Jeff Thesis

Embed Size (px)

Citation preview

  • 8/8/2019 Jeff Thesis

    1/119

    Direct numerical simulation of turbulent,

    chemically reacting flows

    A THESIS

    SUBMITTED TO THE FACULTY OF THE GRADUATE SCHOOL

    OF THE UNIVERSITY OF MINNESOTA

    BY

    Jeffrey Joseph Doom

    IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

    FOR THE DEGREE OF

    DOCTOR OF PHILOSOPHY

    Krishnan Mahesh, Advisor

    January, 2009

  • 8/8/2019 Jeff Thesis

    2/119

    c Jeffrey Joseph Doom 2009

  • 8/8/2019 Jeff Thesis

    3/119

    Acknowledgments

    The person I would like to thank first is my adviser Professor Krishnan Mahesh.

    His experience, intelligence, and patience were instrumental. He provided a

    great graduate experience for which that I am very grateful.

    I am most indebted to my wife, Sarah for her endless support. I would like

    to thank my Mother, Eileen Doom and Sarahs Parents, Roger and Deb Haar

    for their incredible patience and support. I would like to thank my children,

    Kaitlyn and Jonathan. They bring so much joy and happiness to my life. Next,

    I thank Dr. Ken Murphy. He was the one that encouraged me to study science

    and mathematics. I would also like to thank the professors and students at the

    Aerospace Engineering and Mechanics Department for providing an enjoyable

    place to work. In particular, I would like to thank Mr. Michael Mattson,

    Dr. Suman Muppidi and Dr. Yucheng Hou for their enjoyable conversations.

    Finally, I thank my deceased father, Ronald Doom. My father taught me the

    most important skills of all, good work ethic and patience.

    This work was supported by the AFOSR under grant FA95500410341

    and by the Doctoral Dissertation fellowship awarded by the Graduate School

    at the University of Minnesota. Computing resources were provided by the

    Minnesota Supercomputing Institute, the San Diego Supercomputing Center

    and the National Center for Supercomputing Applications.

    i

  • 8/8/2019 Jeff Thesis

    4/119

    To my wife, children, and parents

    ii

  • 8/8/2019 Jeff Thesis

    5/119

    Abstract

    This dissertation: (i) develops a novel numerical method for DNS/LES of com-

    pressible, turbulent reacting flows, (ii) performs several validation simulations,

    (iii) studies autoignition of a hydrogen vortex ring in air and (iv) studies a

    hydrogen/air turbulent diffusion flame.

    The numerical method is spatially non-dissipative, implicit and applicable

    over a range of Mach numbers. The compressible NavierStokes equations are

    rescaled so that the zero Mach number equations are discretely recovered in

    the limit of zero Mach number. The dependent variables are colocated in

    space, and thermodynamic variables are staggered from velocity in time. The

    algorithm discretely conserves kinetic energy in the incompressible, inviscid,

    nonreacting limit. The chemical source terms are implicit in time to allow

    for stiff chemical mechanisms. The algorithm is readily applicable to complex

    chemical mechanisms. Good results are obtained for validation simulations.

    The algorithm is used to study autoignition in laminar vortex rings. A

    nine species, nineteen reaction mechanism for H2/air combustion proposed

    by Mueller et al. [37] is used. Diluted H2 at ambient temperature (300 K)

    is injected into hot air. The simulations study the effect of fuel/air ratio,

    oxidizer temperature, Lewis number and stroke ratio (ratio of piston stroke

    length to diameter). Results show that autoignition occurs in fuel lean, high

    iii

  • 8/8/2019 Jeff Thesis

    6/119

    Abstract iv

    temperature regions with low scalar dissipation at a most reactive mixture

    fraction, MR (Mastorakos et al. [32]). Subsequent evolution of the flame is

    not predicted by MR; a most reactive temperature TMR is defined and shown

    to predict both the initial autoignition as well as subsequent evolution. For

    stroke ratios less than the formation number, ignition in general occurs behind

    the vortex ring and propagates into the core. At higher oxidizer temperatures,

    ignition is almost instantaneous and occurs along the entire interface between

    fuel and oxidizer. For stroke ratios greater than the formation number, ignition

    initially occurs behind the leading vortex ring, then occurs along the length

    of the trailing column and propagates towards the ring. Lewis number is

    seen to affect both the initial ignition as well as subsequent flame evolution

    significantly. Nonuniform Lewis number simulations provide faster ignition

    and burnout time but a lower maximum temperature. The fuel rich reacting

    vortex ring provides the highest maximum temperature and the higher oxidizer

    temperature provides the fastest ignition time. The fuel lean reacting vortex

    ring has little effect on the flow and behaves similar to a nonreacting vortex

    ring.

    We then study autoignition of turbulent H2/air diffusion flames using the

    Mueller et al. [37] mechanism. Isotropic turbulence is superimposed on an

    unstrained diffusion flame where diluted H2 at ambient temperature interacts

    with hot air. Both, unity and nonunity Lewis number are studied. The results

    are contrasted to the homogeneous mixture problem and laminar diffusion

    flames. Results show that autoignition occurs in fuel lean, low vorticity,

    high temperature regions with low scalar dissipation around a most reactive

    mixture fraction, MR (Mastorakos et al. [32]). However, unlike the laminar

    flame where auto-ignition occurs at MR, the turbulent flame autoignites over

  • 8/8/2019 Jeff Thesis

    7/119

    Abstract v

    a very broad range of around MR, which cannot completely predict the onset

    of ignition. The simulations also study the effects of three-dimensionality. Past

    twodimensional simulations (Mastorakos et al. [32]) show that when flame

    fronts collide, extinction occurs. However, our three dimensional results show

    that when flame fronts collide; they can either increase in intensity, combine

    without any appreciable change in intensity or extinguish. This behavior is

    due to the threedimensionality of the flow.

  • 8/8/2019 Jeff Thesis

    8/119

    Table of contents

    Acknowledgments i

    Dedication ii

    Abstract iii

    Table of contents vi

    List of tables ix

    List of figures x

    Nomenclature xiii

    1 Introduction 1

    1.1 Numerical method . . . . . . . . . . . . . . . . . . . . . . . . 1

    1.2 Autoignition of hydrogen vortex ring reacting with hot air . . 3

    1.3 Autoignition in turbulent H2/air diffusion flames . . . . . . . 6

    1.4 Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

    2 Numerical method 11

    2.1 Governing equations . . . . . . . . . . . . . . . . . . . . . . . 11

    2.2 Discretization . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

    2.2.1 Implicit source term . . . . . . . . . . . . . . . . . . . 18

    2.2.2 Zero Mach number limit . . . . . . . . . . . . . . . . . 23

    3 Validation simulations 25

    3.1 Simple chemistry . . . . . . . . . . . . . . . . . . . . . . . . . 263.1.1 Laminar premixed flame . . . . . . . . . . . . . . . . . 26

    vi

  • 8/8/2019 Jeff Thesis

    9/119

  • 8/8/2019 Jeff Thesis

    10/119

    TABLE OF CONTENTS viii

    A.1.2 Momentum . . . . . . . . . . . . . . . . . . . . . . . . 87

    A.1.3 Species . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

    A.1.4 Equation of state . . . . . . . . . . . . . . . . . . . . . 88

    A.1.5 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

    B Flamelet modeling 92

    Bibliography 97

  • 8/8/2019 Jeff Thesis

    11/119

    List of Tables

    3.1 Parameters for laminar diffusion flame with onestep chemistry. 29

    3.2 Parameters for turbulent diffusion flame with onestep chemistry. 34

    3.3 Parameters for diffusion flame with onestep chemistry. . . . . 35

    3.4 Parameters for laminar diffusion flame. . . . . . . . . . . . . . 37

    3.5 Simulation parameters for laminar jet flame. . . . . . . . . . . 38

    3.6 Simulations parameters for lifted jet flame. . . . . . . . . . . . 39

    4.1 Inlet boundary conditions for reacting vortex ring. . . . . . . . 44

    4.2 Mueller Mechanism. . . . . . . . . . . . . . . . . . . . . . . . . 47

    4.3 Global results for homogeneous mixture problem. . . . . . . . 51

    4.4 Parameters used in reacting vortex ring simulations. . . . . . 51

    4.5 Global behavior of reacting vortex ring. . . . . . . . . . . . . . 56

    5.1 References variables for laminar H2/air diffusion flame. . . . . 69

    ix

  • 8/8/2019 Jeff Thesis

    12/119

    List of Figures

    2.1 Storage of variables. . . . . . . . . . . . . . . . . . . . . . . . 15

    2.2 Variation of acoustic CFL and number of outer loop iterations

    with reference Mach number. . . . . . . . . . . . . . . . . . . 22

    3.1 Comparison of computed temperature to analytical solution for

    a laminar premixed flame and spatial order of accuracy. . . . . 26

    3.2 Comparison of asymptotic solution to numerical solution for a

    laminar diffusion flame. . . . . . . . . . . . . . . . . . . . . . . 28

    3.3 Contour plot of temperature for a jet flame and center line pro-

    file of mixture fraction. . . . . . . . . . . . . . . . . . . . . . . 31

    3.4 Scalar dissipation rate and reaction rate for turbulent diffusion

    flame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

    3.5 Isosurface plot of the reaction rate for turbulent diffusion flame

    and contour plot of temperature showing entire computational

    domain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

    3.6 Scatter plot of fuel mass fraction and oxidizer mass fraction. . 34

    3.7 Comparison between Chemkin and present algorithm for a wellstirred reactor. . . . . . . . . . . . . . . . . . . . . . . . . . . 36

    3.8 Grid convergence study showing profiles of a major species and

    a minor species. . . . . . . . . . . . . . . . . . . . . . . . . . . 38

    3.9 Time versus maximum temperature and heat release for a lam-

    inar diffusion jet flame. . . . . . . . . . . . . . . . . . . . . . . 39

    3.10 Contours of a laminar diffusion jet flame. . . . . . . . . . . . . 40

    3.11 Contour plots of a lifted jet flame. . . . . . . . . . . . . . . . . 41

    4.1 Schematic of reacting vortex ring. . . . . . . . . . . . . . . . . 43

    x

  • 8/8/2019 Jeff Thesis

    13/119

    LIST OF FIGURES xi

    4.2 Comparison using onestep chemistry to Hewett and Madnia

    and a comparison of a major species & temperature between

    Chemkin and present algorithm for a wellstirred reactor using

    the Mueller mechanism. . . . . . . . . . . . . . . . . . . . . . 45

    4.3 Grid convergence study for a twodimensional laminar reacting

    vortex ring. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

    4.4 Profile of reaction rates for the species H versus time and tem-

    poral evolution of temperature with all reactions, and reactions

    1, 2, 3, 9, 10 & 11. . . . . . . . . . . . . . . . . . . . . . . . . 48

    4.5 Mixture fraction versus chem for varying temperature and mix-

    ture fraction versus chem for varying fuel/air ratio for the ho-

    mogeneous mixture problem. . . . . . . . . . . . . . . . . . . . 49

    4.6 Domain average of rates of reactions 1,2,3,9,10 and 11 for the

    species H in our reacting vortex ring. . . . . . . . . . . . . . . 52

    4.7 Evolution of autoignition for the reacting vortex ring. . . . . 53

    4.8 Comparison between MR and TMR. . . . . . . . . . . . . . . . 54

    4.9 Time evolution of heat release. . . . . . . . . . . . . . . . . . . 584.10 Contours of heat release and velocity vectors for nonuniform

    and uniform Lewis number. . . . . . . . . . . . . . . . . . . . 59

    4.11 Contours of heat release, temperature, and velocity vectors for

    L/D = 2 and L/D = 6. . . . . . . . . . . . . . . . . . . . . . . 60

    4.12 Time evolution of temperature. . . . . . . . . . . . . . . . . . 62

    4.13 Maximum profiles of heat release, temperature and vorticity

    magnitude versus time. . . . . . . . . . . . . . . . . . . . . . . 64

    5.1 Schematic of initial conditions for the turbulent diffusion flame. 685.2 Comparison between Chemkin and present algorithm for a well

    stirred reactor. Also, grid convergence study of laminar diffu-

    sion flame. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

    5.3 Scatter plot of the laminar diffusion flame with unity and non

    unity Lewis number mixture fractions versus passive scalar. . . 71

    5.4 Evolution of mixture fraction versus heat release and temporal

    evolution of maximum heat release and temperature for the

    laminar diffusion flame. . . . . . . . . . . . . . . . . . . . . . . 72

    5.5 Evolution of temperature from a nondimensional time of 0 to 4. 73

    5.6 Evolution ofYHO2 from a nondimensional time of 0 to 4. . . . 74

  • 8/8/2019 Jeff Thesis

    14/119

    LIST OF FIGURES xii

    5.7 Evolution ofYOH from a nondimensional time of 0 to 4. . . . 74

    5.8 Contours of n, T, , , YHO2, and || for the turbulent H2/airdiffusion flame. . . . . . . . . . . . . . . . . . . . . . . . . . . 75

    5.9 Plot of mixture fraction () versus temperature in Kelvin (T). 76

    5.10 Contour lines of mixture fraction and temperature superposed

    on shaded contours of reaction rate at different instants of time. 78

    5.11 Comparison of temperature for unity Lewis number and non

    unity Lewis number. . . . . . . . . . . . . . . . . . . . . . . . 79

    5.12 Effect of grid resolution for a turbulent diffusion flame. . . . . 80

    5.13 Time evolution of flame fronts. . . . . . . . . . . . . . . . . . . 82

    B.1 Comparison of a major species & temperature between Chemkin

    and flamelet code for a wellstirred reactor. . . . . . . . . . . 95

    B.2 Temporal evolution of mixture fraction versus T, , YOH, and

    YNO. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

    B.3 Contours of||, T, YH2O, and YH. . . . . . . . . . . . . . . . 97

  • 8/8/2019 Jeff Thesis

    15/119

    Nomenclature

    x, y, z Coordinate axes, coordinates

    Density

    u, v, w Velocity in the x, y and z directions

    Yk Mass fraction of species k

    p Pressure

    T Temperature

    t Time

    V Volume

    A Area

    E Total energy per unit mass

    Re Reynolds number

    Sc Schmidt numberLe Lewis number

    P r Prandtl number

    Dynamic viscosity

    W Mean molecular weight

    xiii

  • 8/8/2019 Jeff Thesis

    16/119

    Nomenclature xiv

    Ru Universal gas constant

    Dk Diffusion coefficient of the kth species

    cp Specific heat at constant pressure

    cv Specific heat at constant volume

    Qdk Heat of reaction per unit mass

    k Mass reaction rate for the kth

    speciesn Heat release

    Subscripts and Superscript

    ( )k kth species

    ( )r Reference variable

    ( )face Faces of control volume

    ( )cv Control volume( )d Dimensional value

    Abbreviations

    DNS Direct Numerical Simulation

    LES Largeeddy Simulation

    RANS Reynoldsaveraged NavierStokes

  • 8/8/2019 Jeff Thesis

    17/119

    Chapter 1

    Introduction

    1.1 Numerical method

    The direct numerical simulation (DNS) and large-eddy simulation (LES) of

    turbulent reacting flows are extremely challenging. Combustion involves a

    large number of chemical species, and associated chemical reactions. Different

    chemical reactions possess different time-scales, and different reaction zone

    thicknesses. Turbulence introduces its own range of length and time-scales.

    Furthermore, the nonlinear nature of turbulence can cause errors at the smaller

    scales to corrupt the large scales of the solution. Turbulent combustion can

    occur at very low Mach numbers (e.g. gas-turbine combustors at low pressures)

    or very high Mach numbers (e.g. scramjets). Very low Mach numbers imply

    numerical stiffness because of a large difference between the speed of sound

    and flow velocities, while very high Mach numbers result in shock waves, and

    their attendant problems. Desirable requirements for an algorithm to perform

    DNS/LES of turbulent reacting flows are therefore: (i) to handle high Reynoldsnumbers robustly and accurately, without the use of numerical dissipation,

    1

  • 8/8/2019 Jeff Thesis

    18/119

    1.1. Numerical method 2

    (ii) to handle acoustic stiffness efficiently, (iii) to deal with chemical stiffness,

    and combustion-induced large spatial gradients, and (iv) to deal with shock

    waves/detonation. This dissertation proposes an approach that deals with

    requirements (i) through (iii).

    Most DNS/LES of turbulent reacting flows appear to either use the com-

    pressible NavierStokes equations with Pade spatial discretization, and explicit

    time-advancement (e.g. Poinsot [11], Trouve [50], Lele [27], & Pantano [39]),

    or the zero Mach number equations along with a pressure-projection approach

    (e.g. Majda & Sethian [31], Montgomery & Riley [36], Rutland & Ferziger [43]

    [44], Pember et al [40], and Mahesh et al. [30]). A secondorder, staggered,

    implicit compressible algorithm was proposed by Wall et al. [53]. However, the

    Pade schemes become unstable at high Reynolds numbers; when explicit time-

    advancement is used, they require very small time-step at low Mach numbers,

    and for stiff chemical mechanisms. The zero Mach number equations are very

    efficient at low Mach numbers because they analytically project acoustic waves

    out; also along with implicit time-advancement they can resolve chemical stiff-

    ness efficiently. However, due to the complete absence of acoustic effects, they

    are not applicable to finite Mach number flows.

    The approach proposed in this dissertation is an extension of the algorithm

    developed by Hou & Mahesh [23] for compressible flows without chemical re-

    action. The Hou & Mahesh algorithm is colocated in space, symmetric in

    both space and time, and hence non-dissipative. Motivated by the theoretical

    work of Thompson [49], and similar to the non-dimensionalization used by Bijl

    & Wesseling [1], and van der Heul et al. [52]. The NavierStokes equations

    are non-dimensionalized using an incompressible scaling for pressure, and the

    energy equation is interpreted as an equation for the divergence of velocity. A

  • 8/8/2019 Jeff Thesis

    19/119

    1.2. Autoignition of hydrogen vortex ring reacting with hot air 3

    pressurecorrection approach is used to constrain the divergence of the velocity

    field to satisfy the energy equation. The resulting system of equations ana-

    lytically projects acoustic waves out, in the limit of zero Mach number. The

    discrete equations are fully implicit, and are constrained to discretely conserve

    kinetic energy in the limit of incompressible, inviscid flow. These features

    make the algorithm stable and accurate at high Reynolds numbers, and effi-

    cient and accurate at very low Mach numbers. Hou & Mahesh [23] show results

    for one-dimensional acoustic waves, a periodic shock tube, the incompressible

    Taylor problem, and inviscid isotropic turbulence on very coarse grids. We

    extend the Hou & Mahesh [23] algorithm to include chemical reaction.

    1.2 Autoignition of hydrogen vortex ring re-acting with hot air

    The motion of a fluid column of length L through an orifice of diameter D

    produces a vortex ring. Nonreacting vortex rings have been studied by many

    researchers, and a large volume of work exists (see e.g. the review by Shariff

    and Leonard [47]). Less is known about reacting vortex rings. This dissertation

    is motivated by the relevance of vortex rings to fuel injection, and uses direct

    numerical simulation (DNS) to study the flow generated by a piston pushing

    a ring of gaseous fuel into hot oxidizer.

    Past work on nonreacting vortex rings (Maxworthy [33], [34]; Didden [12];

    Glezer [18]) has studied the process of vortex ring formation, their kinematics,

    and the temporal evolution of vortex circulation. A notable finding is that

    of Gharib et al [17] who show that a vortex ring achieves its maximum cir-

  • 8/8/2019 Jeff Thesis

    20/119

    1.2. Autoignition of hydrogen vortex ring reacting with hot air 4

    culation at the critical stroke ratio (termed formation number) across which

    the flow transitions from a toroidal vortex ring to a vortex ring with trailing

    column of vorticity. Here, the stroke length ratio is defined as the ratio of the

    piston stroke length to the nozzle exit diameter. Sau and Mahesh [46] show

    how the stroke length affects entrainment and mixing. They show that en-

    trainment is highest at the formation number, and decreases when the trailing

    column forms. They explain this behavior by showing that a toroidal vor-

    tex ring entrains ambient fluid efficiently, but the trailing column does not.

    As the stroke length increases towards the formation number, the vortex ring

    remains toroidal, but increases in size and circulation, which increases entrain-

    ment. When the stroke ratio exceeds the formation number, the leading vortex

    ring remains the same, and the trailing column increases in length. The frac-

    tional contribution of the leading vortex ring to overall entrainment therefore

    decreases. Since the trailing column does not entrain ambient fluid efficiently,

    the overall entrainment drops.

    Chen et al. [5][7] have performed experiments on reacting vortex rings of

    (pure/ nitrogen diluted) propane and ethane issuing into air. The fuel and oxi-

    dizer are initially at the same temperature (300 K) in their experiments, and a

    pilot flame is used to initiate combustion. Luminosity measurements are used

    to provide data on flame shape and height. The fuel volume and ring circu-

    lation are varied in a controlled manner. The nozzle exitvelocity is assumed

    tophat (U0), and ring circulation is estimated from hotwire measurements

    as = U20 t where t is the duration of the pulse. The fuel volume is esti-

    mated as VF = U0tA where A is the nozzle crosssectional area. Chen et al.

    also perform simulations of methane/air vortex rings using a steady laminar

    flamelet approach, where the chemical species and temperature are functions

  • 8/8/2019 Jeff Thesis

    21/119

    1.2. Autoignition of hydrogen vortex ring reacting with hot air 5

    of the conserved scalar and scalar dissipation rate. The speed of the vortex ring

    is seen to initially increase, and then decrease at later times. Dilatation due

    to heat release is shown to be responsible for this behavior. The experiments

    show that as the fuel volume or circulation is changed, the flame structure and

    burnout time change significantly. The formation number (which they term

    the overfill limit) does not significantly change from the nonreacting case.

    Fuel volume which exceeds the overfill limit is seen to decrease heat release

    and ring speed. Also, the burnout time and flame shapes are quite different

    for ethane and propane under similar conditions although the stoichiometry

    and diffusivity are nearly identical. Nitrogen dilution (for propane) decreases

    burnout time and flame luminosity, but does not appreciably change flame

    structure.

    Simulations of reacting vortex rings in a slightly different configuration

    were performed by Hewett & Madnia [19] and Safta & Madnia [45]. Hewett

    & Madnia solved the axisymmetric compressible NavierStokes equations and

    used a onestep Arrhenius mechanism to model combustion of a hydrocarbon

    in air. No pilot flame exists; instead the vortex ring of fuel issues into hot

    oxidizer. The stroke ratio of the vortex ring is 2 for all cases considered. The

    simulations study the effect of oxidizer temperature which is controlled such

    that ignition occurs either when the vortex rings forms, or occurs after the

    vortex ring has fully formed. They find that at higher oxidizer temperatures,

    most of the reaction occurs in front of the vortex ring, while at lower temper-

    atures, most of the reaction occurs inside the ring. The heat released during

    reaction is shown to affect the ring vorticity and strain rate. Safta & Madnia

    [45] consider the same problem, but use a methane-hydrogen mixture as fuel.

    They contrast three chemical mechanisms - GRIMech v3.0, and two reduced

  • 8/8/2019 Jeff Thesis

    22/119

    1.3. Autoignition in turbulent H2/air diffusion flames 6

    mechanisms with 11 step, 15 species and 12 step, 16 species respectively.

    The objective of this dissertation is to study the coupling between the

    fluid flow and chemistry with as few simplifying assumptions as possible. We

    therefore use direct numerical simulation to study the flow when a round nozzle

    injects H2 + N2 fuel at room temperature into air at higher temperatures. The

    amount of fuel injected is controlled by varying the stroke length or stroke ratio.

    The governing equations are the threedimensional, unsteady, compressible,

    NavierStokes equations along with conservation equations for the nine species

    described by the Mueller et al. [37] chemical mechanism. The simulations

    consider the effects of Lewis number, stroke length ratio and fuel/air mixtures

    (equivalence ratio) on autoignition and the flow.

    1.3 Autoignition in turbulent H2/air diffu-

    sion flames

    The autoignition of turbulent diffusion flames is central to applications such

    as diesel engines and scramjet engines where fuel is injected into hot oxidizer.

    Fuel and oxidizer mix through convection and diffusion, then autoignite due

    to the high temperatures of the oxidizer. Direct numerical simulation (DNS) is

    used to study a diffusion flame where fuel (hydrogen/nitrogen) reacts with air

    (oxygen/nitrogen). For laminar flames, fuel and oxidizer diffuse, then react,

    yielding products. In turbulent flames, one might see autoignition, extinction

    and even reignition due to the turbulence.

    Mahalingam et al. [29] performed DNS of three-dimensional turbulent

    nonpremixed flames with finite rate chemistry. They incorporated onestep

  • 8/8/2019 Jeff Thesis

    23/119

    1.3. Autoignition in turbulent H2/air diffusion flames 7

    and twostep reaction models for the chemistry. Their results show that the

    intermediate species concentration overshoots the experiments. They suggest

    that this behavior is due to timedependent strain rates in the turbulent flow.

    Montgomery et al. [36] also performed DNS of a three-dimensional turbulent

    hydrogenoxygen nonpremixed flame using a reduced mechanism (7species,

    tenreactions). The simulations were used to suggest a mixture fraction and

    progress variable model for the chemical reactions. Im et al. [24] performed

    DNS of twodimensional turbulent nonpremixed hydrogenair flames to study

    autoignition. They show that peak values ofHO2 align with maximum scalar

    dissipation during ignition. Im et al. also noted that weak and moderate tur-

    bulence enhanced autoignition while stronger turbulence delayed ignition.

    Echekki & Chen [15] used DNS to study autoignition of a hydrogen/air mix-

    ture in two dimensional turbulence. Their results show spatially localized

    sites where autoignition begins, which they define as a kernel. These ker-

    nels have a build up of radicals at high temperatures, fuellean mixtures, and

    low dissipation rates. Mastorakos et al. performed twodimensional simula-

    tion of autoignition of laminar and turbulent diffusion flames with onestep

    chemistry. They found that ignition always occurs at a welldefined mix-

    ture fraction (MR). Hilbert & Thevenin [20] performed similar simulations to

    Mastorakos et al. [32] and Im et al. [24]. The difference between Hilbert &

    Thevenin [21] and others is that the simulation uses multicomponent diffusion

    velocities.

    A significant difference between the past work cited above, and our DNS is

    that we perform fully three dimensional simulations using finite rate chemistry.

    Also, we incorporate the Mueller mechanism for hydrogen/air (nine species

    and nineteen reaction) and not a reduced mechanism. The objectives of our

  • 8/8/2019 Jeff Thesis

    24/119

    1.4. Contributions 8

    simulations are: (i) to study major differences between two dimensional and

    three dimensional simulations of turbulent diffusion flames. In particular,

    what happens to the flame front in three dimensions? (ii) How do the ignition

    kernels evolve in time? (iii) What is the effect of unity Lewis number and

    nonunity Lewis number on ignition, and (iv) How does the mixture fraction

    compare to a passive scalar?

    1.4 Contributions

    The principal contributions of this work are:

    We have developed a novel numerical method for DNS/LES of reactingflow. The algorithm is spatially nondissipative, implicit and applicable

    over a range of Mach number. It discretely conserves kinetic energy in

    the incompressble, invisid, nonreacting limit. The algorithm is readily

    applied to complex chemistry.

    DNS is used to study autoignition of a hydrogen vortex ring in hot air.The most reactive mixture fraction (Mastorakos et al. [32]) is found to

    predict the onset of initial ignition but not the subsequent evolution. A

    most reactive temperature TMR is defined and shown to predict both the

    initial autoignition as well as subsequent evolution.

    The effect of stroke ratio and oxidizer temperature is studied. For strokeratios less than the formation number, ignition in general occurs behind

    the vortex ring and propagates into the core. At higher oxidizer tem-

    peratures, ignition is almost instantaneous and occurs along the entire

    interface between fuel and oxidizer. For stroke ratios greater than the

  • 8/8/2019 Jeff Thesis

    25/119

    1.4. Contributions 9

    formation number, ignition initially occurs behind the leading vortex

    ring, then occurs along the length of the trailing column and propagates

    towards the ring.

    Lewis number is shown to affect both the initial ignition as well as sub-sequent flame evolution significantly. Nonuniform Lewis number simu-

    lations provide faster ignition and burnout time but a lower maximum

    temperature.

    The effect of fuel/air ratio was studied. The fuel rich reacting vortexring provides the highest maximum temperature and the higher oxidizer

    temperature provides the fastest ignition time. The fuel lean reacting

    vortex ring has little effect on the flow and behaves similar to a non

    reacting vortex ring.

    DNS is used to study a turbulent hydrogen/air diffusion flame. Auto-ignition is shown to occur in fuel lean, high temperature regions with

    low scalar dissipation and low vorticity.

    Unlike laminar diffusion flames where auto-ignition occurs at MR, the

    turbulent flame autoignites over a very broad range of around MR.

    The simulations study the effects of three-dimensionality. Twodimensionalsimulations show that when flame fronts collide, extinction occurs (Mas-

    torakos et al. [32]). However, our three dimensional results show that

    when flame fronts collide; they can either increase in intensity or extin-

    guish.

    Comparison between unity Lewis number and nonunity Lewis number

  • 8/8/2019 Jeff Thesis

    26/119

    1.4. Contributions 10

    simulations show that viscous diffusion is important even in the turbulent

    regime.

    It is shown that inadequate grid resolution causes autoignition to occurin fuelrich regions at low temperature. This is a non physical effect of

    grid resolution.

    This dissertation is organized as follows. The numerical method is de-

    scribed in chapter 2. Chapter 3 discusses validation cases of varying com-

    plexity. Chapter 4 considers autoignition of a hydrogen vortex ring reacting

    with hot air. Direct numerical simulation of autoignition in turbulent non

    premixed H2/air flames is discussed in chapter 5.

  • 8/8/2019 Jeff Thesis

    27/119

    Chapter 2

    Numerical method

    This chapter extends the Hou & Mahesh [23] algorithm to include the effects

    of chemical reaction. The resulting algorithm treats the chemical source terms

    implicitly, solves the species equations in a segregated manner, which allows

    easy extension to multiple species and chemical reactions, and reduces to the

    zero Mach number equations in the limit of very small Mach number. The

    chapter is organized as follows. Section 2.1 describes the non-dimensional gov-

    erning equations, and their behavior in the limit of very small Mach number.

    The discrete scheme is described in section 2.2. The positioning of variables,

    and details of the pressurecorrection approach, discretization of the chemical

    source terms are discussed. Also, the behavior of the algorithm in the zero

    Mach number limit is illustrated.

    2.1 Governing equations

    The governing equations are the compressible, reacting Navier-Stokes equation

    for an ideal gas:

    11

  • 8/8/2019 Jeff Thesis

    28/119

    2.1. Governing equations 12

    d

    td+

    udjxdj

    = 0, (2.1)

    dYdktd

    +Ydk u

    dj

    xdj=

    xdj

    dDdk

    Ydkxdj

    + dk, (2.2)

    dudi

    td+

    dudi udj

    xdj=

    pd

    xdi+

    dij

    xdj(2.3)

    dEd

    td+

    dEd + pd

    udjxdj

    =diju

    di

    xdj+

    xdj

    d

    cdpP r

    Td

    xdj

    +

    Nk=1

    Qdkdk (2.4)

    pd = dRdTd = dRuWd

    Td. (2.5)

    where the superscript d denotes the dimensional value. The variables , p, Yk

    and ui denote the density, pressure, mass fraction of species k and velocities

    respectively. E = cvTd + udi u

    di /2 denotes the total energy per unit mass and

    ij = dudixdj

    +ud

    j

    xdi 2

    3

    udk

    xdk

    ij

    is the viscous stress tensor. Ddk, cp and P r

    denote the diffusion coefficient of the kth species, specific heat at constant

    pressure and the Prandtl number. For the source term, Qdk is the heat of

    reaction per unit mass and Qdkdk is the heat release due to combustion for

    the kth species. k is the mass reaction rate for the kth species. Ru is the

    universal gas constant and Wd is the mean molecular weight of the mixture

    defined as: 1Wd

    =N

    k=1YkWk

    . The source term is modeled using the Arrhenius

    law in this paper.

    The reacting governing equation are non-dimensionalized as follows. Let

    r, Yr, L, & Tr denote the reference density, mass fraction, length and tem-perature respectively. The reference velocity, dynamic viscosity and pressure

  • 8/8/2019 Jeff Thesis

    29/119

    2.1. Governing equations 13

    are denoted by ur, r and pr respectively. The ratio =cdpcdv

    is assumed to

    be constant (e.g. Poinsot & Veynante [42]). This yields the following non-

    dimensional variables:

    =d

    r, ui =

    udiur

    , t =td

    L/ur, =

    d

    r, p =

    pd prrur2

    , (2.6)

    T =

    Td

    Tr , Mr =

    urar =

    urRrTr , pr = rRrTr, (2.7)

    Yk =YdkYr

    , Dk =DdkurL

    , k =Ldk

    urrYr, Qk =

    YrQdk

    cp,rTr, (2.8)

    Rr =RuWr

    , and W =Wd

    Wr. (2.9)

    Note that pressure is nondimensionalized using an incompressible scaling.

    This non-dimensionalization is motivated by Thompson [49], Bijl & Wesseling

    [1], Van der Heul et al. [52] and Hou & Mahesh [23]. Therefore, the non-

    dimensional governing equations for reacting flows are:

    t+

    ujxj

    = 0, (2.10)

    Ykt

    +Ykuj

    xj=

    1

    ReSck

    xj

    Ykxj

    + k, (2.11)

    ui

    t+

    uiuj

    xj=

    p

    xi+

    1

    Re

    ij

    xj, (2.12)

    M2r

    t

    p +

    12

    uiui

    +

    xj

    p +

    12

    uiui

    uj

    +

    ujxj

    =( 1)M2r

    Re

    ijuixj

    +1

    RePr

    xj

    W

    T

    xj

    +

    Nk=1

    Qkk (2.13)

    TW = M2rp + 1. (2.14)

  • 8/8/2019 Jeff Thesis

    30/119

    2.2. Discretization 14

    where Sck is the Schmidt number for the kth species. When the reference Mach

    number is zero, the non-dimensional reacting governing equations reduce to:

    t+

    ujxj

    = 0, (2.15)

    uit

    +uiuj

    xj= p

    xi+

    1

    Re

    ijxj

    , (2.16)

    Ykt

    + Ykujxj

    = 1ReSck

    xj

    Ykxj

    + k, (2.17)ujxj

    =1

    RePr

    xj

    W

    T

    xj

    +

    Nk=1

    Qkk, (2.18)

    T

    W= 1. (2.19)

    Notice that the divergence of velocity equals the sum of the terms involving

    thermal conduction and heat release. If the density is constant and there is

    no heat release, the energy equation reduces to the incompressible continuity

    equation. In the presence of heat release, the zero Mach number reacting

    equations (Majda & Sethian [31]) are obtained. Most projection methods for

    the zero Mach number equations project the momentum ui to satisfy the

    momentum equation. Here, the velocity is projected to satisfy the energy

    equation. The reaction source term can be quite complicated for multiple

    species, and is discussed in more detail in section 2.2.1.

    2.2 Discretization

    Density, pressure, and temperature are staggered in time from velocity by

    Hou and Mahesh [23]. The mass fraction of species k are similarly staggered

    in time here. The thermodynamic variables and mass fraction of species k are

  • 8/8/2019 Jeff Thesis

    31/119

    2.2. Discretization 15

    VNt

    VNtVN

    t

    VNt

    VNt

    VNt

    t+ 12

    t+ 12

    1

    2t+

    T1

    2t+ , , ,

    t u pi Y

    k,

    Figure 2.1: Storage of variables.

    advanced in time from t + 12

    to t + 32

    , illustrated in figure 1. The variables are

    colocated in space, to allow easy application to unstructured grids.

    Integrating over the control volume and applying Gausss theorem yields the

    discrete governing equations. The discrete continuity and species equations

    are

    t+ 3

    2

    cv t+1

    2

    cv

    t+

    1

    V

    faces

    t+1facesVt+1N Aface = 0 (2.20)

    and

    (Sk)t+ 3

    2

    cv (Sk)t+1

    2

    cv

    t+

    1

    V

    faces

    (Sk)t+1faces V

    t+1N Aface

    =1

    ReSck

    1

    V

    faces

    1

    Skxj

    Sk2

    xj

    t+1faces

    NjAface + t+1k . (2.21)

    where cv denotes i,j,k and faces denotes summation over all the faces of thecontrol volume. faces and vN denotes the density and normal face velocity at

  • 8/8/2019 Jeff Thesis

    32/119

    2.2. Discretization 16

    each face. Note that Sk = Yk. The variables (Sk)faces denotes the species at

    the face and Nj is the outward normal vector at the face. The variables V and

    Aface denote the volume of the control volume and the area of the face. The

    discrete momentum equation is:

    (gi)t+1cv (gi)tcv

    t

    +1

    V faces (gi)t+ 1

    2

    faces Vt+ 1

    2

    N Aface

    = 1V

    faces

    ptfacesNjAface +1

    Re

    1

    V

    faces

    (ij)t+ 1

    2

    face NjAface. (2.22)

    Here, gi = ui denotes the momentum in the i direction and ij is the stress

    tensor. The discrete energy equation is given by:

    M2r

    tpcv + 1

    2

    uiuit+ 1

    2

    +1

    V facespcv + 1

    2

    uiuit+ 1

    2

    face vt+ 1

    2

    N Aface+

    1

    V

    face

    vt+ 1

    2

    N Aface =( 1)M2r

    Re

    1

    V

    faces

    (ijui)t+ 1

    2

    face NjAface

    +1

    RePr

    1

    V

    faces

    W

    Tt+1

    2

    N

    Aface +

    Nk=1

    Qkt+ 1

    2

    k . (2.23)

    The algorithm solves each equation separately. This allows one to add multiple

    species with relative ease, which allows easy extension to complex chemistry.

    Note that the chemical source term is handled implicitly. Details of the implicit

    procedure are described in section 2.2.1. The algorithm is a pressure-correction

    method. An important feature is that the face-normal velocities are projected

    to satisfy the constraint on the divergence which is determined by the energy

    equation. At small Mach number, this feature ensures that the velocity field is

    discretely divergence free. This is in contrast to most approaches that project

    the momentum which is constrained by the continuity equation. A predictor-

  • 8/8/2019 Jeff Thesis

    33/119

    2.2. Discretization 17

    corrector approach is used to solve the momentum and energy equation:

    pt+3

    2,k+1 = pt+

    3

    2,k + p. (2.24)

    The corrector step is the difference between the predictor equation and the

    exact equation which is defined as:

    gt+1,k+1i,cv = g

    i,cv t

    4

    p

    xi

    cv

    , (2.25)

    ut+1,k+1i,cv = u

    i,cv t

    4t+1cv

    p

    xi

    cv

    . (2.26)

    Substituting equation (2.26) into the nonlinear term uiui converts kinetic en-

    ergy into an equation for p which is:

    (uiui)t+1,k+1 =

    ui

    t

    4t+1cv

    p

    xi

    ui

    t

    4t+1cv

    p

    xi

    = ui u

    i t

    2t+1ui

    p

    xi+ O(p2). (2.27)

    Equation (2.27) is then substituted into equation (2.23), and yields a discrete

    energy equation in terms of p. Note that p is the difference between itera-

    tions, which implies that p converges to zero at each time-step. This allows

    the high order terms in equation (2.27) to be neglected.

    The implementation to solve the following discrete equations are to initial-

    ize the outer loop:

    t+3

    2 ,0 = t+1

    2 , ut+1,0i = uti, Tt+3

    2 ,0 = Tt+1

    2 , vt+1,0N = vtN, St+ 3

    2

    ,0

    k = St+ 1

    2k .

    (2.28)

  • 8/8/2019 Jeff Thesis

    34/119

    2.2. Discretization 18

    The next procedure is to advance the continuity equation (2.20), then advance

    the species equation (2.21). Once the species is advance, advance the momen-

    tum equation (2.22), then obtain vN by interpolation. After obtaining vN, the

    next step is to solve the pressure correction equation (2.23). Use the corrector

    steps to update pressure (2.24), momentum (2.25) and the velocities (2.26),

    then check the convergence for the pressure, momentum, density, and species

    between outer loop iterations.

    2.2.1 Implicit source term

    Consider a chemical system ofN species reacting through M reactions denoted

    as (Poinsot & Veynante [42]):

    Nk=1

    kjk Nk=1

    kjk (2.29)

    for j = 1, M where k is a symbol for species k.

    kj and

    kjare the stoichio-

    metric coefficients of species k for j reactions. The reaction term is defined

    as:

    k = Wk

    Mj=1

    kjQj (2.30)

    where

    Qj = Kfj

    Nk=1

    YkWk

    kj

    forward reaction

    KrjNk=1

    YkWk

    kj

    reverse reaction

    . (2.31)

  • 8/8/2019 Jeff Thesis

    35/119

    2.2. Discretization 19

    Using the empirical Arrhenius law,

    Kfj = AfjTj exp

    Ej

    RT

    = AijT

    j exp

    Taj

    T

    . (2.32)

    Therefore, the source term is:

    k = Wk

    Mj=1

    kj Kfj Nk=1

    SkWk

    kjWk

    Mj=1

    kj

    Krj

    Nk=1

    SkWk

    kj

    . (2.33)

    The backward rates Krj are computed through the equilibrium constants:

    Krj =Kfj

    paRT Nk=1 expS0jR H0jRT (2.34)

    where S0j and H0j are entropy and enthalpy, respectively. The discrete form

    of equation B.11 is:

    t+1k = Wk

    Mj=1

    kj

    Kt+1fj,cv

    Nk=1

    St+ 3

    2

    k,cv + St+ 1

    2

    k,cv

    2Wk

    kj

    WkMj=1

    kj

    Kt+1rj,cv Nk=1

    St+ 32k,cv + St+ 12k,cv2Wk

    kj . (2.35)

    Substituting equation (2.35) into equation (2.21) for t+1k yields:

    (Sk)t+ 3

    2

    cv (Sk)t+1

    2

    cv

    t+

    1

    V

    faces

    (Sk)t+1faces V

    t+1N Aface

    = 1ReSck

    1V

    faces

    1

    Skxj

    Sk2

    xj

    t+1faces

    NjAface

  • 8/8/2019 Jeff Thesis

    36/119

    2.2. Discretization 20

    +Wk

    Mj=1

    kj

    Kt+1fj,cv N

    k=1

    St+ 32k,cv + St+ 12k,cv

    2Wk

    kj

    WkMj=1

    kj

    Kt+1rj,cv N

    k=1

    St+ 32k,cv + St+ 12k,cv

    2Wk

    kj

    . (2.36)

    Equation (2.36) can be represented as:

    apSt+ 3

    2

    cv +nb

    anbSt+ 3

    2

    nb = RHS. (2.37)

    where nb are the neighbors of the cv. A parallel, algebraic multi-grid approach

    is used to solve the system of algebraic equations. The structured grid interface

    of the Hypre library (Lawrence Livermore National Laboratory 2003) is used.

    Three different examples of the implicit source term are described below.

    If N = 1, M = 1 and only forward reaction is assumed, this implies that

    Q1 = A11T1 exp

    Ta1

    T

    Y1W1

    11

    (2.38)

    which implies

    1 = W1A1111T1 exp

    Ta1

    T

    Y1W1

    11. (2.39)

    Note that for a one-step premixed flame:

    = BY exp

    TaTd

    (2.40)

  • 8/8/2019 Jeff Thesis

    37/119

    2.2. Discretization 21

    which upon discretization yields:

    t+1 = BSt+ 3

    2

    cv + St+ 1

    2

    cv

    2exp

    TaTt+1cv

    . (2.41)

    Equation (2.41) is the source term used in section 3.1.1.

    Consider a two-step reaction from Chen [4] and Mahalingam [29] given by:

    A + B I

    A + I P. (2.42)

    Note that A is the fuel, B is the oxidizer, I is the intermediate step, and P is

    the product. The source term for species A is defined as:

    k = B1YAYB exp

    Ta1

    T

    + B2YAYI exp

    Ta2

    T

    . (2.43)

    The discrete form is:

    k = B1St+ 3

    2

    A,cv + St+ 1

    2

    A,cv

    2St+1B,cv exp

    Ta1

    Tt+1cv

    +B2

    St+ 3

    2

    A,cv + St+ 1

    2

    A,cv

    2 St+1

    I,cv exp Ta2Tt+1cv . (2.44)This implicit source term is used in section 3.1.4. Note that the other species

    are handled in a similar manner.

    The final example is a nonlinear source term which is used in 3.1.2 and

    3.1.3:

    k = AF+OYFF YOO expTaT . (2.45)

  • 8/8/2019 Jeff Thesis

    38/119

    2.2. Discretization 22

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-110

    -1

    100

    101

    102

    103

    (a) Mr

    CF L

    10-7

    10-6

    10-5

    10-4

    10-3

    10-2

    10-12

    3

    4

    5

    6

    7

    8

    9

    10

    (b) Mr

    iter

    Figure 2.2: Variation of (a) acoustic CFL and (b) number of outer loop itera-tions with Mr. Taylor problem , Reacting problem .

    Let F = 2 and O = 1. This implies:

    k = A2+1Y2FY1O expTa

    T

    = AS2FS1O exp

    Ta

    T

    (2.46)

    where F is the fuel and O is the oxidizer. The discrete equation for the fuel

    species is:

    t+1

    k

    =

    ASt+ 3

    2

    F,cv + St+ 1

    2

    F,cv

    2 (1)

    St+1

    F,cv(2)

    St+1

    O,cv

    expTa

    Tt+1cv . (2.47)

    The nonlinear source term is linearized by only solving for term (1) and term

    (2) is from previous iterations. The outer loop ensures that the nonlinear

    corrections converge to zero.

  • 8/8/2019 Jeff Thesis

    39/119

    2.2. Discretization 23

    2.2.2 Zero Mach number limit

    The governing equations analytically reduce to the zero Mach number equa-

    tions, as shown in section 2.1. Since the dependent variables are spatially

    colocated, it appears that the pressureprojection step might result in odd

    even decoupling. It is well known that the incompressible equations require

    staggering, temporal dissipation, or low order basis functions for pressure, to

    avoid oddeven decoupling. However, note that the inner loop in the proposed

    algorithm uses nearest neighbors for the pressure equation; it therefore does

    not suffer from oddeven decoupling. The tolerance assigned to the outer loop

    can in practice, be controlled to obtain low Mach number solutions. However,

    a more reliable approach which explicitly introduces temporal biasing is that

    proposed by [53], which is now used here.Recall that the pressure gradient term in the momentum equation is defined

    as:

    p

    xi

    t+ 12

    =1

    4

    p

    xi

    t 12

    +1

    2

    p

    xi

    t+ 12

    +1

    4

    p

    xi

    t+ 32

    . (2.48)

    Temporal biasing can be introduced by computing the pressure gradient as:

    p

    xi

    t+ 12

    =

    1

    4

    p

    xi

    t 12

    +1

    2

    p

    xi

    t+ 12

    +

    1

    4+

    p

    xi

    t+ 32

    . (2.49)

    Taylor expansion of the pressure at time level t + 12

    yields:

    1

    4

    p

    xi

    t 12

    +1

    2

    p

    xi

    t+ 12

    +

    1

    4+

    p

    xi

    t+ 32

    =

    xipt+ 12 + 2tpt

    t+ 12 + O t2 . (2.50)

  • 8/8/2019 Jeff Thesis

    40/119

    2.2. Discretization 24

    i.e.; the leading order term is first order in time, for a given in equation 2.50.

    Alternate biased formulations of the pressure gradient can be used, but as the

    Taylor series expansion shows, using of the order of t would be the most

    timeaccurate.

    Two examples are used to illustrate the low Mach number behavior of the

    algorithm. The first example is the nonreacting Taylor problem discussed in

    more detail in Hou & Mahesh [23], while the second example corresponds to

    a synthetic premixed flame for which an analytical solution is obtained: i.e.

    T = sin x + 2, Y = sin x + 2, u = sin x + 2, p =4

    3cos x sin x,

    = 1/T, = cos x sin x. (2.51)

    For the Taylor problem, t was set to 0.001 and for the reacting case t

    was 0.0001. Figures 2.2 (a) show the variation of acoustic CFL, and number

    of outer loop iterations with Mr respectively. Mr varies from 102 to 105 for

    the Taylor problem, and from 103 to 106 for the reacting case. Solutions

    were also obtained for Mr equal to zero, but are not shown due to the use of

    logarithmic axes. Here, was chosen to be 0.05. Note that the acoustic CFL

    varies inversely with Mr; i.e. low Mach numbers can be computed at fixed

    timestep. Also, note that the number of outer loop iterations at low Mach

    number is only twice that at finite Mach number, and remains constant at

    very low Mach numbers.

  • 8/8/2019 Jeff Thesis

    41/119

    Chapter 3

    Validation simulations

    The algorithm is applied to simulate flames using both simple (3.1) and com-

    plex (3.2) chemical mechanisms. Section 3.1.1 considers a onedimensional

    steady laminar premixed flame while an unsteady laminar diffusion flame is

    considered in section 3.1.2. These two examples illustrate the ability to han-

    dle large heat release at low Mach number. Section 3.1.3 considers a steady

    twodimensional laminar reacting jet flame with large heat release at low Mach

    number. Threedimensional numerical simulations of turbulent nonpremixed

    flames with finite rate chemistry are considered in section 3.1.4. This problem

    illustrates application to turbulence and finite Mach number. The final ex-

    amples in section 3.2 illustrates application to complex chemistry for H2/air

    combustion using a 9 species, 19 reaction mechanism from Mueller et al. [37].

    The wellstirred reactor problem in section 3.2.1 is used to study chemical

    stiffness. A grid convergence study of a laminar diffusion flame is shown in

    section 3.2.2. Section 3.2.3 and 3.2.4 discuss simulations of a laminar jet flame

    and a lifted jet flame, respectively.

    25

  • 8/8/2019 Jeff Thesis

    42/119

    3.1. Simple chemistry 26

    -10 -5 0 5 101

    2

    3

    4

    5

    (a) x

    T

    10-1

    100

    10-3

    10-2

    10-1

    (b) dx

    Error

    Figure 3.1: (a) Comparison of computed temperature to analytical solutionfor a laminar premixed flame. analytic solution, numerical solution.Sc = Re = P r = 1, = 10, = 0.8, Mr = 0.01, (b) spatial order ofaccuracy. Algorithm, second order, Sc = Re = P r = 1, = 0.5, = 2,Mr = 0.01.

    3.1 Simple chemistry

    3.1.1 Laminar premixed flame

    An irreversible, onestep, laminar premixed flame is considered as an example

    of a low Mach number reacting flow. The reaction is given by R P where

    R is the reactant and P is the product. The reaction source term is that

    proposed by Echekki & Ferziger [16]. It approximates the Arrhenius source

    term, and is useful for validation, in that it permits analytical solution. The

    reaction model is defined as:

    =

    0 for < c,

    ( 1)( 1) otherwise .(3.1)

    The corresponding analytical solution is given by:

  • 8/8/2019 Jeff Thesis

    43/119

    3.1. Simple chemistry 27

    =

    (1 1)exp(x) for x 0,1 1 exp [(1 ) x] for x 0 .

    (3.2)

    Generally, is the order of 10 for hydrocarbon combustion (Echekki & Ferziger

    [16]).

    Figure 3.1 (a) shows computed results for a premixed flame where = 10,

    = 0.8, and the inflow Mach number is equal to 0.01. The Echekki & Ferziger

    source term is used. The inflow velocity is set equal to the flame speed, and the

    solution is initialized using a hyperbolic tangent profile for the temperature

    distribution. The inlet boundary condition for temperature, density, species

    and velocity are set to one. Zero derivative boundary conditions are specified

    at the outflow. Boundary conditions based on characteristic analysis are not

    specified, due the simplicity of the solution at the boundaries, and the use of

    sponge boundary conditions at both inflow and outflow boundaries to absorb

    acoustic waves that are generated by the initial transient. A cooling term,

    (U Uref) is added to the right hand side of the governing equations overthe sponge zone, whose length is 10 percent of the domain. Here U and

    Uref denote the vector of conservative variables and the reference solution

    respectively. The coefficient is a polynomial function defined as:

    (x) = As(x xs)n

    (Lx xs)n , (3.3)

    where xs and Lx denotes the start of the sponge and the length of the domain.

    n and As are equal to three. After a brief transient, the solution settles into

    a steady state which is shown in figure 3.1 (a). The computed and analytical

    solutions seem to agree well.

  • 8/8/2019 Jeff Thesis

    44/119

  • 8/8/2019 Jeff Thesis

    45/119

    3.1. Simple chemistry 29

    one-step reaction defined as:

    FYF + OYO PYP (3.4)

    where YF, YO and YP are the mass fractions of the fuel, oxidizer and products.

    The computed solution is compared to an asymptotic solution developed by

    Cuenot & Poinsot [11]. Their analysis is applicable to diffusion flames with

    variable density, non-uniform Lewis number, and finite rate chemistry. Cuenot

    & Poinsot studied unsteady unstrained, steady strained and unsteady strained

    H2O2 flames. Case 2 from their paper corresponds to an unsteady, unstrainedflame, and was chosen for comparison. Table 3.1 lists the conditions for case

    2.

    Table 3.1: Test condition.

    case LeF TF,0 LeO TO,0 F O A Ta Q/cp Re2 1 300 K 1 300 K 8 2 1 108 3600 K 6000 K 10000

    Note that LeF and LeO denote the Lewis number for fuel and oxidizer.

    TF,0 and TO,0 are the initial temperatures of the fuel and oxidizer. is the

    equivalence ratio which is defined as:

    = sYF,0YO,0

    =OWOFWF

    YF,0YO,0

    (3.5)

    where s is the stoichiometric ratio. The equivalence ratio is used to determine if

    the mixture is rich ( > 1), lean ( < 1) or stoichiometric ( = 1). F and O

    denote the stoichiometric coefficients of the species. A is the preexponential

    factor and Ta is the activation temperature. The Arrhenius reaction rate is

  • 8/8/2019 Jeff Thesis

    46/119

    3.1. Simple chemistry 30

    defined by Cuenot & Poinsot as:

    k = kWkA

    YFWF

    F YOWO

    Oexp

    Ta

    T

    . (3.6)

    Figure 3.2 is a comparison of the computed solution to the asymptotic so-

    lution from Cuenot & Poinsot. The asymptotic solution was used to initialize

    the temperature, fuel mass fraction, and oxidizer mass fraction. The initial

    velocity profile was calculated from the energy equation. The initial pressure

    remains uniform in the flame and the initial density is obtained from tem-

    perature. Initially the reaction rate is zero because the asymptotic solution

    assumes infinitely fast chemistry. However, the simulation involves (fast) finite

    rate chemistry. Therefore, the reaction rate rapidly recovers in the simulation

    (Cuenot & Poinsot). The inlet boundary condition for temperature, density,

    and mass fraction of fuel are set to one. Mass fraction of oxidizer and velocity

    are set to zero. Zero derivative boundary conditions are specified at the out-

    flow. The initial time was set at ct/L = 20 and the solution was advanced to

    a time of ct/L = 64. Again, the sponge boundary conditions were used at

    both inflow and outflow boundaries to absorb acoustic waves generated by the

    initial transient. The sponge zone is 10 percent of the domain. Also, the flow

    Mach number is 0.001 and the grid used 512 points. Reasonable agreement

    between computed and asymptotic solution is obtained.

    3.1.3 Laminar twodimensional jet flame

    A steady two dimensional reacting laminar jet flame from Poinsot & Veynante

    [42] is considered. This problem illustrates the ability to handle large heat

    release and nearly incompressible flow for a one-step diffusion flame with a

  • 8/8/2019 Jeff Thesis

    47/119

    3.1. Simple chemistry 31

    (a) x

    y

    0 20 40 60 80 100 120 140

    0.2

    0.4

    0.6

    0.8

    1

    (b) x

    Figure 3.3: (a) Contour plot of temperature for a jet flame. (b) Center line pro-file of the mixture fraction . asymptotic solution from Poinsot & Veynante[11], numerical solution. Mr = 0.01, Re = 500, Sck = P r = 1.

    Damkohler number of 50 106

    . From Poinsot & Veynante [42], if one assumes

    constant mass flow rate (u = constant), v = w = 0 and D = constant, then

    the mixture fraction satisfies:

    Fuf

    x= FDF

    2

    x2.

    Note that is defined as:

    = 2

    YF YO + 1

    +

    1 + 1

    (3.7)

    where is the equivalence ratio. A similarity solution is obtained:

    (x, y) =1

    2

    xexp

    y

    2

    4x

    (3.8)

  • 8/8/2019 Jeff Thesis

    48/119

  • 8/8/2019 Jeff Thesis

    49/119

    3.1. Simple chemistry 33

    (a) (b)

    Figure 3.4: (a) Scalar dissipation rate and (b) reaction rate at z = 4for turbulent diffusion flame. Notice that scalar dissipation is high where thereaction rate is low. Mr = 0.1, Re = 935, Sck = P r = 1.

    (a) (b)

    Figure 3.5: (a) Isosurface plot of the reaction rate for turbulent diffusionflame. (b) Contour plot of temperature showing entire computational domain.

    Mr = 0.1, Re = 935, Sck = P r = 1.

  • 8/8/2019 Jeff Thesis

    50/119

  • 8/8/2019 Jeff Thesis

    51/119

    3.1. Simple chemistry 35

    Table 3.3: Parameters for diffusion flame.

    zst A B A (m3 mol1 s1)

    0.5 8 0.8 1 1 1012

    where k is the wave number, k0 is the wave number at which E(k) is maximum,

    and u0 is the rms velocity.

    The initial diffusion flame is obtained by asymptotic solution (Cuenot &

    Poinsot [11]); table 3 shows the relevant parameters. From table 3, zst is

    defined as zst =1

    1+where is the equivalence ratio. A and B are the sto-

    ichiometric coefficients. A is the preexponential factor. The diffusion flame

    is advanced in time to remove any acoustic transients. After the initial tran-

    sients are removed, the initial turbulent velocity field is superimposed on the

    diffusion flame. The solution is advanced from 0 to 2.0 eddy turnover times

    and analyzed at each 0.1 eddy turnover time. Local extinction is observed to

    occur in the reaction rate. Holes occur in the reaction zones where pure mix-

    ing exists and the hole location corresponds to a high rate of scalar dissipation

    rate as predicted by laminar flamelet theory. This is shown in figures 3.4 and

    3.5 (a). Figure 3.5 (b) is a contour plot of temperature showing the entire

    computational domain. Figure 3.6 shows two-dimensional mass concentration

    distribution of fuel mass fraction and oxidizer mass fraction where good agree-

    ment is obtained with Chen [3] (figure 2 (a) in their paper). The lower bound

    on the species is the location of the flame and the upper bound is influenced

    by extinction. The distribution between the lower and upper bounds is the

    turbulence induced mixing (Mahalingam [29] & Chen [3]).

  • 8/8/2019 Jeff Thesis

    52/119

    3.2. Complex chemistry 36

    (a) time (sec)

    YH2 T

    0 0.02 0.04 0.06 0.08 0.10

    1E-06

    2E-06

    3E-06

    4E-06

    (b) time (sec)

    YHO2

    Figure 3.7: (a) Comparison of a major species & temperature betweenChemkin and present algorithm for a wellstirred reactor. ([YH2 ] Algo-rithm, Chemkin & [T] in kelvin, Algorithm, Chemkin) (b) Illustra-tion of the effect of timestep by comparing the explicit Euler (dt = 1.0e 8)to present implicit source term (dt = 0.001 & dt = 0.0001). implicitdt = 1.0e 3, implicit dt = 1.0d 4 explicit dt = 1.0d 8.

    3.2 Complex chemistry

    This section illustrates application of the algorithm to complex chemistry.

    Combustion of H2 and air is considered, using a 9 species and 19 reaction

    mechanism [37].

    3.2.1 Well stirred reactor

    Figure 3.7 illustrates the ability of the implicit procedure described in section

    2.2 to handle chemical stiffness. A perfectlystirred reactor problem is consid-

    ered, the timestep is varied, and the results are compared to Chemkin [26]

    for accuracy. The implicit treatment of the chemical source terms allows the

    time step to be five orders of magnitude higher (1.0e3) than that possibleusing explicit Euler time advancement (1.0e8). The maximum error of the

  • 8/8/2019 Jeff Thesis

    53/119

    3.2. Complex chemistry 37

    Table 3.4: Parameters for laminar diffusion flame.

    YH2 YO2 YN2,F YN2,O TF TO p u Mr pr Tr0.029 0.233 0.971 0.767 1 4 0 0 0.4 1 atm 300 K

    implicit solution (computed using YHO2) compared to the explicit solution was

    around 5 percent, for dt = 1.0e

    3. For dt = 0.0001, the percent error was

    around 0.01 percent.

    3.2.2 Laminar diffusion flame

    A laminar diffusion flame with complex chemistry was simulated. The param-

    eters are listed in table 3.4. Figure 3.8 shows results from a grid convergence

    study. Uniform grids ranging from 64 to 2048 points were used to obtain

    a gridconverged solution. A nonuniform grid of 128 points that clustered

    points in the flame was also used. The resolution was estimated to accurately

    compute the instant when the flame was thinnest and the heat release was

    highest. Figures 3.8 (a) and (b) show that the solution is stable and qual-

    itatively correct even for the 64 points which only has 2 to 3 points in the

    flame.

    3.2.3 Laminar jet flame

    A three dimensional laminar hydrogen/air round jet flame in heated coflow is

    considered. The jet Reynolds number is 1000, fuel jet width is 1 cm and other

    relevant parameters are listed in table 3.5. The inflow boundary conditions

    are shown in table 4.1. The first derivative is set to zero for thermodynamic

    variables, species, and velocities at the outflow. At the y and z boundaries, the

  • 8/8/2019 Jeff Thesis

    54/119

    3.2. Complex chemistry 38

    (a) x

    YH2O

    (b) x

    YH

    Figure 3.8: Grid convergence study showing profiles of a major species anda minor species. (a) H2O (b) H. N=64, N=128 , N=128 nonuniform grid, N=256, N=512, N=1028, N=2048.

    Table 3.5: Simulation parameters

    YH2 YO2 YN2 T U pFuel 0.029 0 0.971 300 (K) 187 m/s 101325 PaOxidizer 0 0.233 0.767 1200 (K) 75 m/s 101325 Pa

    thermodynamic variables, velocities and species are set to a constant. Sponge

    boundary conditions are used at the transverse boundaries to absorb acoustic

    waves that are generated by the flame. The computational grid is 128 by 64

    by 64 and the domain is 30 by 10 by 10. P r is 0.7, is equal to T0.7 and unity

    Lewis number (Lek = 1) is assumed.

    Figure 3.9 shows the temporal variation of maximum n, T, YOH, YO, YH,

    YH2O, YHO2 and YH2O2 . A steady state solution is obtained at 170 units of

    time and the final maximum temperature is 2370 K. Figure 3.10 illustrates

    the steady state contours of T(K), YOH, YHO2 and YH2O2.

  • 8/8/2019 Jeff Thesis

    55/119

    3.2. Complex chemistry 39

    (a) time

    n T

    (b) time

    Figure 3.9: (a) time versus maximum n and T. (b) time versusmaximum species of YOH, YO, YH, YH2O, YHO2 and YH2O2.

    Table 3.6: Simulation parameters

    YH2 YO2 YN2 T U pFuel 0.11 0 0.89 400 (K) 347 m/s 101325 PaOxidizer 0 0.233 0.767 1100 (K) 100 m/s 101325 Pa

    3.2.4 Sandia lifted jet flame

    Experiments on a lifted turbulent hydrogen/air slot jet flame in a heated coflow

    were performed at Sandia National Laboratories. The jet Reynolds number

    is 11,000 and the simulation parameters are listed in table 3.6. The fuel jet

    width is 2 mm. We perform preliminary two dimensions simulations under

    the same conditions as the Sandia jet flame. The boundary conditions are

    similar to that described in section 3.2.3. The computational grid is 1024 by

    512 by 2 and the domain is 30 by 14 by 1. P r is 0.7, is equal to T0.7 and

    the nonuniform Lewis numbers are from Im et al. [24].

    Figure 3.11 (a) is a contour plot of temperature; note the temperature in-

  • 8/8/2019 Jeff Thesis

    56/119

  • 8/8/2019 Jeff Thesis

    57/119

    3.2. Complex chemistry 41

    (a) (b)

    (c) (d)

    Figure 3.11: Contour plots of (a) T(K), (b) YHO2, (c) z and (d) divergence.

    crease down stream of the jet as expected for a lifted jet. The precursor to

    autoignition is adequate accumulation of the species YHO2. Figure 3.11 (b)

    shows the species YHO2 exists down stream in the jet. Figure 3.11 (c) shows the

    rollup of vorticity as the jet evolves, while the acoustic waves generated by the

    combustion are illustrated in figure 3.11 (d). Note that the acoustic waves ap-

    pear to emanate from the ignition regions. Future simulations should consider

    threedimensionality, larger domain and nonreflecting boundary conditions

    to better represent the acoustic field.

  • 8/8/2019 Jeff Thesis

    58/119

    Chapter 4

    Autoignition of a hydrogen

    vortex ring reacting with hot air

    The objective of this chapter is to study autoignition of fuel injected into

    hot air. We therefore use direct numerical simulation to study the flow when

    a round nozzle injects H2 + N2 fuel at room temperature into air at higher

    temperatures. The amount of fuel injected is controlled by varying the stroke

    length or stroke ratio. The governing equations are the threedimensional, un-

    steady, compressible, NavierStokes equations along with conservation equa-

    tions for the nine species described by the Mueller et al. [37] chemical mecha-

    nism. The algorithm proposed in chapter 2 is used. The simulations consider

    the effects of Lewis number, stroke length ratio and fuel/air mixtures (equiv-

    alence ratio) on autoignition and the flow.

    Chapter 4 is organized as follows. Section 4.1 describes relevant details

    of the simulation. It provides a brief description of the initial and boundary

    conditions. Results from a validation and grid convergence study are shownin section 4.2. The homogeneous mixture problem is discussed in section 4.3

    42

  • 8/8/2019 Jeff Thesis

    59/119

    4.1. Problem statement 43

    Figure 4.1: Schematic of the problem.

    and is used to obtain a chemical time scale and MR. Section 4.4 discusses

    results for the reacting vortex ring. A brief summary in section 4.5 concludes

    this chapter.

    4.1 Problem statement

    Figure 4.1 shows a schematic of the problem and illustrates the ring vortex

    and the trailing column which forms at large stroke ratios. The dimensional

    variables for the reacting vortex ring are chosen such that the equivalence ratio

    () is one, and a stoichiometric mixture is obtained where YH2,0 equals 0.02936

    and YO2,0 equals 0.233. For fuel lean mixtures, equals 0.25 (YH2 = 0.0074)

    and for fuel rich mixtures equals 4 (YH2 = 0.1185). For all simulations, Re is

  • 8/8/2019 Jeff Thesis

    60/119

  • 8/8/2019 Jeff Thesis

    61/119

    4.2. Validation and gridconvergence study 45

    -3 -2.5 -2 -1.5 -1 -0.5 00

    0.2

    0.4

    0.6

    0.8

    1

    (a) r

    YF

    (b) time

    T YH2

    Figure 4.2: (a) Comparison using onestep chemistry to Hewett and Madnia( [19]). Present, Hewett & Madnia. (b) Comparison of a major species& temperature between Chemkin and present algorithm for a wellstirred re-actor using the Mueller mechanism. ([YH2] Algorithm, Chemkin & [T]

    Algorithm, Chemkin).

    the sponge zone whose length is 10 percent of the domain. Here U and Uref

    denote the vector of conservative variables and the reference solution respec-

    tively. The coefficient for the sponge at the outflow, is a polynomial function

    defined as:

    (x) = As(x xs)n

    (Lx xs)n,

    where xs and Lx denotes the start of the sponge and the length of the domain.

    n and As are equal to three.

    4.2 Validation and gridconvergence study

    The simulations are first validated using a onestep chemical mechanism, and

    compared to the one-step simulations of Hewett & Madnia [19]. Our results

  • 8/8/2019 Jeff Thesis

    62/119

    4.2. Validation and gridconvergence study 46

    (a) y

    YH

    (b) y

    YH2O

    Figure 4.3: Grid convergence study for a twodimensional laminar reactingvortex ring. (a) YH (b) YH2O. N=64, N=128, Nx=128, Ny=Nz=128nonuniform in y, Nx = 256, Ny=Nz=128 nonuniform in y, N =256, N = 512.

    are compared to case 5 in their paper where the temperature ratio is 6:1 and

    the stroke length is two. Figure 4.2 (a) shows that good agreement is obtained.

    Next, complex chemistry is validated using the homogeneous mixture or

    wellstirred reactor problem. We use the 9 species (H2, O2, OH, O, H, H2O,

    HO2, H2O2, and N2), 19 reaction Mueller mechanism (Mueller et al. [37])

    which is shown in table 4.2. A detailed experiment and simulation of a lam-

    inar hydrogen jet diffusion flame was performed by Cheng et al. [8]. Their

    work shows that chemical mechanisms from Miller & Bowman [35] , GRIMech

    3.0 [48], Mueller et al [37], Maas & Warnatz [28], and O Conaire et al. [38]

    compare well with experimental results. Therefore, the Mueller mechanism

    should perform adequately for our hydrogen/air reacting vortex ring simula-

    tion. Figure 4.2 (b) is a comparison between Chemkin [26] and our algorithm

    incorporating the Mueller mechanism. Good agreement is observed for the

    homogeneous mixture.

  • 8/8/2019 Jeff Thesis

    63/119

    4.2. Validation and gridconvergence study 47

    Table 4.2: Mueller Mechanism: (cm3moleseckcalK)

    Afj j EjH2/O2 Chain Reactions

    1 H+ O2 O + OH 1.910E14 0.0 16.442 O + H2 H+ OH 5.080E04 2.67 6.293 H2 + OH H2O + H 2.160E08 1.51 3.434 O + H2O OH + OH 2.970E06 2.02 13.4

    H2/O2 Dissociation/Recombination Reactions5 H2 + M H+ H+ M 4.580E19 1.40 104.386 O + O + M O2 + M 6.160E15 0.5 07 O + H+ M OH + M 4.710E18 1.0 08 H+ OH + M H2O + M 2.210E22 2.0 0

    Formation and Consumption of HO29 H+ O2 + M HO2 + M 3.50E16 0.41 1.1210 HO2 + H H2 + O2 1.660E13 0.0 0.8211 HO2 + H OH + OH 7.080E13 0.0 0.312 HO2 + O OH + O2 3.250E13 0.0 0

    13 HO2 + OH H2O + O2 2.890E13 0.0 0.50Formation and Consumption of H2O2

    14 HO2 + HO2 H2O2 + O2 4.200E14 0.0 11.9815 H2O2 + M OH + OH + M 1.20E17 0.0 45.516 H2O2 + H H2O + OH 2.410E13 0.0 3.9717 H2O2 + H H2 + HO2 4.820E13 0.0 7.9518 H2O2 + O OH + HO2 9.550E06 2.0 3.9719 H2O2 + OH H2O + HO2 1.000E12 0.0 0

    The three dimensional computational grid used in the simulations is based

    on a grid refinement study of a twodimensional H2/air reacting vortex ring.

    The computational domain was 10 by 10 in x and y where the nozzle is in the

    y direction. Uniform grids ranging from 64 to 512 points were used to obtain

    a gridconverged solution. Also, a nonuniform grid of 128 points in the y

    direction (to cluster points around the nozzle) and 128 uniform points in the

    x direction was used. Figures 4.3 (a) and (b) show profiles in the middle of

    the reacting vortex ring. Note that grid convergence is obtained for the non

  • 8/8/2019 Jeff Thesis

    64/119

    4.3. Homogeneous problem 48

    (a) time

    H

    (b) time

    Tn

    Figure 4.4: (a) Rates of reactions 1,2,3,9,10, and 11 for species H. (b) Temporalevolution of temperature with all reactions, and reactions 1, 2, 3, 9, 10 & 11.All reactions: (T) , (n) reactions 1, 2, 3, 9, 10, & 11: (T) ,(n) .

    uniform grid of 1282 and uniform 2562 and 5122 grids. To ensure adequate

    resolution for the three dimensional reacting vortex ring, a nonuniform 128

    grid in y, z directions and 256 in the x direction was chosen, and the size of the

    computational domain was 10 by 10 by 10 in the three coordinate directions.

    4.3 Homogeneous problem

    We first consider the homogeneous mixture or wellstirred reactor problem.

    Our objectives are to: (i) obtain an ignition time scale; (ii) obtain an opti-

    mum mixture fraction for autoignition (Mastorakos et al. [32]); (iii) identify

    which reactions contribute to autoignition and which reactions contribute af-

    ter autoignition. Echekki & Chen [15] used DNS to study autoignition of

    a hydrogen/air mixture in two dimensional turbulence. They found that theintermediate species O, H, OH, and HO2 play an important role in the auto

  • 8/8/2019 Jeff Thesis

    65/119

    4.3. Homogeneous problem 49

    (a)

    chem

    (b)

    chem

    Figure 4.5: (a) Mixture fraction versus chem for varying temperature in kelvin. = 1, T = 1100, = 1, T = 1200, = 1, T = 1300, and = 1, T = 1400. (b) Mixture fraction versus chem for varying fuel/air

    ratio, = 1/4, T = 1200, = 1, T = 1200, = 4, T = 1200and = 8, T = 1200.

    ignition process. In the Mueller mechanism, reactions 1,2,3,9,10 and 11 have

    a significant contribution to the minor species O, H, OH, and HO2. Figure

    4.4 (a) shows reaction rates of species H due to reactions 1,2,3,9,10 and 11.

    Figure 4.4 (b) shows temperature and heat release when all the reactions are

    simulated, and when only the autoignition reaction (1,2,3,9,10, and 11) are

    simulated. Note that the maximum heat release for the full reaction set andthe autoignition reaction set occurs at about the same time. Up until this

    instant of maximum heat release, the autoignition reactions are dominant;

    after this instant, the other reactions are also significant. This instant is de-

    fined as chem and will be used to obtain the most reactive mixture fraction,

    MR.

    Mastorakos et al. [32] used onestep chemistry and the homogeneous mix-

    ture problem to obtain chem and the mixture fraction most likely to auto

  • 8/8/2019 Jeff Thesis

    66/119

    4.3. Homogeneous problem 50

    ignite (MR). The following equations are used to set up the wellstirred

    reactor problem in terms of the mixture fraction defined as:

    YH2 = YH2,0, YO2 = YO2,0 (1 ) , YN2 = 1 YH2 YO2, (4.3)

    T = TO (TO TF) .

    For example, if = 0.5, YH2,0 = 0.029, YO2,0 = 0.233, TF = 300 (K) and

    TO = 1200 (K), then YH2 , YO2, YN2 , & T would be 0.5, 0.0145, 0.1165, 0.869, &

    750 (K), respectively. One can now consider different initial mixture fractions,

    obtain the minimum ignition time (chem) and thereby obtain the most reactive

    mixture fraction (MR).

    We obtain MR for the Mueller mechanism. Figure 4.5 (a) shows mixture

    fraction versus chem for initial temperatures that range from 1100 to 1400

    Kelvin and figure 4.5 (b) shows mixture fraction versus chem for fuel/air ratios

    () that range from 0.25 to 8. Note in figure 4.5 that the minimum chem

    corresponds to MR. Figure 4.5 also shows the effect of fuel/air mixture and

    temperature on chem. As temperature increases, chem becomes smaller and

    MR shifts to the right. As increases, chem becomes smaller and MR shifts

    to the left. These results illustrate the balance between adequate fuel and

    high enough temperature for autoignition to occur. Table 4.3 lists the most

    reactive mixture fraction and corresponding values of the initial temperature,

    fuel/air ratio and chem.

  • 8/8/2019 Jeff Thesis

    67/119

    4.4. Results for reacting vortex ring 51

    Table 4.3: Global results for homogeneous mixture problem.

    T0 MR chem TMR1100 1 0.06 3.4e 4 10521200 1 0.08 1.4e 4 11281300 1 0.09 8.5e 5 12101400 1 0.10 5.6e 5 12901800 1 0.13 1.8e 5 16051200 1/4 0.1 3.5e

    4 1110

    1200 4 0.06 8.0e 5 11461200 8 0.04 6.4e 5 1164

    4.4 Results for reacting vortex ring

    Table 4.4 lists the different cases studied. The simulations consider the effect

    of Lewis number, stroke length ratio (L/D), fuel/air ratio and oxidizer tem-

    perature. The nonuniform Lewis numbers are obtained from Im et al [24].

    Two stroke ratios of 2 and 6 are considered. Note that for nonreacting vortex

    rings, a stroke ratio of 2 yields a coherent vortex ring, while a stroke ratio of 6

    yields a vortex ring followed by a trailing column of vorticity (Sau & Mahesh

    [46]). Oxidizer temperatures (TO) of 1200 and 1800 Kelvin are considered. Fi-

    nally, fuel/air ratios were chosen such that the equivalence ratios () are equal

    to 1/4, 1 and 4 yielding lean, stoichiometic and rich mixtures, respectively.

    Table 4.4: Parameters used in reacting vortex ring simulations.

    Lewis number (L/D) TOnonuniform 1 2 1200unity 1 2 1200nonuniform 0.25 2 1200nonuniform 4 2 1200nonuniform 1 2 1800

    nonuniform 1 6 1200

  • 8/8/2019 Jeff Thesis

    68/119

    4.4. Results for reacting vortex ring 52

    time

    YH

    Figure 4.6: Domain average of rates of reactions 1,2,3,9,10 and 11 for thespecies H. nonuniform Lewis reaction 1, reaction 2, reaction3, reaction 9, reaction 10, reaction 11.

    4.4.1 Ignition and flame evolution

    We first consider the nonuniform Lewis number simulations at stroke ratio

    of 2 and oxidizer temperature of 1200 K. Figure 4.6 shows reactions 1,2,3,9,10

    and 11 for the species H. Time is normalized by chem obtained from table

    4.3. Comparison to a corresponding plot for the homogeneous mixture (figure

    4.4 a) shows qualitative similarities along with some differences. The reaction

    rates are seen to peak at time greater than chem for the reacting vortex ring,

    reflecting the finite amount of time taken for fuel and oxidizer to be advected

    and then diffuse to yield regions of MR. Also, the relative importance of

    reactions 9 and 11 is higher for the vortex ring,. The time scale on which

    the reaction rates decay to zero following the initial ignition, is longer for the

    reacting vortex ring, reflecting the propagation of the flame through the vortex

    ring.

  • 8/8/2019 Jeff Thesis

    69/119

    4.4. Results for reacting vortex ring 53

    n

    (a)

    T

    (b)

    n

    T

    (c)

    n

    T

    (d)

    n

    T

    (e)

    n

    T

    (f)

    n

    T

    Figure 4.7: Twodimensional slices showing variation in time of versus n.The color contours correspond to n. On each color contour, a colored contourline is plotted for , T and . The color contour of n ranges from 0 to 1.For the contour lines, purple is (contour lines range from 0.04 to 0.12), T isblack (contour lines range from 1120 to 1140 K) and is red (contour linesrange from 0.0 to 0.15). Nondimensional time: (a) 2.5, (b) 3.0, (c) 3.5, (d)4.0, (e) 4.5 and (f) 5.0.

  • 8/8/2019 Jeff Thesis

    70/119

    4.4. Results for reacting vortex ring 54

    TMR

    MR

    LD

    = 6

    TO = 1800

    t : 0.25 0.75 1.25 1.75 2.25 2.75 3.25

    TMR

    MR

    t : 2.5 3.0 3.5 4.0 4.5 5.0 5.5

    TMR

    MR

    LD

    = 2

    t : 2.5 3.0 3.5 4.0 4.5 5.0 5.5

    Figure 4.8: Time evolution of heat release (n), MR and TMR. On each colorcontour of heat release, a contour line is superimposed for and T. Thecontour of n ranges from 0 to 1. The contour lines of range from 0.04 to0.12 (MR) and temperature range from 1092 to 1164 K (TMR) for L/D = 2and L/D = 6. For oxidizer temperature of 1800 K, ranges from 0.10 to 0.18(MR) and temperature ranges from 1590 to 1650 K (TMR).

  • 8/8/2019 Jeff Thesis

    71/119

    4.4. Results for reacting vortex ring 55

    Figure 4.7 illustrates the process of autoignition in the vortex ring using

    scatter plots of mixture fraction versus heat release at different instances in

    time. At each time instant, heat release contours in a radial crosssection are

    shown. Superposed on the heat release contours is a colored contour line of

    mixture fraction (), temperature (T) and scalar dissipation (), respectively.

    Note that the color contours of heat release (n) range from 0 to 1. Mixture

    fraction and scalar dissipation are defined as (Poinsot and Veynante [?]):

    =

    1

    1 +

    YFY0F

    YOY0O

    + 1

    , (4.4)

    =2

    Re

    xj

    xj

    , (4.5)

    respectively. For the contour lines, mixture fraction is purple (contour lines

    range from 0.04 to 0.12), temperature is black (contour lines range from 1120

    to 1140 K) and scalar dissipation is red (contour lines vary over the entire range

    of 0.0 to 0.15). Note that the ranges for the contour lines of mixture fraction

    and temperature were chosen based on the results of the homogeneous mixture

    problem. Recall MR is 0.08 from the homogeneous problem when T0 = 1200

    K and = 1 (table 4.3).

    Figure 4.7 shows that autoignition occurs in regions of high temperature,

    fuel-lean mixtures, and low scalar dissipation where = MR. Ignition occurs

    at a MR of 0.08 and propagates towards the stoichiometric side ( = 0.5).

    Most of the ignition occurs behind the vortex ring and then moves rapidly

    into the core. The contour line of mixture fraction shows that autoignition

    occurs in fuel lean regions ( = MR = 0.08). Also, the contour lines of scalar

    dissipation show that autoignition occurs in regions of low scalar dissipation.

    The superposed contour lines of temperature are chosen to lie around TMR

  • 8/8/2019 Jeff Thesis

    72/119

    4.4. Results for reacting vortex ring 56

    which is 1130 K (TMR = 1200 MR [1200 300]). Interestingly, maximumheat release follows TMR at every instant of time in figure 4.7. Figure 4.8 is a

    side by side comparison of MR and TMR. This figure shows that the onset of

    ignition occurs at MR and TMR. As ignition continues from behind the vortex

    ring into the core, MR is in the very fuel lean region and does not follow the

    path of ignition. TMR on the other hand, follows the path of ignition all the

    way into the core. This suggests that TMR instead of MR might be better

    in predicting the evolution of autoignition. This behavior is observed in all

    cases simulated (figure 4.8); i.e. in the preignition phase, MR predicts the

    onset of ignition fairly well, however TMR predicts both the onset of ignition

    as well as its subsequent evolution.

    Table 4.5 lists most reactive mixture fraction, burnout time in seconds,

    maximum temperature in Kelvin and ignition time in seconds obtained from

    all the simulations. Note that the onset of ignition largely depends on fuel/air

    ratio and oxidizer temperature. However, in the postignition phase, the evo-

    lution of the flame is greatly affected by other parameters. This behavior is

    discussed in the follow