Вы находитесь на странице: 1из 60

Effects of mesh resolution on Large Eddy Simulation

of reacting ows in complex geometry combustors


G. Boudier
a
, L.Y.M. Gicquel
a,
and T.J. Poinsot
b
a
CERFACS, 42 Avenue G. Coriolis, 31057 Toulouse cedex, France
b
Institut de M ecanique des Fluides de Toulouse, Avenue C. Soula, 31400 Toulouse, France
Abstract
The power of present parallel computers is becoming sufcient to apply Large Eddy Simu-
lation (LES) tools to reacting ows not only in academic congurations but also in real gas
turbine chambers. The most limiting factor to perform LES of real systems is the mesh size
which directly controls the overall cost of the simulation, so that the effects of mesh resolu-
tion on LES results become a key issue. In the present work, an unstructured compressible
LES solver is used to compute the reacting ow in a domain corresponding to a sector of a
realistic helicopter chamber. Three grids ranging from 1.2 to 44 million elements are used
for LES and results are compared in terms of mean and uctuating elds as well as of pres-
sure spectra. Results show that the mean temperature, species and velocity elds are almost
insensitive to the grid size. The RMS eld of the resolved velocity is also reasonably in-
dependent of the mesh while the RMS elds of temperature exhibit more sensitivity to the
grid as expected by the fact that most of the combustion process proceeds at small scales.
The acoustic eld exhibits a limited sensitivity to the mesh, suggesting that LES is adapted
to the computation of combustion instabilities in complex systems.
Key words: Mesh resolution, Industrial GT, Large Eddy Simulations

Corresponding author.
Email address: lgicquel@cerfacs.fr (L.Y.M. Gicquel).
Preprint submitted to Elsevier 2 April 2008
1 Introduction
Ongoing developments for the next generation of gas turbines focus on lean pre-
mixed operating regimes to satisfy emission regulations. The design of these com-
bustion chambers is complex because combustion concepts leading to minimum
emissions are also sensitive to combustion instabilities [13]. These instabilities
are due to a combination of the natural unstable modes of swirling ows (Precess-
ing Vortex Cores or PVCs [48]) with acoustics and unsteady heat release [912].
To study combustion instabilities but also to provide more accurate results for sta-
ble reacting ows, the best numerical technique developed today is Large Eddy
Simulation (LES) [1315]. LES has been used successfully for many academic
ames in simple geometries [1622] but still very rarely for complex realistic
combustion chambers. Multiple issues remain to be investigated before LES can
be used efciently for design of combustion chambers: high-order schemes, Sub-
Grid Scale (SGS) tensor and ux vectors, ame/turbulence interaction, chemical
schemes, boundary conditions, parallel efciency, etc.
In this framework, a fundamental and unresolved question is often neglected: the
effects of mesh resolution on LES results. Multiple authors have underlined the im-
portance of this point for LES [23,24]. Although LES results depend on mesh res-
olution (unlike Reynolds-Averaged Navier-Stokes simulation (RANS) which must
produce grid independent results as soon as the mesh is sufciently rened), they
must satisfy multiple properties: time-averaged values must converge, Root Mean
Square (RMS) resolved values must increase when the mesh cell size decreases
and the SGS turbulence level diminishes, the resolved velocity spectra must ll
towards larger wave-numbers. In practice, these behaviors are expected to be con-
trolled by the LES models, the ow Reynolds number, the grid resolution as well as
the accuracy of the numerical solver (in the context of implicit ltering [23,25,26]).
Mesh dependency analysis of non-reacting LES predictions has recently been ad-
dressed [2729] and quality criteria have been proposed for a posteriori evaluation
2
of the LES ow predictions [24,3033].
For reacting ows, the computer power needed to simulate realistic geometries is
so large that the grids used for LES are usually as large as possible while being still
too coarse to resolve all ow zones: in these circumstances, multiplying the num-
ber of grid points by a signicant factor to verify the effects of grid resolution on
the LES results was impossible until very recent times. Very few LES of reacting
ows have been devoted to mesh dependency in simple congurations [2729] and
none of them has addressed this issue in complex geometry combustors. The situ-
ation has changed in the last two years: porting LES codes on massively parallel
machines in the Top 20 list has allowed a sudden increase of power for combus-
tion computations. For example, Fig. 1 shows the speed-up obtained on such a
parallel architecture for a 40 million cell conguration corresponding to a full gas
turbine combustion chamber with the parallel solver used in the present paper [34].
Speed-ups of nearly 95 percent as obtained here on 4096 processors allow to ad-
dress the problem of mesh dependency by performing one coarse grid simulation
with a reasonable mesh (typically 1.5 million cells) and then comparing it with an
intermediate grid simulation (8 times more cells) and nally with a ne grid
simulation (32 times more cells). In the present work, the mesh dependency of the
LES predictions is studied for a helicopter combustion chamber [35]. Results on the
coarse, intermediate and ne grids for the same regime allow a direct investigation
in terms of mean ow, RMS values, unsteady activity and acoustic mode excitation.
Although this grid dependency exercise must also be performed on simple aca-
demic geometries, using a real Rich-burn, Quick-mix, Lean-burn (RQL) com-
bustor case is an interesting test case because this conguration puts constraints on
meshes which are not found in most academic chambers where simple structured
meshes can be used. Using a real helicopter chamber guaranties that issues relevant
to industrial cases will be taken into account. However a drawback of this choice
is that very limited experimental information is available. Therefore the present pa-
per must be viewed only as a partial response to the problem of LES resolutions
3
in combustion since no experimental result will be used for validation. For exten-
sive comparaisons of the present solver with experimental data, readers are referred
to previous studies [8,3638] where velocity and/or temperature elds have been
compared in various congurations.
The LES solver used for this study and the SGS models required for such a com-
parison are described in Section 2. Details on the combustion model are given in
Section 3, while the target conguration (a sector of a helicopter combustion cham-
ber) is presented in Section 4. Section 5 then discusses results obtained on the three
grids.
2 Massively parallel Large Eddy Simulations of reacting ows
LES for reactive multi-species mixtures involves the spatial ltering operation which
reduces for spatially, temporally invariant and localised lter functions [39,40], to:

f (x, t) =
1
(x, t)
+
Z

(x

, t) f (x

, t) G(x

x) dx

, (1)
where G denotes the lter function and

f (x, t) is the Favre ltered value of the


variable f (x, t) [41].
In the mathematical description of compressible turbulent ows with chemical re-
actions and species transport, the primary variables are the species volumic mass
fractions

(x, t), the velocity vector u


i
(x, t), the total energy E(x, t) e
s
+1/2 u
i
u
i
,
and the density (x, t) =
N
=1

(x, t). Note that

(x, t) is linked to the species


mass fractions Y

(x, t) and mass conservation imposes for a mixture of N species:

N
=1
Y

(x, t) = 1.
The uid follows the ideal gas law, p = r T and e
s
=
R
T
0
C
p
dT p/, where e
s
is the mixture sensible energy, T the temperature, C
p
=
N
=1
C
p,
Y

the uid heat


4
capacity at constant pressure and r is the mixture gas constant, which varies with
composition and is obtained by r =
R
W
= R
N
=1
Y

, where R = 8.314kg m
2
/(s
2
K)
and W

is the molecular weight of the species . The viscous stress tensor, the
heat diffusion vector and the species molecular transport use classical gradient ap-
proaches. The uid viscosity follows Sutherlands law, the heat diffusion coefcient
follows Fouriers law, and the species diffusion coefcients are obtained using a
constant species Schmidt number and diffusion velocity corrections for mass con-
servation [15].
The application of the ltering operation to the instantaneous set of compressible
Navier-Stokes transport equations with chemical reactions yields the LES equa-
tions, Eqs. (2)-(4), which need modelling for the system to be closed [23,42]:
u
i
t
+

x
j
( u
i
u
j
) =

x
j
[P
i j

i j

i j
t
], (2)


E
t
+

x
j
(

E u
j
) =

x
j
[u
i
(P
i j

i j
) +q
j
+q
j
t
] +
T
, (3)


Y
k
t
+

x
j
(

Y
k
u
j
) =

x
j
[J
j,k
+J
j,k
t
] +
k
. (4)
In Eq. (2), the unresolved SGS stress tensor
i j
t
= ( u
i
u
j
u
i
u
j
) is usually ad-
dressed through the concept of SGS turbulent viscosity model and the Boussinesq
assumption [26]. The model henceforth reads (Smagorinsky [43]):

i j
t

1
3

kk
t

i j
=2
t

S
i j
, (5)
with,

S
i j
=
1
2
_
u
i
x
j
+
u
j
x
i
_

1
3
u
k
x
k

i j
. (6)
In Eq. (5) and (6),

S
i j
is the resolved strain tensor and
t
is the SGS turbulent
viscosity. The Smagorinsky model [43] is used here. It expresses
t
as:

t
= (C
S
)
2
S. (7)
5
In Eq. (7), denotes the lter characteristic length and is approximated by the
cubic-root of the cell volume, C
S
is the model constant (C
S
= 0.18) and S =
(2

S
i j

S
i j
)
1/2
.
In Eqs. (3) and (4), the SGS species ux J

i
t
= (

u
i
Y

u
i

Y

), and the SGS energy


ux q
i
t
= (

u
i
E u
i

E), are respectively modelled by use of the species SGS turbu-


lent diffusivity D

t
=
t
/Sc

t
, where Sc

t
is the turbulent Schmidt number (Sc

t
=0.9
for all ). The eddy diffusivity is also used along with a turbulent Prandtl number
Pr
t
= 0.6, so that
t
=
t
C
p
/Pr
t
:
J

i
t
=
_
D

t
W

x
i

V
c
i
_
and q
i
t
=
t

T
x
i
+
N

=1
J

i
t

s
. (8)
In Eq. (8) the mixture molecular weight W and the species molecular weight W

can
be combined with the species mass fraction to yield the expression for the molar
fraction of species : X

=Y

W/W

. In expression (8), V
c
i
is the diffusion correc-
tion velocity resulting from the Hirschfelder Curtis approximation [15] and

T is
the modied ltered temperature which satises the modied ltered state equa-
tion [4446], p = r

T. Finally,

h

s
stands for the enthalpy of species . Although
the performance of the models could be improved through the use of a dynamic
formulation [25,44,4749], they are considered sufcient to address the present
investigation. Note that throughout the work, the variations of the molecular coef-
cients resulting from the unresolved uctuations are neglected so that the various
expressions for the molecular coefcients become only functions of the ltered
eld [15].
3 Combustion modelling
This section presents the models necessary to take into account chemical kinet-
ics and ame/turbulence interactions. For simplicity, a one-step chemistry model,
Eq. (9), is derived for C
10
H
16
based on a detailed model of C
10
H
16
/Air combustion
6
with 43 species and 174 steps (Turbomeca private communication).
C
10
H
16
+14 O
2
10 CO
2
+8 H
2
O (9)
The reduced one-step scheme guaranties proper ame speed predictions only in
the lean regime (i.e. with equivalence ratios, < 1). For the target conguration,
such a chemical scheme is not sufcient to predict proper ame positionning since
the local equivalence ratio reaches a wide range of values. To circumvent such a
shortcoming the pre-exponential constant of the one-step scheme is adjusted versus
local equivalence ratio to reproduce the proper ame speed dependency on the rich
side [50]. The nal expression for the rate of reaction of Eq. (9), is:

Q = A()
_
Y
C
10
H
16
W
C
10
H
16
_
n
1
_
Y
O
2
W
O
2
_
n
2
exp(
T
a
T
) (mol. m
3
. s
1
), (10)
where n
1
= 1.5, n
2
= 0.55, T
a
= 3608.4K and the A() function is:
A() =
3.8410
14
2
(1+tanh(
1.39
0.26
))
+
0.33
4
(1+tanh(
1.6
0.8
))(1+tanh(
1.85
0.8
)). (11)
Figure 2 shows that the adjusted one-step scheme matches the detailed scheme
reasonably well in terms of ame speed and adiabatic temperature for premixed
laminar ames at 8 bar, which is the target pressure for the full combustor.
The ame/turbulence interaction is modelled using the Dynamic Thickened Flame
(DTF) model. Following the theory of laminar premixed ames [51] the ame
speed S
0
L
and the ame thickness
0
L
of a premixed front may be expressed as:
S
0
L

A and
0
L


S
0
L
=
_

A
, (12)
where is the thermal diffusivity and A the pre-exponential constant. Increasing
the thermal diffusivity by a factor F, the ame speed is kept unchanged if the pre-
exponential factor is decreased by the same factor. This operation leads to a ame
7
thickness which is multiplied by F and more easily resolved on a coarser mesh.
While in reacting zones, diffusion and source terms issued from the thickened re-
action are well resolved and turbulence is solely represented by the efciency func-
tion E [52], molecular and thermal diffusion cannot be over-estimated by a factor F
in mixing zones where no combustion occurs (it would yield over-estimated mixing
and wrong ame positions). Dynamic thickening is thus introduced to account for
these points [15,36,53,54]. The baseline idea of the Dynamically Thickened Flame
(DTF) model is to detect reaction zones using a sensor S and to thicken only these
reaction zones, leaving the rest of the ow unmodied. Thickening depends on the
local grid resolution and it locally adapts the combustion process to reach a numer-
ically resolved ame front. The ame sub-grid scale wrinkling and interactions at
the SGS level are supplied by the efciency function [15,36,53,54].
4 LES solver and target conguration
The LES code used here solves the fully compressible, multi-species (variable heat
capacities) Navier-Stokes equations using a nite-volume discretization, on struc-
tured, unstructured and hybrid meshes. Second and third-order temporal and spa-
tial numerical schemes [55,56] offer reliable unsteady solutions for complex ge-
ometries as encountered in the eld of aeronautical gas turbines. In this study the
Lax-Wendroff second-order numerical scheme is chosen for the three meshes.
The conguration (Fig. 3) corresponds to a helicopter combustion chamber where
fuel is injected using an inverted cane injection system, also called pre-vaporizer.
The computational domain focuses on a 36 degree section of a full annular reverse-
ow combustion chamber designed by Turbomeca (Safran group). A premixed
gaseous mixture of C
10
H
16
enters the chamber through the pre-vaporizer, Fig. 3
(a) & (b). Fresh gases are consumed in the primary zone, delimited by the chamber
dilution holes and the liner dome of the combustion chamber, Fig. 3 (a). To ensure
full combustion, this region of the chamber is fed with air by primary jets located
8
on the inner liner, Fig. 3 (b). Burnt gases are then cooled by dilution jets or cooling
lms located on the inner and outer liners as well as on the return bend of the com-
bustion chamber. Multi-perforated plates also ensure local wall cooling in areas of
the chamber shown on Fig. 3 (b).
The combustion regime expected in such burners mixes rich partially premixed
ames in the chamber primary zone (the gases injected in the canes can be con-
sidered as premixed gases at an equivalence ratio of 3.17) and diffusion ames
in the dilution region: i.e. RQL concept. A turbulent combustion model able to
handle both regimes is therefore needed and the DTF model offers this capac-
ity [36,54,57,58]. The three meshes used to assess the impact of the grid resolution
on reacting LES are shown on Fig. 4. Grid characteristics are summarized in Ta-
ble 2 in terms of number of cells, points and minimum/maximum cell volumes. The
meshes are rened in the primary zone, particularly in the lower part where combus-
tion occurs, and in the regions of cooling lms. Due to the restrictive size of cooling
lms, manipulations of the uxes through the faces are needed to specify the cor-
rect Boundary Conditions. This particular approach is applied for the cooling air
effusing from the multi-perforated plates. These plates are also homogeneized:
the ow issuing from the perforations is redistributed over the entire surface and
corrections are performed for momentum uxes and turbulent law-of-the-walls to
account for jet penetrations and wall interactions [59]. The Navier-Stokes Charac-
teristic Boundary Conditions (NSCBC) [56,60] are applied on the other inlet and
outlet boundaries to control the acoustic behavior of the system. Walls are adiabatic
and are treated with a turbulent law-of-the-wall to take into account boundary layer
effects, Table 1. Side boundaries of the computational domain are axi-periodic.
The operating point corresponds to cruising conditions and is the same for the three
grids. The initial condition for the three computations is built as follows: a statisti-
cally stabilized instantaneous solution, obtained with the coarse grid, is interpolated
on the two other meshes. These interpolated solutions are then computed until the
same physical time t
0
to ensure independence on the initial guesses. Finally tem-
9
poral integration is started for all three grids at t
0
until statistical convergence of
the various mean quantities. The averaging time, the time steps and the local CPU
effort are summarized in Table 2. Note that the cost of the three computations for
the same physical time goes from 315 hours on the coarse mesh to 30, 200 on the
ne one (96 times more) so that knowing whether ne grids are really needed and
which additional information they provide is indeed a critical issue to choose a
mesh resolution.
5 Results and discussion
To evaluate the impact of the grid resolution on the LES predictions, the arguments
developed in Pope [24] are addressed. Indeed, it is nowrecognized by the LES com-
munity that two LES instantaneous elds may differ for various reasons [23,25,61].
Figures 5 and 6 illustrate the statement as three instantaneous views obtained from
the three meshes depict degrees of similarity and local differences. The questions of
interest are in this context whether these LES converge towards the same statistics
and whether these statistics are independent of the grid resolution, lter size, LES
model... In this work, grid resolution is investigated as the lter to grid size ratio is
kept xed and equal to one. The LES models are the same for the three simulations:
the Smagorinsky closure for the SGS stress tensor and the DTF model (Section 2)
for the combustion model.
Instantaneous ow eld visualizations are rst presented to illustrate the main ow
features and the general contribution of grid renement when performing reacting
LES of a complex geometry. In the remaining part of the section, temporally av-
eraged elds are presented and analyzed. For that study, the following notation is
adopted to refer to the Reynolds statistical operator:

f (x, t)
_
T
=
1
T
t
0
+T
Z
t
0

f (x, t

) dt

, (13)
10
where T is the time interval used for temporal integration (i.e.: ve ow-through
times, based on the mass ow rate and the primary zone volume). In Eq. (13), the
spatial dependency of
_
T
is omitted but inferred
1
. The Favre notation

f (x, t) is
introduced to emphasize the fact that statistics constructed here are obtained from
Favre ltered quantities and might differ from the true turbulent statistics obtained
in raw experimental measurements for example [15,62]. Furthermore, because of
the ltering procedure introduced in LES, SGS contributions can only be retrieved
based on modeling criteria as underlined by Veynante [63] and Pope [24]. Assess-
ment of the LES resolved rst moment is investigated rst (Section 5.2). Then,
resolved standard deviations are probed in Section 5.3 followed by estimates of the
SGS contributions [24,32,63]. Finally acoustic elds are presented in Section 5.4
using pressure resolved standard deviations and spectra at various locations within
the computational domain.
5.1 Instantaneous ow topology and ame structure
Figure 5 compares instantaneous elds of the axial component of the velocity vec-
tor scaled by the mean velocity at the inlet of the pre-vaporizer for the three com-
putational domains. Although instantaneous elds on the different meshes are dif-
ferent, the main ow structures appear to be very similar: the spreading of the jets
exiting from the pre-vaporizer is well predicted and the high-speed zones induced
by multi-perforated plates are present. Likewise, the impact of the primary jets (bot-
tom liner wall of the main chamber) is equally well captured on all three meshes.
Nevertheless, and as expected, the ne grid computation exhibits more structures
perturbating the primary jet as well as an overall greater amount of small turbulent
structures throughout the computational domain.
A crucial requirement for the LES method when applied in such complex cong-
1
_
T
depends on x and T since the studied ow is not truly stationary. However for long
enough T ,
_
T
can be considered as mainly a function of space.
11
urations is the right prediction of the combustion phenomenon. Modelling, which
is needed to supply proper combustion enhancement due to lack of interactions at
the unresolved scales, is paramount in that context. If improperly parameterized, a
turbulent combustion model can yield different ame positions for LES computed
with different mesh resolutions.
Figure 6 compares instantaneous elds of temperature for the three resolutions.
The cutting plane goes through one of the pre-vaporizer outlets (Plane 1, Fig. 3)
and is colored by the instantaneous eld of temperature scaled by the inlet mean
temperature. The observations drawn from Fig. 5 for the axial component of the
velocity eld also apply to Fig. 6: the temperature elds are clearly enriched
with increasing grid resolution and the impact on the temperature levels seems
reduced.
Figure 7 shows the instantaneous elds of the reaction rate (left column) and
the equivalence ratio in the reaction zone (right column) for the three grids. This
last quantity is the equivalence ratio conditioned by an index equal to zero in
the gases where no combustion takes place and to one in the reaction zones. The
ame position is almost independent of the mesh used and, as expected, the re-
solved wrinkling of the ame front increases with the mesh resolution. These
observations conrm the proper behavior of the turbulence/combustion and the
underlying mechanisms built in the DTF model. Combustion starts close to the
lips of the canes at an equivalence ratio below the value of the one in the canes
(which approaches 3). Indeed, because of the large jet velocities in the cane outlet
regions, combustion can not be sustained immidiately at the cane exits. It allows
partial mixing of the rich premixed fuel mixture with the surrounding fresh air
coming from the upper wall multi-perforated plate (cf. Fig. 3 (b)). Combustion
then occurs for equivalence ratio values approaching 1.7. It then proceeds along
the walls of the chamber (left column). All three LES, even the coarse one, pre-
dict very similar elds of reaction rate.
Figure 8 displays the thickening factor elds (left column) along with the corre-
12
sponding efciency function elds (right column) for the three grids. These re-
sults complement the predictions of the reaction rate elds, Fig. 7 (left column).
An overall agreement is observed between the three simulations. The coarse
mesh solution exhibits less wrinkling and wider ame fronts, compensated by
larger values of the efciency function. Contours of reaction rates coincide with
contours of thickening factor and efciency function on the three meshes. The
grid dependency of the thickening factor accompanied by a proportional adap-
tation of the efciency function leads to similar overall positions of the ame
front. From a statistical point of view, all simulations yield similar mean ame
locations within the combustion chamber and differences may be expected for
higher moments which needs to be determined and quantied for proper assess-
ment. These points are addressed in section 5.2.
As underlined by the right column results of Fig. 7, increasing the ltered eld
content with smaller turbulent structures might also impact the mixture eld to yield
different local ltered compositions. Figure 9 investigates that issue by showing the
Probability Density Function (PDF) of the instantaneous resolved equivalence ratio
in reacting zones, i.e. zones where the local reaction rate is higher than a tenth of
the mean volumetric reaction rate. Combustion mainly occurs in lean premixed
zones, and the tted one-step chemical scheme is seen to operate suitably as very
few ame elements burn in rich conditions ( > 1). The ne mesh results allow
to identify the existence of three peaks which can be identied as follows: (1) the
peak around = 1.7 corresponds to the rich premixed ames created at the cane
outlet, (2) the peak at = 1 is due to the presence of the diffusion amelets and
(3) the lean peak coincides with the mean chamber equivalence ratio ( = 0.33).
The differences in shapes and more specically the presence of a secondary peak
at stoechiometry ( = 1) that is more pronounced as the grid is rened, underline
potential shift in combustion regimes. Further investigations are however needed
as these observations only stem from instantaneous ow studies which will differ
for different grid resolutions as underlined previously. Only temporal statistics are
13
generally of interest to LES predictions.
The preliminary investigation of the instantaneous ow eld conrms the good
quality and behavior of the predictions obtained with all three resolutions. Overall
agreement is clear and differences are only local as expected by the working con-
text of LES. Further analyses are presented below to quantify the impact of mesh
resolution on LES statistics.
5.2 Mean ow results
Spatial elds of the mean temporal statistics, Eq. (13), are good indicators of the
mesh inuence on LES results. In this sub-section, mean elds are investigated for
the three grids for Plane 1, Fig. 3. Figures 10 and 11 respectively show the mean
axial component of the velocity vector and the mean temperature eld scaled by
corresponding mean values at the inlet of the pre-vaporizer.
A good agreement is observed between the three predictions for both quantities,
Figs 10 and 11. The jet spreading at the cane outlet observed on the mean eld of
velocity of Fig. 10 is recovered on the three meshes. High temperature pockets
are located in the same zones and ll similar volumes, assessing the similar mean
behavior of combustion on the three grids (Fig. 11).
Grid resolution has an impact in various highly localized regions and some dis-
crepancies are detected, especially in the mean temperature elds of Fig. 11. The
improved mesh quality of the intermediate and ne grids makes the ow behave
differently in near-wall regions where the chamber ow interacts with the ow
issuing from cooling devices. For example, the thermal boundary layer created
by the multi-perforated plates (Zone 1 and 3 on Fig. 10 (a)) on the intermediate
mesh is thinner than the one on the coarse grid. Likewise, the penetration of the
cooling lm located in the upper part of the liner dome (Zone 2 on Fig. 10 (a)) is
different on the coarse grid. Despite these discrepancies, the agreement between
14
these sets of predictions underlines mesh independence of the rst moments for
these calculations. Convergence in terms of mesh resolution seems to be reached
for the two ner grids at least in terms of rst moments.
More quantitative studies are presented on Fig. 12 where proles in transverse cuts
going through the cane jets and primary jets at different axial positions and for the
same quantities are displayed. Comparisons between the three sets of LES predic-
tions conrm the previous observations on the mean eld maps (Figs. 10 and 11).
Axial velocity proles, Fig. 12 (a), illustrate a slight over-estimation of the axial
velocity magnitude along the center of the cane for the coarse grid predictions.
Likewise, upper boundary layer and recirculating zones located behind the primary
jets seem over-estimated on the coarse grid. The intermediate and ne mesh pro-
les collapse onto the same curves at all stations which indicates that convergence
is reached on the intermediate mesh for that quantity. Furthermore, on the ner
meshes, velocity curves prior to the chamber dome are more diffused on the coarse
grid. Figure 12 presents temperature proles at the same stations: all grid predic-
tions are in very good agreement and discrepancies appear only in the upper part
of the primary zone (Zones 2 and 3) where the stream coming from the top cooling
lm does not penetrate the chamber as much on the ner meshes as on the coarse
grid. Results on the ner meshes predict more local mixing between the chamber
ow and fresh gases coming from the cooling devices. Monotoneous convergence
is achieved for the mean temperature proles: the intermediate mesh predictions
are positioned in-between the coarse and ne grid proles. At other locations, such
a statement is not always true and is explained by the complexity of the ow eld
which is thus very sensitive to various ow and grid parameters.
Spatial distributions of the mean reaction rates in Plane 1 are displayed in Fig. 13.
As underlined from the instantaneous snapshots and the mean eld of temperature,
combustion occurs in the region delimited by the cane outlet and chamber dome
for all meshes. Mean ame anchoring is similar on all three meshes and differences
only appear when localizing the peak reaction rate distributions. Indeed, as the grid
15
resolution increases, the peak value of the mean reaction rate propagates towards
the lips of the cane injector. Convergence for this purely reacting quantity is more
difcult to assess. Such results corroborate the importance of the interaction be-
tween the various models and the local available grid resolution when conducting
LES. They also underline the potential differences in composition as illustrated in
Fig. 9. Such local changes, which may impact mean statistical elds, directly im-
pact Arrhenius-like combustion models such as the DTF model.
5.3 Standard deviations and SGS statistics
One criterion for true LES stems from the observation that at a given value of the
Reynolds number (preferably very large) the total variances, that is to say the sec-
ond central moment based on the unltered variables, should remain independent of
the mesh resolution. The evolution with the grid resolution of the resolved standard
deviation as obtained from the LES eld is henceforth of great interest. Similarly
the SGS contributions need to be evaluated for a proper assessment of the asymp-
totic behavior (if existing) of the total RMS values. The following sub-section dis-
cusses such aspects and presents results obtained from the three grid resolutions for
the resolved RMS values and turbulent kinetic energy for the resolved and residual
motions [24]. The analysis is based on the spatial evolution of the mean temporal
operator, Eq. (13).
Resolved RMS elds for the ltered axial velocity component and temperature are
shown in Figs. 14 and 15 for all meshes.
Figure 14 shows no signicant difference between the three predictions: the RMS
axial velocity elds are similar and values reach approximately the same levels.
This global agreement highlights the proper action of the classical Smagorin-
sky model which results in consistent predictions regardless of the mesh reso-
lution. The SGS velocity model does not infer major changes in the resolved
16
RMS velocity elds. Some local discrepancies can be observed in the wake of
the primary jet and the cane jet, especially on the coarse mesh. These differences
more likely arise from the over-estimated mean velocity eld at these locations
resulting from an under-resolved mesh to allow proper use of the SGS model.
Despite that shortcoming on the coarse mesh, convergence is almost reached on
the intermediate grid and for most of the computational domain.
The resolved RMS temperature, Fig. 15, exhibits larger differences than the re-
solved RMS velocity when the mesh resolution evolves. Higher RMS tempera-
tures are observed on the ne grid, even if levels reached in the primary zone are
very similar between the intermediate and the ne grids. A direct observation of
the ow (Figs. 5 and 6) reveals that these RMS levels are directly due to strong
ame motions which can be captured on the ner grids and not on the coarse
one. Moreover the ner grids are able to capture more ame wrinkling, implying
higher RMS temperature levels. As mentioned before, the local mean reaction
rates are similar on all meshes (Fig. 13) because the DTF model compensates
the reduced resolved ame wrinkling on the coarse mesh by a larger sub-grid
efciency. However, the model cannot explicitely reconstruct the resolved wrin-
kling and RMS temperatures observed in Fig. 15.
Results of Fig. 14 and 15 can be quantitatively conrmed by plotting cuts extracted
from Plane 1, as done on Fig. 16. The proles show that the resolved RMS velocity
depends weakly on the mesh while the resolved RMS temperature grows signi-
cantly when the mesh resolution increases. At this point, it is worth discussing why
these standard deviations behave in such different ways. The main reason has ac-
tually been mentioned many times before in the literature: the velocity eld (and
its standard deviation) is mainly controlled by the large structures and even the
coarse grid is sufcient to capture the RMS levels because the smallest scales are
not modelled by the classical Smagorinsky model (it only models vortex dissipa-
tion beyond a cutting length). On the other hand, combustion is mainly a sub-grid
scale phenomenon, progressing at the smallest scales so that each increase of the
17
mesh resolution will reveal more ame wrinkling and produce higher RMS temper-
ature. The best which can be expected from any ame/turbulence interaction model
is to conserve the zeroth-order moment which is the mean local reaction rate be-
cause it controls the ame position. Predicting the true RMS temperature remains
a challenge for LES of reacting ows
2
which can be tackled only with more so-
phisticated models providing a precise description of the temperature and species
PDF [21,6466] (which is not done with the DTF model).
The quality of a LES computation mostly relies on the scale separation between
the length scales of turbulence captured and the one modeled by the afforded grid
resolution. A quantitative measurement of the turbulence resolution is introduced
in Pope [24] by comparing SGS turbulent kinetic energy k
SGS
with the resolved
turbulent kinetic energy k
resolved
, which reads:
k
resolved
=
1
2
__

U
i

U
i
_
T
__

U
i

U
i
_
T
__
T
(14)
The estimated quantity k
SGS
is derived from the following expression [23]:
SGS
=
C
M

k
SGS
, where
SGS
=
t
is the SGS turbulent viscosity given by the classical
Smagorinsky model [43], C
M
is a constant value and is the lter cutting charac-
teristic length (equal to the cell characteristic length in this study). The resolution
criterion M is thus dened as M = k
SGS
/(k
resolved
+k
SGS
). M varies between 0
(equivalent to DNS where no model is needed) and 1 (corresponding to RANS
where the entire turbulent spectrum is modeled). Figure 17 depicts the spatial evo-
lution of M as obtained from the temporally averaged LES and shows no signicant
discrepancy between the three grids. Regions where M > 0.2 (corresponding to re-
gions where less than 80 % of the turbulent kinetic energy is resolved) are reduced
to the near-wall regions and the potential core of jets on both grids. This criterion
2
Note that eventhough all models for ame/turbulence interactions will exhibit the same
difculties, some models may require less points to resolve the ame front for a given mesh
resolution. For example, the G-equation model [14] will exhibit more wrinkling than the
DTF model for the same resolution.
18
thus guaranties the quality of the results obtained on these three grids, even if care
must be taken when dealing with near-wall regions on the coarse grid. Note that the
Pope criterion only quanties the applicability of SGS modeling for the velocity
eld and its extension to the conserved scalar,

Z, is needed to assess the modeled
to resolved contributions for mixing, which is of great interest from a combustion
point of view. In that case, M
Z
= ZZ
SGS
/(ZZ
resolved
+ZZ
SGS
) is proposed and pre-
sented on Fig. 18. Similarly to Popes criterion, the SGS termis estimated following
the expression of [63]. For mixing, the impact of the grid resolution is observed
in the same regions. Figure 18 also underlines the reduced effect of the grid reso-
lution on M
Z
at the ame front, which is expected since the DTF model provides
a resolved eld with no SGS contribution at that specic location, the interaction
between turbulence and combustion being represented through the efciency func-
tion. Note in that respect that other LES combustion models are expected to behave
differently at the ame front.
5.4 Acoustic elds
The prediction of thermo-acoustic phenomena is of growing interest for industrial
purposes as gas turbine manufacturers are constrained, mostly for environmental
concerns, to design combustion devices able to operate at lean regimes, which are
known to be sometimes unstable. The contribution of LES in this framework is
essential and has already proved its capability to predict ame/acoustics interac-
tions [15,36,54,58,67]. A central question to use LES for combustion instability
studies is to know whether unsteady LES results (pressure elds for example) are
grid-independent. Since the previous section has indicated that the RMS temper-
ature is dependent on grid resolution, checking whether the pressure eld is also
dependent on the mesh resolution is a logical next step.
RMS pressure elds exhibit very similar patterns on all meshes (Fig. 19) even
though the ner meshes lead to slightly higher p

levels. The three simulations iden-


19
tify the same acoustic activity in the combustor as shown by the pressure spectra
displayed in Fig. 20 for two probes: Probe 1 is located near one of the pre-vaporizer
outlets and Probe 2 in the primary zone (Fig. 3). LES reveals two modes around
8700 and 9300 Hz which are recovered on the different meshes with reasonable
accuracy. In such complex geometries, these modes can not be associated to any
simple mode (i.e. quarter wave, etc). The only statement which can be made is that
they are clearly associated to transverse activity between the upper and lower walls
of the chamber.
The fact that all LES provide similar results for the p

elds must be interpreted


with care:
The unsteady pressure eld is the convolution of the source terms (due to un-
steady reaction rate) and of acoustic propagation in the chamber. The acoustic
propagation takes place on long wave lengths (typically tens of centimeters at
10kHz in the burnt gases) and is probably not very sensitive to mesh changes.
The mean value of the source term has also been shown to be insensitive to the
mesh but the dependence of its unsteady component on mesh resolution, on the
other hand, is not studied here. This is obviouly the next step to take.
Predicting unsteady pressure in LES of reacting ows remains a signicant chal-
lenge and the present results only suggest that predicting unsteady pressure in a
real combustor with LES is actually not more difcult than predicting the mean
ame position or the velocity statistics. Other studies [54,6870] have indicated
the same trend and this optimistic outcome is probably justied only for the pre-
diction of the acoustic mode frequencies (which depend mainly on the geometry
and not much on the source terms) but not for the amplitudes of the modes.
In other words, predicting whether a combustor will oscillate and at which fre-
quency is a task which is within the possibilities of LES. However predicting the
exact amplitude of the oscillation might remain more difcult for a long time.
Indeed, this amplitude will depend not only on the mesh and on the prediction of
the unsteady source terms as well as on physical dissipation mechanisms but also
20
on the dissipation of the numerical method, on the acoustic boundary conditions,
etc.
6 Conclusions
Although LES is the most promising tool to study unsteady phenomena in complex
geometry combustion chambers, few studies have addressed the question of the de-
pendence of LES results to mesh resolution in realistic congurations. This paper
considers this problem by performing three LES of the same helicopter combustion
chamber, the rst one with 1, 242, 086 elements, the second one with 10, 620, 245
elements and the third one with 43, 949, 682 elements. Multiple comparisons be-
tween the three sets of LES results are performed. The ame position, its unsteady
behaviour and the mean ow elds (velocity, temperature, reaction rate) are shown
to be reasonably insensitive to mesh resolution. The standard deviation of the re-
solved velocity eld is also shown to depend weakly on resolution, as conrmed by
the Pope criterion, showing that most of the kinetic energy spectrum is resolved on
the different grids. However, the RMS temperature is shown to increase when the
mesh resolution increases, consistently with the fact that combustion is a sub-grid
phenomenon and that its largest part is in the unresolved component. Finally, even
though the RMS temperature depends strongly on mesh resolution, the acoustic ac-
tivity is shown to be less dependent, exhibiting similar acoustic modes (frequency
and amplitudes) on all the three meshes, suggesting that LES is reasonably adapted
to the study of combustion instability in complex geometry combustors.
Acknowledgments
Numerical simulations and visualizations have been conducted on the computers of
the French National Computing Center (CINES) in Montpellier, at the Barcelona
21
Supercomputing Center (BSC) located in Spain and CERFACS in-house computing
facility.
22
References
[1] W. Polifke, A. Fischer and T. Sattelmayer, ASME journal of engineering for gas
turbines and power 125 (2003) 20.
[2] V. Sankaran and S. Menon, J. Turb. 3 (2002) 011.
[3] D. Smith and E. Zukoski, 21st Joint Propulsion Conference, pp. AIAA paper 851248,
Monterey, 1985.
[4] P. Billant, J.M. Chomaz and P. Huerre, J. Fluid Mech. 376 (1998) 183.
[5] A. Gupta, D. Lilley and N. Syred, Swirl ows (Abacus Press, 1984).
[6] M. Hall, Ann. Rev. Fluid Mech. 4 (1972) 195.
[7] O. Lucca-Negro and T. ODoherty, Prog. Energy Comb. Sci. 27 (2001) 431.
[8] S. Roux et al., Combust. Flame 141 (2005) 40.
[9] Y. Mizobuchi et al., Proc. of the Combustion Institute 30 (2004) 611.
[10] L. Mu niz and M. Mungal, Combust. Flame 111 (1997) 16.
[11] H. Pitsch and H. Steiner, Proc. of the Combustion Institute 28 (2000) 41.
[12] L. Vervisch et al., J. Turb. 5 (2004) 004.
[13] N. Peters, Turbulent combustion (Cambridge University Press, 2000).
[14] H. Pitsch, Ann. Rev. Fluid Mech. 38 (2006) 453.
[15] T. Poinsot and D. Veynante, Theoretical and numerical combustion (R.T. Edwards,
2nd edition., 2005).
[16] H. Pitsch and H. Steiner, Phys. Fluids 12 (2000) 2541.
[17] H. Pitsch and L. Duchamp de la Geneste, Proc of the Comb. Institute 29 (2002) 2001.
[18] C. Duwig et al., AIAA Journal 45 (2007) 624.
[19] J. Janicka and A. Sadiki, Proc. of the Combustion Institute 30 (2004) 537.
23
[20] H. El-Asrag and S. Menon, Proc. of the Combustion Institute 31 (2007) 1747.
[21] H. El-Asrag and S. Menon, AIAA/SAE/ASME/ASEE 41st Joint Propulsion
Conference, Tucson, USA, 2005.
[22] O. Stein and A. Kempf, Proc. of the Combustion Institute 31 (2007) 1755.
[23] P. Sagaut, Large Eddy Simulation for incompressible ows (Springer-Verlag, 2000).
[24] S.B. Pope, New Journal of Physics 6 (2004) 35.
[25] S. Ghosal and P. Moin, J. Comput. Phys. 118 (1995) 24 .
[26] S.B. Pope, Turbulent ows (Cambridge University Press, 2000).
[27] B. Vreman, B. Geurts and H. Kuerten, International Journal for Numerical Methods
in Fluids 22 (1996) 297.
[28] J. Meyers, B. Geurts and M. Baelmans, Phys. Fluids 15 (2003) 2740.
[29] C. Pri` ere et al., AIAA Journal 43 (2005) 1753.
[30] A. Vreman, B. Geurts and J. Kuerten, J. Eng. Math. 29 (1995) 299.
[31] B. Geurts and J. Fr ohlich, Physics of Fluids (2002) L41.
[32] M. Klein, Flow Turb. and Combustion 75 (2005) 131.
[33] I. Celik, Z. Cehreli and I. Yavuz, Journal of Fluids Engineering 127 (2005) 949.
[34] N. Allsopp et al., Unfolding the IBM eServer Blue Gene Solution, First ed. (IBM
Redbooks, 2005) pp. 277285.
[35] G. Boudier et al., Proc. of the Combustion Institute 31 (2007) 3075.
[36] P. Schmitt et al., J. Fluid Mech. 570 (2007) 17.
[37] L. Selle et al., Combust. Flame 137 (2004) 489.
[38] C. Pri` ere et al., J. Turb. 5 (2004) 1.
[39] A.A. Aldama, Lecture Notes in Engineering Vol. 49 (Springer-Verlag, New York,
1990).
24
[40] B. Vreman, B. Geurts and H. Kuerten, Phys. Fluids 6 (1994) 4057.
[41] A. Favre, Problems of hydrodynamics and continuummechanics, pp. 231266, SIAM,
Philadelphia, 1969.
[42] J. Ferziger and M. Peri c, Computational Methods for Fluid Dynamics (Springer
Verlag, Berlin, Heidelberg, New York, 1997).
[43] J. Smagorinsky, Monthly Weather Review 91 (1963) 99.
[44] P. Moin et al., Phys. Fluids A 3 (1991) 2746.
[45] G. Erlebacher et al., J. Fluid Mech. 238 (1992) 155.
[46] F. Ducros, P. Comte and M. Lesieur, J. Fluid Mech. 326 (1996) 1.
[47] D.K. Lilly, Phys. Fluids A4 (1992) 633.
[48] M. Germano, J. Fluid Mech. 238 (1992) 325.
[49] C. Meneveau, T. Lund and W. Cabot, J. Fluid Mech. 319 (1996) 353.
[50] A. Linan and F. Williams, Fundamental aspects of combustion (Oxford University
Press, 1993).
[51] F. Williams, Combustion theory (Benjamin Cummings, Menlo Park, CA, 1985).
[52] O. Colin et al., Phys. Fluids 12 (2000) 1843.
[53] J.P. L egier, Simulations num eriques des instabilit es de combustion dans les foyers
a eronautiques, Phd thesis, INP Toulouse, 2001.
[54] C. Martin et al., AIAA Journal 44 (2006) 741.
[55] O. Colin and M. Rudgyard, J. Comput. Phys. 162 (2000) 338.
[56] V. Moureau et al., J. Comput. Phys. 202 (2005) 710.
[57] Y. Sommerer et al., J. of Turbulence 5 (2004).
[58] A. Roux et al., Combust. Flame 152 (2007) 154.
[59] S. Mendez et al., Summer Program, pp. 5772, Center for Turbulence Research,
NASA AMES, Stanford University, USA, 2006.
25
[60] T. Poinsot, T. Echekki and M. Mungal, Combust. Sci. Tech. 81 (1992) 45.
[61] M. Lesieur, New tools in turbulence modelling, edited by O. M etais and J. Ferziger,
pp. 1 28, Les Editions de Physique - Springer Verlag, 1997.
[62] J. Panda and R. Seasholtz, AIAA Journal 44 (2006) 1952.
[63] D. Veynante and R. Knikker, J. Turb. 7 (2006) 1.
[64] S. Pope et al., Proc. of the Summer Program, pp. 283294, Center for Turbulence
Research, Stanford, USA, 2004.
[65] A. Klimenko and R. Bilger, Prog. Energy Comb. Sci. 25 (1999) 595 .
[66] S. James, J. Zhu and M. Anand, Proc. of the Combustion Institute 31 (2007) 1737.
[67] A. Giauque et al., J. Turb. 6 (2005) 1.
[68] A. Sengissen, Simulation aux grandes echelles des instabilit es de combustion : vers le
couplage uide/structure - TH/CFD/06/12, PhD thesis, Universit e de Montpellier II,
2006.
[69] R.A. Huls et al., J. Sound Vib. 304 (2007).
[70] A. Sengissen et al., Combust. Flame 150 (2007) 40.
26
List of Tables
1 BC types and specicities for LES. 29
2 Mesh characteristics used for the coarse, the intermediate and
ne LES meshes. 30
27
TABLES
28
Boundary type Acoustic properties
Specied quantities Relax values
Cane inlet, dilution
and primary holes
NSCBC inlet
u
n
, T, P, Y

50
Outlet NSCBC outlet
P 200
Inlet lms Purely reective inlet
u
n
, T, P, Y

-
Multi-perforated
plates
Purely reective inlet
with momentum
correction
u
n
, T, P, Y

-
Adiabatic walls
Purely reective wall
with law-of-the-wall
- -
Table 1
BC types and specicities for LES.
29
Coarse mesh Intermediate mesh Fine mesh
Total number of points 230, 118 1, 875, 835 7, 661, 005
Total number of cells 1, 242, 086 10, 620, 245 43, 949, 682
Max. cell volume [m
3
] 3.1267110
8
8.9780210
9
4.0574810
9
Min. cell volume [m
3
] 1.81795 10
11
8.29479 10
12
1.1828 10
12
Time step [s] 1.5210
7
0.8810
7
0.4910
7
Averaging time [ms] 10 10 10
Number of iterations during averaging 65, 790 108, 695 204, 082
Total CPU effort for a 10ms(hours) LES 315 4, 550 30, 200
Table 2
Mesh characteristics used for the coarse, the intermediate and ne LES meshes.
30
List of Figures
1 Typical speed-up curve as obtained for the conguration studied in
this work. 34
2 Chemical scheme validation: (a) is ame speed prediction, (b)
is adiabatic ame temperature as functions of the equivalence
ratio. The detailed scheme for C
10
H
16
contains 43 species and 174
chemical steps. 35
3 Combustion chamber: (a) 3D view and (b) side view of the
computational domain. 36
4 Mesh resolution for (a) the coarse mesh, (b) the intermediate mesh
and (c) the ne mesh, as obtained in a transversal cut passing
through one of the canes (Plane 1). 37
5 Instantaneous axial velocity eld scaled by the mean inlet velocity
and obtained on (a) the coarse grid, (b) the intermediate one and
(c) the ne one. 38
6 Instantaneous temperature eld scaled by the mean inlet
temperature and obtained on (a) the coarse grid, (b) the
intermediate one and (c) the ne one. 39
7 Instantaneous elds of reaction rate (a, c and e) and equivalence
ratio conditioned by the local reaction (b, d and f) on Plane 1
obtained on the coarse grid, (a and b), the intermediate grid (c and
d) and the ne one (e and f). 40
31
8 Instantaneous elds of the thickening factor (a, c and e) and the
efciency function (b, d and f) on Plane 1 obtained on the coarse
grid (a and b), the intermediate grid (c and d) and the ne one (e
and f). 41
9 Probability density function of local equivalence ratio in reacting
zones (zones where local reaction rate is higher than a tenth of
the mean volumetric reaction rate) on (a) the coarse grid, (b) the
intermediate one and (c) the ne one. 42
10 Mean axial velocity eld (scaled by the mean velocity at the
pre-vaporizer inlet) obtained on (a) the coarse grid, on (b) the
intermediate grid and on (c) the ne grid. 43
11 Mean temperature eld (scaled by the mean temperature at the
pre-vaporizer inlet) obtained on (a) the coarse grid, on (b) the
intermediate grid and on (c) the ne grid. 44
12 Transverse proles in Plane 1, Fig. 3, passing through one cane
outlet: (a) mean axial velocity proles scaled by the pre-vaporizer
inlet velocity and (b) mean temperature proles scaled by by the
pre-vaporizer inlet temperature. 45
13 Mean reaction rate on Plane 1 obtained on (a) the coarse grid, (b)
the intermediate one and (c) the ne one. 46
14 Fields of RMS resolved axial velocity on (a) the coarse grid, (b)
the intermediate grid and (c) the ne grid. The quantities are
non-dimensionalized by pre-vaporizer inlet conditions. 47
15 Fields of RMS resolved temperature on (a) the coarse grid, (b)
the intermediate grid and (c) the ne grid. The quantities are
non-dimensionalized by pre-vaporizer inlet conditions. 48
32
16 Transverse proles in Plane 1 (cf Fig. 3), passing through one cane
outlet: (a) RMS axial velocity proles scaled by squared inlet
velocity and (b) RMS temperature proles scaled by squared inlet
temperature. 49
17 Spatial evolution of the temporal average of the resolution criterion
M: (a) is the coarse grid, (b) is the intermediate grid and (c) is the
ne grid. Black isoline corresponds to M = 0.2. 50
18 Spatial evolution of the temporal average of the resolution criterion
M
Z
: (a) is the coarse grid, (b) is the intermediate grid and (c) is the
ne grid. Black isoline corresponds to M = 0.2. 51
19 RMS pressure eld obtained on (a) the coarse grid, (b) the
intermediate grid and (c) the ne grid. The quantities are
non-dimensionalized by pre-vaporizer inlet conditions. 52
20 Spectra of the uctuating pressure temporal evolution at (a) Probe
1 and (b) Probe 2 (cf Fig. 3). 53
33
FIGURES
4000
3000
2000
1000
S
p
e
e
d
u
p
4000 3000 2000 1000
Number of processors
Ideal Speedup
40.10
6
cells mesh
Fig. 1. Typical speed-up curve as obtained for the conguration studied in this work.
34
(a)
2.0
1.5
1.0
0.5
F
l
a
m
e

s
p
e
e
d

[
m
.
s
-
1
]
3.0 2.5 2.0 1.5 1.0 0.5
Equivalence ratio [-]
Detailed scheme
1-step scheme
Fitted 1-step scheme
(b)
2600
2400
2200
2000
1800
T
e
m
p
e
r
a
t
u
r
e

[
K
]
3.0 2.5 2.0 1.5 1.0 0.5
Equivalence ratio [-]
Detailed scheme
1-step scheme
Fitted 1-step scheme
Fig. 2. Chemical scheme validation: (a) is ame speed prediction, (b) is adiabatic ame
temperature as functions of the equivalence ratio. The detailed scheme for C
10
H
16
contains
43 species and 174 chemical steps.
35
(a)
Primary holes
Fuel
inlet
Dilution holes
Probe 2
Liner dome
Probe 1
Plane 1
(b)
Multi-perforated plates Cooling
film
Multi-perforated plates
Cooling films
Cooling film for
Nozzle Guide Vane
Primary holes
Dilution holes
Exit
Fuel
inlet
Fig. 3. Combustion chamber: (a) 3D view and (b) side view of the computational domain.
36
(a)
(b)
(c)
Fig. 4. Mesh resolution for (a) the coarse mesh, (b) the intermediate mesh and (c) the ne
mesh, as obtained in a transversal cut passing through one of the canes (Plane 1).
37
(a)
1.422
0.711
0
-0.711
-1.422
!

U /U
inlet
" [ ]
(b)
1.422
0.711
0
-0.711
-1.422
!

U /U
inlet
" [ ]
(c)
1.422
0.711
0
-0.711
-1.422
!

U /U
inlet
" [ ]
Fig. 5. Instantaneous axial velocity eld scaled by the mean inlet velocity and obtained on
(a) the coarse grid, (b) the intermediate one and (c) the ne one.
38
(a)
4.613
3.708
2.803
1.898
1
!

T /T
inlet
" [ ]
(b)
4.613
3.708
2.803
1.898
1
!

T /T
inlet
" [ ]
(c)
4.613
3.708
2.803
1.898
1
!

T /T
inlet
" [ ]
Fig. 6. Instantaneous temperature eld scaled by the mean inlet temperature and obtained
on (a) the coarse grid, (b) the intermediate one and (c) the ne one.
39
(a)
5000
3750
2500
1250
0
Reaction rate [mol.m
-3
.s
-1
]
(b)
3
2.25
1.5
0.75
0
Equivalence ratio []
(c)
5000
3750
2500
1250
0
Reaction rate [mol.m
-3
.s
-1
]
(d)
3
2.25
1.5
0.75
0
Equivalence ratio []
(e)
5000
3750
2500
1250
0
Reaction rate [mol.m
-3
.s
-1
]
(f)
3
2.25
1.5
0.75
0
Equivalence ratio []
Fig. 7. Instantaneous elds of reaction rate (a, c and e) and equivalence ratio conditioned
by the local reaction (b, d and f) on Plane 1 obtained on the coarse grid, (a and b), the
intermediate grid (c and d) and the ne one (e and f).
40
(a)
35
26.5
18
9.5
1
Thickening
(b)
8
6.25
4.5
2.75
1
Efficiency
(c)
35
26.5
18
9.5
1
Thickening
(d)
8
6.25
4.5
2.75
1
Efficiency
(e)
35
26.5
18
9.5
1
Thickening
(f)
8
6.25
4.5
2.75
1
Efficiency
Fig. 8. Instantaneous elds of the thickening factor (a, c and e) and the efciency function
(b, d and f) on Plane 1 obtained on the coarse grid (a and b), the intermediate grid (c and d)
and the ne one (e and f).
41
(a)
1.6
1.2
0.8
0.4
0.0
P
r
o
b
a
b
i
l
i
t
y

d
e
n
s
i
t
y

[

]
3.0 2.5 2.0 1.5 1.0 0.5 0.0
Equivalence ratio []
(b)
1.6
1.2
0.8
0.4
0.0
P
r
o
b
a
b
i
l
i
t
y

d
e
n
s
i
t
y

[

]
3.0 2.5 2.0 1.5 1.0 0.5 0.0
Equivalence ratio []
(c)
1.6
1.2
0.8
0.4
0.0
P
r
o
b
a
b
i
l
i
t
y

d
e
n
s
i
t
y

[

]
3.0 2.5 2.0 1.5 1.0 0.5 0.0
Equivalence ratio []
Fig. 9. Probability density function of local equivalence ratio in reacting zones (zones where
local reaction rate is higher than a tenth of the mean volumetric reaction rate) on (a) the
coarse grid, (b) the intermediate one and (c) the ne one.
42
(a)
1.422
0.711
0
-0.711
-1.422
Zone 1 Zone 2
Zone 3
!

U /U
inlet
" [ ]
(b)
1.422
0.711
0
-0.711
-1.422
!

U /U
inlet
" [ ]
(c)
1.422
0.711
0
-0.711
-1.422
!

U /U
inlet
" [ ]
Fig. 10. Mean axial velocity eld (scaled by the mean velocity at the pre-vaporizer inlet)
obtained on (a) the coarse grid, on (b) the intermediate grid and on (c) the ne grid.
43
(a)
4.464
3.598
2.732
1.866
1
Zone 1 Zone 2
Zone 3
!

T /T
inlet
" [ ]
(b)
4.464
3.598
2.732
1.866
1
!

T /T
inlet
" [ ]
(c)
4.464
3.598
2.732
1.866
1
!

T /T
inlet
" [ ]
Fig. 11. Mean temperature eld (scaled by the mean temperature at the pre-vaporizer inlet)
obtained on (a) the coarse grid, on (b) the intermediate grid and on (c) the ne grid.
44
(a)
-5
0
5
V
e
r
t
i
c
a
l

d
i
s
t
a
n
c
e

y
/
R
c
a
n
e

[
-
]
-2.5 0.0 -0.4 0.4
-2 0
Fine mesh
Intermediate mesh
Coarse mesh
(b)
4 2 0
2.5 0.0 4 2 0
Fine mesh
Intermediate mesh
Coarse mesh
-5
0
5
V
e
r
t
i
c
a
l

d
i
s
t
a
n
c
e

y
/
R
c
a
n
e

[
-
]
Fig. 12. Transverse proles in Plane 1, Fig. 3, passing through one cane outlet: (a) mean
axial velocity proles scaled by the pre-vaporizer inlet velocity and (b) mean temperature
proles scaled by by the pre-vaporizer inlet temperature.
45
(a)
5000
3750
2500
1250
0
Reaction rate [mol.m
-3
.s
-1
]
(b)
5000
3750
2500
1250
0
Reaction rate [mol.m
-3
.s
-1
]
(c)
5000
3750
2500
1250
0
Reaction rate [mol.m
-3
.s
-1
]
Fig. 13. Mean reaction rate on Plane 1 obtained on (a) the coarse grid, (b) the intermediate
one and (c) the ne one.
46
(a)
1.193
0.925
0.656
0.388
0.119
!
U
RMS
/U
inlet
" [ ]
(b)
1.193
0.925
0.656
0.388
0.119
!
U
RMS
/U
inlet
" [ ]
(c)
1.193
0.925
0.656
0.388
0.119
!
U
RMS
/U
inlet
" [ ]
Fig. 14. Fields of RMS resolved axial velocity on (a) the coarse grid, (b) the intermediate
grid and (c) the ne grid. The quantities are non-dimensionalized by pre-vaporizer inlet
conditions.
47
(a)
1
0.775
0.55
0.225
0.1
!
T
RMS
/T
inlet
" [ ]
(b)
1
0.775
0.55
0.225
0.1
!
T
RMS
/T
inlet
" [ ]
(c)
1
0.775
0.55
0.325
0.1
!
T
RMS
/T
inlet
" [ ]
Fig. 15. Fields of RMS resolved temperature on (a) the coarse grid, (b) the intermediate
grid and (c) the ne grid. The quantities are non-dimensionalized by pre-vaporizer inlet
conditions.
48
(a)
-5
0
5
V
e
r
t
i
c
a
l

d
i
s
t
a
n
c
e

y
/
R
c
a
n
e

[
-
]
0.6 0.0
1.0 0.0 0.5 0.0
Fine mesh
Intermediate mesh
Coarse mesh
(b)
1.0 0.0
0.8 0.4 0.0
0.8 0.4 0.0
Fine mesh
Intermediate mesh
Coarse mesh
-5
0
5
V
e
r
t
i
c
a
l

d
i
s
t
a
n
c
e

y
/
R
c
a
n
e

[
-
]
Fig. 16. Transverse proles in Plane 1 (cf Fig. 3), passing through one cane outlet: (a) RMS
axial velocity proles scaled by squared inlet velocity and (b) RMS temperature proles
scaled by squared inlet temperature.
49
(a)
1
0.75
0.5
0.25
0
M criterion
(b)
1
0.75
0.5
0.25
0
M criterion
(c)
1
0.75
0.5
0.25
0
M criterion
Fig. 17. Spatial evolution of the temporal average of the resolution criterion M: (a) is the
coarse grid, (b) is the intermediate grid and (c) is the ne grid. Black isoline corresponds
to M = 0.2.
50
(a)
1
0.75
0.5
0.25
0
M
z
criterion
(b)
1
0.75
0.5
0.25
0
M
z
criterion
(c)
1
0.75
0.5
0.25
0
M
z
criterion
Fig. 18. Spatial evolution of the temporal average of the resolution criterion M
Z
: (a) is the
coarse grid, (b) is the intermediate grid and (c) is the ne grid. Black isoline corresponds
to M = 0.2.
51
(a)
0.0658
0.051
0.0362
0.0214
0.0066
!
P
RMS
/P
inlet
" [ ]
(b)
0.0658
0.051
0.0362
0.0214
0.0066
!
P
RMS
/P
inlet
" [ ]
(c)
0.0658
0.051
0.0362
0.0214
0.0066
!
P
RMS
/P
inlet
" [ ]
Fig. 19. RMS pressure eld obtained on (a) the coarse grid, (b) the intermediate grid and
(c) the ne grid. The quantities are non-dimensionalized by pre-vaporizer inlet conditions.
52
(a)
20000
15000
10000
5000
0
A
m
p
l
i
t
u
d
e

[
P
a
.
s
-
1
]
16000 12000 8000 4000 0
Frequency [Hz]
Coarse mesh
Intermediate mesh
Fine mesh
(b)
30000
20000
10000
0
A
m
p
l
i
t
u
d
e

[
P
a
.
s
-
1
]
16000 12000 8000 4000 0
Frequency [Hz]
Coarse mesh
Intermediate mesh
Fine mesh
Fig. 20. Spectra of the uctuating pressure temporal evolution at (a) Probe 1 and (b) Probe
2 (cf Fig. 3).
53
ADDITIONALS
1 Color version of Fig. 5: i.e. Instantaneous axial velocity eld
scaled by the mean inlet velocity and obtained on (a) the coarse
grid, (b) the intermediate one and (c) the ne one. 56
2 Color version of Fig. 6: i.e. Instantaneous temperature eld scaled
by the mean inlet temperature and obtained on (a) the coarse grid,
(b) the intermediate one and (c) the ne one. 57
3 Color version of Fig. 7: i.e. Instantaneous elds of reaction rate (a,
c and e) and equivalence ratio conditioned by the local reaction
(b, d and f) on Plane 1 obtained on the coarse grid, (a and b), the
intermediate grid (c and d) and the ne one (e and f). 58
4 Color version of Fig. 8: i.e. Instantaneous elds of the thickening
factor (a, c and e) and the efciency function (b, d and f) on Plane
1 obtained on the coarse grid (a and b), the intermediate grid (c
and d) and the ne one (e and f). 59
5 Color views of the instantaneous elds of temperature on Plane 1
along with an isosurface of reaction rate (in blue) obtained on the
coarse grid (a), the intermediate grid (b) and the ne one (c). 60
54
ADDITIONALS
55
(a)
(b)
(c)
Additional 1: Color version of Fig. 5: i.e. Instantaneous axial velocity eld scaled
by the mean inlet velocity and obtained on (a) the coarse grid, (b) the intermediate
one and (c) the ne one.
56
(a)
(b)
(c)
Additional 2: Color version of Fig. 6: i.e. Instantaneous temperature eld scaled by
the mean inlet temperature and obtained on (a) the coarse grid, (b) the intermediate
one and (c) the ne one.
57
(a) (b)
(c) (d)
(e) (f)
Additional 3: Color version of Fig. 7: i.e. Instantaneous elds of reaction rate (a, c
and e) and equivalence ratio conditioned by the local reaction (b, d and f) on Plane
1 obtained on the coarse grid, (a and b), the intermediate grid (c and d) and the ne
one (e and f).
58
(a) (b)
(c) (d)
(e) (f)
Additional 4: Color version of Fig. 8: i.e. Instantaneous elds of the thickening
factor (a, c and e) and the efciency function (b, d and f) on Plane 1 obtained on
the coarse grid (a and b), the intermediate grid (c and d) and the ne one (e and f).
59
(a)
(b)
(c)
Additional 5: Color views of the instantaneous elds of temperature on Plane 1
along with an isosurface of reaction rate (in blue) obtained on the coarse grid (a),
the intermediate grid (b) and the ne one (c).
60

Вам также может понравиться