19.01.2017 Views

PDF-Modeling and Simulation of Turbulent Spray Combustion

resolver

resolver

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>PDF</strong>-<strong>Modeling</strong> <strong>and</strong> <strong>Simulation</strong> <strong>of</strong><br />

<strong>Turbulent</strong> <strong>Spray</strong> <strong>Combustion</strong><br />

Dissertation<br />

zur<br />

Erlangung des Grades<br />

Doktor–Ingenieur<br />

der<br />

Fakultät für Maschinenbau<br />

der Ruhr–Universität Bochum<br />

von<br />

Seung Jin Baik<br />

aus Seoul, Südkorea<br />

Bochum, 2016


Abstract<br />

In this dissertation, based on comprehensive studies <strong>of</strong> the two-phase <strong>PDF</strong> method, a<br />

two-dimensional reactive-spray code has been developed <strong>and</strong> applied. Given two-phase<br />

spray <strong>PDF</strong> transport equation is solved by monte-carlo particle method <strong>and</strong> corresponding<br />

models. The Langevin model is employed to model the gasphase velocity. The molecular<br />

mixing term is closed through mixing models, i.e., Interaction by Exchange with the<br />

Mean (IEM) model, Curls mixing model <strong>and</strong> Euclidian Minimum Spanning Tree (EMST)<br />

model. The exchange terms between two phase are modeled by evaporation models, i.e.,<br />

D 2 -law model <strong>and</strong> film model. Since we will select velocity model <strong>and</strong> mixing models that<br />

include the turbulent kinetic energy k <strong>and</strong> the turbulent dissipation rate ɛ, the st<strong>and</strong>ard<br />

k-ɛ turbulence model is employed.<br />

For discretization <strong>of</strong> the governing equations, the unstructured mesh <strong>and</strong> correspondingly<br />

formed control volumes are defined. In each control volume, the number <strong>of</strong> particles is<br />

monitored <strong>and</strong> controlled by particle cloning <strong>and</strong> annihilation scheme to keep the statistical<br />

errors approximately uniform in the overall computational domain.<br />

The two-dimensional reactive-spray code has been applied to three different experiments,<br />

i.e., a turbulent hydrogen-air flame, a non-premixed, turbulent methane-air counterflow<br />

flame with water droplets, <strong>and</strong> a turbulent ethanol-air spray flame. To reduce the computing<br />

costs, three different parallelization schemes, i.e., OpenMP, MPI, <strong>and</strong> the MPI-<br />

OpenMP hybrid scheme, has been introduced <strong>and</strong> applied to the code. The numerical<br />

results obtained in the present thesis show good agreement with experimental data available<br />

in the literature.<br />

Keywords: turbulent spray flow, Monte Carlo method, two-phase <strong>PDF</strong> method, parallelization<br />

i


Acknowledgements<br />

The present dissertation represents the culmination <strong>of</strong> my work at the chair <strong>of</strong> fluid<br />

mechanics <strong>of</strong> Ruhr-University Bochum.<br />

First <strong>and</strong> foremost I would like to thank my supervisor pr<strong>of</strong>essor Rogg for providing me<br />

with opportunity to work on such a great research team, <strong>and</strong> for his continues encouragement,<br />

motivation, <strong>and</strong> suggestions during my PhD time.<br />

Secondly, I would like to thank for my committee member pr<strong>of</strong>essor Scherer who friendly<br />

took the review <strong>of</strong> my dissertation <strong>and</strong> gave me many fruitful suggestions. I would also<br />

like to thank pr<strong>of</strong>essor Abramovici for being the head <strong>of</strong> my exam committee <strong>and</strong> for his<br />

questions <strong>and</strong> comments during my exam.<br />

I owe many thanks to all my colleagues in the group <strong>of</strong> Pr<strong>of</strong>. Rogg, who have given the<br />

pleasant research atmosphere <strong>and</strong> the countless support. I also give a special thanks to<br />

my friend Dr. Sang-Ho Na who has given me lots <strong>of</strong> comment <strong>and</strong> help for computation<br />

technical point <strong>of</strong> view.<br />

Last but not the least, I would like to thank my parents, family, <strong>and</strong> especially my wife<br />

Nalye Hong. Without her selfless support, great patience <strong>and</strong> constant encouragement,<br />

this work can not be finished.<br />

Finally, I acknowledge Deutsche Forschungsgemeinschaft (DFG) for the financial support<br />

through grant no RO2249/3-1.<br />

ii


Contents<br />

Abstract<br />

Acknowledgements<br />

ii<br />

iii<br />

1 Introduction 1<br />

1.1 Historical Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2<br />

1.2 Outline <strong>of</strong> the Present Thesis . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

2 Governing Equations for Two-Phase Flows 8<br />

2.1 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

2.2 Deterministic Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2.2.1 Gas-Phase Conservation Equations . . . . . . . . . . . . . . . . . . 9<br />

2.2.2 Liquid-Phase Conservation Equations . . . . . . . . . . . . . . . . . 10<br />

2.2.3 Phase Exchange Terms . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

2.2.4 Phase-Indicator Function . . . . . . . . . . . . . . . . . . . . . . . . 12<br />

2.2.5 Field Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14<br />

2.3 Probabilistic Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.3.1 Two-Phase Joint <strong>PDF</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 15<br />

2.3.2 Gas-Phase DF-Transport Equation . . . . . . . . . . . . . . . . . . 17<br />

2.3.3 Liquid-Phase DF-Transport Equation – Williams’ <strong>Spray</strong> Equation . 17<br />

2.4 Pressure-Correction Equation . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

3 Computational Approach 21<br />

3.1 Two-Phase Particle Method . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

3.1.1 Particle Method for Liquid Phase . . . . . . . . . . . . . . . . . . . 22<br />

3.1.2 Particle Method for Gas Phase . . . . . . . . . . . . . . . . . . . . 24<br />

3.2 Expectations <strong>and</strong> Moving Averages . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.2.1 Expectations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3.2.2 Moving Averages . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

4 Models 28<br />

iii


4.1 Velocity Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28<br />

4.2 Mixing Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

4.2.1 IEM-Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

4.2.2 Curl’s Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

4.2.3 EMST-Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

4.3 Evaporation Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

4.3.1 D 2 ´ law Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

4.3.2 Film Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

4.4 Turbulence Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

5 Discretization 39<br />

5.1 Domain <strong>of</strong> Integration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

5.2 Eulerian Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

5.2.1 Poisson Equation for Pressure . . . . . . . . . . . . . . . . . . . . . 41<br />

5.2.2 Turbulence Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

5.3 Lagrangian Particle Equations . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

5.4 Particle Tracing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

5.5 Particle Controlling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46<br />

5.6 Initial <strong>and</strong> Boundary Conditions . . . . . . . . . . . . . . . . . . . . . . . . 48<br />

5.6.1 Initial <strong>and</strong> Boundary Conditions for Particles . . . . . . . . . . . . 48<br />

5.6.2 Boundary Conditions for Eulerian Equations . . . . . . . . . . . . . 48<br />

5.7 Overall Solution Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

6 Parallelization 52<br />

6.1 Parallelization Architecture <strong>and</strong> Strategies . . . . . . . . . . . . . . . . . . 52<br />

6.1.1 Parallel Architecture . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

6.1.2 Distributed Memory System <strong>and</strong> Shared Memory System . . . . . . 54<br />

6.1.3 Parallelization Strategies . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

6.2 Domain <strong>and</strong> Particles Decomposition . . . . . . . . . . . . . . . . . . . . . 56<br />

6.2.1 Schur-Complement Method . . . . . . . . . . . . . . . . . . . . . . 57<br />

6.2.2 Particles Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

6.3 Functional Decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

6.4 MPI-OpenMP Hybrid Approach . . . . . . . . . . . . . . . . . . . . . . . . 66<br />

6.4.1 Performance <strong>and</strong> Limitation . . . . . . . . . . . . . . . . . . . . . . 67<br />

6.4.2 Hybrid Approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

6.5 Heterogeneous Computing with GPU . . . . . . . . . . . . . . . . . . . . . 69<br />

6.5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

6.5.2 GPU Acceleration with OpenCL . . . . . . . . . . . . . . . . . . . . 70<br />

iv


7 <strong>Turbulent</strong> Jet Diffusion Flames (DLR H3 Flame) 75<br />

7.1 Experimental <strong>and</strong> Computational Setup . . . . . . . . . . . . . . . . . . . 75<br />

7.2 Results <strong>and</strong> Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

7.3 Speedup by OpenMP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

8 <strong>Turbulent</strong> Counterflow Flames with Water Droplets 82<br />

8.1 Experimental <strong>and</strong> Computational Setup . . . . . . . . . . . . . . . . . . . 82<br />

8.2 Results <strong>and</strong> Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

8.3 Speedup by MPI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89<br />

9 <strong>Turbulent</strong> Jet Diffusion <strong>Spray</strong> Flames (Sydney flame) 90<br />

9.1 Experimental <strong>and</strong> Computational Setup . . . . . . . . . . . . . . . . . . . 90<br />

9.2 Results <strong>and</strong> Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

9.3 Speedup by the MPI-OpenMP Hybrid Approach . . . . . . . . . . . . . . . 97<br />

10 Conclusions <strong>and</strong> Perspectives 98<br />

Appendices 100<br />

A Derivation <strong>of</strong> the Eulerian Equations 100<br />

A.1 Pressure Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

A.2 <strong>Turbulent</strong> Kinetic Energy <strong>and</strong> Dissipation Rate . . . . . . . . . . . . . . . 103<br />

B Chemical Reaction Mechanism 107<br />

C Mixing Models 109<br />

Bibliography 114<br />

v


List <strong>of</strong> Figures<br />

2.1 Schematic diagram <strong>of</strong> two-phase spray. . . . . . . . . . . . . . . . . . . . . 12<br />

4.1 Euclidian minimum spanning tree in two dimension composition space. . . 32<br />

5.1 Section <strong>of</strong> an unstructured triangular mesh with control volumes. . . . . . 39<br />

5.2 An unstructured triangle with physical coordinates <strong>and</strong> the mapped into<br />

natural coordinate ξ <strong>and</strong> η. . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

5.3 A mapping to determine control volume which the particle is located. . . . 46<br />

5.4 Statistical error <strong>and</strong> speed down in a control volume as a function <strong>of</strong> the<br />

number <strong>of</strong> particles in a control volume. . . . . . . . . . . . . . . . . . . . 47<br />

5.5 Schematic <strong>of</strong> the fully stochastic particle method. . . . . . . . . . . . . . . 50<br />

5.6 Flowchart <strong>of</strong> the overall computation algorithm. . . . . . . . . . . . . . . . 51<br />

6.1 Classification <strong>of</strong> parallel computer architectures by Flynn [22]. . . . . . . . 53<br />

6.2 Schematic <strong>of</strong> distributed memory system. . . . . . . . . . . . . . . . . . . . 54<br />

6.3 Schematic <strong>of</strong> shared memory system. . . . . . . . . . . . . . . . . . . . . . 54<br />

6.4 An example <strong>of</strong> executing time segments. . . . . . . . . . . . . . . . . . . . 55<br />

6.5 Schematic <strong>of</strong> MPI. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

6.6 Schematic <strong>of</strong> domain decomposition. . . . . . . . . . . . . . . . . . . . . . 58<br />

6.7 Schematic <strong>of</strong> particle exchange. . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

6.8 Partial speedup by OpenMP on calculating chemical source term. . . . . . 66<br />

6.9 Partial speedup by OpenMP on expectation values. . . . . . . . . . . . . . 66<br />

6.10 Overall speedup by OpenMP <strong>and</strong> MPI on sprays flame. . . . . . . . . . . . 67<br />

6.11 Schematic <strong>of</strong> MPI-OpenMP hybrid approach. . . . . . . . . . . . . . . . . 68<br />

6.12 Comparing overall speedup between Pure OpenMP <strong>and</strong> Hybrid MPI-OpenMP<br />

(MPI 8 ˆ OpenMP 4). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

6.13 Schematic <strong>of</strong> CPU(left) <strong>and</strong> GPU(right). . . . . . . . . . . . . . . . . . . . 70<br />

6.14 Schematic <strong>of</strong> implementation GPU accelerating with OpenCL. . . . . . . . 71<br />

6.15 Sample calculation for premixed methane flame on 1D code. . . . . . . . . 72<br />

6.16 Comparing speedup <strong>of</strong> parallelized target functions between GPUs; 48 cores<br />

(left), 384 cores(center) <strong>and</strong> 24 cores(right). . . . . . . . . . . . . . . . . . 73<br />

vi


6.17 Schematic <strong>of</strong> MPI-OpenCL hybrid concept. . . . . . . . . . . . . . . . . . . 74<br />

7.1 Schematic <strong>of</strong> the DLR burner [56] <strong>and</strong> <strong>of</strong> the computational domain. . . . . 76<br />

7.2 Axial pr<strong>of</strong>ile <strong>of</strong> the mass fraction <strong>of</strong> H 2 along the centerline. Solid line:<br />

computational result; symbols: experimental data from [1, 21, 65]. . . . . . 76<br />

7.3 Axial pr<strong>of</strong>ile <strong>of</strong> the mass fraction <strong>of</strong> O 2 along the centerline. Solid line:<br />

computational result; symbols: experimental data from [1, 21, 65]. . . . . . 77<br />

7.4 Axial pr<strong>of</strong>ile <strong>of</strong> the mass fraction <strong>of</strong> H 2 O along the centerline. Solid line:<br />

computational result; symbols: experimental data from [1, 21, 65]. . . . . . 77<br />

7.5 Axial pr<strong>of</strong>ile <strong>of</strong> the mass fraction <strong>of</strong> OH along the centerline. Solid line:<br />

computational result; symbols: experimental data from [1, 21, 65]. . . . . . 78<br />

7.6 Axial pr<strong>of</strong>ile <strong>of</strong> temperature along the centerline. Solid line: computational<br />

result; symbols: experimental data from [1, 21, 65]. . . . . . . . . . . . . . 78<br />

7.7 Temperature as a function <strong>of</strong> mixture fraction at x{D “ 5. Red points:<br />

computational result; blue points: experimental data from [1, 21, 65]; line:<br />

chemical equilibrium from [1, 21, 65]. . . . . . . . . . . . . . . . . . . . . . 79<br />

7.8 Temperature as a function <strong>of</strong> mixture fraction at x{D “ 20. Red points:<br />

computational result; blue points: experimental data from [1, 21, 65]; line:<br />

chemical equilibrium from [1, 21, 65]. . . . . . . . . . . . . . . . . . . . . . 79<br />

7.9 Surface plots – from top to bottom – <strong>of</strong> the mass fraction <strong>of</strong> H 2 O, the mass<br />

fraction <strong>of</strong> OH, <strong>and</strong> the temperature, respectively. . . . . . . . . . . . . . . 80<br />

7.10 Marginal density function <strong>of</strong> the mass fraction <strong>of</strong> OH at position x{D “ 5. 81<br />

8.1 Schematic <strong>of</strong> the counterflow burner used in the experiments [107]. . . . . . 83<br />

8.2 Schematic <strong>of</strong> computational domain <strong>and</strong> boundaries. . . . . . . . . . . . . 84<br />

8.3 Pr<strong>of</strong>ile <strong>of</strong> the mean axial velocity component <strong>of</strong> water droplets along the<br />

symmetry line. Solid line: computational result; symbols: experimental<br />

data from [107]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

8.4 Radial pr<strong>of</strong>ile <strong>of</strong> the mean axial velocity component <strong>of</strong> water droplets at<br />

the axial location z “ 1.4 mm. Solid line: computational result; symbols:<br />

experimental data from [107]. . . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

8.5 Phase-plane identifying regimes <strong>of</strong> extinction <strong>and</strong> stable burning. Stars:<br />

computational results, other symbols: experimental data from [107]. . . . . 86<br />

8.6 Maximum mean flame temperature as a function <strong>of</strong> the water mass concentration<br />

Y d for inlet mole fractions X O2 “ 0.21 <strong>and</strong> X CH4 “ 0.23. . . . . . 87<br />

8.7 Surface plots <strong>of</strong> mean temperature with different Y d . A,C: without seeding<br />

<strong>of</strong> water drops; B: stable flame with Y d “ 0.1%; D: extinction with Y d “ 0.9%. 87<br />

8.8 Surface plot <strong>of</strong> mean temperature with streamlines for a case with Y d “ 0.1%. 88<br />

vii


9.1 Schematic <strong>of</strong> the spray burner used in the Sydney experiments [39, 55]. . . 91<br />

9.2 Schematic <strong>of</strong> computational domain <strong>and</strong> boundaries. . . . . . . . . . . . . 92<br />

9.3 Radial pr<strong>of</strong>ile <strong>of</strong> the mean axial velocity component <strong>of</strong> fuel droplets along<br />

at axial station z/D=10. Solid line: computational result; symbols: experimental<br />

data from [39, 55]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

9.4 Radial pr<strong>of</strong>ile <strong>of</strong> the mean axial velocity component <strong>of</strong> fuel droplets along<br />

at axial station z/D=20. Solid line: computational result; symbols: experimental<br />

data from [39, 55]. . . . . . . . . . . . . . . . . . . . . . . . . . . . 93<br />

9.5 Radial pr<strong>of</strong>ile <strong>of</strong> temperature at axial location z/D=10. Solid line: computational<br />

result; symbols: experimental data from [39, 55]. . . . . . . . . . . 93<br />

9.6 Radial pr<strong>of</strong>ile <strong>of</strong> temperature at axial location z/D=20. Solid line: computational<br />

result; symbols: experimental data from [39, 55]. . . . . . . . . . . 94<br />

9.7 Surface plots – from top to bottom – <strong>of</strong> the axial velocity component, the<br />

mass fraction <strong>of</strong> H 2 O, <strong>and</strong> the temperature, respectively. . . . . . . . . . . 95<br />

9.8 Surface plots – from top to bottom – <strong>of</strong> the axial velocity component, <strong>and</strong><br />

the radial velocity component, respectively. . . . . . . . . . . . . . . . . . . 95<br />

9.9 Pr<strong>of</strong>ile along the jet center line <strong>of</strong> the marginal density function f W pR l q as<br />

a function <strong>of</strong> droplets radius. . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

9.10 Droplets radius represented by the liquid Monte-Carlo particles along the<br />

symmetry line <strong>of</strong> the jet. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97<br />

A.1 Integrating points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102<br />

C.1 The convergence comparing between three mixing models. . . . . . . . . . 110<br />

C.2 The convergence shapes <strong>of</strong> IEM mixing model. . . . . . . . . . . . . . . . . 111<br />

C.3 The convergence shapes <strong>of</strong> modified Curl mixing model. . . . . . . . . . . . 112<br />

C.4 The convergence shapes <strong>of</strong> EMST mixing model. . . . . . . . . . . . . . . . 113<br />

viii


List <strong>of</strong> Tables<br />

8.1 Inlet composition for both the experiment [107] <strong>and</strong> the present computations. 82<br />

9.1 Properties <strong>of</strong> liquid ethanol. . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

B.1 Reduced chemical mechanism <strong>of</strong> methane [24]. The activation energies E a<br />

are in cal/mole <strong>and</strong> the pre-exponential constants in cgs units. . . . . . . 108<br />

ix


Nomenclature<br />

Roman Symbols<br />

B M<br />

B T<br />

C D<br />

Spalding mass transfer number<br />

Spalding heat transfer number<br />

drag coefficient<br />

c p specic heat capacity [J/kg K]<br />

D<br />

F D<br />

diffusion coefficient [m 2 /s]<br />

drag force<br />

ɛ dissipation rate <strong>of</strong> turbulent kinetic energy [m 2 {s 3 ]<br />

K<br />

number <strong>of</strong> species<br />

k turbulent kinetic energy [m 2 {s 2 ]<br />

m<br />

n l<br />

P<br />

p<br />

p<br />

P g<br />

P l<br />

R<br />

S<br />

particle mass [kg]<br />

droplet number density<br />

number <strong>of</strong> particles<br />

particle index<br />

pressure [Pa]<br />

number <strong>of</strong> gasphase particles<br />

number <strong>of</strong> liquid phase particles<br />

radius <strong>of</strong> particle [m]<br />

source term<br />

x


T<br />

t<br />

u<br />

V<br />

v<br />

V k<br />

w<br />

w α<br />

X α<br />

Y α<br />

Y d<br />

Nu<br />

P r<br />

Re<br />

Sc<br />

Sh<br />

temperature [K]<br />

time [s]<br />

velocity component along x [m/s]<br />

velocity [m/s]<br />

velocity component along y or r [m/s]<br />

kth control volume<br />

velocity component along z [m/s]<br />

reaction rate <strong>of</strong> species α<br />

mole fraction <strong>of</strong> species α<br />

mass fraction <strong>of</strong> species α<br />

mass concentration <strong>of</strong> water droplets<br />

Nusselt number<br />

Pr<strong>and</strong>tl number<br />

Reynolds number<br />

Schmidt number<br />

Sherwood number<br />

Greek Symbols<br />

α<br />

Γ<br />

index <strong>of</strong> species<br />

interface between two control volumes<br />

λ thermal conductivity [J/m ¨ s ¨ K]<br />

µ viscosity [kg/m ¨ s]<br />

Ω<br />

domain<br />

ρ density [kg/m 3 ]<br />

V<br />

finite volume<br />

xi


φ<br />

ψ<br />

general scalar variable<br />

scalar variable – for <strong>PDF</strong> transport equations<br />

Mathematical Symbols<br />

x P A x is an element <strong>of</strong> A<br />

x :“ A x is defined as A<br />

Superscript<br />

ˆ<br />

sample properties<br />

9 time rate<br />

r<br />

Favre average<br />

Subscript<br />

e<br />

g<br />

l<br />

m<br />

v<br />

energy related<br />

gasphase<br />

liquid phase<br />

mass related<br />

momentum related<br />

Abbreviations<br />

CFD<br />

CFL<br />

Computational Fluid Dynamics<br />

Courant–Friedrichs–Lewy condition<br />

GPGPU General-Purpose Computing on Graphics Processing Unit<br />

LHF<br />

Locally Homogeneous Flow models<br />

CDM Continuous Droplet Model<br />

CFM Continuum Formulation Model<br />

DDM Discrete Droplet Model<br />

DF<br />

Density Function<br />

xii


DNS<br />

DSF<br />

LES<br />

MPI<br />

<strong>PDF</strong><br />

Direct Numerical <strong>Simulation</strong><br />

Deterministic Separated Flow<br />

Large Eddy <strong>Simulation</strong><br />

Message Passing Interface<br />

Probability Density Function<br />

RANS Reynolds Averaged Navier-Stokes<br />

SSF<br />

Stochastic Separated Flow<br />

CUDA Compute Unied Device Architecture<br />

OpenCL Open Computing Language<br />

OpenMP Open Multi-Processing<br />

SF<br />

Separated Flow models<br />

xiii


Chapter 1<br />

Introduction<br />

Notwithst<strong>and</strong>ing that environmental contamination due to pollutants emissions through<br />

combustion has been on the rise for several decades, fossil fuels still cover more than 85<br />

percent <strong>of</strong> the world’s total energy resource requirement. Nuclear resources, which were<br />

considered a substitute for fossil fuels, have continuously shown to give rise to several<br />

critical risks such as radiation leaks due to the disposal <strong>of</strong> the nuclear fuel. After the<br />

Fukushima Daiichi nuclear disaster in Japan in 2011, radioactivity there still is released<br />

into the sea <strong>and</strong> into the air, <strong>and</strong> hence – as a consequence <strong>of</strong> new political aims – the<br />

proportion <strong>of</strong> nuclear systems is in decline. Renewable energy resources from nature<br />

are, definitely, the way to proceed in the future, however their exploitation is not yet<br />

sufficiently developed, neither in terms <strong>of</strong> economy nor in terms <strong>of</strong> efficiency. Therefore,<br />

usage <strong>of</strong> fossil fuels will be unavoidable for decades, <strong>and</strong> hence it is necessary to optimize<br />

the implementation <strong>of</strong> those fuels to be both efficient in energy generation <strong>and</strong> with respect<br />

to environmental aspects.<br />

<strong>Combustion</strong>, however, is a complex phenomenon as it is couples chemistry, fluid dynamics,<br />

<strong>and</strong> heat <strong>and</strong> mass transfer. Moreover, combustion also contains many phenomena that<br />

up to date have not been fully investigated. In contrast to laminar combustion, turbulent<br />

combustion is even more complex due to the inherently large number <strong>of</strong> turbulent time <strong>and</strong><br />

length scales. Further complexities arise through complicated geometries <strong>of</strong> combustion<br />

chambers, say. Consequently, the analysis <strong>of</strong> combustion phenomena requires a deep<br />

theoretical underst<strong>and</strong>ing <strong>of</strong> the fundamental physical processes <strong>and</strong> the development <strong>of</strong><br />

both physical <strong>and</strong> simulation methods that are particularly powerful.<br />

1


1.1 Historical Background<br />

<strong>Spray</strong> combustion plays an important role in many combustion systems such as furnaces,<br />

diesel engines, gas turbines etc. For instance, in internal combustion engines, an implementation<br />

<strong>of</strong> the concept <strong>of</strong> feeding liquid fuels in the form <strong>of</strong> a spray is well known<br />

to result in increased combustion efficiency <strong>and</strong> reduced pollutant formation. Although<br />

combustion <strong>of</strong> liquid fuel sprays up to date has been widely used in praxis, it is still a<br />

challenging field because it inherits all the difficulties from gaseous combustion, <strong>and</strong>, in<br />

addition, encompasses injection <strong>of</strong> liquid fuel, evaporation <strong>of</strong> atomized fuel droplets, as<br />

well as fluid dynamical interaction <strong>and</strong> heat transfer between the liquid <strong>and</strong> the gaseous<br />

phase.<br />

A spray, reacting or non-reacting, can be defined as a special case <strong>of</strong> two-phase flow, that<br />

is composed <strong>of</strong> a dispersed liquid phase in the form <strong>of</strong> droplets or ligaments <strong>and</strong> a continuous,<br />

gaseous or carrier phase [45, 86, 97]. After injection through a nozzle into a gaseous<br />

environment, the liquid fuel splits up into many droplets which are <strong>of</strong> variable size <strong>and</strong>,<br />

have variable velocity as a consequence <strong>of</strong> the atomization process. During the injection<br />

process, depending on the environmental conditions, droplets break up continuously or<br />

collide with one or several other droplets, or with a wall. Depending on the underlying<br />

atomization process, sprays flow can be divided into dense sprays, which are characterized<br />

by a large liquid volume fraction near the exit <strong>of</strong> a nozzle, <strong>and</strong> dilute sprays, which have<br />

a much smaller liquid volume fraction there. After atomization or even during that process,<br />

droplets evaporate into the gaseous flow, <strong>and</strong> chemical reaction occurs when proper<br />

ignition conditions are met. Even though natural spray flow contain variable interactions<br />

between the liquid <strong>and</strong> the gaseous phase, as well between the droplets themselves such as<br />

break-up <strong>and</strong> collision, many previous researches assumed that spray flow has been fully<br />

diluted to circumvent the difficulties <strong>of</strong> spray analyse under dense condition.<br />

Existing numerical methods to solve dilute-sprays problems can be widely separated into<br />

two categories as Locally Homogeneous Flow models (LHF) [17, 18, 19, 45, 97] <strong>and</strong> Separated<br />

Flow models (SF) [45, 84, 85, 86, 97]. LHF models are based on the fundamental<br />

assumption that the flow field locally exists as a homogeneous mixture <strong>of</strong> liquid <strong>and</strong> gas.<br />

Under this assumption, both phases have the same velocity <strong>and</strong> temperature everywhere<br />

in the flow field. The main advantage <strong>of</strong> LHF models is that they require relatively little<br />

information on initial conditions at the injection nozzle, because initial droplet size <strong>and</strong><br />

velocity distribution play no role in the computations. Moreover, since computations for<br />

sprays flow with LHF models are basically identical to single-phase flow computation,<br />

traditional solution methods <strong>and</strong> codes can be used. LHF models, however, can be ap-<br />

2


plied only when the spray flow consists <strong>of</strong> infinitely small droplets because the error due<br />

to fundamental assumptions underlying these models cannot be acceptable [17, 18, 19].<br />

In separated flow or SF models, spray flow is separated basically into a liquid phase (dispersed<br />

phase) <strong>and</strong> a gasphase (carrier phase). Therefore, finite-rate exchange <strong>of</strong> mass,<br />

momentum <strong>and</strong> energy between the two phases can be implemented. Generally, the SF<br />

approach for reacting spray flow can be classified into three sub-models, viz., the Continuum<br />

Formulation Model (CFM), the Continuous Droplet Model (CDM) <strong>and</strong> the Discrete-<br />

Droplet Model (DDM) [17, 45, 84, 85, 86, 97]. CFM employs a continuum formulation <strong>of</strong><br />

the conservation equations for both phases <strong>and</strong> also for evaporation. Since the governing<br />

equations for both phases are similar, numerical implementation is relatively efficient;<br />

however, this formulation has drawbacks for establishing some critical characteristics <strong>of</strong><br />

reacting spray flow such as effects <strong>of</strong> droplet heat-up, turbulent stresses <strong>and</strong> its transport.<br />

The second sub-model in the separated flow approach is the Continuous Droplet Model<br />

(CDM) as described by Williams [45, 97]. In this model, the properties (e.g. radius,<br />

position, velocity, concentration <strong>and</strong> temperature) <strong>of</strong> the liquid phase are described by<br />

a statistical distribution function, the so-called Williams spray equation. The governing<br />

equations for the gasphase involve suitable source terms for interaction effects due to the<br />

presence <strong>of</strong> droplets in the flow. However, most combustion studies have not allowed for<br />

a continuous droplet distribution function. Instead, a finite number <strong>of</strong> different, discrete<br />

droplets has been employed in seeking computer solutions <strong>of</strong> the conservation equations<br />

[45, 97]. In the last model, called the Discrete-Droplet Model (DDM), the liquid phase is<br />

divided into representative samples <strong>of</strong> discrete droplets whose motion <strong>and</strong> transport are<br />

tracked through the flow field based on Lagrangian formulations. This procedure corresponds<br />

to a statistical approach, the so-called Monte Carlo Method, for liquid properties<br />

because a finite number <strong>of</strong> particles is used to represent the entire liquid phase. In fact,<br />

CDM <strong>and</strong> DDM only differ in their approach to the liquid phase in the spray flow. If continuous<br />

distribution or density functions are approximated by a finite number <strong>of</strong> sample<br />

particles in a finite volume, both methods yield approximately identical results [16, 19].<br />

As mentioned above, the liquid or dispersed phase can be subjected to various different,<br />

alternative models to describe the characteristics <strong>of</strong> the droplets. On the other h<strong>and</strong>, the<br />

gaseous carrier phase <strong>of</strong> a reactive turbulent spray flow can, <strong>of</strong>ten, be described sufficiently<br />

accurately by the usual governing equations <strong>of</strong> continuum fluid mechanics with added<br />

source terms that describe the interaction between the droplets <strong>and</strong> the carrier phase.<br />

Direct numerical simulation (DNS), large eddy simulation (LES), <strong>and</strong> simulations based<br />

on Reynolds Averaged Navier-Stokes (RANS) equations for the carrier phase, combined<br />

3


with a probability density function (<strong>PDF</strong>) method to cater for chemistry are widely used<br />

to obtain numerical solutions for the continuous carrier phase.<br />

The respective status <strong>of</strong> studies on spray combustion, both in theory <strong>and</strong> application, has<br />

been reported in [2, 17, 18, 19, 84, 85, 86, 97]. Fundamental models for combustion <strong>of</strong><br />

individual droplets were established in the 1950s by Godsave [27], Spalding [88] <strong>and</strong> Goldsmith<br />

[28] under ideal conditions, e.g, assuming spherically symmetric <strong>of</strong> the droplets. In<br />

the early stages <strong>of</strong> droplet <strong>and</strong> spray research, experiments <strong>and</strong> investigations on single<br />

droplets helped to underst<strong>and</strong> single droplet combustion. For instance, Godsave determined<br />

experimentally the burning rate for benzene, ethyl alcohol, <strong>and</strong> heptane drops<br />

under atmospheric pressure, <strong>and</strong> estimated the diameter <strong>of</strong> flame surrounding the liquid<br />

drop [27]. Spalding [88] introduced the so-called mass <strong>and</strong> heat transfer number B to<br />

express the fluxes <strong>of</strong> mass <strong>and</strong> heat, respectively, at the droplet surface. <strong>Modeling</strong> <strong>of</strong><br />

multi-component fuels began with the work reported by L<strong>and</strong>is <strong>and</strong> Mills [76] in 1970s.<br />

They studied the evaporation <strong>of</strong> heptane-octane two-component drops by means <strong>of</strong> a<br />

spherically symmetric model. Sirignano <strong>and</strong> Law [87] focused on internal heat, mass,<br />

<strong>and</strong> momentum transport <strong>and</strong> other effects related to evaporation. The extreme cases<br />

are the rapid regression or zero-diffusivity limit [49], that were interestedly investigated<br />

in the early 1980s. In [49], the classical evaporation model, so-called D 2 -law [27, 97],<br />

for droplet vaporization <strong>and</strong> combustion which is formulated for a single-component fuel,<br />

was extended for multi-component fuels. Some noteworthy reviews <strong>of</strong> turbulent spray<br />

flows have been reported by Faeth [17] in late 1980s. He compared three types <strong>of</strong> twophase<br />

models for turbulent flows, i.e., locally homogeneous flow (LHF), deterministic<br />

separated flow (DSF), <strong>and</strong> stochastic separated flow (SSF). He recommended the SSF<br />

model for practical dilute sprays. This model considers finite interphase transport rates<br />

<strong>and</strong> uses r<strong>and</strong>om-walk computations to simulate turbulent dispersion for the dispersed<br />

phase. Crowe et al. [9] reviewed both time-averaged <strong>and</strong> time-dependent free-shear flows<br />

<strong>of</strong> two-phase flows. Their time-dependent numerical methods captured the instantaneous<br />

flow <strong>and</strong> let to better prediction <strong>of</strong> the particle trajectories. In the time-averaged method,<br />

a steady flow with gradient diffusion is usually considered. In 1990s, Crowe et al. [10]<br />

reviewed computational approaches to turbulent two-phase flows.<br />

In the early 2000s, Zhu et al. [108] derived a two-phase <strong>PDF</strong> transport equation encompassing<br />

all gaseous phase <strong>and</strong> liquid phase r<strong>and</strong>om variables. A phase-indicator function<br />

[42] was used to capture the interface <strong>of</strong> the multiphase flow. Rumberg <strong>and</strong> Rogg [79, 80]<br />

suggested to describe the multiphase flow with two separate density function. Both phase<br />

density functions <strong>and</strong> their transport equations were defined in the overall two-phase flow<br />

field. Effects <strong>of</strong> interfacial surface interaction, including heat <strong>and</strong> mass transfer, on the<br />

4


overall flow were taken into account by source terms in the transport equation. A few<br />

years later, Naud [60] adopted a joint velocity-composition <strong>PDF</strong> for the continuous phase<br />

<strong>and</strong> the joint <strong>PDF</strong> <strong>of</strong> droplet velocity, droplet temperature, droplet mean gas velocity,<br />

<strong>and</strong> droplet mean gas composition for dispersed phase. After the year 2000, the stochastic<br />

separated-flow (SSF) model [29, 45, 97], which considers the effects <strong>of</strong> turbulent fluctuations<br />

on particle motion, based on the Discrete-Droplet Model or DDM has been widely<br />

employed to numerically simulate turbulent reactive sprays [12, 33, 37, 95].<br />

Following Rumberg <strong>and</strong> Rogg [79, 80], dissimilar to the traditional models CDM <strong>and</strong><br />

DDM, in the present work the gasphase is described by a <strong>PDF</strong> transport equation rather<br />

than by Eulerian equations. Since the <strong>PDF</strong> method has the main advantage that chemical<br />

source terms are closed <strong>and</strong> hence do not require modelling, it is widely used to solve<br />

reacting flow problems. Lundgren [54] studied nonhomogeneous turbulent flow by solving<br />

the velocity <strong>PDF</strong> transport equation. Although the application was limited initially to<br />

simple inert gaseous flow, that is the first well-known application for implementing the<br />

<strong>PDF</strong> approach to turbulent flow. After the <strong>PDF</strong> concept was extended to the reactive<br />

flow problems by Hill [34], Dopazo <strong>and</strong> O’Brien [13] also solved for composition variables,<br />

which is a set <strong>of</strong> scalar quantities, usually encompassing mass fractions <strong>and</strong> temperature,<br />

joint <strong>PDF</strong> transport equation for describing turbulent reactive flows.<br />

The relationship between Monte Carlo particle methods <strong>and</strong> <strong>PDF</strong> methods was worked<br />

out by Pope [67, 68], through whom’s work, Monte Carlo particle methods have become<br />

the key approach to solving <strong>PDF</strong> transport equations. Since Pope’s review article in 1985<br />

[69], numerous applications <strong>of</strong> the <strong>PDF</strong> method to analyse turbulent reactive flow have<br />

contributed to the popularity <strong>of</strong> the <strong>PDF</strong> concept. The Monte-Carlo framework needs<br />

a reasonably large number <strong>of</strong> statistical particles, that means it causes relatively high<br />

computing costs. During the development <strong>of</strong> <strong>PDF</strong> methods, to mitigate the computing<br />

costs, a combination <strong>of</strong> the <strong>PDF</strong> method <strong>and</strong> the traditional Eulerian approach, e.g.,<br />

simulations based on Reynolds Averaged Navier-Stokes (RANS), or, large eddy simulation<br />

(LES) have been introduced [4, 30, 59]. Early steps to implement the hybrid concept were<br />

taken by An<strong>and</strong> [4] in 1989. However, the early work reported that the hybrid concept<br />

had poor efficiency, robustness, <strong>and</strong> accuracy due to lacking numerical consistency [30].<br />

Since the successful resolution <strong>of</strong> the consistency problem by Muradoglu [59], the hybrid<br />

concept has been widely implemented for a number <strong>of</strong> flows [30, 59, 74, 78, 92, 105].<br />

Overview <strong>of</strong> the <strong>PDF</strong> method relating to turbulent single-phase combustion can be found<br />

in [13, 30, 43].<br />

For turbulent spray combustion, the <strong>PDF</strong> methods play an important role. From a<br />

conceptional point <strong>of</strong> view, methods for reacting spray flows can be broadly separated<br />

5


into two categories. These categories can be identified by employing the <strong>PDF</strong> method for<br />

only the dispersed phase or for both the dispersed phase <strong>and</strong> the carrier phase. Approach<br />

<strong>of</strong> the first category is similar to employing the Discrete-Droplet Model (DDM) which uses<br />

William spray equations for modelling <strong>of</strong> the droplet phase <strong>and</strong> the traditional Eulerian<br />

approach for the carrier phase [17, 45, 84, 85, 86, 97]. Approaches <strong>of</strong> the second category,<br />

in which Monte-Carlo particles are used for either phase, can be subdivided further into<br />

two classes. To the first class belong all approaches that, in a so-called fractional step,<br />

use gaseous particles exclusively to describe or model the combustion chemistry but use<br />

traditional formulations such as RANS or LES for the fractional solution step regarding<br />

the accumulative, convective <strong>and</strong> diffusive terms as well as the phase-interaction terms in<br />

the Eulerian governing gasphase equations for overall mass, species mass, momentum <strong>and</strong><br />

energy. Examples are the formulations in, e.g., [12, 26, 60]. To the second class belong<br />

all approaches that use a full <strong>PDF</strong> description for the overall combustion system such as<br />

the approach taken in the present thesis, <strong>and</strong> the approaches presented in [79, 80, 108].<br />

Several models or techniques, respectively, have been developed that help reducing the<br />

costs for turbulent, purely gaseous flames or, alternatively, reactive sprays computed by<br />

<strong>PDF</strong> methods. For instance, so-called flamelet models were developed according to which<br />

the local <strong>and</strong> instantaneous combustion is computed a-priori to the computation <strong>of</strong> the<br />

turbulent reactive flow [50, 64, 97]. Another possibility to reduce computing costs in the<br />

application <strong>of</strong> <strong>PDF</strong>-methods is to employ the so-called in-situ adaptive tabulation (ISAT)<br />

developed by Pope [72]. Using this kind <strong>of</strong> tabulation, for mots particles the composition<br />

variables are obtained by interpolation/extrapolation in a binary-tree database based on<br />

previously stored solutions. In the present work, apart from the above techniques, the<br />

computing costs will be reduced further by applying suitable parallelization schemes.<br />

1.2 Outline <strong>of</strong> the Present Thesis<br />

In the present work, turbulent spray combustion has been modelled <strong>and</strong> – for spatially<br />

two-dimensional geometries – simulated by a full <strong>PDF</strong> approach. Here the adjective full<br />

points to the fact that not only the droplet phase but also the gasphase was subjected to<br />

a full joint-<strong>PDF</strong> formulation that encompasses both phases.<br />

Based on comprehensive studies <strong>of</strong> the two-phase <strong>PDF</strong> method carried out in the framework<br />

<strong>of</strong> the present research, a two-dimensional reactive-spray code has been developed<br />

<strong>and</strong> applied. The remainder <strong>of</strong> the present thesis is structured as follows.<br />

In Chapter 2, the full, overall two-phase <strong>PDF</strong> transport equation encompassing both the<br />

6


gaseous <strong>and</strong> the liquid phase as well as marginal <strong>PDF</strong> transport equations for either spray<br />

phase are derived from the general two-phase governing equations.<br />

In Chapter 3, the Monte-Carlo particle method is introduced for either phases. In Chapter<br />

4, models are presented <strong>and</strong> discussed that are suitable for closure <strong>of</strong> the various originally<br />

unclosed terms in the governing equations presented in Chap. 2.<br />

In Chapter 5, the unstructured mesh <strong>and</strong> correspondingly formed control volumes are<br />

defined to enable discretization <strong>of</strong> the governing equations. To keep the statistical errors<br />

approximately uniform in the overall computational domain, the number <strong>of</strong> liquid <strong>and</strong><br />

gaseous Monte-Carlo particles in each control volume is monitored <strong>and</strong> controlled by<br />

suitable methods.<br />

In Chapter 6, the two-dimensional reactive-spray code is parallelized by applying the<br />

two parallelization methods MPI <strong>and</strong> OpenMP, respectively. 1 An MPI-OpenMP hybrid<br />

scheme is suggested, <strong>and</strong> also implemented in the code. Furthermore, the so-called CPU-<br />

GPU hybrid computing approach 2 is tested to estimate how this so-called heterogeneous<br />

approach to computing can be helpful in applications <strong>of</strong> computational fluid dynamics or<br />

CFD.<br />

In Chaps. 7 to 9, the computer code is applied to three different turbulent problems.<br />

Specifically, in Chap. 7, a turbulent, purely gaseous, hydrogen-air flame is computed to<br />

validate the code, in the first instance, for pure gasphase combustion.<br />

In Chaps. 8 <strong>and</strong> 9, turbulent sprays are computed. Specifically, in Chap. 8 the computer<br />

code is applied to a non-premixed, turbulent methane-air counterflow flame with a water<br />

mist added to the oxidizer jet. In this study, also the extinction behavior <strong>of</strong> the flame, in<br />

dependence <strong>of</strong> the amount <strong>of</strong> water added, is studied.<br />

In Chap. 9, the computer code is applied to a turbulent ethanol-air spray flame in which<br />

the fuel, ethanol, is provided in liquid form.<br />

Finally, in Chap. 10, a summary summary <strong>of</strong> the research work <strong>and</strong> results is given, an<br />

<strong>and</strong> possibly future work on the research subject is outline.<br />

At the end <strong>of</strong> the thesis, three appendices provide additional information.<br />

1<br />

MPI st<strong>and</strong>s for Message Passing Interface, OpenMP for Open Multi-Processing.<br />

2<br />

CPU st<strong>and</strong>s for Central Processing Unit, GPU for Graphics Processing Unit.<br />

7


Chapter 2<br />

Governing Equations for Two-Phase<br />

Flows<br />

2.1 Notation<br />

In the present work, two-phase flows <strong>and</strong> flames are considered that are special in that<br />

they comprise a gasphase (phase 2 or phase g) <strong>and</strong> a liquid phase (phase 1 or phase l).<br />

Specifically sprays are considered, i.e., the gasphase is the so-called carrier phase in which<br />

a huge number <strong>of</strong> small liquid drops – also called droplets – are carried.<br />

Drops are taken as single component drops. The gasphase is taken as a multicomponent,<br />

chemically reactive ideal-gas mixture. In particular, the gasphase comprises K g chemical<br />

species that participate in I chemical reactions. The mass fractions are summarized in a<br />

vector pY 1 , ..., Y K q.<br />

Independent variables are the time, t, <strong>and</strong> the the three components x i <strong>of</strong> the position<br />

vector x “ px 1 , x 2 , x 3 q.<br />

Be the aggregate state either gaseous or liquid, a mass density will be denoted by ρ, <strong>and</strong><br />

a velocity component by V i , i “ 1, 2, 3. Generally, tensor notation will be used for vectors<br />

<strong>and</strong> 2nd-order tensors. Temperature will be denoted by T .<br />

In the entire formulation, effects <strong>of</strong> gravity will be neglected.<br />

8


2.2 Deterministic Formulation<br />

2.2.1 Gas-Phase Conservation Equations<br />

Gasphase overall-mass continuity is governed by<br />

Bρ g<br />

Bt ` B`ρ g V g,i˘<br />

Bx i<br />

“ S m , (2.1)<br />

where S m represents the mass per unit volume <strong>and</strong> time gained by the gasphase from the<br />

liquid phase, or vice versa.<br />

Gasphase species-mass conservation is<br />

ρ g<br />

DY g,α<br />

Dt<br />

“ ´Bj g,α,i<br />

Bx i<br />

` w g,α ` S m,α , (2.2)<br />

where α “ 1, ..., K g . In (2.2) <strong>and</strong> below, j α,i represents the diffusion flux <strong>of</strong> chemical<br />

component α in the i-th coordinate direction, w g,α its net rate <strong>of</strong> production due to<br />

chemical reaction, <strong>and</strong> S m,α its mass per unit volume <strong>and</strong> time gained due to evaporation,<br />

or lost by condensation.<br />

Gasphase-momentum conservation is<br />

ρ g<br />

DV g,j<br />

Dt<br />

“ ´ Bp<br />

Bx i<br />

` Bτ g,ij<br />

Bx j<br />

` S v , (2.3)<br />

where S v denotes the momentum exchange term described in detail further below, in<br />

Chap. 2.2.3, <strong>and</strong> where p <strong>and</strong> τ g,ij are the pressure <strong>and</strong> the viscous part <strong>of</strong> the gas-phase<br />

stress tensor, respectively.<br />

Gasphase-energy conservation can be written as<br />

c g,p ρ DT g<br />

Dt “ ´Bq g<br />

Bx i<br />

´<br />

K<br />

ÿ g<br />

α“1<br />

h g,α w g,α ` S e , (2.4)<br />

where S e denotes the energy exchange term described in detail further below, in Chap.<br />

2.2.3, <strong>and</strong> where h g,α <strong>and</strong> w g,α denote the specific enthalpy <strong>and</strong> mass rate <strong>of</strong> production,<br />

respectively, <strong>of</strong> species α. Dufour effect <strong>and</strong> radiative heat loss [45, 97] are assumed to be<br />

negligibly small. As a consequence, the heat flux q g can be expressed as<br />

q g,i “ ´λ g<br />

BT g<br />

Bx i<br />

`<br />

9<br />

K<br />

ÿ g<br />

α“1<br />

h g,α j g,α,i , (2.5)


with – assuming Fick’s law – j g,α “ ´ρ g D g,α ∇Y g,α . Note that the second term on the<br />

r.h.s. <strong>of</strong> (2.5) represents the energy transported by species mass diffusion.<br />

2.2.2 Liquid-Phase Conservation Equations<br />

Liquid phase overall-mass continuity is governed by<br />

Bρ l<br />

Bt ` B`ρ l V l,i˘<br />

Bx i<br />

“ ´S m , (2.6)<br />

where the r.h.s term represents the mass lost per unit volume an time to the gas phase<br />

due to the evaporation.<br />

Liquid phase species-mass conservation equation is<br />

ρ l<br />

DY l,α<br />

Dt<br />

“ ´ B<br />

Bx i<br />

pρ l D l,α q BY l,α<br />

Bx i<br />

´ S m,α . (2.7)<br />

Liquid phase-momentum conservation is<br />

ρ l<br />

DV l,i<br />

Dt<br />

“ ´ Bp<br />

Bx i<br />

` Bτ l,ij<br />

Bx j<br />

´ S v , (2.8)<br />

where the last term <strong>of</strong> r.h.s represents the momentum lost due to evaporation per unit<br />

volume <strong>and</strong> time <strong>and</strong> F D represents the drag force per unit volume <strong>and</strong> time exerted from<br />

the liquid phase on the gas phase derived as Eq. (2.15).<br />

Liquid phase-energy conservation can be written as<br />

c p,l ρ l<br />

DT l<br />

Dt “ ´ Bq l<br />

Bx i<br />

´ S e , (2.9)<br />

where q l denotes the heat flux for liquid phase <strong>and</strong> S e is expressed in Eq.(2.16).<br />

2.2.3 Phase Exchange Terms<br />

After having presented <strong>and</strong> defined in Chaps. 2.2.1 <strong>and</strong> 2.2.2 the governing equations<br />

for a general two-phase flow, respectively, <strong>and</strong> the associated physical quantities, we now<br />

specify the phase exchange terms S m , S v <strong>and</strong> S e for the special case that the two-phase<br />

flow is a spray.<br />

10


The exchange <strong>of</strong> overall mass between the two phases, S m , is is related to the exchange<br />

<strong>of</strong> species mass, S m,α , by<br />

S m :“<br />

Kÿ<br />

S m,α , (2.10)<br />

α“1<br />

where – for species-mass exchange – S m,α is given by<br />

S m,α “ 9m α<br />

V . (2.11)<br />

Here, <strong>and</strong> below, V denotes a finite volume that is sufficiently small for a differential<br />

formulation <strong>of</strong> the underlying process – here overall mass conservation – to be applied.<br />

The quantities 9m α denote the mass gained per unit time by the gasphase from the liquid<br />

component α, or vice versa, in the finite volume V. Specifically, if 9m α ą 0, evaporation<br />

<strong>of</strong> species α takes place at the liquid surface, <strong>and</strong> if 9m α ă 0, condensation.<br />

For the special case <strong>of</strong> a single-component liquid fuel – this is the case considered in the<br />

present thesis – the overall mass per unit volume <strong>and</strong> time gained by the gasphase from<br />

the liquid phase can be expressed as<br />

S m “ 9m V “ ´ 1<br />

V<br />

ÿN l<br />

d“1<br />

9m d . (2.12)<br />

Here 9m denotes the mass gained per unit time by the gasphase from the liquid phase in<br />

the finite volume V. In the second equality <strong>of</strong> (2.12), <strong>and</strong> below, N l denotes the number<br />

<strong>of</strong> real, physical drops in volume V, i.e.,<br />

N l “ n l V (2.13)<br />

where n l denote the local <strong>and</strong> instantaneous droplet number density.<br />

Momentum exchange between the two phase is described by the exchange term S v which<br />

is the sum <strong>of</strong> the momentum gained by the gasphase from the liquid phase per unit volume<br />

<strong>and</strong> time, 9mV l {V, the drag force per unit volume <strong>and</strong> time exerted from the liquid phase<br />

on the gas phase, F D {V, <strong>and</strong> <strong>of</strong> a term ´ 9mV g {V that results by formulating momentum<br />

conservation in terms <strong>of</strong> a primitive variable rather than a conservative variable, i.e.,<br />

S v “ pV l ´ V g q 9m V ´ FD<br />

V , (2.14)<br />

where 9m{V is given by (2.12). The quantity F D denotes the drag force experienced by<br />

11


the gasphase, exerted from the liquid phase, in the finite volume V. In terms <strong>of</strong> the drag<br />

force experienced by the droplets in V,<br />

ÿN l<br />

d“1<br />

F d D ,<br />

F D can be written as<br />

F D “ ´<br />

ÿN l<br />

d“1<br />

F d D . (2.15)<br />

Energy exchange between the two phase is described by the exchange term S e which<br />

represents the energy gained by the gasphase from the liquid phase. Specifically, for a<br />

spray S e can be written as<br />

S e “ ´<br />

ÿN l<br />

d“1<br />

where 9 Q d is the heat gained by a drop in finite volume V. 1<br />

9Q d<br />

V , (2.16)<br />

2.2.4 Phase-Indicator Function<br />

In separated-flow models <strong>of</strong> two-phase flow, the instantaneous location <strong>of</strong> the interface<br />

is conveniently described by level surfaces [82] βpx, tq is constant with the particular<br />

surface βpx, tq “ β I corresponding to the instantaneous location <strong>of</strong> the interface. For the<br />

two-phase spray which is the special case from general two-phase flow, Fig. 2.1 shows<br />

schematically. Those constant value <strong>of</strong> β I is used to distinguish the fluid phases with<br />

Figure 2.1: Schematic diagram <strong>of</strong> two-phase spray.<br />

1<br />

Thus, the heat lost by the d-th drop is ´ 9Q d .<br />

12


espect to time <strong>and</strong> physical space. At certain time <strong>and</strong> position, the condition <strong>of</strong> β px, tq ă<br />

β I represents liquid phase <strong>and</strong> β px, tq ą β I for gas phase. At the interface, where β “ β I ,<br />

the level surface is governed by The governing equation for phase-indicator function θ k<br />

for each phase k – in the present work, k “ 1 denotes liquid phase <strong>and</strong> k “ 2 denotes<br />

gasphase – can be written as<br />

Bθ k<br />

Bt ` V k ¨ ∇θ k “ Π k , (2.17)<br />

where Π k is defined as<br />

Π k “ `´1˘k`1<br />

Vk,r a , (2.18)<br />

where V k,r denotes the normal component <strong>of</strong> the relative velocity V k,r :“ `V I ´ V k˘<br />

¨ n1<br />

<strong>and</strong> interfacial area concentration a is defined as<br />

a :“ |∇β|δ`β ´ β I˘ . (2.19)<br />

With Eq.(2.17), two-phase overall-mass continuity is governed by<br />

<strong>and</strong> two-phase momentum conservation is<br />

<strong>and</strong> two-phase energy conservation is<br />

B<br />

Bt pθ kρ k q ` B pθ k ρ k V k,j q “ ρ k Π k , (2.20)<br />

Bx j<br />

B<br />

Bt pθ kρ k V k,i q ` B<br />

Bx j<br />

pθ k ρ k V k,i V k,j q<br />

“ θ k<br />

ˆBτk,ij<br />

Bx j<br />

B<br />

Bt pθ kc p ρ k T k q ` B<br />

Bx j<br />

pθ k c p ρ k V k,j T k q<br />

<strong>and</strong> two-phase species conservation is<br />

“ θ k<br />

ˆBqk<br />

Bx i<br />

˙<br />

´ θ k<br />

Kÿ<br />

´ Bp ˙<br />

k<br />

` ρ k V k,i Π k , (2.21)<br />

Bx j<br />

α“1<br />

B<br />

Bt pθ kρ k Y k,α q ` B<br />

Bx j<br />

pθ k ρ k V k,j Y k,α q<br />

h k,α w k,α ` ρ k c p T k Π k , (2.22)<br />

α<br />

˙ ˆBJk,i<br />

“ ´θ k ` θ k ρ k w k,α ` ρ k Y k,α Π k . (2.23)<br />

Bx i<br />

13


In the first term <strong>of</strong> r.h.s <strong>of</strong> conservation equation, θ k is located outside <strong>of</strong> differential term.<br />

For example, the first term <strong>of</strong> r.h.s <strong>of</strong> momentum conservation equation can be expressed<br />

as<br />

θ k<br />

ˆBτk,ij<br />

Bx j<br />

2.2.5 Field Equations<br />

´ Bp ˙<br />

k<br />

“ Bθ kτ k,ij<br />

Bx j Bx j<br />

´ Bθ kp k<br />

Bx j<br />

´ pτ k,ij ´ p k q Bθ k<br />

Bx j<br />

. (2.24)<br />

Heret<strong>of</strong>ore flow field <strong>of</strong> both phases have been considered separately with phase indicator<br />

θ k with same definition as k “ 1 for liquid phase <strong>and</strong> k “ 2 for gasphase. Along with<br />

both phases flow field, whole flow field needs to be defined as<br />

φ :“<br />

2ÿ<br />

θ k pβq φ k . (2.25)<br />

With this general form, the field variables at any `x, t˘ are also defined as<br />

<strong>and</strong><br />

ρ :“<br />

V :“<br />

ψ :“<br />

k“1<br />

2ÿ<br />

θ k pβq ρ k , (2.26)<br />

k“1<br />

2ÿ<br />

θ k pβq V k , (2.27)<br />

k“1<br />

2ÿ<br />

θ k pβq ψ k<br />

. (2.28)<br />

k“1<br />

In terms <strong>of</strong> ρ, V <strong>and</strong> ψ defined in Eqs.(2.26) - (2.28), the conservation equations for<br />

two-phase flow can be deterministically described as following.<br />

Now overall-mass continuity can be simply derived that summation <strong>of</strong> Eq.(2.20) thorough<br />

k “ 1 <strong>and</strong> k “ 2 is expressed as<br />

And momentum conservation is<br />

B<br />

Bt pρV iq ` B pρV i V j q<br />

Bx i<br />

“ Bτ ij<br />

Bx j<br />

´ Bp<br />

Bx j<br />

´<br />

B<br />

Bt ρ ` B pρV j q “ 0 . (2.29)<br />

Bx j<br />

2ÿ<br />

k“1<br />

"<br />

pτ k,ij ´ p k q Bθ k<br />

Bx j<br />

´ ρ k V k,i Π k<br />

*<br />

, (2.30)<br />

where τ ij :“ ř 2<br />

k“1 θ kτ k,ij <strong>and</strong> p :“ ř 2<br />

k“1 θ kp as defined at Eq.(2.25).<br />

14


The energy conservation is<br />

B<br />

Bt pc pρT q ` B pc p ρV j T q<br />

Bx j<br />

“ Bq<br />

Bx i<br />

´<br />

Kÿ<br />

h α w α ´<br />

α“1<br />

2ÿ<br />

k“1<br />

"ˆBqk<br />

Bx i<br />

˙ Bθk<br />

Bx j<br />

´ ρ k c p T k Π k<br />

*<br />

. (2.31)<br />

And the species conservation is<br />

B<br />

Bt pρ kY α q ` B pρ k V j Y α q<br />

Bx j<br />

“ ´BJ α i<br />

Bx i<br />

` ρw α ´<br />

2ÿ<br />

k“1<br />

2.3 Probabilistic Formulation<br />

2.3.1 Two-Phase Joint <strong>PDF</strong><br />

Following Rumberg [79, 108], we introduce the joint r<strong>and</strong>om vector<br />

"ˆBJ<br />

α<br />

k,i<br />

Bx i<br />

˙ Bθk<br />

Bx j<br />

´ ρ k Y k,α Π k<br />

*<br />

. (2.32)<br />

Z :“ pV , ψq , (2.33)<br />

where V denotes the 6-dimensional velocity r<strong>and</strong>om vector pV 1 , V 2 q, <strong>and</strong> ψ the 2 pK ` 1qdimensional<br />

scalar r<strong>and</strong>om vector pψ 1<br />

, ψ 2<br />

q where the phase subscription 1 denotes liquid<br />

phase <strong>and</strong> 2 denotes gasphase. The two-phase joint <strong>PDF</strong> f Z,β <strong>of</strong> Z <strong>and</strong> β can be written<br />

as [79, 108] 2 ˆρ f Z,β pẐ, ˆβ; x, tq “ ρ 1 g˚ppẐ, ˆR, ˆβ; x, tq ` G 2 pẐ, ˆβ; x, tq . (2.34)<br />

In (2.34) <strong>and</strong> below, the subscript 1 identifies the liquid phase, the subscript 2 the gaseous<br />

phase. On particular, where the subscript p is used the indicate that the liquid phase is<br />

present in form <strong>of</strong> a spray. 3 In (2.34), G 2 is the gas-phase density function which, in terms<br />

2<br />

ˆρ f Z,β pẐ, ˆβ; x, tq “ G 1 pẐ, ˆβ; x, tq ` G 2 pẐ, ˆβ; x, tq<br />

“ ρ 1 g˚1 pẐ, ˆr, ˆβ; x, tq ` G 2 pẐ, ˆβ; x, tq<br />

“ ρ 1 g˚p pẐ, ˆR, ˆβ; x, tq ` G 2 pẐ, ˆβ; x, tq<br />

3<br />

A spray is one type <strong>of</strong> two-phase flow. It involves a liquid as the dispersed or discrete phase in the<br />

form <strong>of</strong> droplets or ligaments <strong>and</strong> a gas as the continuous carrier phase – see, e.g., [86], pp. 1.<br />

15


<strong>of</strong> gas-phase averaged density<br />

¯ρ 2 px, tq “ 1<br />

xθ 2 y<br />

8ż<br />

8ż<br />

´8 ´8<br />

θ 2 p ˆβqρpẐ, ˆβqfpẐ, ˆβq dẐd ˆβ “ xθ 2ρy<br />

xθ 2 y , (2.35)<br />

the void fraction<br />

¯θ 2 px, tq “<br />

8ż<br />

8ż<br />

θ 2 p ˆβqfpẐ, ˆβq dẐd ˆβ (2.36)<br />

´8 ´8<br />

<strong>and</strong> the gas-phase Favre <strong>and</strong> phase <strong>PDF</strong><br />

rf 2 pẐ, ˆβ; x, tq “ ρpẐ, ˆβq θ 2 p ˆβq<br />

¯ρ 2 px, tq ¯θ 2 px, tq fpẐ, ˆβq (2.37)<br />

is defined as<br />

G 2 :“ ¯ρ 2 ¯θ2 r f2 . (2.38)<br />

The gasphase density function G 2 is governed by [108]<br />

BG 2<br />

Bt ` ˆV BG 2<br />

i “ ´ B `@<br />

DZ{Dt| Ẑ,<br />

Bx ˆβ D G 2˘<br />

i BẐi<br />

´ B `@<br />

Dβ{Dt|<br />

B ˆβ<br />

Ẑ, ˆβ D G 2˘<br />

` δp ˆβ ´ β I q @ Dβ{Dt|Ẑ, ˆβ D G 2 . (2.39)<br />

The function g˚p denotes the liquid-phase DF. In particular, R denotes the drop radius<br />

which is a r<strong>and</strong>om variable with corresponding sample-space variable ˆR.<br />

It is important to note that (2.34) is valid only for a single-component liquid. 4 In case <strong>of</strong><br />

a multi-component liquid, a suitable generalization <strong>of</strong> (2.34) will have to replace (2.34).<br />

4<br />

If, physically, the liquid consists <strong>of</strong> only a single component, mathematically this situation corresponds<br />

to K liquid components where those K ´ 1 components that, in addition, are present in the<br />

gaseous phase in non-zero concentrations attain strictly zero concentrations in the liquid phase.<br />

16


2.3.2 Gas-Phase DF-Transport Equation<br />

At first, we consider gas-phase for k “ 2 with (2.39).<br />

BG 2<br />

Bt ` ˆV BG 2<br />

i “ ´ B `@<br />

DZ{Dt| Ẑ,<br />

Bx ˆβ D G 2˘<br />

i BẐi<br />

´ B `@<br />

Dβ{Dt|<br />

B ˆβ<br />

Ẑ, ˆβ D G 2˘<br />

Integration <strong>of</strong> (2.40) over the ˆβ for k “ 2 yields<br />

` δp ˆβ ´ β I q @ Dβ{Dt|Ẑ, ˆβ D G 2 . (2.40)<br />

G c pẐq “ ż 8´8<br />

G 2 pẐ, ˆβq dβ , (2.41)<br />

the gas-phase marginal-DF transport equation can be derived as<br />

BG c<br />

Bt ` ˆV i<br />

BG c<br />

Bx i<br />

“ ´ B<br />

B ˆV i<br />

`@<br />

DVi {Dt| ˆV , ˆψ D G c˘<br />

´ B `@<br />

B ˆψ<br />

Dψα {Dt| ˆV , ˆψ D G c˘<br />

α<br />

` ĝ M G c , (2.42)<br />

where<br />

ρ1<br />

ĝ M “ ´<br />

ρ 2 p ˆψq<br />

ż 8<br />

0<br />

B F<br />

DR<br />

Dt | ˆV , ˆψ, ˆR 4π ˆR<br />

` 2 f W ˆR| ˆV L , ˆψ L˘d ˆR . (2.43)<br />

The density function f W occurring in (2.43) is defined in (2.50).<br />

2.3.3 Liquid-Phase DF-Transport Equation – Williams’ <strong>Spray</strong><br />

Equation<br />

5<br />

Starting from a suitably defined liquid-phase density function g˚1 pẐ, ˆβ, ˆr; x, tq – see [79,<br />

5<br />

For a single-component, ideally incompressible liquid, the density is a true constant that is denoted<br />

by ρ 1 . Hence,<br />

G 1 :“ ρ 1 g 1 pẐ, ˆβq , (2.44)<br />

where g 1 :“ ¯θ 1<br />

r f1 . According to Rumberg [108], G k :“ ¯ρ k ¯θk r fk <strong>and</strong> g k :“ ¯θ k<br />

r fk , i.e. g 1 “ ¯θ 1<br />

r f1 .<br />

17


108] for details <strong>of</strong> that definition 6 – <strong>and</strong> defining<br />

g˚p pẐ, ˆβ, ˆR; x s , tq :“<br />

ż 8<br />

0<br />

g˚1 pẐ, ˆβ, ˆr; x, tq δpˆr ´ ˆRq dˆr , (2.46)<br />

where R denotes droplet radius (a r<strong>and</strong>om variable whose sample-space variable is denoted<br />

by ˆR), the DF-transport equation<br />

Bg˚p<br />

Bt ` ˆV i<br />

Bg˚p<br />

Bx i<br />

“ ´ B<br />

"B DZj<br />

BẐj Dt<br />

"B F<br />

Dβ<br />

|Ẑ, ˆβ, ˆR<br />

Dt<br />

"B F<br />

DR<br />

|Ẑ, ˆβ, ˆR<br />

Dt<br />

´ B<br />

B ˆβ<br />

´ B<br />

B ˆR<br />

F *<br />

|Ẑ, ˆβ, ˆR g˚p<br />

*<br />

g˚p<br />

g˚p<br />

*<br />

(2.47)<br />

can be derived. Integration <strong>of</strong> (2.47) over all possible ˆβ yields<br />

Bg p<br />

Bt ` ˆV Bg p<br />

i “ ´ B "B F *<br />

DVi<br />

Bx i B ˆV i<br />

Dt | ˆV , ˆψ, ˆR g p<br />

´ B "B F *<br />

Dψα<br />

B ˆψ α<br />

Dt | ˆV , ˆψ, ˆR g p<br />

´ B<br />

B ˆR<br />

"B DR<br />

Dt | ˆV , ˆψ, ˆR<br />

F<br />

g p<br />

*<br />

,<br />

(2.48)<br />

where<br />

g p pẐ, ˆR; x s , tq :“<br />

ż<br />

g˚p<br />

8´8<br />

pẐ, ˆβ, ˆR; x s , tq d ˆβ “ ¯θ 1 px s , tqf r p pẐ, ˆR; x s , tq . (2.49)<br />

The second equality in (2.49) is derived in detail in [108, 79].<br />

Upon multiplying g p with the local <strong>and</strong> instantaneous number density npx, tq, we obtain<br />

6<br />

This DF is governed by the DF-transport equation<br />

"B F *<br />

Bg˚1<br />

Bt ` ˆV<br />

Bg˚1 B DZj<br />

i “ ´<br />

Bx i BẐj Dt |Ẑ, ˆβ, ˆr g˚1<br />

´ B "B F *<br />

Dβ<br />

B ˆβ Dt |Ẑ, ˆβ, ˆr g˚1<br />

´ B "B F *<br />

Dr<br />

Bˆr Dt |Ẑ, ˆβ, ˆr g˚1<br />

B F B F<br />

´ δp ˆβ Dβ<br />

´ β I q<br />

Dt |Ẑ, ˆβ, Dr<br />

ˆr g˚1 `<br />

Dt |Ẑ, ˆβ, Bˆθ 1<br />

ˆr g˚1<br />

Bˆr .<br />

(2.45)<br />

18


the density function<br />

f W p ˆV , ˆψ, ˆR; x, tq :“ npx, tq g p p ˆV , ˆψ, ˆR; x, tq (2.50)<br />

that obeys the so-called William’s <strong>Spray</strong> Equation – see [97], pp. 449 – viz.,<br />

where<br />

<strong>and</strong><br />

Bf W<br />

Bt<br />

` ˆV Bf W<br />

i “ ´ B "B F *<br />

DVi<br />

Bx i B ˆV i<br />

Dt | ˆV , ˆψ, ˆR f W<br />

´ B "B F *<br />

Dψα<br />

B ˆψ α<br />

Dt | ˆV , ˆψ, ˆR f W<br />

´ B "B F *<br />

DR<br />

B ˆR Dt | ˆV , ˆψ, ˆR f W<br />

` Q ` Γ , (2.51)<br />

Q “ 9n b<br />

n f W , (2.52)<br />

Γ “ 9n c<br />

n f W (2.53)<br />

represent the time rate <strong>of</strong> change <strong>of</strong> formation/desctruction 7 <strong>and</strong> droplet collision, 8 respectively;<br />

9n b <strong>and</strong> 9n c denote the corresponding time rates <strong>of</strong> change <strong>of</strong> the number density<br />

n.<br />

2.4 Pressure-Correction Equation<br />

To calculate pressure, a SIMPLE-like pressure correction algorithm [79, 80] is proposed<br />

<strong>and</strong> taken. The original SIMPLE (Semi-Implicit Method for Pressure Linked Equations)<br />

method is proposed by [63]. Original algorithm is the purpose that the Eulerian pressure<br />

field is corrected through combination <strong>of</strong> momentum equation <strong>and</strong> continuity equation.<br />

The derivation <strong>of</strong> a pressure correction equation is based on the overall mean mass conservation<br />

equations as<br />

where<br />

Bxρy<br />

Bt<br />

` BxρV iy<br />

Bx<br />

“ 0 , (2.54)<br />

xρy “ xθ l yxρ l y ` xθ g yxρ g y , (2.55)<br />

xρV i y “ xθ l yxρ l V l,i y ` xθ g yxρ g V g,i y . (2.56)<br />

7<br />

A droplet may be destroyed, e.g., by coalescence <strong>of</strong> drops. Formation <strong>of</strong> droplets may occur, e.g.,<br />

by droplet breakup or by droplet nucleation.<br />

8<br />

Collision <strong>of</strong> a droplet may occur with one or several other droplets or with a wall.<br />

19


Detailed derivation <strong>of</strong> a pressure equation will be presented below in Chaps. 5.2.<br />

xpy n`1 “ xpy n ` α p`p1n ´ p 1 n´1˘<br />

(2.57)<br />

where α p is a relaxation factor which have been taken in the range 0.1 ď α p ď 0.3 [79, 80].<br />

For a fixed time step pressure corrector p 1 can be separated as<br />

p 1 “<br />

`p<br />

1˘<br />

` `p 1˘S (2.58)<br />

M<br />

where M denotes correction term for unsatisfied mass conservation as usually contained<br />

original SIMPLE method <strong>and</strong> S for statistical error. In the course <strong>of</strong> the iterations for the<br />

time step, p 1 M approaches zero. In contrast, p1 S hardly varies between successive iteration<br />

1˘n 1˘n´1<br />

1˘n´1<br />

<strong>and</strong> time steps. With sufficiently large time step n,<br />

`p «<br />

`p <strong>and</strong><br />

`p « 0,<br />

S<br />

S<br />

M<br />

therefore p 1n ´ p 1n´1 1˘n<br />

«<br />

`p . The pressure correction specified in (2.58) thus corrects<br />

M<br />

towards satisfaction <strong>of</strong> mass conservation <strong>and</strong>, at the same time, filters out the influence<br />

<strong>of</strong> statistical fluctuations between successive iteration <strong>and</strong> time steps [79, 80].<br />

20


Chapter 3<br />

Computational Approach<br />

3.1 Two-Phase Particle Method<br />

For either the gaseous <strong>and</strong> the liquid phase, we distinguish between molecules, fluid particles<br />

<strong>and</strong> particles. In addition, since in the present work a spray is considered that<br />

represents a special form <strong>of</strong> two-phase flow in which the liquid phase is represented exclusively<br />

by a very large number <strong>of</strong> drops, also drops – which due to there small size are<br />

also termed droplets – need to be considered,<br />

The numerical solution method selected in the present is a so-called particle method. In<br />

the context <strong>of</strong> particle methods, a particle is representative for a huge number <strong>of</strong> fluid<br />

particles having – in a numerical sense – nearly the identical properties, Specifically, for the<br />

spray problems considered in the present thesis, both gasphase particles <strong>and</strong> liquid-phase<br />

particles are considered. Thus, a gasphase particle is representative <strong>of</strong> a huge number<br />

<strong>of</strong> gaseous fluid particles having nearly identical velocity <strong>and</strong> scalar properties such as<br />

temperature <strong>and</strong> concentrations. Similarly but yet differently, a liquid-phase particle is<br />

representative <strong>of</strong> a huge number <strong>of</strong> liquid drops having nearly identical velocity <strong>and</strong> scalar<br />

properties such as temperature, concentrations <strong>and</strong> droplet size.<br />

In addition to carrying a velocity (3 components), temperature <strong>and</strong> K concentrations,<br />

a gasphase particle also carries a size. Specifically, since we assume that particles are<br />

spherical, the particle size is taken as a radius. This radius, together with the density<br />

attached to a particle, is used to define the mass attached to a particle. For instance, the<br />

mass attached to to gasphase particle number p is<br />

ż<br />

ż R<br />

p<br />

g<br />

m p gptq “ ρ p gptq dV “ 4 π pRgq p 2 ρ p gptq dR , (3.1)<br />

Vg<br />

p 0<br />

21


<strong>and</strong> ρ p g denote the radius, volume <strong>and</strong> density <strong>of</strong> pasphase particle p, re-<br />

where Rg, p Vg<br />

p<br />

spectively.<br />

Whilst it seems natural to attach to a liquid-phase particle a size (because a droplet has<br />

a size), on first glance its seems surprising to attach a size also to a gasphase.<br />

The essence <strong>of</strong> such a method consists in [69] <strong>and</strong> it is limited for gasphase.<br />

The number <strong>of</strong> liquid particles is denoted P l , the number gaseous particles P g , p “ 1, ..., P l<br />

<strong>and</strong> p “ 1, ..., P g . While the number <strong>of</strong> gaseous particles can be extended up to infinity to<br />

represent every single gaseous molecule, the number <strong>of</strong> liquid particles has upper bound<br />

as number <strong>of</strong> physical drops.<br />

3.1.1 Particle Method for Liquid Phase<br />

Here <strong>and</strong> below, for a liquid particle p, the particle mass is denoted by m p l . Assuming<br />

spherical particles,<br />

m p l<br />

“ 4 3 π ρp l Rp 3 l<br />

. (3.2)<br />

Accordingly, the time rate <strong>of</strong> change <strong>of</strong> the mass <strong>of</strong> liquid particle p is<br />

9m p l<br />

“ 4π ρ p l Rp 2 dR p l<br />

l<br />

dt<br />

, (3.3)<br />

where R p l<br />

<strong>and</strong> ρ p l<br />

denote the radius <strong>and</strong> density <strong>of</strong> liquid phase particle p, respectively.<br />

In terms <strong>of</strong> the particle masses 9m p l , we can write the interfacial mass exchange term S m<br />

specified earlier in Eq. (2.12) as<br />

S m “ ´ 1<br />

V<br />

ÿN l<br />

d“1<br />

9m d « ´ 1<br />

V<br />

ÿP l<br />

p“1<br />

9m p l . (3.4)<br />

Here the first, exact equality has been taken from (2.12), <strong>and</strong> the second, approximate<br />

equality stems by replacing physical drops by liquid particles.<br />

With 9m p l<br />

, the mass <strong>of</strong> species is written as<br />

S m,α “<br />

ÿP l<br />

p“1<br />

Y p<br />

l,α 9mp l , (3.5)<br />

In terms <strong>of</strong> droplet radius rather than diameter, the so-called D 2 -law [28, 49, 93, 97] can<br />

be written as<br />

dR p l<br />

dt<br />

“ ´Kp<br />

R p l<br />

22<br />

, (3.6)


where the evaporation constant K p is given by<br />

K p “<br />

8λ g<br />

ρ p l c ln pB T ` 1q , (3.7)<br />

p,g<br />

where λ g , c p,g <strong>and</strong> B T denote the thermal conductivity, the specific heat capacity <strong>and</strong> the<br />

Spalding heat transfer number, respectively. Here B T is ,<br />

B T “ c p,g pT g ´ T boil q<br />

L<br />

, (3.8)<br />

where L denotes is the latent heat <strong>of</strong> vaporization at mean temperature pT g ` T boil q {2.<br />

The interfacial momentum exchange term S v specified earlier in Eq. (2.14) can be written<br />

as<br />

With<br />

S v “ 1 V<br />

ÿN l<br />

d“1<br />

S d v « 1 V<br />

ÿP l<br />

p“1<br />

S p v . (3.9)<br />

S p v “ pxV l,i y ´ xV g,i yq 9m p l ´ F p D , (3.10)<br />

where<br />

with the drag coefficient<br />

<strong>and</strong> the Reynolds number<br />

F p D “ 3 p<br />

ρ g C D<br />

8 ρ p l<br />

R |xV g,iy ´ V p<br />

p l,i |`xV g,i y ´ V p<br />

l,i<br />

˘mp<br />

l , (3.11)<br />

C p D “ 24<br />

Re p l<br />

˜<br />

Re p l<br />

“ 2xρ g yR p l<br />

1 ` Rep l<br />

6<br />

2{3<br />

¸<br />

, (3.12)<br />

|xV g,i y ´ V p<br />

l,i |<br />

µ g<br />

. (3.13)<br />

In terms <strong>of</strong> the liquid particle, we can write the interfacial energy exchange term S e<br />

specified earlier in Eq. (2.16) as<br />

S e “ 1 V<br />

ÿN l<br />

d“1<br />

9Q d « 1 V<br />

ÿP l<br />

p“1<br />

9Q l<br />

p<br />

, (3.14)<br />

with 9 Q l<br />

p<br />

is defined as<br />

9Q l<br />

p<br />

“ 9m<br />

p<br />

l<br />

˘<br />

p<br />

#¯c p,g`Tg ´ T p<br />

l<br />

B p T<br />

˘+<br />

´ L`T p<br />

l<br />

. (3.15)<br />

23


˘<br />

Here B p p<br />

T<br />

denotes the Spalding heat transfer number <strong>and</strong> L`Tl<br />

is the latent heat <strong>of</strong><br />

vaporization at droplet surface temperature T p<br />

l<br />

. The overbar indicates that the value <strong>of</strong><br />

the specific heat is to be calculated according to the so-called 1/3 rule [2, 36, 102] – see<br />

Chap. 4.3.2.<br />

The equation <strong>of</strong> motion for the p-th liquid particle can be written as<br />

dV p<br />

l,i “ ´ 1<br />

ρ p l<br />

˙ ˆBxpy<br />

dt ´ 1<br />

Bx i ρ p Svdt p , (3.16)<br />

l<br />

where ρ p l<br />

denote the density <strong>of</strong> liquid phase particle p <strong>and</strong> S p v has been defined in Eq.<br />

(3.10).<br />

If we assume that the liquid particle has uniform temperature <strong>and</strong> only single component,<br />

the energy equation for the p-th liquid particle can be reduced as<br />

dT p<br />

l<br />

dt<br />

“<br />

9Q l<br />

p<br />

4<br />

3 πρp l c p,l p R p 3 , (3.17)<br />

l<br />

where the energy exchange term 9 Q l<br />

p<br />

has been defined in Eq. (3.15).<br />

3.1.2 Particle Method for Gas Phase<br />

After formulating the overall exchange terms S m , S v <strong>and</strong> S e in terms <strong>of</strong> liquid-particle<br />

properties, the task now is to define the corresponding gasphase-particle exchange terms<br />

such that in each finite volume V evaporation or condensation, respectively, obey mass,<br />

momentum <strong>and</strong> energy conservation. This task is accomplished by allocating the overall<br />

exchange quantities S m , S v <strong>and</strong> S e to the gasphase-particles, i.e., by defining<br />

S p g,m “ ´ 1<br />

P g V<br />

Sg,m,α p “ 1 ÿP l<br />

P g<br />

p“1<br />

ÿP l<br />

p“1<br />

9m p l , (3.18)<br />

Y p<br />

l,α 9mp l , (3.19)<br />

S p g,v “<br />

S p g,e “<br />

1<br />

P g V<br />

1<br />

P g V<br />

ÿP l<br />

p“1<br />

ÿP l<br />

p“1<br />

S p v , (3.20)<br />

9Q p , (3.21)<br />

24


where the subscript g identifies gasphase particles.<br />

In terms <strong>of</strong> the distributed mass exchange term, the species-mass conservation for the<br />

p-th gas particle can be written as<br />

dY<br />

ρ p g,α<br />

p<br />

g<br />

dt<br />

“ ´Bjp g,α,i<br />

Bx i<br />

` w p g,α ` S p g,m,α , (3.22)<br />

where α “ 1, ..., K g . In (3.22), w p g,α denotes net rate <strong>of</strong> production due to chemical reaction<br />

for the p-th gas particle <strong>and</strong> the diffusion flux j p g,α,i will be modelled by suitable model<br />

represented in Chap. 4.2. And Sg,m,α p has been defined in Eq. (3.19).<br />

In terms <strong>of</strong> the distributed momentum exchange term, The equation <strong>of</strong> motion for the<br />

p-th gas particle can be written as<br />

dV p<br />

g,i “ ´ 1 Bp<br />

ρ p dt ` 1 Bτ p g,ij<br />

g Bx i ρ p dt ` 1<br />

g Bx j ρ p Sg,vdt p , (3.23)<br />

g<br />

where the viscous part <strong>of</strong> the gas-phase stress tensor for the p-th gas particle τ p g,ij will be<br />

modelled by suitable model represented in Chap. 4.1. And Sg,v p has been defined in Eq.<br />

(3.20).<br />

The energy conservation for the p-th gas particle can be written as<br />

dT<br />

c p g,p ρ p g<br />

p<br />

g<br />

dt<br />

“ ´Bqp g<br />

Bx i<br />

´<br />

K<br />

ÿ g<br />

α“1<br />

h p g,αw p g,α ` S p g,e , (3.24)<br />

where q p g, h g,α <strong>and</strong> w p g,α denote the heat flux, the specific enthalpy <strong>and</strong> mass rate <strong>of</strong><br />

production for the p-th gas particle, respectively. And S p g,e has been defined in Eq. (3.21).<br />

25


3.2 Expectations <strong>and</strong> Moving Averages<br />

3.2.1 Expectations<br />

In [30, 69, 79] Reynolds averaged expectations <strong>and</strong> Favre averaged expectations were<br />

defined for single-phase flows. 1<br />

For two-phase flows, for any physical quantity φ, Reynolds averaged expectations can be<br />

written as [79, 80]<br />

xφy “<br />

8ż<br />

8ż<br />

8ż<br />

ˆφf` ˆV , ˆψ, ˆβ˘d ˆV d ˆψd ˆβ<br />

´8 ´8 ´8<br />

“ xθ l yxφ|β ă ˆβ I y ` xθ g yxφ|β ą ˆβ I y<br />

« 1 ÿP l<br />

xθ l yφ p l<br />

P ` 1<br />

l<br />

p“1<br />

P<br />

ÿ g<br />

xθ g yφ p g . (3.27)<br />

P g p“1<br />

1<br />

For single-phase flow, for any physical quantity φ, Reynolds averaged expectations xφy can be<br />

calculated as [30, 69, 79]<br />

xφy “<br />

8ż<br />

« 1 P<br />

8ż<br />

´8 ´8<br />

ˆφf` ˆV , ˆψ˘d ˆV d ˆψ<br />

Similarly, Favre averaged expectations can be calculated as [30, 69]<br />

rφ “<br />

“<br />

8ż<br />

8ż<br />

Pÿ<br />

φ p . (3.25)<br />

p“1<br />

´8 ´8<br />

8ż<br />

« 1 P<br />

8ż<br />

´8 ´8<br />

Pÿ<br />

p“1<br />

ˆρ ˆφ<br />

xρy f` ˆV , ˆψ˘d ˆV d ˆψ<br />

ˆφ r f` ˆV , ˆψ˘d ˆV d ˆψ<br />

ρ p φ p<br />

xρy . (3.26)<br />

26


Similarly, Favre averaged expectations for two-phase flows, can be written as<br />

rφ “<br />

8ż<br />

8ż<br />

8ż<br />

´8 ´8 ´8<br />

ˆρ ˆφ<br />

xρy f` ˆV , ˆψ, ˆβ˘d ˆV d ˆψd ˆβ<br />

“ xθ ly<br />

xρ l y xρφ|β ă ˆβ I y ` xθ gy<br />

xρ g y xρφ|β ă ˆβ I y<br />

« 1 ÿP l<br />

P l<br />

3.2.2 Moving Averages<br />

p“1<br />

xθ l yρ p l φp l<br />

xρ l y<br />

` 1<br />

P g<br />

P<br />

ÿ g<br />

p“1<br />

xθ l yρ p gφ p g<br />

xρ g y<br />

. (3.28)<br />

Generally a distinction is made between ensemble averaging to obtain stochastic mean<br />

values, time averaging to obtain temporal mean values, <strong>and</strong> spatial averaging to obtain<br />

spatial mean values. Spatial averaging leads to LES-like 2 formulations <strong>of</strong> turbulent flows<br />

<strong>and</strong> hence are not considered in the present work.<br />

For flows that are steady in the<br />

temporal mean, ie., for which the mean values do not vary with time, ensemble averaging<br />

<strong>and</strong> temporal averaging leads to identical mean values provided that, in practice, the<br />

number <strong>of</strong> particle as well as the number <strong>of</strong> time steps is sufficiently large.<br />

Since in the current work in the temporal mean unsteady numerical simulations are performed<br />

to approach a statistically <strong>and</strong> hence temporally steady solution, it is advantageous<br />

to monitor that approach by considering a so-called moving average. This averaging procedure<br />

is similar to time averaging but rather than averaging over a large (theoretically<br />

infinite) number <strong>of</strong> time steps averaging is carried out backwards from timestep n only<br />

over the last n back steps, i.e., over the backward period t pnq ´ t pn´n backq . As a numerical<br />

solution evolves, taking moving averages smoothes out temporal fluctuations thereby facilitating<br />

both the display <strong>of</strong> graphical or even animated temporal solutions as well as the<br />

identification <strong>of</strong> approaching an in the mean steady state, see, e.g., [7, 35]<br />

For instance for a non-constant time step size,<br />

∆t pnq “ t pnq ´ t pn´1q (3.29)<br />

a moving average is calculated according to<br />

¯φ pnq “<br />

ř n<br />

i“n´n back<br />

xρy piq ∆t piq φ piq<br />

ř n<br />

i“n´n back<br />

xρy piq ∆t piq (3.30)<br />

In the present work, a typical value is n back “ 100.<br />

2<br />

LES st<strong>and</strong>s for Large-Eddy-<strong>Simulation</strong>.<br />

27


Chapter 4<br />

Models<br />

To evaluate the various conditional means occurring on the right-h<strong>and</strong> sides <strong>of</strong> (2.42) <strong>and</strong><br />

(2.51), respectively, certain local <strong>and</strong> instantaneous terms require modeling. Specifically,<br />

these are the terms DV g,i {Dt, Dψ α {Dt, <strong>and</strong> DR l {Dt, giving rise to a so-called velocity<br />

model, mixing model, <strong>and</strong> evaporation model, respectively. Since we will select a velocity<br />

model that includes the turbulent kinetic energy k <strong>and</strong> the turbulent dissipation rate ɛ, a<br />

suitable two-equations turbulence model will also be required.<br />

Since in the computations a Lagrangian approach will be adopted in which so-called<br />

Monte-Carlo particles are tracked, the Eulerian derivatives following the motion <strong>of</strong> a<br />

fluid particle, viz., DV g,i {Dt, Dψ g,α {Dt, <strong>and</strong> DR l {Dt, are replaced by corresponding Lagrangian<br />

derivatives following individual Monte-Carlo particles, viz., DV ˚<br />

g,i{Dt, Dψ˚g,α{Dt,<br />

<strong>and</strong> DR˚l {Dt.<br />

4.1 Velocity Model<br />

In the present work, as model for DV ˚<br />

g,i{Dt the so-called Langevin equation<br />

dVg,i ˚ “ ´ 1 B xpy<br />

dt `<br />

ρ g Bx rV BV r ´<br />

g,j<br />

r,j dt ` G˚ij Vg,j ˚ ´ g,j¯<br />

i Bx rV dt ` B ij dW j , (4.1)<br />

i<br />

where mean slip velocity r V r “ r V l ´ rV g <strong>and</strong> G˚ij are given by [58]<br />

«<br />

G˚ij “ ´ 1 δ<br />

T ˚L,K<br />

ij ´<br />

1<br />

T ˚L,‖<br />

ff<br />

´ 1 r<br />

T ˚L,K<br />

i r j , (4.2)<br />

28


where the integral timescales T L , T ˚L,K <strong>and</strong> T ˚L,‖<br />

T L “<br />

ˆ1<br />

2 ` 3<br />

˙´1<br />

4 C k<br />

0<br />

ɛ , (4.3)<br />

T ˚L,K “<br />

T ˚L,‖ “<br />

T L<br />

b<br />

, (4.4)<br />

1 ` 4β 2 V r r<br />

2<br />

2k{3<br />

T L<br />

b , (4.5)<br />

1 ` β 2 V r r<br />

2<br />

2k{3<br />

with r i “ U r,i {| r V r |.<br />

And B ij are given by<br />

B ij “<br />

d<br />

ɛ<br />

ˆ<br />

˙<br />

C 0 b i<br />

r 2 k{k `<br />

3 pb r ik{k ´ 1q , (4.6)<br />

with b 1 “ T L {T ˚L,‖ <strong>and</strong> b 2 “ T L {T ˚L,K .<br />

In the present work, the effect <strong>of</strong> slip velocity has been neglected. Hence, (4.1) reduces to<br />

dVg,i ˚ “ ´ 1 ˆBxpy<br />

ˆ1<br />

´ S v˙<br />

dt `<br />

ρ g Bx i<br />

2 ` 3 ˙ ɛ<br />

´<br />

4 C 0 Vg,i ˚ ´ g,i¯<br />

k<br />

rV dt ` pC 0 ɛq 1{2 dW i , (4.7)<br />

which was used earlier for [69]. 1 Here, in addition to the quantities introduced or defined<br />

above, C 0 is a model constant with numerical value 2.1 [3, 69]. The quantity W i denotes<br />

a so-called Wiener process [69, 73], i.e., is a stochastic process, normally distributed with<br />

mean zero <strong>and</strong> variance dt, <strong>and</strong> with the property<br />

dW i ptq “ W i pt ` dtq ´ W i ptq , (4.9)<br />

i “ 1, 2, 3. The quantities k <strong>and</strong> ɛ were introduced above – see page (36). In the present<br />

work, the fields <strong>of</strong> k <strong>and</strong> ɛ will be computed by use <strong>of</strong> a so-called k-ɛ turbulence model –<br />

see Chap. 4.4.<br />

1<br />

The Langevin equation (4.7) is a special form <strong>of</strong> the more general Langevin equation<br />

dV ˚<br />

g,i “ ´ 1<br />

ρ g<br />

B xpy<br />

Bx i<br />

dt ` G ij`V ˚<br />

g,j ´ rV g,j˘dt ` pC0 ɛq 1{2 dW i . (4.8)<br />

For details <strong>and</strong> discussions related to (4.8), references [32, 69] should be consulted.<br />

29


4.2 Mixing Models<br />

For the gaseous phase <strong>of</strong> two-phase flow, locally <strong>and</strong> instantaneously, species mass conservation<br />

is<br />

Dψ α<br />

Dt<br />

“ ´Bj α,i<br />

Bx i<br />

` Spψ α q ` S m,α , (4.10)<br />

α “ 1, ..., K. Assuming in (4.10) Fick’s law <strong>of</strong> diffusion with a constant diffusion coefficient<br />

D α for species α, the composition <strong>PDF</strong><br />

Df ψ<br />

Dt “ ´ B<br />

B ˆψ α<br />

ˆB<br />

D α<br />

B 2 ψ α<br />

Bx 2 i<br />

F<br />

| ˆψ f ψ˙<br />

`<br />

B<br />

B ˆψ α<br />

`Sp ˆψα qf ψ˘<br />

`<br />

B `A<br />

B ˆψ<br />

S m,α | ˆψ<br />

E<br />

f ψ˘<br />

α<br />

(4.11)<br />

is obtained [73]. Since the chemical source term is closed,<br />

AS`ψ α˘|ψα “ ˆψ<br />

E<br />

α “ S` ˆψα˘<br />

. (4.12)<br />

However, the molecular-diffusion term<br />

B<br />

F<br />

B 2 ψ α<br />

D α |<br />

Bx ˆψ 2 α<br />

i<br />

(4.13)<br />

is not closed <strong>and</strong> hence requires modelling by a so-called mixing model. Various mixing<br />

models are available. Specifically, in the present work the so-called IEM model, Curl’s<br />

model, Curl’s modified model, <strong>and</strong> the so-called EMST-model have been employed. Subsequently,<br />

these models are presented <strong>and</strong> discussed.<br />

4.2.1 IEM-Model<br />

IEM st<strong>and</strong>s for Interaction by Exchange with the Mean. In the framework <strong>of</strong> turbulent<br />

combustion, IEM models have been used extensively, see – e.g. – references [13, 94]. In<br />

the present context, the IEM model can be written as<br />

B<br />

D α<br />

B 2 ψ α<br />

Bx 2 i<br />

F<br />

| ˆψ α “ 1 2 C ɛ `<br />

ψ ˆψ ´ xψy˘<br />

. (4.14)<br />

k<br />

Here k <strong>and</strong> ɛ have their usual meaning, <strong>and</strong> C ψ is a model constant with numerical value<br />

2.0, see [73]. With (4.14), the gasphase particle mass conservation equation becomes<br />

Dψ˚α<br />

Dt<br />

“ ´1<br />

2 C ɛ<br />

ψ<br />

`ψ˚α ´ xψ α y˘ ` Spψ˚αq ` S m,α . (4.15)<br />

k<br />

30


4.2.2 Curl’s Model<br />

Curl’s mixing model [11] is based on – from a uniform distribution r<strong>and</strong>omly <strong>and</strong> independently<br />

selected – particles p <strong>and</strong> q, 1 ď p, q ď N p , where N p denotes the number <strong>of</strong><br />

gasphase particles. For these particles, Curl’s mixing model can be written<br />

ψ˚ppq<br />

α pt ` dtq “ ψ˚pqq<br />

α pt ` dtq “ 1 `ψ˚ppq<br />

α ptq ` ψ˚pqq<br />

α ptq˘ . (4.16)<br />

2<br />

The so-called modified Curl model [38] replaces the constant 1/2 on the r.h.s. <strong>of</strong> (4.16) by<br />

a r<strong>and</strong>omly selected weight parameter ξ, 0 ď ξ ď 1, such that the mixing model becomes<br />

ψ˚ppq<br />

α pt ` dtq “ p1 ´ ξqψ˚ppq<br />

α ` ξ 1 `ψ˚ppq<br />

α ptq ` ψ˚pqq<br />

α ptq˘ , (4.17)<br />

2<br />

ψ˚pqq<br />

α pt ` dtq “ p1 ´ ξqψ˚pqq<br />

α ` ξ 1 `ψ˚ppq<br />

α ptq ` ψ˚pqq<br />

α ptq˘ . (4.18)<br />

2<br />

The number <strong>of</strong> r<strong>and</strong>omly selected pairs pp, qq <strong>of</strong> particles is N mix ,<br />

N mix “ ΩP dt , (4.19)<br />

where Ω denotes the mixing frequency,<br />

Ω “ C ψ<br />

ɛ<br />

k . (4.20)<br />

For Curl’s model, the numerical value <strong>of</strong> the model constant C ψ is 2.0; for the modified<br />

Curl model it is 2.3 [51].<br />

4.2.3 EMST-Model<br />

EMST st<strong>and</strong>s for Euclidian Minimum Spanning Tree. EMST-Model [89] is an extension<br />

work for multi-dimension scalar mixing based on the mapping closure model [70] which is<br />

implemented for a single scalar mixing. In this model, particle composition is changed by<br />

particle interactions through the edges <strong>of</strong> a tree which is constructed to have Euclidean<br />

minimum total length. In order to develop the Euclidean Minimum Spanning Tree, it is<br />

required to connect every particle by an edge to at least one neighboring particle. The<br />

spanning property would come by a connection between every particle in the set. And<br />

the sum <strong>of</strong> Euclidean lengths <strong>of</strong> the edges in the tree should be minimum.<br />

31


1<br />

2<br />

0<br />

-1<br />

-1 0 1<br />

Figure 4.1: Euclidian minimum spanning tree in two dimension composition space.<br />

1<br />

Euclidean distance between particle p <strong>and</strong> q in scalar composition is defined as<br />

g<br />

fÿ<br />

dpp, qq “ e K ´<br />

α“1<br />

ψ˚ppq<br />

α<br />

In the present context, the EMST model can be written as<br />

´ ψ˚pqq<br />

¯2<br />

α . (4.21)<br />

dψ˚ppq<br />

α<br />

w ppq<br />

dt<br />

“ ´γ<br />

Nÿ<br />

p´1<br />

ν“1<br />

B ν tpψ˚ppq<br />

α<br />

´ ψ˚pnνq<br />

α q δ p,nν ` pψ˚ppq<br />

α ´ ψ˚pmνq<br />

α q δ p,mν u , (4.22)<br />

where the νth edge <strong>of</strong> the tree connects the particle pair (m ν , n ν ).<br />

δ represents the<br />

Kronecker delta, therefore δ p,mν <strong>and</strong> δ p,nν to be zero except when p “ m ν or n ν that they<br />

are equal to one.<br />

The left h<strong>and</strong> side <strong>of</strong> (4.22) is evaluated as follows.<br />

1. For each particle p, p “ 1, ..., N p , specify a weight w ppq . For instance, a popular<br />

choice is w ppq “ 1{N p . 2<br />

2 The original meaning <strong>of</strong> w is importance weight. In EMST-model, w is represented to particles<br />

weight attributed from particles mass ∆m. A classical way is that ∆m is set as constant [23, 69], then<br />

32


2. For node p, determine a tree consisting <strong>of</strong> all N p nodes <strong>and</strong> N p´1 edges, ν “ 1, ..., N p .<br />

The edge coefficients B ν are calculated by carrying out the following steps [89].<br />

(a) Assume that edge ν connects particles m <strong>and</strong> n. Then the tree is partitioned<br />

into three subtrees, namely, the trivial subtree tνu, the subtree T mν<br />

<strong>and</strong> the<br />

subtree T nν . The subtrees T mν <strong>and</strong> T nν are defined such that they result from<br />

the overall tree by removing edge ν.<br />

(b) First calculate the sums<br />

W Tmν :“ ÿ<br />

<strong>and</strong> then obtain the weight <strong>of</strong> wedge ν,<br />

kPT mν<br />

w pkq , (4.23)<br />

W Tnν :“ ÿ<br />

kPT nν<br />

w pkq , (4.24)<br />

w ν “ minpW Tmν , W Tnν q . (4.25)<br />

(c) Take the edge coefficient B ν as<br />

B ν “ 2w ν . (4.26)<br />

3. The model parameter γ controls the rate <strong>of</strong> variance decay <strong>of</strong> the scalars. It is<br />

obtained through an iterative procedure [89] by solving the following equations.<br />

Here γ is given by<br />

¯M pq “ ´ 1<br />

Nÿ<br />

p´1<br />

B ν tδ p,mν δ q,nν ` δ q,mν δ p,nν u . (4.27)<br />

γ<br />

ν“1<br />

1<br />

γ “ τ ř Np<br />

ř Np<br />

ψ k“1<br />

α<br />

ř K<br />

β“1 C ββ<br />

l“1 w pkq2 ψ˚pkq1<br />

¯Mkl ψ˚plq1<br />

α<br />

, (4.28)<br />

where the C αβ , 1 ď α, β ď, K denote the covariances<br />

C αβ “ x pψ α ´ xψ α yq pψ β ´ xψ β yq y (4.29)<br />

<strong>and</strong> with τ ψ “ τ{C ψ followed as st<strong>and</strong>ard modeling assumptions [69]. C ψ is taken<br />

to be 2.0 [89].<br />

all particles importance becomes equal, w ppq “ 1{N since sum <strong>of</strong> w ppq should be unity in EMST-model<br />

[89, 103].<br />

33


4.3 Evaporation Models<br />

4.3.1 D 2 ´ law Model<br />

One <strong>of</strong> the classic evaporation models is so-called D 2 ´ law [28, 49, 97]. In this model,<br />

the droplet temperature is spatially uniform <strong>and</strong> the temperature is assumed to be the<br />

boiling point <strong>of</strong> the fuel, T d “ T boil [93]. The instantaneous diameter <strong>of</strong> k-th droplet Dl kptq<br />

has a correlation<br />

2<br />

ptq “ D<br />

k2 l pt0 q ´ Kpt ´ t 0 q . (4.30)<br />

D k l<br />

Here Dl k ptq is the instantaneous diameter <strong>of</strong> a droplet that entered physical space at time<br />

to <strong>and</strong> since then has moved along its trajectory. The constant K is given by<br />

K “<br />

8k g<br />

ρ k L c ln pB T ` 1q , (4.31)<br />

p,g<br />

with Spalding heat transfer number B T<br />

B T “ c p,G pT g ´ T boil q<br />

L<br />

, (4.32)<br />

where L is the latent heat <strong>of</strong> vaporization at mean temperature 1 2 pT g ` T boil q. Differentiation<br />

<strong>of</strong> (4.30) yields, in terms <strong>of</strong> the instantaneous droplet radius R l ,<br />

BR k l<br />

Bt<br />

“ ´ K<br />

R k l<br />

. (4.33)<br />

4.3.2 Film Model<br />

Alternatively, to determine the mass vaporization rate <strong>of</strong> droplet, the so-called film model<br />

is provided by [2, 84]. In film model, it is assumed that the mass <strong>and</strong> heat exchange<br />

only occurs at thin layer (so-called film) between the droplet surface <strong>and</strong> the gas. The<br />

instantaneous droplet radius R l which has rate <strong>of</strong> change, 9R l is written as<br />

dR k l<br />

dt<br />

“ ´<br />

9m k l<br />

4πρ k l Rk l<br />

2 . (4.34)<br />

According to film model [2, 84], 9m k l<br />

is defined as<br />

9m k l “ ´2π ¯ρ k ¯Dk Rl k Sh k˚ ln`1 ` BM˘ k , (4.35)<br />

where D denotes the binary diffusion coefficient. The overbar indicates that the value<br />

<strong>of</strong> the specific heat is to be calculated according to the so-called 1/3 rule [2, 36, 102].<br />

For example, the averaged density ¯ρ k <strong>and</strong> the averaged specific heat capacity ¯c k p can be<br />

34


obtained as<br />

¯ρ k “ ρ k l ` 1 ˘ `xρg y ´ ρ k l , (4.36)<br />

3<br />

¯c k p “ c k p,L ` 1 ˘ `cp,g ´ c k p,l . (4.37)<br />

3<br />

To complete Eq. (4.35), the modified Sherwood number Sh˚ is defined as<br />

Sh k˚ “ 2 ` `Sh k 0 ´ 2˘ {F k M , (4.38)<br />

with<br />

<strong>and</strong><br />

where<br />

FM k “ `1 ` BM<br />

k ˘0.7 lnp1 ` BM k q , (4.39)<br />

B k M<br />

Sh k 0 “ 2 ` 0.552 Re k1{2 Sc k1{3 , (4.40)<br />

Re k “ 2xρ g y ˇˇV k<br />

l ´ xV g yˇˇ Rk l {µ g , (4.41)<br />

<strong>and</strong> Schmidt number Sc<br />

Sc k “<br />

The spalding mass transfer number B M is defined as<br />

¯µk g<br />

. (4.42)<br />

¯ρ<br />

k ¯D<br />

k<br />

B k M “ Y F,S ´ Y F,g<br />

1 ´ Y F,S<br />

, (4.43)<br />

where Y F,S denotes the fuel mass fractions <strong>of</strong> vapor at the droplet surface <strong>and</strong> calculated<br />

by Clausis-Clapeyron equation [2, 93] as<br />

Y F,S “<br />

M F<br />

M F ` `xpy{p k F ´ 1˘ ¯Mg<br />

, (4.44)<br />

where M F is the molecular weight <strong>of</strong> the fuel <strong>and</strong> ¯M g is the mean molecular weight <strong>of</strong> gas<br />

phase <strong>and</strong> the partial pressure <strong>of</strong> the fuel vapor p F is given by<br />

p k F “ A exp`´B{ ¯T k˘ , (4.45)<br />

where<br />

¯T k “ Tl k ` 1 ˘ `Tg ´ Tl<br />

k , (4.46)<br />

3<br />

<strong>and</strong> the constants A <strong>and</strong> B are obtained from the Clausis-Clapeyron equation [93].<br />

35


The heat penetrating into liquid phase Q l can be defined as<br />

Q k l “ 9m k l<br />

# ˘<br />

¯cp k`T g ´ Tl<br />

k<br />

BT<br />

k<br />

˘+<br />

´ L`Tl<br />

k , (4.47)<br />

with the spalding heat transfer number B T<br />

B k T “ `1 ` B k M˘φ<br />

´ 1 , (4.48)<br />

where<br />

And the modified Nusselt number Nu˚ is<br />

φ “ c¯<br />

P k ¯ρ k ¯Dk Sh k˚<br />

¯λ k Nu k˚ . (4.49)<br />

Nu k˚ “ 2 ` pNu 0 ´ 2q {F k T , (4.50)<br />

with<br />

¯ Nu 0 “ 2 ` 0.552 Re k1{2 P r k1{3 , (4.51)<br />

where Pr<strong>and</strong>tl number P r defined<br />

P r k “ ¯µk ¯c p<br />

k<br />

¯λ k . (4.52)<br />

Similarly to F M – see Eq. (4.39)<br />

FT k “ `1 ` BT<br />

k ˘0.7 lnp1 ` BT k q . (4.53)<br />

˘<br />

And L`Tl<br />

k is the latent heat <strong>of</strong> vaporization at droplet surface temperature T<br />

k<br />

l .<br />

Finally, the rate <strong>of</strong> change <strong>of</strong> droplet temperature can be written as<br />

B k T<br />

dT k<br />

l<br />

dt<br />

“<br />

Q k l<br />

4<br />

3 πρk l c p,l k R k l<br />

3 , (4.54)<br />

where c p,l is the specific heat for droplet.<br />

4.4 Turbulence Model<br />

Since the mixing model <strong>and</strong> Langevin models are affected by turbulent effect through<br />

the turbulent time scale τ ” k{ɛ, a suitable approach is required to determine turbulent<br />

kinetic energy k <strong>and</strong> its dissipation rate ɛ. For simple flow, τ can be specified by assuming<br />

36


it to be uniform [69]. However, most flows needs a general determination <strong>of</strong> τ. In present<br />

work,the st<strong>and</strong>ard k-ɛ model [40, 47, 48] is selected to obtain k <strong>and</strong> ɛ. The k-ɛ model is<br />

one <strong>of</strong> the most widely applied two-equations turbulence model, in which model transport<br />

equations are solved for two turbulence quantities k <strong>and</strong> ɛ. Since the k-ɛ model is widely<br />

used solve turbulence flow problems, it has been improved <strong>and</strong> modified to overcome its<br />

well known shortcomings, e.g., poor predictions for swirling <strong>and</strong> rotating flows <strong>and</strong> certain<br />

unconfined flows, valid only for fully developed turbulent flows <strong>and</strong> inaccurate near the<br />

wall. Representative variants are The RNG k-ɛ [99] model <strong>and</strong> Realizable k-ɛ model [83].<br />

Instead <strong>of</strong> variants, the st<strong>and</strong>ard k-ɛ model [40, 48, 47] <strong>and</strong> its model constants are<br />

implemented in present work with transport equations for k as<br />

<strong>and</strong> for dissipation rate ɛ<br />

B<br />

Bt pρ Gkq ` B pρ G kV G,i q “ B „ˆ ˙ j<br />

µ ` µt Bk<br />

Bx i Bx j σ k Bx j<br />

B<br />

Bt pρ Gɛq ` B pρ G ɛV G,i q “ B „ˆ ˙ j<br />

µ ` µt Bɛ<br />

Bx i Bx j σ ɛ Bx j<br />

Modelled turbulent viscosity µ t is defined as<br />

And production <strong>of</strong> k is defined as<br />

`<br />

µ t “ ρ G C µ<br />

k 2<br />

` P k ` P b ´ ρ G ɛ ´ Y M , (4.55)<br />

C 1ɛ<br />

ɛ<br />

k pP k ` C 3ɛ P b q ´ C 2ɛ ρ G<br />

ɛ 2<br />

k . (4.56)<br />

ɛ . (4.57)<br />

P k “ ´ρ G VG,i 1 V G,j<br />

1 BV G,j<br />

Bx i<br />

“ µ t S 2 , (4.58)<br />

with the modulus <strong>of</strong> the mean rate-<strong>of</strong>-strain tensor S defined as<br />

S ” a 2S ij S ij , (4.59)<br />

where the mean strain rate S ij<br />

S ij “ 1 2<br />

ˆBVG,j<br />

Bx i<br />

` BV ˙<br />

G,i<br />

. (4.60)<br />

Bx j<br />

37


In the present k-ɛ formulation, effects <strong>of</strong> buoyancy P b <strong>and</strong> effects <strong>of</strong> compressibility on<br />

Turbulence Y m are neglected. And the model constants [48] for st<strong>and</strong>ard k-ɛ model are<br />

provided as<br />

C 1ɛ “ 1.44, C 2ɛ “ 1.92, C µ “ 0.09, σ k “ 1.0, σ ɛ “ 1.3<br />

38


Chapter 5<br />

Discretization<br />

5.1 Domain <strong>of</strong> Integration<br />

In the present work, both Lagrangian particle equations <strong>and</strong> Eulerian equations are solved<br />

employing the so-called method <strong>of</strong> fractional steps [69, 100, 101]. To this end, herein the<br />

spatial computational domain is first triangulized, <strong>and</strong> then control volumes are constructed<br />

from the triangles [63, 75]. Specifically, in the present work control volumes were<br />

constructed on the basis <strong>of</strong> a Delaunay triangulation [77, 96], T , consisting <strong>of</strong> N T triangles<br />

as schematically shown in Fig. 5.1.<br />

Figure 5.1: Section <strong>of</strong> an unstructured triangular mesh with control volumes.<br />

As an example, shown in Fig. 5.1 is a triangle with the vertices i, j, <strong>and</strong> k. Point C marks<br />

the center <strong>of</strong> gravity <strong>of</strong> this triangle. It is seen that the three straight lines running from<br />

39


C to the midpoints <strong>of</strong> the three triangle edges constitute parts <strong>of</strong> the control volumes V i ,<br />

V j , <strong>and</strong> V k around nodes i, j <strong>and</strong> k, respectively. The whole control volumes V i , V j , <strong>and</strong><br />

V k are then made up <strong>of</strong> “pieces <strong>of</strong> cake” stemming from triangles centered about i, j <strong>and</strong><br />

k, respectively [63, 75].<br />

Following the principle <strong>of</strong> construction <strong>of</strong> control volumes just outlined, control volumes<br />

centered about boundary points <strong>of</strong> the domain can be constructed. Naturally, parts <strong>of</strong><br />

such control volumes coincide with particular triangle edges.<br />

5.2 Eulerian Governing Equations<br />

Governing equations in Eulerian form are the gasphase equations governing the pressure<br />

p, <strong>and</strong> the equations <strong>of</strong> the k-ɛ turbulence model. The approach to the discretization <strong>of</strong><br />

these equation is through linear triangle shape functions. Specifically, each dependent<br />

variable φpx, tq – where φ “ p, k, ɛ, ũ, ṽ, ... – is written in terms <strong>of</strong> shape functions α φ , β φ<br />

<strong>and</strong> γ φ , viz.,<br />

φpx, tq “ α φ ptq ` β φ ptqx ` γ φ ptqy , (5.1)<br />

where<br />

α φ ptq “<br />

β φ ptq “<br />

γ φ ptq “<br />

3ÿ<br />

α k φ k ptq , (5.2)<br />

k“1<br />

3ÿ<br />

β k φ k ptq , (5.3)<br />

k“1<br />

3ÿ<br />

γ k φ k ptq , (5.4)<br />

k“1<br />

<strong>and</strong> where φ k denotes the value <strong>of</strong> variable φ at triangle vertex k, k “ 1, 2, 3.<br />

The coefficients α k , β k <strong>and</strong> γ k depend only on the triangle geometry <strong>and</strong> hence are inde-<br />

40


pendent <strong>of</strong> time; they are given by<br />

α 1 “ px 2 y 3 ´ x 3 y 2 q{det , (5.5)<br />

α 2 “ px 3 y 1 ´ x 1 y 3 q{det , (5.6)<br />

α 3 “ px 1 y 2 ´ x 2 y 1 q{det , (5.7)<br />

β 1 “ py 2 ´ y 3 q{det , (5.8)<br />

β 1 “ py 3 ´ y 1 q{det , (5.9)<br />

β 1 “ py 1 ´ y 2 q{det , (5.10)<br />

γ 1 “ px 3 ´ x 2 q{det , (5.11)<br />

γ 1 “ px 1 ´ x 3 q{det , (5.12)<br />

γ 1 “ px 2 ´ x 1 q{det , (5.13)<br />

where<br />

det “ px 2 y 3 ´ x 3 y 2 q ` px 3 y 1 ´ x 1 y 3 q ` px 1 y 2 ´ x 2 y 1 q . (5.14)<br />

From (5.1), the partial derivatives with respect to x <strong>and</strong> y, are obtained as<br />

Bφ<br />

Bx “ βφ ptq “<br />

Bφ<br />

By<br />

“ γ φ ptq “<br />

3ÿ<br />

β k φ k ptq , (5.15)<br />

k“1<br />

3ÿ<br />

γ k φ k ptq . (5.16)<br />

k“1<br />

5.2.1 Poisson Equation for Pressure<br />

Presented in App. A.1 is the detailed derivation <strong>of</strong> the discretized Poisson equation for<br />

pressure. In the appendix, first a differential form <strong>of</strong> the Poisson equation for pressure is<br />

is derived, i.e.,<br />

∆t B2 p 1<br />

Bx 2 i<br />

which in the appendix is labelled (A.9).<br />

“ Bρ<br />

Bt ` BρV i<br />

˚<br />

, (5.17)<br />

Bx i<br />

The result <strong>of</strong> the derivation <strong>of</strong> the discretized Poisson equation for p is<br />

˘ ÿ ÿ`An p 1 n “ Bn , (5.18)<br />

which in the appendix is labelled (A.21). In (5.18), the summation on both the l.h.s. <strong>and</strong><br />

the r.h.s. extends over all neighboring triangles <strong>of</strong> the control-volume defining node n.<br />

The coefficients A n <strong>and</strong> B n are defined in App. A.1.<br />

41


5.2.2 Turbulence Model<br />

Presented in App. A.2 is the detailed derivation <strong>of</strong> the discretized equations for turbulent<br />

kinetic energy k <strong>and</strong> dissipation rate ɛ.<br />

The result <strong>of</strong> the derivation <strong>of</strong> the discretized equations for k <strong>and</strong> ɛ are<br />

ÿ`Apkq<br />

ÿ<br />

n k n˘<br />

“ B<br />

pkq<br />

n , (5.19)<br />

ÿ`Apɛq n ɛ n˘<br />

“<br />

ÿ<br />

B<br />

pɛq<br />

n , (5.20)<br />

which in the appendix are labelled (A.34) <strong>and</strong> (A.35) respectively. In (5.19) <strong>and</strong> (5.20),<br />

the summation on both the l.h.s. <strong>and</strong> the r.h.s. extends over all neighboring triangles <strong>of</strong><br />

the control-volume defining node n. The coefficients A pkq<br />

n , Bn<br />

pkq , A pɛq<br />

n <strong>and</strong> Bn<br />

pɛq are defined<br />

in App. A.2.<br />

5.3 Lagrangian Particle Equations<br />

With particle method, the density functions derived in Chap. 2.3.2 <strong>and</strong> 2.3.3 is represented<br />

by a finite number <strong>of</strong> stochastic particles. Each particle has a set <strong>of</strong> properties. The<br />

properties are mass m p , position x p i , velocity V p<br />

i , mass fraction Y p α , temperature T p .<br />

From a given initial condition at t “ t 0 , the particle properties are varied in time with<br />

the reasonably small time-steps ∆t. To determine the time-step size has used a suitably<br />

formulated Courant–Friedrichs–Lewy (CFL) condition [8] as<br />

„? j<br />

AT<br />

∆t “ C CFL ¨ min , (5.21)<br />

|V i |<br />

where A T denotes area <strong>of</strong> triangle T <strong>and</strong> V i contains both gasphase <strong>and</strong> liquid phase<br />

velocity. In present work, the courant constant C CFL is taken as 0.7. The restricted<br />

time-step size indicates that a fluid particle can not move more than one triangle in single<br />

time-step. The particle properties are updated simultaneously by several different process,<br />

but the method <strong>of</strong> fractional steps [69, 100, 101] is used so that the effects <strong>of</strong> some <strong>of</strong> these<br />

processes can be treated sequentially.<br />

The gas particle position <strong>and</strong> velocity are updated as<br />

x p g,i<br />

`t ` ∆t˘<br />

“ x<br />

p<br />

`t ` ∆t˘<br />

“ V<br />

p<br />

V p<br />

g,i<br />

g,i<br />

g,i<br />

`t˘<br />

` V<br />

p<br />

g,iptq∆t (5.22)<br />

`t˘<br />

` ∆V<br />

p<br />

ptq∆t . (5.23)<br />

g,i<br />

42


∆V p<br />

g,i ptq “ ´ 1<br />

`<br />

ρ p g<br />

„ˆ Bp<br />

Bx i<br />

˙p<br />

ˆ1<br />

2 ` 3<br />

4 C 0<br />

j<br />

p<br />

´ S g,v ∆t<br />

˙ ɛ p<br />

`V g,i<br />

k ´ xV g,iy p˘ ∆t ` pC 0 ɛq 1{2 ∆W i , (5.24)<br />

with S v p denotes momentum exchange between the two phase specific at particle p as<br />

S g,v<br />

p<br />

“<br />

`<br />

1<br />

VP g<br />

ÿP l<br />

p“1<br />

˜<br />

1 ÿ Pl<br />

VP g<br />

p“1<br />

pxV l,i y p ´ xV g,i y p q 9m p l<br />

(5.25)<br />

1<br />

2 ρp l<br />

`xVg,i y p ´ V p<br />

l,i<br />

˘| xVg,i y p ´ V p<br />

l,i |C DA p l<br />

¸<br />

, (5.26)<br />

where mean quantities pBp{Bx i q p , xV g,i y p <strong>and</strong> xV l,i y p are interpolated into particle position<br />

x p g,i as<br />

ˆ Bp<br />

Bx ` Bp<br />

˙p<br />

By<br />

“ β p ` γ p (5.27)<br />

xV g,i y p “ α xV g,iy ` β xV g,iy x p g ` γ xV g,iy y p g (5.28)<br />

xV l,i y p “ α xV l,iy ` β<br />

xV l,iy x<br />

p<br />

l ` γxV l,iy y<br />

p<br />

l , (5.29)<br />

where α φ , β φ <strong>and</strong> γ φ are defined in Chap. 5.2. And the mass exchange rate 9m p g is written<br />

as<br />

9m p l<br />

“ 4π ρ p l Rp 2 dR l<br />

l<br />

dt , (5.30)<br />

where the particle radius time rate <strong>of</strong> change dR l {dt has been modelled suitable evaporation<br />

model considered in Chap. 4.3.<br />

The species mass <strong>of</strong> gas particles is evaluated by mixing models – see Chap. 4.2 – ,<br />

chemical source term <strong>and</strong> phase exchange term as<br />

∆ψg,α p “ ´1<br />

2 C ɛ<br />

ψ<br />

`ψp<br />

k g,α ´ xψ g,α y˘∆t ` Spψg,αq∆t p ` Sg,m,α∆t p , (5.31)<br />

where<br />

S p g,m,α “ 1 P g<br />

ÿP l<br />

p“1<br />

Y p<br />

l,α 9mp l . (5.32)<br />

Chemical source term Spψ p g,αq does not need any model as discussed in Chap. 4.2. To<br />

calculate chemical reaction term, the s<strong>of</strong>tware package COSILAB [77] <strong>and</strong> its API libraries<br />

43


has been used.<br />

The temperature <strong>of</strong> gas particles is evaluated as<br />

with<br />

˜ ¸<br />

Kÿ<br />

∆Tg p “ ´1<br />

2 C ɛ p<br />

ψ<br />

`T<br />

k g ´ xT g y p˘∆t ´ 1<br />

c p ρ p h p αwα p ` Sg,e<br />

p ∆t , (5.33)<br />

g<br />

α“1<br />

˜<br />

Sg,e p “ 1 1<br />

V<br />

ÿP l<br />

Q p l<br />

P g p“1<br />

where the heat exchange term Q p l<br />

is addressed in Chap. 4.3.2.<br />

Similarly, the liquid particle position <strong>and</strong> velocity are updated as<br />

x p l,i<br />

`t ` ∆t˘<br />

“ x<br />

p<br />

l,i<br />

`t ` ∆t˘<br />

“ V<br />

p<br />

V p<br />

l,i<br />

l,i<br />

¸<br />

, (5.34)<br />

`t˘<br />

` V<br />

p<br />

l,iptq∆t (5.35)<br />

`t˘<br />

` ∆V<br />

p<br />

ptq∆t , (5.36)<br />

l,i<br />

with<br />

∆V p<br />

l,i ptq “ ´ 1 „ˆ ˙p j<br />

Bp<br />

ρ p p<br />

´ S v ∆t . (5.37)<br />

g Bx i<br />

Since the liquid particles has single component in the present work, contrary to gas particles,<br />

the species mass <strong>of</strong> liquid particles per volume is constant during computation. The<br />

radius <strong>of</strong> liquid particles is evaluated by suitable evaporation model, for example with<br />

D-square law [28, 49, 93, 97] as<br />

where the evaporation constant K is defined in Chap. 4.3.1.<br />

liquid particles is evaluated as<br />

5.4 Particle Tracing<br />

∆R p l<br />

“ ´ K<br />

R p ∆t , (5.38)<br />

l<br />

And the temperature <strong>of</strong><br />

∆T p<br />

l<br />

“ 1<br />

c p,l ρ p Se p ∆t . (5.39)<br />

l<br />

On their way through the computational domain, particles move through a number <strong>of</strong><br />

control volumes <strong>and</strong> hence triangles. Obviously, the numerical-solution method “needs<br />

to know” where, at any time, each particle is located. To this end, a suitable method <strong>of</strong><br />

particle tracing is required.<br />

44


In the present work, a particle-tracing method consisting <strong>of</strong> two subsequent steps has<br />

been developed. The first step was devised by Löhner [52, 53].<br />

In the first step, for any particle p P t1, 2, ..., P u at time t pn`1q , the host triangle T j ,<br />

j “ 1, ..., N T , is sought. If at time t pnq the host triangle was T i , than at t pn`1q either<br />

T j “ T i or T j P tTn i<br />

i1<br />

, ..., , Tn i<br />

iP<br />

u, where n iP denotes the number <strong>of</strong> direct neighbors <strong>of</strong><br />

triangle T i , provided a suitably formulated Courant–Friedrichs–Lewy (CFL) condition<br />

has been used in the determination <strong>of</strong> the time-step size ∆t :“ t pn`1q ´ t pnq –see Eq.<br />

(5.21).<br />

Specifically, consider the left picture <strong>of</strong> Fig. 5.2, <strong>and</strong> assume – for the moment – that the<br />

Figure 5.2: An unstructured triangle with physical coordinates <strong>and</strong> the mapped into<br />

natural coordinate ξ <strong>and</strong> η.<br />

picture applies to a particle p located at time t pnq at point P “ px, yq inside the depicted<br />

triangle. To establish whether the particle at time t pn`1q is still located at a point P 1<br />

inside this triangle, the coordinates <strong>of</strong> P 1 “ px p , y p q pn`1q are expressed in terms <strong>of</strong> the<br />

natural coordinates<br />

ξ “<br />

η “<br />

1 `yki x pn`1q<br />

p<br />

x ji y ki ´ y ji x ki<br />

1 `´yji xp<br />

pn`1q<br />

x ji y ki ´ y ji x ki<br />

˘<br />

´ x ki yp<br />

pn`1q , (5.40)<br />

˘<br />

` x ji yp<br />

pn`1q , (5.41)<br />

with the notation x ij “ x i ´ x j . The particle is located inside the triangle, if <strong>and</strong> only if<br />

the shape functions [52, 53]<br />

N 1 “ 1 ´ ξ ´ η , N 2 “ ξ , N 3 “ η (5.42)<br />

satisfy<br />

max`N 1 , N 2 , N 3˘<br />

ď 1 <strong>and</strong> min`N1 , N 2 , N 3˘<br />

ě 0 . (5.43)<br />

45


If based on this criterion it is found that the particle has left the triangle, all neighbouring<br />

triangles have to be investigated as possible new host-triangle; the search is stopped once<br />

the new host triangle has been found. That a host triangle is indeed a member <strong>of</strong> set<br />

tT i<br />

n i1<br />

, ..., , T i<br />

n iP<br />

u is ensured by formulating <strong>and</strong> applying a suitably formulated Courant–<br />

Friedrichs–Lewy (CFL) condition in the determination <strong>of</strong> the time-step size ∆t :“ t pn`1q ´<br />

t pnq .<br />

In the second step, for each particle p the new host-triangle T j found in the first step<br />

is investigated to find out to which <strong>of</strong> the possible 3 control volumes P belongs at time<br />

Figure 5.3: A mapping to determine control volume which the particle is located.<br />

t pn`1q . Referring to the right picture in Fig. 5.3, this is accomplished by evaluating the<br />

angle θ between vector ÝÝÝÑ C 1 H jk <strong>and</strong> ÝÝÑ C 1 P 1 . Specifically, if the cross product ÝÝÝÑ C 1 H jk ˆ ÝÝÑ C 1 P 1 is<br />

positive, then θ is measured counterclockwise, otherwise clockwise. Obviously, if θ is in<br />

the range <strong>of</strong> ?H jk C 1 H ki <strong>and</strong> θ is measured counterclockwise, then particle p is located in<br />

control volume k. Similarly, if θ is in the range <strong>of</strong> ?H jk C 1 H ki <strong>and</strong> θ is measured clockwise,<br />

then p is located in j.<br />

5.5 Particle Controlling<br />

Since the local <strong>and</strong> instantaneous (turbulent) expectation values <strong>of</strong> the various physical<br />

quantities are calculated from a finite number <strong>of</strong> particles in each control volume, the<br />

statistical error in forming the expectations is <strong>of</strong> the order P ´1{2 [71, 79, 98] Here P<br />

st<strong>and</strong>s for either P g or P l , respectively. As can be seen from Fig. 5.4, with increasing P<br />

the statistical error decreases, however, the efficiency <strong>of</strong> the numerical algorithm expressed<br />

in terms <strong>of</strong> required CPU time (speed-down time) increases. Referring specifically to the<br />

numbers displayed Fig. 5.4, the optimum number <strong>of</strong> particles lies at the cross-over <strong>of</strong> the<br />

two curves, i.e., at P « 28.<br />

46


1.0<br />

Statistical Error<br />

Speed down<br />

50<br />

40<br />

Statistical Error<br />

0.5<br />

30<br />

20<br />

10<br />

Speed down<br />

0.0<br />

10 30 50 70 90 110<br />

Number <strong>of</strong> Particles<br />

0<br />

Figure 5.4: Statistical error <strong>and</strong> speed down in a control volume as a function <strong>of</strong> the<br />

number <strong>of</strong> particles in a control volume.<br />

In the course <strong>of</strong> a computation, the number <strong>of</strong> particles evolves dynamically. As a<br />

consequence, – assuming that initially particles were distributed spatially in a quasihomogenously<br />

manner 1 – with increasing time certain areas <strong>of</strong> the computational domain<br />

may be thinned out <strong>of</strong> particles whilst in other areas particles may cluster.<br />

To keep during a computation the number <strong>of</strong> particles per unit finite volume approximately<br />

constant, <strong>and</strong> hence to ensure that the statistical error remains approximately<br />

uniformly distributed throughout the physical domain, in the present work the particle<br />

cloning <strong>and</strong> annihilation method [31, 91, 105] is employed. According to this method, particles<br />

in the finite control volumes V are cloned or annihilated such that in each control<br />

volume P varies between an upper <strong>and</strong> lower bound, i.e.,<br />

P min ď P ď P max . (5.44)<br />

Thus, after monitoring each control volume, particles are cloned in a control volumes<br />

having P ă P min by splitting a particle located in the identical control volume.<br />

most popular cloning procedure is that a particle is selected r<strong>and</strong>omly for cloning (with<br />

preference given to high-mass particles), <strong>and</strong> then is divided into two particles having the<br />

same properties as the original particle except particle mass. If the mass <strong>of</strong> the original<br />

1<br />

In this context quasi-homogenous means that for all finite volumes V the number <strong>of</strong> particles per<br />

unit finite volume is approximately the same.<br />

The<br />

47


particle is m, then the mass <strong>of</strong> each cloned particle is m{2. 2<br />

cloning does not affect overall mass conservation.<br />

Hence it is ensured that<br />

Particles are annihilated in control volumes having P ą P max by merging three particles<br />

selected at r<strong>and</strong>om (with preference given to low-mass particles) into two particles<br />

having exactly preserved particle mass <strong>and</strong> mean particle properties. 3 Since this annihilation<br />

method leads to artificial mixing, it does not exactly preserve the particle property<br />

distribution, however it conserves the properties <strong>of</strong> particles at the microscopic level [105].<br />

5.6 Initial <strong>and</strong> Boundary Conditions<br />

5.6.1 Initial <strong>and</strong> Boundary Conditions for Particles<br />

At the inlet, the Monte-Carlo particles are created <strong>and</strong> activated according to the inlet<br />

flow properties. The total mass <strong>of</strong> the new gas particles M P is set to be the mass flux<br />

during the current time step,<br />

M P “ P new m p “ ρ inlet V x,inlet A inlet ∆t . (5.45)<br />

When the particle moves across the axis <strong>of</strong> symmetry, the particle is reflected from the<br />

boundary without change in their properties except for the velocity <strong>and</strong> position in the<br />

direction normal to the boundary. When the particle moves across the open boundary,<br />

the particle is discarded <strong>and</strong> at the wall, the velocity is forced to zero.<br />

5.6.2 Boundary Conditions for Eulerian Equations<br />

The flows under consideration in present work require boundary conditions to be prescribed<br />

along the entire boundary line or surface surrounding the solution domain. No-slip<br />

conditions are applied for the velocities <strong>and</strong> zero normal gradients for scalar variables at<br />

wall boundary. At the symmetry axis, the velocity components normal to the boundary<br />

are set to zero. At the outlet boundary, the pressure is fixed as ambient pressure 1 bar.<br />

At the open boundary, the flow is allowed to adjust its direction into the compute domain<br />

2<br />

Note that after cloning the position <strong>of</strong> the two new particles is identical to the position <strong>of</strong> the<br />

particle cloned.<br />

3<br />

Note that after annihilation the position <strong>of</strong> the two remaining particles could be taken as unchanged<br />

or, alternatively, as the mean position <strong>of</strong> the original three particles. Since in the computations herein<br />

it was found that, within temporal <strong>and</strong> spatial discretization errors neither instantaneous, local values<br />

nor instantaneous expectation values were affected, the first alternative was selected because in terms <strong>of</strong><br />

computational operations it is more efficient.<br />

48


or out <strong>of</strong> it. At inlet boundary, the velocity, composition, <strong>and</strong> turbulent properties <strong>of</strong><br />

gasphase <strong>and</strong> liquid phase are fixed to the initial values.<br />

For the turbulence models also require the specification <strong>of</strong> certain variable values at the<br />

boundaries. There are several methods used for turbulence specification at the boundaries,<br />

<strong>and</strong> normally those methods need the additional information <strong>of</strong> the turbulence intensity,<br />

turbulence length scale or the hydraulic diameter.<br />

In present work, the turbulent quantities at the boundaries, especially inlet, are specified<br />

as [66]<br />

k “ 3 `<br />

¯Vi,inlet I˘2 , (5.46)<br />

2<br />

ɛ “ 0.1643k1.5<br />

l<br />

. (5.47)<br />

The turbulence intensity, I, is defined as the ratio <strong>of</strong> the root-mean-square <strong>of</strong> the velocity<br />

fluctuations, V 1<br />

i , to the mean velocity, xV i y as<br />

I “ V i<br />

1<br />

xV i y . (5.48)<br />

For internal flows the value <strong>of</strong> turbulence intensity can be fairly high with values ranging<br />

from 1% - 10% being appropriate at the inlet <strong>and</strong> for external flows the value <strong>of</strong> turbulent<br />

intensity can be as low as 0.05%.<br />

5.7 Overall Solution Algorithm<br />

In the present work, we combine the advantage <strong>of</strong> a full Monte-Carlo stochastic description<br />

for chemistry, which does not require modeling, with the advantage <strong>of</strong> a full Monte-Carlo<br />

stochastic description <strong>of</strong> both gaseous <strong>and</strong> liquid phase flow. Fundamental idea <strong>and</strong> one<br />

dimensional spray simulation studies was done by Rumberg [79, 80], <strong>and</strong> in the present<br />

work, it has been extended to two-dimensional reactive spray code with various selection <strong>of</strong><br />

fuel mechanism, e.g., detailed chemistry, reduced chemistry or global single-step chemistry<br />

reaction mechanism. In contrast to <strong>PDF</strong>-FVM hybrid concept algorithm [4, 26, 59, 90],<br />

most flow fields are obtained from expectation <strong>of</strong> particle properties except the pressure<br />

p, the turbulent kinetic energy k, <strong>and</strong> the dissipation rate ɛ as schematically shown in<br />

Fig. 5.5. This Figure shows also relationship between three main variable fields. Each<br />

particle has own velocity <strong>and</strong> composition properties. From those particle properties, the<br />

expectations are obtained stochastically. On the other h<strong>and</strong>, xpy, k <strong>and</strong> ɛ are calculated<br />

49


Figure 5.5:<br />

Schematic <strong>of</strong> the fully stochastic particle method.<br />

by solving Eulerian governing equations – see Chap. 5.2. Unlike general FVM methods,<br />

velocity components <strong>and</strong> compositions are known quantities – expectations from particles<br />

– to solve the governing equation. for xpy, k <strong>and</strong> ɛ. Moreover, the particle properties are<br />

updated by those Eulerian fields.<br />

In Fig. 5.6, the flowchart <strong>of</strong> overall algorithm is schematically reported. The overall<br />

algorithm can be divided up two region, i.e., Lagrangian <strong>and</strong> Eulerian part. Detailed for<br />

each task in the algorithm has been considered in the present thesis, e.g., calculate time<br />

step size, calculate the Eulerian variables, trace particles, <strong>and</strong> control number <strong>of</strong> particles<br />

have been introduced Chap. 5.3, App. A, Chap. 5.4 <strong>and</strong> Chap. 5.5, respectively.<br />

50


Figure 5.6:<br />

Flowchart <strong>of</strong> the overall computation algorithm.<br />

51


Chapter 6<br />

Parallelization<br />

The numerical simulation <strong>of</strong> a complex flow field requires a high computing power, especially<br />

for a combustion problem. Since the efficiency <strong>of</strong> computing is mainly dominated by<br />

the performance <strong>of</strong> the central processing units – commonly such a unit is referred to as<br />

CPU – ongoing development <strong>of</strong> CPUs affects the implementation <strong>of</strong> numerical algorithms.<br />

For more than two decades, factually, the density <strong>of</strong> transistors – which is related directly<br />

to the performance <strong>of</strong> a CPU – has doubled every two years. Since the mid 2000s,<br />

however, increasing computing power through more dense transistors had certain reached<br />

limitations <strong>and</strong>, therefore, most <strong>of</strong> the major CPU manufacturers began putting multiple<br />

processors on a single circuit for increasing performance by parallelism. Because <strong>of</strong> the<br />

changed roadmap <strong>of</strong> CPU manufacturers, parallel algorithms have become an essential<br />

point in numerical simulations.<br />

In this chapter, first fundamental parallelization theories, strategies <strong>and</strong> applications are<br />

discussed. Then suitable parallelization methods are selected <strong>and</strong> applied to parallelize<br />

the Monte-Carlo particle method employed in the present work.<br />

6.1 Parallelization Architecture <strong>and</strong> Strategies<br />

6.1.1 Parallel Architecture<br />

According to the so-called Flynn’s taxonomy [22], see Fig.6.1 – classically parallel computer<br />

architectures can be categorized according to the number <strong>of</strong> instruction streams <strong>and</strong><br />

the number <strong>of</strong> data streams which the system can h<strong>and</strong>le at the same time. For example,<br />

a computer program may be executed by following a single series <strong>of</strong> programming statements<br />

(single instruction stream) for one processor or, alternatively, it may be executed by<br />

52


Figure 6.1: Classification <strong>of</strong> parallel computer architectures by Flynn [22].<br />

splitting the overall programming statements into multiple series <strong>of</strong> programming statements,<br />

one series for each processor (multiple instruction streams). On the other h<strong>and</strong>,<br />

input data for a computer program may be provided as a whole to one processor (single<br />

data stream) or, alternatively, overall input data may be split up into multiple input data<br />

(multiple data streams) such that each processor gets its own input data. As specified by<br />

Flynn, the classical von Neumann system has a single instruction stream <strong>and</strong> single data<br />

stream <strong>and</strong>, therefore, can be classified as SISD.<br />

SIMD (single instruction, multiple data) execute identical instructions for all multiple<br />

data sets synchronously. In other words, each processor has the same task with different<br />

data streams. Main characteristics <strong>of</strong> SIMD should execute the same instruction whether<br />

it has some data to h<strong>and</strong>le or not can lower the overall performance <strong>of</strong> the system, therefore<br />

SIMD is difficult to apply on general purpose. However, SIMD concept is ideal for<br />

parallelizing simple loops that operate on large arrays <strong>of</strong> data so called data-parallelism,<br />

<strong>and</strong> with this strength, vector processors which is an application <strong>of</strong> SIMD concept was<br />

used dominantly for supercomputer during the 1990s [62].<br />

One <strong>of</strong> many other SIMD applications is the graphics processing unit (GPU) which was<br />

originally designed to exclusively manipulate computer graphics <strong>and</strong> image processing.<br />

Recently, GPUs have been developed that have many cores which can execute individual<br />

instructions <strong>and</strong>, therefore, GPUs now are also utilized for general high performance<br />

computing. A GPU that also serves such generalized purposes is called a General-Purpose<br />

Computing on Graphics Processing unit (GPGPU).<br />

Different from SIMD systems, multiple instruction, multiple data (MIMD) systems carry<br />

multiple concurrent instructions which operate on multiple data streams. Depending<br />

upon how memory is accessed, MIMD systems can be classified with distributed memory<br />

system <strong>and</strong> shared memory systems as discussed in the next section.<br />

53


6.1.2 Distributed Memory System <strong>and</strong> Shared Memory System<br />

MIMD systems are most popular in parallelization concepts because <strong>of</strong> their versatility.<br />

There are two principal subtypes <strong>of</strong> MIMD systems, viz., distributed memory systems<br />

<strong>and</strong> shared memory systems.<br />

Figure 6.2: Schematic <strong>of</strong> distributed memory system.<br />

Figure 6.3: Schematic <strong>of</strong> shared memory system.<br />

As shown schematically in Fig. 6.2, in a distributed memory system, each processor basically<br />

has its own memory space, <strong>and</strong> consequently can access it independently <strong>of</strong> the<br />

other processors. Thus, each processor modifies data locally, <strong>and</strong> data exchanged between<br />

processors occur explicitly via an interconnection network.<br />

The so-called Message Passing Interface (MPI) is a communication protocol for programming<br />

<strong>of</strong> distributed memory systems. As its name implies, with MPI processor-memory<br />

pairs can communicate by message passing: one pair sends the message (data) by calling a<br />

54


“send” function <strong>and</strong> an other pair or several pairs receive the message (data) by calling a<br />

“receive” function. There are several MPI implemented libraries for C, C++ <strong>and</strong> Fortran,<br />

e.g., MPICH, OpenMPI <strong>and</strong> MSMPI.<br />

In a shared memory system, a group <strong>of</strong> processors is connected to a common memory system<br />

via an internal network as schematically shown in Fig 6.3. Principally, the processors<br />

communicate by accessing common memory with equal priority.<br />

An example <strong>of</strong> sharing memory is Open Multi-Processing (OpenMP) which is a very<br />

popular applications programming interface (API). It can be used in programs written in<br />

C, C++ <strong>and</strong> Fortran, that are intended for use on shared memory systems.<br />

6.1.3 Parallelization Strategies<br />

To obtain reasonable parallelization effects in a CFD code in terms <strong>of</strong> speedup, say,<br />

individual code sections need to be identified <strong>and</strong> analyzed with respect to their suitability<br />

Figure 6.4: An example <strong>of</strong> executing time segments.<br />

for parallelization. Schematically shown in Fig. 6.4 are sections <strong>of</strong> a CFD code such as<br />

– ordered according to their serial CPU time requirement – solution <strong>of</strong> source terms,<br />

solution <strong>of</strong> pressure equation, preparation <strong>of</strong> initial data, <strong>and</strong> so on.<br />

For a variety <strong>of</strong> reasons, not every section in a computer program is suitable to be subjected<br />

to parallelization. Thus, prior to code parallelization, suitable sections need to be<br />

identified. After identification <strong>and</strong> selection <strong>of</strong> meaningful code sections, suitable parallelization<br />

approaches <strong>and</strong> parallel program models should be considered. To this end a<br />

very effective <strong>and</strong> simple approach is to break the most CPU-time intensive section <strong>of</strong><br />

55


the code – in the example <strong>of</strong> Fig. 6.4 this is the section “solution <strong>of</strong> source terms” –<br />

into temporally or spatially discrete segments that can be h<strong>and</strong>led by multiple processes.<br />

Depending on the partitioning target, basically two parallelization methods are available,<br />

i.e., the so-called domain decomposition <strong>and</strong> the so-called functional decomposition.<br />

In the present work, these two methods <strong>and</strong> respective programming models have been<br />

implemented, namely,<br />

ˆ the overall computational domain <strong>and</strong> the total number <strong>of</strong> particles have been decomposed<br />

by employing MPI,<br />

ˆ the most repetitively used functions have been decomposed by employing OpenMP.<br />

Further details will be discussed in the following sections.<br />

6.2 Domain <strong>and</strong> Particles Decomposition<br />

In the domain decomposition approach to a problem partitioning, we first seek to decompose<br />

the data associated with the problem. If possible, we divide these data into small<br />

pieces <strong>of</strong> approximately equal size. Next, we partition the computation that is to be performed,<br />

typically by associating each operation with the data on which it operates. This<br />

partitioning yields a number <strong>of</strong> tasks, each comprising some data <strong>and</strong> a set <strong>of</strong> operations<br />

on that data. An operation may require data from several tasks. In this case, communication<br />

is required to move data between tasks. In the present work, this requirement is<br />

h<strong>and</strong>led by MPI.<br />

MPI is an API library containing functions for communicating so-called messages, or<br />

data, between processors. These exchange messages can be between just two processors<br />

or, alternatively, messages can be collectively sent or received by several processors. MPI<br />

defines a so-called Communicator, which has associated with it a Processor Group <strong>and</strong> a<br />

Context. A Processor Group consists <strong>of</strong> a subset <strong>of</strong> the processors available to the system,<br />

<strong>and</strong> a Context is a tag given by system, whose aim is to distinguish the processors. The<br />

MPI library <strong>of</strong>fers many functions for common or complicated computing assignment.<br />

The most basic functions are<br />

1. MPI INIT, initializing MPI,<br />

2. MPI FINALIZE, finishing MPI,<br />

3. MPI COMM SIZE, get the number <strong>of</strong> available processors,<br />

56


Figure 6.5: Schematic <strong>of</strong> MPI.<br />

4. MPI COMM RANK, get the identity <strong>of</strong> a processor,<br />

5. MPI SEND, send a message,<br />

6. MPI RECV, receive a message.<br />

Since the overall program is divided up into subtasks by MPI, a certain processor should<br />

control the overall procedure. This processor is so-called “master” processor <strong>and</strong> the<br />

others are so-called “slave” processors. A master processor can, eventually, in addition<br />

also work as a slave shown as schematically in Fig. 6.5.<br />

6.2.1 Schur-Complement Method<br />

The Schur-complement method describes a mathematical algorithm that is not limited<br />

to applications in parallel solution <strong>of</strong> CFD methods. However, in the present work we<br />

will apply it to a particular class <strong>of</strong> parallelization methods, viz., the so-called class <strong>of</strong><br />

domain decomposition methods. As the name says, the fundamental idea underlying<br />

domain-decomposition methods is, that the overall domain <strong>of</strong> solution <strong>of</strong> a CFD problem<br />

is split up into subdomains such that (usually) each single processor solves (nearly) the<br />

underlying problem for a single subdomain.<br />

To be specific in the present example, let us assume that we have carried out a discretization<br />

for N “ 9 grid points, i.e., that we have a mesh<br />

M “ tx L “ x 1 , x 2 , x 3 , x 4 , x 5 , x 6 , x 7 , x 8 , x 9 “ x R u , (6.1)<br />

57


Figure 6.6: Schematic <strong>of</strong> domain decomposition.<br />

where x L <strong>and</strong> x R denote the left boundary <strong>and</strong> the right boundary. Let P denote the<br />

number <strong>of</strong> processors which are available to us for simultaneous, parallel use to solve the<br />

problem. To be specific, in the present example let us assume P “ 3. Hence we decompose<br />

the computational domain into P “ 3 parts, e.g. into<br />

M p1q “ tx L “ x 1 , x 2 , x 3 , x 4 u ,<br />

M p2q “ tx 4 , x 5 , x 6 , x 7 u , (6.2)<br />

M p3q “ tx 7 , x 8 , x 9 “ x R u .<br />

Obviously, domain 1 <strong>and</strong> 2 have node x 4 in common, <strong>and</strong> domains 2 <strong>and</strong> 3 node x 7 . At<br />

the nodes x L <strong>and</strong> x R we apply the physical, or natural, boundary conditions mentioned<br />

above. Since node x 4 represents a right boundary to domain 1 <strong>and</strong> a left boundary to<br />

domain 2, x 4 is termed a logical boundary node or, alternatively, a ghost node. For similar<br />

reasons, also node x 7 is a logical boundary or ghost node. Thus the set <strong>of</strong> ghost nodes is<br />

M pGq “ tx 4 , x 7 u . (6.3)<br />

6.2.1.1 Linear System for Subdomain p – General Formulation<br />

For each subdomain p, the vector <strong>of</strong> unknowns is written as<br />

X ppq “<br />

˜<br />

X ppq<br />

N<br />

X ppq<br />

G<br />

¸<br />

. (6.4)<br />

Here, <strong>and</strong> below, the subscripts N <strong>and</strong> G denote the (still unknown) solution vector at a<br />

non-ghost <strong>and</strong> ghost node, respectively. The superscript p indicates, that the respective<br />

quantity refers to processor p, 1 ď p ď P . Analogous sub <strong>and</strong> super scripting applies to<br />

various vectors <strong>and</strong> matrices to be introduced below.<br />

58


For subdomain p we obtain the linear system<br />

A ppq X ppq “ B ppq , (6.5)<br />

which, after introducing the overall ghost-node vector<br />

X G “ ppX pGq<br />

1 q T , ..., pX pGq<br />

N G<br />

q T q T (6.6)<br />

can be partitioned as<br />

˜<br />

A ppq<br />

NN<br />

A ppq<br />

GN<br />

Appq NG<br />

A ppq<br />

GG<br />

¸ ˜<br />

X ppq<br />

N<br />

X G<br />

¸<br />

“<br />

˜<br />

B ppq<br />

N<br />

B ppq<br />

G<br />

¸<br />

. (6.7)<br />

Here N G denotes the number <strong>of</strong> ghost nodes. Recall that for our specific example N G “ 2,<br />

with X pGq<br />

1 “ δ 4 <strong>and</strong> X pGq<br />

2 “ δ 7 ; the example-specific form <strong>of</strong> the various matrices <strong>and</strong><br />

vectors is left as a homework or exercise.<br />

6.2.1.2 The Linear System for the Overall Domain – General Formulation<br />

Assembly <strong>of</strong> the linear system for the overall domain yields – using the principle <strong>of</strong> superposition<br />

–<br />

¨<br />

diagpA p1q<br />

NN , Ap2q NN , ¨ ¨ ¨ , ApP q<br />

NN q ˚<br />

˝<br />

X p1q<br />

N<br />

.<br />

X pP q<br />

N<br />

˛<br />

‹<br />

‚`<br />

Pÿ<br />

p“1<br />

¨<br />

A ppq<br />

NG X G “ ˚<br />

˝<br />

B p1q<br />

N<br />

.<br />

B pP q<br />

N<br />

˛<br />

‹<br />

‚ (6.8)<br />

<strong>and</strong>, for 1 ď p ď P ,<br />

A ppq<br />

GN Xppq N ` Appq GG X G “ B ppq<br />

G . (6.9)<br />

Inspection shows that (6.8) <strong>and</strong> (6.9) can be combined to give<br />

¨<br />

A p1q<br />

NN 0 ¨ ¨ ¨ 0 0 ˛<br />

Ap1q NG<br />

0 A p2q<br />

NN ¨ ¨ ¨ 0 0 ¨<br />

Ap2q NG<br />

. .<br />

. . .<br />

pP ´1q<br />

pP ´1q<br />

0 0 ¨ ¨ ¨ ANN 0 A ˚<br />

NG<br />

˝<br />

˚<br />

˝ 0 0 ¨ ¨ ¨ 0 A pP q<br />

NN<br />

A pP q ‹<br />

NG ‚<br />

A p1q<br />

GN<br />

A p2q<br />

pP ´1q<br />

GN<br />

¨ ¨ ¨ AGN<br />

A pP q<br />

GN<br />

A GG<br />

X p1q<br />

.<br />

X pP q<br />

X G<br />

˛ ¨<br />

‹<br />

‚ “ ˚<br />

˝<br />

B p1q<br />

.<br />

B pP q<br />

B G<br />

˛<br />

‹<br />

‚ , (6.10)<br />

59


where for brevity <strong>of</strong> notation we have defined<br />

<strong>and</strong> where<br />

X ppq :“ X ppq<br />

N , (6.11)<br />

B ppq :“ B ppq<br />

N , (6.12)<br />

A GG :“<br />

B G :“<br />

Pÿ<br />

p“1<br />

Pÿ<br />

p“1<br />

A ppq<br />

GG , (6.13)<br />

B ppq<br />

G . (6.14)<br />

Note that the various matrices <strong>and</strong> vectors introduced so far have the following dimensions:<br />

1. matrix A ppq<br />

NN<br />

is <strong>of</strong> dimension N<br />

ppq<br />

N<br />

ˆ N ppq<br />

N ;<br />

2. matrix A ppq<br />

ppq<br />

NG<br />

is <strong>of</strong> dimension NN ˆ N G;<br />

3. matrix A ppq<br />

GN is <strong>of</strong> dimension N G ˆ N ppq<br />

N ;<br />

4. matrix A ppq<br />

GG is <strong>of</strong> dimension N G ˆ N G ;<br />

5. vectors X p1q to X pP q are <strong>of</strong> dimension N ppq<br />

N ;<br />

6. vectors B p1q to B pP q are <strong>of</strong> dimension N ppq<br />

N ;<br />

7. vectors B p1q<br />

G<br />

q<br />

to BpP<br />

G<br />

, <strong>and</strong> hence vector B G, are <strong>of</strong> dimension N G<br />

8. vector X G is <strong>of</strong> dimension N G .<br />

6.2.1.3 LU Decomposition <strong>of</strong> the Matrix <strong>of</strong> Eq. (6.10)<br />

It is readily shown that the matrix appearing on the left-h<strong>and</strong> side <strong>of</strong> (6.10) has the<br />

LU-decomposition given by<br />

¨<br />

A p1q<br />

NN 0 ¨ ¨ ¨ 0 0 0<br />

˛<br />

0 A p2q<br />

NN ¨ ¨ ¨ 0 0 0<br />

. .<br />

. . .<br />

L “<br />

pP ´1q<br />

0 0 ¨ ¨ ¨ ANN 0 0<br />

˚<br />

˝ 0 0 ¨ ¨ ¨ 0 A pP q ‹<br />

‚<br />

A p1q<br />

GN<br />

A p2q<br />

GN<br />

¨ ¨ ¨ A<br />

pP ´1q<br />

GN<br />

NN<br />

0<br />

A pP q<br />

GN<br />

I<br />

(6.15)<br />

60


¨<br />

U “<br />

˚<br />

˝<br />

I 0 ¨ ¨ ¨ 0 0 A p1q<br />

0 I ¨ ¨ ¨ 0 0 A p2q<br />

NN<br />

.<br />

.<br />

0 0 ¨ ¨ ¨ I 0 A<br />

0 0 ¨ ¨ ¨ 0 I A pP q<br />

NN<br />

0 0 ¨ ¨ ¨ 0 0 S<br />

.<br />

.<br />

.<br />

´1 p1q<br />

NN A<br />

NG<br />

´1 p2q A<br />

NG<br />

pP ´1q´1 pP ´1q<br />

NN A<br />

NG<br />

´1 pP q A<br />

NG<br />

˛<br />

. (6.16)<br />

‹<br />

‚<br />

Also straightforward to show is that<br />

Pÿ<br />

p“1<br />

A ppq ´1<br />

GN Appq ppq<br />

NN A<br />

NG ` S “ A GG “<br />

Since the latter equation can be written as<br />

S “<br />

Pÿ<br />

p“1<br />

Pÿ<br />

p“1<br />

´<br />

¯<br />

A ppq<br />

GG ´ ´1 Appq GN Appq ppq<br />

NN A<br />

NG<br />

A ppq<br />

GG . (6.17)<br />

, (6.18)<br />

it is plausible to define<br />

with<br />

Pÿ<br />

S :“ S ppq (6.19)<br />

p“1<br />

in the second equation <strong>of</strong> (6.20),<br />

S ppq :“ A ppq<br />

GG ´ Appq GN pAppq NN q´1 A ppq<br />

NG<br />

“ A ppq<br />

GG ´ Appq GN Cppq NG ; (6.20)<br />

C ppq<br />

NG<br />

:“ pAppq q´1 NN<br />

A ppq<br />

NG . (6.21)<br />

The usual way to solve a linear system <strong>of</strong> the form<br />

LUX “ B (6.22)<br />

is to use two sweeps in succession, viz., to solve first<br />

LY “ B (6.23)<br />

<strong>and</strong> then<br />

UX “ Y . (6.24)<br />

61


Here Y “ ppY p1q q T , ..., pY pP q q T , pY G q T q T ; similar definitions apply to B <strong>and</strong> X – see<br />

(6.12) <strong>and</strong> (6.14), respectively. Due to the special form <strong>of</strong> L, (6.23) can be solved independently<br />

for each sub-vector Y ppq , where p “ 1, ..., P . Specifically, formally the solution<br />

Y ppq can be written as<br />

<strong>and</strong> hence Y pGq as<br />

Y G “ B G ´<br />

Y ppq “<br />

Pÿ<br />

p“1<br />

´<br />

A ppq<br />

NN¯´1<br />

B ppq , (6.25)<br />

A ppq<br />

GN Y ppq “<br />

Pÿ<br />

p“1<br />

´<br />

ppq¯<br />

B ppq<br />

G ´ Appq GN Y<br />

. (6.26)<br />

If (6.25) <strong>and</strong> (6.26) are substituted into (6.27), after minor manipulation the linear system<br />

¨<br />

I 0 ¨ ¨ ¨ 0 0 A p1q<br />

NN<br />

0 I ¨ ¨ ¨ 0 0 A p2q<br />

NN<br />

. . . . .<br />

0 0 ¨ ¨ ¨ I 0 A NN<br />

˚<br />

˝ 0 0 ¨ ¨ ¨ 0 I A pP q<br />

NN<br />

0 0 ¨ ¨ ¨ 0 0 S<br />

results. The last block in this system,<br />

´1 p1q A<br />

NG<br />

´1 p2q A<br />

NG<br />

pP ´1q´1 pP ´1q A<br />

NG<br />

´1 pP q A<br />

NG<br />

˛<br />

¨<br />

˚<br />

˝<br />

‹<br />

‚<br />

X p1q<br />

.<br />

X pP q<br />

X G<br />

˛ ¨<br />

‹<br />

‚ “ ˚<br />

˝<br />

Y p1q<br />

.<br />

Y pP q<br />

Y G<br />

˛<br />

‹<br />

‚<br />

(6.27)<br />

S X G “ Y G , (6.28)<br />

can be solved first for X G . Then the solution X ppq <strong>of</strong> the remaining blocks are obtained<br />

by solving for each p independently <strong>and</strong> simultaneously<br />

X ppq ` pA ppq<br />

NN q´1 A ppq<br />

NG X G “ Y ppq . (6.29)<br />

or, alternatively,<br />

A ppq<br />

NN Xppq “ B ppq ´ A ppq<br />

NG X G . (6.30)<br />

6.2.1.4 Summary <strong>of</strong> the Steps for the Schur-Complement Method<br />

Here is a summary <strong>of</strong> the individual steps <strong>of</strong> the Schur-complement method to obtain<br />

a solution to the Schur-complement linear system (6.10). It is assumed that to each<br />

slave p ě 1 all its interior nodes <strong>and</strong> the totality <strong>of</strong> ghost nodes are known, as well as a<br />

solution from the previous Picard-iteration step at these nodes. Furthermore, we assume<br />

that processor p “ 1 represents the master, <strong>and</strong> that the master additionally acts as an<br />

ordinary slave.<br />

62


‚ Step 0, in a subroutine or function solve 0 parallel: For each slave p ě 1,<br />

(a) allocate the matrices A ppq<br />

NN , Appq GN , Appq NG , Appq GG , Cppq NG , Sppq <strong>and</strong> S;<br />

(b) allocate the vectors B ppq , B ppq<br />

G , Y ppq , X ppq , D ppq ,<br />

(c) only for the master p “ 1 allocate the vector X global .<br />

‚ Step 1, in a subroutine or function solve 1 parallel: For each slave p ě 1,<br />

(a) assemble the matrices A ppq<br />

NN<br />

(b) solve the linear system (6.25) for Y ppq ,<br />

<strong>and</strong> Appq<br />

NG , <strong>and</strong> the vector Bppq ,<br />

(c) compute the matrix<br />

C ppq<br />

NG “ pAppq NN q´1 A ppq<br />

NG<br />

(6.31)<br />

by solving number <strong>of</strong> ghost nodes linear systems,<br />

(d) assemble the matrices A ppq<br />

GG<br />

<strong>and</strong> Appq<br />

GN<br />

, <strong>and</strong> the vector Bppq<br />

G ,<br />

(e) compute the difference vector<br />

D ppq :“ B ppq<br />

G<br />

then communicate it to the master,<br />

´ Appq GN Y ppq , (6.32)<br />

(f) compute the matrix S ppq from (6.20), then communicate it to the master.<br />

‚ Step 2: The master, p “ 1,<br />

(a) waits until S ppq <strong>and</strong> D ppq have arrived from all processors p,<br />

(b) evaluates S “ ř P<br />

p“1 Sppq ,<br />

(c) evaluates Y pGq “ ř P<br />

p“1 Dppq ,<br />

(d) solves (6.28) for X pGq ,<br />

(e) communicates X pGq to all slaves.<br />

‚ Step 3: Each slave p ě 1,<br />

(a) waits until X pGq has arrived from the master, then<br />

(b) computes A ppq<br />

NG XpGq , <strong>and</strong> then<br />

(c) solves (6.30) to obtain X ppq .<br />

(d) sends X ppq to the master,<br />

(e) the master then assembles the overall solution X global .<br />

63


‚ Step 4: Eventually repeat several times the sequence <strong>of</strong> steps 1, 2 <strong>and</strong> 3<br />

‚ Step 5: Deallocate all arrays allocated in step 0, except, eventually, the global<br />

solution vector X global .<br />

6.2.2 Particles Exchange<br />

In present work, Monte carlo particles method has been implemented. Particles are seeded<br />

in the computation domain <strong>and</strong> they travel through the domain. A simple possibility at<br />

the implementation aspect is that all particles are initially distributed into each processors,<br />

<strong>and</strong> they belong to their pre-introduced processor permanently. However, since the<br />

expectation values, e.g., mean velocity, mean mass fraction <strong>of</strong> species <strong>and</strong> temperature,<br />

are calculated from particles which locate in correspond control volume, the particles<br />

distribution implied above occurs critical difficulties.<br />

For instance, let us imagine a particle P which is initially located in the sub-domain<br />

1 h<strong>and</strong>led by processor 1. While this particle travels in the sub-domain 1, processor 1<br />

can use this particle’s data for getting expectation value for corresponded node. If the<br />

location <strong>of</strong> particle P exceed physical neighbour sub-domain 2, processor 1 does not need<br />

anymore the information <strong>of</strong> particle P but processor 2 needs it. Since each processor has<br />

its own memory space, processor 2 can not access the particle P . Therefore, constantly<br />

distribution <strong>of</strong> particles in each processor is simple <strong>and</strong> efficient only at the beginning <strong>of</strong><br />

calculation.<br />

Presented difficulty can be solved by introducing particles transferring. As shown schematically<br />

in Fig. 6.7, when any particle meets a logical boundary, all properties <strong>of</strong> the particle<br />

Figure 6.7: Schematic <strong>of</strong> particle exchange.<br />

64


is sent to neighbour processor by MPI, <strong>and</strong> that particle becomes inactivated in previous<br />

processor. Then each processor owns only the particles involved in correspond<br />

sub-domain.<br />

6.3 Functional Decomposition<br />

A decomposition at the functional level – also termed a functional decomposition – focuses<br />

on target functions which usually dem<strong>and</strong> high computing power.<br />

Implementation <strong>of</strong><br />

functional decomposition into a computer code appears to be simpler for OpenMP than<br />

for MPI. However, to use OpenMP or any other method <strong>of</strong> the shared memory type,<br />

the code needs to be carefully designed. Besides, functional decomposition is especially<br />

suitable for particle methods such as the ones employed in the present work, since each<br />

particle employs an identical but independent procedure to update particle properties<br />

whilst taking time steps.<br />

OpenMP is provided with most Fortran <strong>and</strong> C++ compilers, <strong>and</strong> it is implemented into<br />

a respective computer code by employing compiler directives, library routines <strong>and</strong> environmental<br />

variables. For instance, a loop structure in a serial code can be parallelized<br />

simply by placing suitable OpenMP compiler directives. The OpenMP library routines<br />

are meant to serve as a control <strong>and</strong> query tool for the parallel execution environment,<br />

which the programmer can use from inside its program. Therefore, the OpenMP runtime<br />

library contains a set <strong>of</strong> external procedures with clearly defined interfaces. For the<br />

detailed description <strong>of</strong> OpenMP directives <strong>and</strong> library routines, [57] should be consulted.<br />

In the present work, OpenMP has been implemented into the numerical solution procedure<br />

that evaluates the chemical source terms for each gasphase particle, <strong>and</strong> into the numerical<br />

solution procedure that evaluates the nodal expectation values from both gasphase<br />

<strong>and</strong> liquid-phase particles properties. To test the effects <strong>of</strong> parallelization by OpenMP,<br />

the two-dimensional reactive-spray code has been used in a 32-processors environment<br />

system. The resulting speedup due to parallelization is shown in Figs. 6.8 <strong>and</strong> 6.9, respectively.<br />

Specifically, to measure parallelization effect solely on the target part <strong>of</strong> the<br />

whole program, an indicator for increasing the performance so-called speedup is defined<br />

as<br />

speedup “<br />

execution time <strong>of</strong> the serial target functions<br />

execution time <strong>of</strong> the parallel target functions . (6.33)<br />

Interestingly, defining number <strong>of</strong> processors increasing beyond 32 slows the code down<br />

again. The reason for this phenomenon probably is that parallelization overheads have<br />

increased to a point where no overall gain in speed can be made.<br />

65


15<br />

speedup<br />

10<br />

5<br />

0<br />

1 2 4 8 16 32 64<br />

number <strong>of</strong> processors<br />

Figure 6.8: Partial speedup by OpenMP on calculating chemical source term.<br />

10<br />

speedup<br />

5<br />

0<br />

1 2 4 8 16 32 64<br />

number <strong>of</strong> processors<br />

Figure 6.9: Partial speedup by OpenMP on expectation values.<br />

6.4 MPI-OpenMP Hybrid Approach<br />

In this section, a comparison between MPI <strong>and</strong> OpenMP is done by employing alternatively<br />

each <strong>of</strong> the two parallel schemes in the code.<br />

66


6.4.1 Performance <strong>and</strong> Limitation<br />

The parallelized code was executed at a PC-cluster installed at in the IT Center <strong>of</strong> RWTH<br />

(Rheinisch-Westfälische Technische Hochschule) Aachen [81] in order to investigate the<br />

effect <strong>of</strong> the number <strong>of</strong> processors on speedup. To this end, the maximum number <strong>of</strong><br />

processors was 32.<br />

To compare different parallelization effects on the serial code, the overall speedup has<br />

been calculated by an indicator similar to that presented in Chap. 6.3, however, now<br />

measuring total computing time instead <strong>of</strong> target function computing time. As shown<br />

in Fig 6.10, generally the performance <strong>of</strong> MPI is better than that <strong>of</strong> OpenMP. This is<br />

easily underst<strong>and</strong>able because OpenMP is applied to certain target parts <strong>of</strong> the whole<br />

program, whilst MPI affects the overall code. However, such performance phenomena<br />

may not be general but rather specific to the underlying computer code. Specifically, the<br />

present two-dimensional simulation code woks with particle methods that obtain bulk <strong>of</strong><br />

variables statistically from particles. Therefore, domain <strong>and</strong> particles decomposition by<br />

MPI results in large parallelization effects.<br />

3 OpenMP<br />

MPI<br />

speedup<br />

2<br />

1<br />

1 2 4 8 16 32 64<br />

number <strong>of</strong> processors<br />

Figure 6.10: Overall speedup by OpenMP <strong>and</strong> MPI on sprays flame.<br />

As considered two different parallel schemes, the main dissimilarity is on a way <strong>of</strong> communication<br />

between processors. While distributed memory programming represented by MPI<br />

involves a communication through message sending <strong>and</strong> receiving explicitly, shared mem-<br />

67


ory represented by OpenMP relies on implicit communication which utilizes the shared<br />

memory space as a passage.<br />

Generally, MPI is <strong>of</strong>ten limited in the number <strong>of</strong> processors that can be used, <strong>and</strong> the<br />

application on lager number <strong>of</strong> processors is <strong>of</strong>ten expensive or hard to tune for communication<br />

between processors. On the other h<strong>and</strong>, OpenMP has strong advantages when<br />

it is applied to loop level parallelization. However, applying OpenMP gets reasonable<br />

performance only on the single computing environment which allows data transferring<br />

implicitly in hardware base.<br />

To overcome above limitation <strong>of</strong> single parallelization scheme, a hybrid concept will be<br />

discussed in the following sections.<br />

6.4.2 Hybrid Approach<br />

A hybrid (MPI-OpenMP) scheme is that MPI <strong>and</strong> OpenMP are combined to compensate<br />

the limitations <strong>of</strong> both <strong>of</strong> parallel scheme. Specifically, in present work, the hybrid<br />

scheme is designed by adding looplevel OpenMP parallelization into the existing domain<br />

decomposition MPI parallelized code as shown schematically in Fig. 6.11.<br />

Figure 6.11: Schematic <strong>of</strong> MPI-OpenMP hybrid approach.<br />

For running the hybrid parallelized two-dimensional code, identical computing environment<br />

as introduced in Chap. 6.4.1 has been used. As shown in Fig. 6.12, the performance<br />

with hybrid algorithm is 2 times better than pure OpenMP algorithm used identical CPU<br />

resources.<br />

68


Figure 6.12: Comparing overall speedup between Pure OpenMP <strong>and</strong> Hybrid MPI-<br />

OpenMP (MPI 8 ˆ OpenMP 4).<br />

6.5 Heterogeneous Computing with GPU<br />

6.5.1 Introduction<br />

In the previous section, the parallelization methods are considered under the homogeneous<br />

computing environment such as CPU. Obviously, better clocks <strong>and</strong> more core integrated<br />

system or cluster will give a guarantee against its performance. However, as discussed in<br />

the beginning <strong>of</strong> previous section, raising performance <strong>of</strong> CPU has reached its limitations,<br />

moreover, not every people can access a cluster system due to the financial reason.<br />

One <strong>of</strong> the potential solutions is to distribute computing load by using GPU. A GPU<br />

was designed originally for h<strong>and</strong>ling the image creations for output to display <strong>and</strong> it has<br />

been still rapidly developing with high dem<strong>and</strong>ing <strong>of</strong> complicated graphics. As shown<br />

schematically in Fig. 6.13, especially GPU has highly parallel structure, therefore for the<br />

purpose <strong>of</strong> manipulating graphics GPU is much efficient than CPU which is designed for<br />

general task. With such unique characteristics <strong>of</strong> GPU, the effort for using GPU at more<br />

general purpose has been arisen by many researchers [14, 44, 46] so called GPGPU.<br />

GPGPU st<strong>and</strong>s for General-Purpose computation on Graphics Processing Units. From<br />

an implementation <strong>of</strong> a matrix multiplication whihc was one <strong>of</strong> the first applications<br />

<strong>of</strong> non-graphical computations on a GPU [14, 46], many attempts <strong>of</strong> using GPU for<br />

general computations is proceeding actively. Since basically GPU should be controlled by<br />

CPU, using GPU for general purpose means a computing on CPU-GPU heterogeneous<br />

environments. Controlling between different devices needs specific program language or<br />

69


Figure 6.13: Schematic <strong>of</strong> CPU(left) <strong>and</strong> GPU(right).<br />

API which allow to port a applications to GPU for return values.<br />

CUDA(Compute Unified Device Architecture) is a parallel computing platform <strong>and</strong> programming<br />

model to use a GPU for general purpose processing invented by NVIDIA. A<br />

main drawback is that CUDA can be used only on NVIDIA’s devices. That means GPUs<br />

<strong>of</strong> Intel <strong>and</strong> AMD which share over then 70 percent <strong>of</strong> the GPUs are not available to use<br />

the CUDA platform.<br />

On the other h<strong>and</strong>, OpenCL(Open Computing Language) is a general interface for generating<br />

programs that are able to run across heterogeneous devices maintained by the<br />

non-pr<strong>of</strong>it technology consortium Khronos Group. OpenCL itself provides only st<strong>and</strong>ard<br />

protocols for parallel computing on heterogeneous computing environments. OpenCL<br />

conformant implementations have been provided by multiple vendors, e.g., AMD, Apple,<br />

Intel, NVIDIA. While OpenCL generally has an advantage that it can be adopted many<br />

other platforms, h<strong>and</strong>ling <strong>and</strong> writing a parallel program is more complicate than CUDA<br />

programming due to cover much wide range <strong>of</strong> devices.<br />

In the present work, relatively simple 1D code is modified for applying OpenCL to know<br />

how a heterogeneous computing is able to be adaptable on computational fluid dynamics<br />

fields <strong>and</strong> how efficient it is.<br />

6.5.2 GPU Acceleration with OpenCL<br />

In principle, OpenCL has been proposed as an open st<strong>and</strong>ard, therefore each GPU vendors<br />

has implemented theirs own way. Nevertheless, OpenCL API is general enough to<br />

run on significantly different architectures while being adaptable enough that each hardware<br />

platform can still obtain high performance. Using the core language <strong>and</strong> correctly<br />

70


following the specification, any program designed for one vendor can execute on anothers<br />

hardware. Because <strong>of</strong> the above characteristics, OpenCL is practical for GPUs performance<br />

comparison.<br />

The OpenCL API is a C with a C++ Wrapper API that is defined in terms <strong>of</strong> the C API.<br />

The code that executes on an OpenCL device, which in general is not the same device as<br />

the host CPU, is written in the OpenCL C language. OpenCL C is a restricted version<br />

<strong>of</strong> the C99 language with extensions appropriate for executing data-parallel code on a<br />

variety <strong>of</strong> heterogeneous devices [25].<br />

Figure 6.14: Schematic <strong>of</strong> implementation GPU accelerating with OpenCL.<br />

In the present work, an OpenCL API which is provided by Intel is used for applying GPU<br />

acceleration on existing one-dimensional reacting code (written in Fortran90). Obviously<br />

the implementation OpenCL one-dimensional code for the reacting flows given here is<br />

straightforwardly extended to the general two or three-dimensional case. Since directly<br />

using OpenCL feature with Fortran is limited to certain Fortran compiler 1 , it is necessary<br />

to divide 1D code into several parts <strong>and</strong> generate DLL (Dynamic-link library) for using<br />

in C++ program as schematically shown in Fig. 6.14.<br />

To construct CPU-GPU combined parallel code with OpenCL, there are certain restrictions,<br />

e.g., calling external functions are not allowed, but functions defined in host side<br />

can be used <strong>and</strong> variable length arrays <strong>and</strong> structures are not supported. As discussed in<br />

Chap. 6.3, most computing work concentrated part <strong>of</strong> present whole code is the calculating<br />

chemical source term. However, since external functions or subroutines are not able<br />

1 Only the PGI, CAPS <strong>and</strong> Cray compiler fully support OpenACC for Fortran <strong>and</strong> partial support<br />

from GNU Fortran. And another possibility, CUDA Fortran is the Fortran version <strong>of</strong> CUDA C <strong>and</strong> is<br />

only supported by the PGI compile.<br />

71


to call from outside <strong>of</strong> device, naturally any subroutine corresponded to chemical source<br />

term can not be used from divided .dll directly. Therefore, relatively simple function has<br />

been implemented on device.<br />

In the present work, the reactive flow simulation code has been implemented by Monte-<br />

Carlo particle method which widely used to solve <strong>PDF</strong> transport equations. Since the<br />

accuracy <strong>of</strong> simulation carried by the Monte-Carlo particle method is affected strongly<br />

the number <strong>of</strong> statistical particles. Each particle has own properties, <strong>and</strong> those properties<br />

are updated through computing time steps. Because those particle properties updating is<br />

simple <strong>and</strong> needs to execute many time, this function is reasonable choice <strong>of</strong> parallelization<br />

target function. To validate the parallelized one-dimensional reactive code by OpenCL,<br />

Figure 6.15: Sample calculation for premixed methane flame on 1D code.<br />

simple premixed flame test case has been constructed. Gasphase combustion is described<br />

by a 10-step reduced chemical mechanism for methane/air systems [41] consisting <strong>of</strong> 14<br />

chemical species <strong>and</strong> 10 chemical reactions. Simulated results is reported in Fig. 6.15,<br />

<strong>and</strong> the results between original Fortran code, ported C++ code <strong>and</strong> C++ with OpenCL<br />

code are identical. Shown in Fig. 6.16, three different GPUs were compared. It should be<br />

72


noted that the speedup shown in figure is measured by computing time <strong>of</strong> only targeting<br />

functions as same strategy in Chap. 6.3. It is interesting to note that the number <strong>of</strong> core<br />

does not guarantee correspond speedup.<br />

Figure 6.16: Comparing speedup <strong>of</strong> parallelized target functions between GPUs; 48 cores<br />

(left), 384 cores(center) <strong>and</strong> 24 cores(right).<br />

In this chapter, CPU-GPU heterogeneous computing with OpenCL has been introduced<br />

<strong>and</strong> considered as an alternative parallelization strategy <strong>of</strong> CFD codes. The general<br />

purpose using on GPUs so-called GPGPU is now developing rapidly on the strength <strong>of</strong><br />

increasing GPU performance. Because <strong>of</strong> that reason, although applying CFD codes<br />

to GPUs has some critical limitations, the potential <strong>of</strong> using GPUs for executing CFD<br />

codes can not be judged at the moment. When the programm technics for computing<br />

on the heterogeneous device environments becomes mature enough, the CFD code can<br />

be parallelized efficiently. For example, host device – multi core CPU – has a roll for<br />

decomposed computational domain <strong>and</strong> particles by MPI, <strong>and</strong> devices – GPU or GPUs –<br />

separate the massive loop functions to small loop by OpenCL as schematically shown in<br />

Fig. 6.17.<br />

73


Figure 6.17: Schematic <strong>of</strong> MPI-OpenCL hybrid concept.<br />

74


Chapter 7<br />

<strong>Turbulent</strong> Jet Diffusion Flames<br />

(DLR H3 Flame)<br />

7.1 Experimental <strong>and</strong> Computational Setup<br />

To validate the computer code developed in the present work for pure gasphase combustion,<br />

the code has been applied to a turbulent hydrogen-air flame investigated at DLR<br />

Stuttgart in 1996 [56, 65]. This flame is non-premixed <strong>and</strong> it was operated without a stabilizing<br />

pilot burner. Experiments were carried out with the burner shown schematically<br />

at the left <strong>of</strong> Fig. 7.1. The burner consists <strong>of</strong> a vertical tube with a nozzle diameter <strong>of</strong><br />

8 mm. The co-flowing air had an average flow velocity <strong>of</strong> 0.2 m/s with an exit diameter<br />

<strong>of</strong> 140 mm. The fuel jet consisted <strong>of</strong> 50 vol.% nitrogen <strong>and</strong> 50 vol.% hydrogen; it was<br />

issued at an average flow velocity <strong>of</strong> 34.8 m/s through a nozzle into the ambient air. The<br />

Reynolds number <strong>of</strong> the fuel jet at the burner exit was 10000, the temperature there was<br />

300 K. The pressure was 1 bar. Details on the experimental arrangement are given in<br />

[56, 65] <strong>and</strong> experimental data for temperature <strong>and</strong> species concentrations are reported<br />

in [1, 21, 65]. On the right <strong>of</strong> Fig. 7.1, the computational domain is shown. Since for<br />

this computation only OpenMP was tested, a domain decomposition was not employed.<br />

Roman numbers I to V indicate the different parts <strong>of</strong> the domain boundary. Specifically,<br />

I indicates the jet inlet, II the co-flow inlet, III the open boundary where entrainment <strong>of</strong><br />

air occurs, IV the outlet boundary, <strong>and</strong> V the symmetry boundary. 1<br />

<strong>Simulation</strong>s are performed on the unstructured mesh shown in Fig. 7.1. The computational<br />

domain extends to 10 <strong>and</strong> 90 diameters in the radial- <strong>and</strong> axial direction, respectively.<br />

At the inlet plane the velocity, composition, <strong>and</strong> turbulent properties <strong>of</strong> the gas<br />

flow are fixed to the respective boundary values. At the symmetry axis, V, the velocity<br />

1<br />

Note that the problem is axisymmetric.<br />

75


Figure 7.1: Schematic <strong>of</strong> the DLR burner [56] <strong>and</strong> <strong>of</strong> the computational domain.<br />

components normal to the boundary are set to zero. At the open boundary, III, the flow<br />

is allowed to adjust its direction into the compute domain or out <strong>of</strong> it.<br />

In the computations 195441 particles were used, <strong>and</strong> the mesh consisted <strong>of</strong> 3409 nodes<br />

<strong>and</strong> 6484 triangles.<br />

Gasphase combustion was described by a detailed chemical mechanism <strong>of</strong> hydrogen [61]<br />

consisting <strong>of</strong> 10 chemical species <strong>and</strong> 21 chemical reactions.<br />

7.2 Results <strong>and</strong> Discussion<br />

Shown in Figs. 7.2 to 7.6 are pr<strong>of</strong>iles in physical space <strong>of</strong> various quantities as obtained<br />

0.06<br />

experiment<br />

simulation<br />

H 2<br />

mass fraction<br />

0.04<br />

0.02<br />

0.00<br />

0 20 40 60 80 100<br />

x/D<br />

Figure 7.2: Axial pr<strong>of</strong>ile <strong>of</strong> the mass fraction <strong>of</strong> H 2 along the centerline. Solid line:<br />

computational result; symbols: experimental data from [1, 21, 65].<br />

76


0.2<br />

experiment<br />

simulation<br />

O 2<br />

mass fraction<br />

0.1<br />

0.0<br />

0 20 40 60 80 100<br />

x/D<br />

Figure 7.3: Axial pr<strong>of</strong>ile <strong>of</strong> the mass fraction <strong>of</strong> O 2 along the centerline. Solid line:<br />

computational result; symbols: experimental data from [1, 21, 65].<br />

0.15 experiment<br />

simulation<br />

H 2<br />

O mass fraction<br />

0.10<br />

0.05<br />

0.00<br />

0 20 40 60 80 100<br />

x/D<br />

Figure 7.4: Axial pr<strong>of</strong>ile <strong>of</strong> the mass fraction <strong>of</strong> H 2 O along the centerline. Solid line:<br />

computational result; symbols: experimental data from [1, 21, 65].<br />

in experiments performed by Neuber et al. [1, 21, 65] <strong>and</strong> from the numerical simulations<br />

carried out in the present work. Specifically, shown in Fig. 7.2 <strong>and</strong> 7.3, respectively, is<br />

the mass fraction <strong>of</strong> hydrogen <strong>and</strong> oxygen on the centerline <strong>of</strong> the turbulent jet.<br />

It can be seen that generally the agreement between the experimental <strong>and</strong> numerical data<br />

is good. The same is the case for the mass fractions <strong>of</strong> water vapor, whose centerline<br />

pr<strong>of</strong>ile is shown in Fig. 7.4. Results showing the same good agreement are shown for the<br />

mass fraction <strong>of</strong> the hydroxyl radical in Fig, 7.5, <strong>and</strong> for the temperature in Fig. 7.6.<br />

77


0.002<br />

experiment<br />

simulation<br />

OH mass fraction<br />

0.001<br />

0.000<br />

0 20 40 60 80 100<br />

x/D<br />

Figure 7.5: Axial pr<strong>of</strong>ile <strong>of</strong> the mass fraction <strong>of</strong> OH along the centerline.<br />

computational result; symbols: experimental data from [1, 21, 65].<br />

Solid line:<br />

Figure 7.6: Axial pr<strong>of</strong>ile <strong>of</strong> temperature along the centerline. Solid line: computational<br />

result; symbols: experimental data from [1, 21, 65].<br />

Plotted in Figs. 7.7 <strong>and</strong> 7.8 is the temperature as a function <strong>of</strong> the mixture faction at two<br />

axial positions, x{D “ 5 <strong>and</strong> x{D “ 20, respectively. In both, the experiments [1, 21, 65]<br />

<strong>and</strong> the computations performed herein, the mixture fraction Z was calculated using the<br />

Bilger’s definition [5], viz.,<br />

Z “<br />

2Y C {W C ` 1<br />

2 Y H{W H ` `Y O,air ´ Y O˘{WO<br />

2Y C,fuel {W C ` 1<br />

2 Y H,fuel{W H ` `Y O,air ´ Y O,fuel˘{WO<br />

, (7.1)<br />

78


where Y α <strong>and</strong> W α denote the mass fractions <strong>and</strong> molecular weights, respectively, for<br />

carbon, C, atomic hydrogen, H, <strong>and</strong> atomic oxygen, O.<br />

Figure 7.7: Temperature as a function <strong>of</strong> mixture fraction at x{D “ 5. Red points: computational<br />

result; blue points: experimental data from [1, 21, 65]; line: chemical equilibrium<br />

from [1, 21, 65].<br />

Figure 7.8: Temperature as a function <strong>of</strong> mixture fraction at x{D “ 20. Red points:<br />

computational result; blue points: experimental data from [1, 21, 65]; line: chemical equilibrium<br />

from [1, 21, 65].<br />

79


It can be seen from the figures that the agreement between experimental data, equilibrium<br />

data <strong>and</strong> computational results is good, although in both the rich <strong>and</strong> the lean part <strong>of</strong><br />

mixture-fraction space the computational results are somewhat closer to the equilibrium<br />

results than to the experimental data.<br />

Shown in Fig. 7.9 are surface plots <strong>of</strong> the expectations <strong>of</strong> the mass fractions <strong>of</strong> H 2 O <strong>and</strong><br />

OH, <strong>and</strong> <strong>of</strong> the temperature as obtained by the computations. It can be seen that the<br />

field <strong>of</strong> all three quantities look reasonable, both in values <strong>and</strong> shape, as can be expected<br />

due to the reasonable agreement between experimental data <strong>and</strong> computational results<br />

shown for the centerline pr<strong>of</strong>iles in Figs. 7.2 to 7.6. The partial lack <strong>of</strong> smoothness in<br />

the fields is due to the statistical errors that are – despite forming moving averages, see<br />

Chap. 3.2.2, – inherently present in Monte Carlo simulations due to the finite number <strong>of</strong><br />

particles employed.<br />

Figure 7.9: Surface plots – from top to bottom – <strong>of</strong> the mass fraction <strong>of</strong> H 2 O, the mass<br />

fraction <strong>of</strong> OH, <strong>and</strong> the temperature, respectively.<br />

80


Shown in Fig. 7.10 is the computed marginal density function G c – see Eq. (2.41) – <strong>of</strong> the<br />

hydroxyl mass fraction as a function <strong>of</strong> the radial coordinate r at the non-dimensional<br />

axial position x{D “ 5. It can be seen that from the centerline up to approximately<br />

r “ 0.015 m practically no gasphase particle carries OH. The mass fraction <strong>of</strong> OH then<br />

increases in the radius range 0.015 m ď 0.03 m up to a value <strong>of</strong>, approximately, 0.003,<br />

however, for not too many gasphase particles. From r “ 0.03 m on, the mass fraction <strong>of</strong><br />

OH then decreases again, taking for many gasphase particles frequently very small values<br />

at the edge <strong>of</strong> the jet, i.e., upon approaching r “ 0.07 m.<br />

Figure 7.10: Marginal density function <strong>of</strong> the mass fraction <strong>of</strong> OH at position x{D “ 5.<br />

7.3 Speedup by OpenMP<br />

The results presented <strong>and</strong> discussed in Chap. 7.2 were obtained by carrying out two<br />

computations on an 8-processor computer. The first computation was carried with a serial<br />

version <strong>of</strong> the code, the second was carried with a version <strong>of</strong> the code parallelized by use <strong>of</strong><br />

OpenMP. For none <strong>of</strong> the two computations domain decomposition was employed. Both<br />

computations produced identical results. However, with the parallelized version <strong>of</strong> the<br />

code a speedup <strong>of</strong> approximately 1.8 could be achieved.<br />

81


Chapter 8<br />

<strong>Turbulent</strong> Counterflow Flames with<br />

Water Droplets<br />

8.1 Experimental <strong>and</strong> Computational Setup<br />

To validate the computer code developed in the present work for turbulent sprays, the<br />

code has been applied to a non-premixed, turbulent methane-air counterflow flame with<br />

water droplets, which was investigated at Cambridge University in 1997 [107]. Specifically,<br />

in the Cambridge experiments the counterflow burner shown schematically in Fig. 8.1 was<br />

used. This burner consisted <strong>of</strong> two opposed nozzles each with an inner exit diameter <strong>of</strong><br />

25 mm. Both nozzles had a honeycomb-flow straightener to ensure a uniform gas flow at<br />

the respective exit plane. Also for both nozzles, the average exit flow velocity was 0.543<br />

m/s. For both nozzles the inlet composition is specified in Table 8.1.<br />

upper nozzle lower nozzle<br />

gaseous phase X CH4 = 0.23 X O2 = 0.21<br />

X N2 = 0.77 X N2 = 0.79<br />

liquid phase Y d =0.906%<br />

Table 8.1: Inlet composition for both the experiment [107] <strong>and</strong> the present computations.<br />

In the table, X α denotes the mole fraction for gasphase species α, α = CH 4 , O 2 , <strong>and</strong> N 2 ;<br />

Y d denotes the mass concentration <strong>of</strong> water droplets,<br />

Y d :“<br />

9m d<br />

9m d ` 9m air<br />

, (8.1)<br />

where 9m d <strong>and</strong> 9m air denote the mass flow rate <strong>of</strong> water droplets <strong>and</strong> the mass flow rate<br />

<strong>of</strong> the carrier gas, respectively. From the definition (8.1) it can be seen that in the<br />

experiments, <strong>and</strong> hence also in the present computations, the water drops were added to<br />

82


Figure 8.1: Schematic <strong>of</strong> the counterflow burner used in the experiments [107].<br />

the oxidizer, not to the fuel.<br />

In the experiments <strong>and</strong> hence the computations, each nozzle was surrounded by an external<br />

nozzle through which nitrogen was fed to form a curtain that reduced the influence <strong>of</strong><br />

the ambient air currents on the flame. The separation distance between the nozzle exits<br />

was 14 mm [106]. The temperature <strong>and</strong> the pressure there at the nozzle exits were 300 K<br />

<strong>and</strong> 1 bar, respectively. Details on the experimental arrangements <strong>and</strong> experimental data<br />

are reported in [107].<br />

83


Figure 8.2: Schematic <strong>of</strong> computational domain <strong>and</strong> boundaries.<br />

On the left side <strong>of</strong> Fig. 8.2 a schematic <strong>of</strong> the Cambridge burner is shown which helps<br />

to underst<strong>and</strong> the particular choice for the computational domain, which – because <strong>of</strong><br />

axisymmetric symmetry – has been taken as a half plane as shown on the right side <strong>of</strong> the<br />

figure. Since for this computation MPI was tested, the domain has been decomposed into<br />

8 sub-domains that in the pdf-version – not the in the printed version – <strong>of</strong> the present<br />

thesis each carry a different background color. Roman numbers I to V denote the different<br />

parts <strong>of</strong> the domain boundary. Specifically, I indicates the upper inlet, II the lower inlet,<br />

III the wall boundary, IV the open boundary where entrainment <strong>of</strong> air occurs, <strong>and</strong> V the<br />

symmetry boundary.<br />

<strong>Simulation</strong>s are performed on the unstructured mesh shown in Fig. 8.2. At the inlets<br />

to the computational domain, labelled I <strong>and</strong> II, respectively, velocity, composition, <strong>and</strong><br />

turbulent properties <strong>of</strong> the gas flow are fixed to the respective boundary values. At the<br />

symmetry axis, V, the velocity components normal to the boundary are set to zero. At<br />

the open boundary, IV, the flow is allowed to adjust its velocity value <strong>and</strong> direction into<br />

or out <strong>of</strong> the computational domain.<br />

In the computations 22400 particles were used, <strong>and</strong> the mesh consisted <strong>of</strong> 4666 nodes <strong>and</strong><br />

9088 triangles.<br />

Gasphase combustion was described by a 2-step reduced chemical mechanism for methane/air<br />

systems [24] consisting <strong>of</strong> 6 chemical species <strong>and</strong> 2 chemical reactions.<br />

84


8.2 Results <strong>and</strong> Discussion<br />

Shown in Figs. 8.3 <strong>and</strong> 8.4 are pr<strong>of</strong>iles in physical space <strong>of</strong> various quantities as obtained<br />

in experiments performed by Zheng [107] <strong>and</strong> from the numerical simulations carried out<br />

Figure 8.3: Pr<strong>of</strong>ile <strong>of</strong> the mean axial velocity component <strong>of</strong> water droplets along the<br />

symmetry line. Solid line: computational result; symbols: experimental data from [107].<br />

Figure 8.4: Radial pr<strong>of</strong>ile <strong>of</strong> the mean axial velocity component <strong>of</strong> water droplets at the<br />

axial location z “ 1.4 mm. Solid line: computational result; symbols: experimental data<br />

from [107].<br />

85


in the present work. Specifically, shown in Fig. 8.3 is the mean axial velocity component<br />

<strong>of</strong> water droplets along the symmetry line <strong>of</strong> the counterflow geometry.<br />

It can be seen that generally the agreement between the experimental <strong>and</strong> numerical data<br />

is good. The same is the case for the mean axial velocity component <strong>of</strong> water droplets as<br />

a function <strong>of</strong> radius taken at axial station z=1.4 mm, which is shown in Fig. 8.4.<br />

Shown in Fig. 8.5 is a phase plane spanned by the methane mole fraction in the fuel jet<br />

lower Jet: O 2<br />

mole fraction<br />

0.22<br />

0.21<br />

0.20<br />

extinction<br />

stable burning<br />

sim. Y d = 0.4%<br />

exp. Y d = 0.0%<br />

exp. Y d = 0.4%<br />

exp. Y d = 0.9%<br />

increasing the mass<br />

concentration <strong>of</strong><br />

water droplets<br />

0.19<br />

0.20 0.23 0.26 0.29 0.32 0.35<br />

upper Jet: CH 4mole fraction<br />

Figure 8.5: Phase-plane identifying regimes <strong>of</strong> extinction <strong>and</strong> stable burning. Stars:<br />

computational results, other symbols: experimental data from [107].<br />

<strong>and</strong> the oxygen mole fraction in the oxidizer jet. In this phase plane, regimes <strong>of</strong> extinction<br />

<strong>and</strong> stable burning are identified for three different values <strong>of</strong> the water mass concentration<br />

Y d . Stars denote computational results obtained herein, the other symbols denote<br />

experimental results [107]. It can be seen that the agreement between the experimental<br />

<strong>and</strong> the computational results is very good. The computed extinction points shown in<br />

this figure have been obtained from plots such as the one shown in Fig. 8.6. Plotted in<br />

the latter figure is the maximum mean flame temperature as a function <strong>of</strong> the water mass<br />

concentration Y d as obtained by computations, with inlet mole fractions as specified in<br />

Table 8.1. It can be seen from the graph that extinction has occurred at Y d « 0.4%.<br />

Shown in Fig. 8.7 are two surface plots on top <strong>of</strong> each other <strong>of</strong> the mean gasphase temperature<br />

as obtained by computations. In each <strong>of</strong> the two surface plots to the left <strong>and</strong> the<br />

86


2000<br />

maximum temperature (K)<br />

1500<br />

1000<br />

500<br />

0<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

mass concentration <strong>of</strong> water droplets Y d(%)<br />

Figure 8.6: Maximum mean flame temperature as a function <strong>of</strong> the water mass concentration<br />

Y d for inlet mole fractions X O2 “ 0.21 <strong>and</strong> X CH4 “ 0.23.<br />

Figure 8.7: Surface plots <strong>of</strong> mean temperature with different Y d . A,C: without seeding <strong>of</strong><br />

water drops; B: stable flame with Y d “ 0.1%; D: extinction with Y d “ 0.9%.<br />

right <strong>of</strong> the symmetry line different conditions – these are labelled A, B, C <strong>and</strong> D – apply.<br />

Specifically, in both the top <strong>and</strong> the bottom surface plot, to the left <strong>of</strong> the symmetry line,<br />

i.e., under conditions A <strong>and</strong> C, respectively, temperature surfaces are displayed that have<br />

been obtained for a purely gaseous counterflow flame, i.e., a flame which was not seeded<br />

with water drops. Naturally, since conditions A <strong>and</strong> C are identical, so are the corresponding<br />

temperature surface plots. The two surface plots to the right <strong>of</strong> the symmetry<br />

87


line, labelled B <strong>and</strong> D, respectively, have been obtained as follow. The plot labelled B<br />

is for a stable burning flame that was seeded with water drops with the under-critical<br />

water mass concentration Y d “ 0.1%. Therefore, this flame pertains to the stable-burning<br />

regime shown in Fig. 8.5. The plot labelled D is for an extinguished flame that was seeded<br />

with water drops with the over-critical water mass concentration Y d “ 0.9%. Therefore,<br />

the latter flame pertains to the extinction regime shown in Fig. 8.5.<br />

Shown in Fig. 8.8 is a surface plot <strong>of</strong> the mean gasphase temperature <strong>and</strong> <strong>of</strong> streamlines<br />

as obtained by a computation for an under-critical case with Y d “ 0.1%. Since the flame<br />

is axisymmetric, so are the temperature surfaces <strong>and</strong> the streamlines shown in the figure.<br />

The temperature surfaces are as expected <strong>and</strong> already discussed in the context <strong>of</strong> Fig.<br />

8.7. In the vicinity <strong>of</strong> the symmetry axis, between the nozzles, the streamlines show a<br />

stagnation point <strong>and</strong> an approximate stagnation plane. Just outside the lower nozzle<br />

through which gaseous oxidizer <strong>and</strong> droplets are supplied, a relatively strong ring-shaped<br />

vortex due to air outflow, caused by expansion, at the bottom <strong>of</strong> the computational<br />

domain can be identified. Outflow occurs everywhere at the computational boundaries<br />

except at the solid outer walls <strong>of</strong> the nozzles where no-slip velocity boundary conditions<br />

apply. Also, several smaller vortices can be identified in the computational domain shown<br />

in the figure.<br />

Figure 8.8: Surface plot <strong>of</strong> mean temperature with streamlines for a case with Y d “ 0.1%.<br />

88


8.3 Speedup by MPI<br />

The results presented <strong>and</strong> discussed in Chap. 8.2 were obtained by carrying out two<br />

computations on an 8-processor computer. The first computation was carried with a serial<br />

version <strong>of</strong> the code, the second was carried with a version <strong>of</strong> the code parallelized by use<br />

<strong>of</strong> MPI. Both computations produced identical results. However, with the parallelized<br />

version <strong>of</strong> the code a speedup <strong>of</strong> approximately 2.3 could be achieved.<br />

89


Chapter 9<br />

<strong>Turbulent</strong> Jet Diffusion <strong>Spray</strong><br />

Flames (Sydney flame)<br />

9.1 Experimental <strong>and</strong> Computational Setup<br />

To validate the computer code developed in the present work for turbulent spray combustion,<br />

the code has been applied to a turbulent ethanol-air spray flame investigated at the<br />

University <strong>of</strong> Sydney in 2009 [39, 55]. Specifically, in the Sydney experiments, the spray<br />

burner was assisted by a stabilizing pilot burner, as shown schematically in Fig. 9.1.<br />

The central jet nozzle diameter was 10.5 mm. The outer diameter <strong>of</strong> the annulus was 25.0<br />

mm. The pilot flame holder was fixed 11 mm from the centre. A co-flow <strong>of</strong> diameter 104<br />

mm surrounded the burner <strong>and</strong> had an average flow velocity <strong>of</strong> 4.5 m/s with temperature<br />

298 K. The pilot stream carried fully burnt combustion products <strong>of</strong> a stoichiometric<br />

mixture <strong>of</strong> ethylene/hydrogen <strong>and</strong> air, with an adiabatic flame temperature <strong>of</strong> 2493 K.<br />

The bulk velocities <strong>of</strong> the pilot stream <strong>and</strong> jet were 11.6 m/s <strong>and</strong> 24 m/s, respectively.<br />

Details on the experimental arrangements <strong>and</strong> experimental data are reported in [39, 55].<br />

The liquid phase properties used in the experiments <strong>and</strong> the present simulations are listed<br />

in Table 9.1.<br />

property <strong>and</strong> unit value<br />

density ρ l (g/cm 3 ) 0.789<br />

latent heat L (kJ/kg) 846<br />

boiling temperature (K) 351.1<br />

Table 9.1: Properties <strong>of</strong> liquid ethanol.<br />

The computational domain is shown in the Fig. 9.2. Since for this computation MPI-<br />

OpenMP hybrid parallelization was tested, the domain has been decomposed into 8 sub-<br />

90


Figure 9.1: Schematic <strong>of</strong> the spray burner used in the Sydney experiments [39, 55].<br />

domains that in the pdf-version – not the in the printed version – <strong>of</strong> the present thesis<br />

each carry a different background color. Roman numbers I to VI indicate the different<br />

parts <strong>of</strong> the domain boundary. Specifically, I indicates the jet inlet, II the pilot flame<br />

inlet, III the co-flow inlet, IV the open boundary where entrainment <strong>of</strong> air occurs, V the<br />

outlet boundary, <strong>and</strong> VI the symmetry boundary. 1<br />

<strong>Simulation</strong>s are performed on the unstructured mesh shown in Fig. 9.2. The computational<br />

domain extends to 8 <strong>and</strong> 40 diameters in the radial <strong>and</strong> axial direction, respectively.<br />

At the inlets to the computational domain, labelled I, II <strong>and</strong> III, respectively, velocity,<br />

composition, <strong>and</strong> turbulent properties <strong>of</strong> the gas flow <strong>and</strong> liquid flow are fixed to the<br />

1<br />

Note that the problem is axisymmetric.<br />

91


Figure 9.2: Schematic <strong>of</strong> computational domain <strong>and</strong> boundaries.<br />

respective boundary values. At the symmetry axis, VI, the velocity components normal<br />

to the boundary are set to zero. At the open boundary, IV, the flow is allowed to adjust<br />

its direction into or out <strong>of</strong> the computational domain.<br />

In the computations 505072 particles were used, <strong>and</strong> the mesh consisted <strong>of</strong> 5156 nodes<br />

<strong>and</strong> 9924 triangles.<br />

Gasphase combustion was described by a global single-step chemical mechanism for ethanol/air<br />

systems [15] comprising <strong>of</strong> 5 chemical species.<br />

9.2 Results <strong>and</strong> Discussion<br />

Shown in Figs. 9.3 <strong>and</strong> 9.4 are pr<strong>of</strong>iles in physical space<br />

<strong>of</strong> the mean axial velocity<br />

Figure 9.3: Radial pr<strong>of</strong>ile <strong>of</strong> the mean axial velocity component <strong>of</strong> fuel droplets along at<br />

axial station z/D=10. Solid line: computational result; symbols: experimental data from<br />

[39, 55].<br />

92


Figure 9.4: Radial pr<strong>of</strong>ile <strong>of</strong> the mean axial velocity component <strong>of</strong> fuel droplets along at<br />

axial station z/D=20. Solid line: computational result; symbols: experimental data from<br />

[39, 55].<br />

component w as obtained in the experiments performed by Gounder et al. [39, 55], <strong>and</strong><br />

from the numerical simulations carried out in the present work. Specifically, Fig. 9.3<br />

refers to the axial station z{D “ 10. It can be seen that generally the agreement between<br />

the experimental <strong>and</strong> numerical data is good. The same is the case for the mean axial<br />

velocity component <strong>of</strong> fuel droplets as a function <strong>of</strong> radius taken at non-dimensional axial<br />

station z{D “ 20, which is shown in Fig. 9.4. Shown in Fig. 9.5 is the radial pr<strong>of</strong>ile<br />

<strong>of</strong> mean temperature at z{D “ 10. It can be seen that for small radii the agreement<br />

Figure 9.5: Radial pr<strong>of</strong>ile <strong>of</strong> temperature at axial location z/D=10. Solid line: computational<br />

result; symbols: experimental data from [39, 55].<br />

93


is good, but that for radii greater than approximately unity temperatures are greatly<br />

over-predicted by the simulation. A similar over-prediction by the simulation can be<br />

seen in Fig. 9.6, which shows the same radial pr<strong>of</strong>ile <strong>of</strong> temperature but at z{D “ 20,<br />

i.e., further downstream. Since on the Sydney flame further experimental data than the<br />

one shown here are not available, the cause <strong>of</strong> the disagreement between experimental<br />

<strong>and</strong> computed radial temperature pr<strong>of</strong>iles can only be speculated about. For instance,<br />

the observed disagreement may, perhaps, be due to an over-prediction <strong>of</strong> the transversal<br />

velocity component or to an over-prediction <strong>of</strong> the rate <strong>of</strong> turbulent mixing. Since in the<br />

turbulent mixing-model employed the ratio ɛ{k appears, either an over-prediction <strong>of</strong> the<br />

turbulent dissipation rate ɛ or an under-prediction <strong>of</strong> the turbulent kinetic energy k could<br />

lead to a mixing rate that is too large.<br />

Figure 9.6: Radial pr<strong>of</strong>ile <strong>of</strong> temperature at axial location z/D=20. Solid line: computational<br />

result; symbols: experimental data from [39, 55].<br />

Shown in Fig. 9.7 are surface plots <strong>of</strong> the expectations <strong>of</strong> the mean axial velocity component,<br />

the mass fraction <strong>of</strong> H 2 O, <strong>and</strong> <strong>of</strong> the temperature as obtained by the computations.<br />

The plots show surfaces <strong>of</strong> the three quantities as it is to be expected on physical grounds.<br />

Specifically, in the plot <strong>of</strong> mean axial velocity, w, the local acceleration <strong>of</strong> the flow in the<br />

vicinity <strong>of</strong> the symmetry axis due to heat release <strong>and</strong> hence expansion is obvious as is, as<br />

a consequence <strong>of</strong> overall mass conservation, the observed slow-down <strong>and</strong> homogenization<br />

sufficiently far downstream. Close to the nozzle exit, the surface plots <strong>of</strong> the mean H 2 O<br />

mass fraction <strong>and</strong> <strong>of</strong> the temperature both clearly show the effect <strong>of</strong> the pilot flame.<br />

Shown in Fig. 9.8 are surface plots <strong>of</strong> the expectations <strong>of</strong> the mean axial velocity component<br />

<strong>of</strong> droplets, w l , <strong>and</strong> at the mean radial velocity component <strong>of</strong> droplets, v l , as<br />

94


Figure 9.7: Surface plots – from top to bottom – <strong>of</strong> the axial velocity component, the<br />

mass fraction <strong>of</strong> H 2 O, <strong>and</strong> the temperature, respectively.<br />

Figure 9.8: Surface plots – from top to bottom – <strong>of</strong> the axial velocity component, <strong>and</strong> the<br />

radial velocity component, respectively.<br />

95


obtained by the computations. It can be seen that the fields <strong>of</strong> both velocity components<br />

look reasonable, both in values <strong>and</strong> shape, as can be expected due to the reasonable agreement<br />

between experimental data <strong>and</strong> computational results shown for the radial pr<strong>of</strong>iles<br />

in Figs. 9.3 <strong>and</strong> 9.4. From axial station z{D « 8, droplets are strongly accelerated into<br />

the radial direction by expansion due to heat release.<br />

Shown in Fig. 9.9 is the computed marginal density function f W pR l q <strong>of</strong> droplet radius<br />

– see Eq. (2.50) – along the jet center line as a function <strong>of</strong> the non-dimensional axial<br />

coordinate z{D. It can be seen that from the nozzle up to approximately z{D “ 5<br />

practically most droplets keep formation as distributed at inlet. The droplet radius R l<br />

then decreases along the axial line to a value <strong>of</strong>, approximately, 7 µm, however, droplets<br />

has been distributed wider than between z{D “ 0 <strong>and</strong> 5. This phenomena shown in this<br />

figure can be confirmed by plotting each single droplet as shown in Fig. 9.10.<br />

Figure 9.9: Pr<strong>of</strong>ile along the jet center line <strong>of</strong> the marginal density function f W pR l q as a<br />

function <strong>of</strong> droplets radius.<br />

Shown in Fig. 9.10 is a scatter plot <strong>of</strong> droplet radius R l along the jet symmetry axis<br />

versus axial distance z from the nozzle exit plane obtained by plotting a point pR l , z{Dq<br />

for each liquid Monte-Carlo particle that – at statistical steady state <strong>of</strong> the reactive flow<br />

– instantaneously is located somewhere on the symmetry axis. It can be seen that the<br />

scattering is relatively small close to the nozzle exit plane <strong>and</strong> that it is increasing with<br />

increasing distance from the nozzle. This is to be expected on physical grounds because<br />

96


Figure 9.10: Droplets radius represented by the liquid Monte-Carlo particles along the<br />

symmetry line <strong>of</strong> the jet.<br />

the initial droplet size distribution, which in the simulations was assumed to be Gaussian,<br />

widens downstream continuously due to differential effects <strong>of</strong> droplet evaporization.<br />

9.3 Speedup by the MPI-OpenMP Hybrid Approach<br />

The results presented <strong>and</strong> discussed in Chap. 9.2 were obtained by carrying out two<br />

computations on a 32-processor cluster. The first computation was carried with a serial<br />

version <strong>of</strong> the code, the second with a version <strong>of</strong> the code parallelized by use <strong>of</strong> the MPI-<br />

OpenMP hybrid scheme. Both computations produced identical results. However, with<br />

the parallelized version <strong>of</strong> the code a speedup <strong>of</strong> approximately 4.2 could be achieved. It<br />

is interesting to note that computations for this case that were carried out applying only<br />

OpenMP – i.e., without MPI – resulted in a speedup <strong>of</strong> approximately only 2.<br />

97


Chapter 10<br />

Conclusions <strong>and</strong> Perspectives<br />

The present thesis has essentially aimed to extend the early research done by Rumberg<br />

[79]. Rumberg’s fully stochastic approach to the formulation <strong>and</strong> solution <strong>of</strong> a turbulent<br />

spray flame in one-dimensional geometry with simplified variables representing evaporation,<br />

mixing, <strong>and</strong> chemical reaction, is now extended to a fully two-dimensional reactive<br />

spray formulation <strong>and</strong> code.<br />

For the special case <strong>of</strong> two-phase spray flows, for both the gaseous phase <strong>and</strong> the liquid<br />

phase <strong>PDF</strong>-transport equations are given in which the terms concerning the local <strong>and</strong><br />

instantaneous interaction between the two phases, <strong>and</strong> certain turbulence terms require<br />

modelling.<br />

To solve the two-phase spray <strong>PDF</strong> transport equation, for both the gaseous <strong>and</strong> the liquid<br />

phase, a particle method has been employed. Specifically, gasphase particles <strong>and</strong> liquidphase<br />

particles, have simultaneously been used to represent the two coexisting phases.<br />

For closure <strong>of</strong> various terms, a velocity model, mixing models <strong>and</strong> evaporation models are<br />

employed. To obtain the turbulent kinetic energy k <strong>and</strong> the turbulent dissipation rate ɛ,<br />

the st<strong>and</strong>ard k-ɛ turbulence model has been used. The latter has been implemented <strong>and</strong><br />

solved for in the code in Eulerian form, as was the Poisson equation for pressure.<br />

To validate both the computer code developed <strong>and</strong> the various physical models for turbulent<br />

spray combustion, the code has been applied to three different experiments, i.e., a<br />

turbulent hydrogen-air flame, a non-premixed, turbulent methane-air counterflow flame<br />

with water droplets, <strong>and</strong> a turbulent ethanol-air spray flame.<br />

In the present thesis, the two-dimensional reactive spray code has been parallelized by<br />

three different parallelization schemes, i.e., OpenMP, MPI, <strong>and</strong> the MPI-OpenMP hybrid<br />

scheme. The parallelized code was executed on a 32-processor cluster. Since OpenMP<br />

<strong>and</strong> MPI have not only strengths but also limitations, the MPI-OpenMP hybrid scheme<br />

98


has been introduced <strong>and</strong> implemented to combine the advantages <strong>and</strong>, simultaneously,<br />

mitigate the disadvantages. The computational results in Chaps. 7 to 9 were obtained<br />

with the three parallelization schemes. Specifically, in Chap. 7, the code parallelized by<br />

OpenMP achieved a speedup <strong>of</strong> 1.8. In Chap. 8, the code parallelized by MPI achieved<br />

a speedup <strong>of</strong> 2.3. In Chap. 9, the code parallelized by MPI-OpenMP hybrid scheme<br />

achieved a speedup <strong>of</strong> 4.2.<br />

The present work has shown that the two-phase <strong>PDF</strong> method is feasible for the simulation<br />

<strong>of</strong> turbulent reactive sprays. There are, naturally, still points that require improvement<br />

to get more accurate results. For instance, using a generalized Langevin model instead <strong>of</strong><br />

the simplified version, would help to predict better the effects <strong>of</strong> turbulence on velocity.<br />

In addition, a more powerful turbulence model could be implemented. Furthermore, the<br />

present work could be extended to a spray simulation with multi-component liquid fuels.<br />

99


Appendix A<br />

Derivation <strong>of</strong> the Eulerian Equations<br />

A.1 Pressure Equation<br />

In the present work, the pressure is calculated by SIMPLE-like algorithm which is modified<br />

from SIMPLE [6, 20, 63]. The main point <strong>of</strong> SIMPLE or SIMPLE liked algorithms is<br />

obtaining the pressure correction field p 1 through combination with the mass continuity<br />

equation <strong>and</strong> the momentum conservation equation.<br />

Start from mass continuity equation which is written as<br />

And applying separated velocity fields leads<br />

Bρ<br />

Bt ` BρV i<br />

Bx i<br />

“ 0 . (A.1)<br />

Bρ<br />

Bt ` BρV i<br />

˚<br />

“ ´BρV i<br />

1<br />

, (A.2)<br />

Bx i Bx i<br />

where V ˚ denotes a velocity prediction for next time step <strong>and</strong> V 1 denotes a velocity<br />

correction related as<br />

V n`1<br />

i<br />

Similar assumption applies to the pressure as<br />

“ V ˚<br />

i ` V 1<br />

i .<br />

(A.3)<br />

p n`1 “ p˚ ` p 1 .<br />

(A.4)<br />

The pressure correction p 1 is related to the velocity correction by approximate momentum<br />

equation [6, 20].<br />

ρ BV i<br />

Bt “ ´ Bp<br />

Bx i<br />

(A.5)<br />

100


ρ`V n`1<br />

i<br />

ρ`V ˚<br />

i<br />

˘<br />

´ Vi<br />

n Bp n`1<br />

“ ´∆t<br />

Bx i<br />

˘<br />

´ Vi<br />

n Bp˚<br />

“ ´∆t<br />

Bx i<br />

ρV 1<br />

i “ ´∆t Bp1<br />

Bx i<br />

.<br />

Substituting (A.8) into (A.2), we obtain the pressure-correction Poisson equation.<br />

∆t B ˆ ˙ Bp<br />

1<br />

“ Bρ<br />

Bx i Bx i Bt ` BρV i<br />

˚<br />

Bx i<br />

(A.6)<br />

(A.7)<br />

(A.8)<br />

(A.9)<br />

Integrating over the domain Ω, (A.9) yields<br />

With applying Green’s theorem,<br />

ż<br />

Ω<br />

∆t B2 p 1 ż<br />

Bρ<br />

dΩ “<br />

Bx 2 i<br />

Ω Bt<br />

żΩ<br />

dΩ `<br />

BρV ˚<br />

i<br />

Bx i<br />

dΩ (A.10)<br />

ż<br />

Γ<br />

ż<br />

∆t Bp1<br />

Bρ<br />

¨ n i dΓ “<br />

Bx i Ω Bt<br />

żΓ<br />

dΩ ` ρVi ˚ ¨ n i dΓ , (A.11)<br />

where n denotes normal vector which is perpendicular to interface Γ between two control<br />

volumes.<br />

As shown Fig. A.1, each fragments <strong>of</strong> triangle has an integral point for surface(ips) <strong>and</strong><br />

for volume(ipv).<br />

The first r.h.s term <strong>of</strong> (A.11) is<br />

where<br />

ż<br />

Ω<br />

ρ ipv “<br />

Bρ<br />

Bt dΩ “<br />

2ÿ<br />

ipv“1<br />

ˆρipv ´ ρ˝ipv<br />

∆t<br />

˙<br />

ipv<br />

Ω ipv ,<br />

3ÿ<br />

pα k ρ k ` β k ρ k x ipv ` γ k ρ k y ipv q .<br />

k“1<br />

(A.12)<br />

(A.13)<br />

The second r.h.s term <strong>of</strong> (A.11) is<br />

101


Figure A.1: Integrating points<br />

ż<br />

Γ<br />

ρV ˚<br />

i ¨ n i dΓ “<br />

“<br />

2ÿ<br />

ips“1<br />

2ÿ<br />

ips“1<br />

ρ ips V ˚<br />

i ips ¨ n iipsL ips<br />

ρ ips<br />

`u˚ips b ips ` v˚ipsa ips˘<br />

,<br />

(A.14)<br />

where<br />

pn x ` n y q L “ b ` a ,<br />

(A.15)<br />

where L denotes the length <strong>of</strong> Γ. And<br />

3ÿ<br />

V i,ips “ pα k V i,k ` β k V i,k x ips ` γ k V i,k y ips q .<br />

k“1<br />

(A.16)<br />

with the shape functions α, β <strong>and</strong> γ introduced in Chaps. 5.2.<br />

The l.h.s term <strong>of</strong> (A.11) is<br />

ż<br />

Γ<br />

∆t Bp1<br />

Bx i<br />

¨ n i dΓ “<br />

2ÿ<br />

ips“1<br />

1<br />

˙<br />

ˆBp<br />

∆t<br />

Bx b ips ` Bp1<br />

By a ips , (A.17)<br />

102


with<br />

We obtain<br />

ż<br />

Γ<br />

∆t Bp1<br />

Bx i<br />

¨ n i dΓ “<br />

Bp 1<br />

Bx “ βp1 “<br />

Bp 1<br />

By “ γp1 “<br />

2ÿ<br />

ips“1<br />

“ ∆t<br />

3ÿ<br />

β k p 1 k ,<br />

k“1<br />

3ÿ<br />

γ k p 1 k .<br />

k“1<br />

˜ 3ÿ<br />

∆t β k p 1 kb ips `<br />

k“1<br />

3ÿ<br />

2ÿ<br />

k“1 ips“1<br />

3ÿ<br />

γ k p 1 ka ips¸<br />

k“1<br />

pb ips β k ` a ips γ k q p 1 k<br />

(A.18)<br />

(A.19)<br />

(A.20)<br />

Finally, the result <strong>of</strong> the derivation <strong>of</strong> the discretized Poisson equation for p is derived as<br />

˘ ÿ ÿ`An p 1 n “ Bn ,<br />

(A.21)<br />

In (A.21), the summation on both the l.h.s. <strong>and</strong> the r.h.s. extends over all neighboring<br />

triangles <strong>of</strong> the control-volume defining node n. Specifically,<br />

2ÿ<br />

A n “ ∆t pb ips β n ` a ips γ n q<br />

ips“1<br />

(A.22)<br />

<strong>and</strong><br />

B n “<br />

`<br />

2ÿ<br />

ipv“1<br />

2ÿ<br />

ips“1<br />

ˆρipv ´ ρ˝ipv<br />

∆t<br />

˙<br />

Ω ipv<br />

ρ ips<br />

`u˚ips b ips ` v˚ipsa ips˘<br />

,<br />

(A.23)<br />

A.2 <strong>Turbulent</strong> Kinetic Energy <strong>and</strong> Dissipation Rate<br />

<strong>Turbulent</strong> kinetic energy k <strong>and</strong> its dissipation rate ɛ are calculated by st<strong>and</strong>ard k-ɛ turbulence<br />

model [40, 47, 48] <strong>and</strong> the transport equations <strong>of</strong> k <strong>and</strong> ɛ are<br />

B<br />

Bt pρkq ` B pρkV i q “ B „ˆ ˙ j<br />

µ ` µt Bk<br />

Bx i Bx j σ k Bx j<br />

` P k ´ ρɛ (A.24)<br />

103


<strong>and</strong><br />

B<br />

Bt pρ Gɛq ` B pρ G ɛV G,i q “ B „ˆ ˙ j<br />

µ ` µt Bɛ<br />

Bx i Bx j σ ɛ Bx j<br />

`<br />

C 1ɛ<br />

ɛ<br />

k P k ´ C 2ɛ ρ G<br />

ɛ 2<br />

k ,<br />

(A.25)<br />

with neglecting the effects <strong>of</strong> buoyancy <strong>and</strong> compressibility on turbulence.<br />

Firstly discretized transport equation <strong>of</strong> k is derived from the equation A.24. Similarly<br />

to discretization <strong>of</strong> pressure equation, integrating over the domain Ω, Eq. (A.24) yields<br />

ż<br />

Ω<br />

B<br />

Bt<br />

żΩ<br />

pρkq dΩ ` B<br />

pρkV i q dΩ “<br />

Bx i<br />

`<br />

ż<br />

ż<br />

Ω<br />

Ω<br />

„ˆ ˙ j<br />

B<br />

µ ` µt Bk<br />

dΩ<br />

Bx j σ k Bx j<br />

pP k ´ ρɛq dΩ<br />

(A.26)<br />

With applying Green’s theorem on the second l.h.s term <strong>of</strong> (A.26) <strong>and</strong> the first r.h.s term<br />

<strong>of</strong> (A.26), these terms can be expressed as<br />

ż<br />

Ω<br />

ż<br />

B<br />

pρkV i q dΩ “ ρkV i ¨ n i dΓ ,<br />

Bx i Γ<br />

(A.27)<br />

ż<br />

Ω<br />

„ˆ ˙ j ż ˆ ˙<br />

B<br />

µ ` µt Bk<br />

dΩ “ µ ` µt Bk<br />

¨ n i dΓ ,<br />

Bx j σ k Bx j Γ σ k Bx j<br />

(A.28)<br />

where n i , Ω <strong>and</strong> Γ are introduced previous discretization in A.1.<br />

With further discreization executed at integration point ipv <strong>and</strong> ips – see Fig. A.1, Eq<br />

(A.26) can be expressed as<br />

2ÿ<br />

ipv“1<br />

ˆρkipv ´ ρk˝ipv<br />

∆t<br />

˙<br />

Ω ipv<br />

`<br />

“<br />

`<br />

2ÿ<br />

ips“1<br />

2ÿ<br />

ips“1<br />

2ÿ<br />

ipv“1<br />

pρkq ips<br />

`u˚ips b ips ` v˚ipsa ips˘<br />

ˆ<br />

µ ` µt<br />

σ k<br />

˙ips<br />

pP k ´ ρɛq ipv<br />

Ω ipv ,<br />

˙ „ˆBk<br />

b ips `<br />

Bx<br />

˙ j ˆBk<br />

a ips<br />

By<br />

(A.29)<br />

104


where<br />

Bk<br />

3ÿ<br />

Bx “<br />

Bk<br />

By<br />

“<br />

k“1<br />

β k k ,<br />

3ÿ<br />

γ k k .<br />

k“1<br />

(A.30)<br />

(A.31)<br />

Production <strong>of</strong> k is<br />

P k “ µ t S 2<br />

“ ρC µ<br />

k 2<br />

ɛ .<br />

(A.32)<br />

Recall linear interpolation variable {phi at point ips or ipv as<br />

φpx, tq “ α φ ptq ` β φ ptqx ` γ φ ptqy .<br />

(A.33)<br />

The discretized equation for k can be written as<br />

<strong>and</strong> discretized equation for ɛ is also derived similarly<br />

ÿ`Apkq<br />

ÿ<br />

n k n˘<br />

“ B<br />

pkq<br />

n , (A.34)<br />

ÿ`Apɛq n ɛ n˘<br />

“<br />

ÿ<br />

B<br />

pɛq<br />

n . (A.35)<br />

In (A.34)<strong>and</strong> (5.20), the summation on both the l.h.s. <strong>and</strong> the r.h.s. extends over all<br />

neighboring triangles <strong>of</strong> the control-volume defining node n.<br />

Specifically,<br />

A pkq<br />

n “<br />

`<br />

´<br />

2ÿ<br />

ipv“1<br />

2ÿ<br />

ips“1<br />

2ÿ<br />

ips“1<br />

pα n ` β n x ipv ` γ n y ipv q ρ nΩ ipv<br />

∆t<br />

pα n ` β n x ips ` γ n y ips q ¨ pu n b ips ` v n a ips q ρ n<br />

pα n ` β n x ips ` γ n y ips q ¨ pβ n b ips ` γ n a ips q ¨<br />

ˆ ˙<br />

µ n ` µt,n<br />

σ k,n<br />

(A.36)<br />

105


<strong>and</strong><br />

B pkq<br />

n “<br />

`<br />

2ÿ<br />

ipv“1<br />

2ÿ<br />

ipv“1<br />

pα n ` β n x ipv ` γ n y ipv q ρk˝nΩ ipv<br />

∆t<br />

pα n ` β n x ipv ` γ n y ipv q ¨ pP k ´ ρɛq n<br />

Ω ipv<br />

(A.37)<br />

are for k, <strong>and</strong><br />

<strong>and</strong><br />

are for ɛ.<br />

A pɛq<br />

n “<br />

`<br />

´<br />

B pɛq<br />

n “<br />

`<br />

2ÿ<br />

ipv“1<br />

2ÿ<br />

ips“1<br />

2ÿ<br />

ips“1<br />

2ÿ<br />

ipv“1<br />

2ÿ<br />

ipv“1<br />

pα n ` β n x ipv ` γ n y ipv q ρ nΩ ipv<br />

∆t<br />

pα n ` β n x ips ` γ n y ips q ¨ pu n b ips ` v n a ips q ρ n<br />

pα n ` β n x ips ` γ n y ips q ¨ pβ n b ips ` γ n a ips q ¨<br />

pα n ` β n x ipv ` γ n y ipv q ρɛ˝nΩ ipv<br />

∆t<br />

ˆ<br />

˙<br />

ɛ<br />

pα n ` β n x ipv ` γ n y ipv q ¨ C 1ɛ<br />

k P ɛ 2<br />

k ´ C 2ɛ ρ G<br />

k<br />

ˆ ˙<br />

µ n ` µt,n<br />

σ ɛ,n<br />

Ω ipv<br />

n<br />

(A.38)<br />

(A.39)<br />

106


Appendix B<br />

Chemical Reaction Mechanism<br />

As discussed in Chap. 4.2, the chemical source term is closed form in gasphase <strong>PDF</strong> transport<br />

equation, therefore, the chemical reaction affects directly in the particle composition<br />

increment [60]. To obtain a composition increment due to a reaction, the reaction rates<br />

should be consulted. In 1889, Svante Arrhenius proposed a formula that can describe the<br />

rate <strong>of</strong> a reaction as a function <strong>of</strong> temperature. This formula is so-called an Arrhenius<br />

equation. Specifically, the Arrhenius equation describes the temperature dependence <strong>of</strong><br />

reaction rate constant k by<br />

ˆ<br />

k “ A exp ´ E ˙<br />

, (B.1)<br />

RT<br />

where A is the pre-exponential factor, E the activation energy, R the universal gas constant,<br />

<strong>and</strong> T the temperature. Equation (B.1) was extended to describe high-temperatures<br />

gasphase kinetic systems. The extended Arrhenius equation describes the temperature<br />

dependence <strong>of</strong> the rate coefficient as<br />

ˆ<br />

k “ A T n exp ´ E ˙<br />

. (B.2)<br />

RT<br />

Chemical reaction <strong>of</strong> fuel <strong>and</strong> oxidizer will lead to several intermediate species <strong>and</strong> require<br />

several or even many reaction steps until final combustion products are obtained.<br />

The production <strong>and</strong> consumption reaction-steps <strong>of</strong> reactants <strong>and</strong> intermediate species are<br />

called elementary reactions, <strong>and</strong> set <strong>of</strong> elementary reactions is called a detailed kinetic<br />

mechanism. Generally, detailed kinetic mechanisms are developed to cover wide ranges<br />

<strong>of</strong> operating conditions in a variety <strong>of</strong> different applications [45, 97, 104]. Thus detailed<br />

kinetic mechanisms contain hundreds to thous<strong>and</strong>s <strong>of</strong> chemical species <strong>and</strong> reactions. In<br />

order to use kinetic mechanisms in CFD, detailed mechanisms should be reduced to a<br />

few numbers <strong>of</strong> species <strong>and</strong> chemical reactions, or even reduced to a single-step global<br />

reaction.<br />

107


In the present thesis, mainly three chemical mechanisms have been used. Pure gaseous<br />

combustion in Chap. 7 has been described by a detailed chemical mechanism <strong>of</strong> hydrogen<br />

comprising <strong>of</strong> 10 chemical species <strong>and</strong> 21 chemical reactions. Detailed reaction information<br />

should be consulted from [61]. For the spray flame with counterflow geometry in Chap.<br />

8, gasphase combustion has been described by a 2-step reduced chemical mechanism for<br />

methane/air systems [24]. The reactions <strong>and</strong> conditions are reported in Table B.1. The<br />

Reaction A E a<br />

1 CH 4 + 1.5O 2 Ñ CO + 2H 2 O 2E15 35000<br />

2 CO + 0.5O 2 Ø CO 2 2E9 12000<br />

Table B.1: Reduced chemical mechanism <strong>of</strong> methane [24]. The activation energies E a are<br />

in cal/mole <strong>and</strong> the pre-exponential constants in cgs units.<br />

reaction order n CH 4<br />

ford <strong>and</strong> nO 2<br />

ford<br />

for first reaction, are 0.9 <strong>and</strong> 1.1, respectively <strong>and</strong> nCO ford , nO 2<br />

ford<br />

<strong>and</strong> n CO 2<br />

rord<br />

for second reaction are 1, 0.5 <strong>and</strong> 1, respectively. Abbreviation ford <strong>and</strong> rord<br />

mean forward <strong>and</strong> reversed order, respectively. For the jet diffusion spray flame in Chap.<br />

9, gasphase combustion has been described by a global single-step chemical mechanism<br />

for ethanol/air systems [15] comprising <strong>of</strong> 5 chemical species. According to [15] The mole<br />

based global single-step reaction rate 9w for ethanol/air, can be written as<br />

˙<br />

ˆ´1.256 ˆ 10<br />

9w “ 1.55 ˆ 10 10 8<br />

exp<br />

rFs 0.15 rOs 1.6<br />

8314.3 ˆ T<br />

with the stoichiometric condition as<br />

kmol<br />

m 3 s ,<br />

(B.3)<br />

C 2 H 2 ` 2H 2 ` 3.5 pO 2 ` 3.76N 2 q ÝÑ 2CO 2 ` 3H 2 O ` 13.16N 2 .<br />

(B.4)<br />

108


Appendix C<br />

Mixing Models<br />

In the present work, effects <strong>of</strong> the molecular diffusion are taken into account through<br />

the mixing models – see Chap. 4.2. In a reactive flow, especially non-premixed flame,<br />

suitable mixing effect is important to get reasonable <strong>and</strong> stable burning, hence the need<br />

to test three mixing models, i.e., IEM mixing model, modified Curl mixing model, <strong>and</strong><br />

EMST mixing model, consulted in the present thesis. These tests have applied on a single<br />

control volume. In order to observe the effect <strong>of</strong> molecular mixing, the chemical reaction<br />

has been eliminated <strong>and</strong> the velocity components in both directions were also neglected so<br />

that the particles remain in the single control volume, therefore, the mixing results would<br />

be not influenced by the velocity. In the tests have used ensemble particles comprising two<br />

composition scalars, i.e., temperature T <strong>and</strong> the mass fraction <strong>of</strong> O 2 . Initially particles<br />

have been distributed with r<strong>and</strong>om properties, for this case, T <strong>and</strong> Y O2 . The value <strong>of</strong><br />

Y O2<br />

has been distributed between zero <strong>and</strong> unity, <strong>and</strong> Temperature T has been set to be<br />

r<strong>and</strong>omly distributed between 300´600 Kelvin. For the St<strong>and</strong>ard test, Ensemble number<br />

<strong>of</strong> particles has been considered to be P “ 1000 <strong>and</strong> the time step is fixed as ∆t “ 10´6 .<br />

Also the turbulent timescale τ (κ{ε) has been set to the value <strong>of</strong> 1.0.<br />

To compare three different mixing model, the so-called convergence parameter σ α for the<br />

component scalar α has been defined as<br />

d řP<br />

p“1 pψp α ´ xψ α y 2 q<br />

σ α “<br />

P<br />

, (C.1)<br />

where p <strong>and</strong> P denote particle index <strong>and</strong> number <strong>of</strong> particles, respectively. This convergence<br />

<strong>of</strong> σ α indicates that value <strong>of</strong> the composition α <strong>of</strong> particles relaxing towards the<br />

mean <strong>and</strong> therefor the decay <strong>of</strong> the variance. Figure C.1 is shown the convergence <strong>of</strong> all<br />

models in comparison to each other. It can be seen that IEM model <strong>and</strong> EMST model<br />

with approximately the same rate, have the fastest convergence among all models. Then,<br />

109


modified Curl model converge, although more slowly. In case <strong>of</strong> modified Curl model, it<br />

can be observed that the convergence have not occurred completely. The reason is in the<br />

nature <strong>of</strong> this model, where a certain number <strong>of</strong> the particles are chosen in each time step<br />

to engage in mixing. Therefore, in some cases, in the final steps one or more particles<br />

would remain out <strong>of</strong> the selection causing a slight derivation from the presumed mean<br />

value <strong>and</strong> also an incomplete convergence. This phenomena could be seen in the last<br />

panels in Fig. C.3. As expected in Chap. 4.2, two mixing models IEM <strong>and</strong> Modified Curl<br />

have straightforward <strong>and</strong> simple implementations <strong>and</strong> approaches. Although they would<br />

not produce results with the highest accuracy in case <strong>of</strong> complicated problems, they have<br />

the advantage <strong>of</strong> the speed <strong>of</strong> computation. This two models would require the least <strong>and</strong><br />

the most effortless calculations, therefor time <strong>and</strong> resources in order to converge. This<br />

test have been performed over the single control volume in several different conditions for<br />

all mixing models mentioned above. It was observed that the change in the number <strong>of</strong><br />

particles P in the control volume would not alter the results significantly. Although the<br />

st<strong>and</strong>ard condition for this test includes 1000 particles, obviously this number would not<br />

be a logical choice for further more complicated tests as discussed in Chap. 5.5.<br />

Figure C.1: The convergence comparing between three mixing models.<br />

110


Figure C.2: The convergence shapes <strong>of</strong> IEM mixing model.<br />

111


Figure C.3: The convergence shapes <strong>of</strong> modified Curl mixing model.<br />

112


Figure C.4: The convergence shapes <strong>of</strong> EMST mixing model.<br />

113


Bibliography<br />

[1] A. Neuber, G. Krieger, M. Tacke, E. Hassel, <strong>and</strong> J. Janicka. 11.tecflam-seminar:<br />

Verbrennungsdiagnostik und prozeßkontrolle, arbeitsgemeinschaft tecflam. pages<br />

83–99, 1995.<br />

[2] B. Abramzon <strong>and</strong> W. A. Sirignano. Droplet vaporization model for spray combustion<br />

calculations. International Journal <strong>of</strong> Heat <strong>and</strong> Mass Transfer, 32(9):1605–1618,<br />

1989.<br />

[3] M. S. An<strong>and</strong> <strong>and</strong> S. B. Pope. Diffusion behind a line source in grid turbulence. In<br />

L. J. S. Bradbury, F. Durst, B. E. Launder, F. W. Schmidt, <strong>and</strong> J. H. Whitelaw,<br />

editors, <strong>Turbulent</strong> Shear Flows 4, pages 46–61. Springer Berlin Heidelberg, Berlin,<br />

Heidelberg, 1985.<br />

[4] M. S. An<strong>and</strong>, S. B. Pope, <strong>and</strong> H. C. Mongia. A pdf method for turbulent recirculating<br />

flows. In C. A. Brebbia, S. A. Orszag, R. Borghi, <strong>and</strong> S. N. B. Murthy,<br />

editors, <strong>Turbulent</strong> Reactive Flows, volume 40 <strong>of</strong> Lecture Notes in Engineering, pages<br />

672–693. Springer US, New York, NY, 1989.<br />

[5] R. W. Bilger, S. H. Stårner, <strong>and</strong> R. J. Kee. On reduced mechanisms for methane-air<br />

combustion in nonpremixed flames. <strong>Combustion</strong> <strong>and</strong> Flame, 80(2):135–149, 1990.<br />

[6] T. J. Chung. Computational fluid dynamics. Cambridge University Press, Cambridge<br />

<strong>and</strong> New York, 2002.<br />

[7] S. M. Correa <strong>and</strong> S. B. Pope. Comparison <strong>of</strong> a monte carlo pdf/finite-volume mean<br />

flow model with bluff-body raman data. Symposium (International) on <strong>Combustion</strong>,<br />

24(1):279–285, 1992.<br />

[8] R. Courant, K. Friedrichs, <strong>and</strong> H. Lewy. Über die partiellen differenzengleichungen<br />

der mathematischen physik. Mathematische Annalen, 100(1):32–74, 1928.<br />

[9] C. T. Crowe, J. N. Chung, <strong>and</strong> T. R. Troutt. Particle mixing in free shear flows.<br />

Progress in Energy <strong>and</strong> <strong>Combustion</strong> Science, 14(3):171–194, 1988.<br />

114


[10] C. T. Crowe, T. R. Troutt, <strong>and</strong> J. N. Chung. Numerical models for two-phase<br />

turbulent flows. Annual Review <strong>of</strong> Fluid Mechanics, 28(1):11–43, 1996.<br />

[11] R. L. Curl. Dispersed phase mixing: I. theory <strong>and</strong> effects in simple reactors. AIChE<br />

Journal, 9(2):175–181, 1963.<br />

[12] S. De, K. N. Lakshmisha, <strong>and</strong> R. W. Bilger. <strong>Modeling</strong> <strong>of</strong> nonreacting <strong>and</strong> reacting<br />

turbulent spray jets using a fully stochastic separated flow approach. <strong>Combustion</strong><br />

<strong>and</strong> Flame, 158(10):1992–2008, 2011.<br />

[13] C. Dopazo <strong>and</strong> E. E. O’Brien. An approach to the autoignition <strong>of</strong> a turbulent<br />

mixture. Acta Astronautica, 1(9-10):1239–1266, 1974.<br />

[14] P. Du, R. Weber, P. Luszczek, S. Tomov, G. Peterson, <strong>and</strong> J. Dongarra. From<br />

cuda to opencl: Towards a performance-portable solution for multi-platform gpu<br />

programming. Parallel Computing, 38(8):391–407, 2012.<br />

[15] R. Dubey, K. Bhadraiah, <strong>and</strong> V. Raghavan. On the estimation <strong>and</strong> validation <strong>of</strong><br />

global single-step kinetics parameters <strong>of</strong> ethanol-air oxidation using diffusion flame<br />

extinction data. <strong>Combustion</strong> Science <strong>and</strong> Technology, 183(1):43–50, 2010.<br />

[16] J. K. Dukowicz. A particle-fluid numerical model for liquid sprays. Journal <strong>of</strong><br />

Computational Physics, 35(2):229–253, 1980.<br />

[17] G. FAETH. Mixing, transport <strong>and</strong> combustion in sprays. Progress in Energy <strong>and</strong><br />

<strong>Combustion</strong> Science, 13(4):293–345, 1987.<br />

[18] G. M. Faeth. Current status <strong>of</strong> droplet <strong>and</strong> liquid combustion. Progress in Energy<br />

<strong>and</strong> <strong>Combustion</strong> Science, 3(4):191–224, 1977.<br />

[19] G. M. Faeth. Evaporation <strong>and</strong> combustion <strong>of</strong> sprays. Progress in Energy <strong>and</strong><br />

<strong>Combustion</strong> Science, 9(1-2):1–76, 1983.<br />

[20] J. H. Ferziger <strong>and</strong> M. Perić. Computational methods for fluid dynamics. Springer,<br />

Berlin <strong>and</strong> New York, 3rd, rev. ed. edition, 2002.<br />

[21] A. Fiolitakis, P. R. Ess, P. Gerlinger, <strong>and</strong> M. Aigner. <strong>Modeling</strong> <strong>of</strong> heat transfer<br />

<strong>and</strong> differential diffusion in transported pdf methods. <strong>Combustion</strong> <strong>and</strong> Flame,<br />

161(8):2107–2119, 2014.<br />

[22] M. J. Flynn. Very high-speed computing systems. Proceedings <strong>of</strong> the IEEE,<br />

54(12):1901–1909, 1966.<br />

115


[23] R. O. Fox. Computational Models for <strong>Turbulent</strong> Reacting Flows. Cambridge University<br />

Press, Cambridge, 2003.<br />

[24] G. Boudier. Methane/air flame with 2-step chemistry: 2s-ch4-cm2. CERFACS<br />

technical report, 2007.<br />

[25] B. Gaster. Heterogeneous computing with OpenCL. Morgan Kaufmann, Waltham,<br />

MA, 2012.<br />

[26] Ge H. <strong>and</strong> Gutheil E. <strong>Modeling</strong> <strong>and</strong> simulation <strong>of</strong> turbulent spray flows using<br />

rans <strong>and</strong> pdf transport equations. interdisziplinäres Zentrum für Wissenschaftliches<br />

Rechnen,Universität Heidelberg, 2005.<br />

[27] G. Godsave. Studies <strong>of</strong> the combustion <strong>of</strong> drops in a fuel spray—the burning <strong>of</strong> single<br />

drops <strong>of</strong> fuel. Symposium (International) on <strong>Combustion</strong>, 4(1):818–830, 1953.<br />

[28] M. Goldsmith. On the burning <strong>of</strong> single drops <strong>of</strong> fuel in an oxidizing atmosphere.<br />

Journal <strong>of</strong> Jet Propulsion, 24(4):245–251, 1954.<br />

[29] A. D. Gosman <strong>and</strong> E. loannides. Aspects <strong>of</strong> computer simulation <strong>of</strong> liquid-fueled<br />

combustors. Journal <strong>of</strong> Energy, 7(6):482–490, 1983.<br />

[30] D. C. Haworth. Progress in probability density function methods for turbulent<br />

reacting flows. Progress in Energy <strong>and</strong> <strong>Combustion</strong> Science, 36(2):168–259, 2010.<br />

[31] D. C. HAWORTH <strong>and</strong> S. H. EL TAHRY. Probability density function approach for<br />

multidimensional turbulentflow calculations with application to in-cylinder flows in<br />

reciprocating engines. AIAA Journal, 29(2):208–218, 1991.<br />

[32] D. C. Haworth <strong>and</strong> S. B. Pope. A generalized langevin model for turbulent flows.<br />

Physics <strong>of</strong> Fluids, 29(2):387, 1986.<br />

[33] E. A. Hennick <strong>and</strong> M. F. Lightstone. A comparison <strong>of</strong> stochastic separated flow<br />

models for particle dispersion in turbulent flows. Energy & Fuels, 14(1):95–103,<br />

2000.<br />

[34] J. C. Hill. Zero diffusivity invariance for turbulent chemical reaction. Physics <strong>of</strong><br />

Fluids, 13(5):1394, 1970.<br />

[35] A. T. Hsu, Y.-L. P. Tsai, <strong>and</strong> M. S. Raju. Probability density function approach<br />

for compressible turbulent reacting flows. AIAA Journal, 32(7):1407–1415, 1994.<br />

116


[36] G. L. Hubbard, V. E. Denny, <strong>and</strong> A. F. Mills. Droplet evaporation: Effects <strong>of</strong><br />

transients <strong>and</strong> variable properties. International Journal <strong>of</strong> Heat <strong>and</strong> Mass Transfer,<br />

18(9):1003–1008, 1975.<br />

[37] L. M. Hudy, A. Naguib, <strong>and</strong> W. M. Humphreys. Stochastic estimation <strong>of</strong> a<br />

separated-flow field using wall-pressure-array measurements. Physics <strong>of</strong> Fluids,<br />

19(2):24103, 2007.<br />

[38] J. Janicka, W. Kolbe, <strong>and</strong> W. Kollmann. Closure <strong>of</strong> the transport equation for the<br />

probability density funcfion <strong>of</strong> turbulent scalar fields. Journal <strong>of</strong> Non-Equilibrium<br />

Thermodynamics, 4(1), 1979.<br />

[39] J.D. Gounder. An Experimental Investigation <strong>of</strong> Non-reacting <strong>and</strong> Reacting <strong>Spray</strong><br />

Jets. Ph.d. thesis, Sydney, 2009.<br />

[40] W. Jones <strong>and</strong> B. Launder. The prediction <strong>of</strong> laminarization with a two-equation<br />

model <strong>of</strong> turbulence. International Journal <strong>of</strong> Heat <strong>and</strong> Mass Transfer, 15(2):301–<br />

314, 1972.<br />

[41] J.Y. Chen. Development <strong>of</strong> reduced mechanisms for numerical modeling <strong>of</strong> turbulent<br />

combustion. Workshop on Numerical Aspects <strong>of</strong> Reduction in Chemical Kinetics<br />

CERMICS-ENPC Cite Descartes - Champus sur Marne, France, 1997.<br />

[42] I. Kataoka. Local instant formulation <strong>of</strong> two-phase flow. International Journal <strong>of</strong><br />

Multiphase Flow, 12(5):745–758, 1986.<br />

[43] W. Kollmann. The pdf approach to turbulent flow. Theoretical <strong>and</strong> Computational<br />

Fluid Dynamics, 1(5):249–285, 1990.<br />

[44] J. Krüger <strong>and</strong> R. Westermann. Linear algebra operators for gpu implementation <strong>of</strong><br />

numerical algorithms. In A. P. Rockwood, editor, ACM SIGGRAPH 2003 Papers,<br />

page 908.<br />

[45] K. K. Kuo. Principles <strong>of</strong> combustion. Wiley, New York, 1986.<br />

[46] E. S. Larsen <strong>and</strong> D. McAllister. Fast matrix multiplies using graphics hardware. In<br />

G. Johnson, editor, the 2001 ACM/IEEE conference, page 55.<br />

[47] B. E. Launder <strong>and</strong> D. B. Spalding. Lectures in mathematical models <strong>of</strong> turbulence.<br />

Academic Press, London <strong>and</strong> New York, 1972.<br />

[48] B. E. Launder <strong>and</strong> D. B. Spalding. The numerical computation <strong>of</strong> turbulent flows.<br />

Computer Methods in Applied Mechanics <strong>and</strong> Engineering, 3(2):269–289, 1974.<br />

117


[49] C. K. Law <strong>and</strong> H. K. Law. A d2-law for multicomponent droplet vaporization <strong>and</strong><br />

combustion. AIAA Journal, 20(4):522–527, 1982.<br />

[50] S. K. LIEW, K. N. C. BRAY, <strong>and</strong> J. B. MOSS. A flamelet model <strong>of</strong> turbulent nonpremixed<br />

combustion. <strong>Combustion</strong> Science <strong>and</strong> Technology, 27(1-2):69–73, 2007.<br />

[51] R. P. Lindstedt, S. A. Louloudi, <strong>and</strong> E. M. Váos. Joint scalar probability density<br />

function modeling <strong>of</strong> pollutant formation in piloted turbulent jet diffusion flames<br />

with comprehensive chemistry. Proceedings <strong>of</strong> the <strong>Combustion</strong> Institute, 28(1):149–<br />

156, 2000.<br />

[52] R. Löhner. Robust, vectorized search algorithms for interpolation on unstructured<br />

grids. Journal <strong>of</strong> Computational Physics, 118(2):380–387, 1995.<br />

[53] R. Löhner <strong>and</strong> J. Ambrosiano. A vectorized particle tracer for unstructured grids.<br />

Journal <strong>of</strong> Computational Physics, 91(1):22–31, 1990.<br />

[54] T. S. Lundgren. Model equation for nonhomogeneous turbulence. Physics <strong>of</strong> Fluids,<br />

12(3):485, 1969.<br />

[55] A. R. Masri <strong>and</strong> J. D. Gounder. Details <strong>and</strong> complexities <strong>of</strong> boundary conditions<br />

in turbulent piloted dilute spray jets <strong>and</strong> flames. In B. Merci, D. Roekaerts, <strong>and</strong><br />

A. Sadiki, editors, Experiments <strong>and</strong> Numerical <strong>Simulation</strong>s <strong>of</strong> Diluted <strong>Spray</strong> <strong>Turbulent</strong><br />

<strong>Combustion</strong>, volume 17 <strong>of</strong> ERCOFTAC Series, pages 41–68. Springer Netherl<strong>and</strong>s,<br />

Dordrecht, 2011.<br />

[56] W. MEIER, S. PRUCKER, M.-H. CAO, <strong>and</strong> W. STRICKER. Characterization <strong>of</strong><br />

turbulent hytvair jet diffusion flames by single-pulse spontaneous raman scattering.<br />

<strong>Combustion</strong> Science <strong>and</strong> Technology, 118(4-6):293–312, 1996.<br />

[57] Miguel Hermanns. Parallel programming in fortran 95 using openmp.<br />

[58] J.-P. Minier, S. Chibbaro, <strong>and</strong> S. B. Pope. Guidelines for the formulation <strong>of</strong> lagrangian<br />

stochastic models for particle simulations <strong>of</strong> single-phase <strong>and</strong> dispersed<br />

two-phase turbulent flows. Physics <strong>of</strong> Fluids, 26(11):113303, 2014.<br />

[59] M. Muradoglu, S. B. Pope, <strong>and</strong> D. A. Caughey. The hybrid method for the pdf equations<br />

<strong>of</strong> turbulent reactive flows: Consistency conditions <strong>and</strong> correction algorithms.<br />

Journal <strong>of</strong> Computational Physics, 172(2):841–878, 2001.<br />

[60] B. Naud. <strong>PDF</strong> modeling <strong>of</strong> turbulent sprays <strong>and</strong> flames using a particle stochastic<br />

approach. s.n.], [S.l., op. 2003.<br />

118


[61] M. Ó Conaire, H. J. Curran, J. M. Simmie, W. J. Pitz, <strong>and</strong> C. K. Westbrook.<br />

A comprehensive modeling study <strong>of</strong> hydrogen oxidation. International Journal <strong>of</strong><br />

Chemical Kinetics, 36(11):603–622, 2004.<br />

[62] P. S. Pacheco. An introduction to parallel programming. Morgan Kaufmann, Amsterdam<br />

<strong>and</strong> Boston, 2011.<br />

[63] S. V. Patankar. Numerical heat transfer <strong>and</strong> fluid flow. Series in computational<br />

methods in mechanics <strong>and</strong> thermal sciences. Hemisphere Pub. Corp. <strong>and</strong> McGraw-<br />

Hill, Washington <strong>and</strong> New York, 1980.<br />

[64] N. Peters. <strong>Turbulent</strong> combustion. Cambridge monographs on mechanics. Cambridge<br />

University Press, Cambridge [Engl<strong>and</strong>] <strong>and</strong> New York, 2000.<br />

[65] D. G. Pfuderer, A. A. Neuber, G. Früchtel, E. P. Hassel, <strong>and</strong> J. Janicka. Turbulence<br />

modulation in jet diffusion flames: <strong>Modeling</strong> <strong>and</strong> experiments. <strong>Combustion</strong> <strong>and</strong><br />

Flame, 106(3):301–317, 1996.<br />

[66] O. Pironneau. Numerical simulation <strong>of</strong> unsteady flows <strong>and</strong> transition to turbulence:<br />

Proceedings <strong>of</strong> the ERCOFTAC Workshop held at EPFL, 26-28 March 1990, Lausanne,<br />

Switzerl<strong>and</strong>. Cambridge University Press, Cambridge [Engl<strong>and</strong>] <strong>and</strong> New<br />

York, NY, USA, 1992.<br />

[67] S. B. Pope. The relationship between the probability approach <strong>and</strong> particle models<br />

for reaction in homogeneous turbulence. <strong>Combustion</strong> <strong>and</strong> Flame, 35:41–45, 1979.<br />

[68] S. B. Pope. Calculations <strong>of</strong> velocity-scalar joint pdf’s. In L. J. S. Bradbury, F. Durst,<br />

B. E. Launder, F. W. Schmidt, <strong>and</strong> J. H. Whitelaw, editors, <strong>Turbulent</strong> Shear Flows<br />

3, pages 113–123. Springer Berlin Heidelberg, Berlin, Heidelberg, 1982.<br />

[69] S. B. Pope. Pdf methods for turbulent reactive flows. Progress in Energy <strong>and</strong><br />

<strong>Combustion</strong> Science, 11(2):119–192, 1985.<br />

[70] S. B. Pope. Mapping closures for turbulent mixing <strong>and</strong> reaction. Theoretical <strong>and</strong><br />

Computational Fluid Dynamics, 2(5-6):255–270, 1991.<br />

[71] S. B. Pope. Particle method for turbulent flows: Integration <strong>of</strong> stochastic model<br />

equations. Journal <strong>of</strong> Computational Physics, 117(2):332–349, 1995.<br />

[72] S. B. Pope. Computationally efficient implementation <strong>of</strong> combustion chemistry using<br />

in situ adaptive tabulation. <strong>Combustion</strong> Theory <strong>and</strong> Modelling, 1(1):41–63, 1997.<br />

119


[73] S. B. Pope. <strong>Turbulent</strong> flows. Cambridge University Press, Cambridge <strong>and</strong> New<br />

York, 2000.<br />

[74] P. P. Popov <strong>and</strong> S. B. Pope. Implicit <strong>and</strong> explicit schemes for mass consistency<br />

preservation in hybrid particle/finite-volume algorithms for turbulent reactive flows.<br />

Journal <strong>of</strong> Computational Physics, 257:352–373, 2014.<br />

[75] C. Prakash <strong>and</strong> S. V. Patankar. A control volume-based finite-element method for<br />

solving the navier-stokes equations using equal-order velocity-pressure interpolation.<br />

Numerical Heat Transfer, 8(3):259–280, 1985.<br />

[76] R. B. L<strong>and</strong>is <strong>and</strong> A. F. Mills. Effect <strong>of</strong> internal diffusion resistance on the evaporation<br />

<strong>of</strong> binary droplets. In 5th. Int. Heat Transfer Conf., 1974, 1974.<br />

[77] Rotexo. Cosilab: <strong>Combustion</strong> simulation laboratory.<br />

[78] D. H. Rowinski <strong>and</strong> S. B. Pope. Computational study <strong>of</strong> lean premixed turbulent<br />

flames using ranspdf <strong>and</strong> lespdf methods. <strong>Combustion</strong> Theory <strong>and</strong> Modelling,<br />

17(4):610–656, 2013.<br />

[79] O. Rumberg. Modellierung und <strong>Simulation</strong> turbulenter reaktiver <strong>Spray</strong>s. PhD thesis.<br />

Shaker, Ruhr-Universität Bochum, 2001.<br />

[80] O. Rumberg <strong>and</strong> B. Rogg. Full pdf modelling <strong>of</strong> reactive sprays via an evaporationprogress<br />

variable. <strong>Combustion</strong> Science <strong>and</strong> Technology, 158(1):211–247, 2000.<br />

[81] RWTH Compute Cluster.<br />

[82] J. A. Sethian. Level set methods <strong>and</strong> fast marching methods: Evolving interfaces in<br />

computational geometry, fluid mechanics, computer vision, <strong>and</strong> materials science,<br />

volume 3 <strong>of</strong> Cambridge monographs on applied <strong>and</strong> computational mathematics.<br />

Cambridge University Press, Cambridge, U.K. <strong>and</strong> New York, 2nd ed. edition, 1999.<br />

[83] T.-H. Shih, W. W. Liou, A. Shabbir, Z. Yang, <strong>and</strong> J. Zhu. A new k-$zepsion$ eddy<br />

viscosity model for high reynolds number turbulent flows. Computers & Fluids,<br />

24(3):227–238, 1995.<br />

[84] W. A. Sirignano. Fuel droplet vaporization <strong>and</strong> spray combustion theory. Progress<br />

in Energy <strong>and</strong> <strong>Combustion</strong> Science, 9(4):291–322, 1983.<br />

[85] W. A. Sirignano. Fluid dynamics <strong>of</strong> sprays—1992 freeman scholar lecture. Journal<br />

<strong>of</strong> Fluids Engineering, 115(3):345, 1993.<br />

120


[86] W. A. Sirignano. Fluid dynamics <strong>and</strong> transport <strong>of</strong> droplets <strong>and</strong> sprays. Cambridge<br />

University Press, Cambridge, U.K., 1999.<br />

[87] W. A. Sirignano <strong>and</strong> C. K. Law. Transient heating <strong>and</strong> liquid-phase mass diffusion<br />

in fuel droplet vaporization. In J. T. Zung, editor, Evaporation—<strong>Combustion</strong> <strong>of</strong><br />

Fuels, volume 166 <strong>of</strong> Advances in Chemistry, pages 3–26. AMERICAN CHEMICAL<br />

SOCIETY, WASHINGTON, D. C., 1978.<br />

[88] D. B. Spalding. The combustion <strong>of</strong> liquid fuels. Symposium (International) on<br />

<strong>Combustion</strong>, 4(1):847–864, 1953.<br />

[89] S. Subramaniam. A mixing model for turbulent reactive flows based on euclidean<br />

minimum spanning trees. <strong>Combustion</strong> <strong>and</strong> Flame, 115(4):487–514, 1998.<br />

[90] S. Subramaniam. Statistical modeling <strong>of</strong> sprays using the droplet distribution function.<br />

Physics <strong>of</strong> Fluids, 13(3):624, 2001.<br />

[91] S. Subramaniam <strong>and</strong> D. C. Haworth. A probability density function method for<br />

turbulent mixing <strong>and</strong> combustion on three-dimensional unstructured deforming<br />

meshes. International Journal <strong>of</strong> Engine Research, 1(2):171–190, 2000.<br />

[92] R. R. Tirunagari <strong>and</strong> S. B. Pope. Les/pdf for premixed combustion in the dns limit.<br />

<strong>Combustion</strong> Theory <strong>and</strong> Modelling, 20(5):834–865, 2016.<br />

[93] S. R. Turns. An introduction to combustion: Concepts <strong>and</strong> applications. McGraw-<br />

Hill series in mechanical engineering. McGraw-Hill, New York, 3rd ed. edition, 2012.<br />

[94] J. Villermaux <strong>and</strong> J. C. Devillon. Représentation de la coalescence et de la redispersion<br />

des domaines de ségrégation dans un fluide per modéle d’interaction<br />

phénoménologique. In Proc. Second Int. Symp. On Chemical Reaction Engineering,<br />

New York: Elsevier, 1972.<br />

[95] B. Wang, H. Q. Zhang, <strong>and</strong> X. L. Wang. A time-series stochastic separated flow<br />

(tsssf) model for turbulent two-phase flows. Numerical Heat Transfer, Part B:<br />

Fundamentals, 55(1):73–90, 2009.<br />

[96] Wikipedia. Delaunay triangulation — Wikipedia, the free encyclopedia.<br />

[97] F. A. Williams. <strong>Combustion</strong> theory: The fundamental theory <strong>of</strong> chemically reacting<br />

flow systems. <strong>Combustion</strong> science <strong>and</strong> engineering series. Addison-Wesley, Reading,<br />

Mass. [u.a.], 2. ed., 6. print edition, 1996.<br />

121


[98] J. Xu <strong>and</strong> S. Pope. Assessment <strong>of</strong> numerical accuracy <strong>of</strong> pdf/monte carlo methods<br />

for turbulent reacting flows. Journal <strong>of</strong> Computational Physics, 152(1):192–230,<br />

1999.<br />

[99] V. Yakhot <strong>and</strong> S. A. Orszag. Renormalization group analysis <strong>of</strong> turbulence. i. basic<br />

theory. Journal <strong>of</strong> Scientific Computing, 1(1):3–51, 1986.<br />

[100] N. N. Yanenko <strong>and</strong> M. Holt. The Method <strong>of</strong> Fractional Steps: The Solution <strong>of</strong><br />

Problems <strong>of</strong> Mathematical Physics in Several Variables. Springer Berlin Heidelberg,<br />

Berlin, Heidelberg, 1971.<br />

[101] B. Yang <strong>and</strong> S. B. Pope. An investigation <strong>of</strong> the accuracy <strong>of</strong> manifold methods <strong>and</strong><br />

splitting schemes in the computational implementation <strong>of</strong> combustion chemistry.<br />

<strong>Combustion</strong> <strong>and</strong> Flame, 112(1-2):16–32, 1998.<br />

[102] M. C. YUEN <strong>and</strong> L. W. CHEN. On drag <strong>of</strong> evaporating liquid droplets. <strong>Combustion</strong><br />

Science <strong>and</strong> Technology, 14(4-6):147–154, 1976.<br />

[103] Z. Ren, S. Subramaniam, <strong>and</strong> S.B. Pope. Implementation <strong>of</strong> the emst mixing model:<br />

http://tcg.mae.cornell.edu/emst, 2002.<br />

[104] Zeiwar Hussein Shaheen. Development <strong>of</strong> detailed <strong>and</strong> reduced bio-diesel kinetic<br />

chemical mechanisms. Dissertation, Bochum, 2015.<br />

[105] Y. Z. Zhang <strong>and</strong> D. C. Haworth. A general mass consistency algorithm for hybrid<br />

particle/finite-volume pdf methods. Journal <strong>of</strong> Computational Physics, 194(1):156–<br />

193, 2004.<br />

[106] R. ZHENG. Experimental <strong>and</strong> numerical studies <strong>of</strong> laminar counterflow flames with<br />

water mist. Dissertation, University <strong>of</strong> Cambridge, 1997.<br />

[107] R. ZHENG, K. Bray, <strong>and</strong> B. Rogg. Short communication. <strong>Combustion</strong> Science <strong>and</strong><br />

Technology, 126(1-6):389–401, 1997.<br />

[108] M. Zhu, K. Bray, O. Rumberg, <strong>and</strong> B. Rogg. Pdf transport equations for two-phase<br />

reactive flows <strong>and</strong> sprays. <strong>Combustion</strong> <strong>and</strong> Flame, 122(3):327–338, 2000.<br />

122

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!