Dissertation Alireza

download Dissertation Alireza

of 248

Transcript of Dissertation Alireza

  • 8/18/2019 Dissertation Alireza

    1/248

    Large Eddy Simulation of Sound Generation by Turbulent Reacting andNonreacting Shear Flows

    Alireza Najafi-Yazdi

    Doctor of Philosophy

    Department of Mechanical Engineering

    McGill University

    Montreal,Quebec

    December 2011

    A dissertation submitted to McGill University in partial fullfilment of therequirements for the degree of Doctor of Philosophy

    Copyright   c

    2011 by Alireza Najafi-Yazdi

  • 8/18/2019 Dissertation Alireza

    2/248

    To my mother, Sharhzad

    ii

  • 8/18/2019 Dissertation Alireza

    3/248

    ABSTRACT

    The objective of the present study was to investigate the mechanisms of sound

    generation by subsonic jets. Large eddy simulations were performed along with

    bandpass filtering of the flow and sound in order to gain further insight into the role

    of coherent structures in subsonic jet noise generation.

    A sixth-order compact scheme was used for spatial discretization of the fully

    compressible Navier-Stokes equations. Time integration was performed through

    the use of the standard fourth-order, explicit Runge-Kutta scheme. An implicit

    low dispersion, low dissipation Runge-Kutta (ILDDRK) method was developed and

    implemented for simulations involving sources of stiffness such as flows near solid

    boundaries, or combustion. A surface integral acoustic analogy formulation, called

    Formulation 1C, was developed for farfield sound pressure calculations. Formulation

    1C was derived based on the convective wave equation in order to take into account

    the presence of a mean flow. The formulation was derived to be easy to implement

    as a numerical post-processing tool for CFD codes.

    Sound radiation from an unheated, Mach 0.9 jet at  ReD  = 400, 000 was consid-

    ered. The effect of mesh size on the accuracy of the nearfield flow and farfield sound

    results was studied. It was observed that insufficient grid resolution in the shear layer

    results in unphysical laminar vortex pairing, and increased sound pressure levels inthe farfield. Careful examination of the bandpass filtered pressure field suggested

    iii

  • 8/18/2019 Dissertation Alireza

    4/248

    iv

    that there are two mechanisms of sound radiation in unheated subsonic jets that

    can occur in all scales of turbulence. The first mechanism is the stretching and the

    distortion of coherent vortical structures, especially close to the termination of the

    potential core. As eddies are bent or stretched, a portion of their kinetic energy is

    radiated. This mechanism is quadrupolar in nature, and is responsible for strong

    sound radiation at aft angles. The second sound generation mechanism appears to

    be associated with the transverse vibration of the shear-layer interface within the

    ambient quiescent flow, and has dipolar characteristics. This mechanism is believed

    to be responsible for sound radiation along the sideline directions.

    Jet noise suppression through the use of microjets was studied. The microjet

    injection induced secondary instabilities in the shear layer which triggered the tran-

    sition to turbulence, and suppressed laminar vortex pairing. This in turn resulted

    in a reduction of OASPL at almost all observer locations. In all cases, the bandpass

    filtering of the nearfield flow and the associated sound provides revealing details of 

    the sound radiation process. The results suggest that circumferential modes are sig-

    nificant and need to be included in future wavepacket models for jet noise prediction.

    Numerical simulations of sound radiation from nonpremixed flames were also

    performed. The simulations featured the solution of the fully compressible Navier-

    Stokes equations. Therefore, sound generation and radiation were directly captured

    in the simulations. A thickened flamelet model was proposed for nonpremixed flames.

    The model yields artificially thickened flames which can be better resolved on the

    computational grid, while retaining the physically currect values of the total heat

  • 8/18/2019 Dissertation Alireza

    5/248

    v

    released into the flow. Combustion noise has monopolar characteristics for low fre-

    quencies. For high frequencies, the sound field is no longer omni-directional. Major

    sources of sound appear to be located in the jet shear layer within one potential core

    length from the jet nozzle.

  • 8/18/2019 Dissertation Alireza

    6/248

    Résumé

    L’objectif de cette étude est d’obtenir la meilleure compréhension des mécanismes

    de géneration de bruit par des jet subsoniques. Cette étude est basée sur simulations

    aux grandes échelles de jets réactifs et sans réactifs.

    Des calculs numériques employant des schéme compacts de sixiéme ordre. L’integration

    temporelle fut éxéciteé à l’aide de schéme Runge-Kutta de de quatrième ordre. Des

    schéme à faible dispersion et dissipation numérique. Un formulation intégrale basée

    sur les analogies acoustiques fut développées pour la prédiction du champ acous-

    tique lointain pour les sources et observateure en mouvement dans un fluide avec

    vitesse uniforme. La formulation fut implémentée à l’aide d’algorithmes facilitant

    l’implémentation pour le traitement de données d’écoulement en haute performance

    utilisant des outils de simiulation á grande échelle.

    Les champs sonore produit par un jet turbulent non-réactif avec nombre de Mach

    de 0.9, et un nombre de Reynolds   ReD   = 400, 000 fut étudié. L’effect de la taille

    du maillage sur la précision de l’écoulement en champs proche et e champs sonore

    loin de source fut analysé. La sous-résolution de la couche decisaillement à la sortie

    du jet méne à l’apparition de structures cohérentes et forte radiation qui no sort pas

    physiquement réalistes. Deux mécanismes principaux de génération sonore par jets

    subsoniques furent identifiés.

    vi

  • 8/18/2019 Dissertation Alireza

    7/248

    vii

    Le premier mécanisme est l’étirement et la distorsion de structures tourbillon-

    naires cohérentes, en particulier prés de la fin du coere potentiel. Ce mécanisme

    est quadripolaire, et émet principalement vers l’arriére du jet dans la direction de

    l’écoulement. Le seconde mécanisme semble être constitué de vibration transversale

    de la couche de cisaillement en réponse á la présemce de structures cohérentes dans

    la jet. Semblable à la radiation d’une plaque à bonds finis, la contribution de ce

    méchanisme est dipolaire et domine la champs sonore dans la direction transversale,

    perpendiculaire au jet.

    L’utilisation de plusieurs microjet fut investiguée pour la réduction du bruit.

    L’injection à l’aide de microjets précipite la transition à la turbulence, favorisent le

    mélange et la destrcutction de structures cohérentes de grande échelle.

    Un filtrage en bandes de étroites fut effectué. Ce traitement des données numérique

    permet de visualiser les relations complexes entre l’écoulement et les onds sonores

    émises. Les résultats démontrent l’importance de modes circumférenciele, ce qui a

    des conśequenecs pour les modiles dits de paquets d’onde pour la preédiction du

    bruit du jet.

    Des simulation numériques d’écoulement et champs sonore d’une flame sans pré-

    mélange furent aussi éxécutées. Les simulations incluent encore une fois l’écoulement

    et le champ sonore associé, obtenus directement des équations de Navier Stokes com-

    pressibles. Un modèle flammelette épaissie fut proposé que donne flammes épaissies

    artificiellement qui peuvent être mieux résolus sur le maillage. Le bruit de combus-

    tion a des caractéristiques monopolaires aux basses fréquences. Principales sources

    de bruit semblent être situé dans la couche de cisaillement.

  • 8/18/2019 Dissertation Alireza

    8/248

    Acknowledgements

    The undertaking of the work presented in this thesis would have not been pos-

    sible without the support, guidance, and encouragement of my advisor, Prof. Luc

    Mongeau. His patience, diligence, willingness, and enthusiasm to explore new ideas

    made my doctoral studies a wonderful experience. His being a mentor passes way

    beyond daily research matters.

    I am also thankful to Prof. Stephane Moreau and Dr. Marlene Sanjose for not

    only our academic discussions, but also for their friendship. Sincere thanks also go

    to Prof. Jeffrey Berghtorson. A great teacher and a true mentor, he has always been

    generous with his time for me. I am also thankful to Prof. Siva Nadarajah who

    kindly served on my thesis advisory committee.

    Special thanks go to Prof. Thierry Poinsot, and Dr. Benedicte Cuenot who

    made my visit to CERFACS possible. They were very generous with their time for

    my various questions on turbulent combustion modeling. Their expertises and kind-ness to share their knowledge helped me to develop a better understanding in the

    field of turbulent combustion.

    I am also grateful to Dr. Ali Uzun, Dr. Christophe Bogey, and Prof. Christophe

    Bailly for sharing their jet results and for their useful comments.

    I would like to thank my friends and colleagues, Dr. Phoi-Tack (Charlie) Lew,

    Kaveh Habibi, Hani Bakhshaei, and others who provided a fun and stimulating envi-

    ronment during the course of my studies at McGill. I am also grateful to my family

    for their unconditional love and support.

    viii

  • 8/18/2019 Dissertation Alireza

    9/248

    Financial support from McGill University, through the McGill Engineering Doc-

    toral Award (MEDA) and Lorne Trottier Fellowship, is gratefully acknowledged.

    My stay at CERFACS was made possible through the financial support of the EC-

    COMET program and the Marie Curie Fellowship. Financial support from the Exa

    Corporation, Green Aviation Research & Development Network (GARDN), Pratt

    & Whitney Canada, and the National Science and Engineering Research Council

    (NSERC) of Canada is also gratefully acknowledged.

    The computational resources were provided by Compute/Calcul Canada through

    the CLUMEQ and the RQCHP High Performance Computing Consortia.

    ix

  • 8/18/2019 Dissertation Alireza

    10/248

    TABLE OF CONTENTS

    Acknowledgements   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . viii

    1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

    1.1 Motivation   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31.2 Background on Unheated Jet Noise   . . . . . . . . . . . . . . . . . 41.3 Background on Reacting Jet Noise   . . . . . . . . . . . . . . . . . 11

    1.4 Objectives   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141.5 Organization of the Thesis   . . . . . . . . . . . . . . . . . . . . . . 151.6 Contributions   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

    I Nonreacting Flows   20

    2 Governing Equations and Numerical Methods   . . . . . . . . . . . . . . . 21

    2.1 Governing Equations   . . . . . . . . . . . . . . . . . . . . . . . . . 222.2 Approximate Deconvolution Model   . . . . . . . . . . . . . . . . . 252.3 Spatial Discretization Scheme   . . . . . . . . . . . . . . . . . . . . 272.4 Spatial Filtering  . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282.5 Temporal Integration Scheme   . . . . . . . . . . . . . . . . . . . . 302.6 Nonreflective Boundary Conditions   . . . . . . . . . . . . . . . . . 312.7 Sponge Zone  . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322.8 Inflow Forcing   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322.9 LES Code Parallelization   . . . . . . . . . . . . . . . . . . . . . . . 34

    3 Convective Ffowcs Williams-Hawkings Equation: Formulation 1C   . . . . 36

    3.1 Originial Ffowcs Williams - Hawkings Acoustic Analogy . . . . . . 363.2 Convective FW-H Equation   . . . . . . . . . . . . . . . . . . . . . 39

    3.3 Formulation 1C   . . . . . . . . . . . . . . . . . . . . . . . . . . . . 453.3.1 Thickness Noise   . . . . . . . . . . . . . . . . . . . . . . . . 453.3.2 Loading Noise   . . . . . . . . . . . . . . . . . . . . . . . . . 49

    x

  • 8/18/2019 Dissertation Alireza

    11/248

    3.3.3 The Special Case of “Wind-tunnel”   . . . . . . . . . . . . . 513.4 Numerical Implementation and Verification  . . . . . . . . . . . . . 52

    3.4.1 Stationary monopole in a moving medium . . . . . . . . . . 52

    3.4.2 Stationary dipole in a moving medium . . . . . . . . . . . . 543.4.3 Rotating monopole in a moving medium . . . . . . . . . . . 55

    4 Large Eddy Simulations of an Isothermal High Speed Subsonic Jet . . . . 60

    4.1 Computational Setup   . . . . . . . . . . . . . . . . . . . . . . . . . 614.2 Computational Grid Setup   . . . . . . . . . . . . . . . . . . . . . . 62

    4.2.1 Grid-30   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624.2.2 Grid-88   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624.2.3 Grid-380   . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634.2.4 Grid Resolution and Subgrid Scales   . . . . . . . . . . . . . 63

    4.3 Nearfield Results   . . . . . . . . . . . . . . . . . . . . . . . . . . . 664.4 Farfield Sound Predictions   . . . . . . . . . . . . . . . . . . . . . . 68

    5 Sound Generation by Subsonic Jets: A Band-Pass Filtering Study   . . . . 82

    5.1 Band-Pass Filtering Procedure  . . . . . . . . . . . . . . . . . . . . 825.2 Band-Pass Filtering of Pressure Field   . . . . . . . . . . . . . . . . 835.3 Sound Generation Mechanism in Subsonic Jets  . . . . . . . . . . . 86

    6 Large Eddy Simulation of Jet Noise Suppression by Impinging Microjets   94

    6.1 Computational Setup   . . . . . . . . . . . . . . . . . . . . . . . . . 95

    6.2 Nearfield Results   . . . . . . . . . . . . . . . . . . . . . . . . . . . 966.3 Farfield Acoustics   . . . . . . . . . . . . . . . . . . . . . . . . . . . 986.4 Bandpass Filter Visualization of Acoustic Nearfield   . . . . . . . . 100

    II Reacting Flows   114

    7 Reacting Flow Simulations   . . . . . . . . . . . . . . . . . . . . . . . . . . 115

    7.1 Governing Equations for Reacting Flows   . . . . . . . . . . . . . . 1167.2 Simplifying assumptions   . . . . . . . . . . . . . . . . . . . . . . . 118

    7.2.1 Diffusion Fluxes   . . . . . . . . . . . . . . . . . . . . . . . . 118

    7.2.2 Unit Lewis Number   . . . . . . . . . . . . . . . . . . . . . . 1197.2.3 Buoyancy effects   . . . . . . . . . . . . . . . . . . . . . . . . 119

    7.3 Mixture Fraction   . . . . . . . . . . . . . . . . . . . . . . . . . . . 119

    xi

  • 8/18/2019 Dissertation Alireza

    12/248

    7.4 Flamelet Modeling   . . . . . . . . . . . . . . . . . . . . . . . . . . 1207.5 Flamelet/Progress Variable Modeling   . . . . . . . . . . . . . . . . 1227.6 The flamelet code   . . . . . . . . . . . . . . . . . . . . . . . . . . . 123

    7.7 Flamelet Modeling for Compressible Flows   . . . . . . . . . . . . . 1237.8 Thickened Flamelet Model   . . . . . . . . . . . . . . . . . . . . . . 1247.9 Thickened Flamelet vs. PDF modeling   . . . . . . . . . . . . . . . 1267.10 Sound Generation by Nonpremixed Flames   . . . . . . . . . . . . . 1287.11 Sound Generation by a Reacting Mixing Layer   . . . . . . . . . . . 130

    7.11.1 Computational Setup   . . . . . . . . . . . . . . . . . . . . . 1307.11.2 Numerical results   . . . . . . . . . . . . . . . . . . . . . . . 133

    7.12 Sound Generation by a Nonpremixed Jet Flame   . . . . . . . . . . 1357.12.1 Computational setup   . . . . . . . . . . . . . . . . . . . . . 1357.12.2 Nearfield results   . . . . . . . . . . . . . . . . . . . . . . . . 1377.12.3 Farfield Sound   . . . . . . . . . . . . . . . . . . . . . . . . . 137

    8 Conclusions & Future Work   . . . . . . . . . . . . . . . . . . . . . . . . . 154

    8.1 Conclusions   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1548.2 Future Work  . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157

    8.2.1 Source location identification   . . . . . . . . . . . . . . . . . 1578.2.2 Heated jets  . . . . . . . . . . . . . . . . . . . . . . . . . . . 1588.2.3 Effect of chevrons and lobed mixers   . . . . . . . . . . . . . 1588.2.4 Wavepacket models   . . . . . . . . . . . . . . . . . . . . . . 1588.2.5 Thickened flamelet model   . . . . . . . . . . . . . . . . . . . 1588.2.6 Combustion noise models   . . . . . . . . . . . . . . . . . . . 159

    8.2.7 Extension of Formulation 1C for diffraction effects   . . . . . 1 5 9

    III Appendices   160

    A A Low-Dispersion and Low-Dissipation Implicit Runge-Kutta Scheme   . . 161

    A.1 Introduction   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161A.2 Dispersion and dissipation of RK schemes . . . . . . . . . . . . . . 162A.3 ILDDRK scheme   . . . . . . . . . . . . . . . . . . . . . . . . . . . 165A.4 Numerical Example   . . . . . . . . . . . . . . . . . . . . . . . . . . 168

    A.4.1 Linear Wave Equation . . . . . . . . . . . . . . . . . . . . . 169A.4.2 Nonlinear Euler equations: One-Dimensional Case   . . . . . 1 7 1

    A.5 Conclusion  . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172

    xii

  • 8/18/2019 Dissertation Alireza

    13/248

    B Systematic definition of progress variables and Intrinsically Low-Dimensional,Flamelet Generated Manifolds for chemistry tabulation   . . . . . . . . 177

    B.1 Introduction   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177B.2 Principal Component Analysis   . . . . . . . . . . . . . . . . . . . . 179

    B.2.1 Background   . . . . . . . . . . . . . . . . . . . . . . . . . . 179B.2.2 Principal direction . . . . . . . . . . . . . . . . . . . . . . . 180B.2.3 Singular Value Decomposition   . . . . . . . . . . . . . . . . 181

    B.3 IL-FGM   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183B.4 Numerical Examples   . . . . . . . . . . . . . . . . . . . . . . . . . 186

    B.4.1 Flamelets with a single-progress variable   . . . . . . . . . . 186B.4.2 Flamelets with multi-progress variables   . . . . . . . . . . . 187

    B.5 Conclusion  . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189

    xiii

  • 8/18/2019 Dissertation Alireza

    14/248

    LIST OF TABLES

    4–1 Reported LES of sound radiation from SP07 jet.   . . . . . . . . . . . . 61

    4–2 Normalized cut-off wavenumbers for the grid spacing in the jet shear

    layer.   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

    4–3 Simulation time steps and runtime. The term AFTT denotes acoustic 

     flow-through time  and is equal to 32  D j/U  j. . . . . . . . . . . . . . . 66

    5–1 Parameters used in the design of the bandpass filters. The frequencies

    are cast in nondimensional form as the Strouhal numbers,  f D j/U  j.   . 84

    A–1 Optimal coefficients for the fourth-order, low-dispersion, low-dissipation,

    implicit Rung-Kutta scheme.   . . . . . . . . . . . . . . . . . . . . . . 168

    A–2 The L2  norm of the error between the numerical results and the exact

    solution at  t  = 300.   . . . . . . . . . . . . . . . . . . . . . . . . . . . 170

    A–3 Error between the numerical results and the exact solution at  t  = 300

    for different CFL numbers.   . . . . . . . . . . . . . . . . . . . . . . 171

    xiv

  • 8/18/2019 Dissertation Alireza

    15/248

    LIST OF TABLES xv

    B–1 The weight coefficients of each species, as obtained from PCA, to define

    the progress variable for a premixed  C H 4-air flame at φ  = 0.85. The

    results are shown to the fourth decimal. The entries for species with

    significant weight are in bold.   . . . . . . . . . . . . . . . . . . . . . 192

    B–2 The weight coefficients of each species, as obtained from PCA, to define

    the progress variable for a premixed  CH 4-air flame at  φ = 1.9. The

    results are shown to the fourth decimal. The entries for species with

    significant weight are in bold.   . . . . . . . . . . . . . . . . . . . . . 193

    B–3 The weight coefficients of each species, as obtained from PCA, to

    define the first progress variable,  c1, for a premixed  CH 4-air flame

    at  φ ∈ [0.6, 1.5]. The results are shown to the fourth decimal. Theentries for species with significant weight are in bold.   . . . . . . . . 198

    B–4 The weight coefficients of each species, as obtained from PCA, to

    define the second progress variable, c2, for a premixed CH 4-air flame

    at  φ ∈   [0.6, 1.5]. The results are shown to the fourth decimal.Theentries for species with significant weight are in bold.   . . . . . . . . 199

  • 8/18/2019 Dissertation Alireza

    16/248

    LIST OF FIGURES

    1–1 Schematic of jet flow field. The jet plume is illustrated with contours of 

    vorticity (hot color scheme) superimposed on the acoustic pressure

    (gray-scale color scheme). Microphone locations are commonly

    specified by their distance,  R, from the jet nozzle, and polar angle,

    Θ, with respect to the jet centerline.   . . . . . . . . . . . . . . . . . 17

    1–2 Wavepacket concept for jet flows. This figure shows how the existence

    of coherent structures results in a wavepacket pressure distribution

    in the shear layer.   . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

    1–3 A one-dimensional wavepacket illustration in (a) the physical and (b)

    the frequency domain. The gray area corresponds to the range of 

    radiating wavenumbers. The wavenumber corresponding to sonic

    propagation is denoted by  ka.   . . . . . . . . . . . . . . . . . . . . . 18

    1–4 Similarity spectra suggested by  Tam   et al.   (1996) for turbulent

    mixing noise: solid line: large-scale similarity spectrum ; dashed

    line: fine-scale similarity spectrum.   . . . . . . . . . . . . . . . . . . 19

    1–5 Sound pressure spectra measured by Bogey et al. (2007) at R  = 100D

    of a subsonic jet (M  j  = 0.9); solid line: measurements at Θ = 30◦;

    dashed line: measurements at Θ = 90

    .   . . . . . . . . . . . . . . . . 192–1 The seven-point overlap at the interface between two adjacent blocks.   35

    xvi

  • 8/18/2019 Dissertation Alireza

    17/248

    LIST OF FIGURES xvii

    2–2 The computation time needed to simulate 30 time steps in a jet flow

    simulation vs. increasing number of processors.   . . . . . . . . . . . 35

    3–1 Flow over a rigid body whose motion is defined by f (x, t).   . . . . . . 57

    3–2 Farfield directivity pattern of a point monopole, measured at r  = 340l,

    radiating in (a) a flow at  M 0  = 0.5 , and (b) a flow at  M 0  = 0.85

    measured at  r  = 340l; solid line: exact solution; symbols: FW-H

    code. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

    3–3 Nearfield directivity pattern of a monopole, measured at   r   = 5l,

    radiating in (a) a flow at  M 0  = 0.5 , and (b) a flow at  M 0  = 0.85

    (test case 1); solid line: exact solution; symbols: FW-H code.   . . . 58

    3–4 The directivity pattern of a point dipole, measured at   r   = 30l,

    radiating in (a) a medium at rest, and (b) a flow at   M 0   = 0.5

    moving in the positive  x1-direction (θ  = 0 [deg]); solid line: exact

    solution; symbols: FW-H code.   . . . . . . . . . . . . . . . . . . . . 58

    3–5 The schematic of a rotating monopole in a moving medium.   . . . . . 59

    3–6 Time history of the sound pressure generated by a rotating monopole

    in a moving medium (test case); (a) at (2l, 0, 0) where solid line

    corresponds to the exact solution and symbols represnt the results

    obtained from the FW-H code; (b) comparison between the results

    at (2l, 0, 0), θ = 0[deg], and (−2l, 0, 0), θ  = 180[deg].   . . . . . . . . 594–1 Grid stretching for Grid-30; left: axial direction; right: transverse

    direction.   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

    4–2 Grid-30 shown in  x-y  plane; every 4th node is shown.   . . . . . . . . . 72

  • 8/18/2019 Dissertation Alireza

    18/248

    LIST OF FIGURES xviii

    4–3 Grid stretching for Grid-380; left: axial direction; right: transverse

    direction.   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

    4–4 Grid-380 shown in x-y  plane; every 6th node is shown.   . . . . . . . . 73

    4–5 Variation of the attenuation due to molecular viscosity and LES filter

    with wavenumber,   k. In (a), solid line: numerical filter; dashed

    line: molecular viscosity (Grid-30); dashed-dotted line: molecular

    viscosity (Grid-88); dotted line: molecular viscosity (Grid-380); the

    symbol ∆ corresponds to the lateral grid spacing at  r  = D j/2. In

    (b), solid line: molecular viscosity; dashed line: LES filter (Grid-

    30); dashed-dotted line: LES filter (Grid-88); dotted line: LES

    filter (Grid-380).   . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

    4–6 Snapshots of vorticity field superimposed on acoustic pressure; (a):

    Grid-30; (b): Grid-88; (c): Grid-380.   . . . . . . . . . . . . . . . . . 75

    4–7 Contours of normalized mean axial velocity,  < U > /U  j, in the  x-y 

    plane; (a): Grid-30; (b): Grid-88; (c): Grid-380.   . . . . . . . . . . . 76

    4–8 Contours of normalized mean axial Reynolds stress, < σxx  >=< uu >

    /U 2 j , in the  x-y  plane; (a): Grid-30; (b): Grid-88; (c): Grid-380.   . . 77

    4–9 Variations of (a) centerline mean axial velocity, and (b) axial tur-

    bulence intensity. Solid line: LES (Grid-380); dashed line: LES

    (Grid-88); dashed-dotted line (Grid-30);  : LES of  Bogey   et al.

    (2011);  :   Arakeri et al.  (2003);  :   Lau et al.  (1979). . . . . . . . . 78

  • 8/18/2019 Dissertation Alireza

    19/248

    LIST OF FIGURES xix

    4–10 Inverse of mean streamwise velocity along the centerline normalized

    by the inflow jet velocity. Legend: solid line, eq. (4.7);  ,  Arakeri

    et al.  (2003);  , LES (Grid-380).   . . . . . . . . . . . . . . . . . . . 79

    4–11 The Ffowcs Williams-Hawkings surface setup for farfield sound pre-

    diction.   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

    4–12 Sound pressure spectra obtained at 52D j  from the jet nozzle: (a) at

    Θ = 30◦ ; (b) at Θ = 90◦; solid line: LES (Grid-380); dashed-line:

    LES (Grid-30);  :Bogey et al.  (2007);  :Tanna (1977).   . . . . . . . 80

    4–13 Directivity of overall sound pressure level (OASPL) in [dB] at 52D j

    from the jet nozzle; (a): LES (Grid-30); (b): LES (Grid-380).

    Legend:   •, LES with Grid-30;   , LES (unfiltered) with Grid-380;   , LES (filtered) with Grid-380;   ,Bogey   et al.   (2007);   ,

    Mollo-Christensen et al.  (1964);  , Lush (1971). . . . . . . . . . . . 81

    5–1 The magnitude response, in dB, of the one-third-octave bandpass

    filter with center frequency equivalent to   Stc   =   f cD j/U  j   = 0.4.

    This frequency band corresponds to  S t =  f D j/U  j ∈ [0.36, 0.45].   . . 885–2 The magnitude response, in dB, of the one-third-octave bandpass

    filter with center frequency equivalent to  Stc  = f cD j/U  j  = 4. This

    frequency band corresponds to  S t =  f D j/U  j ∈ [3.56, 4.48].   . . . . 885–3 Snapshots of the vorticity field superimposed on the acoustic pressure;

    (a): Unfitered field; (b): bandpass filtered around Stc  = 0.4; (c):

    bandpass filtered around  Stc = 4.   . . . . . . . . . . . . . . . . . . . 89

  • 8/18/2019 Dissertation Alireza

    20/248

    LIST OF FIGURES xx

    5–4 Snapshots of the vorticity field superimposed on the bandpass filtered

    acoustic pressure; The band-center frequency was  S tc  =  f D j/U  j  =

    0.4; snapshots are shown in sequence with a time difference of 

    ∆t = 0.84D j/U  j   .   . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

    5–5 Wavepackets and radiated sound obtained from the low-frequency

    bandpass filtering; (a): pressure and vorticity field shown together;

    (b): pressure field only. The potential core length, L, and dominant

    radiation directions are also shown with arrows for reference.  . . . . 91

    5–6 Wavepackets obtained from the bandpass filtering of vorticity and

    pressure fields; the color scheme corrosponds to bandpass filtered

    vorticity field while the gray scale color scheme corrosponds to the

    bandpass-filtered pressure contours; (a): low-frequency passband;

    (b): high-frequency passband.   . . . . . . . . . . . . . . . . . . . . 92

    5–7 Directivity of passband overall sound pressure level (OASPL) in

    [dB] at  R  = 52D j ; Legend:   , low-frequency radiated sound;  ,

    high-frequency radiated sound.   . . . . . . . . . . . . . . . . . . . . 93

    6–1 Contours of instantaneous vorticity superimposed on the pressure field

    (gray-scale colors); (a) Base round jet; (b) with Microjet setup.

    The nozzle is shown only schematically here for comparison, and

    was not included in the computational domain.   . . . . . . . . . . . 101

    6–2 Centerline distribution of (a) mean axial velocity and (b) axial

    turbulence intensity for the base round jet; solid line: present LES;

    :   Zaman (1986),  :   Lau et al.  (1979).   :   Arakeri  et al.  (2003).   . 102

  • 8/18/2019 Dissertation Alireza

    21/248

    LIST OF FIGURES xxi

    6–3 Inverse of mean streamwise velocity along the centerline normalized

    by the inflow jet velocity; solid line: LES of the base round jet;

    dashed dotted line: linear regression.   . . . . . . . . . . . . . . . . . 103

    6–4 Centerline distribution of mean axial velocity; solid line: base round

     jet; dashed dotted line: with microjets.   . . . . . . . . . . . . . . . . 103

    6–5 Contours of normalized mean axial velocity; (a) Base round jet; (b)

    with microjets.   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104

    6–6 The three-dimensional spatial evolution of the mean flow velocity

    with streamwise distance in the presence of microjets.   . . . . . . . 105

    6–7 The effect of microjets on the mean streamwise velocity at 1D from

    the nozzle exit.   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

    6–8 Contours of normalized axial Reynolds stress,  σxx   =  ux u

    x

    U 2j; (a) base

    round jet, (b) with microjets.   . . . . . . . . . . . . . . . . . . . . . 106

    6–9 The streamwise distribution of peak (a) axial turbulence intensity

    and (b) radial turbulence intensity; solid line: base round jet ;

    dashed-dotted line: with microjets   . . . . . . . . . . . . . . . . . . 107

    6–10 Measured and predicted far-field noise directivity; 1: base round jet

    (present LES); 2: with microjets (present LES); 3: base round jet

    (Bogey   et al., 2007); 4: base round jet (Alkislar   et al.,   2007); 5:

    with microjets (Alkislar   et al., 2007). The data are scaled for a

    common distance of  R  = 100D j.   . . . . . . . . . . . . . . . . . . . 108

  • 8/18/2019 Dissertation Alireza

    22/248

    LIST OF FIGURES xxii

    6–11 The Power Spectral Density (PSD) of the farfield acoustic sound

    pressure; (a) radiation angle Θ = 30◦; (b) Θ = 90◦. The data are

    reported for a common distance of  R  = 100D j.   . . . . . . . . . . . 109

    6–12 The band-pass filtered acoustic nearfield of the base setup; the

    frequency band corresponds to   St   = 0.225 to   St   = 0.375; (a)

    tU jDj

    = 438; (b)  tU jDj

    = 443; (c)  tU jDj

    = 448; (d)  tU jDj

    = 453. The arrows

    correspond to dominant radiation directions. . . . . . . . . . . . . . 110

    6–13 The band-pass filtered acoustic nearfield of the base setup; the

    frequency band corresponds to   St   = 0.925 to   St   = 1.075; (a)

    tU jDj

    = 438; (b)  tU jDj

    = 443; (c)  tU jDj

    = 448; (d)  tU jDj

    = 453. . . . . . . . . 111

    6–14 The band-pass filtered acoustic nearfield of the microjet setup; the

    frequency band corresponds to   St   = 0.225 to   St   = 0.375 (a)

    tU jDj

    = 272; (b)  tU jDj

    = 302; (c)  tU jDj

    = 332; (d)  tU jDj

    = 362. . . . . . . . . 112

    6–15 The band-pass filtered acoustic nearfield of the microjet setup; the

    frequency band corresponds to   St   = 0.925 to   St   = 1.075 (a)

    tU jDj

    = 272; (b)   tU jDj

    = 302; (c)   tU jDj

    = 332; (d)   tU jDj

    = 362. . . . . . . . . 113

    7–1 Solution of the steady flamelet equations for partially premixed

    CH 4/air combustion; the flame condition corresponds to the ex-

    periment of   Cabra   et al.   (2005). (a): S-shaped curve showing

    the variation of the temperature at stoichiometry as a function of 

    stoichiometric scalar dissipation rate, χst; (b): the flamelet solution

    in Z  space at  χst = 100s−1

    .   . . . . . . . . . . . . . . . . . . . . . . 140

  • 8/18/2019 Dissertation Alireza

    23/248

    LIST OF FIGURES xxiii

    7–2 A schematic of the effect of flamelet thickening; left: the original

    flamelet; right: the thickened flamelet.   . . . . . . . . . . . . . . . . 141

    7–3 The heat-release source term, ω̇T  obtained from solving the flamelet

    equations.   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 142

    7–4 The heat-release source term, ω̇T   at   χst   = 0.6; solid line: original

    flamelet solution as shown in Fig. 7–3; dashed line: thickened flamelet. 142

    7–5 Vorticity contours of a nonreactive, naturally developing mixing layer

    at  Reδω,0  = 1200.   . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143

    7–6 Vorticity contours of the reacting mixing layer. . . . . . . . . . . . . . 143

    7–7 Mixture fraction contours for the reacting mixing layer.Pure fuel

    corresponds to Z  = 1, while pure oxidizer corresponds to  Z  = 0.   . . 144

    7–8 Mixture fraction contours for the reacting mixing layer.Pure fuel

    corresponds to Z  = 1, while pure oxidizer corresponds to  Z  = 0.   . . 144

    7–9 Temperature contours and dilatation field for the reacting mixing layer.145

    7–10 Locations of the virtual probes for acoustic pressure measurements.   . 146

    7–11 The Fourier transform of the pressure fluctuation,  pac   =  p/p∞ − 1,measured at  R  = 70δ ω,0, from  x  = 92δ ω,0, and  y   = 0; solid line:

    Θ = 10 [deg]; dashed line: Θ = −10 [deg].   . . . . . . . . . . . . . . 1477–12 The Fourier transform of the pressure fluctuation,  pac   =  p/p∞ − 1,

    measured at  R  = 70δ ω,0, from  x  = 92δ ω,0, and  y   = 0; solid line:

    Θ = 50 [deg]; dashed line: Θ = −50 [deg].   . . . . . . . . . . . . . . 147

  • 8/18/2019 Dissertation Alireza

    24/248

    LIST OF FIGURES xxiv

    7–13 The Fourier transform of the pressure fluctuation,  pac   =  p/p∞ − 1,measured at  R  = 70δ ω,0, from  x  = 92δ ω,0, and  y   = 0; solid line:

    Θ = 90 [deg]; dashed line: Θ = −90 [deg].   . . . . . . . . . . . . . . 1487–14 The Fourier transform of the pressure fluctuation,  pac   =  p/p∞ − 1,

    measured at  R  = 70δ ω,0, from  x  = 92δ ω,0, and  y   = 0; solid line:

    Θ = 140 [deg]; dashed line: Θ = −140 [deg].   . . . . . . . . . . . . . 1487–15 The directivity of radiated sound pressure level measured at   R   =

    70δ ω,0, from  x  = 92δ ω,0, and  y   = 0; (a): at frequency f 0; (b): at

    frequency 4f 0. The ambient pressure was assumed to be atmospheric.149

    7–16 Instantaneous snapshots of jet flame flow field; (a): contours of 

    temperature (hot color scheme) superimposed on the acoustic

    pressure (gray-scale color scheme); (b): mixture fraction.   . . . . . . 150

    7–17 Mean and rms statistics of (a) axial velocity, (b) mixture fraction,

    and (c) temperature along the jet flame centerline.   . . . . . . . . . 151

    7–18 Sound pressure level spectra at a distance of  R  = 100D j  from the jet

    flame nozzle; : radiation angle Θ = 30◦; : radiation angle Θ = 90◦.152

    7–19 Directivity of overall sound pressure level (OASPL) in [dB] at 100D j

    from the jet nozzle   . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

    7–20 Temperature field superimposed on the bandpass-filtered pressure field

    of the jet flame; (a): center frequency corresponds to  StD   = 0.4;

    (b): center frequency corresponds to StD  = 1; (c): center frequency

    corresponds to StD  = 4.0   . . . . . . . . . . . . . . . . . . . . . . . 153

  • 8/18/2019 Dissertation Alireza

    25/248

    LIST OF FIGURES xxv

    A–1 Amplification factor (a) and difference in phase of the Runge-Kutta

    schemes: •: the new ILDDRK scheme,  , standard explicit RK4;

    , SDIRK.   . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174

    A–2 (a) Dissipation and (b) dispersion error of the Runge-Kutta schemes

    in logarithmic scales:  •: the new ILDDRK scheme,  , standardexplicit RK4;  , SDIRK   . . . . . . . . . . . . . . . . . . . . . . . . 174

    A–3 Numerical solution of eq. (A.29) at   t   = 300; (a) CFL=0.5; (b)

    CFL=1.0; (c) CFL=1.5; (d) CFL=2.0.   . . . . . . . . . . . . . . . . 175

    A–4 Optional caption for list of figures   . . . . . . . . . . . . . . . . . . . . 176

    B–1 A schematic profile of the mass fraction of a species in a one-

    dimensional freely propagating flame. The flame is sampled at  n

    locations in the physical domain to construct the data set.   . . . . . 1 9 1

    B–2 The structure of a freely propagating flame in a CH 4-air mixture for

    φ = 0.85, T 0 = 473 [K], and P   = 1 [atm] in the physical space; solid

    line: physical space solution; symbols: solution retrieved from the

    previously generated flamelet table. (a): temperature; (b):   CO2

    mass fraction; (c):   CO  mass fraction; (d):   OH  mass fraction.   . . . 194

    B–3 The solution of the flame, shown in Fig.  B–2  versus the progress

    variable; solid line: physical space solution; symbols: flamelet

    table. (a): temperature; (b):   CO2   mass fracion; (c):   CO   mass

    fracion; (d):   OH   mass fraction.   . . . . . . . . . . . . . . . . . . . . 195

  • 8/18/2019 Dissertation Alireza

    26/248

    LIST OF FIGURES xxvi

    B–4 The structure of a freely propagating flame in a CH 4-air mixture for

    φ   = 1.9,   T 0   = 473 [K], and   P   = 1 [atm] in the physical space;

    solid line: physical space solution; circles: solution retrieved from

    the previously generated flamelet table using the new definition of 

    progress variable; triangles: solution retrieved from the previously

    generated flamelet table using  c  =  Y CO2 + Y CO  + Y H 2O + Y H 2. (a):

    temperature; (b):   CO2  mass fraction; (c):   CO  mass fraction; (d):

    OH  mass fraction.   . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

    B–5 The temperature distribution in the progress variable space; solid

    lines: physical space solution for various equivalence ratios ; sym-

    bols: interpolated values for equivalence ratio  φ  = 1.15.  . . . . . . . 197

    B–6 The structure of a freely propagating flame in a CH 4-air mixture for

    φ = 1.15, T 0 = 473 [K], and P   = 1 [atm] in the physical space; solid

    line: physical space solution; symbols: solution retrieved from the

    previously generated flamelet table. (a): temperature; (b):   CO2

    mass fraction; (c):   CO  mass fraction; (d):   OH  mass fraction.   . . . 200

  • 8/18/2019 Dissertation Alireza

    27/248

    NOMENCLATURE

    Roman Symbols

    B   Jet decay rate parameter; see eq. (4.7) on page 68

    D j   Jet diameter

    Dk   Diffusivity of species  k

    DT    Thermal diffusivity

    Dµ   Microjet diamater

    Da   Damköhler number

    F, G, H   Inviscid flux vectors in the Navier-Stokes equations

    Fv, Gv, Hv   Viscous flux vectors in the Navier-Stokes equations

    F r   Froud number

    G(x, y; ∆) LES filter function; see eq. (2.1) on page 21

    g   Body force vector

    H (f ) Heaviside function

    I a   Acoustic intensity

    J    Jacobian of coordinate transformation

    k   Waveknumber

    Le   Lewis number

    M    Mach number

    xxvii

  • 8/18/2019 Dissertation Alireza

    28/248

    NOMENCLATURE xxviii

    P r   Prandtl number

    Q   Vector of flow conserved variables

    q   Heat flux vector

    R   Distance from the jet nozzle exit; see Fig. 1–1 on page 17

    R   Universal gas constant

    R   Ideal Gas constant

    Re   Reynolds number

    S    Entropy

    S ij   Strain rate tensor

    St   Strouhal number

    T    Temperature

    T ij   Lighthill’s stress tensor

    t   Time

    U  j   Jet velocity

    u   = [u v w]T , Velocity vector

    (u,v,w) Velocity components physical coordinates

    V k,j   j-component of diffusion velocity of species k ; see eq. (7.2) on page 117

    x   = [x y z ]T , spatial coordinate vector

    Y    Species mass fraction

    Z    Mixture fractionZ 2 Residual scalar variance of mixture fraction

  • 8/18/2019 Dissertation Alireza

    29/248

    NOMENCLATURE xxix

    Greek Symbols

    αf    Filtering parameter; see eq. (2.36) on page 29

    ∆ Grid spacing

    ∆0   Minimum grid spacing

    δ ij   Kronecker delta

    δ θ   Shear layer momentum thickness

    δ ω   Shear layer vorticity thickness

    εn   Random parameter in forcing procedure; see eq. (2.48) on page 33

    γ    Specific heat ratio

    Π Phillips’ Acoustic parameter; see eq. (1.6) on page 6

    µ   Molecular viscosity

    ω   Angular frequency

    ω̇k   Chemical source term per unit mass; see eq. (7.2) on page 117

    ω̇T    Chemical heat release term per unit mass; see eq. (7.2) on page 117

    ρ   Density

    σij   Shear stress tensor

    σ   Sponge zone damping parameter

    τ e   Emission time; see eq. (3.44) on page 48

    Θ Emission angle; see Fig. 1–1  on page 17

    χ   Scalar dissipation rate; see eq. (7.15) on page 121

    (ξ , η , ζ  ) Generalized curvilinear coordinates

  • 8/18/2019 Dissertation Alireza

    30/248

    NOMENCLATURE xxx

    (ξ x, ξ y, ξ z) Partial derivatives of  ξ  along physical coordinates

    (ηx, ηy, ηz) Partial derivatives of  η  along physical coordinates

    (ζ x, ζ y, ζ z) Partial derivatives of  ζ   along physical coordinates

    Superscripts, Subscripts, and Accents

    (·)0   Ambient condition

    (·)∞   Freestream condition

    (·) j   Jet

    (·)r   Reference condition(·)rms   Root mean square

    (·)sgs   Subgrid scale

    (·)st   Stoichiometric condition

    (·) LES filtered variable, see eq. (2.1) on page (2.1)

    (·) Favre-averaged quantity, see eq. (2.3) on page (2.3)(·)µ   Microjet properties

    < · >   Reynolds averaged property

    Abbreviations

    CFD Computational Fluid Dynamics

    dB Decibel

    DNS Direct Numerical Simulation

    FW-H Ffowcs-Williams & Hawkings

  • 8/18/2019 Dissertation Alireza

    31/248

    NOMENCLATURE 1

    LES Large Eddy Simulation

    LHS Left hand side

    MCAAP McGill Computational Acoustic Analogy Package

    NSCBC Navier-Stokes Characteristic Boundary Conditions

    OASPL Overall sound pressure level

    RHS Right hand side

    RMS Root mean squared

    SPL Sound pressure level

  • 8/18/2019 Dissertation Alireza

    32/248

    “The sensation of sound is a thing  sui generis , not comparable with any of our

    sensations.”

    – Lord Rayleigh in  The Theory of Sound 

    2

  • 8/18/2019 Dissertation Alireza

    33/248

    CHAPTER 1Introduction

    1.1 Motivation

    More than a century after the first powered flight by the Wright brothers, avia-

    tion is a dominant industry and a major job creator, specially in the North American

    economy. With increases in air traffic, aircraft noise has increased and become a nui-

    sance for communities living close to active airports. This has led governments to

    create increasingly stringent regulations for the maximum sound levels allowable dur-

    ing aircraft landing and takeoff. Despite significant improvements over the past few

    decades, aircraft noise remains a significant challenge. Governments and aerospace

    industries have laid out plans to reduce the current aircraft noise level by 20 EPNdB1

    by the year 2020 (Weasoky, 1998).

    One of the most significant contributor to aircraft sound emission is the engine.The sound radiated from the propulsion system consists of fan noise, combustion/core

    noise, and jet exhaust noise. The jet exhaust noise and combustion core noise are

    especially significant during takeoff phase. After more than half a century, sound

    generation by high speed turbulent jet flows remains one of the most recalcitrant and

    challenging problems in aeroacoustics. It is widely accepted that coherent structures

    1 Effective Perceived Noise level dB

    3

  • 8/18/2019 Dissertation Alireza

    34/248

    1.2. BACKGROUND ON UNHEATED JET NOISE 4

    play a significant role in sound generation by turbulent jets. Despite numerous

    experimental and numerical studies, the details of sound generation mechanisms by

    turbulent jet flows are still not well understood, and the nature of the sound sources,

    especially for subsonic jets, is still the subject of debates among researchers. In the

    present study, the role of coherent structures in sound radiation from jet exhaust

    flows is studied. It is expected that the findings of this study be useful to describe

    mechanisms of sound generation by subsonic jets, and develop new phenomenological

    models.

    With the introduction of high-by-pass ratio engines to reduce jet exhaust noise,

    other sources such as combustion and core noise have become more prominent. Sound

    radiation from nonpremixed flames is also studied in the present work.

    1.2 Background and Literature Review: Nonreacting Jet Noise

    The first attempt to develop a theory for jet noise was made by Sir James

    Lighthill (Lighthill,   1952, 1954) who introduced the concept of   acoustic analogy .

    Lighthill’s idea was to recast the exact equations of motion, i.e., the conservation

    of mass and momentum, as an inhomogeneous wave equation where the nonlinear

    terms are moved to the right hand side (RHS) and are treated as sources of sound.

    The problem of calculating the turbulence-generated sound is then equivalent to

    solving the radiation of a distribution of sources into an ideal fluid at rest. Lighthill’s

    inhomogeneous wave equation is

    1

    c

    2

    0

    ∂ 2

    ∂t2

     − ∇2

     p =

      ∂ 2T ij

    ∂xi ∂x j

    ,   (1.1)

  • 8/18/2019 Dissertation Alireza

    35/248

    1.2. BACKGROUND ON UNHEATED JET NOISE 5

    where T ij  is called the Lighthill’s stress tensor and is given by

    T ij  = ρ ui u j + ( p − p0) − c20(ρ − ρ0) δ ij − σij .   (1.2)In eq. (1.2), σij  is the shear stress tensor, ui  and  u j  are the flow velocity components,

    and p0, ρ0, and c0 are the ambient pressure, density, and speed of sound, respectively.

    The first term in Lighthill’s stress tensor, eq. (1.2), is commonly referred to as the

    Reynolds stress, and is a significant source of sound in turbulent flows. The second

    term is due to wave amplitude nonlinearity and entropy variations in the source

    region. The third term represents the attenuation of sound waves due to viscous

    stresses, and is usually neglected. Due to the appearance of the double divergence

    operator   ∂ 2

    ∂xi∂xj, the source of eq. (1.1) is referred to as a quadrupole with strength

    T ij   (Howe, 2003).

    Lighthill’s analogy, eq. (1.1) implies that the problem of turbulence-generated

    sound is equivalent to the radiation of a distribution of quadrupole sources with

    strength T ij   into a stationary, ideal fluid (Howe, 2003). Using his acoustic analogy,

    Lighthill (1954) showed that the acoustic intensity,  I a, of the sound radiated from a

    low Mach number jet is given by

    I a =  K ρ2 jD

    2 j U 

    8 j

    ρ0c50R2

      ,   (1.3)

    where K   is a constant of the order of 10−5.

    The effect of source convection was also discussed by  Lighthill (1952), and then

    investigated in more details by   Ffowcs Williams   (1960, 1963). They showed that

    the Doppler shift due to the convection of quadrupole sources results in stronger

  • 8/18/2019 Dissertation Alireza

    36/248

    1.2. BACKGROUND ON UNHEATED JET NOISE 6

    radiation in the aft direction, such that the acoustic intensity is equal to

    I a =  K ρ2 jD

    2 j U 

    8 j

    ρ0c50 (1 − M c cos Θ)2 +   lrτ rc025/2 R2,   (1.4)

    where M c =  U c/c0  is the convective Mach number,  lr  is the turbulence characteristic

    length scale, and  τ r   is the turbulence characteristic time scale. The parameters  R

    and Θ specify the location of the microphone, as shown in Fig.   1–1. The term

    (1 − M c cos Θ) captures the effects of the Doppler shift, whereas the term   lr/τ rc0takes into account the spatial extent of eddies. The latter term is of significance in

    directions normal to the Mach waves, where  M c cos Θ = 1.

    In Lighthill’s analogy, the convection and refraction of the emitted sound waves

    are neglected, and are lumped into the source term which can lead to inaccurate

    predictions for high speed jet noise. Therefore, Lighthill’s analogy was subsequently

    modified (Phillips, 1960; Goldstein & Howe, 1973; Lilley, 1974;  Goldstein, 2003) to

    take propagation effects into account by including the corresponding terms in the

    wave operator.

    Phillips (1960) derived a convected wave equation,

    D2

    D t2Π −   ∂ 

    ∂xi

    c2

    ∂ Π

    ∂xi

     =

      ∂ui∂x j

    ∂u j∂xi

    +  D

    Dt

    1

    c p

    dS 

    dT 

    −   ∂ 

    ∂x j

    1

    ρ

    ∂σij∂x j

     ,   (1.5)

    where the parameter Π is given by

    Π =  1

    γ  ln

      p

     p0.   (1.6)

  • 8/18/2019 Dissertation Alireza

    37/248

    1.2. BACKGROUND ON UNHEATED JET NOISE 7

    The first term on the RHS of eq. (1.5),   ∂ui∂xj

    ∂uj∂xi

    , represents the aerodynamic sources

    due to velocity fluctuations, while the second and third terms correspond to the

    effects of entropy sources and viscosity fluctuations, respectively.

    In comparison to Lighthill’s analogy, eq. (1.1), a convective term has been moved

    to the left-hand side of eq. (1.5). This leads to the appearance of a second-order total

    time derivative in the wave operator. The spatial dependence of the speed of sound

    has also been included in the wave operator to account for refraction effects.

    Lilley (1974) argued that the first term on the RHS of Phillips’ equation contains

    propagation effects for shear flows that should be included in the convective wave

    operator on the LHS. He then derived a third-order wave equation2 ,

    D

    Dt

    D2Π

    Dt  −   ∂ 

    ∂xi

    c2

    ∂ Π

    ∂xi

    + 2

    ∂u j∂xi

    ∂ 

    ∂x j

    c2

    ∂ Π

    ∂xi

     = −∂u j

    ∂xi

    ∂uk∂x j

    ∂ui∂xk

    + Ψ ,   (1.7)

    where Ψ represents the effects of entropy generation and viscosity fluctuations which

    are generally neglected. Some industrial jet noise prediction tools, such as General

    Electric’s MGB approach (Balsa  et al., 1978), are based on Lilley’s analogy.

    More recently, Goldstein (2003) has proposed a generalized acoustic analogy in

    which the Navier-Stokes equations are recast as a set of linearized inhomogeneous

    2 Lilley’s equation, as published in the original paper (Lilley, 1974), reads slightlydifferent from eq. (1.7). In the original paper, it was assumed that the mean flow isparallel to  x1  axis.

  • 8/18/2019 Dissertation Alireza

    38/248

    1.2. BACKGROUND ON UNHEATED JET NOISE 8

    Euler equations with source terms representing shear-stress and energy-flux pertur-

    bations. This approach requires the proper choice of a base flow around which the

    Navier-Stokes equations are linearized.

    Acoustic analogies have some intrinsic drawbacks. The linearization of the gov-

    erning equations and the dissociation of the propagation effects (i.e. refraction of 

    sound waves) from the source terms are usually somewhat arbitrary without any a

    priori knowledge of the sound field. This means that the source terms in acoustic

    analogy theories are not necessarily true sources of sound. Moreover, the source

    terms are assumed to be known, making acoustic analogies dependent on separate

    experimental measurements or theoretical or numerical calculations. With the ex-

    ception of Lighthill’s equation, almost all acoustic analogies involve nonlinear dif-

    ferential operators which generally cannot be integrated analytically. This implies

    that a quantitative prediction of the farfield sound requires the numerical solution

    of the acoustic analogy partial differential equations, which can be computationally

    intensive.

    As an alternative to acoustic analogy theories, some researchers have investi-

    gated phenomenological models based on the observation of coherent structures in

    turbulent jets.   Mollo-Christensen (1967) suggested a wavepacket concept to model

    sound generation by coherent structures. Figure 1–2  illustrates how the existence of 

    coherent structures in the shear layer results in a pressure distribution that can be

    modeled as a wave packet.

    In supersonic jets, such coherent structures generate sound through Mach wave

    radiation to the farfield, a process analogous to the sound radiation by a supersonic

  • 8/18/2019 Dissertation Alireza

    39/248

    1.2. BACKGROUND ON UNHEATED JET NOISE 9

    wavy wall   (Tam, 1995). The introduction of the wavepacket concept led researchers

    to explore the flow stability theory as a theoretical framework to analyze the nearfield

    dynamics and its relation to farfield sound radiation (Michalke, 1970, 1972; Mankbadi

    & Liu,   1984). Models based on linear convecting instability waves were developed

    by Tam & Morris  (1980) and Tam & Burton (1984a,b). More recently,  Wu (2005)

    investigated the sound radiation from nonlinearly evolving instabilities.

    The relevance of models based on instability wave radiation for subsonic jet noise

    is debatable. This is because the energy in wave-numbers with supersonic phase

    speeds appears only with the growth and decay of subsonically convected instability

    waves. In other words, it can be argued that instability waves do not propagate

    supersonically to radiate Mach waves. The appearance of supersonic phase speeds

    may be explained as an artifact of the Fourier transform (cf. Fig.  1–3). Despite this

    fact, several authors have investigated the sound emissions from wave packet pressure

    fields (Crighton & Huerre, 1990; Avital & Sandham, 1997; Le Dizès & Millet, 2007;

    Obrist, 2009, 2011; Cavalieri  et al., 2011).

    Following an exhaustive review of supersonic jet noise data,  Tam  et al.  (1996)

    showed that far-field sound pressure spectral densities can be fit using two similarity

    spectra. The spectrum shape that dominated the aft angles was associated with

    Mach wave radiation by large scale structures. Hence, this spectrum is commonly

    referred to as the Large Scale Similarity (LSS) spectrum (Morris, 2009). The second

    spectrum was used to fit the sound pressure spectra radiated toward sidelines.   Tam

    et al.  (1996) argued that the sound radiated toward sidelines is generated by small-

    scale structures. Hence, the second spectrum is commonly designated as the Fine

  • 8/18/2019 Dissertation Alireza

    40/248

    1.2. BACKGROUND ON UNHEATED JET NOISE 10

    Scale Similarity (FSS) spectrum. At intermediate angles, a combination of the two

    spectra is needed to match the experimental data.

    The two similarity spectra were used by  Viswanathan   (2002) and   Tam   et al.

    (2008) to fit further experimental results, including the data for subsonic jets.   Tam

    et al. (2008) argued that these results reinforce the idea that the two similarity spectra

    correspond to two separate noise generation mechanisms in jets for all operating

    conditions. This model is similar to the one initially proposed by Schlinker (1975),

    and Laufer  et al.  (1975) for turbulence mixing noise in supersonic jets.

    The large-scale/small-scale description of jet noise radiation is primarily based

    on the fact that a combination of the two similarity spectra can be used to fit a given

    farfield sound pressure spectrum provided that the peak frequencies and spectral

    magnitudes are correctly chosen. However, the decomposition of a given spectrum

    into two or more spectra does not necessarily yield a unique solution. In other words,

    one can choose another set of, say, three ad-hoc similarity spectra and combine them

    such that their sum reproduce a given spectrum. Another important observation

    which seems to be inconsistent with  Tam & Morris  (1980)’s hypothesis is that the

    peak of the supposedly fine scale radiated sound spectrum, measured for example at

    90◦ from the jet axis, is observed at relatively low Strouhal numbers (St ≈ 0.3). Thisvalue of Strouhal number is close to that of the spectrum measured at 30◦ which is

    supposed to be the large scale radiated sound spectrum (c.f. Fig.   1–5). But, one

    would expect that the peak frequency of small-scale radiated sound be at least an

    order of magnitude larger than that of large-scale radiated sound.

  • 8/18/2019 Dissertation Alireza

    41/248

    1.3. BACKGROUND ON REACTING JET NOISE 11

    A more rigorous approach to identify the sound generation mechanism by jets is

    to look for direct relationships between the nearfield flow properties and the radiated

    sound field. For example,   Panda & Seasholtz (2002) and Panda  et al.  (2005) used

    a Rayleigh-scattering technique to measure correlations between the radiated sound

    pressure and nearfield quantities, such as density and velocity fluctuations, for jets at

    Mach numbers 0.8, 0.95, 1.4 and 1.8. They reported significant correlations between

    nearfield density fluctuations and farfield pressure at aft angles where the convective

    velocity was supersonic, while such correlations were negligible for subsonic jets.   Bo-

    gey & Bailly (2007) used a similar causality method to investigate sound generation

    from isothermal jets at Mach numbers 0.6 and 0.9. The nearfield results were ob-

    tained from the large eddy simulations of the jets. The simulations were performed

    at two Reynolds numbers,   ReD   = 1.7 × 103 and   ReD   = 4 × 105, with the aim of studying the effects of both Mach and Reynolds number on the correlations between

    the radiated sound pressure and nearfield flow quantities. They reported significant

    levels of correlation between the centerline turbulence and sound radiated to Θ = 40 ◦

    for all four jets. Maximum correlations were observed at the end of the potential

    core. Based on the analysis,  Bogey & Bailly (2007) suggested the presence of a noise

    generation mechanism near the end of the potential core.

    1.3 Background and Literature Review: Reacting Jet Noise

    Combustion can be a significant contributor to the aero-engine core noise (Smith,

    2004). The interaction of acoustic sound waves with the flame can also lead to

    combustion instability in gas turbine engines and industrial furnaces. Depending

    on the generation mechanism, combustion noise can be characterized as   direct   or

  • 8/18/2019 Dissertation Alireza

    42/248

    1.3. BACKGROUND ON REACTING JET NOISE 12

    indirect  (Strahle, 1971, 1978). Direct noise is caused by volumetric expansion due to

    heat release, and constitutes a monopole source. Indirect sound is generated by the

    differential acceleration of entropy non-uniformities and their interaction with solid

    boundaries. This mechanism constitutes a dipole source (Howe, 1998).

    A review of literature shows that the majority of the combustion noise studies

    are either theoretical or experimental. Numerical simulations of sound radiation by

    open flames can help improve our understanding of broadband combustion noise. In

    the absence of a retro-active feedback leading to instability, the flame can be excited

    in a controlled manner to measure the flame transfer function, a quantity useful

    for combustion instability studies. This approach is called  system identification   in

    numerical studies of combustion instability (Poinsot & Veynante, 2005).

    Zhao & Frankel (2001) preformed the DNS of sound radiation from an axisym-

    metric premixed reacting jet. They used a 6th-order compact scheme (Lele, 1992) for

    spatial discretization, and a 4th-order Runge-Kutta algorithm for temporal integra-

    tion. A generic one-step global reaction was considered to model the chemistry. Their

    computational domain included both the nearfield and the farfield. They observed

    that combustion heat release had a significant effect on the vortical structure of the

     jet, as well as the radiated sound pressure level and directivity. Lighthill’s acoustic

    analogy was used to identify apparent sound source locations. It was concluded that

    heat release stabilized the jet, enhanced sound radiation levels, and altered the fre-

    quency of the most unstable modes to lower values, which led to a broader sound

    spectrum.

  • 8/18/2019 Dissertation Alireza

    43/248

    1.3. BACKGROUND ON REACTING JET NOISE 13

    Flemming et al.  (2007) performed an LES of the H3 flame (Tacke et al., 1998).

    The simulation was based on the low-Mach-number assumption in which the pres-

    sure work is neglected in the energy equation (Poinsot & Veynante, 2005). This

    assumption reduces the computational cost by relaxing the CFL restriction. But,

    the results can no longer be used to directly predict the generation and propagation

    of sound. A linear wave equation with a monopole source term was used to estimate

    the radiated sound pressure level.

    Bui et al. (2007) used a hybrid LES/CAA approach, in which a low-Mach num-

    ber LES was combined with an acoustic perturbation equation modified for reacting

    flows.   Mühlbauer   et al.   (2008) used a RANS/statistical model to simulate broad-

    band combustion noise of the open non-premixed DLR-A jet flame.   Ihme   et al.

    (2009) also used the low-Mach-number assumption to perform an LES of an   N 2-

    diluted C H 4 − H 2/air flame. They developed a modified form of Lighthill’s acousticanalogy to predict combustion generated sound. Their acoustic analogy utilized the

    flamelet/progress-variable model to formulate the excess density.

    So far numerical studies of nonpremixed flames have used the low-Mach-number

    assumption to simulate the hydrodynamic field. The radiated sound has been cal-

    culated based on simplified models. It is interesting to directly simulate the sound

    production and radiation by diffusion flames. Such studies can enhance our cur-

    rent understanding of turbulence/acoustic/flame interactions, and can be used as

    benchmarks to further validate combustion noise models or develop new ones.

  • 8/18/2019 Dissertation Alireza

    44/248

    1.4. OBJECTIVES 14

    1.4 Objectives

    The literature review presented in Sec.  1.2 suggests that further studies on the

    nature of jet noise sources and their dynamics are required. While acoustic analogies

    seem to be incapable of providing more insight into the physics of sound sources,

    phenomenological models, such as wavepacket theory, seem promising as models

    of generation by shear flows. Developing such models, and calibrating them for

    practical noise predictions require a better understanding, both qualitatively and

    quantitatively, of sound radiation by coherent structures at different scales.

    Although two-point space-time correlations provide some information about the

    relation between the nearfield flow and the sound perceived in the farfield, they

    do not provide a comprehensive picture of sound radiation patterns by turbulent

    structures at different scales. The objective of the present study was to investigate

    a new approach, bandpass filtering of the flow field, to provide further insight into

    the dynamics of sound radiation by coherent structures of subsonic jet flows. This

    approach shows how different scales contribute to farfield sound radiation. In the

    present study, Large-Eddy Simulations (LES) of high-speed subsonic jet flows were

    performed to obtain both the nearfield sources and the farfield radiated sound. Band-

    pass filtering was then used as a post-processing tool to visualize and characterize

    the radiated acoustic field in different frequency bands. The result of this study were

    used to identify source mechanisms in subsonic jets. The results suggest that Tam

    et al.  (2008) description of jet noise generation may be oversimplified.

    Similar analysis was also used to perform a preliminary study of the sound

    radiation by a nonpremixed reacting mixing layer, and a diffusion jet flame. The

  • 8/18/2019 Dissertation Alireza

    45/248

    1.5. ORGANIZATION OF THE THESIS 15

    results provide a better understanding of combustion/core noise problem, which has

    not received as much attention as the jet noise problem.

    1.5 Organization of the Thesis

    The thesis is divided into two parts. The problem of sound radiation from

    nonreacting, high speed, subsonic jets is studied in Part I, while combustion noise is

    the subject of Part II. Chapter 2 presents the governing equations, and the numerical

    schemes used in the LES of nearfield flow. The farfield sound prediction method is

    presented in Chapter 3. The LES results for a Mach 0.9 jet at  ReD   = 4 × 105 arepresented in Chapter 4. Chapter 5 describes the bandpass filtering procedure, and the

    major findings obtained from this approach. Using the LES and bandpass filtering,

    the use of mircojets for jet noise suppression is studied in Chapter   6. Numerical

    simulations of sound radiation from a reacting mixing layer, and a jet diffusion flame

    are presented in Chapter   7. Some concluding remarks, and suggestions for future

    work are presented and in Chapter  8.

    1.6 Contributions

    The following list summarizes the major contributions of the present work:

    •  High-fidelity, direct computations of sound radiation from reacting and nonre-acting jet flows.

    •  Investigation of grid resolution effects on the accuracy of jet noise simulations.•  A band-pass filter visualization and analysis of the source region and the acous-

    tic field of subsonic jets.

    •   Development of an Implicit, Low-Dispersion, Low-Dissipation Runge-Kutta(ILDDRK) scheme of fourth order accuracy.

  • 8/18/2019 Dissertation Alireza

    46/248

    1.6. CONTRIBUTIONS 16

    •  Development of a new surface-integral acoustic analogy formulation, Formula-tion 1C , to predict the sound radiated by moving sources in uniformly moving

    media.

    •   Large Eddy Simulation of jet noise suppression by impinging microjets.•  Development of a thickened flamelet model for simulations of turbulent non-

    premixed flames.

    •   Development of a systematic method to define progress variables and the Intrin-sically Low-Dimensional, Flamelet Generated Manifold (IL-FGM) modeling for

    chemistry tabulation in LES of turbulent flames

    •  Direct noise computation of a reacting mixing layer, and a jet diffusion flame.

  • 8/18/2019 Dissertation Alireza

    47/248

    FIGURES OF CHAPTER 1 17

    x

    Θ

    R

    Figure 1–1: Schematic of jet flow field. The jet plume is illustrated with contours of vorticity (hot color scheme) superimposed on the acoustic pressure (gray-scale colorscheme). Microphone locations are commonly specified by their distance,   R, fromthe jet nozzle, and polar angle, Θ, with respect to the jet centerline.

     p(x)

    x

    Figure 1–2: Wavepacket concept for jet flows. This figure shows how the existence of coherent structures results in a wavepacket pressure distribution in the shear layer.

  • 8/18/2019 Dissertation Alireza

    48/248

    FIGURES OF CHAPTER 1 18

    x

    |q (x)|q 0

    2l(a)

    k

    k0

    ka−ka

    downstream   upstream

    q̂ (k)

    (b)

    Figure 1–3: A one-dimensional wavepacket illustration in (a) the physical and (b) thefrequency domain. The gray area corresponds to the range of radiating wavenumbers.The wavenumber corresponding to sonic propagation is denoted by  ka.

  • 8/18/2019 Dissertation Alireza

    49/248

    FIGURES OF CHAPTER 1 19

    -20

    -15

    -10

    -5

     0

     5

     10

     0.1 1 10

       S   P   L

       [   d   B   ]

     f / f  p

    Figure 1–4: Similarity spectra suggested by Tam  et al.  (1996) for turbulent mixingnoise: solid line: large-scale similarity spectrum ; dashed line: fine-scale similarityspectrum.

     80

     85

     90

     95

     100

     105

     110

     115

     120

     0.01 0.1 1

       S   P   L   [   d   B   ]   /   S   t

    St  D = f D / U  j

    Figure 1–5: Sound pressure spectra measured by Bogey  et al.  (2007) at  R  = 100Dof a subsonic jet (M  j   = 0.9); solid line: measurements at Θ = 30

    ◦; dashed line:measurements at Θ = 90◦.

  • 8/18/2019 Dissertation Alireza

    50/248

    Part I

    Nonreacting Flows

    20

  • 8/18/2019 Dissertation Alireza

    51/248

    CHAPTER 2Governing Equations and Numerical Methods

    In the present study, the Navier-Stokes equations are numerically solved in the

    context of Large-Eddy Simulation (LES).The basic idea behind LES is to decompose

    the flow properties into a large-scale or resolved component,  ψ , and a small-scale or

    subgrid component, ψsg. This decomposition is achieved by applying a spatial-filter

    (Pope, 2000),

    ψ(t, x) =

       ψ(t, y)G(x, y;∆)dy ,   (2.1)

    where ∆ is the filter size, and  G   is the filter function satisfying the normalization

    condition    G(x, y; ∆) = 1 .   (2.2)

    For compressible and reacting flows where the density changes significantly, Favre

    filtering is employed. A Favre-filtered quantity, ψ   is defined asψ =  ρψ

    ρ  =

     1

    ρ

       ρ(t, y)ψ(t, y)G(x, y;∆)dy .   (2.3)

    It is customary to filter the Navier-Stokes equations first, and then solve the

    resulting equations for the filtered quantities. This procedure, also known as  implicit 

     filtering , always results in the so-called closure problem, i.e. some terms remain

    unclosed and need modeling. These terms correspond to the effects of subgrid-scale

    dynamics. Commonly used subgrid scale models include the classic Smagorinsky

    21

  • 8/18/2019 Dissertation Alireza

    52/248

    2.1. GOVERNING EQUATIONS 22

    model (Smagorinsky, 1963), and the dynamic Smagorinsky model (Germano et al.,

    1991; Moin  et al., 1991).

    In the implicit filtering, the filter function is not necessarily known as it is defined

    by the numerical grid (Kravchenko & Moin, 1997;   Meyers & Sagaut, 2007;   Lund,

    2003). Hence, the filter effect on energy dissipation cannot be quantified. This can

    be alleviated by explicitly filtering the flow properties at each step as an integral part

    of the numerical simulation (Lund, 2003; Bose et al., 2010). In the present study, the

    explicit filtering technique is adopted in order to quantify the energy dissipation by

    the LES filter and assess the effects of filtering on the quality of the LES results. The

    filtering procedure adopted in this work is based on the approximate deconvolution

    model (ADM) (Stolz & Adams, 1999; Mathew et al., 2003; Bogey & Bailly, 2006) of 

    subgrid scales.

    In this chapter, the governing equations, the LES subgrid modeling through

    ADM, and some details on the numerical schemes and boundary conditions are pre-

    sented.

    2.1 Governing Equations

    The unsteady, non-dimensional, compressible form of the Navier-Stokes equa-

    tions were solved on curvilinear grids. The governing equations, in the generalized

    coordinates, (ξ , η , ζ  ), are given by

    1

    ∂ Q

    ∂t  +

      ∂ 

    ∂ξ 

    F − Fv

    +

      ∂ 

    ∂η

    G − Gv

    +

      ∂ 

    ∂ζ 

    H − Hv

     =

     S

    J  ,   (2.4)

    where,  Q   is the the vector of conserved variables,

    Q =  1

    J [ρ ρu ρv ρw E  ]T  ,   (2.5)

  • 8/18/2019 Dissertation Alireza

    53/248

    2.1. GOVERNING EQUATIONS 23

    and total energy is defined as

    E  =  p

    γ − 1 +

     1

    2

    ρuiui .   (2.6)

    The inviscid flux vectors are

    F =

    ρU 

    ρuU  +  ξ x p

    ρvU  +  ξ y p

    ρwU  +  ξ z p

    (E  + p)U 

    ,   G =

    ρV 

    ρuV   + ηx p

    ρvV   + ηy p

    ρwV   + ηz p

    (E  + p)V 

    ,   (2.7)

    and

    H =

    ρW 

    ρuW  + ζ x p

    ρvW  + ζ y p

    ρwW  + ζ z p

    (E  + p)W 

    ,   (2.8)

    where the variables  U ,  V , and  W  are the transformed velocity vectors in the com-

    putational space given by

    U  = uξ x + vξ y + wξ z ,   (2.9)

    V   = uηx + vηy + wηz ,   (2.10)

    and

    W   = uζ x + vζ y + wζ z .   (2.11)

  • 8/18/2019 Dissertation Alireza

    54/248

    2.1. GOVERNING EQUATIONS 24

    The viscous flux vectors are defined as follows:

    Fv  =

    F v1

    F v2

    F v3

    F v4

    F v5

    =

    ξ xΨxx + ξ yΨxy + ξ zΨxz

    ξ xΨyx + ξ yΨyy + ξ zΨyz

    ξ xΨzx  + ξ yΨzy  + ξ zΨzz

    uF v2 + vF v3 + wF v4 − ξ xq x − ξ yq y − ξ zq z

    ,   (2.12)

    Gv  =

    Gv1

    Gv2

    Gv3

    Gv4

    Gv5

    =

    ηxΨxx + ηyΨxy + ηzΨxz

    ηxΨyx + ηyΨyy  + ηzΨyz

    ηxΨzx + ηyΨzy  + ηzΨzz

    uGv2 + vGv3 + wGv4 − ηxq x − ηyq y − ηzq z

    ,   (2.13)

    and

    Hv  =

    H v1

    H v2

    H v3

    H v4

    H v5

    =

    ζ xΨxx + ζ yΨxy + ζ zΨxz

    ζ xΨyx + ζ yΨyy  + ζ zΨyz

    ζ xΨzx + ζ yΨzy  + ζ zΨzz

    uH v2 + vH v3 + wH v4 − ζ xq x − ζ yq y − ζ zq z

    .   (2.14)

    The shear stress tensor, Ψij  is given by

    Ψij  =  2µ

    Re

    S ij −  1

    3S kkδ ij

     ,   (2.15)

  • 8/18/2019 Dissertation Alireza

    55/248

    2.2. APPROXIMATE DECONVOLUTION MODEL 25

    where S ij  is the strain rate tensor defines as

    S ij  = 1

    2 ∂ui

    ∂x j+

     ∂ u j

    ∂xi  .   (2.16)The heat flux vector,  q i, is defined as

    q i =

      µ

    (γ − 1)M 2r RePr

     ∂ T 

    ∂xi,   (2.17)

    where M r   is the reference Mach number,

    M r  =  U r√ 

    γRT r.   (2.18)

    The Jacobian of coordinate transformation,  J , and transformation metrics are

    calculated in a conservative  form as outlined by Visbal & Gaitonde (2002). If present,

    the source terms are represented by the vector  S  on the right hand side of eq. (2.4).

    2.2 Explicit Filtering and Approximate Deconvolution Model

    Consider a one-dimensional transport equation of the form

    ∂u

    ∂t   +

     ∂ f (u)

    ∂x   = 0 ,   (2.19)

    where f (u) is a nonlinear flux function. Applying a low pass filter,  G, to eq. (2.19)

    yields

    ∂u

    ∂t  + G ∗  ∂ f (u)

    ∂x  = 0 ,   (2.20)

    where   G ∗   (.) denotes low pass filtering by convolution as outlined in eq. (2.1).Eq. (2.20) can be recast as

    ∂u

    ∂t  +

     ∂ f (u)

    ∂x  = Rsgs   (2.21)

  • 8/18/2019 Dissertation Alireza

    56/248

    2.2. APPROXIMATE DECONVOLUTION MODEL 26

    where the subgrid scale residual,

    Rsgs  =

     ∂f (u)

    ∂x   −G

    ∗ ∂ f (u)

    ∂x

      ,   (2.22)

    needs to be modeled. The subgrid model should describe the subgrid scale resid-

    ual, Rsgs  as a function of the filtered solution  u. The approximation deconvolutionmethod models Rsgs  with the following relation

    Rsgs  =  ∂f (u)∂x

      − G ∗ ∂ f (u∗)

    ∂x  ,   (2.23)

    where u∗(x, t) is an approximation of u(x, t) obtained through a deconvolution (Mathew

    et al., 2003),

    u u∗ = Q ∗ u .   (2.24)

    The deconvolution function,  Q, is supposed to be the exact inverse of  G; however, it

    is usually an approximation to the exact inverse of  G such that  QG  is unity for low

    wavenumbers.

    Substitution of eq. (2.24) into eq. (2.20) yields

    ∂u

    ∂t  + G ∗ ∂ f (u

    ∗)

    ∂x  = 0 ,   (2.25)

    which can be recast to give

    G ∗

    ∂u∗

    ∂t  +

     ∂ f (u∗)

    ∂x

     =  G ∗ ∂ u

    ∂t −  ∂ u

    ∂t  .   (2.26)

    Since G ∗ u∗ ≈ G ∗ u, the RHS of eq. (2.26) may be set equal to zero, i.e.

    G ∗∂u∗

    ∂t  +

     ∂ f (u∗)

    ∂x

     = 0.   (2.27)

  • 8/18/2019 Dissertation Alireza

    57/248

    2.3. SPATIAL DISCRETIZATION SCHEME 27

    which implies to

    ∂u∗

    ∂t  +

     ∂ f (u∗)

    ∂x  = 0 .   (2.28)

    Thus eq. (2.28) is solved instead of eq.’s (2.20) or (2.21). The numerical imple-

    mentation is as follows.

    Consider the filtered solution at time step   n  denoted by   u(n). The unfiltered

    solution is obtained from

    u∗ (n) = Q ∗ u(n) ,   (2.29)

    which is used to numerically integrate eq. (2.28) and obtain   u∗ (n+1). The filtered

    solution is then obtained from

    u(n+1) = G ∗ u∗ (n+1) .   (2.30)

    When executed sequentially, the filtering step, eq. (2.30), and the deconvolution

    step, eq. (2.29), can be combined to form a single filtering step,   Q ∗ G ∗ u∗ (n+1).Since Q  is not the exact inverse of  G, the operator  Q ∗ G removes high wavenumber

    components of the solution. In effect, the filter   H   =   Q ∗ G   is a low pass filtersimilar to   G, but with a higher cut-off frequency. In summary, the approximate

    deconvolution model is implemented through explicitly applying a low-pass filter

    during time integration.

    2.3 Spatial Discretization Scheme

    A sixth-order, non-dissipative, central difference compact scheme (Lele, 1992),

    α∂f ∂ξ i−1 + ∂f ∂ξ i + α∂f ∂ξ i−1 = a f i+1 − f i−12∆ξ    + b f i+2 − f i−24∆ξ    ,   (2.31)

  • 8/18/2019 Dissertation Alireza

    58/248

    2.4. SPATIAL FILTERING 28

    was used for spatial discretization where   α   = 1/3,   a   = 14/9, and   b   = 1/9. For

    the boundary points   i = 1 and   i =  N , the following third-order one-sided compact

    scheme equations were used, respectively:∂f 

    ∂ξ 

    1

    + 2

    ∂f 

    ∂ξ 

    2

    =  1

    2∆ξ  (−5f 1 + 4f 2 + f 3)  ,   (2.32)

    and ∂f 

    ∂ξ 

    + 2

    ∂f 

    ∂ξ 

    N −1

    =  1

    2∆ξ  (5f N  − 4f N −1 − f N −2)  .   (2.33)

    For points i  = 2 and i  =  N −1, the following forth-order, central difference, compact

    scheme equations were used, respectively:

    1

    4

    ∂f 

    ∂ξ 

    1

    +

    ∂f 

    ∂ξ 

    2

    + 1

    4

    ∂f 

    ∂ξ 

    3

    =  3

    4∆ξ  (f 3 − f 1)  ,   (2.34)

    and

    1

    4

    ∂f 

    ∂ξ 

    N −2

    +

    ∂f 

    ∂ξ 

    N −1

    + 1

    4

    ∂f 

    ∂ξ 

    =  3

    4∆ξ  (f N  − f N −2)  .   (2.35)

    The above equations form a tri-diagonal system of linear equations which can

    be efficiently solved using the Tridiagonal Matrix Algorithm (TDMA) (Conte &Boor, 1980). The combination of the sixth-order scheme for interior points and the

    third-order, one-sided scheme for boundary points results in a fourth-order global

    accuracy.

    2.4 Spatial Filtering

    Since the central difference compact scheme, eq. (2.31), is non-dissipative, spatial

    filtering of the solution is required at each time step to remove high wavenumber

    components and keep the solution stable. The filtering also serves as the ADM

    subgrid modeling for LES as described in Sec. 2.2.

  • 8/1