You are on page 1of 27

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

net/publication/245312203

Study of flame stabilization in a swirling combustor


using a new flamelet formulation

Article  in  Combustion Science and Technology · August 2005


DOI: 10.1080/00102200590956669

CITATIONS READS

36 147

2 authors, including:

Christophe Duwig
KTH Royal Institute of Technology
70 PUBLICATIONS   932 CITATIONS   

SEE PROFILE

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

GREENEST View project

collar connection View project

All content following this page was uploaded by Christophe Duwig on 10 August 2015.

The user has requested enhancement of the downloaded file.


Combust. Sci. and Tech., 177: 1485–1510, 2005
Copyright Q Taylor & Francis Inc.
ISSN: 0010-2202 print/1563-521X online
DOI: 10.1080/00102200590956669

STUDY OF FLAME STABILIZATION IN A


SWIRLING COMBUSTOR USING A NEW
FLAMELET FORMULATION

CHRISTOPHE DUWIG
LASZLO FUCHS
Division of Fluid Mechanics, Department of Heat and
Power Engineering, Lund Institute of Technology,
Lund, Sweden

The dynamics in a swirl-stabilized flame is studied using large eddy


simulation (LES). We account for the effect of turbulence on the
flame through a model based on a filtered flamelet technique. The
model provides a consistent and robust reaction-diffusion expression
for simulating the correct propagation of premixed flames. The
filtered flamelet formulation has been implemented into a high-
order-accurate LES code and used to study the flame stabilization
and the combustion dynamics in a gas-turbine combustion chamber.
The effects of inlet boundary conditions, in terms of velocity and
equivalence ratio radial profiles, have been studied. The flow is
found to be very sensitive to small changes in terms of flame shapes
and anchoring position. The sensitivity of the results to the subgrid-
scale flame thickness has also been investigated. The influence on
the flame position is not significant. However, a too-large subgrid-
scale flame thickness leads to different flame dynamics.

Keywords: large eddy simulation, premixed combustion, flamelet,


gas turbine, swirling flow

Received 15 November 2004; accepted 21 December 2004.


The authors thank Mr. Dragos Moroianu for his help with some of the data visualiza-
tions. The computations were run on LUNARC and HPC2N computing facilities with the
SNAC allocation 343-2002-2315-14. This work was financially supported by the Swedish
Energy Agency STEM through the TPE program.

Address correspondence to Christophe.Duwig@vok.lth.se

1485
1486 C. DUWIG AND L. FUCHS

INTRODUCTION
Today, the major challenge of energy technologies is to meet the growing
world demand while fulfilling new regulations on pollutant emissions.
Thermal power technologies are widely used for electricity production
but enforcement of new laws on harmful emissions will require new tech-
nologies to make thermal power plants cleaner and more efficient.
Recently, the gas-turbine (GT) industry adopted swirling premixed
combustor that lead to a significant decrease of NOx emissions
(Lefevbre, 1995). However, stability problems arise during the operation
of the device, leading to further studies related to combustion stability
(Correa, 1998).
The dynamics of flames is the result of the interactions among the
fluctuations in fuel=air concentrations, heat release, velocity, and press-
ure (Stone and Menon, 2001). The main difficulty lies in the nonlinear
behavior of chemistry and turbulence phenomena. For example, turbu-
lent mixing instabilities lead to concentration and heat-release fluctua-
tions. Such fluctuations can generate acoustic waves, which might
travel to the fuel injector and generate equivalence ratio fluctuations.
Paschereit et al. (1998) showed the importance of the large-scale flow
structure in the flame=flow interaction. This mechanism is of relevance
for the GT combustors because large scales play an important role in
flame stabilization (e.g., through vortex breakdown).
Vortex breakdown is commonly used to anchor and stabilize flames
(Lefevbre, 1995). A swirling motion is usually added to a jet flow. A
swirling jet is subject to centrifugal forces, leading to a radial expansion
of the jet. A low-pressure region appears around the axis region, close to
the expansion. If the swirling motion is strong enough, the longitudinal
pressure gradient induces an axial backflow and a stagnation point. This
type of recirculation is of great importance for engineering applications
because it might bring burned hot gases toward the fresh gases, thereby
stabilizing the flame. In addition, the expansion enhances mixing and
flame surface area.
Modeling and understanding the vortex breakdown is then a key
issue in flame stabilization. However, despite more than 40 years of
research, the mechanisms of vortex breakdown are only partially under-
stood. The main difficulty of the problem is the unsteady behavior of this
type of flow (cf. Lucca-Negro and O’Doherty, 2001): large structures
resulting from vortex breakdown and the swirling shear layers, affect
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1487

the flame stabilization directly, leading to heat-release fluctuations and


combustion instabilities. The issue is therefore to model and understand
the interaction between the flame and the vortex for answering the ques-
tion: ‘‘How stable will it be?’’ To answer this question, it is necessary to
perform a sensitivity analysis of the stabilization process. Such an analy-
sis is carried out in this paper using an unsteady numerical simulation
tool based on large eddy simulation (LES).
LESs has been shown to be most suitable for handling unsteady
coherent structures in GT combustors (Kim et al., 1999). However, there
are several open issues related to LES. These issues include subgrid-scale
(SGS) models in the momentum and energy equations as well as the
equations related to species transport and equation of state that are
required to close the system. In particular, there is only limited theoreti-
cal foundation for modeling the SGS terms related to combustion. A
central issue is that the flame becomes as thick as the order of the local
grid. Altering the flame thickness, without adjusting the flame speed
appropriately, significantly changes the dynamics of the flame. An
approach that remedies this aspect has been proposed by Chakravarthy
and Menon (2000). The so-called linear eddy model computes the SGS
terms using a one-dimensional (1D) equation. The problem is locally
reduced to a 1D problem oriented in the direction of the progress vari-
able gradient and the turbulence stirring is included using stochastic
rearrangement events. The main drawbacks of this model are the high
CPU costs and the inability to include flame-front curvature. To avoid
a sharp increase of CPU costs, Colin et al. (2000) proposed the ‘‘thick-
ened flame technique.’’ The flame is artificially thickened so that it can
be resolved on the LES grid. The flame propagation speed is not modi-
fied by thickening. Another approach has been used by Kim et al. (1999).
Flame tracking is performed using a level-set technique. Knowing the
flame position enables one to locate the heat release and the density
gradient associated with the flame. However, to stabilize the level-set
equation, a nonphysical diffusion is introduced. It is unclear how this
numerical stabilization affects the flame dynamics.
Recently, Duwig (2002, 2003) proposed a filtered flamelet technique
to account for the SGS in the progress variable. The technique is based
on physical arguments and is presented and used in the paper.
In the next section, the problem of turbulent combustion modeling
using LES is formulated. In the third section, the combustion model
used for this study is described and analyzed. The fourth section presents
1488 C. DUWIG AND L. FUCHS

the industrial burner studied and the operating conditions while the fifth
one describes the numerical methods used. The following section pre-
sents the results of the simulations. First, grid sensitivity is investigated.
Furthermore, the numerical predictions are compared with experimental
data. We study the applicability of the model and the dependence of the
results on in-flow boundary conditions and model parameters. Finally,
the model has been used to study the flame dynamics of the burner.

MODELING OF TURBULENT PREMIXED COMBUSTION


Governing Equations
The basic equations describing the motion of a fluid are the conservation
of momentum, mass, and energy. The system of equations has to be
closed with an equation of state. If the fluid also carries species, one
has to add transport equations for each of the species. When the species
undergo exo- or endothermal chemical reactions, one has to add source=
sink terms to the species transport equations and the energy equation. If
the characteristic speed of the fluid is considerably smaller than the
speed of sound, one may decouple the acoustic effects from the flow,
which simplifies the system of equations. Under this assumption, the
density is a function of temperature alone (i.e., the fluid is assumed to
be ‘‘semicompressible’’). Because the acoustic modes are eliminated
(through this assumption), the short timescales are eliminated. If turbu-
lence is present, the flow is unsteady. Because there is no particular sig-
nificance to an instantaneous turbulent flow field, one seeks statistical
quantities of the dependent variables.
If the basic equations are averaged (independently of the averaging
method), the equations can be written as (Poinsot and Veynante, 2001)
@ q
þ r  ð
qu~Þ ¼ 0 ð1Þ
@t

@ qu~
qu~u~Þ ¼ rP  r  ðquu  qu~u~  lr~
þ r  ð uÞ ð2Þ
@t

@ qY~i  
qu~Y~i Þ ¼ r  quYi  qu~Y~i  qDi rY~i þ x
þ r  ð i ð3Þ
@t

@ qT~  
qu~T~Þ ¼ r  quT  qu~T~  qDT rT~ þ x
þ r  ð T ð4Þ
@t
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1489

 
P
q ¼ ð5Þ
RT

where u is the velocity vector, T the temperature, q the density, l the vis-
cosity, R the specific ideal gas constant, Di the diffusivity of species i, Yi
the mass fraction of species i, and xi the reaction term of species i. The
bar denotes the averaging operation and the tilde the density-weighted
averaging:
qQ
Q~ ¼ ð6Þ
q

In the case of LES, the averaging operation corresponds to spatial


filtering (i.e., applying a low-pass filter that removes all the Fourier compo-
nents that have shorter length scale than the filter size). The averaging=
filtering operator is linear and is assumed commutative with time and space
derivatives. The averaging=filtering operation is not commutative with non-
linear terms. Thus, nonlinear terms lead to expressions that cannot be
expressed in terms of the averaged quantities. These terms are gathered
on the right-hand side in the preceding equations. The SGS terms can be
identified easily in these equations as the difference between the average
of products and the product of averages. Expressing these SGS terms in
terms of averaged quantities is required for ‘‘closing’’ the system. The main
role of the SGS terms is to account for the interaction between the resolved
and the unresolved scales. In the momentum equations, the SGS terms
should account for the dissipative character of turbulence on the small
(unresolved) scales as well as for the transfer of energy among the resolved
and unresolved scales. One may attribute similar SGS properties to the SGS
in the energy- and species-transport equations.
A computational grid can support only Fourier components that
have longer wavelengths than the grid size. Thus, a dependent variable
that is represented on a grid that is used together with a discrete approxi-
mation for the derivatives leads to an implicit filtering. If no explicit SGS
terms are added, then the numerical scheme should at least account for
the small-scale dissipation. This is attained for any numerically stable
scheme. However, too-dissipative numerical schemes are highly inappro-
priate for LES because, in addition to dissipation on small scales, they
may also be dissipative on the larger scale. This effect can be avoided
by choosing appropriate (higher-order) discretizations. In addition, the
spatial resolution has to be fine enough—of the order of magnitude of
1490 C. DUWIG AND L. FUCHS

the Taylor microscale. With such a resolution, the energy transfer among
the scales is dissipation independent and therefore the numerical scheme
may act implicitly as an SGS model. However, one should keep in mind
that this approach requires a better resolution than a corresponding
LES, which uses advanced SGS models. Nevertheless, even with explicit
SGS (such as the dynamic model), the spatial resolution has to be
adequate (usually may be coarser by a factor of at most 2 compared to
the implicit SGS approach).
In addition to the modeling of SGS turbulence (i.e., turbulent trans-
port), an expression for the filtered chemical reaction term is needed. As
pointed out earlier, the filter size that we use is of the order of the Taylor
microscale. During normal operating conditions of gas turbines, chemi-
cal reactions occur at much smaller scales than this scale. Consequently,
combustion requires SGS models. The problem of combustion chemistry
is that a comprehensive mechanism of oxidation involves thousands of
species. It is unrealistic to resolve a thousand additional species trans-
port equations and, therefore, simplifications are needed. Thus, instead
of considering the individual species, we consider a simplification to a
single so-called progress variable, which is related to the temperature
field. The same field can also provide the density field (through the
equation of state).
The present approach is based on the progress variable c; c is inter-
preted as a nondimensional temperature or fuel mass fraction. The range
of the progress variable is 0  c  1. We have c ¼ 0 in the fresh gas and
c ¼ 1 in the burned gas. In the case of nonuniform equivalence ratio in
the device, the progress variable still describes the flame front but is
defined from the local unburned=burned temperature. Temperature (or
the fuel mass fraction) becomes a function of the progress variable and
of the local equivalence ratio. If local variations of the fuel=air equival-
ence ratio are small, the density-filtered c equation becomes

@ qc~  
þ r  ðqu~c~Þ ¼ SGSC þ r  qDth rc þ x
c ð7Þ
@t

SGSC ¼ r  ðquc  qu~c~Þ ð8Þ

This equation is derived directly from the energy equation. One may
observe terms on the right-hand side. The first term is the density-weighted
SGS term (SGSC). The second term is related to the heat diffusivity and the
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1491

 c . The first and the last terms cannot be


last term is the filtered source term x
expressed in terms of the averaged quantities and, hence, are unclosed.
The characteristic length scale of the fuel oxidation is much smaller
than the LES filter size. The filtering will lead to the broadening of the
flame front. In addition, the flame propagation that results from the
reactive-diffusive balance and which is consistent with the SGS transport
should be provided to properly simulate the flame dynamics. In the next
section, the broadening effect of filtering is studied, following Duwig’s
(2003) analysis, and c-equation closure is derived.

SGS MODELING
In the literature one may find several SGS models for the momentum
equations (Gullbrand, 1999). The state of modeling is such that for
wall-free flows, and using adequate spatial resolution, one may get rather
good results for the mean and the root mean square (RMS) of the fluc-
tuations. In fact, as argued earlier, one may skip using an explicit SGS if
the discretization scheme is accurate enough and it is applied on an
adequate spatial resolution. The situation is different for the near-wall
region. In spite of much effort, there is no adequate near-wall SGS model
that can take account for the intermittent processes that take place in
that region. The situation is similar for the transport of scalars (such
as temperature and species). A major issue is, however, the SGS for
the flame-flow interaction (i.e., the reaction rates in the energy and the
species-transport equations). These terms are exponential and there is
little in the way of theory for handling averages of these types of terms.
For oxidation reactions in GT applications, the reactive layer of a
flame is much thinner than the LES filter size (by at least one order of
magnitude). Depending on the combustion regime, the flame’s reactive
layer may be even thinner than Kolmogorov scale. Consequently, we
assume that the thin flame is in the (extended) flamelet regime; that is,
the turbulent eddies are not strong enough to penetrate the reactive layer
of the flame. The limits of the domain of validity of the extended flamelet
regime are not well defined. However, it is believed that most GT oper-
ation conditions are within the flamelet domain. Further discussions
related to premixed flame combustion regimes can be found in the litera-
ture (Poinsot and Veynante, 2001).
To study the filtered flame structure, we consider a modeled 1D
laminar flame (Duwig, 2003). Because the reaction layer is much thinner
1492 C. DUWIG AND L. FUCHS

than the LES filter size, we represent the reaction rate with a Dirac d
function. To have the correct integral reaction rate (i.e., the unburned
density qu multiplied by the laminar flame speed SL in 1D), the Dirac
d function is multiplied by qu SL, leading to
xc ðxÞ ¼ qu SL dðxÞ ð9Þ
Applying a Gaussian filter kernel of size D gives
rffiffiffi  
61 6x 2
xc ðxÞ ¼ qu SL d ¼ qu SL exp  2 ð10Þ
pD D
The filtering operation is equivalent to distributing the reaction rate in a
volume around the flame surface (centered at x ¼ 0) while keeping the
integral reaction rate constant. To obtain the filtered flame structure,
we consider the filtered c equation in 1D with x being a coordinate
normalized by the filter size:
rffiffiffi
d c 1 d 2 c 6  
¼ 2
þ exp 6x 2 ð11Þ
dx a dx p
where a ¼ qu SL D=qD and D is the diffusion coefficient. The equation is
solved using two of the following boundary conditions:

d c d 2 c
cx!1 ! 0 cx!þ1 ! 1 !0 2
!0 ð12Þ
dxx!1 dxx!1

Figure 1 depicts the reaction rate term mapped into the filtered c space.
The nondimensional parameter a determines the filtered flame structure.
It compares the filtered reaction and the diffusion. Written in terms of
length scales, we have
qu SL D q SL D D
a¼  u  ð13Þ
qD qu SL dL dL

where dL is the laminar flame thickness. The parameter a is related to the


ratio between the filtering scale and the flame thickness. For a << 1 the
filtering effect on the flame is negligible, and for a >> 1 the filtering effect
is important and the filtered flame structure no longer depends on the
laminar flame structure. The case a >> 1 recovers the parabola shape
closure of the Bray-Moss-Libby (BML) technique (Boger et al., 1998).
From this 1D filtered flame analysis, it is possible to deduce the filtered
flame structure from the nondimensional number a and to extrapolate it to a
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1493

Figure 1. Production terms PC represented in the density-averaged c space.

three-dimensional (3D) case. Assuming that the filtered flame structure is


known, the source term is directly related to a and to the filtered progress
variable. The c equation reads
qc~
@ q SL D 2 1
qu~c~Þ ¼ u
þ r  ð r c~ þ qu SL Pc ð~
c ; aÞ ð14Þ
@t a D
where Pc ðc; aÞ is obtained from precomputed tables (e.g., Figure 1). The
consistent diffusion term that ensures correct flame propagation speed is
qD ¼ qu SL D=a.
The preceding analysis considers only laminar flames. However, turbu-
lent flames are of major interest. In the thin-flame regime, large turbulent
eddies increase the flame surface by wrinkling the flame front whereas small
eddies lead primarily to the thickening of the preheat zone (Poinsot and Vey-
nante, 2001). The former effect does not modify the filtered flame structure
and is traditionally included using a wrinkling factor N. We have
rffiffiffi
q SL N 6
c ¼ u
x  Pð~
c ; aÞ ð15Þ
D p
1494 C. DUWIG AND L. FUCHS

qu SL ND
qDth ¼ qD ¼ ð16Þ
a
The effect of the small eddies is to modify the flame structure, which is
accounted for through the nondimensional number a. The extension of
the definition of a to turbulent cases is done through a ¼ D=d (Duwig,
2003) with the relation d ¼ dL þ ðN  1ÞdT , where dT denotes the flame
thickness after thickening of the preheat zone. The parameter a sum-
marizes the total thickening, accounting for both the laminar and the tur-
bulent contributions. In the following, we use dT ¼ dL 1:5, assuming that
the local flame thickness is increased by 50% because of turbulent eddies
entering the premixing zone. A rigorous modeling of dT is beyond the
scope of this paper and is left to further investigations.
The equation solved with the filtered momentum and continuity is

qc~
@ q SL DN 2 1
qu~c~Þ ¼ u
þ r  ð r c~ þ qu SL DN Pc ð~
c ; aÞ ð17Þ
@t a D

In the case of nonuniform equivalence ratio in the combustor, the


density qu and the laminar flame speed SL are functions of the local
equivalence ratio Z.
The simplifications made in the preceding paragraphs provide a sim-
ple robust closure ensuring correct flame propagation. However, it is
assumed that the flame structure is similar to an equivalent planar flame.
The flame front resolved by LES is curved and cusps are likely present. It
is clear that the curvature affects the flame structure. Duwig (2003)
notices that the error due to the curvature is relatively low for curvature
radii bigger than the filter size. Consequently, the main curvature effect
should be modeled by the SGS wrinkling. However, the influence of cur-
vature (on resolved or on unresolved levels) on the flame dynamics is
unclear and further investigations are required before stating definitely
that such effects are not affecting the flame dynamic. In addition, some of
the curvature effects are partially accounted for through the diffusion term.
The SGS flame wrinkling is computed using a Damkhöler-like
(1940) correlation N ¼ 1 þ uD =SL , and the unresolved velocity fluctu-
ation uD is assumed to be proportional to the resolved strain rate tensor
norm and the filter length (Flohr and Pitsch, 2000). The evaluation of the
SGS wrinkling through such a correlation is, however, highly question-
able because the expression is purely empirical and is valid for ensemble
or time-averaged (and not spatially averaged) data. This drawback is
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1495

common to most of the LES combustion models. The G equation needs


such an input to model the SGS effects on the propagation rate (Kim
et al., 1999) whereas the thickened flame technique requires a so-called
efficiency function (Colin et al., 2000). Recent work (Charlette et al.,
2002) does improve the SGS wrinkling evaluation and accounts for the
curvature effects. Here, we focus on the effects of large scales on the
flame and on the effect of the in-flow boundary conditions. However,
in the future, the influence of more advanced wrinkling correlations
has to be investigated.
In the present computations, combustion occurs in nonhomoge-
neous and lean fuel=air mixture. To account for the fuel=air mixture fluc-
tuations, the previous formulation has been modified by replacing SL
with SL(Z), where Z is the mixture fraction. The temperature is com-
puted from the progress variable T ¼ Tu(Z) þ (Tb(Z)  Tu(Z))  c. The
density itself is computed from the temperature using the ideal gas law.

Numerical Methods
The present flamelet formulation has been implemented into a Cartesian
finite-difference LES code solving the semicompressible Navier–Stokes
equations with variable density. The code is third-order accurate for
the convective term (Kawamura and Kuwahara, 1984) and fourth-order
accurate for the other terms. The time integration is done by a fully
implicit scheme. The computational grid is composed of a system of
locally refined grids. Local refinements are added gradually in regions
with expected (or computed) large gradients (Gullbrand et al., 2001).
The implicit solver uses a multigrid scheme for enhanced efficiency.
The solver is fast, capable of solving large problems.
In the present computations no explicit SGS models for the momen-
tum and the energy equations are included (referred previously as implicit
SGS model). The time step used is about 4  106 s. The mean values
were computed over 20,000 to 35,000 time steps, except for the fine grid
case, where only 10,000 time steps were used due to higher CPU require-
ments per iteration.
The vortex core has been visualized using criteria based on
the second eigenvalue of the velocity derivatives tensor proposed by
Jeong and Hussain (1995) (the so-called lambda 2 technique). The vortex
core is approximated by the region where the second eigenvalue of the
velocity tensor is negative.
1496 C. DUWIG AND L. FUCHS

Figure 2. Simplified sketch of the burner.

Computational Geometry
Reacting flows in a GT burner have been studied. The geometry, corre-
sponding to an experimental rig, contains a premixing pipe, of diameter
D and length 1.2D, discharging into a rectangular box: the dimensions in
x, y, and z directions are 4D, 3.8D, and 3.8D, respectively. A converging
outlet follows the box. The length of the converging part is 2D and the
outlet area is about four times smaller than the rectangular box cross sec-
tion. Figure 2 depicts the geometry of the combustion chamber. The
coordinate origin is set to be at the pipe exit on its symmetry axis. It
is worth noting the plane symmetry of the geometry. This symmetry is
not utilized in the LESs that are presented here.
The combustion chamber operates in a partially premixed mode.
The inlet fuel=air equivalence ratio varies in the range 0.2–0.8. For the
case considered here, the global equivalence ratio is 0.4. A methane=air
air mixture has been used in both experiments and simulations. The fresh
gas temperature is 680 K and the outlet temperature is about 1650 K.
The Reynolds number based on the pipe diameter is about 92,000, and
the swirl number at the inlet is 0.52.

Boundary Conditions
The boundary conditions are important for LES. Because LES is meant to
resolve the larger structures of the turbulent flow, detailed time-dependent
boundary conditions that have a correct spectral and phase content are
required. Unfortunately, such detailed data are often not available for cases
of practical interest such as GT combustion chambers. For the present
work, only time-averaged data are available. No time-dependent boundary
conditions have been used for the present sensitivity study of the flame
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1497

Figure 3. Inlet boundary profiles: left, normalized radius versus axial velocity; center nor-
malized radius versus tangential velocity; right, normalized radius versus fuel=air equival-
ence ratio.

stabilization, but numerical fluctuations engineered to be close to turbu-


lence are superimposed at the inlet for the comparison with experiments.
To compensate for the uncertainty in the in-flow conditions, the sensitivity
to the inlet conditions has been investigated. Three different types of inlet
velocity profiles have been tested. All are taken from the experimental
results by Bertrams (1996) (denoted fit). Two profiles have been modified
within 5% to test the flow sensitivity to inlet boundaries uncertainty. The
inlet profile is made steeper (denoted þ5%) or flatter (denoted 5%) as
shown in Figure 3. For confidentiality reasons, the velocity scales have
been omitted.
In addition, the sensitivity to the fuel inlet profile has been studied.
Three cases are investigated. The case denoted by H stands for a spatial
homogeneous fuel concentration at the inlet. Figure 3 shows the two
other profiles that are considered: the base case taken from Bertrams
(1996) and case P with a more pronounced central peak. For confidenti-
ality reasons, the velocity scales have been omitted.
It is worth mentioning that all profiles have been assumed to be
axisymmetric. It is probably not the case in reality but no planar
measurements in the premixing tube are available. However, the interac-
tion between the swirl generator and the combustion chamber signifi-
cantly affects the flame stability. To a certain extent, the modification
of the inlet conditions studied in the present paper represents some
effect of the unsteady swirl generation process (modification of the peak
velocity or of the fuel=air mixing efficiency).
At the outlet, the flow is assumed to have a zero gradient. Because
the heat losses at the wall are not known, solid adiabatic walls and non-
slip conditions are imposed. The impact of the adiabatic assumption is
1498 C. DUWIG AND L. FUCHS

expected to be low because we are investigating the flame stabilization in


the central recirculation zone, which is far away from the wall.

RESULTS
Resolution Sensitivity Analysis
The main idea of LES is to resolve important turbulent structures. To
assess the grid influence, three computations have been performed using
a computational grid with 833,000 (coarse), 1.8  106 (medium), and
3  106 (fine) computational cells. Local refinements are used in the
flame region (including the premixing pipe). There are 45 (coarse grid),
60 (medium grid), and 70 (fine grid) grids per diameter of the premixing
pipe. Cubic cells have been used everywhere. Due to too few samples of
results on the finest grid, the RMS values are not statistically fully con-
verged and hence only mean values are compared. These results are
shown in Figure 4.
Results computed on the fine and medium grids agree in terms of the
location of the stagnation point, whereas the coarse grid predicts the
stagnation point farther downstream (Figure 4, left). Also in terms of
the swirl opening angle, the coarse and fine grids agree reasonably well
whereas the medium grid predicts a bigger opening (Figure 4, center).
In addition, some asymmetry is seen in the fine grid result (Figure 4,
left). The reason is that using too-narrow filters sizes will lead to unphy-
sical (resolution-driven) fluctuations. As will be shown later, setting
a ¼ 4 leads to a steep density gradient and turbulence production that
is numerically lower. Consequently, the smoothing effect of turbulence

Figure 4. Influence of the grid on the mean fields: left, normalized axial velocity along the
centerline (U vs. x), center, normalized axial velocity profile 1D downstream of the pipe exit
(U vs. normalized radius); right, temperature profile 0.5D downstream of the pipe exit
(T (K) vs. normalized radius).
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1499

is low and asymmetry is observed. Later in the paper, a sensitivity analy-


sis to the parameter a (or subgrid flame thickness) is presented confirm-
ing the present results.
The medium grid is found to be adequate to conduct sensitivity
analysis of different parameters. The improvements expected by the
use of a finer grid are not significant in this framework.

Comparisons with Experimental Data


The medium-size grid is used to simulate Bertrams’ (1996) experiment.
The geometry of the burner is as explained earlier except that the premix-
ing pipe ends with a longer nozzle. The nozzle penetrates the combustion

Figure 5. Comparisons of normalized axial velocity fields with experiments along three
radial lines located 0.5D, 1D, and 1.5D downstream of the pipe end: left, averaged fields ver-
sus normalized radius; right, RMS fields versus normalized radius; the black line represents
the LES prediction, and symbols denote the experimental data.
1500 C. DUWIG AND L. FUCHS

chamber by a length of  0.1D. Measurements were done using laser


Doppler velocimetry. Axial and tangential velocity components (denoted
U and V, respectively) are compared to predictions. Figures 5 and 6 show
the comparisons on radial lines located at 0.5D, 1D, and 1.5D down-
stream of the premixing pipe end. The overall agreement between the
measurements and the LES predictions is good. The LES captures well
the recirculation where the flame stabilizes as well as the resulting shear
layers. Figure 5 shows that the flame opening angle was underestimated
close to the nozzle but is well captured farther downstream. The discrep-
ancy is observed closer to the in-flow boundary; it might be due to the
imposed in-flow condition. However, the agreement is reasonable and
no data are available to refine the numerical inflow conditions. Figure 5
also shows that the level of turbulence is properly predicted. The magni-
tude of the fluctuations is recovered. The LES data show two peaks in
the inner and outer shear layer, whereas experimental data only show
one peak. Figure 6 presents the comparisons of the predicted tangential
velocity to measurements. As for the axial velocity, the swirl opening
angle is underestimated close to the nozzle but well captured down-
stream. The fluctuation levels are slightly overestimated close to the
burner but correctly captured farther downstream.

Figure 6. Comparisons of normalized tangential velocity fields with experiments along two
radial lines located 0.5D and 1D downstream of the pipe end: left, averaged fields versus
normalized radius; right, RMS fields versus normalized radius; the black line represents
the LES prediction, and symbols denote the experimental data.
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1501

It is worth stressing that the discrepancy between the predictions


and the experiments is rather low compared to the sensitivity to the
in-flow boundary conditions, as will be presented in the following sec-
tions. Globally, the simulations accurately capture the flow field and
the flame stabilization.

Flameback: Effect of the Velocity Peak


To investigate the sensitivity to the inlet velocity profile, the three sets of
velocity profiles are tested keeping the swirl number almost constant
(i.e., 0.52 þ= 0.01); a ¼ 4 has been used for all simulations. These pro-
files do not differ from each other by more than 5%, which is smaller
than experimental uncertainty of such measurements.
Figures 7 and 8 show the mean flow fields. Figure 8 clearly shows
that the location of the stagnation point depends on the inlet velocity
peak. Increasing the velocity peak pushes the stagnation point 0.15D far-
ther downstream while the recirculation is unchanged (Figure 8). Figure 7
shows similar trends. Increasing the velocity peak at the center, the flame
front is also located downstream compared to the base case, because it is
anchored at the stagnation point.
Decreasing the velocity peak at the inlet induces a flameback motion;
the stagnation point moves 0.4D farther upstream (Figures 7 and 8, left).
As the stagnation point enters the premixing pipe, the flame follows it and
the vortex breakdown becomes constrained by the pipe. This changes the
aerodynamics of the combustor. The turbulent flame speed and the recir-
culation velocity increase. The stagnation point is located at the equilib-
rium point between the in-flow conditions and the recirculation. In the
case of 5%, this equilibrium is moved significantly. It is worth noting

Figure 7. Mean temperature (K) field in an axial cut: left, base case; center, case þ5%;
right, case 5%.
1502 C. DUWIG AND L. FUCHS

Figure 8. Mean normalized velocity and temperature (K) profiles: left, along the axis versus
normalized axial coordinate; right, radial profile 1D downstream of the pipe exit versus nor-
malized radius.

that the boundary conditions enforce a steady in-flow=mass flow and


therefore one cannot predict flashback (i.e., when the flame moves inside
the premixing tube up to the swirl generator) but only a limited stagnation
point displacement in the upstream direction (called flameback in the
present paper).
The aerodynamics of the studied configuration and especially the
flame stabilization have been found to be very sensitive to small variation
in the velocity inlet conditions.

Flameback: Effect of Equivalence Ratio


A second sensitivity analysis has been conducted to investigate the
effect of fuel=air mixing on the combustor aerodynamics. Three cases
are studied. First, homogeneous fuel=air mixing is assumed in the
whole device (denoted as case H). Second, the in-flow fuel=air equival-
ence ratio is modified by increasing the fuel concentration at the cen-
terline. The third case starts with the results of the base case. The set of
inlet conditions þ5% (i.e., stagnation point downstream) is used. The
global equivalence ratio is kept constant during the three runs, and
the parameter a is set to 4.
The flow field is initialized with the flow field of the base case. The
inlet conditions are modified to meet the P profile (Figure 3). Figure 9
shows the instantaneous equivalence ratio evolution for four time
instances. On the first frame (Figure 9, panel 1), the stagnation point
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1503

Figure 9. Effect of the fuel=air premixing on the flame stabilization: fuel=air equivalence
ratio is presented, and the flame front is shown by a black dotted line. From 1 to 4: higher
fuel concentration appears in the center of the premixing tube, is convected in the premix-
ing tube until it reaches the stagnation point, and induces an increase of the flame speed
and a flameback into the premixing tube.

is located downstream of the premixing pipe exit. The new fuel=air dis-
tribution affects only the inlet of the premixing tube. On the second
panel, the new fuel=air distribution reaches the stagnation point, the
increase of the flame speed modifies the local equilibrium, and the flame
starts to move backward. The third panel shows this displacement
upstream. The fourth panel shows the end of the movement at the stag-
nation point. It is worth noting that the flame opening has also decreased
as the flame enters the premixing tube. The aerodynamics of the combus-
tor has been changed dramatically as shown in Figure 9 (panel 4); the
stagnation point and the recirculation are totally different in case P.
The flameback phenomenon has been observed experimentally to occur
as the result of such modification (V. Milosaljevic, private communi-
cation, 2003).
The contribution of fluctuation of the flame propagation is similar to
that described in the preceding section and does not contribute to further
understanding of the system. Consequently, these results are not presented.
1504 C. DUWIG AND L. FUCHS

Figure 10. Mean normalized velocity and temperature (K) profiles: left, along the axis ver-
sus normalized axial coordinate; center, radial profile 0.5D downstream of the pipe exit ver-
sus normalized radius; left, radial profile 0.5D downstream of the pipe exit versus
normalized radius.

Comparing the homogeneous case (H) and the base case, Figure 10
(left, center) shows that the stagnation point location and the flame
opening angle are similar. The temperature field differs slightly (Figure 10,
right) but it is due more to the local fuel=air equivalence ratio than to
aerodynamic changes.
From this analysis, we conclude that the swirl stabilization patterns
are mainly determined by the velocity profile as long as the flame is sta-
bilized outside the premixing tube. This includes a broad range of inlet
equivalence ratio but when the inlet conditions impose a too-high equiv-
alence ratio (typically > 0.9) at the center, it leads to a flameback. As
stated in the preceding section, the present in-flow specification does
not allow us to distinguish flameback from flashback.

SGS Flame Thickness and Flame Dynamics


Most of the LES turbulent combustion models consider the subgrid
flame thickness as a way to avoid numerical stiffness and do not address
the problem in a physically consistent manner. In the present section,
two a values are tested, namely a ¼ 4 and a ¼ 1. In addition, a case
where the parameter a is computed from the local flame thickness as
derived in the second section of this paper (i.e., a  2.2–2.8) is also con-
sidered. The strategy is to assess the influence of the parameter a. In all
results, the ‘‘base’’ inlet profile (for velocities and equivalence ratio) is
employed.
In the three cases, the stagnation point is predicted to be at the same
position (Figure 11, left). However, the vortex breakdown pattern differs
significantly depending on the parameter a. Increasing a opens the swirl
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1505

Figure 11. Mean normalized velocity and temperature (K) profiles: left, along the axis ver-
sus normalized axial coordinate; center, radial profile 0.5D downstream of the pipe exit ver-
sus normalized radius; left, radial profile 0.5D downstream of the pipe exit versus
normalized radius.

angle (Figure 11, center and left). This is so because increasing a makes
the flame front steeper and increases the effects of thermal expansion on
the flame stabilization. In addition, increasing a makes the flame less
sensitive to small eddies, explaining a thinner mean flame front (Figure
11, right). This hypothesis is supported by the RMS of the temperature
(Figure 12, right), indicating a broader flame brush when a decreases. In
addition, these fluctuations promote the fuel=air mixing and eliminate
the temperature gradient due to the equivalence ratio distribution
(Figure 11, right).
Figure 12 (left) shows that the turbulent properties along the center-
line are similar in those three cases. This supports the conclusion of
the previous section that the breakdown stabilization is sensitive to the
upstream conditions but less so to the breakdown pattern. Here, the
combustor aerodynamics change significantly (vortex opening angle)

Figure 12. RMS of the normalized velocity and of the temperature: left, along the axis ver-
sus normalized axial coordinate; center, radial profile 1D downstream of the pipe exit versus
normalized radius; left, radial profile 0.5D downstream of the pipe exit versus normalized
radius.
1506 C. DUWIG AND L. FUCHS

Figure 13. Axial planes of the temperature field (left, mean; center, instantaneous) and
the equivalence ratio (right, instantaneous); the arrows show the velocity vector (case a
computed).

but the stabilization occurs at the same location. Figure 12 (center)


shows that the turbulent fluctuations are similar in the case where a is
computed implicitly and when a ¼ 1. However, Figure 12 (left) shows
some differences in the temperature fluctuations.
Figure 13 shows planar cuts of the flame front. Figure 13 (panel 1)
depicts the average temperature field. The flame brush is thick and
smooth. Panels 2 and 3 of Figure 13 show snapshots of the flame. Figure
13 (panel 2) shows that the flame is wrinkled by large turbulent structures;
the flame brush is thinner than in the averages case. Figure 13 (panel 3)
shows that the vortices are wrinkling the flame but at the same time affect
the fuel=air mixing. They generate pockets of rich mixtures leading to
local high-temperature zones. This explains why the maximum tempera-
ture in Figure 13, panel 2, is about 200 K higher than in Figure 13, panel 1.
Figure 14 (right) shows the instantaneous 3D flame front wrinkled
by turbulent vortices. In addition, Figure 14 shows the spiral vortex core
extracted from 3D data (so-called helical mode) responsible for the flame
wrinkling presented previously. The Strouhal numbers (based on the
oscillation frequency, inlet diameter, and mean inlet velocity) associated
with the vortex core movement are 0.9 and 1.8. The vortex core
motion is detected as clear peaks in the azimuthal velocity Fourier spec-
trum (Figure 15, bottom left). The effect is also seen in the axial velocity
spectrum but is weaker. Other peaks have also been observed and are
interpreted as the turbulent crosswise vortices wrinkling the flame.
Extending the analysis frequency to the two other cases (a ¼ 4 and
a ¼ 1), it is found that increasing a decreases the strength of the vortex
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1507

Figure 14. Case a computed: left, instantaneous visualization of the vortex core using the k2
eigenvalue technique; right, instantaneous flame surface (c ¼ 0.7).

core. This effect is also seen in Figure 15; the peak in the spectrum of the
azimuthal frequency is much smaller for a ¼ 4. It has been found that the
turbulence level is also lower in this case. The wrinkling of the flame due
to crosswise vortices is also smaller in the case a ¼ 4 but the frequencies
are in the same range as for the computed a case.
In the case a ¼ 1, the vortex core motion at Strouhal numbers
of  0.9 and  1.8 is also seen but is less clear than in the computed a
case. The effect of the crosswise vortices is more important and they
cover a broader range of frequencies (see Figure 15, left).

Figure 15. Fourier transform of the axial (top) and azimutal (bottom) component of the
velocity, 1D downstream of the premixing pipe exit, 1D from the axis versus the Strouhal
number (frequency Dinlet=Uinlet): left, a computed; center, a ¼ 4; right, a ¼ 1.
1508 C. DUWIG AND L. FUCHS

In all three cases, the important elements of the flame stabilization


(location of the stagnation point and dynamic of the breakdown=of the
vortex core) are found to be similar. However, the subgrid flame thick-
ness is found to have a significant influence on the flame dynamic, on
the turbulent flame-brush thickness, and on the flame position. In
addition, the time derivative of the density (i.e., the source term for ther-
moacoustic instabilities) is very sensitive to the flame dynamic and to the
subgrid flame thickness. The present simulations show the importance of
proper modeling of this thickness to capture the correct flame dynamics.

CONCLUSIONS
In the present paper, a robust and easy-to-implement flamelet model has
been derived and applied to the study of a swirling flame. The SGS dif-
fusion modeling is consistent with the modeling of the reaction rate
regarding the turbulent effects. The problem of SGS flame thickness
has been addressed and an expression for the parameter a has been
proposed.
The model has been used to study the flame stabilization and its
dynamics in a swirl-stabilized combustor. The flame is found to be very
sensitive to variation in several parameters. The case under consider-
ation turned out to be an interesting test case for evaluating models
and assessing parameter sensitivity. The influence of the inlet profile
has been studied by changing the profile obtained from experimental
data by þ=5%. The stagnation point position and the flame dynamics
are found to be very sensitive to the small changes and flameback has
been witnessed. In addition, the influence of the inlet equivalence ratio
has been investigated. Similarly, flameback has been observed when
the fuel=air equivalence ratio at the stagnation point is high enough.
The influence of the SGS thickness on the flame dynamics has been
studied. The key elements of the flame stabilization (location of the stag-
nation point and dynamic of the breakdown=of the vortex core) are
found to be similar. However, the subgrid flame thickness has been found
to have a significant influence on the flame dynamic, on the turbulent
flame brush thickness, and on the flame position. A ‘‘thin’’ SGS flame
is less sensitive to turbulence and damps the fluctuations. In contrast,
a ‘‘thick’’ SGS flame is very sensitive to a broad range of turbulent
frequencies leading to a broad turbulent flame brush and low instan-
taneous heat release. Consequently, the prediction of the flame dynamics
STUDY OF FLAME STABILIZATION IN A SWIRLING COMBUSTOR 1509

(eventually thermoacoustic instabilities) is highly sensitive to the SGS


flame thickness modeling.

REFERENCES
Bertrams, M. (1996) Field Measurements in Turbulent Premixed Flames. M.S.
Thesis=Diplomarbeit, Fachhochschule Aachen & ABB Baden-Dättwil.
Boger, M., Veynante, D., Boughanem, H., and Trouvé, A. (1998) Direct numerical
simulation analysis of flame surface density concept for large eddy simulation
of turbulent premixed combustion. Proc. Combust. Instit., 27, 917–925.
Chakravarthy, V. and Menon, S. (2000) Large-eddy simulation of turbulent
premixed flames in the flamelet regime. Combust. Sci. Technol., 162, 1–50.
Charlette, F., Meneveau, C., and Veynante, D. (2002) A power-law wrinkling
model for LES of premixed turbulent combustion, Part II: Dynamic formu-
lation. Combust. Flame, 131, 181–197.
Colin, O., Ducros, F., Veynante, D., and Poinsot, T. (2000) A thickened flame
model for large eddy simulation of turbulent premixed combustion. Phys.
Fluids, 12, 1843–1863.
Correa, S.M. (1998) Power generation and aeropropulsion gas turbines: From
combustion science to combustion technology. Proc. Combust. Instit., 27,
1793–1807.
Damkhöler, G. (1940) Der Einflub der Turbulenz auf die Flammengeschwindig-
keit in Gasgemischen. Z. Elektrochem., 46, 601–652.
Duwig, C. (2002) A flamelet formulation for large eddy simulation. In Castro,
I.P. and Hancock, P.E. (Eds.) Advances in Turbulence IX, Proceedings of the
Ninth European Turbulence Conference. CIMNE Barcelona, Spain, 801–804.
Duwig, C. (2003) Studies of Gas Turbine Combustion Chambers. Ph.D. Thesis,
Lund Institute of Technology, Lund, Sweden.
Flohr, P. and Pitsch, H. (2000) A Turbulent Flame Speed Closure Model for LES
of Industrial Burner Flows. Proceedings of the Summer Program 2000,
Centre for Turbulence Research, Stanford University, http:==ctr.stanford.
edu=SP00.html (accessed June, 2005).
Gullbrand, J. (1999) Large Eddy Simulation of Turbulent Flows in Combustor
Related Geometries. Ph.D. Thesis, Lund University, Lund, Sweden.
Gullbrand, J., Bai, X.S., and Fuchs, L. (2001). High-order cartesian grid method
for calculation of incompressible turbulent flows. Int. J. Num. Methods
Fluids, 36, 687–709.
Jeong, J. and Hussain, F. (1995) On the identification of a vortex. J. Fluid Mech.,
285, 69–94.
Kawamura, T. and Kuwahara, K. (1984) Computation of High Reynolds Number
Flow Around a Circular Cylinder with Surface Roughness. AIAA paper
84-0340.
1510 C. DUWIG AND L. FUCHS

Kim, W.-W., Menon, S., and Mongia, H.C. (1999) Large-eddy simulation of a gas
turbine combustor flows. Combust. Sci. Technol., 143, 25–62.
Lefevbre, A.H. (1995) The role of fuel preparation in low-emission combustion.
J. Eng. Gas Turb. Power, 117, 617–654.
Lucca-Negro, O. and O’Doherty, T. (2001) Vortex breakdown: A review. Prog.
Energy Combust. Sci., 27, 431–481.
Paschereit, C.O., Gutmark, E., and Weisenstein, W. (1998) Control of thermoacoustic
instabilities and emissions in an industrial-type gas turbine combustor. Proc.
Combust. Instit., 27, 1817–1824.
Poinsot, T. and Veynante, D. (2001) Theoretical and Numerical Combustion, RT
Edwards, Philadelphia, PA.
Stone, C. and Menon, S. (2001) Combustion Instabilities in Swirling Flows.
AIAA paper 2001-3846.

View publication stats

You might also like