You are on page 1of 205

Abstract

Given the importance of fossil fuel combustion for energy supply, ground and air
transport on one hand and problems like pollution and global warming due to CO2
production on the other hand, an improved understanding of combustion is strongly
desired. Although combustion is known to mankind for millennia there are still many
aspects which are not completely understood. The interaction of complex chemical
reactions, transport phenomena, turbulence and radiation effects make combustion
systems difficult and challenging to simulate. However because of the wide range of
application, the reward of design modifications in combustion systems that can lead
to higher efficiency and pollutant reduction is high.
In technical processes, combustion nearly always takes place within a turbulent
rather than a laminar flow field. To improve a given combustion system, possible effects leading to reduced efficiency and mechanisms forming pollutants must
be identified. This requires the knowledge of the flow and species fields which can
be determined through experiments or numerical simulations. An experimental investigation of a combustor requires expensive hardware to extract all the needed
quantities and in a design phase the system would have to be rebuilt after each
modification. Numerical simulations on the other hand can provide more detailed
information and the effects of modifications in geometry and operating conditions
can be investigated much faster, thus reducing the cost and time for design and
production.
This dissertation focuses on gaseous combustion and uses the Large Eddy Simulation technique (LES) to accurately describe the turbulent flow field at an acceptable
cost. Before applying numerical models to practical combustion systems, they have
to be validated on simplified laboratory flames, which are well investigated experimentally. Both flow field data and scalar data is required to adequately validate the
models used for flow and combustion. In the present work such combustion systems
are being simulated numerically and the quality of the models used is assessed by
comparing with the available experimental data. Flames can be classified depending
on the way the reactants are introduced in the combustion zone with two limiting
cases: perfectly premixed flames and non-premixed flames. In the present work both
types of combustion have been investigated.

ii
In this work first a general introduction of the governing equations describing the
fundamental properties of turbulent reacting flows and the simulation techniques to
model these flows are discussed. Then the modeling of turbulent non-premixed and
premixed combustion is discussed in depth, giving an overview of the state of the
art models described in literature. The numerical methods used and some aspects
of their implementation into the two different numerical codes used in the current
work are described. For the simulation of turbulent non-premixed flames a complete
3D LES code was developed at the department of Mechanical Engineering of the
VUB as well as several tools to generate flame tables and inflow boundary conditions.
For the simulation of premixed turbulent combustion the artificially thickened flame
model was implemented in a research code developed at the TU Darmstadt. Results
of the investigation of two non-premixed flames (the piloted flame Sandia D and the
DLR A jet flame) defined in the TNF-workshop of the Combustion Institute are
discussed as well as a premixed combustor named ORACLES, especially designed
for validation of Large Eddy Simulation of turbulent premixed combustion.

Acknowledgements

The present work was performed at the Department of Mechanical Engineering (former Department of Fluid Mechanics) of the Vrije Universiteit Brussel. The author
gratefully acknowledges the financial support of the Flemisch Research Foundation
(FWO-Vlaanderen) to perform this work over the past 4 years as Aspirant Vorser.
I wish to thank the head of the research group Fluid Mechanics and Thermodynamics, Prof. Dr. Ir. Chris Lacor for the opportunity to perform first my Master
thesis and then my PhD under his supervision. Im particularly grateful for the subject he proposed me -which turned out to be a very interesting research topic- for
the opportunities to work together with foreign research groups and for the freedom
to develop my own ideas and vision.
Secondly I would like to thank the people who I cooperated with during my research on Turbulent Combustion. Prof. Dr. Eduardo Brizuela from the Universidad
Nacional de La Plata for his great assistance in the development of the turbulent
non-premixed combustion modeling. Ted, without your long experience and our
many talks over Skype or through email this work would not have been possible. I
hope that I will have the chance to visit you once in the beautiful and sunny Argentina.
I would also like to thank the people of the Institute for Energy and Power Plant
Technology (FG Energie- und Kraftwerkstechnik : EKT) of the Technical University
Darmstadt for their warm hospitality during my stay at the institute. I thank Prof.
Dr.-Ing. Johannes Janicka, head of the institute, to give me the opportunity to
spend two months in his research group. Also I would like to thank the colleagues
of the institute for the many fruitful discussions and personal talks. Especially I
would like to mention Dr.-Ing. Markus Klein and Ing. Martin Freitag with whom
I cooperated the most. Martin, I appreciate a lot the efforts you put in struggling
through the code and papers together with me. I wish you good luck with your PhD
and your further career.
I would also like to thank my colleagues at the Fluid Mechanics research group:
Mark Brouns, Kris Van den Abeele, Sergey Smirnov, Patryk Widera, Santosh
Jayaraju, Ghader Ghorbaniasl and former colleagues Jan Ramboer and Stephan
Geerts. Mark, thanks a lot for sharing an office with me for the past 4 years, you
iii

iv
have become a good friend and we had a lot of fun together in the department. I
wish you the best with your forthcoming PhD. Sergey, Patrick, Santosh and Ghader,
thanks for the many discussions we had on the codes and on LES, I have learnt that
a strong cooperation and team work is essential in research. Jan, thanks for the
fun we had together and for your straightforward view on research and strength
to see things in perspective. Kris, our time together at the department was short
but pleasant and I hope you continue your work with as much enthusiasm. Also
I want to thank Artur Tyliszczak from the university of Czestochowa, who stayed
at our department in the framework of a bilateral project, for his cooperation in
the development of the non-premixed combustion code and the many discussions
we had on turbulent combustion modeling. Last but not least I would like to thank
Alain Wery for his infinite help with all the computer problems I encountered over
the past 4 years and our secretary Jenny Dhaes for her administrative support.
Finally, my main gratitude is to my family and to my girlfriend Elke. Without
their permanent support this work would not have been possible. Elke, thank you
for the moral support you gave me when I came home after a less successful day at
work and for the way you were able to give me the necessary courage to continue
this work till the end. More than you might expect, a good atmosphere at home and
somebody to share your troubles with, is essential for successful scientific research.

Tim Broeckhoven, Mechelen, November 2006

Contents
1 Introduction
1.1 Turbulent combustion modeling . . . . . . . . . . . . . . . . . . . . .
1.2 Overview of this work . . . . . . . . . . . . . . . . . . . . . . . . . .

1
3
5

2 Conservation Equations for Reacting Flow


7
2.1 The species equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 The energy equation . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3 Description of Turbulent Flows
3.1 Large Eddy Simulation . . . . . . . . . . . .
3.1.1 The Smagorinsky model . . . . . . .
3.1.2 The Dynamic Germano Procedure . .
3.1.3 Modeling Scalar Fluxes . . . . . . . .
3.2 Reynolds Averaged Navier-Stokes Equations

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

4 Turbulent Non-premixed Combustion Modeling


4.1 Conserved Scalars and Mixture Fraction . . . . . . . . . . .
4.2 Flame Structure in Mixture Fraction Space . . . . . . . . . .
4.3 Modeling for Diffusion Flames . . . . . . . . . . . . . . . . .
4.3.1 Infinitely Fast Irreversible Chemistry: Mixed is Burnt
4.3.2 Infinitely Fast Reversible Chemistry : Equilibrium . .
4.3.3 Finite Rate Chemistry : Flamelets . . . . . . . . . .
4.4 Turbulent Diffusion Flames in RANS code . . . . . . . . . .
4.4.1 Infinitely Fast Chemistry . . . . . . . . . . . . . . . .
4.4.2 Finite Rate Chemistry . . . . . . . . . . . . . . . . .
4.4.3 Conditional Moment Closure . . . . . . . . . . . . . .
4.5 Turbulent Diffusion Flames in LES code . . . . . . . . . . .
5 Turbulent Premixed Combustion Modeling
5.1 Laminar premixed flames . . . . . . . . . . . . . . . .
5.2 Effects of turbulence and Premixed Regime diagrams
5.3 Turbulent premixed flame models . . . . . . . . . . .
5.3.1 RANS Models . . . . . . . . . . . . . . . . . .
5.3.2 LES Models . . . . . . . . . . . . . . . . . . .
v

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

13
16
18
19
21
21

.
.
.
.
.
.
.
.
.
.
.

25
26
27
29
31
35
36
41
43
46
47
49

.
.
.
.
.

53
53
57
61
61
69

vi
6 Numerical Methods
6.1 Introduction . . . . . . . . . . . . . . . .
6.2 Non premixed Combustion Solver . . . .
6.2.1 Finite Volume Formulation . . . .
6.2.2 Convective Terms . . . . . . . . .
6.2.3 Diffusive Terms . . . . . . . . . .
6.2.4 Time integration . . . . . . . . .
6.3 Pressure correction method . . . . . . .
6.3.1 Solution of the Poisson Equation
6.3.2 Boundary Conditions . . . . . . .
6.4 Premixed Combustion Solver : FLOWSI
6.5 Generation of Combustion Tables . . . .

CONTENTS

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

77
77
77
77
79
84
86
87
90
96
100
101

7 Results Non-premixed Combustion Modeling


7.1 Piloted non-premixed flame (Sandia) . . . . .
7.1.1 RANS simulations . . . . . . . . . . .
7.1.2 LES simulations . . . . . . . . . . . . .
7.2 Non-premixed jet flames (DLR) . . . . . . . .
7.2.1 RANS simulations . . . . . . . . . . .
7.2.2 LES simulations . . . . . . . . . . . . .
7.3 Conclusions . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

105
. 105
. 107
. 121
. 134
. 136
. 137
. 141

8 Results Premixed Combustion Modeling


8.1 Introduction . . . . . . . . . . . . . . . .
8.2 Oracles Configuration . . . . . . . . . . .
8.3 Results and Discussion . . . . . . . . . .
8.3.1 Non-Reacting Flow . . . . . . . .
8.3.2 Reacting Flow . . . . . . . . . . .
8.3.3 Flame Front Wrinkling . . . . . .
8.4 Conclusions . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

147
147
148
151
151
152
156
158

9 Conclusions

159

Appendices

162

A Derivation of balance equation for mixture fraction variance

163

B Flame structure for 1D laminar strained flame

165

C Spatial discretization in a collocated grid system and the odd-even


decoupling phenomenon
167
D List of Publications

171

Bibliography

173

List of Figures
1.1 1973 and 2004 Fuel Shares of World Total Primary Energy Supply . .
1.2 World Energy-Related CO2 emissions by fuel . . . . . . . . . . . . . .
1.3 Schematical representation of a premixed and non-premixed combustion system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

2
2
5

3.1 Leonardo da Vinci: Old Man with Water Studies (c. 1513), Studies
of Water passing Obstacles and falling (c. 1508) . . . . . . . . . . . . 13
3.2 Sketch of a turbulent energy spectrum . . . . . . . . . . . . . . . . . 16
4.1 Diffusion flame structure . . . . . . . . . . . . . . . . . . . . . . . . .
4.2 Definition of coordinate transformation in mixture fraction space . . .
4.3 Diffusion flame problem : decomposition into two problems, mixing
and flame structure . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.4 Flame structure in z-space for a pure CH4 / air flame with inlet
streams at 293K . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.5 Variation of mass heat capacity for some major species . . . . . . . .
4.6 Comparison of temperature in mixture fraction space for irreversible
and reversible fast chemistry . . . . . . . . . . . . . . . . . . . . . . .
4.7 Comparison of composition between mixed is burnt approach (5 species)
and equilibrium approach (17 species) . . . . . . . . . . . . . . . . . .
4.8 Diffusion flame structure for finite rate chemistry . . . . . . . . . . .
4.9 Flamelet concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.10 CH4 and O2 mass fractions for a pure CH4 /air flame in physical space
and in mixture fraction space . . . . . . . . . . . . . . . . . . . . . .
4.11 Some important radicals for a pure CH4 /air flame in physical space .
4.12 Temperature in mixture fraction space for a pure CH4 /air flame for
the entire strain range and comparison with the infinitely fast chemistry limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.13 Normalized -pdf for different values of a,b . . . . . . . . . . . . . . .
4.14 Presumed pdf - mixed is burnt model in RANS code . . . . . . . . . .
4.15 Presumed pdf-flamelet model in RANS code . . . . . . . . . . . . . .

25
28
29
33
34
36
37
37
38
40
40

41
44
46
48

5.1 Structure of a laminar premixed flame . . . . . . . . . . . . . . . . . 54


5.2 Reduced reaction rate vs , total reaction rate and thus the flame
speed are equal for all flames . . . . . . . . . . . . . . . . . . . . . . . 56
5.3 Flame wrinkling by turbulence and definition of turbulent burning
velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
vii

viii

LIST OF FIGURES
5.4
5.5
5.6
5.7
5.8
5.9
6.1
6.2
6.3

6.4
6.5
6.6
6.7
6.8

Turbulent premixed combustion diagram (Borghi) with lines of constant Reynolds, Damkohler and Karlovitz numbers and different regimes
Turbulent premixed combustion regimes: wrinkled flamelet, thickenedwrinkled flame, thickened flame . . . . . . . . . . . . . . . . . . . . .
2D representation of the kinetic balance of the flame front and the
G-field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Comparison between premixed flame thickness L0 and LES mesh size
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Sub Grid Scale wrinkling of the flame front . . . . . . . . . . . . . . .
-function and efficiency function E with = 0.2 . . . . . . . . . . .
Structured 2D Finite Volume Mesh with Cell Centers and Vertices . .
Limiter functions in the Sweby diagram . . . . . . . . . . . . . . . . .
Transport of a top-hat function with different schemes: Central scheme,
1st order upwind scheme and higher order upwind scheme with different limiters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Transport of a top-hat function with higher order upwind scheme with
different limiters, comparing VAN LEER, CHARM and SMART limiter.
Poisson equation stencil for cartesian mesh . . . . . . . . . . . . . . .
One dimensional restriction operator. . . . . . . . . . . . . . . . . . .
One dimensional prolongation operators . . . . . . . . . . . . . . . .
V-cycle and V-sawtooth multigrid cycles with 4 MG-levels . . . . . .

Sandia Piloted Jet Flame. Configuration and picture of the experimental burner . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.2 Typical mesh and computational domain for RANS simulation of Sandia flame . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.3 Mixed-is-burnt table for Sandia flames: Temperature and species
2 plane . . . . . . . . . . . . . . . . . . . .
mass fractions in the ze-zg
7.4 Comparison of mean axial velocity, turbulent kinetic energy k and its
dissipation at the inlets (fuel jet & pilot) for computed inlet and
inlet modeled from measurements. . . . . . . . . . . . . . . . . . . .
7.5 Infinitely fast chemistry results along symmetry axis for flame D
(CASES 1-5 compared with experimental data) . . . . . . . . . . .
7.6 Evolution of k along the symmetry axis for RANS simulation of
Sandia flame D . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.7 Radial profiles for Infinitely fast chemistry results of flame D (CASES
1-3 compared to experimental data) . . . . . . . . . . . . . . . . . .
7.8 Infinitely fast chemistry results along symmetry axis for flame D
(CASES 1, 2, 6, 7 compared with experimental data) . . . . . . . .
7.9 Influence of Grid refinement on RANS simulation of Sandia flame D
7.10 Temperature field and density field in the mixture fraction z - scalar
dissipation rate plane st as obtained by reordering the counterflow
diffusion flame results from Chem1D-simulations . . . . . . . . . . .
2 ,
g
7.11 Flamelet table for Sandia flames as function of ze,zg
st : temperature, OH mass fraction, CO mass fraction, H2 mass fraction . . . .

60
61
67
69
73
75
79
83

84
85
90
94
94
95

7.1

. 106
. 109
. 110

. 111
. 113
. 113
. 114
. 116
. 116

. 118
. 119

LIST OF FIGURES

ix


2 relating
e and
g
7.12 Mapping function F ze, zg
st . . . . . . . . . . . .
7.13 Finite rate chemistry results along symmetry axis for flame D compared with mixed is burnt results and experimental data . . . . . .
7.14 Comparison of laminar flame structure obtained from mixed is burnt
and flamelet model for the Sandia flame . . . . . . . . . . . . . . .
7.15 Radial profiles for finite rate chemistry results of flame D (Comparison
with mixed is burnt case and experimental data) . . . . . . . . . . .
7.16 Computational domain and mesh for LES of Sandia D flame . . . .
7.17 Instantaneous iso-surface of stoichiometric mixture fraction and axial
velocity, temperature, density and turbulent viscosity iso-contours for
the LES of Sandia flame D . . . . . . . . . . . . . . . . . . . . . . .
7.18 Influence of scalar scheme : Sandia D LES results along symmetry
line, cases 1 and 2 compared with experimental data . . . . . . . .
7.19 Influence of scalar scheme: radial profiles, cases 1 and 2 compared
with experimental data . . . . . . . . . . . . . . . . . . . . . . . . .
7.20 Influence of variance model: Sandia D LES results along symmetry
line, cases 1 and 1b compared with experimental data . . . . . . . .
7.21 Influence of variance model: radial profiles, cases 1 and 1b compared
with experimental data . . . . . . . . . . . . . . . . . . . . . . . . .
7.22 Instantaneous inlet velocity field for Sandia D flame obtained with
digital filtering method . . . . . . . . . . . . . . . . . . . . . . . . .
7.23 Influence of inflow boundary condition: Sandia D LES results along
symmetry line, cases 1, 3 and 4 compared with experimental data .
7.24 Influence of inflow boundary condition: radial profiles, cases 1, 3 and
4 compared with experimental data . . . . . . . . . . . . . . . . . .
7.25 Comparison of instantaneous flames for cases 1, 2 and 3: isolines of
mixture fraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.26 Influence of combustion model : Sandia D LES results along symmetry line, cases 1 and 5 compared with experimental data . . . . . .
7.27 Instantaneous scalar dissipation rate field and iso-line of stoichiometric mixture fraction at x/D = 20 . . . . . . . . . . . . . . . . . . .
7.28 Influence of inflow boundary condition: radial profiles, cases 1 and 5
compared with experimental data . . . . . . . . . . . . . . . . . . .
7.29 The DLR flame operated in the lab . . . . . . . . . . . . . . . . . .
7.30 RANS results along symmetry axis for DLR A: comparison of equilibrium model and flamelet model with experimental data . . . . . .
7.31 RANS results: Radial profiles for flame DLR A (comparison of equilibrium model and flamelet model with experimental data) . . . . .
7.32 Computational domain and mesh for LES of DLR A flame . . . . .
7.33 Instantaneous iso-surface of stoichiometric mixture fraction and axial
velocity, temperature, density and turbulent viscosity iso-contours for
the LES of DLR A flame . . . . . . . . . . . . . . . . . . . . . . . .
7.34 Influence of scalar scheme : DLR A LES results along symmetry line,
cases 1 and 2 compared with experimental dat . . . . . . . . . . . .
7.35 Influence of scalar scheme on DLR A flame: radial profiles, cases 1
and 2 compared with experimental data . . . . . . . . . . . . . . .

. 120
. 120
. 121
. 122
. 123

. 124
. 127
. 128
. 129
. 130
. 130
. 131
. 132
. 132
. 133
. 134
. 135
. 135
. 137
. 138
. 139

. 140
. 141
. 142

LIST OF FIGURES
7.36 Influence of variance on DLR A LES results along symmetry line,
cases 1 and 1b compared with experimental data . . . . . . . . . . .
7.37 Influence of variance on DLR A flame: radial profiles, cases 1 and 1b
compared with experimental data . . . . . . . . . . . . . . . . . . .
7.38 Influence of combustion model: DLR A LES results along symmetry
line, cases 1b and 3 compared with experimental data . . . . . . . .
7.39 Influence of combustion model on DLR A flame: radial profiles, cases
1b and 3 compared with experimental data . . . . . . . . . . . . . .
8.1
8.2
8.3

8.4

8.5

8.6
8.7
8.8

. 143
. 144
. 145
. 145

2D sketch of the ORACLES burner . . . . . . . . . . . . . . . . . . .


Instantaneous streamwise velocity field in the burner upstream of the
sudden expansion) . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Transverse profiles of time averaged stream-wise velocities and fluctuations at different axial positions (non-reacting case): comparison
of LES results (solid) and experimental data (points) . . . . . . . . .
Transverse profiles of time averaged mean stream-wise and transverse
velocities, using a constant efficiency function, Colins Model and the
filtering technique for u0e compared to experimental data . . . . . . .
Transverse profiles of the fluctuations of the stream-wise and the
transverse velocities, using a constant efficiency function, Colins Model
and the filtering technique for u0e compared to experimental data . .
Instantaneous flame front (Ecol case) and zoom of the flame front near
the sudden expansion . . . . . . . . . . . . . . . . . . . . . . . . . . .
Mean flame front position and color contours of the rms values of the
progress variable . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Instantaneous snapshot of the flame front with the level of the efficiency function, mapped on a plane . . . . . . . . . . . . . . . . . . .

148
150

151

153

154
154
155
157

B.1 Steady strained counterflow diffusion flame . . . . . . . . . . . . . . . 165


C.1 Collocated (left) and staggered (right) variable arrangement on a 2D
cartesian grid. Arrows denote velocity components and points the
location of scalar variables like pressure. . . . . . . . . . . . . . . . . 168
C.2 Checkerboard pressure field, made of four superimposed uniform fields
on 2-spacing, which is interpreted by CDS as a uniform field . . . . 169

List of Tables
4.1 Stoichiometric ratio s, equivalence ration and stoichiometric mixture fraction for different fuels (hydrogen H2 , methane CH4 , ethane
C2 H6 , ethene C2 H4 , propane C3 H8 , ethyne C2 H2 and butane C4 H10 )
with air or pure O2 as oxidizer . . . . . . . . . . . . . . . . . . . . . . 31
6.1 Expressions for , Dj (), S for the different RANS and LES equation
to be solved . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.2 Limiter Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
7.1 Sandia Flame Series A-F : main dimensions . . . . . . . . . . . . .
7.2 Sandia Flame Series C-F : jet and pilot bulk velocities and Reynolds
number (based on jet properties) . . . . . . . . . . . . . . . . . . .
7.3 Source term models and parameters for used k models . . . . .
7.4 Different RANS cases for Sandia flame D: k- model, Inlet BC, C2
constant and f2 function . . . . . . . . . . . . . . . . . . . . . . . .
7.5 Different LES cases for Sandia flame D: MIB = mixed is burnt, LAM
= laminar inlet, TURB = turbulent inlet, UPW = first order upwind,
TVD = second order TVD with Van Leer limiter . . . . . . . . . .
7.6 Different LES cases for DLR A flame: EQUI = equilibrium model,
LAM = laminar inlet, UPW = first order upwind, TVD = second
order TVD with Van Leer limiter . . . . . . . . . . . . . . . . . . .

. 105
. 106
. 109
. 115

. 125

. 139

8.1 Relevant physical properties for the ORACLES configuration . . . . . 148

xi

xii

LIST OF TABLES

List of symbols
Latin symbols
a

strain rate, scalar convection speed

progress variable

specific enthalpy

turbulent kinetic energy, wavenumber

reference length

pressure

heat flux

radius, radial position

mass stoichiometric ratio

time

characteristic velocity

displacement speed

mixture fraction

fi

external force component

kG

global reaction rate coefficient

mass flow

pk

partial pressure (species k)

heat source

sL

laminar burning velocity

sT

turbulent burning velocity

ui ,uj

velocity vector component

friction velocity

y+

wall coordinate

xi ,xj

position vector component

~n

normal vector

~u

velocity vector
xiii

xiv

LIST OF SYMBOLS

~x

position vector

pre-exponential constant, area

diffusion coefficient, diameter

energy, efficiency function

thickening factor, fuel

flux vector

Gibbs free energy, filter, field variable

total enthalpy

reaction rate coefficient

length scale, characteristic size of domain

oxidizer

production term, products

heat release

gas constant

entropy, strain

temperature

volume

mixture mean molecular weight

AT

(turbulent) flame surface

Cp

mass heat capacity

CS

Smagorinsky constant

Dk

species binary diffusion coefficient

Dt

turbulent diffusion coefficient

Ea

activation energy

h
IhH ,IH

prolongation and restriction operator

Jk

species diffusive flux

Lij

Leonard stress

Markstein length

Rij

Reynolds stress tensor component

Sij

strain rate tensor component

Ta

activation temperature

Vk

diffusion velocity

Vz

SGS scalar variance

progress rate

xv
Wk

species molecular weight

Xk

species mole fraction

[Xk ]

molar concentration

Yk

species mass fraction

Zk

element mass fraction

Greek symbols

thermal diffusivity, heat release factor

pdf function, activation temperature factor

scalar dissipation rate

delta function, flame thickness

turbulent kinetic energy dissipation

Kolmogorov scale

curvature

thermal conductivity

dynamic viscosity, mean value

kinematic viscosity

equivalence ratio, scalar field

limiter function

density

variance, |G| in G-equation context

heat release factor, time scale

progress variable, non-dimensional temperature

ij

Kronecker symbol

turbulent dynamic viscosity

turbulent kinematic viscosity

0
00
kj ,kj
,kj

species reaction coefficients

species chemical source term

chemical heat release

species partial density

ij

stress tensor component

wall shear stress

filter width, cell size

reaction rate

xvi

LIST OF SYMBOLS

flame surface density

subgrid scale wrinkling

Subscripts, Superscripts
b

burnt

i, j

direction

species

time level

st

stoichiometric

unburnt

fuel

left

right

oxidizer

Other symbols
.

filtered, averaged

e.

favre filtered, favre averaged

b.

h.i
0

. ,.

test filtered
mean
00

fluctuation

|.|

magnitude, partial derivative

.|.

conditional

.
.

partial derivative

nabla-operator (gradient,divergence)

, 2

Laplacian

Abbreviations
ATF

Artificially Thickened Flame

BC

Boundary Condition

BML

Bray-Moss-Libby

CFL

Courant-Friedrich-Lewy

xvii
CMC

Conditional Moment Closure

DNS

Direct Numerical Simulation

EBU

Eddy Break-Up

FDF

Filtered Density Function

GS

Gauss-Seidel

LES

Large Eddy Simulation

LIF

Laser Induced Fluorescence

LDV, LDA

Laser Doppler Velocimetry, Anemometry

MIB

Mixed is Burnt

MG

MultiGrid

PDE

Partial Differential Equation

PDF

Probability Density Function

RANS

Reynolds Averaged Navier-Stokes

RHS

Right Hand Side

RK

Runge-Kutta

SGS

Sub-Grid Scale

TVD

Total Variation Diminishing

Dimensionless numbers
Le =

l0 u 0
0

Sc = D
Da = Ct
Ka = C

Re =

Cp D

Lewis number
Reynolds number
Schmidt number
Damkohler number
Karlovitz number

xviii

LIST OF SYMBOLS

Chapter 1
Introduction
Since the industrial revolution most of the energy used by mankind has been provided by burning fossil fuels. Even in the past decades with increasing contributions
of nuclear power, hydro-power, non-fossil fuels and other alternative energy sources
like wind and solar energy, the combustion of fossil fuels (coal, oil and natural gas)
still provides about 80% of the consumed energy. Considering the world total primary energy supply in figure 1.1 one can see that the total share of fossil fuels has
hardly changed over the past three decades. Moreover the International Energy
Agency [1] predicts that by 2030 two-thirds more energy will be consumed. Rising
economies like China show a big hunger for energy and it is expected that the main
part will be provided by fossil fuels since nuclear power rises several environmental
and safety issues while alternative clean energy sources are unable to provide the
needed quantities of energy. Besides the industrys need for energy also the ground
and air transport (cars, trucks, planes) are big consumers of fossil fuels and only
very limited attempts have been made to switch to other energy sources. Because
of this dependency, countries all over the world have a vital interest in accessing
these fossil fuels leading to energy shortages, destabilization of Western economies
and even military conflicts.
These political and economical issues are not the only side-effects caused by
combustion of hydrocarbon fuels. When burnt they form pollutants endangering
the environment and affecting the climate. The emission of CO2 has an undeniable
effect on global warming with enormous consequences. The World Energy Outlook
2004 [2] predicts that the global energy-related CO2 emissions will increase by 1.7%
per year leading to an increase with 62% by 2030 compared to the level of 2002.
More than two-thirds of this increase will come from developing countries. Figure
1.2 shows the world energy-related CO2 emission over the period 1970-2030 and
the distribution over different fuel types. Other pollutants like nitrogen and sulfur oxides play an important role in the atmospheric reactions that create harmful
particulate matter, ground-level ozone (smog), acid rain and eutrophication. The
presence of these and many other species formed by combustion fossil fuels like VOC
(Volatile Organic Compounds) cost society billions of dollars annually from illnesses
and deaths.

CHAPTER 1. INTRODUCTION

Figure 1.1: 1973 and 2004 Fuel Shares of World Total Primary Energy Supply

Figure 1.2: World Energy-Related CO2 emissions by fuel

1.1. TURBULENT COMBUSTION MODELING

Given the importance of combustion for human society and the problems involved, an improved understanding of combustion is required. Although combustion
is known to man for millennia there are still many issues which are not understood.
The interaction of complex chemical reactions, transport phenomena, turbulence
and radiation effects make combustion systems difficult and challenging to simulate.
However because of the wide application the reward of design modifications in combustion systems that can lead to higher efficiency and reduction of pollutants is high.
The author of this dissertation does not claim to provide results that improve
combustion systems and helps resolving above mentioned problems. However this
work hopes to contribute to a better understanding of turbulent combustion and
to help in a later stage the designers of combustion systems - like gas turbines and
internal combustion engines - to improve efficiency and reduce pollutant formation.

1.1

Turbulent combustion modeling

Combustion processes occur in many practical systems to produce heat (furnaces,


heaters), electricity (thermal power plants) for transport (automotive, aeronautical
and rocket engines) and for waste destruction. It can be defined as one or more
irreversible chemical reaction(s) between fuel and oxidizer and can be represented
by a simplified global scheme as:
F uel + Oxidizer P roducts + heat

(1.1)

Combustion phenomena are characterized by an irreversible strong heat release and


a stiff non linear chemical source term (Arrhenius Law). The heat release occurs in
a thin zone - the typical flame thickness is in the order of 0.1 to 1 mm and causes
strong temperature gradients, density variations and intense heat transfer. Fuels
can be gaseous, liquid or solid and those most common are wood, coal, hydrocarbons (methane CH4 , propane C3 H8 , gasoline, kerosene, ...) and hydrogen (H2 ).
The oxidizer is generally the oxygen (O2 ) contained in air but in special systems
like rocket engines, pure oxygen is used to increase burnt gas temperatures and
avoid storage of inert nitrogen. In technical processes, combustion nearly always
takes place within a turbulent rather than a laminar flow field. The reason for this
is twofold. First, turbulence increases the mixing processes and thereby enhances
combustion because mass and heat transfers are higher. Second, combustion releases
heat and thereby generates flow instability by buoyancy and gas expansion, which
then enhances the transition to turbulence. To improve a given combustion system
possible effects leading to reduced efficiency and mechanisms that form pollutants
must be identified. This requires the knowledge of the flow and species fields which
can be determined through experiments or numerical simulations.
An experimental investigation of a combustor requires expensive hardware to
extract all the needed quantities and in a design phase the system would have to
be built after each modification. Numerical simulations can provide more detailed
information and the effects of design modifications can be investigated much faster

CHAPTER 1. INTRODUCTION

thus reducing the cost and time for production. To simulate turbulent combustion
processes the various coupling mechanisms must be investigated and modeled. To
estimate fuel consumption rate, formation of products and pollutants and to predict
ignition, stabilization and extinction the chemical reaction schemes must be known.
The mass transfer of chemical species by molecular diffusion, convection and turbulent transport must be described as well as the conductive, convective and radiative
heat transfer. In gaseous combustion the single phase flow field has to be determined
but in some applications two phase (liquid fuels) or three phase (solid fuels) systems
may be encountered. Phenomena like spray formation, fuel vaporization, droplet
combustion and the formation of soot particles can make the combustion system
very complicated.

In the current work the focus is on gaseous combustion and several of the above
mentioned effects are neglected or simplified. To describe the turbulent flow field
the Large Eddy Simulation technique (LES) is used which can accurately describe
turbulent flows at an acceptable cost and can be combined with relatively simple
combustion models and still provide a good description of the flow in a combustion
system. Before applying the numerical models to practical combustion systems, they
have to be validated on simplified laboratory flames, which are well investigated experimentally. Both flow field data (velocities, turbulent intensities, turbulent scales)
and scalar data (temperature, species concentration) is required to adequately validate the models used for flow and combustion. In the present work such combustion
systems will be investigated numerically and the quality of the models used will be
assessed by comparing with the available experimental data.

Flames can be classified depending on the way the reactants are introduced in the
combustion zone. The two limiting cases are perfectly premixed and non-premixed
flames. In premixed combustion the reactants (fuel and oxidizer) are perfectly mixed
at the molecular level before entering the reaction zone. Since there is no more mixing required this situation is favorable in terms of burning efficiency but on the other
hand there is a risk that the flame will propagate in the premixed reactants upstream
of the burner and cause damage to the combustor. In non-premixed combustion the
reactants are introduced separately into the combustion chamber and the mixing
of the reactants, controlled by convection, molecular diffusion and turbulent transport, limits the reaction rate. In figure 1.3 both cases are schematically represented.
Because of the different characteristics of these flames, a lot of models to describe
turbulent combustion have been developed for one of these two limiting cases, while
others can describe both and the intermediate situation of partially premixed combustion. In the present work both types of combustion have been investigated using
different models and thus the description of the models and evaluation of the results
will be presented in separate chapters.

1.2. OVERVIEW OF THIS WORK

Figure 1.3: Schematical representation of a premixed (a) and non-premixed (b)


combustion system

1.2

Overview of this work

After this general introduction, chapter 2 will introduce the governing equations describing the fundamental properties of turbulent reacting flows. Further in chapter
3 the nature of turbulent flows will be described and the simulation techniques used
for modeling these flows - focussing on the Large Eddy Simulation technique - will
be discussed.
In chapter 4 the modeling of turbulent non-premixed combustion is discussed
while in the next chapter the same is done for premixed combustion. Chapter 6
presents the numerical methods used for the simulations and some aspects of their
implementation into the code. Two different numerical codes are used in the current
work. For the simulation of turbulent non premixed flames a complete 3D LES code
was developed at the department of Mechanical Engineering of the Vrije Universiteit
Brussel (Fluid Mechanics and Thermodynamics Research Group) as well as several
tools to generate flame tables and inflow boundary conditions. For the simulation
of premixed turbulent combustion the artificially thickened flame model was implemented in a research code developed at the TU Darmstadt. The discretization
in space and the time advancement will be discussed as well as the importance of
adequate boundary conditions. The coupling of the modeled flow and mixing fields
and the modeled combustion will be described as well as some numerical aspects
regarding the generation of the combustion lookup tables. A short description of
the parallelization of the code will close this chapter.
In chapter 7 the results of the investigation of two non-premixed flames defined
in the TNF-workshop [3] will be discussed. First results of a piloted flame known as

CHAPTER 1. INTRODUCTION

Sandia1 D will be presented followed by results from a jet flame known as DLR2
A. Chapter 8 will present the results for the investigation of a premixed combustor
named ORACLES especially designed for validation of Large Eddy Simulation of
turbulent premixed combustion.
Finally chapter 9 will summarize and conclude the work presented in this dissertation and will give some suggestions for further improvement of the current work.

The Combustion Research Facility CRF is part of a larger division at Sandia National Laboratories (California, USA) called the Combustion and Physical Sciences Center. In addition to the
CRFs combustion-related research, it also conducts research on broader energy and homeland security issues, including hydrogen energy and microfluidics research. http://www.ca.sandia.gov/crf/
2
Institute of Combustion Technology (Institut f
ur Verbrennungstechnik) of the DLR (Deutsches
Zentrum f
ur Luft-und Raumfahrt) research institute in Stuttgart, Germany. The institute performs
research on the design of technical burning systems to obtain a decrease of pollutants such as
soot, nitrogen oxides and unburned hydrocarbons, to increase reliability, particularly regarding
igniting, quenching and thermo-acoustic oscillations with intermittent burning and the effective
and environmentally friendly use of new fuels http://www.dlr.de/vt/institut

Chapter 2
Conservation Equations for
Reacting Flow
Flames and other combustion processes are chemically reacting flows, which are
governed by a set of conservation equations applying to the flow, the chemical species
mass fractions and the energy.
The continuity and momentum equations are given by
uj
=0
+
t
xj
ui ui uj
p
ij
+
=
+
+ fi
t
xj
xi
xj

(2.1)
i = 1, 2, 3

(2.2)

, p and ui are respectively the mass density, pressure and the flow velocity components, is the viscous stress tensor and the last term in (2.2) represent external
forces (e.g. gravity).
For a mixture of N chemically reacting species, the conservation equation for the
mass fraction of species k is written:
Jkj
Yk uj Yk
+
=
+ k
t
xj
xj

(2.3)

with Yk = k the mass fraction of species k. The diffusive flux in the first term on
the right hand side is denoted by Jk (Jkj is the j-th component of Jk ) and the last
term k is the chemical source term (the reaction rate of species k).
The energy equation for combustion processes can be written as,
h uj h
p
p
qj
ui
+
=
+ uj

+ q ij
t
xj
t
xj
xj
xj

(2.4)

where h is the specific enthalpy, q is the heat flux and the q the heat source due to
radiation (or another heat source term like an electric spark for ignition). The last
term in (2.4) is the viscous heating source term.
The system is closed with an expression for the stress tensor ij in terms of velocity
gradients and an equation of state.
7

CHAPTER 2. CONSERVATION EQUATIONS FOR REACTING FLOW

For a mixture of N perfect gases, the total pressure is the sum of the partial
pressures:
N
N
X
X
R
p=
pk =
k
T
(2.5)
Wk
k=1
k=1
where T is the temperature, R is the perfect gas constant, k = Yk and Wk are
respectively the density and the molecular weight of species k. Since the density
of the multi-species gas is
=

N
X

k =

k=1

N
X

Yk

(2.6)

k=1

the state equation can be compactly written as


p=

R
T
W

(2.7)

where W is the mean molecular weight of the mixture given by:


N
X
1
Yk
=
W
k=1 Wk

or W =

N
X

Xk W k

(2.8)

(2.9)

k=1

with Xk the species mole fraction.


The viscous stress tensor ij is defined by:


2
ij = 2Sij ij Skk
3
and

1
Sij =
2

ui uj
+
xj
xi

(2.10)

is the rate of strain tensor and is the dynamic viscosity; represents the Kronecker
delta.

2.1

The species equation

The molecular transport processes that cause the diffusive fluxes are quite complicated and should normally involve the full array of diffusion coefficients. Since in
models of turbulent combustion molecular transport is less important than turbulent transport, it is useful to consider simplified versions of the diffusive fluxes.
The species flux Jk can be expressed with a diffusion velocity Vk by
Jkj = Vkj Yk

(2.11)

but it is a very hard task to determine these velocities [4, 5] and generally a simplified
approach based on Ficks law is used:
Jkj = Dk

Yk
xj

(2.12)

2.1. THE SPECIES EQUATION

where Dk is the binary diffusion coefficient, or mass diffusivity, of species k. For the
combustion in air, where nitrogen occurs in large amounts, all other species may
be treated as trace species. In (2.12) the binary diffusion coefficient with respect
to the abundant component N2 may be used as an approximation (Dk = Di,N2 ).
It should be noted that in a multi-component system this approximation violates
mass conservation, if non-equal diffusivities Dk are used, since the sum of all mass
fractions has to be unity. Indeed if one sums equation (2.3) over all species k and
uses

N
X

k = 0 (see further) and (2.6) the global mass conservation is recovered:

k=1

uj
+
=
t
xj
xj
and thus

N
X

N
X

Yk Vkj

k=1

=0

Yk Vkj = 0

(2.13)

(2.14)

k=1

If Ficks law is used, the RHS term of equation (2.13) is not necessarily zero unless
all diffusion coefficients are equal (Dk = D). The equation for the mass fraction of
species k with binary diffusion flux becomes:
Yk uj Yk

+
=
t
xj
xj

Yk
Dk
xj

+ k

(2.15)

In this case the species diffusion is a result of concentration gradients only, which
is termed ordinary diffusion. Real mixtures of interest in combustion contain many
components, not just two. Also gradients of temperature and pressure can produce
species diffusion, i.e., the thermal diffusion (Soret) and pressure diffusion effects,
respectively. For most combustion processes these effects are small and can safely
be neglected.
For simplicity it will also be assumed that all mass diffusivities Dk are proportional
to the thermal diffusivity , such that the Lewis numbers (Lek )
Lek =

=
Dk
Cp Dk

(2.16)

are constant. Here is the thermal conductivity (W m1 K 1 ), Cp the mass heat


capacity (Jkg 1 K 1 ) and the thermal diffusivity (m2 s1 ). Lek is a local quantity,
but in most cases it changes very little from one point of the flame to another. For
all species, except for atomic and molecular hydrogen (because their low molecular
weight), it is assumed that diffusion can be described with a single diffusion coefficient D = , resulting in the approximation of Le = 1. The diffusion coefficient
D can be obtained from the Schmidt number (Sc) giving the ratio of momentum
diffusivity (viscosity) and mass diffusivity
D =

Sc

(2.17)

10

CHAPTER 2. CONSERVATION EQUATIONS FOR REACTING FLOW

Chemistry in reacting flow is brought into the model equations through the
chemical source term k , which is the mass of species k produced per unit volume
and unit time, that appear in (2.15). The overall reaction of one mole fuel with
a moles oxidizer to form b moles of combustion products, can be expressed by the
global reaction mechanism:
(2.18)
F + aO *
) bP
From experimental measurements, the rate at which fuel is consumed can be expressed as
d[XF ]
F =
= kG (T ).[XF ]m [XO ]n
(2.19)
dt
where [Xk ] is used to denote the molar concentration (mole/m3 ) of the k-th species
in the mixture ([Xk ] = Yk /Wk ). The minus sign indicates that the fuel concentration decreases with time. The global rate coefficient kG is in general strongly
depending on temperature. The exponents n and m are related to the reaction order. The reaction is m-th order with respect to the fuel, n-th order with respect to
the oxidizer and (m + n)-th order overall. For global reactions n and m are not necessarily integers and have to be derived from experimental data. The use of global
reactions to express the chemistry in a specific problem is a black box approach.
Although this approach may be useful in solving problems, it does not provide a
basis for understanding what is actually happening chemically in a system. It is
totally unrealistic to believe that a oxidizer molecules simultaneously collide with a
single fuel molecule to form b product molecules, since this would require breaking
several bonds and subsequently forming many new bonds. In reality, many sequential processes, elementary reactions, occur involving many intermediate species. The
collection of elementary reactions necessary to describe an overall reaction, is called
a reaction mechanism.
Consider a set of chemical reactions
N
X

k=1

0
kj
Mk *
)

N
X

k=1

00
kj
Mk

j = 1, . . . , M

(2.20)

with N species and M chemical reactions (k denotes species, j denotes reaction).


00
0
Then with kj = kj
kj
one can write the source term in the species equation as
k =

M
X

kj = Wk

j=1

M
X

j=1

kj Qj

(2.21)

where Wk is the molecular weight of species k (mostly in kg/kmol or g/mol) and


Qj is the rate of progress of reaction j. Because elements are conserved, following
relations are also valid
N
X

kj Wk = 0 j = 1, . . . , M

(2.22)

k=1
N
X

k=1

k = 0

(2.23)

2.2. THE ENERGY EQUATION

11

The rates of progress can generally be written as


Qj = K f j

 0
N 
Y
Yk kj

Kbj

Wk

k=1

 00
N 
Y
Yk kj

Wk

k=1

(2.24)

where the first term contains the forward reaction rate and the second one the
backward reaction rate. If the reaction is a fundamental step, the rate constants
Kf j and Kbj are expressed in units (g/cm3 )mj 1 s1 and (g/cm3 )nj 1 s1 where mj =
N
X

0
kj
and nj =

k=1

N
X

00
kj
are the overall orders of forward and backward reactions.

k=1

The reaction rates are usually expressed using the empirical Arrhenius law
Kf j

Ea
= Af j T exp j
RT
j

(2.25)

where Af j is the pre-exponential constant (also called frequency factor because it


depends on the number of collisions leading to chemical reaction) of the forward
reaction j, j is the temperature exponent which is often set to zero and Eaj is the
activation energy of reaction j.

2.2

The energy equation

The balance equation for the enthalpy is given by (2.4) where h is the specific
enthalpy defined as the sum of the specific enthalpies hk :
h=

N
X

hk Y k

(2.26)

k=1

hk =

h0k

T
T0

Cp,k ()d

(2.27)

where Cp,k is the specific heat capacity and h0k is the specific enthalpy of formation
at temperature T0 of species k which can be determined from tables or polynomials
[6, 7, 8].
The heat capacity at constant pressure of the mixture, is:
Cp =

N
X

Cp,k Yk

(2.28)

k=1

The heat capacity Cp is a function of both temperature T and composition Yk . It


may change significantly from one point to another. However, in most hydrocarbon/air flames, the properties of nitrogen dominate and the mass heat capacity of
the mixture is close to that of nitrogen. Furthermore, this value moderately changes
when the temperature goes from 300K to 3000K, so that Cp is often assumed to be
constant in many theoretical approaches and some combustion codes.
The heat flux vector qj includes the heat diffusion due to a temperature gradient

12

CHAPTER 2. CONSERVATION EQUATIONS FOR REACTING FLOW

(Fouriers Law ) and the effect of enthalpy transport by the diffusion of species fluxes
Jk :
N
X
T
qj =
+
hk Jkj
(2.29)
xj k=1
where is the thermal conductivity. In (2.29) the transport of heat due to gradients
in the concentration of species has been neglected (Dufour effect). The term q in
(2.4) represents heat sources due to radiation and must be retained when radiation
effects are important but is often neglected. Equations (2.12), (2.16), (2.27) and
(2.29) can be inserted in the energy equation (2.4). Neglecting pressure and viscous
effects, because they are small compared to the reaction terms, a simplified energy
equation is obtained:

h uj h
+
=
t
xj
xj

xj

N
X

Cp,k Vkj Yk

k=1

T
+ q
xj

(2.30)

If in addition heat capacity is assumed to be equal for all species Cp,k = Cp (applying
equation (2.14)) and radiative heat transfer can be neglected, the energy equation
contains no source terms. In that case, the equation reduces to an equation for a
conserved scalar:
!
h uj h

h
+
=
(2.31)
t
xj
xj Cp xj
The energy equation can also be written in terms of the temperature:
Cp T
uj Cp T
p

+
=
+
t
xj
t xj

xj

N
X

Cp,k Jkj

k=1

N
X
T

hk k + q (2.32)
xj k=1

If the specific heat capacities Cp,k are all assumed equal and constant Cp,k = Cp (using (2.14)), the pressure is constant and heat transfer due to radiation is neglected,
the temperature equation becomes:
uj T

T
+
=
t
xj
xj

T
Cp xj

+ T

(2.33)

The heat release due to chemical reactions is written as


T =

N
M X
N
1 X
1 X
hk k =
Wk kj Qj hk
Cp k=1
Cp j=1 k=1

where Qj is the progress rate of the reaction j and


from reaction j.

PN

k=1

(2.34)

Wk kj hk is the heat release

Chapter 3
Description of Turbulent Flows
In many flows one can observe complicated, chaotic behavior and although the
fluid particles all obey the governing equations (2.1,2.2,2.4) simple flow patterns
are rarely encountered. Looking at clouds in the sky or water flowing in a river,
scientists have tried to explain this behavior for a long time. Figure 3.1 shows two
drawings of Leonardo da Vinci expressing the search for an explanation for this flow
phenomena.

Figure 3.1: Leonardo da Vinci: (left) Old Man with Water Studies (c. 1513)
(right) Studies of Water passing Obstacles and falling (c. 1508)
These pseudo-random, unsteady, three-dimensional and dissipative flows are
called turbulent flows. This turbulence is inherently included in the Navier-Stokes
equations 1 . The non-linear inertia term ui uj in the momentum equation amplifies
the perturbations while a stabilizing term, the friction or stress term ij inhibits the
1

Named after Claude-Louis Navier ( 1785 - 1836) - a French engineer and physicist who
specialized in mechanics - and Sir George Gabriel Stokes ( 1819 1903) - an Irish mathematician
and physicist, who at Cambridge made important contributions to fluid dynamics, optics, and
mathematical physics (including Stokes theorem).

13

14

CHAPTER 3. DESCRIPTION OF TURBULENT FLOWS

system from diverging. As the stress term is a diffusive term, momentum is transported in the gradient direction and thus the momentum field is smoothed. The
strength of turbulence will thus depend of the ratio of inertial forces to friction
forces. Writing a typical inertia term 0 u0 u0 and friction term 0 0 u0 /l0 with reference density 0 , viscosity 0 and reference velocity u0 and length l0 the following
non-dimensional parameter can be found to describe the intensity of turbulence:
Re

l0 u0
0 u0 u0
u0 =
0 0 l0
0

(3.1)

with Re the Reynolds number 2 . Numerous examples of turbulent flows arise in


aeronautics, hydraulics, chemical engineering, meteorology, astrophysics and biological flows.
In the past, because of the complexity of the governing equations, analytical
solutions were restricted to very simple cases and an experimental approach was the
only method to tackle more complex flow problems. The experimental approach
is mainly based on similitude so that measurements made on one system, which
is generally in the laboratory, can be used to describe the behavior of other systems. Still experimental analysis is a very valuable tool to study turbulent flows
but the rapid growth of computer power 3 and the price reduction of computer
hardware, have made numerical approaches a valuable alternative. Furthermore, in
many situations like hypersonic flows or combustion problems with high temperature, measurements are very difficult or even impossible. In the numerical approach
the governing partial differential equations of the flow are discretized in space and
a time marching method is used to deal with the time derivative term. Three main
procedures are now in use for the computation of turbulent flows, Direct Numerical
Simulation (DNS), Large-Eddy Simulation (LES), and a statistical approach based
on the Reynolds Averaged Navier-Stokes (RANS) equations.
The most accurate approach is DNS where the Navier-Stokes equations are solved
without averaging or approximation other than numerical discretization errors which
can be estimated and controlled. All motions contained in the flow are resolved and
none are modeled. The results of a DNS contain very detailed information about
the flow physics and DNS is used as a research tool for understanding the mechanisms of turbulence production, energy transfer, dissipation, noise production, drag
reduction, and many other physical aspects of turbulence. In a DNS all the kinetic
energy dissipation, which occurs on the smallest scales, has to be captured. Therefore the grid size must not be larger than a viscously determined length scale, called
the Kolmogorov 4 length scale, . The Kolmogorov length scale can be determined
2

Named after Osborne Reynolds( 1842 - 1912) a British fluid dynamics engineer who introduced the number now having his name in 1893
3
In 1965 Moore, co-founder of the Intel company, observed that computer power grows exponentially with a doubling period of 24 months which has been corrected to 18 months in 1970 and
this prediction has proven to be correct over the past 3 decades
4
Named after Andrey Nikolaevich Kolmogorov ( 1903 - 1987) a Soviet mathematician who
made major advances in the fields of probability theory and topology.

15
by assuming that the production of turbulence balances its dissipation. One finds
= ( 3 /)1/4 where (m2 /s) is the kinematic viscosity and (m2 /s3 ) is the dissipation per unit mass. The dissipation per unit mass can be approximated as
u3 /l where u is a characteristic velocity of the flow and l a reference length
scale. If L = nl is the characteristic size of the computational domain, the number
of points in one direction should thus be of the order,
N

nl
L
=

and the number of points for a resolved DNS in three dimensions can be estimated
as,
!3
!9/4
nl
n1/3 uL
= (n1/3 Re)9/4
N

A certain amount of time has to be simulated and thus a certain amount of time
steps is required. Using the stability criterion for convection (Courant-FriedrichLevy number) the allowed time step scales with 1/N and thus the computational
effort can be estimated by
EDN S (n1/3 Re)9/4 (n1/3 Re)3/4 = nRe3
For high Reynolds numbers the number of points enormously increases and obviously
DNS is restricted to relatively low Reynolds numbers.
Large-Eddy Simulation (LES) tries to overcome this problem by directly simulating only the low-frequency modes in space (large eddies) and by modeling the energy
exchange with the high-frequency modes (small eddies). The small eddies which are
not explicitly simulated are modeled using a subgrid scale model. The justification
for such a treatment is that the larger eddies contain most of the energy and are
clearly depending on the application whereas the smaller eddies are more universal
and easier to model. Looking at the turbulent energy spectrum (turbulent kinetic
energy E versus wave number k) one can distinguish three ranges of length scales:
the energy containing range with largest scales the size of which is determined by the
geometry, the dissipation range where the smallest eddies dissipate the energy and
in between the inertial range where mid-sized eddies transfer energy from the large
to the small scales. For a good LES the inertial subrange must be resolved while the
dissipation range can be modeled since turbulence is almost isotropic in the smallest
scales and even simple models work well to account for their contribution.
To separate the large from the small scales, LES is based on the definition of a
filtering operation. The Navier-Stokes equations are filtered and the effect of small
scales appears through a subgrid-scale (SGS) stress term. To close the system, an
expression for the SGS stresses must be obtained. Similarly to DNS, LES provides a
three-dimensional, time dependent solution of the Navier-Stokes equations. Thus, it
still requires fairly fine meshes. LES is thus a valuable tool to study more complex
configurations - closer to those of engineering interest - and at Reynolds numbers
beyond the reach of DNS.
If one is only interested in knowing the averaged quantities, the Reynolds Averaged Navier-Stokes (RANS) approach can be used where all of the unsteadiness is

16

CHAPTER 3. DESCRIPTION OF TURBULENT FLOWS

Figure 3.2: Sketch of a turbulent energy spectrum (log-log representation)


averaged out. If the flow is steady, time averaging is used as a statistical averaging.
Although the small scales are universal, the large ones are affected very strongly
by the geometry and it is difficult to develop models that can represent the global
effect of the wide range of scales correctly. The complexity of turbulence makes it
impossible for a single RANS model to represent all turbulent flows and thus some
adjustment of model parameters is often required. This makes the approach less
suited to be used as a predictive tool since one does not know a priori how these
parameters have to be tuned. On the other hand the approach is less expensive in
terms of computational effort since the solution is no longer time dependent and
many problems can be reduced to 2D configurations.
Other approaches such as Unsteady Reynolds Averaged Navier-Stokes (URANS),
Very Large-Eddy Simulation (VLES), Semi-Deterministic Simulation (SDS), and
Detached Eddy Simulation (DES) can be used and an extensive review can be found
in [9, 10].

3.1

Large Eddy Simulation

The governing equations for LES are obtained after filtering the Navier-Stokes equation. The filtering operation can be written in terms of a convolution integral:
(xj , t) =

Z Z Z

G xj x0j x0j , t dx01 dx02 dx03

(3.2)

Where is a turbulent field, G is some spatial filter and D is the flow domain.
The difference between the filtered field and the original field is called the fine
structure or small scale portion
0 =

= + 0

(3.3)

3.1. LARGE EDDY SIMULATION

17

A typical example for the filter is the so-called top-hat filter




G xj

x0j

3
Y

j=1

1/j for x0j



0
xj

0 for

>

j
2

(3.4)

j
2

In finite volume methods often implicit or Schumann filtering is used which is a


top-hat filter based on the local mesh cell having the filter widths j equal to the
local mesh size xj . Thus the integration interval can be narrowed to xj /2
+
x0j xj /2 and using the notation x
j = xj /2 and xj = xj /2 one can write
3
Y

1
(xj , t) =
k=1 xk

!Z

x+
1
x
1

x+
2
x
2

x+
3
x
3

x0j , t dx01 dx02 dx03

(3.5)

where the product in front of the integral is the inverse of the volume of the cell.
Thus Schumann-filtering reduces the filtering process to volume averaging over a
CFD-cell in a finite volume approach and thus is denoted as implicit filtering since
it is the discretization on the grid itself that acts as a filter. For variable density flow
the Favre averaging is used rather than the standard one. A Favre-filtered variable
is defined as,
e = /
(3.6)
Filtering the Navier-Stokes equations one obtains
uei
=0
+
t
xi

(3.7)

ue i ug
p
ij
i uj
+
=
+
t
xj
xi
xj

(3.8)

Since in many combustion models other forms of the energy equation are used (in
terms of temperature, enthalpy, energy, ...) the filtering of equation (2.4) will not
be discussed here but instead the filtered general transport equation for a conserved
scalar will be used:
e ug

i
+
=
t
xj
xj

D
xj

(3.9)

The momentum equation can be further worked out by approximating the filtered
diffusive term
ui uj
2 uk
ij =
+

ij
xj
xi
3 xk

uei uej 2 uek


e

ij
xj
xi
3 xk

(3.10)

and by defining the subgrid scale stress as

ei u
ej )
ijsgs = (ug
i uj u

(3.11)

The filtered momentum equation then reads


"

uei uej
ue i ue i uej
p

2 uek
e
+
=
+

ij ijsgs
t
xj
xi xj
xj
xi
3 xk

(3.12)

18

CHAPTER 3. DESCRIPTION OF TURBULENT FLOWS

The only remaining term that has to be modeled is the sub-grid stress ijsgs representing the influence of the shear stress due to unresolved turbulent motion on
the resolved velocity field. The only difference with the unfiltered equation is the
addition of the sub-grid stress to the diffusion term. For the scalar equation the
same splitting in a resolved and a sub-grid part can be applied
e + F sgs
ej
ug
j = u
j

(3.13)

and approximating the diffusive term the equation can be written as

e ue i e
+
=
t
xj
xj

e
f + F sgs
D

j
xj

(3.14)

Notice that the molecular correlation term between the velocity and scalar has been
neglected which is in general valid for high Reynolds-number flows.
To model the sub-grid stresses one can assume that the influence of the small
scale structures on the large structures can be described by large-scale features. A
simple class of models are the so-called linear models based on an eddy viscosity
approach using the similarity between the random motion of fluid particles with
the Brownian-motion of molecules and associated molecular viscosity. The most
well-known model of this group was developed by Smagorinksy [11]. The model is
simple and has good numerical properties but tends to overestimate the dissipation of
kinetic energy. Another popular model by Bardina [12] is the scale-similarity model
yielding accurate sub-grid scale stresses but this model does not dissipate sufficient
kinetic energy thus making simulations unstable. Both models can be combined
to form so-called mixed models. Moreover these models contain parameters that
have to be set but the ideal values depend on the configuration and the location
in the flow field. Therefore dynamic procedures have been developed [13, 14] to
determine appropriate model-parameter fields. With this procedure the tunable
parameters vanish and the model becomes self-consistent. Further non-linear models
have been developed like the algebraic stress models [15] and models based on the
renormalization group theory [16]. Another approach to model the sub-grid stresses
ijsgs is to solve a transport equation for the turbulent kinetic energy [17, 18, 19]. A
more complete review of sub-grid scale models is given in [9]. In the present work
the Smagorinsky model and its dynamic counterpart will be used.

3.1.1

The Smagorinsky model

Using an eddy viscosity approach the sub-grid stresses ijsgs is expressed by


1 sgs
1
ijsgs kk
ij = 2t (Seij Semm ij )
3
3

e
u

(3.15)

e
ui
where Seij = 12 ( x
+ xji ) is the large-scale strain rate tensor. In the Smagorinksy
j
model the eddy viscosity is based on a typical length scale CS proportional to the
local cell size and a typical time scale related to the contraction of the deformation
velocity tensor,


(3.16)
t = (CS )2 Se

3.1. LARGE EDDY SIMULATION

19

1/2

. This work will rely


with = (x y z )1/3 the filter width and Se = 2Seij Seij
on an pressure correction scheme to determine the value of pressure so that the
continuity equation is satisfied (see chapter 6). Therefore the pressure parameter P
is introduced as
1 sgs
(3.17)
P = p kk
3
One should be aware that solving the resulting momentum equation will yield the
pressure parameter P and not the physical pressure p
"

uei uej 2 uek


ue i ue i uej
e + t )
=
+
(
+

ij
+
t
xj
xi xj
xj
xi
3 xk

!#

(3.18)

The main drawback of this model is that the model parameter CS varies significantly
in the flow field. For example in a channel flow a value of CS = 0.2 is known to
give reasonable results in the center of the channel while close to the wall CS should
be reduced to 0.06. A near wall damping function by van Driest [20] can be used
y+
(1 e 25 ) so that t vanishes near the solid boundaries. The near wall damping is a
function of the wall coordinate y + = uyw with u = (w /)1/2 the friction velocity.
w is the wall shear stress and yw is the distance to the closest wall. In complex
configurations it is not straightforward to apply this damping. It is thus desirable
to have an automatic determination of the model parameter for each location in the
flow field and this is obtained by dynamic procedures.

3.1.2

The Dynamic Germano Procedure

Originally, this model was developed by Germano et al. [13]. Later, Moin et al. [21]
generalized the model for the large-eddy simulation of variable density flows. Lilly
[14] proposed using a least squares technique to minimize the difference between the
closure assumption and the resolved stresses. The formulation used is based on the
one described by Moin et al. [21] combined with the least square approach of Lilly
[14].
The key element of the dynamic model concept is to use the spectral information
contained in the resolved field. This information is obtained by introducing a test
filter with a larger filter width than the resolved grid filter, which generates a second
field with scales larger than the resolved field. Assume that for a quantity , the
b The width of the test
spatially test-filtered quantity is denoted by a caret as .
b The test-filtered stresses T are defined by direct analogy to
filter is denoted by .
ij
subgrid-scale stresses ij , equation (3.11). Using the standard averaging equation
and omitting the superscript sgs for clarity (3.11) can be written as
ij = ui uj

ui uj

(3.19)

The test-filtered stresses Tij are defined in a similar way as ij ,


d
Tij = u
i uj

dj
di u
u
b

(3.20)

20

CHAPTER 3. DESCRIPTION OF TURBULENT FLOWS

The Leonard stresses5 Lij are given by subtracting the test filtered subgrid-scale
stress (3.19) from Tij
1dd
d
eiu
ej
Lij = Tij c
u
ue i ue j
ij =

(3.21)

and are completely computable from the resolved variables. By applying the test
filter to ij defined in equation (3.15) and substituting expression (3.16) for t , the
test-filtered values c
ij may be written as,
with ij given as,

1
d
d
c
kk ij = 2CS ij
ij
3

(3.22)

1
ij = 2 g
|S|(Seij Semm ij )
3

(3.23)

d
Approximating Cd
S ij = CS
ij which is equivalent to considering that CS is constant
over an interval at least equal to the test filter cutoff length one obtains

1
d
d
c
ij
ij
kk ij = 2CS
3

(3.24)

The same model of (3.15) can also be used for the test field stresses,
1
Tij Tkk ij = 2CS ij
3

(3.25)

with,
1 e
e
g
2 |S|(
ij =
S ij S
mm ij )
3
Now by inserting Tij from (3.25) and c
ij from (3.24) into (3.21) one has,
1
d
Lij Lkk ij = 2CS (
ij ij ) = 2CS Mij
3

(3.26)

(3.27)

with Lkk = Tkk bkk .


Equation (3.27) constitutes a system of five independent equations for CS as
both sides are symmetric tensors and their trace is zero. To determine a single value
for CS , Lilly [14] suggested minimizing the following square error:


1
E = Lij Lkk ij 2CS Mij
3

2

(3.28)

Originally, these stresses were introduced by Leonard [22] in the context of an incompressible
flow. If the subgrid scale velocity u0i = ui u
i is defined, the SGS stresses can be decomposed into
three parts,
i u
j = (
ui + u0i )(
uj + u0j ) u
i u
j = Lij + Cij + Rij
ij = ui uj u
i u
j u
i u
j are the Leonard stresses, Cij = u
i u0j + u0j u
i are the cross terms and
where Lij = u
0
0
Rij = uj ui are the SGS Reynolds stresses.

3.2. REYNOLDS AVERAGED NAVIER-STOKES EQUATIONS

21

Setting the first derivative of (3.28) to CS equal to zero and since the trace of the
tensor Mij is zero, the following expression for the dynamic Smagorinsky model is
obtained:
Lij Mij
CS =
(3.29)
2Mij Mij
Calculating CS from (3.29) no more tuning of model parameters is required. However
this procedure may sometimes yield negative values of CS which is inconsistent
with the Smagorinksy model. Some authors argue that negative values should be
allowed to represent back-scatter. Statistically, energy transfer occurs down the
energy cascade but instantaneously energy may be transferred from small to large
scales. Since the eddy-viscosity approach is based on statistics it should only transfer
energy to the smaller scales. Moreover a negative CS results in a negative t and
possibly in a negative effective viscosity. This destabilizes the numerical simulation
and will result in counter-gradient transport if a gradient-flux approach is used
for scalar transport. The calculated value for CS may also fluctuate strongly in
space and time which may destabilize the simulation. Therefore some averaging
or smoothing is usually used (averaging in homogeneous direction(s), filtering in
space, time averaging the local value (Eulerian) or time averaging along streamlines
(Lagrangian)) and the obtained Germano parameter CS is clipped to positive values.

3.1.3

Modeling Scalar Fluxes

In the filtered scalar equation (3.14) the unknown term Fjsgs must be modeled.
Assuming that turbulence contributes to mixing like additional diffusion, Fjsgs is
modeled with an eddy diffusivity approach similar to the eddy viscosity model for
the subgrid scale stress. Thus Fjsgs is given by the product of a turbulent diffusivity
D,t and the gradient of the filtered scalar
Fjsgs = D,t

e
xj

(3.30)

The diffusion coefficients D and D,t are proportional to the viscosities and t
and are scaled by the Schmidt-Numbers and t describing the ratio of momentum
transport due to viscosity to the scalar transport due to diffusion. The filtered
transport equation is then given by

e uei e
+
=
t
xj
xj

3.2

"

t
+
t

e
xj

(3.31)

Reynolds Averaged Navier-Stokes Equations

While LES is based on spatial filtering to separate small scales and to reduce the
computational cost, RANS models discard the transient behavior of the turbulent
field and only mean values are considered. Such time averaging was already suggested by Reynolds in 1895 and thus Reynolds-averaging was known long before

22

CHAPTER 3. DESCRIPTION OF TURBULENT FLOWS

LES. Time averaging f (xj , t) the mean value f (xj , t) is obtained


1 Z +T /2
f (xj , t) dt
T T T /2

f (xj , t) = lim

(3.32)

The fluctuation is obtained by subtracting the mean f (xj , t) from its instantaneous
value
f 0 (xj , t) = f (xj , t) f (xj , t)
(3.33)
and thus the mean value of the fluctuation is zero f 0 (xj , t) = 0. Again the solution
of the resulting equations can be further simplified by Favre-averaging
fe =

(3.34)

Applying Reynolds averaging to the governing equations one obtains equations very
similar to the filtered equations in LES
fj
u
=0
+
t
xj

(3.35)

fj
u0i u0j
f
ui f
ui u
p
ij
+
=
+

t
xj
xi
xj
xj

(3.36)

f
fj Y

Yfk u
k
+
=
t
xj
xj
e
fj

e u
+
=
t
xj
xj

Yfk
D
xj

e
D
xj

Yk0 u0j
xj

(3.37)

0 u0j
xj

(3.38)

e the favre averaged scalar, is


fi is the favre averaged velocity component,
where u
the mean density and p the mean pressure. The unknown correlations of density,
velocity and scalar emerge from averaging the non-linear convection term. These
correlations can be calculated from their transport equations but any transport
equation for unclosed terms will lead to further unclosed terms. This closure problem
of turbulence can be overcome by modeling the unknown terms based on known
mean quantities. Several classes of models exist:

Algebraic models describing the unknown terms directly e.g. by a mixing


length model like the Prandtl
model [23] linking the turbulent viscosity to the
2 e
velocity gradient: t = lm S with Se the mean strain tensor and lm a mixing
length.
One Equation Models
solvingan additional transport equation for turbulent

kinetic energy k t = C lk k or directly for turbulent viscosity like in the
Spalart-Allmaras model [24].
Two Equation Models solving two transport equations, generally for the turbulent kinetic energy k and its dissipation rate [25, 26, 27]. These are among

3.2. REYNOLDS AVERAGED NAVIER-STOKES EQUATIONS

23

the most popular turbulence models and combine reasonable accuracy with
good numerical properties. The turbulent kinematic viscosity is defined as
t = C f kt

(3.39)

where k is the turbulent kinetic energy, t the turbulent time scale (depending
on k and ), C a constant set to 0.09 and f a wall damping function. The
standard transport equation for the k model are given by

t k
k uj k
=
( + )
+ P
+
t
xj
xj
k xj

(3.40)

uj

t
1
+
=
( + )
+ (C1 f1 P C2 f2 ) + E
t
xj
xj
xj t

(3.41)

where and t are respectively laminar and turbulent dynamic viscosities.


The turbulent time scale t for the standard k model is equal to
 k/, and
uj ui
ui
0.5
in the Yang-Shih model [28] to k/ + (/) . P = t xj + xi xj is the


2

ui
is a term
production term of turbulence kinetic energy and E = t xj x
k
reproducing near wall anisotropy. C1 and C2 are two model parameters and
f1 and f2 two functions that have to be set (e.g. f2 = 1 0.22 exp (Ret /36)
[29]). k and are the Schmidt numbers for k and .

Reynolds-stress models solving transport equations for each component of


the Reynolds-stress tensor u0i u0j . Since these models are able to deal with
anisotropic turbulence they are able to model e.g. swirling flows which are
common in combustion applications. The drawback of the models is the increased computational effort and the not so good convergence properties.

24

CHAPTER 3. DESCRIPTION OF TURBULENT FLOWS

Chapter 4
Turbulent Non-premixed
Combustion Modeling
In many combustion problems the reactants enter the combustion chamber in separate streams, with oxidizer and fuel not mixed and mixing must bring the reactants
into the reaction zone fast enough for combustion to proceed. Figure 4.1 presents
the typical structure of a diffusion flame with two states, the fuel (which can be
diluted with other gases like N2 ) on the left and the oxidizer (diluted or not) on the
right side. Oxidizer and fuel will diffuse towards the reaction zone where they burn
and heat will be released. In this reaction zone the temperature reaches a maximum
and diffuses away from the flame front heating up the fuel and oxidizer streams.
Far away from the flame front the gas mixture is either too lean (oxidizer side) or
too rich (fuel side) to burn and thus chemical reaction can only occur in a limited
region where fuel and oxidizer are adequately mixed. The most favorable mixing is
obtained when fuel and oxidizer are in stoichiometric proportions. The flame front
in figure 4.1 can only persist if fuel and oxidizer streams flow against each other
and reactants are fed at a certain speed. An unstrained flame will spread in time
and gets choked by its own combustion products and the reaction rate will decrease
until extinction occurs. On the other hand, the speed bringing the streams together

Figure 4.1: Diffusion flame structure


25

26

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

cannot be too high, since then the fuel and oxidizer cannot react (blowing a fire will
improve combustion but blowing to hard will extinguish the flame). Compared to
premixed flames, diffusion flames do not have a reference speed (flame speed) nor
a reference thickness (flame thickness). Indeed the flame cannot propagate to the
fuel or oxidizer side and so the reaction zone will not move significantly relative to
the flow field. Because of this, diffusion flames are more sensitive to velocity perturbations and turbulence than premixed flames. An unstretched flame will grow
indefinitely with time while in a stretched flame the thickness will depend essentially
on the stretch value (i.e. the speed of the incoming fuel and oxidizer streams) and
cannot be determined from fluid properties and flame speed like in a premixed flame.
Since mixing is the key characteristic of diffusion flames the solution procedure is
usually decoupled into two problems: the mixing process between fuel and oxidizer
and the combustion process itself.

4.1

Conserved Scalars and Mixture Fraction

To fix thoughts, consider a simple reaction involving only fuel (F ), oxidizer (O) and
products (P ). The single step reaction can then be written as:
F F + O O *
) P P

(4.1)

The mass fraction of each species Yk (F ,O and P ) follows the balance equation
Yk ui Yk

Yk
Dk
+
=
t
xi
xi
xi

+ k

(4.2)

The reaction rates w i are all related to the single-step reaction rate Q by equation
(2.21)
k = Wk k Q
(4.3)
and so the oxidizer reaction rate is linked directly to the fuel reaction rate by
O = s F


(4.4)

WO
where s = OF W
= YYOF
is the mass stoichiometric ratio. Also the reaction rate
F
st
for temperature can be linked to the fuel reaction rate by equation (2.34)

T = Q F

(4.5)

where Q (J/kg) is the heat released by the combustion per mass unit. Using these
relations and assuming the same diffusion coefficient for all species one can rewrite
the balance equations for fuel, oxidizer and temperature:
ui YF

YF
YF
+
=
D
t
xi
xi
xi

+ F

(4.6)

YO ui YO

YO
+
=
D
t
xi
xi
xi

+ s F

(4.7)

4.2. FLAME STRUCTURE IN MIXTURE FRACTION SPACE


T
ui T

+
=
t
xi
xi

T
Cp xi

Q
F
Cp

27
(4.8)

Combining these equations and assuming unity Lewis numbers (Le = 1 or D =


/Cp ) shows that the three quantities:
Z1 = sYF YO

Z2 =

Cp T
+ YF
Q

Z3 = s

Cp T
+ YO
Q

(4.9)

follow the same balance equation without source terms:

Z
Z ui Z
+
=
D
t
xi
xi
xi

(4.10)

Z is a passive or conserved scalar that changes only due to diffusion and convection
and does not depend on the reaction. Of course reaction still indirectly affects Z
through the temperature, density and velocity fields. The three variables Z1 ,Z2 ,Z3
follow the same equation but have different boundary conditions but normalizing
them as
Zj ZjO
(4.11)
zj = F
Zj ZjO

they all have the same boundary conditions zj = 1 in the fuel stream and zj = 0
in the oxidizer stream and one can write the transport equation for the mixture
fraction which expresses the local fuel/oxidizer ratio as
z ui z

z
+
=
D
t
xi
xi
xi

(4.12)

Combining equations (4.11) and (4.9) and applying the boundary conditions for the
passive scalars Zj one obtains
sYF YO + YO,2
z=
=
sYF,1 + YO,2

Cp
Q
Cp
Q

(T TO,2 ) + YF

(TF,1 TO,2 ) + YF,1

sCp
(T TO,2 ) + YO YO,2
Q
sCp
(TF,1 TO,2 ) YO,2
Q

(4.13)
where subscripts 1 and 2 refer to the fuel and oxidizer streams. In literature many
different symbols are used for the mixture fraction (Z,f ,,. . .) but in this work z
will be used.

4.2

Flame Structure in Mixture Fraction Space

Using the mixture fraction, the number of variables can be reduced since equation
(4.13) shows that all species are a function of mixture fraction and temperature. If
it is assumed that the structure of the diffusion flame only depends on the mixture
fraction z and the time t one can write T = T (z, t) and Yk = Yk (z, t) defining the
complete flame structure. Formally this variable change can be seen as a variable
change in the species equation (4.2) from (x1 , x2 , x3 , t) to (z, y1 , y2 , t) where y1 and
y2 are spatial coordinates in planes parallel to iso-z surfaces as shown in figure 4.2.

28

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

Figure 4.2: Definition of coordinate transformation in mixture fraction space


In the equations obtained after this transformation the gradients along the flame
front (in y1 and y2 directions) can be neglected compared to the terms normal to the
flame front surface (z-gradients). Physically this implies that the flame structure is
locally one-dimensional and only depends on time and a flame normal coordinate.
In a general multi-dimensional flow this requires the flame to be thin compared to
flow and wrinkling scales. Each part of the flame front can then be viewed as a
small laminar flame called flamelet. The species mass fraction balance can then be
rewritten as
"

"

ui
z

z
Yk
Yk z
+Yk
+

+ ui
D

+
t
t
xi
z
t
xi xi
xi
|

{z

Continuity

{z

Mixture fraction balance

!#
}

z z
D
xi xi

2 Yk
= k
z 2

(4.14)
and using continuity equation (2.1) and the balance equation for mixture fraction
(4.12) this can be simplified to

1 2 Yk
Yk
= k + 2
t
2
z

(4.15)

where the scalar dissipation rate has been introduced defined as


z z
= 2D
xi xi

(4.16)

The temperature equation can be recast in the same way:

T
1 2T
= T + 2
t
2 z

(4.17)

Equations (4.15) and (4.17) are called the flamelet equations. More complex extended flamelet equations can be derived including effects of non-unity Lewis number, variable molecular weights and radiative heat losses [30, 31, 32, 33]. The only
term depending on spatial variables xi is the scalar dissipation rate which controls

4.3. MODELING FOR DIFFUSION FLAMES

29

Figure 4.3: Diffusion flame problem : decomposition into two problems, mixing and
flame structure
mixing since it determines the gradients of z. Once is known the flamelet equations
can be solved in mixture fraction space to provide the flame structure T = T (z, t)
and Yk = Yk (z, t). Although not explicitly mentioned Yk and T are parameterized
by the scalar dissipation rate. The dimensions of are those of an inverse time
and it can be shown for a steady strained one-dimensional diffusion flame that
is proportional to the strain [34]. The scalar dissipation measures the z-gradients
and thus the molecular fluxes of species towards the flame. If the structure of the
flamelet can be assumed to be steady, even though the flow itself and the z-field
depend on time, steady flamelets are obtained and T = T (z, t) and Yk = Yk (z, t) are
determined by
1 2 Yk
1 2 Tk
k = 2 and T = 2
(4.18)
2
z
2
z
Reaction rates only depend on z and and all flow information (mixing) is incorporated in while chemical effects are described by the flame structure in mixture
fraction space.

4.3

Modeling for Diffusion Flames

Diffusion flame computations can thus be decoupled into two problems as schematically shown in figure 4.3. The mixing problem described by equation (4.12) is
solved to obtain the mixture fraction field as a function of spatial coordinates and
time including all aspects related to the flow (geometry, BC). The flame structure
problem determines the relation between the flame variables (Yk and T ) and the
mixture fraction by solving equations (4.15) and (4.17) and all chemistry information is gathered in this part of the solution. Regarding the speed of the chemistry,

30

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

one can assume fast chemistry or equilibrium when chemical time scales are shorter
than all other flow characteristic times. This simplification implies that equilibrium
is instantaneously reached. If this assumption is not fulfilled, finite rate chemistry
effects have to be taken into account. A second assumption can be made on the reversibility of the chemistry. If reaction (4.1) proceeds only from left to right, i.e. the
reverse equilibrium constant is zero the description of the chemistry can be further
simplified. Combining these assumptions, four cases can be identified:
Infinitely fast and irreversible chemistry (also denoted mixed is burnt): Fuel
and oxidizer cannot coexist and T (z) and Yk (z) do not depend on . The
2
zero reaction rate k in equation (4.18) is obtained because zY2k = 0 and this
corresponds to an infinitely thin flame separating fresh and burnt gases. On
both sides of the flame front, temperature and species profiles are linear versus
z with a discontinuous slope at the flame front.
Infinitely fast reversible chemistry (equilibrium) : Fuel, oxidizer and products
are in equilibrium and are linked by
YFF YOO
= K(T )
YPP

(4.19)

where K(T ) is the reaction equilibrium constant at temperature T . Fuel and


oxidizer can be found at the same place but T (z) and Yk (z) do not depend on
. All molecular fluxes are negligible compared to reaction rates and chemical
reactions take place without mixing and reach a local equilibrium.
Irreversible finite rate chemistry : Fuel and oxidizer can exist simultaneously
and depend on flow time scales, especially on scalar dissipation and thus T (z)
and Yk (z) must be parameterized by .
Reversible finite rate chemistry : No simple model can express the structure of
Yk and T in mixture fraction space and their dependence on scalar dissipation.
In a two-feed system with fuel stream m
1 and oxidizer stream m
2 the mixture fraction
can also be defined as
m
1
z=
(4.20)
m
1+m
2
It is easy to show that both definitions are identical. Starting from equation (2.19)
and (4.4) one can write dYO = sdYF where dYk denotes the change in mass fraction
of species i and integrate this equation from the unburnt state
sYF YO = sYF,u YO,u

(4.21)

The local mass fraction of the fuel is YF,u = z.YF,1 where YF,1 denotes the mass
fraction of the fuel in the fuel stream and similarly YO,u = (1 z).YO,2 where YO,2
denotes the mass fraction of the oxygen in the oxidizer stream (e.g. YO2 ,2 = 0.232
for air). Introducing these relations into (4.20) one obtains
z=

sYF YO + YO,2
sYF,1 + YO,2

(4.22)

4.3. MODELING FOR DIFFUSION FLAMES


Fuel
H2
H2
CH4
CH4
C 2 H6
C 2 H4
C 3 H8
C 2 H2
C4 H10

YF,1
1
1
1
1
1
1
1
1
1

YO,2
1
0.23
1
0.23
0.23
0.23
0.23
0.23
0.23

s
8
8
4
4
3.73
3.43
3.64
3.07
3.59

31

8.0
34.8
4.0
17.4
16.2
14.9
15.8
13.4
15.6

zst
0.111
0.028
0.200
0.055
0.059
0.063
0.059
0.070
0.060

Table 4.1: Stoichiometric ratio s, equivalence ration and stoichiometric mixture


fraction for different fuels (hydrogen H2 , methane CH4 , ethane C2 H6 , ethene C2 H4 ,
propane C3 H8 , ethyne C2 H2 and butane C4 H10 ) with air or pure O2 as oxidizer
which is identical to (4.13). The stoichiometric mixture is defined by sYF = YO so
"

sYF,1
zst = 1 +
YO,2

#1

1
1+

(4.23)

F,1
where = s YO,2
is the equivalence ratio. This equivalence ratio is an important
chemical parameter in diffusion flames and corresponds to the equivalence ratio
obtained when premixing the same mass of fuel and oxidizer streams (see chapter
5). It does not correspond to the overall or global equivalence ratio inside the
burner which also depends on the oxidizer and fuel flow rates. For some pure fuels
(YF,1 = 1) s, and zst are given in table 4.1. For undiluted hydrocarbon fuels
and hydrogen one can notice that the stoichiometric mixture fraction is small and
thus the flame is located close to the oxidizer stream. This also means that the
temperature is strongly changing from the cold oxidizer stream temperature to the
flame temperature over a small z-range and density will drop significantly over this
small interval.

4.3.1

Infinitely Fast Irreversible Chemistry: Mixed is Burnt

The most simple approach to determine the flame structure is to assume infinitely
fast irreversible chemistry. The species mass fractions and temperature are then
piecewise linear functions of the mixture fraction and fuel and oxidizer cannot coexist. To fix thoughts consider the global reaction for the complete combustion of a
hydrocarbon fuel Cm Hn


C m Hn + m +

n
4

O2 m CO2 +

n
H2 O
2

(4.24)

If z < zst , fuel is deficient and the mixture is called fuel lean. Combustion will then
terminate when all fuel is consumed, and the mass fraction of fuel is zero in the
burnt gas. The remaining oxygen mass fraction in the burnt gas is then calculated

32

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

from (4.22) and (4.23)


YF,b =
YO2 ,b =

0
1

z
zst

YO2 ,2 if z zst

(4.25)

Similarly if z > zst , oxygen is deficient and the mixture is called fuel rich. Combustion will then terminate when all oxygen is consumed
YO2 ,b =
YF,b =

0
zzst
1zst

YF,1 if z zst

(4.26)

For a hydrocarbon fuel Cm Hn the element mass fractions in the unburnt mixture
are
WC
ZC = m W
YF,u
F
H
ZH = n W
Y
WF F,u
ZO =
YO2,u

(4.27)

since C is only present in the fuel and O is only present in the oxidizer. Because the
elements are conserved during combustion one can also determine the mass fractions
of the products from the element mass fractions in the burnt mixture
WC
mW
YF,b +
F

ZC =

WC
Y
WCO2 CO2 ,b

H
nW
Y + 2 WWHHO YH2 O,b
WF F,b

ZH =

(4.28)

WO
O
ZO = 2 W
YO2 ,b + 2 WWCO
YCO2 ,b +
O
2

WO
Y
WH2 O H2 O,b

Combining equations (4.27) and (4.28) and introducing the expressions for YO2 ,b
and YF,b from (4.25) and (4.26) the piecewise linear relations of the products mass
fractions in terms of z are obtained.
If z zst :
z
z
YH2 O,b = YH2 O,st
(4.29)
YCO2 ,b = YCO2 ,st
zst
zst
If z zst :
1z
1z
YCO2 ,b = YCO2 ,st
YH2 O,b = YH2 O,st
(4.30)
1 zst
1 zst
where
mWCO2
nWH2 O
YCO2 ,st = YF,1 zst
YH2 O,st = YF,1 zst
(4.31)
WF
2WF
The so-called Burke-Schumann [35] profiles of fuel, oxidizer and the combustion
products are shown in figure 4.4. Substituting the expressions for YF (z) and YO (z)
in the definition of mixture fraction (4.13) based on the conserved scalars Z2 or Z3
the flame structure for T is obtained
T (z) =

QYF,1
z
Cp

z < zst

QYF,1
1z
zst 1z
Cp
st

z > zst

zTF,1 + (1 z)TO,2 +

T (z) = zTF,1 + (1 z)TO,2 +

(4.32)

4.3. MODELING FOR DIFFUSION FLAMES

33

Figure 4.4: Flame structure in z-space for a pure CH4 / air flame with inlet streams
at 293K (a) Profiles of species mass fractions Yi (b) Temperature profile with
constant and variable Cp (c) Mass heat capacity for the obtained mixture and
temperature Cp (Yi , T ) (d) Molecular weight of the mixture
The maximum value of temperature occurs at z = zst and is called the adiabatic
flame temperature
Tad = zst TF,1 + (1 zst ) TO,2 +
|

{z

M ixing

QYF,1
zst
Cp

{z

Reaction

(4.33)

The first part of the expression is given by the pure mixing of fuel and oxidizer stream
while the second part is the temperature increase caused by burning a fuel amount
zst YF,1 to equilibrium at constant pressure. In figure 4.4b the temperature profile is
given for a pure CH4 /air flame with a heat release per unit mass, Q = 50100kJ/kg
and a mass heat capacity, Cp = 1400J/(kgK). Since Cp is significantly changing
with temperature and is depending on the mixture compositions, the temperature
from equation (4.32) will differ from the real one. To take this variable Cp into
account one can determine the enthalpy of the mixture. The enthalpy is also a
conserved scalar if the flow is adiabatic and radiation is neglected. Enthalpy is then
linearly related to the mixture fraction
h(z) = zhF,1 + (1 z) hO,2

(4.34)

The enthalpy is related to the temperature through


h(T ) =

T
Tref

Cp (T )dT =

T
Tref

a1 + a 2 T + a 3 T 2 + a 4 T 3 + a 5 T 4

R dT

34

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

Figure 4.5: Variation of mass heat capacity for some major species

T3
T4
T5
T2
+ a3
+ a4
+ a5
+ a6
a1 T + a 2
2
3
4
5

(4.35)

where ak are polynomial coefficients known from a thermodynamical database [8] for
all species Yi in a low and high temperature range. The heat capacity of the mixture
P
is the given by Cp = N
the same way the polynomial coefficients
i=1 Cpi Yi and in P
for the mixture can be calculated ak = K
i=1 aki Yi for k = 1, 6. Figure 4.5 shows
the species heat capacities as a function of temperature and one can see that they
differ significantly. For a given mixture the coefficients ak are known and h can
be calculated for 6 different temperatures in the considered range. A system of
equations can then be solved to determine the coefficients relating T to h
T = b 1 + b 2 h + b 3 h2 + b 4 h3 + b 5 h4 + b 6 h5

(4.36)

and thus from h(z) given by (4.34) the flame structure T (z) can be determined.
Figures 4.4 (c) and (d) show the heat capacity and molecular weight of the mixture
and figure 4.4 (b) shows the temperature obtained by taking into account the variable
Cp . One can observe that the difference with the constant Cp solution is small at
the lean side but is significant at the rich side due to the much higher heat capacity
of methane. The heat capacity of the mixture increases with z, first because of
the temperature increase and above zst because of the increase of methane in the
mixture. At high z, Cp decreases again because the effect of temperature decrease
is stronger than the effect of increasing YCH4 .
The mixed is burnt approach provides a relatively simple flame structure that
provides values for T and from the mixture fraction z needed by the flow solver.
Only major species concentrations are known since simple one step reactions have

4.3. MODELING FOR DIFFUSION FLAMES

35

to be considered to obtain the above mentioned expressions for T (z) and Yk (z).
Important intermediate species that occur in real flames like CO and H2 are not
taken into account. It is difficult to derive the flame structure if the fuel stream
contains more then one fuel type (e.g. a mixture of CH4 and H2 ) but dilution of the
fuel stream with nitrogen or oxygen (and thus air) can be treated in the same way.

4.3.2

Infinitely Fast Reversible Chemistry : Equilibrium

When chemical reactions controlling a diffusion flame are not irreversible, fuel, oxidizer and products may coexist at the same location and at each point in the
z-diagram the equilibrium condition (4.19) must be satisfied. Even though the
chemistry proceeds fast (no finite rate effects) the flame structure is more complex
than in the mixed is burnt case. Chemical equilibrium is usually described by either
of two equivalent formulations: equilibrium constants or minimization of free energy.
If a generalized solution method is used the two methods require about the same
computational effort [36]. However with the method of minimization of free energy
each species can be treated independently without specifying a set of reactions a priori as is required with equilibrium constants. Therefore the minimization method
is most often used in numerical programs for chemical equilibrium computations
[37, 38, 39, 40].
The code used is based on CHEMKIN [40] A General- Purpose, Problem- Independent, Transportable, Fortran Chemical Kinetics Code Package developed at
Sandia Laboratories by Kee, Mitchell, Miller and Jefferson. The condition of equilibrium can be stated in terms of several thermodynamic functions, such as the minimization of Gibbs or Helmholtz energy or the maximization of entropy. The choice
of the thermodynamic function depends on the kind of problem. If one wants to use
temperature and pressure to characterize the thermodynamic state, Gibbs energy is
most easily minimized since p and T are its natural variables (G(T, p) = H T S).
The numerical method to obtain the equilibrium composition of the mixture is based
on the minimization of free energy with certain constraints because of the conservation of elements. The well-known Lagrange method is used for the minimization and
the chemical potentials for the species are determined from thermochemical tables
[41, 42]. The equations required to obtain equilibrium composition are not all linear
in the composition variables and thus an iteration procedure is generally required.
Often a Newton-Raphson method is used to solve for corrections to initial estimates
of species mass fractions, Lagrange multipliers and temperature. The reader is referred to [4, 39, 40] for more details of the method.
For constant pressure combustion the thermodynamic state is specified by pressure and enthalpy. The enthalpy is determined from the mixing of the reactants as
described by equation (4.34). The output of the code is the mass or mole fractions
of the considered species and the temperature. Figures 4.6 and 4.7 compare the
composition and temperature in mixture fraction space for a pure CH4 /air flame
(zst = 0.055) with the results for irreversible infinitely fast chemistry. For the equilibrium computation 17 species have been taken into account. On the lean side the

36

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

Figure 4.6: Comparison of temperature in mixture fraction space for irreversible


and reversible fast chemistry
flame structure is almost identical but on the rich side a clear difference can be
observed. There appears a significant drop in temperature for z > zst and only at
z close to 1 the irreversible solution is recovered. The drop in temperature corresponds to the formation of intermediate species CO and H2 , i.e. the endothermic
dissociation of CO2 and H2 O and the change in the heat capacity of the mixture.
In the rich side the mass fractions of CO2 and H2 O are smaller than in the mixed
is burnt profiles because of the formation of the intermediates. The method thus
provides a flame structure including reversible chemistry effects and can also deal
with mixtures of fuels.

4.3.3

Finite Rate Chemistry : Flamelets

In many circumstances it is inadequate to assume infinitely fast chemistry and one


needs to incorporate more chemistry into the turbulent combustion model. The link
between flow variables like species mass fraction or temperature and the mixture
fraction is no longer unique if finite rate chemistry is introduced. A point with a
certain mixture fraction can correspond to a zone of pure mixing (A), ignited flame
(B) or quenched flame (C) as shown in figure 4.8 (showing an ideally lifted flame).
Additional information is required to choose the correct flame structure and this
can be achieved using flamelet models. The basic idea is to assume that a small
instantaneous flame element embedded in the turbulent flow has the structure of
a laminar flame (see figure 4.9). This assumption requires that the reaction zones
are thin compared to the turbulent flow scales i.e. the Damkohler number is large.
To represent correctly the turbulent flamelets using laminar diffusion flames some
parameters must be preserved: the pressure should be the same as the conditions at
infinity (mass fractions YF0 ,YO0 and temperature TF0 ,TO0 ) but other parameters may
be important like strain, curvature and flamelet age and are retained or not de-

4.3. MODELING FOR DIFFUSION FLAMES

37

Figure 4.7: Comparison of composition between mixed is burnt approach (5 species


CH4 ,O2 , N2 , CO2 , H2 O) and equilibrium approach (17 species CH4 ,O2 , N2 , CO2 ,
H2 O, CO, H2 , OH, CH3 , CH3 O, CH2 O, HCO, HO2 , H2 O2 , H, O, C)

Figure 4.8: Diffusion flame structure for finite rate chemistry

38

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

Figure 4.9: Flamelet concept

pending on the complexity of the used flamelet model. In the model used, only the
scalar dissipation rate at the stoichiometric surface is retained which is enough to
account for stretch effects and quenching but cannot be used e.g. to predict flame
stabilization. The flame structure is then given by a set of functions Yk (z, ) and
T (z, ) which are the solution of the flamelet equations (4.15) and (4.17).
In these equations the only term depending on the spatial variables (xi ) is the scalar
dissipation rate which controls the mixing (gradients of z) and once it is known the
flamelet equation can be entirely solved in mixture fraction space. The value used
for is usually assumed to be the value at the stoichiometric plane (z) = (zst ).
Using the steady flamelet approach (assuming a steady flamelet structure although
the flow itself, and so the z-field, depends on time) one can compute the functions
Yk (z, st ) and T (z, st ) independently of the flow code and store them in flamelet
libraries.
Another approach can be used to generate these flamelet libraries and will be used
in the current work. The original set of transport equations (continuity, momentum,
N species mass fractions) can be solved for a 1D counterflow laminar diffusion flame.
Since chemistry is stiff special computer codes [40, 43, 44] have been developed to
solve the discretized equations. In these codes a set of N+1 strongly non-linear coupled PDEs are solved. When a mesh is generated to discretize the equations, one
has to account for steep gradients in the field variables which occur in flames. The
chemical active layer can be very thin compared to the computational domain depending on the chemical composition and the flame stretch. To avoid unreasonable
computational effort some mesh clustering has to be applied but since the location
of gradients is not known a priori an adaptive gridding technique is inevitable. A set
of species that are taken into account and a set of chemical reactions has to be chosen.
For hydrocarbon fuels commonly used mechanisms are the Smooke mechanism
[45] with 16 species - 25 reactions and the family of GRI-mechanisms [46]: GRI
1.2 with 32 species - 177 reactions, GRI 2.1 with 49 species - 279 reactions and
GRI 3.0 with 53 species - 325 reactions. In the Smooke and GRI 1.2 mechanism
only intermediate species containing the elements C,H and O are considered while
in the other two mechanisms also species containing nitrogen (e.q. NOx , NHx ) are
taken into account as well as the chemical reactions describing their formation and

4.3. MODELING FOR DIFFUSION FLAMES

39

destruction. For all considered species thermodynamic and transport properties have
to be known. Due to the non-linear form of the discretized PDE, linearization is
performed and an implicit solution strategy is used to solve the stiff set of equations.
In this work the 1D flame solver Chem1D developed at the TU Eindhoven [43] has
been used. More details about this solver and a description of the used models and
numerics can be found in [47]. The solver will provide the flame structure in physical
space (x) which can be easily recast to mixture fraction space to obtain Yk (z) and
T (z). Several input parameters are required for the flame solver:
Boundary conditions: the composition of fuel and oxidizer streams as well as
their temperatures have to be set.
Chemical mechanism : several mechanisms describing the combustion of hydrocarbon fuels can be used. Available schemes are Smooke, and GRI 2.1 and
3.0. Through input databases with thermodynamic data, transport properties
and reactions other mechanisms can be introduced by the user. As an example
the GRI 1.2 mechanism was introduced.
Strain or stretch : which will determine the scalar dissipation rate. Changing
this value corresponds to varying the inlet stream velocities of the counterflow
diffusion flame.
Solver settings : parameters for the numerical solution of the discretized equations (differential scheme, number of iterations, time integration, time step
control, grid settings, regridding parameters)
Other parameters allow to include more detailed effects (such as radiation, flame
acoustics, variable Le number, ...) that will not be considered in the present work.
Figure 4.10 shows the flame structure of a pure CH4 /air flame (zst = 0.055)
for different strain values, the lowest is close to equilibrium and the highest one is
close to extinction, obtained with Chem1d. On the left, the fuel and oxidizer mass
fractions are shown in physical space for 3 different strain levels. The x = 0 plane
corresponds to the position of the stoichiometric mixture fraction. As one can see, if
the strain increases, both methane and oxygen mass fractions start decaying further
from the flame front. In the low strain limit the solution approaches the infinitely
fast chemistry solution where fuel and oxidizer cannot coexist. In mixture fraction
space shown on the right of figure 4.10 it can be seen that the oxidizer extends well
into the fuel side (z > zst ) for high strain values. In figure 4.11 some important
radicals are shown for the same flame. It can be seen that intermediate species like
CH3 and CH2 O do not extend far into the oxidizer side because they are rapidly
consumed at the reaction layer. Radicals like OH do not extend far into the fuel
side since they are rapidly consumed as methane is attacked and converted to CO.
The OH radical shows a narrow peak near the reaction zone and can thus be used
to track the flame front (OH-PLIF measurement technique).
Finally figure 4.12 shows the temperature in mixture fraction space for different
strain rates. In the left figure the temperature is shown for the entire strain range the

40

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION


0.25

0.8

0.2

0.6

0.6

0.15

0.4

0.4

0.1

0.2

0.2

0.05

Y
0
0.5

Yi

0.8

YCH

0
0.5

0
0

0.05

0.1

0.15

0.2

0.25

Figure 4.10: CH4 and O2 mass fractions for a pure CH4 /air flame: (left) in physical
space (
low strain, medium strain, high strain) ; (right) in mixture fraction
space (zoom around zst ,
low strain, high strain, infinitely fast chemistry
limit)
3

x 10

0.06

x 10

x 10
1

0.8
1.5

CH O

CH

0.02

0.6

YCO

0.04

OH

0.4
0.5
0.2

0
0.5

0
0.5

0
0.4

0.3

0.2

0.1

Figure 4.11: Some important radicals for a pure CH4 /air flame in physical
space:(left) OH and CO (
low strain, medium strain, high strain); (right)
CH3 and CH2 O (
low strain, medium strain, high strain)
1D laminar flame solver can handle. If the strain becomes to low, which is equivalent
to a very low velocity of the incoming fuel and oxidizer streams, the flame structure
is mainly determined by the diffusive terms in the governing equations. The flame
becomes very thick and starts spreading over the entire calculation domain. Indeed
in the limit of an unstrained flame, the flame structure will not converge to a steady
state, since the flame will grow infinitely and the flame solver cannot converge. In
the high strain limit the flame becomes very thin. For sufficiently large strain rates,
chemistry becomes a limiting factor: chemical reactions have more difficulty keeping
up with the rate at which fuel and oxidizer enter the reaction zone. Eventually

4.4. TURBULENT DIFFUSION FLAMES IN RANS CODE

41

chemistry becomes too slow to burn the incoming reactants and quenching occurs.
2500
MIB Variable C

Flamelet low strain


Flamelet medium strain
Flamelet high strain

2000

T(K)

1500

1000

500

0
0

0.25

0.5

0.75

Figure 4.12: Temperature in mixture fraction space for a pure CH4 /air flame for the
entire strain range a (1/s) (left) and comparison with the infinitely fast chemistry
limit (right)

4.4

Turbulent Diffusion Flames in RANS code

To fix thoughts the general favre averaged (RANS) transport equations are recalled:
the continuity or mass conservation equation (3.35), momentum conservation equation (3.36), the species mass fraction equation (3.37) and the mixture fraction equation which is described as a general conserved scalar (3.38)
uei
+
=0
t
xi

(4.37)


fj
ij ug
p
ue i ue i u
i uj
+
=
+
t
xj
xi
xj

Yfk uei Yfk

+
=
t
xj
xj

Yk

D
+ k
ug
i Yk
xj

ze ue i ze
+
=
t
xj
xj

for k = 1, . . . , N

z

ug
D
iz
xj

(4.38)
(4.39)
(4.40)

This set of equations contains several unclosed terms that have to be modeled:

The Reynolds stresses ug
i uj which are generally closed using the turbulent
viscosity assumption by Boussinesq with an expression similar to the viscous
stress tensor ij for Newtonian fluids

ui uj = ui uj = t

uei uej 2 uek


2
+

ij + kij
xj
xi
3 xk
3

(4.41)

42

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION


with t the turbulent viscosity (t = t ). The last term has been added to
P

recover the right definition of the turbulent kinetic energy k = 21 3k=1 ug
k uk .
Models for the turbulent viscosity are discussed in section 3.2.
g


The turbulent scalar transport terms ug
i Yk and ui z usually modeled using
t Yfk
t ze


and ug
where Sct
a gradient assumption ug
Y
=

i k
iz =
Sckt xi
Sct xi
is a turbulent Schmidt number. The closed transport equation for ze is then
given by
!


ze ue i ze

t ze
(4.42)
+
=
+
t
xj
xj
Sc Sct xj

The laminar diffusive fluxes for species or mixture fraction can be neglected
against turbulent transport assuming a large turbulence level (large Re number
limit) or can be retained by adding a laminar diffusivity to the turbulent e.g.
ze
ze
z
D

for the mixture fraction D x


=
=
j
xj
Sc xj
The species chemical reaction rate k which is hard to model because of the
non linearity of the source term.
As discussed in section 4.3 diffusion flame computations may be split into two
problems: a mixing problem providing the mixture fraction field z (xi , t) and a
flame structure problem where species mass fraction Yk , temperature T and reaction
rates k are expressed as functions of the mixture fraction. This remains true for
turbulent flames but the averaging adds a complexity compared to laminar flames.
To determine average values the mean value of mixture fraction ze is not sufficient and
higher z moments are needed or if possible the entire probability density function
(pdf) of z. When the pdf of mixture fraction p(z) is known, mean species mass
fractions, temperature and reaction rates are given by
Yek =
Te

k =

Z
Z

0
1
0

Yk |z p (z ) dz


T |z p (z ) dz

k |z p (z ) dz

(4.43)
(4.44)
(4.45)

where Q|z denotes the conditional average of quantity Q for a given value of the
mixture fraction z = z depending on z and other quantities such as the scalar
dissipation rate and p (z ) is the z-probability density function. Thus two problems
have to be solved: a mixing problem providing the mean mixture fraction ze (xi , t)
2 ) and a flame structure problem
and some of its higher moments for the pdf (e.g. zg
expressing Yk , T and k as a function of mixture fraction. These functions are given
as conditional expressions Yk |z and T |z , since mass fractions and temperature
may depend on multiple parameters. Two possibilities now remain. In the primitive
variable method, assumptions are made on the flame structure to provide conditional

4.4. TURBULENT DIFFUSION FLAMES IN RANS CODE

43

quantities for species and temperature e.g. from flamelet libraries. No species mass
balance equations or energy/temperature equations are solved and the mean reaction
rates k do not have to be modeled. In the reaction rate approach, balance equations
for species mass fractions and temperature are solved and reaction rates k have to be
modeled. The primitive variable method is clearly less time consuming because less
balance equations have to be solved but in the reaction rate method more additional
effects like heat release, compressibility effects or secondary reactant injection can
be modeled. In this work only the primitive variable approach will be used to model
turbulent non-premixed combustion.

4.4.1

Infinitely Fast Chemistry

For infinitely fast chemistry the instantaneous mass fractions Yk and temperature T
are directly linked to the mixture fraction z (either through the Burke-Schumann relations (4.25,4.26,4.32) for irreversible chemistry or through the equilibrium relation
(4.19) for reversible chemistry) and thus the conditional averages can be simplified
to




Yk |z = (z ) Yk (z ) and
T |z = (z ) T (z )
(4.46)

or in terms of a mass weighted pdf

pe (z ) =

(z ) p (z )

(4.47)

equations (4.43) and (4.44) can be rewritten as


Yfk =
Te

1
0
1
0

(z ) Yk (z ) pe (z ) dz

(4.48)

(z ) T (z ) pe (z ) dz

(4.49)

e
Determining Yfk and Te is now reduced to finding the pdf p(z)
of mixture fraction.
It can either be presumed or determined by solving a transport equation (the so
called transported pdf method [48, 49, 50, 51, 52, 53, 54, 55, 56]). A widely used
approximation is to presume the shape of the pdf using simple analytical functions.
The most popular is the -function depending only on two parameters, the mean
2 :
mixture fraction ze and its variance zg
e
p(z)
= R1
0

(a + b) a1
z a1 (1 z)b1
z a1 (1 z)b1
=
z (1 z)b1
=
a1
b1
B(a, b)
(a)(b)
z (1 z) dz

with B(a, b) the normalization factor, the -function defined by (x) =


and the pdf parameters a and b determined by
a = z

z(1 z)

and
b=

2
zg

(1 z)a
z

+
0

(4.50)

et tx1 dt

(4.51)

(4.52)

44

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

Typical -pdf shapes are shown in figure 4.13. One can observe that the -pdf is
symmetric if a = b since then z = 0.5 and takes a gaussian shape for small values
g
2 (a = b = 50 corresponds to z
2 0.0025). If a = b < 1 the pdf has a
of zg
singularity at both ends of the z-interval and if only a or b is smaller than 1 the
pdf has one singularity at z = 0 or z = 1 respectively. If a 6= b and a, b > 1 the
pdf is an asymmetric function. An estimate of the mixture fraction variance is still

Figure 4.13: Normalized -pdf for different values of a,b


required to determine the -functions and in RANS it is generally found by solving
a transport equation. This equation can be derived from the balance equations for
z and ze as shown in Appendix A. The exact equation obtained is
2
2
zg
ue i zg
+
=
t
xi

 2 
ui z
xi

{z

z 2
+
D
xi
xi
|

Turbulent transport

2z

{z

ze

D
xi
xi

Molecular diffusion
ze
z z
2ui z
2D
xi
xi xi
|

{z

Production

{z

!
}

(4.53)

Dissipation

where the last term is the scalar dissipation rate ep of the perturbations of the
2 . In
mixture fraction field z. This scalar dissipation rate measures the decay of zg
z z
fact the total scalar dissipation rate e defined as 2D x
can be written as
i xi
e = 2D

ze ze
z ze
z z
+ 4D
+ 2D
xi xi
xi xi
xi xi

(4.54)

and neglecting density fluctuations the total scalar dissipation rate can be written
as the sum of the scalar dissipation rate em due to mean ze field and the scalar
dissipation rate ep due to fluctuations of z
e = e m + ep

(4.55)

4.4. TURBULENT DIFFUSION FLAMES IN RANS CODE

45

Usually in RANS mean gradients are neglected against fluctuation gradients and
2 :
thus e ep . Some unclosed terms remain in the balance equation for zg
The turbulent transport is modeled using a classical gradient assumption
ui z 2 =

2
t zg
Sct1 xi

(4.56)

with t the turbulent viscosity and Sct1 a turbulent Schmidt number.


The molecular diffusion can be neglected against turbulent transport assuming
large Reynolds numbers or can be modeled using an expression similar to (4.56)
using the molecular viscosity .
The production term is also modeled using a gradient assumption
ui z

ze
t ze ze
=
xi
Sct2 xi xi

(4.57)

with Sct2 a second turbulent Schmidt number.


The scalar dissipation rate ep measuring the decay of the mixture fraction
2 and thus plays a similar role as the
fluctuations estimated by the variance zg
dissipation rate of kinetic energy for the velocity field. Assuming the simplest
2
dzg
case of homogeneous flows equation (4.53) can be reduced to
= ep .
dt
The scalar dissipation rate may thus be modeled using a turbulent mixing time
t
2
zg
2
(4.58)
ep = c
= c zg
t
k
with c a model constant. This relation expresses that the scalar dissipation
time and turbulence dissipation are proportional.

2 can then be written as


The closed equation for zg

2
2
zg

t

ue i zg
=

+
+
t
xi
xi
Sc1 Sct1

2
t
zg

+2
+
xi
Sc2 Sct2

2
ze ze
c zg
xi xi
k
(4.59)
It is also possible to determine ep by solving a balance equation using several approximations for unclosed terms as given in [57]:
!

e
e
e
e ue i e
e
e
e g

e2
z
ui
D
2Cp1 g ug
z
C

u
u
C
+
=
C
p2
d1
d2
i
i j
2
t
xi
xi
xi
k
xj
k
z 2 | {z xi}
zg
{z
}
|
Pz
Pk
(4.60)
g
2
with Pz the production of scalar fluctuations z and Pk the production of turbulent
kinetic energy. Cp1 , Cp2 , Cd1 and Cd2 are empirical model constants.

46

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

Figure 4.14 shows how the model is integrated in the solver. The RANS solver
provides pressure p, velocity components u
, v, w,
mixture fraction z and variance
g
2
of mixture fraction z . Based on the expressions for species mass fraction Yk and
temperature T as a function of mixture fraction and the presumed beta shape pdf
2 , tables are generated. The code enters these tables with z
e and
based on ze and zg
g
z 2 and returns Te and Yfk . Density is then calculated from state equation using p,
!1
N
X
Y
k
Te and the molecular mass of the mixture given by W =
.
Wk
k=1

Figure 4.14: Presumed pdf - mixed is burnt model in RANS code

4.4.2

Finite Rate Chemistry

The flame structure is given by a set of functions Yk (z, st ) and T (z, st ) as a


solution of the flamelet equations (4.15 and 4.17) or from a 1D laminar counterflow
flame calculation. They can be determined independently of the turbulent code and
stored in flamelet libraries using the mixture fraction z and the scalar dissipation at
stoichiometry st as parameters. The mean species mass fractions are then obtained
from the integration of the flame structure with the joint-pdf of the mixture fraction
and scalar dissipation p (z, st )
Yf

+
0

1
0

Yk (z, st ) p (z, st ) dzdst

(4.61)

T (z, st ) p (z, st ) dzdst

(4.62)

and the temperature is given by:


Te

+
0

1
0

Often statistical independence between the mixture fraction z and its scalar dissipation rate is assumed :
p (z, st ) = p (z) p (st )
(4.63)

4.4. TURBULENT DIFFUSION FLAMES IN RANS CODE

47

For the mixture fraction pdf generally a -function (4.50) is presumed based on ze
g2 as for infinitely fast chemistry. For the pdf of the scalar dissipation a delta
and z
function (st est ) can be used but a log-normal distribution is more physical
1
(lnst )2
exp
p (st ) =
2 2
st 2

(4.64)

where the parameter is linked to the mean value of st :


e st =

+
0

2
st p (st ) dst = exp +
2

(4.65)

and is the variance of ln() so that




g
2
2
g
exp( 2 ) 1

st
st =

(4.66)

g
and is often assumed to be constant ( = 2). In these formulations
st is still
missing but one can use the analytical expression for the scalar dissipation field in
a 1D steady strained laminar flames (see appendix B) to relate the mean scalar
g
e :
dissipation at stoichiometry
st and the mean scalar dissipation

with

g
e =
st



F (z)
g
2
e
g
z
,
z
p(z)dz =
F
st
F (zst )


F (z) = exp 2 erf (2z 1)

i2 

(4.67)

(4.68)

and the mean scalar dissipation rate e can be determined from (4.58) or (4.60).
In figure 4.15 the implementation of the presumed pdf - flamelet approach in the
RANS code is shown. The flame structure is pre-computed giving the temperature
and species mass fraction in the z,st plane. The integration with the pdfs is also
performed in preprocessing and thus Te and Yek are tabulated in a 3D table with
g
2 and
2 and
g
e, z
e from which
g
parameters ze, zg
st . The flow solver provides z
st
can be calculated through (4.67) and thus all necessary parameters to perform the
lookup in the flamelet table. Using the mean species mass fraction and temperature
the density is calculated and is transferred to the flow solver.

4.4.3

Conditional Moment Closure

The CMC-technique was independently proposed for modeling turbulent reacting


flows in a RANS context by Klimenko [58] and Bilger [59] . The idea is, instead of
getting species mass fractions from e.g. flamelet tables, to solve balance equations
for the mean value of mass fraction Yk for a given value of mixture fraction z = , denoted as hYk |i (these are also called conditional averages or conditional moments).
The main idea is that most of the fluctuations in mass fractions and temperature,
that make closure of chemical reaction rates difficult, can be associated with the fluctuations of one scalar variable (mixture fraction). The assumption is then that the

48

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

Figure 4.15: Presumed pdf-flamelet model in RANS code


fluctuations of scalars (species mass fractions, temperature etc.) around their conditional moments are relatively small. If the conditional averages of mass fractions
and temperature are available the conditional average of the reaction rate (i.e. the
reaction rate appearing in the equation for the conditional mass fraction) can then
be closed according to the following first-order simplification (first-order closure)
hk |i = k (hY1 |i, hY2 |i, . . . , hYN |i, hT |i)

(4.69)

with hk |i the conditional reaction rate appearing in equation for hYk |i. This
simple closure which is not possible for the reaction rate in the equations for the
unconditional mass fractions is due to the fact that the fluctuations of the conditional averages are much smaller. The closure can also be extended to second-order.
The unconditional mean mass fractions can be obtained from the conditional ones
through a (presumed) pdf (e.g. the same as in presumed pdf flamelet models)
hYk i =

1
0

hYk |ip()d

(4.70)

Hence, to calculate the integral numerically, conditional averages are needed at different values of z. In practice about 50 values are used. This relatively large number
is also required as the equations for conditioned mass fractions contain second-order
derivatives in z-space which have to be estimated with sufficient precision. Compared to transported pdf methods, CMC has the advantage that the dimensions

4.5. TURBULENT DIFFUSION FLAMES IN LES CODE

49

remain fixed for any number of reacting components. In full pdf methods the dimensional space is proportional to the number of components. The CMC equation
reads (assuming unity Lewis number)
2 hYk |i
hYk |i
+ h|ihv|ihYk|i = h|ih|i
+ h|ih k |i + eQ + ey
t
2
(4.71)
with the value of mixture fraction, k the rate of formation of species k and
the scalar dissipation. The term eQ is assumed to be negligible at high Re [58].
Following the same reference, the first and last term of ey can also be neglected and
the remaining term in ey results in an expression for a conditional turbulent flux
h (v00 Yk00 ) |i where denotes the turbulent fluctuations about the conditional
average. For more information the reader is referred to the abundant literature on
CMC.
h|i

4.5

Turbulent Diffusion Flames in LES code

Probability density functions (PDF) introduced in RANS context can be easily


extended to LES where they are sometimes called filtered density functions (FDF)
[60, 61], large eddy pdf (LEPDF) [62] or subgrid scale pdf [63, 34]. The (favre)
filtered species mass fraction can be written as
Yk =
=

1
0

(x0 , t) Yk (z (x0 , t)) G(x0 x)dx0

(z )Yk (z )dz

(z z(x0 , t))G(x0 x)dx0

(4.72)

with G the LES-filter and (z z(x0 , t)) the fine-grain function. The filtered density
function is then

p(z ; x, t) =

(z z(x0 , t))G(x0 x)dx0

(4.73)

(z )p(z ; x, t)
the

same expression for the favre filtered species mass fraction is obtained

e
Defining again a mass weighted filtered density function, p(z)
=

Yk =

1
0

e )dz
Yk (z )p(z

(4.74)

One can show that the FDF should have all the properties of a PDF. In [64] it is
proposed to use a presumed -shape for the pdf of mixture fraction based on the
2 as given in equation
filtered mixture fraction ze and its unresolved fluctuations zg
2 =
(4.50). Similarly as in RANS an exact but unclosed balance equation for zg
(z g
ze)2 can be derived but usually in LES other models are used. Appealing to
the fractal nature of turbulence and assuming that the small scale statistics can be
inferred from somewhat larger scale structures (as in the scale similarity subgrid
scale models) a scale similarity model is proposed in [64]


 2 

d
be
2 = C
e2 ) z
zg
z (z

(4.75)

50

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

b > and C is a model


where the caret b. denotes a test filter with filter size
z
parameter that has to be determined (and depends on the chosen test filter). Usually
the test filter width is chosen as twice the LES filter width[65]. The model constant
can also be determined from the turbulent spectrum [66] or by a dynamic procedure
[67]. In [68] and [65] a dynamic model based on a scaling law is proposed
2
2 = C2 |z
e|
zg

(4.76)

relating the unresolved mixture fraction fluctuation to the gradient of the resolved
mixture fraction field ze and the local filter width serves as the length scale of
the subgrid turbulence. This relation can be obtained by considering the transport equation for the subgrid scale variance and assuming local equilibrium. The
constant C can be determined dynamically through a procedure similarly as in the
Germano model but to have a stable value averaging in statistically homogeneous
directions is desired. Similarly as for the standard Smagorinksy model a constant
value throughout the domain can also be used. Once the resolved mixture fraction
2 are determined the same procedure as in
field ze and its subgrid scale variance zg
RANS can be applied. Using these two scalars the lookup tables for infinitely fast
chemistry (mixed is burnt or equilibrium) can be used to find the filtered values of
temperature Te and species mass fractions Yek and thus the density to be returned
to the flow solver. Apart from the turbulence modeling and the model for mixture
fraction variance the scheme given in figure 4.14 also applies to LES.
To include finite rate effects the subgrid scale probability density function concept can be easily extended to include the dependence of species mass fractions and
temperature on both mixture fraction z and the scalar dissipation rate [69, 70].
The filtered species mass fraction is the given by
Yk =

1
0

e , )d dz
Yk (z , )p(z

(4.77)

The dynamics of the local strain-diffusion competition involved in scalar mixing


suggests that must be concentrated in locally one dimensional layer like structures
[71]. The -structure in z-space can therefore be estimated from the solution of a
1D counterflow diffusion flame as given in appendix B
= 0 F (z) = st

F (z)
F (zs t)

with F (z) = exp 2 erf1 (2z 1)

i2 

(4.78)

with 0 the local peak value of within the reaction layer and st the value at
the stoichiometric plane which can be significantly lower than 0 (especially if zst
is low). Thus flamelet libraries may be parameterized in terms of either 0 as in
[70] or st which has been used already in the RANS context. Assuming statistical
independence between the mixture fraction and the stoichiometric scalar dissipation
rate one can write
Yk =

1
0

e , x, t)p(
e st , x, t)dst dz
Yk (z , st )p(z

(4.79)

e st , x, t) one can use the log normal pdf as in RANS or assume that species
For p(
mass fractions and temperature weakly depend on the scalar dissipation rate and

4.5. TURBULENT DIFFUSION FLAMES IN LES CODE

51

keep only the first two terms in the Taylor series (which is in fact equivalent to
e st ) = (st
est ))
assume a delta-shaped pdf for the scalar dissipation p(
"

Yk
Yk (z, st ) Yk (z, e st ) +
st

Inserting this into (4.79) yields


Yk =

1
0

st =
est

(st e st )

e , x, t)dz
Yk (z , e st )p(z

(4.80)

(4.81)

The mean scalar dissipation rate at stoichiometry is related to the mean scalar
dissipation rate by
Z 1
F (z )
e
e
e )dz
= st
p(z
(4.82)
0 F (zst )

A model is then required for e and in [70], [68] and [72] a simple equilibrium hypothesis is used to obtain


e
t

e =
+
|ze|2
(4.83)
Sc Sct
with Sc the Schmidt number (sometimes also called Peclet number P e), Sct a ture the filtered dynamic viscosity and t the subgrid scale
bulent Schmidt number,
turbulent viscosity (provided by the subgrid turbulence models). Jimenez et al. [73]
use an extension of the RANS model for e by defining a SGS scalar mixing time as
1

= 2
=
z
Vz
z z2

(4.84)

with Vz the SGS scalar variance and assume that this characteristic mixing time is
proportional to the turbulent characteristic time t :
z2

=C
=
2
z
t
k
z

where k is the SGS kinetic energy k =


energy dissipation =

ui ui
xj xj

(4.85)

1
(ui ui ui ui ) and the filtered kinetic
2

and estimate the model constant from DNS

1
1
C=
=
. Unfortunately, unlike in RANS k- calculations, quantities such as
Sc
0.75
the SGS kinetic energy and its dissipation are not available in a practical LES and
have to be approximated based on the SGS turbulence model. Using a Smagorinksky
type eddy viscosity model and Yoshizawas model for the SGS kinetic energy Jimenez
et al. [73] propose to use


= 2 + CS 2 |S| S ij S ij

and k = 2CI 2 S ij S ij

(4.86)

and the filtered scalar dissipation is then given by


=


+ CS 2 |S|  2
2
z

z
ScCI 2

(4.87)

52

CHAPTER 4. TURBULENT NON-PREMIXED COMBUSTION

CS and CI are assumed constant or can be determined through a dynamic procedure.


The kinetic energy can also be available from a transport equation [18]. Note that
in [73] a transport equation similar to (4.59) is solved for Vz = z 2 z 2 instead of
using (4.75) or (4.76)
Vz uj Vz

+
=
t
xj
xj

Vz
(D + Dt )
xj

+ 2(D + Dt )

z z
xj xj

(4.88)

In this dissertation the -pdf for mixture fraction and log-normal pdf for scalar
dissipation is used. Solving the balance equation for ze and using (4.83) for e and
e and Te from the lookup tables
2 all parameters needed to determine Y
(4.76) for zg
k
are available. The same scheme as in RANS, as shown in figure 4.15, is applied to
couple the flame structure, which is pre-computed, with the flow solver.

Chapter 5
Turbulent Premixed Combustion
Modeling
5.1

Laminar premixed flames

A premixed flame is a propagation of a localized combustion zone (the flame front) at


subsonic or supersonic velocity. In the first case the flame is a deflagration wave, in
the second case a detonation wave. An important difference is that where the pressure increases significantly across the flame in a detonation flame, in a deflagration
flame there is often a negligible pressure difference. Only deflagration type premixed
flames will be further considered. Figure 5.1 schematically shows the structure of a
1D premixed flame. In a laminar premixed flame the fresh gases (fuel and oxidizer)
and burnt gases are separated by a thin reaction zone (the flame thickness lF is
typically 0.1-1 mm) where a strong T-gradient is observed (temperature ratio 5-7).
The premixed flame propagates towards the fresh gases and the thermal flux corresponding to the temperature gradient preheats the fresh gases before they start
burning. The propagation speed sL depends on various parameters like fuel and
oxidizer compositions and temperature, and is typically 0.1-1.0 m/s. Based on the
flame thickness lF , laminar flame speed sL and kinematic viscosity of the fresh gases
the flame Reynolds number can be defined as ReF = lFsL and is usually about 4.
Consider a one-step irreversible chemical scheme :
Reactants Products

(5.1)

and assume :
constant pressure (sL : 0.1-1 m/s  sound speed : 300-600 m/s)
all species have equal molecular weight Wk = W , heat capacities Cpk = Cp
and diffusion coefficients Dk = D (and thus Lek = Le)
unity Lewis numbers (Le =

Cp D

= 1 : molecular and heat diffusion are equal)

adiabatic conditions
53

54

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

Figure 5.1: Structure of a laminar premixed flame (Tu unburnt fresh gas temperature, Tb burnt gas temperature)
Under these conditions one can simplify the governing equations (2.1),(2.3),(2.33) to
obtain the equations describing the 1D irreversible 1-step laminar premixed flame.
Assuming lean combustion - i.e the reaction rate is only limited by the fuel mass
fraction and YO remains approximately constant - one obtains in the flame reference
frame:
u =
constant = 1 sL
!
d
dYF
dYF
=
D
+ F
1 sL
dx
dx
dx
!
d
dT
dT
=

QF
1 Cp sL
dx
dx
dx

(5.2)
(5.3)
(5.4)

where Q(J/kg) is the reaction heat per unit mass defined in (2.34) and (4.5) and the
subscript 1 denotes fresh gas conditions (the burnt gases are denoted by subscript
2). If one integrates (5.3) and (5.4) between x = and x = + one obtains
1 sL YF,1 =

1 Cp sL (T2 T1 ) = Q

Z +

= F

F = QF

(5.5)
(5.6)

since the diffusive terms are zero on both ends of the domain and the inlet speed
must be equal to the flame speed. F is the total fuel consumption rate and equation
(5.5) shows that all the fuel entering the domain is burnt downstream the flame front
and equation (5.6) states that the power released by the combustion of this fuel is
entirely converted to sensible energy to heat up the mass flux 1 Cp sL from T1 = Tu
to T2 = Tb i.e. the adiabatic flame temperature is given by
T2 = T 1 +

QYF,1
Cp

(5.7)

5.1. LAMINAR PREMIXED FLAMES

55

The same relation can also be derived starting from the conservation of enthalpy
between fresh and burnt gases
Cp (T1 T0 ) +

N
X

h0f,k Yk1

k=1

N
X

h0f,k Yk2

(5.8)

Wk k 1
Y = QYF1
WF F F

(5.9)

= Cp (T2 T0 ) +

k=1

and thus using (4.3)


N
X

Cp (T2 T1 ) +

k=1

h0f,k Yk1 Yk2 =

N
X

h0f,k

k=1

Introducing the reduced reactant mass fraction and reduced temperature


Y =

YF
YF1

and =

Cp (T T1 )
T T1
=
1
QYF
T2 T 1

(5.10)

where Y goes from 1 in the fresh gases to 0 in the burnt gases (assuming lean
combustion) and goes from 0 to 1 in these zones, equations (5.3) and (5.4) can be
written as
!

dY
d
dY
F
1 sL
=
D
+ 1
dx
dx
dx
YF
!
d
d
d
F
1 sL
=
1
dx
dx Cp dx
YF

(5.11)
(5.12)

Summing these equations and assuming unity Lewis number (/Cp = D) leads to:
d
d(Y + )
d(Y + )
=
D
1 sL
dx
dx
dx

(5.13)

so Y + is a conserved scalar and because in the fresh and burnt gases Y = 1, = 0


or Y = 0, = 1 the only solution is
Y + =1

(5.14)

In fact this equation expresses that the total enthalpy in the system is constant: if the
fuel mass fraction goes down i.e. the chemical enthalpy decreases, the temperature
goes up i.e. the sensible enthalpy increases. From a numerical point of view one can
only solve for the -equation and Y is obtained from (5.14):
d
d
=
1 sL
dx
dx

d
Ta
B (T1 + (T2 T1 ))1 (1 ) exp
Cp dx
T1 + (T2 T1 )
(5.15)
where the Arrhenius law (see equations (2.24) and (2.25)) has been used to express
the fuel reaction rate F (assuming a 1-step irreversible reaction). The reduced
temperature is also called the progress variable c (c = 0 in fresh gases and c = 1 in
fully burnt ones) which is thus defined either as a reduced temperature or a reduced
fuel mass fraction:
YF YFu
T Tu
(5.16)
or c = b
c=
Tb T u
YF YFu

56

CHAPTER 5. TURBULENT PREMIXED COMBUSTION


9
8

Reduced reaction rate

=4
=8
=13
=20

6
5
4
3
2
1
0
0

0.2

0.4

0.6

0.8

Figure 5.2: Reduced reaction rate F /(1 YF1 B) vs with 1 = 0, the heat release
factor = 0.75 and B1 adjusted so that the total reaction rate (area under the
curve) and thus the flame speed are equal for all flames
where T, Tu , Tb are respectively the local, unburnt gases and burnt gases temperature
and YF , YFu , YFb the corresponding fuel mass fractions (YFb 6= 0 for rich combustion).
These two definitions are equivalent for unity Lewis number (same molecular and
thermal diffusivity) and without heat losses (adiabatic) and compressibility effects.
The reduced reaction rate is usually written in a more convenient form by Williams

F
(1 )
= B (T )1 (1 )e exp
1
YF
1 (1 )

(5.17)

where - measuring the heat release - and - measuring the activation temperature
- are given by
QYF1
Ta
T2 T 1
=
=
(5.18)
=
T1
C p T2
T2
Figure 5.2 shows the variations of the reduced reaction rate versus the progress
variable for various values of . One can see that for increasing (i.e. when the
activation temperature is increasing) the problem becomes stiffer. The zone where
the reaction rate is non-zero in a premixed flame is small compared to the diffusion
length scales and roughly has a width of r 2/. The maximum reaction rate is
given by (assuming 1 = 0)
F =

1
1 YF1 B

exp 1

at max = 1

1
+

(5.19)

Several expressions have been derived for the laminar flame speed (see e.g. [34])
and in general the flame speed varies like the square root of the (thermal) diffusion
1
coefficient Dth
(of the fresh gases) and the square root of the reaction rate
s0L

1
Dth

(5.20)

5.2. EFFECTS OF TURBULENCE AND PREMIXED REGIME DIAGRAMS 57


The flame speed is also depending on the pressure (since Dth and b depend on p)
and on the fresh gas temperature T1 (because T2 depends on T1 and strongly
depends on T2 due to the exponential term).
The flame thickness can be introduced based on scaling laws to be
=

1
D1
= th
1 Cp sL
sL

(5.21)

This diffusive thickness can be determined once the flame speed is known but in
practice for example when a mesh has to be generated to resolve the flame the
obtained thickness from (5.21) tends to be too small. A more useful thickness can
be obtained from the temperature profile
L0 =

T2 T 1


0.7


max T
x

(5.22)

and is called thermal thickness. The drawback of this formulation is that the
temperature profile has to be known to compute L0 and thus this approximation
cannot be used e.g. to determine mesh constraints. Therefore Blint proposed the
following evaluation which is close to L0 in practice
Lb = 2

1
Dth
sL

T2
T1

(5.23)

The reaction rate zone r which is significantly smaller than L0 can then be estimated
from the activation parameter

5.2

Effects of turbulence and Premixed Regime


diagrams

Damkohler (1940) was the first describing turbulent combustion and introduced
wrinkling as the main mechanism controlling turbulent flames. He defined a turbulent flame speed by equating the mass flux m
through the instantaneous turbulent
flame surface AT to the mass flux through a laminar plane flame of cross-sectional
area A of the considered control volume as illustrated in figure 5.3. The turbulent
flame front is locally moving at the laminar flame speed sL and the turbulent flame
speed is defined as the speed the plane flame should have to consume the same
amount of fresh gases:
m
= 1 sL AT = sT A
(5.24)
If the density is assumed to be constant one can see that the increase of the turbulent
flame speed compared to the laminar burning velocity is due to the increase of the
total flame surface AT allowing a higher fuel consumption rate for the same crosssectional area A:
AT
sT
=
=
(5.25)
sL
A
The ratio is the flame front wrinkling factor corresponding to the available flame
surface divided by its projection in the mean propagation direction. The wrinkling

58

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

Figure 5.3: Flame wrinkling by turbulence and definition of turbulent burning velocity
will depend on the Reynolds number and thus on the fresh gas rms velocity u0 . Many
semi-phenomenological models relating sT to u0 can be found in literature [4] but
experimental and theoretical results show considerable scattering and it is doubtful
that a unique relation between the turbulent flame speed and the rms velocity exists.
Damkohler argued that the turbulent flame speed could be expressed in analogy with


Dt 1/2
and since the turbulent diffusivity Dt
the laminar burning velocity as sT
c
is proportional to u0 lt and D is proportional to sL lF :
sT

sL

u0 lt
s L lF

!1
2

(5.26)

showing that the turbulent flame speed is also depending on the length scale ratio
lt /lF . Often relations of the form
u0
sT
=1+C
sL
sL

!n

(5.27)

are used to express the turbulent flame speed sT , where 0.5 n 1.0 is a scaling
exponent and C a constant depending on the length scale ratio.
The nature of a combustion process depends strongly on the length and time scales
in the flame front and in the (turbulent) flow field. The turbulent and chemical time
scales can be defined as :
lt
lF
t = 0
C =
(5.28)
u
sL
where lt , u0 , lF and sL are respectively the integral length scale, the rms velocity
fluctuation, the flame thickness and the flame burning velocity. The turbulent scales
u0 and lt can be related to the turbulent kinetic energy and its dissipation
u0 k 1/2

lt =

u03

(5.29)

5.2. EFFECTS OF TURBULENCE AND PREMIXED REGIME DIAGRAMS 59


k
. The Kolmogorov length, time and velocity scales are (with the

kinematic viscosity):
and so t =

!1/4

 1/2

u = ()1/4

(5.30)

Theflame thickness
is the ratio of the diffusivity D and the laminar


 burning velocity
D
sL lF = sL and in the assumption of unity Schmidt number Sc = D = 1 the
turbulent Reynolds number (comparing the turbulent transport with viscous forces)
can be rewritten as :
u0 lt
u0 lt
Re =
=
(5.31)

s L lF
The turbulent Damkohler number compares the turbulent and chemical time scales:
Da =

s L lt
lt /lF
t
= 0 = 0
C
u lF
u /sL

(5.32)

In the limit of high Damkohler number (Da  1) the chemical time is short compared to the turbulent corresponding to a thin reaction zone distorted and convected
by the flow field. The internal structure of the flame is not affected by the turbulence
and may be described as a laminar flame element called flamelet. On the other
hand a low Damkohler number (Da  1) corresponds to slow chemical reaction
where reactants and products are mixed by the turbulent structures before reaction.
Most practical situations correspond to high or medium values of Da. It is worth
noting that various chemical time scales may be encountered in the same combustion process: fuel oxidation generally corresponds to short chemical time scales while
pollutant production or destruction (CO oxidation, NO formation) is slower.
The Karlovitz number is defined as
Ka = Da1
=
with
=

1/2
2
=
,
2
1/2

sL =

u2
l2
C
= F2 = 2

sL

(5.33)

,
lF

(5.34)

and u2 = ()1/2

and uses the Kolmogorov scale rather than the integral length scale because the
interaction between chemistry and turbulence occurs at the smallest scales only.
The different combustion regimes can be clearly visualized in a combustion diagram where the velocity scale ratio u0 /sL is plotted logarithmically over the length
scale ratio lt /lF . Using (5.33), (5.32) and (5.31) one can show that following relations
between these ratios are valid 1
1

Ka2/3 =

so Ka2/3 =

2
l2F u
2 s2L
1/3

lF

1/3
lt

1/3

u0
sL

2/3

lF

2/3
sL


lF sL 1/3

u2
2

1/3

and

u2
2

1
2

and since lF sL =

u03
lt

one obtains

u0
sL

= Ka2/3

lF
lt

1/3

and

thus lines of constant Karlovitz number are straight lines in the Borghi diagram with positive
slope 1/3

60

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

,
Figure 5.4: Turbulent premixed combustion diagram (Borghi) with lines of constant
Reynolds, Damkohler and Karlovitz numbers and different regimes

u0 /sL = Ka2/3 (lt /lF )1/3 = Re (lt /lF )1 = Da1 (lt /lF )

(5.35)

so in the diagram lines of constant Reynolds, Damkohler and Karlovitz can be easily
constructed.
Different zones can be observed :
In the laminar zone (1) the fluctuations in the flow field u0 are small compared
to the adiabatic burning velocity sL and the length scales of the largest flow
structures lt are small compared to the flame thickness lF . The line Re = 1
separates the laminar flame regime from the turbulent combustion regime
(Re > 1).
The flamelet regime or thin wrinkled flame regime is characterized by the
inequalities Re > 1 (turbulent) and Ka < 1 (fast chemistry). In this regime
the small turbulence scales (i.e. the Kolmogorov scales) have a turbulent time
larger than C and so the turbulent motions are too small to affect the flame
structure(see figure (5.5a). Two zones can be distinguished:
(u0 /sL ) < 1 : wrinkled flamelets (zone 2). In this zone u0 which may be
viewed as the rotation speed of the larger turbulent motions is unable
to wrinkle the flame surface up to flame front interaction. The laminar
propagation is dominant and the interactions between turbulence and
combustion remain limited.
(u0 /sL ) > 1 : corrugated flamelets or wrinkled flame with pockets(zone 3).
The larger structures are able to induce flame front interactions leading
to pockets. In this zone u0 sL u .
The Ka = 1 limit where chemical length, time and velocity scales equal the
Kolmogorov scales (Klimov- Williams criterion) separates the flamelet regime
from the thin reaction zone regime.

5.3. TURBULENT PREMIXED FLAME MODELS

61

Figure 5.5: Turbulent premixed combustion regimes proposed in [74]: (a) wrinkled flamelet, (b) thickened-wrinkled flame, (c) thickened flame ( T denotes mean
temperature, T instantaneous)
1 < Ka < 100 : Thickened wrinkled flame regime or thin reaction zone (zone
4, see figure (5.5b). . Since < lF the smallest eddies can enter into the
flame structure but they are not necessarily able to affect the inner layer or
reaction zone. This zone where heat is released has a thickness lr quite lower
than the thermal thickness of the flame lF typically lr 0.1lF . We may
therefore introduce a Karlovitz number based on the thickness of this layer
2
Kalr = lr2 0.01Ka. Therefore the value Kalr = 1 corresponds to Ka = 100.
Ka > 100 : Thickened flame regime, well-stirred reactor or broken reaction
zone regime (zone 5). The system is intensively disturbed by the small flow
structures so that the reaction may take place in regions instead of layers (see
figure (5.5c).

5.3

Turbulent premixed flame models

In the next two sections a short overview of the most common models for turbulent
premixed combustion will be given. Although this dissertation does not include
RANS simulations of premixed combustion, some well-known models will be described for completeness and because some concepts introduced in a RANS-context
have recently been extended to LES. The reader is referred to a number of reference
works [34, 63] and overview papers [75, 76] for more detailed information. For LES
of premixed combustion the three most used methods will be described, focussing
on the Artificially Thickened Flame model used in this dissertation.

5.3.1

RANS Models

Assuming constant pressure, unity Lewis numbers and adiabatic conditions, the
relation for the reduced temperature (or the progress variable c) defined in section

62

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

5.1 can also be used for a turbulent low-speed premixed flame. The system of
governing equations to solve is then given by:

fi ) =
+
(
u
t xi
fi
u

fi u
fj ) =
(
u
+
t
xi
 e
e
f
ui =
+
t
xi

0
p
x
+
j

xi

xi

(5.36)


ij ug
i uj


e ug
+
D
i
xi

(5.37)
(5.38)

.
= D
xi
xi
This system of equations has to be solved for , f
ui and e and contains three unclosed
terms which have to be modeled using the mean quantities or by solving additional
conservation equations:

where the averaged thermal diffusion term has been modeled as D

turbulent stress tensor (Reynolds stress) ui uj


turbulent scalar transport ui
mean reaction rate
The first term (momentum equation) is usually viewed as additional stress and
modeled through a turbulent viscosity t (Boussinesq assumption) as in non reacting
flow. The other two unclosed terms are specific for the introduction of combustion
into the Navier-Stokes equations. The most difficult term to model is the filtered
chemical reaction term .
Arrhenius approach
The simplest approach is to neglect turbulent effects and to assume that the mean
reaction rate corresponds to the reaction rate obtained using mean local values

and :
!


 
T
a
= = B 1 exp
(5.39)
2 T1 )
T1 + (T
In most situations this model is completely inadequate and is only relevant when
chemical time scales are larger than turbulent time scales corresponding to the
well stirred reactor regime where reactants mix rapidly and burn slowly (C  t ,
Da  1).
Eddy Break Up model (EBU)
The EBU model proposed by Spalding assumes high Reynolds and Damkohler numbers (Re  1,Da  1). The idea is to consider that turbulent motions control the
reaction rate and that chemistry does not play an explicit role. The reaction zone

5.3. TURBULENT PREMIXED FLAME MODELS

63

is viewed as a collection of fresh and burnt gas pockets transported by turbulent


eddies. The mean reaction rate is expressed as
= CEBU

2
g

(5.40)

EBU

2 are the temperature fluctuations


where CEBU is a model constant of order unity, g
and the turbulent time is estimated from the turbulent kinetic energy and its dissi2 can be made by
pation rate EBU = t = k . An estimation for the fluctuations g
assuming an infinitely thin flame and that the temperature can only take two values
= 0 or = 1 so that 2 = :

2

2 =

= 2 2 = 1

(5.41)

The reaction rate is written as a simple function of the mean quantities without
needing additional transport equations. It does not include any effects of chemical
kinetics (no pre-exponential constant B or activation temperature Ta appear in
(5.40)). Some models try to incorporate the chemical features in the model constant
CEBU or couple the EBU model to the Arrhenius law. In highly strained regions
where the ratio k is large the model tends to overestimate the reaction rate. The
estimation of the characteristic time EBU is quite arbitrary and could be modeled
using other assumptions leading to better predictive capacities (e.g. ITNFS [77]
model).
Bray Moss Libby model (BML)
The BML model combines a statistical approach using probability density functions
and a physical analysis. The pdf of the progress variable at a given location x is
expressed as the sum of fresh, fully burnt and burning gas contributions
p(, x) = (x)() + (x)(1 ) + (x)f (, x)

(5.42)

with , and the probability to have respectively fresh, burnt and reacting mixture
at location x. Assuming large Re and Da numbers the flame front is thin and the
probability to be in the burning gas is low  1. For an infinitely thin flame the
reduced temperature only takes two values = 0 and = 1 and thus
p() = () + (1 )

(5.43)

with + = 1. The mean reduced temperature is then given by


=
and

1
0

p()d = u + b = .0 + .1 =
= e = b = b

with b the density of the burnt gas and thus =


density is then given by

(5.44)
(5.45)

and = 1

= u + b = (1 )u + b

e
.
b

The mean
(5.46)

64

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

with u the density of the fresh gases. Defining the heat release factor as
=

Tb
u
1=
1
b
Tu

(5.47)

one can rewrite (5.46) as


e
u = (1 + )b = (1 + )

(5.48)

Combining equations (5.48) and (5.45) one can determine and


=

1 e
1 + e

and =

(1 + )e
1 + e

(5.49)

and thus the pdf is completely determined and depends only on the mean progress
variable e and the heat release factor . However the mean reaction rate cannot
be determined since the probability has been neglected 2 and another approach
is needed to model . Since the origin of the BML model different approaches
have been proposed to close the mean reaction rate. Three of them will be briefly
discussed.
Starting from the instantaneous balance equation for a balance equation for (1)
can be derived. Since under the BML model assumptions is either zero or unity one
has that (1 ) = 0 everywhere and the remaining terms in the balance equation
give after averaging
1

=
2D
2m 1
xi xi

2m 1

(5.50)

where m characterizes the chemical reaction


R

1

f ()d
m =
= R0 1

f ()d
0

(5.51)

f describes the turbulent mixing. A transport equation


and the scalar dissipation
f [78] or a simple model based on the fluctuations of the progress
can be derived for
variable (similar to the model for of equation (4.58) for non-premixed combustion)
f =

2
g
t

(5.52)

2 a balance equation may be derived equivalent to the transport equation for


For g
g
z 2 in appendix A or assuming intermittency between fresh and burnt gas equation
(5.41) can be used. Then the mean reaction rate can be written as

1
1e
e
)
(1
2m 1 t

(5.53)

and the expression of the EBU model is recovered incorporating chemistry effects
through m .
2

since (x) =

R1
0

()p(,

x)d = (x)

R1
0

()f

(, x)d 0

5.3. TURBULENT PREMIXED FLAME MODELS

65

A second model is based on the dependence of the mean reaction rate on the passage
frequency of the flame front. If one considers the turbulent flame brush - which is
the spatial zone wherein the instantaneous flame front is moving in time - a location
close to the fresh gas side as well as a location close to the burnt gas side exhibit
few passages of the flame front. Thus in both locations the mean reaction rate is
low. However the mean temperature in the point close to the burnt gas is much
higher ( 1) than the point close to the fresh gas ( 0) and so an Arrhenius
based model like equation (5.39) based on the mean progress variable would yield
completely wrong mean reaction rates. The mean reaction rate can be expressed
as the product of the flame crossing frequency fc and the reaction rate per flame
crossing c
0 s0 2(1 )
(5.54)
= c fc = 0 L
L /tt
Tb
where Tb is the mean period of the telegraphic temperature signal ( = 0 or = 1)
that can be estimated from the turbulent time scale. The flame transit time tt
measures the time to cross a flow front ( : 0 7 1) and can be estimated from
the timescale L0 /s0L . Estimating the reaction rate per flame crossing is difficult in
practice (while fc can be measured relatively easy) and the model can better be
rewritten in terms of the flame surface density:
= 0 hsc is

(5.55)

where is the flame surface density (i.e. the available flame surface area per unit
volume) and hsc is is the average flame consumption speed along the surface. The
main advantage of this approach is that the complex chemistry effects (incorporated
in hsc is ) are being separated from the interaction between turbulence and combustion (modeled through ). The turbulent flame can be viewed as an ensemble of
small laminar flame elements (flamelets) having a structure similar to a laminar
stagnation point flame. The consumption rate sc of these flamelets can be computed including complex chemistry effects using a simple model of a laminar planar
stagnation point flame. The result can be stored in flamelet libraries tabulating sc
as a function of the flame stretch . The mean consumption speed is then given by
hsc is =

sc ()p()d

(5.56)

with p() the probability to have stretch on the flame surface. Often one assumes
p() = ( ) with the mean local stretch generally estimated as the inverse
of the turbulent time scale = 1t = k and thus hsc is sc () = I0 s0L where I0
is the stretch factor introduced by Bray [79]. From DNS it can be shown that I0
is of the order of unity and therefore strain effects are neglected in some models
setting hsc is = s0L . The challenge is then to provide an adequate model for the
flame surface density (FSD). This can be done using simple algebraic expressions or
solving a balance equation [80, 81, 34, 76, 75]. Models based on flame crossing [82]
and fractal theories have been applied [83, 84] and different closures for the balance
equation of have been proposed [85, 86, 87, 88].

66

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

Probability Density Function model (PDF)


Introducing a probability density function p( ) measuring the probability that the
reduced temperature takes values between and + d . The information contained in a pdf can be used to extract for example the mean temperature and its
variance
Z 1
Z 1
2
(5.57)
p( )d
=
02 =
p( )d
0

and for an adiabatic single reaction with unity Lewis number the mean reaction rate
is given by
Z
=

( )p( )d

(5.58)

When Lewis numbers are not unity or complex chemistry has to be taken into
account these expressions can be easily extended by considering a joint probability
p (1 , 2 , . . . , N ) where 1 , 2 , . . . , N are thermochemical variables such as mass
fractions and temperature. Any favre averaged quantity is then given by
fe =

1 ,2 ,...,N

f (1 , 2 , . . . , N ) pe (1 , 2 , . . . , N ) d1 , d2 , . . . , dN

(5.59)

where pe is the favre pdf pe = p. This stochastic description has many theoretical
advantages and pdfs provide all the required information to describe unsteady reacting flow fields. Quantities such as ( ) or f (1 , 2 , . . . , N ) are provided from
chemistry and laminar flame studies. The difficulty is to determine the pdf changing
at every point in the flow field. Two main descriptions have been used: to presume
the shape of the pdf or to solve a balance equation [48, 51]. In general the pdf
can take any shape and exhibit multiple extrema and contains information on the
mean value of the variable, its first moment (the variance) and all higher moments.
For many applications the pdf functions present common features and thus one can
presume the shape of the pdf parameterized by one or two parameters. The BML
model described previously is a presumed pdf model assuming a bimodal shape for
the pdf (only two flow states = 0 and = 1). More sophisticated shapes have been
used and the most popular is the -pdf which was already extensively described in
the context of non-premixed combustion (see chapter 4):
e
p()
= R1
0

a1 (1 )b1
a1 (1 )b1
(a + b) a1
=
=
(1 )b1
a1
b1
B(a,
b)
(a)(b)
(1 ) d

(5.60)

Besides the determination of the mean progress variable also its variance has to be
known to determine the pdf depending on parameters a and b
)

(1
a =
1
2
g

and b =

(1 )a

(5.61)

2 a balance equation may be derived equivalent to the transport equation


For g
2 in appendix A. The presumed pdf method yields good results when only
for zg
one parameter, such as the reduced temperature, is required to describe chemical
reaction. When more than one variable is needed for chemistry, constructing a

5.3. TURBULENT PREMIXED FLAME MODELS

67

Figure 5.6: 2D representation of the kinetic balance of the flame front and the G-field
presumed multi-dimensional pdf becomes more difficult. Sometimes one assumes
that variables are statistically independent and then the joint pdf can be split as
a product of single variable pdfs : p (1 , 2 , . . . , N ) p(1 )p(2 ) . . . p(N ). This
assumption does not hold in practical situations because species mass fractions and
temperature are closely related and thus not statistically independent.
An exact balance equation for a multi-species, mass-weighted probability density
function pe (1 , 2 , . . . , N ) can be derived. This derivation is beyond the scope of
this work and the reader is referred to [51, 48, 80] for a more elaborate description.
Pdf balance equations are not solved directly and generally Monte-Carlo methods
are used where stochastic fluid particles are introduced to describe the chemical
composition. Although this method is general and powerful, its application to practical cases remains difficult and time consuming.
The G-field equation
In this approach the flame surface is described as an infinitely thin propagating
surface (flamelet). If chemistry is sufficiently fast, it occurs in layers that are thin
compared to the length scales of the flow. The key idea is to track the position of
the flame front using a field variable G. The field equation for G does not explicitly
contain a chemical source term and can be derived starting from the kinetic balance
for the laminar flame front :
x~F
= ~u + ~n.sL
(5.62)
t
with on the left hand side the propagating velocity of the front v~F , and on the right
hand side the flow velocity ~u and the burning velocity normal to the front sL (see
figure 5.6). The normal vector depends on the gradient of the G-field :
~n =

G
|G|

(5.63)

There is an iso-scalar surface G(~x, t) = G0 that divides the flow field in two main
regions where G > G0 is the region of burnt gas and G < G0 that of the unburnt

68

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

mixture. Differentiating with respect to t

G
~x

+ G
=0
t
t G=G0

(5.64)

and substituting this expression together with the definition of the normal vector in
equation (5.62) leads to the kinematic G-equation :
G
+ ~uG = sL .|G|
(5.65)
t
This equation contains a convective term on the lhs and an Eikonal 3 term with
the burning velocity on the rhs but no diffusion term. Although G represents an
arbitrary scalar, it is convenient to interpret it as the distance from the flame front
by imposing |G| = 1 for G 6= G0 . Introducing the following expression for the
laminar burning velocity
sL = s0L s0L L LS
(5.66)
with s0L the laminar burning velocity of the unstretched flame, L the Markstein
Length (of the order of the flame thickness), the flame front curvature
"

G
= .~n =
|G|

!#



(5.67)
G=G0

and S the strain rate defined by the velocity gradient S = (~n.~u.~n)|G=G0 . The
transport equation for the instantaneous level set G is then given by
G
+ .(~uG) = s0L DL LS
(5.68)
t
where DL = s0L L is the Markstein diffusivity and = |G|. To leading order
the curvature and strain term can be neglected. For turbulent combustion, making
several assumptions, a set of closed transport equations for the favre mean flame
front location and its variance can be derived [89, 90]:
e
G
e G)
e = (s0 )|G|
e D
e
+ .(~u
t e |G|
T
t

(5.69)

g
2
eg
G
g
g
eG
e 2 c G
2 ) = 2D (G)
2 )
2 + .(D G
(5.70)
+ .(~u
t
e
t k
k
t
k
e is the curvature of the mean flame front, c a model constant in the kinewhere
matic restoration term reducing fluctuations of the instantaneous flame front (playing a similar role as the scalar dissipation for diffusive scalars) and k denotes the
gradients in the mean flame front tangential direction.
A similar equation can be derived in the thin reaction zone regime where the smallest eddies (of the Kolmogorov size) can enter in the preheat zone but are too large
to influence the chemical reactions taking place in the inner layer. Since the inner
layer which is responsible for sustaining the reaction processes is still laminar, a level
set formulation for an iso-temperature surface can be derived.
3

The Eikonal equation is given by |u(x)| =


N

PN  u 2
i=1

xi

= F (x)

x where in an open

set in < Physically, the solution u(x) is the shortest time needed to travel from the boundary
to x inside , with F (x) being the speed at x (http://en.wikipedia.org/wiki/Eikonal equation
and http://mathworld.wolfram.com/EikonalEquation.html)

5.3. TURBULENT PREMIXED FLAME MODELS

69

Figure 5.7: Comparison between premixed flame thickness L0 and LES mesh size
. The flame front separates fresh and burnt gases.

5.3.2

LES Models

In the past decade Large Eddy Simulation (LES) has become a mature approach
for non-reacting flows. The capabilities of LES to predict unsteady turbulent flows
make the approach very well suited for reactive flows [91, 92]. The classical concept
of LES is to solve spatially filtered transport equations, by explicitly computing the
large structures of the flow field, whereas the effect of the smaller ones, the sub
grid scales, is modeled. This idea makes the computational cost acceptable for high
Reynolds number flows compared to DNS. Improved modeling concepts have made
it possible to perform Large Eddy Simulations of real combustion systems recently
[93, 94, 95, 96]. It turns out that LES provides an adequate description of turbulencecombustion interaction because the large structures are explicitly computed and
instantaneous fresh and burnt gas zones, where turbulence characteristics are quite
different, are clearly identified.
A difficult problem encountered in LES of premixed flames is that the typical flame
thickness L0 is generally much smaller than the mesh size as shown in figure 5.7.
Thus the flame front cannot be resolved on the computational mesh leading to
numerical problems. To overcome this, several approaches have been described in
the literature: the flame front tracking technique (G-equation) [97], the use of a filter
larger than the grid resolution [98] (leading to a flame surface density formulation)
and the artificial thickening of the flame [93, 99, 100]. Some concepts encountered
in the description of the RANS models have been extended to LES.
The G-equation in LES context
The G-equation formalism, viewing the flame as an infinitely thin propagating surface tracked by a field variable G, as described in the RANS context, can be extended
e = G and the
to LES. The resolved flame brush is then related to the iso-level G
0
filtered G-equation is written as [101, 102]
e
e
fi G
u
G
+
= (0 sT )|G|
t
xi

(5.71)

70

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

The challenge is then to find a model for the subgrid scale turbulent flame speed sT
generally based on a closure using the subgrid scale turbulence level u0 according to
equation (5.27) are used
!n
sT
u0
(5.72)
=1+ 0
s0L
sL
q

e = 2g
where one can estimate u0 |S|
Sij g
Sij . The constants and n have to be
specified or dynamically determined [103]. Equation (5.71) corresponds to a simple
physical analysis, displacement of the resolved flame front with displacement speed
sT , but the turbulent flame speed is not a well-defined quantity and no universal
model is available. The G-equation is also known for numerical difficulties and
induces flame cusps, generally avoided by adding a diffusive term in equation (5.71).
Nevertheless the model is widely used and has been successfully applied to several
practical applications [96, 102, 104]. Smith and Menon [105] use the G-equation in
combination with their linear eddy model (LEM) for LES. An advantage of the LEM
model is that it handles scalar transport in a Lagrangian manner, circumventing
problems of unphysical oscillations arising in finite volume methods as a result of
the steep gradients in flow properties caused by the thin flames [106].

Flame surface density LES formulation


This model was proposed by Boger et al. [98]. The flame front is too thin to be
resolved on an LES mesh but by filtering the progress variable equation with a filter
larger than the mesh size, the filtered flame front can be numerically resolved. The
turbulent dynamic premixed flame is smeared out sufficiently to be resolved on the
LES grid. Topological features of the flame with sufficiently large length scales are
resolved and SGS topological features are absorbed in SGS models. Assuming a
single step reaction and unity Lewis number the balance equation for the progress
variable c is given by
c
+ .(uc) = .(Dc) + c = w|c|
(5.73)
t
where w is the displacement speed of the iso-surface c. Applying the LES filter to
this equation yields

with

f u
c)) = w|c|
+ (
uc) + ((uc
t
w|c| = .(Dc) + c = hwis |c| = hwis

(5.74)
(5.75)

with hwis , , respectively the surface-averaged mass-weighted displacement


speed, the SGS flame surface density (i.e. subgrid scale flame surface per unit
volume) and the flame front wrinkling factor (i.e. the projection of the subgrid scale
surface in the propagation direction). The surface-averaged displacement speed can
be approximated by hwis u SL whereas or can be found from algebraic
relations or balance equations. Boger & Veynante [98] propose
=4

6 c(1 c)

(5.76)

5.3. TURBULENT PREMIXED FLAME MODELS

71

with assumed constant in a first step (usually between 1.0 and 1.5) or estimated
from the subgrid turbulence velocity u0 as in the G-equation context (see also next
section). In [107] it is argued that expression (5.76) is insufficiently precise and a
dynamic formulation is presented for a coupled fractal-similarity model for the SGS
flame surface density.
If a filtered transport equation is solved for the flame surface density, the modeling
complexity is significantly increased. Starting from equation (5.73) an exact but
unclosed balance equation for |c| can be derived and since = |c| the balance
equation for is obtained after LES filtering

+ . [huis ] + [hwnis ] = h.u nn : .uis + hw.nis


t

(5.77)

The LHS terms in this equation correspond, respectively, to unsteady effects, convection by the flow field (this term may be split into contributions from the resolved and the unresolved flow motions) and laminar flame propagation. The RHS
terms denotes the evolution of flame area because of the strain rate induced by
the flow field and combined propagation and diffusion effects. In [108] the effect of
the different terms in the the balance equation for are investigated for a simple
three-dimensional flame propagation test problem. An LES FSD-PDF model with
tabulated chemistry was used in [109] to study a ducted flame.
Artificially Thickened Flame Model
The artificially thickened flame (ATF) model, first introduced by Butler et al [110]
is very attractive due to its inherent numerical stabilizing features. The basic idea
of the approach is to consider a flame thicker than the actual one, with the same
laminar flame speed, so that the thickened flame front can be resolved on the computational mesh. The major drawback of this approach is the reduction of the
Damkohler number, altering the interaction between the turbulence and the flame.
To counteract for the reduction of the flame surface wrinkling a Subgrid Scale Wrinkling Model is introduced. To this day different model approximations have been
proposed in the literature. A large number of reported works concentrate rather on
a priori and a posteriori tests of the ATF model using DNS data [99, 111, 112] of
relative low Reynolds numbers.
To describe the turbulent premixed combustion the ATF flame model start from
the balance equation for the progress variable (representing the non-dimensional
u
temperature, = TTbT
, where subscripts b and u respectively denote burnt and
Tu
unburnt quantities). Note that for laminar premixed flames the flame speed s0L and
flame thickness L0 may be expressed as
s0L

L0

D
s0L

(5.78)

where D is the molecular diffusivity and the reaction rate. In the frame of the ATF
model, the thickening of the flame is achieved by decreasing the reaction rate with
a factor F whereas the diffusion is increased by the same factor F . This operation
leaves the laminar flame speed s0L unchanged but increases the flame thickness i.e.

72

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

L1 = F L0 so it can be resolved on the LES computational mesh. The progress


variable can then be described by [99]
ui


+
=
F D
t
xi
xi
xi


.
F

(5.79)

The source term for the progress variable is modeled using Arrhenius law so various
phenomena like ignition, flame-stabilization and flame wall interaction are taken
into account
!


Ta

(5.80)
= A 1 exp
b Tu )
Tu + (T

In (5.80), A is the pre-exponential constant (taken independent of the temperature)


and Ta the activation temperature. Thickening the flame modifies the interaction between turbulence and chemistry because the Damkohler number Da which describes
the relation between turbulent and chemical time scales
Da =

t
lt s 0
= 0 L0
c
u L

(5.81)

is reduced by a factor F . To compensate for this reduction of flame surface wrinkling


due to the thickening, an efficiency function E is introduced [113, 99]. This efficiency
function measures the subgrid scale wrinkling as a function of the local subgrid
turbulent velocity u0e and a length scale e . The diffusion coefficient in (5.79) is
replaced by the product EDF while the pre-exponential constant is replaced by
AE/F so that the conservation equation for the progress variable now reads
ui

EF D
+
=
t
xi
xi
xi


AE 
Ta
+
1 exp

F
Tu + (Tb Tu )

(5.82)

Thus the turbulent flame propagates at a turbulent flame speed sT = Es0L (where
E goes to unity in laminar regions) and keeps a thickness of the order of L1 = F L0 .
Since the efficiency function describes the underestimation of the actual wrinkling of
the flame front one can express E as the ratio between the subgrid wrinkling factors
of the actual flame and the thickened flame.
Consider a box of size (the grid size) inside which a thin flame of thickness L0
displays unresolved wrinkles down to an inner cut-off scale c as shown in figure 5.8.
This scale can be defined as the inverse mean curvature of the flame
c = |h.nis |1

(5.83)

and its ratio to L0 (the laminar flame thickness), (the Kolmogorov length scale)
and G (the Gibson length scale (u0 /s0L )3 corresponding to the vortices in the
Kolmogorov scale having speed s0L ) depends on the turbulence and flame properties.
The subgrid turbulent flame speed can then be determined from the total flame
surface area (as in equation 5.25):
ASGS
s T
=
=
0
sL
2

(5.84)

5.3. TURBULENT PREMIXED FLAME MODELS

73

Figure 5.8: Sub Grid Scale wrinkling of the flame front


The subgrid scale wrinkling can be written in terms of a power-law depending
on the inner (c ) and outer () cut-off scales:

= 1 +
c

(5.85)

and thus

s T
= (1 + |h.nis |)
(5.86)
s0L
In general can depend on scales , c or parameters like the Reynolds number.
If the flame front is not wrinkled |h.nis | = 0 and c and thus = 1 and
sT = s0L i.e. the laminar flame speed is recovered. If is assumed to be independent
of scales and 0 < < 1 one recovers the fractal models [83] with = D 2, D
being the fractal dimension of the flame. In [99] a space-filling flame is assumed
(D=3) and the subgrid averaged curvature is estimated as
|h.nis |

1 1
e

(5.87)

with e a filter size and a parameter to be determined. Relations (5.85) and


(5.87) are equivalent for = 1 and = 1. Recalling the balance equation for the
SGS flame surface density (5.77) and assuming an equilibrium between production
and destruction of SGS flame surface density one obtains [113, 99, 112]
h.u
nn
: .u}is = hw.nis
|
{z

(5.88)

aT

with aT the strain rate induced by the flow field and acting on the flame front. This
simplification assumes that the time scales associated with the SGS flame dynamics
are much shorter than the unsteadiness in the resolved field and that the flame
can adjust instantaneously to the imposed fluctuations. In transient conditions this
assumption is not valid and either the full balance equation (5.77) has to be solved or
dynamic models can be used [114]. The destruction term in (5.88) is approximated
by setting the flame front displacement speed equal to the unstrained laminar flame
speed
hw.nis s0L |h.nis |
(5.89)

74

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

Using this expression and estimating the subgrid strain rate haT is from the ratio
of the subgrid scale turbulent velocity u0e (i.e. the square root of the SGS kinetic
energy on the filter size level) and the filter size e one can rewrite (5.87) as
1 +

u0e
e
ha
i
=
1
+

T s
s0L
s0L

(5.90)

recovering an expression similar to (5.27) and (5.72). To take into account the net
straining effect of all relevant turbulent scales smaller than e the subgrid strain
rate haT is is written as
haT is

e /L , u0e /s0L , Re

u0e
e

(5.91)

where the function corresponds to the integration of the effective strain rate induced by all scales affected by the artificial thickening and L is the actual thickness
of the flame. The function obtained from DNS results of flame-vortex interaction
experiments [113, 99, 115] is written as

e /L , u0e /s0L

= 0.75

exp


1.2
u0e /s0L

0.3

e
L

2/3

(5.92)

Substituting this expression in (5.91) and (5.90) the wrinkling of the subgrid scale
flame front is given by


1 + e /L , u0e /s0L

u0e
s0L

(5.93)

With a thickened flame L1 = F L0 , the actual wrinkling of the flame front is


underestimated by an efficiency factor E being the ratio of the wrinkling factor of
the real flame and the thickened flame:
E=

(L0 )
(L1 )

1 + e /L0 , u0e /s0L u0e /s0L




1 + e /L1 , u0e /s0L u0e /s0L


1/2

is a model parameter [99] that scales as Ret


= 0

2 ln(2)

1/2

3cms Ret

(5.94)

:
(5.95)

where 0 is a model constant of the order of unity and cms = 0.28 is a model constant
from the DNS results of Yeung et al. [116, 99, 115, 112]. Since it is difficult and
costly to determine the local Ret in LES a constant value for is used in this work.
Figure 5.9 shows the function and the efficiency function as given by (5.92) and
(5.94).
The subgrid scale velocity u0e in (5.94) remains to be estimated. This can be
achieved through the subgrid scale turbulent viscosity provided by the flow-field
sgs-model
q
t
= CS e 2Sij Sij
(5.96)
u0e =
C S e

5.3. TURBULENT PREMIXED FLAME MODELS

75

,
Figure 5.9: -function and efficiency function E with = 0.2
However, there are two problems with this formulation. Firstly the model constant
is defined by Smagorinsky to obtain the right amount of kinetic energy dissipation
and not to obtain the right SGS turbulent velocity at any scale e > . To obtain
a better estimate for
r the SGS velocity at scale e one could use (5.96) at the filter
g g

level u0e = CS e 2Sije Sije . The second problem is that in the local absence of
q

turbulence the laminar flame speed s0L is not recovered since the strain 2Sij Sij is
influenced by the thermal expansion. In Colin et al [99] it is therefore proposed to
use an operator which naturally subtracts the dilatational part of the velocity field
h

u0e = 23x 2 ( u)

(5.97)

In this work the expressions given in (5.96) and (5.97) will be used and their influence
on the prediction of the flow and field and the flame structure will be examined.

76

CHAPTER 5. TURBULENT PREMIXED COMBUSTION

Chapter 6
Numerical Methods
6.1

Introduction

Two different numerical codes are used in the current work. For the simulation of
turbulent non-premixed flames a complete 3D LES code was developed at the department as well as several tools to generate flame tables and inflow boundary conditions.
This code will be described in detail in the next section. For the simulation of premixed turbulent combustion the artificially thickened flame model was implemented
in a research code named FLOWSI developed at the TU Darmstadt1 . The main
numerical aspects of this code are briefly discussed in the last section of this chapter
and for a more detailed description the reader is referred to several PhD-dissertations
[117, 118, 119, 102] and publications [120, 121, 122, 123, 124, 125, 126, 127].

6.2
6.2.1

Non premixed Combustion Solver


Finite Volume Formulation

To solve the partial differential equations describing the physics of the flow the problem must be split into elements of finite extension in space and time. In this work
the Finite Volume method is used, discretizing the entire region of interest (the computational domain) in finite volume cells. The governing equations are integrated
in each of these control volumes. In this section the finite volume discretization is
explained for a general balance equation for the scalar

uj
Dj ()
+
=
+ S
|{z}
t
xj
xj
|{z}
| {z }
| {z }
source
accumulation convection diffusion

(6.1)

This equation describes the accumulation of the scalar at any point in space
x
at any time t. The scalar is convected by the velocity uj and diffused at rate Dj ().
The source term S describes the production or destruction of . Table 6.1 gives
1

Institute for Energy and Power Plant Technology (FG Energie- und Kraftwerkstechnik : EKT),
Technical University Darmstadt,Petersenstr. 30, 64287 Darmstadt, Germany, http://www.ekt.tudarmstadt.de/home.php

77

78

CHAPTER 6. NUMERICAL METHODS

the expression for , Dj () and S for the balance equations from chapter 4 that
have to be solved to describe turbulent non-premixed combustion.
Equation

Dj ()

Continuity

Momentum

ui

ij

p
+ fi
x
i

Mixture Fraction

z
D x
j

Sc

Mixture Fraction Variance

z 2

Turbulent Kinetic Energy

Turbulent Energy Dissipation

z 2
xj

t
k

k
xj

xj

t
Sct

Pz
P
1
T

(C1 P C2 )

Table 6.1: Expressions for , Dj (), S for the different RANS and LES equation
to be solved
Integrating the equation for the scalar over a volume V (e.g. the quadrilateral
ABCD of figure 6.1) of arbitrary shape one obtains
Z

dV +
t

uj
dV =
xj

Dj ()
dV +
xj

S dV

(6.2)

Using the Gauss-theorem the volume integrals of the convection and diffusion terms
can be transformed into surface integrals
Z

dV =
xj

S=V

nj dS

(6.3)

where dS is a surface element and nj the outward unit vector normal to the surface.
Bringing the time derivative out of the integral one can write the integral form of
the general transport equation for

Z

dV

uj nj dS =

Dj () nj dS +

S dV

(6.4)

Using a structured grid with hexahedral cells the control volumes are surrounded
by six neighboring cells in 3D (four in 2D). Assuming that the values for and S

6.2. NON PREMIXED COMBUSTION SOLVER

79

Figure 6.1: Structured 2D Finite Volume Mesh with Cell Centers and Vertices
are constant over the cell and using a cell centered approach the integral form of the
balance equation in discretized form can be written as

V +

faces

Ff,C =

Ff,D + SP V

(6.5)

faces

where Ff,C and Ff,D are respectively the convective and diffusive fluxes across the
face f given by
Ff,C = (uj )f nj Sf
Ff,D = (Dj ())f nj Sf

(6.6)
(6.7)

where the diffusion term depends proportional on the gradient of with a corresponding diffusion coefficient

Dj () = D
(6.8)
xj

6.2.2

Convective Terms

The convective terms physically correspond to the propagation of information in


certain predefined directions and this property should be reflected somehow into the
scheme. A straightforward expression for the convective flux through face f = BC
between cells P and Q on figure 6.1 is


F C
n

BC

1 


n
F CP + F CQ
2

(6.9)

This central discretization is second order accurate, computationally efficient and


simple to implement. However although the accuracy is high the truncation errors
introduce oscillations. If the diffusion is small, these oscillations are not dissipated
and cause instability. Hence this method is hardly suited for scalar transport especially when the scalar has to remain bounded as is the case for example with the

80

CHAPTER 6. NUMERICAL METHODS

mixture fraction = z. Values outside the allowed interval [0, 1] have no physical meaning. Clipping the value of mixture fraction to its allowed limits is not an
appropriate solution since oscillations remain and further nonlinearities are introduced. For the momentum equations where = ui the presence of a turbulent
viscosity in the diffusive terms can be sufficient to suppress the oscillations. If this
is not the case a dissipation term can be added to stabilize the scheme. The central
discretization with artificial dissipation can then be written in 1D as


F C
n

i+ 21





1 


n
di+ 1
F C
n + F C
2
i+1
i
2

(6.10)

Most schemes were originally developed for a 1D equation and further in this
section the 1D scalar convection equation will be used to explain the schemes. The
extension to 3D is mostly based on an ad-hoc treatment by applying the scheme in
the 3 directions. This implies that the performance of the schemes will depend on
the quality of the mesh.
Consider the 1D scalar convection equation (convection speed a)
u au
+
= 0 with flux f = au
t
x

(6.11)

Assuming a Euler backward scheme for the time discretization, the discretization
for cell i in a cell centered control volume method is given by
un+1
= uni
i

i
t h
fi+ 1 fi 1
2
2
x

(6.12)

A central discretization is obtained with


fi+ 1 =
2

1
[(au)i+1 + (au)i ]
2

(6.13)

A Von Neumann stability analysis learns that this scheme is not stable and in order
to stabilize the scheme some artificial dissipation has to be added. Instead of solving
the original equation (6.11) one then solves
u au
2u
+
= x 2
t
x
x

(6.14)

The term in the right hand side does not change the order of the PDE but introduces
an error term of order 1 making the resulting scheme only first order accurate. The
resulting scheme is (assuming a constant)
un+1
= uni
i

t
t
a (ui+1 + ui ) +
(ui+1 2ui + ui1 )
2x
x

(6.15)

which can also be written in the form of equation (6.12) with a numerical flux given
by
a

(ui+1 + ui ) (ui+1 ui ) = fi+ 1 (ui+1 ui )


(6.16)
fi+
1 =
2
2
2

6.2. NON PREMIXED COMBUSTION SOLVER

81

In order not to destroy the second order accuracy of the original scheme one can
add artificial dissipation under the form of a fourth order derivative
4u
u au
+
= x3 4
t
x
x

(6.17)

The resulting numerical flux is given by

fi+
1 = fi+ 1 + (ui+2 3ui+1 + 3ui ui1 )
2

(6.18)

This type of fourth order dissipation is for example used for the momentum equations
in the RANS simulations of the current work. For more details the reader is referred
to [128, 129, 130]. The scalar convection equation represents a wave propagation in
positive x-direction if a > 0 or negative x-direction if a < 0. This means that if
a > 0 the solution at a point i is only influenced by upwind points i 1, i 2,...
This physical interpretation shows that the central scheme is a bad choice (and thus
unstable unless artificial dissipation is added) and the use of upwind schemes for
convective problems is more appropriate. Assuming a > 0 the 1st order upwind
discretization is given by
un+1
= uni
i

at
(ui ui1 )
2x

(6.19)

which can be written also as a central scheme plus a correction term


un+1
= uni
i

at
at
(ui+1 ui1 ) +
(ui+1 2ui + ui1 )
2x
2x

(6.20)

Comparing this expression with (6.15) shows that the first order upwind scheme
can be interpreted as a central scheme with artificial dissipation coefficient given by
= a2 . The numerical flux is

fi+
1 =
2

1
1
(fi + fi+1 ) |a| (ui+1 ui )
2
2

(6.21)

This scheme is only first order accurate in space and a higher accuracy can be
achieved by involving more points. Equation (6.21) is then rewritten as

fi+
1 =
2



1 L
1 
L
R
fi+1/2 + fi+1/2
|a| uR
i+1/2 ui+1/2
2
2

(6.22)

L
where uR
i+1/2 and ui+1/2 are more accurate representations of the solution left and
right of the cell face than respectively ui+1 and ui that can be obtained by linear
interpolation of the solution left and right of the cell face:

uLi+1/2 =

ui + 21 (ui ui1 )

1
uR
i+1/2 = ui+1 + 2 (ui+1 ui+2 )

(6.23)

This technique of extrapolation of the solution to increase the accuracy of the scheme
is known as the MUSCL approach introduced by van Leer [131]. However the higherorder upwind schemes, just like the central scheme, are not monotonic. A scheme is

82

CHAPTER 6. NUMERICAL METHODS

said to be monotonicity preserving if no new local extrema are created and the value
of a local minimum/maximum is non-decreasing/non-increasing in time. According
to Godunovs theorem linear schemes for convection equations can only be monotone
if they are first order accurate. This non-monotone behavior is clearly unwanted if
the scheme is to be used for a bounded scalar such as the mixture fraction. The key
word in Godunovs theorem is linear : building in some non-linearity in the scheme
can make a high order scheme monotonic. Such non-linearity is provided by limiters:
the solution is locally monitored and as soon as there is tendency to non-monotonic
behavior the limiter will react. The action of the limiter will basically be to reduce
locally the order of the scheme such that it will become monotone again. Since this
is a local action the scheme will formally still keep its higher order accuracy. Van
Leer [131] shows that the MUSCL scheme will behave non-monotonic as soon as one
L
of the extrapolated values uR
i+1/2 and ui+1/2 is either smaller than min(ui , ui+1 ) or
greater than max(ui , ui+1 ). Another approach to derive monotonicity conditions is
due to Harten [132] who introduces the Total Variation Diminishing concept (TVD).
Defining the total variation TV of a discrete solution to a scalar conservation law
as
X
T V (u)
|ui+1 ui |
(6.24)
i

a numerical scheme is said to be Total Variation Diminishing if


T V (un+1 ) T V (un )

(6.25)

All monotone schemes are TVD and all TVD schemes are monotonicity preserving.
To introduce the limiters one can rewrite the higher order upwind scheme in a
flux difference notation as
uLi+1/2 =

+
(ui+1 ui )
ui + 21 ri+1/2

1
uR
i+1/2 = ui+1 2 ri+1/2 (ui+1 ui )

with

ui ui1
ui+1 ui
The limiter is then introduced
+
ri+1/2
=

uLi+1/2 =

ri+1/2
=

ui+2 ui+1
ui+1 ui

(6.26)
(6.27)

+
(ui+1 ui )
ui + 21 ri+1/2

1
uR
i+1/2 = ui+1 2 ri+1/2 (ui+1 ui )

(6.28)

Substitution of these expressions in the expression of the higher order upwind flux
and assuming a constant, one obtains the following numerical flux

fi+
1 =
2




i
1
1h
+

+ a ri+1/2
ui+1/2
(fi + fi+1 ) |a| a+ ri+1/2
2
2

(6.29)

1
where ui+1/2 = ui+1 ui and a = (a |a|) or more compactly written
2

fi+
1 =
2


i
1
1 h
(fi + fi+1 ) |a| 1 ri+1/2 ui+1/2
2
2

(6.30)

6.2. NON PREMIXED COMBUSTION SOLVER


Limiter
MINMOD
VAN LEER
VAN ALBADA
CHARM

OSPRE
SMART
SUPERBEE

83

(r)
max(0, min(r, 1))
2r
if r 0
1+r
0 if r < 0
r(r + 1)
if r 0
r2 + 1
0 if r < 0
r(3r + 1)
if r 0
(r + 1)2
0 if r < 0
1.5r(r + 1)
if r 0
r2 + r + 1
0 if r < 0
max(0, min(2r, 0.75r + 0.25, 4))
max(0, min(2r, 1), min(r, 2))

Table 6.2: Limiter Functions


2
1.8
1.6

(r)

1.4
1.2
1

MinMod
Van Leer
Van Albada
Charm
Ospre
Smart
SuperBee

0.8
0.6
0.4
0.2
0
1

Figure 6.2: Limiter functions in the Sweby diagram


with r = r + if a > 0 and r = r if a < 0. Depending on the ratio r the extrapolation
in (6.26) will violate monotonicity or not. Several limiters have been introduced in
literature [131, 133, 129, 134] and in table 6.2 some of the most encountered limiters
have been summarized and the behavior of the limiters is given in the Sweby [135]
diagram in figure 6.2. As a test a top-hat function has been transported with the
above mentioned schemes and results are shown in 6.3.
Note that in the current FDS formulation the CHARM and SMART limiter
are not monotone. The Superbee limiter (biggest possible stable (r)) seems to
perform the best but it tends to sharpen gradients and is thus inapplicable for
smoother functions. The reason that the CHARM and SMART limiter do not
preserve monotonicity in the FDS formulation is that they violate the monotonicitycriterion: 0 (r) 2. This can be clearly observed in the Sweby diagram of figure
6.2. All other mentioned limiters do respect the condition. The numerical flux can

84

CHAPTER 6. NUMERICAL METHODS


1st O UPWIND
EXACT
CENTRAL

0.8

0.6

0.6

Exact
Van Leer
Charm
MinMod
Exact
Van Albada
Smart
SuperBee

0.8

0.4

0.4

0.2

0.2

0.2

0.4

0.6

0.8

0.2

0.4

0.6

0.8

Figure 6.3: Transport of a top-hat function with different schemes: (left) Central
and 1st order upwind scheme, (right) Higher order upwind scheme with different
limiters in FDS formulation of (6.30).
be rewritten in another form (see [129])

fi+
1 =
2





1
1
1
1
+

(fi + fi+1 ) |a|i+1/2 + a+ Ri+1/2


ui1/2 a Ri+1/2
ui+3/2
2
2
2
2
(6.31)

with
+
Ri+1/2
=

ui+1 ui
ui ui1

Ri+1/2
=

ui+1 ui
ui+2 ui+1

(6.32)

(R)
2. This
R
criterion is satisfied by the CHARM and SMART limiters. Expressions (6.30) and
 
1
(R)
=
which
(6.31) are identical only if the limiter is symmetric i.e. if
R
R
is not the case for the two above mentioned limiters. The top hat convection test
is repeated with the formulation of (6.31) and one can see in figure 6.4 that now
both the CHARM and SMART limiter do preserve monotonicity as well as the Van
Leer limiter which satisfies the symmetry condition and can thus be used in both
formulations.
and the monotonicity criterion for this formulation reads 0

6.2.3

Diffusive Terms

The diffusive or viscous flux is easier to determine than the convective flux since it
represents propagation into all directions simultaneously. Therefore a straightforward expression for the flux on a cell face f = BC between cells P and Q on figure
6.1 is given by

1 P
Dj () + DjQ () nj
(6.33)
(Dj ()nj )f =
2
This central scheme is in agreement with the physical property of the diffusive term
and thus does not cause problems as for the convective terms. But the viscous flux is
not only function of the unknowns but also of its gradients. Therefore the expression

6.2. NON PREMIXED COMBUSTION SOLVER

85

Figure 6.4: Transport of a top-hat function with higher order upwind scheme
with different limiters using the numerical flux of (6.31). Comparing VAN LEER,
CHARM and SMART limiter.
for the viscous flux is usually written as


(Dj ()nj )f = Dj f ,
nj

xj f

(6.34)

The cell face value of the scalar is determined from a central approximation
1
(P + Q )
(6.35)
2
and for the determination of the gradients on the cell faces different approaches are
possible. Gradients can be determined in the cell centers (e.g. P) by integrating the
gradients over the same control volume V=ABCD and applying the Gauss-theorem.
Then the cell face gradients can be interpolated from the cell center values (e.g.
1
BC = (P + Q ) ). The problem with this approach is that it causes
2
odd-even decoupling leading to convergence problems or oscillatory solutions. The
gradient on the face can also be determined by averaging the gradient values in the
1
cell vertices (e.g. BC = (B + C ) ) and the cell vertex values are then
2
determined by integrating over a control volume containing the vertex (e.g. the
volume PQRS for vertex C). But this approach suffers from the same problems as
the cell center approach. To avoid this problem gradients can be directly calculated
on the cell faces. The drawback of this approach is that the computation cost is
much higher since there are 3 times more cell faces then cell centers or vertexes in a
hexahedral mesh. The gradient on face BC is then determined by integrating over
the control volume abcd
Z
I
1
1
1

dS
(ab
n
dV =
BC =
ab + bc nbc + cd ncd + da nda )
Vabcd Vabcd
Vabcd
Vabcd
(6.36)
f =

86

CHAPTER 6. NUMERICAL METHODS

and to determine the scalars on the faces of the new control volume abcd the following
approximations are used
da = P

bc = Q

cd =

1
(P + Q + R + S )
4

(6.37)

To save computational cost the normals of the new control volume are approximated
from the normals of the mesh cells instead of exactly calculating them (which only
has to be done once but then a large amount of data would have to be stored)

n
bc =

n =

1
2
1
2

ab

6.2.4


(n
BC + nEF )

(
n
AB + nBE )

(6.38)

Time integration

The study of accuracy of time integration is beyond the scope of this thesis. Here
a brief description of time integration methods is given to complete the picture. In
all the simulations presented here the time step was sufficiently small to ensure that
the errors due to time integration are much smaller than those arising from spatial
discretization.
Equation (6.5) can be written more compactly as

V = Res
t

(6.39)

where the residual Res is given by


Res =

Ff,C +

faces

Ff,D + S V

(6.40)

faces

This equation can be treated as an ordinary differential equation that has to be


integrated in time (an initial value problem)
dU
= Res(U )
dt

(6.41)

where U denotes the vector of unknowns. To advance in time a multistage low


storage Runge-Kutta (RK) scheme was used
U1 =
U0 + tC1 Res(U0 )
U2 =
U0 + tC2 Res(U1 )
...
UN = U0 + tCN Res(UN 1 )

(6.42)

where U0 is the solution at time level tn and UN is the solution at the new time level
tn+1 = tn + tn . For linear problems with constant propagation coefficients an N
stage RK scheme is N-th order accurate provided the RK coefficients are chosen as
Ck =

1
N k+1

(6.43)

6.3. PRESSURE CORRECTION METHOD

87

where k is the RK stage number. For non-linear problems, however, the formal
accuracy of the scheme is only of order 2. Like the discretization in space the discretization in time affects the accuracy of the simulation. Just as smaller mesh cells
will improve the precision, smaller time steps will yield a more accurate solution
but at the expense of a more time consuming simulation. Since one is dealing with
unsteady flows and evaluates time averaged quantities a sufficiently long integration time is needed to capture also the low frequency phenomena. With explicit
time integration methods as used here the time step width is even more limited by
stability requirements. According to Courant, Friedrich and Lewy the convective
transport within one time step has to be limited so that information only travels
to the neighboring cell but not further. Neglecting acoustic phenomena (low Mach
number approach) the CFL-number can be written as
CF L =

t 




u .SI +
u .SJ +
u .S K
V

(6.44)

with S the normal surfaces expresssed in the cell center for the I,J and K direction
of the grid. For an explicit Euler time integration the CFL-number should be smaller
than 1. For a multistage RK-scheme the CFL-limit is slightly higher but difficult
to determine exactly in 3D. To advance in time as fast as possible - to obtain the
needed total integration time in a reasonable computational time - the time step is
always chosen as high as possible. In fact also the diffusive fluxes have to be limited
to ensure stability. A viscous CFL-number can be determined from


h
t 8ef f
2
2 2


i

CF LV IS =
S I + S J + S K + 2 S I . S J + S I .S K + S K . S J
V V
(6.45)
with ef f the effective viscosity (+t ). The allowed time step is then the minimum
of the inviscid and viscous time step.

6.3

Pressure correction method

Although in low Mach number flows pressure variations are typically small, density and temperature gradients can be large due to heat transfer and/or chemical
reactions. These flows can be treated by implicitly discretizing the compressible
governing equations in time and to add a pseudo-time derivative to the resulting
discrete system of equations [136]. The solution at the new time level is then a
steady state solution of the equations in pseudo-time and can be obtained with an
iterative method. For low speed flows several convergence-acceleration techniques
such as preconditioning, multigrid and residual smoothing can be used to increase
the efficiency [137], [138], [139, 140], [141]. Due to the overhead of the iterative
method that is necessary to solve the implicit, discretized system of equations, this
approach is still computationally expensive.
If the Mach number of the entire flow field is low the so called Low Mach Number
approximation can be used. In this approach the acoustic modes are decoupled from
the velocity and entropy modes. The methods are called fractional step or projection

88

CHAPTER 6. NUMERICAL METHODS

method, and were developed independently by Chorin [142] and Temam [143, 144].
In the projection method one takes the divergence of the momentum equation leading to a Poisson equation for pressure. In constant density (incompressible) flows
the velocity field is divergence free at any time level n (i.e. .un+1 = .un = 0)
but in a variable density flow the divergence of the momentum field at time level
n + 1 (.(un+1 )) appears in the right hand side of the Poisson equation. Generally
this term is replaced by the time derivative of density at time level n + 1 but the
discrete form of this term remains a controversial issue. When used in combination
with a collocated grid, the pressure correction method can produce spurious oscillations of the pressure field known in literature as odd-even decoupling. In appendix
C it will be clarified how this phenomenon is avoided for the collocated cell centered
arrangement used in the current work.
The fractional step method used in this work is slightly different from the original pressure correction method in a way that the pressure term is removed in the
intermediate step. For the sake of clarity, it is assumed that the time derivative is
discretized with an explicit forward Euler method. The extension to a multi-stage
Runge-Kutta method is straightforward as the same procedure is followed at every
stage.
The semi-discretized form of the momentum equation may be written as,
ijn
n uni unj
p
(ui )n+1 n uni
=
+
+
t
xj
xi
xj

(6.46)

Dropping the gradient of pressure from equation (6.46) an intermediate momentum


field ui is obtained,
n uni unj
ijn
(ui ) n uni
+
=
t
xj
xj

(6.47)

and the intermediate velocities can be written as


ui =

(ui )
n+1

(6.48)

The velocity field at time level n + 1 is then given by


p
(ui )n+1 (ui )
=
t
xi

(6.49)

where the pressure is chosen so that the continuity equation is satisfied at time level
n + 1 i.e.
!n+1

n+1 un+1
i
+
=0
(6.50)
t
xj
Taking the divergence of equation (6.49) one can write
t

n+1 un+1
ui
2p
i
=

xi xi
xi
xi

(6.51)

6.3. PRESSURE CORRECTION METHOD

89

and replacing the divergence of the momentum field at time level n + 1 by the
time derivative of density from equation (6.50) a Poisson equation for pressure is
obtained:
!n+1
2p
ui

t
=

(6.52)
xi xi
xi
t
So equation (6.47) is first integrated in time to obtain the intermediate momentum field and velocity field. Boundary conditions are applied for the intermediate
velocity field using the same expressions as for the final velocity field.2 Then an
approximation is chosen for the time derivative of density at time level n + 1 so that
the complete RHS of (6.52) is known and the constant coefficient Poisson equation
can be solved to determine the pressure field. Once the pressure field is calculated
the final velocity at time level n + 1 can be obtained from equation (6.53)
t p
(6.53)
n+1 xi
For the time derivative of density one can use a simple first order approximation
un+1
= ui
i

!n+1

n+1 n
t

(6.54)

Cook & Riley [146] reported that this time derivative was the most destabilizing
part in their computations. They also stated that odd-ordered approximations for
the density derivative were less stable than the even-ordered. However Charentenay
et al. [147] used the following third order backward-difference formulation

!n+1

11n+1 18n + 9n1 2n2


6t

(6.55)

for the simulation of a premixed ozone flame with density ratio up to 3.64 and did
not report any stability problems. Lessani & Papalexandris [148] use a second order
approximation for this derivative

!n+1

3n+1 4n + n1
2t

(6.56)

in combination with a two-stage predictor-corrector scheme using an Adams-Bashfort


method for the momentum equations. Pierce [149] applies a spatial filter to the
time derivative of density using the test-filter operator of the dynamic procedure
and states that this improves the stability but does not guarantee it completely.
Using an iterative semi-implicit scheme (treating only the stiffest terms implicitly)
he reports that for weak density variations only 2 or 3 iterations are needed while
large density ratios require more.
In this work both the first order derivative (6.54) and the second order derivative
(6.56) have been used in combination with the spatial filtering using the test filter
of the dynamic model [149].
2

In [145] it is argued that for the intermediate velocity field a more appropriate boundary
conditioned can be derived setting the non dimensional intermediate velocity on the boundary
n
where is a field related to the pressure field by p = +
according to ui = un+1
+
t
i
xi
2
t/(2 Re)

90

CHAPTER 6. NUMERICAL METHODS

Figure 6.5: Poisson equation stencil for cartesian mesh

6.3.1

Solution of the Poisson Equation

Effective solution of the Poisson equation for pressure is one of the most important
aspects of low-Mach number flow simulations. The most straightforward approach
is the use of a direct solver. This approach can be used for the investigated type of
flow problems for the following cases
For 2D-axisymmetric RANS simulations with rotational periodicity (only 1
cell in circumferential direction).
For 3D-LES simulations in a cylindrical grid system with a homogeneous circumferential direction.
For 3D-LES simulations in a cartesian grid system if at least one periodic
direction is available (e.g. when the geometry is essentially 2D)
If the flow is periodic in one or more directions and the mesh is uniform in those
directions, one can benefit by solving the Poisson equation partially or completely
in the Fourier space. Applying a (Fast) Fourier Transform to the right hand side of
the Poisson equation makes this procedure very efficient. The obvious advantage of
this method is that the discretized Poisson equation is solved exactly in the most
direct way. In general, however, if the flow does not have periodic features, the
direct solver becomes extremely expensive, both in terms of CPU and data storage.
In this case an iterative method is more efficient.
2D Direct Poisson Solver
Considering a cartesian non-uniform grid in 2D the discretized form of the Poisson
equation with the stencil given in figure 6.5 can be written as
ij pi1,j + ij pi+1,j + ij pi,j1 + ij pi,j+1 + ij pi,j = qi,j

(6.57)

where the coefficients depend on the local mesh. In case of a uniform mesh the

6.3. PRESSURE CORRECTION METHOD

91

coefficients are constant and given by


ij =

1
1
1
2
2
1
, ij =
, ij =
, ij =
, ij = 2
(6.58)
2
2
2
2
x
x
y
y
x
y 2

The system of equations can written in matrix form as


Ai Pi1 + Bi Pi + Ci Pi+1 = Qi

(6.59)

with Pi and Qi the following vectors (i = 0, 1, . . . , N I)


Pi = (p0 p1 . . . pN J )Ti

Qi = (q0 q1 . . . qN J )Ti

(6.60)

and Ai and Ci are diagonal matrices

Ai =

0 0 ...
0 ij
..
..
.
.

ij
0

..
.
0
0

and Bi is a tridiagonal matrix

Ci =

Bi =

ij ij ij

..
.
0
0

0 0 ...
0 ij
..
..
.
.

ij ij
.. .. ..
.
.
.

...

ij
0

...

0
0
..
.

ij ij 0
ij ij ij
0

..
.
0
0

(6.61)

(6.62)

The two elements in the upper left and lower right corner of matrix Bi depend on
the boundary conditions for pressure at i = 0 and i = N I. If one has a Dirichlet
boundary condition at i = 0 the two elements in the upper left corner are 1 and 1
while in case of Neumann boundary conditions they are 1 and 1. Also the elements
q0 and qN J depend on the boundary condition for pressure in j-direction. If one has
p
a Neumann BC for p at i = 0 (e.g. an inlet as in considered applications), x
=0
has to be satisfied and A0 = 0, Q0 = 0 and B0 and C0 are diagonal matrices (with
respectively 1 and 1 on the diagonal). If one has a Dirichlet BC for p at i = N I
(e.g. an outlet as in the considered case) pressure is imposed and CN I = 0, QN I = 0
and AN I and BN I are diagonal matrices (with 1 on the diagonal).
If one writes the solution of this system of equations in matrix form as
Pi = Mi Pi+1 + Ni

(6.63)

with Mi a matrix and Ni a vector and substitutes this expression in (6.59) one
obtains
Ai (Mi1 Pi + Ni1 ) + Bi Pi + Ci Pi+1 = Qi
(6.64)

92

CHAPTER 6. NUMERICAL METHODS

or
Pi = (Ai Mi1 + Bi )1 Ci Pi+1 + (Ai Mi1 + Bi )1 (Qi Ai Ni1 )

(6.65)

Comparing (6.63) and (6.65) one can see that Mi and Ni are given by
Mi =
(Ai Mi1 + Bi )1 Ci
Ni = (Ai Mi1 + Bi )1 (Qi Ai Ni1 )

(6.66)
(6.67)

The matrices (Ai Mi1 + Bi )1 Ci and (Ai Mi1 + Bi )1 do not depend on the right
hand side Qi but only on the mesh and can thus be precalculated and stored to save
computation time. The system of equations is then solved starting at i = 0 where
M0 = 0 and N0 = 0 and equations (6.66) and (6.67) are solved to find Mi and Ni for
increasing i. Then if all Mi matrices and Ni vectors are known one can solve (6.63)
starting at i = N I since PN I = NN I (because CN I = 0 and thus MN I = 0) and solve
backward till all Pi are determined and the complete pressure field is obtained.
3D Multigrid Poisson Solver
The most efficient iterative method to solve discretized steady equations known
today is a multigrid approach. It was first introduced and applied by Fedorenko [150]
for solving elliptic equations, such as the Poisson equation, and further developed by
Brandt [151]. There is by now a fairly well-developed theory of multigrid methods
for elliptic equations [151], [152]. The method can also be extended to hyperbolic
systems such as the Euler equations [153], [154], [155], [156].
The idea of the multigrid approach emerges from the fact that most iterative
methods eliminate the high frequency errors very fast, but fail to eliminate the low
frequency errors at the same rate. In addition, the notion of high and low frequency
error is related to the coarseness of the grid, in that any low frequency error on a
certain grid will become a high frequency error on another sufficiently coarser grid.
Therefore, the idea is to solve the equations on a sequence of grids ranging from
fine to coarse, so that the complete spectrum of errors can be eliminated with a rate
proper to the elimination of the high frequency errors.
Consider the following linear problem (e.g. the discretized Poisson equation):
Lu=f

(6.68)

Suppose v is an approximation to the solution of (6.68). Defining the residual as


r = f Lv and the error (i.e. the difference between the exact solution and its
approximation) as u = u v, the error will satisfy the residual equation
Au = r

(6.69)

After applying a few relaxation sweeps of an iterative method to (6.68), removing


effectively the high frequency components of the error, the approximation to the
solution of (6.68) v and the error u will be sufficiently smooth (i.e. the error contains
only low frequency components). On a coarser grid, however, the error u will appear
more oscillatory and, therefore, relaxation on this coarser grid will be more effective

6.3. PRESSURE CORRECTION METHOD

93

in removing this part of the error. To accelerate the convergence one can go to
a coarser grid and relax the residual equation Au = r, with an initial solution
u = 0. Once the residual equation is converged, u can be interpolated on the fine
grid (the so called prolongation) and added to v to obtain a new approximation to
the solution of (6.68). After this the whole procedure can be repeated.
This two grid algorithm is the basic building block of any multigrid method.
Before the general multigrid algorithm is described, let us first give a more detailed
account of how the two grid method works. The multigrid procedure in the cell
centered finite volume context is slightly different from that in the finite difference
approach. In the finite difference method the coarser grid is formed by removing
every second point of the fine mesh in every grid direction. Since the discrete
solution is considered in grid points, the points in which the solution is defined on
the coarser mesh will coincide with the points in which solution is defined on the
fine mesh. In the finite volume context, though, the solution is considered in the
centers of the cells. The cells of the coarser grid are formed by merging 2 (in 1D),
4 (in 2D) or 8 (in 3D) neighboring cells of the fine mesh (all of them having one
common vertex). Therefore, the discrete solution on the coarse grid is not defined
in the same points as the solution on the fine grid and one has to use adequate
restriction and prolongation operators.
Consider a discretization of a linear equation Lu = f on mesh h with spacing h
L h uh = f h

(6.70)

After applying one or more relaxation sweeps of the iterative method to (6.70), a
relatively smooth approximation vh to the solution is obtained. To obtain a coarse
grid correction to this approximation, first the residual rh is calculated
rh = f h L h vh

(6.71)

The residual equation for the error is then given by


Lh uh = rh

(6.72)

Instead of solving this equation on the fine grid h , one solves the following equation
on the coarser grid H :
LH uH = IhH rh
(6.73)
where LH is the operator L discretized on the coarse grid H and IhH is the restriction operator that brings fine grid information onto the coarse grid. In the
finite volume approach, the restriction of the residual is achieved by summing the
residuals rh of the fine mesh cells (weighted by the volumes, areas or lengths in 3D,
2D and 1D respectively and divided the sum by the volume of the coarse grid cell).
E.g. a 1D case on a uniform grid is shown in figure 6.6. It is also possible to use a
bigger stencil on the fine grid.
Solving equation (6.73) is fundamentally different from solving (6.68) on the
coarse mesh since the RHS of (6.73) ensures that the coarse grid solution will have
the accuracy of the fine grid. uH is an approximate correction of the fine grid

94

CHAPTER 6. NUMERICAL METHODS

Figure 6.6: One dimensional restriction operator.

Figure 6.7: One dimensional prolongation operators


solution and not an approximation of the solution itself. When a solution for (6.73)
is obtained, it is interpolated on the fine grid and added to vh
h
vh + I H
uH 7 vh

(6.74)

h
IH
is called a prolongation operator. The simplest prolongation is the piecewise
constant prolongation which is a first order operator. The correction in the fine
mesh cell is simply the correction in the coarse mesh cell of which the fine mesh
cell is a part. To improve the accuracy of the prolongation one can apply a more
accurate second order prolongation that involves a bigger stencil . Both prolongation operators are shown for the 1D case on a uniform mesh in figure 6.7. If the
mesh used is not uniform a volume-weighted average can be used. Based upon the
experience from a large number of numerical simulations [157, 158] a constant linear
prolongation works well for both uniform and non-uniform meshes. The weighted
linear prolongation does not give significant improvement in convergence, and it is
rather complicated and costly.
In the two grid approach, the coarse grid can still have a large amount of points,
which can make a solution of (6.73) still very expensive. In the multigrid approach
the two grid procedure is extended to still coarser grids. E.g. in a three grid method,
instead of solving equation (6.73) exactly, a two grid method is used, involving a still
coarser grid. Basically the exact solution of the residual equation is only required
on the coarsest grid, which normally requires a small computational effort, as the
number of points is very limited. Either a direct solver or many sweeps of an iterative
solver are used on the coarsest grid. A typical multigrid procedure is given by:

Perform NM L1 relaxation sweeps on the fine mesh (M L1), calculate residuals


and restrict them to a coarser grid (M L2).

6.3. PRESSURE CORRECTION METHOD

95

Figure 6.8: V-cycle (left) and V-sawtooth (right) multigrid cycles with 4 MG-levels;
denotes the smoother is applied on the level and denotes only prolongation is
performed
Perform NM L2 relaxation sweeps on the next mesh for an equation that has
the restricted residual from the previous mesh as a right hand side. Calculate
residuals of this equation and restrict them on yet a coarser grid (M L3).
Repeat this procedure till the coarsest mesh is reached. Solve the equation on
this mesh exactly.
Interpolate the correction to the previous mesh (prolongation), add it to the
correction on that mesh and perform one or more relaxation sweeps.
Repeat this procedure till the finest mesh is again reached.
When going back from the coarse to the fine grids one can either perform one or
more relaxation sweep or immediately prolongate the correction to further finer
grids. The first case is denoted V-cycle multigrid strategy while the second is called
the V-sawtooth multigrid strategy. Both strategies are schematically represented for
4 multigrid levels in figure 6.8. More sophisticated strategies, which have not been
investigated in the current work, can also be employed. For the investigated test
cases a V-sawtooth MG strategy with 4 or 5 MG-levels has shown to be the most
optimal in terms of computational effort. Typically few sweeps are performed on the
fine grids (since they are expensive) while a lot of sweeps are used on the coarsest
grids (on the coarsest, one wants an exact solution).
For a successful multigrid calculation, one needs an iterative method (a smoother)
that is efficient in removing high frequency components of the error. The alternating
line Gauss-Seidel method, which has been employed in the present work, is a much
better smoother than for example the simple point-Jacobi method.
2u 2u
Consider a discretization of a 2D Poisson equation ( 2 + 2 = q) on a cartesian
x y
uniform mesh (the treatment of non-uniform and 3D meshes is similar):
ui+1,j 2ui,j + ui1,j ui,j+1 2ui,j + ui,j1
+
= qi,j
x2
y 2

(6.75)

96

CHAPTER 6. NUMERICAL METHODS

The point Jacobi method determines the n + 1-th approximation of the solution
from
n
n
uni+1,j 2un+1
uni,j+1 2un+1
i,j + ui1,j
i,j + ui,j1
+
= qi,j
(6.76)
x2
y 2
If an approximation un is known, the next approximation un+1 can be directly
calculated from (6.76). This is typically done in a 2D loop over indices i and j. In
this loop the solution in points (i, j 1) and (i 1, j) is already updated on level
n + 1, when un+1
i,j is calculated. One therefore can use these updated values instead
n
of ui1,j and uni,j1 , which leads to the classical Gauss-Seidel (GS) method:
n+1
n+1
uni,j+1 2un+1
uni+1,j 2un+1
i,j + ui,j1
i,j + ui1,j
+
= qi,j
x2
y 2

(6.77)

This method has better convergence properties than the point Jacobi method. It
can, however, still be improved by adding one extra point at which an already
updated solution is used:
n+1
n+1
n+1
uni,j+1 2un+1
un+1
i,j + ui,j1
i+1,j 2ui,j + ui1,j
+
= qi,j
x2
y 2

(6.78)

Here the values of the solution lying on the grid lines j = const must be updated
by solving a system of linear equations with a tridiagonal matrix. This is done by
means of the Thomas algorithm or LU decomposition. This approach is called the
n+1
line GS method. If an updated value of solution un+1
i,j+1 is used instead of ui+1,j , one
gets the line GS scheme for another direction in which the Thomas algorithm must
be applied:
n+1
n+1
n+1
uni+1,j 2un+1
un+1
i,j + ui1,j
i,j+1 2ui,j + ui,j1
+
= qi,j
x2
y 2

(6.79)

By alternating (6.78) and (6.79) one gets a method with good smoothing properties
of high frequency components of the error. The method is called Alternating Line
Gauss-Seidel and is used as a smoother in the current work (in 3D so in 3 alternating
directions i,j and k).

6.3.2

Boundary Conditions

RANS inlet BC
For RANS simulations the mean velocity, turbulent kinetic energy k and its dissipation have to be imposed at the inlet boundary. One can use the available
experimental data for velocity and estimate the turbulent kinetic energy from the
available turbulent intensities (since all considered cases are axisymmetric configurations, a cylindrical coordinate system notation is used :u streamwise, v radial and
w circumferential). If only streamwise fluctuations are available from experiments
one can assume that
g + ww
g uu
g
vv

g
k = uu

(6.80)

6.3. PRESSURE CORRECTION METHOD

97

which is more or less true for a fully developed pipe flow. If also radial turbulent
intensity is known from experiment one can assume
g vv
g
ww

g
g
k = uu/2
+ vv

(6.81)

Determining the dissipation rate is more difficult. In [159] the following relation
is suggested
C 3/4 k 3/2
(6.82)
=
lm
with lm the Prandtl mixing length and C = 0.09. The major drawback of this
relation is that lm has to be estimated. From the DNS data of fully developed pipe
flow it can be derived that lm R/10 in the center of the pipe with R the pipe
radius but the mixing length decreases towards the wall. In [160] it is suggested to
use for internal flow the following estimation for based on an estimation of the
ratio of turbulent and laminar viscosity
=

C k 2
t

(6.83)

In [26] it is suggested to determine the profile of by solving the transport equation


(3.41) with frozen profiles for mean velocity and kinetic energy from measurements
(as discussed above). A fully developed flow is assumed by setting the axial derivatives equal to zero. In [26] the influence of the inlet conditions for have been
studied and a strong dependence of the solution has been observed. Another possibility which will be used in the current work is to solve the streamwise momentum
equation and k equations for a fully developed pipe flow. The only parameter that is set according to the experiment is the mean bulk velocity (or the flow
rate). Assuming a steady state (/t = 0) fully developed pipe flow (all derivatives
in stream wise and circumferential directions are zero and v and w are zero), the
streamwise momentum and k equation can be written as
u
1
( + t ) r
r r
r
1
r r
1
r r

t
k
+
r
k
r

+
r
r

p
z

=
=

= C2 f2

(6.84)
u
r

{z

Pk

!2

+ C1 f1 E
k
k

(6.85)

(6.86)

Where C2 , f2 , C1 , f1 and E depend on the chosen model (Chien, Standard


k2
or Yang-Shih k -model). The turbulent viscosity is given by t = C f .

Starting from an initial estimation of the fields u(r), k(r) and (r) all the terms in
the RHS are calculated. Then discretizing the diffusive term with a central finite
difference formulation, a system of equations is obtained for uj , kj and i , which
is solved iteratively (implicitly in pseudo time using a Thomas algorithm at each
pseudo time step). A new profile for u(r), k(r) and (r) is then obtained and the

98

CHAPTER 6. NUMERICAL METHODS

process can be repeated until required convergence is obtained (i.e. until the profiles
are not changing anymore). Since this method is extremely fast a very fine grid can
be used and later for the inlet profiles an interpolation to the required points can
be performed. Values for flow density and viscosity can be set according to the
composition and temperature of the considered inlet streams. The pressure gradient
which is the driving force for the flow can be estimated from the force equilibrium
and the Blasius solution of a turbulent Pipe Flow:
dp
2wall
=
,
dz
R

1
wall = u2 ,
8

= 0.3164Re1/4 ,

Re =

u2R

(6.87)

where u is the bulk velocity. Integrating the obtained velocity field u(r) over the
pipe section the bulk velocity can be calculated and so the pressure gradient can be
modified to finally obtain the correct mass flow. This procedure can also be applied
to an annular channel.
LES inlet BC
Where in RANS the inlet boundary conditions are determined by the mean velocities
and turbulent kinetic energy and dissipation (hui i,k,), which can be taken from
experimental results or simplified calculations, in LES the setting of the turbulent
inflow conditions is not trivial.
The most simple approach is just to use local mean velocities hui i, neglecting all turbulent fluctuations. This may be reasonable if the inflow is far upstream
of the area of interest or if the relevant fluctuations develop inside the computational domain. However this approach may yield large coherent structures instead
of turbulence leading to a totally wrong prediction of the mixing.
If the configuration upstream of the inflow is known an additional simulation can be performed to compute the necessary velocity fields in the inflow plane
(simultaneously or pre-computed). This method yields of course a very accurate
description of the inflow velocity fields but the additional computational cost can be
very high (LES resolution is required for the mesh upstream).
A transient inflow-velocity can be obtained by superimposing a random velocn
0n
ity fluctuation u0n
i , at each time step n, to the mean velocity hui i: (ui = hui i+ui ).
The level of the noise can be adjusted to match the variance hu0i u0i i obtained e.g.
from experiments. However this approach cannot represent other statistical properties like energy spectra or cross-correlations hu0i u0j i (i 6= j). The length scales present
in the flow field when using this approach are very small and so there is a relative
lack of turbulent kinetic energy in the low wave number range. The dissipation is
too strong and hence most of the kinetic energy will be damped to zero right after
the inflow plane as shown in [161], [162]. To overcome this problem Lee [163] suggested to apply an inverse Fourier transformation using an energy spectrum of real
turbulence. This method has been tested in [164, 165] but is rather complicated,
hard to implement and requires equidistant cartesian grids and three-dimensional
energy spectra which are hard to determine experimentally.
Although random noise is not appropriate to describe turbulence, the artificial
generation of turbulence-like velocity fields in the inflow plane is an interesting

6.3. PRESSURE CORRECTION METHOD

99

concept. Many statistical properties can be reproduced at a relatively low cost


if one uses a suitable procedure. Such a procedure has been proposed by Klein
[162] and has been successful used in different cases. The procedure starts with
creating a 3D random scalar field Ui which has no turbulent spectrum or desired
length scales. Convolution with an appropriate filter results in a turbulence-like
spectrum and creates the desired length scales. Assuming a Gaussian two-point
correlation function Klein deduced a relation between the filter coefficients and the
length scale. The turbulent velocity field ui is then computed from these scalar fields
Ui by conditioning the mean value hUi i = 0 and correlations hUi Uj i = ij . Then the
velocity field is constructed as
ui = hui i + aij Uj

(6.88)

where the tensor of coefficients aij is computed using a relation due to Lund [166]
from the Reynolds stress tensor Rij :


R11 q 0
0

R21
2

a
0
(6.89)
aij =
22
21
a

11
q
R32 a21 a31
R31
2
2
R33 a31 a32
a11
a22
This procedure creates a pseudo turbulent velocity field and slices of this 3D field
can be applied as inflow data for the LES computation. Although the created field
does not satisfy the continuity equation since it is only based on statistical properties
it creates turbulence like structures at little computational cost. The procedure can
be summarized as follows [162]:
1. Choose a length scale for each direction corresponding to the inflow plane
(Ly = ny y, Lz = nz z) and a time scale (or by Taylors hypothesis a length
scale Lx in the inflow direction).
2. Choose the filter widths according to the condition N 2n , = x, y, z.
3. Create 3 normalized random scalar fields R ( = x, y, z). The dimensions of
the fields are determined by the number of cells in the inflow plane (My Mz )
and the filter widths : [Nx : Nx , Ny + 1 : My + Ny , Nz + 1 : Mz + Nz ]
4. Apply the digital filter to R to obtain a 2D field of spatially correlated data
U (j, k)
5. Construct the fluctuation fields u0 from the given Reynolds stress tensor Rij .
To only satisfy the trace elements it is sufficient to use u0 = R U but if
one uses the procedure of equation (6.88) described by Lund [166] the crosscorrelations can also be set.
6. Store the obtained inlet slice u
7. Discard the first y,z plane of the R field, shift the data R (i, j, k) = R (i +
1, j, k) and fill the plane R (Nx , j, k) with new random numbers.

100

CHAPTER 6. NUMERICAL METHODS

8. Repeat steps 4-7 for each time step till the desired number of inlet data slices
are obtained
A drawback of this digital filtering procedure is that it can be used only on cartesian
equidistant grids. In the current work the inflow data in the non-uniform inlet plane
has been obtained by interpolation from the data generated by the above mentioned
procedure. Since in axisymmetric geometries velocity fields and Reynolds-stress
components are usually given in a cylindrical coordinate system, the obtained cartesian velocity fields have been transformed to their cylindrical counterparts before
performing the Lund transformation (using the available components Rzz , Rrr and
Rrz and estimating R Rrr ).
Instead of using low-pass filtering one can use a diffusion process to generate the
desired length scales in the random fields [167]. Since diffusion is applied in physical
space it can be applied to arbitrary grids and the implementation in a CFD-code
is straightforward since diffusive transport is already available. The typical length
scale in the random fields is approximately half the extent of a cell and to remove
the small structures diffusion is applied:
2 Ui
Ui
=D 2
t
xi

(6.90)

For a given coefficient of diffusion D, the length scale L increases with the time
diffusion is applied according to [167]:
L

2Dn

t
x
x2

(6.91)

where t is the time step, x the width of a cell and n the total number of time
steps. It can be shown that diffusion is equivalent to convoluting the original signal with a Gauss-Filter yielding autocorrelation functions of Gaussian shape. This
approach can be extended to create inhomogeneous length scale fields by applying
a inhomogeneous diffusion coefficient field. This is particularly interesting close to
walls where the length scale is proportional to the distance from the wall (mixing
length assumption). This procedure requires the selection of a desired length scale
field (e.g. a linear function of the distance to the wall) and the Reynolds stresses
(from experiments) in the points of the inlet plane. More recently some even more
sophisticated methods have been introduced including also cross-correlation length
scales into the inlet flow field [168, 169]. Unfortunately the necessary data for these
advanced methods (length scales, Reynolds-stress components) are usually not all
available from experiments and some assumptions and estimations have to be made
to obtain the necessary information.

6.4

Premixed Combustion Solver : FLOWSI

This section gives a brief overview of the TU-Darmstadt research code FLOWSI.
The governing equations in a low-Mach number approximation are solved using a
finite volume technique on a Cartesian mesh with velocities located on a staggered

6.5. GENERATION OF COMBUSTION TABLES

101

grid whereas the progress variable as well as the density are stored in the cell center.
For spatial discretization second order central differences are used for the momentum
equation. The convective fluxes of the scalar equation are discretized using a nonlinear TVD-scheme (CHARM-limiter) [134], to keep the solution bounded, and to
avoid unphysical oscillations in the progress variable field. The overall temporal discretization is an explicit third order Runge-Kutta method and the Poisson equation
for the pressure is inverted using a direct fast elliptic solver. Solid boundaries are included in the computational domain using the immersed boundary technique [170].
The presence of walls is expressed by an artificial force added to the momentum
equation. This force has been added to the velocities after the pressure correction
at every Runge-Kutta time step. At the outflow Neumann boundary conditions
for the velocity and the pressure are prescribed, negative velocities are clipped. At
the inlet boundary, mean velocity profiles with superimposed fluctuations, generated by a method presented by Klein [162] are imposed. The code is parallelized
using MPI. More detailed information can be found in [117, 118, 119, 102]. The
code has been used for LES simulations of non-reactive flow [127, 126], two-phase
flows [121], non-premixed [120, 122, 123, 171, 124, 125] and premixed combustion
[172, 173, 104].

6.5

Generation of Combustion Tables

Apart from ordering the data generated by the Equilibrium Code (CHEMKIN)
or Steady Laminar 1D Counterflow Flame Code (CHEM1D) and interpolating it
to the desired mixture fraction and variance grid, the most difficult part of the
combustion table generation is the pdf-integration. In the numerical modeling of
turbulent flames it is common to adopt the -pdf as assumed probability distribution
function.
z a1 (1 z)b1
(6.92)
p(z) =
B(a, b)
Where

(a)(b)
B(a, b) =
=
(a + b)

1
0

z a1 (1 z)b1 dz

(6.93)
Z

is the normalization factor and the -function is defined by (x) =


et tx1 dt.
0
This pdf is defined by only two parameters, the two moments, the mean value
2 < z
e(1 ze):
0 ze 1 and its variance 0 zg
a = z

z(1 z)

and

2
zg

(6.94)

(1 z)a
(6.95)
z
Numerical computation of the beta-pdf covering the full range of means and
variances is difficult because in several limit cases overflows or underflows can occur.
It is difficult to calculate the normalization factor B(a, b) if a and/or b are below
b=

102

CHAPTER 6. NUMERICAL METHODS

unity since the integrand is badly conditioned for numerical integration because the
integrand and its slope are infinite at one or both ends of the integration domain.
In [174] an accurate method is described to compute the normalization factor for
low values of the parameters a and b using a finite integration interval. Using the
recursive property of the gamma-function (1 + x) = x(x) one can write
() =

( + 2)
( + 1)
=

( + 1)

(6.96)

and thus
B(a, b) =

(a + b)(a + b + 1)(a + b + 2)(a + b + 3) (a + 2)(b + 2)


=B
ab(a + 1)(b + 1)
(a + b + 4)

{z

1
0

z a+1 (1z)b+1 dz

(6.97)
The last integral is much better suited for numerical integration since the integrand
is zero at both ends with zero slope. The high accuracy of the method is shown
on several examples in [174]. Apart from the normalization also the calculation of
2 [175]. To
the pdf p(z) itself can be a problem for some combinations of ze and zg
determine the mean species and temperature the pdf has to be known over the entire
range of the mixture fraction space 0 z 1. In the extreme cases for the mean
2 must be 0 and the pdf is restricted to a Delta
ze = 0 and ze = 1 the variance zg
function at respectively z = 0 and z = 1. The variance must range between 0 and
2 = 0 the pdf is a delta function at z = z
e and the laminar flame
ze(1 ze). If zg
solution is obtained (in fact for this case no integration is needed and the laminar
flame solution is used in the table). The upper limit ze(1 ze) can not be reached
since for a = b = 0 the -pdf is not defined. If a 1 and b 1 the computation
of the pdf shows no singularities. If a < 1 and b 1 the pdf shows a singularity at
z = 0. To account for the steep slope of the pdf near z = 0 the pdf at the lowest
point of the integration domain that is used can be calculated from
p(z1 ) =

A A0
z1

(6.98)

where A is the total area B(a, b) calculated with the method described above and
A0 is the partial area calculated in the same way but without the first integration
interval z1 . A similar procedure maybe applied for p(zN ) with N the last point
in the integration domain if If b < 1 and a 1. When both a < 1 and b < 1 the
pdf shows a singularity at both ends. The pdf in the first and last point is then
obtained from
p(z1 )z1 + p(zN )zN = A A0
(6.99)
where A0 is now the area obtained by integration over the domain except for the
first (z1 ) and (zN ) last intervals. To partition the area difference the ratio of the
pdf at the end intervals is used
p(z1 )
z1a1 (1 z1 )b1
= a1
p(zN )
zN (1 zN )b1

(6.100)

6.5. GENERATION OF COMBUSTION TABLES

103

Finally when one or both of the parameters a and b is much greater then unity (which
2 0), very large or small numbers exceeding the computer range may
occurs if zg
occur and special care has to be taken (e.g. using double precision). The pdf can
also be prescaled with the peak value. If also the normalizing area is prescaled in
the same manner both scaling factors cancel and no underflow will be encountered
[175].

104

CHAPTER 6. NUMERICAL METHODS

Chapter 7
Results Non-premixed
Combustion Modeling
7.1

Piloted non-premixed flame (Sandia)

The first test case is a turbulent diffusion flame that has been studied experimentally
by Masri et al. [176] and Barlow et al. [177]. The inflow consists of an axisymmetric
jet, surrounded by a pilot gas stream and a co-flow air stream. Figure 7.1 shows the
configuration and the burner dimensions are given in table 7.1.
Main jet inner diameter DF
Pilot annulus inner diameter
Pilot annulus outer diameter DP
Burner outer wall diameter

7.2 mm
7.7 mm (wall thickness is 0.25 mm)
18.2 mm
18.9 mm (wall thickness is 0.35 mm)

Table 7.1: Sandia Flame Series A-F : main dimensions


The jet fluid is a mixture of three parts air and one part methane (CH4 ) by
volume. This mixture significantly reduces the problem of fluorescence interference
from soot precursors, allowing improved accuracy in the scalar measurements. Partial premixing with air also reduces the flame length and produces a more robust
flame than pure CH4 or nitrogen-diluted CH4 . Consequently, the flames may be
operated at reasonably high Reynolds number with little or no local extinction for
moderate jet velocities, even with a modest pilot. The mixing rates are high enough
so that the flames burn as diffusion flames, with a single reaction zone near the
stoichiometric mixture fraction. There is no indication of significant premixed reaction in the fuel-rich CH4 /air mixtures. Therefore although the fuel jet contains
air one can treat this flame as a purely non-premixed flame. The total flame series
contains 6 different flames, named Sandia A-F. Flame A is a laminar flame while
flame B is in the transitional regime. The most investigated flame from the series
is Flame D (Re = 22, 400) which has a small degree of local extinction [177]. It was
selected as the initial target for TNF1 [3] comparisons since high Reynolds number is desirable for model validation while the small probability of local extinction
1

This workshop is an open and ongoing international collaboration among experimental and
computational researchers in turbulent non-premixed and partially premixed combustion. The

105

106

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

Figure 7.1: Sandia Piloted Jet Flame. Configuration and picture of the experimental
burner: Central fuel jet, pilot flame (in fact a set of little stabilization flames in
annular ring) and co-flow.
allows for useful comparisons with models that do not include extinction. Flames
E and F have significant and increasing probability of local extinction above the
pilot region, with flame F being close to global extinction of the downstream part
of the flame. The pilot is a lean ( = 0.77) mixture of C2 H2 , H2 , air, CO2 , and
N2 with the same nominal enthalpy and equilibrium composition as methane/air
at this equivalence ratio. The flow rates for the main jet and the pilot stream are
scaled in proportion for the C-F series, so that the energy release of the pilot is
approximately 6% of the main jet for each flame. Table 7.2 gives the bulk velocities
for the fuel jet and pilot stream as well as the Reynolds number based on the fuel
jet diameter (D = 2RF = 0.0072m), fuel jet bulk velocity and kinematic viscosity
of the methane/air mixture ( = 1.58E 05m2 /s). The burner is positioned in a
vertical wind tunnel providing a co-flow air stream at 0.9 m/s.
Flame
C
D
E
F

Bulk velocity Jet


29.7 m/s ( 2 m/s)
49.6 m/s ( 2 m/s)
74.4 m/s ( 2 m/s)
99.2 m/s ( 2 m/s)

Bulk velocity Pilot


6.8 m/s ( 0.30 m/s)
11.4 m/s ( 0.50 m/s)
17.1 m/s ( 0.75 m/s)
22.8 m/s ( 1.00 m/s)

Re
13,400
22,400
33,600
44,800

Table 7.2: Sandia Flame Series C-F : jet and pilot bulk velocities and Reynolds
number (based on jet properties)
current emphasis is on fundamental issues of turbulence-chemistry interactions in gaseous flames.
The objectives are to establish an internet library of well-documented flames that are appropriate
for model validation and the advancement of basic scientific understanding of turbulent combustion
and to provide a framework for collaborative comparisons of measured and modeled results.

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

107

The flame series has been investigated experimentally at the Combustion Research Facility, Sandia National Laboratories where scalar measurements have been
performed [177, 178, 179] and at the FG Energie-und Kraftwerkstechnik, TU Darmstadt, where velocity fields have been measured [180]. Scalar measurements include
Raman/Rayleigh/LIF measurements of mixture fraction (z), temperature T and
species mass fractions N2 , O2 , CH4 , CO2 , H2 O, H2 , CO, OH, and NO with a spatial
resolution of 0.75 mm. Results include axial profiles in each flame (x/d = 5, 10,
15, ... , 80) and radial profiles (at x/d = 1, 2, 3, 7.5, 15, 30, 45, 60, 75). Averaged
results (both Favre and Reynolds averages), rms-fluctuations and single-shot data
are included in the data archives [3]. Two-component LDV-measurements were performed at the Technical University of Darmstadt [180] and include axial profiles and
radial profiles (x/d = 0, 7.5, 15, 30, 45, 60, 75) of axial and radial velocity (mean
and rms fluctuations) and turbulent shear stress (u0 v 0 ).

7.1.1

RANS simulations

To test the flow solver and combustion modeling, the Sandia flame has first been
investigated using RANS. The computational cost of a RANS simulation is much
lower than for an LES of the same configuration since one is looking for a steady state
solution instead of a temporally evolving unsteady flow field. Moreover in RANS one
can use the axisymmetry of the studied configuration to reduce the number of grid
points drastically. In circumferential direction only one physical cell is considered
and the axisymmetry of the configuration is taken into account in the discretization
of the equations. Because of this quasi-2D approximation a direct solver (section
6.3.1) can be used for the Poisson equation having a better performance than the
multigrid solver (section 6.3.1) for a moderate number of points (especially in the
radial (j-) direction the number of points should not be too big). Since the spatial
resolution needed for a RANS simulation is significantly smaller than for an LES,
also in radial and axial direction a lower number of points can be used (or a bigger
computational domain extending further in axial or radial direction). The average
computation time to obtain a fully converged solution is typically in the order of
several hours on a modern high performance computing system while for LES several
weeks of cpu-time are required to obtain fully developed and statistically averaged
results. This explains why it is much more straightforward to test different models
and boundary conditions for RANS than for LES. Moreover it is sometimes difficult
to observe the effects of model modifications in the instantaneous flow field and to
predict the effect on the time averaged results.
Figure 7.2 shows a typical mesh and computational domain for the RANS simulations of the Sandia flame. The computational domain is 100D in axial direction,
25D in radial direction and has an opening angle in circumferential direction of
1o . The normal grid consist of 88 cells in radial direction of which 16 are uniformly distributed in the fuel jet (r = 0.225mm) and 24 in the pilot stream
(r = 0.23mm) and the remaining cells in the co-flow are clustered towards the
pilot stream (rin = 0.23mm and rout = 10mm). In axial direction the grid
consists of 80 cells clustered towards the inlet of the computational domain so that

108

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

the first cells in the fuel stream are square (r = x). The domain dimensions
and number of cells are identical to the ones used in [26]. To test the quality of
the mesh a grid refinement study has been performed using a mesh which is twice
as fine as the mesh described above but with a smaller computational domain to
restrict the number of cells (as mentioned before the performance of the direct solver
deteriorates significantly if the number of points becomes too big).
The numerical methods used are described in detail in chapter 6. The set of equations solved are: the momentum equations (3.36) for the mean velocity components
e , a transport equation for the mean mixture fraction ze (4.42) and its variance
(ue,ve,w)
g
z 2 (4.59) and transport equations for the turbulent kinetic energy ke (3.40) and its
dissipation e (3.41). The scalar dissipation rate e is modeled using the turbulent
mixing time (4.58). The laminar and turbulent Schmidt numbers (Sc and Sct ) are
2 . Settings for the two different k
both set to 0.7 in the equations for ze and zg
models used can be found in table 7.3. Note that the Yang-Shih model uses the modified constant C2 as suggested in [27, 26] and that terms in (3.40, 3.41) - taking the
variable density effect on turbulence into account - have been neglected. Moreover
all functions f1 ,f2 ,f have been set to 1 except for f2 in case 1b (see further). Since
all considered cases are unconfined the effect of walls is supposed to be negligible.
All equations have been integrated in time using a 4-stage RK scheme (6.42) and
convective fluxes have been discretized with a second order central scheme for the
momentum equations and a first order upwind scheme for the scalar equations (see
section 6.2.2). A small amount of 4-th order artificial dissipation (6.18) was added
to the momentum equation to stabilize the simulation. The boundary conditions
are given by:
At x = 0 inlet BC where imposed i.e. all quantities where imposed (u,v,w,k,,z,z 2 )
except for the pressure which is extrapolated from inside the domain. All other
relevant quantities (T ,,) are calculated in the dummy cell from the combustion variables (z,z 2 ). As discussed in section 6.3.2 different profiles for the
axial velocity u, kinetic energy k and dissipation can be used, and their
impact will be discussed further. The mixture fraction is set to 1 in the fuel
stream, 0.27 in the pilot stream (to obtain the same temperature as the real
pilot flame) and 0 in the co-flow and the variance is set to a very small value
(1E-04).
At r = 0 corresponding to the symmetry line, an axisymmetric BC is imposed.
At x = 100D and r = 25D a Neumann boundary condition is used, extrapolating all variables except for pressure which is imposed to the atmospheric
pressure.
In circumferential direction rotational periodicity is applied.
The coupling of the flow code and the combustion tables is performed as described in figures 4.14 for the infinitely fast chemistry model and 4.15 for the flamelet
model.

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)


k model

Standard

1.3

Yang-Shih

1.3

k
q

2
2 ui
xj xk

C1

109
C2

1.44 1.92

t
2

C k

1.44 1.83 C k

q 

Table 7.3: Source term models and parameters for used k models

Figure 7.2: Typical mesh and computational domain for RANS simulation of Sandia
flame
Infinitely fast chemistry
The mixed-is-burnt tables are pre-computed by performing the -pdf integration
starting from the laminar Burke-Schumann relations with variable Cp effect (4.25,
4.26, 4.32). The tables have 200 points in mixture fraction space (ze) clustered
around the stoichiometric value (ze = zst ) and a variable number of points in the
2 ) uniformly distributed between 0 and z
e(1 ze). The obtained
variance space (zg
temperature field and species mass fractions are shown in figure 7.3. In the flow
solver a linear interpolation is used to obtain the temperature and molecular weight
of the mixture from the flame table.
First results of the RANS simulations using the infinitely fast chemistry model
will be discussed. Five different simulations have been performed denoted CASE
1-5. Cases 1 and 2 use the inlet profiles obtained by solving the set of equations
(6.84-6.86) both in the fuel pipe and the pilot annulus (in the co-flow a constant
axial velocity of 0.9 m/s and constant values for k and are set). The pressure
gradient in (6.84) is adjusted so that the bulk velocity corresponds exactly to the
ones given in table 7.2. Case 3 uses the experimental profile [180] for the axial
g
g with uu
velocity u and turbulent kinetic energy (assuming k = uu/2
+ vv
and vv from experimental database). For , equation (6.86) is then solved with
fixed u and k fields as done in [26]. Cases 4 and 5 also use the experimental profile
for the axial velocity u and turbulent kinetic energy, but determine respectively
3/4

from =

C k 3/2
lm

[159] with C = 0.09 and lm = R/10 (R being the fuel jet radius

110

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

Figure 7.3: Mixed-is-burnt table for Sandia flames: Temperature and species mass
2 plane (Boundary conditions for temperature and composition
fractions in the ze-zg
are TF = 291K, TO = 294K, YF,1 = 0.156, YO,1 = 0.196, YO,2 = 0.232

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)


70

40

Simulation fuel pipe


Profile from exp.
Simulation pilot annulus

60

111
Simulation fuel pipe
Profile from exp.
Simulation pilot annulus

35

Measurements

Measurements
30

50

2 2

k(m /s )

u(m/s)

25

40

30

20

15

20
10

10

0
0

0.002

0.004

0.006

0.008

0.01

0.012

0.014

0
0

0.016

0.002

0.004

0.006

0.008

r (m)

0.01

0.012

0.014

0.016

r (m)

x 10

4.5
4

5000

Simulation fuel pipe


Simulation fixed u,k
Modeled (Jones)
Modeled (Numeca)

4000

3500

(m2/s3)

(m2/s3)

3.5

2.5
2

3000

2500

2000

1.5

1500

1000

0.5
0
0

Simulation pilot annulus


Simulation fixed u,k
Modeled (Jones)
Modeled (Numeca)

4500

500

0.5

1.5

r (m)

2.5

3.5

4
3

x 10

r (m)

10
3

x 10

Figure 7.4: Comparison of mean axial velocity, turbulent kinetic energy k and its
dissipation at the inlets (fuel jet & pilot) for computed inlet and inlet modeled
from measurements.
2

or the annulus width for the pilot) and from = t C k [160] with t 50 in the
fuel jet and t 20 in the pilot (based on the full computations of case 1). As
one can see the computed velocity profiles for a fully developed pipe and annulus
flow are a little below the experimental values with a maximum deviation of 3 m/s
(Note in table 7.2 that the measurement error in the bulk velocity is in the order of
2 m/s). The computed profile for k in the fuel and pilot corresponds in shape to the
experimental profile but the computed values are clearly higher (about 5 m2 /s2 on
the jet axis to 12.5 m2 /s2 close to the wall of the fuel pipe. The dissipation shows
a strong peak near the walls of the pipe and annulus both in the full computation
and in the computation based on the experimental u and k profiles. The estimation
from [159] or [160] are of the good order of magnitude away from walls but close to
the walls there is a strong underestimation. The difference between the two models
is not very big. The different inlet profiles for u,k and are shown in figure 7.4. All
cases use the Yang-Shih k- model except for case 2 which uses the standard k-
model (see table 7.3).
Figure 7.5 shows the evolution of the mixture fraction and its variance, the temperature and density and the mean axial velocity and the turbulent kinetic energy
along the symmetry axis. The axial distance from the nozzle has been made dimensionless with the nozzle diameter D and the axial velocity with the exit velocity on
the symmetry axis. The density from experiments is determined from the exper-

112

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

imental species mass fractions Yk , temperature T and atmospheric pressure. The


g
g One observes that all cases,
turbulent kinetic energy is given by k = uu/2
+ vv.
except the standard k- run of Case 2, show reasonably good agreement with experimental results. One can see that axial velocity decays too rapidly and too close
to the nozzle for the standard k- model while for all other cases the correspondence with the experiments is very good. The standard k- model fails to predict
the correct turbulent shear stress which is the most important component in the
x-momentum equation on the symmetry line. This can be observed by looking at
the evolution of k along the symmetry axis in figure 7.6. Since turbulent shear
stress is the most important production term of k, peaks in turbulent shear stress
correspond to zones of positive k
[26]. One can observe that cases 1 and 2 have
x
the same inlet condition for k (both computed profiles) and both initially decay
from their inlet value. Around x/D = 10 the standard k- model shows a strong
increase in k corresponding to a big shear stress and thus to a too strong decay in
axial velocity. One can also notice that all Yang-Shih cases show a similar behavior
for k after x/D = 15 with a slightly smaller value for case 1 with computed inlet
profiles. The difference in mean axial velocity between cases 3-5 is barely visible
and only a small difference in k for x/D < 10, caused by the different values for
at the inlet, can be observed in 7.6. The mean velocity decay has an important
impact on the mixture fraction and thus on the temperature evolution and thus
the results for the standard k- model deviate significantly from the experimental
results. Therefore for all further simulations the Yang-Shih model was chosen. One
can clearly observe for cases 3-5 that for all quantities along the symmetry axis in
figure 7.5 the difference is negligible and thus the impact of the model used for at
the inlet is small at least for this configuration.
Comparing the results obtained with the computed inlet (case 1) and with velocity and turbulent kinetic energy from the experiment (cases 3-5) one can see that
the dimensionless axial velocity is almost identical but that there is a small difference in k (figure 7.5(f)). This is reflected in a little smaller decay of the mixture
fraction for x/D > 50 and thus a better correspondence of the temperature and
density field (figure 7.5(a-c-d)). The flame length (x/D|(z=zst ) ) is 41.5 compared to
the experimental value of 47 and thus there is a too strong decay of the mixture
fraction. The mixture fraction variance peaks too close to the nozzle (x/D 20).
For case 1 the variance value is bigger in the far field (x/D > 60) resulting together
with the higher value of z in this zone in a temperature that stays higher and closer
to the experimental value till the end of the computational domain.
Figures 7.7 show the mean mixture fraction, axial velocity, temperature and kinetic
energy in radial cuts of the domain at x/D = 7.5, 15, 45, 60. At x/D = 7.5 the velocity field is in good agreement with the experimental values but the turbulent kinetic
energy is strongly overestimated. In case 2 the turbulent kinetic energy dissipation
is clearly underestimated leading to the high values of k. Comparing cases 1 and 3
it is clear that the higher k in case 1 is due to the higher value at the inlet as shown
in figure 7.4. The mixture fraction field is clearly too wide for all cases and thus for
r/D > 1 the temperature is strongly overestimated. These observations can also be
made further downstream where it is even more clear that the jet spreading (i.e. the

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

0.15

(a)

0.5

(b)

0.1

z"2

113

0.05
0

20

40 x/D 60

80

100

(c)

2000

20

40 x/D 60

80

100

80

100

(d)

1500
1000

0.5

500
0

20

40 x/D 60

80

100

(e)

20

40 x/D 60

(f) CASE1
CASE3
CASE4
CASE5
CASE2
EXP

60

u/u

0.5

40
20

20

40 x/D 60

80

100

20

40 x/D 60

80

Figure 7.5: Infinitely fast chemistry results along symmetry axis for flame D (CASES
1-5 compared with experimental data from [178, 3, 180]): (a) mean mixture fraction
(b) variance of mixture fraction (c) temperature (d) density (e) axial velocity (f)
turbulent kinetic energy
15
CASE 1
CASE 3
CASE 4
CASE 5
CASE 2

10

0
0

10

20

30

40

50

x/D

60

70

80

90

100

Figure 7.6: Evolution of k along the symmetry axis for RANS simulation of Sandia
flame D

100

114

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

,
Figure 7.7: Radial profiles for Infinitely fast chemistry results of flame D (CASES
1-3, experimental data from [178, 3, 180]) at x/D = 7.5, x/D = 15, x/D = 45 and
x/D = 60: (a) Mixture fraction, (b) Axial velocity, (c) Temperature, (d) Turbulent
kinetic energy
jet half width increases too fast) is too strong. However at x/D = 45 and x/D = 60
where the overall mixture fraction is low (z < zst ) the predicted temperature field is
in good agreement with the experiment especially for case 1. Moreover, while close
to the nozzle the kinetic energy is overpredicted, further downstream it is underpredicted. Overall cases 1 and 3 do not differ too much with slightly better results for
the computed inlet case.
The main reason for the bad performance of the standard k- model is the
influence of the C2 constant in the equation for the dissipation of kinetic energy
since it affects the spreading rate of the jet significantly and thus the decay of the
centerline velocity. This can be clearly seen in figure 7.8 where the results for two
extra cases have been added. Case 6 uses the same settings as case 1 except for
the f2 function in the dissipation term of equation (3.41) 1t C2 f2 which is set
to the expression proposed in [29]: f2 = 1 0.22 exp (Ret /36). Case 7 is identical
to case 2 except for the value of C2 which is set to 1.83 as in the Yang-Shih cases.

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

115

CASE
1
2
3

k- model
Yang-Shih
Standard
Yang-Shih

Inlet
Computed u,k- (6.84)-(6.86)
Computed u,k- (6.84)-(6.86)
Experimental u,k + from (6.86)

C2
1.83
1.92
1.83

f2
1
1
1

Yang-Shih

1.83

5
6
7

[160] 1.83
Yang-Shih Experimental u,k + from
1
t
)
Yang-Shih
Computed u,k- (6.84)-(6.86)
1.83 1 0.22 exp( Re
36
Yang-Shih
Computed u,k- (6.84)-(6.86)
1.83
1

3/4

Experimental u,k + from

C k 3/2
lm
C k 2
t

[159]

Table 7.4: Different RANS cases for Sandia flame D: k- model, Inlet BC, C2
constant and f2 function

Comparing cases 2 and 7 one can observe that C2 has a strong influence on the
decay of the mean axial velocity on the centerline and with the use of the modified
constant the agreement with the experiments and Yang-Shih cases is good. Starting
at x/D = 25 the turbulent kinetic energy starts deviating from the Yang-Shih cases
and for x/D > 40 a stronger decay of the mixture fraction and lower value of
its variance is observed and thus the temperature starts decreasing more rapidly.
Comparing cases 1 and 6 one can see that the f2 function has a small effect on the
velocity field - the mean axial velocity is a bit higher in the far field - resulting in a
little less decay of the mixture fraction for x/D > 40. The temperature and density
are therefore in better agreement with the experimental values.

Finally a grid refinement has been performed with the infinitely fast chemistry
model. The cell size in both axial and radial direction has been chosen as half the
cell size of the normal grid and the computational domain has been slightly reduced
to limit the computational cost (since the performance of the direct Poisson solver
strongly depends on the number of points in radial direction). The solution differs
only slightly indicating that the normal mesh is sufficiently fine. As an example the
velocity and kinetic energy along the symmetry line are shown in figure 7.9.
It is thus clear that both the inlet boundary conditions and the turbulence model
have a significant effect on the obtained results. Since the modeling of turbulent
non-premixed combustion is split in a mixing and a combustion problem it is in
first instance important to get the flow field and turbulence properties k and
as close as possible to the experimental values to get a correct prediction of the
flame. All considered cases predict a too strong decay of mixture fraction along
the symmetry line and thus underpredict the flame length. More sophisticated
turbulence models are able to improve the prediction of the flow field and flame
properties [26, 181, 182] but the dependence on several model parameters, makes
the general predictive capabilities of the RANS turbulence models for turbulent
combustion doubtful.

116

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

Figure 7.8: Infinitely fast chemistry results along symmetry axis for flame D (CASES
1, 2, 6, 7 compared with experimental data from [178, 3, 180]): (a) mean mixture
fraction (b) variance of mixture fraction (c) temperature (d) density (e) axial velocity
(f) turbulent kinetic energy

Figure 7.9: Influence of Grid refinement on RANS simulation of Sandia flame D


(settings of case 1): axial velocity (left) and turbulent kinetic energy (right)

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

117

Finite rate chemistry


To include finite rate chemistry effects, a steady flamelet model as discussed in
sections 4.4.2 and 4.3.3, can be used. In this work the 1D flame solver Chem1D
developed at the TU Eindhoven [43] has been used to generate the flamelet tables.
A set of laminar counterflow diffusion flame calculations has been performed for the
strain range allowed by the flame solver. The resulting flame structure in physical
space has been converted to mixture fraction space using the Bilger definition [183]
for mixture fraction
z=

2 (YC YC,2 ) /WC + 0.5 (YH YH,2 ) /WH (YO YO,2 ) /WO
2 (YC,1 YC,2 ) /WC + 0.5 (YH,1 YH,2 ) /WH (YO,1 YO,2 ) /WO

(7.1)

where YC , YH and YO are the element mass fractions of carbon, hydrogen and oxygen,
WC , WH and WO their atomic weights and subscripts 1 and 2 denote the fuel
stream and oxidizer (co-flow) values. Note that in the experimental determination
of mixture fraction for the Sandia flame, the oxygen terms terms are excluded since
the jet and co-flow boundary conditions for YO are relatively close and shot noise
in the measurements of elemental oxygen mass fraction causes additional noise in
the measured mixture fraction field. The fuel jet and co-flow air is presumed to
be dry i.e. YH2 O is neglected. The used mechanism is GRI 1.2 with 32 species
and 177 reactions [46], and does not take the nitrogen chemistry into account. The
species mass fraction of the nitrogen containing species is so small that they have
no significant influence on the mixture molecular weight and thus on the density,
which is coupled back to the flow solver. Therefore neglecting their presence hardly
affects the results. The obtained flame structures in mixture fraction space are then
parameterized with the scalar dissipation rate at stoichiometric mixture fraction
provided by the flame solver: Yk (z, st ) and T (z, st ). In mixture fraction space
200 points are used clustered around the stoichiometric mixture fraction and 100
points in stoichiometric scalar dissipation space clustered towards the minimum and
maximum value. The resulting temperature and density field are shown in figure
7.10. Notice that for increasing scalar dissipation the maximum flame temperature
decreases and shifts to little higher values of the mixture fraction. The density
field shows a strong gradient in mixture fraction space at both the fuel side and
oxidizer side which is even more pronounced for low strain values. To obtain the full
flame tables for the turbulent combustion simulation the pdf-integration has to be
performed. A -pdf was presumed for the mixture fraction (4.50) and a log-normal
for the scalar dissipation (4.64). As discussed in the chapter on numerical methods
special care was taken to avoid numerical problems arising from the normalization
of the -pdf or over-/underflows during the calculation of the pdf (see section 6.5).
A 3D table is then obtained to determine the temperature T , species mass fractions
Yk and molecular weight of the mixture Wm from the mean mixture fraction ze, its
2 and the mean stoichiometric scalar dissipation
g
variance zg
st provided by the flow
solver (figure 4.15). Figure 7.11 gives an impression of the final flamelet table for
temperature and some intermediate species. Finally to obtain the scalar dissipation
g
e
at stoichiometric mixture fraction
 st from , equation (4.67) can be used, where
g
2
the mapping function F ze, z
is precalculated and tabulated as a function of ze

118

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

,
Figure 7.10: Temperature field and density field in the mixture fraction z - scalar
dissipation rate plane st as obtained by reordering the counterflow diffusion flame
results from Chem1D-simulations
2 as shown in figure 7.12.
and zg

These flamelet tables were used to simulate the Sandia D flame with identical
settings as the mixed is burnt simulations. Figure 7.13 compares the obtained results along the symmetry line. One can observe that there is a postponed mixture
fraction decay for x/D < 40 in the flamelet case as a consequence of the postponed velocity decay. The temperature is higher and thus the density is lower in
the flamelet case which has to be compensated by a higher velocity. For x/D > 10
there is a clear difference in the turbulent kinetic energy and mixture fraction variance for the flamelet and mixed is burnt case, and the peak in both quantities is
positioned further downstream for the flamelet case. Note the strong resemblance
2 . Overall it is difficult to determine which model gives
in the behavior of k and zg
the best correspondence with the experimental results. In the Sandia D flame there
are no important finite rate chemistry effect (this is more the case for flames E and
F with higher Re number). The influence of the chemistry model on the flow field
is thus small. This can also be observed in the radial profiles shown in figure 7.15
for x/D = 7.5 and x/D = 30. There is a clear overprediction of the temperature
for both chemistry models which is caused by an underestimation of the mixture
fraction variance and because the radiation is neglected. One can see in the radial
profiles that for z = zst the temperature obtained from the mixed is burnt and
flamelet tables is too high. The maximum temperature from the experiments is
about 1945K while for the mixed is burnt the maximum temperature is 2085K and
for the flamelet model 2130K. The flame is unconfined, can be regarded as optically
thin and hardly any soot is formed and thus the self-absorption of the flame can
be neglected. As a result radiation can be seen as a pure loss of heat and thus a
reduction of the overall temperature field. The total radiative flux has been measured to be about 5% of the total power of the flame (including the pilot). In [26] a
temperature difference in the order of 100K is observed when a radiation model is
included in the simulation. Although the lower temperature will affect the density
field, the flow field does not seem to be strongly affected. One can also observe a

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

119

,
2 ,
g
Figure 7.11: Flamelet table for Sandia flames as function of ze,zg
st : temperature,
OH mass fraction, CO mass fraction, H2 mass fraction

120

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

2 relating
e and
g
Figure 7.12: Mapping function F ze, zg
st

Figure 7.13: Finite rate chemistry results along symmetry axis for flame D compared
with mixed is burnt results and experimental data from [178, 3, 180]: (a) mean
mixture fraction (b) variance of mixture fraction (c) temperature (d) density (e)
axial velocity (f) turbulent kinetic energy

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

121

bump in the temperature profile of case 5 around x/D = 20, where the temperature difference between both combustion models is the highest. The reason for this
difference can be explained by looking at the laminar flame structure for the mixed
is burnt and flamelet models in figure 7.14. One can see that for low strain (i.e. low
scalar dissipation rate) at the rich side of the flame a much higher temperature is
obtained from the 1D counterflow flame simulation This is caused, for the Sandia
flame, by the presence of oxygen in the fuel stream. This is also why the equilibrium
chemistry model cannot be applied to the Sandia flame since the code would never
allow fuel and oxidizer to coexist without reacting. Since around x/D = 20 the
mixture fraction is about 0.8 a significant difference in temperature is thus obtained
when retrieving the temperature from the flamelet table with low scalar dissipation
rate.
2500

2000

T(K)

1500

1000
MIB
FLAM LOW STRAIN
FLAM MEDIUM STRAIN
FLAM HIGH STRAIN

500

0
0

0.2

0.4

0.6

0.8

Figure 7.14: Comparison of laminar flame structure obtained from mixed is burnt
and flamelet model for the Sandia flame: flamelet with low strain (a = 0.6, st =
0.125), high strain (a = 1250.4, st = 70.0) and medium strain (st = 34.4)
,
Overall the correspondence between the experiments and the numerical simulations is acceptable. The flame structure is the most sensitive to the turbulence
model and the chemistry effects are small because no important finite rate effects
or extinction occur in this flame. Including the radiation in the simulations - which
can be performed by adjusting the temperature in the lookup tables - would yield a
better correspondence in the temperature field.

7.1.2

LES simulations

The numerical methods used are described in detail in chapter 6. The set of equation solved are the momentum equations (3.18) for the filtered velocity compoe and a transport equation (3.31) for the filtered mixture fraction ze. A
nents (ue,ve,w)
Smagorinsky eddy viscosity model is used to model the sub-grid stresses and the
subgrid scalar flux is closed using a gradient assumption with Sc = Sct = 0.7.
All equations have been integrated in time using a 4-stage RK scheme (6.42) and
convective fluxes have been discretized with a second order central scheme for the
momentum equation and a first order upwind or a second order TVD scheme with

122

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

,
Figure 7.15: Radial profiles for finite rate chemistry results of flame D (Comparison
with mixed is burnt case and experimental data from [178, 3, 180]) at x/D = 7.5 and
x/D = 30: (a) Mixture fraction, (b) Axial velocity, (c) Temperature, (d) Turbulent
kinetic energy
Van Leer limiter for the mixture fraction equation (see section 6.2.2). A small
amount of 4-th order artificial dissipation (6.18) was added in the zone close to
the nozzle (exponentially decaying value from x/D = 0 to x/D = 2) to suppress
unphysical oscillations in the velocity field arising from the high velocity gradients
occurring at the edge of the fuel jet and pilot stream. These oscillations do not
destabilize the simulation since the turbulent viscosity is high enough to prevent
them from growing, but unphysical values of the velocity fluctuations close to the
nozzle are predicted. Their occurrence could be avoided by clustering the grid to
resolve the boundary layer of the fuel pipe and pilot annulus completely but this
would increase the computational cost significantly. The computational domain and
mesh are shown in figure 7.16. The mesh used is a cartesian, non-orthogonal, nonuniform, structured multi-block mesh. The computational domain is 60D long and
192 mesh cells are used in axial direction clustered towards the inlet so that the
first cells in the fuel jet are squared. The domain in y- and z-direction is diverging
form 4D 4D at the inlet to 8D 8D at the outlet and 80 cells are used in
both directions. The computational domain is thus significantly smaller than in the
RANS simulation to allow a sufficient mesh resolution and an acceptable computational effort. At the inlet the mesh in y- and z-directions is uniform in the fuel jet
and the pilot and in the co-flow it is clustered towards the pilot. The mesh is split
in 4 blocks in axial direction to allow parallelization of the computation.
A multigrid solver (section 6.3.1) is used for the Poisson equation using 5 multigrid levels with a number of smoothing sweeps determined to obtain an optimal
performance. The time derivative for density occurring in the rhs of the Poisson
equation has been spatially filtered using the test filter of the dynamic model [149].
At the inlet of the domain Dirichlet boundary conditions have been used imposing
the velocity and mixture fraction and extrapolating the pressure from within the
domain. All other relevant parameters are calculated from the mixture fraction.
The imposed instantaneous velocities are either constant in time (laminar inlet) and

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

123

Figure 7.16: Computational domain and mesh for LES of Sandia D flame

taken from the experimental database [180] or they are time varying according to
the method of Klein et al. [162] described in section 6.3.2. At the outlet Neumann
boundary conditions have been applied for all quantities except for pressure which
is imposed (atmospheric pressure). To avoid reflections the axial velocity has been
clipped at the outlet to prevent negative axial velocities. The combustion models
used for LES are the same as for RANS as described in the previous sections i.e.
the same tables for infinitely fast chemistry (mixed is burnt tables) and finite rate
chemistry (flamelet tables) have been used. However the determination of the parameters used to perform the look-up in these tables is different as already discussed in
section 4.5. The mixture fraction variance is not obtained from a transport equation
but from a scaling law model (4.76). The scalar dissipation is modeled according
to (4.83) derived from a simple equilibrium hypothesis with Sc = Sct = 0.7 and t
from the SGS model.
Figure 7.17 gives a qualitative idea of some of the important flow field features.
In the left picture an instantaneous iso-surface of stoichiometric mixture fraction is
shown together with the contours of axial velocity and temperature in the center
plane (instantaneous fields in a plane through the centerline, not averaged in circumferential direction). One can observe that the flame surface is strongly wrinkled
due to the turbulent flow field. The picture also reveals that the computational
domain is fairly narrow and the lateral boundary conditions might slightly influence
the flow results (this has not been checked with a bigger computational domain
due to the high computational cost). The temperature is strongly varying along
the flame surface. In the right picture the turbulent viscosity and density fields are
shown.

124

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

Figure 7.17: Instantaneous iso-surface of stoichiometric mixture fraction and (left)


axial velocity U (m/s) and temperature T (K), (right) density (kg/m3 ) and turbulent viscosity t (kg/ms) iso-contours for the LES of Sandia flame D

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

125

In total five different cases are considered as summarized in table 7.5. The cases
differ in:
the used combustion model: mixed-is-burnt or flamelet approach
the used inlet boundary condition for the velocity field: laminar constant
velocity field from experiment or turbulent velocity field generated using the
method of Klein et al. [162]
the scheme for the convective fluxes of the mixture fraction equation: first
order upwind or second order TVD with Van Leer limiter
Cases 3 and 4 differ by the used length scale in the turbulent inlet flow (see
section6.3.2) using 1.5 times shorter length scales in case 4 than in case 3 where
they are set to 0.2D. Most simulations have been time averaged over 75,000 CASE
1
2
3
4
5

Combustion Model
MIB
MIB
MIB
MIB
FLAMELET

Inlet
LAM
LAM
TURB
TURB
LAM

Convective scalar flux


UPW
TVD
TVD
TVD
TVD

Table 7.5: Different LES cases for Sandia flame D: MIB = mixed is burnt, LAM =
laminar inlet, TURB = turbulent inlet, UPW = first order upwind, TVD = second
order TVD with Van Leer limiter
100,000 time steps corresponding to about 4 - 6 through-flow times based on the
jet bulk velocity. For the mean quantities this can be regarded as sufficient but for
the rms values some more time averaging should be performed to obtain statistically
converged results (i.e. the mean rms values are not changing anymore when the time
averaging is continued). The reason for the limited total physical time is the allowed
time step which is determined by the convective speed and the cell size at the inlet.
All results shown along the symmetry line are only time averaged while the radial
plots are also averaged in circumferential direction (because of the axisymmetry of
the configuration). All simulations have been performed either on four CPUs of a
Transtec 8000L Itanium2 1.5 Ghz cluster or on four CPUs of a Hewlett-Packard
XC4000 Linux cluster with AMD Opteron dual core CPUs.
Figure 7.18 compares the time averaged LES results for the Sandia D flame along
the symmetry line for the infinitely fast chemistry combustion model with different
schemes for the scalar fluxes (cases 1 and 2). As one can see the potential core of
the jet is much too long for case 1 i.e. the velocity on the symmetry line starts
decaying only at x/D = 15. Because of the laminar inflow it takes a long distance
in axial direction before the jet core breaks up and the velocity starts decaying. As
already mentioned in section 6.3.2 this problem can only be solved by applying a
turbulent inflow with the same properties as the experimental inlet i.e. the same
turbulent intensities, length scales and energy spectra. An approximation for the

126

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

real turbulent inlet using artificially generated turbulence will be discussed further.
The rms2 of the axial velocity is almost zero for x/D < 10 and then strongly peaks
for x/D 20 to a value of about 3 times the experimental level, causing the strong
decay of mean axial velocity between x/D = 15 and x/D = 20. In figure 7.19
one can see that the overestimation of the rms of U is high at x/D = 15 but
further downstream (x/D = 30 and x/D = 45) the rms level is in good agreement
with the experimental values. For x/D > 20 the mean axial velocity is in good
agreement with the experiment. The delayed velocity decay is also immediately
visible in the mean mixture fraction which stays 1 (pure fuel) for x/D < 15 and then
starts decaying at the correct rate but with a mean mixture fraction which is above
the experimental value for the entire computational domain. The rms of mixture
fraction is zero for x/D < 10 and is overall too low compared to the experiment.
The behavior of the mixture fraction is reflected in the temperature which starts
increasing too far from the nozzle and the predicted flame length is clearly too big.
For case 2 a less diffusive scheme (2nd order TVD) was used for the convective fluxes
of the scalar equation and as can be observed this has a significant effect on the flame
structure. The rms level of mixture fraction is clearly in much better agreement
with the experiment and start increasing about 5D closer to the nozzle than in
case 1. Therefore the mean mixture fraction starts decaying closer to the nozzle
resulting in an underestimation of the mixture fraction for 20 < x/D < 45. This
can also be observed in the mean temperature, which is higher than the experimental
value for x/D < 40 but shows a better agreement than case 1. Also the rms level
of temperature is in better agreement and shows the two peaks which are also
observed in the experiment (the low rms point corresponds to the mean flame front
postion). Although the momentum equations are treated identically as in case 1 the
effect of the scalar scheme is also reflected in the velocity field. The jet break-up
occurs about 5D closer to the nozzle resulting in a mean velocity lower than the
experimental value for x/D > 15. In the radial plots of figure 7.19 one can see that
close to the nozzle (at x/D = 7.5) the mean velocity profile from both simulations
shows a strong deviation around r/D = 0.75 where the effect from the laminar inlet
profile is still visible. The peak in the rms of velocity for both cases is located at
r/D = 0.5 corresponding to the shear layer between the jet and pilot stream where
the rms of velocity is high in a wider r/D-range in the experiment and is maximal
for r/D = 1. The mixture fraction profile is also to narrow corresponding to the
persisting potential core. Overall the rms values of case 2 using the TVD scheme
are a little higher then case 1 using the more diffusive upwind scheme for the scalar
convective fluxes.
For the cases 1 and 2 a low value (0.09) for the constant C in (4.76) was used. In
[68] it is proposed to determine this constant dynamically. Although the dynamic
procedure could be applied locally it is mentioned that averaging in statistically
homogeneous directions should be used to obtain an accurate and stable value.
Because of the used cartesian type grid it is difficult to average in the homogeneous
circumferential direction and therefore the dynamic model has not been applied in
2

When referred to rms in this work, only the resolved fluctuations are meant.

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

127

(a)

(b)
0.15

rms z

1
0.5

0.1

0.05

0
0

20

(c)

40

0
0

60

1000
0
0

20

(e)

40

rms U

U/Uo

40

60

40

60

0.5

20

40
x/D

60

200
0
0
300

60

0
0

(d)

400

rms T

T(K)

2000

20

20

(f)

CASE 1
CASE 2
Exp.

200
100
0
0

20

40

60

x/D

Figure 7.18: Influence of scalar scheme : Sandia D LES results along symmetry
line, cases 1 and 2 compared with experimental data [178, 3, 180] (a) Mean Mixture
Fraction (b) Mixture Fraction rms (c) Mean Temperature (d) Temperature rms (e)
Mean Axial Velocity (f) Axial Velocity rms
this work. A discretization in a cylindrical type of grid would be more appropriate
to apply the dynamic procedures. To test the effect of the constant in (4.76), it was
increased so that a higher variance was obtained for the same mixture fraction field
(i.e. the mixture fraction gradients). This case is denoted 1b since all other settings
are identical to case 1. The results are compared in figures 7.20 and 7.21 . The rms
of mixture fraction is clearly higher than for case 1 and the rms of temperature is
in good agreement with the experimental level showing a maximum at both sides
of the mean flame position. The rms of velocity is even more overestimated then in
case 1 and starts increasing much closer to the nozzle. The jet breaks up closer to
the nozzle leading to a better prediction for the mean velocity on the symmetry line
but also to a much better agreement for the mixture fraction along the symmetry
line. Also the temperature is in good agreement on the rich side of the flame but
the decrease in temperature for x/D > 45 is underestimated. Clearly the variance
model has a significant effect and thus it would be interesting to investigate the
effect of a dynamic model or a model based on the scale similarity (4.75).
To investigate the effect of the inlet boundary conditions cases 1, 3 and 4 are
compared. The only difference between cases 3 and 4 are the length scales in the
introduced flow field i.e. the filter size used in the digital filtering procedure (section
6.3.2). Figure 7.22 shows a typical instantaneous axial velocity field at the inlet of
the domain. Note that in the figure, the cell averaged velocities are plotted. One
can immediately see that the description of the inlet velocity profile on the cartesian
mesh is not very accurate.

128

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

(a)

(b)

(a)

0.5

1 (e) 1.5

0.5

1.5
r/D

0
0

0.5

1 (f) 1.5

0
0

0.5

1 (e) 1.5

0.5

1.5
r/D

2.5

0
0

0.5

1.5
r/D

rms U
2 (e) 3

1000

3
r/D

z
2 (d) 3

0.4

0
0
30

50

0
0

rms T

2 (f) 3

rms z

0
0

200

3
r/D

2.5

0.5

1 (f) 1.5

2.5

0.5

2.5

400
200
0
0

2.5

(c)4

10
2

(e)4

0.1

0
0

4
r/D

(d)4

(f)4

4
r/D

40
20
0
0

1000
0
0

1.5
r/D

0.05

2000

(b)

20

0
0

400

0
0

1 (d) 1.5

100

2.5

CASE 1
CASE 2
Exp.

0.2
0
0
100

20

0.1

U(m/s)

2 (c) 3

0.5

(a)

T(K)

1000

0.6

rms z

z
U(m/s)

0
0
2000

2.5

0.1
0
0
200

2.5

20

(b)

40

T(K)

200

2.5

CASE 1
CASE 2
Exp.

40

0.2

0.5

0
0

1 (c) 1.5

60

(a)

0
0
2000

0.5

400

0
0

0
0

2.5

0.2

rms U

0
0

50

2.5

1000

1 (d) 1.5

rms U

20

0.5

rms T

T(K)

rms U

40

rms T

U(m/s)

60

0
0
2000

z
0
0
100

2.5

rms T

0.5

rms z

1 (c) 1.5

0.1

U(m/s)

0.5

CASE 1
CASE 2
Exp.

0.2

T(K)

CASE 1
CASE 2
Exp.

0.5
0
0

(b)

rms z

400
200
0
0

Figure 7.19: Influence of scalar scheme: radial profiles, cases 1 and 2 compared with
experimental data [178, 3, 180] at x/D = 7.5, x/D = 15, x/D = 30 and x/D = 45
(a) Mean Mixture Fraction (b) Mixture Fraction rms (c) Mean Axial Velocity (d)
Axial Velocity rms (e) Mean Temperature (f) Temperature rms

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

129

(a)

(b)
0.15

rms z

1
0.5
0
0

20

(c)

40

T(K)

rms T
20

(e)

40

(d)

40

60

40

60

rms U

0.5

20

40
x/D

60

200
0
0
300

60

U/Uo

20

400

1000

0
0

0.05
0
0

60

2000

0
0

0.1

20

(f)

CASE 1
CASE 1b
Exp.

200
100
0
0

20

40

60

x/D

Figure 7.20: Influence of variance model: Sandia D LES results along symmetry
line, cases 1 and 1b compared with experimental data [178, 3, 180] (a) Mean Mixture
Fraction (b) Mixture Fraction rms (c) Mean Temperature (d) Temperature rms (e)
Mean Axial Velocity (f) Axial Velocity rms
Looking at the mean velocity along the symmetry line in figure 7.23 one can see
that the effect of the turbulent inflow conditions is strong and the jet breaks up
much closer to the nozzle than in the experiment. Further downstream however
the mean axial velocity is too low compared to the experiment. The zone of high
velocity rms is much longer than for case 1 with a lower maximum value. This
effect of the turbulent inlet on the instantaneous flame structure can also be seen
in the right picture of figure 7.25 showing clearly that the jet break up occurs much
closer to the nozzle. However the mixture fraction decays to fast, resulting in an
underestimation of the mixture fraction almost in the entire flow field. One can see
in the temperature profile that now the flame length is underestimated. The strong
deviation in axial velocity at x/D = 7.5 and r/D = 0.75 observed for cases 1 and
2 does not occur for cases 3 and 4. The radial profile of mixture fraction clearly
shows the higher spreading of the jet corresponding to the stronger axial decay of
velocity and mixture fraction. Further downstream the jet spreading seems to be
too strong resulting in a strong underestimation of the axial velocity and mixture
fraction in the center of the jet. The difference between cases 3 and 4 is very small
so that it seems with the current implementation the length scales have little effect
on the flame (at least for the two investigated cases on a mesh which is rather coarse
to represent the structures introduced in the flow field). Although the method used
provides a more realistic inlet flow field it seems there is still some improvement
needed to obtain the correct development of the jet. The use of a cylindrical mesh
that can be clustered towards the walls of the fuel pipe and annulus would allow for

130

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING


(a)

(b)

(a)

0.5

1 (e) 1.5

0.5

1.5
r/D

0
0

0.5

1 (f) 1.5

0
0

1 (e) 1.5

0.5

1.5
r/D

2.5

0
0

0.5

1.5
r/D

rms U

20

2 (e) 3

1000

3
r/D

z
1

0.4

2 (d) 3

0
0
30

50

2 (f) 3

(c)4

rms z

0
0

200

3
r/D

0.5

1 (f) 1.5

2.5

0.5

2.5

200
0
0

2.5

10
2

(e)4

0.05

4
r/D

(d)4

(f)4

4
r/D

40
20
0
0

1000
0
0

1.5
r/D

0.1

0
0

2000

2.5

400

20

0
0

400

0
0

(b)

0.2

0
0

rms T

0.1

0
0
100

1 (d) 1.5

100

2.5

CASE 1
CASE 1b
Exp.

0.6

0.5

(a)

U(m/s)

2 (c) 3

1000

(b)

rms z

z
U(m/s)

0.5

0.1
0
0
200

2.5

20

0
0
2000

2.5

T(K)

40

0.2

40

T(K)

200

2.5

CASE 1
CASE 1b
Exp.

0.5

0
0

1 (c) 1.5

60

(a)

0
0
2000

0.5

400

0
0

0
0

2.5

0.2

rms U

0
0

50

2.5

1000

1 (d) 1.5

rms U

20

0.5

rms T

T(K)

rms U

40

rms T

U(m/s)

60

0
0
2000

z
0
0
100

2.5

rms T

0.5

rms z

1 (c) 1.5

0.1

U(m/s)

0.5

CASE 1
CASE 1b
Exp.

0.2

T(K)

0.5
0
0

(b)

1
CASE 1
CASE 1b
Exp.

rms z

400
200
0
0

Figure 7.21: Influence of variance model: radial profiles, cases 1 and 1b compared
with experimental data [178, 3, 180] at x/D = 7.5, x/D = 15, x/D = 30 and
x/D = 45 (a) Mean Mixture Fraction (b) Mixture Fraction rms (c) Mean Axial
Velocity (d) Axial Velocity rms (e) Mean Temperature (f) Temperature rms

Figure 7.22: Instantaneous inlet velocity field for Sandia D flame obtained with
digital filtering method [162] (axial velocity m/s)

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

131

(a)

(b)
0.15

rms z

1
0.5
0
0

20

(c)

40

T(K)

rms T
20

(e)

40

(d)

40

60

40

60

rms U

0.5

20

40
x/D

60

200
0
0
300

60

U/Uo

20

400

1000

0
0

0.05
0
0

60

2000

0
0

0.1

20

(f)

CASE 1
CASE3
CASE 4
Exp.

200
100
0
0

20

40

60

x/D

Figure 7.23: Influence of inflow boundary condition: Sandia D LES results along
symmetry line, cases 1, 3 and 4 compared with experimental data [178, 3, 180]
(a) Mean Mixture Fraction (b) Mixture Fraction rms (c) Mean Temperature (d)
Temperature rms (e) Mean Axial Velocity (f) Axial Velocity rms
a better representation of the inlet flow field. Instead of using the method by Klein
et al. [162] a method based on applying diffusion to a initial random turbulent field,
derived by Kempf et al. [167], could be used It allows to generate the turbulent field
directly on the used mesh (instead of an equidistant cartesian mesh) and to vary
the length scale within the pipe or annulus (in practice the length scales are smaller
close to the wall than in the center). Moreover a the use of a cylindrical grid allows
to simulate a short part of the upstream part of the fuel pipe and pilot annulus, so
that the artificial turbulence can develop even better to a realistic turbulent field.
Finally some off the Reynolds-stress components needed to scale the turbulent inlet
fields (Lund transformation (6.88)) are not available and thus have to be estimated
or set to zero.
Finally the effect of the combustion model on the flame is investigated in case 5.
All results are compared with case 1b since also for this run the higher constant in
the variance model has been used although the model for the scalar fluxes is identical
as in case 2 (unfortunately no results of the TVD scheme with higher variance and
infinitely fast chemistry are available to compare with). As one can see in figures 7.26
and 7.28 the mean axial velocity is in good agreement with the experimental values.
Again the rms levels of velocity are too high and rms levels of mixture fraction and
temperature are in much better agreement with the experiment. Although the mean
mixture fraction profile along the symmetry line is only slightly differing from case
1b, a much bigger difference is observed in the mean temperature.

132

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

(a)

(a)

(b)

0
0
2000

T(K)

0.5

1 (e) 1.5

1000

0
0

0.5

1.5
r/D

2.5

1 (d) 1.5

0
0

2.5

2 (c) 3

0.5

1 (f) 1.5

rms z

20

0
0
2000

2.5

2 (e) 3

200
0
0

0.5

1.5
r/D

2.5

1000

0
0

3
r/D

2 (d) 3

2 (f) 3

50

0
0

400

0.1

0
0
100

40

100

0
0

2.5

0.5

rms U

rms U

20

rms T

U(m/s)

60
40

z
0
0
200

2.5

0.5

rms T

0.1

U(m/s)

1 (c) 1.5

CASE 1
CASE 3
CASE 4
Exp.

0.2

T(K)

0.5

rms z

CASE 1
CASE 3
Case 4
Exp.

0.5
0
0

(b)
0.2

400
200
0
0

3
r/D

Figure 7.24: Influence of inflow boundary condition: radial profiles, cases 1, 3 and
4 compared with experimental data [178, 3, 180] at x/D = 7.5 and x/D = 30 (a)
Mean Mixture Fraction (b) Mixture Fraction rms (c) Mean Axial Velocity (d) Axial
Velocity rms (e) Mean Temperature (f) Temperature rms

Figure 7.25: Comparison of instantaneous flames for cases 1, 2 and 3: isolines of


mixture fraction (z = 0.05)

7.1. PILOTED NON-PREMIXED FLAME (SANDIA)

133

(a)

(b)
0.15

rms z

1
0.5
0
0

20

(c)

40

T(K)

rms T
20

(e)

40

(d)

40

60

40

60

rms U

0.5

20

40
x/D

60

200
0
0
300

60

U/Uo

20

400

1000

0
0

0.05
0
0

60

2000

0
0

0.1

20

(f)

CASE 1b
CASE 5
Exp.

200
100
0
0

20

40

60

x/D

Figure 7.26: Influence of combustion model : Sandia D LES results along symmetry
line, cases 1 and 5 compared with experimental data [178, 3, 180] (a) Mean Mixture
Fraction (b) Mixture Fraction rms (c) Mean Temperature (d) Temperature rms (e)
Mean Axial Velocity (f) Axial Velocity rms
For 10 < x/D < 40 the temperature predicted by the flamelet model is significantly higher than for the mixed is burnt model. Looking at the mixture fraction
rms one can see that it is overall a bit lower and thus a higher temperature may
be expected (see figures 7.11 and 7.3 where for increasing variance and fixed mean
mixture fraction the temperature will drop). However, the higher temperature obtained for the flamelet model is mainly caused by an underestimation of the scalar
dissipation rate. If one looks at the temperature in mixture fraction space for the
laminar flame solution in figure 7.14 one can see that if the scalar dissipation rate is
underestimated, i.e. a flamelet with low strain is extracted from the table, a much
higher temperature than for the mixed is burnt model will be obtained for a certain mixture fraction on the rich side (z > zst ). Figure 7.27 shows an instantaneous
scalar dissipation field together with the iso-line of stoichiometric mixture fraction at
x/D = 20. One can see that the highest values of occur around the stoichiometric
mixture fraction but that strongly varies along this iso-line. Thus the temperature
obtained from the flamelet table will strongly vary along the flame front. However
the value of the scalar dissipation is rather low and thus mainly flamelets from the
low strain side of the table will be accessed in the lookup-table. As seen from 7.14,
for rich mixtures (inside the zst isoline on the picture), the temperature will be higher
than with the mixed-is-burnt approach. To validate the used model for the scalar
dissipation rate one should compare statistically averaged values or conditional averages (conditioned on mixture fraction) of with experimental values. Unfortunately

134

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

Figure 7.27: Instantaneous scalar dissipation rate field and iso-line of stoichiometric
mixture fraction at x/D = 20
these quantities are not available yet from measurements and the models described
in section 4.5 are mainly validated based on DNS results. Recently laser Rayleigh
measurements of thermal gradient structures have been performed in the near-field
of a jet flame by Frank and Kaiser [184, 185] which could be of great value to validate the scalar dissipation modeling.
One can see that the influence of the model used for scalar fluxes, the inlet
velocity fields applied and the combustion model all have a significant impact on
the simulation results for Sandia flame D. A good representation of the turbulent
inlet velocity field seems to be essential to predict the correct spreading of the
turbulent jet and to obtain the correct velocity and mixture fraction decay on the
centerline. Although the mean velocity fields are in reasonably good agreement with
the experimental results, the rms of axial velocity is strongly overestimated in all
simulations between x/D = 10 and x/D = 30. The flame length is very sensitive to
the prediction of the mixture fraction field and thus to changes in inlet conditions
and numerical modeling.

7.2

Non-premixed jet flames (DLR)

The second case considered is a simple jet flame denoted DLR flame, which is also
one of the well documented target flames of the TNF workshop [3]. The burner
consists of a 35 cm long straight stainless steel tube with an inner diameter of 8
mm and a thinned rim at the exit. The tube was surrounded by a contoured nozzle
with an inner diameter of 140 mm supplying co-flowing dry air at an exit velocity
of 0.3m/s.
The fuel is a mixture of 22.1% CH4 , 33.2% H2 and 44.7% N2 by volume. The
stoichiometric mixture fraction is 0.167 which is significantly lower than for the Sandia flame series and therefore the temperature and density variations with mixture
fraction are large close to the air side (0 < z < zst ). Methane was used as a fuel
component in order to study the interaction of its chemistry with turbulent flow
and hydrogen was added to stabilize the flame, without changing the simple flow

7.2. NON-PREMIXED JET FLAMES (DLR)


(a)

(b)

(a)

0.5

1 (e) 1.5

0.5

1.5
r/D

0
0

0.5

1 (f) 1.5

0
0

0.5

1 (e) 1.5

0.5

1.5
r/D

2.5

0
0

0.5

1.5
r/D

rms U

20

2 (e) 3

1000

3
r/D

z
1

2 (d) 3

2 (f) 3

(c)4

200

3
r/D

0.5

1 (f) 1.5

2.5

0.5

2.5

200

rms z

0
0

10
2

(e)4

0.05

4
r/D

(d)4

(f)4

4
r/D

40
20
0
0

1000
0
0

1.5
r/D

0.1

0
0

2000

2.5

400

20

0
0

400

0
0

(b)
CASE 1b
CASE 5
Exp.

0.4

0
0
30

50

0
0

2.5

0.2

0
0

rms T

0.1

0
0
100

1 (d) 1.5

(a)
0.6

0.5

100

rms U

0.1

2.5

1000

U(m/s)

2 (c) 3

0.2

0
0
200

2.5

20

0
0
2000

2.5

T(K)

40

(b)

rms z

z
U(m/s)

1 (c) 1.5

0.2

40

T(K)

200

2.5

CASE 1b
CASE 5
Exp.

0.5

0
0

0.5

60

(a)

0
0
2000

0
0

2.5

400

0
0

50

2.5

1000

1 (d) 1.5

rms U

20

0.5

rms T

T(K)

rms U

40

rms T

U(m/s)

60

0
0

z
0
0
100

2.5

rms T

0.5

rms z

1 (c) 1.5

0.1

U(m/s)

0.5

CASE 1b
CASE 5
Exp.

0.2

T(K)

0.5

0
0
2000

(b)

1
CASE 1b
CASE 5
Exp.

rms z

0
0

135

400
200
0
0

Figure 7.28: Influence of inflow boundary condition: radial profiles, cases 1 and 5
compared with experimental data [178, 3, 180] at x/D = 7.5, x/D = 15, x/D = 30
and x/D = 45 (a) Mean Mixture Fraction (b) Mixture Fraction rms (c) Mean Axial
Velocity (d) Axial Velocity rms (e) Mean Temperature (f) Temperature rms

Figure 7.29: The DLR flame operated in the lab

136

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

field of the round jet. In other configurations reported in the literature, methane
flames with high Re numbers were prevented from lifting or extinguishing by pilot
flames or recirculation zones which have the drawback of complicating the flow field
or chemistry and, hence, the mathematical simulation. Finally, dilution by nitrogen
was used to decrease thermal radiation, inhibit soot formation and to improve the
signal quality of the spontaneous Raman scattering technique.
Two different cases, denoted DLR A and DLR B, have been experimentally investigated with a jet bulk velocity of respectively Ujet = 42.2m/s and Ujet = 63.2m/s
and corresponding Reynolds numbers of 15, 200 and 22, 800. The higher Re case is
close to the blow-off condition and has some localized extinction.
Scalar Point Measurements were performed at Sandia laboratories: Simultaneous
Raman/ Rayleigh/ LIF measurements of T , N2 , O2 , CO2 , H2 O, H2 , CO, OH, and
NO were obtained with a spatial resolution of 0.75 mm. Results include axial profiles (x/D = 2.5 up to x/d = 120) and radial profiles (x/d = 5, 10, 20, 40, 60, 80)
of mean and rms values, conditional statistics, and single-shot data for each flame.
Typically, 800-1000 samples were acquired at each location. Scalar Imaging was
performed at DLR including selected planar images of temperature (Rayleigh scattering), OH, CH, and NO. These provide information on the spatial structure of the
flames. Two-component LDV measurements of velocities were conducted at the TU
Darmstadt at locations in the flames corresponding to the scalar measurements and
at additional locations closer to the nozzle. In this work only the DLR A case has
been investigated.

7.2.1

RANS simulations

All the settings for the flow solver are identical to the ones used for the Sandia flame
(see section 7.1.1). The computational domain is 120D in axial direction, 30D in
radial direction and has a opening angle in circumferential direction of 1o . The grid
contains 80 cells in radial direction of which 20 are uniformly distributed in the
fuel jet (r = 0.2mm) and in the co-flow the remaining 60 cells are clustered so
that the first cell of the co-flow has the same r as in the fuel jet. In axial direction 100 cells are used clustered towards the inlet so that the first cells in the fuel
stream are square (r = x). The inlet profiles are fully computed with the set
of equations (6.84-6.86). Two combustion models are used: the equilibrium model
(infinitely fast reversible chemistry) and the flamelet model (finite rate chemistry).
The flamelet table has a low strain limit of a = 1.0 (st = 0.04) and a high strain
limit of a = 1585.0 (st = 37.0). The coupling with the flow and lookup in the
combustion tables is performed as described for the Sandia flame.
Figures 7.30 and 7.31 present the results for the RANS simulations of the DLR A
flame for both chemistry models. As one can see the decay of the centerline velocity
is too strong and as a result also the mixture fraction along the centerline decays too
fast. The spreading of the jet is too strong as can be clearly seen in the radial velocity
profiles (figures 7.31 a and b) at x/D = 5 and x/D = 20. This results in a maximum
temperature occurring too close to the nozzle, and far downstream the nozzle in a

7.2. NON-PREMIXED JET FLAMES (DLR)

137

(a)

(b)

"2

0.5

0
0

0.1

50

(c)

0.05

0
0

100

1000

0
0

50

(e)

100

50

100

0.5
0
0

100

(f)

60

EQUI
FLAM
EXP

40

u/uo

50

2000

(d)

0.5

20
0
0

50

x/D

100

0
0

20

40

60

80

x/D

Figure 7.30: RANS results along symmetry axis for DLR A: comparison of equilibrium model and flamelet model with experimental data from [3, 186]): (a) mean
mixture fraction (b) variance of mixture fraction (c) temperature (d) density (e)
axial velocity (f) turbulent kinetic energy
strong underestimation of the mixture fraction and thus temperature. The strong
drop in temperature for x/D > 50 is caused by the steep gradient of temperature in
mixture fraction space at the lean side of the flame. Indeed, because of the low value
of zst , a change of 0.01 in mixture fraction at the lean side corresponds to a change
of about 100K in temperature while at the rich side it corresponds to a change of
less than 20K. Since the mixture fraction seems to decay too fast for x/D > 50 the
temperature drops strongly. The use of the flamelet model does not significantly
change the results and a better prediction of the mixture fraction field is needed to
improve the results. A mesh refinement and improvement of the turbulence model
could result in an improved prediction of the velocity field and thus of the scalar
fields but this has not been investigated in this work. One should also remember
that both the radial gradients of velocity and mixture fraction close to the nozzle
are much greater than for the Sandia flame because no pilot is present. Therefore
the numerical solution is more sensitive to the mesh and numerical scheme, and the
combination with the strong temperature gradient in mixture fraction space on the
lean side, results in an overall worse comparison with experimental results.

7.2.2

LES simulations

The numerical methods used are the same as for the Sandia D flame (see section
7.1.2). The computational domain is 80D long and 192 mesh cells are used in axial
direction clustered towards the inlet so that the first cells in the fuel jet are squared.

138

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING


(a)

(a)

(b)
EQUI
FLAM
EXP

0.8

40

0.6

30

0.4

20

0.4

0.2

10

0.2

1.5

0
0

2.5

0.5

(c)

1.5

0
0

2.5

0
0

(d)
60

40

1000

20
500

1.5

0
0

2.5

0.5

(a)

1.5

2.5

0
0

(b)

0.5
0.4

0
0

10

10

0.1

5
2

0
0

0
0

(d)

1000

0
0

10

(c)
20

15

1500

15

1000

10

10

0
0

500

0
0

10

(d)

2000

500

20

1500

0.05

(c)
2000

0.15

0.2
0.1

(b)
EQUI
FLAM
EXP

0.2

15

0.25

0.3

2
(a)

20
EQUI
FLAM
EXP

0.5

20

500

1500

40

0
0

2000

1500

0
0

(c)

60

0
0

10

(d)

1000

20

30

0.6

0.5

EQUI
FLAM
EXP

50

0.8

0
0

(b)

5
5

10

0
0

Figure 7.31: RANS results: Radial profiles for flame DLR A (equilibrium model
and flamelet model with experimental data from [3, 186]) at x/D = 5, x/D = 20,
x/D = 40 and x/D = 60: (a) Mixture fraction, (b) Axial velocity, (c) Temperature,
(d) Turbulent kinetic energy
The domain in y- and z-direction is diverging form 6D 6D at the inlet to
16D 16D at the outlet and 80 cells are used in both directions. The domain is
thus a bit longer and significantly wider than for the Sandia D case, to prevent a too
strong influence of the boundary conditions on the simulation results (for the Sandia
case a test with an extended computational domain should be performed to address
the impact of the domain boundaries). At the inlet the mesh in y- and z-directions
is uniform in the fuel jet (20 cells in the fuel jet diameter) and in the co-flow it is
clustered towards the fuel jet so that the first cell has the same size as the fuel jet
cells. The mesh is split in 4 blocks in axial direction to allow parallelization of the
computation.
For the DLR A case only 3 different cases have been investigated to address the
effect of the scalar convective scheme and the influence of the combustion model.
Since the turbulent inflow generation still needs some improvement as discussed
earlier for the Sandia flame, it has not been used for this case and a laminar inlet

10

7.2. NON-PREMIXED JET FLAMES (DLR)

139

Figure 7.32: Computational domain and mesh for LES of DLR A flame
velocity profile with experimental values from [186] for axial velocity has been used.
The settings for the 3 cases have been summarized in table 7.6.
CASE
1
2
3

Combustion Model
EQUI
EQUI
FLAMELET

Inlet
LAM
LAM
LAM

Convective scalar flux


UPW
TVD
UPW

Table 7.6: Different LES cases for DLR A flame: EQUI = equilibrium model, LAM
= laminar inlet, UPW = first order upwind, TVD = second order TVD with Van
Leer limiter
Figure 7.33 gives a qualitative idea of the flame, showing an instantaneous isosurface of stoichiometric mixture fraction together with the contours of axial velocity
and temperature in the center plane. One can observe that the flame surface is
strongly wrinkled due to the turbulent flow field. The picture also reveals that the
computational domain is sufficiently wide to capture the entire flame and thus the
influence of the lateral boundary will be much smaller than for the Sandia case.
First the influence of the scalar scheme is investigated, comparing cases 1 and
2. As one can see from figure 7.34 the results are very similar to the Sandia case.
Because of the laminar inflow, the core of the jet persists for more than 10D, and the
centerline velocity start decaying to far from the nozzle. This is also clearly observed
in the radial plots of axial velocity in figure 7.35 for x/D = 5 and x/D = 10.
Further downstream the radial velocity profiles are in better agreement with the
experimental results. For 10 < x/D < 20 the rms of velocity is much too high and
the axial velocity decays rapidly along the centerline resulting in an underestimation
of the axial velocity from x/D = 20 on. This is also reflected in the mixture fraction
decay. As was the case for the Sandia flame the results improve when a less diffusive

140

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

,
Figure 7.33: Instantaneous iso-surface of stoichiometric mixture fraction and (left)
axial velocity U (m/s) and temperature T (K), (right) density (kg/m3 ) and turbulent viscosity t (kg/ms) iso-contours for the LES of DLR A flame
TVD scheme is applied resulting in a more realistic level of fluctuations of mixture
fraction (rms of z). In the radial plots of figure 7.35 it can be seen that there is
a strong underestimation of the temperature rms close to the flame (were z zst
and T Tmax ). Looking at the mean temperature one can observe that for case 2
a very good agreement with the experiment is found. For case 1 a stronger decay
in mixture fraction is observed for x/D > 50 resulting in a lower temperature and
more important a very strong peak in the rms of temperature (and also in the rms
of z). This is caused by the strong gradient of temperature in mixture fraction space
as already discussed in section 7.2.1. At the lean side of the flame the results are
far more sensitive to the prediction of mixture fraction than on the rich side.
As in the case of the Sandia flame, the influence of the constant used in the
variance model was investigated and this case is denoted 1b. It should be noted
that the presented results have only been time averaged over 50,000 time steps and
thus cannot be regarded as statistically converged. Figures 7.36 and 7.37 present
respectively the comparison of case 1 and 1b along the centerline and in the radial
cuts where experimental data is available. It can be observed that the mixture
fraction along the centerline is in better agreement with the experiment but that
the axial velocity is even more strongly underestimated than for case 1. The mean
temperature shows a strange bump between x/D = 20 and x/D = 50 which cannot be explained immediately from the mean mixture fraction field and its rms.
Possibly with a longer time averaging the results would improve and some transient
effects might have had an influence on the current results (i.e. the time averaging
was started before the flow and flame were completely adapted to the new model).
However the too strong peak in temperature rms for x/D > 50 observed in case 1 is
not present anymore and in the radial plots at x/D = 5, x/D = 10 and x/D = 20 a
much better agreement of the temperature rms with experimental values is observed.

7.3. CONCLUSIONS

141
(a)

(b)
0.1

rms z

1
0.5
0
0

20

(c)

40

0
0

60

rms T

T(K)

1000

20

(e)

40

rms U

U/Uo

(d)

40

60

40

60

0.5

20

40
x/D

60

200
0
0

60

0
0

20

400

2000

0
0

0.05

20

(f)

CASE 1
CASE 2
Exp.

100
50
0
0

20

40

60

x/D

Figure 7.34: Influence of scalar scheme : DLR A LES results along symmetry line,
cases 1 and 2 compared with experimental data [3, 186] (a) Mean Mixture Fraction
(b) Mixture Fraction rms (c) Mean Temperature (d) Temperature rms (e) Mean
Axial Velocity (f) Axial Velocity rms
Finally the influence of the combustion model is investigated comparing the
results for the equilibrium chemistry model and the flamelet model. Comparison is
made with case 1b since for the flamelet computations also the modified variance
model has been used. Again the number of iterations over which the averaging has
been performed is limited (50,000) so that results might change if further averaging
is performed. One can see in figure 7.38 that with the flamelet model the centerline
velocity decay is a little smaller corresponding to the lower velocity rms. The mixture
fraction fraction is a little higher for x/D < 30 and a bit smaller than for the
equilibrium case for x/D > 30. The temperature along the centerline for the flamelet
case is in better agreement with the experiment. In the radial plots of figure 7.39
one can see that close to the nozzle (x/D = 5 and x/D = 10) the difference between
the two cases is rather small for the mean quantities but the rms values differ
significantly. However the differences caused by the model might be overwhelmed
by the effect of insufficient averaging since this has a more pronounced effect of the
rms values than on the mean values.

7.3

Conclusions

A low Mach-number approximation solver for RANS simulations and LES of nonpremixed turbulent combustion has been developed and successfully applied to two
target flame test cases defined in the framework of the TNF-workshop [3]. To model
the turbulent combustion three different models have been used: infinitely fast irre-

142

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

(a)

(b)

(a)

0.5

1 (e) 1.5

0.5

1.5
r/D

0
0

0.5

1 (f) 1.5

0
0

0.5

1 (e) 1.5

0.5

1.5
r/D

2.5

0
0

0.5

1.5
r/D

rms U
2 (e) 3

1000

3
r/D

z
2 (d) 3

0.4

0
0
30

50

0
0

rms T

2 (f) 3

rms z

0
0

200

3
r/D

2.5

0.5

1 (f) 1.5

2.5

0.5

2.5

400
200
0
0

2.5

(c)4

10
2

(e)4

0.1

0
0

4
r/D

(d)4

(f)4

4
r/D

40
20
0
0

1000
0
0

1.5
r/D

0.05

2000

(b)

20

0
0

400

0
0

1 (d) 1.5

100

2.5

CASE 1
CASE 2
Exp.

0.2
0
0
100

20

0.1

U(m/s)

2 (c) 3

0.5

(a)

T(K)

1000

0.6

rms z

z
U(m/s)

0
0
2000

2.5

0.1
0
0
200

2.5

20

(b)

40

T(K)

200

2.5

CASE 1
CASE 2
Exp.

40

0.2

0.5

0
0

1 (c) 1.5

60

(a)

0
0
2000

0.5

400

0
0

0
0

2.5

0.2

rms U

0
0

50

2.5

1000

1 (d) 1.5

rms U

20

0.5

rms T

T(K)

rms U

40

rms T

U(m/s)

60

0
0
2000

z
0
0
100

2.5

rms T

0.5

rms z

1 (c) 1.5

0.1

U(m/s)

0.5

CASE 1
CASE 2
Exp.

0.2

T(K)

CASE 1
CASE 2
Exp.

0.5
0
0

(b)

rms z

400
200
0
0

Figure 7.35: Influence of scalar scheme on DLR A flame: radial profiles, cases 1
and 2 compared with experimental data [3, 186] at x/D = 5, x/D = 10, x/D = 20
and x/D = 40 (a) Mean Mixture Fraction (b) Mixture Fraction rms (c) Mean Axial
Velocity (d) Axial Velocity rms (e) Mean Temperature (f) Temperature rms

7.3. CONCLUSIONS

143
(a)

(b)
0.1

rms z

1
0.5
0
0

20

(c)

40

0
0

60

rms T

T(K)

1000

20

(e)

40

rms U

U/U

(d)

40

60

40

60

0.5

20

40
x/D

60

200
0
0

60

0
0

20

400

2000

0
0

0.05

20

(f)

CASE 1
CASE 1b
Exp.

100
50
0
0

20

40

60

x/D

Figure 7.36: Influence of variance: DLR A LES results along symmetry line, cases
1 and 1b compared with experimental data [3, 186] (a) Mean Mixture Fraction (b)
Mixture Fraction rms (c) Mean Temperature (d) Temperature rms (e) Mean Axial
Velocity (f) Axial Velocity rms
versible chemistry (mixed is burnt), infinitely fast reversible chemistry (equilibrium
model) and finite rate chemistry (flamelet model). Tools have been developed to
generate the combustion lookup tables which are accessed by the flow solver with the
parameters describing the combustion (mixture fraction, mixture fraction variance
and for finite rate effects the scalar dissipation rate).
For the RANS simulations the effect of the inlet boundary conditions for velocity, turbulent kinetic energy and its dissipation have been investigated. For the
Sandia D flame the best results were obtained with inlet profiles which had been
separately computed for a fully developed turbulent pipe and annulus. Therefore all
other RANS simulations have been performed with these inlet profiles. Secondly the
effect of some parameters of the used k- turbulence model have been investigated
and it was confirmed that mainly the C2 parameter in the transport equation for
the dissipation of kinetic energy has an important influence. Since the investigated
flames are unconfined no attention was paid to correct wall modeling (this would
certainly be important for bluff body flames or confined burners). The effect of the
combustion model has been investigated and seems not to be very strong in the
investigated flames. The reason is that no important finite rate chemistry effects
are present in the Sandia D flame or in the DLR A flame. For the higher Reynolds
number cases (Sandia E and F, DLR B) the combustion model will have a more
significant influence but this has not been investigated. Where for the Sandia D
flame the results are overall in good agreement with the experimental results, the
RANS simulations of the DLR A flame show a strong deviation for x/D > 50. This
is mainly caused by the deviation of the mixture fraction field far downstream the

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING


(a)

(b)

(a)

0.5

1 (e) 1.5

0
0

0.5

1.5
r/D

0
0

0.5

1 (f) 1.5

0
0

0.5

1 (e) 1.5

0.5

1.5
r/D

2.5

0
0

0.5

1.5
r/D

rms U

20

2 (e) 3

1000

3
r/D

z
1

2 (d) 3

0.4

0
0
30

50

2 (f) 3

200

3
r/D

2.5

0.5

1 (f) 1.5

2.5

0.5

2.5

400
200

rms z

0
0

(c)4

10
2

(e)4

0.1

0
0

4
r/D

(d)4

(f)4

4
r/D

40
20
0
0

1000
0
0

1.5
r/D

0.05

2000

(b)

20

0
0

400

0
0

0
0

2.5

CASE 1
CASE 1b
Exp.

0.2

0
0

rms T

0.1

0
0
100

1 (d) 1.5

(a)
0.6

0.5

100

2.5

1000

U(m/s)

2 (c) 3

0.1
0
0
200

2.5

20

0
0
2000

2.5

T(K)

40

(b)

rms z

z
U(m/s)

1 (c) 1.5

0.2

40

T(K)

200

2.5

CASE 1
CASE 1b
Exp.

0.5

0
0

0.5

60

(a)

0
0
2000

0
0

2.5

400

0
0

50

2.5

1000

1 (d) 1.5

rms U

20

0.5

0.2

rms U

T(K)

rms U

40

rms T

U(m/s)

60

0
0
2000

z
0
0
100

2.5

rms T

0.5

rms z

1 (c) 1.5

0.1

U(m/s)

0.5

CASE 1
CASE 1b
Exp.

0.2

T(K)

0.5
0
0

(b)

1
CASE 1
CASE 1b
Exp.

rms z

rms T

144

400
200
0
0

Figure 7.37: Influence of variance on DLR A flame: radial profiles, cases 1 and 1b
compared with experimental data [3, 186] at x/D = 5, x/D = 10, x/D = 20 and
x/D = 40 (a) Mean Mixture Fraction (b) Mixture Fraction rms (c) Mean Axial
Velocity (d) Axial Velocity rms (e) Mean Temperature (f) Temperature rms
nozzle and the strong gradient of the temperature at the lean side in mixture fraction space.
For the LES of Sandia flame D the influence of the scalar convection scheme,
the inlet boundary condition, the model constant in the variance model and the
combustion model have been investigated. With the second order TVD scheme for
the convective fluxes of mixture fraction, better results are obtained than with the
more diffusive first order upwind scheme. Combining a laminar inflow boundary
condition and the upwind scheme for scalar fluxes results in break-up of the potential core of the jet much too far from the nozzle. Therefore the jet spreading
is wrongly predicted and the flame length does not correspond to the experimental
results. This can be clearly observed in the mean results and in the instantaneous
structure of the the flame and flow field. A more adequate representation of the
inlet velocity field has been applied using artificially generated turbulence (to avoid
the high computational cost of separately simulating the upstream geometry). The
effect of a turbulent inflow field on the jet break-up is immediately observed in the

7.3. CONCLUSIONS

145

(a)

(b)
0.1

rms z

1
0.5
0
0

20

(c)

40

0.05
0
0

60

rms T

T(K)

(d)

40

60

40

60

400

2000
1000
0
0

20

(e)

40

0
0

rms U

o
0.5
0
0

20

200

60

U/U

20

40

20

(f)

CASE 1b
CASE 3
Exp.

100
50
0
0

60

20

40

x/D

60

x/D

Figure 7.38: Influence of combustion model: DLR A LES results along symmetry
line, cases 1b and 3 compared with experimental data [3, 186] (a) Mean Mixture
Fraction (b) Mixture Fraction rms (c) Mean Temperature (d) Temperature rms (e)
Mean Axial Velocity (f) Axial Velocity rms

(a)

(b)

(a)

0
0
2000

T(K)

rms U

20
0.5

1 (e) 1.5

2.5

1000

0
0

0.5

1.5
r/D

2.5

0.5

1 (d) 1.5

0
0

2.5

0.5

1 (c) 1.5

0.5

1 (f) 1.5

rms z

40
20

0
0
2000

2.5

0.5

1 (e) 1.5

200
0
0

0.5

1.5
r/D

2.5

1000

0
0

0.5

1.5
r/D

2.5

0.1

0.5

1 (d) 1.5

2.5

0.5

1 (f) 1.5

2.5

0.5

2.5

100
50
0
0

2.5

400

0.2

0
0
150

2.5

60

50

0
0

rms T

U(m/s)

60
40

z
0
0
100

2.5

rms U

0.5

rms T

1 (c) 1.5

0.1

U(m/s)

0.5

CASE 1b
CASE 3
Exp.

0.2

T(K)

0.5
0
0

(b)

1
CASE 1b
CASE 3
Exp.

rms z

400
200
0
0

1.5
r/D

Figure 7.39: Influence of combustion model on DLR A flame: radial profiles, cases
1b and 3 compared with experimental data [3, 186] at x/D = 5 and x/D = 10 (a)
Mean Mixture Fraction (b) Mixture Fraction rms (c) Mean Axial Velocity (d) Axial
Velocity rms (e) Mean Temperature (f) Temperature rms

146

CHAPTER 7. RESULTS NON-PREMIXED COMBUSTION MODELING

instantaneous flow field and results in a much faster decay of the velocity along
the centerline. However no satisfying correspondence of the flow field with the
experimental results is obtained. The use of a cartesian type mesh, a too coarse
mesh, a non dynamic SGS model and an inadequate setting of the length scales in
the inflow stream (which can not be varied with the radius in the currently used
method) might all contribute to the strong deviation with the experiment. More
detailed testing (maybe for a non-reacting jet flow) of the methods used should be
performed to address the separate effect of the above mentioned shortcomings. The
effect of the model constant in the variance model seems also to be significant and
a more appropriate dynamic model should be implemented so that it is automatically adjusted to provide the correct variance of the mixture fraction field. The
effect of the combustion model on the velocity and mixture fraction fields seems to
be small but a strong difference in temperature is observed. This effect is caused
by the flamelet table which predicts a much higher temperature at the rich side of
the flame for low values of the scalar dissipation rate. Since the mixture fraction
and variance fields are not very different it is likely that the strong deviation of the
temperature is caused by an underestimation of the scalar dissipation rate. Also
here the non-dynamic SGS model might have an influence since the scalar dissipation rate depends on the turbulent viscosity. It would be interesting to compare the
obtained scalar dissipation field with experimental values to refine the models. The
simulations of the DLR A case reveal similar effects of the investigated parameters
but some results have been insufficiently time averaged due to the small time step
and high computational cost.

Chapter 8
Results Premixed Combustion
Modeling
8.1

Introduction

A difficult problem encountered in LES of premixed flames is that the typical flame
thickness L0 is generally a lot smaller than the mesh size . Thus the flame front
can not be resolved on the computational mesh leading to numerical problems. To
overcome this, several approaches have been described in the literature: the flame
front tracking technique (G-equation) [97], the use of a filter larger than the grid
resolution [98] and an artificial thickening of the flame [93, 99, 100]. The three models have been discussed in detail in section 5.3.2.
In this dissertation a Large Eddy Simulation together with the Artificially Thickened Flame approach is used to study a well known experimental set-up consisting of
a rectangular dump combustor - ORACLES. The artificially thickened flame (ATF)
model, first introduced by Butler et al [110] is very attractive due to its inherent numerical stabilizing features. The major drawback of artificially thickening the flame
is that the interaction between turbulence and flame is altered. To compensate for
the inability of small vortices to wrinkle the flame a subgrid scale wrinkling model
has to be introduced.
The ORACLES configuration has been investigated by Pitsch et al [187], under
partially premixed conditions, by Domingo et al [188] , testing a new model for PDF-sgs-closure and by D
using [189], who used the G-equation approach to predict
the flame and flow field and also by Fureby [190], who tested a recently proposed
flame-wrinkling model. In this work the influence of different approximations for
the subgrid scale velocity, used to locally estimate the subgrid scale wrinkling and
thus the turbulent burning velocity, on the prediction of the flow and combustion
field, is studied.
Simulations were performed with the FLOWSI-code developed at the TU Darmstadt and the main characteristics of this code have been described in section 6.4.
147

CHAPTER 8. RESULTS PREMIXED COMBUSTION MODELING

130, 6

70, 4

z
y

29, 9

39, 7

14

30, 4

30, 4

29, 9

148

Figure 8.1: 2D sketch of the ORACLES burner, the channel width in y direction is
150, 5mm (all dimensions are in mm)
Properties
Re
Ubulk
u
b
Tu
Tb
u
b
s0L

Value
Unit
20000
11.0
m/s
1.296
kg/m3
0.166
kg/m3
276
K
1980
K
1.66 E-5 m2 /s
6.40 E-5 m2 /s
0.27
m/s

Table 8.1: Relevant physical properties for the ORACLES configuration

8.2

Oracles Configuration

The configuration investigated is documented in the literature as ORACLES (One


Rig for Accurate Comparisons with Large Eddy Simulations) burner [111, 191, 192,
193, 194]. It consists of two separate mixing chambers, from where a propane-air
mixture enters the combustion chamber as two fully developed turbulent channel
flows. Behind the two meter long separated channels the splitter plate narrows with
an angle of 14 . Further downstream the flow is exposed to a sudden expansion,
with a step height of D= 29.9 mm. For a better understanding, a sketch of the
ORACLES combustion chamber is presented in Fig. 8.1.
For the isothermal and for the reacting case stream-wise and transverse velocity
components have been measured using LDV. At position 5D and 5.8D upstream
of the backward facing step the data set of the upper and lower channel exhibit a
slight difference, which has to be considered when generating inflow data for LES.
For the combustion chamber mean velocities and fluctuations have been reported
between 1D and 10D downstream from the step [111, 193].
Commercial propane was used as fuel. The global, average composition can be
expressed by C3.01 H7.94 . A more detailed itemization including the exact species

8.2. ORACLES CONFIGURATION

149

concentration is given in [194] 1 . The equivalence ratio in the two channels can be
varied independently within the limits of = 0.0 and = 0.85. A variation of the
equivalence ratio allows to study combustion under partially premixed conditions
and the occurrence of extinction.
The present simulation deals with a constant equivalence ratio of = 0.75 for
both inflow channels. In this case the flame will anchor in the shear layer at the edge
of the expansion. The Reynolds number is calculated based on the flow properties of
the unburnt mixture, the bulk velocity of the channel and the height of the sudden
expansion (29.9 mm). In the investigated case the Reynolds number is set to be
equal 20, 000, assuming a bulk velocity of 11 m/s for both channels. More details of
the experimental set-up can be found in Besson et al [111] and Bruel et al [191].
Using CHEM1D the physical properties of a 1D laminar flame were calculated
with the GRI3.0 mechanism. All relevant physical properties, used in the simulation
are summarized in table 8.1.
To classify the quality of the combustion LES the computational filter size has to
be compared with flame and turbulence properties, as recommended by the TNF7
workshop [195]. To satisfy this requirement one can use the experimental classification of the burner in the Borghi-Peters [196, 97] Diagram presented in the report of
Bruel et al [193] and discussed in section 5.2. The ratio of the integral length scale
lt over the flame thickness lf is around 500, and the turbulence level u0 over laminar
flame speed sL is given to be approximately 9. Therefore the region of interest in
figure 5.4 is located close to a Karlovitz number of unity, which describes the border between the corrugated flamelets and the thin reaction zone. The fundamental
requirement for LES is that the filter length is larger than the Kolmogorov scale. A
further requirement of Pitsch et al [197] is that the filter size is larger than the laminar flame thickness. Taking the Karlovitz number of one, and assuming the filter
size to be in the order of 1 mm, the ratio of filter size over laminar flame thickness
is calculated to be approximately 7. A classification into the Regime Diagram of
Pitsch et al [197] or D
using et al [198], extended to numerical requirements, shows
that the chosen filter size is adequate [197, 198].
Upstream of the combustion chamber the backward facing step is included in
the computational domain using the immersed boundary technique [170]. Hereby,
the walls of the step and the splitter plate are expressed by an artificial force added
to the momentum equation. This force has been added to the velocities after the
pressure correction at every Runge-Kutta time step. The splitter plate between the
two separated channels, in the upstream direction, induces a periodic instability
of Helmholtz type, which can not be captured if the computational domain starts
at the backward facing step. This instability is expected to influence the behavior
of the premixed flame and thus this part should be included in the computational
domain.
1

Composition of the mixture is 0.8% C2 H6 , 86.4% C3 H8 , 10.55% C3 H6 , 2.25 % C4 H10

150

CHAPTER 8. RESULTS PREMIXED COMBUSTION MODELING

Figure 8.2: Instantaneous streamwise velocity field in the burner upstream of the
sudden expansion)

The total extension of the computational domain in axial (x), transverse (z) and
the periodic (y) direction is 23.7D 4.35D 2.34D where D represents the height of
the backward facing step. It has to be mentioned that in stream-wise direction 5D
of the domain are located upstream of the sudden expansion. The computational
domain is resolved with 712 130 64 6 106 grid points. Setting the velocities
to zero at the upper and lower wall of the combustion chamber and using Neumann
boundary conditions for the pressure mimics the walls of the combustion chamber.
At the outflow Neumann boundary conditions for the velocity and the pressure are
prescribed, negative velocities are clipped. The inlet boundary conditions for the
two channels are fully developed channel mean velocity profiles with superimposed
fluctuations, generated by a method presented by Klein et al [162]. The method uses
digital filtering of random data to achieve prescribed two point correlations together
with components of the Reynolds stress tensor. Experimental data [111] has been
used to generate the inflow data 5D upstream of the step. These measurements
contain axial and transverse fluctuations and cross-correlations. The length scales
at the inflow for both channels have been set to 0.4 times the width of the supplying
channels in streamwise direction and 0.125 times in the two other directions. Since
the computational domain starts more than 150 mm before the area of interest, it
is expected that the inflow conditions can be considered as reasonable. Figure 8.2
gives an idea of the turbulent flow field upstream of the sudden expansion. One
can observe the turbulent structures from the artificially generated turbulence and
one can see that they still develop in the supplying channels so that this upstream
part is needed to obtain the correct velocity field at the sudden expansion where the
domain of interest starts. The periodic instability introduced by the splitter plate
can also be observed.
All simulations run 60,000 time steps which is equivalent to approximately five flow
through times, based on the bulk velocity, so the number of collected samples can
been considered as sufficient. The computations run parallel for approximately 80h
on four CPUs of a Transtec 8000L Itanium2 1.5 Ghz cluster.

12.0
8.0
4.0
0.0
-4.0

x = 8D

12.0
8.0
4.0
0.0
-4.0

x = 7D

12.0
8.0
4.0
0.0
-4.0

x = 4D

12.0
8.0
4.0
0.0
-4.0

x = 3D

12.0
8.0
4.0
0.0
-4.0

x = 2D

12.0
8.0
4.0
0.0
-4.0

x = 1D

151

4.0

x = 8D

2.0
0.0
4.0

x = 7D

2.0
0.0
4.0

uax' [m/s]

uax [m/s]

8.3. RESULTS AND DISCUSSION

x = 4D

2.0
0.0
4.0

x = 3D

2.0
0.0
4.0

x = 2D

2.0
0.0

Exp
LES

4.0

Exp
LES

x = 1D

2.0
0.0
-50

-25

25

50

z [mm]

-50

-25

25

50

z [mm]

Figure 8.3: Transverse profiles of time averaged stream-wise velocities and fluctuations at different axial positions (non-reacting case): comparison of LES results
(solid) and experimental data (points)

8.3

Results and Discussion

In the following sections the results of the investigation of the ORACLES configuration are presented. Initially results of the isothermal flow case are presented
and discussed. Furthermore three simulations of the reacting flow field have been
conducted with different approximations for the SGS-velocity, which has been used
to model the efficiency function. Results are compared to experimental data and
discussed.

8.3.1

Non-Reacting Flow

Initially the non-reacting flow is simulated to ensure the capability of the chosen grid
resolution. The agreement between the results of the simulation and the experimental data can be considered to be very good in terms of the mean streamwise velocity
component. The recirculation zone attached to the lower wall of the channel is predicted very well, however at the upper wall the numerical simulation over-predicts
the length of recirculating fluid slightly. The corresponding fluctuations yield, although not perfect an encouraging agreement, as can be seen in Fig. 8.3. Interesting
to note, both experimental and former numerical investigations [187, 188, 189, 190]
have shown a non-symmetric behavior of the isothermal flow field, which is also identified in the current simulation. Abbot and Kline investigated a large series of single
and double backward facing steps, where several influencing variables were investigated independently [199]. They found that an asymmetry of the flow pattern can
be observed for expansion ratios bigger then 1.5 which is evidently the case in the

152

CHAPTER 8. RESULTS PREMIXED COMBUSTION MODELING

current configuration. Another conclusion of their study was that the re-attachment
length is less influenced by the level of fluctuations or the mean upstream velocity
profiles, but strongly depends on the expansion ratio of the channel.
Two shear layers 35 mm away from the centerline of the combustion chamber
are clearly visible in the profiles of the velocity fluctuations, together with a mixing
layer of lower intensity level at the centerline. Starting the computational domain at
the backward facing step, the simulation failed to predict the level of the turbulent
fluctuations adequately.

8.3.2

Reacting Flow

To resolve the density jump across the flame front, the product of the thickening
factor and the laminar flame thickness must become resolvable with sufficient grid
points. In the present case the thickened flame front, L1 = F L0 = 200.2mm 4mm,
is resolved using roughly about five grid points. To guarantee the ability of the thickened flame to interact with the integral turbulent scales, the thickened flame must be
considerably smaller than the integral length scales. Considering the integral scale of
the streamwise velocity component to be about 0.4 times the width of the supplying
channels, one ends up with a dominant length scale of approximately 12mm. Hence,
the ability of large scales, to interact with the thickened flame front is conserved.
Therefore the chosen thickening factor can be addressed as a reasonable choice.
The subgrid scale velocity u0e needed to model the subgrid scale wrinkling, was
calculated with two different approaches, expressed by (5.96) and (5.97). These two
simulations, together with an additional third calculation, where the efficiency function E was set to a constant value of 1.5, which is approximately the average value
of E in the entire un-burnt region of the domain, are compared to experimental
findings [111]. To simplify the discussion of the results the following abbreviations
will be used: using the model presented by Colin et al [99] for u0e is named Ecol ,
using the filtering technique for u0e is named Ef il and using a constant value for
the mean efficiency function is denoted Econ .
The flame structure is strongly three dimensional for all considered models as
can be seen in figure 8.6 showing an instantaneous flame front for the Ecol case. The
flame has a V-shape and is attached to the edge of the sudden expansion. Large
scale wrinkles in streamwise direction can be observed corresponding to the vortex
separation from the backward facing steps of the sudden expansion. Also in the
spanwise direction a strong wrinkling of the flame front by turbulent vortices is
observed, showing that it is necessary to use a big enough computational domain in
this direction to capture the turbulent structures and flame structure adequately.
Under reacting conditions the average flow field becomes much more symmetric,
which can be seen in the comparison between the experiments and the numerical
simulations, plotted in Figs. 8.4 and 8.5. Also the time averaged flame front position is found to be symmetric with respect to the center of the channel, in all three

8.3. RESULTS AND DISCUSSION

153

2.0
20.0

0.0

Exp
Colin
E= c
Filt

10.0
x = 8D
0.0
-5.0

-1.0
-2.0
2.0

20.0

0.0
x = 7D

-1.0

0.0
-5.0

-2.0
2.0

x = 4D

utrans [m/s]

uax [m/s]

x = 4D

1.0

10.0
0.0
-5.0
20.0

x = 7D

1.0

10.0

20.0

x = 8D

1.0

x = 3D

0.0
-1.0
-2.0
2.0

x = 3D

1.0
0.0

10.0

-1.0
0.0
-5.0
20.0

-2.0
2.0

x = 2D

x = 2D

1.0
0.0

10.0

-1.0
0.0
-5.0
20.0

-2.0
2.0

x = 1D

x = 1D

1.0
0.0

10.0

-1.0
0.0
-5.0

-2.0
-50

-25

z [mm]

25

50

-50

-25

25

50

z [mm]

Figure 8.4: Transverse profiles of time averaged mean stream-wise and transverse
velocities, using a constant efficiency function E = 1.5 (dotted), using Colins Model
for u0e (dashed-dotted) and using the filtering technique for u0e (solid) compared
to experimental data (points)

154

CHAPTER 8. RESULTS PREMIXED COMBUSTION MODELING

6.0

4.0

x = 8D

x = 8D

4.0
2.0
2.0
0.0
6.0

0.0
4.0

x = 7D

x = 7D

4.0
2.0
2.0
0.0
6.0

0.0
4.0

x = 4D

x = 4D

utrans' [m/s]

uax' [m/s]

4.0
2.0
0.0
6.0

x = 3D

2.0

0.0
4.0

x = 3D

4.0
2.0
2.0
0.0
6.0

0.0
4.0

x = 2D

x = 2D

4.0
2.0
2.0
0.0
6.0

0.0
4.0

x = 1D

x = 1D

Exp
Colin
E= c
Filt

4.0
2.0
2.0
0.0

0.0
-50

-25

z [mm]

25

50

-50

-25

25

50

z [mm]

Figure 8.5: Transverse profiles of the fluctuations of the stream-wise and the transverse velocities, using a constant efficiency function E = 1.5 (dotted), using Colins
Model for u0e (dashed-dotted) and using the filtering technique for u0e (solid) compared to experimental data (points)

Figure 8.6: Instantaneous flame front (Ecol case) and zoom of the flame front near
the sudden expansion

8.3. RESULTS AND DISCUSSION

155

Figure 8.7: Mean flame front position and color contours of the rms values of the
progress variable (maximum level is 0.4). A constant efficiency function E = 1.5 is
used in the upper plot, Colins Model for u0e in the middle plot and the filtering
technique for u0e in the lower plot
simulations, Ecol , Econ and Ef il . This can be seen in figure 8.7 which compares the
average flame surface ( = | = max ) for the investigated cases and superimposed
the color contour of the rms of (which measures the fluctuations of and thus gives
an idea of the turbulent flame brush). In the central region of the channel, where the
fluid can be assumed to be generally un-burnt, the numerical and experimental data
of the mean velocities (Fig. 8.4) are found to be in very good agreement, but closer
to the wall, where the fluid is sometimes burnt and sometimes un-burnt some discrepancies between experimental results and the simulations can be identified. All
three simulations fail to predict the recirculation zone, at x = 1D, independently of
the efficiency function modeling. Further downstream the channel, the Econ simulation over-predicts the streamwise velocity component, in the same amount as the
other two, Ef il and Ecol , under-predict the streamwise velocity of the experimental
investigation.
For the transverse velocity component the difference between the simulations is
more evident. The large discrepancy at x = 1D, z = 50mm can be explained
with the inability to capture the recirculation zone properly. Taking into account
that adiabatic boundary conditions have been used for the walls of the channel, the
flame will always be attached to the edge of the sudden expansion. The experimental data-set gives no further insight if the flame is attached to the wall, or burns
slightly lifted. This might explain the disagreement between experimental and numerical results. Comparing the runs with different approximations for the efficiency
functions one can clearly see that Ef il and Ecol behave similarly well, while Econ
shows a strong deviation, especially visible at z = 50mm in the profiles of the
transverse velocity distribution.

156

CHAPTER 8. RESULTS PREMIXED COMBUSTION MODELING

The time averaged fluctuations of the stream-wise velocity component, plotted


in Fig. 8.5 seem to be almost identical for all three calculations. Comparing the
results to the experimental findings, yields again relatively good agreement in the
central region of the channel, but especially at x = 2D and at x = 8D the absolute
level of fluctuations is not predicted properly in the close to wall regions. Beside the
fact that the resolution is insufficient to capture the viscous sub-layer of the channel,
another possible explanation for the under-prediction is given in the following. It
is reported in the literature that under reacting conditions an oscillatory motion
can be observed in the combustion chamber. The impact of this oscillations can be
identified up to 5D upstream of the splitter plate. The corresponding acoustic frequency, about 50Hz might result in a temporarily uneven mass flux, approaching the
combustion chamber. Clearly, this deviates from the constant incoming mass flux,
used in the current simulation. Unfortunately no detailed information about the
amplitude of this oscillation can be found in the literature. In the work of Domingo
et al the effort was made to include this frequency in the simulation, by adding a
50Hz oscillation to the incoming mass flux [109].
For the fluctuations of the transverse velocity component (Fig. 8.5) an overall
encouraging agreement is found at all stream-wise positions presented. But again,
all simulations fail to predict the shear layers at z = 50 (x = 1D), which might be
contributed to the fact that the recirculation zone was not captured precisely.
Comparing the three calculations, it can be concluded that the inability of the
filtering method to correct for the dilatational part of the SGS-fluctuations is less
important in high Reynolds number flows, but remains significant for moderate and
low Reynolds number flows. Removing the local information of the SGS-velocities
and taking the efficiency function as constants leads to results, which are less precise,
but also end up in a similar level of prediction quality. Although one expects the
SGS-fluctuations to be more or less uniform in a channel, apart from the wall, this
seems not to be the case under reacting conditions in the investigated configuration.

8.3.3

Flame Front Wrinkling

In Fig. 8.8 instantaneous snapshots of the flames together with the approximated
efficiency function mapped on a plane are presented. The top figure (Econ ), with
a constant efficiency function of 1.5 in the entire domain, shows a shorter flame
compared to the other two figures (Ecol , Ef il ) and additionally the flame is located
further from the bottom and top wall of the combustion chamber. Introducing the
efficiency function locally increases the flame speed from s0L to sT = Es0L which
shortens the flame considerably. The flame fronts are clearly attached to the corners of the sudden expansion due to the assumption of adiabatic walls, and show
especially in the Ecol and Ef il simulations large scale wrinkles caused by the separation of the flow from the steps.
Assuming a large wrinkling at the subgrid scale level, the maximum value for the

8.3. RESULTS AND DISCUSSION

157

Figure 8.8: Instantaneous snapshot of the flame front (black) with the level of the
efficiency function, mapped on a plane (for comparison step height=29.9 mm). Light
colors mean small fluctuation level and dark color high fluctuations. A constant
efficiency function E = 1.5 is used in the upper plot, Colins Model for u0e in the
middle plot and the filtering technique for u0e in the lower plot

158

CHAPTER 8. RESULTS PREMIXED COMBUSTION MODELING

efficiency function can be estimated to be E max = F 2/3 7 with the used thickening
factor. Locally the turbulent flame speed can thus become significantly higher than
the laminar flame speed in zones of high subgrid scale velocities u0e . The mean
efficiency function along the flame front is a lot lower than this maximum. From
Fig. 8.8 it can be seen that E is locally high in the shear layers and in the central
zone of the combustion chamber due to the splitter plate.
Comparing the snapshots between the two different approximated efficiency functions the fluctuations seem to look considerably different, but the impact on the
flame front seems to be rather weak, which might come from the relatively high
Reynolds number of the investigated problem. The statistical behavior of both approximations (Fig. 8.4 and Fig. 8.5) with respect to the mean velocity components
is almost identical. Downstream the channel the mean streamwise velocities, using
Colins model for u0e , are slightly higher than with the filtering approach. Although
locally the efficiency function in the Ecol and Ef il simulations is a lot larger than the
constant value of 1.5 used in the Econ simulation the flames are considerably longer
than in the latter case.

8.4

Conclusions

A code for large eddy simulation of turbulent premixed combustion based on the
concept of the artificially thickened flame has been developed and tested on the
ORACLES burner. The test case has been investigated in first instance on the inert
case. The result demonstrate good agreement with the experiment [111]. To evaluate
several ways of modeling the subgrid scale velocity u0e for the calculation of the
efficiency function, a systematic comparison in three dimensional LES calculations
has been performed. The results show that these approaches have a relatively small
impact on the behavior of the flame front. In contrast to DNS investigations done
at lower Reynolds numbers it can be concluded that for relatively high Reynolds
numbers, the absolute level of the fluctuation is the dominant factor towards flame
wrinkling and flame length.

Chapter 9
Conclusions
In this work the numerical simulation of turbulent non-premixed and premixed combustion has been studied. For the simulation of non-premixed combustion a complete flow solver has been developed able to perform RANS and LES simulations of
combustor configurations with different models. The solver is based on a low-Mach
number approximation and uses cartesian, non-orthogonal and non-uniform grids.
For LES the code was parallelized using MPI (Message Passing Interface) to reduce
the computational time. Combustion is modeled with a conserved scalar approach
and a presumed pdf is used to describe the effect of turbulence. Preprocessing tools
have been developed to generate combustion tables for infinitely fast chemistry (irreversible or reversible) and finite rate chemistry, which are accessed by the flow
solver with two or three combustion parameters. The code has been validated on
two well documented target flames from the TNF-workshop: Sandia D and DLR A.
First less computationally expensive RANS simulations were performed, to test the
code and the implementation of the combustion models. The results of the RANS
simulations are in good agreement with the experimental data for the Sandia D case.
In the DLR A case however, the simulation results deviate strongly from the experiment far from the nozzle where a wrong prediction of the mixture fraction field
causes a strong deviation in the temperature field. At the lean side of this flame the
temperature is very sensitive to the mixture fraction since the stoichiometric mixture fraction is low. The importance of some parameters in the turbulence model
have been investigated and the influence of the inlet boundary conditions has been
studied. The mixed is burnt and flamelet combustion model have been used for
the Sandia D flame which has a single fuel (methane) but diluted with air. For the
DLR flame the equilibrium model was used since the fuel is a mixture of methane
and hydrogen but the fuel stream contains no air. The influence of the combustion
model on the flow field is not very strong since no important finite rate chemistry
effects are present in the investigated cases. Then the same test cases were investigated using the LES approach which is known to provide an adequate description
of the turbulent flow field and the turbulence-combustion interaction. Compared
to RANS, the large structures are explicitly computed, while only the small scales
are modeled using a subgrid scale model. The obtained simulation results are in
reasonably good agreement with the experimental data but some of the investigated
models still need some improvement. The computational cost of an LES is much
159

160

CHAPTER 9. CONCLUSIONS

higher than for a RANS simulation and thus it is more difficult to investigate the
effect of many parameters. Moreover jet-flames are not really worthy targets for
LES, since most RANS models have been developed and scaled for this kind of
flows. Hence RANS models perform well for this kind of flows at a much lower cost
than LES. The influence of the scalar convection scheme, the inlet boundary condition, the model constant in the variance model and the combustion model has been
investigated. Both the scheme for convective scalar fluxes and the variance model
have a significant effect on the flow field. A dynamic model for the coefficient in the
variance model should be investigated so that this parameter does not have to be
tuned for every test case and a more general model is obtained. The velocity field at
the inlet has a strong influence on the development of the jet and thus on the flame
structure. Therefore a velocity field with correct turbulent intensities and length
scales is desired to obtain results in good agreement with the experimental findings.
A method to artificially generate turbulent flow fields has been applied but needs
some further improvement. As for the RANS simulation the effect of the combustion model is small but an underestimation of the scalar dissipation rate results in a
higher temperature at the rich side of the flame. This does not seem to correspond
to the experimental findings since for the flamelet model the temperature field is
in worse agreement with the experiment than the more simple mixed is burnt model.
To investigate turbulent premixed combustion, the Artificially Thickened Flame
model has been implemented in the research code FLOWSI of the Department of
Energy and Powerplant technology of the TU Darmstadt. A Large Eddy Simulation together with this approach is used to study a well known experimental set-up
consisting of a rectangular dump combustor - ORACLES. The major drawback of
artificially thickening the flame is that the interaction between turbulence and flame
is altered. To compensate for the inability of small vortices to wrinkle the flame a
subgrid scale wrinkling model has to be introduced. In this work the influence of
the subgrid scale wrinkling on the flame front in a high Reynolds number flow has
been investigated. Moreover the influence of different approximations for the subgrid scale velocity on the prediction of the flow field and flame structure has been
studied. Initially the non-reacting flow has been simulated to ensure the capability
of the chosen grid resolution. The agreement between the results of the simulation
and the experimental data can be considered to be very good in terms of the mean
streamwise velocity component and the corresponding fluctuations yield, although
not perfect, an encouraging agreement. Both experimental and former numerical investigations have shown a non-symmetric behavior of the isothermal flow field which
is also identified in the current simulation. Under reacting conditions the average
flow field becomes much more symmetric. In the central region of the channel, where
the fluid can be assumed to be generally un-burnt, the numerical and experimental
data of the mean velocities are found to be in very good agreement, but closer to the
wall, where the fluid is sometimes burnt and sometimes un-burnt some discrepancies
between experimental results and the simulations can be identified. The time averaged fluctuations of the stream-wise velocity component seem to be almost identical
for all three calculations. Comparing the results to the experimental findings, yields
again relatively good agreement in the central region of the channel, but especially

161
the absolute level of fluctuations is not predicted properly in the close to wall regions. This is probably caused by an insufficient mesh resolution near the walls of
the combustion chamber and the use of adiabatic wall boundary conditions. Moreover, it is reported in the literature that under reacting conditions an oscillatory
motion can be observed in the combustion chamber and the corresponding acoustic
frequency, about 50Hz, might result in a temporarily uneven mass flux, approaching
the combustion chamber. This effect can not be predicted with the current simulation strategy and thus this phenomenon is not captured. It would therefore be
interesting to investigate the same configuration under slightly different conditions.
The above mentioned oscillation does not seem to occur if different equivalence ratios are used for both supply channels and thus this effect could be eliminated.
Unfortunately the model used with a single progress variable and 1 step chemistry
can not account for the variable equivalence ratio. Therefore the extension to a
multispecies solver based on the ATF approach has been made but no simulation
results are yet available. Comparing the three calculations, it can be concluded
that the inability of the filtering method to correct for the dilatational part of the
SGS-fluctuations is less important in high Reynolds number flows. Removing the local information of the SGS-velocities and taking the efficiency function as constants
leads to results, which are less precise, but also end up in a similar level of prediction
quality. It would also be interesting to investigate other premixed burners, such as
the triangular flame holder, where apart from velocity fields also scalar field measurements (such as the progress variable) are available. This would yield a better
understanding of the influence of e.g. the subgrid scale wrinkling model of the flame.

162

CHAPTER 9. CONCLUSIONS

Appendix A
Derivation of balance equation for
mixture fraction variance
The exact transport equation for the variance of mixture fraction in a RANS context can be derived in the following way. Starting from the instantaneous balance
equation for mixture fraction:
z ui z

z
D
=
+
t
xi
xi
xi

(A.1)

and multiplying each term with 2z and using the following relations

xi

z 2
z
2
2z
=
2z t + 2z t
+ z2
(A.2)
=
t
t
t
ui
ui z 2
ui z
2 ui
iz
= 2z u
+ z2
(A.3)
+
2z
=
2z
xi
xi
xi
xi
xi
!
!


z 2
z
z z

z
(A.4)
D
=
D
+ 2D
2Dz xi
= 2z
xi
xi
xi
xi
xi xi

one obtains a transport equation for z 2


ui
z 2 ui z 2
+
+ z2
+
t
xi
t
xi
|

{z

=0(continuity)

z 2
=
D
xi
xi

2D

z z
xi xi

(A.5)

A similar equation can be obtained for ze2 by multiplying the balance equation for
the mean mixture fraction ze by 2ze
ze2 ue i ze2

z

+
= 2ze
ug
D
iz
t
xi
xi
xi

(A.6)

Averaging equation (A.5) and subtracting this equation from (A.6) and using the
f = 0)
following relations and properties of the averaging procedure (
ui z 2 =

uei + ui (ze2 + zez 2 + z 2 )


163

164

APPENDIX A. BALANCE EQ. FOR MIXTURE FRACTION VARIANCE


h

g
e
2
2 + ug
uei ze2 + uei zg
i z z + ui z

e

2 + ug
uei ze2 + ue i zg
i z z + ui z

z 2
2

2
e
z
+ 2Dz x
+ 2D ze z
= D xez i + D z
xi
xi
i
xi
z z
z
e
z z
z z
e
z e
+
4D
+
2D
= 2D x
2D
x
x
x
xi xi
i
i
i
i
xi xi
D

(A.7)

f2 ze2
2 = z
one obtains a balance equation for the variance of mixture fraction zg
2
2
ue i zg
zg
+
=
t
xi

 2 
ui z
xi

{z

z 2
+
D
xi
xi
|

Turbulent transport

2z

{z

ze

D
xi
xi

Molecular diffusion
ze
z z
2ui z
2D
xi
xi xi
|

{z

Production

{z

Dissipation

!
}

(A.8)

Appendix B
Flame structure for 1D laminar
strained flame
For a one dimensional steady strained diffusion flame with infinitely fast chemistry
and constant density an analytical solution for the mixture fraction and scalar dissipation rate field can be derived. Figure B.1 shows schematically the structure of
such a flame. Assuming constant density the imposed velocity field corresponds to

Figure B.1: Steady strained counterflow diffusion flame

u1 = ax1

and u2 = ax2

(B.1)

which satisfies the continuity equation. The stagnation plane is located at x1 = 0 and
a is the strain rate (s1 ). In a real burner such a stagnation point flow configuration
is achieved by placing fuel and oxidizer jets face to face. The strain rate in such a
configuration is not known a priori and the strain is not constant along the flame
normal. However the magnitude of the flame strain can be estimated from the
velocities of the jets UF and UO and the distance between the two jets h: a
165

166

APPENDIX B. FLAME STRUCTURE 1D FLAME

|UF | + |UO |
. The z balance equation can be written as
h
!

z
z

D
+
(u1 z) +
(u2 z) =
t
x1
x2
x1
x1

z
+
D
x2
x2

(B.2)

(B.3)

or using the continuity equation in a non-conservative form


z
z
z

z
D
+ u2
=
+ u1
t
x1
x2
x1
x1

z
+
D
x2
x2

On the axis normal to the steady strained flame assuming constant density and
diffusion coefficient the z equation reduces to
ax1

z
2z
=D 2
x1
x1

(B.4)

Performing the substitution = x1 a/2D this equation can be rewritten as


2z
z
+ 2
=0
2

(B.5)

with boundary conditions z(+) = 0 and z() = 1. The solution is given by


z=

1
[1 erf()]
2

with = x1

a
2D

(B.6)

Using the definition of the error function erf


2
erf() =

exp t2 dt

(B.7)

one can write


z
=
21 x 1 erf()
x1

= 12 2 exp ( 2 ) x
1
=

a
2D

x2 a

i
exp 2D

(B.8)

The scalar dissipation rate is then given by


z
= 2D
x1

!2

a
a
exp x21

(B.9)

a
depending on the location x1 and with a maximum in the stagnation plane. The

scalar dissipation can be written in the mixture fraction space as


=

h
i2
h
i2
a
exp 2 erf1 (1 2z)
= 0 exp 2 erf1 (1 2z)
= 0 F (z) (B.10)

Appendix C
Spatial discretization in a
collocated grid system and the
odd-even decoupling phenomenon
Discretizing the computational domain, one has to choose where the different variables - scalars like pressure, density or mixture fraction and velocity vector components - are being stored. The obvious choice is to store all the variables at the
same set of grid points and to use the same control volumes for all variables; such
a grid is called collocated and the programming is simplified by this choice. Another possibility is to use a staggered arrangement introduced by Harlow and Welsh
(1965) offering several advantages over the collocated arrangement. Both arrangement are shown in figure C.1. Several terms that require interpolation with the
collocated arrangement, can be calculated (to a second-order approximation) without interpolation using a staggered arrangement. Perhaps the biggest advantage
of the staggered arrangement is the strong coupling between the velocities and the
pressure which helps to avoid some types of convergence problems and oscillations
in pressure and velocity fields. The advantages of collocated grids over staggered
ones are computational simplicity and straightforward extension of the method to
curvilinear coordinate systems. Furthermore, when multigrid procedures are used,
the same restriction and prolongation operators for transfer of information between
the various grids can be used for all variables. On the other hand, particular attention should be paid to the fact that the pressure correction method on collocated
grids can produce spurious oscillations of the pressure field. This problem is known
in the literature as odd-even decoupling. In this appendix it will be clarified how
this phenomenon is avoided for the collocated cell centered arrangement used in the
current work.
To fix thoughts consider again the set of equations to be solved in a 1D formulation
!n+1
(u)n+1

+
=0
(C.1)
t
x
(u)n+1 (u)
p
=
t
x
167

(C.2)

168

APPENDIX C. COLLOCATED GRID AND ODD-EVEN DECOUPLING

,
Figure C.1: Collocated (left) and staggered (right) variable arrangement on a 2D
cartesian grid. Arrows denote velocity components and points the location of scalar
variables like pressure.
Discretizing the continuity equation (C.1) using a second order central scheme in
point i one obtains
!n+1

n+1
(u)n+1
i+1 (u)i1
=0
+
2x

(C.3)

and the momentum equation written in point i with a second order central discretization for the pressure gradient is given by
(u)i
(u)n+1
pi+1 pi1
i
=
t
2x

(C.4)

Writing equation (C.4) in points i + 1 and i 1

(u)n+1
pi+2 pi
i+1 (u)i+1
=
t
2x

(C.5)

pi pi2
(u)n+1
i1 (u)i1
=
(C.6)
t
2x
and subtracting equation (C.6) from (C.5) and dividing by 2x one obtains the
poisson equation for pressure
n+1
(u)i+1 (u)i1
(u)n+1
pi+2 2pi + pi2
i+1 (u)i1

= t
2x
2x
4x2

(C.7)

which is by substitution of equation (C.3) equivalent to (6.52)


1
p=
t
2

!n+1

1
(u)
t

(C.8)

with in the left hand side the laplace operator and in the right hand side the time
derivative of density and the divergence of the intermediate momentum field. One
can see that the discretization of the laplacian of pressure in (C.7) is written on an

169
extended stencil and involves nodes which are 2x apart! It is a discretized Poisson
equation on a grid twice as coarse as the basic one and the equations are split into
four unconnected systems, one with i and j both even, one with i even and j odd,
one with i odd and j even, and one with both odd. Each of these systems gives a
different solution. For a flow the checkerboard pressure distribution shown in figure
C.2 satisfies these equations and could be produced. However, the pressure gradient
is not affected and the velocity field may be smooth. There is also the possibility that
one may not be able to obtain a converged steady-state solution. To avoid the odd-

Figure C.2: Checkerboard pressure field, made of four superimposed uniform fields
on 2-spacing, which is interpreted by CDS as a uniform field
even decoupling problem one can discretize the poisson equation on a compact stencil
() which can be obtained by using a flux interpolation method [200, 148]. Using
the momentum equation in semi-discretized form (C.2) and substituting (u)n+1
in
i
the continuity equation (C.1) one obtains

!n+1

"

p
+
(u) t
=0
x
x
|

{z

(C.9)

where F is an auxiliary momentum flux which is staggered with respect to other


variables in space. This implies that each auxiliary flux Fj is defined on the center
of the cell interface that is normal to the xj direction, whereas all flow quantities
are evaluated at the centers of the computational cells. The auxiliary fluxes are
calculated via interpolation and in a finite volume formulation one then obtains

with

Fi+1/2
Fi1/2
n+1
+
=0

t
x
i

Fi+1/2
=

(u)i+1 (u)i
pi+1 pi
t
2
x

(C.10)

(C.11)

170

APPENDIX C. COLLOCATED GRID AND ODD-EVEN DECOUPLING

Fi1/2
=

(u)i (u)i1
pi pi1
t
2
x

(C.12)

and thus the equivalent poisson equation is given by


(u)i+1 (u)i1
pi+1 2pi + pi1
n+1
t
+

=


x2
t
2x
i

(C.13)

where compared to equation (C.7) the laplacian of the pressure field has is written
on a compact stencil and thus the odd-even decoupling is avoided.

Appendix D
List of Publications
Dispersion and dissipation properties of the 1D spectral volume method and
application to a p-multigrid algorithm VAN DEN ABEELE Kris, BROECKHOVEN Tim, LACOR Chris, Journal of Computational Physics, accepted
Generation of Turbulent Inflow and Initial Conditions based on Multi-Correlated
Random Fields FATHALI Mani, KLEIN Markus, BROECKHOVEN Tim,
LACOR Chris, BAELMANS Martine, Journal of Computational Physics, in
review
Optimization of time integration schemes coupled to spatial discretization for
use in CAA applications, RAMBOER Jan, BROECKHOVEN Tim, SMIRNOV
Sergey, LACOR Christian, Journal of Computational Physics, Volume 213 ,
Issue 2, p.777 - 802, 2006
Large eddy simulation of particle-laden flows using a high-order cell-averaged
finite volume method, BROECKHOVEN Tim, LESSANI Bamdad, SMIRNOV
Sergey, LACOR Christian, Computers and Fluids, second review
Finite volume formulation of compact upwind and central schemes with artificial selective damping, BROECKHOVEN Tim, RAMBOER Jan, SMIRNOV
Sergey, LACOR Christian, SIAM Journal on Scientific Computing, 21(3), pp.
341-367, 2004
Investigation of Subgrid Scale Wrinkling Models and their Impact on the
Artificially Thickened Flame Model in Large Eddy Simulations, paper presentation BROECKHOVEN Tim, FREITAG Martin, LACOR Chris, SADIKI
Amsini, JANICKA Johannes, Complex Effects in Large Eddy Simulations,
Lecture Notes in Computational Science and Engineering Vol. 56 Kassinos,
S.; Langer, C.; Iaccarino, G.; Moin, P. (Eds.) 2007, ISBN-10: 3-540-34233-8,
ISBN-13: 978-3-540-34233-5, Available: January 3, 2007 p365-382
Numerical Methods, BROECKHOVEN Tim, RAMBOER Jan, SMIRNOV
Sergey, LACOR Christian, in Large Eddy Simulation For Acoustics,WAGNER
Claus, HTTL Thomas, SAGAUT Pierre, Cambridge University Press, 2005
171

172

APPENDIX D. LIST OF PUBLICATIONS

PIV Measurements of a Double Annular Jet for validation of numerical Simulations BROECKHOVEN Tim, BROUNS Mark, VANHERZEELE Joris,
VANLANDUIT Steve, LACOR Chris, 13th Int Symp on Applications of Laser
Techniques to Fluid Mechanics Lisbon, Portugal, 26-29 June, 2006
Investigation of Subgrid Scale Wrinkling Models and their Impact on the
Artificially Thickened Flame Model in Large Eddy Simulations, paper presentation, BROECKHOVEN Tim, FREITAG Martin, LACOR Chris, SADIKI
Amsini, JANICKA Johannes ,The Cyprus International Symposium on Complex Effects in Large Eddy Simulations (CY-LES 2005) , 23-26 sept 2005,
Cyprus
Large Eddy Simulation of Piloted and Bluff-Body Diffusion Flame, BROECKHOVEN Tim, SMIRNOV Sergey, LACOR Christian, BRIZUELA Eduardo,
Proceedings of the 21st International Congress of Theoretical and Applied
Mechanics (ICTAM04), 15-21 August Warsaw, Poland, 2004
Finite Volume Formulation of Compact Upwind and Central Schemes with Artificial Selective Damping, SMIRNOV Sergey, BROECKHOVEN Tim, RAMBOER Jan, LACOR Christian Second MIT Conference on Computational
Fluid and Solid Mechanics, June 17-20, 2003, Massachusetts Institute of Technology, Boston, USA
Parallel Computations using a Finite Volume Compact Scheme Formulation,
BROECKHOVEN Tim, LACOR Christian, SMIRNOV Sergey, 6th National
Congress on Theoretical and Applied Mechanics, Ghent, May 26-27, 2003
Large Eddy Simulations of a Premixed Combustor with the Artificially Thickened Flame Model, BROECKHOVEN Tim, FREITAG Martin, LACOR Chris,
SADIKI Amsini, JANICKA Johannes, Ercoftac Annual Seminar Belgian Pilot
Center, CENAERO, December 12, 2005
Large Eddy Simulation of Piloted and Bluff-Body Diffusion Flame, BROECKHOVEN Tim, SMIRNOV Sergey, LACOR Christian, BRIZUELA Eduardo,
ECCOMAS European Congress on Computational Methods in Applied Sciences and Engineering, 24-28 July, Jyvaskyla, Finland, 2004

Bibliography
[1] The International Energy Agency.
http://www.iea.org/, 2006.

Key World Energy Statistics.

[2] The International Energy Agency.


World Energy Outlook 2004.
http://www.worldenergyoutlook.org/, 2006.
[3] TNF. International workshop on measurement and computation of turbulent
nonpremixed flames. http://www.ca.sandia.gov/TNF.
[4] K.K. Kuo. Principles of Combustion. John Wiley New York, 1986.
[5] A. Ern and V. Giovangigli. Multicomponent Transport Algorithms. Springer
Verlag Heidelberg, 1994.
[6] Ideal gas thermodynamic data in polynomial form for combustion and air pollution use. http://garfield.chem.elte.hu/Burcat/burcat.html. Prof. Alexander
Burcat, Faculty of Aerospace Eng. ,Technion - Israel Inst. of Technology.
[7] Ideal gas thermodynamic data for combustion and air- pollution use. Technion
Report TAE 697, 1993. A. Burcat and B. McBride.
[8] Coefficients for calculating thermodynamic and transport properties of individual species. NASA Report TM-4513, 1993. B.J. McBride, S. Gordon and
M.A. Reno.
[9] Pierre Sagaut. Large-Eddy Simulation for Incompressible Flows . Scientific
Computation. Springer, 2002.
[10] Edited by C. Wagner, T.H
uttl, and P.Sagaut. Large-Eddy Simulation For
Acoustics. Cambridge University Press, 2006.
[11] J. Smagorinsky. General circulation experiments with the primitive equations.
Mon. Weather Rev., 91:99164, 1963.
[12] J. Bardina, J.H. Ferziger, and W.C. Reynolds. Improved subgrid models for
large eddy simulation. AIAA paper 80-1357, 1980.
[13] M. Germano, U. Piomelli, P. Moin, and W. H. Cabot. A dynamic subgrid-scale
eddy viscosity model. Physics of Fluids A, 3(7), jul 1991.
173

174

BIBLIOGRAPHY

[14] D. K. Lilly. A proposed modification of the Germano subgrid-scale closure


method. Physics of Fluids, A(4):633635, 1992.
[15] C. G. Speziale. Turbulence modeling for time-dependant rans and vles: A
review. AIAA Journal, 36:173184, 1998.
[16] V. Yakhot and S. Orszag. Renormalization group (rng) methods for turbulence
closure. Journal of Scientific Computing, 1:352, 1986.
[17] D. Carati, G. S. Winckelmans, and H. Jeanmart. On the modelling of the
subgrid-scale and filtered-scale stress tensors in large-eddy simulation. J. Fluid
Mech., 2001.
[18] S.Menon, P.K. Yeung, and W.W. Kim. Effect of subgrid models on the computed interscale energy transfer in isotropic turbulence. Computers and Fluids,
25(2):165180, 1996.
[19] V.K. Chakravarthy and S. Menon. Subgrid modeling of turbulent premixed
flames in the flamelet regime. Flow, Turbulence and Combustion, 65:133161,
2000.
[20] E. van Driest. On turbulent flow near a wall. Journal of aeronautical science,
23:10071011, 1956.
[21] P. Moin, K. Squires, W. Cabot, and S. Lee. A Dynamic Subgrid-Scale model
for Compressible Turbulence and Scalar Transport. Physics of Fluids, 3, 1991.
[22] A. Leonard. Energy cascade in large-eddy simulations of turbulent fluid flows.
Adv. in Geophys., A-18:237248, 1974.
[23] L. Prandtl. Bericht u
ber die entstehung der turbulenz. Zeitschrift f
ur Angewadte Mathematik und Mechanik, 5:136139, 1925.
[24] P. R. Spalart and S. R. Allmaras. A one equation turbulence model for aerodynamic flows. La Recherche A`erospatiale, 1:521, 1994.
[25] W. Jones and B. Launder. The prediction of laminarization with a twoequation model of turbulence. Int. Heat and Mass Transfer, 15:301314, 1972.
[26] B. Merci. Numerical Simulation and Modelling of Turbulent Combustion.
Proefschrift tot het verkrijgen van de graad van doctor in de toegepaste wetenschappen, Vakgroep Mechanica van Stroming, Warmte en Verbranding, Universiteit Gent, 2001.
[27] M. Hallback, D.S. Henningson, A.V. Johansson, and P.H. Alfredsson. Turbulence and Transition Modelling. Springer-Verlag Berlin, 1995.
[28] Z.Y. Yang and T.H. Shih. A new time scale based k- model for near-wall
turbulence. AIAA Journal, 31(7):11911198, 1993.

BIBLIOGRAPHY

175

[29] K. Hanjalic and B.E. Launder. Contribution towards a reynolds-stress closure


for low-reynolds-number turbulence. J. of fluid Mechanics, 74(4):593610,
1976.
[30] H. Pitsch and N. Peters. A consistent flamelet formulation for nonpremixed combustion considering differential diffusion effects. Combust.
Flame, (114):2640, 1998.
[31] H. Pitsch. Extended flamelet model for les of non-premixed combustion. Annual Research Briefs, Center for Turbulence Research:149158, 2000.
[32] H. Pitsch. Unsteady flamelet modeling for differential diffusion in turbulent
jet diffusion flames. Combust. Flame, (123):358374, 2000.
[33] M. Ihme and H. Pitsch. LES of a non-premixed flame using an extended
flamelet/progress variable model. AIAA 2005-0558, 2005.
[34] T. Poinsot and D. Veynante. Theoretical and Numerical Combustion. Edwards,
2001.
[35] S.P. Burke and T.E.W. Schumann. Diffusion flames. Industrial and Engineering Chemistry, 20:9981005, 1928.
[36] An analytical investigation of three general methods of calculating chemicalequilibrium compositions. NASA TN D-473, 1960. F.J. Zeleznik and S. Gordon.
[37] Computer program for computation of complex chemical equilibrium compositions, rocket performance, incident and reflected shocks, and chapman-jouguet
detonations. NASA SP-273, 1971. S. Gordon and B.J. McBride.
[38] Nasa computer program cea (chemical equilibrium with applications).
http://www.grc.nasa.gov/WWW/CEAWeb/ceaWhat.htm.
[39] Computer program for calculation of complex chemical equilibrium compositions and applications i. analysis. NASA RP 1311, 1994. S. Gordon and B.J.
McBride.
[40] Chemkin-iii: A fortran chemical kinetics package for the analysis of gas-phase
chemical and plasma kinetics. http://www.ca.sandia.gov/chemkin/. Robert
J. Kee, Fran M. Rupley and Ellen Meeks (Thermal and Plasma Processes
Department) and James A. Miller (Combustion Chemistry Department Sandia
National Laboratories).
[41] Transport coeffcients for the nasa lewis chemical equilibrium program. NASA
Report TM-4647, 1995. R.A. Svehla.
[42] Nist-janaf thermochemical tables : Fourth edition. Journal of Physical and
Chemical Reference Data , monograph no. 9, 1998. Chase, M.W. Jr.

176

BIBLIOGRAPHY

[43] Chem 1d. http://www.combustion.tue.nl/chem1d. A one-dimensional laminar


flame code, Eindhoven University of Technology.
[44] Oppdif: A fortran program for computing opposed-flow diffusion flames.
SAND96-8243 Sandia National Laboratories, 1996. A. E. Lutz, R. J. Kee,
J. F. Grcar and F. M. Rupley.
[45] Reduced kinetic mechanisms and asymptotic approximations for methane-air
flames. Lecture Notes in Physics Springer-Verlag, 1991. M.D. Smooke.
[46] Gri-mech (releases 1.2, 2.11, 3.0). http://www.me.berkeley.edu/gri mech/.
University of California at Berkeley, Stanford University, The University of
Texas at Austin, and SRI International.
[47] B. Somers. The Simulation of Flat Flames with Detailed and Reduced Chemical
Models. Proefschrift tot het verkrijgen van de graad van doctor, Technische
Universiteit Eindhoven, 1994.
[48] S.B. Pope. Pdf methods for turbulent reacting flow. Progr. Energy Comb.
Sci., 19:119192, 1985.
[49] W. Kollmann J.-Y. Chen and R. W. Dibble. Pdf modeling of turbulent nonpremixed methane jet flames. Combustion Science and Technology, 64:315
346, 1989.
[50] J.-Y. Chen and W. Kollmann. Pdf modeling and analysis of thermal no formation in turbulent nonpremixed hydrogen-air jet flames. Combust. Flame,
88:397412, 1992.
[51] C. Dopazo. Recent developments in pdf methods. In Turbulent Reacting Flows,
Academic Press, Ed. P.A.Libby and F.A.Williams, pages 375474, 1994.
[52] Chong M. Cha. Transported pdf modeling of turbulent nonpremixed combustion. Center for Turbulence Research Annual Research Briefs, pages 7986,
2001.
[53] B. NAUD. PDF modeling of turbulent sprays and flames using a particle
stochastic approach. Proefschrift ter verkrijging van de graad van doctor,
Technische Universiteit Delft, 2003.
[54] B. Merci, B. Naud, D. Roekaerts, and E.Dick. Combination of a transported
scalar pdf method with a non-linear k-e model. application to piloted jet diffusion flames. In Proceedings of the European Combustion Meeting (ECM2003),
Orlans, France, pages CDrom, 6 pages, 2003.
[55] B. Merci, B. Naud, and D. Roekaerts. low and mixing fields for transported
scalar pdf simulations of a piloted jet diffusion flame (delft flame iii). Flow,
Turbulence and Combustion, 74(3):239272, 2005.

BIBLIOGRAPHY

177

[56] B. Merci, D. Roekaerts, B. Naud, and S. B. Pope. Comparative study of


micromixing models in transported scalar pdf simulations of turbulent nonpremixed bluff body flames. Combust. Flame, 146:109130, 2006.
[57] J.P.H. Sanders and I. Gokalp. Scalar dissipation rate modelling in variable
density turbulent axisymmetric jets and diffusion flames. Physics of Fluids,
10(4):938948, 1998.
[58] A.Y.Klimenko. Multicomponent diffusion of various admixtures in turbulent
flow. Fluid Dyn. (USSR), 25:327, 1990.
[59] R.W. Bilger. Conditional moment closure for turbulent reacting flows. Phys.
Fluids A, 5:436, 1993.
[60] V. Raman and H. Pitsch. Les/filtered-density function simulation of turbulent
combustion with detailed chemistry. Center for Turbulence Research Annual
Research Briefs, pages 297309, 2005.
[61] V. Raman, H.Pitsch, and R.O. Fox. Hybrid large-eddy simulation/lagrangian
filtered-density-function approach for simulating turbulent combustion. Combust. Flame, 143:5678, 2005.
[62] F. Gao and E.E. OBrien. A LES scheme for turbulent reacting flows. Physics
of Fluids, 5:12821284, 1998.
[63] L. Vervisch and D. Veynante andD. Olivari. Introduction to Turbulent Combustion. VKI Lecture Series 99-04, 1999.
[64] A.W. Cook and J.J. Riley. A subgrid Model for equilibrium chemistry in
turbulent flows. Phys. Fluids, 6:28682870, 1994.
[65] C. Wall, B.J. Boersma, and P. Moin. An evaluation of the assumed beta
probability density function subgrid-scale model for large eddy simulation of
nonpremixed, turbulent combustion with heat release. Phys. Fluids, 12:2522
2529, 2000.
[66] A.W. Cook and W.K. Bushe. A subgrid-scale Model for the scalar dissipation
rate in nonpremixed combustion. Phys. Fluids, 11:746748, 1999.
[67] J. Reveillon and L. Vervisch. Response of the dynamic LES model to heat
release induced effects. Phys. Fluids, 8:22482250, 1996.
[68] C.D. Pierce and P. Moin. A dynamic model for SGS variance and dissipation
rate of a conserved scalar. Phys. Fluids, 10:30413044, 1998.
[69] A.W. Cook and J.J. Riley. Subgrid-Scale Modeling for Turbulent Reacting
Flows. Combust. Flame, 112:593606, 1998.
[70] S.M. De Bruyn Kops, J.J. Riley, G. Kosaly, and A.W. Cook. Investigation
of Modeling for Non-Premixed Turbulent Combustion. Flow, Turbulence and
Combustion, 60:105122, 1998.

178

BIBLIOGRAPHY

[71] N. Peters. Laminar diffusion flamelet models in non-premixed turbulent combustion. Prog. Energy Comb. Sci., 10:319339, 1984.
[72] S.S. Girimaji and Y. Zhou. Analysis and modeling of subgrid scalar mixing
using numerical data. Phys. Fluids A, 8:12241236, 1996.
[73] C. Jimenez, F. Ducros, B. Cuenot, and B. Bedat. Subgrid sclae variance
and dissipation of a scalar field in large eddy simulations. Physics of Fluids,
13(6):17481754, 2001.
[74] R. Borghi and M. Destriau. Combustion and flames, chemical and physical
principles. Editions TECHNIP, 1998.
[75] S. Candel, D. Thevenin, N. Darabiha, and D. Veynante. Problems and perspectives in numerical combustion. In Computational Methods in Applied Sciences,
pages 4862. John Wiley and Sons Ltd, 1996.
[76] S. Candel, D. Thevenin, N. Darabiha, and D. Veynante. Combust. Sci. and
Tech., 149:297337, 1999.
[77] C.Meneveau and T.Poinsot. Stretching and quenching of flamelets in premixed
turbulent combustion. Combust. Flame, 86:311332, 1991.
[78] T.Mantel and R.Borghi. A new model of premixed wrinkled flame propagation
based on a scalar dissipation equation. Combust. Flame, 96:443457, 1994.
[79] K.N.C. Bray. Studies of the turbulent burning velocity. Proc. R. Soc. London
A., 431:315335, 1990.
[80] L. Vervisch, E. Bidaux, K.N.C. Bray, and W. Kollmann. Surface density function in premixed turbulent combustion modelling, similarities between propability density function and flame surface approaches. Phys. Fluids, 7(10):2496
2503, 1995.
[81] A.Trouve and Poinsot. The evolution equation for the flame surface density.
J. Fluid Mech., 278:131, 1994.
[82] K.N.C. Bray, M.Champion, and P.A.Libby. The interaction between turbulence and chemistry in premixed turbulent flames. Turbulent Reactive Flows,
Lecture Notes in Engineering, pages 541563, 1989.
[83] F. Gouldin, K. Bray, and J.Y. Chen. Chemical closure model for fractal
flamelets. Combust. Flame, 77:241, 1989.
[84] C. Fureby. A fractal flame-wrinkling large eddy simulation model for premixed
turbulent combustion. Proceedings of the Combustion Institute, 30:593601,
2005.
[85] R.O.S. Prasad and J.P. Gore. An evaluation of flame surface density models
for turbulent premixed jet flames. Combust. Flame, 116:114, 1999.

BIBLIOGRAPHY

179

[86] J.-M. Duclos, D. Veynante, and T. Poinsot. A comparison of flamelet models


for premixed turbulent combustion. Combust. Flame, 95:101117, 1993.
[87] D. Veynante, J. Piana, J.M. Duclos, and C.Mantel. Experimental analysis of
flame surface density model for premixed turbulent combustion. Proceedings
of the Combustion Institute, 26:413420, 1996.
[88] T. Mantel, J.M. Samaniego, and C.T. Bowman. Fundamental mechanisms in
premixed turbulent flame propagation via vortex flame interactions - part ii :
numerical simulation. Combust. Flame, 118:557582, 1999.
[89] M. Herrmann. Numerical Simulation of Premixed Turbulent Combustion
Based on a level Set Flamelet Model. Dissertation zur erlangung des akademischen grades eines doktors der ingenieurswissenschaften, RTWH Aachen, 2001.
[90] Peters N. Turbulent combustion. Cambridge University Press, 2000.
[91] J. Janicka and A. Sadiki. Large eddy simulation of turbulent combustion
systems. In 30th Int. Symp. on Combustion, pages 537548, Chicago, 2004.
[92] D. Veynante. Large eddy simulation for turbulent combustion. In Proc. European Combustion Meeting, pages 311332, Louvain-la-Neuve, 2005.
[93] L. Selle, G. Lartigue, T. Poinsot, R. Koch, K.U. Schildmacher, W. Krebs,
B. Prade, P. Kaufmann, and D. Veynante. Compressible large-eddy simulation of turbulent combustion in complex geometry on unstructured meshes.
Combust. Flame, 137:489505, 2004.
[94] P. Moin. Advances in large eddy simulation methodology for complex flows.
Heat and Fluid Flow, 23:710720, 2002.
[95] P. Moin and S. Apte. Large-eddy simulation of realistic gas turbines combustors. In 42nd AIAA Aerospace Sciences Meeting 2004-0330, pages 110, Reno,
2004.
[96] S. Menon, C. Stone, and N. Patel. Multi scale modeling for LES of engineering
designs of large-scale combustors. In 42nd AIAA Aerospace Sciences Meeting
2004-0157, pages 115, Reno, 2004.
[97] N. Peters. Turbulent Combustion. Cambridge University Press, 2000.
[98] M. Boger, D. Veynante, H. Boughanem, and A. Trouve. Direct numerical
simulation analysis of flame surface density concept for large eddy simulation
of turbulent premixed combustion. In 27th Symposium on Combustion, pages
917925, 1998.
[99] O. Colin, F. Ducros, D. Veynante, and T. Poinsot. A thickened flame model for
large eddy simulations of turbulent premixed combustion. Physics of Fluids,
12(7):18431863, 2000.

180

BIBLIOGRAPHY

[100] C. Nottin, R. Knikker, M. Boger, and D. Veynante. Large eddy simulation


of an acoustically excited turbulent premixed flame. In 28th Int. Symp. on
Combustion, pages 6773, 2000.
[101] A.R. Kertstein, W. Ashurst, and F.A. Williams. Field equation for interface
propagation in an unsteady homogeneous flow field. Phys. Rev. A, 37:2728
2731, 1988.
[102] M. D
using. Large-Eddy Simulation turbulenter Vormischflammen. Fortschrittberichte vdi, reihe 6 energietechnik, nr. 517, EKT TU-Darmstadt, 2004.
[103] H.G. Im, T.S. Lund, and J.H. Ferziger. Large eddy simulation of turbulent
front propagation with dynamic subgrid models. Physics of Fluids A, 9:3826
3833, 1997.
[104] M. Freitag and J. Janicka. Investigation of a strongly swirled unconfined
premixed flame using les. In Proc. Comb. Inst. 31, 2006.
[105] T.M. Smith and S. Menon. LES of Turbulent Reacting Stagnation Point Flows.
AIAA paper 97-0372, 1997.
[106] V.K. Chakravarthy and S. Menon. Characteristics of a subgrid model for
Turbulent Premixed Flames. AIAA paper 97-****, 1997.
[107] R. Knikker, D. Veynante, and C. Meneveau. A dynamic flame surface density
model for large eddy simulation of turbulent premixed combustion. Physics
of Fluids, 16(11):9194, 2004.
[108] E. R. Hawkes and R.S. Cant. A flame surface density approach to large-eddy
simulation of premixed turbulent combustion. Proceedings of the Combustion
Institute, 28:5158, 2000.
[109] P. Domingo, L. Vervish, S. Payet, and R. Hauguel. DNS of a premixed turbulent V flame and LES of a ducted flame using a FSD-PDF subgrid scale
closure with FPI-tabulated chemistry. Combust. Flame, 143(4):566586, 2005.
[110] T.D. Buttler and P.J. ORourke. A numerical method for two dimensional
unsteady reacting flows. In 16th Int. Symp. on Combustion, pages 15031515,
1977.
[111] M. Besson, P. Bruel, L. Champion, and B. Deshaies. Experimental analysis
of combustion flows developing over a plane-symmetric expansion. Journal of
Thermophysics and Heat Transfer, 14(1):5968, 2000.
[112] F. Charlette, C. Meneveau, and D. Veynante. A power-law flame wrinkling
model for les of premixed turbulent combustion part 1: Non-dynamic formulation and initial tests. Combust. Flame, 131:155180, 2002.
[113] C. Angelberger, D.Veynante, F. Egolfopoulos, and T. Poinsot. Large eddy simulation of combustion instabilities in premixed flames. Center for Turbulence
Research Processings of the Summer Program, pages 6182, 1998.

BIBLIOGRAPHY

181

[114] F. Charlette, C. Meneveau, and D. Veynante. A power-law flame wrinkling


model for LES of premixed turbulent combustion part 2: Dynamic formulation
and initial tests. Combust. Flame, 131:181197, 2002.
[115] C. Meneveau and T. Poinsot. Stretching and quenching of flamelets in premixed turbulent combustion. Combust. Flame, 86:311332, 1991.
[116] P.K. Yeung, S.S. Girimaji, and S.B. Pope. Straining and scalar dissipation on
material surfaces in turbulence: implications for flamelets. Combust. Flame,
79:340365, 1990.
[117] M. Klein. Direkte Numerische Simulation des prim
aren Strahlzerfalls in Einstoffzerst
auberd
usten. Fortschritt-berichte vdi, reihe 6 energietechnik, nr.,
EKT TU-Darmstadt, 2002.

[118] H. Forkel.
Uber
die Grobstruktur-simulation turbulenter WasserstoffDiffusionsflammen. Fortschritt-berichte vdi, reihe 6 energietechnik, nr. 428,
EKT TU-Darmstadt, 2000.
[119] A. Kempf. Large-Eddy Simulation of Non-Premixed Turbulent Flames.
Fortschritt-berichte vdi, reihe 6 energietechnik, nr. 513, EKT TU-Darmstadt,
2003.
[120] H. Forkel and J. Janicka. Large-Eddy Simulation of a Turbulent Hydrogen
Diffusion Flame. Flow, Turbulence and Combustion, 65:163175, 2000.
[121] M. Klein. Direct numerical simulation of a spatially developing waterfilm at
moderate reynolds number. International Journal of Heat and Fluid Flow,
26:722731, 2005.
[122] A. Kempf, H. Forkel, A. Sadiki, and J. Janicka andJ. Chen. Large-eddy simulation of a counterflow configuration with and without combustion. Proc.
Comb. Inst., 28:3540, 2000.
[123] A. Kempf, R.P. Lindstedt, and J. Janicka. Large-eddy simulation of a bluffbody stabilized nonpremixed flame. Combust. Flame, 144:170189, 2006.
[124] F. Flemming, M. Freitag, A. Sadiki, and J. Janicka. Time resolved combustion
noise investigation based on a wave equation approach. In 12th AIAA/CEAS
Aeroacoustics Conference 8-10 May, Cambridge, MA, USA, 2006. AIAA-20062615.
[125] F. Flemming, A. Sadiki, and J. Janicka. Investigation of combustion noise
using a les/caa hybrid approach. In Proc. 31st Comb. Inst., accepted, 2006.
[126] M. Freitag, M. Klein, M. Gregor, A. Nauert, D. Geyer, C. Schneider, A. Dreizler, and J. Janicka. Mixing analysis of a swirling recirculating flow using dns
and experimental data. In TSFP, Williamsburg, pages 491496, 2005.
[127] M. Freitag and M. Klein. Direct numerical simulation of a recirculating,
swirling flow. Flow Turbulence and Combustion, 75:5166, 2005.

182

BIBLIOGRAPHY

[128] A. Jameson, W. Schmidt, and E. Turkel. Numerical Solutions of the


Euler equations by Finite Volume Methods with Runge-Kutta Time Stepping
Schemes. AIAA Paper 81-1259, January 1981.
[129] C.Hirsch. Numerical computation of internal and external flows, Computational Methods for Inviscid and Viscous Flows, Vol. 2. Wiley Interscience,
1990.
[130] Lacor C. Solution of time dependent reynolds averaged navier-stokes equations
with the finite volume method. Course of GrasMech, 1998.
[131] B. Van Leer. Towards the ultimate conservative difference scheme.V.A second
order sequel to Godunovs method. J. Comput. Phys., 32:101136, 1979.
[132] A.Harten. On a class of high resolution total-variation -stable finite-difference
schemes. SIAM J.Num.Analysis, 21:1, 1984.
[133] G.D.Van Albada, B.Van Leer, and W.W.Roberts. A comparative study of
computational methods in cosmic gas dynamics. Astron. Astrophysics, 108:76,
1982.
[134] PHOENICS.
/d lecs/numerics/scheme.htm, 2002.

www.cham.co.uk/phoenics/d polis

[135] P.K.Sweby. High resolution schemes using flux-limiters for hyperbolic conservation laws. SIAM J.Num.Anal., 21:995, 1984.
[136] A. Jameson. Time Dependent Calcualtions Using Multigrid, with Applications
to Unsteady Flows past Airfoils and Wings. AIAA Paper, 91-1596, June 24-26
1991.
[137] S. Venkateswaran and L. Merkle. Analysis of Preconditioning Methods for
the Euler and Navier-Stokes Equations. Von Karman Lecture Series 1999-03,
1999.
[138] T. Broeckhoven. Studie van externe stromingen bij stall condtities. Masters
thesis, Dept. Stromingsmechanica, Faculteit Toegepaste Wetenschappen, Vrije
Universiteit Brussel, 2002.
[139] B. Lessani, J. Ramboer, and C. Lacor. Efficient Large-Eddy Simulations of
Low Mach number Flows Using Preconditioning and Multigrid. In Proc. Third
AFOSR Int. Conf. on DNS/LES, 2001.
[140] B. Lessani, J. Ramboer, and C. Lacor. Efficient Large-Eddy Simulations of
Low Mach number Flows Using Preconditioning and Multigrid. International
Journal of Computational Fluid Dynamics, 18:221, 2004.
[141] N.Alkishriwi, M.Mainke, and W.Schroder. A large-eddy simulation method
for low mach number flows using preconditioning and multigrid. Computers
and Fluids, 35:11261136, 2006.

BIBLIOGRAPHY

183

[142] A. J. Chorin. Numerical Solution of the Navier-Stokes Equations. Mathematics


of Computations, 23, 1968.
[143] R. Temam. Sur lapproximation de la solution des equations de Navier-Stokes
par la method des pas fractionnaires. (I): Arch. Rational Mech. Anal., 32:135
153, 1969.
[144] R. Temam. Sur lapproximation de la solution des equations de Navier-Stokes
par la method des pas fractionnaires. (II): Arch. Rational Mech. Anal., 33:377
385, 1969.
[145] J.Kim and P.Moin. Application of a fractional-step method to incompressible
navier-stokes equations. Journal Comp. Phys., 59:308323, 1985.
[146] A.W.Cook and J.J.Riley. Direct numerical simulation of a turbulent reactive
plume on a parallel computer. Journal Comp. Phys., 129:263283, 1996.
[147] J. de Charentenay, D. Thvenin, and B. Zamuner. Comparison of direct numerical simulations of turbulent flames using compressible or low-mach number formulations. International Journal for Numerical Methods in Fluids, 39(6):497
515, 2002.
[148] B. Lessani and M. V. Papalexandris. Time-accurate calculation of variable
density flows with strong temperature gradients and combustion. Journal
Comp. Phys., 212(1):218 246, 2006.
[149] C. D. Pierce. Progress-Variable Approach for Large-Eddy Simulation of Turbulent Combustion. Ph.D. thesis, Stanford University, 2001.
[150] R. P. Fedorenko. The Speed of Convergence of one iterative process. USSR
Comp. Math. and Math. Phys., 4:227235, 1964.
[151] A. Brandt. Multi-level Adaptive Solutions to Boundary Value Problems. Mathematical Computation, 31, 1977.
[152] W. Hackbusch. Multi-Grid Methods and Applications. Springer-Verlag Berlin,
Heidelberg, 1985.
[153] A. Jameson. Multigrid Algorithms for Compressible Flow Calculations. Multigrid Methods, II-Proceedings of the Second European Conference, pages 166
201, 1986.
[154] A. Jameson. Multigrid Solution of the Euler Equations Using Implicit
Schemes. AIAA-85-0293, 1985.
[155] W. K. Anderson, J. L. Thomas, and D. L. Whitfield. Multigrid Acceleration
of the Flux Split Euler Equations. AIAA-86-0274, 1986.
[156] W. Mudler. A High Resolution Euler Solver. AIAA-89-1949, 1989.

184

BIBLIOGRAPHY

[157] Z. W. Zhu. Multigrid Operation and Analysis for Complex Aerodynamics.


Ph.D. Thesis, Vrije Universiteit Brussel, 1996.
[158] P. Eliasson. Dissipation Mechanism and Multigrid Solutions in a Multiblock
Solver for Compressible Flow. In PhD Thesis, Dept. Numerical Analysis and
Computer Science, KTH, 1993.
[159] W.P.Jones. Turbulence Modelling and Numerical Solution Methods: Turbulent
Reacting Flows. Academic Press, 1994.
[160] NUMECA International. FINETM Numecas Flow Integrated Environment,
User Manual FINE Version 6.1-1. 5 Avenue Franklin Roosevelt, 1050 Brussels,
Belgium, Web : http://www.numeca.com, 2003.
[161] C.Mengler, C.Heinrich, A.Sadiki, and J.Janicka. Numerical prediction of momentum and scalar fields in a jet in cross flow: comparison of les and second
order turbulence closure calculations. In TSFP2 Vol II, pages p425430, 2001.
[162] M. Klein, A. Sadiki, and J. Janicka. A digital filter based generation of inflow
data for spatially developing direct numerical or large eddy simulations. J.
Comp. Physics, 186:652665, 2003.
[163] S. Lee, S. K. Lele, and P. Moin. Simulation of spatially evolving turbulence
and the applicability of taylors hypothesis in compressible flow. Physics of
Fluids A 4, 7:15211530, 1992.
[164] H.Le and P.Moin. Direct numerical simulation of turbulent flow over a backward facing step. Technical Report TF-58, Stanford University, 1994. Technical Report.
[165] S. A. Stanley, S. Sarkar, and J. P. Mellado. A study of the flow-field evolution
and mixing in a planar turbulent jet using direct numerical simulation. J.
Fluid Mech., 450:377407, 2002.
[166] T. S. Lund, X. Wu, and K. D. Squires. Generation of turbulent inflow data
for spatially-developing boundary layer simulations. Journal Comp. Phys.,
140:233258, 1998.
[167] A.Kempf, M.Klein, and J.Janicka. Efficient generation of initial- and inflow
conditions for transient turbulent flows in arbitrary geometries. Flow, Turbulence and Combustion, 74(1):6784, 2005.
[168] M.Fathali, M.Klein., T. Broeckhoven, C.Lacor, and M.Baelmans. Generation
of turbulent inflow and initial conditions based on multi-correlated random
fields. In ICCFD, 2006.
[169] M.Fathali, M.Klein., T. Broeckhoven, C.Lacor, and M.Baelmans. Generation
of turbulent inflow and initial conditions based on multi-correlated random
fields. Journal Comp. Phys. in review.

BIBLIOGRAPHY

185

[170] E.M. Saiki and S. Biringen. Numerical simulation of a cylinder in uniform flow:
Application of a virtual boundary method. J. Comp. Phys, 123(2):450465,
1996.
[171] M. Freitag and M. Klein. An improved method to assess the quality of large
eddy simulations in the context of implicit filtering. Journal of Turbulence, 7,
2006.
[172] T. Broeckhoven, M. Freitag, C. Lacor, A. Sadiki, and J. Janicka. Investigation of subgrid scale wrinkling models and their impact on the artificially
thickened flame model in large eddy simulations. In The Cyprus International
Symposium on Complex Effects in Large Eddy Simulations, Cyprus, 2005.
[173] T. Broeckhoven, M. Freitag, C. Lacor, A. Sadiki, and J. Janicka. Investigation
of subgrid scale wrinkling models and their impact on the artificially thickened
flame model in large eddy simulations. Lecture Notes in Computational Science
and Engineering (LNCSE) Kassinos, S.; Langer, C.; Iaccarino, G.; Moin, P.
(Eds.), 56:365382, 2006.
[174] E. Brizuela. Note : Accurate normalisation of the beta-function pdf. Journal
Comp. Phys., 119:385387, 1995.
[175] E. Brizuela. Accurate numerical computation of the beta pdf, 2006. private
communication.
[176] Masri A. R., Dibble R. W., and Barlow R. S. The structure of turbulent nonpremixed flames revealed by raman-rayleigh-lif measurements. Prog. Energy
Combust. Sci., 22:307362, 1996.
[177] R.S. Barlow and J.H. Frank. Effects of turbulence on species mass fractions
in methane/air jet flames. Proc. Combust. Inst., 27:10871095, 1998.
[178] R.S. Barlow and A.N. Karpetis. Measurements of scalar variance, scalar dissipation, and length scales in turbulent piloted methane/air jet flames. Flow
Turb. Combust., 72:427448, 2004.
[179] R.S. Barlow, J.H. Frank, and J.Y. Chen. Scalar structure and transport effects
in piloted methane/air jet flames. Combust. Flame, 143:433449, 2005.
[180] Ch. Schneider, A. Dreizler, J. Janicka, and E.P. Hassel. Flow field measurements of stable and locally extinguishing hydrocarbon-fuelled jet flames. Combust. Flame, 135:185190, 2003.
[181] B. Merci, E.Dick, J. Vierendeels, D.Roekaerts, and T.W.J. Peeters. Application of a new cubic turbulence model to piloted and bluff-body diffusion
flames. Combust. Flame, 126(1-2):15331556, 2001.
[182] B. Merci, E.Dick, and C. De Langhe. Application of an improved epsilonequation to a piloted jet diffusion flame. Combust. Flame, 131(4):465468,
2002.

186

BIBLIOGRAPHY

[183] R.W. Bilger, S.H. Starner, and R.J.Kee. On reduced mechanisms for methane
air combustion in nonpremixed flames. Combust. Flame, 80:135149, 1990.
[184] J. H. Frank and S. A. Kaiser. High-resolution rayleigh imaging of dissipative
structures in a turbulent jet flame. In 13th Int Symp on Applications of Laser
Techniques to Fluid Mechanics, Lisbon, Portugal, 2006.
[185] S. A. Kaiser and J. H. Frank. Imaging of dissipative structures in the near field
of a turbulent non-premixed jet flame. In Proc. Combust. Inst., Heidelberg,
Germany, 2006.
[186] W. Meier, R.S. Barlow, Y.-L. Chen, and J.-Y. Chen. Raman/rayleigh/lif measurements in a turbulent ch4/h2/n2 jet diffusion flame: Experimental techniques and turbulencechemistry interaction. Combust. Flame, 123:326343,
2000.
[187] L. Duchamp de Lageneste and H. Pitsch. Progress in large-eddy simulation of
premixed and partially premixed turbulent combustion. In Center for Turbulent Research, Annular Research Briefs, pages 97107, Stanford, USA, 2001.
[188] P. Domingo, L. Vervish, S. Payet, and R. Hauguel. Analysis of scalar dissipation rate and turbulent transport in LES of premixed turbulent combustion. In
TSFP4, International Symposium on Turbulence and Shear Flow Phenomena,
Williamsburg, VA, USA, 2005.
[189] M. D
using. Large-Eddy Simulation turbulenter Vormischflammen. PhD thesis,
Darmstadt, 2004.
[190] C. Fureby. A fractal flame-wrinkling large eddy simulation model for premixed
turbulent combustion. 30th Symposium on Combustion, pages 593601, 2004.
[191] P. Bruel and P.D. Nguyen. Generic 2-d combustor: Determination of lean
extinction limits (1/2). WP3 6M, Modeling of Low Emmisions Combustion
Chambers Using Large Eddy Simulations (MOLECULES), CNRS-ENSMA,
2001. Technical Report.
[192] P. Bruel and P.D. Nguyen. Generic 2-d combustor: Determination of lean
extinction limits (2/2). WP3 12M, Modeling of Low Emmisions Combustion
Chambers Using Large Eddy Simulations (MOLECULES), CNRS-ENSMA,
2002. Technical Report.
[193] P. Bruel and P.D. Nguyen. Report and data files of the combustion flow field.
WP3 18M, Modeling of Low Emmisions Combustion Chambers Using Large
Eddy Simulations (MOLECULES), CNRS-ENSMA, 2002. Technical Report.
[194] P. Bruel and P.D. Nguyen. Report and data files of the combustion flow field.
WP3 24M, Modeling of Low Emmisions Combustion Chambers Using Large
Eddy Simulations (MOLECULES), CNRS-ENSMA, 2003. Technical Report.

BIBLIOGRAPHY

187

[195] TNF7. Seventh international workshop on measurement and computation of


turbulent nonpremixed flames. Chicago, 2004.
[196] R. Borghi. On the structure and morphology of turbulent premixed flames,
chapter 6.3 Jets, pages 117138. Recent advances in Aerospace Science.
Plenum Publishing Corporation, 1985.
[197] H. Pitsch and L. Duchamp de Lageneste. Large-eddy simulation of premixed
turbulent combustion using a level-set approach. In 29th Symposium on Combustion, pages 20012008, Sapporo, Japan, 2002.
[198] M. D
using, A. Sadiki, and J. Janicka. Towards a classification of models for the
numerical simulation of premixed combustion based on a generalised regime
diagram. Combust. Theory Modeling, 10(1):105132, 2006.
[199] D.E. Abbot and S.J. Kline. Experimental investigation of a subsonic turbulent flow over single and double backward-facing steps. Journal of basic
engineering, 84:317325, 1962.
[200] Y. Morinishi, T.S. Lund, and O.V. Vasilyevand P. Moin. Fully conservative
higher order finite difference schemes for incompressible flow. Journal Comp.
Phys., 143(1):90124, 1998.

You might also like