ABSTRACT Title of dissertation: DEVELOPMENT OF A LAGRANGIAN -EULERIAN MODELING FRAMEWORK TO DESCRIBE THERMAL DEGRADATION OF POROUS FUEL PARTICLES IN SIMULATIONS OF WILDLAND FIRE BEHAVIOR AT FLAME SCALE Mohamed Mohsen Ahmed Doctor of Philosophy, 2023 Dissertation Directed by: Professor Arnaud Trouvé Department of Mechanical Engineering The dynamics of wildland fires involve multi-physics phenomena occurring at multiple scales ranging from sub-millimeter scale representative of small vegetation particles to several kilometers representative of meteorological scales. The objective of this research is to develop an advanced physics-based computational tool for detailed modeling of the coupling between the solid-phase and the gas-phase processes that control the dynamics of flame spread in wildland fire problems. This work focuses on a modeling approach that resolves processes occurring at flame and vegetation scales, i.e., the formation of flammable vapors from the porous biomass vegetation due to pyrolysis, the subsequent combustion of these fuel vapors with ambient air, the establishment of a turbulent flow because of heat release and buoyant acceleration, and the thermal feedback to the solid biomass through radiative and convective heat transfer. A modeling capability called PBRFoam is developed in this dissertation based on the general-purpose Computational Fluid Dynamics (CFD) library OpenFOAM and an in-house Lagrangian Particle Burning Rate (PBR) model that treats drying, thermal pyrolysis, oxidative pyrolysis, and char oxidation using a one-dimensional porous medium formulation. This modeling capability allows the description of fire spread in vegetation fuel beds comprised of mono- or poly- dispersed porous particles including thermal degradation processes occurring during both flaming and smoldering combustion. The modeling capability is calibrated for cardboard and pine wood using available micro- and bench-scale experimental data obtained. Then it is applied to simulate the fire spread across the idealized fuel beds made of laser-cut cardboard sticks that have been studied experimentally at the Missoula Fire Sciences Laboratory. The simulations are conducted with prescribed particle and environmental properties (i.e., fuel bed height, fuel bed packing, particle size, moisture content, and wind velocity) that match the experimental conditions. The model is first validated against experimental measurements and observations such as the rate of spread of the fire and the flame residence time. The modeling capability is then used to provide insights into local as well as global behavior at the individual particle level and the fuel bed level with variations of the fuel packing. The modeling capability is also applied to simulations of fire spread across idealized vegetation beds corresponding to mixed-size cylindrical-shaped sticks of pine wood under prescribed wind conditions. Depending on the particle size distribution, the simulations feature complete fuel consumption with a successful transition from flaming to smoldering combustion or partial fuel consumption with no or limited smoldering. These simulations show the existence of either a mixed mode of heat transfer through convection and radiation for small particles or a radiation- dominant heat transfer mode for larger particles. The results are interpreted using a novel diagnostic called the Pseudo Incident Heat Flux (PIHF) and 2-D maps that characterize single particle response as a function of the PIHF and the flame residence time. DEVELOPMENT OF A LAGRANGIAN-EULERIAN MODELING FRAMEWORK TO DESCRIBE THERMAL DEGRADATION OF POROUS FUEL PARTICLES IN SIMULATIONS OF WILDLAND FIRE BEHAVIOR AT FLAME SCALE by Mohamed Mohsen Ahmed Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2023 Advisory Committee: Professor Arnaud Trouvé, Chair/Advisor Professor James Baeder, Dean’s Representative Dr. Mark Finney Professor Johan Larsson Professor Stanislav Stoliarov Professor Peter Sunderland © Copyright by Mohamed Mohsen Ahmed 2023 Dedication To my dear parents, my lovely sister, my darling Radwa, and our beautiful Nelli. ii Acknowledgments The completion of this Ph.D. dissertation would not have been possible without the generous support and encouragement of so many people, for which I am deeply grateful. First, I would like to express my deepest gratitude to my advisor, Dr. Arnaud Trouvé. His guidance, mentorship, and expertise have been invaluable throughout my research journey. His insightful comments and feedback on my work were instrumental in refining my ideas and improving the quality of my research. I am also grateful for the opportunities he provided me to present my research at conferences and workshops, which helped me gain exposure and build my professional network. His support in securing funding for my research projects was also critical in enabling me to complete my research. I am also grateful to the faculty members at the University of Maryland who dedicated part of their precious time to serve as an advisory committee for my dissertation. Thanks to Dr. James Baeder, Dr. Johan Larsson, Dr. Stanislav Stoliarov, and Dr. Peter Sunderland for their insightful comments and constructive discussion, which helped me to refine my research. Special thanks to Dr. Mark Finney from the U.S. Forest Service for allocating some of his valuable time to be on my advisory committee as well as for his ultimate support. I am very thankful to him for sharing the data of his experiments and for the invaluable comments and feedback that he provided at different phases during this project. iii I would also like to extend special thanks to Mr. Jason Forthofer from the U.S. Forest Service for the thoughtful and deep discussions we exchanged about model implementation and code development. His guidance was very critical in improving the quality of the computational tools developed in this project. I would like to acknowledge all my former and present colleagues at the ME and FPE departments especially: Salman Verma, Rui Xu, Yiren Qin, Yujeong Kim, Bouaza Lafdal, Zishanul Hussain, Anagh Dave, Kelliann Lee, Jeffrey Bors and Sofia Le Braddock. I am also grateful to the administrative staff of the FPE department: Christine O’Brien and Nicole Hollywood. I am thankful for the financial support from the USDA Forest Service, Missoula Fire Sciences Laboratory, with Dr. Mark Finney as Program Manager. I am also thankful for the partial financial support provided by the U.S. National Science Foundation (NSF, CBET Program, Award #1604907) and by FM Global. I would like also to acknowledge the supercomputing resources made available by UMD (https://www.glue.umd.edu/hpcc) and by NSF (ACCESS Program, Grant #TG- CTS140046, https://access-ci.org). Finally, I would like to express my deep appreciation to my family. Their ultimate love, encouragement, and support have been my source of strength and inspiration throughout my academic career. Their sacrifices and unwavering belief in my potential have made this accomplishment possible. iv https://www.glue.umd.edu/hpcc https://access-ci.org Table of Contents Dedication ii Acknowledgements iii List of Tables x List of Figures xi Nomenclature xxiii 1 Introduction 1 1.1 Wildland fire behavior modeling . . . . . . . . . . . . . . . . . . . . . 2 1.1.1 Scales of wildland fire behavior dynamics . . . . . . . . . . . . 3 1.1.2 Semi-empirical approach to wildfire behavior modeling . . . . 5 1.1.3 Physics-based approach to wildfire behavior modeling . . . . . 6 1.2 Motivation and scope of this dissertation . . . . . . . . . . . . . . . . 9 1.3 Literature review of relevant experimental/modeling work in the vegetation-flame scale . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 1.4 Objectives and contributions . . . . . . . . . . . . . . . . . . . . . . . 13 1.4.1 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 1.4.2 Author’s contributions . . . . . . . . . . . . . . . . . . . . . . 15 1.4.3 Organization . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 2 The Computational Modeling Framework 22 2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 2.2 Particle Burning Rate (PBR) Model . . . . . . . . . . . . . . . . . . 24 2.2.1 Conservation equations for species/total mass inside the solid phase of an individual particle . . . . . . . . . . . . . . . . . . 29 2.2.2 Rate of change of the particle volume (constant/shrinking/swelling) 32 2.2.3 Conservation equations for species/total mass inside the gas phase of an individual particle . . . . . . . . . . . . . . . . . . 33 2.2.4 Conservation equation for energy inside an individual particle 35 2.2.5 Outputs of the PBR model . . . . . . . . . . . . . . . . . . . . 39 v 2.2.6 Particle burning state . . . . . . . . . . . . . . . . . . . . . . . 40 2.3 Fuel bed model and the coupled Lagrangian-Eulerian formulation . . 42 2.3.1 Conservation equations for species/total mass in the gas-phase 43 2.3.2 Conservation equations for momentum in the gas-phase . . . . 46 2.3.3 Conservation equation for energy in the gas-phase . . . . . . . 46 2.3.4 Radiation transfer in the multiphase fuel bed . . . . . . . . . . 47 2.4 Convective transfer model . . . . . . . . . . . . . . . . . . . . . . . . 49 2.5 Drag model and particle motion . . . . . . . . . . . . . . . . . . . . . 52 2.6 Gas-phase homogeneous combustion model . . . . . . . . . . . . . . . 54 2.7 Sub-grid scale turbulence model . . . . . . . . . . . . . . . . . . . . . 57 2.8 Gas-phase radiation absorption and emission model . . . . . . . . . . 59 2.8.1 The Prescribed Global Radiant Fraction (PGRF) approach . . 60 2.8.2 The Weighted-Sum-of-Gray-Gases (WSGG) approach . . . . . 61 2.9 Soot modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 2.10 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 3 Verification and Evaluation of the Computational Framework 67 3.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 3.2 Verification tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68 3.2.1 In-depth heat conduction of solid particles . . . . . . . . . . . 68 3.2.2 Momentum exchange between the solid- and gas-phases . . . . 69 3.2.3 Pressure-velocity coupling inside the fuel bed . . . . . . . . . . 72 3.2.4 Radiation absorption and emission inside the fuel bed . . . . . 74 3.2.5 Smoldering propagation due to radiation penetration . . . . . 78 3.3 Calibration of the thermal degradation model . . . . . . . . . . . . . 80 3.3.1 Micro-scale thermal degradation of pine wood . . . . . . . . . 80 3.3.2 Micro-scale thermal degradation of cardboard . . . . . . . . . 84 3.4 Evaluation of the PBR model against bench-scale experiments . . . . 88 3.5 Representative results of complete consumption and the issue of vol- ume change . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 3.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 4 Thermal Feedback in Canonical Pool Fire Configurations 101 4.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 4.2 Medium-scale methanol pool fire . . . . . . . . . . . . . . . . . . . . . 102 4.2.1 Numerical configuration . . . . . . . . . . . . . . . . . . . . . 103 4.2.2 Flame structure . . . . . . . . . . . . . . . . . . . . . . . . . . 105 4.2.3 Thermal feedback . . . . . . . . . . . . . . . . . . . . . . . . . 111 4.3 Turbulent ethylene burner with controlled coflow (FM-Burner) . . . . 115 4.3.1 Numerical configuration . . . . . . . . . . . . . . . . . . . . . 116 4.3.2 Flame structure . . . . . . . . . . . . . . . . . . . . . . . . . . 119 4.3.3 Radiant emissions . . . . . . . . . . . . . . . . . . . . . . . . . 123 4.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 vi 5 Simulations of Thermal Degradation of Biomass Particles under Prescribed External Gas Conditions 130 5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 5.2 Pyrolysis of biomass particles under steady and oscillatory heating conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 5.2.1 Particle response time . . . . . . . . . . . . . . . . . . . . . . 134 5.2.2 Response of thick charring particles to radiative heating . . . 134 5.2.3 Response of thick non-charring (shrinking) particles to radia- tive heating . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 5.2.4 Response of thin non-charring (shrinking) particles to radiative heating . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139 5.2.5 Response of thin particles to convective heating . . . . . . . . 140 5.2.6 Response of thick particles to convective heating . . . . . . . . 142 5.3 Maps of the pyrolysis and oxidation of porous biomass particles . . . 145 5.3.1 Effect of the particle size . . . . . . . . . . . . . . . . . . . . . 151 5.3.1.1 Maps of small particles (2 mm diameter) heated for a long duration (τPIHF = 120 s) . . . . . . . . . . . . 151 5.3.1.2 Maps of large particles (20 mm diameter) heated for a long duration (τPIHF = 120 s) . . . . . . . . . . . . 154 5.3.2 Effect of the residence time (τPIHF) . . . . . . . . . . . . . . . 158 5.3.3 Threshold value of the particle core temperature for complete smoldering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160 5.4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165 6 Simulations of Fire Spread in Engineered Cardboard Fuel Beds 168 6.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168 6.2 Numerical configuration . . . . . . . . . . . . . . . . . . . . . . . . . 169 6.2.1 Computational domain and fuel bed setup . . . . . . . . . . . 169 6.2.2 Computational mesh resolution . . . . . . . . . . . . . . . . . 175 6.2.3 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . 177 6.2.4 Sub-models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178 6.2.5 Computational cost . . . . . . . . . . . . . . . . . . . . . . . . 178 6.3 Fire behavior in Burn 53 and Burn 67 . . . . . . . . . . . . . . . . . 179 6.3.1 Flow field at cold flow . . . . . . . . . . . . . . . . . . . . . . 179 6.3.2 Flow field during combustion . . . . . . . . . . . . . . . . . . 182 6.3.3 Flame structure . . . . . . . . . . . . . . . . . . . . . . . . . . 184 6.3.4 Fire intensity and rate of spread . . . . . . . . . . . . . . . . . 187 6.3.5 Local particle viewpoint . . . . . . . . . . . . . . . . . . . . . 190 6.4 The Pseudo Incident Heat Flux (PIHF) . . . . . . . . . . . . . . . . . 193 6.5 The flame residence time . . . . . . . . . . . . . . . . . . . . . . . . . 194 6.6 Fire behavior variation with the fuel bed packing . . . . . . . . . . . 197 6.6.1 Global fire behavior . . . . . . . . . . . . . . . . . . . . . . . . 197 6.6.2 Local particle viewpoint . . . . . . . . . . . . . . . . . . . . . 200 6.7 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203 vii 7 Simulations of Fire Spread in Idealized Uniform and Mixed Size Pine Wood Fuel Beds 204 7.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204 7.2 Numerical configuration . . . . . . . . . . . . . . . . . . . . . . . . . 205 7.2.1 Computational domain and fuel bed setup . . . . . . . . . . . 207 7.2.2 Computational mesh resolution . . . . . . . . . . . . . . . . . 209 7.2.3 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . 211 7.2.4 Sub-models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 212 7.2.5 Computational cost . . . . . . . . . . . . . . . . . . . . . . . . 212 7.3 Fire spread in the uniform fuel bed . . . . . . . . . . . . . . . . . . . 212 7.3.1 Global fire behavior . . . . . . . . . . . . . . . . . . . . . . . . 213 7.3.2 The structure of the fire . . . . . . . . . . . . . . . . . . . . . 216 7.3.3 Sensitivity to the grid resolution . . . . . . . . . . . . . . . . . 217 7.4 Fire spread in the mixed-size fuel patch . . . . . . . . . . . . . . . . . 224 7.4.1 Variations with the fuel bed height . . . . . . . . . . . . . . . 225 7.4.2 Local gas conditions at mid-elevation of the fuel bed . . . . . 228 7.4.3 Particle behavior in the wind-dominated fire . . . . . . . . . . 231 7.4.4 Particle behavior in the buoyancy-dominated fire . . . . . . . 237 7.4.5 The Pseudo Incident Heat Flux (PIHF) . . . . . . . . . . . . . 238 7.4.6 Interpretation of the particles’ behavior using the quasi-steady 2-D maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 243 7.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247 8 Conclusion and Future Directions 249 A Numerical Algorithms of the Particle Burning Rate Model (PBR) 257 A.1 Discretization of the governing equations . . . . . . . . . . . . . . . . 260 A.1.1 Solid-phase mass conservation equations . . . . . . . . . . . . 260 A.1.2 The temperature equation . . . . . . . . . . . . . . . . . . . . 262 A.1.3 The gas-phase oxygen mass equation . . . . . . . . . . . . . . 267 A.1.4 The pressure equation . . . . . . . . . . . . . . . . . . . . . . 272 A.2 Iterative scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275 A.3 Temporal discretization . . . . . . . . . . . . . . . . . . . . . . . . . . 276 A.4 Deforming mesh and re-meshing capabilities . . . . . . . . . . . . . . 278 B Modified Eddy Dissipation Model for Diffusion Controlled Flames 280 B.1 Model derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280 B.2 Simulations of two-dimensional laminar counter-flow methane-air flames284 C Extension of the LSP Soot Model to LES Framework 287 C.1 Conservation of soot mass . . . . . . . . . . . . . . . . . . . . . . . . 287 C.2 Soot formation and oxidation terms . . . . . . . . . . . . . . . . . . . 288 C.3 Closure model for turbulent soot formation and oxidation rates . . . . 289 C.4 Integration of the probability density function . . . . . . . . . . . . . 291 viii D Verification of the WSGG Model Implementation in OpenFOAM 292 D.0.1 Radiation configuration 1 . . . . . . . . . . . . . . . . . . . . 293 D.0.2 Radiation configuration 2 . . . . . . . . . . . . . . . . . . . . 294 D.0.3 Radiation configuration 3 . . . . . . . . . . . . . . . . . . . . 295 D.0.4 Radiation configuration 4 . . . . . . . . . . . . . . . . . . . . 297 E Installation of PBRFoam on HPC Clusters and its Parallel Performance 299 E.1 Installation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299 E.2 Parallel Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . 301 Bibliography 303 ix List of Tables 3.1 Analytical solutions of heat conduction inside solid fuel particles of different geometries, where τ = αst/ζ 2, αs = ks/ρscs, Tg the external gas temperature, and Ts,i the initial particle temperature. . . . . . . . 69 3.2 Calibrated model parameters for pine wood. . . . . . . . . . . . . . . 85 3.3 Calibrated chemical kinetic parameters for pine wood. . . . . . . . . . 85 3.4 Thermo-physical properties of cardboard. . . . . . . . . . . . . . . . . 87 3.5 Chemical kinetic parameters for thermal degradation of cardboard. . 87 5.1 Input parameters of the biomass particles used in the oscillatory heating study. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 5.2 Ignition and burnout times of particles exposed to quasi-steady (QS) radiative heating and oscillatory radiative heating at 0.1 and 1 Hz. The percentage difference in burnout time is calculated relative to the QS case. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 5.3 Ignition and burnout times of particles exposed to quasi-steady (QS) convective heating and oscillatory convective heating at 0.1 and 1 Hz. The percentage difference in burnout time is calculated relative to the QS case. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143 6.1 Conditions of Burn 53 and Burn 67. . . . . . . . . . . . . . . . . . . . 175 6.2 Characteristics of Burn 53 and Burn 67. A comparison between observed and simulated quantities. . . . . . . . . . . . . . . . . . . . 187 7.1 Design parameters of the buoyancy-dominated and wind-dominated line fires in the uniform pine wood fuel bed. . . . . . . . . . . . . . . 206 7.2 Simulated parameters of the buoyancy-dominated and wind-dominated line fires in the uniform pine wood fuel bed. . . . . . . . . . . . . . . 216 7.3 Summary of the simulation results corresponding to particles in the mixed-size patch located at x = 40 m, y = 0.2 m, z = 0 position. . . 242 A.1 Definition of the mesh parameters. In these expressions, Arect is the exposed surface area of rectangular particles, and Lcyl is the length of the cylindrical-shaped particles. The subscript C refers to the center of the cell; the subscript R refers to its right-boundary. . . . . . . . . 259 x List of Figures 1.1 Scales of the wildland fire problem. Images sources: Smithsonian’s National Zoo & Conservation Biology Institute [3]; Josh Edelson via Getty Images; Greenpeace International [4]; and NASA Earth Observatory [5]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.2 The different classes of CFD models used for simulating wildland fire behavior: combustion solvers resolve dynamics at the vegetation and flame scales; wildfire solvers resolve dynamics at the fireline and geographical scales; atmospheric boundary layers (ABL) solvers resolve dynamics at the meteorological scales. . . . . . . . . . . . . . . . . . 7 1.3 The larger framework of this project: chain of model development. . . 10 2.1 A schematic of fire spread in a fuel bed showing the characteristic length-scales of the flame and the different zones identified based on the particle burning state. . . . . . . . . . . . . . . . . . . . . . . . . 42 2.2 Schematic side view of a typical fuel bed showing fuel patches separated by a given distance (left). The computational grid is shown in the background. Each fuel patch is modeled by a collection of mono- or poly-dispersed particles described at the sub-grid level (right). The thermal degradation of each particle is tracked using the PBR solver through the computational grid shown in the inset. . . . . . . . . . . 44 2.3 Variation of the convective heat transfer coefficient with: a) wind velocity for particles with different sizes; and b) particle’s effective diameter at a wind velocity of 1 m/s. . . . . . . . . . . . . . . . . . . 51 2.4 Variation of drag coefficient with Reynolds number for cylindrical- and rectangular-shaped particles. . . . . . . . . . . . . . . . . . . . . 53 2.5 Temperature profile of laminar counter-flow diffusion flames under adiabatic conditions obtained from well-resolved simulations using the modified EDM. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 2.6 Flowchart of the computational modeling framework: the chart of the PBRFoam solver is highlighted in yellow; the chart highlight in gray depicts an interface; the chart of the PBR solver is highlighted in blue. 65 3.1 Temperature evolution inside a semi-infinite rectangular-shaped solid particle. The exposed surface is at ζ = 0. . . . . . . . . . . . . . . . . 70 xi 3.2 Temperature evolution inside a) cylindrical- and b) spherical-shaped solid particles. The exposed surface is at ζ = 0.01 m. . . . . . . . . . 70 3.3 Simulations (solid lines) and analytical solution (symbols) of the mo- mentum exchange between solid and gas-phases: a) transient velocity of the solid and the gas-phases, b) momentum of the gas-phase and total momentum of the system (gas+solid). . . . . . . . . . . . . . . . 72 3.4 Comparison between simulations with different grid resolutions (solid lines) and the analytical solution (symbols) of the pressure variation inside a fuel bed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 3.5 Simulations with different grid resolutions (solid lines) and analytical solution (symbols) of the radiation absorption by the fuel bed: a) radiation intensity variation with different packing ratios, b) spatial variation of the radiation intensity along the fuel bed length. . . . . . 76 3.6 Spatial variation of the radiation intensity inside a fuel bed that emits and self-absorbs radiation. The solid lines refer to the numer- ical simulations, while the symbols refer to the analytical solution. The emission contribution obtained from a case where absorption is neglected is also shown. . . . . . . . . . . . . . . . . . . . . . . . . . . 77 3.7 Smoldering propagation in idealized charred pine wood bed. The gas-phase is colored by the mass fraction of CO2. The fuel bed is colored by a) the percentage mass with red indicating 100% char and blue indicating 100% ash, b) the positive net radiative heat flux indicating heating zones. . . . . . . . . . . . . . . . . . . . . . . . . . 79 3.8 Smoldering propagation in idealized charred pine wood bed: a) time evolution of the back and front edges of the smoldering zone; b) smoldering intensity. . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 3.9 Variation of the mass loss rate of a micro-scale pine wood sample with the particle temperature at a) heating rates of 5 and 10 K/min in the inert environment; and b) heating rate of 10 K/min in oxidative environments. Lines: PBR simulations using the kinetic parame- ters of Lautenberger and Fernandez-Pello [44]. Symbols: the TGA measurements of Ref. [80]. . . . . . . . . . . . . . . . . . . . . . . . . 83 3.10 Variation of the mass loss rate of a micro-scale pine wood sample with the particle temperature at a) heating rates of 5 and 10 K/min in the inert environment; and b) heating rate of 10 K/min in oxidative environments. Lines: PBR simulations using the kinetic parameters of Anca-Couce et al. [80]. Symbols: the TGA measurements of Ref. [80]. 84 3.11 Evolution of the cardboard particle normalized mass loss rate (top row) and normalized mass (bottom row) with temperature. Comparison of the simulated thermal degradation of cardboard using the calibrated PBR model (solid lines) and the TGA measurements of Refs. [81, 82] (symbols) at a heating rate of 10 K/min in a) 100% N2 environment, b) 10% O2 − 90% N2. . . . . . . . . . . . . . . . . . . . . . . . . . . . 89 xii 3.12 Comparison of the simulated heat flow (solid line) and the DSC measurements of Ref. [81] collected at a heating rate of 10 K/min in an inert environment. . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 3.13 Time evolution of the mass loss rate of the 3.8 cm pine wood particle in inert and oxidative environments at a) 25 kW/m2 and b) 40 kW/m2 irradiation. Comparison between simulations of the PBR model (lines) and experimental data of Kashiwagi et al. [85] (symbols). . . . . . . . 91 3.14 Spatial variation of exposed surface and in-depth temperatures of the 3.8 cm pine wood particle at 40 kW/m2 irradiation in: a) 100% N2, b) 89.5% N2 − 10.5% O2, and c) air. Comparison between simulations from the PBR model (lines), experimental data of Kashiwagi et al. [85] (symbols). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 3.15 Spatial variation of exposed surface temperature and in-depth tem- peratures of the 3.8 cm pine wood particle at 25 kW/m2 irradiation in 89.5% N2 − 10.5% O2. Comparison between simulations using the PBR model (lines), experimental data of Kashiwagi et al. [85] (symbols). 94 3.16 Predictions, using different grid resolutions, of a) MLR and b) the temperature at 5 mm depth inside the particle under 40 kW/m2 irradiation in ambient air. . . . . . . . . . . . . . . . . . . . . . . . . 95 3.17 Variation of a 10 mm diameter pine wood particle mass loss rate (a) and mass (b) with time obtained using PBR simulations. The particle is exposed to 60 kW/m2 radiative flux for 60 s in ambient air flowing at 1 m/s. The vertical dashed line indicates the end of the external heating period. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96 3.18 Two snapshots of spatial profiles of temperature (a) and volume fraction of char and ash (b) inside a 10 mm diameter pine wood particle exposed to 60 kW/m2 radiative flux for 60 s in ambient air flowing at 1 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 3.19 Two snapshots of spatial profiles of oxygen mass fraction (a) and char oxidation reaction rate (b) inside a 10 mm diameter pine wood particle exposed to 60 kW/m2 radiative flux for 60 s in ambient air flowing at 1 m/s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 3.20 Temporal evolution of the volume of 10 mm diameter pine wood particle exposed to 60 kW/m2 radiative flux for 60 s in ambient air flowing at 1 m/s. The dashed line represents simulations conducted using the char and ash yields listed in Table 3.2. The solid line represents simulations conducted with modified char and ash yields of 0.3 and 0.005, respectively. . . . . . . . . . . . . . . . . . . . . . . . . 99 4.1 A 3-D view of the computational domain of the methanol pool fire (left), and 2-D sections showing an enlarged view of the medium mesh resolution near the fuel pan (right). . . . . . . . . . . . . . . . . . . . 103 xiii 4.2 A view of the virtual heat flux gauges used in the numerical config- uration: a) side view of the computational domain near the burner lip and the gauges; b) top view of the burner surface and the gauges. The gauges are colored in black. The computational domain and the burner surface are colored in red. . . . . . . . . . . . . . . . . . . . . 106 4.3 Sequence of instantaneous snapshots of the methanol pool flame taken at selected times during a representative instability cycle. The flame is visualized using volume rendering of the high-temperature region (defined as the region where temperatures are larger than 800 K). The time between successive snapshots is 0.06 s and the total duration of the sequence is 0.42 s. The fuel pan and liquid pool surface are colored grey. The simulation was performed with the fine mesh and with PGRF. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 4.4 Sequence of instantaneous snapshots of the flame-flow configuration taken at selected times during a representative instability cycle and plotted in a central vertical plane. The flame is visualized using isocontours of temperature; the flow is visualized using velocity vectors. The fuel pan and liquid pool are colored white. See the caption of Fig. 4.3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109 4.5 The power spectrum of the temporal variations of the simulated heat release rate. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 4.6 Centerline vertical variations of: mean temperature, T ; mean vertical velocity, u; mean radial velocity, v; rms temperature, Trms; rms vertical velocity, urms; rms radial velocity, vrms; cross-correlation between vertical velocity and temperature, u′T ′. Comparisons between experimental data (symbols) and simulation results obtained with a coarse mesh (dashed blue line), a medium mesh (dash-dotted black line), and a fine mesh (solid red line). Simulations performed with PGRF. The vertical bars denote the uncertainties in the experimental data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 4.7 Radial variations of the mean gauge heat flux at 13 mm above the pool surface: (Left) radiative heat flux, q̇′′rad; (Right) total heat flux, q̇′′tot. Comparisons between experimental data (square symbols) and simulations performed with the medium mesh and with PGRF (dashed red line with diamond symbols) or with WSGG (dashed blue line with circle symbols). The vertical bars denote the uncertainties in the experimental data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 4.8 Radial variations of the mean radiative heat flux at 13 mm above the pool surface. Comparisons between experimental data (square symbols) and simulations performed with the medium mesh and with the WSGG-Bordbar model (dashed blue line with circle symbols), WSGG-Cassol (solid red line with square symbols), and PMC-LBL (dotted purple line with x symbols). The vertical bars denote the uncertainties in the experimental data. . . . . . . . . . . . . . . . . . 115 xiv 4.9 Comparisons between experimental data (square symbols) and simula- tions performed with the WSGG-Bordbard model with 36 solid angles (dashed red line with diamond symbols), 64 solid angles (dashed blue line with circle symbols), or 100 solid angles (dotted purple line with square symbols). See the caption of Fig. 4.7. . . . . . . . . . . . . . . 116 4.10 Photographs of the FM-Burner taken during experiments conducted at different oxygen concentrations [88]. . . . . . . . . . . . . . . . . . 117 4.11 A 3D representation of the computational domain of the FM-Burner (left), and a slice showing the computational mesh near the burner (right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 4.12 Contours of the instantaneous and mean values of: the temperature (T and T )), the mixture fraction (Z and Z)), and the soot volume fraction (fv and fv)). The white isocontour represents fv = 10−6. The results correspond to the FM-Burner at 20.9% O2 coflow. . . . . . . . 120 4.13 Contours of the instantaneous and mean values of: the volumetric soot formation rate (ω̇′′′ sf and ω̇′′′ sf )), and the volumetric soot oxidation rate (ω̇′′′ sf and ω̇′′′ sf)). The results correspond to the FM-Burner at 20.9% O2 coflow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 4.14 Radial profiles of mean and rms temperature at different elevations above the fuel surface in the FM-Burner at 20.9% O2 coflow. Compar- ison between experiments (symbols) and simulations with PGRF (red dash-dot line), WSGG with gas and soot absorption/emission (blue dashed line), and WSGG with gas radiation only (solid black line). . 124 4.15 Radial profiles of mean and rms soot volume fraction at different elevations above the fuel surface of the FM-Burner at 20.9% O2 coflow. Comparison between experiments (symbols) and simulations with the WSGG model (blue dashed line). . . . . . . . . . . . . . . . . . . . . 125 4.16 Axial variations of the mean radiative power of the FM-Burner at 20.9% O2 and 15.2% O2 concentrations in the coflow. Comparisons between experimental data (symbols) and simulations performed with the PGRF model (dotted red line), the WSGG model with gas and soot contributions (dashed blue line), and the WSGG with gas contribution only (solid black line). . . . . . . . . . . . . . . . . . . . . . . . . . . 127 4.17 Axial variations of the radiation absorption and emission of the FM- Burner at 20.9% O2 concentrations in the coflow. The results are estimated using the WSGG model accounting for gas and soot radiation.127 5.1 Response of a 20 mm-thick charring particle to radiative heating: a) particle surface and core temperatures; b) particle mass loss rate. Comparison between particle response to quasi-steady forcing (QS) (lines with symbols) and oscillatory forcing at f = 0.1 Hz or f = 1 Hz (lines). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136 5.2 Response of a 20 mm thick non-charring particle to radiative heating: a) particle size; b) particle mass loss rate. See caption of Fig. 5.1. . . 138 xv 5.3 Response of the 2 mm non-charring particle to radiative heating: a) particle surface and core temperatures; b) particle mass loss rate. See caption of Fig. 5.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140 5.4 Response of a 2 mm non-charring particle to convective heating: a) particle surface and core temperatures; b) particle mass loss rate. Comparison between the particle response to quasi-steady forcing (QS) (lines with symbols) and oscillatory forcing by oscillations in Tg at f = 0.1 Hz or f = 1 Hz (lines). . . . . . . . . . . . . . . . . . . . . 142 5.5 Response of a 2 mm non-charring particle to convective heating: a) particle surface and core temperatures; b) particle mass loss rate. Comparison between the particle response to quasi-steady forcing (QS) (lines with symbols) and oscillatory forcing by oscillations in both Tg and ug at f = 0.1 Hz or f = 1 Hz (lines). . . . . . . . . . . . 144 5.6 Mass loss rate response of the a 20 mm thick particle to radiative heating: a) non-charring; b) charring. See caption of Fig. 5.5 . . . . . 145 5.7 Response of a 10 mm particle to an external irradiation of G = 60 kW/m2 for 60 s, and ambient air flow at ug = 1 m/s: a) the net total heat flux; b) the surface and core temperatures; c) the convective heat transfer coefficient; d) the surface emissivity. The duration of the external irradiation is highlighted in red. . . . . . . . . . . . . . . 148 5.8 Response of a 10 mm particle to an external irradiation of G = 60 kW/m2 for 60 s, and ambient air flow at ug = 1 m/s: a) the reactions rates of Rd-Rop; b) the normalized solid constituent mass. The duration of the external irradiation is highlighted in red. . . . . . 149 5.9 Maps of a 2 mm pine wood particle for τPIHF = 120 s: a) degree of completion of the drying reaction Rd; b) degree of completion of the pyrolysis reaction Rp+Rop; c) degree of completion of the char oxidation reaction Rco. The gray zone indicates PIHF < PIHF0. . . . 153 5.10 Maps of a 2 mm pine wood particle for τPIHF = 120 s: a) peak value of the pyrolysis rate; b) pyrolysis duration. The red solid iso-contour depicts the flaming threshold of 1 g/s/m2. The gray zone indicates PIHF < PIHF0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154 5.11 Maps of a 2 mm pine wood particle for τPIHF = 120 s: a) the peak particle temperature Rd; b) the temperature at the core of the particle at t = τPIHF. The gray zone indicates PIHF < PIHF0. . . . . . . . . . 155 5.12 Maps of a 20 mm pine wood particle for τPIHF = 120 s: a) degree of completion of the drying reaction Rd; b) degree of completion of the pyrolysis reaction Rp+Rop; c) degree of completion of the char oxidation reaction Rco. . . . . . . . . . . . . . . . . . . . . . . . . . . 156 5.13 Maps of a 20 mm pine wood particle for τPIHF = 120 s: a) peak value of the pyrolysis rate; b) pyrolysis duration. The red solid iso-contour highlights the flaming threshold of 1 g/s/m2. . . . . . . . . . . . . . . 157 5.14 Maps of a 20 mm pine wood particle for τPIHF = 120 s: a) the peak particle temperature Rd; b) the temperature at the core of the particle at t = τPIHF. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158 xvi 5.15 Maps of the degree of completion of pyrolysis (left) and the degree of completion of char oxidation (right) for a 2 mm pine wood particle heated for different periods: a) τPIHF = 5 s, b) τPIHF = 15 s, and c) τPIHF = 30 s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161 5.16 Maps of the particle peak temperature (left), and the particle core temperature at the end of the heating period (right) for a 2 mm pine wood particle heated for different periods: a) τPIHF = 5 s, b) τPIHF = 15 s, and c) τPIHF = 30 s. . . . . . . . . . . . . . . . . . . . . 162 5.17 Maps of the degree of completion of pyrolysis (left) and the degree of completion of char oxidation (right) for a 20 mm pine wood particle heated for different periods: a) τPIHF = 15 s, b) τPIHF = 30 s, and c) τPIHF = 60 s. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 5.18 Maps of the particle peak temperature (left), and the particle core temperature at the end of the heating period (right) for a 20 mm pine wood particle heated for different periods: a) τPIHF = 15 s, b) τPIHF = 30 s, and c) τPIHF = 60 s. . . . . . . . . . . . . . . . . . . . . 164 5.19 Scatter plot of the char oxidation degree of completion against the particle core temperature achieved at the time corresponding to the end of the heating duration τPIHF. . . . . . . . . . . . . . . . . . . . . 165 6.1 Images from the experimental burns of fire spread in cardboard fuel beds conducted at the Missoula Fire Sciences Laboratory (courtesy of Dr. Mark Finney): a) top view of an experiment conducted on an inclined surface, featuring flaming and glowing zones; b) back view of a buoyancy-dominated fire experiment in the wind tunnel (Burn 53), featuring a uniform flame front with peak and trough structures; c) isometric view of the wind-dominated fire in the wind tunnel at high speed (Burn 67), featuring flaming and glowing zones as well as a curved flame front. . . . . . . . . . . . . . . . . . . . . . . . . . . . 170 6.2 Schematic of the wind tunnel facility at Missoula Fire Sciences Labo- ratory (adapted from Ref. [105]). The dimensions on the image are obtained from private communication with Dr. Mark Finney. The blue dashed box contours the simulated region. . . . . . . . . . . . . 171 6.3 The computational domain used in the simulations of the cardboard fire spread. The fuel bed is colored in yellow. . . . . . . . . . . . . . . 172 6.4 Variation of the fuel bed packing ratio with the axial spacing between particle arrays. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174 6.5 Contours of the velocity magnitude at the vertical center-plane of the wind tunnel (top slice) and at a horizontal plane at y = 0.025 m eleva- tion (bottom slice). The contours correspond to cold flow simulations of a) Burn 53 and b) Burn 67. . . . . . . . . . . . . . . . . . . . . . . 180 6.6 Profiles of the mean velocity at different locations in the streamwise direction in a) the low-velocity case (Burn 53), and b) the high-velocity case (Burn 67). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181 6.7 Profiles of the root-mean-square velocity. See the caption of Fig. 6.6. 181 xvii 6.8 Streamwise variation of ∆y+w in: a) the low-velocity case; b) the high- velocity case. The dashed line represents instantaneous values and the solid line represents mean values. The cardboard fuel bed starts at x = 0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182 6.9 Tunnel flow field in the case of wind-dominated fire (Burn 67). Con- tours of the axial velocity in the tunnel mid-plane. The velocity vectors are scaled by velocity magnitude. . . . . . . . . . . . . . . . . 183 6.10 Tunnel flow field in the case of a buoyancy-dominated fire (Burn 53) with a) a prescribed inlet velocity, b) a prescribed total pressure. See the caption of Fig. 6.9. . . . . . . . . . . . . . . . . . . . . . . . . . . 184 6.11 A 3-D rendering of the simulated fires: a) back view of Burn 53 showing the iso-surface of hot gases at 1000 K and the cardboard particles colored by the mass loss rate (fire spreads into the page); b) isometric view of Burn 67 showing the iso-volume of hot gases > 550 K and the cardboard particles colored by their surface temperature (fire spreads from left to right). . . . . . . . . . . . . . . . . . . . . . . . . 186 6.12 Global fire behavior of Burn 53 and Burn 67 in terms of a) rate spread, and b) fire line intensity. The dashed lines represent the slope from the measured rate of spread. . . . . . . . . . . . . . . . . . . . . . . . 188 6.13 Time evolution of the pyrolysis front obtained from simulations with different boundary conditions: a) effect of inlet boundary in Burn 53; b) effect of the side walls on Burn 67. . . . . . . . . . . . . . . . . . . 189 6.14 Time evolution of the pyrolysis front obtained from simulations with the baseline mesh and a refined mesh. . . . . . . . . . . . . . . . . . . 190 6.15 Local viewpoint of the gas conditions in Burn 67 at the top of the particle located at x = 3 m, z = 0: a) local gas temperature, b) local gas velocity, c) local oxygen mass fraction; and d) local irradiation. . 192 6.16 The behavior of the particle located at x = 3 m, z = 0 in Burn 67: a) particle surface and core temperature, b) particle constituents masses, c) particle mass loss rate; and d) net heat flux at the particle’s exposed surface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 193 6.17 The convective and radiation contributions to the pseudo incident heat flux (PIHF) extracted at the top of the fuel bed at x = 3 m, z = 0 in a) Burn 53, and b) Burn 67. . . . . . . . . . . . . . . . . . . . . . . . 195 6.18 Flame residence time based on a) gas temperature, b) particle flaming duration, and c) duration of intense PIHF. The red solid line is a spline fit of the raw data. The vertical dashed blue lines delimit the start and the end of the residence time. The data correspond to the particle located at x = 3 m, z = 0 in Burn 67. . . . . . . . . . . . . . 196 6.19 A 3-D rendering of the flaming zone variation with the packing ratio βs. The flaming particles are colored by the gaseous fuel release rate (GFRR) (i.e., the production rate of fuel from the pyrolysis reactions Rp and Rop). The hot gases at 1000 K are represented by the white iso-surface. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 198 xviii 6.20 Variations of the a) fire intensity, b) the flame zone depth, c) the fire rate of spread, and d) the flame height with the fuel bed packing ratio. The symbols represent the simulated cases and the solid line represents a linear fit to the simulated data. . . . . . . . . . . . . . . 199 6.21 The behavior of 36 particles in Burn 67, starting from x = 3 m, z = 0 and separated by 0.2 m in the x−direction and 0.25 m in the z− direction: a) particle surface temperature, b) normalized particle masses, c) particle state, and d) particle mass loss rate. . . . . . . . . 201 6.22 Variation of the residence time with the packing ratio: a) the particle flaming and smoldering times, b) the PIHF time, and c) the flame residence time based on the gas temperature. The symbols represent the average of 36 particles, the bars represent the standard deviation, and the solid lines represent linear fit. The dashed blue line represents the experimental fit of Finney et al. [2]. . . . . . . . . . . . . . . . . . 202 7.1 A 3-D view of the computational domain used in the simulations of fire spread in pine wood fuel beds: a) the uniform fuel bed with 2 mm particles (βs = 0.005); b) the fuel bed with a patch of mixed small/large particles (Dp = 2 and 10 or 20 mm) located at 38 ≤ x ≤ 42 m (βs = 0.0068 or 0.012). . . . . . . . . . . . . . . . . . . . . . . . 208 7.2 Propagation of the wind-dominated fire in the uniform bed: a) time evolution of the simulated flame, pyrolysis and smoldering zones (solid and dashed lines) and the estimated spread from Rothermel model (dotted line); b) time evolution of the simulated fire intensity from homogeneous gas-gas reactions (black line) and from heterogeneous solid-gas reactions (blue line). . . . . . . . . . . . . . . . . . . . . . . 214 7.3 Global behavior of the buoyancy-dominated fire in the uniform bed at 1 m/s external wind speed. See the caption of Fig. 7.2 . . . . . . . 215 7.4 Visualization of the wind-dominated fire in the uniform bed: a) iso- metric view (fire spreads from left to right) of; b) back view of the hot gases only (fire spreads into the page). The hot gases in (a) and (b) are rendered by an iso-volume of T > 550 K and colored by the gas temperature. The glowing solid particles in (a) are colored by their surface temperature. . . . . . . . . . . . . . . . . . . . . . . . . . . . 218 7.5 Visualization of the buoyancy-dominated fire in the uniform bed: a) isometric view (fire spreads from left to right) of; b) back view of the hot gases only (fire spreads into the page). The hot gases in (a) and (b) are rendered by an iso-volume of T > 550 K and colored by the gas temperature. The glowing solid particles in (a) are colored by their surface temperature. . . . . . . . . . . . . . . . . . . . . . . . . 219 7.6 Visualization of the solid particles colored by their respective burning state. The unburned particles are colored in green, the burned particles are colored in black, the flaming particles are colored red, and the smoldering particles are colored in yellow. . . . . . . . . . . . . . . . 220 xix 7.7 Global fire behavior in the buoyancy-dominated fire predicted using three grid resolutions of the gas-phase: a) pyrolysis front, b) fire intensity from homogeneous gas phase combustion, and c) fire intensity from heterogeneous particle reactions. The particle resolution is fixed at 25 µm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221 7.8 Global fire behavior in the wind-dominated fire predicted using three grid resolutions of the gas-phase. See the caption of Fig. 7.7 . . . . . 222 7.9 Global fire behavior in the buoyancy-dominated fire predicted using three mesh resolutions of the particles. The gas-phase resolution is 10 cm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 223 7.10 Behavior of the particle located 20 m downstream of the ignition burner in the buoyancy-dominated fire in terms of a) the temperature at the core of the particle; b) the particle mass. The results are obtained using three mesh resolutions of the particles, and a gas-phase resolution of 10 cm. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224 7.11 Snapshots of the wind-dominated fire while spreading across the mixed- size patch of 2 and 10 mm particles. The flame is rendered by an iso-volume of gases at Tg > 550 K. The solid particles are colored by their surface temperature. . . . . . . . . . . . . . . . . . . . . . . . . 226 7.12 The variation of a) the core temperature and b) the normalized mass of the lower, middle and upper parts of the 20 mm diameter particle located at x = 40 m, z = 0 in the mixed-size patch of the wind- dominated fire. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227 7.13 The variation of a) the core temperature and b) the normalized mass of the lower, middle and upper parts of the 10 mm diameter particle located at x = 40 m, z = 0 in the mixed-size patch of the wind- dominated fire. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228 7.14 External conditions around the particle evaluated at x = 40 m, y = 0.2 m, z = 0 position in the wind-dominated fire: a) local gas temperature; b) local irradiation; c) local oxygen mass fraction; d) local gas velocity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230 7.15 Time variation of the surface and core temperatures of the particles in the mixed-size patch with 2 mm and 10 mm particles (a, b), and in the mixed-size patch with the 2 mm and 20 mm particles (c, d). The data correspond to particles located at x = 40 m, y = 0.2 m, z = 0 in the wind-dominated fire. . . . . . . . . . . . . . . . . . . . . . . . . 232 7.16 Time variation of the normalized mass loss rate of particles in mixed- size patch of the wind-dominated fire. See caption of Fig. 7.15. . . . . 233 7.17 Time variation of the normalized solid constituents masses of particles in mixed-size patch of the wind-dominated fire. See caption of Fig. 7.15.234 7.18 Spatial distribution of: a) temperature; b) volume fraction of char and ash; c) oxygen mas fraction; d) char oxidation reaction rate inside the 10 mm particle located at x = 40 m, y = 0.2 m, z = 0 in the mixed-size patch of the wind-dominated fire. . . . . . . . . . . . . . . 236 xx 7.19 Spatial distributions inside the 20 mm particle. See the caption of Fig. 7.18. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237 7.20 Time variation of the normalized solid constituents masses of particles in the mixed-size patch with 2 mm and 10 mm particles (a, b), and in the mixed-size patch with the 2 mm and 20 mm particles (c, d). The data correspond to particles located at x = 40 m, y = 0.2 m, z = 0 in the buoyancy-dominated fire. . . . . . . . . . . . . . . . . . . . . . 239 7.21 Time variation of the a) total heat flux, b) convective heat flux, and c) radiative heat flux at the surface of the 2 and 10 mm particles located at x = 40 m, z = 0 in the mixed-size patch of the wind-dominated fire. The heat fluxes are evaluated at mid-elevation. . . . . . . . . . . 240 7.22 Time variation of the instantaneous values of the PIHF (black solid line), the moving average (red solid line), and the time integral of the intense PIHF of the 10 mm particle located at x = 40 m, y = 0.2 m, z = 0 position in the wind-dominated fire. . . . . . . . . . . . 242 7.23 Time variation of the convective and the radiative components of the PIHF of particles in mixed-size patch of the wind-dominated fire. See caption of Fig. 7.15. . . . . . . . . . . . . . . . . . . . . . . . . . . . 244 7.24 Time variation of the convective and the radiative components of the PIHF of particles in mixed-size patch of the buoyancy-dominated fire. See caption of Fig. 7.20. . . . . . . . . . . . . . . . . . . . . . . . . . 245 7.25 Maps of Rco degree of completion for 2 mm particles (a,b), 10 mm particles (c,d), and 20 mm particles (e,f). The cross symbol represents conditions from the buoyancy-dominated fire listed in Table 7.3. The circles represent similar conditions from the wind-dominated fire. . . . 246 A.1 One-dimensional computational mesh for particles with: (a) rectangu- lar slab geometry; (b) cylindrical or spherical geometry . . . . . . . . 258 A.2 Thermal boundary condition at the exposed surface of the particle. . 267 A.3 3D rendering of a fuel bed with discrete fuel particles colored by their corresponding time-step during fire spread. . . . . . . . . . . . . . . . 277 B.1 Predictions of the heat release rate per unit surface area of counter-flow methane/air diffusion flames under different mixing conditions. . . . . 285 B.2 Profiles of mixture fraction (left) and temperature (right) at the centerline distance between the fuel and oxidizer nozzles for an under- resolved flame using EDM and a well-resolved flame using single-step Arrhenius chemistry. . . . . . . . . . . . . . . . . . . . . . . . . . . . 286 B.3 Scatter of the estimated values of the HRRPUA using a representative value of χst = 1 s−1 in Eq. B.13 . . . . . . . . . . . . . . . . . . . . 286 D.1 Comparison between the numerical simulations (lines) and the LBL solution (symbols) of radiation configuration 1. . . . . . . . . . . . . . 294 D.2 Comparison between the numerical simulations (lines) and the LBL solution (symbols) of radiation configuration 2. . . . . . . . . . . . . . 295 xxi D.3 Comparison between the numerical simulations (lines) and the LBL solution (symbols) of radiation configuration 3. . . . . . . . . . . . . . 296 D.4 Comparison between the numerical simulations (lines) and the LBL solution (symbols) of radiation configuration 4. . . . . . . . . . . . . . 298 E.1 Test of the scaling performance of PBRFoam on the HPC cluster Expanse. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302 xxii Nomenclature Abbreviations ABL Atmospheric Boundary Layer CFD Computational Fluid Dynamics DL Deep Learning DSC Differential Scanning Calorimetry EDM Eddy Dissipation Model FMC Fuel Moisture Content GFRR Gaseous Fuel Release Rate GIS Geographic Information System HPC High Performance Computing HRR Heat Release Rate HRRPUA Heat Release Rate Per Unit Area LBL Line-by-line LES Large Eddy Simulation LSP Laminar Smoke Point MaCFP Measurement and Computation of Fire Phenomena MLR Mass Loss Rate MPI Message Passing Interface MRTE Multiphase Radiative Transfer Equation PBR Particle Burning Rate PDF Probability Density Function PGRF Prescribed Global Radiant Fraction PIHF Pseudo Incident Heat Flux PMC Photon Monte Carlo QS Quasi-steady RHS Right Hand Side ROS Rate of Spread TGA Thermogravimetric Analysis UMD University of Maryland WALE Wall Adapting Local Eddy-viscosity WSGG Weighted-Sum-of-Gray-Gases xxiii Symbols α heat diffusivity β packing ratio χrad radiant fraction ∆H heat of reaction/combustion ∆y+w vertical wall unit δ size (half-thickness) of rectangular-shaped particles δik Kronecker symbol ∆LES LES filter size ω̇′′′ volumetric mass reaction rate of gas-phase species/soot ḣ′′′th volumetric rate of heat transfer ṁ′′′ solid-phase volumetric mass reaction rate ṁ′′ mass flux q̇′′′comb heat release rate of the gas-phase combustion q̇′′′rad net radiative power density of the gas-phase q̇′′ heat flux Q̇′′ rad radiative heat flux Q̇′ fire line intensity ϵ emissivity ϵTurb subgrid-scale turbulent dissipation rate η mass yield γ effective heat conductivity due to pore radiation κ radiation absorption coefficient µ dynamic viscosity ν kinematic viscosity Ω radiation solid angle ω inverse of periodic time ϕ coordinate factor ψ porosity ρ density σ Stefan-Boltzmann constant σp particle surface area to volume ratio τ characteristic time-scale/residence time τw wall shear stress xxiv Θ normalized temperature h̃th thermal enthalpy of the gas ζ general coordinate A pre-exponential factor A⋆ projected surface area of the particle Ap the area of the particle’s exposed surface Arect surface area of rectangular particles Bi Biot number c heat capacity CD aerodynamic drag coefficient CL aerodynamic lift coefficient cp specific heat at constant pressure CA shape factor CFL Courant–Friedrichs–Lewy number D mass diffusivity Deff effective diameter E activation energy Em radiation emission of the gas-phase mixture Es radiation emission of the solid-phase ERco degree of completion of the char oxidation reaction ERd degree of completion of the drying reaction ERp+Rop degree of completion of the pyrolysis reactions f frequency f− i , f + i flux across cell’s upstream or downstream faces F ′′′ i aerodynamic force vector component (per unit volume) fv soot volume fraction FD,j component of the aerodynamic drag force FL,j component of the aerodynamic lift force FO Fourier number G irradiation gi gravity vector component in index notation h heat transfer coefficient h0f enthalpy of formation I radiation intensity xxv K permeability k thermal conductivity kTurb subgrid-scale turbulent kinetic energy Lcyl length of cylindrical particles lsp laminar smoke point height Le Lewis number m mass m′′ f fuel loading n reactant exponent NC Byram convection number Np number of particles NuD Nusselt number based on effective diameter P presumed probability density function p pressure Pr Prandtl number R radius of cylindrical and spherical particles/ideal gas constant r radial coordinate ReD Reynolds number based on effective diameter S− i , S + i surface area of cell’s upstream or downstream faces Sc Schmidt number T temperature t time u velocity uj velocity vector component in index notation V volume X volume fraction x rectilinear coordinate xj coordinate in index notation Y mass fraction y rectilinear coordinate Z mixture fraction z rectilinear coordinate Zst stoichiometric mixture fraction xxvi Subscripts ∞ freestream or ambient a ash ad adiabatic bed a control volume of solid vegetation and the gas-phase C evaluated at the cell center c char conv convective ds dry solid g gas g→s gas to solid hrr heat release rate i cell or face index k species/reaction index L evaluated at the cell’s upstream face m mixture of gases p particle R evaluated at the cell’s downstream face rad radiative Rco char oxidation reaction Rd drying reaction Rop oxidative pyrolysis reaction Rp pyrolysis reaction s solid sf soot formation sg solid to gas so soot oxidation surf evaluated at the particle surface tot total Turb subgrid-scale turbulent quantity w evaluated at the wall ws wet solid xxvii xxviii 1 Introduction Wildland fires are a natural phenomenon that can occur in any region of the world where there is sufficient vegetation (e.g., forest fires, grassland fires, etc.) and a favorable climate (e.g., hot dry weather and high wind conditions). These fires can be caused by unintended actions such as lightning strikes and human activities. Wildland fires also include prescribed burns that intend to clean dense forests and manage the wildland. Uncontrolled wildland fires can spread rapidly, driven by strong winds and/or inclined terrains, and can quickly consume large areas of forest, grassland, or other types of vegetation, and sometimes lead to catastrophic disasters. The impacts of wildland fire disasters can be severe, including damage to property, loss of wildlife habitat, the displacement of people from their homes, as well as economic losses. According to the National Interagency Fire Center (NIFC) [1], wildfires have increased in occurrence by 223% since 1983, with 68,988 wildfires that burned 7,577,183 acres in the United States (U.S.). In 2022 alone, 4.5 million U.S. homes were identified to be at high or extreme risk of wildfire. Moreover, approximately 4.4 billion U.S. dollars were spent on fire suppression costs in 2021 [1]. In addition to these immediate impacts, wildland fires can also have long-term effects on the environment, including changes to soil fertility and water quality. 1 The increase in wildland fire risk in recent years is due mainly to: 1) climate change: the rising temperatures and prolonged drought conditions make it easier for fires to ignite and spread quickly; 2) improper land management practices: the accumulation of fuel in many forests makes fires more intense when they occur; and 3) urbanization: the expansion of urban communities to the wildland increases the risk of fires caused by human activities as well as exposes communities at the wildland- urban interface (WUI) to fire risk. As a result of the increasing wildland fire risk, advanced wildland fire behavior modeling tools are critical to help land managers, firefighters, and other responders plan and implement strategies to effectively predict, control, and mitigate wildland fire risks. This dissertation work is part of a larger research program sponsored by the U.S. Forest Service that aims at developing an advanced dynamical fire spread model in replacement of classical models and an operational tool for the real-time prediction of wildland fires. We describe in the following section the categories of wildland fire behavior models and how the current work compares to existing models and further model developments. 1.1 Wildland fire behavior modeling Wildland fire behavior refers to the physical and chemical processes that govern the ignition, growth, and spread of wildfires [2]. These processes include the heat transfer from the flame to the vegetation fuel or from a vegetation particle to another within the vegetation fuel bed; the production of flammable vapors due to thermal 2 degradation of individual fuel particles (known as pyrolysis); the reaction of these flammable vapors with ambient air to release heat, produce luminous turbulent buoyant flames and smoke (known as combustion); the reaction of the charred particles with ambient oxygen (known as glowing or smoldering); and the interaction of the turbulent buoyant flames with the surrounding environment, including the vegetation fuel, topography, and weather. Wildland fire behavior is affected by various factors, including the characteristics of the fuels, such as fuel material, fuel moisture content, and fuel arrangement, as well as weather conditions, such as temperature, humidity, wind speed and direction, and atmospheric turbulence. The topography of the landscape (e.g., flat or sloped terrain) also affects fire behavior, as it can significantly influence the rate of spread of the fire and its direction. Wildland fire behavior research is critical to understand the complex behavior of wildfires and their interactions with the surrounding environment. The modeling of wildland fire behavior combines various disciplines, including atmospheric science, ecology, physics, and computer science, to develop comprehensive models that can simulate the behavior of wildfires in different environments. These models are essential for developing effective wildfire management strategies, reducing the risk of wildfire damage, and protecting the safety of firefighters and local communities. 1.1.1 Scales of wildland fire behavior dynamics Before reviewing the approaches to model wildland fire behavior, it is important first to recognize that the dynamics of wildland fires involve multi-physics phenomena 3 occurring at multiple scales and different length scales are believed to play a role in fire behavior. As shown in Fig. 1.1 the wildland fire behavior can be studied at different scales: the vegetation scale characterized by the geometry of the biomass fuel particles; the flame scales represented by a characteristic flame height and width, as well as the length of the fire line that characterize the combustion and heat transfer processes; the geographical scales that characterize the terrain topography and land cover; and the meteorological scales represented by the depth of the atmospheric boundary layer that characterizes atmospheric conditions. Vegetation scales O(1 mm – 1 cm) Flame scales O(1 m) Geographical scales O(10s-100s m) Meteorological scales O(1 km) Figure 1.1: Scales of the wildland fire problem. Images sources: Smithsonian’s National Zoo & Conservation Biology Institute [3]; Josh Edelson via Getty Images; Greenpeace International [4]; and NASA Earth Observatory [5]. In wildland fire problems, the vegetation scale is on the order of a few millimeters or centimeters; the flames scales are the order of a few meters; the geographical scales 4 are typically on the order of a few tens or hundreds of meters; and the atmospheric boundary layer is on the order of kilometers. We focus in this dissertation on the wildland fire behavior at the vegetation-to-flame scales. 1.1.2 Semi-empirical approach to wildfire behavior modeling The first approach to wildland fire behavior modeling uses a semi-empirical technique that combines empirical observations with simplified mathematical equa- tions to describe fire behavior based on input data such as weather conditions, terrain, and fuel load. Details of this approach are outside the scope of this dissertation, but interested readers are invited to review the description of relevant tools such as BEHAVE [6], FARSITE [7], ELMFire [8], and FlamMap [9]). In short, these models rely on a classical empirical model that was proposed by Rothermel in 1972 [10] to estimate the rate of spread of the fire. These models can also use Geographic Information System (GIS) to create maps at regional scales that show areas at risk of wildfire. There are some limitations and drawbacks to the semi-empirical approach which makes its application in a variety of scenarios questionable [11]. For example, The Rothermel model is based on laboratory-scale observations of certain classes of fuels and may not accurately predict fire behavior under extreme conditions, such as very high winds, extreme slopes, or unusual fuel structures. Moreover, the Rothermel model treats the fuel bed as a uniform entity that burns at a constant rate to complete consumption. However, fire behavior, in reality, can be highly variable 5 depending on the characteristics of the fuel particles and the spatial heterogeneity of the fuel bed. Therefore, there is a strong interest to replace the Rothermel model with an advanced dynamical model based on physical principles [12]. 1.1.3 Physics-based approach to wildfire behavior modeling The second approach to wildland fire behavior modeling relies on the phys- ical description of the governing processes of fluid dynamics, heat transfer, and combustion. This physics-based approach has potential advantages over the semi- empirical modeling approach, including reducing model-related errors in predictions (by contrast to errors introduced from data or users) and explaining rather than just correlating fire behaviors. Physics-based models are also expected to better describe phenomena outside the range of current semi-empirical models, such as spread/no spread thresholds, acceleration/deceleration, fluctuating wind conditions, discontinuous and heterogeneous fuel beds, and conditions outside current laboratory and field data sets (very strong winds, large flame dimensions, etc.). The physics-based modeling has the potential to provide detailed information on the interactions between physical phenomena occurring at any of the wildland fire scales by means of Computational Fluid Dynamics (CFD). However, because of computational cost, the domain of application of 3D CFD-based tools is limited to a particular range of scales. Thus, current 3D CFD-based wildland fire models are scale-specific and belong to one of the following three classes (see Fig. 1.2): combustion solvers aimed at describing the coupling between pyrolysis, combustion, 6 radiation, and flow occurring at the vegetation and flame scales; wildfire solvers aimed at describing the coupling between combustion and flow occurring at the fire line scales and/or geographical scales; and atmospheric boundary layer (ABL) solvers aimed at describing the coupling between combustion and flow occurring at the meteorological scales. W flame L flame L fireline L topography L ABL O(1 km)O(1 m) O(10s-100s m) ABL solvers: WRF-SFIRE, WRF-Fire MESO-NH/ForeFire Combustion solvers: FDS, FIRESTAR3D, OpenFOAM WildfIre solvers: FIRETEC, WFDS L vegetation O(1 mm) O(1 cm) L land_cover CFD A multi-scale problem Figure 1.2: The different classes of CFD models used for simulating wildland fire behavior: combustion solvers resolve dynamics at the vegetation and flame scales; wildfire solvers resolve dynamics at the fireline and geographical scales; atmospheric boundary layers (ABL) solvers resolve dynamics at the meteorological scales. Examples of combustion solvers that have been developed for wildland fire dynamics applications include a group of models known as multiphase models [13–16]. These solvers provide a relatively fine-grained treatment of the fuel bed, the combustion, and the heat transfer processes that are responsible for flame spread. Simulations with these solvers are typically performed in small domains equivalent to a field scale. There are, however, simplifications in existing solvers which we would 7 like to improve in our modeling approach. Particularly, we would like to include a more sophisticated description of the in-depth thermal degradation processes through a porous medium treatment of the solid fuel. Other examples of combustion solvers include FDS [17] and FireFoam [18]; these solvers are well-established fire modeling tools that are primarily used for building fire applications. Examples of wildfire solvers include FIRETEC (see Refs. [19–21]) and WFDS (see Ref. [22] and the recent development in Refs. [23,24]). These solvers provide a coarse-grained treatment of unresolved vegetation-scale and flame-scale processes through a simplified (but physics-based) combustion model. Simulations with these solvers are typically performed in relatively larger domains (e.g., 1 km in size). Examples of atmospheric boundary layer (ABL) solvers that have been de- veloped for wildland fire dynamics applications include WRF-SFIRE and WRF- Fire [25–28], and MESO-NH/ForeFire [29]. These ABL solvers typically provide a macroscopic level treatment of unresolved vegetation-scale, flame-scale, fireline-scale, and geographical-scale processes through a parameterized rate of spread model based on the Rothermel model or simplified physical models of surface fire spread (e.g. [30]), which may have uncertainty in their application to a variety of scenarios. Simulations with ABL solvers are typically performed in arbitrary-size field-scale domains (from a few kilometers to several tens of kilometers and beyond). A strength of ABL solvers is that they are integrated with research-level or operational-level numerical weather prediction capabilities (i.e. WRF and MESO-NH) and therefore incorporate detailed descriptions of the fuel maps, topographic maps, and weather conditions. 8 1.2 Motivation and scope of this dissertation As described in the previous section, the existing operational tools rely on sim- plified semi-empirical models that are questionable, while there is a major challenge in the ability to span a wide range of scales with physics-based models due to the computational burden. Motivated by these challenges, this dissertation work is part of a larger project sponsored by the Missoula Fire Science Laboratory that aims to produce a dynamic physics-based dynamical fire spread model as a replacement for the classical Rothermel model and to deploy this model into an operational tool for real-time wildland fire simulations using Deep Learning (DL) techniques. Figure 1.3 shows the chain of model development in this project. As described in Fig. 1.3, there are three pieces in this project that are connected together. The first piece is a high-fidelity physics-based 3-D solver called PBRFoam that resolves the dynamics of wildland fire behavior at vegetation-flame scale using the Large Eddy Simulations approach (LES). The development of PBRFoam is the main objective of this dissertation and will be discussed in detail in the following chapters. In short, the tool PBRFoam supports the development of the second piece by providing simulation data and detailed diagnostics of the parameters needed to drive the dynamical 1-D fire spread model. For example, we would like to understand how the local gas condition around an individual particle inside a fuel bed varies with the packing of the fuel bed so that a reasonable boundary condition can be set in the dynamic 1-D model. The second piece is a dynamic 1-D fire spread model called LIHTFire [31]. This model aims at providing burning rates, heat transfer, and ignition with explicit 9 accounting of heterogeneous fuels and time-varying weather. This tool uses the Particle Burning Rate (PBR) model that we developed throughout this dissertation work to describe the thermal degradation of individual fuel particles. It accounts for the coupling between the solid particles and the gas-phase through a set of boundary conditions imposed at the fuel particle surface given estimates of the conditions of the gas-phase around the fuel bed. LIHTFire solves a single 1-D fire line spanning domains of about 50− 100 m in less than a minute, but this is still considered too slow to be implemented directly in real-time simulations of large fires. This takes us to the third piece which aims at using advanced Artificial Intelligence techniques to embed the fire model LIHTFire inside a Deep Learning (DL) model that accounts for combinations of fuel, weather, and topographical conditions (see Finney et. al. [32]). Boundary conditions Data High-resolution 3-D physics- based tool (PBRFoam) Dynamical 1-D fire spread model (LIHTFire) Deep learning operational tool Figure 1.3: The larger framework of this project: chain of model development. 1.3 Literature review of relevant experimental/modeling work in the vegetation-flame scale As we are interested in developing a computational tool for detailed modeling of wildland fire behavior at the vegetation-flame scale, we present here a literature review of the experimental and modeling efforts that have been carried out to study wildland fire behavior in this regime. 10 Vogel and Williams [33] found that the rate of flame propagation over horizontal arrays of vertically oriented matchsticks in quiescent air depends on the spacing, the length of the matchsticks, and whether the matchsticks are ignited from the bottom or the top. Wolff et al. [34] investigated fire spread across fuel beds made of arrays of toothpicks of uniform or mixed sizes in a movable ceiling wind tunnel. They identified a relationship between the rate of spread of the fire and the ratio between the wind speed to the fuel loading (i.e., the mass of fuel particles per unit bed area). They also highlighted some important effects in wind tunnel testing of fire spread such as the effect of flow circulation in closed ceiling tunnels, the edge effects, and the inability to obtain an asymptotic one-dimensional fire front. More importantly, they found the burnout time of individual fuel particles to vary linearly with the fuel loading and to be independent of the wind speeds considered in their tests. More recently, Di Cristina et al. [35] identified three regimes of flame spread across discrete wooden dowels under different particle spacing and wind speeds in the range of 2.2 to 3.4 m/s: 1) a continuous flame spread regime at a low spacing in which a flame covers multiple fuel particles; 2) a discrete flame spread regime at a high spacing in which a flame is only attached to an individual particle; 3) a quenching regime at large spacing at which a flame could not be sustained. A more practical technique to experimentally study flame spread in fuel beds comprised of discrete particles at a larger scale was introduced by Finney et al [36]. In these experiments, the fire spread over fuel beds made of laser-cut cardboard of different arrangements was tested in the Missoula wind tunnel at wind speeds up to 1.34 m/s. Similar to the observations made by Wolff et al. [34], these fire spread experiments 11 showed a linear relationship between the fuel packing ratio and the flame residence time. Using the same experimental technique, He et al. [37,38] conducted fire spread experiments in a smaller wind tunnel and at wind speeds between 1 and 3.8 m/s. They found that the flame residence time increases linearly with the fuel packing ratio. However, their data suggest that there is also a dependence of the residence time on wind speed. On the computational modeling side, the studies concerning the fire behavior in fuel beds comprised of discrete particles with different fuel bed characteristics are very limited. For example, Frangieh et al. [39] studied the flame structure of the cardboard fires that have been studied experimentally at the Missoula Fire Sciences Laboratory [36]. They were mainly interested in simulating the coherent flame structure observed in line fires (i.e., the peaks and troughs) using numerical simulations in a wind tunnel and in an open field. Awad et al. [40] conducted 2-D simulations to investigate the fuel moisture content (FMC) threshold for spread and no-spread conditions in grassland fires under different fuel bed loads (i.e., packing) and wind speeds. They found a correlation between the threshold FMC and the Byram number. In summary, there are a number of studies that aimed at characterizing the burning behavior of discrete fuel particles through experiments of fire spread in idealized fuel beds of thin discrete particles. However, the behavior of fire spread in mixed-size class fuel beds has received little to no examination to this point in the wildland fire literature. Moreover, existing experimental diagnostics and modeling efforts are very limited and there is a need for more studies concerning the burning 12 behavior in terms of the flame residence time, individual particle response, the degree of fuel consumption, the dependence on fuel bed characteristics such as fuel particle size and packing, the respective weights of radiative and convective heating, and the modes of flaming and smoldering combustion. 1.4 Objectives and contributions 1.4.1 Objectives The general objective of this project is to develop an advanced computational modeling capability that simulates wildland fire behavior at vegetation and flame scales with detailed descriptions of thermal degradation processes occurring during both flaming and smoldering combustion. Our main focus here is on the development, validation as well as testing of the computational modeling capability in canonical configurations. As described in the previous section, this modeling capability will serve in two ways: 1) providing data for the development of operational tools, and 2) developing our fundamental understanding of the wildland fire behavior through setting the basis for parametric studies and sensitivity analyses. Our specific objectives in this dissertation are to: 1. develop a particle burning rate (PBR) model that describes the thermal degradation of individual solid biomass/vegetation fuel particles and implement it as an object-oriented C++ tool 2. couple the model into a multiphase solver using an open-source library called 13 OpenFOAM [41] which allows the utilization of advanced models previously developed in our group at the University of Maryland, as well as an easy implementation of new models 3. validate the model against available experimental measurements and observa- tions 4. simulate flame spread across surrogate vegetation beds comprised of mono- /poly-dispersed particles with prescribed particle and environmental properties (i.e., bed height, surface-to-volume ratio, packing ratio, moisture content, and wind velocity) 5. study the global fire behavior in terms of flame structure, fire intensity, global fuel consumption, and the individual particle response as a function of the fuel packing and the duration of the thermal loading process, and the flame residence time. On the route to achieve these objectives, there are important intermediate steps discussed in the following chapters. These steps focus primarily on evaluating the ability of the models adopted and developed in this work to accurately describe two main drivers of fire spread: 1) pyrolysis and oxidation of isolated biomass particles at both micro- and bench-scales; 2) heat feedback in canonical fire flames in terms of thermal radiation and flame structure. 14 1.4.2 Author’s contributions The present work is part of a project sponsored by the U.S. Forest Service that aims at two objectives: 1) developing a dynamical 1-D fire spread model, and 2) studying the wildland fire behavior dynamics through high-fidelity 3-D simulations of spreading line fires in vegetation fuel beds. Past work conducted at the University of Maryland by Dr. Salman Verma [42] on high-fidelity 3-D simulations was focused on studying the structure of non-spreading laboratory-scale line fires produced from stationary gaseous burners in the presence of external wind or on inclined surfaces. On the other hand, the previous version of the dynamical 1-D fire spread model developed by Mr. Jason Forthofer at the Missoula Fire Sciences Laboratory included a particle burning rate (PBR) model that describes only the drying and the pyrolysis of thermally-thick non-porous particles, and it did not account for smoldering combustion. The present study extends the previous work by developing a model of thermal degradation of porous biomass fuel particles, which includes a description of both flaming and smoldering combustion, that can be implemented in the 1-D dynamical fire spread model as well as in the high-fidelity 3-D simulations of spreading fires. Several implementations and code development were conducted during this Ph.D. work. The candidate started first by developing an early MATLAB version of the PBR model that was originally created by Dr. Arnaud Trouvé to describe the drying and pyrolysis of thermally-thin as well as thermally-thick rectangular- shaped particles. The candidate extended this MATLAB code to account for the 15 thermal degradation of 1-D axisymmetric cylindrical- and spherical-shaped particles. The candidate then used this MATLAB code as a stand-alone solver to simulate the unsteady response of charring and non-charring biomass particles (in terms of ignition, mass loss rate, and burnout) when exposed to conditions of the surrounding gas that are fluctuating in time (i.e., surrounding gas temperature, velocity, and irradiance) [43]. Following the work of Lautenberger and Fernandez-Pello [44], the MATLAB version of the PBR model was then further developed by Dr. Arnaud Trouvé and the candidate to account for smoldering combustion by treating the particles as a porous-medium that includes in-depth oxygen diffusion. The candidate used this new solver to generate 2-D maps that describe the burning behavior of individual particles under prescribed thermal loads (Chapter 5). The candidate created an object-oriented C++ code as stand-alone software for the PBR model of porous biomass particles. The candidate also assisted with the coupling of the C++ PBR code with the dynamical 1-D fire spread model developed by Mr. Jason Forthofer at the Missoula Fire Science Laboratory. The candidate is also the developer of the 3-D multi-phase fluid dynamics solver PBRFoam that is based on the PBR code and the multi-purpose open-source C++ library OpenFOAM [41] (Chapter 2 and Appendix A). The candidate constructed a two-way coupling interface between the Lagrangian particle tracking library of OpenFOAM and the newly developed PBR solver. This interface exchanges information that allows the solid particles to evolve under the given gas-phase conditions and in turn modify the gas-phase by exchanging information about the contribution of the 16 solid-phase to the mass, momentum, energy, and radiation of the gas-phase. To construct the gas-phase solution algorithm in the PBRFoam solver, the candidate adopted a similar approach to that has been adopted in the well-established fire modeling tool FireFoam developed at FM Global [18], particularly, the description of the pressure-velocity coupling using the PIMPLE algorithm [45]. Additionally, the candidate conducted intermediate model developments and code implementations using the OpenFOAM library toward more accurate fire modeling as part of the Measurement and Computation of Fire Phenomena (MaCFP) working group. These efforts aimed at evaluating the accuracy of different fire modeling approaches in predicting thermal feedback in non-sooting and sooting buoyant turbulent diffusion flames. The candidate developed a modified version of the Eddy Dissipation Model (originally proposed by Magnussen and Hjertag [46]) using laminar diffusion theory to dynamically estimate the rate of fuel consumption based on laminar diffusion in fire regimes where laminar diffusion is dominating over turbulent mixing such as near the flame base (see Appendix B). The candidate also developed and implemented an extension of the Laminar Smoke Point (LSP) soot model (originally proposed by Yao et al. [47]) to account for subgrid-scale turbulent fluctuations in LES description of soot formation and oxidation rates using a probability density function called β−PDF (Appendix C). Furthermore, based on an original code implementation of a Weighted-Sum-of-Gray-Gases (WSGG) framework implemented in the OpenFOAM library by Sikic et al. [48], the candidate implemented a new version that uses a WSGG model proposed by Cassol et al. [49] that accounts for soot radiation (Appendix D). 17 In terms of model verification and calibration, the candidate conducted a series of benchmark tests to evaluate the accuracy of the implemented models and to check the establishment of a reliable computational framework (Chapter 3). As part of the MaCFP workshop, the candidate performed Large Eddy Simula- tions (LES) using FireFoam solver in two configurations that have been identified as target experiments by the MaCFP working group for validation of fire models. The emphasis in these simulations was on evaluating the ability of the EDM combustion model, the Prescribed Global Radiant Fraction model (PGRF), and the WSGG radiation modeling approach to predict thermal feedback in, first, a non-sooting methanol pool fire, then in a sooting ethylene burner (Chapter 4). Using the newly developed modeling capability (PBRFoam), the candidate performed LES simulations of fire spread in fuel beds made of vertically-oriented engineered cardboard sticks that have been studied experimentally at the Missoula Fire Sciences Laboratory [36] (Chapter 6). The candidate created a C++ script to generate the population of the cardboard particles inside the fuel beds using the given information about the separation distances between arrays of particles. The candidate constructed 3-D configurations of two experimental burns called Burn 53 and Burn 67, as well as 7 configurations corresponding to fuel beds with different particle separation and fuel bed packing. The emphasis in these simulations was on validating the model by comparing the simulated fire behavior in terms of ROS, flame structure, and residence time with available experimental measurements and observations. The candidate also performed LES simulations of fire spread in surrogate fuel 18 beds comprised of cylindrical-shaped pine wood particles using PBRFoam. The candidate constructed configurations that have either a uniform fuel bed of the same- size particles or a fuel bed that features a patch of mixed-size particles (Chapter 7). The focus of this study was on studying the degree of consumption of the particles, the respective weights of convective and radiative heating, and the transition from flaming to smoldering combustion. 1.4.3 Organization In chapter 2, the computational framework is described in detail. The governing equations for the thermal degradation of porous biomass particles are first introduced, followed by a description of the multi-phase formulation that we adopted in our implementation to describe the coupling between the solid biomass particles and the gas-phase. The chapter also provides details of the physical sub-models adopted to describe the convective heat transfer at the surface of the solid particles, the multi-phase radiation transport, the gas-phase combustion, as well as the sub-grid scale turbulence. This chapter is concluded with a summary of the implementation of the computational framework. In chapter 3, a series of benchmark test cases are presented to verify, calibrate and evaluate the computational modeling framework. This chapter first discusses verification test cases through simplified configurations where analytical solutions can be derived at both the particle level as well as the fuel bed level. Then, the calibration of the thermal degradation model with micro-scale experiments is 19 discussed. A discussion about the ability of the stand-alone PBR solver to predict thermal degradation of bench-scale experiments is presented, followed by a discussion for representative cases of particles that feature complete consumption (i.e. through smoldering) and particles that feature volume-change (i.e., shrinking or swelling). Chapter 4 presents numerical studies of flame structure and thermal feedback in two canonical pool fire configurations. This chapter first presents LES results from a medium-scale pool fire that is fueled by a non-sooting ethylene fuel, followed by LES results from a turbulent buoyant flame produced from a circular burner (called the FM-burner) that is fueled by sooting ethylene fuel under controlled coflow conditions. In chapter 5, a numerical study of the response of the biomass vegetation particles to external heating conditions imposed at their exposed surface is presented. The chapter discusses first the response of a selection of particles, including thin, thick, charring, and/or non-charring particles, in terms of ignition and mass loss rate due to pyrolysis to either steady or oscillatory external gas conditions. Then the chapter presents 2-D maps that characterize the thermal degradation of porous flaming and smoldering particles in terms of the degree of completion of the drying, the pyrolysis, and the char oxidation processes as a function of the external thermal load and the ambient wind. Chapter 6 focuses on LES simulation results of fire spread in engineered fuel beds made of cardboard sticks that have been studied experimentally at the Missoula Fire Sciences Laboratory. The chapter first discusses the model predictions of the fire behavior against experimental observations of two burns corresponding to a 20 buoyancy-dominated or a wind-dominated fire regime. The rest of the chapter focuses on fire diagnostics related to the flame residence time and the rate of spread and their variation with the packing of the fuel bed. Chapter 7 presents LES simulation results of fire spread in idealized fuel beds comprised of vertically-oriented cylindrical-shaped pinewood sticks under two wind speeds corresponding to either a buoyancy-dominated fire or a wind-dominated fire. The chapter discusses first the fire propagation in uniform fuel beds comprised of small particles that are on the order of 1 mm-scale. Then, results from fire spread simulations in fuel beds that feature a patch with larger particles (on the order of 1 cm-scale) that are added to the smaller ones. The last part of the chapter focuses on an interpretation of the LES results regarding the degree of completion of thermal degradation of the fuel bed using the 2-D maps constructed in chapter 5. Chapter 8 presents a summary of the work conducted in this dissertation and provides some concluding remarks. The chapter also discusses recommendations and suggestions for a continuation of this work in the future. 21 2 The Computational Modeling Framework 2.1 Overview We model the fuel bed as a population of porous particles with different arrangements and with either unique or mixed sizes and geometries. Each porous particle is described as a system with a solid-phase and a gas-phase, which allows for a detailed treatment of the particle-to-external-gas outflow of volatile mass and the external-gas-to-particle diffusion of oxygen mass; this detailed treatment is required to account for in-depth oxidative pyrolysis and char oxidation. We consider here thermally-thick and composition-thick particles featuring in-depth variations of temperature and composition. The particle’s geometry can be modeled as rectangular (leaves or sticks), cylindrical (needles or stems), or spherical (embers or firebrands). This chapter discusses the development and implementation of a Lagrangian Particle Burning Rate (PBR) solver to describe the thermal degradation of individual solid fuel particles, and a coupled multiphase solver to describe the interaction between a solid vegetation fuel bed and the surrounding gases in simulations of wildland fire spread at flame scale. In terms of exchange of information, the PBR solver uses estimates of the 22 gas-to-solid heat flux onto a population of particles as input quantities and produces in return estimates of the particle mass, size, and energy as well as the production rate of combustible gaseous volatiles and other gaseous products inside the fuel bed as output quantities to the coupled multiphase solver. More precisely, the input quantities to the PBR model are: q̇′′g→s = h(Tg,∞ − Tp,surf ) + ϵp,surf (G− σT 4 p,surf ) (2.1) and ṁ′′ g→s,O2 = hmass(Yg,O2,∞ − Yg,O2,surf ) (2.2) where h is the convective heat transfer coefficient, Tg,∞ the external gas temperature in the vicinity of the particle under consideration, Tp,surf the temperature of the exposed surface of that particle, ϵp,surf the particle surface emissivity (treated as an opaque solid), G the averaged radiation heat flux incident on the particle external surface (G designates the irradiation due to distant hot sources, e.g., the flame, the plume, other particles, etc., as well as the irradiation due to ambient gas), σ the Stefan-Boltzmann constant, hmass is the convective mass transfer coefficient hmass = (h/c̄p,g) with c̄p,g the heat capacity of the gas at constant pressure, where Yg,O2,∞ is the mass fraction of oxygen in the external gas in the vicinity of the particle, and where Yg,O2,surf is the mass fraction of oxygen gas at the exposed surface of the particles. In the expression for q̇′′g→s, the input quantities are the external gas temperature Tg,∞, the convective heat transfer coefficient h (treated through Nusselt-number- correlations as a function of the external flow velocity and temperature, ug,∞ and 23 Tg,∞ and the irradiation G. In the expression for ṁ′′ g→s,O2 , the input quantity is the external oxygen mass fraction in the gas in the vicinity of the particle Yg,O2,∞. These quantities are functions of spatial location inside the fuel bed, i.e., presumably functions of both the distance along the direction of fire spread and the vertical elevation. We refer to these quantities as "external gas conditions" and we describe their variations using the coupled multiphase formulation described in section 2.3. 2.2 Particle Burning Rate (PBR) Model Thermal degradation of individual solid fuel particles exposed to the external gas conditions is tracked individually in space and time following a Lagrangian viewpoint. In a one-way coupled formulation, the external gas conditions are imposed, and the model calculates the time-dependent response of the fuel particle to the heat loading from the gaseous environment. In a two-way coupled formulation, the external gas conditions (i.e., composition, velocity, temperature, and radiation) are updated with terms representing the effects of the fuel particles. The description of the fuel particle model presented in this section applies to both one-way and two-way coupled formulations. We follow the work of Lautenberger and Fernandez-Pello [44, 50] by describing the fuel particle as a matrix of pores featuring a solid-phase and a gas-phase. We consider each porous particle to experience four heterogeneous (gas/solid) reactions: a drying reaction (Rd); a thermal pyrolysis reaction (Rp); an oxidative pyrolysis reaction (Rop); and a char oxidation reaction (Rco): 24 [wet solid] −→ ηH2O,Rd [H2O] + ηds,Rd [dry solid] (Rd) [dry solid] −→ ηv,Rp [fuel1] + ηc,Rp [char] (Rp) [dry solid] + ηO2,Rop O2 −→ ηv,Rop [fuel2] + ηc,Rop [char] (Rop) [char] + ηO2,Rco O2 −→ ηCO2,Rco CO2 + ηa,Rco [ash] (Rco) where ηH2O,Rd and ηds,Rd are the mass yields of water vapor and dry solid in reaction Rd, respectively, ηH2O,Rd + ηds,Rd = 1, where ηv,Rp and ηc,Rp are the mass yields of volatile and char in reaction Rp, respectively, ηv,Rp+ηc,Rp = 1, where ηO2,Rop is the oxygen-to-dry-solid mass ratio, and ηv,Rop and ηc,Rop are the mass yields of volatile and char in reaction Rop, respectively, ηv,Rop + ηc,Rop = 1 + ηO2,Rop, and where ηO2,Rco is the oxygen-to-char mass ratio, and ηCO2,Rco and ηa,Rco are the mass yields of CO2 and ash in reaction Rco, respectively, ηCO2,Rco + ηa,Rco = 1 + ηO2,Rco. Note that these reactions are written per unit kg of the reactant. Consistent with the model proposed in Ref. [44], we do not differentiate between the char products of the thermal and oxidative reactions Rp and Rop. While the gaseous fuel produced from Rp and Rop may have different chemi