Está en la página 1de 74

Turbulent combustion modeling

Denis Veynante
a,
*
, Luc Vervisch
b
a
Laboratoire E.M2.C., CNRS et Ecole Centrale Paris, Grande Voie des Vignes, 92295 Chatenay-Malabry Cedex, France
b
Institut National des Sciences Appliquees de Rouen, UMR CNRS 6614/CORIA, Campus du Madrillet,
Avenue de l'Universite BP 8, 76801 Saint Etienne du Rouvray Cedex, France
Received 4 November 2000; accepted 12 October 2001
Abstract
Numerical simulation of ames is a growing eld bringing important improvements to our understanding of combustion. The
main issues and related closures of turbulent combustion modeling are reviewed. Combustion problems involve strong coupling
between chemistry, transport and uid dynamics. The basic properties of laminar ames are rst presented along with the major
tools developed for modeling turbulent combustion. The links between the available closures are illuminated from a generic
description of modeling tools. Then, examples of numerical models for mean burning rates are discussed for premixed turbulent
combustion. The use of direct numerical simulation (DNS) as a research instrument is illustrated for turbulent transport
occurring in premixed combustion, gradient and counter-gradient modeling of turbulent uxes is addressed. Finally, a review
of the models for non-premixed turbulent ames is given. q 2002 Published by Elsevier Science Ltd.
Keywords: Premixed ames; Non-premixed ames; Turbulent combustion; Scalar turbulent transport; Direct numerical simulation; Reynolds
averaged NavierStokes modeling
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
2. Balance equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
2.1. Instantaneous balance equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196
2.2. Reynolds and Favre averaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 197
2.3. Favre averaged balance equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
2.4. Filtering and Large Eddy Simulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198
3. Major properties of premixed, non-premixed and partially premixed ames . . . . . . . . . . . . . . . . . . . . 200
3.1. Laminar premixed ames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
3.2. Laminar diffusion ames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201
3.3. Partially premixed ames . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
4. A direct analysis: Taylor's expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
5. Scales and diagrams for turbulent combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
5.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
5.2. Turbulent premixed combustion diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
5.2.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
5.2.2. Combustion regimes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
5.2.3. Comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209
5.3. Non-premixed turbulent combustion diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
5.3.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
Progress in Energy and Combustion Science 28 (2002) 193266 PERGAMON
www.elsevier.com/locate/pecs
0360-1285/02/$ - see front matter q 2002 Published by Elsevier Science Ltd.
PII: S0360-1285(01)00017-X
* Corresponding author. Tel.: 133-1-41-13-10-80; fax: 133-1-47-02-80-35.
E-mail address: denis@em2c.ecp.fr (D. Veynante).
6. Tools for turbulent combustion modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
6.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212
6.2. Scalar dissipation rate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
6.3. Geometrical description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
6.3.1. G-eld equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
6.3.2. Flame surface density description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
6.3.3. Flame wrinkling description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
6.4. Statistical approaches: probability density function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
6.4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218
6.4.2. Presumed probability density functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
6.4.3. Pdf balance equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
6.4.4. Joint velocity/concentrations pdf . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
6.4.5. Conditional moment closure (CMC) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
6.5. Similarities and links between the tools . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221
7. Reynolds-averaged models for turbulent premixed combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
7.1. Turbulent ame speed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
7.2. Eddy-break-up model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223
7.3. BrayMossLibby model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
7.3.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
7.3.2. BML model analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
7.3.3. Recovering mean reaction rate from tools relations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
7.3.4. Reynolds and Favre averaging . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
7.3.5. Conditional averagingcounter-gradient turbulent transport . . . . . . . . . . . . . . . . . . . . . 227
7.3.6. Extensions to partially premixed combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
7.4. Models based on the ame surface area estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
7.4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
7.4.2. Algebraic expressions for the ame surface density S . . . . . . . . . . . . . . . . . . . . . . . . . . 229
7.4.3. Flame surface density balance equation closures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
7.4.4. Analysis of the ame surface density balance equation . . . . . . . . . . . . . . . . . . . . . . . . . 234
7.4.4.1. Turbulent transport. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
7.4.4.2. Strain rate induced by the mean ow eld, A
T
. . . . . . . . . . . . . . . . . . . . . . . . 234
7.4.4.3. Strain rate a
T
due to turbulent motions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
7.4.4.4. Propagation and curvature terms. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
7.4.5. Flame stabilization modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
7.4.6. A related approach: G-equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
8. Turbulent transport in premixed combustion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
8.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
8.2. Direct numerical simulation analysis of turbulent transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
8.2.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
8.2.2. Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
8.3. Physical analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
8.3.1. Comments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241
8.4. External pressure gradient effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
8.5. Counter-gradient transportexperimental results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
8.6. To include counter-gradient turbulent transport in modeling . . . . . . . . . . . . . . . . . . . . . 243
8.7. Towards a conditional turbulence modeling? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
9. Reynolds averaged models for non-premixed turbulent combustion . . . . . . . . . . . . . . . . . . . . . . 245
9.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
9.2. Fuel/air mixing modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
9.2.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
9.2.2. Balance equation and simple relaxation model for ~ x . . . . . . . . . . . . . . . . . . . 246
9.3. Models assuming innitely fast chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
9.3.1. Eddy dissipation model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
9.3.2. Presumed pdf: innitely fast chemistry model . . . . . . . . . . . . . . . . . . . . . . . . 247
9.4. Flamelet modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
9.4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 194
9.4.2. Flame structure in composition space, Y
SLFM
i
(Z
p
; x
p
) . . . . . . . . . . . . . . . . . . . 249
9.4.3. Mixing modeling in SLFM . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
9.4.4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
9.5. Flame surface density modeling, coherent ame model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
9.6. MIL model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
9.7. Conditional moment closure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
9.8. Pdf modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
9.8.1. Turbulent micromixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
9.8.2. Linear relaxation model, IEM/LMSE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
9.8.3. GIEM model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 256
9.8.4. Stochastic micromixing closures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
9.8.5. Interlinks PDF/ame surface modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
9.8.6. Joint velocity/concentrations pdf modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
10. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
1. Introduction
The number of combustion systems used in power
generation and transportation industries is growing
rapidly. This induces pollution and environmental problems
to become critical factors in our societies. The accurate
control of turbulent ames therefore appears as a real
challenge.
Computing is now truly on par with experiment and
theory as a research tool to produce multi-scale information
that is not available by using any other technique. Computa-
tional uid dynamics (CFD) is efciently used to improve
the design of aerodynamical systems, and today no real
progress in design can be made without using CFD. With
the same objectives, much work has been devoted to turbu-
lent combustion modeling, following a variety of
approaches and distinct modeling strategies. This paper is
intended to provide a generic review of these numerical
models.
A wide range of coupled problems are involved in turbu-
lent ames:
The uid mechanical properties of the combustion system
must be well known to carefully describe the mixing
between reactants and, more generally, all transfer
phenomena occurring in turbulent ames (heat transfer,
molecular diffusion, convection, turbulent transport,
etc.).
Detailed chemical reaction schemes are necessary to esti-
mate the consumption rate of the fuel, the formation of
combustion products and pollutant species. A precise
knowledge of the chemistry is absolutely required to
predict ignition, stabilization or extinction of reaction
zones together with pollution.
Two (liquid fuel) and three (solid fuel) phase systems may
be encountered. Liquid fuel injection is a common pro-
cedure and the three-dimensional spatial distribution of
gaseous reactants depends on complex interactions
between the breakdown of the liquid sheets, the vapor-
ization of the liquid, turbulent mixing, and droplet
combustion.
Radiative heat transfer is generated within the ame by
some species and carbon particles resulting from soot
formation and transported by the ow motion. In
furnaces, walls also interact with combustion through
radiative transfer.
Turbulent combustion modeling is therefore a very
broad subject. All the aspects of the problem are not
addressed in the present review. We will only focus on
the closure schemes developed and used to understand
and calculate turbulent transport and mean burning rates
in turbulent ames. The detail of chemistry, its reduc-
tion, tabulation, etc. are not considered. However, the
links existing between the models are identied, showing
similarities which are sometimes much stronger than is
usually thought.
Numerical modeling of ames is developed from the
following steps (Fig. 1):
Under assumptions such as the high activation energy
limit, asymptotic analysis [14] allows the analytical
determination of ame properties in well-dened model
problems (ignition, propagation of ame front, instabil-
ities and acoustics, etc.). This approach, limited to simpli-
ed situations, leads to analytical results exhibiting
helpful scale factors (dimensionless numbers) and
major ame behaviors. Asymptotic analysis is particu-
larly well suited to perform quantitative comparison
between various phenomena.
Simplied experiments are useful to understand the basic
properties of combustion (laminar ames, ame/vortex
interactions, etc.) [5,6]. These experiments are accom-
panied by numerical simulations of laminar ames incor-
porating complex chemistry and multi-species transports
along with radiative heat losses [7].
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 195
For given chemistry and transport model, in direct
numerical simulation (DNS) all the scales of the turbu-
lence (time and length) are calculated without resorting to
closures for turbulent uxes and mean burning rate.
Turbulent ames are analyzed in simple congurations
to extract data impossible to measure in experiments, and
to isolate some specic phenomena (heat release, Lewis
number) [812].
Because of the large number of degrees of freedominvolved
in turbulent combustion, a full DNS of a practical system
cannot be performed and averaging techniques leading to
unclosed equations are necessary. Models for turbulent ames
are then developed: closure techniques are proposed for
unknown terms found in exact averaged balance equations.
Once the models have been implemented in numerical
codes, validation procedures are required. The numerical
modeling is validated against measurements obtained from
experiments. Congurations as close as possible to actual
industrial systems are chosen for these tests. Then, the ultimate
step is the simulation of a real combustion device.
The decomposition discussed in Fig. 1 is quite formal.
Turbulent combustion modeling is actually a continuous
ring between theoretical studies to analyze combustion,
understand ames and improve models, implementation of
these models into CFD, experimental measurements
and comparison between these experimental data and the
numerical results.
Following a short presentation of the balance equa-
tions for reactive ows (Section 2), a rst part is
devoted to a brief description of laminar ames (Section
3). After a presentation of the unsuccessful Taylor's
expansion for closing the mean burning rate (Section
4), the physical analysis leading to turbulent combustion
diagrams is developed (Section 5). Then, modeling tools
available to derive turbulent combustion models are
described and the relations between a priori quite differ-
ent formalisms are established (Section 6). The next
three sections are devoted to combustion modeling in
the context of Reynolds Averaged NavierStokes
(RANS) equations. For premixed turbulent combustion,
we review the available closures for the mean reaction
rate (Section 7) and turbulent transport (Section 8). In a
subsequent section (Section 9), the modeling of the
mean burning rate in non-premixed turbulent ames is
addressed.
2. Balance equations
2.1. Instantaneous balance equations
The basic set of balance equations comprises the classical
NavierStokes, species and energy transport equations.
These instantaneous local balance equations are, using the
classical lettering [1315]:
Mass:
2r
2t
1
2ru
j
2x
j
= 0 (1)
Momentum (i = 1,2,3):
2ru
i
2t
1
2ru
j
u
i
2x
j
= 2
2p
2x
i
1
2t
ij
2x
j
1F
i
(2)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 196
Fig. 1. Combustion modeling steps.
where t
ij
denotes the viscous force tensor and F
i
a body
force.
Species (N species with k = 1; ; N):
2rY
k
2t
1
2ru
j
Y
k
2x
j
= 2
2J
k
j
2x
j
1 _ v
k
(3)
where J
k
j
is the molecular diffusive ux of the species k
and _ v
k
the mass reaction rate of this species per unit
volume.
Total enthalpy h
t
= h 1u
i
u
i
=2 :
2rh
t
2t
1
2ru
j
h
t
2x
j
=
2p
2t
1
2
2x
j
(J
h
j
1u
i
t
ij
) 1u
j
F
j
(4)
where u
i
t
ij
and u
i
F
j
denote respectively the power due to
viscous and body forces.
These equations are closed by expressions for the species
molecular uxes and the viscous forces. In practical situa-
tions, all uids are assumed to be Newtonian, i.e. the viscous
tensor is given by the Newton law:
t
ij
= m
l
2u
i
2x
j
1
2u
j
2x
i
_ _
2
2
3
m
l
d
ij
2u
k
2x
k
_ _
(5)
where the molecular viscosity m
l
depending on the uid
properties is introduced. d
ij
is the Kronecker symbol.
Species molecular diffusivities are generally described
using the Fick law, assuming a major species:
J
k
j
= 2
m
l
Sc
k
2Y
k
2x
j
(6)
Sc
k
is the Schmidt number of the species k, dened as:
Sc
k
=
m
l
rD
k
(7)
D
k
is the molecular diffusivity of the species k relative to the
major species.
More complex expressions may be used to describe multi-
species molecular diffusion. Soret effect (species diffusion
under temperature gradients) and molecular transport due to
pressure gradients are usually neglected. Enthalpy diffusion
is described according to the Fourier law:
J
h
j
= 2
m
l
Pr
2h
2x
j
1

N
k=1
Pr
Sc
k
21
_ _
h
k
2Y
k
2x
j
_ _
(8)
The Prandtl number Pr compares the diffusive transport
of momentum (viscous forces) and temperature. In the
previous expressions, radiative heat transfer and Dufour
effect (enthalpy diffusion under mass fraction gradients)
are neglected. The Prandtl number is written as a function
of the thermal diffusivity l and the constant pressure
specic heat C
p
:
Pr =
m
i
C
p
l
_ _
(9)
Then, the Lewis number Le
k
of the species k, comparing
thermal and mass diffusivities is introduced:
Le
k
=
Sc
k
Pr
_ _
=
l
rC
p
D
k
_ _
(10)
Under the assumption of unity Lewis number, the enthalpy
diffusive ux (Eq. (8)) is simplied and mass fraction and
enthalpy balance equations are formally identical if 2P=2t;
u
i
t
ij
and u
i
F
j
are negligible (low Mach number assumption)
[16]. This assumption is generally made to simplify tur-
bulent ame modeling, especially in premixed ames
when species mass fractions and temperature are assumed
to be equivalent variables. Nevertheless, thermo-diffusive
instabilities occur in premixed systems when the Lewis
number is lower than unity (for example for hydrogen).
One direct consequence of these instabilities is an increase
of the premixed ame area and of the global reaction rate
[14,17].
2.2. Reynolds and Favre averaging
Unfortunately, the full numerical solution of the instanta-
neous balance equations is limited to very simplied cases
(DNS [9,11,12]), where the number of time and length
scales present in the ow is not too great. To overcome
this difculty, an additional step is introduced by averaging
the balance equations to describe only the mean ow eld
(local uctuations and turbulent structures are integrated in
mean quantities and these structures have no longer to be
described in the simulation). Each quantity Q is split into a
mean

Q and a deviation from the mean denoted by Q
/
:
Q =

Q 1Q
/
with

Q
/
= 0 (11)
Then, the previous instantaneous balance equations may be
ensemble averaged to derive transport equations for the
mean quantity

Q: This classical Reynolds averaging tech-
nique, widely used in non-reacting uid mechanics, brings
unclosed correlations such as u
/
Q
/
that are unknown and
must be modeled. The numerical procedure is called
Reynolds Averaged NavierStokes (RANS) modeling.
In turbulent ames, uctuations of density are observed
because of the thermal heat release, and Reynolds averaging
induces some additional difculties. Averaging the mass
balance equation leads to:
2 r
2t
1
2
2x
i
ru
i
1r
/
u
/
i
_ _
= 0 (12)
where the velocity/density uctuations correlation r
/
u
/
i
appears. To avoid the explicit modeling of such correlations,
a Favre (mass weighted) [18] average
~
Q is introduced and
any quantity is then decomposed into Q =
~
Q 1Q
//
:
~
Q =
rQ
r
;

Q
//
=
r Q 2
~
Q
_ _
r
= 0 (13)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 197
The Favre averaged continuity equation:
2 r
2t
1
2 r ~ u
i
2x
i
= 0 (14)
is then formally identical to the Reynolds averaged conti-
nuity equation for constant density ows. This result is true
for any balance equations (momentum, energy, mass frac-
tions, etc.). Nevertheless, Favre averaging is only a math-
ematical formalism:
There is no simple relation between Favre,
~
Q and
Reynolds,

Q; averages. A relation between
~
Q and

Q
requires the knowledge of density uctuations correla-
tions r
/
Q
/
remaining hidden in Favre averaging (see
Section 7.3.4):
r
~
Q = r

Q 1r
/
Q
/
(15)
Comparisons between numerical simulations, providing
Favre averaged quantities
~
Q; with experimental results
are not obvious. Most experimental techniques determine
Reynolds averaged data

Q and differences between
~
Q and

Q may be signicant (Section 7.3.4 and Fig. 17).


2.3. Favre averaged balance equations
Averaging instantaneous balance equations yields:
Mass:
2 r
2t
1
2 r ~ u
j
2x
j
= 0 (16)
Momentum (i = 1,2,3):
2 r ~ u
i
2t
1
2 r ~ u
j
~ u
i
2x
j
= 2
2 r

u
//
i
u
//
j
2x
j
2
2 p
2x
i
1
2 t
ij
2x
j
1

F
i
(17)
Chemical species (for N species, k = 1; ; N):
2 r
~
Y
k
2t
1
2 r ~ u
j
~
Y
k
2x
j
= 2
2 r

u
//
j
Y
//
k
2x
j
2
2J
k
j
2x
j
1

_ v
k
(18)
Total enthalpy
~
h
t
:
2 r
~
h
t
2t
1
2 r ~ u
j
~
h
t
2x
j
= 2
2 r

u
//
j
h
//
t
2x
j
1
2 p
2t
1
2
2x
j
J
h
j
1u
i
t
ij
_ _
1u
j
F
j
(19)
The objective of turbulent combustion modeling is to
propose closures for the unknown quantities appearing in
the averaged balance equations, such as:
Reynolds stresses

u
//
i
u
//
j
: The turbulence model provides
an approximation for this term. The closure may be done
directly or by deriving balance equations for these
Reynolds stresses. However, most combustion works
are based on turbulence modeling developed for non-
reacting ows, such as k1; simply rewritten in terms
of Favre averaging, and heat release effects on the
Reynolds stresses are generally not explicitly included.
Species (

u
//
j
Y
//
k
) and temperature (

u
//
j
T) turbulent uxes.
These uxes are usually closed using a gradient transport
hypothesis:
r

u
//
j
Y
//
k
= 2
m
t
Sc
kt
2
~
Y
k
2x
j
(20)
where m
t
is the turbulent viscosity, estimated from the
turbulence model, and Sc
kt
a turbulent Schmidt number
for the species k.
Nonetheless, theoretical and experimental works have
demonstrated that this assumption may be wrong in
some premixed turbulent ames and counter-gradient
turbulent transport may be observed [19,20] (i.e. in an
opposite direction compared to the one predicted by Eq.
(20), see Sections 7.3.5 and 8).
Laminar diffusive uxes J
k
j
; J
h
j
; etc. are usually small
compared to turbulent transport, assuming a sufciently
large turbulence level (large Reynolds numbers limit).
Species chemical reaction rates

_ v
k
: Turbulent combus-
tion modeling generally focuses on the closure of these
mean burning rates.
These equations, closed with appropriate models, allow
only for the determination of mean quantities, that may
differ from the instantaneous ones. Strong unsteady mixing
effects, resulting from the rolling up of shear layers, are
observed in turbulent ames, and the knowledge of steady
statistical means is indeed not always sufcient to describe
turbulent combustion. An alternative is to use large eddy
simulation (LES).
2.4. Filtering and Large Eddy Simulation
The objective of Large Eddy Simulation (LES) is to ex-
plicitly compute the largest structures of the ow (typically
the structures larger than the computational mesh size),
while the effects of the smaller one are modeled. LES is
widely studied in the context of non-reactive ows [21
24], its application to combustion modeling is still at an
early stage [25]. As in RANS, the complex coupling
between micromixing and chemical reactions occurring at
unresolved scales needs models, however, LES possesses
some attractive properties:
Large structures in turbulent ows generally depend on
the geometry of the system. On the contrary, smaller
scales feature more universal properties. Accordingly,
turbulence models may be more efcient when they
have to describe only the smallest structures.
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 198
Turbulent mixing controls most of the global ame prop-
erties. In LES, unsteady large scale mixing (between
fresh and burnt gases in premixed ames or between
fuel and oxidizer in non-premixed burners) is simulated,
instead of being averaged.
Most reacting ows exhibit large scale coherent struc-
tures [26], which are also especially observed when
combustion instabilities occur. These instabilities result
from the coupling between heat release, the hydrody-
namic ow eld and acoustic waves. They need to be
avoided because they induce noise variations of the
main properties of the system, large heat transfers
and, even in some extreme cases, the destruction of
the device.LES may be a powerful tool to predict the
occurrence of such instabilities [27] and consequently
improve passive or active control systems.
With LES, large structures are explicitly computed and
instantaneous fresh and burnt gases zones, with different
turbulence characteristics (Section 8.7) are clearly iden-
tied. This may help to describe some properties of the
ame/turbulence interaction.
In LES, the relevant quantities Q are ltered in the spec-
tral space (components greater than a given cut-off
frequency are suppressed) or in the physical space (weighted
averaging in a given volume). The ltered operation is
dened by:

Q(x) =
_
Q(x
p
)F(x 2x
p
) dx
p
(21)
where F is the LES lter. Standard lters are:
A cut-off lter in the spectral space:

F(k) =
1 if k # p=D
0 otherwise
_
(22)
where k is the spatial wave number. This lter preserves
the length scales greater than the cut-off length scale 2D:
A box lter in the physical space:
F(x) = F(x
1
; x
2
; x
3
) =
1=D
3
if ux
i
u # D=2; i = 1; 2; 3
0 otherwise
_
(23)
where (x
1
; x
2
; x
3
) are the spatial coordinates of the loca-
tion x. This lter corresponds to an averaging of the
quantity Q over a box of size D.
A Gaussian lter in the physical space:
F(x) = F(x
1
; x
2
; x
3
)
=
6
pD
2
_ _
3=2
exp 2
6
D
2
x
2
1
1x
2
2
1x
2
3
_ _
_ _
(24)
All these lters are normalized:
_11
21
_11
21
_11
21
F(x
1
; x
2
; x
3
) dx
1
; dx
2
; dx
3
= 1 (25)
In combusting ows, a mass-weighted, Favre ltering, is
introduced as:
r
~
Q(x) =
_
rQ(x
p
)F(x 2x
p
)dx
p
(26)
Instantaneous balance equations (Section 2) may be ltered
to derived balance equations for the ltered quantities

Q or
~
Q: This derivation should be carefully conducted:
Any quantity Q may be decomposed into a ltered
component

Q and a `uctuating' component Q
/
; according
to: Q =

Q 1Q
/
: But, in disagreement with classical
Reynolds averaging (ensemble average), Q
/
may be non-
zero:
Q
/
(x) =
_
Q(x
p
) 2

Q(x
p
)
_ _
F(x 2x
p
)dx
p
=
_
Q(x
p
)F(x 2x
p
)dx
p
2
_

Q(x
p
)F(x 2x
p
)dx
p
=

Q(x) 2Q(x) (27)
where
Q(x) =
_ _
Q(x
1
)F(x
p
2x
1
)dx
1
_ _
F(x 2x
p
)dx
p
=
__
Q(x
1
)F(x
p
2x
1
)F(x 2x
p
)dx
1
dx
p


Q(x)
(28)
To summarize:
Q

Q; Q
/
0;
~
~
Q
~
Q;

Q
//
0 (29)
The relations used in RANS Q =

Q; Q
/
= 0;
~
~
Q =
~
Q;

Q
//
= 0
are true when a cut-off lter in the spectral space is chosen
(Eq. (22)). Then, all the frequency components greater than
a cut-off wave number k
c
= p=D vanish.
The derivation of balance equations for the ltered quan-
tities

Q or
~
Q requires the exchange of ltering and differ-
entiation operators. This exchange is theoretically valid only
under restrictive assumptions and is wrong, for example,
when the lter size varies (lter size corresponding to the
mesh size, depending on the spatial location). This point has
been carefully investigated [28]. In most simulations, the
uncertainties due to this operator exchange are neglected
and assumed to be incorporated in subgrid scale modeling.
Filtering the instantaneous balance equations leads to
equations formally similar to the Reynolds averaged balance
equations given in Section 2.3:
mass:
2 r
2t
1
2 r ~ u
j
2x
j
= 0 (30)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 199
momentum (for i = 1; 2; 3):
2 r ~ u
i
2t
1
2 r ~ u
j
~ u
i
2x
j
= 2
2
2x
j
r u
i
u
j
2 ~ u
i
~ u
j
_ _ _ _
2
2 p
2x
i
1
2 t
ij
2x
j
1

F
i
(31)
Chemical species (N species, k = 1; ; N):
2 r
~
Y
j
2t
1
2 r ~ u
j
~
Y
k
2x
j
= 2
2
2x
j
r

u
j
Y
k
2 ~ u
j
~
Y
k
_ _ _ _
1

_ v
k
(32)
Total enthalpy h
t
= h 1u
i
u
i
=2
2 r
~
h
t
2t
1
2 r ~ u
j
~
h
t
2x
j
= 2
2
2x
j
r

u
j
h
t
2 ~ u
j
~
h
t
_ _ _ _
1
2 p
2t
1
2
2x
j
J
h
j
1u
i
t
ij
_ _
1u
j
F
j
(33)
where

Q and
~
Q denote LES ltered quantities instead of
ensemble means.
The unknown quantities are:
Unresolved Reynolds stresses
_
u
i
u
j
2 ~ u
i
~ u
j
_
; requiring a
subgrid scale turbulence model.
Unresolved species uxes
_

u
j
Y
k
2 ~ u
j
~
Y
k
_
and enthalpy
uxes
_

u
j
h
t
2 ~ u
j
~
h
t
_
:
Filtered laminar diffusion uxes J
k
j
; J
h
j
:
Filtered chemical reaction rate

_ v
k
:
These ltered balance equations, coupled to subgrid scale
models may be numerically solved to simulate the unsteady
behavior of the ltered elds. Compared to direct numerical
simulations (DNS), part of the information contained in the
unresolved scales is lost (and should be modeled).
Compared to RANS, LES provides valuable information
on the large resolved motions.
Either using RANS or LES, combustion occurs at the
unresolved scales of the computations. Then, the basic
tools and formalism of turbulent combustion modeling are
somehow the same for both techniques. Most of the RANS
combustion models can be modied and adapted to LES
modeling. The scope of this review is limited to RANS.
3. Major properties of premixed, non-premixed and
partially premixed ames
3.1. Laminar premixed ames
The structure of a laminar premixed ame is displayed in
Fig. 2. Fresh gases (fuel and oxidizer mixed at the molecular
level) and burnt gases (combustion products) are separated
by a thin reaction zone (typical thermal ame thicknesses,
d
l
, are about 0.11 mm). A strong temperature gradient is
observed (typical ratios between burnt and fresh gases
temperatures are about 57). Another characteristic of a
premixed ame is its ability to propagate towards the
fresh gases. Because of the temperature gradient and the
corresponding thermal uxes, fresh gases are preheated
and then start to burn. The local imbalance between diffu-
sion of heat and chemical consumption leads to the propa-
gation of the front. The propagation speed S
L
of a laminar
ame depends on various parameters (fuel and oxidizer
compositions, fresh gases temperature, etc.) and is about
0.11 m/s. There is an interesting relation between the ther-
mal ame thickness, d
l
, the laminar ame speed, S
L
and the
kinematic viscosity of the fresh gases, n:
Re
f
=
d
l
S
L
n
< 4 (34)
where the thermal thickness d
l
corresponds to a temperature
jump of 98% of the temperature difference between fresh
and fully burnt products. The ame Reynolds number, Re
f
,
is then almost constant. This relation, derived, for example,
from the Zeldovich/Frank-Kamenetskii (ZFK) theory
[14,15] is often implicitly used in theoretical derivation of
models for premixed turbulent combustion.
For a simple one-step irreversible chemical scheme:
reactants - products
the ame is described using a progress variable c, such as
c = 0 in the fresh gases and c = 1 in the fully burnt ones.
This progress variable may be dened as a reduced tempera-
ture or a reduced mass fraction:
c =
T 2T
u
T
b
2T
u
or c =
Y
F
2Y
u
F
Y
b
F
2Y
u
F
(35)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 200
Fig. 2. Structure of a laminar plane premixed ame.
where T, T
u
and T
b
are respectively the local, the unburnt gas
and the burnt gas temperatures. Y
F
, Y
u
F
and Y
b
F
are respec-
tively the local, unburnt gas and burnt gas fuel mass frac-
tions. Y
b
F
is non-zero for a rich combustion (fuel in excess).
For a unity Lewis number (same molecular and thermal
diffusivities), without heat losses (adiabatic combustion)
and compressibility effects, the two denitions (35) are
equivalent and mass and low Mach number energy balance
equations reduce to a single balance equation for the
progress variable:
2rc
2t
17 ruc
_ _
= 7 rD7c
_ _
1 _ v (36)
The previous Eq. (36) may be recast in a propagative form,
introducing the displacement speed v of the iso-c surface:
2c
2t
1u7c =
1
r
7 rD7c
_ _
1 _ v
u7cu
_ _
.........,,.........
displacement speed
7c } } = wu7cu (37)
Eq. (36) then describes the displacement of an iso-c surface
with the displacement speed w measured relative to the ow.
Introducing the vector n normal to the iso-c surface and
pointing towards fresh gases (n = 27c=u7cu); the displace-
ment speed may be split into three contributions:
w =
1
r 7c } }
nn : 7 rD7c
_ _
2D7n 1
1
r 7c } }
_ v (38)
w = 2
1
ru7cu
2
2n
rDu7cu
_ _
........,,........
w
n
2 D7n
..,,..
w
c
1
1
ru7cu
_ v
...,,...
w
r
where 2=2n = n7 denotes a normal derivative. w
n
corre-
sponds to molecular diffusion normal to the iso-c surface,
w
c
is related to the curvature 7n of this surface and corre-
sponds to tangential diffusion. w
r
is due to the reaction rate
_ v: In a rst approximation, w
n
1w
r
may be modeled with
the laminar ame speed, S
L
, whereas w
c
incorporates wrink-
ling surface effects and may be expressed using Markstein
lengths [29].
The propagation of reactive fronts has been the subject of
various developments and more discussion may be found in
Ref. [2] and references therein.
3.2. Laminar diffusion ames
In laminar diffusion ames, fuel and oxidizer are on both
sides of a reaction zone where the heat is released. The
burning rate is controlled by the molecular diffusion of the
reactants toward the reaction zone (Figs. 3 and 4). In a
counter-owing fuel and oxidizer ame (Fig. 4), the amount
of heat transported away from the reaction zone is exactly
balanced by the heat released by combustion. A steady
planar diffusion ame with determined thickness may be
observed in the vicinity of the stagnation point. Increasing
the jet velocity, quenching occurs when the heat uxes
leaving the reaction zone are greater than the chemical
heat production. The structure of a steady diffusion ame
therefore depends on ratios between characteristic times
representative of molecular diffusion and chemistry [30].
The thicknesses of the mixing zone and of the reaction
zone vary with these characteristic times. In opposition
with premixed ames:
Diffusion ames do not benet from a self-induced propa-
gation mechanism, but are mainly mixing controlled.
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 201
Fig. 3. Generic structure of a laminar diffusion ame.
Fig. 4. Sketch of a counter-owing fuel and oxidizer diffusion
ame.
The thickness of a diffusion ame is not constant, but
depends on the local ow properties.
Let us consider the irreversible single step chemical reac-
tion between fuel and oxidizer:
F 1sO - (1 1s)P
where s is the mass stoichiometric coefcient.
In term of mass fraction, this chemical reaction may be
written:
n
P
dY
P
= n
F
dY
F
1n
o
dY
o
where dY
F
, dY
O
and dY
P
are the variations of fuel, oxidizer
and product mass fractions. n
i
are the stoichiometric molar
coefcients of the reaction, W
i
denotes the species molar
weight and _ v is the reaction rate. The balance equations
for mass fractions and temperature are necessary to identify
the properties of the ame:
2rY
F
2t
17(ruY
F
) = 7(rD
F
7Y
F
) 2n
F
W
F
_ v
2rY
O
2t
17(ruY
O
) = 7(rD
O
7Y
O
) 2n
O
W
O
_ v
2rT
2t
17(ruT) = 7
l
C
P
7T
_ _
1n
F
W
F
Q
C
p
_ _
_ v
The molecular diffusion is expressed using the Fick law,
the chemical rate of fuel and oxidizer are respectively _ v
F
=
n
F
W
F
_ v and _ v
O
= n
O
W
O
_ v: Q is the amount of heat released
by the combustion of an unit mass of fuel.
The internal structure of diffusion ames is usually
discussed using the extent of mixing between fuel and oxidi-
zer. It is rst assumed that fuel and oxidizer molecular
diffusivities are equal (i.e. D
F
= D
O
= D). Combining the
transport equation for Y
F
and Y
O
, a conserved scalar (quan-
tity that is not inuenced by the chemical reaction, a
SchwabZeldovitch variable) w(Y
F
; Y
O
) = Y
F
2Y
O
=s is
introduced, with the mass stoichiometric coefcient s =
(n
O
W
O
=n
F
W
F
): The mixture fraction Z is then dened by
normalizing w using values in the fuel and oxidizer streams.
Z evolves through the diffusive layer from zero (oxidizer) to
unity (fuel):
Z =
f
Y
F
Y
F;o
2
Y
O
Y
O;o
11
f 11
(39)
Y
F,o
is the fuel mass fraction in the fuel feeding stream.
Similarly, Y
O,o
is the oxidizer mass fraction in the oxidizer
stream (for instance, in air, Y
O,o
<0.23), f is the chemical
equivalence ratio:
f =
sY
F;o
Y
O;o
(40)
The mixture fraction follows the balance equation:
2rZ
2t
17(ruZ) = 7(rD7Z) (41)
Other SchwabZeldovitch variables w(Y
F
, T) and w(Y
O
, T)
(conserved scalars) may be derived by combining the
variables (Y
F
, T) and (Y
O
, T). The mixture fraction and these
additional conserved scalars are linearly related and one
may write:
Y
O
(x; t) = Y
O;o
(1 2Z(x; t))
.......,,.......
Mixing
1
n
O
W
O
n
F
W
F
C
p
Q
_ _
[Z(x; t)(T
F;o
2T
O;o
) 1(T
O;o
2T(x; t))]
.......................,,.......................
Combustion
(42)
Y
F
(x; t) = Y
F;o
Z(x; t)
....,,....
Mixing
1
C
p
Q
[Z(x; t)(T
F;o
2T
O;o
) 1(T
O;o
2T(x; t))]
...................,,...................
Combustion
(43)
where T
O,o
and T
F,o
are the temperatures of the fuel and
oxidizer streams respectively. Using these algebraic re-
lations, the diffusion ame is fully determined when the
mixture fraction Z and any one of T, Y
F
, or Y
O
is known.
The conserved scalar approach may still be useful when
fuel and oxidizer molecular diffusivities differ, but an ad-
ditional mixture fraction:
Z
L
=
F
Y
F
Y
F;o
2
Y
O
Y
O;o
11
F11
(44)
should be introduced, satisfying [31]:
r
DZ
Dt
=
1
L
7
l
C
p
7Z
L
_ _
(45)
where:
L= Le
O
(1 1f)=(1 1F) with F = (Le
O
=Le
F
)f
(46)
where Le
i
is the Lewis number of the species i. The relations
between Z and Z
L
are given in Table 1. When Le
O
= Le
F
;
Z
L
= Z: In experiments or in simulations involving complex
chemistry, the mixture fraction is dened from mass frac-
tions of atomic elements [32].
Mass fractions and temperature balance equations may be
reorganized into a new frame where Z is one of the coordi-
nates (see for instance Ref. [14] or Ref. [33]). A local or-
thogonal coordinate system attached to the surface of
stoichiometric mixture is introduced and the derivatives in
the stoichiometric plane are denoted '. For unity Lewis
number and using Eq. (41), the species transport equation
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 202
may be written:
r
2Y
i
2t
1ru
'
7
'
Y
i
= rx
2
2
Y
i
2Z
2
17
'
(rD7
'
Y
i
)
2rD7
'
(lnu7Zu)7
'
Y
i
1 _ v
i
(47)
In Eq. (47), x is the scalar dissipation rate of the mixture
fraction Z:
x = D
2Z
2x
j
2Z
2x
j
_ _
= Du7Zu
2
(48)
measuring the inverse of a diffusive time t
x
= x
21
: As this
time decreases, mass and heat transfers through the stoichio-
metric surface are enhanced.
When iso-Z surface curvatures are not too strong, the
gradients measured along the stoichiometric surface are
smaller than the gradients in the direction Z perpendicular
to the stoichiometric surface, the balance equation for the
mass fractions reduces to:
r
2Y
i
2t
= rx
2
2
Y
i
2Z
2
1 _ v
i
(49)
Neglecting unsteady effects, the time derivative vanishes
and for unity Lewis numbers, the ame structure is fully
described by:
rx
2
2
Y
i
2Z
2
1 _ v
i
= 0 and rx
2
2
T
2Z
2
1 _ v
T
= 0 (50)
showing that the chemical reaction rate is directly related to
the function T(Z, x). Under these hypothesis, the diffusion
ame is completely determined as a function of the mixture
fraction Z and the scalar dissipation rate x (or 7Z):
Y
i
= Y
i
(Z; x); T = T(Z; x)
An expression for x(Z,t) and full solutions for various lami-
nar ames may be derived from asymptotic developments
[30,34], or solving Eq. (50) leading to Fig. 5.
A coordinate j is dened across the one-dimensional
amelet such as:
dZ
dj
= 2n
Z
7Z = u7Zu (51)
where n
Z
denotes the normal vector to the iso-Z surfaces,
pointing towards Z = 0: The reaction rate may be integrated
in the direction j through the ame, using Eq. (50):
_
V
i
=
_11
21
_ v
i
(j)dj =
_1
0
_ v
i
(x; Z)
u7Zu
dZ
= 2
_1
0
rDu7Zu
2
2
Y
i
2Z
2
dZ (52)
Assuming that r, D and u7Zu do not vary across the ame
(this is typical of a amelet assumption used in turbulent
combustion modeling where the ame is assumed very
thin),
_
V
i
becomes:
_
V
i
< 2rDu7Zu
2Y
i
2Z
_ _
Z=1
Z=0
= 2 rD
2Y
i
2j
_ _
Z=1
Z=0
(53)
This last relation illustrates how the integrated reaction
rate of a species i is directly related to the molecular diffu-
sion ux of that species through the ame.
Diffusion combustion is limited by two regimes corre-
sponding to pure mixing of the reactants and innitely fast
chemistry (Fig. 5). When the chemistry is innitely fast, the
temperature depends on mixing through Z, but not on the
rate of mixing x [35]. Then, piecewise relationships exist
between Z, Z
L
, species mass fractions and temperature,
summarized in Table 1. Eq. (43) provides the maximum
ame temperature T
f
obtained when Y
F
= Y
O
= 0 and Z =
Z
st
= 1=(1 1f)
T
f
=
T
F;o
1T
O;of
1Y
F;o
Q
C
p
1 1f
In many combustion systems, the innitely fast chemistry
hypothesis cannot be invoked everywhere. For example in
ignition problems or in the vicinity of stabilization zones,
and more generally when large velocity gradients are found.
The characterization of diffusion ames from the innitely
fast chemistry situation to the quenching limit is therefore of
fundamental interest for turbulent combustion. The counter-
ow diffusion ame (Fig. 4) is a generic conguration well
suited to reproduce and to understand the structure and the
extinction of laminar diffusion ames. These extinction
phenomena have been theoretically described using asymp-
totic developments [30,34,36]. A diffusive time t
x
< x
21
st
=
(Du7Zu
2
)
21
Z=Z
st
and a chemical time t
c
are combined to build a
Damkohler number Da
p
= (t
x
=t
c
) < (t
c
x
st
)
21
: The response
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 203
Table 1
Piecewise relations for innitely fast chemistry including non-unity Lewis number. Z
st
= 1=(1 1f) and Z
L
st
= 1=(1 1F) (see Eq. (46)). The
subscript
o
denotes a quantity measured in pure fuel or oxidizer, T
f
is the ame temperature
Oxidizer side
Z , Z
st
and Z
L
, Z
L
st
Fuel side
Z . Z
st
and Z
L
. Z
L
st
Z = Z
L
(1 1F)=(1 1f) Z = (f(Z
L
(1 1F) 21)=F11)=(1 1f)
Y
F
= 0 Y
O
= 0
Y
O
= Y
O;o
(1 2Z(1 1f)) Y
F
= Y
F;o
(Z(1 1f) 21)=f
T = (T
f
2T
O;o
)Z(f 11) 1T
O;o
T = (T
F;o
2T
f
)(Z(f 11) 21)=f 1T
f
of the burning rate to variations of Da
p
leads to the so-called
`S' curve (Fig. 6) [14]. Starting from a situation where the
chemistry is fast, decreasing Da
p
(increasing x) makes the
burning rate and transport through the stoichiometric
surface greater, until chemistry cannot keep up with the
large heat uxes. Then, extinction develops. The value of
the Damkohler Da
p
q
at the extinction point may be estimated
by quantifying the leakage of fuel (or oxidizer) through the
stoichiometric surface [37].
Two limit cases are thus important for non-premixed
turbulent combustion modeling: pure mixing without
combustion (Da
p
- 0) and innitely fast chemistry
(Da
p
-1). These cases delineate the domain where ames
may develop in planes (Z, Y
F
), (Z, Y
O
) and (Z, T) (Fig. 5).
Moreover, for a given location within a diffusion ame, by
traveling along the normal to the stoichiometric surface,
T(Z) can be constructed and characterizes the combustion
regime (i.e. fast or slow chemistry, Fig. 5). Many turbulent
combustion models are based on this description of diffusion
ame; when the ow is turbulent, T(Z) is replaced by the
mean temperature calculated for a given value of Z, i.e. for a
given state in the mixing between fuel and oxidizer.
3.3. Partially premixed ames
In non-premixed combustion, some partial premixing of
the reactants may exist before the reaction zone develops.
Then, the pure diffusive/reactive layer, as observed in a
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 204
Fig. 5. Inner structure of non-premixed ames. The distribution in mixture fraction space of fuel, oxidizer and temperature lies between the
innitely fast chemistry limit and the pure mixing case. The thickness of the diffusive zone l
d
is estimated from the scalar dissipation rate x at
the stoichiometric surface, whereas the characteristic thickness of the reaction zone l
r
depends on both l
d
and the Damkohler number [12].
laminar diffusion ame, may not be the unique relevant
model problem. Furthermore, many ames in burners are
stabilized by the recirculation of burnt gases, leading to
stabilization mechanisms controlled by the mixing between
fuel, oxidizer, and burnt gases. The mixtures feeding the
reaction zone are then not always pure fuel and pure
oxidizer. There are situations where partial premixing is
clearly important:
Auto-ignition in a non-homogeneous distribution of fuel
and oxidizer, where the reactants can be mixed before
auto-ignition occurs.
Laminar or turbulent ame stabilization, when combus-
tion does not start at the very rst interface between fuel
and oxidizer in the vicinity of burner exit, so that fuel and
oxidizer may mix without burning.
After quenching of the reaction zone, the reactants may
mix leading to possibility of re-ignition and combustion
in a partially premixed regime [38].
The triple ame is an interesting model problem to
approach partially premixed combustion. In a laminar
shear layer where the mixing between cold fuel and oxidizer
develops, a diffusion ame may be stabilized at the splitter
plate by the combination of heat losses with viscous ow
effects, or, further downstream [39]. In this latter case,
combustion starts in a region where fuel and oxidizer have
been mixed in stoichiometric proportion. The resulting
premixed kernel tends to propagate towards fresh gases
and contributes to the stabilization of the trailing diffusion
ame. In a mixing layer conguration, the stoichiometric
premixed kernel evolves to a rich partially premixed ame
in the direction of the fuel stream, while a lean partially
premixed ame develops on the air side (Fig. 7). These
two premixed ames are curved because their respective
propagation velocities decrease when moving away from
the stoichiometric condition. The overall structure,
composed of two premixed ames and of a diffusion
ame, is usually called `triple ame'. Such triple ames
have been rstly experimentally observed by Phillips [40].
Since this pioneer work, more recent experiments have
conrmed the existence of triple ames in laminar ows
[4143]. Theoretical studies [4448] and numerical simula-
tions [4953] have been devoted to triple ames. The propa-
gation speed of triple ames is controlled by two
parameters: the curvature of the partially premixed front,
increasing with the scalar dissipation rate imposed in front
of the ame, and the amount of heat release. The effect of
heat release is to deect the ow upstream of the triple
ame, making the triple ame speed greater than the propa-
gation speed of a planar stoichiometric ame. This deec-
tion also induces a decrease of the mixture fraction gradient
in the trailing diffusion ame. The triple ame velocity
decreases when increasing the scalar dissipation rate at the
ame tip. Triple ame velocity response to variations of
scalar dissipation rate may be derived by approximating
the ame tip by a parabolic prole and using results from
expansions in parabolic-cylinder coordinates. This analysis
was used by Ghosal and Vervisch to include small but nite
heat release and gas expansion, the triple ame velocity U
TF
may be written [47]:
U
TF
< S
L
(1 1a) 2
b
Z
st
(1 1a)

4n
F
22
_

l
rC
p
x
st
_
(54)
where a = (T
burnt
2T
fresh
)=T
burnt
is dened from the
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 205
Fig. 6. Generic response of the heat released by a one-dimensional strained diffusion ame versus Damkholer number. The dashed line denotes
innitely fast chemistry. Da
p
q
and Da
p
i
are the critical values of Da
p
= (t
c
x
st
)
21
at quenching and ignition, respectively. t
c
is a given chemical
time and x
st
= Du7Zu
2
Z=Z
st
is the scalar dissipation rate under stoichiometric conditions.
temperatures on both sides of a stoichiometric premixed
ame for the same mixture, b = a(T
A
=T
burnt
) is the Zeldo-
vitch number [14], T
A
is the activation temperature, n
F
the stoichiometric coefcient of the fuel and x
st
is measured
far upstream in the mixing layer where the triple ame
propagates. The value of the scalar dissipation rate at the
triple point is of the order of x
st
=(1 1a)
2
[47]. These re-
lations are valid for small values of a and moderate, but
non-zero, values of x
st
. The triple ame velocity given by
Eq. (54) may be combined with LandauSquire solution for
non-reacting laminar round jet to construct a stability
diagram for lift-off and blowout of jet laminar diffusion
ames [54].
A variety of studies suggest that nite rate chemistry and
quenching in non-premixed combustion are somehow
linked to partially premixed combustion [55].
4. A direct analysis: Taylor's expansion
A direct approach to describe turbulent combustion is rst
discussed in this section. This simple formalism, based on
series expansion, illustrates the difculties arising from the
non-linear character of chemical sources.
Consider a simple irreversible reaction between fuel (F)
and oxidizer (O):
F 1sO - (1 1s)P
where the fuel mass reaction rate _ v
F
is expressed from the
Arrhenius law as:
_ v
F
= 2Ar
2
T
b
Y
F
Y
O
exp 2
T
A
T
_ _
(55)
where A is the pre-exponential constant, and T
A
is the acti-
vation temperature.
As the reaction rate is highly non-linear, the averaged
reaction rate

_ v
F
cannot be easily expressed as a function
of the mean mass fractions
~
Y
F
and
~
Y
O
the mean density r
and the mean temperature
~
T: The rst simple idea is to
expand the mean reaction rate

_ v
F
as a Taylor series:
exp 2
T
A
~
T
_ _
= exp 2
T
A
~
T
_ _
1 1

11
n=1
P
n
T
//n
~
T
n
_ _
;
T
b
=
~
T
b
1 1

11
n=1
Q
n
T
//n
~
T
n
_ _
(56)
where P
n
and Q
n
are given by:
P
n
=

n
k=1
(21)
n2k
(n 21)!
(n 2k)![(k 21)!]
2
k
T
A
~
T
_ _
k
;
Q
n
=
b(b 11)(b 1n 21)
n!
(57)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 206
Fig. 7. Schematic of a freely propagation triple ame.
The mean reaction rate,

_ v
F
becomes [56]:

_ v
F
= 2A r
2
~
T
b
~
Y
F
~
Y
O
exp
_
2
T
A
~
T
_

_
1 1

Y
//
F
Y
//
O
~
Y
F
~
Y
O
1(P
1
1Q
1
)
_

Y
//
F
T
//
~
Y
F
~
T
1

Y
//
O
T
//
~
Y
O
~
T
_
1(P
2
1Q
2
1P
1
Q
1
)
_

Y
//
F
T
//
2
~
Y
F
~
T
2
1

Y
//
O
T
//
2
~
Y
O
~
T
2
_
1

_
(58)
Eq. (58) leads to various difculties. First, new quantities
such as

Y
//
k
T
//n
have to be closed using algebraic expressions
or transport equations. Because of non-linearities, large
errors exist when only few terms of the series expansion
are retained. Expression (58) is quite complicated, but is
only valid for a simple irreversible reaction and cannot be
easily extended to realistic chemical schemes (at least 9
species and 19 reactions for hydrogen combustion, several
hundred species and several thousand reactions for hydro-
carbon combustion, etc.). For these reasons, reaction rate
closures in turbulent combustion are not based on Eq. (58).
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 207
Fig. 8. Turbulent premixed combustion regimes as identied by Borghi and Destriau [62]: (a) amelet (thin wrinkled ame), (b) thickened
wrinkled ame regime, and (c) thickened ame regime.
Models are rather derived from physical analysis as discuss
below.
Nevertheless, this approach is used in some simulations
of supersonic reacting ows [57] or to describe reaction in
atmospheric boundary layer where the temperature T may
be roughly assumed to be constant [58]. In these situations,
only the rst two terms in the series expansion are kept.
A segregation factor, a
FO
, is then introduced:
a
FO
= 2

Y
//
F
Y
//
O
~
Y
F
~
Y
O
= 2 1 2

Y
F
Y
O
~
Y
F
~
Y
O
_ _
(59)
to characterize the mixing between the reactants F and O. If
they are perfectly separated

Y
F
Y
O
= 0 and a
FO
= 21: On
the other hand, a perfect mixing (

Y
//
F
Y
//
O
= 0) leads to a
FO
=
0: This segregation factor may be either postulated or
provided by a balance equation (see Ref. [59] in a large
eddy simulation context). Then, the mean reaction rate
becomes:

_ v = 2A(1 1a
FO
) r
2
~
T
b
~
Y
F
~
Y
O
exp 2
T
A
~
T
_ _
(60)
5. Scales and diagrams for turbulent combustion
5.1. Introduction
As the mean burning rate

_ v cannot be found from an
averaging of Arrhenius laws, a physical approach is required
to derive models for turbulent combustion. Turbulent
combustion involves various lengths, velocity and time
scales describing turbulent ow eld and chemical reac-
tions. The physical analysis is mainly based on comparison
between these scales.
The turbulent ow is characterized by a Reynolds number
comparing turbulent transport to viscous forces:
Re =
u
/
l
t
v
(61)
where u
/
is the velocity rms (related to the square root of the
turbulent kinetic energy k), l
t
is the turbulence integral
length scale and n the kinematic viscosity of the ow.
The Damkohler number compares the turbulent (t
t
) and
the chemical (t
c
) time scales:
Da =
t
t
t
c
(62)
In the limit of high Damkohler numbers (Da q1); the
chemical time is short compared to the turbulent one, corre-
sponding to a thin reaction zone distorted and convected by
the ow eld. The internal structure of the ame is not
strongly affected by turbulence and may be described as a
laminar ame element called a `amelet'. The turbulent
structures wrinkle and strain the ame surface. On the
other hand, a low Damkohler number (Da p1) corresponds
to a slow chemical reaction. Reactants and products are
mixed by turbulent structures before reaction. In this
perfectly stirred reactor limit, the mean reaction rate may
be expressed from Arrhenius laws using mean mass frac-
tions and temperature, corresponding to the rst term of the
Taylor's expansion (58).
In turbulent ames, as long as quenching does not occur,
most practical situations correspond to high or medium
values of the Damkohler numbers. It is worth noting that
various chemical time scales may be encountered: fuel
oxidation generally corresponds to short chemical time
scales (Da q1) whereas pollutant production or destruc-
tion such as CO oxidation or NO formation are slower.
5.2. Turbulent premixed combustion diagram
5.2.1. Introduction
The objective is to analyze premixed turbulent combus-
tion regimes by comparing turbulence and chemical char-
acteristic length and time scales. This analysis leads to
combustion diagrams where various regimes are presented
as function of various dimensionless numbers [14,23,29,60
62]. These diagrams could be a support to the selection and
development of the relevant combustion model for a given
situation. A formalism combining recent analysis [29,62] is
retained here.
For turbulent premixed ames, the chemical time scale,
t
c
, may be estimated as the ratio of the thickness d
l
and the
propagation speed S
L
of the laminar ame.
1
Estimating the
turbulent time from turbulent integral scale characteristics
(t
t
= l
t
=u
/
); the Damkohler number becomes:
Da =
t
t
t
c
=
l
t
d
l
S
L
u
/
(63)
where a velocity ratio (u
/
/S
L
) and a length scale ratio (l
t
/d
l
)
are evidenced.
5.2.2. Combustion regimes
For large values of the Damkohler number (Da q1); the
ame front is thin and its inner structure is not affected by
turbulence motions which only wrinkle the ame surface.
This amelet regime or thin wrinkled ame regime (Fig. 8a)
occurs when the smallest turbulence scales (i.e. the Kolmo-
gorov scales), have a turbulent time t
k
larger than t
c
(turbulent motions are too slow to affect the ame structure).
This transition is described in term of the Karlovitz
number Ka:
Ka =
t
c
t
k
=
d
l
l
k
u
k
S
L
(64)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 208
1
This chemical time t
c
corresponds to the time required for the
ame to propagate over a distance equal to its own thickness. This
time may also be viewed as a diffusive time scale, using Eq. (34)
t
c
= d
l
=S
L
= (1=Re
f
)(d
2
l
=n):
The size l
k
and the velocity u
k
of Kolmogorov structures
are given by [64]:
l
k
=
v
3
1
_ _
1=4
; u
k
= (v1)
1=4
(65)
where 1 is the dissipation rate of the turbulent kinetic energy
k. The integral length scale lt may be written:
l
t
=
u
/3
1
_ _
(66)
using v = d
l
S
L
; corresponding to a unity ame Reynolds
number Re
f
(Eq. (34)), yields
Ka =
u
/
S
L
_ _
3=2
l
t
d
l
_ _
21=2
(67)
Reynolds, Re, Damkohler, Da, and Karlovitz, Ka, numbers
are related as:
Re = Da
2
Ka
2
(68)
and a set of two parameters (Re, Da), (Re, Ka) or (Da, Ka) are
necessary to discuss regimes in the case of premixed reactants.
The Karlovitz number also compares the ame and the
Kolmogorov length scales according to:
Ka =
d
l
l
k
_ _
2
(69)
The Karlovitz number is used to dene the Klimov
Williams criterion, corresponding to Ka = 1; delineating
between two combustion regimes. This criterion was rst
interpreted as the transition between the amelet regime
(Ka , 1); previously described, and the distributed combus-
tion regime where the ame inner structure is strongly modi-
ed by turbulence motions. A recent analysis [29] has
shown that, for Karlovitz numbers larger than unity (Ka .
1); turbulent motions become able to affect the ame inner
structure but not necessarily the reaction zone. This reaction
zone, where heat is released, has a thickness d
r
much lower
that the thermal thickness d
l
of the ame (d
r
< 0:1d
l
): The
Karlovitz number based on this reaction thickness is:
Ka
r
=
d
r
l
k
_ _
2
=
d
r
d
l
_ _
2
d
l
l
k
_ _
2
<
1
100
d
l
l
k
_ _
2
<
Ka
100
(70)
Then, the following turbulent premixed ame regimes are
proposed [29]:
Ka ,1: Flamelet regime or thin wrinkled ame regime
(Fig. 8a). Two subdivisions may be proposed depending
on the velocity ratio u
/
/S
L
:
(u
/
/S
L
) ,1: wrinkled ame. As u
/
may be viewed as
the rotation speed of the larger turbulent motions,
turbulent structures are unable to wrinkle the ame
surface up to ame front interactions. The laminar
propagation is predominant and turbulence/combus-
tion interactions remain limited.
(u
/
/S
L
) .1: wrinkled ame with pockets (`corrugated
ames'). In this situation, larger structures become able
to induce ame front interactions leading to pockets.
1 , Ka # 100 (Ka
r
, 1) : Thickened wrinkled ame
regime or thin reaction zone. In this case, turbulent
motions are able to affect and to thicken the ame preheat
zone, but cannot modify the reaction zone which remains
thin and close to a laminar reaction zone (Fig. 8b).
Ka .100 (Ka
r
.1): Thickened ame regime or well-
stirred reactor. In this situation, preheat and reaction
zones are strongly affected by turbulent motions and no
laminar ame structure may be identied (Fig. 8c).
These various regimes are generally displayed on a log-
arithmic diagram (u
/
/S
L
; l
t
/d
l
), similar to the one presented
on Fig. 9.
5.2.3. Comments
This analysis, leading to a rough classication of combus-
tion regimes as a function of characteristic numbers, has
been developed as a support to derive and choose turbulent
combustion models. Following this classication, most
practical applications correspond to amelet or thickened
wrinkled ame regimes. Nevertheless, such analyses are
only qualitative and should be used with great care. A
diagram such as the one displayed on Fig. 9 cannot be
readily used to determine the combustion regime of a prac-
tical system from (u
/
/S
L
) and (d
l
/l
t
) ratios:
The analysis is based on the assumption of a homo-
geneous and isotropic turbulence unaffected by heat
release, which is not the case in combustion systems.
Some quantities used are not clearly dened. For
example, the ame thickness d
l
may be based on the
thermal thickness or on the diffusive thickness. Accord-
ingly, the limits between the various regimes may notice-
ably change.
All regime limits are based on order of magnitude esti-
mations and not on precise derivations. For example, the
amelet regime limit could correspond to a Karlovitz
number Ka = 0:1 or Ka = 10; rather than Ka = 1:
Various effects are not taken into account. Unsteady and
curvature effects play an important role neglected here.
Turbulent premixed combustion diagrams were analyzed
using direct numerical simulations of ame/vortex inter-
actions [65]. Results show that the amelet regime seems
to extend above the KlimovWilliams criterion (see Fig.
9). DNS has revealed that small turbulent scales, which
are supposed in classical theories to have the strongest
effects on ames, have small lifetimes because of viscous
dissipation and therefore only limited effects on combus-
tion, results recovered experimentally [66]. Peters [29]
shows that the criterion Ka = 100 (i.e. Ka
r
= 1) is in
quite good agreement with the transition proposed in
Ref. [65], at least when the length scale ratio, l
t
/d
l
, is
sufciently large.
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 209
Additional length scales have been introduced in the
literature. For instance the Gibson scale l
G
, to character-
ize the size of the smaller vortex able to affect the ame
front was used [63]. This length was dened as the size of
the vortex having the same velocity as the laminar ame
speed S
L
.
All these analyses are implicitly based on a single step
irreversible reaction. In actual turbulent combustion, a
large number of chemical species and reactions are
involved (several hundred species and several thousand
reactions for propane burning in air). These reactions
may correspond to a large range of chemical time scales.
For example, the propane oxidation may assumed to be
fast compared to turbulent time scale. On the other hand,
the CO
2
formation from carbon monoxide (CO) and OH
radical in the burnt gases is much slower with a chemical
time of the same order as turbulent times.
5.3. Non-premixed turbulent combustion diagram
5.3.1. Introduction
Two numbers, a length ratio and a velocity ratio, have
been used to identify premixed turbulent combustion
regimes. The problem is more difcult in non-premixed
turbulent combustion because diffusion ames do not propa-
gate and, therefore, exhibit no intrinsic characteristic speed.
In addition, the thickness of the ame depends on the aero-
dynamics controlling the thickness of the local mixing
layers developing between fuel and oxidizer (Section 3.2)
and no xed reference length scale can be easily identied
for diffusion ames. This difculty is well illustrated in the
literature, where various characteristic scales have been
retained depending on the authors [33,6771]. These
classications of non-premixed turbulent ames may be
organized in three major groups:
The turbulent ow regime is characterized by a Reynolds
number, whereas a Damkohler number is chosen for the
reaction zone [72].
The mixture fraction eld is retained to describe the
turbulent mixing using

Z
//2
and a Damkohler number
(ratio of Kolmogorov to chemical time) characterizes
the ame [33].
A velocity ratio (turbulence intensity to premixed lami-
nar ame speed) and a length ratio (integral scale to
premixed laminar ame thickness) may be constructed
[68] to delineate between regimes.
Additional lengths have also been introduced, using for
instance thicknesses of proles in mixture fraction space
[67].
A laminar diffusion ame is fully determined from a
Damkohler number Da
p
= (t
c
x
st
)
21
; where the value of
the chemical time t
c
depends on the fuel chemistry [30]
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 210
Fig. 9. Turbulent premixed combustion diagram [29,62]. Combustion regimes are identied using the length scale (l
t
=d
l
) and the velocity
(u
/
=S
L
) ratios. The KlimovWilliams criterion (Ka = 1) corresponds to a ame thickness d
l
equal to the Kolmogorov scale l
k
. Below this line,
the ame is thinner than any turbulent scale. Below the line delineating the Peters criterion (Ka = 100 or Ka
r
= 1), the reaction zone thickness,
d
r
, is thinner than any turbulent scale and is not affected by turbulent motions (the criterion is plotted assuming d
r
< 0:1d
l
). The amelet regime
limit devised by Poinsot et al. [65] from direct numerical simulations is also displayed. The criterion proposed in Ref. [140] to delineate
between gradient (above) and counter-gradient (below) turbulent transport is displayed assuming a heat release factor t = T
b
=T
u
21 = 6 where
T
u
and T
b
are, respectively, the fresh and the burnt gases temperature (see Section 8).
(Section 3.2). In this number, the scalar dissipation rate
under stoichiometric condition (Z = Z
st
); x
st
= Du7Zu
2
st
;
measures at the same time a mechanical time, t
x
= x
21
st
;
and, a characteristic mixing length, l
d
= (D=x
st
)
1=2
: Accord-
ing to asymptotic developments [30], the reaction zone
thickness is of the order of l
r
< l
d
(Da
p
)
21=(a11)
; where a is
the order of a global one-step reaction.
Because diffusion ames do not feature a xed reference
length, a main difculty arises when effects of unsteadiness
need to be quantied. In a steady laminar ame the local rate
of strain is directly related to x
st
(and to a ame thickness),
however, when the velocity eld uctuates, unsteadiness in
diffusion ames develops at two levels [73]:
The mixture fraction eld Z does not immediately
respond to velocity uctuations, leading to a distribution
of x
st
for given rates of strain. Because a strong correla-
tion exists between x
st
and velocity gradients taken along
the stoichiometric line [74], this effect is not the dominant
one when nite rate chemistry occurs.
For nite rate chemistry, the burning rate does not im-
mediately follow variations of x
st
, leading to a second
level of unsteadiness, modifying the burning rate (Eq.
(71)).
u
/
- [Unsteadiness in mixing] - x
s
- [Burning rate unsteadiness(for Da
p
, 1)] - _ v
i
(71)
Summarizing these effects in a generic diagram is an
arduous task.
A diagram for laminar ames submitted to curvature
associated to a time varying strain rate was obtained by
Cuenot and Poinsot from DNS results of ame/vortex inter-
action [70]. In this diagram presented on Fig. 10, the ame
thickness is d
i
<l
d
, whereas r and u
/
denote respectively the
characteristic size and velocity of the vortex pair. This
analysis identies two limiting Damkohler numbers,
Da
LFA
and Da
ext
. When Da
p
is larger than Da
LFA
, the ame
front may be viewed as a steady laminar ame element and
its inner structure is not affected by vortices. On the other
hand, when Da
p
# Da
ext
; ame extinction occurs. In the
intermediate Damkohler number range (i.e. Da
ext
,Da
p
,
Da
LFA
), strong unsteadiness effects are observed.
In a non-premixed turbulent ame, the reaction zones
develop within a mean mixing zone whose thickness l
z
is
of the order of the turbulent integral length scale l
t
(Fig. 11):
l
z
< u7
~
Zu
21
< l
t
<
k
3=2
1
_ _
(72)
Turbulent small scale mixing mainly depends both on
velocity uctuations, transporting the iso-Z surfaces (stir-
ring), and diffusion between these iso-surfaces that compose
the mixing layer of thickness l
d
, with
l
d
<
D
~ x
st
_ _
1=2
(73)
where ~ x
st
denotes the conditional mean value of the scalar
dissipation rate x for Z = Z
st
:
When transport of species and heat by velocity uctua-
tions is faster than transfer in the diffusion ame, a departure
from laminar amelet is expected. Also, when the Kolmo-
gorov scale l
k
is of the order of the ame thickness, the inner
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 211
Fig. 10. Non-premixed ame/vortex interaction regimes by Cuenot and Poinsot [70]. This diagram delineates the steady laminar amelet
assumption (LFA) validity regions, the quenching limits and the zone where unsteady and curvature effects are important during ame vortex
interaction. u
/
is the level of velocity uctuations, d
i
= l
d
is the ame thickness (l
d
< u7Zu
21
), t is a chemical time and r the characteristic size
of the vortices.
structure of the reaction zone may be modied by the turbu-
lence. As diffusion ame scales strongly depend on the local
ow motions, one may write:
l
d
< a
1
l
k
and ~ x
st
<
a
2
t
k
(74)
where a
1
$ 1 and a
2
# 1 (the maximum local strain rate
would correspond to l
d
= l
k
).
Then using t
t
=t
k
=

Re
_
; the Damkohler number com-
paring turbulent ame scale and chemical ame scale is
recast as:
Da =
t
t
t
c
=
t
t
t
k
t
k
t
c
<
t
t
t
k
a
2
~ x
st
t
c
< a
2

Re Da
p
_
(75)
Constant Damkohler numbers Da
p
correspond to lines of
slope 1/2 in a loglog (Da, Re) diagram. When the chemistry
is sufciently fast (large Da values), the ame is expected to
have a laminar ame structure. This condition may be
simply expressed as Da
p
$ Da
LFA
On the other hand, for
large chemical times (i.e. when Da
p
# Da
ext
), extinction
occurs. Laminar ames are encountered for low Reynolds
numbers (Re , 1): Results are summarized in Fig. 12.
In a practical combustion devices, a
1
and a
2
would
evolve in space and time according to ow uctuations,
velocity and scalar energetic spectra. In a given burner, it
is likely that one may observe at different locations amelet
behavior and strong unsteadiness, or even quenching.
As the classication of premixed turbulent ames, these
considerations are limited by the numerous hypothesis
necessary to derive the regimes.
6. Tools for turbulent combustion modeling
6.1. Introduction
The mean heat release rate is one of the main quantities of
practical interest that should be approximated by turbulent
combustion models. The simplest and most direct approach
is to develop the chemical rate in Taylor series as a function
of species mass fractions and temperature (Eq. (58)). This
analysis is limited by its low accuracy and by the rapidly
growing complexity of the chemistry (Section 4). It is then
concluded that the non-linear character of the problem
requires the introduction of new tools. These new tools
must be designed to describe turbulent ames and have to
provide an estimation of mean production or consumption
rates of chemical species. They also need to be based on
known quantities (mean ow characteristics, for example)
or on quantities that may be easily modeled or obtained from
closed balance equations. In this section, a generic descrip-
tion of the main concepts used to model turbulent combus-
tion is proposed. Relations between the various approaches
are also emphasized, but the discussion of the closure
strategy is postponed to subsequent sections.
The basic ingredients to describe turbulent ames remain
the quantities introduced for laminar ame analysis: the
progress variable c for premixed combustion (c = 0 in
fresh gases and c = 1 in burnt gases, see Section 3.2),
and, the mixture fraction Z for non-premixed ames (Z is
a passive scalar, with Z = 0 in pure oxidizer and Z = 1 in
pure fuel, see Section 3.2). The ame position would corre-
spond to values of the progress variable c lying between 0
and 1, or, to Z taking on values in the vicinity of Z = Z
st
:
Three main types of approaches are summarized in
Fig. 13:
The burning rate may be quantied in terms of turbulent
mixing. When the Damkohler number Da = t
t
=t
c
; com-
paring turbulent (t
t
) and chemical (t
c
) characteristic times,
is large (a common assumption in combustion modeling),
the reaction rate is limited by turbulent mixing, described in
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 212
Fig. 12. Schematic of non-premixed turbulent combustion regimes
as function of the Damkohler number Da = t
t
=t
c
(constructed from
the turbulent integral time scale t
t
and chemical time t
c
) and Re the
turbulent Reynolds number.
Fig. 11. Sketch of a non-premixed turbulent ame. Z is the mixture
fraction, l
d
the diffusive thickness, l
r
the reaction zone thickness, l
t
the turbulence integral length scale.
terms of scalar dissipation rates [75]. The small scale dissi-
pation rate of species controls the mixing of the reactants
and, accordingly, plays a dominant role in combustion
modeling, even for nite rate chemistry.
In the geometrical analysis, the ame is described as a
geometrical surface, this approach is usually linked to a
amelet assumption (the ame is thin compared to all
ow scales). Following this view, scalar elds (c or Z) are
studied in terms of dynamics and physical properties of iso-
value surfaces dened as ame surfaces (iso-c
p
or iso-Z
st
).
The ame is then envisioned as an interface between fuel
and oxidizer (non-premixed) or between fresh and burnt
gases (premixed). A ame normal analysis is derived by
focusing the attention on the structure of the reacting ow
along the normal to the ame surface. This leads to amelet
modeling when this structure is compared to one-dimen-
sional laminar ames. The density of ame surface area
per unit volume is also useful to estimate the burning rate.
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 213
Fig. 13. Three types of analysis of premixed or non-premixed turbulent ames.
The statistical properties of scalar elds may be collected
and analyzed for any location within the ow. Mean values
and correlations are then extracted via the knowledge of
one-point probability density functions (pdf). The determi-
nation of these pdfs leads to pdf modeling. A one-point
statistical analysis restricted to a particular value of the
scalar eld is related to the study of conditional statistics.
Conditional statistics which are obviously linked to the
geometrical analysis and to ame surfaces when the con-
ditioning value is c
p
or Z
st
.
6.2. Scalar dissipation rate
In a rst step, the transport equation for

c
//2
or

Z
//2
is
derived, these uctuations characterize non-homogeneities
and intermittencies. In the case of the progress variable, the
variance

c
//2
is dened as:
r

c
//2
= r (c 2 ~ c)
2
= r(

c
//2
2 ~ c
2
) = rc
2
2 r ~ c
2
(76)
Starting from the balance equation for the progress vari-
able (Eq. (36)), c is decomposed into c = ~ c 1c
//
; then the
new equation is multiplied by c
//
and averaged. After
straightforward manipulations, the exact transport equation
for

c
//2
reads:
2 r

c
//2
2t
17( r ~ u

c
// 2
) 17( r

u
//
c
// 2
) = (77)
7(rD7c
// 2
) 12c
//
7(rD7~ c)
22 r

u
//
c
//
7~ c
.....,,.....
Production
22rD7c
//
7c
//
.....,,.....
Dissipation
1 2 _ vc
//
..,,..
Source
In addition to the two diffusive terms 7(rD7c
// 2
) and
2c
//
7(rD7~ c); which are non-zero, but expected small for
large Reynolds number ows (especially the second one),
two important terms are found: The uctuating part of the
scalar dissipation rate 2rD7c
//
7c
//
and a correlation _ vc
//
involving the chemical source.
In the literature, various expressions have been associated
to the terminology scalar dissipation rate (in laminar ame
theory, it actually quanties a diffusion speed, Section 3.2).
It may include the density r, a factor 2 and be written in
term of instantaneous (c) or uctuating (c
//
) values of the
concentration species. Hereafter:
r ~ x = rD7c7c = rD7~ c7~ c 12rD7c
//
7~ c 1rD7c
//
7c
//
leading to, when mean gradients are neglected:
r ~ x < rD7c
//
7c
//
(78)
Then, r ~ x is the dissipation rate of the uctuations of the
scalar eld.
In the simplied case of homogeneous ames (no ~ c or
~
Z
gradient), the time evolution of the scalar variances are
governed by:
Premixed combustion:
2 r

c
// 2
dt
= 22rD7c
//
7c
//
12 _ vc
//
Non-premixed combustion:
2 r

Z
// 2
dt
= 22rD7Z
//
7Z
//
These equations have important implications:
The scalar dissipation rate directly measures the decaying
speed of uctuations via turbulent micromixing. Since the
burning rate depends on the contact between the reactants,
in any models, the scalar dissipation rate enters directly or
indirectly the expression for the mean burning rate. For
instance, when assuming very fast chemistry and a combus-
tion limited by mixing, the mean burning rate is proportional
to the scalar dissipation rate of Z or c.
Within a premixed system, turbulent mixing occurs
between fresh and burnt gases. One may then expect a
very strong coupling between mixing phenomena and
chemical reaction. This is observed in the equation for

c
// 2
where, at the same time, ~ x and the chemical source _ vc
//
are
involved.
In a non-premixed ame, fresh fuel and fresh oxidizer
have to be mixed at the molecular level for reacting
and the ame is mainly controlled by turbulent mixing
occurring between the fresh gases. In consequence, there
is no chemical source acting on the evolution of

Z
//2
:
The mixture fraction Z is sensitive to chemistry only via
density change, making the coupling between chemistry
and mixing different than in the case of premixed
combustion.
This preliminary analysis shows that dissipation rate of
scalars is a very key concept of turbulent combustion and,
directly or indirectly, x appears in any tools used to model
ames. The main stumbling block in turbulent combustion
modeling and bridges between the various modeling
concepts emerge through the scalar dissipation rate.
6.3. Geometrical description
The ame front is here described as a geometrical entity.
This analysis is generally linked to the assumption of a
sufciently thin ame, viewed as an interface between
fresh and burnt gases in premixed combustion or as an inter-
face between fuel and oxidizer in non-premixed situations.
Two formalisms have been proposed: eld equation or ame
surface density concept.
6.3.1. G-eld equation
The balance equation for the progress variable may
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 214
be written:
2c
2t
1u7c =
1
r
(7(rD7c) 1 _ v) = wu7cu
or
2c
2t
1u
c
7c = 0
(79)
where u
c
is the absolute propagation velocity of the progress
variable eld. This velocity is easily decomposed into the
ow velocity u and w the relative velocity of the iso-c
surface measured with respect to the ow, leading to u
c
=
u 1wn; where n = 2(7c=u7cu) is the local normal vector to
the iso-c surface, pointing toward the fresh gases. The rela-
tive displacement speed w is given by Eq. (37).
In laminar ames, a G-eld whose level G = G
0
repre-
sents the ame surface, was introduced to simulate the
propagation of fronts [76]. The `G-equation' may be written:
2G
2t
1u
c
7G = 0 or
2G
2t
1u7G = wu7Gu (80)
This kinematic description of premixed combustion
possesses some attractive aspects:
The internal ame front structure does not need to be
resolved on the computational mesh. Only the G-eld,
generally much thicker than the ame front, needs to be
resolved. Usually, G(x) is dened as the distance of the
given location x to the ame front [77].
Under the assumption of constant density (thermodiffu-
sive assumption), the G-equation may be used for low
cost direct numerical simulations. Each G iso-surface is
then related to a ame front, and one may argue that a
single simulation corresponds to the computation of
several ames.
In non-constant density ows where thermal heat release
is included, the displacement speed w of the G-iso-surface is
affected by thermal expansion and should be corrected for
density variations, even in the case of a steady laminar ame
propagating at the constant laminar ame speed S
L
, dened
relatively to the unburnt gases. This correction is:
w =
r
u
r
w
u
(81)
where w
u
is the ame displacement speed relative to the
unburnt gases of density r
u
.
A more difcult point is the coupling required between
the G-equation and the species or energy balance equations.
The G-equation provides a kinematic description of the
ame front and involves its displacement speed w. The
reactant consumption and the heat release rate are controlled
by the consumption speed S
c
:
r
u
S
c
=
_1
21
_ vdj (82)
where r
u
is the fresh gases density and j the spatial coordi-
nate along the normal direction to the ame front. Of course,
w and S
c
are related but may be quite different, especially in
high ame front curvature zones [78]. The displacement
speed w may also be quite different from the laminar
ame speed S
L
.
The coupling between the consumption speed S
c
and the
displacement speed w is a very key point in G-eld model-
ing. Three approaches have been proposed to overcome this
difculty:
Flame front tracking technique: The displacement of the
ame front is evaluated from the displacement speed w,
leading to an estimation of the volume of burnt gases
produced along with the thermal heat release [79,80].
Based on a purely geometric approach, this technique is
well suited to two dimensional simulations, but its exten-
sion to 3D cases may not be straightforward.
Temperature (or energy) reconstruction: The temperature
eld is directly estimated from the G-eld as [81]:
T = T
u
1
Q
C
p
H(G 2G
p
) (83)
where T
u
is the temperature of unburnt gases, Q the reac-
tion heat release and H denotes the Heaviside function
(the Heaviside function is smoothed on the mesh of the
simulation). This approach does not require a balance
equation for the energy but is not applicable when heat
losses or compressibility effects (for example in an in-
ternal combustion engines) occur.
Estimation of the heat release rate from the G-eld: The
G-eld is used to estimate the heat release rate to be
incorporated in the balance energy equation from a
formulation similar to Eq. (83) [82]. Accordingly, any
other effects (heat losses, compressibility) may be
included.
This formalism may be extended to turbulent ames.
Averaging the progress variable Eq. (79) leads to:
2~ c
2t
1 ~ u7~ c =
ru
r
27( r

u
//
c
//
) 1rwu7cu
u7~ cu

_
............,,............
S
T
u7~ cu (84)
where the turbulent ame speed, S
T
, has to be modeled. The
G-equation becomes:
2
~
G
2t
1 ~ u7
~
G =
ru
r
S
T
u7
~
Gu
The mean turbulent ame brush is then located at the
points where
~
G = G
0
: The G-equation does not required
thin ame elements per se. The overall turbulent ame is
only viewed as a propagating surface without solving for the
internal ame structure. This formalism is therefore a good
candidate for the numerical simulation of large systems
where the knowledge of the internal structure of the ame
brush is not required [79]. Nevertheless, a model has to be
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 215
supplied for S
T
, when doing so, turbulent transport and
propagation may be separated to carefully model the effect
of turbulence on w, the instantaneous displacement speed of
the premixed front [29,33].
6.3.2. Flame surface density description
The ame is identied as a surface and the ame surface
density S is introduced, S measures the available ame area
d A per unit volume d V. The mean burning rate of a species
i is then modeled as:

_ v
i
=
_
V
i
S (85)
where V

i
is the mean local burning rate per unit of ame area
integrated along the normal direction to the ame surface.
V

i
is related to the properties of the local ame front and is
generally estimated from a prototype laminar ame, incor-
porating more or less complexity. For instance, one may
consider a planar laminar ame, submitted or not to a steady
strain, a laminar ame where curvature effects have been
introduced, or even a laminar unsteady strained and curved
ame. The main advantage of this formulation, summarized
on Fig. 14, is to decouple the chemical description (V

i
) from
the ame/turbulence interaction (S). The ame surface is
convected, diffused, curved and strained by the velocity
eld [17,83].
The ame surface density S may be estimated either from
algebraic relations (see Section 7.4.2) or as a solution of a
balance equation. Using a phenomenological analysis, this
balance equation was rst proposed by Marble and
Broadwell [84] for non-premixed turbulent combustion.
More rigorous derivations were obtained from geo-
metrical considerations [17,85] and from a statistical
description [8688] leading to an exact, but unclosed,
balance equation. The derivation using statistical tools is
now briey summarized.
In premixed combustion, the ame surface density S of
the iso-c
p
surface is estimated from the conditional gradient
of the progress variable c [86]:
S(c
p
) = u7cud(c 2c
p
) = u7cuuc = c
p
_ _

P(c
p
) (86)
where d(c 2c
p
) is a local measure of the probability (see
Section 6.4.3),
_
u7cuuc = c
p
_
is the conditional average of
u7cu for c = c
p
and

P(c
p
) is the probability to nd c = c
p
at
the given location. From this denition and the balance
equation for the progress variable c, an exact equation for
the ame surface density S may be derived according to the
following steps [87]:
1. Derivation of an equation for u7cu from the equation for
the instantaneous progress variable c.
2. Derivation of an equation for u7cu d(c 2c)
p
by condi-
tioning the previous u7cu balance equation.
3. Averaging the u7cu d (c 2c)
p
balance equation leading
to an exact equation for the ame surface density S =
u7cud(c 2c
p
):
This derivation is valid for any iso-scalar surface (c
p
can
take any values between zero and unity) and S(c
p
; x; t) is also
called a surface density function, the derivation of its balance
equation is quite tedious [87] and similar to the derivation of
a balance equation for the probability density function
(Section 6.4.3), details are not given here. Two equivalent
forms of the progress variable equation may be used:
A classical reaction/diffusion formulation:
2c
2t
1u7c =
1
r
[7(rD7c) 1 _ v] (87)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 216
Fig. 14. Flame surface density modeling.
A propagative form:
2c
2t
1u7c = wu7cu (88)
where the displacement speed w of the ame front rela-
tively to the ow was introduced (Eq. (37)). This equa-
tion corresponds to the G-equation [77] (Section 6.3.1).
Two different types of equation for S are associated to
these two formulations of the progress variable balance
equation.
A reaction/diffusion formulation:
2S
2t
17(kul
s
S) = k7u 2nn : 7ul
s
S
2
_
1
u7cu
2
2n
_
1
r
[7(rD7c) 1 _ v]
__
s
S
2
2
2c
p
__
1
r
[7(rD7c) 1 _ v]
_
s
S
_
(89)
A propagative formulation:
2S
2t
17(kul
s
S)
= k7u 2nn : 7ul
s
S 27 kwnl
s
S
_ _
1kw7nl
s
S
(90)
where n is the unit vector normal to the c = c
p
surface and
pointing towards the fresh gases (n = 27c=u7cu): 7n corre-
sponds to the ame front curvature. 2=2n = n7 is a deriva-
tive normal to the ame front. 2/2c
p
is a derivative in the
sample space c
p
. The surface average of Q, kQl
s
, is dened
as:
kQl
s
=
Qu7cud(c 2c
p
)
u7cud(c 2c
p
)
=
_
Qu7cu c } = c
p
_
_
u7cu c } = c
p
_ (91)
Here the notation kQl
s
implicitly indicates that the mean
is taken for c = c
p
; and kQl
s
is a function of c
p
.
The propagative formulation is often written in term of
total stretch A
c
p
of the iso-c
p
surface
1
S
2S
2t
17(ku 1wnl
s
S)
_ _
= A
c
p
= k 7u 2nn : 7u
......,,......
tangential strain rate
1 w7n
,....,
Curvature
l
s
(92)
The LHS terms in the two balance Eqs. (89) and (90) corre-
spond to unsteady effects and to the ame surface convec-
tion. The rst term in the RHS expresses the action of the
tangential strain rate on the ame surface. The last two terms
in Eq. (89) describe respectively reaction/diffusion effects
along the ame normal direction and uxes in the sample
space c
p
. The last two terms in the RHS of the propagative
Eq. (90) correspond to front convection due to a normal
propagation and combined propagation/curvature effects.
These two formulations of the S balance equation induce
the following comments:
The two balance Eqs. (89) and (90) are mathematically
strictly equivalent but the problem is not expressed in the
same way. In the propagative form (Eq. (90)), many
effects are incorporated in the ame front propagation
speed, w, that may differ from the laminar ame speed
S
L
[78]. On the other hand, a ux term in the sample space
is found in the reaction/diffusion derivation. In Eq. (90)
the imbalance between diffusion and reaction is cast in
the form of the propagation velocity w, Eq. (89) recalls
that transfer phenomena between iso-surfaces are
involved in this propagation.
The derivation of the balance Eqs. (89) and (90) impli-
citly forgets some mathematical singularities that may
become important in particular situations. For example,
the normal vector n is assumed to be well dened and
having a nite derivative. This may not be the case when
two close ame fronts interacts.
In combustion modeling, a single iso-c
p
surface is
assumed to correspond to the ame front. This is a priori
true when the ame is innitely thin, assuming that the
local reaction rate per unit ame area, V

i
, describes
whether the ame is actually burning or not. However
in real turbulent ames, the local burning zone is not
innitely thin and the ame front, identied as the
location of the maximum reaction, may differ from the
c
p
iso-surface. This may be for instance the case when
analyzing data of direct numerical simulations where the
ame front has to be resolved on the computational grid.
Using the Favre decomposition (u = ~ u 1u
//
); the
convection ux term may be decomposed into mean
and turbulent components:
kul
s
S = ~ uS 1ku
//
l
s
S (93)
The strain rate term is also split into a contribution due to
the mean ow and a contribution due to turbulent velo-
city uctuations:
k7u 2nn : 7ul
s
S = (7 ~ u 2knnl
s
: 7~ ul
s
........,,........
S
A
T
1k7u
//
2nn : 7~ u)S
........,,........
a
T
(94)
All these denitions may be extended to ame fronts which
are not innitely thin. Integrating Eq. (86) across iso-surface
levels leads to:
_1
0
S(c
p
)dc
p
=
_1
0
_
u7cuuc = c
p
_

P(c
p
)dc
p
= u7cu (95)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 217
that is a direct estimation of the inverse of the mean local
ame thickness d (d = (1=u7cu)
21
) and may be viewed as a
`generalized ame surface density', u7cu follows a balance
equation similar to the S(c
p
) balance Eq. (90), where
surface averages kQl
s
are replaced by `generalized surface
averages' kQl
s
which do not depend on c
p
:
kQl
s
=
1
u7cu
_1
0
kQl
s
S(c
p
)dc
p
(96)
In ame surface density models, under a amelet assump-
tion, for any values of c
p
, the ame surface is assumed to
behave such as S(c
p
) < u7cu = S:
All balance equations remain formally the same but have
to be closed. The same modeling issues emerge using the G-
equation formalism [89]. One needs to develop closures for
the turbulent ux of ame surface, for the propagation vel-
ocity of the surface, as well as for the effects of curvature
and strain rate.
6.3.3. Flame wrinkling description
The previous formalism may be recast in term of ame
surface wrinkling. The basic idea is to introduce the ratio
J(c
p
) of the ame surface to its projection in the direction of
propagation n
p
:
J(c
p
) =
S(c
p
)
n
p
(nu7cud(c 2c
p
))
=
_
u7cuuc = c
p
_

P(c
p
)
n
p

_
nu7cud(c 2c
p
)
_
=
_
u7cuuc = c
p
_
n
p

_
nu7cuuc = c
p
_ =
1
n
p
knl
s
(97)
where n and n
p
are the unit vectors normal to the instan-
taneous ame front and to the mean propagating direction
respectively. These vectors are chosen pointing towards the
fresh gases and are given by:
n = 2
7c
u7cu
; n
p
= 2
_
7cuc = c
p
_
_
7cuc = c
p
_ _
_
_
_
_
_
(98)
Then:
J(c
p
) =
_
u7cuuc = c
p
_
_
7cuc = c
p
_ _
_
_
_
_
_
=
S(c
p
)
_
7cuc = c
p
_ _
_
_
_
_
_

P(c
p
)
(99)
As in Section 6.3.2, a generalized ame surface wrinkling,

J; is introduced:

J =
_1
0
S(c
p
)dc
p
_1
0
_
7cuc = c
p
_ _
_
_
_
_
_

P(c
p
)dc
p
=
u7cu
u7 cu
=

S
u7 cu
(100)
For an innitely thin ame front, J(c
p
) =

J for any c
p
value.
Flame surface density and ame wrinkling factor are
closely related. Balance equations may also be derived
and closed for J(c
p
) or

J: These equations are considerably
more complicated than ame surface density balance equa-
tions, but the wrinkling factor may be more convenient for
initial and/or boundary conditions (

J $ 1 everywhere).
This approach has been explored by Weller et al. [90,91].
6.4. Statistical approaches: probability density function
6.4.1. Introduction
Predictions of radicals and intermediate species such as
OH, or pollutants like CO, require the description of the
ame front internal structure, for intermediate states
between fresh and burnt gases in premixed ames or
between fuel and oxidizer in non-premixed ames. Even
though G-eld and density of ame surface S need some
statistical treatments, they are initially based on a geo-
metrical view describing the ame as a thin interface. In
probability density function methods, one wishes to relax
this assumption by focussing on the statistical properties of
intermediate states within the ame front.
The probability density function (pdf)

P(Y
p
; x; t) quanti-
es the probability to nd, for a given location x and a time t,
a variable Y (mass fraction, temperature, velocity, etc.)
within the range [Y
p
2DY=2; Y
p
1DY=2]: This probability
is equal to

P(Y
p
; x; t)DY [9296]. The pdf satises the
following simple relations:
_
Y

P(Y
p
; x; t)dY
p
= 1
_
Y

P(Y
p
; x; t)dY
p
=

Y(x; t)
_
Y
(Y
p
2

Y)
2

P(Y
p
; x; t)dY
p
= Y
/2
(x; t)
where Y
p
is the sample space variable corresponding to the
random variable Y. When more than one variable is required
to capture the ame structure, a joint probability density
function

P(Y
p
1
; ; Y
p
N
; x; t) is introduced [97]. The mean
burning rate (or any mean quantity) is then estimated as:
_ v
Y
1
(x; t)
=
_
Y
1

_
Y
N
_ v
Y
1
(Y
p
1
; ; Y
p
N
)

P(Y
p
1
; ; Y
p
N
; x; t)dY
p
1
dY
p
N
(101)
Conditional statistics have been used to dened the ame
surface density (Section 6.3.2), these conditional means are
also useful in a pdf context. Consider a non-premixed ame
where the chemistry is reduced to a single step reaction, and
where radiative heat losses are neglected. Laminar combus-
tion would be parameterized with two variables, for
example, fuel mass fraction Y
F
and mixture fraction Z (see
Section 3.2). The turbulent ame is then fully described by
the joint pdf of mixture fraction and fuel mass fraction,

P(Y
p
F
; Z
p
; x; t): For such ames, it is interesting to focus on
the statistical properties of the fuel mass fraction Y
F
for a
given value of the mixture fraction Z (Section 3.2 and
Fig. 5).The conditional pdf

P
c
(Y
p
F
uZ
p
; x; t) is introduced
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 218
and dened for a given value of Z
p
as:

P(Y
p
F
; Z
p
; x; t) =

P
c
(Y
p
F
uZ
p
; x; t)

P(Z
p
; x; t)
The conditional mean is readily dened from:
_
Y
F
uZ(x; t) = Z
p
_
=
_1
0
Y
p
F

P
c
(Y
p
F
uZ
p
; x; t)dY
p
F
In this decomposition,

P(Z
p
; x; t) describes the statistical
properties of fuel/air mixing, whereas

P
c
(Y
p
F
uZ
p
; x; t) and
_
Y
F
uZ(x; t) = Z
p
_
are linked to the internal structure of the
ame front.
Two main directions may be chosen to built numerical
models from pdf:
To presume the pdf shape from available mean quantities.
To solve a balance equation for the pdf.
6.4.2. Presumed probability density functions
The simplest approach to the mixing between fuel and
oxidizer in a non-premixed ame is to presume the shape
of the pdf
~
P(Z
p
; x; t); usually with a Beta function [72]:
~
P Z
p
; x; t
_ _
=
(Z
p
)
a21
1 2Z
p
_ _
b21
_1
0
(Z
1
)
a21
1 2Z
1
_ _
b21
dZ
1
This presumed pdf should reproduce the mean of the
mixture fraction
~
Z and its variance

Z
//2
:
~
Z =
_1
0
Z
p
~
P Z
p
; x; t
_ _
dZ
p
;

Z
//2
=
_1
0
Z
p
2
~
Z
_ _
2
~
P Z
p
; x; t
_ _
dZ
p
and using the relation [98]:
_1
0
(Z
p
)
n
~
P Z
p
; x; t
_ _
dZ
p
=
a(a 11)(a 1(n 21))
(a 1b)(a 1b 11)(a 1b 1(n 21))
The two-parameters a and b are determined as:
a =
~
Z
~
Z(1 2
~
Z)

Z
//2
21
_ _
$ 0; b = a
1
~
Z
21
_ _
$ 0
This technique requires the solving of a balance equation for
the mean and the variance of Z (Section 9.2).

Z
//2
vanishes
when reactants are perfectly mixed and reaches its maxi-
mum value,
~
Z
_
1 2
~
Z
_
; when fuel and oxidizer are comple-
tely segregated (Section 5.3), then the pdf takes the formof a
double peak function, with peaks located at Z = 0 and Z = 1:
This approach is also used in premixed ames replacing the
mixture fraction Z by the progress variable c [99].
6.4.3. Pdf balance equation
The pdf balance equation is rst written for the progress
variable c. The time evolution of a probability density func-
tion is easily derived by expressing the pdf as the average of
a function d(c
p
2c(x; t)) dened as [100]:
d c
p
2c(x; t)
_ _
= 1=Dc
p
if c
p
2Dc
p
=2 , c , c
p
1Dc
p
=2
d(c
p
2c(x; t)) = 0 otherwise
By denition:

P(c
p
; x; t) = d(c
p
2c(x; t)) (102)
The time evolution of the reactive species c(x; t) is given by:
2c
2t
= 2u7c 1
1
r
[7(rD7c(x; t)) 1 _ v] (103)
and d(c
p
2c(x; t)) satises:
2
2C
[d(c
p
2c(x; t))] =
2
2c
[d(c
p
2c(x; t))]
2c(x; t)
2C
= 2
2
2c
p
[d(c
p
2c(x; t))]
2c(x; t)
2C
(104)
2
2
2C
2
[d(c
p
2c(x; t))]
= 2
2
2c
p
[d(c
p
2c(x; t))]
2
2
c(x; t)
2C
2
1
2
2
2c
p2
[d(c
p
2c(x; t))]

_
2c(x; t)
2C
_
2
(105)
where C can either be time or any spatial coordinates. These
relations (104) and (105) and Eq. (102) are key relations,
useful to obtain all equations discussed in this section.
After simple manipulations combining Eqs. (102) and
(103) with Eq. (104) the transport equation for the pdf
may be written:
2
2t
[

P(c
p
; x; t)] = 2
2
2c
p
2c(x; t)
2t
uc(x; t) = c
p
_ _

P(c
p
; x; t)
_ _
(106)
where conditional averaging
Q(c)d(c
p
2c(x; t)) =
_
Q(c)uc(x; t) = c
p
_

P(c
p
; x; t)
is introduced.
Eq. (106) shows that the time evolution of the pdf is
controlled by a ux in the sample space c
p
. This ux is
driven by a velocity equal to the conditional mean of the
time evolution of the progress variable c. In other words,
when the mean of (2c(x; t)=2t) is non-zero for the value c =
c
p
; the probability of nding the occurrence c = c
p
is modi-
ed. Because the probability to nd all the possible values is
constant and equal to unity:
_1
0

P(c
p
; x; t)dc
p
= 1
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 219
an increase or a decrease of the density of probability for
c = c
p
; implies a modication of this density for other
values of c
p
, justifying the convective term in sample
space. This term insures that the density of probability is
transported from point to point in the sample space c
p
in a
conservative manner. The ux depends on the conditional
mean of the time evolution of the progress variable given by
Eq. (103):
2c
2t
uc(x; t) = c
p
_ _
= 2u7cuc(x; t) = c
p
_ _
1
1
r
[7(rD7c(x; t))]uc(x; t) = c
p
_ _
1 _ vuc(x; t) = c
p
_ _
(107)
In Eq. (107), the conditional value of the source term is a
function of c dened in one-point and is exactly known
since:
_ v(c)uc(x; t) = c
p
_ _
= _ v(c
p
)
The main advantage of pdfs in turbulent combustion lies in
this availability to deal with chemistry, any term dened in
one-point (as chemical source) is closed. Nonetheless, this
advantage is offset by the fact that reactants are brought to
the reaction zone by diffusion, and the conditional mean of
2c(x; t)=2t also includes a conditional diffusive term:
1
r
(7(rD7c))uc(x; t) = c
p
_ _
(108)
named the micromixing term and remaining unclosed (as
any term involving spatial derivatives). This micromixing
term may be rewritten with the scalar dissipation rate,
x = Du7cu
2
; using Eq. (102) with Eq. (105) and assuming
rD < cst :
2
2c
p
D7
2
cuc(x; t) = c
p
_ _

P(c
p
; x; t)
_ _
= 2D7
2

P(c
p
; x; t) 1
2
2
2c
p2
xuc(x; t) = c
p
_ _

P(c
p
; x; t)
_ _
The rst term in the RHS, D7
2

P(c
p
; x; t) is usually neg-
ligible compared to the transport due to velocity uctua-
tions. The total dissipation rate, x is recovered as:
x =
_1
0
xuc(x; t) = c
p
_ _

P(c
p
; x; t)dc
p
Using diffusion or scalar dissipation rate, the probability
density function balance equation may be organized in
two different forms:
In terms of molecular diffusion:
2
2t
[

P(c
p
; x; t)]
= 2
2
2c
p
___
2u7c 1
1
r
7(rD7c)uc(x; t) = c
p
_
1 _ v(c
p
)
_

P(c
p
; x; t)
_
(109)
In terms of scalar dissipation rate:
2
2t
[

P(c
p
; x; t)]
= 2
2
2c
p
2u7cuc(x; t) = c
p
_ _
1 _ v(c
p
)
_ _

P(c
p
; x; t)
_ _
2
2
2
2c
p2
xuc(x; t) = c
p
_ _

P(c
p
; x; t)
_ _
The convective term may be split into mean,
u7

P(c
p
; x; t)
and uctuating components. Using an eddy viscosity model,
the uctuating part becomes:
2n
t
7

P(c
p
; x; t)
leading to the pdf balance equation:
2
2t
[

P(c
p
; x; t)] 1 u7

P(c
p
; x; t)
= n
t
7

P(c
p
; x; t) 2
2
2c
p
[ _ v(c
p
)

P(c
p
; x; t)]
2
2
2
2c
p2
xuc(x; t) = c
p
_ _

P(c
p
; x; t)
_ _
(110)
Weighted, or Favre, averages are also introduced in pdfs, for
instance when r = r(c) :
r
~
P(c
p
; x; t) = r(c
p
)d(c(x; t) 2c
p
) = r(c
p
)

P(c
p
; x; t)
When more than one species is taken into account, the pdf
balance equation is derived using the same formalism with
r
~
P(Y
p
1
; ; Y
p
N
) = r(Y
1
; ; Y
p
N
)d(Y
1
2Y
p
1
)d(Y
N
2Y
p
N
) [93]:
r
2
2t
~
P(Y
p
1
; Y
p
N
; x; t)
= 2 r

N
i=1
2
2Y
p
i
___
2u7Y
i
uY(x; t) = Y
p
_
1 _ v
p
i
_

~
P
_
Y
p
1
; ; Y
p
N
; x; t
__
2 r

N
i=1

N
j=1
2
2Y
p
i
2
2Y
p
j
__
D
2Y
i
2x
k
2Y
j
2x
k
uY(x; t) = Y
p
_

~
P
_
Y
p
1
; ; Y
p
N
; x; t
_
_
(111)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 220
The same terms are identied on the RHS: convection by the
conditional velocity and by the chemical source, and micro-
mixing. As was done in Eq. (110), the convective term
in physical space can be decomposed into mean and
uctuating parts.
6.4.4. Joint velocity/concentrations pdf
To avoid the gradient transport assumption for the
turbulent ux, (u
/
7cuc(x; t) = c
p
); the joint velocity/
concentration pdf is introduced. Once this joint pdf is
known, turbulence models, such as k1, are, a priori, no
longer required for the mean ow. However, an equation
for 1 is still needed to estimate a characteristic mixing time,
except when the frequency of mixing is also included in the
joint pdf [101]. The transport equation for this joint compo-
sition/velocity pdf is given below for the progress variable c,
where the RHS contains the unclosed terms.
2
2t
[

P(u
p
; c
p
; x; t)] 1u
p
7

P(u
p
; c
p
; x; t)
1 n7
2
u
i
2
1
r
2 p
2x
i
_ _
2
2u
p
i
[

P(u
p
; c
p
; x; t)]
1
2
2c
p
[ _ v(c
p
)

P(u
p
; c
p
; x; t)]
=
2
2u
p
i
1
r
2p
/
2x
i
_
_
_
_
u = u
p
; c = c
p
_ _

P(u
p
; c
p
; x; t)
_ _
....................,,....................
Pressure fluctuations
2
2
2u
p
i
n7
2
u
/
i
uu = u
p
; c = c
p
_ _

P(u
p
; c
p
; x; t)
_ _
...................,,...................
Viscous dissipation
2
2
2c
p
D7
2
cuu = u
p
; c = c
p
_ _

P(u
p
; c
p
; x; t)
_ _
...................,,...................
Molecular diffusion
(112)
The LHS terms are closed and represent respectively accu-
mulation, convection in physical space by the random vel-
ocity eld (incorporating turbulent transport), convection of
the pdf in velocity space, here the convective velocity is the
mean of the viscous dissipation and the mean pressure gradi-
ent, and nally the closed chemical source. The unclosed
terms (RHS) are the pressure uctuations, the uctuating
part of the viscous dissipation and micromixing. All these
phenomena remain to be closed.
6.4.5. Conditional moment closure (CMC)
Conditional moment closure modeling was rst proposed
in Refs. [102,103]. As with the pdf, the idea is to focus on
particular states between fresh gases and fully burnt product
in premixed ames, or, between fuel and oxidizer in non-
premixed combustion. However, here only conditional
moments (rY
i
uc = c
p
) are considered. In premixed ames,
the conditional quantity is the progress variable c, whereas
for non-premixed combustion, the mixture fraction is used.
In premixed ames, the mean value,
~
Y
i
of Y
i
may be esti-
mated as:
r
~
Y
i
=
_1
0
rY
i
uc = c
p
_ _

P(c
p
; x; t)dc
p
(113)
One may solve a balance equation for the conditional quan-
tities Q
i
(c
p
) dened as:
Q
i
(c
p
) =
rY
i
uc = c
p
_ _
ruc = c
p
_ _
This balance equation is [104,105]:
ruc = c
p
_ _
2Q
i
2t
= 2 ru
i
uc = c
p
_ _
2Q
i
2x
i
1 rxuc = c
p
_ _
2
2
Q
i
2c
p2
1 _ v
i
uc = c
p
_ _
1E
Q
i
1E
Y
i
(114)
The two last terms of Eq. (114) are usually neglected,
E
Q
i
appears from molecular diffusion along with differ-
ential diffusion effects across the iso-c surface, E
Y
i
represents the effects of turbulent uctuations on the
deviation from the conditional mean. The three rst
terms on the right hand side are unclosed, they describe
convective transport, micromixing (x enters this term) and
chemical source. Closures for the conditional values of the
scalar dissipation rate, but also for the conditional value of
the chemical source of Y
i
, calculated for a given value of the
progress variable c
p
or the mixture fraction Z, are required
[106,107]. One equation has to be solved for each value of
c
p
retained. The number of these values is determined from
the accuracy required to estimate both the mean from Eq.
(113) and the second order derivative in the sample space
(i.e. 2
2
/2c
p2
) found in Eq. (114). On the other hand, the
probability density function entering expression (113),

P(c
p
; x; t) is presumed.
CMC may also be viewed as a multi-surface description,
any conditional quantity
_
rY
i
uc = c
p
_
corresponding to the
conditional average of Y
i
along the iso-surface c = c
p
:
6.5. Similarities and links between the tools
Major links between the tools used in turbulent combus-
tion modeling are now developed. Without loss of the
generic character of the discussion, we consider the case
of a turbulent premixed ame represented with a progress
variable c. To describe this turbulent ame, three quantities
are useful:
The scalar dissipation rate of the progress variable
r ~ x = rD7c7c.
The pdf of the progress variable

P(c
p
; x; t):
The ame surface density S or the mean eld

G:
Simple links exist between these quantities:
The conditional value of the scalar dissipation rate enters
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 221
the pdf transport equation (Eq. (110)), therefore there is a
direct connection between

P(c
p
; x; t) and ~ x:
The surface density function S(c
p
; x; t); or ame surface,
is related to the pdf via the conditional value of u7cu; S =
_
u7cuuc = c
p
_

P(c
p
) (Eq. (86)).
Since x is proportional to u7cu; the ame surface S, the
pdf

P(c
p
) and the dissipation rate ~ x are very strongly
related. Using the joint pdf of c and x, the ame surface
density may be written:
S(c
p
; x; t) =

x
D
_
uc = c
p
_ _

P(c
p
; x; t)
=
_
x

x
p
D
_

P(c
p
; x
p
; x; t)dx
p
(115)
and characteristic length scales d
c
=

D=x
_
of the iso-c
distribution are embedded within S(c
p
).
Combining Eq. (86) and Eq. (91) with Eq. (96), the mean
c-scalar dissipation rate is also a function of surface
densities:
r ~ x = rDu7cu
2
=
_1
0
krDu7cul
s
S(c
p
)dc
p
= krDu7cul
s
u7cu
(116)
where the generalized surface average kQl
s
is dened by
Eq. (96). Relations between ame surface densities and
scalar dissipation rates were anticipated by Borghi [108].
In premixed combustion, using a G-eld equation, the
ame front is identied to a given level G = G
0
[29].
The ame surface density is then:
S(G
0
) = u7GuuG = G
0
_ _

P(G
0
) (117)
Hence, previous relations may be recast in terms of the G-
equation.
The CMC formalism (see Section 6.4.5) may be re-
organized in term of ame surface density:
r
~
Y
i
=
_1
0
rY
i
uc = c
p
_ _

P(c
p
; x; t)dc
p
=
_1
0
rY
i
u7cu
_ _
s
S(c
p
)dc
p
(118)
where the conditional mean Q
i
(c
p
) may be written:
Q
i
(c
p
) =
rY
i
uc = c
p
_ _
ruc = c
p
_ _
=
krY
i
=u7cul
s
kr=u7cul
s
(119)
and appears as directly related to c = c
p
surface averaged
quantities.
The links between the combustion modeling tools are
summarized in Table 2. The mean burning rate is given by:

_ v =
_1
0
_ v(c
p
)

P(c
p
)dc
p
(120)
Using relations (86) and (91) and Eq. (116), this expression
becomes:

_ v =
_1
0
_ vuc = c
p
_ _

P(c
p
)dc
p
=
_1
0
_ v
u7cu
_ _
s
S(c
p
)dc
p
=
_ v
u7cu
_ _
s
u7cu =
_ v

x=D
_
_ _
s
S =
1
krDu7cul
s
_ v
u7cu
_ _
s
r ~ x
(121)
Models based on probability density functions, conditional
means, ame surface density function and generalized ame
surface density are then related via the scalar dissipation rate.
These expressions are extended to burning rate depending
on many species using multi-dimensional pdfs and con-
ditional averaging. When the local reaction rate is a function
of various quantities (species mass fractions, temperature,
etc.), Y
i
and a sampling scalar c:

_ v =
_
Y
1

_
Y
N
_1
0
_ v
i
(Y
p
1
; ; Y
p
N
; c
p
)

P(Y
p
1
; ; Y
p
N
; c
p
)dY
p
1
; ; dY
p
N
dc
p
=
_1
0
_
Y
1

_
Y
N
_ v
i
(Y
p
1
; ; Y
p
N
; c
p
)

P
c
(Y
p
1
; ; Y
p
N
uc
p
)dY
p
1
; ; dY
p
N
_ _
..........................,,..........................
_ v;(Y
1
;;Y
N
;c)uc=c
p
_ _

P(c
p
)dc
p
(122)
decomposing the joint-pdf as

P(Y
p
1
; ; Y
p
N
; c
p
) =

P
c
(Y
p
1
; ; Y
p
N
uc
p
)

P(c
p
): Note that such a relation may be
used when in premixed ames the progress variable, based
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 222
Table 2
Exact relations between mean scalar dissipation rate, ~ x, probability density function,

P(c
p
), or surface density, S(c
p
), of iso-surface c = c
p
and
generalized ame surface density S = u7cu: Connections with the G-equation are readily obtained with Eq. (117). kQl
s
is a surface average (Eq.
(91)), kQl
s
corresponds to generalized surface average and is dened in Eq. (96). c is the conditional variable (progress variable in premixed
ames, mixture fraction in non-premixed combustion)
Scalar dissipation rate r ~ x Probability density function

P(c
p
)
Flame surface density S(c
p
)
r ~ x
_1
0
_
rDu7cu
2
uc = c
p
_

P(c
p
)dc
p
_1
0
krDu7cul
s
S(c
p
)dc
p

P(c
p
)
_
xuc = c
p
_
via PDF Eq: S(c
p
)=
_
u7cuuc = c
p
_
S = u7cu
r ~ x
krDu7cul
s
_1
0
_
u7cuuc
p
_

P(c
p
)dc
p
_1
0
S(c
p
)dc
p
on reactant mass fractions, and the temperature are indepen-
dent (compressibility effects, heat losses, non-unity Lewis
numbers, etc.).
The fundamentals of turbulent combustion modeling
clearly rely on pdf, ame surface density, G-eld and scalar
dissipation rate concepts. The previous relations may be
used to carefully compare the proposed closure schemes
and Table 3 summarizes the tools and their related modeling
issues. Various expressions for mean reaction rates are
displayed in Table 4. Major differences between the various
approaches only appear when closing the unknown quanti-
ties, but at the light of these relations, many closures are
essentially equivalent.
7. Reynolds-averaged models for turbulent premixed
combustion
7.1. Turbulent ame speed
Turbulent premixed ames may be described in terms of a
global turbulent ame speed S
T
. From experimental data
[109,110] or theoretical analysis (Renormalization group
theory [111]), the following expression has been proposed:
S
T
S
L
= 1 1a
u
/
S
L
_ _
n
(123)
where a and n are two model constants of the order of unity.
u' is the turbulent velocity (i.e the RMS velocity).
Unfortunately, the turbulent ame speed S
T
is not a fully
well dened quantity [112]. Experimental data exhibit a
large scatter because they depend on various parameters
(chemistry characteristics, turbulence scales, ow geometry,
etc.). While this global approach is not particularly well suited
to close Favre averaged transport equations, it may be of
interest in the context of Large Eddy Simulation [79,113].
7.2. Eddy-Break-Up model
Devised in Ref. [114], this model is based on a phenom-
enological analysis of turbulent combustion assuming high
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 223
Table 3
Tools for turbulent combustion modeling
Description Tools Modeling issues
Geometrical G-eld with G = G
p
at the ame
2
~
G=2t 1 ~ u7
~
G = (r
u
= r)S
T
u7
~
Gu
Propagation speed S
T
Flame surface S
Algebraic closure or transport equation

_ v
i
= Combustion
_
V
i
S
Turbulence
Total stretch = curvature 1strain rate
Small scales Scalar dissipation rate, r ~ x = rDu7Yu
2
Fast chemistry
~
_ v < ~ x
Algebraic closure or transport equation Provide the rate of micromixing
Statistical Probability density function
~
P(Y
p
F
; Z
p
; x; t) Micromixing
(1) Presumed 1amelets (1) Strategy to generate the pdf
(2) Transport equation (2) Solve 2
~
P=2t =
Conditional mean, CMC, presumed PDF
_
Y
F
uZ(x; t) = Z
p
_
=
_
1
0
Y
p
F

P
c
(Y
p
F
uZ
p
; x; t)dY
p
F
Micromixing and conditional source
Fundamental links: S(x; t) =
_
u7cuuc = c
p
_

P(c
p
; x; t) and r ~ x = krDu7cul
s
u7cu
Table 4
Exact expressions for the averaged reaction rate as a function of the mean scalar dissipation, ~ x; probability density function,

P(c
p
) or surface
density, S(c
p
) of iso-surface c = c
p
: Connections with the G-equation are readily obtained with Eq. (117). kQl
s
is a surface average (Eq. (91)),
kQl
s
corresponds to generalized surface average and is dened in Eq. (96)
Tools Averaged reaction rate
Scalar dissipation r
_
1
0
_
_ v
u7cu
_
s
S(c
p
)dc
p
_
1
0
krDu7cul
s
S(c
p
)dc
p
~ x = r
_
_ v
u7cu
_
s
krDu7cul
s
~ x
Probability density function
_
1
0
_ v(c
p
)

P(c
p
)dc
p
Flame surface density
_
1
0
_
_ v
u7cu
_
s
S(c
p
)dc
p
=
_
_ v
u7cu
_
s
u7cu
Reynolds (Re @ 1) and Damkohler (Da @ 1) numbers. The
reaction zone is viewed as a collection of fresh and burnt
gases pockets. Following the Kolmogorov cascade, turbu-
lence leads to a break down of fresh gases structures.
Accordingly, the mean reaction rate is mainly controlled
by the turbulent mixing time t
t
. When oxidizer is in excess,
the mean reaction rate is expressed as:

_ v
F
= 2C
EBU
r

Y
//2
F
_
t
t
(124)
where Y
//
F
denotes the fuel mass fraction uctuations and
C
EBU
is a model constant of the order of unity. The turbulent
mixing time, t
t
is estimated from the turbulence kinetic
energy k and its dissipation rate 1 according to:
t
t
=
k
1
as an approximation of the characteristic time of the integral
length scales of the turbulent ow eld.
The reaction rate may be recast in terms of progress vari-
able, c, as:

_ v = 2C
EBU
r

c
//2
_
t
t
(125)
Mass fraction uctuations

Y
//2
F
(or progress variable uctua-
tions

c
//2
) must be modeled and may be estimated from a
balance equation (see Eq. (77)). Assuming an innitely thin
ame front (i.e. c = 0 or c = 1),

c
//2
is easily estimated
because c
2
= c :
r

c
//2
= r(c 2 ~ c)
2
= r(
`
c
2
2 ~ c
2
) = r ~ c(1 2 ~ c) (126)
The square root has been introduced for dimensional reasons
in Eqs. (124) and (125) but, unfortunately, Eqs. (125) and
(126) lead to inconsistencies because the ~ c derivative of

_ v;
d

_ v=d~ c; is innite both when ~ c = 0 and when ~ c = 1 (Borghi,
1999, private communication). Then, a corrected version of
the Eddy-Break-Up model, without the square root, is used
for practical simulations:

_ v = C
EBU
r
1
k
~ c(1 2 ~ c) (127)
or, in terms of fuel mass fraction:

_ v
F
= 2C
EBU
r
1
k
~
Y
F
Y
0
F
1 2
~
Y
F
Y
0
F
_ _
(128)
where Y
0
F
is the initial fuel mass fraction in the reactant
stream, assuming excess of oxidizer.
The EBU model was found attractive because the reaction
rate is simply written as a function of known quantities
without any additional transport equation and is available
in most commercial CFD codes. The modeled reaction rate
does not depend on chemical characteristics and assumes a
homogeneous and isotropic turbulence. Some adjustments
of the model constant C
EBU
have been proposed to mimic
chemical features [115]. Eddy-Break-Up modeling tends to
overestimate the reaction rate, especially in highly strained
regions, where the ratio 1/k is large (ame-holder wakes,
walls, etc.).
7.3. BrayMossLibby model
7.3.1. Introduction
Known under the initials of its authors, Bray, Moss and
Libby, or, from the involved physical hypothesis, BiModal
Limit, this model, rst proposed in 1977 [116], has been the
subject of a large amount of work leading to many improve-
ments (see papers by Bray, Moss and Libby, and then by
Bray, Champion and Libby). Combining a statistical
approach using probability density functions and a physical
analysis, this model has evidenced some special features of
turbulent premixed combustion (counter-gradient turbulent
transport, ame turbulence generation, etc.). The presenta-
tion is mainly limited here to basic concepts of the Bray
MossLibby (BML) formulation.
A one-step, irreversible chemical reaction between two
reacting species, fresh gases (R) and combustion products
(P) is considered:
R - P
Classical assumptions are made to simplify the formula-
tion: perfect gases, incompressible ows, constant chemical
properties, unity Lewis numbers, etc. A progress variable, c,
of the chemical reaction is introduced where c = 0 in fresh
gases and c = 1 in fully burnt gases, as described in Section
3.1.
The basic idea of the BML formulation is to presume the
probability density function of the progress variable c at a
given location (x; t) as a sum of fresh, fully burnt and burn-
ing gases contributions (Fig. 15):

P(c
p
; x; t) = a(x; t)d(c
p
)
....,,....
fresh gases
1b(x; t)d(1 2c
p
)
......,,......
burnt gases
1g(x; t)f (c
p
; x; t)
......,,......
burning gases
(129)
where a, b and g respectively denote the probability to
have, at location (x; t); fresh gases, burnt gases and reacting
mixture. d(c
p
) and d(1 2c
p
) are respectively the Dirac delta
functions corresponding to fresh gases (c = 0) and fully
burnt ones (c = 1):
Normalization of the probability density function:
_1
0

P(c
p
; x; t)dc
p
= 1 (130)
leads to the following relations:
a 1b 1g = 1 (131)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 224
_1
0
f (c
p
; x; t)dc
p
= 1 (132)
with f (0) = f (1) = 0:
The balance equation for the progress variable c may be
written:
2rc
2t
17(ruc) = 7(rD7c) 1 _ v (133)
This equation is averaged and the mean reaction rate, at the
location (x; t) is:

_ v(x; t) =
_1
0
_ v(c
p
)

P(c
p
; x; t)dc
p
(134)
leading to Eq. (129):

_ v(x; t) = g(x; t)
_1
0
_ v(c
p
)f (c
p
; x; t)dc
p
(135)
All studies devoted to this line of models are based on such a
formulation. The objective is now to determine the unknown
functions a, b, g and the probability density function f.
Using the Damkohler number, Da, comparing turbulence
and chemical time scales, and, the turbulent Reynolds
number Re, we will focus only on the case where
Re @Da @1. In this situation, the combustion is controlled
by the turbulent transport and the reaction zone may be
assumed to be innitely thin. Accordingly, g !1 (i.e.
a @g and b @g).
7.3.2. BML model analysis
This model is developed under the assumption that
Re @Da @1, corresponding to g !1. At a given location
in the ow, an intermittency between fresh gases (c = 0)
and fully burnt ones (c = 1) is observed and the probability
density function of the progress variable c reduces to:

P(c
p
; x; t) = a(x; t)d(c
p
) 1b(x; t)d(1 2c
p
) 1O(1=Da)
(136)
which is `quasi-bimodal' for large Damkohler numbers.
Then at the point (x; t) inside the reaction zone, c looks as
a telegraphic signal as displayed in Fig. 16. This signal also
satises:
c
2
= c 1O(1=Da) and c(1 2c) = O(1=Da) (137)
Under this assumption, a and b are easily determined as a
function of the Favre average progress variable ~ c :
r ~ c = r ~ c =
_1
0
rc
p

P(c
p
)dc
p
= r
b
b (138)
where r
b
is the burnt gases density.
b =
r ~ c
r
b
and a = 1 2
r ~ c
r
b
(139)
The mean density r may be written:
r =
_1
0
r

P(c
p
)dc
p
= ar
u
1br
b
= 1 2
r`c
r
b
_ _
r
u
1
r`c
r
b
r
b
(140)
One may also introduce the reaction heat release factor t,
dened as:
t =
r
u
r
b
21 =
T
b
T
u
21 (141)
leading to:
r
u
= (1 1t)r
b
= r(1 1t~ c) (142)
corresponding to the perfect gases state law, assuming a
constant pressure P. Then, the probabilities a and b
become:
a =
1 2 ~ c
1 1t~ c
; b =
(1 1t) ~ c
1 1t~ c
(143)
The probability density function

P(c) is determined and
depends only on the mean progress variable ~ c (and on the
heat release factor, t, which is xed for a given chemical
reaction). The BML model involves a presumed pdf (Eq.
(135)), however, the mean reaction rate cannot be calculated
from the pdf since g was neglected in Eq. (129)).
Starting from the conservative and non-conservative
forms of the progress variable c balance equations:
2rc
2t
17(ruc) = 7(rD7c) 1 _ v
r
2c
2t
1ru7c = 7(rD7c) 1 _ v
and multiplying by c and adding these equations:
2rc
2
2t
17(ruc
2
) = 7(rD7c
2
) 22rD7c7c 12c _ v (144)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 225
Fig. 15. Probability density function in premixed turbulent
combustion.
Fig. 16. Intermittency between fresh and fully burnt gases at a given
location in the reaction zone. This signal corresponds to a bimodal
(c = 0 and c = 1) probability density function.
Subtracting the balance equation for c
2
(Eq. (144)) to the
balance equation for the progress variable c (Eq. (133))
leads to a balance equation for c(1 2c) :
2
2t
[rc(1 2c)] 17[ruc(1 2c)]
= 7(rD7[c(1 2c)]) 12rD7c7c 22c _ v 1 _ v (145)
Under the assumption of the BML model, the progress vari-
able c is equal to zero or to unity. Accordingly, c(1 2c) = 0
and the balance Eq. (145) reduces to:
2rD7c7c = 2c _ v 2 _ v (146)
leading to, after averaging:
2rD7c7c = (2c
m
21)

_ v (147)
where a progress variable c
m
, dened as:
c
m
=
_1
0
c _ vf (c)dc
_1
0
_ vf (c)dc
(148)
has been introduced and characterizes the chemical reaction.
The mean reaction rate

_ v becomes:

_ v = 2
rx
2c
m
21
(149)
where rx is the scalar dissipation rate of the progress
variable c.
rx = r` x = rD7c7c = rD
2c
2x
i
2c
2x
i
(150)
The mean reaction rate

_ v is then related to the dissipation
rate ~ x; describing the turbulent mixing, and to c
m
, character-
izing the chemical reaction. A transport equation for the
scalar dissipation rate may be written and solved [117], or
one may postulate a linear relaxation of the uctuations
generated by micromixing (Section 9.2.2), leading to:
rx =
rc
//2
t
t
(151)
where a turbulent time scale, t
t
is introduced. Assuming an
intermittency between fresh and burnt gases (c = 0 or
c = 1),

c
//2
is given by Eq. (126). Then:

_ v =
2
2c
m
21
r ~ c(1 2 ~ c)
t
t
(152)
The Eddy-Break-Up model expression (Eq. (127)) is
recovered (Section 7.2). The BML model may then be
viewed as a theoretical derivation, where the assumptions
made are clearly stated, of the Eddy-Break-Up (EBU)
model, initially based on a phenomenological approach.
7.3.3. Recovering mean reaction rate from tools relations
The analysis developed in Section 7.3.2 is the usual
derivation of the BML model. However, expression (149)
linking mean reaction rate and scalar dissipation may also be
recovered from the general relations between modeling
tools (Eq. (121)):

_ v =
1
krDu7cul
s
_ v
u7cu
_ _
s
r ~ x (153)
To link mean reaction rate and scalar dissipation rate using
the amelet analysis, Eq. (153) requires estimates for _ v=u7cu
and rDu7cu; averaging of these quantities along iso-surface
c = c
p
and integration over all the possible c
p
values.
Assuming that innitely thin ame elements may be viewed
as one-dimensional, steady state, premixed laminar ame
propagating at a given laminar ame speed S
L
, the balance
equation for the progress variable c in this 1D ame is:
r
0
S
L
7c = 7(rD7c) 1 _ v (154)
where r
0
denotes the density in the fresh gases. Integrating
Eq. (154) across all the ame front from fresh gases to burnt
gases (21 # x # 11) and up to a given location x
0
, corre-
sponding to a progress variable c
0
(21 # x # x
0
); leads to:
r
0
S
L
=
_11
21
_ vdx =
_1
0
_ v
u7cu
dc (155)
(rD7c)
x
0
= r
0
S
L
c
0
2
_x
0
21
_ vdx
=
_1
0
_ v
u7cu
dc
_ _
c
0
2
_c
0
0
_ v
u7cu
dc (156)
As the ame front is supposed to be thin and planar, all c
p
iso-surfaces have the same surface and all quantities are
constant along these surfaces, moreover the ame is
supposed locally one-dimensional and u7cu = 7c: Then:
_1
0
k _ v=u7cul
s
dc
p
_1
0
krDu7cul
s
dc
p
=
_1
0
_ v=u7cudc
_1
0
rDu7cudc
=
_11
21
_ vdx
_11
21
_ vdx
_1
0
c dc 2
_1
0
_c
0
0
( _ v=u7cu)dc dc
0
=
2
_11
21
_ v dx
2
_11
21
c _ vdx 2
_11
21
_ v dx
=
2
2c
m
21
(157)
where c
m
is dened as:
c
m
=
_11
21
c _ vdx
_11
21
_ vdx
(158)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 226
Then, from Eq. (153):

_ v = 2
rx
2c
m
21
(159)
recovering Eq. (149). This second analysis leads to the
following comments:
The denition of c
m
is slightly different in Eqs. (149) and
(159). In the rst one, the integration is performed over
all possible values of the progress variable c at a given
location (x; t): In the second analysis, the integration is
performed along a normal direction to the ame front,
assumed to be a laminar one-dimensional premixed
ame. The expressions (148) and (158) are identical
when:
f (c) =
1
d
l
dc
dx
_ _
21
f
(160)
where (dc=dx)
f
describes the inner structure of the ame-
let.
This second derivation clearly exhibits how a amelet
model works (the turbulent ame is viewed as a collec-
tion of thin laminar ame elements). The c
m
parameter is
related to the inner structure of the ame front and to the
properties of the chemical reaction.
Making use of the ame surface density S as dened by
Eq. (116), the expression for the mean burning rate
becomes:

_ v = r
0
S
L
S (161)
where
S =
2
2c
m
21
1
r
0
S
L
r ~ x <
2
2c
m
21
1
r
0
S
L
r
~ c(1 2 ~ c)
t
t
(162)
The BML model has been developed starting from a
statistical analysis (two peaks pdf), the mean reaction
rate was written in terms of the scalar dissipation rate
of the progress variable c, which may be recast in terms
of ame surface density, exhibiting simple links between
modeling tools.
7.3.4. Reynolds and Favre averaging
Assuming a bimodal distribution of the progress variable
c, Reynolds ( c) and Favre ( ~ c) averages are easily related.
From:
c =
_1
0
c

P(c)dc = b (163)
rc = r ~ c =
_1
0
rc

P(c)dc = r
b
b = r
b
c (164)
together with expressions (142) and (143), one easily
obtains:
c =
(1 1t) ~ c
1 1t~ c
(165)
corresponding to a model for the density/progress variable
correlations (Eq. (15)):
r
/
c
/
= 2 r
t~ c(1 2 ~ c)
1 1t~ c
(166)
Reynolds ( c) and Favre ( ~ c) averages of the instantaneous
progress variable c are compared on Fig. 17 for various
values of the heat release factor t. The discrepancy between
the two quantities strongly increases with t. r
/
c
/
is plotted
as a function of ~ c on Fig. 18.
7.3.5. Conditional averagingcounter-gradient turbulent
transport
The analysis of intermittency between fresh gases (c = 0)
and fully burnt gases (c = 1) leads to the introduction of
conditional averaging. The Favre average
~
Q of any quantity
Q may be expressed as a function of the conditional
averages for fresh gases (

Q
u
) and fully burnt products (

Q
b
) :
r
~
Q = rQ =
_1
0
rQ

P(c)dc = ar
u

Q
u
1br
b

Q
b
(167)
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 227
Fig. 17. The Reynolds average c of the progress variable cis plotted as a function of the Favre average ~ c for various values of the heat release
factor t, assuming a bimodal distribution (c = 0 or c = 1) of c (Eq. (165)).
using Eqs. (142) and (143):
~
Q = (1 2 ~ c)

Q
u
1 ~ c

Q
b
(168)
where

Q
u
and

Q
b
are dened as:

Q
u
=
_11
21
Q

P
c
(Quc = 0)dQ

Q
b
=
_11
21
Q

P
c
(Quc = 1)dQ

P
c
(Quc) is the pdf of Q for the given value c of the progress
variable (conditional pdf, see Section 6.4.1).
The components ~ u
k
of the mean velocity vector ~ u may be
written as a linear combination of their conditional fresh and
burnt gases averages:
~ u
i
= (1 2 ~ c) u
u
i
1 ~ c u
b
i
(169)
Then:
r

u
//
i
c
//
= r( u
i
c 2 ~ u
i
~ c) = r( ~ c u
b
i
2 ~ u
i
~ c) = r ~ c(1 2 ~ c)( u
b
i
2 u
u
i
)
(170)
which is the scalar turbulent ux, generally modeled using a
gradient assumption:
r

u
//
i
c
//
= 2
m
t
S
c
2~ c
2x
i
_ _
(171)
The two expressions (170) and (171) may describe opposite
uxes: consider a left-traveling one dimensional turbulent
ame, because of thermal expansion the conditional velo-
city in the burnt gases, u
b
i
; is expected to be larger than the
conditional velocity in the fresh gases, u
u
i
: According to Eq.
(170), the turbulent ux,

u
//
i
c
//
is expected to be positive. On
the other hand, as the mean progress variable gradient is also
positive, Eq. (171) leads to a negative value of

u
//
i
c
//
: This
situation, called `counter-gradient turbulent transport', is a
key point of the BML analysis and will be further discussed
in Section 8.
The Reynolds stresses

u
//
i
u
//
j
may also be decomposed
using the same formalism:

u
//
i
u
//
j
= (1 2 ~ c)u
/
i
u
/
j
u
1 ~ cu
/
i
u
/
j
b
1 ~ c(1 2 ~ c)( u
b
i
2 u
u
i
)( u
b
j
2 u
u
j
)
............,,............
intermittency
(172)
where one may note a weighted mean between the Reynolds
stresses in the fresh
_
u
/
i
u
/
j
u
_
and in the burnt gases
_
u
/
i
u
/
j
b
_
representative of turbulent motions. The additional term
represents the intermittency between fresh and burnt gases.
7.3.6. Extensions to partially premixed combustion
Some attempts have been made to extend BML modeling
concepts to partially premixed combustion, i.e. when reac-
tants are not perfectly premixed before burning [118]. Two
difculties are then encountered. First, in the denition of
the progress variable c (Eq. (35)), unburnt and burnt gas
temperatures and fuel mass fractions T
u
, T
b
, Y
u
F
and Y
b
F
are
no longer constant and depend on local equivalence ratio. A
balance equation may be still derived (see, for example, Ref.
[119]) but it incorporates additional terms and is not obvious
to close, especially when rich and lean zones coexist in the
same ow eld. Mixing should also be taken into account
through a mixture fraction Z. The challenge is then to model
the joint probability density function

P(c
p
; Z
p
; x; t): Lahjaily
et al. [118] write:

P(c
p
; Z
p
; x; t)
= a(x; t)d(c
p
)

P
u
(Z
p
; x; t) 1b(x; t)d(1 2c
p
)

P
b
(Z
p
; x; t)
1g(x; t)F(c
p
; Z
p
; x; t)H(Z
p
2Z
min
)
..............,,..............
burning zones
1g
m
(x; t)F
m
(c
p
; Z
p
; x; t)[1 2H(Z
p
2Z
min
)]
...................,,...................
mixing without burning
(173)
where

P
u
(Z
p
; x; t) and

P
b
(Z
p
; x; t) represent the mixture frac-
tion distributions in fresh and burnt gases respectively. F
and F
m
are the distributions within the reaction zones,
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 228
Fig. 18. Correlation 2r
/
c
/
= r as a function of the Favre averaged progress variable ~ c for various values of the heat release factor t, assuming a
bimodal distribution (c = 0 or c = 1) of c (Eq. (166)).
assuming to be amelets, and within the non-reactive
mixing zone between fresh and burnt gases. Combustion
is supposed to occur when Z
p
$ Z
min
(H is the Heaviside
function verifying H(z , 0) = 0 and H(z $ 0) = 1). Under
BML assumptions, the two last contributions are neglected
(g p1 and g
m
p1) and dilution effects are incorporated
through

P
u
(Z
p
; x; t) and

P
b
(Z
p
; x; t):
7.4. Models based on the ame surface area estimation
7.4.1. Introduction
Several ame surface density models are now discussed.
Their derivation and their histories differ, but theyare all based
on similar concepts, described in Section 6.3.2. These models
assume that the chemical reaction occurs in thin layers
separating fresh gases from fully burnt ones (high Dam-
kohler number limit). The reaction zone may then be viewed
as a collection of laminar ame elements called amelets.
The ame surface density is here introduced at the light of
experimental data from Refs. [120,121]. The experimental
burner is displayed on Fig. 19. A turbulent premixed
propane/air ame is stabilized behind a small cylinder
(blockage ratio of 6%). Flow rates are about 35100 g/s,
corresponding to inlet velocities between 10 and 30 m/s
(turbulence levels from 5 to 10%). Equivalence ratios f
are in the range 0.71.1. Velocity (laser Doppler veloci-
metry), CH and C
2
radical emission (reaction rate estima-
tion) and high-speed laser tomography (ame front
characteristics) measurements have been performed and
are described in Refs. [120123].
In Fig. 20 (half burner), ame surface density proles are
plotted as a function of the transverse coordinate for various
downstream locations and for two equivalence ratios.
7.4.2. Algebraic expressions for the ame surface density S
Assuming intermittency between fresh and burnt gases
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 229
Fig. 19. Experimental burner. A propane/air premixed ow is
injected in a rectangular burner through a grid. The turbulent
ame is stabilized behind a small cylinder (blockage ratio of 6%)
[120,121].
Fig. 20. Transverse ame surface density (S) proles (m
21
) plotted as a function of the transverse location for various downstream locations
(mm downstream the rod). (a) f = 0:78; (b) f = 0:9: Flow rate: 35 g/s [121].
(Section 7.3), Bray, Moss, Libby and their coworkers have
proposed to describe the mean reaction rate

_ v as the product
of a ame crossing frequency n and a local reaction rate per
ame crossing, w
F
:

_ v = w
F
n (174)
Because c(t) may be viewed as a telegraphic signal (Fig. 16),
this crossing frequency is derived from a statistical analysis
of the telegraph equation, leading to:
n =
2 c(1 2 c)
^
T
(175)
where
^
T is time scale for the uctuations of c.
This analysis is attractive because the crossing frequency
n may be easily obtained in experiments, for example from
time-resolved local temperature measurements (thermo-
couple). On the other hand, the ame surface density S
and the local reaction rate per ame crossing (w
F
) are not
obvious to estimate. Eq. (175) is generally closed estimating
^
T from a characteristic turbulent time t
t
. The reaction rate
per crossing ame, w
F
(Eq. (174)), is usually modeled as:
w
F
=
r
0
S
L
d
t
=t
t
(176)
where r
0
is the unburnt gases density, S
L
and d
l
are respec-
tively the speed and the thickness of the laminar ame. The
ame transit time t
t
measures the averaged time spend by a
point in the ow to cross a ame front and corresponds to the
mean transition time between c = 0 and c = 1 levels of the
progress variable, as shown in Fig. 21 (in practice, the c-
signal is not exactly bimodal).
This model has been latter rewritten in term of ame
surface density, leading to the algebraic expression [124]
S = g
c(1 2 c)
s
y
L
y
=
g
s
y
L
y
1 1t
(1 1t~ c)
2
~ c(1 2 ~ c) (177)
where g is a constant of order unity. s
y
is a amelet orienta-
tion factor measuring the mean angle of the instantaneous
ame front with the c surface and assumed to be an universal
model constant (s
y
< 0:5): L
y
is a ame front wrinkling
length scale and Eq. (165) has been used to replace c as a
function of ~ c: A submodel is required to describe the wrink-
ling length scale L
y
, generally assumed to be proportional to
the integral length scale l
t
:
L
y
= C
l
l
t
S
L
u
/
_ _
n
(178)
where C
l
and n are two constants of the order of unity [75].
According to Gouldin [125,126], the amelet orientation
factor s
y
is directly linked to the vectors normal to the
instantaneous ame front, n, and to the mean ame brush,
n
p
(see Section 6.3.3):
1
s
y
=
1
un
p
nu
_ _
s
(179)
Despite of the same normal vectors n and n
p
involved, the
amelet orientation factor s
y
(Eq. (179)) and the ame
surface wrinkling factor J (Eq. (97)) are different quanti-
ties. These relations, combined with Eqs. (100) and (177),
may relate the wrinkling length scale L
y
to the thickness d
B
of the mean ame brush, under the assumption of a innitely
thin ame front:
S = g
c(1 2 c)
s
y
L
y
= Ju7 cu < aJ
c(1 2 c)
d
B
(180)
where a is a model constant of the order of unity. Then:
L
y
<
g
a
_ _
1
un
p
nu
_ _
s
n
p
n ( )
s
d
B
(181)
Assuming that a, g and s
y
are constant, d
B
/J measures the
wrinkling length scale L
y
of the ame front.
The agreement between this BML model and the exper-
imental data is very good, as displayed on Fig. 22 where the
ratio S=[ c(1 2 c)]; corresponding to g=s
y
L
y
and assumed to
be constant, is displayed. Nevertheless, a submodel is
required for the wrinkling length scale which increases
with the downstream location in our experiment.
Estimating the local reaction rate per unit ame area
_
V
c
from the laminar ame speed S
L
(
_
V
c
= r
0
S
L
where r
0
is the
fresh gases density) the mean reaction rate becomes, when
n = 1 :

_ v = r
0
g
s
y
C
l
u
/
l
t
c(1 2 c) = r
0
g
s
y
C
l
u
/
l
t
(1 1t) ~ c(1 2 ~ c)
(1 1t~ c)
2
(182)
As t
t
= l
t
=u
/
is a turbulence characteristic time scale, the
mean reaction rate

_ v is proportional to the intermittency
between fresh and burnt gases, determined from ~ c(1 2 ~ c)
or c(1 2 c); and inversely proportional to t
t
. The physical
analysis leading to the Eddy-Break-Up model is recovered
and a similar expression for the reaction rate is found. The
previous expression is slightly different than the one
proposed in Section 7.3.3. The discrepancies are easily
explained by the crude model used for the scalar dissipation
of the progress variable to derive Eq. (162).
In Eq. (182), u
/
=l
t
is generally modeled from the
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 230
Fig. 21. Denition of the ame transit time t
t
is the ame crossing
frequency BML model.
turbulence k1 model as 1=k: The ITNFS efciency function
G
k
[127] may be introduced in the turbulent time scale to
account for the reduced ability of small turbulent structures
to wrinkle the ame front. ITNFS is one of the closure
schemes for the ame surface density balance equation
(see Section 7.4.3 and Table 5). The mean reaction rate is
then [118,128,129]:

_ v = ar
0
G
k
l
t
d
t
;
u
/
S
L
_ _
1
k
c(1 2 c) (183)
where a is a model constant and the efciency function G
K
depends on the length scale (l
t
/d
l
) and the velocity (u
/
=S
l
)
ratio comparing the turbulence and the laminar ame char-
acteristics. The efciency function has been tted from DNS
data [127,130]:
log
10
(G
k
) = 2
1
s 10:4
exp[2(s 10:4)]
1
_
1 2exp[2(s 10:4)]) s
1
u
/
S
L
_ _
s 20:11
_ _
(184)
where
s = log
10
l
t
d
l
_ _
and
s
1
u
/
S
L
_ _
=
2
3
1 2
1
2
exp 2
u
/
S
L
_ _
1=3
_ _ _ _ (185)
When the length scale ratio l
t
/d
t
tends towards zero, G
K
also
decreases, reducing the effective ame strain rate, as
displayed on Fig. 23. G
K
only slightly depends on the vel-
ocity ratio u
/
=S
L
: The efciency function does not reach a
constant level when l
t
/d
l
increases since increasing l
t
/d
l
,
keeping u
/
=S
L
constant, corresponds to an implicit increase
of the turbulent Reynolds number Re.
Because the efciency function G
K
tends to counter-
balance the known trend of Eddy-Break-Up modeling to
overestimate the mean reaction rate in highly strained
regions, this simple approach improves the accuracy of the
numerical predictions.
The BML model proposes a simple algebraic expression
to estimate the ame surface density s and the correspond-
ing reaction rate, but Bray, Moss, Champion and Libby
have mainly focused their attention on a careful description
of the turbulent uxes using the balance equations for the
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 231
Fig. 22. S=( c(1 2 c)) transverse proles in mm
21
corresponding to g=(s
y
L
y
) in the Bray, Champion, Libby model (Eq. (177)), plotted as the
function of the mean progress variable c for several locations downstream from the rod: (a) f = 0:78; (b) f = 0:9 [121].
Reynolds stresses r

u
//
i
u
//
j
and the scalar uxes r

u
//
i
c
//
to take
into account the possible occurrence of counter-gradient
transport and ame turbulence generation [131,132].
The ame surface density may also be derived from frac-
tal theories, leading to [133]
S =
1
L
outer
L
outer
L
inner
_ _
D22
(186)
where L
inner
and L
outer
are respectively the inner and outer
cut-off length scales (the ame surface is assumed to be
fractal between these two scales). D is the fractal dimension
of the ame surface. The cut-off scales are generally esti-
mated from the turbulence Kolmogorov l
k
and the integral l
t
length scales.
7.4.3. Flame surface density balance equation closures
The previously described balance equations for S (Eqs.
(89) and (90)) are unclosed and require modeling. In Table
5, various closures found in the literature are compared
where S
1
is the strain rate acting on the surface and induced
by the mean ow eld. S
2
is the strain rate due to the turbu-
lent motions (Eq. (94)) and the third source term, S
3
, occurs
only in the derivation proposed in Ref. [117]. D describes
the consumption of ame area. The modeled balance equa-
tion is rewritten as:
2S
2t
17(
~
US) = 7
n
t
s
S
7S
_ _
1S
1
1S
2
1S
3
2D (187)
In this expression, the turbulent ux of ame surface
density is expressed using a classical gradient assumption,
n
t
is the turbulent viscosity and s
S
a ame surface turbulent
Schmidt number. Five main closures are summarized:
The CPB model [134] is derived from the exact transport
equation for the ame surface density. The strain rate due
to the turbulent uctuations is estimated from the time
scale

n=1
_
of the Kolmogorov structures. The turbulent
strain rate is probably overestimated. Despite the fact that
the Kolmogorov structures contain the highest energy,
their lifetime is too short (because of viscous effects) to
actually affect the ame front [65].
The coherent ame model (CFM), developed by Candel
and his coworkers following the initial work of Marble of
Broadwell [84]. Three versions are presented in Table 5.
In the initial version (CFM1), the strain rate due to the
turbulent uctuations is estimated from the character-
istic time of the integral length scale (k/1). In the two
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 232
Table 5
Comparison of source (S
i
) and consumption (D) terms in the ame surface density balance equation in different turbulent premixed combustion
models (details in text). k and 1 denote, respectively, the turbulent kinetic energy and its dissipation. Re is the turbulent Reynolds number. a
0
,
b
0
, g, l, C
A
, a, d, C, E and K are model constants, G
k
is the efciency function in the ITNFS model [127] and depends on the length scale (l
t
/d
l
)
and the velocity (u
/
/S
L
) ratios comparing the turbulence and the laminar ame characteristics. In the Choi model (CH) [137], u
/
denotes the rms
turbulent velocity and l
tc
is an arbitrary length scale introduced for dimensional consistency and combined to a
0
as a single arbitrary constant
Model S
1
= k
m
S S
2
= k
t
S S
3
D
CPB [134] A
ik
2 u
k
2x
i
S a
0
C
A
_
1
n
_
1=2
S a
0
S
L
2 1e
2aR
3(1 2 c)
S
2
R =
(1 2 c)1
SS
L
k
CFM1 A
ik
2 u
k
2x
i
S a
0
1
k
S b
0
S
L
1C

k
_
1 2 c
S
2
CFM2-a A
ik
2 u
k
2x
i
S a
0
G
K
1
k
S b
0
S
L
1C

k
_
1 2 c
S
2
CFM2-b [130] A
ik
2 u
k
2x
i
S a
0
G
K
1
k
S b
0
S
L
1C

k
_
c(1 2 c)
S
2
MB [117] E
u
/
i
u
/
k
k
2 u
k
2x
i
S a
0

Re
_
1
k
S
F
S
L
1
k
u
/
i
c
/
2 c
2x
i
b
0
S
L

Re
_
c(1 2 c)
_
1 1d
S
L

k
_
_
2g
S
2
CD [136] a
0
l
1
k
S for k
t
# a
0
K
S
L
d
L
b
0
S
L
1 2 c
S
2
CH1 a
0
_
1
15n
_
1=2
S b
0
S
L
c(1 2 c)
S
2
CH2 [137] a
0
u
/
l
tc
S b
0
S
L
c(1 2 c)
S
2
succeeding formulations (CFM2), the expression of the
turbulent strain rate acting on the ame front is improved
from direct numerical simulations and multi-fractal
analysis (ITNFS model, Eq. (184) [127]). The destruction
term differs in these two last formulations (CFM2a and
CFM2b).
The MB model [117] is based on an exact equation for the
scalar dissipation rate x :
r ~ x = rD
2c
//
2x
i
2c
//
2x
i
(188)
assuming a constant density r. The transport equation is
rewritten as a ame surface density transport equation
under the amelet assumption (see Section 6.5, Eq.
(116)). This approach leads to a different expression for
the source term S
1
and an additional source term S
3
is
found. In a rst analysis, this term S
3
, which does not
depend on the available ame surface density S; might
be viewed as an ignition term [135] involving a gradient
of the mean fuel mass fraction or of the mean temperature
(i.e. fresh gases are ignited by heat transfers). However,
this analysis does not hold because the ame surface
density balance equation is derived assuming an
established ame surface. In fact, as shown below, S
3
corresponds to an anisotropic contribution of the tur-
bulent strain rate a
T
(see Eq. (205)). S
3
seems to be negli-
gible in practical simulations, at least to describe the
ame front propagation in a homogeneous and isotropic
turbulent ow eld.
The CD model [136] is similar to the rst version of the
coherent ame model (CFM1). An additional term is
proposed to take into account ame extinction under
excessively high strain rates. Such a term was tested in
the coherent ame model but the choice of the critical
strain rate is somewhat arbitrary. Moreover, this critical
strain rate cannot be deduced from planar strained
laminar ame studies because, due to curvature and
unsteady effects, a ame is able to sustain higher strain
rates than expected.
The CH model [137] has been devised for spark-ignited
engines to recover experimental data obtained in a closed
vessel [138]. The consumption term D is similar to the
one proposed in CFM2-b model whereas two formula-
tions of the strain rate induced by turbulent motions are
proposed. The rst expression (CH1) corresponds to the
closure in CPB model, based on the Kolmogorov turbu-
lent time scale. In CH2, the strain rate is only propor-
tional to the turbulence intensity u
/
and an arbitrary
length scale, l
tc
, is incorporated in the model constant.
Despite these comments, all these closures exhibit strong
similarities. For example, the consumption term D is always
proportional to S
2
. A comparison between these models to
predict turbulent ame speed S
T
may be found in Ref. [130].
In the case of a statistically one-dimensional turbulent ame
propagating in a frozen turbulence, a KPP (Kolmogorov
PetroskiPiskunov) analysis was used to analytically deter-
mine the turbulent ame speed S
T
as a function of the model
parameters.It was found that only the CFM-2 formulation is
able to predict the so-called bending phenomenon, where
the turbulent ame speed decreases before the occurrence of
ame extinction when the turbulence level increases, as
experimentally evidenced [110].
A recent work, [135] has compared CPB, CFM1, MB and
CD models to predict a turbulent premixed jet ame. The
CD predicts extremely high temperatures whereas CFM1,
MB and CPB provide reasonable predictions of mean vel-
ocities and temperatures. A slight overestimate (respectively
underestimate) of temperature is pointed out for CPB
(respectively MB) and is probably due to the expression
for the strain rate. The MB closure [117] is found to be
more sensitive to the inlet turbulent quantities than CFM1
but CFM2 models, incorporating an efciency function
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 233
Fig. 23. ITNFS efciency function G
K
(Eq. (184)) as a function of the length scale ratio l
t
/d
l
for several values of the velocity ratio u
/
/S
L
(0.1;
1.0; 10; 11).
including the inability of small vortices to wrinkle the ame
surface, have not been tested. Flame surface density models
are extended to non-isenthalpic ows (premixed reactants
injected in a co-ow of air [139]). A pioneering work has
been rst conducted [118] extending the algebraic ame
surface density BML model to non-isenthalpic ows,
using similar concepts. A presumed pdf for a mixture frac-
tion Z, determined from mean and rms values of Z, is
incorporated to account for dilution phenomena.
7.4.4. Analysis of the ame surface density balance equation
Following the description of the exact balance equation
for the ame surface density S (Section 6.3.2) and the
summary of the most popular closures (Section 7.4.3), the
aim of this section is to carefully analyze this balance
equation. This analysis may be based on direct numerical
simulations [17] or on experimental data [120,121].
Starting from the `propagative' form (Eq. (90)):
2S
2t
17( ~ uS) 17(ku
//
l
s
S)
= (7 ~ u 2knnl
s
: 7 u)
........,,........
A
T
S 1k 7u
//
2nn : 7u
//
l
s
........,,........
a
T
S
27[kwnl
s
S]
....,,....
propagation
1kw7nl
s
S
....,,....
curvature
(189)
each unclosed term may be investigated as follows:
7.4.4.1. Turbulent transport. The turbulent ux of ame
surface density is generally expressed using gradient
transport:
ku
//
l
s
S = 2
n
t
s
S
7S (190)
Bidaux and Bray (1994, unpublished work, see for example
[75,140]) have shown from a simple BML-type approach
that the turbulent uxes of the mean progress variable ( ~ c)
and of the ame surface density (S) are closely related.
Assuming that the conditional velocity on the ame
surface, k ~ ul
s
is a linear function of the conditional fresh
( u
u
) and burnt gases velocities:
k ~ ul
s
= u
u
1c
p
( u
b
2 u
u
) (191)
where c
p
denotes the c-level chosen to dene the ame front.
The BML relation (Eq. (169)):
~ u = (1 2 ~ c) u
u
1 ~ c u
b
(192)
leads to:
ku
//
l
s
= kul
s
2 ~ u = (c
p
2 ~ c)( u
b
2 u
u
) (193)
From Eq. (170) the ame surface density turbulent ux
becomes:
k

u
//
l
s
S =
(c
p
2 ~ c)
~ c(1 2 ~ c)

u
//
c
//
S (194)
This relation, conrmed by direct numerical simulations
[140], shows that turbulent uxes of ~ c and S are closely
related. A counter-gradient turbulent transport will be
observed simultaneously for these two scalar elds.
7.4.4.2. Strain rate induced by the mean ow eld, A
T
. The
only unclosed quantities in the strain rate due to the mean
ow eld, A
T
, are the orientation factors knnl
s
. Following
[134], the vector normal to the ame front, n may be split
into a mean component, M, and a uctuation, m:
n = M1m with knl
s
= M and kml
s
= 0 (195)
Then, the orientation factors become:
knnl
s
= MM1kmml
s
(196)
Using the denition nu7cu = 27c and assuming an
innitely thin ame front leads after averaging to [134,141]:
knl
s
S = MS = 27 c (197)
where c and ~ c are related using the BML relation r ~ c = r
b
c
(Eq. (138)). Then, only the uctuation cross products kmml
s
remain unclosed. Several closure schemes have been
proposed:
In Ref. [134] one assumes an isotropic distribution of the
uctuating components of the normal vector n:
km
i
m
j
l
s
=
d
ij
3
(1 2M
k
M
k
) (198)
In Ref. [117], a relation between orientation factors and
Reynolds stresses is indirectly proposed:
kn
i
n
j
l
s
=

u
//
i
u
//
j
2k
(199)
where k is the turbulent kinetic energy.
From experimental data [120], a much more complicated
model is proposed that will not be described here because
its practical implementation is probably not so easy.
These authors have also shown that the isotropic
assumption made in Ref. [134] is clearly wrong for the
turbulent ame stabilized downstream a small rod. On the
other hand, a slight modication of Ref. [117] leads to
very good results [121]:
kn
i
n
i
l
s
=

ki

u
//
k
2
4k
; kn
i
n
ji
l
s
=

u
//
i
u
//
j
2k
(200)
These orientation factors have very important effects and
may lead to surprising results. In the ame-holder stabilized
turbulent ame investigated in Ref. [121], the main velocity
gradient is the transverse gradient of the mean axial velocity,
corresponding to the mixing layer shear stress. But, because of
the low value of the corresponding orientation factor, its
contribution to the strain rate A
T
induced by the mean ow
eld is not that important. The strain rate A
T
is dominated by
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 234
the axial gradient of the mean axial velocity, induced by the
thermal expansion due to the combustion heat release, as
illustrated in Fig. 24. Notice also that the strain rate A
T
cannot be reduced to the simplied expression rstly
proposed [84]:
A
T
= u7
~
Uu (201)
Recent works [142] have shown that u7
~
Uu may probably be
viewed as a model for the total strain rate A
T
1a
T
and not only
for the strain rate A
T
due to the mean ow. This nding
explains why, in previous versions of the coherent ame
model where the orientation factors kn
i
n
j
l
s
were not
incorporated, the mean strain rate A
T
was not included
because the simple expression (201) clearly overestimates A
T
.
7.4.4.3. Strain rate a
T
due to turbulent motions. The source
term due to the strain rate a
T
is:
a
T
S = k7u
//
2nn : 7u
//
l
s
S = d
ij
2u
//
i
2x
j
2n
i
n
j
2u
//
i
2x
j
_ _
s
S
(202)
In most models, this term is generally assumed to be
proportional to the inverse of a turbulent time scale, either
the Kolmogorov time scale (CPB model) or the integral time
scale 1/k (CFM, MB and CD models). This turbulent time
may corrected with the efciency function, G
k
, of the ITNFS
closure (Eq. (184)). Nevertheless, a
T
is always modeled
being isotropic despite the orientation factors nn occurring
in expression (202).
Using the previous splitting of the normal vector to the
ame front, n, combined with the geometrical relation
(197), leads to:
a
T
S =
_
d
ij
2u
//
i
2x
j
_
s
S 2
1
S
_
2u
//
i
2x
j
_
s
2 c
2x
i
2 c
2x
j
1
_
m
j
2u
//
i
2x
j
_
s
2 c
2x
i
1
_
m
i
2u
//
i
2x
j
_
2 c
2x
j
2
_
m
i
m
j
2u
//
i
2x
j
_
s
S (203)
This interesting relation decomposes the source term due to
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 235
Fig. 24. (a) Transverse proles of the mean velocity gradients and the corresponding strain rate A
T
. (b) Components of the source term A
T
S due
to the strain rate induced by the mean ow eld. Data are plotted as a function of the transverse coordinate y for x = 80 mm and f = 0:9 [121].
the strain rate induced by turbulent motions, a
T
S, into two
parts: a contribution depending only on turbulent character-
istics that may be assumed to be isotropic (rst and last
terms) and an anisotropic contribution, where the mean
ame surface orientation occurs through the gradient of
the mean progress variable (second, third and fourth
terms). A priori, this last part is not proportional to the
available ame surface density S.
When velocity and normal vector uctuations are
supposed uncorrelated (this assumption is well veried in
direct numerical simulation data), the previous expression
may be reduced using:
m
j
2u
//
i
2x
j
_ _
s
= km
j
l
s
2u
//
i
2x
j
_ _
s
= 0 (204)
from the m
j
denition. Then:
a
T
S = d
ij
2u
//
i
2x
j
_ _
s
S 2 m
i
m
j
2u
//
i
2x
j
_ _
s
S2
...............,,...............
isotropic contribution
1
S
2u
//
i
2x
j
_ _
s
2 c
2x
i
2 c
2x
j
........,,........
anisotropic contribution
(205)
This simple analysis explains the third source term, S
3
found
in Ref. [117] from the derivation of a balance equation for
the scalar dissipation rate ~ x of the progress variable c. This
term S
3
corresponds to an anisotropic contribution depend-
ing on 7 c and on the strain rate due to turbulent motions.
Assuming a gradient type closure for the turbulent transport
of the fresh gases, S
3
is proportional to 7 c7 c and appears as
a model for the anisotropic contribution of the turbulent
strain rate. Nonetheless, Mantel and Borghi [117] have
shown that this term is negligible in numerical simulations
of a ame propagating in an homogeneous and isotropic
turbulent ow eld. This nding is probably questionable
in other congurations and further investigations are
required.
7.4.4.4. Propagation and curvature terms. These two terms
are analyzed together because they derive from the laminar
ame propagation and are related to the ame front
displacement speed w. The modeling of these terms
requires the description of the displacement speed w that
may have values far from the laminar ame speed S
L
[78].
This stands as the main difculty of the propagative
approach for the ame surface density balance equation
because various effects are incorporated in w (strain rate,
curvature effects, etc.).
Assuming a constant displacement speed w equal to the
laminar ame speed S
L
, the normal propagation term
becomes:
7[kwnl
s
S] = S
L
7[knl
s
S] = S
L
7[27 c] = 2S
L
7
2
c (206)
where the geometrical relation (197) is used.
This term is generally neglected in models (see Table 5).
Experimental data from Refs. [120,121] in a V-shape turbu-
lent ame show that it is not always negligible and is of the
same order as the curvature term.
The propagation/curvature term becomes:
kw7nl
s
S = S
L
k7nl
s
S (207)
where the only unknown is the mean ame front curvature
k7nl
s
The curvature is positive when the ame front is
convex towards the fresh gases, which is, a priori, the case
when the mean progress variable c is close to zero (see Fig.
25). On the other hand, curvatures are probably negative
(ame convex towards the burnt gases) when c < 1: As-
suming that the mean curvature is of the order of the inverse
of the wrinkling length scale L
y
, we have:
lim
c-0
k7nl
s
=
1
L
y
and lim
c-1
k7nl
s
= 2
1
L
y
Then, a simple linear model may be proposed:
k7nl
s
<
c
p
2 c
L
y
(208)
Replacing L
y
by its value as a function of the ame surface
density S from BML modeling (Eq. (177)) leads to the
following model for the curvature term in the S-equation:
kw7nl
s
S < bS
L
c
p
2 c
c(1 2 c)
S
2
(209)
where b is a model constant.
This term is positive in the fresh gas side of the turbulent
ame brush and becomes negative in the burnt gas side. This
trend is in agreement with the ndings of direct numerical
simulations [17] and experimental measurements [120,121],
as displayed in Fig. 26. The expression differs from classical
closure where kw7nl
s
is always negative (term D in
Table 5).
A linear relaxation is retained to model the mean curva-
ture k7nl
s
(Eq. (208)). This type of closure is common in
turbulent reacting ows and possesses similarities with the
relaxation model used for the scalar dissipation rate (Section
9.2) or for the pdf balance equation (Section 9.8.2). The
links between these closures are further discussed in Section
9.8.5.
The previous analysis was based on both a theoretical
analysis of the exact S-balance equation and experimental
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 236
Fig. 25. Flame front curvature analysis.
data [120,121]. An improved version of the closed balance
equation may be proposed:
2S
2t
17( ~ uS) 17(ku
//
l
s
S)
= (7 ~ u 2knnl
s
: 7~ u)S 1G
k
u
/
S
L
;
l
t
d
l
_ _
1
k
S 1S
L
7
2
c
1bS
L
(c
p
2 c)
S
2
c(1 2 c)
(210)
where G
k
is the ITNFS efciency function [127]. Orientation
factors, kn
i
n
j
l
s
may be described using a formulation
previously discussed. The turbulent ux ku
//
l
s
S is described
using a classical gradient expression.
7.4.5. Flame stabilization modeling
The ame surface density elds obtained from laser
tomography and the mean reaction rate estimated from
CH radical emission are compared in Fig. 27. According
to Eq. (85), reaction rate and ame surface density are
roughly proportional, excepted close to the stabilization
rod, the ame surface density S is high, whereas the mean
reaction rate remains low. In this zone, fresh and burnt gases
are separated by an interface (high surface densities) where
combustion has started but is not yet fully established. This
nding displays one of the difculties of ame surface
density models. In their present formulations, these models
are unable to describe the ame stabilization because the S-
equation (Eq. (189)) is derived assuming that the ame does
exist. All source terms in this equation are proportional to S
or to S
2
and the equation cannot generate ame surface
when there is no initial ame surface. In addition, an initia-
tion effect must also be incorporated to account for ignition
time delay in the local reaction rate per unit surface, V

i
(Eq.
(85)), to recover the observed results.
7.4.6. A related approach: G-equation
As described in Section 6.3.1, the premixed turbulent
ame may be described using a level set approach. Most
of the modeling issues discussed above are then recast in
terms of G-equation and modeling for the turbulent burning
velocity S
T
[89].
8. Turbulent transport in premixed combustion
8.1. Introduction
Turbulent uxes of the progress variable c, r

u
//
i
c
//
; are
generally modeled using a gradient transport hypothesis as
for inert scalars (Section 7):
r

u
//
i
c
//
= 2
m
t
Sc
t
2~ c
2x
i
_ _
(211)
where m
t
is the turbulent viscosity given by the turbulence
model and Sc
t
is a turbulent Schmidt number. Theoretical
[143] and experimental studies [19,20] have evidenced in
some turbulent premixed ames counter-gradient turbulent
transport where the turbulent uxes r

u
//
i
c
//
and the mean
progress variable ~ c gradient, 2~ c=2x
i
; have the same sign in
some regions and cannot be approximated with Eq. (211).
This phenomenon is known as counter-gradient turbulent
transport or counter-gradient turbulent diffusion.
This surprising nding was explained by the work of
Bray, Moss and Libby discussed in Section 7.3. In their
formulation, the turbulent uxes of the progress variable c
are directly connected to the conditional mean velocities in
fresh ( u
u
i
) and burnt gases ( u
b
i
) (Eq. (170) Section 7.3.5):
r

u
//
i
c
//
= r ~ c(1 2 ~ c) u
b
i
2 u
u
i
_ _
(212)
Even though conditional velocities are not obvious quanti-
ties, this expression is useful to explain counter-gradient
turbulent transport: Because of the thermal expansion due
to combustion heat release, the burnt gas conditional vel-
ocity, u
b
i
; is likely to be greater than the fresh gas conditional
velocity, u
u
i
: Then, in opposition with the modeled expres-
sion (Eq. (211)), the turbulent uxes of the progress variable
c have the same sign as the mean gradient (2~ c=2x
i
): Counter-
gradient transport also increases when the heat release factor
t, dened by Eq. (141), increases [144].
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 237
Fig. 26. Transverse proles of propagation (27[knl
s
S]), curvature
(k7nl
s
S) and combined (k7nl
s
S 27[knl
s
S]) terms (m
22
) in the
ame surface density balance equation plotted as a function of the
mean progress variable c: The consumption term modeled as 2S
2
/
(1 2 c) and the new proposed closure, 7
2
c 1b(c
p
2 c)S
2
= c(1 2 c)
where b = 0:4 and c
p
= 0:5 are also displayed. f = 0:90; x =
80 mm [121].
Fig. 27. Flame surface density (half top) and mean reaction rate,
estimated form CH radical emission (half bottom) elds are
compared for f = 0:90: Flame surface density data, extracted
from two different data sets, are not available from 30 to 70 mm
downstream the rod [121].
The various versions of this BML model propose an alge-
braic closure for the mean reaction rate of the progress
variable c, but focus attention on the scalar turbulent trans-
port (closure schemes for the turbulent uxes r

u
//
i
c
//
balance
equations), whereas other models may lead to more sophis-
ticated reaction rate formulations, retaining a simple gradi-
ent closure (Eq. (211)) for these uxes.
The occurrence of counter-gradient turbulent transport
have been analyzed using direct numerical simulation
(DNS) [140]. The results demonstrate the power of DNS
to help in the modeling of turbulent combustion. This
section is devoted to these results, obtained from a direct
solution without any closure models, of the instantaneous
balance equations.
8.2. Direct numerical simulation analysis of turbulent
transport
8.2.1. Introduction
Gradient and counter-gradient turbulent scalar transport
was observed in DNS of premixed ame/three-dimensional
turbulence interactions [145,146]. Then, simulations of two-
dimensional ame/turbulence interactions were reported
[140]. The reduced costs of 2D simulations allow the inves-
tigation of a large range of ame and turbulence parameters.
In these simulations, a planar laminar premixed ame is
rstly superimposed on a homogeneous isotropic turbulent
ow eld (Fig. 28). The ame front is progressively
wrinkled by turbulent motions, and the turbulence decays
in time. After a time of the order of the eddy-turnover time
of the largest turbulence structures, the ame may be
assumed to be in equilibrium with the turbulent ow eld
and relevant modeling information is extracted from DNS.
As the numerical conguration is statistically one-
dimensional in the propagating direction, quantities such
as mean progress variable or mean turbulent uxes may
be extracted from averaging in the perpendicular direction.
8.2.2. Results
Turbulent uxes ru
//
c
//
extracted from the direct numer-
ical simulations conducted by Trouve [145] (denoted CTR)
and Rutland [146] are displayed in Fig. 29. The rst data-
base clearly exhibits gradient turbulent transport, whereas
the second one corresponds to a counter-gradient situation.
The main discrepancy between the two databases lies in the
initial turbulence level which is higher in the CTR simula-
tion (u
/
0
=S
L
= 10) than in the Rutland database (u
/
0
=S
L
= 1):
Mean and conditional average velocities across the turbu-
lent ame brush are displayed in Fig. 30 for the two DNS
simulations. These results lead to the following comments.
As expected, in the Rutland database, the burnt gas con-
ditional velocity, u
b
; is higher than the fresh gas condi-
tional velocity, u
u
; leading to a counter-gradient turbulent
transport in agreement with expression (212).
On the other hand, in the CTR database, the fresh gas
conditional velocity is higher than the burnt gas con-
ditional velocity. This result is rst surprising: u
b
is
expected to be larger than u
u
because of thermal expan-
sion due to the heat release. However, this observation is
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 238
Fig. 28. Numerical conguration. A plane laminar premixed ame
is initially superimposed to an homogeneous and isotropic turbu-
lence.
Fig. 29. Turbulent ux ru
//
c
//
displayed as a function of the mean progress variable ~ c (bold lines) in the tri-dimensional direct numerical
simulations from CTR () and Rutland (- - -). Positive (respectively negative) values of ru
//
c
//
denote a counter-gradient (respectively
gradient) turbulent transport. The corresponding turbulent uxes estimated using the BML expression (Eq. (212)) are displayed for comparison
using thin lines. Velocities are made dimensionless using the laminar ame speed S
L
[140].
in agreement with expression (212) and a gradient turbu-
lent transport. Note that Eq. (212) overestimates the
turbulent ux

u
//
c
//
(see Fig. 29) because, due to the
ame thickness that must be resolved in the numerical
simulation, the progress variable c is not fully bimodal
(c = 0 or c = 1). In both cases, the BrayMossLibby
model remains able to predict the turbulent ux type
(gradient or counter-gradient).
The two-dimensional simulations conducted in Ref.
[140], referenced as CRCT in Fig. 31 have been used to
analyze the occurrence of counter-gradient turbulent
transport. The gradient turbulent transport is clearly
enhanced by an increase in the turbulence level u
/
/S
L
and
decreasing values of the heat release factor t. The
`increase', in terms of velocity ratio u
/
/S
L
, of the line
delimiting gradient and counter-gradient turbulent transport
when l
t
/d
l
decreases is due to the reduced ability of small
scale turbulence motions to wrinkle the ame front. This
phenomenon has already been discussed [65] and included
in the ITNFS model [127]. Three- and two-dimensional
direct numerical simulations lead to very similar results
for this problem.
8.3. Physical analysis
A simple physical analysis is now reported to derive a
criterion predicting the occurrence of counter-gradient
turbulent diffusion [140].
Following Bidaux and Bray (1994) (unpublished work
already presented in the BML model context Section
7.4.4), turbulent uxes of the progress variable c,

u
//
i
c
//
;
are directly connected to the surface-averaged uctuating
velocity, ku
//
i
l
s
(Eq. (194)). Thus, a model for

u
//
i
c
//
may be
deduced from a model for ku
//
i
l
s
involving the conditional
fresh and burnt gases mean velocities:
ku
//
i
l
s
= ku
i
l
s
2 ~ u
i
= (c
p
2 ~ c) u
b
i
2 u
u
i
_ _
(213)
In the following, the ow eld is assumed to be statistically
one-dimensional and only the turbulent transport in the
propagating direction,

u
//
c
//
will be described. Our analysis
is based on the two limiting cases pictured in Fig. 32:
Low turbulence level: The ame front remains smooth
and the velocity jump between fresh and burnt gases, u
b
2
u
u
; is determined primarily by thermal expansion and its
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 239
Fig. 30. Conditional velocities across the turbulent ame brush displayed as a function of the mean progress variable ~ c in the Rutland (top) and
the CTR (bottom) databases. Favre averaged velocity, ~ u (
), fresh gases conditional velocity, u
u
( ), burnt gases conditional velocity, u
b
()
and ame front conditional velocity, kuls
(- - -). Velocities are made dimensionless using the laminar ame speed S
L
[140].
value is close to the one obtained in a plane and laminar
ame ( u
b
2 u
u
< tS
L
): Eq. (213) becomes:
ku
//
l
s
= (c
p
2 ~ c)tS
L
(214)
High turbulence level: Due to strong viscous dissipation
of turbulent eddies in the hot burnt gas, the ame front
motions are assumed to be dominated by the turbulence
properties taken upstream of the ame.
At the leading edge of the turbulent ame (near ~ c = 0),
the ame front is convected towards the fresh gas with a
mean velocity estimated by 2u
/
(see Fig. 32). Then,
u
b
2 u
u
< 2u
/
(215)
where u
/
denotes the rms velocity in the fresh gases.
At the trailing edge of the ame brush ( ~ c < 1); the ame
front is convected by turbulent motions towards the burnt
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 240
Fig. 31. Premixed turbulent combustion diagram. The DNS ame-ow conditions are plotted as a function of the velocity ratio, u
/
/S
L
, and
length-scale ratio, l
t
/d
L
. The Classical KlimovWilliams criterion and the criterion from Ref. [65] are given to show the domain of validity of
amelet combustion. Also plotted are the DNS conditions of the Rutland (CR, t = 2:3) and CTR (t = 3) simulations. As the turbulence is
decaying in the CTR simulation, CTR conditions are displayed as an almost vertical line. The symbols W (t = 3) and A (t = 6) correspond to
the CRCT DNS. In 2D DNS, the turbulence decay is smaller and is not represented. Filled (open) symbols denote gradient (counter-gradient)
turbulent diffusion. The transition criterion, N
B
; ts
L
=2au
/
= 1 (Eq. (220)), separating CGD (below) from GD (above) is plotted for (t = 3)
and (t = 6) [140].
Fig. 32. Two limiting cases: counter-gradient transport promoted by thermal expansion (left); gradient transport due to turbulent motions
(right).
gases with a mean speed estimated by 1u
/
(see Fig. 32):
u
b
2 u
u
< 1u
/
(216)
leading to the simple linear model:
ku
//
l
s
= 22(c
p
2 ~ c)au
/
(217)
where a is an efciency function, similar to the ITNFS
model [127], accounting for the weak ability of small
scale turbulent motions to wrinkle and convect the
ame front. a is expected to be of order unity for large
turbulent length scales and vanishes when turbulent
eddies are too small to affect the ame front. The factor
2 has been introduced assuming c
p
< 0:5:
Then, modeling ku
//
l
s
as a sum of these two contributions
leads to:
ku
//
l
s
= (c
p
2 ~ c)(tS
L
22au
/
) (218)
and the turbulent ux becomes:

u
//
c
//
= ~ c(1 2 ~ c)(tS
L
22au
/
) (219)
This simple model is well veried in direct numerical simu-
lations [140] and has also been recovered when applying a
second order modeling (i.e. balance equations for turbulent
scalar uxes) to stagnating ames in the limit of small turbu-
lence intensities [131]. The turbulent ux may be viewed as
the sum of two contributions acting in opposite directions,
one induced by turbulent motions and the other by thermal
expansion. Then, the turbulent transport is analyzed as
follows: for a sufciently high turbulence level, the ame
is unable to impose its own dynamics to the ow eld and
the turbulent transport is of the gradient type, as for any
inert scalar. On the contrary, when the turbulence level
remains low, the thermal expansion due to heat release
dominates and the ame is able to impose its own dynamics
leading to a counter-gradient turbulent transport. Counter-
gradient turbulent diffusion occurs when

u
//
c
//
is positive
and expression (219) may be used to derive a criterion deli-
neating gradient and counter-gradient regimes. The Bray
number:
N
B
=
tS
L
2au
/
(220)
is greater (respectively lower) than unity when a counter-
gradient (respectively gradient) turbulent transport is
expected. This criterion is well veried by direct numerical
simulation results, as shown in Fig. 31. The efciency func-
tion a has been estimated from DNS (Fig. 33). Recent
experimental results [147] have conrmed these ndings.
8.3.1. Comments
Following Fig. 31 and criterion (220), in practical appli-
cations turbulent transport may be counter-gradient, or
close to the transition between gradient and counter-
gradient regimes. In many cases, the heat release factor
t is about 57 leading to a transition between gradient
and counter-gradient situations when u
/
=S
L
is of the order
of 3. Nevertheless, the possible occurrence of counter-
gradient transport is generally neglected in modeling.
The mean progress variable gradient may be estimated
as:
2~ c
2x
<
~ c(1 2 ~ c)
d
b
(221)
where a length scale d
b
characterizing the ame brush is
introduced. Then, the gradient type contribution in Eq.
(219) corresponds to a PrandtKolmogorov turbulence
modeling:
2~ c(1 2 ~ c)d
b
au
/
< 2ad
b

k
_
2~ c
2x
(222)
where k is the turbulent kinetic energy and u
/
=

k
_
.
Recent works [148,149] have reported regimes, corre-
sponding to low turbulence levels, where combustion
instabilities occur and wrinkle the ame front, acting in
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 241
Fig. 33. DNS estimated efciency function a (Eq. (21)) as a function of the length scale ratio l
t
/d
l
comparing the integral turbulent length scale
and the laminar ame thickness [140].
an opposite direction than counter-gradient diffusion
reducing the ame front wrinkling. These instabilities
should also be included in combustion models [150].
Expression (219) has been derived to analyze the occur-
rence of counter-gradient transport, but is not suited for
numerical simulations. For instance, such a simplied
formulation is unable to predict a gradient transport in
the vicinity of the leading edge ( ~ c < 0) always observed,
even in counter-gradient situations.
8.4. External pressure gradient effects
The previous analysis is completed by investigating the
effects of externally imposed pressure gradients on turbulent
premixed ames. Counter-gradient turbulent transport was
rst explained [19,151] by differential buoyancy effects of
pressure gradients on pockets of heavy and cold fresh gases
and on pockets of light and hot burnt gases. In many
combustion systems, ames are ducted and submitted to
strong pressure gradients due to thermal expansion and lead-
ing to ow accelerations. In Ref. [20], it is experimentally
shown that counter-gradient transports are enhanced in
ducted ames.
In Refs. [148,152], the same type of DNS as [140] was
conducted, introducing an externally imposed pressure
gradient (in fact, due to technical reasons, a constant
acceleration). The main conclusions, displayed in Figs. 34
and 35, are:
A favorable pressure gradient, i.e. a pressure decrease
from unburnt to burnt gases, is found to decrease the
ame wrinkling (see Fig. 34), the ame brush thickness,
and the turbulent ame speed S
T
(Fig. 35). It also
promotes counter-gradient turbulent transport.
On the other hand, adverse pressure gradients tend to
increase the ame brush thickness and turbulent ame
speed (Fig. 35), and enhance classical gradient turbulent
transport. As proposed in Ref. [153], the turbulent ame
speed is modied by a buoyancy term linearly dependent
on both the imposed pressure gradient and the integral
length scale l
t
.
A corrected Bray criterion (Eq. (220)) has been proposed
to account for the pressure gradient effects. We do not
give more details because this modied criterion needs to
be validated and improved (length scale effects are not so
clear and various analyses are possible, see [152]).
All these results suggest that counter-gradient turbulent
transport should be expected in most ducted turbulent ames
and some experimental observations are now reported.
8.5. Counter-gradient transportexperimental results
The V-shape turbulent ame was described in Section
7.4.1. The thermal expansion modies the ame dynamics
as is clearly apparent on high speed tomography lms
(scheme in Fig. 36, see also Ref. [121]): in the rst part
of the chamber (region 1), because of the rod wake, the
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 242
Fig. 34. Superimposed instantaneous temperature and vorticity
elds at the same time. (a) No imposed pressure gradientgradient
turbulent transport; (b) favorable imposed pressure gradient
counter-gradient turbulent transport. A planar laminar ame
separating fresh gases (left) from burnt gases is initially super-
imposed to an homogeneous and isotropic turbulent ow eld
(u
/
0
=S
L
= 5) [152].
Table 6
Geometrical analysis of the scalar turbulent transport in a V-shaped premixed ame (see Fig. 36). G and CG denote, respectively, a gradient and
a counter-gradient turbulent diffusion
Zone u
b
2 u
u
2~ c
2x

u
//
c
//
v
b
2 v
u
(when y .0)
2~ c
2y
(when y .0)

v
//
c
//
1 ,0 .0 G .0 ,0 G
2 .0 .0 CG .0 ,0 G
coherent structures embedding the ame front turn clock-
wise (counter-clockwise) in the upper (lower) ame sheet,
as in classical Von Karman vortex streets, except for
their symmetry due to pressure waves. When the center-
line velocity increases, because of thermal expansion in
burnt gases (region 2), the upper (lower) coherent struc-
tures start to turn counter-clockwise (clockwise). This
phenomenon may be recast in terms of turbulent trans-
port using a simple geometrical analysis based on the
BrayMossLibby relation (212) and summarized in
Table 6. Accordingly, the transverse turbulent ux

v
//
c
//
is always of gradient type, but the change in rota-
tion of the coherent structures corresponds to a transi-
tion between gradient and counter-gradient transport for
the downstream turbulent ux

u
//
c
//
: The turbulent uxes
are of gradient type just behind the rod, as expected to
ensure the stabilization of the ame, and becomes of
counter-gradient type further downstream.
8.6. To include counter-gradient turbulent transport in
modeling
Except for the work of Bray, Moss, Libby and their
co-workers [19,128,129,132,151], very few works have
been devoted to the effective modeling of counter-
gradient transport. The approach of Bray et al. is mainly
based on second order modeling, using balance equations
for the turbulent uxes

u
//
i
c
//
: These equations are easily
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 243
Fig. 35. Reduced turbulent ame speed S
T
/S
L
plotted as a function of the reduced time t=t
f
= S
L
t=d
L
; where t
f
= d
L
=S
L
is a ame time, for
different values of the externally imposed pressure gradient. (a) initial turbulent level u
/
0
=S
L
= 5; no pressure gradient () and increasing
favorable pressure gradient ( and - - -); (b) initial turbulent level u
/
0
=S
L
= 2; no pressure gradient () and increasing adverse pressure
gradient ( and - - -). Markers correspond to the Libby theory [152,153].
Fig. 36. A rough scheme of the coherent structures dynamics
observed using laser tomography [121].
derived from momentum and progress variable balance
equations:
2 r

u
//
i
c
//
2t
...,,...
(I)
1
2 r ~ u
j

u
//
i
c
//
2x
j
....,,....
(II)
= 2
2 r

u
//
j
u
//
i
c
//
2x
j
....,,....
(III)
2 r

u
//
i
u
//
j
2~ c
2x
j
....,,....
(IV)
2r

u
//
j
c
//
2~ u
i
2x
j
....,,....
(V)
2 c
//
2 p
2x
i
..,,..
(VI)
2 c
//
2p
/
2x
i
..,,..
(VII)
2u
//
i
2J
k
2x
k
...,,...
(VIII)
c
//
2t
ik
2x
k
..,,..
(IX)
1 r

u
//
i
_ v
..,,..
(X)
(223)
where the RHS terms correspond, respectively, to turbulent
transport of r

u
//
i
c
//
(III), ~ c-gradient effects (IV), mean vel-
ocity gradient effects (V), the action of mean (VI) and uc-
tuating (VII) pressure gradients, r

u
//
i
c
//
turbulent dissipation
(VIII and IX) and reaction rate (X). This balance equation is,
of course, unclosed and each term may be extracted from
direct numerical simulations [132,140,152].
The discussion of the closure schemes for this equation is
beyond the scope of the present review and the reader may
nd relevant information in Refs. [128,132,151]. Some
comments may be made:
The

u
//
i
c
//
turbulent transport (III) is generally modeled
using a classical gradient expression (gradient turbulent
transport at the third order).
Mean progress variable gradient terms (IV) needs
Reynolds stress modeling for

u
//
i
u
//
j
: Then, new balance
equations are derived, and closed, for these quantities
(second order turbulence model).
The mean velocity gradient term (V) is closed because
turbulent uxes

u
//
j
c
//
are provided by their balance
equations.
The mean pressure gradient term (VI) is easily known
under a BML assumption (Section 7.3). Making use of
Eq. (165) yields:
c
//
= c 2 ~ c = t
~ c(1 2 ~ c)
1 1t~ c
(224)
The uctuating pressure term (VII) is more difcult to
understand and to model. In Ref. [151], this term is
neglected, however this assumption is not supported by
DNS results [152]. The mean pressure term (VI) and the
uctuatingpressureterm(VII) probablyneedtobe modeled
together as c
//
2p=2x
i
: In Ref. [128], the pressure gradient
termclosure is carefullydiscussed. Recently, a model based
on a partitioning of each pressure uctuation covariance
into contributions from reactants, products and thin ame-
lets was proposed [132]. The comparison of this new
model with DNS results are encouraging and conrm the
importance of the intermittency between the conditional
mean pressure gradients in reactants and products.
Turbulent dissipation terms (VIII and IX) are generally
expressed together using small scale dissipation rate
assumptions [151].
Second order closures for turbulent uxes and Reynolds
stresses require nine additional balance equations in 3D simu-
lations (three for progress variable uxes,

u
//
i
c
//
; and six for
Reynolds stresses,

u
//
i
u
//
j
). Because of very high computa-
tional costs, model closures and implementation difculties,
very few simulations have been conducted using the second
order formulation. Recently [129], very promising results
were obtained, particularly in predicting experimental results
[20].
8.7. Towards a conditional turbulence modeling?
The BML formulation (Section 7.3.5) directly distin-
guishes between the properties of fresh and burnt gases,
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 244
Fig. 37. Axial velocity u histograms for four transverse locations 80 mm downstream the rod in the V-shape turbulent ame (Fig. 19): burnt
gases (y = 0 mm; --); reaction zone (y = 5 mm; and y = 12 mm;
); unburnt gases (y = 18 mm; ) [121].
this was achieved using conditional averaging (Eq. (168)).
This approach is very attractive because turbulence charac-
teristics may differ in fresh and in burnt gases as shown in
Fig. 37, where velocity histograms obtained using laser
Doppler velocimetry in the V-shape turbulent ame
(Section 7.4.1) are displayed.
The velocity distribution is almost Gaussian in the
unburnt and the burnt gases, but becomes clearly bimodal
in the reaction zone denoting an intermittency between fresh
and burnt gases, according to Eq. (172). This bimodal distri-
bution does not correspond to the assumptions made in the
derivation of most turbulence models, such as k1.
A conditional approach to determine the conditional
averages of a quantity Q in fresh (

Q
u
) and burnt (

Q
b
)
gases is probably a promising way leading to a straight-
forward description of turbulent transport (Eqs. (170) and
(172)). The objective is then to derive balance equations for
quantities such as:
rcQ = r ~ c

Q
b
; r(1 2c)Q = r(1 2 ~ c)

Q
u
which is easy from c and Q balance equations [141]. These
equations remain to be closed and a few attempts have been
conducted in this direction [91,148,154], but no clear
conclusions can be found in the literature.
All these works devoted to turbulent transport suggest
that turbulent combustion modeling might probably be
greatly improved by advancing the description of turbulent
transport itself. This point motivates large eddy simulation
(LES) for turbulent combustion modeling [155].
9. Reynolds averaged models for non-premixed
turbulent combustion
9.1. Introduction
Much work has been devoted to the numerical modeling
of non-premixed combustion systems, mainly assuming a
chemistry much faster than mixing and molecular diffusion.
This `mixed is burnt' regime is easily described from the
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 245
Table 7
Modeling strategies for non-premixed turbulent ames. Z is the mixture fraction, x = Du7Zu
2
its scalar dissipation rate,
~
P denotes pdfs, V

, an
integrated amount of heat release, S, the density of ame surface and

Y
C
i
; a conditional moment. PaSR stands for partially stirred reactor,
turbulent micromixing was expressed with a linear relaxation (IEM-LMSE) D7
2
Y
i
< (
~
Y
i
2Y
i
)=t
t
, where t
t
< (k=1) is an eddy break-up mixing
time; other formulation exist [96] (from Ref. [55])
Major assumptions Flame structure in mixture
fraction space
Turbulent combustion model
Innitely fast chemistry Species are function of Z:
Y
i
= Y
i
(Z
p
)
Presumed Pdf:
~
P(Z
p
) from
~
Z and

Z
//2
Chemical equilibrium [35]

Y
i
=
_
Z
Y
i
(Z
p
)
~
P(Z
p
)dZ
p
1D strained reaction zones Species are functions of Z and x:
Y
i
= Y
i
(Z
p
; x
p
(t))
Flamelet modeling:
~
Y
i
=
_
Z
_
x
Y
i
(Z
p
; x
p
)
~
P(Z
p
; x
p
)dZ
p
dx
p
[165]
Diffusion balances chemistry
as in a 1D counter-ow ame
Steady [63] or unsteady [177]
Single-step or complex
chemistry
Total heat release is a function of x:
_
V(x
st
) =
_
11
21
_ v
T
(j; x
st
)dj
Coherent ame model:
Solve for the density of ame surface S
Constant (Le
i
= 1) or
complex transport properties
(Le
i
1)
j coordinate across the amelet [88]

_ v
T
=
_
V( x)S [84]
Diffusion is captured via
micromixing modelling
Solve for conditional moments:
Y
C
i
(Z
p
) =
_
Y
i
uZ = Z
p
_
Conditional moment closure:
presumed pdf

P(Z
p
) [102,103]
~
Y
i
=
_
1
0

Y
C
i
(Z
p
)
~
P(Z
p
)dZ
p
Simple transport properties
(Le
i
= 1)
Solve for representative PaSR:
dY
i
=dZ = (
~
Y
i
2Y
i
1t
t
_ v
PaSR
i
)=(
~
Z 2Z)
to get _ v
i
(Z; t
t
)
PaSR
Presumed pdf 1PaSR modelling:
MIL/PEUL [68,191]

_ v
i
=

k
_
Z
( _ v
PaSR
i
(Z
p
; t
p
t
))
k

P(Z
p
; t
p
t
)dZ
p
dt
t
Single-step or complex
chemistry
Solve for Monte-Carlo particles:
dY
k
i
=dt = (
~
Y
i
2Y
k
i
)=t
t
1 _ v
k
i
to get
~
P(Y
p
1
; ; Y
p
n
)
Pdf methods:
~
Y
i
=
_
Y
1

_
Y
n
Y
p
i
~
P(Y
p
1
; ; Y
p
n
)dY
p
1
dY
p
n
[93]
Chemical source is closed
(only for pdf methods)
turbulent mixing of conserved scalars (Section 3.2) and
mixing problems have been the subject of many discussions
[156158].
There exist strong motivations for improving non-premixed
and partially premixed turbulent combustion modeling:
The development of new combustion technologies for
aircraft engines, and more generally for gas turbines
operating in the non-premixed regime, implies the accu-
rate determination of the position in the ow where
combustion starts and the control of pollutants emission.
Crucial points which cannot be addressed using the in-
nitely fast chemistry hypothesis.
Many practical systems include liquid injection of the
fuel, followed by non-premixed and partially premixed
combustion.
Even in burners operating in the premixed regime, the
premixing of the reactants is not always complete at the
molecular level and some partial premixing may be
observed. Sometimes, partial premixing is even desirable
to limit pollutant emissions (stratied charge engines).
As for premixed combustion (Sections 7 and 8), the
modeling of turbulent diffusion ames relies on simplifying
assumptions for both chemistry and transport. Depending on
the simplications made for these mechanisms, various
approaches for laminar ames and models for turbulent
ames are obtained (see Table 7). Hypotheses formulated
to construct models for non-premixed turbulent ames may
be organized into three major groups:
Assumption of innitely fast chemistry (mixed is burnt).
Finite rate chemistry assuming a local diffusivereactive
budget similar to the one observed in laminar ames
(amelet assumption).
Finite rate chemistry with treatment of molecular and
heat transport separated from chemical reaction (CMC,
pdf method; Section 6). Diffusion is then addressed using
turbulent micromixing modeling, while chemical sources
can be dealt with in an exact and closed form (only for
pdf).
The proposed closures are all based on a particular
description of turbulent mixing, hence the basic concepts
useful to capture fuel/air turbulent mixing are rst
presented.
9.2. Fuel/air mixing modeling
9.2.1. Introduction
The mean value of the mixture fraction
~
Z gives an indica-
tion of the local mean fuel/air mixing in turbulent ows. In
addition, the structure and the properties of the ame depend
on

Z
//2
; measuring the degree of mixing between reactants.
A simple description of turbulent mixing is thus obtained
from the two elds:
~
Z and

Z
//2
: Introducing the classical
gradient transport closure for the turbulent uxes (Section
6), one may write:
r
2
~
Z
2t
1 r ~ u7
~
Z = rn
t
7
2
~
Z (225)
r
2

Z
//2
2t
1 r ~ u7

Z
// 2
= rn
t
7
2
Z
//2
12 rn
t
u7
~
Zu
2
....,,....
Production
2 2 r ~ x
.,,.
Dissipation
(226)
The rst term on the RHS is the turbulent transport, the
second in Eq. (226) is the production of uctuations by
the mean gradients, the last is the scalar dissipation rate ~ x
remaining unclosed. r ~ x = rDu7Zu
2
was discussed for
premixed turbulent ames, in the EBU, BML and S models
(Section 7).
9.2.2. Balance equation and simple relaxation model for ~ x
A transport equation may be derived for ~ x from the Z
balance equation (Eq. (41)):
2 r ~ x
2t
1
2 r ~ u
j
~ x
2x
j
1
2
2x
j
rDu
//
j
2Z
//
2x
i
2Z
//
2x
i
_ _
= 22
2
2x
j
rDu
//
j
2Z
//
2x
i
2
~
Z
2x
i
_ _
.........,,.........
(I)
22
2
~
Z
2x
i
rD
2Z
//
2x
j
2u
//
j
2x
i
_ _
........,,........
(II)
22
2
~
Z
2x
j
rD
2Z
//
2x
i
2u
//
j
2x
i
_ _
........,,........
(III)
22
2~ u
j
2x
i
rD
2Z
//
2x
i
2Z
//
2x
j
_ _
.........,,.........
(IV)
22 rD
2u
//
j
2x
i
2Z
//
2x
i
2Z
//
2x
j
.......,,.......
(V)
22 D
2
r
2
2x
j
2Z
//
2x
i
_ _
2
2x
j
2Z
//
2x
i
_ _
............,,............
(VI)
2
2
r
D
2Z
2x
i
2r
2x
i
2
2x
j
rD
2Z
2x
j
_ _
...........,,...........
(VII)
(227)
The term (I) on the RHS describes curvature effects of the
mean mixture fraction eld, and as (II) and (III), contains
correlations between the Z eld and the velocity eld.
Compared to other terms, these correlations vanish for suf-
ciently large Reynolds numbers. (IV) corresponds to the
correlations between mean ow motion and uctuations.
(V) is the strain rate of the scalar Z eld, already discussed
for ame surface density modeling in premixed ames
(Section 7.4.4). (VI) is the dissipation rate of the scalar
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 246
dissipation rate and (VII) is negligible when density gradi-
ents are kept small.
To derive a linear relaxation closure for the scalar dissi-
pation rate, a homogeneous Z steady eld in equilibrium
(production = dissipation) is rst considered. Then,
Eq. (227) reduces to a balance between the straining rate
of Z and the dissipation rate of the scalar dissipation rate:
22rD
2u
//
j
2x
i
2Z
//
2x
i
2Z
//
2x
j
= 2D
2
r
2
2x
j
2Z
//
2x
i
_ _
2
2x
j
2Z
//
2x
i
_ _
(228)
the uctuating velocity gradient is assumed to be of the
order of the inverse of the small scales characteristic time,
i.e. (2u
//
j
=2x
i
) , (1=t
k
) , (1=n)
1=2
: The remaining part of the
strain rate term is proportional to ~ x; then,
22rD
2u
//
j
2x
i
2Z
//
2x
i
2Z
//
2x
j
,
r ~ x
t
k
, R
l
r ~ x
(k=1)
where R
l
is a Reynolds number based on the Taylor
microscale. A quadratic behavior in ~ x; with a direct depen-
dence on (

Z
//2
)
21
is anticipated for the dissipation rate of the
scalar dissipation rate:
2rD
2
2
2x
j
2Z
//
2x
i
_ _
2
2x
j
2Z
//
2x
i
_ _
, R
l
r ~ x
2

Z
//2
The equilibrium condition (Eq. (228)) leads to:
~ x =

Z
//2
(k=1)
(229)
recovering the widely used linear relaxation model.
Eq. (229) is a very simplied description of micromixing.
The scalar dissipation rate depends on the detail of the char-
acteristics of the turbulence (velocity and scalar energetic
spectrum), the linear relaxation is thus a rst approximation
that may be improved, for instance, by solving a modeled
balanced equation for ~ x: Such a closed equation may be
found in Ref. [159].
The knowledge of
~
Z and

Z
//2
from their closed transport
equation may be used to presume the pdf of Z,
~
P(Z
p
; x; t);
with a b-shape (see Section 6.4.2). Because the internal
structure of diffusion ames is easily obtained from condi-
tional statistics as
_
Y
i
uZ = Z
p
_
(Section 3.2), many models
(amelet, CMC) incorporate a b-pdf to calculate means
quantities:
~
Y
i
(x; t) =
_1
0
(Y
i
uZ = Z
p
)
....,,....
Flame structure
~
P(Z
p
; x; t)
,........,
Mixing
dZ
p
(230)
The simplest of these models invokes the innitely fast
chemistry assumption.
9.3. Models assuming innitely fast chemistry
9.3.1. Eddy dissipation model
The eddy dissipation model (EDC) is a direct extension to
non-premixed ame of the eddy break up (EBU) closure,
initially devoted to turbulent premixed combustion [160]
(Section 7.2). The fuel burning rate is calculated according
to:
r _ v
F
= a r
1
k
min
~
Y
F
;
~
Y
O
s
; b
~
Y
P
(1 1s)
_ _
(231)
where a and b are adjustable parameters of the closure. In
Eq. (231), the reaction rate is limited by a decient species.
To account for the existence of burnt gases bringing the
energy to ignite the fresh reactants, this species may be
the reaction products. A priori, this model does not respect
the response of diffusion combustion in mixture fraction
space and may generate mean fuel mass fraction values
lower than
~
Y
IFCM
F
: Actually, for large a, Eq. (231) is difcult
to justify.
9.3.2. Presumed pdf: innitely fast chemistry model
One of the rst descriptions of non-premixed combustion
was given in Ref. [35] which assumed an innitely fast
single step chemical reaction (Section 3.2). Considering
the piecewise relations (Fig. 5, Table 1):
Y
F
= Y
IFCM
F
(Z); Y
O
= Y
IFCM
O
(Z); T = T
IFCM
(Z) (232)
relating the fuel and oxidizer mass fractions, and the
temperature to the mixture fraction in the case of innitely
fast chemistry, mean quantities may be directly obtained
with a b-pdf (Section 6.4.2) and Eq. (230):
~
Y
IFCM
F
=
_1
0
Y
IFCM
F
(Z
p
)
~
P(Z
p
; x; t)dZ
p
(233)
~
Y
IFCM
O
=
_1
0
Y
IFCM
O
(Z
p
)
~
P(Z
p
; x; t)dZ
p
(234)
~
T
IFCM
=
_1
0
T
IFCM
(Z
p
)
~
P(Z
p
; x; t)dZ
p
(235)
The mean ame structure is then known from an innitely
fast chemistry model (IFCM) without solving balance equa-
tions for mean thermochemical quantities. The inputs of this
`mixed is burnt' closure are
~
Z and

Z
//2
determining the b-
pdf chosen to capture the detail of fuel/air mixing. IFCM is
therefore a two-equations model for non-premixed turbulent
combustion. It is very popular and is usually coupled with a
low Mach number solution of the turbulent ow [161], so
that the thermodynamics of the ow is fully known from Eq.
(235). IFCM is an interesting rst guess to provide the
global ame structure for the maximum of heat that can
be released, a valuable result for certain aspects of design.
Note that the piecewise relations are exact when the chemi-
cal time is innitely small. Therefore, for very large
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 247
Damkohler numbers (Section 3.2), IFCM is an accurate
description of a turbulent diffusion ame. Unfortunately
such a zero order model does not exist for premixed turbu-
lent combustion. To handle multi-step chemistry, innitely
fast chemistry may be replaced by a chemical equilibrium
condition [162].
This model lacks any prediction capacities when ignition,
quenching or even small nite rate chemistry effects exist.
9.4. Flamelet modeling
9.4.1. Introduction
Experiments in jets ames and direct numerical simula-
tions suggest that there exist situations in burners where the
chemistry is fast, but not innitely fast (see for example
[163,164]). In these measurements and calculations, the
response of the ame in mixture fraction space lies in the
vicinity of the curves given by Y
IFCM
i
(Fig. 38).
Models have been proposed for such ames [63,165]. For
a given state of mixing in the turbulent ow, thus given
values of Z and x, amelet models are derived assuming
that the local balance between diffusion and reaction is
similar to the one found in a prototype laminar ame for
the same values of Z and x. Flamelet models are therefore
constructed from an asymptotic view of diffusive-reaction
layers as given by Fig. 6. The two control parameters of
planar and steady laminar strained ames are used: the
mixture fraction Z and its scalar dissipation rate x (Section
3.2). In a turbulent ow, these two quantities uctuate in
space and time, but when the joint pdf
~
P(Z
p
; x
p
; x; t) is
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 248
Fig. 38. Scatter plots of major species mole fractions and temperature as functions of mixture fraction: (a) Data from three-dimensional direct
numerical simulations of Montgomery et al. [164]; (b) Raman scattering measurements of Barlow et al. [163] in H
2
/argonair ame. Solid lines
indicate chemical equilibrium.
known, the mean properties of the ame may be calculated
as:
~
Y
i
=
_
Z
p
_
x
p
Y
SLFM
i
(Z
p
; x
p
)
~
P(Z
p
; x
p
; x; t)dx
p
dZ
p
(236)
Y
SLFM
i
(Z
p
; x
p
) is the local ame structure in mixture fraction
space and
~
P(Z
p
; x
p
; x; t) captures the statistics of fuel/air
mixing. SLFM stands for steady laminar amelet model.
This model may be viewed as a direct improvement of the
innitely fast chemistry model (IFCM), since it uses the
same formalism, but with an additional parameter: the scalar
dissipation rate x, thereby including nite rate chemistry
effects. For a given chemistry, and therefore a given chemi-
cal time t
c
, when the Damkohler number Da
p
= (t
c
x
st
)
21
is
large, IFCM is recovered. An increase in x is followed by
nite rate chemistry effects, or even quenching when x
becomes too large (Fig. 6), then, Y
SLFM
i
(Z
p
; x
p
) features
mixing without reaction (Fig. 5).
The inputs of SLFM are similar to those of IFCM:
~
Z;

Z
//2
to which ~ x is added. Two issues emerge:
1. Y
SLFM
i
(Z
p
; x
p
) must be determined and tabulated under
particular hypothesis, choosing a given laminar ame
prototype.
2.
~
P(Z
p
; x
p
; x; t) must be presumed using the mean values
available to quantify fuel/air mixing (i.e.
~
Z;

Z
//2
and ~ x).
9.4.2. Flame structure in composition space, Y
SLFM
i
(Z
p
; x
p
)
Y
SLFM
i
(Z
p
; x
p
) may be tabulated from solutions of coun-
ter-ow diffusion ames (Fig. 4 [63]), a ame conguration
widely studied experimentally [166,167]. Assuming within
the turbulent ow thin quasi-one-dimensional structures
convected and stretched by the uid motions, and neglecting
higher order terms, the equations for the species and
temperature become (Section 3.2):
2Y
i
2t
= _ v
i
1
x
Le
i
_ _
2
2
Y
i
2Z
2
_ _
2T
2t
= 2

N
n=1
h
n
_ v
n
C
p
1x
2
2
T
2Z
2
_ _
where x = (l=rC
p
)u7Zu
2
:
These equations were used to discuss laminar diffusion
ames (see Eqs. (47) and (50) and Section 3.2). Omitting the
time derivative (steady amelet), for a given value of x
corresponding to local micromixing conditions, one has to
solve for:
_ v = 2
x
Le
i
2
2
Y
i
2Z
2
(237)
The solution of this equation for given concentrations and
temperatures boundary conditions, and various x provides a
Flamelet Library Y
i
(Z,x). A variety of techniques are avail-
able to build these libraries [168].
In SLFM, the characteristic time required to balance diffu-
sion and reaction is assumed to be much smaller than any
other owtime scale of the problem. The hypotheses involved
in SLFM can be disputed, leading to various improvements:
Eq. (237) has been obtained neglecting diffusion in the
direction tangential to the iso-Z
st
surface, arguing that
when the mixing element is sufciently thin and features
weak curvature, the gradients measured along the stoi-
chiometric surface are much smaller than those in the
perpendicular direction (Section 3.2). Thus, when using
Eq. (237) to describe a prototype ame of turbulent
amelets, one supposes that the mixing eld may be
reduced to a steady one-dimensional structure. In conse-
quence, the validity of Eq. (237) in a turbulent ow also
depends on the properties of micromixing, and up to now,
it is not obvious how to draw conclusions since one
would need to measure the scalar dissipation rate in
turbulent ames. Some experimental results are available
[169], but more works are required to conclude on the
dimensionality of scalar micromixing in ames.
There are other issues related to the multi-dimensional
character of diffusion ames. Straining cannot be
uniformly distributed along the ame sheet, leading to
amelet interactions when the distribution of x is non-
uniform on the iso-Z
st
. This transverse loss or gain of heat
modies the structure of the amelet in the normal direc-
tion to the stoichiometric surface [55].
Reference states at innity used to tabulate the amelets
may have to account for partial premixing [170,171].
Consider the simple case of a jet ame, where close to
the nozzle inlet, pure fuel and pure air react to form
products. Moving downstream, turbulent diffusion
mixes these products with air, on the air side, and with
fuel, on the fuel side. Further downstream, the reactants
feeding the reaction zone are not likely to be either pure
fuel or pure air. This situation is strongly enhanced in
ows where recirculation zones are found to stabilize
combustion. In Ref. [171], this shortcoming of SLFM is
overcome introducing transient amelets, for which
reference states at innity vary according to the value
of a progress variable ~ c:
Flamelet libraries can be calculated in physical space or
in mixture fraction space. It is usually observed that
the decay of OH towards equilibrium is predicted by
the amelet solution in mixture fraction space using the
scalar dissipation rate as a control parameter sensitive to
species boundary conditions, a trend that is not fully
reproduced by the amelet solution in physical space
where the input parameter is the strain rate [171].
Unsteadiness is also an important aspect. Time-depen-
dent amelets have been used to include unsteady effects
[172174]. As shown in Ref. [175], when a high value of
the scalar dissipation rate is imposed to the ame for a
sufciently short period of time, extinction may not be
completed, with a limiting frequency at which the ame
almost behaves like a steady state ame. Unsteady ame-
lets were used [176] to simulate extinction and re-ignition
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 249
in a turbulent jet ame, history effects were included
using a Lagrangian time measured along the stoichio-
metric line. This work was pursued [177] introducing
as an additional control parameter, the diffusion time
needed to exchange mass and energy over a distance
DZ in mixture fraction space.
This is more generally related to questions arising
concerning the determination of quenching limits along
with the accuracy of quenching predictions using amelet
theory a problem which can be addressed numerically,
for instance by post-processing DNS databases to study
the reactive/diffusive layer in terms of amelets
[70,74,178181]. Numerical simulations of ame/vortex
interaction, used to construct a combustion diagram
([70], Fig. 10), show that when strong unsteadiness
and/or curvature effects appear, the laminar amelet
assumption does not always predict quenching (Fig.
10). Two critical Damkohler numbers, Da
LFA
and Da
ext
,
are easily derived from the simulations to nd the limit
conditions where unsteady (Da
LFA
) and extinction (Da
ext
)
effects become important. These numbers are then
compared with Da
p
q
; the quenching Damkohler derived
from asymptotic analysis [30]. It is found that
(Da
LFA
=Da
p
q
) < 2 while (Da
ext
=Da
p
q
) < 0:4: Therefore,
Eq. (237) is an interesting approximation for Da
p
larger
than Da
p
q
; twice the asymptotic quenching value in these
simulations. Below this value, even in the simple cong-
uration of ame/vortex interaction, the time evolution of
the reaction zone is fully dominated by unsteady effects.
Another related point of interest in amelet theory is the
response of the turbulent ame when quenching zones
develop. In other words, one may discuss the assumption
that the occurrence of local extinction at some points
does not prevent the use of a amelet model for the
remaining part of the turbulent ame. Again using
DNS, constant density ames near extinction were
studied [182]. A critical Damkohler number at which
extinction occurs is determined as a function of a ame
thickness parameter, dened as the ratio between the rms
mixture fraction and the reaction zone width in mixture
fraction space. As expected for these ames featuring
strong unsteadiness effects, the value of the critical
extinction Damkohler number was different from the
one predicted by laminar ame theory. This observation
was explained by statistical variability. Consequently, an
extinction may be observed with Damkohler numbers
larger or smaller than Da
p
q
: In this last study, the lower
value of Da
ext
differs from that of steady amelets and
exceeds the value Da
p
q
given by amelet theory.
Other DNS results have shown that the scalar dissipation
rate controlling the growth of the ame hole is lower than
the one that should be applied to rst quench the ame
[38]. In Fig. 39, a DNS database is further analyzed [53]
in which a diffusion ame is pinched by a pair of vortices.
The value of x when the ame extinguishes is in perfect
agreement with x
q
measured in a laminar amelet library
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 250
Fig. 39. Three successive times of two-dimensional unsteady quenching of a planar diffusion ame pinched by a pair of vortices, conguration
studied in Ref. [53]. Left, bold: iso-reaction rate, dotted: vorcity, zoom: region of the ow where the centerline response of the ame versus
inverse mixture fraction dissipation rate is plotted. Right, solid line: 1D amelet library, diamond: top left 2D ame, square: middle left 2D
ame, triangle: bottom left 2D ame.
(Fig. 39, top). But once a hole exists in the ame, the
scalar dissipation rate x
q
Ed
at the quenching location is
smaller than x
q
and varies depending on the structure of
the edge ame. The uxes of heat along the stoichio-
metric surface are responsible for this departure from
one-dimensional amelet behavior. When the diffusion
ame and its extremity have reaction zones of the same
thickness, for x .x
q
the response of the reaction rate
follows the laminar amelet one (Fig. 39, middle). Latter
after quenching, x
q
Ed
decreases and a partially premixed
front develops at the edge of the burning zone, then the
reaction rate reaches values above the laminar amelet
behavior (Fig. 39, bottom). These values are representa-
tive of the existence of burning in a premixed regime at
the extremity of the diffusion ame. Therefore, the
quenching scalar dissipation rate x
q
is the relevant quan-
tity to describe the quenching of a burning amelet, but
diffusion ame quenching leads to edge ame combus-
tion (partially premixed amelets).
When partially premixed amelets are expected in the
combustion system, for instance at the base of a lifted
turbulent jet-ame, it may be interesting in the modeling
to use premixed amelets instead of diffusion amelets.
Bradley and coworkers have proposed amelet closures
along these lines where chemical sources are tabulated
using one-dimensional premixed ames [183]. The
chemical sources are parametrized in terms of the
mixture fraction Z, that determines the equivalence
ratio of the amelet whose progress of reaction is given
by a progress variable c. The mean burning rates are then
obtained by averaging the source terms with a presumed
form for the joint-pdf of Z and c.
Within the family of closures for non-premixed turbulent
ames, models based on Eq. (237) represent great progress
compared to the innitely fast chemistry hypothesis. When
the chemistry is fast enough with mixing elements featuring
sheet-like properties, they provide interesting results
(Fig. 40) and further renements of amelet models are
under development [184]. It is also worthwhile to note
that with SLFM, the turbulent micromixing of chemical
species does not need any particular treatment, since diffu-
sion is directly included and coupled with chemistry in Eq.
(237).
9.4.3. Mixing modeling in SLFM
In steady laminar amelet models (SLFM), the fuel/air
turbulent mixing is captured via the joint pdf
~
P(Z
p
; x
p
; x; t):
Most of the amelet models explicitly suppose that the
mixture fraction and its dissipation rate are two uncorrelated
quantities:
~
P(Z
p
; x
p
; x; t) <
~
P(Z
p
; x; t)
~
P(x
p
; x; t)
A b-function is assumed to presume
~
P(Z
p
; x; t) (see Section
6.4.2) and a log-normal distribution is used for
~
P(x
p
; x; t)
[63]:
~
P(x
p
; x; t) =
1

2p
_
x
p
s(x; t)
exp 2
1
2s
2
(x; t)
(ln x
p
2g(x; t))
2
_ _
The two parameters s and g are provided by the rst and
second moments of the scalar dissipation rate:
~ x = exp g 1
s
2
2
_ _
and

x
//2
= ~ x
2
(exp(s
2
) 21
In Ref. [63], s
2
= 2 is proposed.
Another alternative is to estimate a mean scalar dissipa-
tion rate under stoichiometric condition [89]. Starting from
the solution of a steady strained planar counter-ow ame:
Z(h) =
1
2
erfc(h=

2
_
); where h(j) =

a
D
_
_j
0
r(j
p
)
r
0
dj
p
a is the strain rate, j the coordinate normal to the stoichio-
metric plane and the assumption r
2
D < cst = r
2
0
D
0
was
used. The corresponding scalar dissipation rate is given by:
x(Z) =
a
2p
exp(22[erfc
21
(2Z)]
2
)
= x
0
exp(22[erfc
21
(2Z)]
2
) = x
0
F(Z) (238)
erfc
21
denotes the reciprocal of the complementary error
function and x
0
is the maximum value of the scalar dissipa-
tion rate. Mean and conditional scalar dissipation rate are
then related by:
~ x
st
= ~ xF(
~
Z;

Z
//2
) (239)
where F is given by:
F(
~
Z;

Z
//2
) =
F(Z
st
)
_1
0
F(Z
p
)
~
P(Z
p
)dZ
p
(240)
and to retain in the amelet library the proles
Y
SLFM
i
(Z
p
; ~ x
st
) for averaging with
~
P(Z
p
) (Eq. (236)).
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 251
Fig. 40. Top: Favre-averages mean values of temperature and OH
mole fraction along the centerline of a jet ame. Bottom: radial
proles of mean temperature and OH mole fraction. Comparison
between experiments and amelet modeling [177].
Here again, the stumbling block is the estimation of the
mean scalar dissipation rate ~ x: The linear relaxation model
may be written (Eq. (229)):
~ x = C
x

Z
//2
(k=1)
where C
x
is a constant depending on the topological and
spectral properties of the mixture fraction eld. More
recently, it was proposed to choose [33]:
~ x = C
x
(DZ
F
)
2
(k=1)
C
x
< 2; DZ
F
is the thickness of the reaction zone of these
laminar ame measured in mixture fraction space. The reac-
tion zone is centered at the location Z = Z
st
: As the stoichio-
metric surface is close to the air side for hydrocarbon ame
(Z
st
p1); one may write:
DZ
F
< 2Z
st
which provides an interesting approximation of this reaction
thickness. This last expression was successful to model
turbulent ame lift-off [185].
9.4.4. Conclusion
The laminar amelet model presumes the conditional
mean of species
_
rY
i
uZ = Z
p
_
averaging over the response
of laminar prototype ames:
_
rY
i
uZ = Z
p
_
= r
_
x
p
Y
SLFM
i
(Z
p
; x
p
)
~
P(x
p
)dx
p
In our generic classication of turbulent combustion model-
ing (Section 6) and Eq. (230), SLFM may be viewed as a
presumed pdf technique involving conditional mean values
determined from laminar ame solutions. Another alterna-
tive is to build models where quantities are estimated intro-
ducing a direct treatment of micromixing and small scale
diffusion (see Sections 9.69.8).
9.5. Flame surface density modeling, coherent ame model
Flame surface density concepts were rstly introduced for
the coherent ame model (CFM) [84]. The balance equation
for the ame surface density S was based on phenomeno-
logical considerations starting from balance equation for a
material surface where combustion effects have been in-
tuitively added [84,186]. Recent works [88,187,188] have
provided an exact balance equation, identifying the ame
surface to the stoichiometric iso-surface Z = Z
st
: The ame
surface density is then dened as:
S = u7Zud(Z 2Z
st
) =
_
u7ZuuZ = Z
st
_

P(Z
st
) (241)
where

P(Z
st
) denotes the probability of nding Z = Z
st
:
Starting from this denition and the balance equation for
the mixture fraction Z (Eq. (41)), an exact but unclosed,
balance equation for the ame surface density S may be
derived (Section 6.3.2).
In ame surface density models, under a amelet assump-
tion, it is assumed that the interface between fuel and
oxidizer in non-premixed burners behaves, for any values
of Z, as S(Z
p
) < u7Zu = S (Section 6.3.2). The mean burn-
ing rate

_ v
i
is then expressed as the product of S by
_
V
i
the
local reaction rate per unit of ame area. S accounts for
ame turbulence interaction and
_
V
i
for the chemistry.
According to Eq. (121)
_ v
i
=
_ v
i
u7Zu
_ _
s
u7Zu =
_
V
i
S (242)
The local burning rate
_
V
i
is estimated from a one-dimen-
sional laminar ame as:
_
V
i
=
_ v
i
u7Zu
_ _
s
<
_1
0
_ v
i
u7Zu
dZ =
_11
21
_ v
i
dj (243)
where j is the coordinate along the normal to the ame
front.
In diffusion amelets, where _ v
i
depends on the mixture
fraction Z and the scalar dissipation x, expressions (96), (91)
and the relation:
_
_ v
i
uZ
p
_
=
_
x
_ v
i
(Z
p
; x
p
)

P
c
(x
p
uZ
p
)dx
p
(244)
lead to:
_
V
i
=
1
u7Zu
_1
0
_
x
_ v
i
(Z
p
; x
p
)

P
c
(x
p
uZ
p
)

P(Z
p
)dx
p
dZ
p
(245)
Then, using Eq. (243):
_
V
i
<
_11
21
_
x
_ v
i
(x
p
; j)

P
c
(x
p
uZ
p
)dx
p
dj
<
_
x
_11
21
_ v
i
(x
p
; j)dj
_ _
........,,........
_
V
i
(x
p
)

P(x
p
)dx
p
(246)
The reaction rate per unit of ame area should, therefore, be
estimated from the integrated reaction rate of a amelet
submitted to a scalar dissipation rate x with the distribution

P(x
p
): In practice,
_
V
i
is estimated as
_
V
i
=
_
V
i
( ~ x); where ~ x
is the averaged scalar dissipation rate. The coherent ame
model, then, corresponds to a simplied version of the
steady laminar amelet model. The integrated burning rate
_
V
i
may also be expressed as a function of the species mol-
ecular uxes, see Eq. (53) in Section 3.2.
The introduction of effects of unsteadiness was realized in
Ref. [186], following the analysis proposed in Ref. [172] in
the context of SLFM, a ame time response is then intro-
duced when calculating
_
V
i:
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 252
9.6. MIL model
MIL (`Modele Intermittent Lagrangien' or Lagrangian
Intermittent Model) was proposed by Borghi in Ref. [68].
As in the steady laminar amelet model (SLFM) and the
conditional moment closure (CMC), MIL incorporates the
ame structure in Z space. Arguing that mixing occurs with-
out reaction for large values of the scalar dissipation rate,
corresponding to fast mixing, the ame structure is
constructed dynamically from two possibilities of diffusion
combustion: Mixing without reaction (before ignition) and
innitely fast chemistry (after ignition) (Fig. 41). The tran-
sition between mixing without combustion and a `mixed is
burnt' regime is controlled by the position of an ignition
time delay, t
ig
, within the distribution of fuel/air turbulent
mixing times t
k
, t
t
, 11, where t
k
. (n=1)
1=2
is the
Kolmogorov time.
The ignition delay t
ig
is associated to a scalar dissipation
rate x
ig
(t
ig
< x
21
ig
): The main objective of MIL is to account
for unsteadiness in the coupling between small scale diffu-
sion and chemistry, by means of the spectral distribution of
micromixing times. For a given value of t
ig
< x
21
ig
; mixing
times smaller than t
ig
(t
t
,t
ig
or x .x
ig
) prevent ignition
and contribute to pure mixing. Whereas for x ,x
ig
, corre-
sponding to t
t
.t
ig
, mixing is slower than chemistry and
combustion develops.
The inputs of MIL are:
1. A table of ignition delay times t
ig
depending on mixture
fraction, fuel mass fraction and mean temperature, t
ig
=
t
ig
(
~
T;
~
Y
F
; Z): This look-up table is easily derived from
partially stirred reactor (PaSR) calculations, where
~
T
and
~
Y
F
are the mean conditions in the PaSR. Detailed
chemistry can be used to construct such a t
ig
table
[189].
2. The pdf of turbulent mixing times

P(t
t
; x; t) (equiva-
lent to the pdf of the inverse of x).
3. The pdf of the mixture fraction
~
P(Z
p
; x; t):
4. The mean concentration of fuel
~
Y
F
:
The output is the mean burning rate
~
_ v
F
to calculate
new
~
Y
F
values, accounting for auto-ignition and transi-
ent burning in turbulent ames. Assuming that combus-
tion occurs only when ignition is faster than turbulent
fuel/air mixing, the probability to nd ignition at a
particular location is:
a(Z; x; t) =
_11
t
ig
(Z;x;t)

P(t
t
; x; t)dt
t
where

P(t
t
; x; t) is the pdf of the micromixing times
having a mean value equal to the integral length scale
time (k/1). Since t
ig
depends on position in mixture
fraction space and mean temperature, a is also a func-
tion of Z and of the local ow conditions through posi-
tion and time. Once

P(t
t
; x; t) and
~
P(Z
p
; x; t) are known,
the mean burning rate of fuel is estimated as:
~
_ v
F
(x; t) =
_1
0
_11
t
ig
(Z
p
;x;t)
_ v
MIL
F
(Z
p
; t
p
t
)
~
P(Z
p
; x; t)

P(t
p
t
; x; t)dt
p
t
dZ
p
(247)
A partially stirred reactor (PaSR), combining mixing
and chemical effects, is introduced to calculate
_ v
MIL
F
(Z
p
; t
p
t
) [68], leading to:
dY
p
F
dZ
=
(
~
Y
p
F
2Y
F
) 1t
p
t
_ v
MIL
F
(
~
Z 2Z
p
)
(248)
This technique is close to the one chosen in Ref. [190]
for modeling NO
x
production in turbulent jet ames. Eq.
(248) describes a local ame element, where molecular
diffusion, D7
2
Y
F
, is modeled as a linear relaxation term
(
~
Y
F
2Y
F
)=t
t
: Solving Eq. (248) with boundary condi-
tions similar to the ones of amelet libraries provides
_ v
MIL
F
(Z
p
; t
p
t
):
A generic shape for

P(t
t
; x; t) was proposed [189,191]:

P(t
t
; x; t) = (q 21)
t
(q21)
t
q
k
exp 2
(q 21)
q
t
t
t
k
_ _
q
_ _
q(x; t) is determined so that
~
Y
MIL
F
calculated along the MIL
trajectory in mixture fraction space (Fig. 41), weighted by
the b-pdf, reproduces the mean
~
Y
F
:
Fig. 42 shows an example of numerical simulation of a
lean premixed prevaporized (LPP) combustion chamber
using MIL. The LPP system is divided into three parts:
rst, liquid fuel is injected and vaporized (prevaporized
tube), then is mixed with air in the premixing tube before
entering the main combustion chamber. Combustion is
expected to reduce pollutant emission in a lean premixed
regime. The premixing tube operates for a large range of
inlet temperatures (from 300 to 800 K) and a large range of
pressures (from 1 to 40 bar). For the highest temperature and
pressure conditions, some undesirable effects, like auto-
ignition or ashback, may be observed in the premixed
tube. These ndings are reproduced with turbulent combus-
tion closures like MIL describing strongly coupled micro-
mixing and ignition phenomena [191].
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 253
Fig. 41. Sketch of MIL in mixture fraction space.
9.7. Conditional moment closure
Innitely fast chemistry (IFCM), steady laminar amelets
(SLFM), and MIL suggest that non-premixed turbulent
ames may be conveniently studied using conditional aver-
aging in the mixture fraction space. Instead of getting Q
i
=
_
rY
i
uZ = Z
p
_
=
_
ruZ = Z
p
_
from laminar amelets, the CMC
approach proposes to solve a balance equation for Q
i
[102,103,105], which is written, neglecting two terms
(Eq. (114)):
_
ruZ = Z
p
_
2Q
i
2t
= 2
_
ru
i
uZ = Z
p
_
2Q
i
2x
i
.......,,.......
(I)
1
_
rxuZ = Z
p
_
2
2
Q
i
2Z
p2
........,,........
(II)
1
_
_ v
i
uZ = Z
p
_
.....,,.....
(III)
(249)
In addition to turbulent transport included in (I), micro-
mixing (II) and the chemical source (III) are unclosed. In
the simplest version of CMC, the uctuations of chemical
source in mixture fraction space are neglected leading to:
_
_ v
i
uZ = Z
p
_
= _ v
i
__
Y
1
uZ = Z
p
_
; ;
_
Y
N
uZ = Z
p
__
(250)
As for any combustion models, the scalar dissipation rate is
unknown (II). With CMC, the conditional average
_
rxuZ = Z
p
_
needs closure. This conditional mean is an
important ingredient of the pdf transport equation and
micromixing models are available (Section 9.8.1). Simi-
larly, when the pdf is known via its presumed b-shape
(Section 6.4.2), the conditional value of x may be approxi-
mated from
~
P(Z
p
) to obtain
_
rxuZ = Z
p
_
[192].
CMC has been the subject of many works and renements
during the past years [105]. The following points have been
addressed:
The approximation of the conditional chemical source
using Eq. (250) is much less restrictive than to write
~
_ v
i
<
_ v
i
(
~
Y
1
; ;
~
Y
N
) but implies a weak level of uctuations for
a given value of Z. When calculating the chemical source,
the conditional pdf

P
c
(Y
i
uZ
p
) is approximated by a peak
located at
_
Y
i
uZ = Z
p
_
: Accordingly, this method is
expected to provide interesting results for conditions
not too far from innitely fast chemistry. The uctuations
of species concentrations and temperature for a given
level of mixture fraction may be included in the condi-
tional burning rate by using second order CMC. This
leads to a variety of modeling renements where more
transport equations for conditional quantities are studied,
modeled and solved [105].
The estimation of
_
rxuZ = Z
p
_
from a b-pdf is feasible
for simple shapes of
~
P(Z
p
); but as shown from the pdf
balance equation (Eq. (110)), the detail of
_
rxuZ = Z
p
_
is
linked to the detail of the shape of the pdf, that may not be
correctly captured with a presumed pdf.
Within a homogeneous eld, the CMC equation reduces
to:
_
ruZ = Z
p
_
2Q
i
2t
=
_
rxuZ = Z
p
_
2
2
Q
i
2Z
p2
1
_
_ v
i
uZ = Z
p
_
(251)
This last formulation may be compared with the amelet
equation (Eq. (237)) replacing Q
i
by Y
i
[107]. These two
turbulent combustion models have similarities in term of
their general formalism, but their underlying physical
assumptions strongly differ. CMC incorporates diffusion
as turbulent micromixing and decouples it from chemis-
try. On the other hand, the amelet assumption deals with
a coupled representation of diffusion and reaction as in a
laminar ame for a given value of x, and therefore, a
given thickness of the laminar diffusive zone.
Once the number M of conditional values required to
capture the conditional means of the N chemical species is
determined, CMC modeling requires the solution of N M
balance equations. CMC was also extended to reacting
particles [193].
9.8. Pdf modeling
In premixed turbulent combustion modeling, BML and
related closures assume a bimodal probability density func-
tion (pdf) of the progress variable. In non-premixed ames,
using innitely fast chemistry (IFCM), steady laminar
amelet (SLFM), or conditional moment closure (CMC),
the pdf of the mixture fraction was assigned a b-shape.
The objective of PDF modeling is to relax all hypotheses
concerning the shape of pdfs. Once a methodology has been
developed to calculate pdfs, it is possible to construct turbu-
lent combustion closures in which all the values taken by
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 254
Fig. 42. Snapshot of temperature in a TURBOMECA-ONERA LPP engine computed with MIL [191]. 1: Vaporized tube, 2: premixed tube,
3: ame tube. Combustion starts in the premixed tube.
species and temperature in mixture fraction space may be
reached (Figs. 5 and 43). As in amelet modeling and CMC,
a detailed description of chemistry may be incorporated in
the modeling. Differential diffusion is easily introduced in
SLFM when tabulating counter-ow ames, these prefer-
ential diffusion effects may be more difcult to introduce
in CMC, in the context of pdfs, some works have been made
in this direction [194,195].
Two levels of complexity may be chosen:
One may solve the balance equation for the joint pdf of
the thermochemical variables, species, temperature, etc.
(Eq. (111) in Section 6):
r
2
2t
~
P(Z
p
; x; t)
= 2 r

N
i=1
2
2Y
p
i
2u7Y
i
uY(x; t) = Y
p
_ _
1 _ v
p
i
_ _
~
P(Z
p
; x; t)
_ _
2 r

N
i=1

N
j=1
2
2Y
p
i
2
2Y
p
j
D
2Y
i
2x
k
2Y
j
2x
k
_
_
_
_
Y(x; t) = Y
p
_ _
~
P(Z
p
; x; t)
_ _
(252)
Velocity eld statistics may also be included into the pdf,
then the balance equation for the joint pdf of velocity and
species is resolved (Eq. (113)). Recent developments
have also included time scale information into the pdf
[196,197].
There are two major issues with pdf modeling:
Models must be proposed for capturing micromixing and
pressure uctuation and viscous effects (the latter only
when the velocity eld is incorporated into the pdf).
A new numerical treatment is required and leads to the
decomposition of the pdf into a large number of com-
putational particles required to succeed in the computa-
tional implementation of pdf methods [93,198,199].
The chemical composition and velocities of these par-
ticles evolve according to simple differential equations,
resulting from a discrete Monte Carlo simulation of the
continuous pdf equation.
The main advantage of the pdf method lies in the
possibility of treating complex chemical sources directly
(Section 6).
9.8.1. Turbulent micromixing
From the discussion of fundamentals properties of ames
and of turbulent combustion models, the scalar dissipation
rate appears as a key quantity in both premixed and non-
premixed turbulent systems. Using pdfs and CMC, one
needs to express the conditional mean value of the scalar
dissipation rate (or the conditional mean value of diffusion).
This term represents turbulent micromixing [95,96].
Using turbulence modeling, such as the k1 closure, a
mean mechanical time t
t
< (k=1) and a microscale time
t
k
< (n=1)
1=2
may be estimated. A global picture of mixing
should therefore distinguish between macro and micro-
mixing [200]:
The macromixing is organized from a cascade mechanism
between large and small scales. The study of this cascade in
a fully developed turbulence shows that the mixing
frequency f
Z
is a function of the characteristic length l
Z
[201]:
f
Z
(l
Z
) ,
1
n
_ _
1=2
h
k
l
z
_ _
2=3
The speed at which Z is mixed increases when the charac-
teristic length l
Z
decreases, suggesting that the cascade
mechanism is limited by the mixing frequency f
Z
(l
Z
) of the
largest scales. When the Schmidt number of Z is greater than
unity (n .D
i
), the impact of the molecular dissipation is
less than the effect of viscous dissipation, then the smallest
scales of Z are smaller than the Kolmogorov scales. For
those small scales (l
Z
,h
k
) the mixing frequency is of the
order of a constant f
Z
(l
Z
) = (1=n)
1=2
: This cascade process
from large to small scale does not change the energy in the
system and therefore does not affect

Z
//2
:
The molecular diffusion acting at smaller scales makes
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 255
Fig. 43. Two-dimensional pdf of mixture fraction and reaction
progress variable for a round jet non-premixed ame [104].
the gradients smoother and then induces the micromixing,
phenomenon required to put the reactants in contact with the
reaction zone. When n = D; this process starts at the
Kolmogorov scale, when n D; the Batchelor scale
becomes the relevant limiting length:
l
B
, (nD
2
=1)
1=4
= Sc
21=2
h
k
All these approximations are given for non-reactive ows to
exactly quantify the effect of combustion is an arduous task.
This difculty is also encountered in amelet or ame
surface modeling, but appears in a different form.
In pdf modeling, the micromixing closure must estimate
the term
_
D7
2
Y
i
uY = Y
p
_
describing the molecular diffu-
sion of each chemical species, or mimic its impact on the
pdf. Various techniques have been proposed [95].
9.8.2. Linear relaxation model, IEM/LMSE
In b-pdf modeling (Section 9.2.2), the scalar dissipation
rate of the mixture fraction was expressed as ~ x = (

Z
//2
=t
t
):
The extension of this linear relaxation to conditional diffu-
sion provides the linear mean square estimation (LMSE) or
interaction by exchange with the mean (IEM) model. First
proposed in the pdf context by Dopazo and O'Brien [202],
IEM was also developed and used in chemical engineering
applications by Villermaux [203].
Fig. 44 shows the conditional mean of the diffusive
budget of the mixture fraction Z measured in a DNS of a
mixing layer. A quasi-linear response is observed close to
the mean, with a slope that can be approximated using the
inverse of t
t
< (k=1): Since the gradient of Z must relax
to zero for Z = 0 and Z = 1; this quasi-linear response
is combined with the conditions
_
D7
2
ZuZ = 0
_
=
_
D7
2
ZuZ = 1
_
= 0: For a `Gaussian-type pdf' (no peaks
at Z = 0 and Z = 1), the behavior at the edges of the sample
space is not of prime importance to capture the major prop-
erties of micromixing, and a rst approximation is given by:
_
D7
2
ZuZ = Z
p
_
<

Z 2Z
p
t
t
_ _
<

Z 2Z
p
k=1
_ _
(253)
This simple model cannot capture the time evolution of the
pdf in homogeneous mixing problems and has many short-
comings, for instance it relaxes the uctuations without
modifying the shape of the pdf [9496,204].
9.8.3. GIEM model
An improvement of the IEM model was recently
discussed [205], a b-pdf is used to reproduce the correct
behavior of conditional diffusion. The mixture fraction is
chosen as a shadow eld to estimate micromixing, this
closure is also called a `Beta-Mapping-Closure'. In a homo-
geneous eld, the time evolution of the mixture fraction pdf
is:
2
~
P(Z
p
; x; t)
2t
= 2
2
2Z
p
__
D7
2
ZuZ = Z
p
_
~
P(Z
p
; x; t)
_
since
_
D7
2
ZuZ = 0
_
= 0; one may write:
_
D7
2
ZuZ = Z
p
_
= 2
1
~
P(Z
p
; x; t)
2
2t
_Z
p
0
~
P(Z
1
; x; t)dZ
1
_ _
(254)
When
~
P(Z
p
; x; t) is given by a b-function parameterized
with
~
Z and

Z
//2
; this equation determines (D7
2
ZuZ = Z
p
).
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 256
Fig. 44. Left: Scatter plot of D7
2
Y. Right: (D7
2
YuY(x; t) = Y
p
): These two terms were measured in DNS of an inert mixing layer [234].
The LMSE/IEM model leads to:
_
D7
2
ZuZ = Z
p
_
<
1
t
t
(
~
Z 2Z
p
) (255)
Making a direct analogy GIEM proposes:
_
D7
2
ZuZ = Z
p
_
< C(Z
p
; t)(B
Z
(t) 2Z
p
) (256)
where C(Z
p
, t) is a pseudo-mixing frequency depending both
on the position in mixture fraction space and time. It allows
for reproducing the correct behavior at the edges Z
p
= 0 and
Z
p
= 1 and its time dependence gives the correct behavior
during the relaxation of the pdf. B
Z
(t) denotes the point
where the diffusive ux is zero, relaxing the constraint
B
Z
(t) =
~
Z of IEM, C(Z
p
, t) and B
Z
(t) are determined from
Eqs. (254) and (256).
The same treatment is applied to any reactive species Y
i
,
using the same pseudo-mixing frequency (an assumption
made in most micromixing models):
_
D7
2
Y
i
uY
i
= Y
p
i
; Z = Z
p
_
= C(Z
p
; t)(B
Y
i
(t) 2Y
p
i
)
The parameter B
Y
i
(t) is then estimated from the fact that
micromixing by itself should not change the mean:
2Y
i
2t
_ _
= (2D7
2
Y
i
) < 0 - B
Y
i
(t) =
C(Z; t)Y
i
C(Z; t)
This attractive model conserves the simplicity of linear
relaxation modeling, but overcomes some of its drawbacks.
As

Z
//2
is an input of the closure, the mean scalar dissipation
rate ~ x should be previously modeled. This approach has
similarities with the estimation of the conditional values
of x in CMC using b-pdf [192] and is not a full closure
of micromixing per se.
9.8.4. Stochastic micromixing closures
Micromixing modeling may also be developed by choos-
ing an ad hoc stochastic process mimicking the relaxation of
pdfs due to small scale diffusion. These closures are usually
presented in the context of non-continuous interactions of
volumes of uid. When only diffusion is acting, the generic
form of these models is written [94]:
2
~
P(Z
p
; t)
2t
=
2b
t
Z
_
__
~
P(Z
1
; t)
~
P(Z
11
; t)
T(Z
1
; Z
11
- Z
p
)dZ
1
dZ
11
2
~
P(Z
p
)
_
b is a parameter of the model and T(Z
1
, Z
11
- Z
p
) is the
probability that a volume of concentration Z
1
interacts with
a volume at Z
11
to evolve into Z
p
. The difcult point lies in
the choice of the transition probability T. Multiple choices
are possible and have been made (see Ref. [95] for a
review), the simplest one consists of writing:
T(Z
1
; Z
11
- Z
p
) = d Z
p
2
Z
1
1Z
11
2
_ _
(257)
meaning that the two volumes interact to generate a volume
with Z
p
= (Z
1
1 Z
11
)=2: Eq. (257) was rst proposed by
Curl [206] to simulate the production of droplets of a given
diameter by coalescence of two droplets of various sizes and
re-dispersion of the two droplets.
These closures are clearly linked to a purely stochastic
view of micromixing and are well suited to Monte Carlo
simulations [93]. Their applications to ames can be disputed
easilysince mixing described via stochastic processes does not
always account for the presence of reaction zones. However,
much work has been done to improve these models which
have then been quite successful in the calculation of jet
ames [97]. Fig. 45 shows pdfs of the temperature calcu-
lated from Eq. (110) [207] using the coalescence-dispersion
model [206] and the k1 closure for the velocity eld [208].
For the same calculations, Fig. 46 illustrates the impact of
micromixing modeling on temperature and CO concentra-
tions predictions, the reduced chemical scheme of Ref.
[209] has been utilized with the LMSE (IEM) and Curl's
coalescence re-dispersion formulation.
Micromixing modeling is obviously one of the greatest
challenges of turbulent combustion modeling, and strong
efforts have been made in this direction. Many types of mixing
models were proposed, a full reviewis beyond the scope of this
paper. Detailed discussion of mixing models in the context
of pdfs may be found in the following list of references
[93,9597,104,194,196,205,206,208,210236].
9.8.5. Interlinks PDF/ame surface modeling
The unclosed molecular diffusion term in the pdf balance
Eq. (109) may be recast as:
1
r
(7(rD7Z))uZ = Z
p
_ _
= 2
1
ru7Zu
2
2n
(rDu7Zu)u7ZuuZ = Z
p
_ _
2
_
D7nu7ZuuZ = Z
p
_
= 2
1
ru7Zu
2
2n
(rDu7Zu)
_ _
s
1kD7nl
s
_ _
_
u7ZuuZ = Z
p
_
= 2[kv
n
l
s
1kD7nl
s
]
_
u7ZuuZ = Z
p
_
(258)
The molecular diffusion is decomposed into a normal contri-
bution, that may be expressed in terms of a normal displace-
ment speed w
n
and a curvature term D7n; according to
Eq. (38). The diffusion coefcient D is assumed constant
and kD7nl
s
= Dk7nl
s
: The ame surface averaged curva-
ture may be modeled from geometrical considerations
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 257
[88,121] as in Section 7.4.4 (Eq. (208)):
k7nl
s
< 2

Z 2Z
1
L
y
(259)
where L
y
is a wrinkling length scale of the ame surface
and Z
1
is the

Z level where mean curvatures are equal
to zero. Mean curvatures are supposed positive (i.e.
convex toward

Z = 0 region) and negative for

Z = 1.
The comparison of Eqs. (258) and (259) shows that
the geometrical assumption used to model the mean
curvature corresponds to an LMSEIEM model of the
molecular diffusion, when neglecting the normal displace-
ment speed kw
n
l
s
(Section 9.8.2):
2
1
r
(7(rD7Z))uZ = Z
p
_ _
=

Z 2Z
p
t
t
(260)
It was shown in (Section 9.8.2) that for a non-reacting
scalar, the IEM model provides an interesting estimation
of the micromixing term around

Z = Z
p
when the turbulent
mixing time is correctly calculated [234] but becomes de-
cient when Z
p
tends towards 0 or 1, where u7Zu is expected to
relax to zero. This shortcoming is attenuated in the ame
surface density context where the surface averaged curva-
ture k7nl
s
is multiplied by the ame surface density (Eq.
(90)), ensuring zero values at either side of the ame brush.
The turbulent mixing time t
t
and the ame wrinkling length
scale, L
y
, are then related as L
y
< D
_
u7ZuuZ = Z
p
_
t
t
; where
_
u7ZuuZ = Z
p
_
characterizes laminar ame elements.
Expressing the mixing time from linear relaxation hypo-
thesis, t
t
<

Z
//2
= ~ x and using Eq. (116) under a amelet
assumption yields L
y
<

Z
//2
=u7Zu: Wrinkling length scales
depend on scalar uctuations given by Eq. (226), eviden-
cing, once again, the need for modeling scalar dissipation
rates.
9.8.6. Joint velocity/concentrations pdf modeling
When the joint pdf of velocity and species enters the
problem, one needs to model additional viscous and
pressure terms related to velocity uctuations (see Eq.
(113)). In turbulent non-reactive ows, experiments have
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 258
Fig. 45. Comparison between measurements (left) and numerical predictions (right) of unweighted temperature pdf at different locations in the
Masri et al. [207] `L' jet ame (location x=D = 20). Pdf calculations by Jones and Kakhi [208] using the Curl mixing model [206].
2
In non-premixed turbulent ames, the mean normal displace-
ment speed kw
n
l
s
is expected to be negligible against curvature
effects [88] because such ames are unable to propagate. This is
not the case in premixed combustion, where the IEMLMSE model
requires a correction to account for kw
n
l
s
that is of the order of the
laminar ame speed [237].
shown that the velocity uctuations have pdfs that may be
approximated with Gaussian functions. Closures for the
velocity pdf may then be obtained using a Langevin type
model, a simple stochastic model reproducing this velocity
distribution, as done in Ref. [238].
Basic properties of stochastic methods [239] show that an
equation with a diffusive term for a pdf

P:
2

P(Z
p
)
2t
= 2
2
2Z
p
(A(Z
p
; t)

P(Z
p
)) 1
1
2
2
2
2
2
Z
p
(B(Z
p
; t)

P(Z
p
))
is well reproduced by a stochastic process, Z(t), similar to
the Langevin process:
dZ(t) = A(Z(t); t)dt 1

B(Z(t); t)
_
dW(t)
where dW(t) is a white noise (Wiener process). Using this
equivalence, the Langevin model [238] for the pressure and
viscous unclosed terms of Eq. (113) may be written:
__
1
r
2p
/
2x
i
uu = u
p
; Z = Z
p
_
2
_
n7
2
u
/
i
uu = u
p
; Z = Z
p
_
_

P(u
p
; Z
p
; x; t)
= G
ij
(u
p
j
2 u
j
)

P(u
p
; Z
p
; x; t) 1C
0
1
2
2u
p
i
[

P(u
p
; Z
p
; x; t)]
(261)
The physical properties of this closureare includedinthe tensor
G
ij
. The simplest model considers a linear relaxation for the
velocity eld, then G
ij
= (1=2 13C
0
=4)(1=k)d
ij
; where C
0
is a
constant. More sophisticated methods have been proposed
[240], where a parallel is drawn between Reynolds stress
closures (modeled equations for u
/
i
u
/
j
) and the tensor G
ij
[241].
As already indicated, a numerical solution for the pdf
implies the development of accurate Monte Carlo solutions
[93]. The pdf eld is decomposed into a set of stochastic
particles whose evolution simulates the turbulent ame.
When a Lagrangian context is adopted [242], each particle
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 259
Fig. 46. Radial proles of CO mass fraction and temperature at different locations in Masri et al. [207] `B' jet ame, pdf calculations [208]. Two
mixing models have been used: lled circles: measurements; solid line: coalescencedispersion micromixing Curl model [206]; dashed line:
LMSE micromixing model (IEM). Reduced chemical scheme of Jones and Lindstedt [209].
n is characterized by its position x
p
n
(t); its velocity u
p
n
(t) and
concentration(s) Y
p
n
(t): The Langevin model given in Eq.
(261) corresponds to the system for the discrete particles:
dx
in
dt
= u
in
du
in
dt
= 2
1
r
2 p
2x
i
_ _
1n
2
2
u
i
2x
j
2x
j
_ _
1G
ij
(u
j
n
2 u
j
) 1(C
0
1)
1=2
dW
i
dt
_ _
dY
p
i
dt
=
(
~
Y
i
2Y
p
i
)
(k=1)
1 _ v
i
(Y
p
i
)
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
_
(262)
The last difculty is the determination of the mean pressure
eld (2 p=2x
i
): Methods have been proposed to calculate it
from Monte Carlo simulations [101]. In most of Monte
Carlo simulations, the Lagrangian simulation is coupled
with Eulerian calculations of means values providing
the mean pressure eld and the dissipation rate of the
velocity uctuations. However, full pdf methods includ-
ing a model for the turbulent frequency following a
uid particle have also been developed with success [197].
Fully detailed reviews on pdf methods may be found in
Refs. [9395].
10. Conclusion
Modeling turbulent combustion is a challenging task. In
turbulent ames, various difculties (strong heat release,
complex chemistry, large range of time and length scales,
etc.) are added to the complexity of constant-density turbu-
lent ows.
A review of the most classical Reynolds (or Favre) aver-
aged NavierStokes (RANS) models has been proposed.
Three main ingredients must be modeled:
Reynolds stresses: u
i
u
j
2 ~ u
i
~ u
j
:
Turbulent transport of species mass fractions:

u
i
Y
k
2 ~ u
i
~
Y
k
:
Mean reaction rate of species:

_ v
k
:
Most works have focus on rened descriptions of the
mean reaction rate

_ v
k
: To provide this reaction rate, tools
have been proposed, which are based on a pure statistical or
more oriented geometrical view of turbulent ames. All
these various models are related via a scalar dissipation
rate, a key quantity representing micromixing of the reac-
tants occurring at the small scales where chemical reactions
develop. The links and the similarities between combustion
models have been demonstrated here. The precise knowl-
edge of these relations may be useful to understand the exact
implications of the underlying physical hypothesis. Major
models are then directly derived from the proposed expres-
sions, which illuminate unexpected links between well-
known closure schemes.
Few studies have focussed on species turbulent transport,
generally modeled using a gradient expression, whereas
counter-gradient turbulent transport is known to appear in
some situations. Reynolds stresses are generally described
from well established turbulent models (mostly k, 1), simply
re-written in terms of Favre averaged quantities. So far,
combustion effects on the ow (ame induced turbulence
generation, higher viscous dissipation, etc.) are not inten-
tionally included in the calculations.
There are many open questions and challenges left. One
of them is the understanding of the coupling between spray
and combustion including detailed chemistry. DNS with a
two-way coupling between a dilute spray and a carrier phase
were recently performed to progress in this direction [243]
and closures exist [243245], but more work is needed to
reach a level where liquid atomization, vaporization
together with ames are properly described.
RANS combustion models will remain useful for the next
few years. However, large eddy simulations (LES) stands as
very promising technique for turbulent combustion:
Combustion ow elds generally exhibit large scale
motions [26].
LES appear as a promising tool to capture combustion
instabilities.
Flames are mainly driven by mixing. LES is a good
candidate to capture unsteady turbulent mixing.
LES may directly provide part of the description of turbu-
lence/combustion interactions because zones of fresh and
burnt gases, having different turbulence characteristics, are
instantaneously identied at the level of the resolved grid.
LES is at a very early stage for combustion applications and
only few works have been done in this direction, mainly
devoted to feasibility tests (two-dimensional simulations,
constant density ows, etc.). Nonetheless, the results of these
preliminary tests suggest that LES will rapidly become a
complementary way to carefully simulate and understand
turbulent combustion systems. So far LES modeling is essen-
tially based on the same modeling strategies as the ones
described here for RANS turbulent combustion modeling.
References
[1] Linan A, Williams FA. Fundamental aspects of combustion.
Oxford: Oxford University Press, 1993.
[2] Clavin P. Premixed combustion and gasdynamics. Annu Rev
Fluid Mech 1994;26:32152.
[3] Buckmaster J, Ludford G. Theory of laminar ames.
Cambridge: Cambridge University Press, 1993.
[4] Clavin P. Dynamics of combustion fronts in premixed gases:
from ames to detonations. Proc Comb Inst 2000;28:56985.
[5] Roberts WL, Driscoll JF. A laminar vortex interacting with a
premixed ame: measured formation of pockets of reactants.
Combust Flame 1991;87:24556.
[6] Mueller CJ, Driscoll JF, Reuss DL, Drake MC, Rosalik ME.
Vorticity generation and attenuation as vortices convect
through a premixed ame. Combust Flame 1998;112:34258.
[7] Smooke MD, Mitchel RE, Keyes DE. Numerical solution of
two-dimensional axisymmetric laminar diffusion ames.
Combust Sci Technol 1989;67:85122.
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 260
[8] Givi P. Model free simulations of turbulent reactive ows.
Prog Energy Combust Sci 1989;15:1107.
[9] Poinsot T. Using direct numerical simulations to understand
premixed turbulent combustion. Twenty-Sixth Symposium
(International) on Combustion. Pittsburgh: The Combustion
Institute, 1996 p. 21932.
[10] Baritaud T, Poinsot T, Baum M. Direct numerical simulation
for turbulent reacting ows. Paris: Editions Technip, 1996.
[11] Poinsot T, Trouve A, Candel S. Applications of direct
numerical simulations of premixed turbulent combustion.
Prog Energy Combust Sci 1996;21:53176.
[12] Vervisch L, Poinsot T. Direct numerical simulation of non-
premixed turbulent ame. Annu Rev Fluid Mech
1998;30:65592.
[13] Barrere M, Prud'homme R. Equations fondamentales de
l'aerothermochimie. Paris: Masson, 1973.
[14] Williams FE. Combustion theory. 2nd ed. Reading, MA:
Addison-Wesley, 1985.
[15] Kuo KK. Principles of combustion. New York: Wiley, 1986.
[16] Poinsot T, Veynante D. Theoretical and numerical combus-
tion. Philadelphia, PA: Edwards, 2001.
[17] Trouve A, Poinsot T. The evolution equation for the ame
surface density. J Fluid Mech 1994;278:131.
[18] Favre A. Statistical equations of turbulent gases. In: SIAM,
editor. Problems of hydrodynamics and continuum
mechanics. Philadelphia: SIAM, 1969. p. 23166.
[19] Bray KNC, Libby PA, Masuya G, Moss JB. Turbulence
production in premixed turbulent ames. Combust Sci Tech-
nol 1981;25:12740.
[20] Shepherd IG, Moss JB, Bray KNC. Turbulent transport in
conned premixed ame. 19th Symposium(International) on
Combustion. Pittsburgh: The Combustion Institute, 1982.
p. 42331.
[21] Piomelli U, Chasnov JR. Large eddy simulations: theory and
applications. In: Hallback H, Henningson DS, Johansson
AV, Alfredsson PH, editors. Turbulence and transition
modelling, Dordrecht: Kluwer Academic Publishers, 1996.
p. 269336.
[22] Ferziger J. Large eddy simulation: an introduction and
perspective. In: Metais O, Ferziger J, editors. New tools in
turbulence modelling. Berlin: Les Editions de Physique
Springer, 1997. p. 2947.
[23] Lesieur M. Recent approaches in large-eddy simulations of
turbulence. In: Metais O, Ferziger J, editors. New tools in
turbulence modelling. Berlin: Les Editions de Physique
Springer, 1997. p. 128.
[24] Lesieur M, Metais O. New trends in large-eddy simulations
of turbulence. Annu Rev Fluid Mech 1996;28:4582.
[25] Veynante D, Poinsot T. Reynolds averaged and large
eddy simulation modeling for turbulent combustion. In:
Metais O, Ferziger J, editors. New Tools in Turbulence
Modelling. Berlin: Les Editions de PhysiqueSpringer, 1997.
p. 10540.
[26] Coats CM. Coherent structures in combustion. Prog Energy
Combust Sci 1996;22:427509.
[27] Menon S, Jou WH. Large eddy simulations of combustion
instability in an axisymmetric ramjet combustor. Combust
Sci Technol 1991;75(13):5372.
[28] Ghosal S, Moin P. The basic equations for the large eddy
simulation of turbulent ows in complex geometry. J Comp
Phys 1995;118:2437.
[29] Peters N. The turbulent burning velocity for large-scale and
small-scale turbulence. J Fluid Mech 1999;384:10732.
[30] Linan A. The asymptotic structure of counterow diffusion
ames for large activation energies. Acta Astronaut
1974;1007(1).
[31] Linan A, Orlandi P, Verzicco R, Higuera FJ. Effects of non-
unity Lewis numbers in diffusion ames. Studying turbulence
using numerical databasesV. In: Moin P, Reynolds WC,
1994. Stanford, CA: CTR, Stanford University. p. 518.
[32] Barlow RS, Fiechtner GJ, Carter CD, Chen J-Y. Experiments
on the scalar structure of turbulent CO/H
2
/N
2
jet ames.
Combust Flame 2000;120(4):54969.
[33] Bray KNC, Peters N. Laminar amelets in turbulent ames.
In: Libby PA, Williams FA, editors. Turbulent reacting
ows. London: Academic Press, 1994. p. 63113.
[34] Cuenot B, Poinsot TJ. Asymptotic and numerical study of
diffusion ames with variable Lewis number and nite rate
chemistry. Combust Flame 1996;104(1):11137.
[35] Burke SP, Schumann TEW. Diffusion ames. Ind Engng
Chem 1928;20:9981004.
[36] Kim JS, Williams FA. Extinction of diffusion ames with
nonunity Lewis numbers. J Engng Math 1997;31:10118.
[37] Clavin P, Linan A. Theory of gaseous combustion. An intro-
ductive course. In: Verlarde MG, editor. Non equilibrium
cooperative phenomena in physics and related elds. New
York: Plenum Press, 1983.
[38] Favier V, Vervisch L. Edge ames and partially premixed
combustion in diffusion ame quenching. Combust Flame
2001;125(1/2):788803.
[39] Linan A. Ignition and ame spread in laminar mixing layer.
In: Buckmaster J, Jackson TL, Kumar A, editors. Combus-
tion in high speed ows. Dordrecht: Kluwer Academic
Publishers, 1994. p. 461.
[40] Phillips H. Flame in a buoyant methane layer. Tenth Sym-
posium (International) on Combustion. Pittsburgh: The
Combustion Institute, 1965.
[41] Kioni PN, Rogg B, Bray KNC, Linan A. Flame spread in
laminar mixing layers: the triple ame. Combust Flame
1993;95:276.
[42] Plessing T, Terhoeven P, Peters N, Mansour MS. An exper-
imental and numerical study on a laminar triple ame.
Combust Flame 1998;115(3):33553.
[43] Kioni PN, Bray KNC, Greenhalgh DA, Rogg B. Experimen-
tal and numerical studies of a triple ame. Combust Flame
1998;116:192206.
[44] Dold JW. Flame propagation in a nonuniform mixture:
analysis of a slowly varying triple ame. Combust Flame
1989;76:7188.
[45] Hartley LJ, Dold JW. Flame propagation in a nonuniform
mixture: analysis of a propagating triple-ame. Combust
Sci Technol 1991;80:2346.
[46] Buckmaster J. Edge-ames and their stability. Combust Sci
Technol 1996;115:4168.
[47] Ghosal S, Vervisch L. Theoretical and numerical investiga-
tion of a symmetrical triple ame using a parabolic ame tip
approximation. J Fluid Mech 2000;415:22760.
[48] Bourlioux A, Cuenot B, Poinsot T. Asymptotic and nu-
merical study of the stabilization of diffusion ames by hot
gas. Combust Flame 2000;120(1/2).
[49] Ruetsch GR, Vervisch L, Linan A. Effects of heat release on
triple ame. Phys Fluids 1995;6(7):144754.
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 261
[50] Domingo P, Vervisch L. Triple ames and partially premixed
combustion in autoignition of nonpremixed mixtures. Twenty-
Sixth Symposium (International) on Combustion. Pittsburgh:
The Combustion Institute, 1996. p. 23340.
[51] Plessing T, Terhoven P, Peters N, Mansour MS. An experi-
mental and numerical study on a laminar triple ame.
Combust Flame 1998;115(3):33553.
[52] Echekki T, Chen JH. Structure and propagation of methanol
air triple ames. Combust Flame 1998;114(1/2):23145.
[53] Favier V, Vervisch L. Effects of unsteadiness in edge-ames
and liftoff in non-premixed turbulent combustion. Twenty-
Seventh Symposium (International) on Combustion. Pitts-
burgh: The Combustion Institute, 1998. p. 123945.
[54] Ghosal S, Vervisch L. Stability diagram for lift-off and blow-
out of a round jet laminar diffusion ame. Combust Flame
2001;124(4)64655.
[55] Vervisch L. Using numerics to help understanding of nonpre-
mixed turbulent ames. Proc Combust Inst 2000;28:1124.
[56] Borghi R. Reactions chimiques en milieu turbulent. PhD
Thesis. Universite Pierre et Marie Curie, Paris 6, 1978.
[57] Villasenor R, Pitz RW, Chen JY. Interaction between chemi-
cal reaction and turbulence in supersonic non-premixed
H
2
air combustion. In: 29th Aerospace Science Meeting,
Reno (Nevada), AIAA paper 91-0375, 1991.
[58] Nieuwstadt FTM, Meeder JP. Large-eddy simulation of air
pollution dispersion: a review. In: Metais O, Ferziger J,
editors. New tools in turbulence modelling. Berlin: Les
Editions de PhysiqueSpringer, 1997. p. 26480.
[59] Meeder JP, Nieuwstadt FTM. Subgrid-scale segregation of
chemically reactive species in a neutral boundary layer. In:
Chollet JP, Voke PR, Kleiser L, editors. Direct and large
eddy simulation II. Dordrecht: Kluwer Academic Publishers,
1997. p. 30110.
[60] Bray KNC. Turbulent ows with premixed reactants in
turbulent reacting ows. In: Libby PA, Williams FA, editors.
Topics in applied physics, vol. 44. New York: Springer,
1980. p. 115.
[61] Borghi R. On the structure and morphology of turbulent
premixed ames. Rec Adv Aerosp Sci 1985:11738.
[62] Borghi R, Destriau M. Combustion and ames, chemical and
physical principles. Paris: Editions Technip, 1998.
[63] Peters N. Laminar amelet concepts in turbulent combustion.
Twenty-First Symposium (International) on Combustion.
Pittsburgh: The Combustion Institute, 1986. p. 123150.
[64] Tennekes H, Lumley JL. A rst course in turbulence.
Cambridge: MIT Press, 1972.
[65] Poinsot T, Veynante D, Candel S. Quenching processes and
premixed turbulent combustion diagrams. J Fluid Mech
1991;228:561606.
[66] Roberts WL, Driscoll JF, Drake MC, Goss LP. Images of the
quenching of a ame by a vortex: to quantify regimes of
turbulent combustion. Combust Flame 1993;94:5869.
[67] Bilger RW. The structure of turbulent non premixed ames.
Twenty-Second Symposium (International) on Combustion.
Pittsburgh: The Combustion Institute, 1988.
[68] Borghi R. Turbulent combustion modelling. Prog Energy
Combust Sci 1988;14:24592.
[69] Y.Y. Lee, Nonpremixed reacting ows near extinction. PhD
Thesis. Cornell University, 1994.
[70] Cuenot B, Poinsot TJ. Effects of curvature and unsteadiness
in diffusion ames, implications for turbulent diffusion
combustion. Twenty-Fifth Symposium (International) on
Combustion. Pittsburgh: The Combustion Institute, 1994.
[71] Cook A, Riley JJ. Direct numerical simulation of a turbulent
reactive plume on a parallel computer. J Comp Phys
1996;129(2):26383.
[72] Libby PA, Williams FA. Turbulent combustion: fundamental
aspects and a review. In: Libby PA, Williams FA, editors.
Turbulent reacting ows. London: Academic Press, 1994.
p. 261.
[73] Cuenot B, Egolfopoulos FN, Poinsot T. An unsteady laminar
amelet model for nonpremixed combustion. Combust
Theor Model 2000;4(1):77.
[74] Mahalingam S, Chen JH, Vervisch L. Finite-rate chemistry
and transient effects in direct numerical simulations of turbu-
lent nonpremixed ames. Combust Flame 1995;102(3):285
97.
[75] Bray KNC. The challenge of turbulent combustion. Twenty-
Sixth Symposium (International) on Combustion. Pittsburgh:
The Combustion Institute, 1996.
[76] Williams FA. Turbulent combustion. In: Buckmaster JD,
editor. The mathematics of combustion. Philadelphia:
SIAM, 1985. p. 11651.
[77] Kerstein AR, Ashurst W, Williams FA. Field equation for
interface propagation in an unsteady homogeneous ow
eld. Phys Rev A 1988;37(7):272831.
[78] Poinsot T, Echekki T, Mungal MG. A study of the laminar
ame tip and implications for premixed turbulent combus-
tion. Combust Sci Technol 1991;81(13):45.
[79] Smiljanovski V, Moser V, Klein R. A capturingtracking
hybrid scheme for deagration discontinuities. Combust
Theor Model 1997;1(2):183215.
[80] Bourlioux A, Moser V, Klein R. Large eddy simulations of
turbulent premixed ames using a capturing/tracking hybrid
approach. Sixth International Conference on Numerical
Combustion, New Orleans, Louisiana 1996.
[81] Kim W-W, Menon S, Mongia HC. Large-eddy simulation of
a gas turbine combustor ow. Combust Sci Technol
1999;143:2562.
[82] Piana J, Veynante D, Candel S, Poinsot T. Direct numerical
simulation analysis of the G-equation in premixed combus-
tion. In: Chollet JP, Voke PR, Kleiser L, editors. Direct and
large eddy simulation II. Dordrecht: Kluwer Academic
Publishers, 1997. p. 32130.
[83] Fichot F, Lacas F, Veynante D, Candel S. One-dimensional
propagation of a premixed turbulent ame with a balance
equation for the ame surface density. Combust Sci Technol
1993:90.
[84] Marble FE, Broadwell JE. The coherent ame model of non-
premixed turbulent combustion. Project Squid TRW-9-PU,
Project Squid Headquarters, Chaffee Hall, Purdue Univer-
sity, 1977.
[85] Candel S, Poinsot T. Flame stretch and the balance equation
for the ame area. Combust Sci Technol 1990;70:115.
[86] Pope SB. The evolution of surfaces in turbulence. Int J
Engng Sci 1988;26(5):44569.
[87] Vervisch L, Bidaux E, Bray KNC, Kollmann W. Surface
density function in premixed turbulent combustion model-
ing, similarities between probability density function and
ame surface approach. Phys Fluids A 1995;7(10):2496
503.
[88] Van Kalmthout E, Veynante D. Direct numerical simulations
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 262
analysis of ame surface density models for non premixed
turbulent combustion. Phys Fluids A 1998;10(9):234768.
[89] Peters N. Turbulent combustion. Cambridge: Cambridge
University Press, 2000.
[90] Weller HG, Marooney CJ, Gosman AD. A new spectral
method for calculation of the time-varying area of a laminar
ame in homogeneous turbulence. 23rd Symposium (Inter-
national) on Combustion. Pittsburgh: The Combustion
Institute, 1990. p. 62936.
[91] Weller HG. The development of a new ame area combus-
tion model using conditional averaging. Thermo-Fluids
Section Report TF/9307. Imperial College of Science Tech-
nology and Medicine, 1993.
[92] O'Brien EE. The probability density function (pdf) approach
to reacting turbulent ows. In: Williams FA, Libby PA,
editors. Turbulent reacting ows. London: Academic Press,
1980. p. 185.
[93] Pope SB. Pdf method for turbulent reacting ows. Prog
Energy Combust Sci 1985;11:11995.
[94] Kollmann W. The pdf approach to turbulent ow. Theor
Comp Fluid Dynam 1990;1:285349.
[95] Dopazo C. Recent developments in pdf methods. In: Libby
PA, Williams FA, editors. Turbulent reacting ows. London:
Academic Press, 1994. p. 375474.
[96] Dopazo C, Valino L, Fuego F. Statistical description of the
turbulent mixing of scalar elds. Int J Mod Phys B
1997;11(25).
[97] Chen JY, Kollmann W. Pdf modeling of chemical nonequi-
librium effects in turbulent non-premixed hydrocarbon
ames. Twenty-Second Symposium (International) on
Combustion. Pittsburgh: The Combustion Institute, 1988.
[98] Abramowitz M, Stegun IA. Handbook of mathematical func-
tions. NY, USA: Dover, 1970.
[99] Bradley D, Lawes M, Scott MJ. Combust Flame
1994;99:58190.
[100] Lundgren TS. Distribution function in the statistical theory of
turbulence. Phys Fluids 1967;10(5):96975.
[101] Pope SB. Computations of turbulent combustion: progress
and challenges. Twenty-Third Symposium (International)
on Combustion. Pittsburgh: The Combustion Institute, 1990.
[102] Klimenko AY. Fluid Dynam 1990;25:327.
[103] Bilger RW. Conditional moment closure for turbulent react-
ing ow. Phys Fluids 1993;5(2):32734.
[104] Chen J-Y, Kollmann W. Comparison of prediction and
measurement in non-premixed turbulent ames. In: Libby
PA, Williams FA, editors. Turbulent reacting ows. London:
Academic Press, 1994. p. 213308.
[105] Klimenko AY, Bilger RW. Prog Energy Combust Sci
1999;25(6):595687.
[106] Smith NSA. Conditional moment closure of mixing and reac-
tion in turbulent nonpremixed combustion. In: Moin P,
Reynolds WC, editors. Annual research briefs. Stanford,
CA: CTR, Stanford University, 1996. p. 8599.
[107] Swaminathan N, Bilger RW. Assessment of combustion
submodels for turbulent nonpremixed hydrocarbon ames.
Combust Flame 1999;116(4):51945.
[108] Borghi R. Turbulent premixed combustion: further discus-
sions on the scales of uctuations. Combust Flame
1990;80:30412.
[109] Abdel-Gayed RG, Bradley D, Hamid MN, Lawes M. Lewis
number effects on turbulent burning velocity. 20th Sympo-
sium (International) on Combustion. Pittsburgh: The
Combustion Institute, 1984. p. 50512.
[110] Abdel-Gayed RG, Bradley D. Combustion regimes and the
straining of turbulent premixed ames. Combust Flame
1989;76:2138.
[111] Yakhot V, Orszag CG, Thangam S, Gatski TB, Speziale
CG. Development of turbulence models for shear ows by
a double expansion technique. Phys Fluids 1992;4(7):
1510.
[112] Gouldin FC. Combustion intensity and burning rate integral
of premixed ames. Twenty-Sixth Symposium (Interna-
tional) on Combustion. Pittsburgh: The Combustion Insti-
tute, 1996. p. 3818.
[113] Im HG, Lund TS, Ferziger JH. Large eddy simulation of
turbulent front propagation with dynamic subgrid models.
Phys Fluids A 1997;9(12):382633.
[114] Spalding DB. Mixing and chemical reaction in steady
conned turbulent ames. 13th Symposium (International)
on Combustion. Pittsburgh: The Combustion Institute,
1971. p. 64957.
[115] Said R, Borghi R. A simulation with a cellular automaton for
turbulent combustion modelling, Twenty-Second Sym-
posium (International) on Combustion. Pittsburgh: The
Combustion Institute, 1988 p. 56977.
[116] Moss JB, Bray KNC. A unied statistical model of the
premixed turbulent ame. Acta Astronaut 1977;4:291319.
[117] Mantel T, Borghi R. A new model of premixed wrinkled
ame propagation based on a scalar dissipation equation.
Combust Flame 1994;96(4):443.
[118] Lahjaily H, Champion M, Karmed D, Bruel P. Introduction
of dilution in the BML model: application to a stagnating
turbulent ame. Combust Sci Technol 1998;135:15373.
[119] Poinsot T, Veynante D, Trouve A, Ruetsch G. Turbulent
ame propagation in partially premixed ames. Proceedings
of the 1996 Summer Program. Center for Turbulence
Research, 1996. p. 11136.
[120] Veynante D, Duclos JM, Piana J. Experimental analysis of
amelet models for premixed turbulent combustion, Twenty-
Fifth Symposium (International) on Combustion. Pittsburgh:
The Combustion Institute, 1994.
[121] Veynante D, Piana J, Duclos JM, Martel C. Experimental
analysis of ame surface density model for premixed turbu-
lent combustion. Twenty-Sixth Symposium (International)
on Combustion. Pittsburgh: The Combustion Institute,
1996. p. 41320.
[122] Duclos JM. Etude theorique et experimentale d'une amme
turbulente stabilisee par un barreau cylindrique. PhD Thesis.
Ecole Centrale Paris, 1997.
[123] Martel C. Etude experimentale de la combustion turbulente
premelangee. Analyse de modeles. PhD Thesis. Ecole
Centrale Paris, 1998.
[124] Bray KNC, Champion M, Libby PA. The interaction
between turbulence and chemistry in premixed turbulent
ames. In: Borghi R, Murphy SN, editors. Turbulent reacting
ows, Lecture Notes in Engineering, vol. 40. Berlin:
Springer, 1989. p. 54163.
[125] Gouldin FC. Review of closure models for the amelet
regime of premixed turbulent combustion. Joint Meeting of
the British and German Sections of the Combustion Institute,
Queens' College, Cambridge, UK, 1993. p. 525.
[126] Gouldin FC, Miles PC. Chemical closure and burning rates in
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 263
premixed turbulent ames. Combust Flame 1995;100(1/
2):201.
[127] Meneveau C, Poinsot T. Stretching and quenching of ame-
lets in premixed turbulent combustion. Combust Flame
1991;86:31132.
[128] Bailly P. Contribution l'etude de l'interaction turbulence
combustion dans les ammes turbulentes de premelange a
l'aide de modeles du second ordre. PhD Thesis. Universite de
Poitiers, 1996.
[129] Bailly P, Garreton D, Simonin O, Bruel P, Champion M,
Deshaies B, Duplantier S, Sanquer S. Experimental and
numerical study of a premixed ame stabilized by a rectan-
gular section cylinder, Twenty-Sixth Symposium (Interna-
tional) on Combustion. Pittsburgh: The Combustion
Institute, 1996. p. 92330.
[130] Duclos JM, Veynante D, Poinsot T. A comparison of amelet
models for premixed turbulent combustion. Combust Flame
1993;95(1/2):10118.
[131] Bray KNC, Champion M, Libby PA. Premixed ames in
stagnating turbulence part iv: a new theory for the Reynolds
stresses and Reynolds uxes applied to impinging ows.
Combust Flame 2000;120(1/2):118.
[132] Domingo P, Bray KNC. Laminar amelet expressions for
pressure uctuation terms in second moment models of
premixed turbulent combustion. Combust Flame
2000;121(4):55574.
[133] Gouldin FC, Bray KNC, Chen JY. Chemical closure model
for fractal amelets. Combust Flame 1989;77:24159.
[134] Cant RS, Pope SB, Bray KNC. Modelling of amelet surface
to volume ratio in turbulent premixed combustion. 23rd
Symposium (International) on Combustion, Orleans. Pitts-
burgh: The Combustion Institute, 1990. p. 80915.
[135] Prasad ROS, Gore JP. An evaluation of ame surface density
models for turbulent premixed jet ames. Combust Flame
1999;116(12):114.
[136] Cheng WK, Diringer JA. Numerical modelling of SI engine
combustion with a ame sheet model. SAE paper 910268.
International Congress and Exposition, Detroit, 1991.
[137] Choi CR, Huh KY. Development of a coherent amelet
model for a spark ignited turbulent premixed ame in a
closed vessel. Combust Flame 1998;114(3/4):33648.
[138] Checkel MD, Thomas A. Turbulent combustion of premixed
ames in closed vessels. Combust Flame 1994;96(4):351
70.
[139] Prasad ROS, Paul RN, Sivathanu YR, Gore JP. An evaluation
of combined ame surface density and mixture fraction
models for nonisenthalpic premixed turbulent ames.
Combust Flame 1999;117(3):51428.
[140] Veynante D, Trouve A, Bray KNC, Mantel T. Gradient and
counter-gradient scalar transport in turbulent premixed
ames. J Fluid Mech 1997;332:26393.
[141] Dopazo C. On conditioned averages for intermittent turbu-
lent ows. J Fluid Mech 1977;81:4338.
[142] Trouve A. The production of premixed ame surface area in
turbulent shear ow. Combust Flame 1994;99:68796.
[143] Libby PA, Bray KNC. Countergradient diffusion in premixed
turbulent ames. AIAA J 1981;19:20513.
[144] Masuya G, Libby P. Non gradient theory for oblique turbulent
ame with premixed reactants. AIAA J 1981;19:20513.
[145] Trouve A, Veynante D, Bray KNC, Mantel T. The coupling
between ame surface dynamics and species mass conserva-
tion in premixed turbulent combustion. In: Center for Turbu-
lence Research, editor. Proceedings of the Summer Program,
Stanford, Stanford: Center for Turbulence Research, 1994.
p. 95124.
[146] Rutland C, Cant RS. Turbulent transport in premixed ames.
In: Moin P, Reynolds WC, editors. Studying turbulence
using numerical databasesV. Stanford, CA: CTR, Stan-
ford University, 1994. p. 7594.
[147] Kalt PAM. Experimenal investigation of turbulent scalar ux
in premixed ames. PhD Thesis. University of Sydney, 1999.
[148] Boughanem H. Evaluation des termes de transport et de
dissipation de surface de amme par simulation numerique
directe de la combustion turbulente. PhD Thesis. University
of Rouen, France, 1998.
[149] Boughanem H, Trouve A. The occurence of ame instabil-
ities in turbulent premixed combustion. 27th Symposium
International on Combustion 1998.
[150] Paul RN, Bray KNC. Study of premixed turbulent combus-
tion including LandauDarrieus instability effects. Twenty-
Sixth Symposium (International) on Combustion. Pittsburgh:
The Combustion Institute, 1996. p. 25966.
[151] Bray KNC, Moss JB, Libby PA. Turbulent transport in
premixed turbulent ames. In: Zierep J, Oertel H, editors.
Convective transport and instability phenomena. Germany:
University of Karlsuhe, 1982.
[152] Veynante D, Poinsot T. Effects of pressure gradients on
turbulent premixed ames. J Fluid Mech 1997;353:83114.
[153] Libby PA. Theoretical analysis of the effect of gravity on
premixed turbulent ames. Combust Sci Technol
1989;68:1533.
[154] Chen JY, Lumley JL, Gouldin FC. Modeling of wrinkled
laminar ames with intermittency and conditional statistics.
21st Symposium (International) on Combustion. Pittsburgh:
The Combustion Institute, 1986. p. 148391.
[155] Boger M, Veynante D. Large eddy simulations of a turbulent
premixed V-shape ame. Eighth European Turbulence
Conference, Barcelona 2000.
[156] Koochesfahani MM, Dimotakis PE. Mixing and chemical
reactions in turbulent liquid mixing layer. J Fluid Mech
1986;170:83112.
[157] McMurty PA, Riley JJ, Metcalfe RW. Effects of heat release
on the large-scale structure in turbulent mixing layer. J Fluid
Mech 1989;199:297332.
[158] Hermanson JC, Dimotakis PE. Effects of heat release in
turbulent, reacting shear layer. J Fluid Mech 1989;199:
33375.
[159] Jones WP. Turbulence modelling and numerical solution
methods for variable density and combusting ows. In:
Libby PA, Williams FA, editors. Turbulent reacting ows,
London: Academic Press, 1994. p. 30968.
[160] Magnussen BF, Hjertager BH. On the mathematical model-
ing of turbulent combustion with special emphasis on soot
formation and combustion. Sixteenth Symposium (Interna-
tional) on Combustion. Pittsburgh: The Combustion Insti-
tute, 1976. p. 71929.
[161] Ferziger JH, Peric M. Computational methods for uid
dynamics. Berlin: Springer, 1996.
[162] Coupland J, Priddin CH. Modelling of ow and combustion
in a production gas turbine combustor. In: Durst F, Launder
BE, Lumley JL, Schmidt FW, Whitelaw JH, editors. Turbu-
lent shear ows 5th. Berlin: Springer, 1987. p. 31023.
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 264
[163] BarlowRS, Dibble RW, Chen JY, Lucht RP. Combust Flame
1990(82).
[164] Montgomery CJ, Kosaly G, Riley JJ. Direct numerical simu-
lation of turbulent H
2
O
2
combustion using reduced chem-
istry. In: AIAA Paper 93-0248. 31st Aerospace Sciences
Meeting and Exhibit, Reno, NV, 1993. p. 110.
[165] Peters N. Prog Energy Combust Sci page 1984:319.
[166] Sun CJ, Sung CJ, Wang H, Law CK. On the structure of
nonsooting counterow ethylene and acetylene diffusion
ames. Combust Flame 1996;107(4):32135.
[167] Sung CJ, Liu JB, Law CK. Structural response of counter-
ow diffusion ames to strain rate variations. Combust
Flame 1995;102:48192.
[168] Rogg B. The Cambridge universal laminar amelet computer
code. In: Peters N, Rogg B, editors. Reduced kinetic mechan-
isms for applications in combustion systems, Appendix C.
Berlin: Springer, 1993.
[169] Nandula SP, Brown TM, Pitz RW. Measurements of scalar
dissipation in the reaction zones of turbulent nonpremixed
H
2
air ames. Combust Flame 1994;99:77583.
[170] Bish ES, Dahm WJA. A strained dissipation and reaction
layer formulation for turbulent diffusion ames. Western
States Section Fall Meeting. Pittsburgh: The Combustion
Institute, 1993.
[171] Ferreira JC. Flamelet modelling of stabilization in turbulent
non-premixed combustion. PhD Thesis. Swiss Federal Insti-
tute of Technology, ETH, Zurich, 1996.
[172] Haworth DC, Drake MC, Pope SB, Blint RJ. The importance
of time-dependent ame structures in stretched laminar
amelet models for turbulent jet diffusion ames. Twenty-
Second Symposium (International) on Combustion. Pitts-
burgh: The Combustion Institute, 1988. p. 58997.
[173] Pitsch H, Wan YP, Peters N. SAE Paper, 952357, 1995.
[174] Zhang Y, Rogg B, Bray KNC. 2-d Simulation of turbulent
autoignition with transient laminar amelet source term
closure. Combust Sci Technol 1995;105:21127.
[175] Darabiha N. Transient behaviour of laminar counterow
hydrogenair diffusion ames with complex chemistry.
Combust Sci Technol 1992;86:16381.
[176] Mauss F, Keller D, Peters N. Alagrangian simulation of ame-
let extinction and re-ignition in turbulent jet diffusion ames.
Twenty-Third Symposium (International) on Combustion.
Pittsburgh: The Combustion Institute, 1990. p. 6938.
[177] Pitsh H, Chen M, Peters N. Unsteady amelet modeling of
turbulent hydrogenair diffusion ames. Twenty-Seventh
Symposium (International) on Combustion. Pittsburgh: The
Combustion Institute, 1998. p. 105764.
[178] Mell WE, Nilsen VN, Kosaly G, Riley JJ. Direct numerical
simulation investigation of the conditional moment closure
model for nonpremixed turbulent reacting ows. Combust
Sci Technol 1991;91:17986.
[179] Delhaye B. Etude des ammes de diffusion turbulentes.
Simulations directes et modelisation. PhD Thesis. Ecole
Centrale Paris, 1994.
[180] Delhaye B, Veynante D, Candel S. Simulation and modeling
of reactive shear layers. Theoret Comput Fluid Dynam
1994;6:6787.
[181] Chang CHH, Dahm WJA, Trygvason G. Lagrangian model
simulations of molecular mixing, including nite rate chemi-
cal reactions, in a temporally developing shear layer. Phys
Fluid 1991;3(5):130011.
[182] Lee YY, Pope SB. Nonpremixed turbulent reacting ow near
extinction. Combust Flame 1995;101:50128.
[183] Bradley D, Gaskell PH, Gu XJ. The mathematical modeling
of liftoff and blowoff of turbulent non-premixed methane jet
ames at high strain rate. Proc Combust Inst 1998;27:1199
206.
[184] Cuenot B, Egolfopoulos FN, Poinsot T. An unsteady laminar
amelet model for non-premixed combustion. Combust
Theor Mod 2000;4:7797.
[185] Muller CM, Breitbach H, Peters N. Partially premixed turbu-
lent ame propagation in jets ames. Twenty-Fifth Sympo-
sium (International) on Combustion. Pittsburgh: The
Combustion Institute, 1994. p. 1099106.
[186] Fichot F, Delhaye B, Veynante D, Candel S. Strain rate
modelling for a ame surface density equation with applica-
tion to non-premixed turbulent combustion, 25th Symposium
(International) on Combustion. Pittsburgh: The Combustion
Institute, 1994.
[187] van Kalmthout E, Veynante D, Candel S. Direct numerical
simulation analysis of ame surface density equation in non-
premixed turbulent combustion, Twenty-Sixth Symposium
(International) on Combustion. Pittsburgh: The Combustion
Institute, 1996.
[188] Van Kalmthout E, Veynante D. Analysis of ame surface
density concepts in non-premixed turbulent combustion
using direct numerical simulation. Eleventh Symposium on
Turbulent Shear Flows, Grenoble, France 1997.
[189] Obounou M, Gonzalez M, Borghi R. A lagrangian model for
predicting turbulent diffusion ames with chemical kinetic
effects. Twenty-Fifth Symposium (International) on Combus-
tion. Pittsburgh: The Combustion Institute, 1994. p. 110713.
[190] Broadwell JE, Lutz A. Aturbulent jet chemical reaction model:
NO
x
production in jet ames. Combust Flame 1998:114.
[191] Ravet F, Vervisch L. Modeling non-premixed turbulent
combustion in aeronautical engines using pdf-generator.
36th Aerospace Sciences Meeting and Exhibit AIAA paper
98-1027, Reno, NV 1998.
[192] Girimaji SS. Phys Fluids 1992;4:2529.
[193] Smith NSA, Ruetsch GR, Oefelein J, Ferziger JH. Simulation
and modeling of reactive particles in turbulent nonpremixed
combustion. In: Moin P, Reynolds WC, editors. Studying
turbulence using numerical databasesVII. Stanford, CA:
CTR, Stanford University, 1998. p. 3960.
[194] Kerstein AR. A linear-eddy model for turbulent scalar trans-
port and mixing. Combust Sci Technol 1988;60:391421.
[195] Fox RO. The lagrangian spectral relaxation model for differ-
ential diffusion in homogeneous turbulence. Phys Fluids
1999;11(6):155071.
[196] Fox RO. The spectral relaxation model of the scalar dissipa-
tion rate in homogeneous turbulence. Phys Fluids
1995;7(5):108294.
[197] Van Slooten Jayesh PR, Pope SB. Advances in pdf modeling
for inhomogeneous turbulent ows. Phys Fluids 1998;10(1):
24665.
[198] Spielman LA, Levenspiel O. A Monte Carlo treatment for
reacting and coalescing dispersed phase systems. Chem
Engng Sci 1965;20:24754.
[199] Valino L. Aeld Monte Carlo formulation for calculating the
probability density function of a single scalar in turbulent
ow. Flow Turb Combust 1998;60(2):15772.
[200] Fox RO. Computational methods for turbulent reacting ows
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 265
in the chemical process industry. Revue de L'Institut Fran-
cais du Petrole 1996;51(2).
[201] Corrsin S. The isotropic turbulent mixer. Part II. Arbitrary
Schmidt number. AIChE J 1964;10:8707.
[202] Dopazo C. Non-isothermal turbulent reactive ows: stochas-
tic approaches. PhD Thesis. University of New York, Stony
Brook, 1973.
[203] Villermaux J. Micromixing phenomena in stirred reactors.
Encycl Fluid Mech 1986:707.
[204] Eswaran V, Pope SB. Direct numerical simulation of the
turbulent mixing of a passive scalar. Phys Fluids
1988;31(3):50620.
[205] Tsai K, Fox RO. The bmc/giem model for micromixing in
non-premixed turbulent reacting ows. Ind Engng Chem Res
1998:6.
[206] Curl RI. Dispersed phase mixing. Theory and effects in
simple reactors. AIChE J 1963;175(9).
[207] Masri AR, Bilger RW, Dibble RW. Turbulent non-premixed
ames of methane near extinction. Combust Flame
1988;73:2618.
[208] Jones WP, Kakhi M. Pdf modeling of nite rate chemistry
effects in turbulent nonpremixed jet ames. Combust Flame
1998;115(1/2):21029.
[209] Jones WP, Lindstedt RP. Global reaction schemes for hydro-
carbon combustion. Combust Flame 1988;73:23349.
[210] Dopazo C. Relaxation of initial probability density functions
in the turbulent convection of scalar elds. Phys Fluids
1979;22(20).
[211] Janicka J, Kolbe W, Kollmann W. Closure of the transport
equation for the probability density function of turbulent
scalar elds. J Non-Equilib Thermodyn 1979;4(47).
[212] Pope SB. An improved turbulent mixing model. Combust Sci
Technol 1982;28:1315.
[213] Pope SB. Consistent modeling of scalars in turbulent ows.
Phys Fluids 1982;26:4048.
[214] Leonard A, Hill J. Direct numerical simulation of turbulent
ows with chemical reaction. J Scient Comput
1988;3(1):2543.
[215] Kraichnan RH. Closures for probability distributions. Bull
Am Phys Soc 1989;34:2298.
[216] Dutta A, Tarbell JM. Closure models for turbulent reacting
ows. AIChE J 1989;35(12).
[217] Gao F. An analytical solution for the scalar probability
density function in homogeneous turbulence. Phys Fluids
1991;3(4):5113.
[218] Valino L, Dopazo C. A binomial sampling model for scalar
turbulent mixing. Phys Fluids A 1990;2(7):120412.
[219] Norris AT, Pope SB. Turbulent mixing model based on
ordered pairing. Combust Flame 1991;83:2742.
[220] Valino L, Dopazo C. A binomial langevin model for turbu-
lent mixing. Phys Fluids A 1991;3(12):30347.
[221] Valino L, Ros J, Dopazo C. Monte Carlo implementation and
analytic solution of an inert-scalar turbulent-mixing test
problem using mapping closure. Phys Fluids A
1991;3(9):21918.
[222] Pope SB. Mapping closure for turbulent mixing and reaction.
Theoret Comput Fluid Dynam 1991;2:25570.
[223] Gao F. Mapping closure and non-gaussianity of the scalar
probability density functions in isotropic turbulence. Phys
Fluids A 1991;3(10):243844.
[224] Gao F, O'Brien EE. Joint probability density function of a
scalar and its gradient in isotropic turbulence. Phys Fluids
1991;3(6).
[225] Gao F, O'Brien EE. A mapping closure for multispecies
ckian diffusion. Phys Fluids 1991;3(3).
[226] Kerstein AR. Linear-eddy modelling of turbulent transport.
Part 6. Microstructure of diffusive scalar mixing elds. J
Fluid Mech 1991;231:36194.
[227] Chen J-Y, Kollmann W. Pdf modeling and analysis of ther-
mal NO formation in turbulent nonpremixed hydrogenair
jet ames. Combust Flame 1992;88:397412.
[228] Jiang T-L, Givi P, Gao F. Binary and trinary scalar mixing by
Fickian diffusionsome mapping closure results. Phys
Fluids A 1992;4(5):102835.
[229] Kerstein AR. Linear-eddy modelling of turbulent transport.
Part 7. Finite rate chemistry and multi-stream mixing. J Fluid
Mech 1992;240:289313.
[230] Fox RO. Improved FokkerPlanck model for the joint scalar,
scalar gradient pdf. Phys Fluids 1994;1(6):33448.
[231] Masri AR, Subramaniam S, Pope SB. A mixing model to
improve the pdf simulation of turbulent diffusion ames.
Twenty-Sixth Symposium (International) on Combustion.
Pittsburgh: The Combustion Institute, 1996.
[232] Fox RO. On velocity-conditioned scalar mixing in homo-
geneous turbulence. Phys Fluids 1996;8(10):267891.
[233] Fox RO. On the relationship between lagrangian micromix-
ing models and computational uid dynamics. Chem Engng
Process 1998;37:52135.
[234] Reveillon J, Vervisch L. Subgrid mixing modeling: a
dynamic approach. AIAA J 1998;36(3):33641.
[235] Subramaniam S, Pope SB. A mixing model for turbulent
reactive ows based on Euclidean minimum spanning
trees. Combust Flame 1998;115(4):487514.
[236] Subramaniam S, Pope SB. Comparison of mixing model
performance for nonpremixed turbulent reacting ows.
Combust Flame 1998;117(4):73254.
[237] Anand MS, Pope SB. Calculations of premixed turbulent
ames by pdf methods. Combust Flame 1987:67.
[238] Haworth DC, Pope SB. A generalized Langevin model for
turbulent ows. Phys Fluids 1986;29:20828.
[239] Gardiner CW. Handbook of stochastic methods. Berlin:
Springer, 1997.
[240] Pope SB. On the relationship between stochastic Lagrangian
models of turbulence and second-moment closures. Phys
Fluids 1994;6:97385.
[241] Wouters HA, Peeters TWJ, Roekaerts D. On the existence of
a stochastic Lagrangian model representation for second-
moment closures. Phys Fluids 1996;8(7):17024.
[242] Pope SB. Lagrangian pdf methods for turbulent ows. Annu
Rev Fluid Mech 1994;26:23.
[243] Reveillon J, Vervisch L. Spray vaporization in non-premixed
turbulent ames: a single droplet model. Combust Flame
2001;121(1/2):7590.
[244] Hollmann C, Gutheil E. Modeling of turbulent spray diffu-
sion ames including detailed chemistry. Twenty-Sixth
Symposium (International) on Combustion. Pittsburgh: The
Combustion Institute, 1996. p. 17318.
[245] Calimez X. Simulation a petite echelle par une methode VOF
d'ecoulement diphasiques reactifs. PhD Thesis. Ecole
Centrale Paris, 1998.
D. Veynante, L. Vervisch / Progress in Energy and Combustion Science 28 (2002) 193266 266

También podría gustarte