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.

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

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

An investigation of the performance of turbulent mixing models

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

PDF calculations of turbulent lifted flames of H 2 /N 2 fuel issuing into a vitiated co-flow


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

RESOLVING TURBULENCE- CHEMISTRY INTERACTIONS IN MIXING-CONTROLLED COMBUSTION WITH LES AND DETAILED CHEMISTRY

Capturing localised extinction in Sandia Flame F with LES-CMC

Transported PDF modelling with detailed chemistry of pre- and auto-ignition in CH 4 /air mixtures

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

HYBRID RANS/PDF CALCULATIONS OF A SWIRLING BLUFF BODY FLAME ( SM1 ): INFLUENCE OF THE MIXING MODEL

Lifted methane air jet flames in a vitiated coflow

In Situ Detailed Chemistry Calculations in Combustor Flow Analyses

MODELLING MIXING AND REACTION IN TURBULENT COMBUSTION

LES of the Sandia Flame D Using an FPV Combustion Model

A Method for the Prognosis of Turbulent Combustion

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

Sensitivity analysis of LES-CMC predictions of piloted jet flames

REDIM reduced modeling of quenching at a cold inert wall with detailed transport and different mechanisms

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

1D Raman/Rayleigh/CO-LIF line measurements of major and temperature in turbulent DME/air jet flame

K. A. Kemenov, H. Wang, S. B. Pope Sibley School of Mechanical and Aerospace Engineering, Cornell University, Ithaca, NY 14853

PDF Simulations of a Bluff-Body Stabilized Flow

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

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

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

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

A Unified 3D CFD Model for Jet and Pool Fires

HYBRID RANS/PDF CALCULATIONS OF A SWIRLING BLUFF BODY FLAME ( SM1 ): INFLUENCE OF THE MIXING MODEL

HIGH-FIDELITY MODELS FOR COAL COMBUSTION: TOWARD HIGH-TEMPERATURE OXY-COAL FOR DIRECT POWER EXTRACTION

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

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

Modeling laminar partially premixed flames with complex diffusion in OpenFOAM R

JOINT SCALAR VS. JOINT VELOCITY-SCALAR PDF MODELLING OF BLUFF-BODY STABILISED FLAMES WITH REDIM

A NUMERICAL ANALYSIS OF COMBUSTION PROCESS IN AN AXISYMMETRIC COMBUSTION CHAMBER

Fuel 105 (2013) Contents lists available at SciVerse ScienceDirect. Fuel. journal homepage:

Soot formation in turbulent non premixed flames

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

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

A Jet-Stirred Apparatus for Turbulent Combustion Experiments

Analysis of lift-off height and structure of n-heptane tribrachial flames in laminar jet configuration

Hybrid finite-volume/transported PDF method for the simulation of turbulent reactive flows

Assessment of turbulence-chemistry interaction models in the computation of turbulent nonpremixed

Overview of Turbulent Reacting Flows

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

Large eddy simulation/probability density function modeling of a turbulent CH 4 =H 2 =N 2 jet flame

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

The development of advanced combustion systems is mainly controlled by the objective to increase fuel

Approximation of chemical reaction rates in turbulent combustion simulation

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

NUMERICAL ANALYSIS OF TURBULENT FLAME IN AN ENCLOSED CHAMBER

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

LARGE-EDDY SIMULATION OF PARTIALLY PREMIXED TURBULENT COMBUSTION

PDF modeling and simulation of premixed turbulent combustion

A PDF/PHOTON MONTE CARLO METHOD FOR RADIATIVE HEAT TRANSFER IN TURBULENT FLAMES

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

Algebraic Turbulence-Chemistry Interaction Model

Numerical Simulations of Hydrogen Auto-ignition in a Turbulent Co-flow of Heated Air with a Conditional Moment Closure

Numerical Simulation of Entropy Generation in Hydrogen Enriched Swirl Stabilized Combustion

LES-CMC of a dilute acetone spray flame with pre-vapor using two conditional moments

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

Comparison of Mixing Model Performance for Nonpremixed Turbulent Reactive Flow

Numerical Modeling of Laminar, Reactive Flows with Detailed Kinetic Mechanisms

CHARACTERISTICS OF ELLIPTIC CO-AXIAL JETS

Reacting Flow Modeling in STAR-CCM+ Rajesh Rawat

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

Asymptotic Structure of Rich Methane-Air Flames

PDF Modeling and Simulation of Premixed Turbulent Combustion

Process Chemistry Toolbox - Mixing

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

Unsteady Flamelet Modeling of Soot Formation in Turbulent Diffusion Flames

Asymptotic Analysis of the Structure of Moderately Rich Methane-Air Flames

Thermal NO Predictions in Glass Furnaces: A Subgrid Scale Validation Study

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

Simulation of Turbulent Lifted Flames and their Transient Propagation

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

Yiguang Ju, Hongsheng Guo, Kaoru Maruta and Takashi Niioka. Institute of Fluid Science, Tohoku University, ABSTRACT

Coupling of ChemApp and OpenFOAM

CFD and Kinetic Analysis of Bluff Body Stabilized Flame

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

Universities of Leeds, Sheffield and York

Investigation of ignition dynamics in a H2/air mixing layer with an embedded vortex

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

Erratum to: High speed mixture fraction and temperature imaging of pulsed, turbulent fuel jets auto igniting in high temperature, vitiated co flows

REACTION-RATE, MIXTURE-FRACTION, AND TEMPERATURE IMAGING IN TURBULENT METHANE/AIR JET FLAMES

LES Approaches to Combustion

Detailed Chemical Kinetics in Multidimensional CFD Using Storage/Retrieval Algorithms

Numerical simulation of oxy-fuel jet flames using unstructured LES-CMC

High-Repetition Rate Measurements of Temperature and Thermal Dissipation in a Nonpremixed Turbulent Jet Flame

The Effect of Mixture Fraction on Edge Flame Propagation Speed

A novel transient turbulent jet flame for studying turbulent combustion

Numerical evaluation of NO x mechanisms in methane-air counterflow premixed flames

LES of the Cambridge Stratified Swirl Burner using a Sub-grid pdf Approach

Scalar dissipation rate at extinction and the effects of oxygen-enriched combustion

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

Large Eddy Simulation of Piloted Turbulent Premixed Flame

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

Advanced Turbulence Models for Emission Modeling in Gas Combustion

Modeling Turbulent Combustion

Transcription:

The Influence of Chemical Mechanisms on PDF Calculations of Nonpremixed Piloted Jet Flames Renfeng Cao and Stephen B. Pope Sibley School of Mechanical and Aerospace Engineering, Cornell University, Ithaca, NY 14853, USA 1. Introduction There are several previous successful PDF modeling studies [1-3] of the Barlow & Frank nonpremixed piloted jet flames (flames D, E and F) [4]. While there is at least partial understanding of the influence of turbulent mixing models on these calculations [5,6], there has been no systematic study of the influence of the chemical mechanism that is employed. In this work, focusing on flame F (which has significant local extinction and exhibits the strongest sensitivity), we present PDF calculations using six different mechanisms. These range from a 5-step reduced mechanism [7], to the GRI3.0 detailed mechanism [8] which involves 53 species. The principal results considered (which are compared to the experimental data [9]) are means of temperature and species mass fraction conditional on mixture fraction. More details of the current work are provided in [10]. 2. The joint velocity-turbulence frequencycomposition PDF method There are three different kinds of PDF methods [11]. The simplest one is the composition PDF method, in which only the modelled joint PDF equation for composition is solved: the mean flow field equations are solved by a standard CFD code, and the PDF provides the chemistry closure. The second kind is based on the joint velocity-composition PDF method, in which a separate model is required for the time or length scale of the turbulence. The most complete PDF method is the joint velocity-turbulence frequencycomposition PDF method: this is used in the current work and is subsequently referred as the joint PDF method. The particle implementation of the joint PDF method requires models for mixing, velocity and turbulent frequency following a fluid particle [11]. The SLM (Simplified Langevin Model) is used for the evolution of the particle velocity. The stochastic frequency model of Van Slooten et al. [12] is used for the turbulent frequency of particles, which provides the time scale of turbulence. The model constants used are the same as those used in many previous studies, e.g. [6,13,14]. In PDF methods, the effect of molecular diffusion on the composition is represented by a mixing model. The mixing model constant C ϕ is traditionally set to 2.0; but different values have also been used in previous PDF calculations. The EMST [15] mixing model is used in the present work to investigate the effect of different chemical mechanisms. 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. The chemical mechanisms considered in this paper are listed in Table 1. The GRI detailed kinetic mechanism provides the most comprehensive and standardized set of mechanisms for methane combustion [8]. The detailed versions of 2.11 and 3.0 are investigated and denoted as GRI2.11 and GRI3.0, respectively. The 15-step reduced GRI2.11 mechanism [16] is denoted by ARM2 following Tang et al. [22]. This mechanism has been successfully used in the joint PDF calculations of flame F performed by Xu, Tang and Pope [1,2]. The 5-step reduced GRI2.11 mechanism[7] is denoted as S5G211 in current paper. The detail description of the skeletal mechanism is provided by James et al. [17]. The Smooke mechanism is described in [18], but three reactions have been updated [19] and are shown in Table 2. Table 1. Chemical mechanisms used in the present comparisons Mechanism # of # of species steps NO species Reference GRI 2.11 49 277 With NO [8] GRI 3.0 53 325 With NO [8] ARM2 19 15 With NO [16] S5G211 9 5 With NO [7] Skeletal 16 41 Without NO [17] Smooke 16 46 Without NO [18,19] Table 2. Three updates to the Smooke mechanism REACTIONS k = A T b exp(-e/rt) A b E CH 4 +H=CH 3 +H 2 2.2 10 4 3.0 8750. H+O 2 =OH+O 2.0 10 14 0.0 16800. 2.1 10 18-1.0 0. H+O 2 +M=HO 2 +M H 2 O/21./ CO 2 /5.0/ H 2 /3.3/ CO/2.0/ O 2 /0.0/ N 2 / 0.0/ 3. Numerical solutions There are several implementations of particle-mesh methods to solve the modelled joint PDF equations. All computations presented here use a code named HYB2D [14] which implements a hybrid FV/particle algorithm. In the hybrid algorithm, the PDF/particle method (particle part) is coupled with a finite volume solver (FV part). The FV part solves the mean conservation equations for mass, 4 th Joint Meeting of the U. S. Sections of the Combustion Institute Page 1 of 6

momentum, energy and the mean equation of state; and the particle part solves the fluctuating velocity-turbulent frequency-compositions PDF transport equations. The FV part provides mean fields of velocity, density and pressure to the particle part and obtains the turbulent fluxes and reaction source term from the particle part. The flow considered here is statistically steady 2D axisymmetric, and non-swirling. A polar-cylindrical (z,r) coordinate system is used with the origin at the center of the fuel jet at its exit plane. The computational domain is rectangular, of extent (0,80D) in the axial (z) direction, and (0,20D) in the radial (r) direction, where D is the diameter of the jet (D=3.6mm). The inlet velocity profile is interpolated from measurements [9,20]. The ratio of production to dissipation of turbulent kinetic energy is specified as unity which, together with the specified profiles, determines the inlet profile of mean turbulence frequency. The temperature, compositions, and density are specified as being uniform in each stream in accord with the experimentally-determined values [9]. The co-flow boundary (r=20d) is treated as a perfect-slip wall. Symmetry conditions are applied on the centerline (r=0). At the exit plane, in the FV part, the mean density and the axial and radial mean velocities are extrapolated from the interior, and the pressure is specified. Systematic tests have been performed on the following numerical parameters to determine appropriate values used in the current calculations: (i) the ISAT (In-Situ-Adaptive Tabulation) [21] error tolerance (2 10-5 ), (ii) the number of cells in the domain (96 96), (iii) the number of particles per cell (100), and (iv) the coefficients of time averaging (over 1000 particle time steps). These values ensure that numerical errors are no greater than 2% (with respect to the peak value) for the mean temperature and major species, and 5% for the minor species. 4. Results and discussion It has been shown in previous studies that both radiation and the pilot temperature T p can significantly affect the calculations of flame F [2,22]. In the current work, the effect of T p has been tested using the skeletal mechanism and the effect of radiation has been tested using the skeletal and GRI3.0 mechanisms. For more details refer to [6]. All of the calculations presented in this paper are obtained using the pilot temperature T p =1880K and without radiation. 4.1 Calculations of the velocity field and the mixture fraction Correct representation of the velocity field and mixture fraction is essential for the calculations of turbulent reactive flows. The radial profiles of the mean and rms axial velocities obtained using the S5G211, Smooke and GRI3.0 mechanisms with C ϕ =2.0, together with the calculations using GRI3.0 mechanism with C ϕ =1.5, are shown in Fig.1. Although there are some differences between these calculations (especially calculations using the Smooke mechanism), they are generally very close to each other, and agree with the experimental data reasonably well. Calculations using other mechanisms (i.e., ARM2, skeletal, and GRI211) yield similar velocity profiles (not shown) as Fig. 1. Figure 2 shows the radial profiles of the mean and rms mixture fraction obtained using the skeletal, Smooke and GRI3.0 mechanisms with C ϕ =2.0, together with the calculations using the GRI3.0 mechanism with C ϕ =1.5. It can be seen that while the mean radial profiles of mixture fraction are almost identical for all these calculations shown in Fig. 2, the rms of mixture fraction has some obvious differences between the different mechanisms. For the GRI3.0 mechanism, calculations using a smaller value of C ϕ (1.5) yield higher values of the rms of the mixture fraction, which is as expected. The same observations can be made for other calculations using other mechanisms (S5G211, ARM2, and GRI2.11, not shown). 4.2 Calculations of the conditional mean and rms The conditional mean and rms of all species in flame F at z/d=1, 2, 3, 7.5, 15, 30, 45, 60, and 75 are calculated in the current work. It is found that that the conditional mean and rms are very sensitive to different mechanisms and different values of C ϕ at z/d=7.5, 15, and 30, where significant local extinction and reignition are observed [1,4,9], but not so sensitive at other locations. Especially, the conditional means of the temperature and major species are almost identical at the downstream locations (z/d=45, 60, and 75). Significant differences can be observed in some minor species like CO and NO. To illustrate this point, Figure 3 shows the conditional mean temperature obtained using the S5G211 and GRI3.0 mechanisms, which are the most simple mechanism and most comprehensive mechanism used in the current work, respectively. It can be seen from Fig. 3 that the conditional mean temperatures of all three calculations are very close to each other and agree very well with the experimental data at z/d=1, 2, 3, 45, 60 and 75, although there are small discrepancies between calculations using the GRI3.0 and S5G211 mechanisms at z/d=1, 2 and 3. On the other hand, these calculation are very different at z/d=7.5, 15, and 30. Among these locations, z/d=15 has the most significant local extinction and hence is most sensitive to different mechanisms and different values of C ϕ. A similar tendency has been observed for other calculations (not shown). 4 th Joint Meeting of the U. S. Sections of the Combustion Institute Page 2 of 6

Fig. 1. Computed and measured radial profiles of mean and rms axial velocity. Symbols, measurements [20]; lines, PDF calculations using: the S5G211 mechanism with C ϕ =2.0 (blue solid), the Smooke mechanism with C ϕ =2.0 (green solid), the GRI3.0 mechanism with C ϕ =2.0 (red dashed), and the GRI3.0 mechanism with C ϕ =1.5 (black dotted). Fig. 2. Computed and measured radial profiles of mean and rms mixture fraction. Symbols, measurements [9]; lines, PDF calculations using: the skeletal mechanism with C ϕ =2.0 (blue solid), the Smooke mechanism with C ϕ =2.0 (green solid), the GRI3.0 mechanism with C ϕ =2.0 (red dashed), and the GRI3.0 mechanism with C ϕ =1.5 (black dotted). 4.3 Comparison of different mechanisms Since the calculations are very sensitive to the mechanisms and the value of C ϕ at z/d=15, it is natural to chose this location to investigate the performances of different mechanisms. The temperature, one characteristic major species (CH 4 ) and one characteristic minor species (CO) are studied at this location. Six mechanisms and four values of C ϕ have been used for these calculations, as listed in Table 3 and the results are shown in Figs. 4 and 5. Fig. 3. Measured and computed conditional mean temperature. Symbols, measurements [9]; lines, PDF calculations using: the S5G211 mechanism with C ϕ =2.0 (green dash dotted), the GRI3.0 mechanism with C ϕ =2.0 (blue dashed), and the GRI3.0 mechanism with C ϕ =1.5 (red solid). Several observations can be made based on Figs. 4 and 5. First, for all mechanisms, the conditional rms decreases with increasing C ϕ, which is as expected. In Fig. 4, for the skeletal, ARM2 and S5G211 mechanisms, the differences between the calculations obtained using smaller values of C ϕ (e.g., 1.5 and 2.0) are much larger than those obtained using larger values of C ϕ (e.g., 2.0 and 3.0). This indicates that these calculations become more and more sensitive to the change of the mixing rate as the calculations get closer to global extinction. Second, in the current test cases, each 4 th Joint Meeting of the U. S. Sections of the Combustion Institute Page 3 of 6

C ϕ =1.2 C ϕ =1.5 C ϕ =2.0 C ϕ =3.0 Fig. 4. Effect of different mechanisms (skeletal, ARM2 and S5G211) and different values of the mixing model constant C ϕ on calculations of conditional mean and rms of: (a) the temperature; (b) the mass fraction of CH 4 ; and (c) the mass fraction of CO. Symbols, measurements [9]. The values of C ϕ used for each mechanism are listed in Table 3. Table 3. List of mechanisms and the values of C ϕ used in current work (* denotes the case that yield the closest agreement with the measurements for tested values of C ϕ ) Mechanisms C ϕ Shown 1.2 1.5 2.0 3.0 in Skeletal Yes * Yes Yes 4 Fig. ARM2 Yes * Yes Yes Fig. 4 S5G211 Yes * Yes Yes Fig. 4 GRI 2.11 Yes * Yes Fig. 5 GRI 3.0 Yes * Yes Fig. 5 Smooke Yes * Yes Fig. 5 mechanism has a value of C ϕ (which is marked by * in * Table 3, and denoted as C ϕ in the following) that yields closest agreement with the measurements. In other words, all of these mechanisms are capable of representing the local extinction of flame F by using appropriate values of C ϕ. Based on test cases in the current work, these values are: C * ϕ =1.5 for the ARM2, GRI2.11, GRI3.0 and skeletal mechanisms; C * ϕ =1.2 for the S5G211 mechanism and C * ϕ =2.0 for the Smooke mechanism. It should be noted that, because discrete values of C ϕ are used in the current work, there might exist different values of C * ϕ that could yield even better agreement with the experimental data. For example, the calculations using GRI2.11 with C * ϕ =1.5 over-predict the measured conditional mean temperature a little bit, which indicates that calculations using a smaller value of C ϕ (e.g., 1.4) may be even closer to the measurements. Third, the closest agreement with a conditional mean corresponds to closest agreement with its conditional rms. For these closest-agreement calculations, they agree reasonably well with the measured conditional rms's around stoichiometric (ξ st =0.351) and on the lean side. They generally under-represent the rms s on the fuel rich side. 4 th Joint Meeting of the U. S. Sections of the Combustion Institute Page 4 of 6

C ϕ =1.5 C ϕ =2.0 C ϕ =3.0 Fig. 5. Effect of different mechanisms (GRI2.11, GRI3.0 and Smooke) and different values of the mixing model constant C ϕ on calculations of conditional mean and rms of: (a) the temperature; (b) the mass fraction of CH 4 ; and (c) the mass fraction of CO. Symbols, measurements [9]. The values of C ϕ used for each mechanism are listed in Table 3. 4.4 Performance of GRI mechanisms and calculations of the NO It has been shown in previous studies [23] that calculations of laminar opposed-flow partially premixed methane/air flames using ARM2, GRI2.11 and GRI3.0 yield similar results for major species, while there may be significant differences for minor species. The current calculations indicate a similar tendency. While all of these three mechanisms yield closest agreement with the experimental data with C ϕ * =1.5, the S5G211 calculations have different behavior and obtain closest agreement with C ϕ * =1.2. Figure 6 shows the mass fraction of NO conditional on mixture fraction at different axial locations for ARM2, GRI2.11, GRI3.0 and S5G211, all with their closestagreement cases. It may be seen that the calculations using GRI3.0 yield significantly higher levels of NO than those obtained using the ARM2 and GRI2.11 mechanisms. The ARM2 calculations and GRI2.11 calculations have some differences at z/d=15, but they are very close to each other at other locations. This is consistent with the laminar flame Fig. 6. Comparisons of different GRI mechanisms in their closest-agreement cases using the conditional mean mass fraction of NO in flame F. Symbols, measurements [9]; lines, PDF calculations using the GRI3.0 mechanism (red solid line), the GRI2.11 mechanism (blue dashed line), the ARM2 mechanism (green dash dotted line) and S5G211 mechanism (black dotted line). The EMST mixing model is used for all of these calculations with C ϕ =1.5 (except C ϕ =1.2 for the S5G211 mechanism). 4 th Joint Meeting of the U. S. Sections of the Combustion Institute Page 5 of 6

studies performed by Barlow et al. [23] and the composition PDF calculations of flame D performed by Raman et al. [24]. The calculations using the S5G211 mechanism yield significantly higher levels of NO at z/d=7.5 and 15, but yield the same level as that of ARM2 and GRI2.11 at other locations. It should be noticed that all the calculations presented in the current paper are obtained without considering the effect of radiation. Since radiation has a significant effect on the mass fraction of NO, the agreement with the experimental data will change when radiation is taken into consideration. Presumably, the observations about the relative tendency will still be valid. 5. Conclusions In the current work, joint PDF calculations with six different chemistry mechanisms (i.e., ARM2, GRI2.11, GRI3.0, S5G211, skeletal, and Smooke) have been performed on Flame F. This flame is well known to have significant local extinction and exhibits the strongest sensitivity. Hence it is very useful for investigating the performance of different chemistry mechanisms, especially the capability to capture location extinction and reignition. The EMST mixing model with a series values of mixing model constant C ϕ is used for these calculations. It is found that all six mechanisms are capable of capturing the local extinction by using appropriate values of C ϕ. While calculations using the ARM2, GRI2.11, GRI3.0 and skeletal mechanisms with C ϕ =1.5 yield closest agreement with the experimental data, the S5G211 and Smooke calculations behavior differently and yield closest agreement with the experimental data by using C ϕ =1.2 and 2.0, respectively. The calculations using the GRI3.0 mechanism generally yield significantly higher levels of NO in flame F than those of ARM2 and GRI2.11 mechanisms. The calculations using the S5G211 mechanism yield higher levels of NO at z/d=7.5 and z/d 15 than those of ARM2 and GRI2.11, but almost the same level at z/d=30 and after. Acknowledgements This work is supported by Air Force Office of Scientific Research under grant No. F-49620-00-1-0171. The computations were conducted using the resources of the Cornell Theory Center, which receives funding from Cornell University, New York State, federal agencies, foundations, and corporate partners. References: 1. Xu, J. and S.B., Pope, Combust. Flame, 123 (2000) 281-307 2. Tang, Q., J. Xu and S.B. Pope, Proc. Combust. Inst., 28 (2000) 133-139 3. Lindstedt, R. P., S.A. Louloudi and E.M. Vaos, Proc. Combust. Inst, 28 (2000) 149-156 4. Barlow, R. S. and J. H., Frank, Proc. Combust. Inst, 27 (1998) 1087-1095 5. Ren, Z. and S.B. Pope, Combust. Flame, 136 (2004) 208-216 6. Cao, R., S.B. Pope, and A.R. Masri, Turbulent Lifted Flames in a Vitiated Coflow Investigated Using Joint PDF Calculations, Combust. Flame, accepted, 2005 7. Mallampalli, H. P., T. H., Fletcher and J.Y. Chen, Paper 96F-098, Presented at the Fall Meeting of the Western States Section of the Combustion Institute University of Southern California, Los Angeles, CA October 28-29, 1996 8. GRI-Mech Web site, http://www.me.berkeley.edu/grimech 9. Web site for the International Workshop on Measurement and Computaion of Turbulent Nonpremixed Flames (TNF), R. S. Barlow, Ed., Sandia National Laboratories, http://www.ca.sandia.gov/tdf/workshop. 10. Cao, R. and S.B. Pope, Combust. Flame, The Influence of Chemical Mechanisms on PDF calculations of Nonpremixed Piloted Jet Flames, (in preparation), 2005 11. Pope, S.B., Prog. Energy Combust. Sci., 11 (1985) 119-192 12. Van Slooten, P.R., Jayesh and S.B. Pope, Phys. Fluids, 10 (1998) 246-265 13. Jenny, P., M. Muradoglu, K. Liu, S. B. Pope and D. A. Caughey, J. Comput. Phys., 169 (2001) 1-23 14. Muradoglu, M., S.B. Pope and D.A. Caughey, J. Comp. Phys., 172 (2001) 841-878 15. Subramaniam, S. and S.B. Pope, Combust. Flame, 115 (1998) 487-514 16. Sung, C.J., C.K. Law and J.Y. Chen, Proc. Combust. Inst., 27 (1998) 295-304 17. James, S., M.S., Anand, M.K., Razdan and S.B., Pope, in Proceeding of ASME Turbo Expo Land, Sea & Air 1999, The Forty-Fourth ASME Gas Turbine and AeroEngine Technical Congress, Indianapolis, IN, 18. Smooke, M.D., I.K., Puri and K., Seshadri, Proc. Combust. Inst., 21, (1986) 1783-1792. 19. Bennett B.A.V. and M.D., Smooke, (private communication). 20. Schneider, Ch., A. Dreizler, J. Janicka and E.P. Hassel, Combust.Flame, 135, (2003) 185-190 21. Pope, S.B., Combust. Theory Modelling, 1 (1997) 41-63 22. Tang, Q., Ph.D. Thesis, Cornell University, 2003 23. Barlow, R.S., A.N. Karpetis, J.H. Frank and J.Y. Chen, Combust. Flame 127 (2001) 2102-2118 24. Raman, V., R.O. Fox and A.D. Harvey, Combust. Flame, 136, (2004) 327-350 4 th Joint Meeting of the U. S. Sections of the Combustion Institute Page 6 of 6