Simulation of The Ignition Process in An Annular Multiple-Injector Combustor and Comparison With Experiments

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/270284309

Simulation of the Ignition Process in an


Annular Multiple-Injector Combustor and
Comparison With Experiments

Article in Journal of Engineering for Gas Turbines and Power · March 2014
DOI: 10.1115/1.4028265

CITATIONS READS

4 136

7 authors, including:

Maxime Philip Matthieu Boileau


CentraleSupélec University of Strasbourg
3 PUBLICATIONS 21 CITATIONS 31 PUBLICATIONS 524 CITATIONS

SEE PROFILE SEE PROFILE

Ronan Vicquelin Thomas Schmitt


CentraleSupélec CNRS
31 PUBLICATIONS 248 CITATIONS 41 PUBLICATIONS 404 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Implicit relaxation methods for fluid models View project

Control of Combustion Instabilities View project

All content following this page was uploaded by Maxime Philip on 13 January 2015.

The user has requested enhancement of the downloaded file.


Maxime Philip
CNRS, UPR 288,
Laboratoire d’Energetique Moleculaire
et Macroscopique Combustion,
Ecole Centrale Paris,
Grande Voie des Vignes,
Ch^atenay-Malabry 92295, France
e-mail: maxime.philip@ecp.fr

Matthieu Boileau
CNRS, UPR 288,
Laboratoire d’Energetique Moleculaire
et Macroscopique Combustion,
Ecole Centrale Paris,
Grande Voie des Vignes,
Ch^atenay-Malabry 92295, France
e-mail: matthieu.boileau@ecp.fr Simulation of the Ignition
Ronan Vicquelin
CNRS, UPR 288,
Process in an Annular
Laboratoire d’Energetique Moleculaire
et Macroscopique Combustion,
Ecole Centrale Paris,
Multiple-Injector Combustor
Grande Voie des Vignes,
Ch^atenay-Malabry 92295, France
and Comparison With
e-mail: ronan.vicquelin@ecp.fr

Thomas Schmitt1
Experiments
CNRS, UPR 288,
Ignition is a problem of fundamental interest with critical practical implications. While
Laboratoire d’Energetique Moleculaire
there are many studies of ignition of single injector configurations, the transient ignition
et Macroscopique Combustion,
of a full annular combustor has not been extensively investigated, mainly because of the
Ecole Centrale Paris,
added geometrical complexity. The present investigation combines simulations and
Grande Voie des Vignes,
experiments on a complete annular combustor. The setup, developed at EMC2 (Ener-
Ch^atenay-Malabry 92295, France
getique Moleculaire et Macroscopique Combustion) Laboratory (Mesa, AZ), features six-
e-mail: thomas.schmitt@ecp.fr
teen swirl injectors and quartz walls allowing direct visualization of the flame. High
speed imaging is used to record the space time flame structure and study the dynamics of
Daniel Durox the light-round process. On the numerical side, massively parallel computations are car-
CNRS, UPR 288, ried out in the large eddy simulation (LES) framework using the filtered tabulated (F-
Laboratoire d’Energetique Moleculaire TACLES) flamelet model. Comparisons are carried out at different instants during the
et Macroscopique Combustion, light-round process between experimental data and results of calculations. It is found
Ecole Centrale Paris, that the simulation results are in remarkable agreement with experiments provided that
Grande Voie des Vignes, the thermal effects at the walls are considered. Further analysis indicate that the flame
Ch^atenay-Malabry 92295, France burning velocity and flame front geometry are close to those found in the experiment.
e-mail: daniel.durox@ecp.fr This investigation confirms that the LES framework used for these calculations and the
selected combustion model are adequate for such calculations but that further work is
Jean-François Bourgouin needed to show that ignition prediction can be used reliably over a range of operating
CNRS, UPR 288, parameters. [DOI: 10.1115/1.4028265]
Laboratoire d’Energetique Moleculaire
et Macroscopique Combustion,
Ecole Centrale Paris,
Grande Voie des Vignes,
Ch^atenay-Malabry 92295, France
e-mail: jean-francois.bourgouin@ecp.fr

Sebastien Candel
CNRS, UPR 288,
Laboratoire d’Energetique Moleculaire
et Macroscopique Combustion,
Ecole Centrale Paris,
Grande Voie des Vignes,
Ch^atenay-Malabry 92295, France
e-mail: sebastien.candel@ecp.fr

1
Corresponding author.
Contributed by the Combustion and Fuels Committee of ASME for publication in
the JOURNAL OF ENGINEERING FOR GAS TURBINES AND POWER. Manuscript received July
11, 2014; final manuscript received July 15, 2014; published online September 30,
2014. Editor: David Wisler.

Journal of Engineering for Gas Turbines and Power MARCH 2015, Vol. 137 / 031501-1
C 2015 by ASME
Copyright V

Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms


1 Introduction representative of that found in an actual aeronautical engine com-
bustor. Experimental data on such configurations are rare but use-
Ignition is a critical process in the continuous combustion sys- ful for code validation. The study is carried out under pure
tems used in gas turbines. This needs to be completed safely and premixed conditions to allow further simplification. This defines a
reliably leading to a stabilized flame while producing a minimum relatively simple configuration but already features the complex-
pressure peak. In general, one wishes to avoid accidental extinc- ities of a turbulent reacting flow with swirl thus representing the
tion during ignition and make sure that all injectors are fully initi- first step in the simulation of ignition in a full scale engine. To our
ated by the light-round process. In aeronautical applications, the knowledge, this is a unique transparent wall annular combustor.
capacity to re-ignite at high altitude is considered in the engine Another annular system has been developed independently by
certification. In aero-engines, ignition is usually achieved with a Worth and Dawson [22] to study self-excited azimuthal combus-
couple of spark plugs or torches, located at two opposite azi- tion instabilities but it provides a more limited optical access.
muthal positions in the chamber, leading to a process which can Ignition has already been systematically investigated on MICCA
be decomposed into three phases: [23] and shown to be repeatable and to provide useful informa-
tion. Some ignition results are also reported in Ref. [24]. In con-
(1) Energy deposition: a discharge in the form of a spark pro-
trast with the present experiments, it is found in that reference that
duces an initial flame kernel.
the ignition delay increases as the flow rate is augmented. MICCA
(2) Kernel expansion: this kernel expands and creates a flame
features sixteen swirling injectors fed by a plenum, and a full
that stabilizes in the neighborhood of the injector located in
chamber formed by transparent quartz walls allowing direct opti-
the vicinity of the igniter.
cal access. The flame dynamics can then be characterized using a
(3) Flame propagation—light-round: The flame front propa-
high speed intensified camera enabling to perform comparisons
gates within the whole annular space and progressively
between calculated and experimental flames. In order to follow
ignites all injectors arranged in the chamber backplane.
the flame propagation over a longer period of time, a single elec-
When these three steps are completed, the engine can be accel- tric spark igniter out of the two available systems is used in this
erated to reach nominal operating conditions. study. The challenge is to accurately predict the interaction
It is well known that the physical mechanisms involved in phase between the reacting front and the local turbulent fluctuations and
(1) are quite complex because they involve an energy deposition dynamics in order to recover phases (2) and (3) of the light-round
through a spark and a subtle competition between convective, diffu- process observed in the experiment. After a brief presentation of
sive and reactive processes. The complexity is already present in the experimental configuration (Sec. 2), numerical aspects are
laminar flow ignition, where the influence of spark energy [1–3], described in Sec. 3. Simulation results are presented in Sec. 4 and
mean flow at the spark location [4], detailed kinetics [5], or pressure compared with experimental data.
effects [6] has been the subject of many studies. Ignition models
derived for laminar situations have been extended to deal with tur-
bulent flow conditions, like those prevailing in internal combustion 2 Experimental Configuration
(IC) engines [7–9], but the resulting models only partially capture
the interactions between chemistry and turbulent fluctuations. 2.1 Main Features. The experimental setup MICCA (Fig. 1)
Issues associated with phases (2) and (3) are perhaps easier to comprises an annular chamber which is dimensionally similar to
examine using the novel high performance computing resources that of a helicopter combustor [23]. A mixture of propane/air is
and can be investigated within the recently developed large-eddy injected through eight tubes into a plenum. The mixture is deliv-
simulation (LES) framework for reacting turbulent flows [10]. LES ered by sixteen swirl injectors forming a periodic pattern in the
have proved to provide reliable and accurate results in many config- chamber backplane and defining sixteen equal angular sectors
urations [11–19], including ignition or flashback cases where large (Dh ¼ 22.5 deg) delineated in Fig. 2. Two concentric cylindrical
scale flame motion must be tracked [20]. tubes made of quartz form the chamber side walls. Burnt gases are
The ignition of a single sector of a gas turbine is reported in exhausted to the atmosphere and evacuated through a heat
Ref. [19] while the special case of ignition of an annular combus- exchanger by a hood.
tor was recently simulated at Cerfacs [21] showing that the calcu- The inner and outer quartz tubes diameters are, respectively,
lation of a full transient ignition is feasible in a geometry which 300 mm and 400 mm (Fig. 1) and their height is 200 mm. Injectors
was close to that of a helicopter gas turbine combustor, taking into
account the injection of fuel in liquid form and the subsequent
chemical reactions in the chamber. The calculation was validated
qualitatively, but it relied on a set of assumptions and simplifica-
tions, and experimental data were not available for direct compari-
sons between the space time flame structures determined
numerically and the real flame structure.
The objective of the present investigation is to carry out the
same type of LES using up-to-date improved combustion models,
in a well controlled configuration fed by premixed gases. The
experimental system used for detailed comparisons with the simu-
lations comprises an annular chamber designated as MICCA
(standing for “multiple-injector combustor for combustion dynam-
ics analysis”) and operated at EM2C laboratory. The geometry,
which is dimensionally similar to existing industrial combustors,
includes multiple swirling injectors and fully turbulent combus-
tion. In this respect, it is idealized but representative of an actual
engine. It allows easy optical visualization through fully transpar-
ent walls. The burner is thus operating at ambient pressure and the
thermal power is limited. However, the thermal power corre-
sponding to one of the cases is only a few times lower than the
power of a typical helicopter combustor of similar size at starting Fig. 1 Direct view of the MICCA combustion chamber. The
conditions [21]. As the flow is fully turbulent and injection is swirler geometry appears as an inset on the right side of this
swirled, the flame propagation mechanism is expected to be photograph.

031501-2 / Vol. 137, MARCH 2015 Transactions of the ASME

Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms


Table 1 Physical parameters corresponding to three experi-
mental operating points. Bulk velocity Ubulk is based on the
inner tube of the swirlers.

Physical parameter #1 #2 #3

Mass flow rate m_ air (g  s1) 16.7 23.4 30.1


m_ C3 H8 (g  s1) 0.81 1.13 1.45
Bulk velocity Ubulk (m  s1) 12.2 17.1 22.0
Power P (kW) 37 52 67
Injection temperature Tinj (K) 298
Ambient pressure pres (Pa) 101,325
Equivalence ratio / 0.74
Swirl number S 0.82

preheated conditions, and will be presented in Sec. 4.2. The differ-


ent experimental operating conditions are gathered in Table 1.

3 Numerical Setup
Fig. 2 Schematic top view of the MICCA combustor providing This section provides indications on the flow solver, combus-
the position of the swirlers, pressure taps, and spark plug tion model, computational domain, boundary conditions, meshing,
and numerical procedure employed.
have an inner tube diameter of 10 mm, each featuring six 3-mm in
diameter tangential inlets (Fig. 1). The present geometry of the 3.1 LES Flow Solver and Numerical Integration Methods.
combustion chamber slightly differs from that used in Ref. [23]. Simulations rely on the AVBP flow solver developed by Cerfacs
The quartz tubes are shorter (they were 400 mm long in Ref. [23]) and IFP Energies Nouvelles, a parallel computational fluid
to reduce the computation time and the swirlers have a simpler dynamics (CFD) code that integrates the three-dimensional com-
design. However, the measured swirl numbers are very close in pressible Navier–Stokes equations on unstructured and hybrid
both cases: 0.82 in the present study and 0.7 in Ref. [23]. The meshes [25]. To fulfill the requirement of low-dissipation schemes
Reynolds numbers at the exit of the swirler and at the tangential for LES [26,27], AVBP is based on a centered scheme and uses a
inlets are, respectively, Re ¼ 1.14  104 and Re ¼ 6.3  103. Taylor–Galerkin weighted residual central distribution scheme for
When operating at full blast, MICCA reaches a maximum power time integration, third-order in time and space [28]. Subgrid scale
of 100 kW. viscosity is provided by the wall adapting local eddy model [29].

2.2 Instrumentation and Data Acquisition. To allow direct 3.2 Physical Model for Combustion. The combustion model
comparisons between experimental data and calculations, a high is an essential element in LES calculations. This model has to
intensified complementary metal-oxide semiconductor (CMOS) suitably represent the chemical kinetics and its interaction with
camera is used. Its resolution is 512  512 px2. The frame rate and the subgrid scale turbulence. It is here based on the F-TACLES
shutter speed are, respectively, set to 6000 Hz and 166 ls, thus en- model [11,30], recently developed at EM2C. This relies on tabu-
abling to properly resolve the flame front evolution during the lated chemistry in terms of filtered flamelets, in order to retrieve
process of ignition. The camera is sensitive to radiation in the visi- the right laminar burning velocity degenerescence when the flame
ble and UV range, down to a wavelength k ’ 200 nm. An intensi- wrinkling is fully resolved on the LES grid, and all intermediate
fier, whose gain remains constant during the light-round process, species at a low computational cost. This is combined with a wrin-
is used to magnify the signal. The camera faces the spark plug in kling function which describes effects of subgrid scale turbulence
a diametral plane passing through this device at a distance of [31]. The nonuniformity of the grid, and hence of the LES filter
2.9 m away from the chamber axis, and at an elevation of 1.15 m size, is here taken into account in the F-TACLES model formula-
above the chamber backplane. MICCA also features five pressure tion: the entries of the thermochemical database are then ð~ c; DÞ,
taps placed every two injectors (Fig. 2), thus covering half of the where c~ is the mass-weighted filtered progress variable defined as
eq eq
full chamber. Each pressure tap is connected to a microphone c ¼ YCO2 =YCO 2
where YCO 2
is the equilibrium mass fraction of
flush mounted in a wave guide (see Ref. [23]). The acquisition fre- CO2, and D is the flamelet filter size. The filtered flame resolution
quency is 32,768 Hz. The pressure signals recorded by the micro- is dynamically controlled setting D ¼ nxDx where Dx is a typical
phones can be compared to numerical probes placed at the exact cell size calculated as the cubic root of the local cell volume and
same location in the calculation (see Sec. 4.2). nx ¼ 5 is the minimum resolution requirement to properly capture
the flame propagation on the grid [30]. The model was imple-
mented in the compressible solver AVBP using the tabulated ther-
2.3 Experimental Procedure. It is important in this analysis mochemistry for compressible flows approach [32].
to match experimental and numerical conditions. One of the issues
in this respect concerns the boundary treatment at the combustor 3.3 Computational Domain and Boundary Conditions.
walls. It is convenient in the present simulations to consider that The computational domain (Fig. 3) exactly corresponds to the
the walls are adiabatic (see Sec. 3.3). In order to match this condi- MICCA setup, it includes the plenum, swirlers, and chamber
tion as closely as possible (i.e., to minimize heat losses at the walls. Navier–Stokes characteristic conditions are used at the
walls), the MICCA combustor is operated during about 10 min at boundaries. The eight channels delivering the propane/air mixture
the operating point, so that the tubes may reach the thermal steady impose momentum, mass fractions and temperature at their inlet.
state. Then, the injection is stopped in a sudden manner and A large volume is added downstream the chamber to represent the
re-ignited after a short delay corresponding to the time required to free exhaust in the ambient atmosphere. In order to avoid negative
evacuate hot gases from the chamber and replacing them with a velocities at the outlet of the domain, a slow coflow (set to
cool fresh stream. To quantify the importance of this thermal 1 m  s  1) is imposed in this volume. The pressure at the outlet is
effect, experimental tests have been carried out in cold and set to 101,325 Pa. The inlet and outlet boundaries require

Journal of Engineering for Gas Turbines and Power MARCH 2015, Vol. 137 / 031501-3

Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms


velocity profiles determined experimentally by laser Doppler
velocimetry. Finally, the number of tetrahedra in the domain is
310 million (corresponding to 55 million nodes). A summary of
mesh and computational features is given in Table 2. Only case #2
(see Table 1) is discussed in what follows, but case #3 has also
been simulated and the results obtained confirm the present
analysis.

3.4 Simulation Procedure. The simulation is started on a


double-sector geometry, featuring two injectors, by injecting the
Fig. 3 Axial slice of the computational domain. The black mixture of propane and air. This is not ignited, but serves to estab-
arrow symbolizes the air coflow. lish the turbulent flow delivered by the swirlers. After conver-
gence the result is replicated on the whole domain and the full-
scale simulation is continued a little longer to cancel any flow
relaxation coefficients. At the inlet, this coefficient has been set to correlation between the sectors. A 3-mm in diameter spherical ker-
1000 s1 in order to impose the mass flowrate, hence generating a nel of hot gases is then placed at the exact location of the spark.
partially reflective boundary condition at the supply channels This kernel is well resolved on the LES mesh and ensures a suc-
inlets. The pressure relaxation coefficient at the nonreflective out- cessful ignition in phase (1). This initial hot region is rapidly carved
let of the large volume representing the ambient atmosphere has up by the flow at the outlet of the swirlers producing an initial flame
been set to 100 s1. This value will have little influence on the kernel igniting the first injector. The flame is then conveyed to the
process being simulated because it is set sufficiently far from the neighboring swirlers and the light-round is in progress.
combustor exhaust. No slip and adiabatic walls have been chosen In order to save central processing unit (CPU) time associated
for the combustor lateral walls while the boundaries of the large with the large amount of small cells needed to properly compute
atmospheric volume are treated with slip conditions and are also the flame in the large volume used to represent the ambient atmos-
assumed to be adiabatic. Because the flow at the plenum inlet and phere, the reaction rate is set to zero at the chamber outlet.
in the swirler guide vanes is naturally turbulent, it is not necessary
to use stochastically excited inlet or outlet boundaries to get the
proper turbulence level in the region of interest, i.e., at the swirler 4 Results and Discussion
exit. Label definitions used to discuss results are given in Fig. 2. The
Test calculations have been performed on several grids (i.e., chamber is divided in two main parts, namely Hþ and H where
showing different levels of refinement, namely Dx ¼ 1 mm, the flame front, respectively, propagates clockwise or counter-
Dx ¼ 0.5 mm and Dx ¼ 0.375 mm at the outlet of the swirler). clockwise. Each of these parts features seven sectors (numbered
Mesh convergence was found for Dx ¼ 0.5 mm in a double-sector from S1 to S7 for Hþ and from S-1 to S-7 for H), and share sec-
geometry. The entire mesh has been constructed on this basis by tors S0 and S8.
replicating seven times the double-sector grid (Fig. 4). The In order to evaluate the simulation fidelity, the computational
regions of high velocity gradients, turbulence, and/or combustion results are first compared to the experimental images captured
are adequately refined (with the smallest cells size reaching
0.15 mm in the pipes of the swirlers, and 0.5 mm in the chamber Table 2 Mesh characteristics and computational features
backplane). To capture the light-round in the whole annulus, the
cell size growth factor has been limited to 2, so that the largest Number of cells 310  106
cell at the exit of the chamber is 1 mm. Preliminary nonreactive Number of points 55  106
and reactive tests carried out on a single injector grid have shown Time step (s) 4.7.10  8
that mesh convergence is achieved when these conditions are ful- CPU time (h) 1.5  106
filled. Moreover, in order to check the ability of the solver and of
the mesh resolution to properly retrieve the cold flow, results of
numerical simulations were compared with axial and azimuthal

Fig. 5 Time evolution of the numerical integrated heat release


(solid line) and the experimental integrated light intensity (sym-
bols) normalized by their respective maximum. The light gray
area corresponds to reacting material outside the chamber, pro-
ducing additional light intensity in the experiment. This is illus-
Fig. 4 A double-sector domain and matching cylindrical mesh trated in the subfigure appearing as an inset. (a) t 5 10 0 ms, (b)
slice. Dx corresponds to the size of the cell. t 5 20 ms, and (c) t 5 30 ms.

031501-4 / Vol. 137, MARCH 2015 Transactions of the ASME

Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms


from the CMOS camera. The mutual initial time for both experi- 4.1 Flame Geometry. It appears clearly from the direct com-
ment and computation was fitted based on a match of their respec- parison carried out in Figs. 6 and 7 that the simulation suitably
tive integrated heat release (see Fig. 5). A comparison of various retrieves many of the features observed experimentally. One
instants of the light-round process in case #2 is carried out in Figs. observes a great similarity between the flame shapes at the largest
6 and 7. The experimental images show light intensity emitted by scales found in the experimental and numerical data. The initial
the flame during the process of ignition, and represented in false propagation in the form of an arch observed during the first
colors on a yellow to red scale to improve visualization. Regard- instants of ignition is well obtained. This is followed by two tur-
ing the numerical snapshots, an isosurface of the progress variable bulent fronts progressing in clockwise and counterclockwise
highlights the flame front, colored by axial velocity levels. Note directions. The foot of each of these fronts is seen to be ahead of
that the value of this isosurface has a limited influence on the the top during the whole process in both experiment and calcula-
flame position and geometry as the flame thickness is relatively tion. The F-TACLES model retrieves a relatively high level of
compact compared to the overall flame brush. In the present case, flame surface wrinkling at the smallest resolved scales.
the axial velocity color map on the flame front is more contrasted The paths followed by the two flame fronts (respectively, in
when one considers the isosurface corresponding to c ¼ 0.9. This Hþ and H, see Fig. 2) differ in geometrical configuration, a dif-
choice also serves to highlight the interaction between the flow ference which is also found in the experiment. The Hþ flame front
field and the flame. The flows originating from the swirlers are moves preferentially along the external wall, while the H flame
exhibited by an isosurface of the velocity field. It is worth remem- front travels along the internal wall. This can be explained by the
bering that, in the calculations, the flame is artificially extin- dynamics generated by the swirler, which tends to accelerate the
guished outside the chamber by setting the reaction rate to zero in flow along the external wall for Hþ and along the internal wall
the large volume representing the ambient atmosphere.

Fig. 6 Three instants in the ignition sequence, respectively corresponding to t 5 10 (a), 20


(b), and 30 ms (c) for the operating point #2. Left: experimental data in the form of light inten-
sity emitted by the flame during the light-round process (plotted in false colors, yellow and
black corresponding, respectively, to the highest and lowest value of light intensity). Right:
computation results for the same physical time. The flame front is outlined by an isosurface of
the progress variable c 5 0.9, and colored by the axial velocity (light yellow: 230 m  s21; black:
115 m  s21). Blue isosurfaces correspond to the velocity field U 5 25 m  s21. Dashed lines rep-
resent the edges of the quartz walls. (a) t 5 10 ms, (b) t 5 20 ms, and (c) t 5 30 ms.

Journal of Engineering for Gas Turbines and Power MARCH 2015, Vol. 137 / 031501-5

Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms


Fig. 7 Three instants of an ignition sequence, respectively 40 (a), 50 (b), and 65 ms (c). Same
caption as in Fig. 6.

for H. Correspondingly one finds that the flame is slowed down constant rate between 20 and 50 ms when the two fronts
along the internal tube for Hþ and along the external wall for merge (around 50 ms, see image (b) in Fig. 7).
H. This behavior was already mentioned by Bourgouin et al. (iv) The remaining fresh gases are burnt, decreasing the heat
[23] in a previous set of ignition experiments carried out on this release rate, while hot gases are evacuated.
particular combustion chamber, with slightly different geometrical (v) Finally, the steady state is reached (image (c) in Fig. 7) at
conditions. A similar behavior was also found by Cordier et al. the nominal power corresponding to the selected injection
[33] in experiments on a linear five-injector configuration. conditions (52 kW in the present case).
Figure 5 compares the time evolution of the normalized values
of the numerical integrated heat release and the experimental inte- Although the numerical and experimental signals correspond to
grated light intensity. Both signals feature the five phases of the different quantities, flame light intensity is often considered as a
light-round process: fair indicator of local heat release. In that respect, normalized pro-
files in Fig. 5 show a close agreement during phases (i), (ii), and
(i) During the first instants, the energy provided for ignition (iii). The light gray area corresponds to the plume formed by the
generates a small flame kernel, which is rapidly distorted flame outside the chamber (see inset in Fig. 5) which is not
by the flow coming out of the swirl injector. Since the accounting for in the simulation. The difference between the nu-
chamber is initially filled with fresh gases, this kernel pro- merical and the experimental steady state levels can be explained
duces a sudden initial expansion, and as a result the heat by the light radiated by the flames located in the back, which have
release increases sharply. a reduced contribution because they are fainted by the quartz tubes
(ii) The flame brush takes the form of an arch which expands and present a smaller apparent surface due to perspective effects.
outward (image (a) in Fig. 6) within the height of the com-
bustor, and starts igniting the surrounding injectors. In that
phase, the integrated heat release grows nearly linearly 4.2 Burning Velocity and Pressure Signals. Figure 8 shows
until the arch reaches the exit of the chamber. the flame fronts merging time (i.e., the time when the two flame
(iii) The process of light-round progresses with burner-to- fronts meet) measured for the three experimental conditions men-
burner ignition by the two separate flame fronts at a fairly tioned in Table 1, as well as the merging time deduced from the

031501-6 / Vol. 137, MARCH 2015 Transactions of the ASME

Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms


Fig. 9 Volume integral of heat release rate per sector in the H1
Fig. 8 Flame fronts merging time as a function of the bulk ve- part of the chamber (see Fig. 2 for designation of the sectors)
locity Ubulk. Open circle symbols (case #1), open square and
open diamond symbols (case #2), and open triangle symbols
(case #3) represent experimental data at three various condi-
tions. The stars stand for the times obtained from the LES cal-
culations, case #2 and case #3, respectively, (case #3 is not
presented in this article).

simulations in case #2 and case #3 (not presented in this article).


These times are estimated from the images recorded by the
CMOS camera, when the two flame fronts are merging. “COLD”
means that the experiment was carried out without any preheating
of the combustor and the ignition takes place in the presence of
cold boundaries; “PREHEATED” indicates that the combustion
chamber has been operated for about 10 min prior to the ignition
experiment. This type of operation defines conditions at the boun-
daries which are closer to the adiabatic conditions used in the nu-
merical simulation. To evaluate the variability of the experiments,
several tests have been carried out for each operating point, and
more specifically for case #2 that corresponds to the simulation.
In that particular condition, a dispersion of 10% is observed, but Fig. 10 Transit time of the flame fronts as a function of azi-
muthal angle. Circle and plus symbols represent, respectively,
the major part of the points is around a value of 50 ms. One can experimental and numerical times; black and red colors stand,
note the large discrepancy between the experimental tests carried respectively, for H1 and H2 (see Fig. 2 for designation of the
out in cold and preheated conditions. In cold conditions the merg- sectors).
ing time is of the order of 80 ms for the same flow condition, i.e.,
60% larger than the merging time in preheated conditions. The
LES merging time calculated by assuming that the chamber walls from a complex process involving heat release rate variations, per-
are adiabatic closely matches the time taken by the flame to cross turbations induced by the turbulent flow and multiple reflections
the whole annular space obtained under “PREHEATED.” on the combustor boundaries. One can expect that experimental
To get a more precise quantification of the transit time of the and numerical signals, which are a single realization of a stochas-
flame front when it passes by the flow established by the swirl tic process, will not exactly match, but the comparison is never-
injectors it is convenient to plot the heat release rate per sector. theless instructive. In addition the pressure signals provide a
This is done in Fig. 9 for the Hþ chamber part only as similar quantitative measurement of the delay separating the first sector
results can be obtained for the other half chamber H. The igni- ignition and the merging of the flame fronts. Figure 11 shows two
tion time of each sector is defined by the instant when the sector- different pressure signals, respectively, retrieved from sensors P1
averaged heat release reaches its maximum. In the experiment, the and P4 and low pass filtered at 1 kHz (similar results were
transit times are estimated from the CMOS camera snapshots. A obtained for P2, P3, and P5, see Fig. 2 for pressure taps locations).
comparison between LES and experiment is given in Fig. 10. The The vertical dashed lines correspond to the ignition of the first
Hþ flame brush appears to be faster until it reaches an azimuthal three sectors (top subfigure) and the merging of the flame fronts
angle of 90 deg. The initially slower H flame reaches this azi- (bottom subfigure), and are obtained from the integrated heat
muthal angle in about the same time and flame brush merging release rate per sector (see Fig. 9). The numerical signal P1 fea-
takes place in the S8, opposite to the spark plug. The Hþ flame tures three initial peaks corresponding to the ignition time of sec-
brush is favored by the flow circulation generated by the swirlers tors S0, S1, and S2. These peaks do not appear as clearly in the
during the first instants of ignition, since the direction of propaga- experimental signal: only a single pressure peak corresponding to
tion is the same as that of the flow induced by the swirlers. But as the ignition of S0 is observed. In the LES, a sudden pressure drop
this flame progresses along the external tube, the path traveled is occurs at t ¼ 15 ms when the initial arch flame front reaches the
effectively longer, and the H flame brush catches up. Finally, exit of the chamber where the reaction rate is set to zero (see Sec.
one finds that in both half chambers the transit time is a quasi lin- 3.4). The merging of the flame fronts can clearly be identified on
ear function of the azimuthal angle, corresponding to an absolute both numerical and experimental signals, showing a sharp drop of
propagation velocity of 12 m  s1. This characterizes both the the pressure around 47 ms. For 15 ms < t < 47 ms, it is not possible
experiment and the simulation. to identify the ignition times of individual sectors from the pres-
It is also interesting to see if the simulation is able to reproduce sure signals because of the shifts in H and Hþ propagation times
the experimental pressure signals. The pressure fluctuations result and the multiple acoustic wave reflections on the combustor walls.

Journal of Engineering for Gas Turbines and Power MARCH 2015, Vol. 137 / 031501-7

Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms


viewpoint for plotting three-dimensional visualizations of the cal-
culation. It is found that the computations match the data recorded
in the experiment in terms of instantaneous flame geometry, flame
brush spatial location and flame merging time, provided that the
experimental chamber is preheated to match the adiabatic condi-
tion in LES. The structure of the flame brush in its largest scales is
very close to that observed experimentally indicating that the sim-
ulation retrieves experimental features with great fidelity. This
indicates that the F-TACLES combustion model in combination
with the AVBP flow solver is able to properly represent the flame
brush geometry and suitably retrieve the burning velocity and the
absolute flame velocity. The transit time of the flame front per
sector compares well with the experimental observation, and the
differences with the experimental values are quite limited. Five
phases in the light-round process have been identified, from the
initial expansion of the kernel in the first instants to the steady
state, passing by a formation of an arch followed by the propaga-
tion of two distinct flame brushes sweeping the annular space.
The importance of the aerodynamic field induced by the swirl
injectors is also investigated. Experiments and calculations indi-
cate that the path followed by the flame depends on the direction
of propagation of the flame brush (in the clockwise or counter-
clockwise direction). Further analysis will be performed on these
particular simulations to reach a better understanding of the light-
round process in such a configuration.

Acknowledgment
We acknowledge Partnership for Advanced Computing in
Fig. 11 Two pressure signals retrieved from respectively sen- Europe (PRACE) for awarding us access to resource Curie-TN
sors P1 and P4 (see Fig. 2 for designation of the sectors). Thick based in France at TGCC. The PhD fellowship provided to M.
solid line: LES; thin solid line: experiment. Philip by the Initiative d’Excellence (IDEX) Paris-Saclay is also
gratefully acknowledged. This work is also partially supported by
Snecma (Safran group). Thanks are due to Cerfacs for sharing the
The experimental signal amplitudes are generally lower than those AVBP solver. We also wish to thank the anonymous reviewers for
obtained in the simulation. There are a few possible reasons for their helpful comments.
this. On the experimental side, one may note that the waveguide
microphones act as a low pass filter with a cutoff frequency of the
order of 1 kHz. This will reduce the amplitude levels of fast pres- References
sure perturbations and thus reduce the experimental signal level. [1] Beduneau, J., Kim, B., Zimmer, L., and Ikeda, Y., 2003, “Measurements of
One may also note that the numerical damping of acoustic pertur- Minimum Ignition Energy in Premixed Laminar Methane/Air Flow by Using
Laser Induced Spark,” Combust. Flame, 132(4), pp. 653–665.
bations is lower than the real value because viscous effects associ- [2] Champion, M., Deshaies, B., Joulin, G., and Kinoshita, K., 1986, “Spherical
ated with the acoustic boundary layers at the walls are Flame Initiation—Theory Versus Experiments for Lean Propane–Air
underestimated as these layers are not suitably resolved in the Mixtures,” Combust. Flame, 65(3), pp. 319–337.
LES. A third possibility may be that the sources of acoustic per- [3] Kurdyumov, V., Blasco, J., Sanchez, A., and Linan, A., 2004, “On the Calcula-
tion of the Minimum Ignition Energy,” Combust. Flame, 136(3), pp. 394–397.
turbations which are linked to the time derivative of the heat [4] Baum, M., and Poinsot, T., 1995, “Effects of Mean Flow on Premixed Flame
release rate are not suitably predicted and are overestimated to Ignition,” Combust. Sci. Technol., 106(1–3), pp. 19–39.
some extent by the F-TACLES model. There is in fact no experi- [5] Ko, Y., Arpaci, V., and Anderson, R., 1991, “Spark-Ignition of Propane Air
ence in the application of this model to combustion noise from Mixtures Near the Minimum Ignition Energy. 2. A Model Development,” Com-
bust. Flame, 83(1–2), pp. 83–88.
confined flames (and more generally the experience in the applica- [6] Sloane, T., 1992, “Numerical-Simulation of Electric Spark-Ignition in Methane
tion of LES to the calculation of combustion noise from confined Air Mixtures at Pressures Above One Atmosphere,” Combust. Sci. Technol.,
flames is still quite limited). Thus the model may well provide a 86(1–6), pp. 121–133.
suitable representation of the heat release rate but may perhaps [7] Granet, V., Vermorel, O., Lacour, C., Enaux, B., Dugue, V., and Poinsot, T.,
2012, “Large-Eddy Simulation and Experimental Study of Cycle-to-Cycle Var-
not yield sufficiently accurate values of the temporal variation of iations of Stable and Unstable Operating Points in a Spark Ignition Engine,”
the local values of this quantity. Nevertheless, the signals show a Combust. Flame, 159(4), pp. 1562–1575.
degree of similarity. [8] Kravchik, T., and Sher, E., 1994, “Numerical Modeling of Spark-Ignition and
Flame Initiation in a Quiescent Methane–Air Mixture,” Combust. Flame,
99(3–4), pp. 635–643.
[9] Vermorel, O., Richard, S., Colin, O., Angelberger, C., Benkenida, A., and Vey-
5 Conclusions nante, D., 2009, “Towards the Understanding of Cyclic Variability in a Spark
Ignited Engine Using Multi-Cycle LES,” Combust. Flame, 156(8), pp.
The ignition sequence of a full annular combustion chamber is 1525–1541.
investigated in this article by combining experimental data and [10] Mastorakos, E., 2009, “Ignition of Turbulent Non-Premixed Flames,” Prog.
results of LES. The experimental system is dimensionally similar Energy Combust. Sci., 35(1), pp. 57–97.
to that existing in helicopter gas turbine combustors. It is equipped [11] Auzillon, P., Gicquel, O., Darabiha, N., Veynante, D., and Fiorina, B., 2012, “A
Filtered Tabulated Chemistry Model for LES of Stratified Flames,” Combust.
with 16 swirl injectors delivering a perfectly premixed stream of Flame, 159(8, SI), pp. 2704–2717.
reactants. This allows a unique comparison between experimental [12] Franzelli, B., Riber, E., Gicquel, L. Y. M., and Poinsot, T., 2012, “Large Eddy
data and results of calculation. The combustion chamber features Simulation of Combustion Instabilities in a Lean Partially Premixed Swirled
quartz side walls allowing direct optical access to the combustion Flame,” Combust. Flame, 159(2), pp. 621–637.
[13] Freitag, M., and Janicka, J., 2007, “Investigation of a Strongly Swirled Uncon-
region. The flame brush dynamics is observed with a high speed fined Premixed Flame Using LES,” Proc. Combust. Inst., 31(1), pp. 1477–1485.
imaging camera. The data recorded by the camera is directly com- [14] Menon, S., and Patel, N., 2006, “Subgrid Modeling for Simulation of Spray
pared with simulation results by making use of the same Combustion in Large-Scale Combustors,” AIAA J., 44(4), pp. 709–723.

031501-8 / Vol. 137, MARCH 2015 Transactions of the ASME

Downloaded From: http://gasturbinespower.asmedigitalcollection.asme.org/ on 09/30/2014 Terms of Use: http://asme.org/terms


[15] Moureau, V., Domingo, P., and Vervisch, L., 2011, “From Large-Eddy Simula- [25] Moureau, V., Lartigue, G., Sommerer, Y., Angelberger, C., Colin, O., and Poin-
tion to Direct Numerical Simulation of a Lean Premixed Swirl Flame: Filtered sot, T., 2005, “Numerical Methods for Unsteady Compressible Multi-
Laminar Flame-PDF Modeling,” Combust. Flame, 158(7), pp. 1340–1357. Component Reacting Flows on Fixed and Moving Grids,” J. Comput. Phys.,
[16] Pitsch, H., 2006, “Large-Eddy Simulation of Turbulent Combustion,” Ann. 202(2), pp. 710–736.
Rev. Fluid Mech., 38, pp. 453–482. [26] Gullbrand, J., and Chow, F., 2003, “The Effect of Numerical Errors and Turbu-
[17] Selle, L., Lartigue, G., Poinsot, T., Koch, R., Schildmacher, K., Krebs, W., lence Models in Large-Eddy Simulations of Channel Flow, With and Without
Prade, B., Kaufmann, P., and Veynante, D., 2004, “Compressible Large Eddy Explicit Filtering,” J. Fluid Mech., 495, pp. 323–341.
Simulation of Turbulent Combustion in Complex Geometry on Unstructured [27] Kravchenko, A., and Moin, P., 1997, “On the Effect of Numerical Errors in
Meshes,” Combust. Flame, 137(4), pp. 489–505. Large Eddy Simulations of Turbulent Flows,” J. Comput. Phys., 131(2),
[18] Gicquel, L. Y. M., Staffelbach, G., and Poinsot, T., 2012, “Large Eddy Simula- pp. 310–322.
tions of Gaseous Flames in Gas Turbine Combustion Chambers,” Prog. Energy [28] Colin, O., and Rudgyard, M., 2000, “Development of High-Order Taylor–
Combust. Sci., 38(6), pp. 782–817. Galerkin Schemes for LES,” J. Comput. Phys., 162(2), pp. 338–371.
[19] Jones, W. P., and Tyliszczak, A., 2010, “Large Eddy Simulation of Spark Ignition [29] Nicoud, F., and Ducros, F., 1999, “Subgrid-Scale Stress Modelling Based on
in a Gas Turbine Combustor,” Flow Turbul. Combust., 85(3–4), pp. 711–734. the Square of the Velocity Gradient Tensor,” Flow Turbul. Combust., 62(3), pp.
[20] Sommerer, Y., Galley, D., Poinsot, T., Ducruix, S., Lacas, F., and Veynante, D., 183–200.
2004, “Large Eddy Simulation and Experimental Study of Flashback and Blow- [30] Fiorina, B., Vicquelin, R., Auzillon, P., Darabiha, N., Gicquel, O., and
Off in a Lean Partially Premixed Swirled Burner,” J. Turbul., 5(37), pp. 1–3. Veynante, D., 2010, “A Filtered Tabulated Chemistry Model for LES of Pre-
[21] Boileau, M., Staffelbach, G., Cuenot, B., Poinsot, T., and Berat, C., 2008, “LES of an mixed Combustion,” Combust. Flame, 157(3), pp. 465–475.
Ignition Sequence in a Gas Turbine Engine,” Combust. Flame, 154(1–2), pp. 2–22. [31] Charlette, F., Meneveau, C., and Veynante, D., 2002, “A Power-Law Flame
[22] Worth, N. A., and Dawson, J. R., 2013, “Self-Excited Circumferential Instabil- Wrinkling Model for LES of Premixed Turbulent Combustion. Part II: Dynamic
ities in a Model Annular Gas Turbine Combustor: Global Flame Dynamics,” Formulation,” Combust. Flame, 131(1–2), pp. 181–197.
Proc. Combust. Inst., 34(2), pp. 3127–3134. [32] Vicquelin, R., Fiorina, B., Payet, S., Darabiha, N., and Gicquel, O., 2011,
[23] Bourgouin, J.-F., Durox, D., Schuller, T., Beaunier, J., and Candel, S., 2013, “Coupling Tabulated Chemistry With Compressible CFD Solvers,” Proc. Com-
“Ignition Dynamics of an Annular Combustor Equipped With Multiple Swirling bust. Inst., 33(1), pp. 1481–1488.
Injectors,” Combust. Flame, 160(8), pp. 1398–1413. [33] Cordier, M., Vandel, A., Renou, B., Cabot, G., Boukhalfa, M. A., Esclapez, L.,
[24] Bach, E., Kariuki, J., Dawson, J. R., and Mastorakos, E., 2013, “Spark Ignition Barre, D., Riber, E., Cuenot, B., and Gicquel, L., 2013, “Experimental and
of Single Bluff-Body Premixed Flames and Annular Combustors,” AIAA Paper Numerical Analysis of an Ignition Sequence in a Multiple-Injectors Burner,”
No. 2013-1182. ASME Paper No. GT2013-94257.

Journal of Engineering for Gas Turbines and Power MARCH 2015, Vol. 137 / 031501-9

DownloadedViewFrom:
publicationhttp://gasturbinespower.asmedigitalcollection.asme.org/
stats on 09/30/2014 Terms of Use: http://asme.org/terms

You might also like