Sie sind auf Seite 1von 12

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/274641191

Implementation of a Steady Laminar Flamelet


Model for nonpremixed combustion in LES and
RANS simulations

Conference Paper June 2013

CITATION READS

1 534

3 authors:

Michael Pfitzner Hagen Mller


Universitt der Bundeswehr Mnchen MTU Aero Engines
211 PUBLICATIONS 749 CITATIONS 19 PUBLICATIONS 34 CITATIONS

SEE PROFILE SEE PROFILE

Federica Ferraro
Universitt der Bundeswehr Mnchen
6 PUBLICATIONS 7 CITATIONS

SEE PROFILE

All content following this page was uploaded by Hagen Mller on 30 July 2015.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Ju n e 1 1 - 1 4 , 2 0 1 3
Je j u , Ko re a

Implementation of a Steady Laminar Flamelet Model for


non-premixed combustion in LES and RANS simulations
Hagen Muller 1 , Federica Ferraro1 and Michael Pfitzner1
1 Thermodynamics Institute, Department of Aerospace Engineering,
Bundeswehr University Munich, 85577 Neubiberg, Germany

Abstract

As part of the development of a numerical tool to simulate non-premixed combustion in liquid rocket
engines by means of LES, a steady laminar flamelet model has been implemented in OpenFOAM.
In the current contribution, the implementation details are presented. Furthermore, the code is
used to simulate non-premixed methane/air combustion with both LES and RANS methods. A
comparison with available experimental data shows good agreement and proves the capability of the
new implementation to model turbulent combustion at a reasonable computational effort.

1. INTRODUCTION
The majority of energy conversion technologies, such as diesel engines and gas turbines, are based on
turbulent non-premixed combustion. Naturally, there is great interest in modeling the prevailing physical
phenomena by means of Computational Fluid Dynamics (CFD). Research in this field started off about
sixty years ago and has undergone remarkable improvements. In this course many models of varying
complexity and accuracy have been proposed. An excellent overview of state-of-the-art combustion
modeling and new trends is given in the recent volume edited by Echekki and Mastorakos [1].
An established model for turbulent non-premixed combustion is the flamelet approach which was
mainly advanced by Peters [2]. It is based on the view of a turbulent flame as an ensemble of many laminar
diffusion flames, usually referred to as flamelets. Today several extensions of the flamelet approach exist,
taking into account effects which are neglected in the fundamental formulation. Details can be found in the
books by Peters [3] as well as Poinsot and Veynante [4]. The version presented in the current contribution
is referred to as the steady laminar flamelet model (SLFM) and takes a steady laminar counterflow
diffusion flame as a basis for the flamelet library. Its structure is defined by the scalar dissipation rate
which accounts for strain effects and is connected to the velocity gradients of the turbulent flame. In
terms of computational efficiency, the main advantage of the flamelet approach is the decoupling of the
chemistry structure from the turbulent flow. This is achieved by introducing the passive scalar mixture
fraction which is transported in the CFD and denotes the mass fraction of fuel-generated species in a
mixture. The influence of turbulence on combustion is taken into account using a presumed probability
density function (PDF). As commonly accepted, a -distribution is used for the mixture fraction PDF

Corresponding author: Hagen Muller (hagen.mueller@unibw.de)


and the Dirac-function for the scalar dissipation rate. In this formulation the SLFM has been applied
to several configurations with turbulent non-premixed combustion. The flamelet concept, however, has
mainly been developed in the RANS context where it has shown its capability for instance in the studies
by Zimmermann [5]. Recently, the approach has been extended to be used in Large-Eddy simulations
(LES) as shown by Pitsch [6] and Kempf [7].
The flamelet approach is not yet among the chemistry models available in the OpenFOAM software
package as provided by OpenCFD Ltd. Therefore, Cuoci et al. [8] have recently developed an extension
that provides all steps necessary to use a SLFM with OpenFOAM in the RANS context. Their software
package also contains various PDF shapes and the possibility to model soot formation as well as heat
losses. Our long-term goal, however, is to develop LES methods to simulate combustion in high-thrust
rocket engines which operate at high pressures and cryogenic injection temperatures. In that context,
real-gas thermodynamic effects have to be taken into account when calculating the flamelets as shown by
Pohl et al. [9]. With the flamelet calculation being closed-source in the extension by Cuoci et al. [8] and
the solver limited to RANS simulations, the need to perform real-gas LES motivated the implementation
of a new SLFM OpenFOAM extension. The objective hereby is to consistently include the new libraries
in the existing OpenFOAM structure in order to guarantee full flexibility.
In the next section, the formulation of the implemented SLFM and a brief outline of the code structure
is shown. In addition, the new implementation is validated for two test cases. A 2-dimensional laminar
counter flow diffusion flame is simulated and the results are compared with the according flamelet solution.
In a second step, the well-documented non-premixed methane/air Sandia Flame D, which has been
investigated experimentally by Barlow et al. [10], is simulated using both RANS and LES methods.

2. MODEL FORMULATION
The SLFM concept is based on the assumption that the flame in a turbulent flow can at any time
be regarded as an ensemble of small laminar diffusion flames, generally referred to as flamelets. This
assumption is justified when the Damkohler number is large, i.e. when the chemical reaction zone is
thin compared to the turbulent length scales. Within this limit the SLFM is a strong and widely-used
approach that allows the user to account for finite-rate chemistry effects at a reasonable computational
effort. Its main advantage is that the flamelets, which describe the local structure of the turbulent flame,
are coupled to the turbulent flow by only a few parameters. In practice, this fact is used to calculate the
flamelets independently from the turbulent flow in a pre-processing step and store them in a flamelet
library. Thermodynamic properties and species mass fractions can then be extracted from these tables
using representative parameters which are transported in the turbulent code.
In this work, the flamelets are generated using a counterflow diffusion flame configuration (see figure
1). Oxidizer and fuel flow through opposed nozzles and form a diffusion flame in their midst which, along
its centerline, can be reduced to a one-dimensional problem. The temperature, the species composition
at the nozzles as well as the operating pressure is hereby chosen to match the configuration used in the
turbulent code. This type of flame is typically described using the mixture fraction Z which is a measure
for the amount of fuel atoms in a given mixture. By definition, it is 1 at the fuel and 0 at the oxidizer inlet.
The governing equations for temperature and species mass fractions of the flamelet in mixture fraction
space can then be obtained applying a coordinate transformation, as shown by Peters [2]. For uniform
diffusion (Le = 1) the flamelet equations can be written as

Yk 1 2Yk
k = 0 (1)
t 2 Z2
1 n
 2 
T 1 T 1 cp T
+ + hk k = 0 (2)
t 2 Z2 cp Z Z c p k=1
where is the thermodynamic density, Yk is the chemical species mass fraction, T denotes the temperature,
c p and hk are the specific isobaric heat capacity and the specific enthalpy of species k, respectively.
Figure 1: Schematic view of a counterflow diffusion flame configuration.

The chemical species source terms k are calculated with a chemistry reaction mechanism. The scalar
dissipation rate can be thought of as an inverse diffusion time scale and has the dimension 1/s.
 2
Z
= 2D (3)
y

is a function of the mixture fraction, but can, as shown by Peters [3], be parametrized by its value at
stoichiometric mixture st . Therefore, st acts as an external parameter that imposes strain on the flamelet
and defines its structure. In the current work, the flamelets are calculated with the open-source chemistry
software Cantera [11] where the scalar dissipation rate is adjusted by defining the mass flow at the two
inlets. Starting with a low mass flow, several flamelets can be calculated by incrementally increasing
the mass flow and also the imposed strain until the extinction limit is reached. Once the solution of the
flamelet equations is known for a sufficient number of st , the temperature as well as the species mass
fractions can be tabulated as a function of mixture fraction and scalar dissipation rate at stoichiometry
(T (Z, st ), Yk (Z, st )).
The mean values for species mass fraction and temperature are then calculated using a Favre presumed
probability density function (PDF) to integrate the flamelets.
Z Z 1
Yek = Yk (Z, st )P(Z, st )dZdst (4)
0 0
Z Z 1
Te = T (Z, st )P(Z, st )dZdst (5)
0 0
This step is realized with the new OpenFOAM utility canteraToFoam which at large follows the
procedure described by Echekki and Mastorakos [1]. Hereby, the joint PDF P(Z, st ) is decomposed
assuming statistical independence.
P(Z, st ) = P(Z)P(st ) (6)
The shape of the PDF for the scalar dissipation rate is modeled with a simple Dirac function P(st ) =
(st
fst ). For the mixture fraction, the shape of the PDF is approximated using a presumed -PDF.

( + )
P(Z) = Z 1 (1 Z) 1 (7)
() ( )

is the gamma-function, and are the -PDF parameters defining its shape depending on the mean
002 .
mixture fraction Ze and its variance Zf
!
e Z)
Z(1 e
= Ze 1 (8)
Zf002
Cantera: New OF-utility canteraToFoam:

Solve the Flamelet-equations -PDF integration

Y k (Z , st )
T (Z , st )

Flamelet library: Yk ( Z , st , Z' '2 )


PRE-PRO- hs( Z , st , Z
' ' 2)
CESSING

st , Z
' '2

Z
Yk , Z ,
New OF-solver flameletFoam: st ,
hs
Z ' '2
Pressure-Based RANS/LES CFD solver

CFD-
solver

Figure 2: Schematic of the SLFM procedure as implemented in the current work.

!
  e Z)
Z(1 e
= 1 Ze 1 (9)
Zf002

Large parts of the thermodynamic calculation in OpenFOAM are based on the sensible enthalpy hs , from
which the temperature and other thermodynamic quantities are derived. Thus, in order to be consistent
with the OpenFOAM standard and to be flexible in terms of thermodynamic modeling, the utility
canteraToFoam calculates the enthalpy from the given temperature. The whole set of thermodynamic
models of OpenFOAM is available for this step. After the PDF-integration and the calculation of hs , the
mean values of species mass fraction Yek and enthalpy hes are stored in the flamelet library as function
of the mean values of mixture fraction Z, e mixture fraction variance Zf 002 and scalar dissipation rate at

stoichiometric mixture
fst . These three quantities have to be provided by the turbulent CFD code, such that
the species composition and enthalpy can be obtained from the library by interpolation. This functionality
has been implemented in the new pressure-based OpenFOAM solver flameletFoam for both LES and
RANS simulations. The details of the modeling approach depends on the turbulence closure.
In the RANS context two additional transport equations for the mean mixture fraction and its variance
have to be solved. !
Ze uei Ze Ze
+ = e f f (10)
t xi xi xi
! !2
002
Zf 002
uei Zf Zf002 Ze
+ = e f f + 2e f f e (11)
t xi xi xi xi
Here, the effective viscosity is composed of a laminar and a turbulent contribution (e f f = + t ). The
turbulent or eddy viscosity is calculated with a turbulence model that can be chosen from the variety
of models available in OpenFOAM. Using the assumption that the mixture fraction fluctuations decay
in a manner proportional to the turbulent fluctuations, the scalar dissipation rate in equation 11 can be
expressed as follows
002
e = C Zf (12)
k
where k and denote the turbulent kinetic energy and its dissipation, respectively. The constant C is set
to the value 2.0 as proposed by Janicka and Peters et al. [12].
In the LES context, the filtered transport equation for the mixture fraction formally looks identical
to the Reynolds-averaged transport equation (see equation 10). However, the effective viscosity is now
composed of a resolved contribution and a part which is due to the unresolved subgrid scale (SGS)
fluctuations (e f f = + sgs ). The latter needs modeling and can be described by the SGS turbulence
models available in OpenFOAM. Under the assumption of local equilibrium, the SGS mixture fraction
variance Zf002 and the filtered scalar dissipation rate
e are modeled according to the approach which has
been proposed by Pierce and Moin [13].

Ze 2

002 = C 2
Zf

(13)
Z
xi

Ze 2

e f f
e = C (14)


xi
Moin et al. [14] have proposed a dynamic procedure to determine the constant CZ . However, for the
current implementation a constant value of 1.0 has been found to be sufficient. A further simplification
in the current implementation is made concerning the conditioning of the scalar dissipation rate on
stoichiometric mixture. The flamelets in the library are characterized by st , while only the unconditioned
value can be calculated with equation 12 and equation 14. Among others, Pitsch et al. [15] demonstrated
a model which makes use of a presumed shape PDF to connect the unconditioned scalar dissipation rate
to its value at the point of stoichiometry st . However, it is reported [16] that in practice the approximation
= st is acceptable since the error is small around the stoichiometric mixture, which is the focus of
interest. Furthermore modeling the conditioned scalar dissipation rate introduces additional uncertainties.
For these reasons, the unconditioned scalar dissipation rate is used in the current implementation to extract
the species mass fractions and the sensible enthalpy from the flamelet library. The SLFM procedure as
implemented in the current work is also illustrated in figure 2.

3. VERIFICATION AND VALIDATION


Two test cases have been considered in order to verify the new implementation. In a first step, an
axisymmetric laminar counterflow diffusion flame configuration (see figure 1) has been simulated on a
2-dimensional grid using the new solver flameletFoam. A comparison with the results from the chemical
kinetic software Cantera for the same configuration, allows for a validation of the correct flamelet library
0.25 2500

T
0.2 2000

Temperature [K]
mass fraction
0.15 1500

0.1 1000

O2 CH4
0.05 500

0 0
0 0.25 0.5 0.75 1
Z

Figure 3: Temperature and mass fraction profiles in mixture fraction space along the centerline of a
counterflow diffusion flame: Cantera, flameletFoam.

setup and table interpolation. Furthermore the new solver has been used to simulate a turbulent diffusion
flame (Sandia Flame D) which has been investigated experimentally by Barlow et al. [17]. Both RANS
and LES turbulence closures have been used for the validation.
3.1 Laminar Counterflow Diffusion Flame
As mentioned in the previous section, a laminar counterflow diffusion flame problem can be described
with the one-dimensional flamelet equations (see equation 1 - 2). It is therefore expected that a CFD code
which solves for the full compressible Navier-Stokes equations, reproduces the flamelet solution if the
configuration and modeling is chosen properly. This is particularly true for the SLFM since the flamelet
solution is used to construct the table library which is then used for the interpolation of species mass
fraction and enthalpy. However, for a first verification of the new solver, a comparison of the flamelet
solution as obtained from the chemical kinetics software Cantera with the simulation results of the new
solver is necessary to validate the pre-processing and the interpolation at each time step. The boundary
conditions for the fuel and oxidizer inlets of the tested counterflow diffusion flame have been chosen
to match the boundary conditions of the diffusion flame which is investigated in the next section. This
particular configuration therefore represents a single flamelet of the flamelet library which is needed to
subsequently simulate the turbulent flame. Thus, temperature, pressure and species mass fractions at the
nozzle boundaries have been chosen according to the experimental configuration by Barlow et al. [17]
which is summarized in table 1. The coflow in the experiment corresponds to the oxidizer nozzle. Further
details of the turbulent diffusion flame setup are discussed in the next section. For this first verification,
the effect of strain has not been considered and the velocity at either side is therefore set to a value
small enough to guarantee that strain is negligible (Uin j = 0.025 m/s). The computational domain is a
2-dimensional wedge with inlets at the front as well as at the rear end and an outlet at the upper boundary.
The grid has a uniform distribution in both directions and contains 100 20 cells.
In figure 3 the temperature and the methane (CH4 ) as well as oxygen (O2 ) mass fraction profiles are
shown in mixture fraction space. The results show that the flamelet solution matches the results of the new
solver exactly. This shows that the flamelet library is set up properly and the solver extracts the correct
values at each time step.
Table 1: Boundary conditions for Sandia Flame D.

Fuel Pilot Coflow


U [m/s] 49.6 11.4 0.9
T [K] 294 1888 294
Z 1 0.271 0

3.2 Turbulent Diffusion Flame (Sandia Flame D)


For further validation, a piloted diffusion flame has been chosen (Sandia Flame D). In this configura-
tion, the fuel is a 25%/75% methane-air mixture by volume and is injected through an axisymmetric pipe
with a diameter of D = 7.2 mm. It is enclosed by a pilot nozzle with an outer diameter of DP = 18.2 mm.
The pilot flow is a pre-burned mixture which has been adjusted to match the species composition and
temperature of the main fuel/oxidizer flow with the mixture fraction Z = 0.271. The pilot is surrounded
by a slow (UCF = 0.9 m/s) air coflow. The bulk inflow velocities of fuel and pilot nozzle are U = 49.6 m/s
and UP = 11.4 m/s, respectively. Table 1 summarizes the boundary conditions. The experiments for this
configuration have mainly been performed by Barlow et al. [10, 17, 18]. They have measured the mixture
composition by means of Raman and LIF spectroscopy, the temperature profiles have been obtained by
Rayleigh measurements.
The numerical setup for the configuration differs according to the turbulence closure. For the LES, a 3-
dimensional axisymmetric grid with 4.4 106 grid cells have been used which has been stretched in radial
and axial direction. Turbulence on the subgrid scales has been modeled with the algebraic Smagorinsky
model and a white-noise boundary condition has been used to account for velocity fluctuations at the
fuel inlet. Fewer cells are needed for the RANS-simulation, where a 2-dimensional wedge is sufficient to

Figure 4: Instantanious temperature (upper plot) and mixture fraction (lower plot) LES distribution as
obtained from flameletFoam. The black isolines denote stoichiometric mixture Zst = 0.351
2500 1.2

2000 1

0.8
1500
Te[K ]

0.6

Ze
1000
0.4

500 0.2

0 0
0 20 40 60 80 0 20 40 60 80
x/D x/D

Figure 5: Axial mean temperature (left) and mean mixture fraction (right) profiles: experiments Barlow
et al., flameletFoam LES, flameletFoam RANS

reflect the axisymmetric configuration. In the current work, a grid with 3.8 104 cells has been used to
discretize the domain. Turbulent fluctuations have been modeled with the unmodified - model.
In figure 4, snapshots of the temperature and the mixture fraction in a plane perpendicular to the
centerline are shown. Naturally, the results refer to the LES. The black isolines denote the location of
stoichiometric mixture fraction (Zst = 0.351) and therefore mark the most reactive layer in the flow field.
The snapshots show that in the region just downstream of the nozzle (x/D < 5), the fuel and pilot stream
show hardly any turbulent fluctuations. This short laminar section is followed by a transitional area where
vortical structures evolve on the outer side of the pilot stream and the shear layer between coflow and fuel
starts to wrinkle. Depending on the realization, the flame becomes fully turbulent in the section between
25 < x/D < 45. It is interesting to observe the reaction zone around Zst in the temperature plot. In the
transitional zone, the flame thickness clearly varies, i.e. narrow reaction zones are followed by broad ones
and vice-versa. This is an effect, that has previously been observed by Pitsch and Steiner [19] and can be
attributed to the local scalar dissipation rate. The narrow reaction zones correspond to spots of high local
mixture fraction gradients and therefore scalar dissipation rate (see equation 14). Thus, the flamelets that
are used to represent these regions correspond to counterflow diffusion flames which were generated with
high inlet mass flows. This is also reflected in the decreased maximum temperature in this region.
A quantitative comparison with the measurements is possible for the averaged temperature and mixture
fraction profiles along the flame centerline as well as along the radial axis at different axial positions.
Figure 5 compares the averaged LES results and the RANS simulations with the experiments on the
symmetry axis. The LES results are in excellent agreement with the experiments. The mixture fraction
profile is particularly well reflected in the fully turbulent region of the flame. In the transitional region,
the deviation could be attributed to the velocity boundary condition where white-noise has been used
to reflect the velocity fluctuations. It is, however, known that this method induces fluctuations which
damp too quickly and lead to a delayed jet-breakup. Simulations with more sophisticated methods are
planned. Although the delayed break-up can also be observed in the temperature profile, the transitional
region is captured satisfactorily. Within the limits of the turbulence closure, the RANS results also
match the measurements in good agreement. The shape of the temperature profile is, although slightly
shifted downstream, well reproduced. The offset can be explained by the underestimated turbulence in
the first section of the flame. Jet break-up is shifted downstream and the mixture fraction is therefore
overpredicted.
The same observations can be made in figure 6, where the radial temperature and mixture fraction
profiles at three axial positions are compared. The radial simulation results at the location closest to the
nozzle (x/D = 15) are much the same for the LES and the RANS closure. Both are in excellent agreement
2500 1

x / D = 15 x / D = 15
2000 0.8

1500 0.6
Te[K ]

Ze
1000 0.4

500 0.2

0 0
0 1 2 3 0 1 2 3
2500 0.8

x / D = 30 x / D = 30
2000
0.6

1500
Te[K ]

0.4
1000 Ze
0.2
500

0 0
0 1 2 3 4 0 1 2 3 4
2500 0.5

x / D = 45 x / D = 45
2000 0.4

1500 0.3
Te[K ]

Ze

1000 0.2

500 0.1

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
r/D r/D

Figure 6: Radial mean temperature (left) and mean mixture fraction (right) profiles at three different
axial positions: experiments Barlow et al., flameletFoam LES, flameletFoam RANS

for the mixture fraction, but slightly overestimate the maximum temperature. This can be attributed to an
inherent weakness in the SLFM formulation. It is reported that minor local extinction occurs in this area
of the flame [10] - an effect which can not be reproduced sufficiently with the SLFM. Further downstream,
the difference between RANS and LES becomes more apparent. The deviation between experiment and
RANS simulation in the temperature profiles, can fully be attributed to the discrepancies in the transport
of mixture fraction. The LES reproduces the experiments quite well. The discrepancy in the outer section
of the LES profiles at the location x/D = 45 are very likely due to insufficient averaging time.

4. CONCLUSION
A new OpenFOAM solver for reacting, pressure-based LES and RANS simulations using the steady
laminar flamelet model has been implemented. The flamelets are generated with the open-source chemical
kinetics software Cantera and are subsequently integrated with a new utility which uses a presumed
-shape PDF. The integrated flamelets are arranged in a flamelet library which is accessed by the new
solver to retrieve the species composition and the enthalpy at each time step. This procedure allows to
include finite-rate chemistry in a turbulent CFD simulation at a reasonable computational effort. The new
code has been tested for a laminar counterflow diffusion flame configuration as well as for a turbulent,
jet flame (Sandia Flame D) and reproduced the reference data in excellent agreement. The code can be
downloaded from our website (http://www.unibw.de/thermo/mitarbeiter-en/mueller) and comes with a
manual as well as with a tutorial.

ACKNOWLEDGEMENTS
Financial support has been provided by the German Research Foundation (Deutsche Forschungs-
gemeinschaft DFG) in the framework of the Sonderforschungsbereich Transregio 40. Computational
resources have been provided by the Leibniz-Rechenzentrum Munchen (LRZ).

REFERENCES
[1] T. Echekki and E. Mastorakos. Turbulent Combustion Modeling. Springer, 2011.

[2] N. Peters. Laminar diffusion flamelet models in non-premixed turbulent combustion. Progress in
Energy and Combustion Science, 10(3):319339, 1984.

[3] N. Peters. Turbulent Combustion. Cambridge University Press, 2000.

[4] T. Poinsot and D. Veynante. Theoretical and Numerical Combustion. 3rd edition, 2012.

[5] I. Zimmermann. Modeling and numerical simulation of partially premixed flames. PhD thesis,
University of the Federal Armed Forces Munich, 2009.

[6] H. Pitsch. Large-Eddy Simulation of Turbulent Combustion. Annu. Rev. Fluid Mech., 38:453482,
2006.

[7] A.M. Kempf. Large-Eddy Simulation of Non-Premixed Turbulent Flames. PhD thesis, Technische
Universitat Darmstadt, 2003.

[8] A. Cuoci, A. Frassoldati, T. Faravelli, and E. Ranzi. http://creckmodeling.chem.polimi.it/.

[9] S. Pohl, M. Jarczyk, M. Pfitzner, and B. Rogg. Real gas effects in hydrogen/oxygen counterflow
diffusion flames. 4th European Conference for Aerospace Sciences, 2011.

[10] R.S. Barlow and J.H. Frank. Effects of turbulence on species mass fractions in methane/air jet flames.
Proc. Combust. Inst., 27(1):10871095, 1998.

[11] URL http://cantera.github.io/docs/sphinx/html/index.html.

[12] J. Janicka and N. Peters. Prediction of Turbulent Jet Diffusion Flame Lift-Off Using a PDF Transport
Equation. In Symposium (International) on Combustion, volume 19, Haifa, Israel, 1982.

[13] C.D. Pierce and P. Moin. Progress-variable approach for large-eddy simulation of non-premixed
turbulent combustion. J. Fluid Mech, 504:7397, 2004.

[14] P. Moin, K. Squires, W. Cabot, and S. Lee. A dynamic subgrid-scale model for compressible
turbulence and scalar transport. Phys. Fluids A, 3(11):27462757, 1991.

[15] H. Pitsch, M. Chen, and N. Peters. Unsteady Flamelet Modeling of Turbulent Hydrogen-Air
Diffusion Flames. 21st Symposium (International) on Combustion / The Combustion Institute, pages
10571064, 1998.

[16] Ansys CFX-solver Theory Guide. Ansys CFX Release 11.0, 2006.
[17] R.S. Barlow, J.H. Frank, A.N. Karpetis, and J.-Y. Chen. Piloted methane/air jet flames: Transport
effects and aspects of scalar structure. Combustion and Flame, 143:433449, 2005.

[18] R.S. Barlow and J.H. Frank. URL: http://www.sandia.gov/TNF/DataArch/FlameD.html.

[19] H. Pitsch and H. Steiner. Large-Eddy Simulation of a Turbulent Piloted Methane/Air Diffusion
Flame (Sandia Flame D). Physics of Fluids, 12(10):25412554, 2000.

View publication stats

Das könnte Ihnen auch gefallen