You are on page 1of 15

Combustion and Flame 162 (2015) 10631077

Contents lists available at ScienceDirect

Combustion and Flame


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / c o m b u s t fl a m e

Time-resolved PIV of lean premixed combustion without and with


porous inert media for acoustic control
Joseph Meadows, Ajay K. Agrawal
Department of Mechanical Engineering, University of Alabama, Tuscaloosa, AL 35487, United States

a r t i c l e

i n f o

Article history:
Received 29 April 2014
Received in revised form 28 August 2014
Accepted 30 September 2014
Available online 21 October 2014
Keywords:
Lean premixed combustion
Thermo-acoustic instability
Passive control
Porous inert media
Combustion noise
Combustion dynamics

a b s t r a c t
In the past, we have documented the efcacy of the ring-shaped porous inert media (PIM) to passively
mitigate combustion noise and thermo-acoustic instabilities in lean premixed (LPM) swirl-stabilized
combustion systems, although the underlying mechanisms involved have not been understood. The present study utilizes time-resolved particle image velocimetry (PIV) to measure the turbulent ow eld in a
LPM combustion system without and with PIM placed on the dump plane. Measurements of instantaneous and average ow elds, vorticity, and turbulent kinetic energy are presented to gain insight into
the ow structure. Results are analyzed using the proper orthogonal decomposition (POD) to identify
the effect of PIM on the coherent turbulent structures in the ow eld. The addition of PIM removes
the recirculation zones and the dominant coherent turbulent structures formed in the shear layers are
convected downstream of the reaction zone. Harmonic reconstruction of the ow eld is performed at
the frequency of thermo-acoustic instability to identify the relationship between the ow and acoustic
elds of the system. Without PIM, the vortical modes in the corner recirculation zone are shown to be
the driving source for the instability. PIM alters the ow eld from global instability to convective instability, which effectively eliminates the feedback mechanism for thermo-acoustic instabilities in the present LPM swirl-stabilized combustion system.
2014 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

1. Introduction
Due to stringent emission regulations, lean premixed (LPM)
combustion has gained increased utility for power generating gas
turbines and related applications. LPM combustion strategies
reduce the ame temperature to decrease the thermal NOx formation. However, LPM combustion is more susceptible to thermoacoustic instabilities [1]. Recirculation zones are commonly found
in LPM swirl-stabilized combustion, and are known as sources of
combustion instabilities arising from vortex interactions with the
ame front [2]. For example, the precessing vortex core (PVC)
develops when a central vortex core precesses around the axis of
symmetry at a well-dened frequency, and it usually contributes
to vortex breakdown and the associated recirculation zone in a high
Reynolds number ow [3,4]. Another important feature of swirl stabilized ows is the formation of shear layers at the interface of uid
streams with different velocities [5]. The hydrodynamic instability
of shear layers is known as the KelvinHelmholtz instability, and it
is usually convectively unstable [5].
Corresponding author.
E-mail addresses: jwm3320@bellsouth.net (J. Meadows), AAgrawal@eng.ua.edu
(A.K. Agrawal).

One of the most challenging aspects of LPM combustor design is


the consideration of combustion noise and thermo-acoustic instabilities. Combustion noise is generated by heat release rate uctuations in a turbulent ow eld [6]. Putnam [7] and Strahle [8]
developed analytical models and empirical data to correlate sound
pressure level (SPL) with operating parameters such as airfuel
ratio, fuel type, reactants ow rate, and geometry in non-premixed
combustion systems. Under certain conditions, sound waves can
couple with the natural acoustic modes of the combustor and constrain the operational limits or cause catastrophic fatigue failure
[5]. According to the Rayleighs criterion, thermo-acoustic instabilities develop if the heat release rate uctuations are in phase with
the pressure uctuations [9] since the acoustic energy is added to
the system faster than it is dampened. In this case, the oscillation
amplitude will initially increase exponentially until it saturates
at some limit cycle [10].
The mechanisms responsible for exciting thermo-acoustic
instabilities are numerous and some are still unknown. Thermoacoustic instabilities often involve interactions between several
physical phenomena such as unsteady ame propagation leading
to unsteady ow velocities, acoustic wave propagation, and hydrodynamic instabilities [11]. Vortex formation in combustion chambers is also one of the main sources of ame/acoustic coupling [12].

http://dx.doi.org/10.1016/j.combustame.2014.09.028
0010-2180/ 2014 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

1064

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

The time delay between vortex formation and the instant of heat
released by combustion in the vortex can provide the phase relationship between the oscillating pressure eld and unsteady heat
release rate to drive the instability [11]. As such, uctuating vortical ow structures tend to excite pre-existent pressure oscillations,
thereby increasing the likelihood of combustion instabilities [13].
In general, these instabilities occur near frequencies associated
with the combustors natural longitudinal, radial, azimuthal, or
bulk modes [14]. The acoustic eld can also excite hydrodynamic
ow instabilities, which leads to large organized vortical structures
near the ame front to further introduce uctuations in the heat
release rate [15,16]. The coupling between acoustic and hydrodynamic modes can occur through non-linear mechanisms, and the
acoustic and hydrodynamic frequencies do not have to match to
create thermo-acoustic instabilities [17,18].
A number of active and passive control techniques have been
utilized to mitigate thermo-acoustic instabilities [1923]. Active
methods usually consist of actuation of the fuel or air delivery system in response to the ame behavior while passive methods seek
to modify the combustor geometry, for example, by bafes, resonators, and acoustic liners [22]. When instability occurs at one dominant frequency, passive techniques such as Helmholtz resonators,
matching upstream and downstream lengths, gradual diameter
changes, and locating an orice at the antinodes of the quarter
wave have all proven to be moderately effective at reducing
thermo-acoustic instabilities, however, their practical implementation in actual combustor design is often difcult [20]. Hermann
et al. [24] attached cylindrical extensions to the burner nozzle
and inclined them by 10 degree with respect to ow axis to reduce
the formation of coherent structures, and thus, displace the combustion zone downstream of its former position to increase the
time lag to passively control the thermo-acoustic instabilities in
a 260 MW heavy duty gas turbine.
Agrawal and Vijaykant [25] developed a passive technique to
mitigate combustion noise and thermo-acoustic instabilities in
swirl-stabilized, LPM combustion systems. This passive technique
involves placing an open-cell structure of porous inert material
(PIM) at the dump plane of the combustor. PIM is a ceramic annular
ring alloyed with hafnium carbide/silicon carbide (HfC/SiC) layered
coating to resist high temperature oxidation in combustion environments, and can withstand operating temperatures up to
1800 C [25]. PIM could be designed to t a wide range of dimensions for conventional combustors using advanced additive manufacturing techniques, and as such can potentially extend the
operational limits of current gas turbine engines without costly
design modications. The porous material itself is characterized
by porosity (percentage of void volume) and pore density expressed
in terms of pores per cm (ppcm). Both porosity and pore density of
the insert affect the ow structure and pressure drop across the PIM
[26]. Marbach and Agrawal [26] have shown that PIM can be used to
extend the lean blow-off limit of LPM combustion. Since, a high
porosity PIM is desired to minimize the pressure drop, this and subsequent studies have utilized PIM inserts with 85% porosity, i.e.,
only 15% of the ow area of the insert is blocked to the ow.
Sequera and Agrawal [1] used simplied computational uid
dynamics analysis to gain preliminary understanding of the ow
eld without and with porous inserts. The combustor was modeled
using axisymmetric geometry with swirling ow in conjunction
with a turbulent premixed combustion model based on the work
of Zimont [27]. The 2D axisymmetric model revealed that the porous insert eliminates the corner recirculation zone, strengthens the
swirling ow and central recirculation zone, and directs some
combustion products through the porous insert. A series of experiments were performed to assess the efcacy of PIM at mitigating
combustion dynamics in swirl-stabilized LPM combustion [1]. It
was shown that a divergent PIM annulus with pore density of 18

ppcm (45 ppi) provided the greatest reduction in the total SPLs.
Pore density of less than 18 ppcm caused the ame to stabilize
within the porous structure, which decreased the ability of the
PIM to mitigate combustion noise and instabilities. Sequera and
Agrawal [6] and Williams and Agrawal [28] demonstrated the
effectiveness of porous insert to reduce combustion noise and/or
instabilities for a range of operating conditions. Smith [29] and
Borsuk et al. [30] conducted experiments using a diffuser shape
porous annular ring insert with pore density of 18 ppcm. Experiments were conducted by varying the ow operating conditions
as well as the axial location of the swirler upstream of the combustor dump plane. In these studies, thermo-acoustic instabilities
were observed without PIM in the combustor. In all cases, the
thermo-acoustic instabilities were mitigated in the presence of
PIM inserts with peak SPL reduction of up to 30 dB.
While the PIM inserts are effective in mitigating combustion
dynamics, and the responsible mechanisms have been hypothesized and discussed in the previous literature, a detailed understanding of the underlying mechanisms remains unknown.
Meadows and Agrawal [31] determined the absorption coefcient
of PIM as a function of temperature and frequency, and for lean
direct injection (LDI) combustion found that only a small fraction
of the acoustic energy mitigated by porous media could be contributed to the acoustic absorption of the material. The study hypothesized that other mechanisms such as, alteration of the turbulent
ow eld, redistribution of the heat release rates and changes in
fuel atomization and evaporation processes in the presence of
PIM inserts help to mitigate combustion instabilities. Meadows
and Agrawal [32] utilized time-resolved PIV to investigate the
non-reacting swirl-stabilized ow eld without and with PIM for
acoustic control. POD analysis of data showed that PIM results in
a more uniform distribution of turbulent energy in the ow eld.
The precessing vortex core and corner recirculation zone were also
eliminated with PIM.
The present study seeks to quantify the effects of PIM on the
turbulent structures in a LPM swirl-stabilized combustion system,
and to identify the coherent uctuations in the ow eld at the frequency at which thermo-acoustic instabilities are observed. The
overall goal is to identify and explain the mechanisms responsible
for mitigating combustion noise and thermo-acoustic instabilities
in LPM swirl-stabilized combustion with PIM. First, the static stability and dynamic stability of the ame are discussed by analyzing
ame photographs and dynamic pressure data. Next, time-resolved
PIV technique is used to provide insight into instantaneous and
time-averaged ow elds. POD analysis is used to quantify the
effect of PIM on the dominant coherent turbulent structures, and
harmonic reconstruction of the ow eld is performed at the instability frequency to explore the coupling between the ow eld and
thermo-acoustics of the system.
2. Experimental setup
Figure 1 shows the swirl-stabilized combustor setup oriented
vertically. The experiment is operated at atmospheric conditions.
After passing through a pressure regulator, dehumidier, and
water trap, the air ow is split into primary air and seeded air ow
lines. The air ow is seeded with approximately 2 lm diameter
titanium dioxide particles produced by a custom designed solid
particle seeder located 2.0 m upstream of the test setup to allow
for proper mixing. The seeded air rst ows through a plenum
lled with marbles to breakdown the large vortical structures
and to further homogenize the seed particles in the air ow. Then,
the air ows through the mixing tube and enters the cylindrical
combustor via a swirler in an annulus with outer diameter of
4 cm and inner diameter of 2 cm. The swirler consists of six vanes
at 28 to the horizontal plane to produce a swirl number of 1.5. The

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

D = 8.0 cm
Exit
Plane

Quartz
Combustor

L = 30.5 cm

D = 4.8 cm

Dump
Plane

PIM

Swirler

Mixing
Tube

Reactants
Fig. 1. Schematic of combustor setup.

quartz cylindrical enclosure is 30.5 cm long with 8.0 cm inner


diameter. PIM consists of two porous rings stacked on top of each
other, with inner and outer diameters of 48 mm and 78 mm,
respectively. Each PIM ring has height of 25.4 mm and the pore
density of 18 pores per cm.
The primary air ow rate is controlled by a manual valve and
measured using a laminar ow element (LFE) with reported calibration error of 5 liters per a minute (lpm). The total pressure
and pressure drop in the LFE are measured with an absolute pressure transducer and a differential pressure transducer respectively.
The measured air ow rate is corrected for temperature as specied by the LFE manufacturer. The seeded air ow rate is controlled
by the supply pressure and measured using a sonic nozzle (FlowMaxx SN08-SA-062), with a 60/40 percent split between the ow
rates of primary air and seeded air. Methane fuel ow rate is measured and controlled by mass ow controller with measurement
uncertainty of 1.5 liters per a minute. The total air ow rate entering the combustor was set to 3.12 g/s (150 SLPM). Methane enters
the mixing tube approximately 61.0 cm upstream of the dump
plane at a ow rate set to obtain an equivalence ratio of 0.79.
Sound measurements are acquired by Bruel & Kjaer Model 8149
condenser microphone. The condenser microphone was placed at
the exit plane of the combustor, and 30.5 cm away from the combustor outer wall to prevent the overheating of the sensor. Microphone output is converted to pressure uctuations using the
sensitivity (45.8 mV/Pa) of the condenser microphone with a
dynamic range of 16.5134 dB and an uncertainty of 0.1 dB. The
condenser microphone is calibrated by a piston-phone generating
pure tone of 114 dB at 251.2 Hz. The sound pressure measurements are sampled at 40 kHz for one second resulting in a frequency resolution of 1 Hz. Sampled data are processed using
LabView embedded fast Fourier transform (FFT) function to obtain
the sound pressure spectra. The sound measurements were used
mainly to isolate the acoustic frequency and to quantify the

1065

difference in the SPL without and with PIM. The pressure uctuations within the combustor are not measured in this study, but
they are expected to be much greater.
Swirl-stabilized combustor ow consists of several complex
phenomena which all can contribute to the system bifurcating to
an unstable operating regime. Experimental techniques such as
Particle Image Velocimetry (PIV) have been utilized to improve
the understanding of these interactions. For example, Wicksall
[3336] investigated the effects of fuel composition on the ow
eld of ames using PIV and OH Planar Laser Induced Fluorescence
(OH PLIF) techniques and observed differences in the ow eld of
different fuels. Experimental investigation of detailed turbulence
characteristics has been difcult in the past, because of the limited
temporal resolution of non-intrusive optical diagnostic capabilities. However, recent advancements in high-speed lasers and digital imaging systems make it possible to simultaneously resolve
turbulent ow structures with a wide range of length and time
scales. Computational and numerical simulations/models such as
Large Eddy Simulations (LES) and thermo-acoustic modeling are
active research areas [3739], and time resolved PIV has proven
as an effective technique to provide extensive experimental validation data leading to increased fundamental understanding of the
complex physical processes.
OConnor and Lieuwen analyzed the multidimensional disturbance eld caused by transverse acoustic excitation of a swirling
annular nozzle ow and a premixed-swirl stabilized ame using
time-resolved PIV at framing rate of 10 kHz [40]. They showed that
the ow eld near the nozzle is superposition of acoustic and vortical disturbances, and that different disturbances are observed in
different portions of the ow eld. OConnor and Lieuwen also
investigated the vortex breakdown bubble in a transversely excited
swirl ow [15]. In both of these studies [15,40], the non-reacting
and reacting ow elds were compared to delineate the effects
of combustion. The dominant coherent structures, in particular
the center recirculation zone and inner and outer shear layers,
were observed in both non-reacting and reacting ow elds. The
ame mainly affected the size and aspect ratio of the center recirculation zone, the shear layer spreading angle, and the magnitude
of the velocity perturbations and vortical disturbances. More
recently, Steinberg et al. [41] utilized stereoscopic PIV, OH PLIF,
and OH chemiluminescence to investigate the vortex structure
and their interactions with the ame region. The ow eld was
found to contain either periodically shed toroidal vortices or helical
precessing vortex cores that excited the thermo-acoustic instabilities. Caux-Brisebois et al. [42] used proper orthogonal decomposition (POD) analysis to identify helical precessing vortex cores in a
swirl-stabilized combustor experiencing thermo-acoustic instability. Terhaar et al. investigated the key parameters governing the
PVC both experimentally and analytically, and found that the excitation or suppression of the instability is related to the backow
velocity and density gradient in the shear layer [43]. Stopper
et al. quantied the effect of pressure on the ow eld using
time-resolved PIV [44]. Temme et al. used pressurespatial correlations and phase averaged PIV measurements to identify an equivalence-ratio oscillation instability mechanism [45]. These
advanced laser diagnostic techniques have proven ideal for investigating combustion dynamics and thermo-acoustic instabilities.
In this study, the velocity measurements in the ow eld are
obtained using the time-resolved PIV technique. Quantronix
Hawk-Duo 532-120-M Nd:YAG laser with wavelength of 532 nm
and 18 mJ/pulse at the 4.2 kHz repetition rate is used for the experiments. The time between the two laser pulses was set to 50 ls.
Schematic of the PIV experimental setup and laser/camera timing
diagram are shown in Fig. 2. TSI divergent sheet optic, with
f = 25 mm cylindrical lens, combined with a 500 mm spherical
lens, was used to create 1 mm thick laser sheet. Photron SA5

1066

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

Camera
Combustor

Laser
Laser
90
Synchronizer
Computer
Camera

Front View

Top View

 = 1/ 

Camera

Laser Pulse
Delay
Laser 1
Pulse

Laser 2
Pulse

Fig. 2. Schematic and timing diagram for the PIV experimental setup.

Fastcam camera with Sigma 105 mm focal length lens and


1024  1024 pixel sensor was used at framing rate of 8.4 kHz to
image eld of view of 65 mm by 70 mm. A 532 nm bandpass optical lter is used to lter the undesirable light. The spatial resolution for the experiment is 98 lm per pixel, which corresponds to
minimum velocity resolution of 1.96 m/s.
Velocity eld calculations were performed using Insight 4G
Data Acquisition, Analysis, and Display Software from TSI. The
velocity was computed using a two pass Recursive Nyquist Grid.
The initial interrogation window size of 64  64 pixels with 50%
overlap grid spacing was used. The computed velocity vectors then
proceed through local median vector validation with a reference
vector used as the median value of all vectors in the neighborhood.
The results from the previous pass are used to optimize the spot
offset for the next pass and the interrogation window is reduced
by one half the size of the previous pass. FFT is used to compute
the correlation function, and the location of the correlation peak
is determined by tting a Gaussian curve to the highest pixel and
its four nearest neighbors. The measured data were rejected based
on three criterions: passing the median test as mentioned above,
peak to noise ratio of 1.75, and six-sigma validation. On average
about 13% of the vectors are rejected mainly because of the
unwanted reections in the quartz cylinder, which were subtracted out during preprocessing. A recursive lling using the local
mean was used to ll the holes in the vector eld. The lling procedure sorts the holes by the number of valid neighbors found initially. It rst lls the holes with the most valid neighbors since they
have the best chance to be lled; it then lls the holes with the second most valid neighbors, in which the holes lled in the previous
pass are also treated as valid neighbors. The processing technique
provided 1.568 mm spacing between vectors. Based on the minimum velocity resolution, the spacing between velocity vectors
and the viscosity of air at 1500 C, the minimum turbulent Reynolds

number is 10.5, which signies scale resolution of about an order


of magnitude greater than the Kolmogorov scale.
3. Results and discussion
3.1. Flame stability
This section discusses the static ame stability and acoustic
behavior of the system. Figure 3(a) and (b) show photograph and
schematic of LPM swirl-stabilized combustion. The ame stabilizes
immediately downstream of the dump plane, see Fig. 3(a). The corner recirculation zone is formed between the combustor wall and
the outer shear layer, and it recirculates hot gas products to ignite
the incoming reactants. The ame in Fig. 3(a) has a light blue1
appearance, which is characteristic of a lean low NOx emissions
ame. The swirl effect is seen visually in the photograph as the ame
inclines from the dump plane/swirler exit toward the combustor
wall. The schematic in Fig. 3(b) highlights the typical ow structures
in a swirl-stabilized LPM combustor [3]. Reactants entering the combustor from the annular swirler produce inner and outer shear layers. The vortices shed from the outer shear layer are located in the
corner recirculation zone, and the vortices shed from the inner shear
layer reside in the central recirculation zone. The ow structures
formed from inner and outer shear layers have been observed experimentally using time-resolved PIV [15] and computationally using
LES [46].
A photograph and schematic of LPM combustion with PIM are
shown in Fig. 3(c) and (d) respectively. The photograph in
Fig. 3(c) depicts amelets stabilizing on the downstream surface
of the PIM. The bulk of the reactant ow passes through the
1
For interpretation of color in Fig. 3, the reader is referred to the web version of
this article.

1067

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

Vortex
Breakdown
Bubble

Inner and
Outer Shear
Layers
Combustor
Liner

Corner
Recirculation
Zone

Swirler

Reactants
Reactants

(a)

(b)
Core Flame
Region
Surface
Flame

PIM

Reactants

Reactants

(c)

(d)

Fig. 3. Photographs and schematics of swirl-stabilized LPM combustion without and with PIM.

opening in the center, and the downstream portion of the ame in


the core region is observed in the photograph. The core ame
stabilizes partially upstream and partially downstream of the PIM
surface as illustrated in the schematic in Fig. 3(d). The amelets
on the surface of the PIM help stabilize the core ame. The static
ame stability is favorably affected by the amelets since any large
scale turbulent uctuations would be eliminated as the reactant
ow penetrates into the PIM. A non-porous structure of the same
geometry was also tested; however the static stability of the ame
was rather poor. With a non-porous structure, the ame would
either shift back and forth within the core region and downstream
of the solid structure indicating an imbalance of ow velocity and
turbulent ame speed, or it would stabilize downstream of the solid
structure. Thus, the ability of the reactants to ow through the porous insert and formation of amelets on the downstream surface of
the PIM are the essential features of the present concept.
Acoustic measurements are acquired as described in the
experimental setup, and the sound pressure levels (SPL) as function
of frequency are determined using Eq. (1), and the total SPL is
determined from Eq. (2).

SPLf 10  log 10

P2rms f

!
1

P 2ref

where Pref = 20 lPa

SPLtotal 10  log 10

n
X
100:1SPLi

!
2

i1

Figure 4 shows the sound pressure spectra with a distinct peak


at 531 Hz. The longitudinal natural frequency of a similar experimental setup has been determined by Meadows and Agrawal
[31] using acoustic wave equation with pressure antinode and
node boundary conditions at the inlet and outlet of the combustor
respectively. The longitudinal modes of the combustor is a function
of the combustor length (L), gas constant (R), ratio of specic heats
p
(c), and the temperature (T); f l n cRT =4L where n is the mode
number. Assuming air as the working uid and temperature range
of 10001400 K yields the natural frequency range of the rst longitudinal mode between 520 Hz and 615 Hz. Thus, the observed
peak SPL at frequency of 531 Hz is most likely an excitation of

1068

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077


0.03

Without PIM (Total SPL = 92.0 dB)


With PIM (Total SPL = 87.6 dB)

Pressure Variance

0.025

0.02

0.015

0.01

0.005

200

400

600

800

1000

Frequency (Hz)
Fig. 4. Pressure variance spectrum.

the rst longitudinal mode. With PIM, SPL peak occurs essentially
at the same frequency, but the P2rms value decreases by 74% compared to the case without PIM. Note that the SPL peak without
and with PIM occurs at the same frequency which indicates that
the PIM has a negligible effect on the natural frequency of the rst
longitudinal mode of the combustor. The total SPL without and
with PIM is 92.0 dB and 87.6 dB, which corresponds to reduction
in SPL of 4.4 dB by the PIM.
The addition of PIM in LPM combustion alters the fundamental
static ame stability mechanism and signicantly decreases the
SPL at the frequency of the thermo-acoustic instability. Without
PIM, the corner and center recirculation zones ignite the incoming
reactants, and with PIM, the amelets on the downstream annular
surface of the PIM ignite and help stabilize the reactants owing
through the core region. The corner recirculation zone is a known
source of ame-acoustic coupling and PIM effectively eliminates
this region of the ow eld.
3.2. Flow eld
According to the Rayleigh criterion [9], if the pressure uctuations are in phase with the heat release rate uctuations,
thermo-acoustic instability will occur. Thus, the region where heat
release rate uctuations occur is the dominant source of thermoacoustic instabilities, and this region is analyzed without and with
PIM in this study. Without PIM, the ow eld is measured immediately downstream of the dump plane, and the ame stabilizes
in this region. With PIM, the ow eld is measured immediately
downstream of the PIM where the ame stabilizes. The eld of
view is approximately 65 mm (transverse) by 70 mm (axial). Axial
location, z = 0 mm corresponds to the combustor dump plane, and
z = 50 mm corresponds to the exit plane of the PIM. The transverse
location, r = 40 mm corresponds to the combustor wall; however,
velocity measurements were not acquired within 5 mm of the
combustion wall because of the distortion of the PIV image caused
by the curvature of the quartz cylinder. A total of 1000 instantaneous vector elds are obtained at a frequency of 4.2 kHz. The data
are processed using an ensemble average, and all data analyses are
performed using MATLAB. The out of plane vorticity is approximated using forward differencing. The turbulent kinetic energy
(TKE) is determined by multiplying 0.5 the sum of the variance
of the velocity components. TKE does not distinguish between
the periodic and turbulent uctuations rather it is a superposition

of the two. The SPL spectra revealed a small peak at the instability
frequency, and thus is expected that the periodic uctuations have
a small impact on the magnitude of the TKE for the present test
conditions.
Figure 5(a)(c) show the instantaneous velocity vector eld
superimposed on the axial velocity contour plots without PIM.
The succession of images from top to bottom corresponds to time
step of 6.0 ms. The inner and outer diameter of the swirler is
20 mm and 40 mm respectively. Thus, no axial ow is observed
in the central region between r = 10 mm at z = 0 mm. Also, there
is no axial ow at the dump plane in the corner region beyond
the swirlers outer diameter. The reactant ow entering the combustor accelerates as it tilts toward the combustor wall. A large
recirculation zone is present in the central region, which effectively
decreases the ow area downstream of the dump plane. In
Fig. 5(a)(c), the central recirculation zone and the shear layers
appear wrinkled indicating turbulent nature of the ow eld.
Figure 5(d)(f) shows instantaneous ow eld of LPM swirlstabilized combustion with PIM. Similar to the case without PIM,
the time step between the ow elds is 6 ms. The ow eld shows
higher axial velocity in the center of the combustor indicating that
the majority of the ow passes through the central void, and only a
small portion of the ow enters the annular PIM. Combustion takes
place in the core region both upstream and downstream of the PIM
surface. Connement of the ow area by the PIM and heat release
in the core region cause signicant increase in the axial velocity.
The thermo-acoustic oscillations of the rst longitudinal mode will
generate pressure waves traveling in the positive and negative
axial direction. The pressure waves cause velocity oscillations, evident in Fig. 5 where shear layers downstream of the PIM appear
wrinkled, and the axial velocity contours reveal regions of both
high and low axial velocities.
The time-averaged axial velocity contour without PIM in Fig. 6(a)
shows the jet ow entering the combustor. The regions of high
velocity gradients on either side of the jet ow are the inner and
outer shear layers. Figure 6(a) reveals a large central recirculation
zone and a relatively small corner recirculation zone. The ow eld
appears symmetric with well-dened shear layers. Without PIM,
the maximum average axial velocity in the eld of view is approximately 6.5 m/s. With PIM, Fig. 6(b), the core ow at much higher
velocity is observed at the exit of the PIM (z = 50 mm). The maximum average axial velocity is about 11.5 m/s, which represents an
increase of over 75% compared to the case without PIM. The outer
shear layer between the core ow and PIM inside diameter is
observed at the PIM exit, while the inner shear layer is formed farther downstream. Figure 6(b) clearly demonstrates a decrease in
the shear layer spreading angle with the addition of PIM.
Vorticity represents the local spinning of the uid at a particular
location. Average vorticity contours without and with PIM are
shown in Fig. 6(c) and (d) respectively. Figure 6(c) shows that
without PIM, the vorticity is generated in the outer shear layer
adjacent to the corner recirculation zone, which is a known source
of thermo-acoustic instabilities [2]. The inner shear layer adjacent
to the central recirculation zone also generates vorticity. With PIM,
Fig. 6(d), a signicant increase in vorticity is observed in both inner
and outer shear layers. However, the vorticity generated in these
shear layers is convected downstream because of the high axial
velocity of the core ow. Figure 6(e) and (f) show the Turbulent
Kinetic Energy (TKE) contours without and with PIM respectively.
Without PIM, Fig. 6(e) shows that the TKE is highest in the central
recirculation zone and the inner shear layer. With PIM, Fig. 6(f),
TKE in the core region is much higher because of the interactions
between inner and outer shear layers downstream of the PIM. In
the corner region downstream of the PIM, the TKE is relatively
negligible indicating that the surface stabilized amelets could
be laminar.

1069

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

V m/s
7.5

60

V m/s
110

10

100

7.5

90

Y mm

2.5

Y mm

40

20

80

70

2.5

60
0
-40

-20

X mm

20

50
-40

40

-20

(a)

X mm

20

40

(d)
V m/s
7.5

60

V m/s
110

10

100

7.5

90

Y mm

2.5

Y mm

40

20

80

70

2.5

60
0
-40

-20

X mm

20

50
-40

40

-20

(b)

X mm

20

40

(e)
V m/s
7.5

60

V m/s
110

10
5

100

40

7.5

Y mm

20

Y mm

90

2.5

80

70

2.5
60
0
-40

-20

X mm

20

40

(c)

50
-40

-20

X mm

20

40

(f)

Fig. 5. Instantaneous axial velocity contours without (left) and with (right) PIM.

Figure 7 shows transverse proles of average axial velocity (left)


and TKE (right) at several axial locations, y = 1 mm, 20 mm, and
40 mm, where y represent the axial distance from the reference
plane, z = 0 without PIM and z = 50 mm with PIM. At y = 1 mm,
where the shear layers are initially formed, the average axial velocity proles for the two cases are fundamentally different. Without
PIM, Fig. 7(a), the axial velocity prole shows peaks on either side
of the centerline representing inner and outer shear layers on
either side of the swirling jet ow. The corner recirculation zone
is indicated by the negative axial velocities for |r| > 30 mm. With
PIM, Fig. 7(a) reveals a nearly top-hat prole of high axial velocity

in the core region, and an outer shear layer at the intersection of


the core ow and inside diameter of the PIM. Further downstream
at y = 20 mm, negative axial velocity is observed in the center
region without PIM, Fig. 7(b). With PIM, the axial velocity is positive at all transverse locations, which indicates that the bulk ow
in this region is owing downstream. The steep velocity gradient
in the center signies the formation of an inner shear layer in this
region. For y = 40 mm, without PIM, Fig. 7(c), the average axial
velocity is slightly negative at |r| = 10 mm signifying the end of
the center recirculation zone located upstream. TKE represents
the velocity uctuations, and the highest velocity perturbations

1070

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

V m/s

V m/s

60

110

10

10

100

7.5

7.5

90

Y mm

Y mm

40

20

80

70

2.5

2.5

60

0
-40

-20

20

40

50
-40

X mm

-20

(a)
Vorticity

-200

600
90

200
80

-200
70

-600

-600
60

-1000
20

Vorticity
1000

Y mm

Y mm

200

20

-1000
50
-40

40

X mm

-20

(c)
TKE

0.4

20

20

40

(e)

TKE
8
6

90

Y mm

Y mm

0.8

40

100

1.2

40

X mm

20

110

1.6

-20

X mm

(d)

60

0
-40

100

600
40

40

110

1000

-20

20

(b)

60

0
-40

X mm

80

70

60
50
-40

-20

X mm

20

40

(f)

Fig. 6. Average axial velocity (top), vorticity (middle), and turbulent kinetic energy (bottom) without (left) and with (right) PIM.

are observed in the shear layer for all transverse proles, except
immediately downstream of the PIM in Fig. 7(d). In general TKE
in the center region is much greater with PIM than that without
PIM.
Typical ow features in swirl-stabilized combustion systems
including the inner and outer shear layers, and the central and corner recirculation zones have been reported in several experimental
[15,40,41,4347] and computational [3,38] studies. The axial
velocity proles in this study clearly demonstrate that the PIM
eliminates the corner recirculation zone and the central recirculation zone is moved farther downstream and away from the ame.
The central recirculation zone present without PIM is replaced by a

core ow of high axial velocity with PIM. The axial velocity uctuations in the shear layers are higher with PIM. However, these
velocity perturbations are convected downstream because of the
high, and positive, axial velocity of the bulk ow. Thus, the perturbations do not interact with the incoming ow, and the possibility
for velocity perturbation feedback to the upstream is eliminated.
Furthermore, the corner recirculation zone present without PIM
is changed to an annular region of amelets with PIM. These amelets help stabilize the ame in the core region, and convect, in the
downstream direction, any disturbances generated at the outer
shear layer between the core ow and inside diameter of the PIM
to prevent upstream feedback of the ow disturbances.

1071

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

40 mm Downstream of Reference Plane

40 mm Downstream of Reference Plane


12

Axial Velocity (m/s)

10

Without PIM
With PIM

Turbulent Kinetic Energy

10

Without PIM
With PIM

0
-30

-20

-10

10

20

30

-20

-10

10

20

Radial Location (mm)

(a)

(b)

20 mm Downstream of Reference Plane

30

20 mm Downstream of Reference Plane

12

10

10

Without PIM
With PIM

Turbulent Kinetic Energy

Axial Velocity (m/s)

-30

Radial Location (mm)

Without PIM
With PIM

0
-30

-20

-10

10

20

30

-20

-10

10

20

Radial Location (mm)

(c)

(d)

1 mm Downstream of Reference Plane

30

1 mm Downstream of Reference Plane

12

10

10

Without PIM
With PIM

Turbulent Kinetic Energy

Axial Velocity (m/s)

-30

Radial Location (mm)

Without PIM
With PIM

0
-30

-20

-10

10

20

30

Radial Location (mm)

(e)

-30

-20

-10

10

20

30

Radial Location (mm)

(f)

Fig. 7. Transverse proles of average axial velocity (left) and turbulent kinetic energy (right) without and with PIM.

3.3. POD analysis


Instantaneous and average ow elds offer insight into the turbulent structures present; however a method to quantify the

energy contributions of different coherent structures to the ow


eld will provide a systematic approach to compare these
structures. Proper orthogonal decomposition (POD) is an efcient
analysis technique to capture the dominant components of an

1072

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

innite-dimension process with only nitely many, and often surprisingly few, modes [48]. POD analysis was introduced in the
context of turbulence by Lumley [49]. POD analysis decomposes
a series of velocity elds into a set of deterministic functions and
time coefcients. An understanding of the coherent structures
present in the turbulent ow eld can be inferred from the POD
modes provided the user has an intuitive idea of the structures
which are present in the ow eld. The time-resolved data can also
be used to determine the frequency of modes. POD analysis can be
performed using two methods: the direct method [50] or the snapshot method [51]. Both methods lead to the same solution; however, the direct method requires more computational time and
memory. In this study, the snapshot method based on Chen et al.
[52] is used to compute the POD modes, and for each POD mode,
fast Fourier transform (FFT) on the time coefcients is performed.
POD analysis is performed on the uctuating velocity eld
obtained by subtracting the average eld from the instantaneous
eld. The velocity eld can be reconstructed by summing the average ow eld, hui, and a linear combination of all orthogonal
n
modes (Eigen functions), /i weighted by the time coefcient, an,
as shown in Eq. (3).

^ i hui
u

M
X
n
an /i

n1

Energy Contribution

0.25

0.2

1
No PIM
PIM
No PIM (Cummulative)
PIM (Cummulative)
0.8

0.15

0.6

0.1

0.4

0.05

0.2

0
10

Modes
Fig. 8. Energy contribution of POD modes 110.

Cummulative Energy Contribution

where M is the number of instantaneous snapshots.


In this study, 1000 instantaneous snapshots are used to decompose the ow eld into spatially-dependent POD modes and
time-dependent coefcients. The modes represent the turbulent
structures in the ow eld, and each mode has an associated Eigen
value that quanties the energy contribution of the mode. The
energy contribution of each mode and the cumulative energy contribution of modes 110 are shown in Fig. 8. Without PIM, mode 1
accounts for 22% of the turbulent energy. About 61% of the total turbulent energy is present in modes 110. With PIM, mode 1 accounts
for 12.5% of the total turbulent energy, and 56.8% of the turbulent
energy is present in modes 110. A signicant reduction in cumulative contribution is observed in the rst two modes which contribute to 32.8 and 22.8% of total turbulent energy without and with
PIM respectively. Thus, PIM helps to distribute the turbulent energy
across a larger number of different turbulent structures.
Eq. (3) shows that every mode is present at each instance in
time, and is weighted by a time coefcient. Thus, the FFT on the
time coefcients for a particular mode can yield the frequency content of the mode. Figures 9 and 10 show POD modes 13 and the

associated spectra without and with PIM respectively. Without


PIM, mode 1 in Fig. 9 clearly shows a large turbulent structure in
the central recirculation zone rotating about the axis of symmetry,
which is characteristic of the precessing vortex core (PVC). The
rotation of the PVC is difcult to observe in a stationary mode,
but the frequency spectra of mode 1 reveal a well-dened frequency of 78 Hz. The PVC represented by mode 1 contributes to
22% of the turbulent energy in the ow eld as shown in Fig. 8.
The PVC is commonly observed in swirl stabilized ow and has
been veried both experimentally and computationally [3,4,41
44]. Detailed analysis of heat release rate uctuations is necessary
to quantify the impact of the PVC on the thermo-acoustic behavior
of the system, and it is beyond the scope of the present work. In
this study, pressure measurements taken outside the combustor
did not reveal any peak at 78 Hz. In Fig. 9, modes 2 and 3 represent
turbulent structures mainly in the central recirculation zone, but
unlike mode 1, the frequency of occurrence is not well dened.
Note that modes 2 and 3 only contribute to respectively 10.0 and
7.5% of total turbulent energy.
With PIM, modes 13 in Fig. 10 show that the coherent turbulent structures are conned to the inner shear layer formed downstream of the PIM, and a well-dened frequency of occurrence is
not observed. Figure 8 shows, with PIM, 12.0, 10.0, and 5.0% of
the total turbulent energy is contributed by modes 1, 2, and 3
respectively. The PIM changes the ow eld such that the PVC is
no longer present. When compared to the case without PIM, PIM
breaks up the turbulent structures such that the highest energy
modes (i.e., modes 13) contain less of the total turbulent energy.
Recall from the ow eld data in Figs. 57, PIM alters the ow
eld such that the corner recirculation zone is eliminated. Moreover, with PIM, the vortical structures formed in the shear layers
are convected out of the ow. Accordingly, the turbulent structures
represented by modes 13 in Fig. 10 will be convected out of the
ow domain; thus preventing the possible feedback mechanism
for thermo-acoustic instability. The POD analysis provides a
method to quantify the energy contribution of turbulent structures; however, it does not reveal the relationship between the
velocity eld and acoustic characteristics of the system. Timeresolved data allow for spectral analysis of the velocity eld, which
is utilized in the next section to quantitatively show the feedback
mechanisms responsible for thermo-acoustic instabilities.
3.4. Harmonic reconstruction
Time-resolved data can be used to identify the coherent uctuations at a particular frequency using the harmonic reconstruction
technique outlined by OConnor and Lieuwen [15,40]. Harmonic
reconstruction of the velocity eld is performed at the frequency
of instability. A FFT at every point in the ow eld is performed
on both components of the velocity eld including the average
^~
eld. The magnitude, A
x, and the phase, u~
x, of the FFT are used
to harmonically reconstruct the ow eld at the frequency of
thermo-acoustic instability, see Eq. (4).

h
i
^~
^ ~
u
x; t Re A
xeixtu~x

The term, xt, represents different phases in the harmonic wave.


The analysis is a quantitative tool to identify the uctuating
component of the velocity eld in the entire domain at a particular
frequency, and is used to visualize the global instability and the
convectively unstable ow eld without and with PIM respectively. By denition, a global instability or absolute unstable system occurs if a perturbation travels both upstream and
downstream to affect the entire ow, and a convectively unstable
system occurs if a perturbation introduced by a given ow is convected downstream by the mean ow [53].

1073

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

Mode 1

Mode 2
60

40

40

40

Y mm

Y mm

60

Y mm

60

20

20

0
-40

-20

20

20

0
-40

40

-20

20

0
-40

40

22

20

20

20

18

18

18

16

16

16

14

14

14

12
10

|Y(f)|

24

22

|Y(f)|

24

22

12

10

100

200

300

400

500

600

700

800

Frequency (Hz)

100

200

300

400

20

40

12

10

X mm

24

-20

X mm

X mm

|Y(f)|

Mode 3

500

600

700

800

100

200

Frequency (Hz)

300

400

500

600

700

800

Frequency (Hz)

Fig. 9. POD modes 13 (top) without PIM and the associated frequency spectrum (bottom).

Mode 1

Mode 2

80

60

80

60

-40

-20

20

X mm

40

80

60

-40

-20

20

X mm

40

-40

16

16

14

14

14

12

12

12

10

10

10

|Y(f)|

16

|Y(f)|

|Y(f)|

100

Y mm

100

Y mm

Y mm

100

Mode 3

8
6

100

200

300

400

500

Frequency (Hz)

600

700

800

100

200

300

400

500

600

700

800

20

X mm

40

-20

100

200

Frequency (Hz)
Fig. 10. POD modes 13 (top) with PIM and the associated frequency spectrum (bottom).

300

400

500

Frequency (Hz)

600

700

800

1074

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

0 Deg

180 Deg

60

60

20000

20000

40

-10000

-20000

20

0
-40

10000

Y mm

Y mm

10000

-20

X mm

20

40

-10000

-20000

20

0
-40

40

-20

60 Deg

X mm

20

40

240 Deg

60

60

20000

20000

40

-10000

-20000

20

0
-40

10000

Y mm

Y mm

10000

-20

X mm

20

40

-10000

-20000

20

0
-40

40

-20

120 Deg

X mm

20

40

300 Deg

60

60

20000

20000

40

-10000

-20000

20

0
-40

-20

X mm

20

40

10000

Y mm

Y mm

10000

40

-10000

-20000

20

0
-40

-20

X mm

20

40

Fig. 11. Harmonically reconstructed velocity and vorticity uctuations for finst = 531 Hz without PIM.

The harmonically reconstructed velocity and vorticity uctuations for finst = 531 Hz without and with PIM are shown in Figs. 11
and 12, respectively. The vorticity quantities are derived from the
velocity eld reconstructions in Eq. (4). Without PIM, Fig. 11, the
majority of the velocity vectors with large magnitude and associated vorticity are located in the corner recirculation zone and outer
shear layers. The ow eld is oscillating at 531 Hz, and the different phases correspond to different times during one oscillation
period. The excitation of the rst longitudinal mode is observed,

and forward and backward waves travel in the longitudinal (axial)


direction [54]. The plane wave incident on the dump plane boundary can excite vorticity because of the transfer of energy from the
acoustic to vortical modes [5], which can further excite vorticity in
the corner recirculation zone. The propagation of the disturbances
in both the downstream and upstream direction is indicative of an
absolutely unstable system. The coupling of the corner recirculation zone with the natural frequency associated with the rst
longitudinal mode creates the global ow instability, which is

1075

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

0 Deg

180 Deg

120

120

40000

Y mm

20000

80

100

20000

Y mm

100

40000

80

-20000

-20000

-40000

-40000

60

-40

60

-20

X mm

20

40

-40

-20

60 Deg

X mm

20

40

240 Deg

120

120

40000

Y mm

20000

80

100

20000

Y mm

100

40000

80

-20000

-20000

-40000

-40000

60

60

-40

-20

X mm

20

40

-40

-20

120 Deg

X mm

20

40

300 Deg

120

120

40000

Y mm

20000

80

-20000

100

20000

Y mm

100

40000

80

-20000

-40000

60

-40

-40000

60

-20

X mm

20

40

-40

-20

X mm

20

40

Fig. 12. Harmonically reconstructed velocity and vorticity uctuations for finst = 531 Hz with PIM.

the primary driving mechanism for thermo-acoustic instability


without PIM.
With PIM, Fig. 12, the oscillating velocity eld is conned to the
inner shear layer downstream of the PIM. The vorticity generated
in this region is convected downstream by the high velocity core
ow, which can be visualized by examining the contours in
Fig. 12 in sequential order relative to phase angle. The high region
of vorticity at z = 75 mm and r = 5 mm moves approximately
2.5 mm downstream as phase angle changes from 60 to 120
degrees. The time between the two phase angles is 0.3 ms; thus

the vortical structure is convected downstream at 8.33 m/s. The


average ow eld in the same region in Fig. 6(b) is approximately
8 m/s; which matches with the vortex convection velocity determined from Fig. 12. Thus, the perturbations introduced in the ow
eld are convected downstream by the mean ow and the system
is convectively unstable.
The harmonic reconstruction of the ow eld with PIM demonstrates coupling of thermo-acoustic instability with the vortices
shed in the inner shear layer downstream of the PIM, but these vortices are convected downstream by the core ow. PIM decreases the

1076

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077

driving force for the instability since the system changes from a global instability without PIM to a convective instability with PIM.
Lieuwen [5] discusses the difference between convectively and
globally unstable systems. A convective instability propagates only
in the ow direction while a global instability can initiate at one
location and then propagate in all directions of the ow eld. Previous work by Meadows and Agrawal has shown that at finst the PIM
would absorb approximately 30% of the acoustic energy [31]. Thus,
in terms of the Rayleigh criterion, with PIM, thermo-acoustic instability is mitigated by simultaneously increasing the dampening
force and decreasing the driving force in the system.

because PIM decreases the driving force and increases the acoustic
damping of the system. Overall, with PIM, the present combustor
design changes from a globally unstable system to a convectively
unstable system.

4. Conclusions

References

The ow eld for a LPM swirl-stabilized combustor without and


with porous inert media is analyzed. Thermo-acoustic instability is
observed at the frequency of the rst longitudinal mode of the
combustor, finst = 531 Hz, and a signicant decrease in SPL at this
frequency and in total SPL is observed with the addition of PIM
to the system. PIM has negligible effect on the frequency of the rst
longitudinal acoustic mode of the combustor. The combustor ow
eld is analyzed using time-resolved PIV, and the following conclusions explain the effects of PIM:

[1] D. Sequera, A.K. Agrawal, Numerical Simulations of Swirl-Stabilized


Combustion Coupled with Porous Inert Medium, in: Proceedings of the 6th
U.S. National Combustion Meeting, Paper 11C3, 2009.
[2] P.-H. Renard, D. Thvenin, J.C. Rolon, S. Candel, Dynamics of ame/vortex
interactions, Prog. Energy Combust. Sci. 26 (2000) 225282.
[3] Y. Huang, V. Yang, Dynamics and stability of lean-premixed swirl-stabilized
combustion, Prog. Energy Combust. Sci. 35 (2009) 293364.
[4] N. Syred, A review of oscillation mechanisms and the role of the precessing
vortex core (PVC) in swirl combustion systems, Prog. Energy Combust. Sci. 32
(2006) 93161.
[5] T.C. Lieuwen, Unsteady Combustor Physics, Cambridge University Press, New
York, NY, 2012.
[6] D. Sequera, A.K. Agrawal, Passive control of noise and instability in a swirlstabilized combustor with the use of high-strength porous insert, J. Eng. Gas
Turb. Power 134 (5) (2012). 051505/1-11.
[7] A. Putnam, Combustion Driven Oscillations in Industry, Elsevier, New York,
1971.
[8] W.C. Strahle, Combustion noise, Prog. Energy Combust. Sci. 4 (3) (1978) 157
176.
[9] J.W.S. Rayleigh, R.B. Lindsay, The Theory of Sound, Dover Publications, New
York, NY, 1945.
[10] T.C. Lieuwen, V. Yang, Combustion instabilities in gas turbine engines:
operational experience, fundamental mechanisms and modeling, Progress in
Astronautics and Aeronautics, Reston, VA, 2005.
[11] K.R. McManus, T. Poinsot, S.M. Candel, A review of active control of
combustion instabilities, Prog. Energy Combust. Sci. 19 (1) (1993) 129.
[12] T. Poinsot, D. Veynante, Theoretical and Numerical Combustion, second ed., RT
Edwards Inc., Philadelphia, PA, 2005.
[13] K.C. Schadow, E. Gutmark, K.J. Wilson, R.A. Smith, Multistep dump combustor
design to reduce combustion instabilities, J. Propul. Power 6 (1990) 407411.
[14] T. Lieuwen, Modeling premixed combustion-acoustic wave interactions: a
review, J. Propul. Power 19 (2003) 765781.
[15] J. OConnor, T. Lieuwen, Recirculation zone dynamics of a transversely excited
swirl ow and ame, Phys. Fluids 24 (7) (2012) 075107.
[16] K. Kashinath, S. Hemchandra, M.P. Juniper, Nonlinear thermoacoustics of
ducted premixed ames: The inuence of perturbation convection speed,
Combust. Flame 160 (2013) 28562865.
[17] D.E. Rogers, A mechanism for high-frequency oscillation in ramjet combustors
and afterburners, J. Jet Propul. 26 (1956) 456462.
[18] A.M. Steinberg, I. Boxx, M. Sthr, C.D. Carter, W. Meier, Flowame
interactions causing acoustically coupled heat release uctuations in a
thermo-acoustically unstable gas turbine model combustor, Combust. Flame
157 (2010) 22502266.
[19] T.X. Yi, D.A. Santavicca, Combustion instability and ame structure of
turbulent swirl-stabilized liquid-fueled combustion, J. Propul. Power 28
(2012) 10001014.
[20] C.M. Jones, J.G. Lee, D.A. Santavicca, Closed-loop active control of combustion
instabilities using subharmonic secondary fuel injection, J. Propul. Power 15
(1999) 584590.
[21] J.Y. Lee, E. Lubarsky, B.T. Zinn, Slow active control of combustion instabilities
by modication of liquid fuel spray properties, Proc. Combust. Inst. 30 (2005)
17571764.
[22] S.M. Candel, Combustion instabilities coupled by pressure waves and their
active control, Proc. Combust. Inst. 24 (1992) 12771296.
[23] A. Coker, Y. Neumeier, T. Lieuwen, B.T. Zinn, S. Menon, Studies of active
instability control effectiveness in a high pressure, liquid fueled combustor,
AIAA, Paper 20031009 (2003).
[24] J. Hermann, A. Orthmann, S. Hoffmann, P. Berenbrink, Combination of active
instability control and passive measures to prevent combustion instabilities in
a 260 mW heavy duty gas turbine, 2001, DTIC ADP011146.
[25] A.K. Agrawal, S. Vijaykant, Passive Noise Attenuation System, U.S. Patent No.
8,109,362, 2012.
[26] T.L. Marbach, A.K. Agrawal, Experimental study of surface and interior
combustion using composite porous inert media, J. Eng. Gas Turbines Power
127 (2005) 307313.
[27] V.L. Zimont, Gas premixed combustion at high turbulence. Turbulent ame
closure combustion model, Exp. Therm. Fluid Sci. 21 (1) (2000) 179186.

1. The static ame stability method without PIM is the recirculation of hot gas products to ignite the incoming reactants. With
PIM, the static ame stability mechanism is altered, and small
amelets with low velocity perturbations stabilize on the
downstream surface of the annular PIM. Flamelets ignite the
reactants passing through the center and the ame stabilizes
in the central void partially below and partially above the
downstream surface of the PIM.
2. Without PIM the vorticity is generated in the inner and outer
shear layers and corner recirculation zone, and the velocity perturbations occur mainly in the center and corner recirculation
zones. With PIM the vorticity is generated mainly in the inner
shear layer, and the shear layer spreading angle is reduced signicantly. With PIM, the TKE in the central zone of the combustor is much higher, the corner recirculation zone is eliminated,
and velocity perturbations are greatest in the inner shear layer
located downstream of the PIM.
3. Without PIM, the rst POD mode of turbulent structures has an
energy contribution of 22%. A PVC rotating at a frequency of
78 Hz is associated with mode one, and modes 2 and 3 represent turbulent structures in the central recirculation zone. With
PIM, the rst mode has much smaller energy contribution of
only 12.5%, and the PVC is eliminated. Without PIM, energycontaining turbulent structures are located mainly in the recirculation zones. With PIM, the turbulent structures are most
prominent in the inner shear layer, which is also the region of
high axial velocity.
4. The ow eld is harmonically reconstructed at finst = 531 Hz,
and without PIM the velocity perturbations in the corner recirculation zone couple with the rst longitudinal acoustic mode,
and a global ow instability is observed. With PIM, the coupling
of the velocity eld with thermo-acoustics is observed in the
inner shear layer, but the vorticity generated in this region is
convected downstream by the high velocity of the core ow.
Without PIM, the corner recirculation zone is the primary
source for driving the thermo-acoustic instability in this study.
Elimination of the corner recirculation zone by PIM effectively
removes the feedback mechanism for the instability. PIM also
absorbs a fraction (about 30%) of the acoustic energy. In terms of
the Rayleigh criterion, thermo-acoustic instability is mitigated

Acknowledgments
Joseph Meadows was supported by the Department of Education Graduate Assistance in Areas of National Needs (GAANN) Fellowship program. This research was supported in part by NASA
Grant NNX13AN14A.

J. Meadows, A.K. Agrawal / Combustion and Flame 162 (2015) 10631077


[28] L. Williams, A. Agrawal, Acoustic effects of porous inert media on lean
premixed combustion at elevated pressures, AIAA, Paper 20120207 (2012).
[29] Z. Smith, Passive Control of Combustion Noise and Thermo-acoustic Instability
with Porous Inert Media, M.S. Thesis, University of Alabama, 2011.
[30] A. Borsuk, J. Meadows, J. Williams, A.K. Agrawal, The Effect on Swirler Geometry
and Swirl Number on Passive Control of Combustion Noise and Instability, J.
Eng. Gas Turb. Power (2014), http://dx.doi.org/10.1115/1.4028613. in press.
[31] J. Meadows, A.K. Agrawal, Porous inserts for passive control of combustion
noise in liquid fuel combustion, Combust. Sci. Technol. (2014) (submitted for
publication).
[32] J. Meadows, A.K. Agrawal, Time-resolved PIV measurements of non-reacting
ow eld in a swirl-stabilized combustor without and with porous inserts for
acoustic control, ASME Paper GT2014-27203, J. Eng. Gas Turb. Power (2014),
http://dx.doi.org/10.1115/1.4028381. in press.
[33] R.W. Schefer, D.M. Wicksall, A.K. Agrawal, Combustion of hydrogen-enriched
methane in a lean premixed swirl-stabilized burner, Proc. Combust. Inst. 29
(2002) 843851.
[34] D.M. Wicksall, A.K. Agrawal, Acoustics measurements in a lean premixed
combustor operated on hydrogen/hydrocarbon fuel mixtures, Int. J. Hydrogen
Energy 32 (2007) 11031112.
[35] D.M. Wicksall, A.K. Agrawal, R.W. Schefer, J.O. Keller, The interaction of ame
and ow eld in a lean premixed swirl-stabilized combustor operated on H2/
CH4/air, Proc. Combust. Inst. 30 (2005) 28752883.
[36] D.M. Wicksall, A.K. Agrawal, R.W. Schefer, J.O. Keller, Inuence of hydrogen
addition on ow structure in enclosed swirling methane ame, J. Propul.
Power 21 (2005) 1624.
[37] N.A. Worth, J.R. Dawson, Modal dynamics of self-excited azimuthal
instabilities in an annular combustion chamber, Combust. Flame 160 (2013)
24762489.
[38] S. Hermeth, G. Staffelbach, L.Y.M. Gicquel, V. Anisimov, C. Cirigliano, T. Poinsot,
Bistable swirled ames and inuence on ame transfer functions, Combust.
Flame 161 (2014) 184196.
[39] M. Bauerheim, J. Parmentier, P. Salas, F. Nicoud, T. Poinsot, An analytical model
for azimuthal thermoacoustic modes in an annular chamber fed by an annular
plenum, Combust. Flame 161 (2014) 13741389.
[40] J. OConnor, T. Lieuwen, Disturbance eld characteristics of a transversely
excited burner, Combust. Sci. Technol. 183 (2011) 427443.

1077

[41] A.M. Steinberg, C.M. Arndt, W. Meier, Parametric study of vortex structures
and their dynamics in swirl-stabilized combustion, Proc. Combust. Inst. 34
(2013) 31173125.
[42] V. Caux-Brisebois, A.M. Steinberg, C.M. Arndt, W. Meier, Thermoacoustic
coupling in swirl-stabilized ames with helical vortices, AIAA Paper 2013
3650 (2013).
[43] S. Terhaar, K. Oberleithner, C.O. Paschereiti, Key parameters governing the
precessing vortex core in reacting ows: an experimental and analytical study,
in: Proceedings of the Combustion Institute, 2014 (in press).
[44] U. Stopper, W. Meier, R. Sadanandan, M. Sthr, M. Aigner, G. Bulat,
Experimental study of industrial gas turbine ames including quantication
of pressure inuence on ow eld, fuel/air premixing and ame shape,
Combust. Flame 160 (2013) 21032118.
[45] J.E. Temme, P.M. Allison, J.F. Driscoll, Combustion instability of a lean
premixed prevaporized gas turbine combustor studied using phase-averaged
PIV, Combust. Flame 161 (2014) 958970.
[46] Y. Huang, S. Wang, V. Yang, Systematic analysis of lean-premixed swirlstabilized combustion, AIAA J. 44 (2006) 724740.
[47] R. Sadanandan, M. Sthr, W. Meier, Simultaneous OH-PLIF and PIV
measurements in a gas turbine model combustor, Appl. Phys. B 90 (2008)
609618.
[48] P. Holmes, J.L. Lumley, G. Berkooz, Turbulence coherent structures dynamical
systems and symmetry/Philip Holmes, in: John L. Lumley, Gal. Berkooz (Eds.),
Cambridge University Press, Cambridge, New York, 1996.
[49] J.L. Lumley, The structure of inhomogeneous turbulent ows, atmospheric
turbulence and radio wave propagation, 1967, pp. 166178.
[50] G. Berkooz, P. Holmes, J.L. Lumley, The proper orthogonal decomposition in the
analysis of turbulent ows, Annu. Rev. Fluid Mech. 25 (1993) 539.
[51] L. Sirovich, Turbulence and the dynamics of coherent structures. I Coherent
structures. II Symmetries and transformations. III Dynamics and scaling,
Quant. Appl. Math. 45 (1987) 561571.
[52] H. Chen, D.L. Reuss, V. Sick, On the use and interpretation of proper orthogonal
decomposition of in-cylinder engine ows, Meas. Sci. Technol. 23 (2012)
085302.
[53] J.E. Portillo, Nonparallel Analysis and Measurements of Instability Waves in a
High-Speed Liquid Jet, PhD Dissertation, Purdue University, 2008.
[54] L.E. Kinsler, A.R. Frey, Fundamentals of Acoustics, Wiley, New York, 1962.

You might also like