Thesis Liangyu

download Thesis Liangyu

of 210

Transcript of Thesis Liangyu

  • 7/30/2019 Thesis Liangyu

    1/210

    The Pennsylvania State University

    The Graduate School

    Department of Mechanical and Nuclear Engineering

    DETAILED CHEMISTRY, SOOT, AND RADIATION

    CALCULATIONS IN TURBULENT REACTING FLOWS

    A Thesis in

    Mechanical Engineering

    by

    Liangyu Wang

    c 2004 Liangyu Wang

    Submitted in Partial Fulfillmentof the Requirements

    for the Degree of

    Doctor of Philosophy

    May 2004

  • 7/30/2019 Thesis Liangyu

    2/210

    The thesis of Liangyu Wang has been reviewed and approved* by the following:

    Daniel C. HaworthAssociate Professor of Mechanical EngineeringThesis Co-AdviserCo-Chair of Committee

    Stephen R. TurnsProfessor of Mechanical EngineeringThesis Co-AdviserCo-Chair of Committee

    Andre L. BoehmanAssociate Professor of Fuel Science

    Robert J. SantoroProfessor of Mechanical Engineering

    Richard C. BensonProfessor of Mechanical EngineeringHead of the Department of Mechanical and Nuclear Engineering

    *Signatures are on file in the Graduate School.

  • 7/30/2019 Thesis Liangyu

    3/210

    iii

    Abstract

    The present work aims at a comprehensive approach for the simulation of tur-

    bulent reacting flows. In particular, it focuses on the modeling of detailed chemistry,

    detailed soot formation and oxidation, and the modeling of detailed radiative heat trans-

    fer in gas-phase turbulent flames. In addition, the present work centers on numerical

    investigations of oxygen-enriched turbulent nonpremixed flames.

    Issues that arise in calculating detailed chemistry, soot formation and oxidation,

    and thermal radiation in turbulent reacting flows are reviewed and discussed. Two

    detailed models of turbulent combustion are developed using state-of-the-art models of

    detailed chemistry, soot, and radiation calculations in turbulent flames. One of the

    models is based on an empirical description of the turbulent flow field and the other isbased on CFD modeling of the flow field.

    The empirical-description-based model is an extension of Two-Stage Lagrangian

    (TSL) model of turbulent jet flames. This extension includes the incorporation of a

    detailed soot model and the improvement of the radiation model. The soot model is a

    detailed one adopted from Appel-Bockhorn-Frenklachs soot model. The dynamics of

    soot particles are described by the method of moments adapted to the TSL formulation.

    The original constant-emissivity radiation model is improved by solving the radiative

    transfer equation on the spatial configuration of the TSL model using the spherical

    harmonic P1 method and the discrete ordinate S2 method. The gray medium assumption

    is employed and the Planck-mean absorption coefficient is used to determine the radiative

    properties of both gas-phase species and soot particles. With the extended TSL model,

    the characteristics of soot, radiation and NOx emissions in oxygen-enriched flames are

    studied.

    The CFD-based model is based on an engineering CFD code (GMTEC) and it

    solves the compressible flow equations on unstructured meshes. GMTEC is extended byincorporating a detailed chemistry model, a detailed soot model, and a detailed radiation

    model. The detailed chemistry model is based on the use of the CHEMKIN libraries,

  • 7/30/2019 Thesis Liangyu

    4/210

    iv

    and the calculations of chemistry are accelerated by using the ISAT software. The

    effective use of ISAT for detailed chemistry in nonhomogeneous systems is outlined. The

    detailed soot model is adopted from Frenklachs detailed soot model with the method of

    moments. It is coupled to the three-dimensional CFD code through transport equations

    of soot moments. Two detailed radiation models are implemented, the P1-gray model

    and the P1-FSK model. Both models employ the spherical harmonic P1 method for the

    solution of the radiative transfer equation on three-dimensional unstructured meshes.

    The P1-gray model employs the gray medium assumption and Planck mean absorption

    coefficient for radiative property evaluations. The P1-FSK model addresses the nongray

    nature of the radiative heat transfer by using the full-spectrum k-distribution method.

    The CFD-based comprehensive model is then exercised to simulated an oxygen-enriched

    flame.

    The two detailed models developed have proven to be successful in the simulation

    of oxygen-enriched turbulent flames. The advantage of the TSL model is its compu-

    tational economy. It is shown to be capable of predicting the general trends of soot,

    radiation, and NOx emission with oxygen index, fuel type, and initial jet velocity, but it

    failed to provide quantitative predictions of flame structure due to its simplistic treat-

    ment of the hydrodynamics.

    The advantage of the CFD-based model is its capability of performing detailed,

    quantitative predictions and of capturing the strong couplings among soot, radiation,

    flame structure, and NOx emissions in oxygen-enriched flames. It can be used to identify

    the key sensitivities in soot and NOx formations, to study the effects of nongray gas-

    phase and soot radiation, and to study the influence of mixing, fuel type, and oxygen

    index on the soot formation, NOx emission, and thermal radiation characteristics of

    oxygen-enriched turbulent flames. The deficiencies of the CFD-based model include the

    simple turbulent combustion model, neglect of turbulent fluctuations in composition and

    temperature, and the P1 approximation used to solve the RTE. These are the subjects

    of ongoing research.

  • 7/30/2019 Thesis Liangyu

    5/210

    v

    Table of Contents

    List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

    List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . x

    Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xii

    Chapter 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    1.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

    1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

    1.3 Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

    Chapter 2. Detailed Chemistry Calculations . . . . . . . . . . . . . . . . . . . . 7

    2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

    2.2 Expenses in Chemistry Calculations . . . . . . . . . . . . . . . . . . 10

    2.3 Strategies in Detailed Chemistry Calculations . . . . . . . . . . . . . 13

    2.3.1 Simplification of the Flow Field Physics . . . . . . . . . . . . 13

    2.3.2 Reduction of Chemistry . . . . . . . . . . . . . . . . . . . . . 13

    2.3.3 Storage/Retrieval Scheme . . . . . . . . . . . . . . . . . . . . 15

    2.4 In Situ Adaptive Tabulation (ISAT) . . . . . . . . . . . . . . . . . . 17

    2.5 Turbulence/Chemistry Interactions . . . . . . . . . . . . . . . . . . . 21

    2.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

    Chapter 3. Soot Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

    3.2 Soot Formation and Oxidation: Current Understanding . . . . . . . 30

    3.2.1 Soot Particle Inception . . . . . . . . . . . . . . . . . . . . . . 31

    3.2.2 Soot Surface Growth . . . . . . . . . . . . . . . . . . . . . . . 32

    3.2.3 Soot Particle Coagulation . . . . . . . . . . . . . . . . . . . . 32

    3.2.4 Soot Particle Oxidation . . . . . . . . . . . . . . . . . . . . . 33

  • 7/30/2019 Thesis Liangyu

    6/210

    vi

    3.3 Soot Formation and Oxidation: Modeling . . . . . . . . . . . . . . . 34

    3.4 Method of Moments . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

    3.5 Soot Calculations in Turbulent Flames . . . . . . . . . . . . . . . . . 41

    3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

    Chapter 4. Radiation Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . 45

    4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

    4.2 Governing Equations for Radiative Heat Transfer . . . . . . . . . . . 46

    4.3 Solution Methods for the Radiative Transfer Equation . . . . . . . . 48

    4.3.1 Optically Thin Approximation . . . . . . . . . . . . . . . . . 49

    4.3.2 Spherical Harmonic Method . . . . . . . . . . . . . . . . . . . 49

    4.3.3 Discrete Ordinate Method . . . . . . . . . . . . . . . . . . . . 50

    4.3.4 Zonal Method . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

    4.3.5 Statistical Method . . . . . . . . . . . . . . . . . . . . . . . . 52

    4.3.6 Hybrid Methods . . . . . . . . . . . . . . . . . . . . . . . . . 52

    4.4 Radiative Properties of Participating Media . . . . . . . . . . . . . . 53

    4.4.1 Radiative Properties of Gas-Phase Species . . . . . . . . . . . 53

    4.4.2 Radiative Properties of Soot Particles . . . . . . . . . . . . . 57

    4.5 Turbulence-Radiation Interactions . . . . . . . . . . . . . . . . . . . 61

    4.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

    Chapter 5. Two-Stage Lagrangian Simulations of Oxygen-Enriched flames . . . 65

    5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

    5.2 The Two-Stage Lagrangian Model . . . . . . . . . . . . . . . . . . . 66

    5.3 Detailed Chemical Kinetics . . . . . . . . . . . . . . . . . . . . . . . 68

    5.4 Soot Calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

    5.5 Radiation Calculation . . . . . . . . . . . . . . . . . . . . . . . . . . 72

    5.6 Results and Discussions . . . . . . . . . . . . . . . . . . . . . . . . . 74

    5.6.1 Performence of Radiation Sub-Models . . . . . . . . . . . . . 76

    5.6.2 NOx Emissions . . . . . . . . . . . . . . . . . . . . . . . . . . 76

    5.6.3 Soot Volume Fractions . . . . . . . . . . . . . . . . . . . . . . 80

  • 7/30/2019 Thesis Liangyu

    7/210

    vii

    5.6.4 Flame Radiation . . . . . . . . . . . . . . . . . . . . . . . . . 87

    5.6.5 Artificially Enhanced In-Flame Soot . . . . . . . . . . . . . . 93

    5.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

    Chapter 6. CFD Modeling of Oxygen-Enriched Flames . . . . . . . . . . . . . . 97

    6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

    6.2 The CFD model of flow fields . . . . . . . . . . . . . . . . . . . . . . 98

    6.3 Detailed Chemistry Modeling . . . . . . . . . . . . . . . . . . . . . . 100

    6.3.1 Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

    6.3.2 Turbulent Combustion Model . . . . . . . . . . . . . . . . . . 101

    6.3.3 Effective Use of ISAT . . . . . . . . . . . . . . . . . . . . . . 102

    6.4 Detailed Soot Modeling . . . . . . . . . . . . . . . . . . . . . . . . . 111

    6.4.1 Soot Moment Transport Equations . . . . . . . . . . . . . . . 111

    6.4.2 Soot Moment Source Terms . . . . . . . . . . . . . . . . . . . 116

    6.4.3 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117

    6.5 Detailed Radiation Modeling . . . . . . . . . . . . . . . . . . . . . . 117

    6.5.1 P1-Gray model . . . . . . . . . . . . . . . . . . . . . . . . . . 118

    6.5.2 P1-FSK model . . . . . . . . . . . . . . . . . . . . . . . . . . 120

    6.5.3 Validation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

    6.6 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . 125

    6.6.1 On Radiation Calculations . . . . . . . . . . . . . . . . . . . . 126

    6.6.1.1 Effects of Nongray Radiation . . . . . . . . . . . . . 129

    6.6.1.2 Effects of Soot Radiation . . . . . . . . . . . . . . . 134

    6.6.1.3 Applicability of the P1 method to Jet Flames . . . . 136

    6.6.2 Simulation of an Oxygen-Enriched Flame . . . . . . . . . . . 140

    6.6.2.1 Key Sensitivities in Soot Predictions . . . . . . . . . 140

    6.6.2.2 Key Sensitivities in Radiation Predictions . . . . . . 145

    6.6.2.3 NOxemissions . . . . . . . . . . . . . . . . . . . . . 148

    6.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149

  • 7/30/2019 Thesis Liangyu

    8/210

    viii

    Chapter 7. Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . 151

    7.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151

    7.2 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153

    7.3 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155

    References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158

    Appendix A. Detailed Reaction Mechanism . . . . . . . . . . . . . . . . . . . . . 172

    Appendix B. Radiation Submodels in the TSL Model . . . . . . . . . . . . . . . 184

    Appendix C. Evaluation of Planck-Mean Absorption Coefficients . . . . . . . . . 195

  • 7/30/2019 Thesis Liangyu

    9/210

    ix

    List of Tables

    5.1 Test Conditions of Oxygen-Enriched Flames . . . . . . . . . . . . . . . . 75

    5.2 Calculated Peak Temperatures by TSL . . . . . . . . . . . . . . . . . . . 78

    6.1 Statistics of Using ISAT . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

    6.2 Comparisons Between ISAT and Direct Integration . . . . . . . . . . . . 109

    6.3 Statistics of the Two Approaches to the Mapping Gradient Matrix Cal-

    culation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

    6.4 Summary of Turbulence Model Parameters . . . . . . . . . . . . . . . . 127

    6.5 Case specifications for soot sensitivity study. . . . . . . . . . . . . . . . 142

    6.6 Comparisons of global quantities soot models. . . . . . . . . . . . . . . 144

    6.7 Comparisons of global quantities radiation models. . . . . . . . . . . . 148

  • 7/30/2019 Thesis Liangyu

    10/210

    x

    List of Figures

    2.1 Interconnection between simulation code and ISAT algorithm . . . . . . 19

    5.1 TSL two-reactor and reactor/diffusion-flame models. . . . . . . . . . . . 675.2 The evolution of GRI-Mech. . . . . . . . . . . . . . . . . . . . . . . . . . 70

    5.3 The performance of radiation models. . . . . . . . . . . . . . . . . . . . 77

    5.4 NOx emission indices from TSL model and experiments: fuel, velocity,

    and oxygen index effects. . . . . . . . . . . . . . . . . . . . . . . . . . . 79

    5.5 Temperature profiles from TSL model for propane flames with 40% oxy-

    gen index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

    5.6 Axial profile of equivalent soot volume fraction from experiments for

    propane flames with v0 = 21.8 m/s. . . . . . . . . . . . . . . . . . . . . . 82

    5.7 Axial profile of equivalent soot volume fraction from TSL model for

    propane flames with v0 = 21.8 m/s. . . . . . . . . . . . . . . . . . . . . . 84

    5.8 Normalized peak equivalent soot volume fraction from experiments (curves

    with symbols) and TSL model: fuel and oxygen index effects. . . . . . . 85

    5.9 Normalized peak equivalent soot volume fraction from experiments (curves

    with symbols) and TSL model: fuel jet velocity and oxygen index effects. 86

    5.10 Radiant fractions from experiments: fuel, velocity, and oxygen index effects. 885.11 Radiant fractions from TSL model: fuel, velocity, and oxygen index effects. 89

    5.12 Calculated global residence times as functions of oxygen index for differ-

    ent fuels. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

    5.13 Comparisons between calculated soot and gas-phase radiation heat flux

    for propane and natural gas flames with v0 = 21.8 m/s, 40% oxygen

    index, and soot volume fractions increased to match experiments. . . . . 92

    5.14 Calculated soot contribution to total radiation for propane flames with

    peak soot volume fractions increased to match experiments. . . . . . . . 94

    5.15 Effects of artificially enhanced soot on NOx emission and radiation for a

    propane flame with 30% oxygen index and v0 = 21.8 m/s. . . . . . . . . 95

  • 7/30/2019 Thesis Liangyu

    11/210

    xi

    6.1 Performance of ISAT in terms of CPU time and CPU time ratio . . . . 110

    6.2 A control volume for derivation of moment transport equation . . . . . . 112

    6.3 Computational domain for modeling oxygen-enriched turbulent jet flames 127

    6.4 An artificial absorption coefficient and its planck mean . . . . . . . . . . 130

    6.5 Axial profile of radiation heat flux and temperature calculated by the

    two radiation models: gray vs. nongray . . . . . . . . . . . . . . . . . . 132

    6.6 Radial profiles of temperature calculated by the two radiation models at

    four axial locations: gray vs. nongray . . . . . . . . . . . . . . . . . . . 133

    6.7 (Manipulated) distribution (axial profile left, contour right) of soot vol-

    ume fraction for a propane flame with 40% oxygen index and 21.8 m/s

    jet velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135

    6.8 Axial profile of radiation heat flux and temperature calculated by the

    two radiation models: soot effects . . . . . . . . . . . . . . . . . . . . . . 137

    6.9 Radial profiles of temperature calculated by the two radiation models at

    four axial locations: soot effects . . . . . . . . . . . . . . . . . . . . . . . 138

    6.10 Computational configuration to test the P1 approximation . . . . . . . . 139

    6.11 Axial profiles of radiation heat flux: study of the P1 approximation . . . 141

    6.12 Axial profiles of equivalent soot volume fraction. . . . . . . . . . . . . . 143

    6.13 Axial profiles of radiant heat flux. . . . . . . . . . . . . . . . . . . . . . 146

    6.14 Computed contours of flame temperature, soot volume fraction, and

    species mass fractions of CO2, H2O, and NO. . . . . . . . . . . . . . . . 147

  • 7/30/2019 Thesis Liangyu

    12/210

    xii

    Acknowledgments

    I am most grateful and indebted to my thesis advisors, Dr. Dan Haworth, for

    his guidance and for his profound knowledge in turbulent combustion modeling and in

    numerical methods, and Dr. Steve Turns, for his guidance and for his profound knowledge

    in combustion theory and combustion diagnostics. I am also grateful and indebted to Dr.

    Michael Modest, for his profound knowledge in radiative heat transfer and in numerical

    methods. I thank my other committee members, Dr. Andre Boehman and Dr. Robert

    Santoro, for their insightful commentary on my work. Finally, I thank my family for

    their support and encouragement.

  • 7/30/2019 Thesis Liangyu

    13/210

    1

    Chapter 1

    Introduction

    1.1 Motivation

    Combustion has had significant impact on our daily life since the beginning of

    human history. Today we depend heavily on the combustion processes that transform

    the chemical energy in fossil fuels into the thermal energy that powers our society. Auto-

    mobiles, aircraft, power plants, and furnaces are only a few examples where combustion

    plays an important role. However, in addition to making our lives easier and better,

    combustion also negatively affects our society. It threatens human lives by generating

    environmental pollutant such as oxides of nitrogen (NOx), and by changing the global

    climate pattern via greenhouse effects. Considering the importance of combustion and

    the decreasing resource of fossil fuels, it becomes more and more critical to obtain com-

    bustion processes of high fuel efficiency but low pollutant emissions.

    Most combustion processes occur in a turbulent flow environment: this includes

    automotive engines, gas turbine combustors, and industrial burners, among other de-

    vices. Turbulent combustion is among the most challenging and important subjects in

    theoretical and engineering sciences. It involves a range of complex physical and chemical

    phenomena that interact, strongly and nonlinearly with one other. The major physical

    processes include turbulent transport, finite-rate chemical reaction, radiative heat trans-

    fer, and multiphase flow. Each process individually represents a challenging subject, let

    alone the nonlinear interactions among them. The difficulties of understanding turbulent

    combustion are further compounded by three-dimensional flows in complex geometries,

    such as the combustion chamber of an automotive engine.

    Until about 30 years ago, the study of combustion processes and the development

    of combustion technologies relied almost exclusively on experimental methods. The ex-

    perimental approach is to analyze and optimize the performance of a real combustion

  • 7/30/2019 Thesis Liangyu

    14/210

    2

    process by conducting experiments usually on an abstracted or small-scale version of the

    real process or device, since the actual operating conditions may not be accessible for

    repeated and controlled evaluations. While capable of providing the most realistic an-

    swers to many combustion problems, experimental methods suffer from scaling problems,

    measurement difficulties, operating costs, and time consumption [1].

    Today numerical modeling plays an increasingly important role in the design

    and optimization of turbulent combustion processes. In comparison with experimental

    methods, numerical modeling and numerical experiments may be less expensive and take

    less time than experimental programs. More importantly, they can provide information

    that cannot be obtained from experiments.

    Because of practical limitations of computer storage and speed, and our inability

    to understand and describe mathematically the complex phenomena involved in turbu-

    lent combustion, simplifications of varying degree must be employed in constructing a

    turbulent combustion model. Current simulation codes for single-phase turbulent com-

    bustion may be computational fluid dynamics (CFD) based, with a turbulence sub-

    model such as k-, and a probability density function (PDF) based approach for turbu-

    lence/chemistry interactions. A reduced mechanism for the chemistry of interest is used,

    because even a moderately detailed mechanism that considers 50 chemical species would

    be computationally intractable. Radiative heat transfer should be considered in many

    combustion analysis, but radiation is such a complicated phenomenon that it is either

    treated simply or ignored all together in current combustion simulation codes.

    Detailed chemical mechanisms are required in combustion simulations to address

    issues such as extinction and ignition phenomena, pollutant emissions including NOx,

    soot, CO and unburned hydrocarbons, and unsteady phenomena including combustion

    instabilities. Radiative heat transfer needs to be considered and treated accurately in

    combustion simulations, because it often is the dominant heat transfer mode due to its

    fourth power dependence on temperature. Besides being a primary mechanism for heat

    transfer, radiation changes the flame temperature, affects the flow field through densitychanges, and affects NOx emissions through the thermal NO mechanism. Soot is a major

    pollutant and its formation represents incomplete combustion. Soot is also an important

  • 7/30/2019 Thesis Liangyu

    15/210

    3

    industrial product and it is a strong source of radiation from flames. The prediction of

    soot formation is therefore of interest in many situations.

    With continuous improvements in physical understanding, numerical methods,

    and computer capabilities, combustion modeling is becoming more and more sophisti-

    cated. The incorporation of detailed chemical mechanisms and sophisticated radiation

    models into the current combustion simulation codes is a timely research subject.

    1.2 Objectives

    Development of a comprehensive combustion simulation tool

    Investigation of oxygen-enriched combustion

    Over the past 30 years, advances in computer science and numerical methods have

    made CFD, in place of wind tunnels, the primary analytical tool in solving many aero-

    dynamic problems [1]. At present, however, CFD-based combustion simulation tools

    remain limited in their capacity to deal simultaneously with three-dimensional time-

    dependent turbulent flow, realistic chemistry and turbulence-chemistry interaction, mul-

    tiphase/heterogeneous systems, radiation heat transfer and turbulence/radiation inter-

    action [2, 3]. The need for experiments will remain for the foreseeable future, not only

    to play a primary role in combustion investigation, but also to provide valuable data

    and physical insights for model development and model validation. Nevertheless, it isexpected that in the near future combustion simulation will follow the same pattern as

    observed for aerodynamics where CFD has become a dominant tool in the analysis and

    design of processes and devices.

    For numerical simulations to be truly useful and productive in the design and

    analysis of combustion processes, a comprehensive approach has to be taken, that is, a

    combustion simulation tool must take into account all the pertinent chemical-physical

    phenomena and must incorporate the appropriate corresponding sub-models into the

    overall solution to a combustion process of interest [3, 4]. A comprehensive combustion

    simulation code should be capable of providing detailed information on a combustion

    process and properties, such as temperature and pressure distributions, velocity fields,

  • 7/30/2019 Thesis Liangyu

    16/210

    4

    chemical species compositions, NOx formation, soot formation and oxidation, radiative

    heat loss, and so on.

    Combustion models also serve as a way of organizing our physical understanding.

    A good model should capture our physical understanding of the underlying processes.

    One objective of the present research is to extend the scope of current combustion

    simulation tools by incorporating calculations of detailed chemistry, soot formation and

    oxidation, and radiative heat transfer.

    Another objective of this research is to explore oxygen-enriched combustion. The

    use of oxygen-enriched air or pure oxygen as an oxidizer offers a number of advantages

    in many combustion applications, such as metal heating and melting, glass melting, and

    waste incineration [5]. In glass melting, for example, use of oxygen can result in reduced

    particulate emissions, decreased NOx emissions, increased productivity, and fuel sav-

    ings [6]. Soot formation and thermal radiation are closely coupled factors in determining

    flame structure, temperature, and pollutant emissions, and have particular significance in

    common oxy-fuel combustion applications. Industrial oxy-fuel burners have been devel-

    oped that create low-momentum, highly luminous flames [7]. Studies have demonstrated

    the benefits of this technology in glass manufacturing processes [8]. In other applications

    such as aluminum melting, maintaining a relatively large convective heat transfer com-

    ponent, in addition to increased luminosity, can be advantageous. In such applications,

    an air-oxygen oxidizer is sometimes used rather than pure oxygen. A goal of this research

    is to use the newly developed comprehensive simulation tools to investigate numerically

    the effects of process variables such as oxidizer oxygen content, fuel composition, and

    fuel jet velocity on soot formation, radiation, and pollutant emissions in turbulent jet

    flames.

    1.3 Approach

    If the fuel stream and oxidizer stream are initially separated, a nonpremixed flame

    can then be formed as the two streams mix and react, with the rate of reaction being

    controlled by the rate of mixing. Turbulent nonpremixed flames are employed in many

    practical combustion systems, principally because of the ease with which such flames

  • 7/30/2019 Thesis Liangyu

    17/210

    5

    can be controlled [9]. Examples of such combustion systems include diesel engines, gas

    turbine combustors, and industrial burners. The combustion processes in these systems

    can be idealized as a turbulent nonpremixed jet flame, which is an example of a canoni-

    cal turbulent reacting flow. The jet flame retains the essential physical features of these

    practical combustions systems while allowing detailed quantitative measurements includ-

    ing initial and boundary conditions, and parametric variations of key global parameters,

    such as Reynolds and Damkohler numbers [3]. Measurement and modeling of turbulent

    non-premixed jet flames have been the subject of a biennial international workshop [10].

    Turbulent nonpremixed jet flames are the focus of this research effort, not only

    because of their importance to practical combustion systems, but also because of the

    availability of a large amount of high quality experimental data. Model validation and

    development will be based on two sets of experimental data: measurements for piloted

    methane-air turbulent jet diffusion flames from Sandia National Laboratory [11], and

    measurements for oxygen-enriched turbulent jet flames from the Propulsion Engineering

    Research Center at Penn State University [12].

    Two numerical simulation codes are used for this research effort: the Two-Stage

    Lagrangian (TSL) model developed at Sandia National Laboratory [13] and the GMTEC

    CFD code developed at General Motors [14]. The TSL model has proved to be a useful

    and computationally efficient model for the representation turbulent jet diffusion flames

    [15]. The most significant advantage of the TSL model is its computational efficiency,

    even with detailed chemical mechanisms of more than 50 species. This advantage derives

    from fact that turbulent mixing in the TSL model is treated simply by using experimental

    correlations. In this research, we extend the ability of the TSL code to deal with detailed

    chemistry by incorporating a soot model and by improving its original simple radiation

    model.

    The GMTEC code is a finite-volume CFD code that solves a system of cou-

    pled nonlinear partial differential equations (pdes) for a compressible multi-component

    turbulent flow. Principle pdes correspond to conservation of mass (continuity), mo-mentum, absolute enthalpy and species mass fractions. GMTEC employs an iterative

    time-implicit pressure-based sequential (segregated) procedure for the solution of the

  • 7/30/2019 Thesis Liangyu

    18/210

    6

    coupled pdes. Favre-averaged dependent variables are calculated and a standard two-

    equation k- model is used for turbulent closure. The conservation equations are solved

    on an unstructured mesh using cell-centered variables. The discretization accuracy is

    first-order in time and up to second-order in space. GMTEC was designed for hydrody-

    namics of chemically reacting flows, with ports for incorporation of additional combustion

    sub-models. In this research, we incorporate detail chemistry, soot, and radiation models

    into GMTEC.

  • 7/30/2019 Thesis Liangyu

    19/210

    7

    Chapter 2

    Detailed Chemistry Calculations

    2.1 Introduction

    The mathematical starting point for most nonreacting flow problems is the Navier-

    Stokes equations. For compressible flows, an energy equation is required. The energy

    variable often is taken to be internal energy. Supplemental to this set of conservation

    equations (pdes) are the equations of state and fluid property specification. For real or

    ideal gases, these equations serve as the link between fluid dynamics and thermodynamic

    aspects of the flow. Together with appropriate initial and boundary conditions, these

    equations describe completely a nonreacting flow.

    The mathematical starting point for turbulent reacting flow problems also resides

    on the Navier-Stokes equations. Reacting flows involve frequently large heat release and

    transformation of chemical species. These features of reacting flows require the addition

    of additional terms and equations to the standard set of Navier-Stokes equations. Using

    Cartesian tensor notation, the set of conservation equations for chemically reacting flows

    can be expressed as follows (e.g. [16, 17]):

    mass

    t+

    uixi

    = 0 (2.1)

    momentum

    ujt

    +ujui

    xi=

    ijxi

    p

    xj+ gj for j = 1, 2, 3 (2.2)

    speciesY

    t+

    Yuixi

    = Jixi

    + S for = 1, 2, . . . , N s (2.3)

  • 7/30/2019 Thesis Liangyu

    20/210

    8

    absolute enthalpy

    h

    t+

    huixi

    = Jhixi

    +Dp

    Dt+ ij

    uj

    xi+ Qrad. (2.4)

    In these governing equations, a Roman index denotes a component of a three-dimensional

    vector(e.g., i = 1, 2, 3), a Greek index denotes a chemical species (e.g., = 1, 2, . . . , N s),

    and the usual summation convention applies over repeated Roman indices within a term.

    Here u denotes velocity, Y denotes the mass fractions of the Ns chemical species, and h

    the absolute enthalpy. Mixture mass density is , pressure is p, body force (per unit mass)

    is g, and , J,

    Jh, denote, respectively, the viscous stress tensor and the molecular fluxes

    of species and enthalpy. The chemical source term, S, equals the product of the molar

    chemical production rate, , and molecular weight, W, for species , S = W.

    The volume rate of heating due to radiation is Qrad.

    Compared to the nonreacting case, this set of conservation equations includes

    species transport equations with chemical source terms to address chemical transforma-

    tions, and it employs absolute enthalpy instead of internal energy as the energy variable.

    Chemical heat release produces high temperature in the flow field and variable density

    effects. The high temperature results in thermal radiation playing a dominant role in

    heat transfer. These are a few complexities arising from chemical reactions that com-

    pound the already difficult solutions of the Navier-Stokes equations. Detailed discussions

    on these issues can be found in references [3, 9, 17].

    All of these chemistry-related computations, including radiation properties for a

    participating medium, are based on the information provided by a chemical reaction

    mechanism of interest, which works with a thermodynamic database and a transport

    property database for the chemical species involved. These databases usually organize

    the thermodynamic and transport data in terms of polynomials as functions of temper-

    ature: for example, CHEMKIN database [18], JANAF Tables [19], and NASA database

    [20]. A reaction mechanism is a collection of elementary (or global) reactions necessary to

    describe the chemical transformation and reaction rate coefficients for each reaction; usu-

    ally the latter are given in Arrhenius form. The rate coefficients are derived mainly from

  • 7/30/2019 Thesis Liangyu

    21/210

    9

    experimental measurements in shock tubes and flow reactors, from quantum-mechanical

    calculations based on different theoretical approximations, or from rough estimates based

    on simple collision theories. Dedicated software such as the CHEMKIN package [21] has

    been developed to facilitate chemistry-related computations. Given a reaction mecha-

    nism, a thermodynamic database and a transport database as inputs, the CHEMKIN

    package can return information on chemical production rates, thermodynamic properties,

    and molecular transport coefficients.

    The number of species and the number of elementary reactions in a mechanism

    determine the level of complexity in chemistry calculations. In reality, a chemical mech-

    anism for the transformation of hydrocarbon fuels can include as many as thousands

    of elementary reactions involving hundreds of intermediate species. The combustion of

    a simple fuel like methane in air requires 325 reactions and 53 species (GRI Mech 3.0

    [22]) for a satisfactory chemical description. In the case of auto-ignition of a Diesel fuel

    with the typical fuel cetane, several thousand elementary reactions are required to de-

    scribe the overall process [3]. In practice, the number of species tracked in combustion

    simulations impacts the computer memory usage and CPU time. To make the chem-

    istry calculations tractable in three-dimensional time-dependent CFD simulations even

    with the largest available supercomputers, only a moderate level of chemistry (tens of

    species and hundreds of elementary reactions) generally can be allowed. Broadly used in

    current combustion simulations are mechanisms reduced from their detailed versions, or

    even simpler one-step global mechanisms that have only three generic species: reactant,

    oxidizer, and product.

    However, a detailed reaction mechanism is a prerequisite for a realistic and accu-

    rate prediction of chemically reacting flows. Predictions that require detailed chemistry

    include flame propagation speed, flame location and size, flame extinction and ignition

    phenomena, and pollutant emissions. Many important issues that arise in practical com-

    bustion applications and environmental concerns can be addressed numerically only with

    detailed chemistry. The simulation of engine combustors requires detailed mechanismsto address kinetically controlled phenomena [23] such as low-temperature auto-ignition,

  • 7/30/2019 Thesis Liangyu

    22/210

    10

    and to address time-dependent phenomena such as combustion instability [24] and vari-

    ability [25, 26]. The use of alternative fuels and fuel additives require detailed mecha-

    nisms to address fuel composition issues [27]. The application of turbulent combustion

    models in the chemical industry requires detailed mechanisms to study operating char-

    acteristics of chemical reacting systems [3, 28]. Stringent pollutant emission regulations

    require detailed mechanisms to predict trace pollutant species such as NOx, unburned

    hydrocarbons, and soot [29]. With advances in computer technology and numerical al-

    gorithms, detailed chemistry calculations will eventually become an intrinsic part of any

    comprehensive, reliable, and credible combustion simulation tool.

    2.2 Expenses in Chemistry Calculations

    It is computationally demanding to solve the complete conservation equations

    including detailed chemistry even for simple two-dimensional laminar flames [30]. For

    more complex flows, especially turbulent flames, the inclusion of a detailed chemistry

    calculation is even more challenging. For the foreseeable future, CPU time and com-

    puter memory limitations will prohibit implementations of fully detailed descriptions of

    chemistry into three-dimensional time-dependent CFD simulations of combustion appli-

    cations [31]. This statement is based on three factors that make chemistry calculations

    expensive in CPU time and demanding in memory requirement.

    The first factor is that the chemical system is usually described in terms of themass fractions of each chemical species as functions of space and time. Because of

    the large differences in species diffusivities due to large differences in their molecular

    weight, and because of the large differences in species chemical source terms due to

    large differences in their chemical bonds, each species requires one pde (equation 2.3) for

    its conservation. Therefore, if there are Ns species in the system under consideration,

    usually Ns1 species conservation equations (the sum of species mass fraction is unity)

    are needed to determine the chemical state of the reacting system. For even a moderately

    sized chemical mechanism, this number far exceeds the number of pdes required for mass,

    momentum and energy conservation.

  • 7/30/2019 Thesis Liangyu

    23/210

    11

    The differences in species molecular diffusivities mentioned above often can be

    neglected in turbulent flows because of the relatively large effect of turbulent mixing.

    However, the differences in chemical source terms result in great complexity and intro-

    duce fundamental difficulties in the simulation of turbulent reacting flows.

    The second factor that makes chemistry calculations expensive is the nonlinearity

    and complexity of the chemical source terms. A chemical reaction mechanism involving

    Ns species and L elementary reactions can be written as [9]

    Ns=1

    lX

    Ns=1

    lX for l = 1, 2, . . . , L (2.5)

    where l and

    l are the stoichiometric coefficients for the th species X and l

    th

    reaction. The net molar production rate, i.e., the chemical source term , for each

    species can be written as

    =

    Ll=1

    lql for = 1, 2, . . . , N s, (2.6)

    where,

    l = (

    l

    l), (2.7)

    and,

    ql = kflNs=1

    [X]l krl

    Ns=1

    [X]l . (2.8)

    Here [X] denotes the molar concentration of species X, kfl and krl are, respectively,

    the forward and reverse rate coefficient for the lth reaction and they are related through

    an equilibrium constant. The rate coefficients are commonly expressed in Arrhenius form

    as,

    kfl = AflTbfl exp

    EA,flRuT

    , (2.9)

    where Afl is the pre-exponential factor, bfl is the temperature exponent, and EA,fl is the

    activation energy. These parameters usually are determined by experiments. Inspection

    of the rate expression shows that in order to compute the chemical source term of each

  • 7/30/2019 Thesis Liangyu

    24/210

    12

    species, all the elementary reactions in the mechanisms have to be considered, in general.

    This is tedious for a mechanism with hundreds or thousands of reactions.

    The third expense in chemistry calculations results from the fact that there is a

    large range of chemical time scales inherent in a reaction mechanism, typically ranging

    from 1010 to 1 second or more in combustion problems [32] . Mathematically, chemical

    time scales can be defined by the inverses of the absolute values of the eigenvalues of

    the Jacobian matrix dS/dY. Conceptually, these time scales correspond to the time

    required for the species concentrations to fall from their initial values to a value equal

    to 1/e times their initial value [9]. Since the chemical time scales are different from

    typical mixing time scales in a combustion system, species conservation equations often

    are solved by an operator-splitting method [33, 34], where the the change in composition

    resulting from chemical reaction is determined for every time step by integrating a set

    of odesdY

    dt= S(

    Y , T , p) for = 1, 2, . . . , N s. (2.10)

    The ratio between the largest and smallest time scales characterizes the degree of stiff-

    ness of the ode set. The smallest time scales have to be resolved in the numerical

    solution, even if one is interested only in the slow processes. Otherwise the numerical

    solution tends to become unstable. The wide variation in time scales makes the above

    odes very stiff and severely increases the cost of solving the equations.

    These three observations together illustrate why in a practical combustion calcu-lation with a moderately detailed chemical mechanism, over 90 percent of the CPU time

    may be spent on chemistry calculations. This poses a serious obstacle to combustion

    simulations. For example, in a finite-volume simulation of an axisymmetric laminar jet

    flame with a 122-species mechanism, the CPU time for the integration of equations (2.10)

    is about 0.06 second per time step on a 2.8 GHz Intel Xeon processor. If we use a mod-

    erately fine mesh of 5,000 cells, then 2,000 time steps would require 10 7 integrations,

    which corresponds to seven days of CPU time for just the chemistry. This illustrates

    that detailed chemistry calculations are impossible or extremely expensive to implement

    in three-dimensional time-dependent CFD simulations of practical turbulent combustion

    systems.

  • 7/30/2019 Thesis Liangyu

    25/210

    13

    2.3 Strategies in Detailed Chemistry Calculations

    Over the past decade, significant progress has been made in the ability of compu-

    tational models to model both chemical kinetics and fluid mechanics in turbulent reacting

    flows. Describing turbulent flames using full chemical kinetics and fully resolved fluid

    dynamics is still computationally infeasible due to time and memory constraints. Various

    strategies have been developed to address this issue and these strategies can be classified

    into three categories:

    1. Simplification of the flow field physics;

    2. Reduction of chemistry;

    3. Storage/retrieval schemes for using both detailed descriptions of chemistry and

    fluid dynamics.

    2.3.1 Simplification of the Flow Field Physics

    Simplifications of the flow field physics in turbulent flames have been made possi-

    ble by the dramatic developments in experimental combustion diagnostics in recent years,

    particularly in the use of lasers. These developments have led to an ability to make de-

    tailed measurements within turbulent flames, from which it is possible to construct a

    detailed picture of the structure of the flames [35]. Based on insightful experimental

    observations, models that capture the salient features of combustion fluid dynamics have

    been proposed. One example is the Two-Stage Lagrangian (TSL) model [15] of turbulent

    jet flames. The advantages of such models reside in the fact that they permit the use of

    arbitrarily complex chemistry while retaining the main features of the flow fields. Dis-

    advantages are that although the use of experimental correlations relax the constraints

    imposed by computer time and memory, it precludes using these models as quantitative

    simulation tools for detailed predictions of flame structure.

    2.3.2 Reduction of Chemistry

    The reduction of complex chemical systems has long been desired in combus-

    tion simulations. Besides the driving force obtained from computational complexity in

  • 7/30/2019 Thesis Liangyu

    26/210

    14

    combustion simulations, there are the large uncertainties in determining some elemen-

    tary chemical kinetic rates, which may exceed three orders of magnitude [36]. Reduced

    chemistry enables the simulations of the flames that could not be simulated otherwise,

    and therefore enhances the understanding of these flames and provides more accurate

    predictions or more educated guesses of the flame structures. By introducing approxi-

    mations, however, reduced chemistry looses the ability to describe combustion chemistry

    in a full, accurate, and general manner. Therefore they fail to address many subtle is-

    sues in combustion applications and sometimes can only be applied to a limited range

    of thermo-chemical conditions. Nevertheless, reduced chemistry is used predominately

    in current combustion simulations.

    Numerous approaches have been devised for the reduction of complex chemical

    systems. Here, the reduction of complex chemical systems refers not only to the re-

    duction of reaction mechanisms and number of the chemical species, but also to other

    methods that lower the dimension, or number of degrees of freedom, in the description of

    the time evolution of a chemical system. These other methods include the rate-controlled

    constrained-equilibrium (RCCE) method [37, 38], the computational singular perturba-

    tion (CSP) method [39, 40], and the intrinsic low-dimensional manifolds (ILDM) method

    [41, 42, 43]. Compared to reaction mechanism reductions, which are applied through-

    out the computational domain and time, these other methods can be viewed as reduced

    mechanisms with time- and space-varying stoichiometric coefficients, rate parameters,

    and even number of reactions, although they usually do not provide reduced reaction

    mechanisms in an explicit form. Therefore, mechanism reduction can be termed as

    global chemistry reduction, and the other methods as local chemistry reduction.

    All reduction methods aim at reducing the number of differential equations re-

    quired to solve the chemical system; this usually is the bottleneck in turbulent com-

    bustion calculations. The techniques frequently employed include sensitivity analysis

    (used to identify the rate-limiting reaction steps), reaction flow/flux analysis (used to

    determine the characteristic reaction path), and eigenvalue/vector analysis (used to de-termine the characteristic time scales and directions of the chemical reactions) [32]. In

  • 7/30/2019 Thesis Liangyu

    27/210

    15

    reaction mechanism reductions, quasi-steady-state and/or partial equilibrium assump-

    tions [9] commonly are employed to identify species in steady state and reactions in

    partial equilibrium; these are then used to eliminate species and/or reactions that can

    be represented by algebraic expressions. Reviews on mechanism reduction can be found

    in [44, 45], and the general procedure of applying this approach is described in [46, 36, 31].

    In local reduction methods, CSP and ILDM are based on approaches from dynamic sys-

    tem theory, while RCCE is based on the maximum entropy principle of thermodynamics.

    Both CSP and ILDM employ eigenvalue analysis to identify the fast time scales of the

    chemical reacting systems, the dimensions of which can then be reduced. However, CSP

    aims at decoupling the slow and fast time scales to remove the stiffness of the odes (equa-

    tion 2.10) of the chemical system, while ILDM aims at describing the chemical system

    with only a small number of reaction progress variables. RCCE assumes that a non-

    equilibrium reacting system will relax to its final equilibrium state through a sequence

    of rate-controlled constrained equilibrium states that can be determined by maximizing

    the entropy subject to the instantaneous values of the constraints. Thus only the rate

    equations for the constraints, the number of which is much smaller than the number of

    species, must be integrated. The accuracy of RCCE can be systematically improved by

    adding constraints imposed by increasingly faster reactions. All these reduction methods

    take a detailed mechanism as input. CSP and ILDM can be automated. For mecha-

    nism reduction and RCCE, a computer program can be written to realize automated

    chemistry reduction [47, 38].

    2.3.3 Storage/Retrieval Scheme

    The storage/retrieval schemes [48] comprise the third category of strategies in de-

    tailed chemistry calculations. The basic idea of storage/retrieval schemes is that during

    the course of solving the time evolution of the chemical system represented by equa-

    tions (2.10), the information generated at each integration time step (such as the initial

    conditions and the solutions), are stored in a specially organized table. When similar

    initial conditions are encountered, instead of performing an expensive direct integration

    of the odes, a retrieval or interpolation of the data in the table is performed with a CPU

  • 7/30/2019 Thesis Liangyu

    28/210

    16

    time that is much less than that of the direction integration. For this approach to be

    workable, table reuse must be sufficiently high. Intuitively, this should be the case when

    the system approaches equilibrium or steady state. Mathematically, this has been shown

    to be true for many situations in combustion simulations with the aid of the concepts

    introduced in the ILDM method [41, 48, 32]. ILDM is a useful concept for understanding

    the evolution of reacting systems, in addition to providing a method for chemistry reduc-

    tion. Its conclusions regarding chemical composition space establish the foundation for

    various storage/retrieval methods. Together with the understanding obtained via ILDM

    for reacting flows, the storage/retrieval scheme constitutes an effective solution to the

    problem of using detailed chemistry in CFD simulations.

    Several methods that have been developed recently fall into the storage/retrieval

    category. These methods can be grouped into methods based on tabulation, such

    as structured look-up table (LUT) [49], in situ adaptive tabulation (ISAT) [48], and

    database on-line for function approximation (DOLFA) [50]; methods based on neural

    networks [51, 52]; and methods based on orthogonal polynomials [53, 54, 55]. General

    criteria by which storage and retrieval methods can be judged are the CPU time re-

    quired to create the store; the memory required for the store; inaccuracy in the retrieval

    (or interpolation error); the CPU time required for the retrieval/interpolation; and the

    degree to which the method is generally applicable and automated [48].

    Among these storage/retrieval methods, one of the most successful and promis-

    ing is in situ adaptive tabulation. ISAT achieves an efficient solution of the reaction

    equations through the dynamic creation of a look-up table based on direct integration

    results and an accuracy controlled retrieval based on eigenvector analysis of the reacting

    system, thus allowing for the implementation of detailed chemistry in turbulent reacting

    flow calculations. Successful application of ISAT to a detailed chemical mechanism of

    160 species and 1540 reactions has been reported [56]. ISAT will be discussed further in

    section 2.4.

  • 7/30/2019 Thesis Liangyu

    29/210

    17

    2.4 In Situ Adaptive Tabulation (ISAT)

    The primary concern of ISAT for reacting flow calculations is the efficient solution

    of the reaction equations (2.10). The essential features of ISAT that set it apart from

    other storage/retrieval methods are on-the-fly (or dynamic) tabulation of reaction data

    as needed (in situ), unstructured and adaptive tabulation of data of unknown topology,

    and explicit control of retrieval/interpolation error. It is these features that make ISAT

    suitable to facilitate implementation of detailed chemistry, or chemistry acceleration, in

    turbulent reacting flows. The basic idea and operations of ISAT are described in the

    paper by Pope [48].

    At any spatial location and time in a reacting flow, the thermo-chemical state, or

    the composition, of the gas mixture can be characterized by the mass fractions Y( =

    1, 2, . . . , N s) of Ns species, the enthalpy h or temperature T, and the pressure P. These

    are called the composition variables and can be written in vector form as

    = {1, 2, . . . , Ns+2}. (2.11)

    There are dependences among the components of ; for example, the mass fractions sum

    to unity. Therefore, if the number of independent components is assumed to be D, then

    the thermo-chemical state of fluid is completely determined by

    = {1, 2, . . . , D}. (2.12)

    One set of values of can be considered as one point, or a vector, in the D-

    dimensional composition space. All physically possible values of define the realizable

    region of the composition space, which is a convex polytope in D-dimensional space. The

    accessed region of the composition space is defined as the set of all compositions that

    occur in a particular flame calculation; it is much smaller than the realizable region.

    To illustrate this, consider a location where the temperature and pressure are 1,800 K

    and 1 atm, respectively. In principle, the mass fraction of fuel species can be any value

    between zero and unity; however, under these conditions, its value is almost certain to

  • 7/30/2019 Thesis Liangyu

    30/210

    18

    be zero. Therefore, the accessed region is a small subset of the realizable region, and

    is an intrinsic low-dimensional manifold (ILDM). This observation forms the foundation

    on which various storage/retrieval methods are based.

    ISAT addresses the efficient solution of the reaction equations (2.10) over a time

    step starting from the initial condition (t0) =0, where temperature and pressure are

    taken to be constant without loss of generality. The composition at time t0 + t is a

    unique function of 0; this is called the reaction mapping and is denoted byR(0). The

    conventional approach to determine the mapping is to integrate the reaction equation

    numerically using a stiff ode solver such as DVODE [57]. That approach is referred to

    as direct integration, or DI. For a detailed mechanism, DI is computationally expensive

    for the reasons that have been discussed in section 2.2. ISAT alleviates this problem

    by building up a table during the course of the reacting flow simulation. Each table

    entry corresponds to a solution evolved from a certain initial condition. As the table,

    or as the intrinsic low-dimensional manifold, becomes populated, computation efficiency

    is realized by solving the reaction equation though retrieving and/or interpolating from

    the existing solutions instead of performing expensive DI.

    The coupling of ISAT with a reacting flow simulation code is illustrated in Fig-

    ure 2.1.

    The ISAT algorithm works as follows. Every time the flow code needs a solution,

    or a mapping, of the reaction equation, a query is sent to the ISAT module. At the

    beginning, the ISAT table is empty, so a DI is performed and the resulted mapping is

    forwarded to the flow code and is stored in the table together with its initial condition.

    The store corresponds to a generation of a table entry, or a record, and it is termed an

    add. During an add, computed and stored as well is the mapping gradient A(0), which

    is the Jacobian of the mapping R(0) and defined as

    Aij =dRi(

    0)

    dj. (2.13)

    The mapping gradient contains information about the sensitivity of R to the variations

    in and this information is used for two purposes. First, it establishes a neighborhood

  • 7/30/2019 Thesis Liangyu

    31/210

    19

    Reacting

    Flow

    Simulation

    Code

    ISAT

    Module

    Reaction

    Mapping

    Calculation

    t, tol

    0

    q

    0

    A( ) 0R( )qR( )

    0R( )

    Fig. 2.1. Interconnection between simulation code and ISAT algorithm

    centered on 0, where a linear interpolation suffices to provide for any point (initial

    condition) in the neighborhood the reaction mapping to an accuracy specified by the flow

    code. This neighborhood is called the region of accuracy (ROA). Second, it provides the

    coefficients necessary to perform the linear interpolation.

    For subsequent queries by the flow code with an initial condition q, ISAT at-

    tempts to find a ROA that contains the query point q. Three situations can occur: if

    such a ROA is found, then linear interpolation is performed and the resulting mapping

    is returned. This outcome is termed a retrieve. If such a ROA is not found, then a

    direct integration is performed. Based on the mapping R(q) from DI and the mapping

    gradient, some ROAs are examined to determine if the linear approximation in a certain

    ROA for R(q) is sufficiently accurate. If it is, then that ROA is grown to include the

    query point q and this outcome is termed a grow. If not, the a new ROA is added and

    the outcome is an add.

    The CPU time savings of ISAT comes from mainly the economy of the retrieval/

    interpolation operations compared to the direct integration. One important aspect is

    how fast a ROA can be found that contains the query point q. The organization of the

  • 7/30/2019 Thesis Liangyu

    32/210

    20

    record in the table plays an important role in this regard. ISAT adopted a binary tree

    [58] structure for its record organization. Each leaf of the tree is a record, which consists

    of the tabulation point 0, the reaction mappingR(0), the mapping gradient

    A(0),

    and information on the shape of region of accuracy (consisting of a unitary matrix and

    a vector). Each node of the tree is a cutting plane, which is defined by a vector and a

    scalar such that as the tree is traversed from the top for a query composition q, the

    information given by the vector and the scalar directs the search to either the left branch

    or the right branch of the tree. In this manner, the tree is traversed quickly to find a

    leaf that can be used to approximate the query point q.

    The CPU time savings of ISAT is counterbalanced by the cost of add and grow

    operations, since both operations include direct integration in addition to other manip-

    ulations. Specifically, the fraction of total queries that result in adds and grows is a

    key parameter that determines the gain of ISAT in computational efficiency over direct

    integration. This fraction depends both on the problem at hand and the algorithm used

    to determine the shape and the extent of a ROA added to the table. In general, ISAT

    should be more efficient for state-steady cases than for transient cases, for premixed

    problems than for nonpremixed problems, and for homogeneous reactants than for non-

    homogeneous reactants. There are parameters that control the total number of adds and

    grows, the number of trees used in the table, the error tolerance for linear interpolation,

    and the scaling and transformation of the tabulation points and their mappings.

    ISAT was developed originally for Lagrangian Monte Carlo PDF-based turbulent

    combustion modeling, and most applications to date have been in that context. Most

    applications have been limited to homogeneous systems or to statistically stationary

    configurations with small to moderate-sized mechanism (fewer than 50 species) [48, 59].

    In such cases, speedups compared to direct integration of up to a factor of 1,000 have

    been reported. In other cases [60, 61], speedups of 10 to 100 have been found. ISAT

    also can be used for grid-based CFD applications, such as finite-volume calculations of

    turbulent reacting flow. A difference with respect to a particle PDF method is that thenumber of computational cells used in grid-based CFD calculations is smaller than the

    number of particles used in PDF-based calculations by about two orders of magnitude.

  • 7/30/2019 Thesis Liangyu

    33/210

    21

    ISAT can also be used for large mechanisms containing more than 100 species. The

    difficulty is that the table size becomes quite large and memory usage becomes an issue

    (memory scales as square of the number of species). Special measures have to be taken

    to work around this issue for large mechanisms. Therefore, the benefits of using ISAT for

    grid-based CFD calculations and for large mechanisms are expected to be lower than for

    PDF-based calculations and for medium-sized mechanisms. Still, examples with various

    degrees of success have been reported [62, 56].

    2.5 Turbulence/Chemistry Interactions

    Most combustion applications take place in a turbulent environment. Turbulent

    flow and chemical kinetics are among of the most challenging problems in non-linear

    physics. The strong nonlinear interactions between turbulence and chemistry make tur-

    bulent combustion even harder to understand. Turbulence/chemistry interactions have

    been a central subject in the research of turbulent reacting flows.

    Turbulence/chemistry interactions (TCI) arise from the fact that mixing processes

    in turbulent flow are not fast compared with the rates of chemical reactions. The time

    scales of chemical reactions can range from 1010 s to more than 1 s [32], while the

    time scale of turbulent mixing typically is no smaller than 103 s or 104 s [48]. TCIs

    arise also from the fact that large spatial and temporal variations in species composition

    and temperature occur in turbulent combustion. In nonpremixed combustion, turbulentmixing creates pockets of fuel-rich and fuel-lean mixture, while in premixed combustion,

    turbulent mixing creates pockets of cold unreacted and hot reacted mixture. The mean

    chemical reaction rate can not be evaluated directly from the averaged values of species

    composition and temperature [63].

    The effect of turbulence on chemical reactions takes place through the large-scale

    stirring motions of turbulence. By stretching and curvature, the large-scale motions of

    turbulence enhance greatly the molecular diffusion rates of chemical species and heat,

    and therefore enhance greatly chemical reactions, which occur at molecular scales and

    must be balanced by molecular diffusion. The differential diffusion of chemical species

    also can become important. In addition, the turbulent motions produce large variations

  • 7/30/2019 Thesis Liangyu

    34/210

    22

    in species composition and temperature, which cause the mean reaction rates to be

    strongly coupled to molecular diffusion at the smallest scales of the turbulence.

    The effect of chemical reaction on turbulence takes place through the large heat

    release from chemical reactions. On one hand, the large heat release produces large

    density variations, which influence greatly the solution of Navier-Stokes equations.The

    resultant large density gradients produce a source term in the vorticity transport equa-

    tion and therefore enhance the intensity of turbulence. On the other hand, the heat

    release produces local dilatation, which acts as a sink term in the vorticity equation and

    therefore reduces the intensity of turbulence. Furthermore, the heat release produces

    high temperature regions and therefore high viscosity regions, which enhance the dif-

    fusion of the vorticity in the turbulent flow field. The net effect of heat release on the

    turbulent intensity depends on the specific conditions in which chemical reactions occur

    in turbulent flows.

    Great success has been achieved in applying the conservation equations (2.12.4)

    and detailed chemistry to laminar flames. This demonstrates that our knowledge of

    chemical kinetics and molecular transport processes is sufficient for accurate predictions

    from first principles. Turbulent reacting flows are governed by the same conservation

    equations and the same chemical kinetics (a chemical mechanism for turbulent flames

    should not be different from one for laminar flames, since chemical reactions occur at

    molecular level [64]). However, turbulent reacting flows are characterized by a broad

    spectrum of length and time scales and a complete numerical simulation of such flows

    must resolve the smallest and the largest of all these scales. Direct numerical simulation

    (DNS) is an established technique for that purpose.

    DNS involves solving the conservation equations directly with extremely fine

    meshes to fully resolve all relevant length and time scales and to provide a sequence

    of realizations of the flow that contains a vast amount of details. DNS has been ap-

    plied to simple turbulent flows with simple reactions and to the study of fundamental

    processes involved in turbulent flames, e.g. [65, 66]. It is a powerful research tool fromwhich much can be learned about the physics of turbulent combustion and from which

  • 7/30/2019 Thesis Liangyu

    35/210

    23

    turbulent combustion models can be developed and calibrated. However, even in nonre-

    acting flows, DNS can provide useful information only for simple geometry flows at low to

    moderate Reynolds numbers. In reacting flows, the composition fields introduce length

    and time scales that may be much smaller than those of the velocity fields. The species

    conservation equations with highly nonlinear reaction source terms are more difficult

    to solve than the incompressible Navier-Stokes equations. Therefore for the foreseeable

    future, DNS will not be feasible for accurate predictions of practical turbulent flames.

    Even if a DNS solution of a practical turbulent flame could be obtained, the large

    amount of detail in time and space would be overwhelming and of little practical interest.

    In practice, it is the mean quantities, such as the mean fuel consumption rate and mean

    pollutant formation rate, that typically are desired. This requires limiting the dynamic

    range of length and time scales in the problem and that is accomplished by applying

    various averaging techniques to the conservation equations. Favre averaging (density-

    weighted averaging) is usually employed because of large fluctuations in density due to

    chemical heat release. Averaging reduces greatly the number of degrees of freedom in the

    problem, but it introduces unclosed terms that need to be modeled. The Favre-averaged

    conservation equation are written as follows [16]:

    mass

    t+

    uixi

    = 0 (2.14)

    momentum

    ujt

    +uj ui

    xi=

    uju

    i

    xi+

    ij

    xi

    p

    xj+ gj for j = 1, 2, 3 (2.15)

    species

    Yt

    +Yui

    xi

    = Y u

    ix

    i

    Ji

    xi

    + S for = 1, 2, . . . , N s (2.16)

  • 7/30/2019 Thesis Liangyu

    36/210

    24

    enthalpy

    h

    t+

    huixi

    = hu

    ixi

    Jh

    i

    xi+

    Dp

    Dt+ ij

    uj

    xi+ Qrad. (2.17)

    Here the angle bracket and the tilde operators denote the conventional Reynolds

    averaging and Favre averaging, respectively. The superscript denotes fluctuations from

    the Favre-averaged mean.

    The effects of turbulence on chemistry calculations are embodied in the unclosed

    terms: turbulent fluxes of species (Y

    ui

    ) and energy (hui

    ), and the mean species

    chemical reaction rates (S). The turbulent transport terms for species and enthalpy

    are usually modeled based on the gradient diffusion and turbulent viscosity hypothesis

    [67] for non-reacting flows and are expressed as:

    species

    Yu

    i = t

    Sct

    Yxi

    (2.18)

    enthalpy

    hui = t

    P rt

    h

    xi. (2.19)

    Here t is the turbulent viscosity estimated from a turbulence model, and Sct and

    P rt are the turbulent Schmidt number for the species and turbulent Prandtl number,

    respectively.

    Theoretical [68] and experimental work [69] have shown that turbulent diffusiv-

    ity for reacting scalars is quite different from that for nonreacting cases and further

    investigations are required. Nevertheless, these hypotheses are employed ubiquitously

    in practice. Alternative treatments for the turbulent fluxes that provide more accurate

    results include second-order transport models and PDF methods [3].

    The determination of the mean reaction rates is the central problem of practical

    interest in turbulent combustion simulations. The various approaches for modeling themean reaction rates are generically referred to as turbulent combustion models.

  • 7/30/2019 Thesis Liangyu

    37/210

    25

    Turbulent combustion models can be divided into two major categories according

    to the relative time scales for chemistry compared to other physical processes: fast-

    or-slow-chemistry models and finite-rate-chemistry models. In each category, turbulent

    combustion models are further divided into models for premixed, nonpremixed, or par-

    tially premixed reactants. Although quite different in appearance, the various models

    essentially attempt to represent the same physical processes. Common links among these

    models have been explored, for example, by Veynante and Vervisch [63]. The wide vari-

    ety of combustion model reflects the difficulties that arise in averaging the reaction rate

    terms in the species equations.

    Fast-or-slow-chemistry combustion models essentially ignore the interactions be-

    tween turbulence and chemistry by assuming that the time scales associated with chem-

    ical reactions are very short or very long in comparison with time scales for turbulent

    mixing. That is, mixing or chemistry is the rate-determining process. In the fast- or

    slow-chemistry limit, models are developed based on different concepts. First, a simple

    and widely used approach is to express the mean reaction rate in the same functional

    form as that of the unaveraged reaction rate. Because the latter usually are expressed

    in Arrhenius form, this is sometimes referred to as the Arrhenius model [3]. Second,

    based on the concept of a turbulence energy cascade in nonreacting flows, eddy-breakup

    (EBU) models [70] and variants, such as the eddy-dissipation model [71], link the mean

    reaction rate to the rate of turbulent mixing and hence to a turbulence time scale. Third,

    based on the concept of laminar flamelets [72], a variety of flamelet models have been

    formulated for premixed, nonpremixed, and partially premixed combustion regimes [68].

    Classical flamelet models that fall into this category include the conserved scalar equilib-

    rium models (CSEM) [68] for nonpremixed flames, the Bray-Moss-Libby (BML) model

    [73, 74] and the coherent flame model (CFM) [75, 76] for premixed flames. These fast-or-

    slow-chemistry models sometimes provide plausible predictions, but they cannot capture

    the finite-reaction-rate effects that are often the focus of interest.

    Finite-rate-chemistry combustion models treat chemistry more realistically byconsidering the full range of chemical time scales and by dealing explicitly with the in-

    teractions between turbulence and chemistry. However, special mathematical approaches

  • 7/30/2019 Thesis Liangyu

    38/210

    26

    have to be employed to handle the strong nonlinearities in determining the mean reac-

    tion rates. Statistical approaches are most suitable to this task. These include models

    based on PDF methods and models based on conditional moment closure (CMC) meth-

    ods. The PDF-based models can be further divided into transported PDF models, where

    a modeled PDF transport equation is solved, usually by means of a Lagrangian Monte

    Carlo particle-based method [77, 78, 14], and presumed PDF models, where the shape or

    form of the PDF is parameterized and modeled equations are solved for the parameters

    [3, 17]. Here we shall use the term PDF method to refer to a transported PDF method.

    The CMC method [79, 80] considers conditional averages and higher moments of quan-

    tities such as species mass fractions and enthalpy, conditional on mixture fraction (for

    nonpremixed combustion) or reaction progress variable (for premixed combustion). The

    concept of CMC and its applications have been reviewed in a recent paper by Klimenko

    and Bilger [81].

    Laminar flamelet concepts also can be applied to the modeling of finite-rate chem-

    istry. Flamelet models based on a scalar G-equation for premixed combustion and

    based on mixture fraction for nonpremixed flames have been developed and widely used

    [68]. Another approach to account for finite-rate chemistry is the linear eddy model

    (LEM), which was first formulated for nonreacting flows and later was extended to react-

    ing flows [82, 83]. LEM is a method of simulating molecular mixing on a one-dimensional

    domain embedded in a turbulent flow and therefore is capable of capturing molecular

    effects in turbulent reacting flows.

    Finite-rate-chemistry combustion models have been a key research subject in the

    simulation of turbulent combustion, and the above overview offers only a brief catego-

    rization of the most popular models currently in use. More detailed descriptions and

    formulations can be found in many dedicated books and monographs, such as Libby and

    Williams [17], Peters [68], and Haworth [3].

    The above discussion is based on invoking a statistical averaging technique to re-

    duce the dynamic range of scales in turbulent reacting flows. This approach is commonlyreferred to as Reynolds-averaged Navier-Stokes (RANS) modeling. With appropriate clo-

    sure models, the RANS approach allows only for the determination of mean quantities,

  • 7/30/2019 Thesis Liangyu

    39/210

    27

    which may differ largely from the instantaneous values. Strong unsteady mixing effects

    are observed in turbulent flames and the knowledge of steady statistical means may not

    always be sufficient to describe the turbulent combustion phenomena of interest. An

    alternative approach to reducing the dynamic range of scales while accounting explicitly

    for the unsteady effects is a spatial filtering technique. This is the approach that is taken

    in large eddy simulation (LES).

    In large eddy simulation [67, 68, 84, 85], the larger energy-containing scales are

    resolved explicitly while the effects of unresolved smaller scales are modeled. The distinc-

    tion between the resolved large scales and the modeled small scales usually is determined

    by the grid resolution that is affordable in the computation domain. The small scales

    are the subgrid scales and the models sometimes are referred to as subgrid models.

    Mathematically, the large and small scales are separated by filtering the instantaneous

    governing equations. The resulting filtered equations contain unclosed terms that need

    to be modeled. The modeled equations are solved numerically to simulate the unsteady

    behavior of the large-scale motions. Compared to RANS, LES provides information on

    the large resolved scales, which is valuable in many practical applications such as com-

    bustion stabilities in a gas turbine combustor. As in RANS, the interactions between

    turbulence and chemistry occur at unresolved scales of computation. Therefore, the

    basic tools and formalism of RANS-based turbulent combustion modeling can carried

    directly to LES. Most of the RANS combustion models discussed above can be modified

    and adapted to LES subgrid-scale modeling [63].

    2.6 Summary

    Issues that arise in calculating detailed chemistry in turbulent reacting flows have

    been discussed in this chapter. Detailed chemistry is critical for realistic and accurate

    predictions of turbulent combustion. However, the characteristics of detailed chem-

    istry make the calculations quite difficult. This is mainly due to the complexity of the

    chemistry, e.g., large number of species and elementary reactions required, and the broad

  • 7/30/2019 Thesis Liangyu

    40/210

    28

    range of time scales involved, from 1010 s to more than 1 s. Different strategies to over-

    come these difficulties have been developed, such as simplification of flow field descrip-

    tion, chemistry reduction, and storage/retrieval schemes. One of the storage/retrieval

    schemes, the in situ adaptive tabulation, has been described in detail. This is one of the

    most promising strategies for implementing detailed chemistry in turbulent combustion

    calculations and is the approach that has been adopted for this study. The nature of

    turbulence-chemistry interactions also has been discussed. Direct numerical simulation

    can provide a complete description of a turbulent reacting flow, but is not feasible in

    the foreseeable future for practical combustion systems because of computational power

    limitations and the huge amount of data that is generated. Averaging and filtering tech-

    niques are employed to reduce the range of scales to be resolved in the flows to make

    the simulation tractable. LES is capable of capturing the effects of unsteady mixing by

    explicitly resolving large energy-containing scales and modeling the more homogeneous

    small scales. RANS models can only provide mean quantities but are sufficient for many

    practical applications. Averaging and filtering introduce unclosed terms that need to

    be modeled in the governing equations. The various combustion models that have been

    developed can be divided into fast-or-slow chemistry and finite-rate-chemistry models.

    Fast-or-slow chemistry models are used widely in combustion simulations due to their

    simplicity. Finite-rate-chemistry models provide more realistic results but need statisti-

    cal tools to accommodate the random nature of turbulent flames. The PDF method with

    a Monte Carlo solution scheme is considered to be one of the most promising approaches

    for the modeling of finite-rate turbulent combustion.

  • 7/30/2019 Thesis Liangyu

    41/210

    29

    Chapter 3

    Soot Calculations

    3.1 Introduction

    Most combustion processes that involve hydrocarbons produce soot. Under ideal

    conditions, combustion of hydrocarbons leads mainly to carbon dioxide and water. Under

    practical conditions, in locally fuel-rich regions, the combustion or pyrolysis of hydrocar-

    bons generates intermediate species and radicals that lead eventually to the appearance

    of soot particles.

    Soot consists mainly of carbon. Other elements such as hydrogen and oxygen

    are usually present in small amounts. For example, soot emitted from long-residence-

    time turbulent nonpremixed flames, including toluene, benzene, acetylene, propylene,

    and propane flames burning in air, have the following elemental mole ratio ranges: C:H

    of 8.318.3, C:O of 58109, and C:N of 292976 [86]. Soot density is less than that of

    carbon black and usually in the range of 17001800 kg/m3, depending on the porosity

    of soot [87]. Soot particles are generally small, ranging in size from 5 nm to 80 nm, but

    may be up to several micron in extreme cases [88]. While mostly spherical in shape, soot

    particles may also appear in agglomerated chunks and even long agglomerated filaments

    [86]. Experiments in diffusion flames of hydrocarbon fuels have shown that the soot

    volume fraction generally lies in the range of 106108 [88]. These physical properties

    of soot particles affect their optical properties, which in turn affect the accuracy of

    determining soot quantities and radiation in both experiments and simulations.

    Soot formation is a complicated phenomenon that involves highly coupled chem-

    ical and physical processes. It is remarkable that hydrocarbon fuel molecules containing

    only a few carbon atoms transform into soot particles containing millions of carbon

    atoms. The study of soot processes in combustion systems has drawn great attention.

    Progress has been achieved in understanding the essential features of the chemistry and

  • 7/30/2019 Thesis Liangyu

    42/210

    30

    physics. However, the understanding is incomplete and many questions persist and de-

    bates continue regarding the details of soot nucleation, growth, and oxidation. This

    reflects the difficulty of the problem that soot formation poses for combustion systems.

    The prediction of soot formation is of interest for the following four reasons:

    The formation of soot stems from incomplete combustion, which reduces the com-bustion efficiency.

    Soot is a ma jor pollutant. It contains trace elements that have hazardous effects

    on human health. The particles themselves also are an issue.

    Industrial applications, such as furnace and heat generators, require formation

    of soot to enhance the heat transfer via radiation. However, the soot has to be

    oxidized before these devices release the exhaust into the environment.

    Soot is an important industrial product refered to as carbon black. It finds wide

    application, such as filler in tires or other materials, toner in copiers, and black

    pigment in color printings.

    3.2 Soot Formation and Oxidation: Current Understanding

    The study of soot formation and oxidation includes experimental, theoretical, and

    computational efforts.