PDF Modeling and Simulation of Premixed Turbulent Combustion

Similar documents
PDF modeling and simulation of premixed turbulent combustion

A Method for the Prognosis of Turbulent Combustion

Turbulent Premixed Combustion

Lecture 14. Turbulent Combustion. We know what a turbulent flow is, when we see it! it is characterized by disorder, vorticity and mixing.

Lecture 8 Laminar Diffusion Flames: Diffusion Flamelet Theory

Lecture 9 Laminar Diffusion Flame Configurations

A G-equation formulation for large-eddy simulation of premixed turbulent combustion

IMPROVED POLLUTANT PREDICTIONS IN LARGE-EDDY SIMULATIONS OF TURBULENT NON-PREMIXED COMBUSTION BY CONSIDERING SCALAR DISSIPATION RATE FLUCTUATIONS

LES of the Sandia Flame D Using an FPV Combustion Model

D. VEYNANTE. Introduction à la Combustion Turbulente. Dimanche 30 Mai 2010, 09h00 10h30

Laminar Premixed Flames: Flame Structure

Flow and added small-scale topologies in a turbulent premixed flame

Exercises in Combustion Technology

Lecture 10 Turbulent Combustion: The State of the Art

Towards regime identification and appropriate chemistry tabulation for computation of autoigniting turbulent reacting flows

A comparison between two different Flamelet reduced order manifolds for non-premixed turbulent flames

Numerical Investigation of Ignition Delay in Methane-Air Mixtures using Conditional Moment Closure

Introduction to Turbulence and Turbulence Modeling

Probability density function (PDF) methods 1,2 belong to the broader family of statistical approaches

Consistent turbulence modeling in a hybrid LES/RANS PDF method for non-premixed flames

Process Chemistry Toolbox - Mixing

Modeling Turbulent Combustion

Modeling flame brush thickness in premixed turbulent combustion

Best Practice Guidelines for Combustion Modeling. Raphael David A. Bacchi, ESSS

Tutorial School on Fluid Dynamics: Aspects of Turbulence Session I: Refresher Material Instructor: James Wallace

Extinction and reignition in a diffusion flame: a direct numerical simulation study

Subgrid-scale mixing of mixture fraction, temperature, and species mass fractions in turbulent partially premixed flames

Lecture 15. The Turbulent Burning Velocity

Mass Transfer in Turbulent Flow

Flamelet modelling of non-premixed turbulent combustion with local extinction and re-ignition

LES Approaches to Combustion

Large-eddy simulation of an industrial furnace with a cross-flow-jet combustion system

CFD and Kinetic Analysis of Bluff Body Stabilized Flame

Simulation of Turbulent Lifted Flames and their Transient Propagation

Published in: Proceedings of the Fluent Benelux User Group Meeting, 6-7 October 2005, Wavre, Belgium

The Hybrid Method for the PDF Equations of Turbulent Reactive Flows: Consistency Conditions and Correction Algorithms

Large-Eddy Simulation of Turbulent Combustion

Fluid Dynamics and Balance Equations for Reacting Flows

Advanced Turbulence Models for Emission Modeling in Gas Combustion

The influence of C ϕ is examined by performing calculations with the values C ϕ =1.2, 1.5, 2.0 and 3.0 for different chemistry mechanisms.

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

TURBULENT COMBUSTION NORBERT PETERS. Institut für Technische Mechanik Rheinisch-Westfälische Technische Hochschule Aachen, Germany

Overview of Turbulent Reacting Flows

S. Kadowaki, S.H. Kim AND H. Pitsch. 1. Motivation and objectives

LARGE-EDDY SIMULATION OF PARTIALLY PREMIXED TURBULENT COMBUSTION

A NUMERICAL ANALYSIS OF COMBUSTION PROCESS IN AN AXISYMMETRIC COMBUSTION CHAMBER

RANS-SLFM and LES-SLFM numerical simulations of turbulent non-premixed oxy-fuel jet flames using CO2/O2 mixture

Capturing localised extinction in Sandia Flame F with LES-CMC

Before we consider two canonical turbulent flows we need a general description of turbulence.

Comparison of Turbulence Models in the Flow over a Backward-Facing Step Priscila Pires Araujo 1, André Luiz Tenório Rezende 2

NONLINEAR FEATURES IN EXPLICIT ALGEBRAIC MODELS FOR TURBULENT FLOWS WITH ACTIVE SCALARS

Examination of the effect of differential molecular diffusion in DNS of turbulent non-premixed flames

Budget analysis and model-assessment of the flamelet-formulation: Application to a reacting jet-in-cross-flow

Large-eddy simulation of a bluff-body-stabilized non-premixed flame using a recursive filter-refinement procedure

Dr.-Ing. Frank Beyrau Content of Lecture

ANSYS Advanced Solutions for Gas Turbine Combustion. Gilles Eggenspieler 2011 ANSYS, Inc.

DARS overview, IISc Bangalore 18/03/2014

Steady Laminar Flamelet Modeling for turbulent non-premixed Combustion in LES and RANS Simulations

Lecture 6 Asymptotic Structure for Four-Step Premixed Stoichiometric Methane Flames

Soot formation in turbulent non premixed flames

M U S C L E S. Modelling of UnSteady Combustion in Low Emission Systems. Project No. GRD Contract No. G4RD-CT

A Priori Model for the Effective Lewis Numbers in Premixed Turbulent Flames

Comparison of Mixing Model Performance for Nonpremixed Turbulent Reactive Flow

Evaluation of a Scalar Probability Density Function Approach for Turbulent Jet Nonpremixed Flames

Premixed and non-premixed generated manifolds in large-eddy simulation of Sandia flame D and F

A Priori Testing of Flamelet Generated Manifolds for Turbulent Partially Premixed Methane/Air Flames

Well Stirred Reactor Stabilization of flames

Introduction Flares: safe burning of waste hydrocarbons Oilfields, refinery, LNG Pollutants: NO x, CO 2, CO, unburned hydrocarbons, greenhouse gases G

Turbulence Resolution Scale Dependence in Large-Eddy Simulations of a Jet Flame

Turbulence: Basic Physics and Engineering Modeling

2. Conservation Equations for Turbulent Flows

NUMERICAL ANALYSIS OF TURBULENT FLAME IN AN ENCLOSED CHAMBER

Pressure and Fuel Effects on the Flame Brush Thickness of H 2 /CO Flames

An Introduction to Theories of Turbulence. James Glimm Stony Brook University

LES of an auto-igniting C 2 H 4 flame DNS

LES-PDF SIMULATION OF A HIGHLY SHEARED TURBULENT PILOTED PREMIXED FLAME

hydrogen auto ignition in a turbulent co flow of heated air with LES and CMC approach.

Explicit algebraic Reynolds stress models for boundary layer flows

An investigation of the performance of turbulent mixing models

An Unsteady/Flamelet Progress Variable Method for LES of Nonpremixed Turbulent Combustion

7. Turbulent Premixed Flames

A validation study of the flamelet approach s ability to predict flame structure when fluid mechanics are fully resolved

Experimental study of velocity filtered joint density function for large eddy simulation

Scalar gradient and small-scale structure in turbulent premixed combustion

Turbulent Mixing in a Tubular Reactor: Assessment of an FDF/LES Approach

Unified turbulence models for LES and RANS, FDF and PDF simulations

Representing Turbulence/Chemistry interaction with strained planar premixed

DNS of Reacting H 2 /Air Laminar Vortex Rings

Calculations of bluff-body stabilized flames using a joint probability density function model with detailed chemistry

Structures of Turbulent Bunsen Flames in the Corrugated-Flamelet Regime

Turbulent Boundary Layers & Turbulence Models. Lecture 09

Simulation and modeling of reacting particles in turbulent nonpremixed combustion

Effects of Damköhler number on flame extinction and reignition in turbulent nonpremixed flames using DNS

Eddy viscosity. AdOc 4060/5060 Spring 2013 Chris Jenkins. Turbulence (video 1hr):

2.2 The Turbulent Round Jet

Reacting Flow Modeling in STAR-CCM+ Rajesh Rawat

Assessment of a flame surface density-based subgrid turbulent combustion model for nonpremixed flames of wood pyrolysis gas

Impact of numerical method on auto-ignition in a temporally evolving mixing layer at various initial conditions

Large Eddy Simulation of Piloted Turbulent Premixed Flame

Modeling of turbulence in stirred vessels using large eddy simulation

Transcription:

Monte Carlo Methods Appl. Vol. No. (), pp. 43 DOI 5 / MCMA.7. c de Gruyter PDF Modeling and Simulation of Premixed Turbulent Combustion Michael Stöllinger and Stefan Heinz Abstract. The use of probability density function (PDF) methods for turbulent combustion simulations is very attractive because arbitrary finite-rate chemistry can be exactly taken into account. PDF methods are well developed for non-premixed turbulent combustion. However, many real flames involve a variety of mixing regimes (non-premixed, partially-premixed and premixed turbulent combustion), and the development of PDF methods for partially-premixed and premixed turbulent combustion turned out to be a very challenging task. The paper shows a promising way to overcome this problem by extending existing PDF methods such that a variety of mixing regimes can be covered. The latter is done by a generalization of the standard scalar mixing frequency model. The generalized scalar mixing frequency model accounts for several relevant processes in addition to velocity-scalar correlations that are represented by the standard model for the mixing of scalars. The suitability of the new mixing frequency model is shown by applications to several premixed turbulent Bunsen flames which cover various regimes ranging from flamelet to distributed combustion. Comparisons to existing concepts focused on the inclusion of reaction effects in mixing frequency models for nonreacting scalars reveal the advantages of the new mixing frequency model. It is worth noting that the same methodology can be used in corresponding filter density function (FDF) methods. Keywords. Premixed turbulent combustion, Premixed stoichiometric methane-air flames, Transported PDF methods, Scalar mixing frequency.. Introduction The extension of Reynolds-averaged Navier-Stokes (RANS) equations by probability density function (PDF) methods [3,, 3, 9, 4, 45, 48, 5] has significant advantages for turbulent reacting flow simulations due to the fact that there is no need to involve approximations of uncertain generality to close chemical reaction rates. The same advantage is given by the generalization of large eddy simulation (LES) methods by filter density function (FDF) methods [9, 4, 5, 6, 8, 4, 56]. FDF methods are usually more accurate but also more expensive than PDF methods [6, ]. Due to the exact treatment of chemical reactions, the performance of PDF and FDF methods is essentially determined by the modeling of the transport of scalars (e.g., species mass fractions and temperature). Such scalar transport models involve two ingredients: a scalar mixing frequency model that determines the characteristic time scale of mixing, and a scalar mixing model that describes the change of the PDF of a scalar [3, 6, 9, 5].

Michael Stöllinger and Stefan Heinz Most of the previous applications of PDF and FDF methods were related to simulations of non-premixed turbulent combustion. In this case, the characteristic length and time scales of scalar fields are usually larger than the characteristic length and time scales of turbulent motions. Correspondingly, the scalar mixing frequency can be assumed to be controlled by the frequency of large-scale turbulent motions. The performance of scalar mixing models for non-premixed turbulent combustion is relatively well investigated [3, 7, 36, 37, 59]. Mitarai et al. [37] compared predictions of different mixing models to the results of direct numerical simulation (DNS). Merci et al. [3, 33, 34, 35] and Xu and Pope [6] studied various scalar mixing models in turbulent natural gas diffusion flames. The FDF method also has been successfully applied to non-premixed flames by Raman et al. [53] and Sheiki et al. [55]. Applications of PDF and FDF methods to premixed turbulent combustion are more complicated than calculations of non-premixed turbulent combustion. The appearance of fast flamelet chemistry may result in very thin reaction zones such that scalar mixing can take place on scales which are much smaller than all scales of turbulent motions [, 9, 39, 49]. Correspondingly, there exist only a few applications of PDF methods to premixed flames. To overcome problems of earlier approaches, Mura et al. [39] recently suggested a PDF model where the outer parts of the flame structure (reactant side and product side) are described by a standard scalar mixing model whereas the inner part (the reaction zone) is described by a flamelet model. However, this approach is complicated and related to several questions (e.g. regarding the matching of both combustion regimes). The scalar mixing frequency is provided via a transport equation for the scalar dissipation rate which involves effects of chemical reactions in corresponding equations for non-reacting scalars. The model of Mura et al. [39] requires the adjustment of six parameters to the flow considered, the modeling of the dissipation rate of non-reacting scalars does not agree with many other models [54], and there are questions regarding the inclusion of chemical reaction effects (see section 7). More recently, Lindstedt and Vaos [9] and Cha and Trouillet [6, 7] developed models that relate the scalar mixing frequency of reacting scalars to the characteristic frequency of turbulent motions and the scalar mixing frequency of non-reacting scalars, respectively. Both approaches use flamelet ideas in conjunction with several other assumptions: e.g., the assumption of a local equilibrium between production and dissipation in the scalar dissipation rate equation [9], and the assumption of local homogeneity and isotropy [6, 7]). The investigations reported here reveal that these assumptions are not satisfied in general. The paper will address the questions described in the preceding paragraph in the following way. Section deals with a description of the PDF modeling approach. The modeling of molecular scalar mixing is the concern of section 3. The problems related to the use of the standard model for scalar mixing will be described, and a generalization of the standard mixing frequency model will be presented. The following sections describe applications of the generalized mixing frequency model to simulations of

PDF Modeling of Premixed Turbulent Combustion 3 three premixed turbulent Bunsen flames which cover various regimes ranging from flamelet to distributed combustion. The flames considered are discussed in section 4, and the flame simulations are described in section 5. Section 6 presents simulation results obtained with the scalar mixing frequency model introduced in section 3 in comparison to measurements. Comparisons with modeling approaches applied previously will be discussed in section 7. Section 8 deals with conclusions of this work.. The modeling approach The modeling approach will be developed on the basis of PDF equations (which can be used as FDF equations if the ensemble averages involved are replaced by corresponding spatially filtered variables [9, ]). The derivation of the transport equation for the joint PDF F θ (θ; x, t) of a number N s of scalars (e.g., species mass fractions and temperature) can be found in a variety of textbooks. The PDF transport equation obtained reads [, 3, 9, 5] ρ F θ t + ρ u if θ x i = ρ u i θf θ x i N s θ = ( ρ Dφ ) Dt θ F θ. () Here, ρ is the ensemble averaged mass density. A Favre decomposition u i = u i +u i is used for velocities. Favre decompositions of other variables will be represented in the same way, this means Q refers to the mass-density averaged mean of any variable Q, and the corresponding Favre fluctuation is Q = Q Q. The sum convention is used throughout this paper for repeated Roman subscripts but not for Greek subscripts. Dφ θ and Dt θ denote conditional Favre averages, where θ is the sample space scalar vector. According to the Navier-Stokes equations, the substantial derivatives Dφ Dt of scalars are given by Dφ = ( ) ρν φ + S. () Dt ρ x k Sc x k Here, ν denotes the kinematic viscosity, Sc is the Schmidt number of scalar, and S is a source rate due to chemical reactions. By adopting relation (), the scalar PDF transport equation () can be written u i ρ F θ + ρ u if θ = [ ] ν F θ ρ u i t x i x k Sc x θf θ k N s = [ ] ρ M + S F θ. θ (3) The only assumptions used in the derivation of the latter equation are ν = ν and Q/ x k = Q/ x k, where Q refers to any fluctuating variable. The molecular mixing term M in (3) is given by M = F θ N s β= θ β ( ν Sc φ x k φ β x k θf θ ). (4)

4 Michael Stöllinger and Stefan Heinz M M is written as a Favre fluctuation because expression (4) reveals that the mean of disappears, M =. (5) By multiplying equation (3) with the corresponding variables and integrating over the scalar sample space, one finds for scalar means φ and scalar variances φ the transport equations Dφ Dt = ( ) ν φ ρ u ρ x k Sc x k φ + S, (6) k and Dφ Dt ( = ν φ ρ u ρ x k Sc x k k φ ) φ u k φ ε + S x φ. (7) k Here, D/ Dt = / t + u k / x k refers to the mean Lagrangian time derivative, and the scalar dissipation rate ε is given by ε = ν φ φ. (8) Sc x k x k The definition (4) of M determines a relationship between the molecular mixing term M and scalar dissipation rate ε, M φ = ε ν φ φ. (9) Sc x k x k By replacing ε in the scalar variance equation (7) according to relation (9) one obtains a relation between M φ and the transport of scalar variances which will be used below for the modeling of M. The approach applied implies the need to parametrize the conditional turbulent scalar flux u i θ. This flux will be modeled by the usual gradient diffusion assumption [, 3, 9, 5], u i θ F θ = ν t F θ. () Sc t x i Here, Sc t refers to a constant turbulent Schmidt number and ν t = C µ kτ denotes the turbulent viscosity: k is the turbulent kinetic energy, τ is the dissipation time scale of turbulence, and C µ is a dimensionless parameter. Expression () for the conditional turbulent scalar flux implies the usual gradient-diffusion models for the turbulent scalar flux and turbulent scalar variance flux, u i φ = ν t φ, u i Sc t x φ = ν t φ. () i Sc t x i

PDF Modeling of Premixed Turbulent Combustion 5 For simplicity, we will assume that Sc = Sc t. The resulting transport equation for the scalar PDF is then given by ρ F θ t + ρ u i F θ = ρ ν + ν t F θ x i x i Sc t x i Equation () implies for scalar means the equation Dφ Dt N s = ( ) ρ M + S F θ. () θ = ( ) ν + νt φ ρ + S. (3) ρ x k Sc t x k By adopting relation (9) for ε, the scalar variance equation implied by equation () reads ( ) Dφ = ρ ν + ν t φ + ν + ν t φ φ + M Dt ρ x k Sc t x k Sc t x k x φ + S φ. (4) k 3. Molecular mixing modeling The scalar PDF transport equation () is unclosed as long as the molecular mixing term M is not specified. The latter can be done in a variety of ways [, 3, 9, 5]. We will use a parametrization of M according to the interaction-by-exchange-withthe-mean (IEM) mixing model [, 6], M = C τ φ. (5) The model (5) represents the standard mixing model which is used in most simulations of turbulent reacting flows [3, 5]. The reason for that is given by the simplicity of the model (5). The model (5) contains two ingredients: a scalar mixing model that describes changes of scalar values, and a scalar mixing frequency model that determines the characteristic time scale of mixing. The scalar mixing is modeled by φ, which corresponds to the idea that scalar fluctuations φ tend to disappear (scalar values φ tend to relax to their mean value φ ). The scalar mixing frequency ω = C /(τ) is modeled in proportionality to the mixing frequency /τ of largescale turbulent motions, where C is a nondimensional parameter. The consideration of the relatively simple standard mixing model (5) represents a natural first step of investigations of the performance of improved models for the mixing frequency ω in M models []. Justification for this approach arises from the observation that the problem considered requires, first of all, an improved modeling of the scalar mixing frequency ω = C /(τ) (see the discussion in the introduction). The question of whether the use of other scalar mixing models may further improve simulation results will be the concern of future efforts.

6 Michael Stöllinger and Stefan Heinz Expression (5) requires the definition of C in order to specify the scalar mixing frequency ω = C /(τ). The scalar mixing frequency ω determines, first of all, the evolution of the scalar variance. Thus, the equation for scalar variances represents the natural basis for the definition of C. By adopting expression (5) in equation (4), the scalar variance equation can be written e = P C B. (6) Here, B is given by B = φ /τ, and e and P are given by the expressions e = Dφ ( ) ρ ν + ν t φ S Dt ρ x k Sc t x φ, (7) k P = ν + ν t Sc t φ x k φ x k. (8) Equation (6) can be used to find an appropriate constant value for C on the basis of DNS data. The latter can be done, for example, by adopting Overholt and Pope s DNS data of passive scalar mixing in homogeneous isotropic stationary turbulence with imposed constant mean scalar gradient [4]. For the flow considered one finds that e =, this means there is a local equilibrium between the production P and dissipation C B of scalar variances, P = C B. Overholt and Pope s DNS data then provide C = (.8;.;.) for Taylor-scale Reynolds numbers Re λ = (8; 5; 84), respectively (these C values were obtained without using the gradient-diffusion approximation () for the turbulent scalar flux). An analysis of the Reynolds number dependence of C indicates a value C =.5 for infinitely high Taylor-scale Reynolds numbers [9]. Based on these findings, the standard model for C is given by the use of a constant value C for C [3, 9, 5]. However, a constant value of C means that the turbulent velocity and scalar fields are strongly correlated such that the scalar mixing frequency ω = C /(τ) is completely controlled by the frequency /τ of large-scale turbulent motions. There is, however, no guarantee that a constant value for C represents a reasonable approximation for flows which are more complex than homogeneous isotropic and stationary turbulence. For example, the use of a constant value for C appears to be inappropriate if the characteristic time and length scales of scalar fields are found below the corresponding scales of turbulent motions, which is a characteristic feature of premixed turbulent combustion. Correspondingly, several attempts have been made to overcome the limitations of using a constant value for C [7, 9, 38]. Unfortunately, all these approaches presented previously are faced with significant problems: see the detailed discussion provided in section 7. To overcome the problems related to the use of a constant value for C, an attempt to generalize the standard model for C will be made in the following.

PDF Modeling of Premixed Turbulent Combustion 7 The basic idea of the approach applied is the following one. In extension of the standard model for C one may assume that e is small compared to P and C B in (6), this means one assumes that relation (6) can be basically characterized by a local equilibrium between the production P and dissipation C B of scalar variances. Such an approach is often used, for example, for the derivation of algebraic stress models [3, 9, 5]. In other words, e may be seen as a relatively small deviation from a local equilibrium between P and C B. On this basis, C can be calculated by the constraint to minimize the deviation e from the local equilibrium. What is the most convenient way to formulate such a constraint for the calculation of C? C depends on e, P, and B according to relation (6). The calculation of e, P, and B in terms of simulation results will be affected by numerical errors due to the discretization of equations and the finite number of particles involved in the Monte Carlo simulation. Thus, e, P, and B calculated in this way represent fluctuating variables. The most convenient way to calculate C under these conditions is to minimize the averaged squared deviations e from the local equilibrium between production T and dissipation of scalar variances (see the detailed explanations below). Here, Q T refers to a temporal average of any variable Q. This idea of calculating C has some similarity with the calculation of the dynamic Smagorinsky coefficient in LES [5]. However, there is a problem related to this approach. The production P results from the modeling of velocity-scalar correlations u k φ : see equations (7) and (). Thus, the assumption of minimal deviations from a local equilibrium between the production and dissipation of scalar variances means that the mixing of scalars is controlled by the turbulent velocity field. This condition cannot be considered to be satisfied for premixed combustion where the characteristic mixing time scale may well be smaller than the characteristic time scales of the velocity field. Indeed, the application of this approach to the flame simulations described below leads to poor results for the F3 and F flames as compared to measurements, and it leads to a global extinction of the F flame in simulations. Thus, the idea discussed in the preceding paragraph will be modified by considering a generalized local equilibrium between the production and dissipation of scalar variances. The term S φ due to chemical reactions accounts for a correlation between scalar mixing and chemical reactions which may well be relevant in addition to velocity-scalar correlations. Therefore, the term S φ will be added to the relevant terms P and C B on the right-hand side of (6). To evaluate the relevance of the second term on the left-hand side of (6), it is helpful to split this term into two contributions, ( ) ρ ν + ν t φ = ν + ν t φ ρ x k Sc t x k Sc t x k φ x k + ρ Sc t ρ (ν + ν t ) x k. (9) The first contribution on the right-hand side of (9) can be combined with the production P on the right-hand side of (6). However, the relevance of the last term in

8 Michael Stöllinger and Stefan Heinz (9) seems to be low because there is no reason to expect a systematic effect on the turbulent mixing of scalars. Indeed, the simulation results reported below show that the modification of the first term on the right-hand side of (9) due to the last term is less than 5%. Thus, the last term of (9) will be added to the terms of minor relevance on the left-hand side of (6). Correspondingly, the scalar variance equation (6) can be written E = A C B. () Here, B is again given by B = φ /τ, and E and A are given by the expressions E = Dφ Dt A = ν + ν t Sc t ρ (ν + ν t ) φ, () ρ Sc t x k x k ( ) φ φ + φ x k x k x + S φ. () k It is worth noting that equation () represents an exact rewriting of the scalar variance equation (6). The notion of representing the scalar variance equation in terms of relation () is to combine the relevant processes on the right-hand side. The term E then represents deviations from the local equilibrium between production processes and dissipation. As explained above, the calculation of E, A, and B by simulation data results in fluctuating variables E, A, and B due to the effects of the discretization of equations and the finite number of particles involved in the Monte Carlo simulation. Equation () represents the basis for the calculation of C. In particular, C will be chosen such that the averaged quadratic error E T = A T C A B T + C B (3) T will become minimal. Q T refers again to the temporal average of any variable Q. By analyzing E T as a function of C one finds that C = A B T B (4) T minimizes the averaged quadratic error E. The difference between the standard T model assuming a constant value for C and the C model (4) is given by the fact that the model (4) provides local C values which minimize the deviations from the local equilibrium between production processes and dissipation of scalar variances. In addition to velocity-scalar correlations, the model (4) also accounts for velocityvariance correlations (given by the second term in the parentheses expression of ()) and scalar-reaction correlations (given by the S φ term in ()). Thus, the model (4) will be referred to below as generalized correlation (GC) model. The performance of this modeling concept in simulations will be discussed in the following sections in comparison to other C models.

PDF Modeling of Premixed Turbulent Combustion 9 4. The flames considered The turbulent premixed F3, F, and F flames studied experimentally by Chen et al. [8] are considered to investigate the performance of the PDF modeling approach described in sections and 3. The three highly stretched stoichiometric methane-air flames cover a range of Reynolds and Damköhler numbers. Based on an order of magnitude analysis Chen et al. [8] found that all three flames are located in the distributed reaction zones regime [44]. In particular, the F flame is located at the borderline to the well stirred reactor regime, and the F3 flame is located at the borderline to the flamelet regime. This wide range of combustion conditions is often found in spark ignition engines [] which present one of the most important applications involving premixed turbulent combustion. Due to the simple configuration, the broad range of combustion conditions, and the high quality experimental database [8], the three flames considered are well appropriate to investigate the performance of PDF methods for premixed turbulent combustion. The three flames are generated with the same burner. Table presents the mean nozzle exit velocities, the corresponding Reynolds numbers, and the centerline turbulent kinetic energy values. The burner design is shown schematically in figure. The burner consists of a nozzle with diameter D = mm for the main stream which is surrounded by a large pilot stream to stabilize the turbulent main jet flame. The laminar pilot stream is generated by an array (65 holes of diameter mm) of small jets issued through a cooled perforated plate. Both streams consist of a stoichiometric methane-air mixture (equivalence ratio Φ = ). The burner is surrounded by air at rest. The outer air is entrained into the three flames at axial positions 3D-5D, changing the flame brush to non-perfectly premixed. The experimental database includes radial profiles of the mean velocity, the turbulent kinetic energy, mean and variance of the temperature, and the mean mass fractions of the major species CH 4, O, CO, H O and minor species CO, H, OH. The error in the measurements of the mean velocity is estimated to be less than %, and the error of the mean temperature is expected to be less than %. The error in the measurements of the major species is between 8% to 5%, and the error regarding the minor species is within % to 5%. The flames considered have been studied numerically by using different combustion models. Prasad and Gore [5] solved RANS equations. The averaged reaction rate was closed with a flame surface density model which is valid under flamelet conditions [47]. Therefore, only the F3 flame has been investigated. Herrmann [] investigated all three flames using a modified level set approach [45] that accounts for the effects of the instantaneous flame structure and the entrainment of cold ambient air. The model is strictly valid only under flamelet conditions. Pitsch and De Lageneste [46] formulated the level set concept for LES of premixed flames and performed a F3 flame simulation. They reported issues regarding the strong influence of boundary conditions in their LES. Concerns include the significant heat losses due to the cooled burner surface, which are estimated to reach up to % in the pilot flame.

Michael Stöllinger and Stefan Heinz The PDF method was applied first by Mura et al. [39] to simulate the F3 and F flames. They showed that the IEM mixing model [, 6] combined with the standard model ω = /τ for the scalar mixing frequency is inapplicable to the two premixed flames close to the flamelet regime. The use of the standard model corresponds to the idea that the scalar mixing frequency is controlled by the inertial range turbulent motions. This assumption is valid for distributed combustion, but in the flamelet regime the mixing process is controlled by the laminar flame length scale. Mura et al. [39] suggested a PDF model where the outer parts of the flame structure (reactant side and product side) are described by a standard mixing model and the inner part (reaction zone) by a flamelet model. The scalar mixing frequency is provided by means of a transport equation for the scalar dissipation rate. The latter equation was obtained under the assumption of infinitely high Reynolds number and flamelet combustion conditions. More recently, Lindstedt and Vaos [9] applied the transported PDF model approach to F, F, and F3 flame simulations. They demonstrated the applicability of the transported PDF model approach to premixed turbulent combustion in the flamelet regime under the condition that the modeling of the scalar mixing frequency is modified. They closed the molecular mixing term by using the modified coalesencedispersion (CD) mixing model [3]. Comparisons of results obtained with different constants C =, 4, 6, 8 and a scalar mixing frequency model depending on the ratio of the laminar flame velocity to the Kolmogorov velocity were performed. The latter model will be referred to as Lindstedt-Vaos (LV) model. By assuming % heat loss for the pilot flame, a stably burning F flame was obtained by adopting a constant C > 4. However, by assuming % heat loss for the pilot flame, flame extinction occured over a wide range of constant C values. The latter problem did not appear by using the LV model, this means the LV model enabled stable F flame simulations for the case of % heat loss. Although the results obtained by Lindstedt and Vaos are encouraging, one has to see that the LV model involves flamelet ideas in conjunction with assumptions that are not satisfied in general (for example, the assumption of a local equilibrium between production and dissipation in the scalar dissipation rate equation). Thus, the generality of the closures applied remains to be established [9]. 5. Flame simulations A hybrid PDF-RANS approach was used to simulate the turbulent premixed flames described in section 4. RANS equations were used for the calculation of the velocity field. The conservation of mass and momentum equations applied are given by [5] D ρ u i Dt = p x i + x j ρ t + ρ u i x i =, (5) [ ( ui ρ ν + u j x j x i 3 )] u k δ ij x k ρ u i u j x j. (6)

PDF Modeling of Premixed Turbulent Combustion The Reynolds stress required to close equation (6) is given by a linear stress model, ). (7) u i u j = ( 3 kδ ui ij ν t + u j x j x i 3 u k δ ij x k The turbulent viscosity ν t is given by the usual parametrization ν t = C µ k ε. (8) According to the realizable k ε turbulence model [57], the turbulent kinetic energy k and dissipation rate ε of turbulent kinetic energy are calculated by the equations D ρ ε Dt D ρ k Dt = [ ( ρ ν + ν ) ] t k + ρ C µ S k ρ ε, (9) x j σ k x j ε = [ ( ρ ν + ν ) ] t ε ε + ρ C Sε ρ C x j σ ε x j k + νε. (3) S = S ij S ij is the magnitude of the mean rate-of-strain tensor. The model parameters are given in Shih et al. [57] and table. It is worth noting that C µ and C are no constants here. The scalar PDF transport equation () was closed with the mixing model (5) combined with the model (4) for C. The PDF transport equation () was solved by Monte Carlo simulation. The corresponding equations for particle positions x i and particle scalars φ (species mass fractions and sensible enthalpy) are given by [3, 9] ( dx i = ū i + ) ρ (ν t + ν) (ν t + ν) dt + dw i, (3) ρ Sc t x i Sc t dφ = ω (φ φ )dt + S (φ)dt. (3) Here, dw i denotes the increment of the i th component of a vectorial Wiener process. The increments dw i are Gaussian random variables which are determined by their first two moments, dw i =, dw i dw j = δ ij dt. (33) The scalar mixing frequency is defined by ω = C /τ. C is calculated according to the GC model (4) for the fuel (CH 4 ) mass fraction because combustion takes place only if fuel is available. A numerical limit φ 6 was applied to avoid the calculation of unphysically high values for C in regions where the scalar variance φ becomes very small. Temporally averaged variables involved in the calculation of C (4) were obtained by a moving average over 5 iterations. In this way, the solution of the RANS equations (5)-(3) provides the mean velocity u i, turbulent viscosity ν t and turbulence time scale τ = k/ε. These variables are then used to solve the particle

Michael Stöllinger and Stefan Heinz equations (3)-(3). The mean density ρ and mean viscosity ν are calculated from the particle compositions and temperatures and fed back into the RANS equations. The steady-state RANS equations were discretized by the finite volume method. Details can be found in table. The particle equations (3)-(3) are solved numerically by a mid-point rule [4] in order to achieve second order accuracy in time. The time step is determined from a local time stepping procedure [4]. Mean values of scalar quantities are calculated by a weighted summation over particles in a cell. For example, the term S φ was calculated by a summation N p i= S (φ i )(φ i φ ) over all N p particles in a cell. The number of particles per cell is set to 5. A higher number of particles per cell was found to have no effect on simulation results. The statistical error is further reduced by averaging over the last iterations [4]. All computations presented have been performed by using the FLUENT code []. The chemical reaction rates S (φ) were provided by a skeletal chemical mechanism DRM [6] consisting of 3 species (H, H, O, O, OH, H O, HO, H O, CH, CH (S), CH 3, CH 4, CO, CO, HCO, CH O, CH 3 O, C H, C H 3, C H 4, C H 5, C H 6, N ) and 4 elemental reaction. The suitability of the DRM mechanism will be demonstrated in section 6 by comparisons with results obtained with the full GRI-. mechanism [58]. The composition change due to chemical reactions was treated by the in situ adaptive tabulation (ISAT) method developed by Pope [5]. The equations were solved on a -dimensional axisymmetric domain. The domain extends up to D downstream (axial direction) from the nozzle exit plane and 6.5D in radial direction to allow entrainment of the ambient air. Here, D refers to the nozzle diameter. The domain is discretized into 7 (axial by radial) cells. The grid is non-uniform to improve the accuracy of computations in the flame region. The grid independence of the solution has been checked by comparison with results obtained on a 6 grid. The boundary conditions applied are described in table 3. The profiles for the axial velocity and turbulent kinetic energy at the jet inlet have been taken from the experimental database of Chen et al. [8]. The profile for the turbulent dissipation rate has been calculated from the profile of the turbulent kinetic energy and measurements of the lateral length scale l lat by adopting the relation ε = (k/3) 3/ /l lat [8]. The pilot composition was calculated from the chemical equilibrium of a stoichiometric methane-air mixture with % heat loss. To reduce the computational time, the simulations were performed in two steps. First, a laminar flame model [] was used to generate realistic initial conditions for PDF simulations. The PDF simulations were then initialized with the results from the laminar flame model. Such a use of realistic initial conditions increases significantly the convergence rate of PDF simulations. The computations have been performed parallel on four.8ghz Opteron processors each equipped with 4GB of SDRAM. The computational time required for a converged solution (involving approximately 3 iteration steps) was about 5 hours.

PDF Modeling of Premixed Turbulent Combustion 3 6. Simulation results Radial profiles of the normalized mean axial velocity U/U are presented in figure at different axial positions h = x/d. The mean axial velocity U = u is normalized by the bulk velocity U = 3, 5, 65 m/s for the F3, F, and F flames, respectively. The overall agreement between simulation results and measurements is excellent. The thermal expansion within the turbulent jet can be recognized by the increase of the axial velocity at radial positions r/d > along the x-axis for all three flames. As a result of this expansion, the shear layer (which is roughly located at the position of the maximum gradient of the mean axial velocity) is pushed outward in radial direction. This trend can also be seen in figure 3 where radial profiles of the normalized turbulent kinetic energy k/k (k = 3.8,.8,.7 m /s for the F3, F, and F flames, respectively) are shown. The peaks of the turbulent kinetic energy k are shifted outward for increasing axial positions. The results for the higher Reynolds number F and F flames agree very well with the measurements whereas an overprediction of the turbulent kinetic energy can be seen regarding the F3 flame, especially close to the burner head. Similar overpredictions have been reported by Lindstedt and Vaos [9]. The F3 flame was also studied by Pitsch and de Lageneste [46] by using LES in combination with a level set approach. Their turbulent kinetic energy results show a better agreement at h =.5 but a similar disagreement at h = 6.5. The F3 flame has the lowest axial velocity and the highest temperature. Thus, low Reynolds number effects which are not accounted for in the k ε model applied may be the reason for the turbulent kinetic energy overprediction. Radial profiles of the turbulence Reynolds number Re t = kτ/ν are shown in figure 4 at two axial positions for the three flames considered. Here, τ = k/ɛ is the characteristic time scale of large-scale turbulent motions, and ν is the mean molecular viscosity. The results are shown in radial direction until the flow becomes laminar. The profiles of Re t are qualitatively very similar for all three flames since they are generated with the same burner. Quantitatively, the turbulence Reynolds number of the F flame is about three times larger than that of the F3 flame. For a cold round jet one would expect a peak of Re t in the shear layer. Regarding the flames considered, the high temperature in the shear layer drastically increases the molecular viscosity. The relatively high viscosity then implies the monotonic decrease of Re t at h =.5 and the local minima observed in the Re t profiles at h = 8.5. The appearance of small Re t values in the flame brush due to a high viscosity (relaminization) is a well known phenomenon. These small Re t values may explain the differences between simulation results and turbulent kinetic energy measurements regarding the F3 flame: most turbulence models have been developed under the assumption of high Re t values. Figure 5 shows radial profiles of the mean reaction progress variable C = (T T u )/(T b T u ) [8] at different axial positions for the three flames considered. Here, T is the mean temperature, T b = 48 K is the adiabatic flame temperature and T u = 98 K is the temperature of the surrounding air. The simulation results of

4 Michael Stöllinger and Stefan Heinz the F3 flame agree very well with the measurements. This agreement indicates that the new GC frequency model is well applicable to flamelet conditions. The F and F flame simulation results show an overprediction of the progress variable at h =.5. Lindstedt and Vaos [9] found a similar overprediction in their F flame simulations using the same pilot inlet conditions. A reason for the observed overprediction of the temperature close to the burner exit regarding the F and F flames could be given by the complex interaction between the turbulent jet and laminar pilot stream. Such flow conditions are rather difficult to predict within the RANS framework. LES results for the F and F flames could clarify whether this is indeed the reason for the observed overprediction. However, such LES results have not been reported so far. Possibly, a better agreement between simulation results and measurements may be obtained by adopting refined inlet conditions (e.g., a reduction of the pilot inlet temperature regarding the F and F flames). Figure 6 shows simulation results of the mean oxygen mass fraction Y O, and figure 7 shows simulation results of the mean fuel (methane) mass fraction Y CH4 for the three flames considered. Both oxygen and fuel concentrations are well predicted in all three flames. The entrainment of surrounding air is clearly visible in figure 6. It is most intense in the highest Reynolds number F flame due to the high turbulence intensity. Mean mass fractions of the product species Y H O and Y CO are shown in figures 8 and 9, respectively. The H O concentration results agree very well with the measurements for all three flames. The CO concentration of the F3 flame is also well predicted whereas the F and F flame results display a slight CO underprediction. Figure presents radial profiles of the mean CO mass fraction Y CO. This figure shows for all three flames a significant overprediction of CO which increases downstream. This finding could explain the underprediction of the CO concentration in the F and F flames: the slightly too high temperature levels in the F and F flame simulations imply high CO levels and a corresponding slower oxidation of CO to CO. The CO overestimation in the F3 flame cannot be explained in such a way since the temperature and CO levels are well predicted. Similar high CO levels have been reported by Lindstedt and Vaos [9]. The estimated error in the measurements of Chen et al. [8] regarding the minor species CO and OH is between % and 5%. Thus, the errors of measurements cannot explain the overpredictions of up to % which are observed in the simulations. Simulations with the full GRI. mechanism [58] do not improve the CO predictions (see the discussion of the influence of differently complex chemical mechanisms in the following paragraph). Given the relatively well predictions of all the other species it is unclear which reason may cause the observed discrepancy between CO measurements and simulation results. Figure shows results of the mean OH mass fraction Y OH. The OH levels are well predicted for all the three flames for axial positions h 6.5. This finding is somewhat unexpected regarding the F flame: its temperature is overpredicted for h 6.5, and the OH-radical concentration is known to be very sensitive to temperature levels. Moreover, all the measurements of

PDF Modeling of Premixed Turbulent Combustion 5 OH seem to indicate a broader flame region, which implies significant OH amounts in flame regions with very low mean temperatures. Similar results have been found by Lindstedt and Vaos [9]. They concluded that some caution may be required regarding the interpretation of the experimental data. The influence of differently complex chemical mechanisms [5, 3] and finite-rate chemistry effects [8] on the predictions of transported PDF methods have been studied in previous investigations. Two different chemical mechanisms have been applied here to study the influence of the chemistry scheme: simulation results obtained with the DRM [6] skeletal mechanism are compared in figure with results obtained with the full GRI. mechanism [58]. CO and OH mass fractions of F3 flame simulations are shown to address the overestimation of CO predictions and sensitivity of the OH-radical concentration: see the discussion in the preceding paragraph. In both F3 flame simulations, the mixing model (5) was used in combination with the GC model (4) for C. One observes that the results obtained with the different mechanisms are almost identical. This fact indicates that the simplifications used to obtain the skeletal mechanism DRM do not affect the simulation results. 7. Comparison with other scalar mixing frequency models After demonstrating the good performance of the GC model in flame simulations, let us compare the GC model with other ω models for premixed turbulent combustion. First, previously developed ω models [6, 9, 38] will be described to explain the underlying assumptions of these models. Second, flame calculations involving such other ω models will be compared to predictions based on the GC model. An alternative approach to the direct derivation of models for the scalar mixing frequency ω is to consider a transport equation for the scalar dissipation rate ε. The relation between ε and ω provided by the expressions (9) and (5) is given by ω φ = ε + ν Sc φ x k φ x k. (34) Usually, the relatively small last contribution in (34) is neglected in these approaches (which corresponds to the assumption of high Reynolds numbers) such that ω is given by ω = ε /φ. Transport equations for the scalar dissipation rate have been proposed, for example, by Zeman and Lumley [63], Jones and Musonge [5], Mantel and Borghi [3] and Mura and Borghi [38]. The structure of scalar dissipation rate models [3, 54] can be illustrated by the model of Mura and Borghi [38], which is applicable to premixed turbulent combustion. The latter model is given by ρ ε MB t = + ρ (u k U Lk )ε MB x k x k ( c ε k P ε P ε MB + c U k ε + ε k ε ( ρ MB D t x k εmb β φ ) ) ρ ε MB. = (35)

6 Michael Stöllinger and Stefan Heinz The model (35) was derived on the basis of flamelet assumptions for infinitely high Damköhler numbers. The source terms on the right-hand side are related to the production and dissipation terms that appear in transport equations for the turbulent kinetic energy and scalar variances. In particular, the production P by large-scale velocity fields and the production P by large-scale scalar fields are given by P = ν t u k x k u k x k, P = ν t Sc t φ x k φ x k. (36) The diffusion coefficient is given by D t = ν t /Sc ε, where Sc ε refers to the turbulent Schmidt number in the ε equation. The propagation term U Lk of flamelets and the parameter β are given by U Lk = ( ) ( S L + φ φ φ, β = β ) k/3/s L x k x j x j 3 c S L ε. (37) k Here, S L is the laminar burning velocity of the mixture. The model constants are given by =.9, β = 4., c U = c =., c ε =. and Sc ε =.3 [6]. The model (35) will be referred to below as Mura and Borghi (MB) model. It is worth noting, however, that there are some questions related to the MB model. The reaction contributions in the MB model are given by the flamelet propagation velocity U Lk and the term /3c ε S L / k in the expression for the β coefficient. To see the importance of the flamelet propagation as compared to the fluid velocity we consider the ratio Z = U Lk U Lk / u k u k between the magnitude of the flamelet propagation velocity and the magnitude of the fluid velocity. Figure 3 shows the values of the ratio Z and the term /3c ε S L / k along the mean flame front position (see the definition below) for the F3 flame. The ratio Z remains less than. for most parts of the flame. Hence, the influence of the flamelet propagation velocity is negligible. The factor /3c ε S L / k is less than., which implies that β β. The two reactive contributions Z and /3c ε S L / k are even smaller for the F and F flames (not shown) since the fluid velocity and the turbulence levels are much higher in these flames. Thus, the MB model (35) reduces to a transport equation for a passive scalar regarding the flames considered. The latter cannot be seen to be justified since the fuel mass fraction (which was used for the calculation of the scalar dissipation rate) is a highly reactive scalar. It is interesting to compare the MB model for non-reacting scalars (this means for U Lk = and β = β) with the model of Jones and Musonge [5] and a variety of other scalar dissipation rate equations studied by Sanders and Goekalb [54]. This comparison reveals a significant difference between the MB model and other models for the scalar dissipation rate: the ε/k term acts as a production term in equation (35) whereas the corresponding term in the model of Jones and Musonge [5] represents a sink term. Thus, the MB model disagrees with many other models for the scalar dissipation rate of non-reacting scalars.

PDF Modeling of Premixed Turbulent Combustion 7 The underlying assumptions related to the algebraic scalar frequency model of Lindstedt and Vaos [9] can be seen by taking reference to the algebraic version of the MB model. By assuming that the evolution of the scalar dissipation rate is dominated by the source terms on the right-hand side, equation (35) implies an algebraic expression for the scalar mixing frequency ω = ε /φ, ( ) ω = ( β 3 c ε S L / P ) P c + c U k ε ε + ε k. (38) The use of the first-order Taylor series approximation for the chemical reaction contribution then results in ω = C ( + ) 3 c S L ε ε k k, (39) where C = (c P /ε + c U P/ε + )/β denotes a mixing parameter. Kuan et al. [7] and Lindstedt and Vaos [9] derived an expression for ω which is similar to expression (39). Their scalar mixing frequency model is given by ω LV = C ( + C ρ u ρ ) S L ε v k k. (4) Here, C = 4 and C =. are model constants [9]. ρ u denotes the density of the unburned mixture, ρ is the mean mass density, and v k is the Kolmogorov velocity. The essential assumption related to the Lindstedt and Vaos (LV) model (4) is given by the consideration of C as a constant, which corresponds to the idea of velocity and scalar fields that are characterized by a local equilibrium between production and dissipation (P /ε and P/ε are constant). This condition is not satisfied regarding the flames considered here. Figure 4 shows radial profiles of the scalar production-todissipation ratio P /ε and the turbulence production-to-dissipation ratio P/ε in F3 flame simulations. As can be seen, the profiles of the production-to-dissipation ratios vary significantly at both axial positions. Hence, the scalar production-to-dissipation ratio P /ε and the turbulence production-to-dissipation ratio P/ε are far from being constants. Similar results have been found regarding the F and F flames (not shown). It is also worth noting that the inclusion of the effect of chemical reactions differs from expression (39), in particular, by the use of the Kolmogorov velocity v k instead of k. An alternative suggestion for the algebraic modeling of the scalar mixing frequency was made by Cha and Trouillet [6]. Their model relies on the assumption that the largescale turbulent motions control the scalar mixing frequency (the mixing frequency of reacting scalars is related to the mixing frequency of non-reacting scalars). However, Cha and Trouillet s model was developed for non-premixed turbulent flames. Thus, it will be not considered for comparisons here.

8 Michael Stöllinger and Stefan Heinz The simplest ω model is recovered by neglecting chemistry effects in the LV model, ω ST = C ε k. (4) The model (4) will be referred to below as standard (ST-*) scalar mixing frequency model, where * refers to the constant value of C applied. The basic assumption related to the model (4) is that the scalar mixing frequency ω is controlled by the mixing frequency ω = ε/k = /τ of large-scale turbulent motions. Let us have a closer look at the validity of this assumption. Figure 5 shows radial profiles of the normalized scalar mixing frequency ω τ c, the normalized turbulence frequency ωτ c, and the normalized Kolmogorov frequency ω η τ c at axial positions h = 4.5 and h = 8.5 for the F3 and F flames. Here, the Kolmogorov frequency ω η is defined by ω η = (ε/ν) /. The frequency values have been normalized by the chemical time scale τ c =.44 ms which is defined as ratio of the laminar flame thickness and laminar burning velocity [8]. High values of the normalized scalar mixing frequency are found within the turbulent flame brush. The reason for that is given by the three terms involved in A (see relation ()). The correlation S φ between scalar mixing and chemical reactions increases the scalar mixing frequency directly. The rapid consumption of the fuel in the premixed flame steepens the gradients of the scalar mean and variance. The steep gradients then enhance the scalar mixing by increasing the two contributions in addition to S φ in A. Figure 5 shows that the Kolmogorov frequency is larger than the scalar mixing frequency in the F flame at h = 4.5. Therefore, Kolmogorov scale eddies are able to enter and thicken the flame. This broadening was also reported in the experimental study of Chen et al. [8]. In the F3 flame the Kolmogorov frequency is found to be smaller than the scalar mixing frequency, which indicates a flamelettype combustion. The turbulence frequency is much smaller than the scalar mixing frequency in the turbulent flame brush of both flames. The latter finding clearly shows that large-scale turbulent motions do not control the scalar mixing in these regions, this means the basic assumption of the model (4) is not satisfied for the flows considered. This conclusion is in consistency with O Young and Bilger s [43] experimental investigations of premixed flames. The authors found that the ratio between the turbulence frequency and scalar mixing frequency varies over almost an order of magnitude. To see the differences to predictions based on the GC model, simulations of the F3, F, and F flames have been performed by adopting the MB model, the LV model, and the ST-6 model (in accord with the observations of Lindstedt and Vaos [9] it was found that the combination of the ST-* model with C = 6 represents an optimal choice). It turned out that the MB model combined with β = 4. did not allow calculations of stably burning F3, F, and F flames. Thus, the MB model was used in conjunction with β =.3, which provided the best agreement between simulation results and measurements. Results obtained with the different frequency models are compared with experimental data in figure 6. This figure shows the mean flame front position which is defined as iso-line of the mean temperature. The iso-line temperatures are