175 280 57 8.74 0.37 37.7
the ame region, just downstream of the plate. From this curve,
the mean pressure in the compact combustion zone can be evalu-
ated. It is shown, in Ref. [49], that the ame height is 12 mm and
that the maximum reactive surface variation takes place at a dis-
tance of 8 mm from the plate. One obtains in this way a mean rms
pressure of 57 Pa at this distance.
The quality factor of the acoustic OPR resonator, 1/(2 ), was
determined previously (Ref. [43]). A width f at half-height of the
resonance peak of 6 Hz was obtained, which corresponds to a rel-
atively small damping. This value gives 2
0
= 2f 37.7 s
1
.
The power released in the burner is in the steady state

Q
0
=
13.3 kW. At the limit cycle, the relative uctuation in heat release
rate is

Q

/

Q
0
= 13.6%. In Section 1, corresponding to the P1 probe,
the rms pressure value is 175 Pa, the mean velocity in the duct
is 1.19 ms
1
, and the velocity rms uctuation is 0.37 ms
1
. The
sound pressure level measured at 34.5 cm is (p
rms
)
M
8.74 Pa.
Directivity measurements, not shown in this article, indicate that
emission is almost isotropic around the burner (less than 1 dB
variation).
The values used to evaluate the various terms in the balance of
energy are gathered in Table 1. The rst term in Eq. (39) corre-
sponds to a net outow of power:
A
(p
2
rms
)
M

0
c
=0.260 W.
In the case under investigation, pressure perturbations in the tube
are nearly in phase with heat release rate uctuations (Fig. 8) and
the source term corresponds to a power of

Q
0
1

(p

rms
)
F
p
0

Q

rms

Q
0
= 0.294 W,
where (p

rms
)
F
is the rms pressure in the ame zone.
The dissipation term is easy to estimate:
2
0
_
1
2
(p
2
rms
)
1

0
c
2
+
1
2

0
_

2
rms
_
1
_
V
b
=0.009 W.
This indicates that the power dissipated in the system is negligible
when compared to the other terms in the balance of energy. This
is due to the fact that friction losses at the duct walls are small
D. Durox et al. / Combustion and Flame 156 (2009) 106119 117
Fig. 15. On the left: Time signals of the pressures p
H
measured at the level of the piston head and p
M
measured by a microphone placed downstream the perforated plate.
This microphone is moved along the axis at various distances from the plate (3 mm up to 100 mm). On the right: Time signals of p
H
and p
r
.
and that very little energy is dissipated in the relatively thin perfo-
rated plate. The source term and the radiated acoustic power have
the same order of magnitude and the balance of energy is reason-
ably well fullled. An energy balance is obtained to within 10%,
which is probably the greatest accuracy that could be expected. It
is dicult to get a more precise result, mainly because the value
of the pressure perturbation inside the ame can only be deduced
from measurements inside the duct. The pressure perturbation in
the ame region is determined by extrapolating the standing wave
inside the tube, and this pressure varies substantially in the region
under consideration. A variation of a few millimeters in the curve
displayed in Fig. 14 induces important variations of the pressure
and modies the source term. While a perfect balance is not ob-
tained, the previous estimates indicate that the energy budget can
be derived from the balance of acoustic energy which includes the
Rayleigh source term. The energy provided by this term is evac-
uated in the form of acoustic radiation with only small acoustic
losses due to dissipation in the resonator.
At this point it is worth indicating that closing the energy bud-
get in an experiment is dicult and not often attempted in other
investigations, but that this can be done numerically, as exempli-
ed in [18], where the instantaneous terms are estimated and the
acoustic energy balance is closed without approximation.
6. Conclusion
The Rayleigh criterion is examined in two cases of thermoa-
coustic self-sustained oscillations involving unconned ames. In
the rst case the burner behaves like a Helmholtz resonator and
responds to pressure perturbations radiated by a V- or an M-
ame. In the second case the burner equipped with a perforated
plate resonates like an organ pipe and features longitudinal acous-
tic modes. In these two generic cases, it is found that pressure
perturbations inside the burner just upstream of the ame region
are nearly in phase with perturbations in heat release rate in the
ame. The pressure radiated by the system is evaluated and it is
shown that this component of the waveeld is in phase with heat
release uctuations in the ame region. This conclusion applies
in the two experimental situations considered in this study and
probably holds for other unconned geometries where the ame
is established at the open end of a burner. In conned cases this
may not be applicable, but it is often observed that pressure on
the upstream side and in the vicinity of the injection plane is in
phase with heat release uctuations. Energy is fed into the self-
sustained thermoacoustic instability, as implied by the standard
version of the Rayleigh criterion. This criterion must be fullled
to obtain self-sustained oscillations, but this is only a necessary
condition. Experimental data show that the actuation of the sys-
tem by the external pressure perturbations emitted by the ame
must be in quadrature with the internal pressure. The mechanism
is most effective when the maximum of the external pressure co-
incides with the maximum of growth rate of the pressure inside
the burner. Theoretical predictions also indicate that a time delay
is necessary to obtain an instability between the velocity uctua-
tion at the burner exhaust and heat release rate uctuations. This
delay must be such that <
0

c
< 2 modulo 2, with a maxi-
mum of eciency for
0

c
= 3/2 modulo 2. This relation, well
demonstrated in the unconned ame experiments, is also valid
in conned situations [59]. When the system operates on a limit
cycle, the power fed into the resonant mode must balance the
losses associated with damping and radiation to the fareld. This
is well veried here, in the case of the organ pipe burner, by es-
timating the Rayleigh source term, the power dissipation, and the
integrated energy ux radiated away from the ame. A balance
is approximately obtained in which the power generated by the
Rayleigh source term, minus a small amount of damping losses
in the system, determines the level of acoustic radiation to the
fareld.
Appendix A. The pressure radiated by a resonator
It is interesting to illustrate the process that gives rise to the
radiated pressure component of the sound eld. This is done here
by submitting the OPR (Fig. 2) to an external excitation produced
by a loudspeaker set at 35 cm from the outlet, in the exit plane.
A microphone is placed near the piston head and measures p
H
;
another microphone is placed near the perforated plate. This mi-
crophone measures the sum of the incident eld generated by the
loudspeaker and the radiated pressure: p
M
= p
i
+ p
r
. The incident
eld p
i
can be measured by displacing the resonator and recording
the signal detected by the same microphone. In this experiment
the duct length is 340 mm. The resonance takes place at a quarter-
wave mode frequency f =239 Hz.
Fig. 15a displays the signal detected at the piston head p
H
to-
gether with signals measured by microphone M at various axial
distances from the perforated plate. It is found that near the plate
the signal p
M
is almost in phase with the pressure inside the duct
characterized by p
H
. As microphone M is displaced away from
the duct end, the radiated pressure component diminishes, the di-
rect signal p
i
takes over, and the signal detected by M comes to
be in quadrature with the internal pressure p
H
. This is consis-
tent with resonance conditions and with the eld decomposition
118 D. Durox et al. / Combustion and Flame 156 (2009) 106119
in terms of an excitation component and a radiated pressure. It is
also interesting to extract p
r
from the measurements by substract-
ing the direct eld p
i
from the pressure signal measured close to
the perforated plate. The direct eld is obtained by removing the
resonator and recording the signal in this conguration. The ex-
tracted radiated pressure, p
r
= p
M
p
i
and the pressure at the
piston head are plotted in Fig. 15b, which also shows the theoret-
ical estimate ( p
r
)
th
= (1)
n
(ka

) p
H
for n = 0 and for a duct radius
is a = 35 mm, so that one may estimate ( p
r
)
th
0.13 p
H
. As ex-
pected, the three signals p
H
, p
r
and (p
r
)
th
are in phase. The small
difference between p
r
and (p
r
)
th
can be due to the inuence of
the perforated plate. This is not taken into account in the present
analysis.
This experiment supports the modeling adopted for the sound
eld in the vicinity of the duct end.
Appendix B. An acoustic source in a semi-innite duct
To interpret results obtained in the OPR case it is interesting
to consider a simplied model problem in which a heat release
source term is placed in a semi-innite duct. This problem serves
to show that when the duct situated on the upstream side of the
ame operates at resonance, the pressure eld in the vicinity of
the source is in phase with heat release. This can then be used
to explain the more complex situation described in the OPR case
analyzed in this article. In this analysis, it is assumed that pressure
perturbations are harmonic. The duct has a rigid termination at
x = 0 so that ( p/x)
0
= 0. The duct extends to +. An acoustic
point source is located at x = x
0
. The pressure eld satises the
Helmholtz equation
d
2
p
dt
2
+k
2
p = (x x
0
). (B.1)
This equation, together with the boundary conditions, yields the
following pressure eld. In the upstream duct (for x x
0
),
p =
coskx
ik coskx
0
+k sinkx
0
. (B.2)
On the downstream side of the source (for x > x
0
),
p =
coskx
0
ik coskx
0
+k sinkx
0
exp
_
ik(x x
0
)
_
. (B.3)
Assuming a resonance condition sinkx
0
= 0, one nds that on the
upstream side of the source,
p =
coskx
ik coskx
0
. (B.4)
If one now considers harmonic heat release rate uctuations local-
ized in the source section, the pressure eld is given by
d
2
p
dt
2
+k
2
p =
1
c
2
i

Q

0
A
(x x
0
), (B.5)
where

Q

0
is the total unsteady heat release and A is the duct cross
section. Using expression (B.4), it is easy to express the pressure
eld on the upstream side of the source,
p =
coskx
coskx
0
1
c

Q

0
A
, (B.6)
which indicates that, at resonance, the pressure eld is in phase
with heat release uctuations. This result is in agreement with
what is observed in the OPR system, indicating that the duct re-
sponse to heat release uctuations in the source region is essen-
tially determined by the combination of the incident and reected
waves giving rise to a resonant mode in the section between the
duct origin and the source.
References
[1] J.W.S. Lord Rayleigh, Nature 18 (1878) 319321.
[2] A.D. Pierce, Acoustics: An Introduction to Its Physical Principles and Applica-
tions, McGrawHill, New York, 1981.
[3] R.H. Cantrell, R.W. Hart, J. Acoust. Soc. Am. 36 (4) (1964) 697706.
[4] B.-T. Chu, Acta Mech. 3 (1) (1965) 215234.
[5] M.K. Myers, J. Fluid Mech. 226 (1991) 383400.
[6] S. Candel, J. Sound Vibrat. 41 (1975) 207232.
[7] F. Nicoud, T. Poinsot, Combust. Flame 142 (2005) 153159.
[8] L. Crocco, Proc. Combust. Inst. 12 (1968) 8599.
[9] A.A. Putnam, Combustion Driven Oscillations in Industry, Elsevier, New York,
1971.
[10] T. Poinsot, A. Trouv, D. Veynante, S. Candel, E. Esposito, J. Fluid Mech. 177
(1987) 265292.
[11] T.C. Lieuwen, B.T. Zinn, Proc. Combust. Inst. 27 (1998) 18091816.
[12] C.O. Paschereit, E. Gutmark, W. Weisenstein, Proc. Combust. Inst. 27 (1998)
18171824.
[13] F.E.C. Culick, Dynamics of combustion systems: Fundamentals, acoustics and
control, in: Proceedings of the RTO/VKI Special Course Active Control of Engine
Dynamics (VKI), Rhode Saint Genese, Belgium, 2001, pp. 6.16.133.
[14] S. Candel, Proc. Combust. Inst. 29 (2002) 128.
[15] A.P. Dowling, S.R. Stow, J. Propuls. Power 19 (2003) 751764.
[16] T. Lieuwen, K. McManus, J. Propuls. Power 19 (5) (2003) 721.
[17] J. Lee, D. Santavicca, in: T. Lieuwen, V. Yang (Eds.), Combustion Instabilities in
Gas Turbine Engines: Operational Experience, Fundamental Mechanisms, and
Modeling, AIAA, Reston, VA, 2005, pp. 481529.
[18] C.E. Martin, L. Benoit, Y. Sommerer, F. Nicoud, T. Poinsot, AIAA J. 44 (4) (2006)
741750.
[19] D.M. Kang, F.E.C. Culick, A. Ratner, Combust. Flame 151 (3) (2007) 412425.
[20] B. Bellows, T. Lieuwen, Nonlinear response of a premixed combustor to forced
acoustic oscillations, in: 42nd AIAA Aerospace Science Meeting, 58 January
2004, Reno, Nevada, AIAA Paper 2004-455 (2004).
[21] M. Ohtsuka, S. Yoshida, S. Inage, N. Kobayashi, Combustion oscillation analysis
of premixed ames at elevated pressures, ASME Turbo Expo 98, ASME Paper
98-GT-581 (1998).
[22] C.J. Lawn, W. Polifke, Combust. Sci. Technol. 176 (2004) 13591390.
[23] R. Balachandran, B.O. Ayoola, C.F. Kaminski, A.P. Dowling, E. Mastorakos, Com-
bust. Flame 143 (2005) 3755.
[24] C.A. Armitage, R. Balachandran, E. Mastorakos, R.S. Cant, Combust. Flame 146
(2006) 419436.
[25] G.A. Richards, D.L. Straub, E.H. Robey, J. Propuls. Power 19 (2003) 795810.
[26] T. Poinsot, D. Veynante, Theoretical and Numerical Combustion, Edwards,
Philadelphia, 2001.
[27] C. Stone, S. Menon, J. Turb. 4 (2003) 114.
[28] Y. Huang, H.-G. Sung, S.-Y. Hsieh, V. Yang, J. Propuls. Power 19 (2003) 782794.
[29] L. Selle, G. Lartigue, T. Poinsot, R. Koch, K.-U. Schildmacher, W. Krebs, B. Prade,
P. Kaufmann, D. Veynante, Combust. Flame 137 (2004) 489505.
[30] Y. Matsui, Combust. Flame 43 (1981) 199209.
[31] M. Fleil, A.M. Annaswamy, Z.A. Ghoneim, A.F. Ghoniem, Combust. Flame 106
(1996) 487510.
[32] A. Dowling, J. Fluid Mech. 394 (1999) 5172.
[33] S. Ducruix, D. Durox, S. Candel, Proc. Combust. Inst. 28 (2000) 765773.
[34] K.R.A.M. Schreel, E.L. van den Tillart, R.W.M. Janssen, L.P.H. de Goey, in: G.D.
Roy, K.H. Yu, J.H. Whitelaw, J.J. Witton (Eds.), Advances in Combustion and
Noise Control, Craneld Univ. Press, Bedford, 2005, pp. 247258.
[35] A.L. Birbaud, D. Durox, S. Ducruix, S. Candel, Proc. Combust. Inst. 31 (2007)
12571265.
[36] C. Klsheimer, H. Bchner, Combust. Flame 131 (2002) 7084.
[37] C.A. Armitage, A.J. Riley, R.S. Cant, A.P. Dowling, S.R. Stow, Flame transfer func-
tion for swirled LPP combustion from experiments and CFD, in: Proc. ASME
Turbo Expo 2004, Vienna, Austria, June 1417, 2004.
[38] A.N. Lipatnikov, P. Sathiah, Combust. Flame 142 (2005) 130139.
[39] S.H. Preetham, T.C. Lieuwen, Proc. Combust. Inst. 31 (2006) 14271434.
[40] D. Durox, T. Schuller, S. Candel, Proc. Combust. Inst. 29 (2002) 6975.
[41] T. Schuller, D. Durox, S. Candel, Combust. Flame 135 (2003) 525537.
[42] D. Durox, T. Schuller, S. Candel, Proc. Combust. Inst. 30 (2005) 17171724.
[43] N. Noiray, D. Durox, T. Schuller, S. Candel, Combust. Flame 145 (2006) 435
446.
[44] E.C. Fernandes, R.E. Leandro, Combust. Flame 146 (2006) 674686.
[45] R. Price, I. Hurle, T. Sugden, Proc. Combust. Inst. 12 (1969) 10931102.
[46] T. Schuller, D. Durox, S. Candel, Combust. Flame 128 (2002) 88110.
[47] S. Candel, D. Durox, T. Schuller, Flame interactions as a source of noise and
combustion instabilities, AIAA Paper 2004-2928, 2004.
[48] T. Schuller, N. Noiray, D. Durox, S. Candel, On mechanisms of intense combus-
tion noise emission, in: 13th Int. Congr. Sound and Vibration (ICSV13), Vienna,
Austria, July 26, 2006.
D. Durox et al. / Combustion and Flame 156 (2009) 106119 119
[49] N. Noiray, D. Durox, T. Schuller, S. Candel, Proc. Combust. Inst. 31 (2007)
12831290.
[50] B.D. Bellows, M.K. Bobba, A. Forte, J. Seitzman, T. Lieuwen, Proc. Combust.
Inst. 31 (2007) 31813188.
[51] Y. Huang, V. Yang, Combust. Flame 136 (3) (2004) 383389.
[52] I. Hurle, R. Price, T. Sugden, A. Thomas, Proc. R. London Ser. Soc. A 303 (1968)
409427.
[53] P. Clavin, E. Siggia, Combust. Sci. Technol. 78 (1991) 147155.
[54] S. Rienstra, A. Hirschberg, An introduction to acoustics, Report IWDE 92-06,
Eindhoven University of Technology (2005).
[55] P.M. Morse, K.U. Ingard, Theoretical Acoustics, McGrawHill, New York, 1968.
[56] D.W. Bechert, J. Sound Vibrat. 70 (3) (1980) 389405.
[57] L. Crocco, S.I. Cheng, Theory of Combustion Instability in Liquid Propellant
Rocket Motors, Butterworths Science Publication, London, 1956.
[58] T. Schuller, D. Durox, S. Candel, Combust. Flame 134 (2003) 2134.
[59] N. Noiray, D. Durox, T. Schuller, S. Candel, A method for estimating the
noise level of unstable combustion based on the ame describing func-
tion, in: 11th CEAS-ASC Workshop on Experimental and Numerical Analy-
sis and Prediction of Combustion Noise, Lisbon, Portugal, 2728 September
2007.

S-ar putea să vă placă și