LIBRARY Michigan State University PLACE iN RETURN BOX to remove We checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE WI, Ii “—7! MSU is An Affirmative ActioniEquai Opportunity Institution emails-pt THEORETICAL AND EXPERIMENTAL STUDY OF WIND—AIDED FLAME SPREAD OVER VAPORIZING SOLIDS By SANJAY AGRAWAL A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 1992 g.) u, o / n)?“ 3.x / >\ ABSTRACT THEORETICAL AND EXPERIMENTAL STUDY OF WIND-AIDED FLAME SPREAD OVER VAPORIZING SOLIDS By SAN J AY AGRAWAL This study presents a theoretical and experimental investigation of wind-aided flame spread over vaporizing solids. Effects of free stream velocity and oxygen mass frac- tion, solid-phase unsteadiness, blowing and surface radiative heat loss on pyrolysis and flame front speed (location) are investigated. Investigations of the thermal degradation of vaporizing solids and the stability of wind-aided flame spread across a ceiling are also performed. Pyrolysis and flame from location and production rates of major chemical species were measured as functions of time for flames spreading over PMMA samples for various coflow speeds and oxygen mass fractions. The surface heat flux distribution in the burning zone is determined from burned samples using a regression model. Mass loss rate and in-depth temperature measurements were obtained during pyrolysis of PMMA samples in nitrogen environments for various heat fluxes. Explicit expressions are deduced for the pyrolysis front speed, flame tip location, surface regression rate, mass loss rate and surface heat flux under the flame. Numeri- cal results are correlated to provide expressions for the surface heat flux in the preheat zone and the excess pyrolyzate. The pyrolysis front speed varies linearly with both free stream velocity and oxy- gen mass fraction, comparing favorably with experiments. The solid is always transient during the flame spread process, as confirmed by degradation experiments. Inclusion of surface radiative heat loss is necessary to produce a phenomena observed during the experiments, flame spread termination before the end of the sample. Using burned samples and a model it was determined that the net surface heat flux in the burning zone decreases with the streamwise distance, and increases with the free stream velocity. The influence of the gas-phase reaction rate and solid-phase cooling is lim- ited to the first 3 cm of the sample. This, coupled with surface radiative heat losses and solid-phase unsteadiness are the main reasons for the shorter experimental flame overhang. An inviscid stability analysis suggests that in the ceiling configuration the flame is stable because of forced convection. For small Froude numbers it is necessary to include finite reaction rates in the analysis. Copyright by SANJAY AGRAWAL 1992 To my family ACKNOWLEDGEMENTS I am grateful to my advisor Dr. Indrek Wichman, without whose support and gui- dance this completed work would not have been possible. I am especially grateful to him for providing financial support, encouragement and the means to attend numerous professional meetings through the course of my doctoral program. Dr. Arvind Atreya has also been instrumental in helping me achieve this goal. He provided great moral and professional support. I am especially grateful to him for his help in the experimental portion of my research and for letting me use his lab. The long nightly discussions with him will be fondly remembered. I am thankful to Drs. John Lloyd, Manoochehr Koochesfahani and Milan Mik- lavcic, who formed the rest of my committee, for carefully reading my work and giv- ing useful suggestions. The use of G-splines, suggested by Dr. Miklavcic is greatly appreciated. I am thankful to my colleagues and friends, especially Dr. Kamel Mekki, for his help in conducting flame spread experiments. I dedicate this thesis to my family: my parents - who have been my strongest source of inspiration kept me constantly motivated and provided invaluable support through regular long-distance phone-calls and letters; my sister and brother-in-law, brother and Iother family members for their abiding confidence in my abilities; and my wife - Madhu, for her constant love, understanding, companionship and support. Funding for this research was provided by U.S Department of Agriculture under contract numbers 86-FSTY-9-90192 and 90-37291-5768. Equipment was funded by NSF under the grant number CBT-8415423. vi TABLE OF CONTENTS List of Tables ............................................................................................................. List of Figures ............................................................................................................ Nomenclature .............................................................................................................. Chapter I: INTRODUCTION .................................................................................... Synopsis .......................................................................................................... 1.1. Introduction ............................................................................................. 1.2. Wind-Aided Flame Spread ..................................................................... 1.2.1. Statement of the Problem ........................................................ 1.2.2. Literature Review ..................................................................... 1.2.3. Theoretical Framework ............................................................ 1.3. Objectives of the Present Work ............................................................. 1.4. Organization of the Thesis ..................................................................... Chapter 2: A PRELIMINARY MODEL OF FLAME SPREAD ............................. Synopsis .......................................................................................................... 2.1. Introduction .......... '. .................................................................................. 2.2. Problem Formulation ......................................... 2.3. Solution ................................................................................................... 2.3.1. Pyrolysis Zone (0 < g < ép) ................................................... 2.3.2. Preheat Zone (ép < 5 < 1) ...................................................... 2.4. Results ..................................................................................................... 2.4.1. Pyrolysis Front Speed .............................................................. 2.4.2. Flame Front Speed ................................................................... 2.4.3. Flame Location ........................................................................ 2.5. Discussion ............................................................................................... 2.6. Conclusions ............................................................................................. Appendix 2.A: Green’s Function Solution of Gas-Phase Preheat Zone Equation ................................................................................. vii xi xii xvi 15 15 16 18 27 31 33 35 37 37 38 39 49 51 Appendix 23: Laplace Transform Solution of Gas-Phase Preheat Zone Equation ....................................................................... 53 Appendix 2.C: Green’s Function Solution of Solid—Phase Preheat Zone Equation ....................................................................... 54 Appendix 2.D: Evaluation of the Integral I (A ,B) ........................................ 55 Chapter 3: HEAT FLUX DISTRIBUTIONS ............................................................ 57 Synopsis .......................................................................................................... 57 3.1. Introduction ............................................................................................. 58 3.2. The Model ............................................................................................... 59 3.3. Solutions for Surface Heat Flux and Surface Temperature .................. 65 3.3.1. Heat Flux ....................... 65 3.3.2. SurfaceTemperature ................................................................ 70 3.4. Solutions for Normalized Fluxes and Overall Fluxes ........................... 72 3.4.1. Excess Pyrolyzate .................................................................... 81 3.5. Conclusions ............................................................................................. 86 Chapter 4: AN EXPERIMENTAL INVESTIGATION ............................................ 90 Synopsis .......................................................................................................... 90 4.1. Introduction ............................................................................................. 91 4.2. Experimental Setup ................................................................................. 92 4.2.1. Combustion Wind Tunnel ....................................................... 92 4.2.1.1. Inlet Section .............................................................. 92 4.2.1.2. Turbulence Manipulation Section ............................ 94 4.2.1.3. Test Section ............................................................... 95 4.2.1.4. Exhaust Section ......................................................... 96 4.2.2. Gas Analysis Equipment ......................................................... 96 4.2.2.1. Total Hydrocarbon Analyzer .................................... 98 4.2.2.2. Water Analyzer ......................................................... 98 4.2.2.3. CO-C02 Analyzer .................................................... 98 4.2.2.4. Oxygen Analyzer ...................................................... 98 4.2.3. Data Acquisition ...................................................................... 99 4.2.4. Sample Preparation .................................................................. 99 4.2.5. Experimental Procedure ........................................................... 99 4.3. Results ..................................................................................................... 99 4.3.1. Temperature Measurements ..................................................... 101 4.3.2. Flame Spread Rates ............... , .................................................. 103 4.3.3. Species Production Rates ....................................................... .. 108 viii 4.4. Discussion ..................................................................................... 111 4.5. Conclusions ............................................................................................. 112 Chapter 5: THERMAL DEGRADATION OF A VAPORIZING SOLID .............. 114 Synopsis .......................................................................................................... 1 14 5.1. Introduction ............................................................................................. 114 5.2. Goveming Equations .............................................................................. 117 5.3. Solution ................................................................................................... 120 5.3.1. Preheating Stage ....................................................................... 121 5.3.2. Vaporizing Stage ...................................................................... 121 5.3.2.1. In-depth Temperatures Profiles at the Vaporization Temperature ....................................... 123 5.3.2.2. Transient Mass Loss Rate ........................................ 123 5.4. Experimental Setup .................................................................... . ............ 125 5.4.1. Effect of Sample Thickness and In-depth Absorption 131 5.5. Comparison with the Literature .............................................................. 133 Chapter 6: FLAME SPREAD OVER AN UNSTEADILY VAPORIZING SOLID ...................................................................................................... 137 Synopsis .......................................................................................................... 137 6.1. Introduction ............................................................................................. 138 6.2. Model Problem ........................................................................................ 139 6.3. Solution ................................................................................................... 143 6.3.1. Surface Heat Flux under the Flame ........... 143 6.3.2. Mass Flux in the Pyrolysis Zone ............................................ 146 6.3.3. Flame Front Location .............................................................. 147 6.4. Results and Discussion ........................................................................... 148 6.4.1. Effect of Blowing .................................................................... 151 6.4.2. Effect of Radiant Losses from the Surface ............................. 155 6.4.3. Influence of Unsteady Pyrolysis .............................................. 156 6.6. Conclusions ............................................................................................. 158 Appendix 6.A: Blowing Correction ............................................................... 159 Appendix 68: Turbulent Wind-Aided Flame Spread .................................. 162 Chapter 7: SURFACE HEAT FLUX IN THE BURNING ZONE .......................... 164' Synopsis .......................................................................................................... 164 ix 7.1. IntroduCtion ............................................................................................. 165 7.2. Net Surface Heat Flux in the Burning Zone ......................................... 167 7.3. Regressed Surface Location ................................................................... 148 7.4. Pyrolysis Experiments: Determination of so ......................................... 170 7.5. Flame Spread Experiments: Determination of to .................................. 171 7.6. Results and Discussion ........................................................................... 173 7.6. Conclusions ............................................................................................. 176 Chapter 8: STABILITY ANALYSIS ........................................................................ 177 Synopsis .......................................................................................................... 177 8.1. Introduction ............................................................................................. 178 8.2. Steady Flame Spread .............................................................................. 181 8.2.1. Basic-State Velocity Profile ..................................................... 185 8.2.2. Basic-State Temperature Profile .............................................. 185 8.3. Stability Analysis .................................................................................... 193 8.4. Results and Discussion ........................................................................... 195 8.5. Conclusions ............................................................................................. 205 Appendix 8.A: Reaction Zone Solution for Temperature ............................ 208 Appendix 83: Asymptotic Approximation for T, as n—mf ....................... 210 Appendix 8.C: Analysis of the Disturbance Equation .................................. 211 Chapter 9: CONCLUSIONS ........................................ , ............................................. 214 Synopsis .......................................................................................................... 214 9.1. Conclusions ............................................................................................. 215 9.2. Suggestions for Future Work ................................................................. 220 LIST OF REFERENCES ........................................................................................... 221 LIST OF TABLES Table 3.1 Ratios of the heat released by the flame, or received by the surface in each zone, to the total heat released, as a function of Y0... Here tire is the excess pyrolyzate, "'2’, is the total mass lost from the solid, 9" and gm are the total flame heat release in the pyrolysis zone and the preheat region, respectively, and g-py, Em: and E” are the total surface heat flux in the pyrolysis zone, the under-flame preheat zone and the preflame preheat zone, respectively. The quantity 3707 is the total heat released by the flame. ........................................................................................ 85 Table 6.1 Property values ..................................................................................... 154 Table 8.1 Table of parameters used. Parameters a -b and QYO .Jv are obtained from Eq. (8.A.2) and (8.17) respeCtively ....................... 192 Table 8.2 Values of the nondimensional activation energy [3 used for the different B numbers ................................................................. 192 xi Figure 1.1 Figure 1.2 Figure 1.3 Figure 2.1 Figure 2.2 Figure 2.3 Figure 2.4 Figure 2.5 Figure 2.6 Figure 2.7 Figure 2.8 Figure 3.1 Figure 3.2 Figure 3.3a Figure 3.3b LIST OF FIGURES Modes of flame spread ......................................................................... Schematic diagram of wind-aided flame spread .................................. Thesis flow chart .................................... Schematic diagram of wind-aided flame spread; Pyrolysis zone extends till xp (1 ) and flame reattaches at xf(t ). Superscript * signifies a dimensional quantity ........................................................... Equations and boundary conditions in the pyrolysis zone .................. Equations and boundary conditions in the preheat zone .................... Variation of mass flux parameter (M) and surface fuel mass fraction (Yps) as a function of QYO .JL .............................................. Variation of pyrolysis front speed (5p) and flame front speed (fif) as a function of free stream oxygen mass fraction (Y0 -) .......... Spread rates versus mass flux parameter. Here é is the non- dimensional pyrolysis front speed, Q is the nondimensional flame front speed, HP is the approximate (zero-M assumption) nondimensional pyrolysis front speed and if is the approximate (zero-M ) nondimensional flame front 5 eed. The reason the curves do not coincide is that E and f are evaluated for M = 0 whereas in the present eory the minimum value of M is 0.0636 ................................................................................................ Ratio of spread rates (ép/ E”) versus mass flux parameter M ........... Flame locations at different nondimensional time instants. The x and z coordinates are nondimensional Howarth coordinates .............. Model configuration .............................................................................. Plot of h(§,§p ), given by Eq. (3.3) et seq, versus é. Observe that h is piecewise continuous at if ................................................... Plot of the normalized nondimensional heat flux gT(§) versus 5, obtained from the solution of Eq, (3.4) using G-splines. The stepsize A: is increased by the factor 10 at 3Q. The maximum of the curve represents the location 9 ............................................... Plot of the nondimensional heat flux, g :00; /2)gT/‘5c_ , versus é. Here g (Q) is always less than g (ép), even though g (Q) is always greater than gfifip) .................................................................. xii 3 6 13 19 23 24 4o 41 43 44 46 6O 64 68 69 Figure 3.4 Figure 3.5a Figure 3.5b Figure 3.6a Figure 3.6b Figure 3.7 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Figure 4.6 Figure 4.7 Figure 4.8 Figure 4.9 Figure 5.1 Figure 5.2 Plot of the reduced temperature I: , given by Eq. (3.10), versus E. The first kink (from left) in the graph (for Y0..=1) represents the location of the flame front. The second and third kinks (numerical error) occur because of step size changes that make the computation faster. This is also seen in Fig. 3.3a for Y0..=1 OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO Heat flux correlation for the under-flame preheat zone, showing the rise of the heat flux as the downstream flame tip (reattach- ment point) is approached .................................................................... Heat flux correlation for the pre-flame preheat zone. Note how rapidly the decay occurs ....................................................................... Heat flux of Figure 3. 3a for fixed x (xo=1) versus time t ................. Surface temperature of Figure 3.4 for fixed x (x0 =1) versus time t. Plot of the ratio gig/g};- given by Eq. (3.13). Also shown is the graph of 5/90! (dotted line), for comparison. In this figure the parameters are for PMMA with a wind speed of 1 m/s . In our model, because of the Oseen flow assumption, the wind speed appears as a non-dimensional quantity, and hence, will not change the above ratio. Schematic of the experimental apparatus ............................................ Schematic of the gas analysis equipment ............................................ Schematic of wind-aided flame spread in the tunnel over a ceiling-mounted sample of PMMA ...................................................... Measured surface temperatures and their rate of change with time for PMMA during flame spread (U .. = 0.9 m/s , Y0... = 0.331). Here surface temperatures are plotted for different streamwise locations on the sample ...................................... Flame front location as a function of time. Note the slight acceleration for xf 2.40 cm. Also note the negligible error [(O(mm)], encountered in determination of flame front location using VCR ............................................................................................ Pyrolysis-front location as a function of time ..................................... Dependence of the pyrolysis and flame front speeds on the free stream velocity. ...................................................................................... Dependence of pyrolysis and flame front speeds on the free stream oxygen mass fraction (blank symbols represent Up and crossed symbols represent U f) ............................................................ Dependence of the species production rates on pyrolysis front- location (U, = 0.9 m/s; 1 - Y0, = 0.23; 2 — Y0“ = 0.43; 3 — Y0... = 1.0) ...................................................................................... Schematic of the solid phase model .................................................... Surface temperature and total mass loss as a function of time for PMMA. Heat flux = 2.1 W/cm2 OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO ................................................................. OOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOOO xiii 73 74 77 93 97 100 102 104 105 106 107 109 118 Figure 5.3 Figure 5.4 Figure 5.5 Figure 5.6 Figure 5.7 Figure 5.8 Figure 5.9 Figure 5.10 Figure 5.11 Figure 5.12 Figure 5.13 Figure 6.1 Figure 6.2 Figure 6.3 Figure 6.4 Figure 6.5 Figure 6.6 Figure 7.1 Figure 7.2 Figure 7.3 Figure 7.4 Figure 7.5 Figure 8.1a Figure 8.1b Figure 8.2 Figure 8.3a Surface temperatures as a function of time, during preheating of the solid ................................................................................................. In-depth temperatures as a function of time, during preheating of the solid ............................................................................................ Mass loss rate as a function of time .................................................... ' Varying external heat flux .................................................................... - Mass loss rate as a function of time under varying heat flux ............ Effect of sample thickness on mass loss rate ...................................... Schematic of the apparatus .................................................................. In-depth temperatures as a function of time, during vaporization of the solid under q =3. 4W /cm Mass loss rate as a function of time. Comparison of integral model with the data of Vovelle et a1. (1987) ...................................... Mass loss rate as a function of time. Comparison of numerical model with the data of Vovelle et al. (1984b) .................................... Mass loss rate as a function of time. Comparison of integral model with the data of Vovelle et al. (1984a) .................................... Schematic of the model ........................................................................ Dependence of total oxygen depletion (total fuel production x v) on the pyrolysis front location. Symbols represent experi- mental results. 1—Y0”=0.233; Z-Yo..=0-43; 3—Yo,,=0.61 .................. Blowing correction ............................................................................... Comparison between the convective heat flux and the net heat flux ........................................................................................................ Dependence of oxygen depletion rate on the pyrolysis front location (Parametric dependence). Yo ..=O.233. Symbols represent experimental results. Solid curve represents the model predictions. Calculated Curves (1a) to (1e) are for different conditions as explained in Secs. 6.4.1-643 ........................................ Effect of blowing correction with transient mass flux on total oxygen depletion rate. Solid curve represents the model predictions. Calculated curves (1a), (1b) and (1f) are for different blowing corrections as explained in Appendix 6.A ............. Schematic of the gas-phase model ....................................................... Schematic of the solid-phase model .................................................... Regressed surface location as a function of time ................................ Net surface heat flux in the burning zone ........................................... Net surface heat flux near the leading edge ........................................ Physical configuration of the wind-aided flame spread ...................... Schematic diagram of the model ......................................................... Basic- state temperature and velocity profiles. Note that T—profile plotted here is for the infinite-reaction-rate ......................... Inner zone (reaction region) temperature profiles, .8 =0.275. Curves 1-6 represent profiles for different [3’s .................................... ......................................................... xiv 122 124 126 127 128 129 130 134 135 136 140 150 152 153 154 161 166 169 172 174 175 179 166 186 189 Figure 8.3b Figure 8.3c Figure 8.4 Figure 8.5a Figure 8.5b Figure 8.5c Figure 8.6a Figure 8.6b Figure 8.6c Figure 8.7 Figure 9.1 Inner zone (reaction region) temperature profiles, B =1.125. Curves 1-6 represent profiles for different B’s .................................... 190 Inner zone (reaction region) temperature profiles, B =1.6. Curves 1-6 represent profiles for different B’s .................................... 191 Froude number F r0 as a function of free stream velocity and streamwise distance ....................................................................... 196 Stability curve; non—dimensional frequency versus Froude number, B =0.275. Curves 1-6 are for different inner tempera— ture profiles of Fig. 8.3a ....................................................................... 199 Stability curve; non-dimensional frequency versus Froude number, B =1.125. Curves 1-8 are for different inner tempera- ture profiles of Fig. 8.3b ...................................................................... 200 Stability curve; non-dimensional frequency versus Froude number, B =1.6. Curves 1-6 are for different inner temperature profiles of Fig. 8.3c .............................................................................. 201 Frequency curve; neutral frequency versus streamwise distance, B =0.275. Curves 1-6 are for different inner temperatru'e profiles of Fig. 8.3a .............................................................................. 202 Frequency curve; neutral frequency versus streamwise distance, B =1.125. Curves 1-8 are for different inner temperature profiles of Fig. 8.3b .............................................................................. 203 Frequency curve; neutral frequency versus streamwise distance, B =1.6. Curves 1-6 are for different inner temperature profiles of Fig. 8.3c ............................................................................................ 204 Variation of frequency (0)0) with —Im(ao ). Curves plotted are for temperature profile #5 for F r0 equal to 2 and 4; B =1 . 125. .............................................................................................. 206 Variation of kpCp as a function of temperature ................................ 219 XV NOMENCLATURE CHAPTERS 2-3 mmmgwmenmowb 3“) 3? v N «sax g < q: H”2JI>UIQ$§5-5?F‘ml Pre-exponential factor Mass transfer number Specific heat Specific heat ratio (gas to solid) Coefficient matrix (Eq. 3.5) Mass diffusion coefficient Activation energy Heat flux to the interface Normalized heat flux Total heat flux (Eq. 3.13) Total heat released by flame Transformed heat flux defined before Eq. (3.4) Function defined as h (x )= ‘11—: xexp (x 2)erfc (-x) Variable defined in Eq. (3.4) Latent heat of vaporization; reference lengths in gas and solid Lewis number, Le =0t/D Mass flux from solid to gas (dimensional) Mass flux parameter Prandtl number - Heat of combustion per unit mass of fuel vapor Thermal responsivity ratio (gas to solid); reaction term Universal gas constant Time Temperature Streamwise velocity; variable defined after Eq. (3.11) Streamwise velocity (free stream, pyrolysis front, flame front) Transverse velocity Transverse mass flux Streamwise coordinate Transverse coordinate Mass fraction Mass based transverse coordinate xvi O -t 8 =- d a eNDW §- E“ Variable defined after Eq. (5.7) Mass transfer number, Eq. (6.19) Specific heat Heat of combustion per unit mass of fuel vapor Blowing correction Convective heat transfer coefficient Thermal conductivity Heat of gasification Lewis number Mass loss rate; Mass flux Nondimensional mass loss Prandtl number xvii Heat flux Reynolds number Regresged surface location Time, t=t-tv defined after Eq. (5.6) Temperature, T=T -T,, Streamwise velocity Streamwise velocity (free stream, pyrolysis front, flame front) Transverse velocity Sample width Streamwise coordinate Transverse coordinate Mass fiaction 0 ~<~<><§*€-‘.:3.mm-<'cag Acceleration due to gravity Reference length in gas Heat of combustion per unit mass of fuel vapor Universal gas constant Reaction term Time Temperature Streamwise velocity Free stream velocity Transverse velocity Transverse mass flux Streamwise coordinate Transverse coordinate Mass fraction Mass based transverse coordinate Wave number Nondimensional activation energy Coupling function Disturbance amplitude Modified inner variable for 11 Similarity variable in gas Thermal conductivity Dynamic viscosity Kinematic viscosity; mass based stoichiometric coefficient Density Stream function Frequency Vorticity Disturbance stream function Modified inner variable for temperature Subscripts *8 $9100: “:1“. Flame; inflection point Fuel vapor Gas Oxygen; modified variable for a, B and Fr Solid; surface Solid surface (wall) Free stream Inflection point Superscripts A * Modified variable Eq. (8.26) Disturbance variable Eq. (8.26) Dimensional quantity xix CHAPTER 1 INTRODUCTION SYNOPSIS In this chapter the problem is introduced and the relevant literature is reviewed. The objectives and organization of the thesis are discussed. 1.1. INTRODUCTION The present study involves investigation of wind-aided flame spread across a vaporizing solid in the ceiling. configuration. A typical practical scenario is a corridor ceiling fire. The threat of uncontrolled fire has been a formidable problem to mankind for centuries, yet a complete understanding of the factors affecting its growth has not been achieved. The investigation into flame spread over solids is motivated by the need to prevent and control fire growth in 5 buildings and other confined spaces. Spread of the flame results from the complex interaction of transport processes in the gas and the condensed phases, the vaporization of the fuel and the chemical reaction of the fuel vapors with the gaseous oxidizer. The spread can occur in horizontal, vertical or any other flow configuration (Fernandez-Pello and Mao, 1981); can be assisted or opposed by the natural convec- tion (wall fires) or forced convection (ceiling or floor fires) or both (Fig. 1.1). Natural convection is generated by the density stratification in the vicinity of the flame. Forced convection occurs because of the wind (corridor flow). In flame spread processes where the oxidizing gas flows in the direction of flame propagation, the environment is said to "favor" the propagation of the flame. As a result the spread process is usually more rapid than other modes (and consequently more hazardous) and has been proven to be extremely difficult to understand and quan- tify (Fernandez-Pello and I-Iirano, 1983; Fernandez-Pello, 1984). This kind of spread can be further classified as buoyancy-controlled upward flame spread (wall fires) or wind-aided flame spread (ceiling or floor fires). The interest that this mode of flame propagation represents for the fire-safety field has motivated the development of the ASTM E-84 (Anonymous, 1977) tunnel test of flammability. This test, which simu- lates wind-aided flame spread across a ceiling (Fig. 1.1) in a corridor, uses a forced flow of hot oxidizing gas to enhance the rate of flame spread, thus subjecting the N N —+ <— —> -—> —> —> —> —> —> —> Opposed Convection (ceiling mode) Forced Convection (ceiling mode) _> —> __..> —> _’ —> —> <— —> —> .x\ R Opposed Convecrion (floor mode) Forced Covection (floor mode) N , i \\‘ Natural Convection Natural Convection OPPOSED FLOW FLAME SPREAD WIND-AIDED FLAME SPREAD Figure 1.1 Modes of flame spread flammability properties of the material to stringent conditions. The ten minute test principally involves simple observation of flame spread rate and how far flame con- verge (Parker, 1977). The first attempt to develop a comprehensive model of wind-aided flame spread was made by Carrier et al. (1980). In fact this model was developed with the aim of describing the ASTM E—84 tunnel test. The complicated analytical techniques needed to reduce the problem to a manageable computational form suffice to demonstrate the difficulty of the problem. On the experimental side there were few studies because of the relative hazard.» This prevented the experimental study for long time (Femandez— Pello and Hirano, 1983; Femandez-Pello, 1984). Largely because of unavailability of experimental information, the development of theoretical models for flame spread in forced concurrent flow started only in the last decade (Femandez-Pello, 1979; Annamalai and Sibulkin, 1979b; Carrier et al., 1980, 1983, DiBlasi et al., 1987, 1988a. 1988b). We note, however, that the development of theoretical models for buoyancy- controlled upward spread began in early seventies (Markstein and deRis, 1972, 1973; Orloff et al., 1974; Sibulkin and Kim, 1977; Fernandez-Pello 1978; Annamalai and Sibulkin 1979b; Fernandez-Pello and Quintiere, 1982; Hasemi, 1986; Saito et al., 1986; Kulkarni and Fisher, 1988). The theoretical model development for opposed flow flame spread began even before this (deRis, 1969; Sirignano, 1974, Fernandez-Pello and Williams, 1975; Wichman and Williams, 1983). Because of the objectives of the present study and limited literature, the remainder of this section is confined to wind- aided flame spread. Several reviews exist which discuss various other modes and scales of flame (or fire) spread (Friedman 1968; Magee and McAlevy 1971; Sirignano, 1972; Williams, 1976; Hirano and Akita, 1979; Emmons, 1980; Femandez-Pello and Hirano, 1983; Fernandez-Pello, 1984). 1.2. WIND-AIDED FLAME SPREAD 1.2.1. STATEMENT OF THE PROBLEM For the flame to spread the solid ahead of the flame should pyrolyze to provide fuel to react with the oxidizer. The hot still-reacting post-combustion gases move ahead of the pyrolysis zone and enhance the rate of heat transfer to the virgin solid. Heat transfer is also augmented by the radiation from the flame. This results in a tem- perature rise at the surface. ‘The fuel produced in the pyrolysis zone is not completely consumed in the bunting zone. The excess pyrolyzate move ahead of the pyrolysis front (though still reacting with the oxidizer) and extends the diffusion flame. Thus we have a flame front (flame tip, xf) moving ahead, followed by a pyrolysis front (xp) (Fig. 1.2). Trailing both of them is a burnout front which signifies unavailability of the fuel; this could occur because of charting or total fuel consumption (vaporizing or non-charting solid). For the typical downstream fuel element the series of events is initial heating, pyrolysis, ignition, flame spread and extinction. A systematic study is performed for a fire that is spreading in this mode over a continuous solid fuel slab. Specific objectives of the present work are discussed later in Sec. 1.3. 1.2.2. LITERATURE REVIEW All of the theoretical studies of wind-aided flame spread assume the flow is lam- inar (Carrier et al., 1980, 1983; Fernandez-Pello, 1979; DiBlasi et al., 1987, 1988a, 1988b). Most of the models (except DiBlasi et al., 1987, 1988a, 1988b) can be classified as thermal models. In these models the rate of flame spread is determined by the rate at which the surface temperature of the combustible is raised to a . prescribed value (vaporization temperature), through heat transfer from the flame and external heat sources. The combustible is assumed to evaporate as soon as it reaches PMMA sample ‘ Gravity l Figure 1.2 Schematic diagram of wind-aided flame spread this prescribed temperature. In all of these models a one step infinitely fast gas-phase reaction (va+v00—> products) is assumed.“ DiBlasi et al., (1987, 1988a, 1988b) assumed a zeroth order Arrhenius type solid fuel pyrolysis rate. They also took into account the finite-rate gas-phase chemical kinetics (one-step second-order irreversible Arrhenius reaction). It was found that effect of finite-rate chemistry is limited to the flame tip and the upstream leading edge, and is prominent for reduced oxygen concen- trations (e.g. the extinction limit). They also found that for wind-aided flame spread streamwise diffusion is negligible, and the gas and solid-phases can be accurately represented by boundary layer equations. The experimental study of Loh and Fernandez-Pello (1984) shows the unimportance of finite-rate chemical kinetics. The possibility of developing a theoretical model of wind-aided flame spread that considers only heat transfer processes represents a great simplification; it is not clear that this will always be a good approximation. The momentum equation introduces strong nonlinearities into the problem. Gen- erally, in all theoretical models developed, the energy equation is decoupled from the momentum equation by assuming a uniform flow velocity (Oseen approximation) (Car- rier et al., 1980, 1983) or a parabolic channel-flow profile (given by Hagen-Poiseuille formula) (DiBlasi et al., 1987, 1988a, 1988b). Carrier et al. (1983) employed the boundary-layer approximation, but their results were not much different from those obtained by using the Oseen approximation. We observe here that there have been no systematic studies to assess the validity and limitations of the Oseen approximation. Despite its relatively common use, its origin remain unexplored: recent work in micro-gravity (rig) research is changing this, however. The flame length depends on the total amount of fuel produced. In all of the models (except Carrier et al., 1980) the solid phase in the pyrolysis zone is assumed steady. Although DiBlasi et al. (1987) assumed the solid as transient, but their assumption of mass loss occurring at the surface with zeroth order Arrhenius-type dependence on surface temperature, made the .mass flux constant since during vapori- zation the surface temperature is relatively constant (Vovelle et al., 1987). Most of the theoretical models predict spread rates that are in qualitative agree- ment with the experimental observations. Based on their experimental results Loh and Fernandez-Pello (1984) produced an expression for flame spread (pyrolysis front speed): 2 _ pg cps Atr Tf'Tv Up _ U” pscsls [TV-Too (1.1) The experiments were done on thick PMMA sheets in the floor configuration. Equa- tion (1.1) shows that for thick non-charring materials, flame spread is linearly depen- dent on flow velocity, proportional to the square power of the oxygen mass fraction and steady. ' Theoretical models of wind-aided flame spread over vaporizing solids predict a linear dependence of spread rate on free stream gas velocity (Fernandez-Pello, 1979; Annamalai and Sibulkin, 1979b; Carrier et al., 1980, 1983). For non-charting materials the flame spread is steady whereas for charting materi- als it accelerates before reaching steady state (DiBlasi et al., 1987, 1988b; Loh and Fernandez-Pello, 1984; Saito et al., 1986). In all of the studies flame (soot) radiation and surface heat losses (reradiaton) are neglected. Carrier et al. (1980) included flame radiation in a very approximate form. The emissivity was assumed constant. Most of the models developed consider the material as non-charting because of difficulty in analyzing solid-phase processes. Non-charring solids vaporize completely without leaving any residue (char), e. g. plastics. Charring makes problem inherently transient. Carrier et al., (1983) treated char and virgin material as solids with different densities. The volatiles in the virgin material vaporize at constant temperature and char erodes at some different constant temperature. Solutions are provided only for steady-state case, although the formulation is transient. DiBlasi et al. (1988b) developed a theoretical model of the spread of a flame over thin charting combustible solids. Tire charting of the solid is modeled by treating the solid as a two-component material; the combustible component that gasifies, and the char that remains occupying the initial solid volume. In the'gas phase, the combustion reaction is modeled with a one-step irreversible second order Arrhenius chemical reaction. The results of the analysis suggest that simplified model of Saito et al. (1986) could predict certain aspects of the flame spread process without much error. In conclusion, based on .the available literature the present understanding of wind-aided flame spread can be expressed as follows: "Wind-aided flame spread can be treated as the spread of a pyrolysis front which varies linearly ,with the free stream velocity and as square of the free stream oxygen mass fraction, and is steady for non-charring materials" 1.2.3. THEORETICAL FRAMEWORK A simple model is presented in this section for wind-aided flame spread that gives us the essential parametric dependencies. Let us assume that the solid is only heated under the flame, i.e, before the flame tip (xf) the surface temperature of the solid is ambient (Tn). This assumption is used by (Sibulkin and Kim, 1977; Fernandez-Pello, 1979; Quintiere, 1981; Loh and Fernandez-Pello, 1984). If the gas is assumed steady and the flame radiation and surface heat losses are neglected, then the net heat flux into the solid is just the convective heat flux qczm. By further assuming that the solid is thick (semi-infinite) the problem is reduced to the solution of the following equation: dl' d2T pscs‘gt' ="- As _dy2 , (1.2) 10 and the boundary conditions are dT -13 [_] =qconv ’ T (°°,t )=T.. Here p, c and 7t. are the density, specific heat and thermal conductivity respectively. Subscripts s , g and co represent conditions in the solid, gas and the ambient, respec- tively. The onset of the pyrolysis front is characterized by the surface reaching its vaporization temperature, Tv . The time required 'to do so, At , can be found easily from Eq. (1.2). Also the pyrolysis front in this time has traveled from xp to x,, thus Up =(xf -Jtp )/At. The location xf has faced 4,2,, for the time interval At. Thus one obtains (Sibulkin and Kim, 1977; Quintiere, 1981 ; Annamalai and Sibulkin, 1979b) U __4_ deliflxfflxf-xp] * p n Mean-T42 (1.3) In the boundary layer the convective heat flux goes as x'l’z. Now if we assume (1,2,, (x )=(pg cm, 1.8 UJPn)la(Tf -T,,) (Loh and Fernandez-Pello, 1984), then Eq. (1.3) becomes 7. T -T 2 Upszpgcpg g( f V) 1 x" (1.4) ps cs As (Tv —Too)2 xf If xf >> xp one obtains Eq. (1.1). The flame length xf primarily depends on the total fuel available, which depends on the heat flux distribution in the pyrolysis zone and 11 the solids’ degradation behavior. Thus, the important parameters to study are the heat flux distribution, Up , Jcp , xf , and the solid degradation behavior. Equation (1.3) is only valid when there is no preheating of the solid upstream of xf. In the next chapter we analyze this in greater detail and formulate the problem with preheating extending all the way to U mt. Note that locations x > U m,t are undis- turbed. 1.3. OBJECTIVES OF THE PRESENT WORK A systematic study is undertaken in order to determine the major variables (fac- tors) that affect wind-aided flame spread. Other than the usual environmental variables like free-stream velocity and oxygen mass fraction, the effects of unsteadiness of the behavior of the solid, heat flux near the leading edge, surface heat loss and stability of the flame on wind-aided flame spread phenomena are also examined. In doing so, explicit expressions for the pyrolysis front speed, estimates for the realistic flame from location and transient mass loss rate are sought. No particular emphasis is placed on the solution approach, i.e., whether experimental or theoretical. This is a debatable philosophical point that has spawned libraries full of weighty expostulations by many wise savants. I would say: "The difficulty of the problem dictates that the method of solution vary according to the particular question being examined; it is not possible to force this complex problem to yield to a certain solution method, or even to a single restricted class of solution methods". The goal is to develop a foundation for the wind-aided flame spread problem that will allow major tasks, like modeling of turbulent wind-aided flame spread over char- ring solids, to be successfully undertaken. 12 1.4. ORGANIZATION OF THE THESIS This thesis is organized into 7 chapters other than Introduction and Conclusion. It is best understood in the context of Fig. 1.3. Chapters 2, 4 and 6 deals with the theoretical description and experimental determination of pyrolysis front movement and flame front speeds and locations. Chapters 3 and 7 deal with the determination of the surface heat fluxes theoretically and experimentally, respectively. Chapter 5 deals exclusively with the solid phase. The transient effects in the solid are analyzed theoretically and experimentally. Chapter 8 deals exclusively with the gas-phase. Here the stability of the flame is analyzed and characteristic pulsation frequency of the flame is determined. The specific objectives and the main conclusions of each chapter are: 1. Based on the need, present understanding of the problem, a theoretical model which assumes a steady pyrolysis zone is developed. This results in an explicit expression for pyrolysis fi'ont speed. The expression suggests linear variation with free stream velocity. It also suggests linear variation with free stream oxygen mass fraction, disagreeing with the experimental results of Loh and Femandez- Pello (1984). Analysis of model 1 is extended to obtain heat flux distributions in the pyrolysis region and preheat region. The numerical results for surface heat flux in the pyrolysis zone and in the preheat zone can be reduced to expressions which depends only on flame front speed and pyrolysis front speed. An expression for excess pyrolyzate is also obtained. The results indicate that underflame preheat- ing is the major contributor towards preheating of the solid. An experimental investigation is performed in order to understand the reasons for the disagreement between our model and experimental results, and also to under- stand the controlling mechanisms better. The experimental results suggest a linear dependence of pyrolysis front speed on free stream velocity. It was found that for higher oxygen concentrations there is substantial flame radiation; this is l4 reflected from the walls of the combustion tunnel and strikes the sample surface, increasing the flame spread rate. For cases with no reflection the spread rate agrees with the expression obtained previously. For cases with some reflection it agrees with the previous experimental results (Loh and Fernandez-Pello, 1984). Moreover it was found that solid is always transient during the experiments, sug- gesting a much shorter flame tip (which primarily depends on amount of fuel) than predicted by the assumption of a steady solid. A simple one-dimensional model for a vaporizing solid is developed in order to determine its transient response. An explicit expression is obtained which gives the uansient mass loss as a function of time and the applied heat flux. The expression agrees favorably with the concurrently-performed experiments and numerical results. A model is developed for wind-aided flame spread over a transient solid. The model predicts a realistic flame tip location as a function of the pyrolysis front location (speed) and time for both laminar and turbulent flows. The predictions agree with the experiments. They also. suggest that surface heat losses cannot be neglected. Surface heat fluxes in the burning zone are obtained from the burned PMMA sample by use of a simple model. The results suggest that the net surface heat flux in the burning zone increases with the free stream and decreases with the streamwise distance, though it does not vary as x"1’2. In order to check for the stability of the flame, a linear inviscid parallel stability analysis is performed. The results indicate that in the ceiling configuration the flame is stable because of forced convection (i.e, the Froude number is high). It also suggests that for lower Froude numbers the stability of the problem could be lost to infinite-reaction-rate kinetics. l4 reflected from the walls of the combustion tunnel and strikes the satnple surface, increasing the flame spread rate. For cases with no reflection the spread rate agrees with the expression obtained previously. For cases with some reflection it agrees with the previous experimental results (Loh and Fernandez-Pello, 1984). Moreover it was found that solid is always transient during the experiments, sug- gesting a much shorter flame tip (which primarily depends on amount of fuel) than predicted by the assumption of a steady solid. A simple one-dimensional model for a vaporizing solid is developed in order to determine its transient response. An explicit expression is obtained which gives the transient mass loss as a function of time and the applied heat flux. The expression agrees. favorably with the concurrently-perforrned experiments and numerical results. A model is developed for wind-aided flame spread over a transient solid. The model predicts a realistic flame tip location as a function of the pyrolysis front location (speed) and time for both laminar and turbulent flows. The predictions agree with the experiments. They also. suggest that surface heat losses cannot be neglected. Surface heat fluxes in the burning zone are obtained from the burned PMMA sample by use of a simple model. The results suggest that the net surface heat flux in the burning zone increases with the free stream and decreases with the streamwise distance, though it does not vary as x‘m. In order to check for the stability of the flame, a linear inviscid parallel stability analysis is performed. The results indicate that in the ceiling configuration the flame is stable because of forced convection (i.e, the Froude number is high). It also suggests that for lower Froude numbers the stability of the problem could be lost to infinite-reaction-rate kinetics. CHAPTER 2 A PRELIMINARY MODEL OF FLAME SPREAD SYNOPSIS This chapter theoretically analyzes flame spread over thick non-charting (vaporiz- ing) solid fuels in concurrent oxidizer flows. Emphasis is placed on making rationally justifiable physical approximations that produce useful results. Exact solutions for the pyrolysis and flame front speeds are obtained. These quantities are both linearly dependent on the free stream velocity. The pyrolysis front speed is also linearly dependent on the free stream oxygen mass fraction. 15 16 2.1. INTRODUCTION Wind-aided flame spread is a hazardous process that is difficult to quantify both in the wildland and in controlled small-scale laboratory experiments. The wildland fire poses an exceedingly formidable problem whose complications we do not presently wish to examine. This chapter focuses instead on the simpler small-scale problem. Even for this case, however, the problem is quite complex, since flame spread usually occurs under turbulent conditions, buoyancy and radiation often play major roles in the heat transfer mechanism, the solid fuel may char or display other complex behavior, and the detailed gas-phase chemical processes are unknown. These difficulties have, generally speaking, produced two general classes of wind-aided flame spread models, (A) the simplified ones that derive practical correla- tion formulae and (B), the comprehensive numerical ones that describe flame spread in terms of tables and graphs. Before describing some of their differences we emphasize first their similarities: they always assume laminar flow; they usually do not include radiation; they usually do not describe charrin g or other complicated solid behavior. If they include any of the latter two it is an approximate or simplified way. Models of class (A) have their origin in a sequence of phenomenological assump- tions that the authors do not attempt to rigorously justify. They are based more on practical experience and qualitative order-of-magnitude estimates than on deductive reasoning floor the first principles. Examples are a steady-state gas phase (Fernandez- Pello, 1979), a constant difference between flame and pyrolysis zone lengths (Quin- tiere, 1981), and a thermal balance at the pyrolyzing surface between only the incom- ing gas phase heat flux and the energy for vaporization (i.e., neglect conduction into the solid) (Fernandez-Pello, 1979), etc. Nevertheless, they produce useful formulae for practical engineering data correlations while providing insight into the many complexi- ties of this problem. Happily enough, these assumptions can in fact be justified by the rigorous analysis of the governing equations. 17 Our theory, however, is not general enough to discuss the validity of various recent models of class (A) that have undertaken a rather novel approach (Saito et al., 1986; Kulkarni and Fisher, 1988). As their starting point they take the flame spread formula of a highly simplified qualitative analysis (Quintiere, 1981). They then later artificially reintroduce physical complexity by substituting formulae describing the external radiant and piloted heat fluxes (Saito et al., 1986), the mass flux distribution from charting solids (Kulkarni and Fisher, 1988), etc. We may, with confidence, describe these as purely phenomenological models of wind-aided flame spread. By contrast, models of class (B) attempt a comprehensive description of wind— aided flame spread. These analyses employ the governing conservation equations, usu- ally without imposing too many restrictive modeling assumptions. Thus, they retain streamwise and transverse diffusion (DiBlasi et al, 1987, 1988a, 1988b), radiation (Carrier et al., 1980, 1983), charring (DiBlasi et al., 1988b; Carrier et al., 1983), solid-phase pyrolysis kinetics (DiBlasi et al., 1987, 1988a, 1988b), and finite-rate gas- phase chemistry (DiBlasi et al., 1987, 1988a, 1988b). Their drawbacks are (1) they are difficult to understand and interpret because results are presented in graphical or tabu- lar form and (ii) they do not produce definite and usable data correlations since depen- dencies on certain expected dominant parameters (e.g., the free-stream oxygen concen- tration) are usually all that computational time and cost constraints allow them to determine. Thus the influences of more routine parameters (such as the thermal responsivity ratio) usually go unstudied even though they form an important part of a successful correlation. Present analysis attempts to follow a middle course between those of types (A) and (B). We do this by on the one hand analyzing the conservation equations them- selves, not a set of postulated phenomenological relationships, and on the other hand by making enough justifiable approximations So that resorting to a numerical solution becomes unnecessary. In short, we will use analysis to extract as much worthwhile 18 information about this physical problem as possible. We wish to simultaneously deepen our physical understanding and improve our theoretical predictions. The model problem we study closely resembles that of Carrier, et al. (1980); in fact, it is identical except we introduce one additional modeling assumption that is based on their numerical results and on the analysis of a related, simplified model problem (Wichman and Baum, 1988). It is this assumption that makes the difference between numerical and analytical solution. 2.2. PROBLEM FORMULATION The gas is initially quiescent with ambient temperature T; and oxygen mass frac- tion Y5... The solid fuel also has initial temperature TL. At time t' = 0" the gas is swept into motion with free stream velocity U 2, in the direction parallel to the solid surface while the fuel is simultaneously ignited at the leading edge. A thermally undisturbed region, a flame front, and a pyrolysis zone each propagate downstream thereafter with velocities U 1,, U; and U; respectively. We adopt the idealized two dimensional geometry of Figure 2.1. The coordinate system is attached to the leading edge, past which a uniform flow of gas into the region x' > 0, y' > 0 above the solid begins at t" = 0. The solid occupies the region x’ > 0, y,’ < 0, where the subscript s here denotes the solid (it may also designate surface). At any instant the pyrolysis zone extends to xp“ (t'). The flame reattaches at xf " (t' ), as will be discussed later. It is now necessary to state some of our major assumptions for simplifying the governing equations (Williams, 1985). Concerning diffusion, we assume here that high rates of , streamwise convection produce much larger gradients in the direction per- pendicular to the free stream than along it, allowing the boundary-layer assumption to be invoked, thereby enabling streamwise diffusion of mass, momentum, energy and species to be neglected. 19 _ Pyrolysis zone Preheat zone Dead zone Ix. "P C Q C X 4! C2 8 '9 Solid Figure 2.1 Schematic diagram of wind-aided flame spread. Pyrolysis zone extends till - x‘, ”(t ) and flame reattaches at xf(t ). Superscript * signifies a dimen- sronal quantity 20 The gas-phase chemical reaction is assumed to occur through the simple one-step irreversible reaction F + v0 —) products , where v is the mass-based stoichiometric coefficient. A Schvab-Zeldovich analysis will be performed for the spreading diffusion flame with absolutely no regard for finite-rate chemical processes, whether near the inception point or anywhere downstream. In this study gas-phase radiation is assumed negligible. This is a reasonable approximation for small laboratory scale experiments for some polymeric materials such as polymethylmethacrylate (PMMA), although recent experiments with wood have shown that a minimum external radiant flux level is necessary for spread to occur (Saito et al., 1987). Consequently, attention is focused on simple vaporizing solids such as PMMA; the influences of charring for cellulosic materials and external radia- tion under large-scale conditions are left for future consideration. Finally the convective terms in the transport equations are simplified by assuming that the gas flows over the surface with uniform and constant velocity U f. This assumption is entirely consistent with the first two because diffusion flames with infinitely fast chemistry can logically be expected to locate themselves near the exter- nal fi'ee stream far from the surface, where the convective supply of oxidizer is greatest and the influences of finite-rate chemistry, flame quenching (and therefore streamwise diffusion) are least. Tire equations for conservation of species and energy in the gas and the equation for conservation of energy in the solid therefore become eDYi‘ a e ani. e p , - , pD , =—v,—R ,i=0,F, (2.1) Dt 3y 3y .DT‘ a . .31" Q‘ . - D = R , 2.2 p Dt' 31'") 8f] 6,? ( ) and 21 ,‘ - j . =0, (2.3) 3 Dr cs 32 where D“E 8‘+U; a. +ve a“ Dr 3: 8x and D: a +ni*(x‘,:‘) a Dr’ ar‘ 9; Byi' are the transient-convective operators in the gas and solid and R‘ =A‘p'2Y0’YF’exp(-E’/R'*T*) is the reaction term. Here m'(x‘,t*) = p' (x*,0,t*)v‘(x*,0,t') is the mass flux from the solid to the gas, v; = 1 and v0 = v. The following additional minor simplifying assumptions have been made in deducing Eqs. (2.1)-(2.3): equal binary diffusion coefficients, constant mixture specific heat, unity Lewis number (Le =0t/D ), low Mach number, and. negligible solid-phase regression rate. The boundary and initial conditions for Eqs. (2.l)-(2.3) are most easily discussed with reference to Fig. 2.1, where it is shown that there are three primary zones, (i) a pyrolysis zone, 0 < x’ < x;(t' ), where the solid is ejecting volatile combustible into the gas, (ii) a preheat zone, x;(t') < x‘ < U Lt' , where the fuel surface is being heated prior to pyrolysis, and (iii) a dead zone, x‘ > U Lt. , where there is no influence of the upstream flow. The term "dead zone" is understood by observing that the wavelike transient-convective operator of Eqs. (2.1) and (2.2) limits streamwise signals (originating at the leading edge) to the region 0 S x’ S U it" because stream- wise communication by diffusion has been eliminated by the boundary layer 22 approximation. The initial, free stream and upstream boundary conditions are assumed identical (see Fig. 2.1). Thus. the ambient temperature is T L , the ambient oxidizer concentra- - tion is Y5... , and the fuel concentration is zero, Y; = 0. The ambient temperature of the solid is assumed identical to that of the gas. In the pyrolysis zone we assume that the solid pyrolyzes at the constant vaporization temperature T3, a conventional and well-justified approximation. In chapter 5 we will discuss more on this approximation. When 7" < T: no pyrolysis can occur. Furthermore the solid fuel regression rate, as already mentioned, is neglected, as is the depletion of fuel (semi-infinite solid). The boundary conditions for the temperature and fuel concentration fields are shown in Figs. (2.2) and (2.3), where L‘ is the endothermic specific heat of pyrolyzation and x is the mass fraction of pure fuel in the mass flux from the surface. In the preheat zone the mass flux from the solid vanishes and the surface temperature is therefore always less than T5. The flame reattaches at x' = x; because the assumption of infinitely fast gas-phase reaction requires all fuel entering the gas to react - mixing with the oxidizer stream is impossible. This condition is obviously never satisfied in practice but experi- ence in the laboratory shows that under some conditions (e.g., flame spread in a ceiling configuration, where buoyancy pushes the flame closer to the surface near the tip) it is reasonable. In the preheat zone the heat flux from the flame to the surface must equal that into the Solid since there is no vaporization. Also, the gas and solid surface tem- peratures are identical, the normal gradient of Y I; vanishes under the flame and the normal gradient of Y5 vanishes ahead of the flame. At the reattachment point we have Y5 = Y I; = 0. All of these conditions are also shown in Figs. (2.2) and (2.3). The governing equations (2.1)-(2.3) and the initial and boundary conditions of Figs. 2.2 and 2.3 are now put into simpler form. First we introduce the time- 3 dependent Howarth transformation, with the mass coordinate z I = I p' dy " and the o =1 5 OBY Y0” 5v ywfikrt—v'fi] . Q 23 PYROLYSIS ZONE at) am 1 alt-1 = .— + w— - —— ox dz Pr 3:2 an L as,_ '52" 113% R 8;, +Per , PRE-I-IEAT ‘- ZONE -— Bs=0 [35 = [33: LV- Prkfi .— C 3C. 3255 __ BC? ‘ Bs=T-1 Bs=0 Figure 2.2 Equations and boundary conditions in the pyrolysis zone 24 PRE—HEAT ZONE ‘ Z Br=Br=0 BFBYP L=§(—-+--a:(l---1—820 at dx Fr 322 PYROLYSIS DEAD ZONE L{sy}= 0 : ZONE Br B E [31‘5” : a a , a a [31°an %=0. (27%; [37%. Taurus) KP tss=ess g=t : 212-2111. - Bs=° a: at} _ 55:0 C. Figure 2.3 Equations and boundary conditions in the preheat zone 25 - y' 1" modified transVerse mass flux w’ = p'v’ + I-afidy‘ + Uijgerdy‘. This, along 0 ‘ o x with the assumption p'zD * = constant, eliminates the density from the convection and diffusion terms. The equations and boundary conditions are then nondimensionalized by defining p=p’/p..'. D =D'ID..'. L =L'/cp..‘,T..'. Q =Q‘lc ..'T..'. t =2'U,‘/Lg‘, r =T“/T..“, u =u’/U..*, v =v’/U..*, w =w’/p..’U..’. x =x*/Lg‘, z =z'/p.,’Lg’, z, =y_.*/L,*. Y0 = Y5. Yp =Y;-. where the subscript co denotes free stream conditions, L; = v: / U ”I is the reference length in the gas and L; = a;/U:, is the reference length in the solid (as. = As'lpfcs‘). Note that the nondimensional streamwise velocity u is unity. The Schvab—Zeldovich variables . YOTYo» 1) Br=——'YF’ v _ Y-Y,, iilflr=TQ1+ ovo, (2.4) are then used to eliminate the reaction terms in two of the gas phase equations. We do not need to consider the third nonlinear equation (for T, say) because infinite-rate chemistry has been assumed. In the solid we define, for notational consistency, %=R-L ' as Thus, in the gas we obtain 26 93.; 93.; ab.- 1 325,. a: + dx +w dz Pr 322 =0, i=Y,T, (2.6) where Pr is the Prandtl number. Note that in this transformed coordinate system, the continuity is satisfied by Bw/Bz=0 since u=1. In the solid we define a new transverse coordinate, t z: _ ys x/L;/L,’ [war afag’xu 1,] ’ so that Eq. (2.3) reduces to 2 fi+ —~IPrR 3‘35 J36“ 4:0, (2.7) at BC. at;,2 where C = CI, ,.;/c is the specific heat ratio and R= p“, cpflAngfi ';A is the thermal responsivity (Wichman and Williams, 1983). Tire initial and boundary conditions for Eqs. (2.6) and (2.7) are BT=BY=BS=0att=0,x=0,andz=oo,C,=-oo. (2.8) At the interface we have aBr _ You 0> x' IU; , when the initial transients have disappeared and a steady flow has been established in the tunnel. Since we have invoked the boundary layer approximation, this steady flow must be of the Blasius type , meaning that for t’ >> x‘lU; the observer sees a steady Blasius boundary layer whose leading edge is at x = 0. In sum, for x‘lULt’ << 1 the flow is of Blasius boundary layer type, steady but dependent upon x, while for 1 < x'IULt' < no it is a Rayleigh boundary layer, tran- sient but independent of x . In between there must be a transition of some sort which, however, is of no concern to us at present. What is important is the approximation that this reasoning suggests: in the pyro- lysis zone near the leading edge (see Figs. 2.1-2.3) the unsteady terms in the equa- tions may be discarded, whereas in the preheat zone, where there is no pyrolysis, they must be retained because the surface temperature in this zone is a function of time and space (T3’ET;(x* ,t‘ )). Consequently we are led to postulate a steady state in the pyrolysis zone. Once this has been done Eqs. (2.6)-(2.10) can be solved exactly after substituting 11, 0xp(t) , (2'11) W: . w—fi—r . for the transverse convective mass flux. Here M = w*\/x IplrraU, In terms of the dimensional quantities. It is well known that for constant velocity streaming over a flat plate this is the only surface blowing function that allows self-similar solutions of Eqs. 29 (2.6)—(2.10). This distribution is consistent with the boundary layer approximation when we assume a direct proportionality between the gas-phase heat flux to the surface and the ensuing mass flux. Note that w = M N}- for all z. This approximation (w=M/‘1)? ) is quite reasonable for many materials, since it implies the heat flux (which turns out to have the boundary layer form, 8T /82 (x ,0) ~Q0 N; ) is directly propor- tional to the mass flux. For analytical solution, Eqs. (2.6) and (2.7) may be written in a simpler and more convenient form. Substitute Eqs. (2.11) for w , then introduce the similarity variables §=xh, n =(‘1P_r/2)z/‘/; (gas), (2.12) n. = (VPrR /C)€.N§ (solid) , to find 313: BB.- 325i . 4§(1-E.)-a—§-+2(M‘1P_r—n -5n—-—an-2-=O,I=Y,T, (2.13) in the gas and 2 2 a 2 _§C £+M8BS_aBs "O (2.14) Fri? 8:. an, 3113' in the solid. Note that these equations can be used to justify the steady-state approxi- mation in the pyrolysis zone, since for small é (i.e., x/t << 1 , fip ~ 0(10‘3)) all terms containing the time t vanish. This assumption may be questionable in the solid for some values of C and R; however, the assumption of constant surface heat flux,‘ temperature (T v) and Constant blowing velocity M (all in the E, — n coordinate sys- tem), leads to a steady-state solution as the only bounded solution of Eq. (2.14). In other words, under these conditions an unsteady bounded solution of Eq. (2.14) in the pyrolysis zone does not exist. 30 The initial and boundary conditions of Eqs. (2.8) become B1 = By = B5 = 0 at g = 0 and at n = co , ns = —oo , while the interface conditions reduce to Pyrolysis zone (0 < g < S, 2: Y i) %=2WM(By+—9vi+x), .. 3131 _ 2W7 35s L .11) w— QC an: +25M‘jfi, __ Tv ‘1 You. (2'15) iii) B13- — Q - -V— , iV) B33 = Tv -1 Preheat zonejg‘IL < t; < 1): an , (216) ii) ESL = 2‘17”? 313s ° 3n QC ans ' Observe that M = 0 in the preheat zone, lip < l; < 1. Equations (2.13)-(2.16) are now solved separately in the pyrolysis and preheat zones (in the dead zone the solutions are simply By = B1 = B5 = 0). Because of the step change in the mass flux at Q the solutions for BY and BT will have continuous values but discontinuous derivatives of all orders. Thus, the gas phase solutions are "patched" at 5,, , not matched in the asymptotic sense (Van Dyke, 1975). Note finally . that the solutions for any dependent variable (such as By , BT, BS , or the pyrolysis front speed EP) will depend on t linearly through the similarity transformation of Eqs. (2.12). Therefore, acceleratory flame spread is impossible in this self-similar version of the flame spread problem. 31 The mathematical solutions of Eqs. (2.13)-(2.16) [or (2.6)-(2.10)] are now written in the pyrolysis and preheat zones. 2.3.1. PYROLYSIS ZONE (0 < 5 < ép): We note that in this region the gas and solid phase problems are mathematically uncoupled. By now putting a(-)/a§ = o Eqs (2.13) and (2.14) reduce to simple ordi- nary differential equations that are easily solved subject to the far-field boundary con- ditions [Eqs. (2.8)] and, at the interface, Eqs. (2.15.i) and (2.15.iii) (for the gas) and Eq. (2.15.iv) (for the solid). Thus, erfc( 11 -M‘1P_r ) 0 BY = BYS erfc( -M\/P7) .. ___ erfc( 11 - ME ) "”31 B” erfc(—MVP?) ’ (2'17) 17035 = 1335 CXP(MT13 ), where B15 and B55 are given by Eqs. (2.15.iii) and (2.15.iv) respectively, and Bys- [ v fx]l+h(M\/ITr-) ,h(¢)- “(be erfc(-¢). (2-18) One can now use Eqs. (2.17) and (2.18) to derive expressions for the gas-phase surface fuel mass fraction Yps and the mass flux eigenvalue M. Since Bys = -Y,.-S - YOJV one finds, from Eq. (2.18), _ xh(M~17>7)-Y0../v F3" 1+h(M~fP—r) ’ (2.19) 32 which shows that a minimum non-zero mass flux is necessary to produce a non- negative value of Yrs ; it is found by solving the transcendental equation h( M \lP_r ) = Y0 m,/(vx). The requirement of a minimum positive M states a simple physical fact: flame spread is not possible unless the surface is pyrolyzing at a rate sufficient to sustain it. The eigenvalue M is found by substituting Eqs. (2.17.ii) and (2.17.iii) into the surface energy balance of Eq. (2.15.ii). This gives QY eo/v - (TV _1) h(M\/P7 )-_- L0+(T -1)/c as, (2.20) where B is the mass transfer number defined by Spalding. Note that B must satisfy the condition B > Y0 _,/(vx); without this flame spread is not possible. Equation (2.20) suggest two limiting cases, one neglecting (T, — 1)/C, the other neglecting the vaporization enthalpy L . Both will obviously overpredict M . In the first case the solid-phase temperature gradient term in Eq. (2.15.ii) is discarded. This is reasonable when Pr, M, and C are of order unity, Q is large, and L/Q is of order unity, and suggests that the energy balance occurs between the incoming gas phase heat flux and surface vaporization enthalpy, as for flame spread over a thin fuel. This procedure was utilized by Fernandez-Pello (1979). On the other hand, the second case has zip—r/QC ~ 0(1), Pr ~ 0(1) and ML/Q << 0 (1). These conditions are satisfied when the vaporization enthalpy flux ML is much smaller than the heat flux Q from the flame. Since so little energy is required to vaporize the fuel most of the incident heat flux is conducted into the solid. Our result is of course the composite of these two cases. 33 2.3.2. PREHEAT ZONE (5,], < E, < 1): Before deriving the solutions in this zone we note that even though the governing equations in the gas change discontinuously across §= §P we can still apply Eqs. (2.17.i) and (2.17.ii) as "initial" conditions at g, for the solutions to By and B1 in §> tip. This is achieved by writing n = (Way/45;? in Eqs. (2.17.i-ii), and it means that By and Br will in fact be continuous across the line i = 54,, but that the streamwise derivatives of all orders will be discontinuous. The further requirement of continuous heat flux into the solid at g (i.e., lim 3 Ian = lim 8 Ian) allows us to ” H; B1 t—rc '37 A derive the pyrolysis zone spread rate formula (Sec. 2.4]: this condition is necessary because the - heat flux from the flame to the surface cannot have a step change at any value of §. What this requirement produces, unfortunately, is a solid phase preheat zone solution that cannot be patched to the pyrolysis zone solution at Zip. Although this result is somewhat unsatisfactory it is reasonable (given the assumption of a steady pyrolysis zone) because a step change in the mass flux across the pyrolysis front, cou- pled with a continuous normal gas phase heat flux there, requires a step change in the heat flux entering the solid [compare boundary conditions (2.15.ii) and (2.16.ii)]. This in turn produces the discontinuity in B; at é = 5p . We do not expect this to produce any difficulties, however, since the solid phase solution only enters the problem through the boundary conditions. As long as they are consistent with the gas-phase processes the resulting predictions will be accurate enough. To solve for By we put w = 0 in Eq. (2.6) and apply the initial condition, By (x ,z ,0) = 0, the far-field boundary condition, By (x ,oo,t) = 0, the interface condition, 313,, (x ,0,t )/82 = 0, and the patching condition obtained by evaluating the upstream solution given by Eq. (2.17.i) at xp = épt. The solution is found by using Green’s functions after a suitable change of variable (Appendix 2.A): 34 = _ - S + u _ . By (§,11) erfc (—M m {1? 0e cosh2su erfc [ms M VPr ]ds , (2 21) 1 1 where u = 71150 - €11)“: - QB; and (0 = [(53 — €p)l(§p(1 ‘ 5.0)]? i The procedure for finding B1 is slightly more complicated because the boundary condition at the interface is inhomogeneous. But since the equation and boundary con- ditions are linear we use superposition and write B1 = B11 + B12, where the initial and boundary conditions for B11 are identical to those for BY (witir BYS replaced by B13) and for B12 we impose the conditions B12(x,z,0) = B12(x,oo,t) = BT(xp,z,t) = 0 and 8Br2(x,0,t ya: = g (x ,t), where g (x ,t) is the unknown heat flux distribution from the flame. The solution is (see Appendix 2.A for B11, Appendix 2.3 for B12) 314%“) = WfilfiJ‘T—fé e‘“ + “2)cosh23u erfc[0)s — MWT’FMs o ‘11 g: g (S) 25 _ - T .. —n (1 —s) 24; 1(1—s)m-,——.r)“""[ e—s it“ (2'22) where u and to are defined after Eq. (2.21) and 327(5) = aBT(§,0)/an = 2g (x ,t NW? . Finally for the solid phase we put w = 0 in Eq. (2.7) and apply the conditions BS (x ,C, ,0) = B5 (x ,oo,t) = B; (x =t,§_,,t) = 0 and 8B; (x ,0,t )IBC, = g (x ,t ). The solution (Appendix 2.C) is 1 _ rig 83(5) .. 413:5] 35 where gs (§) = BBS (gm/ans, ”(13 = Cs/Z‘I; = [C l2‘lPrR 111,. For future reference we note here that g (x ,t) and g“ (x ,t) are related to one another [see Eq. (21011)] by T. Q g = is . (2.24) while g7 (é) and gs (§) obey the relationship (2.16.ii), giving _ __1_ gr - Q figs - (2.25) We are now able to deduce formulas for the pyrolysis front and flame front spread rates, the flame location, the heat flux to the interface, etc. 2.4. RESULTS We begin by deriving a complicated integral equation giving the heat flux to the surface in terms of the various thermophysical parameters. This is done by requiring that the temperature across the interface be continuous. The following relationships are obtained at the surface of the preheat zone under and in front of the flame, respec- tively; ,0 Y i) was): as: ’- 3°“. a, sisif. .. . BS (QC) (2'26) u)Br(§.0)= Q was). t, sisr. By substituting for By, B1 and BS and gs from Eqs. (2.21), (2.22), (2.23) and (2.25) we find 36 § 1 i _ ‘11 i 81(3) ill;- 8T(S) ”“9 ‘ 2t; 10 -s))/'s—\/E——sds +m2fils-WJE—JEds’ (2'27) where Yo... v +BTSI(§);§p <§<§f ”(’9 = (Bu— Bys)1(§);§f < i <1 , (228) with _ _2_” .32 erfc((ns -M1/fi) Hg)— if e erfc(-MW) d5 0 (2.29) being the integral of Eqs. (2.21) and (2.22) evaluated at 1] = 0. Equation (2.27) is a type of Fredholm-Volterra integral equation of the first kind for the unknown heat flux distribution gT (é), which is related to the physical heat flux [see Eq. (2.22) et. seq.) by g = \[P—rgT/Z‘J; . We expect g to decrease in the streamwise direction, eventually to zero at g = 1. This decrease however, may not be monotonic since local increases in gT can produce local relative maxima of g. We may expect a relative maximum to appear only at the flame reattachment point, there being no physical reason for finding one anywhere else. In any case, Eq. (2.27) is exceedingly difficult to solve, because very little is known about integral equations of this kind (Linz, 1989). An asymptotic analysis (coordinate expansion) about the points x = Q, and i = 1 suggests that the second term on the right-hand side of Eq. (2.27) produces near Q the growth, g7. 8: (2(Tv - 1)/Q WHEN—17:5), while the first term is responsible for the decay near (5: 1, g,- = 1 -— 5,. Whether or not the growth in gT is strong enough to produce a local maximum in g is still an open question that we do not wish to explore here. Chapter 3 deals with the solution of Eq. (2.27). It will suffice for our purposes to evaluate Eq. (2.27) at certain choice points where the solution for g1 (é) is easy. 37 2.4.1. PYROLYSIS FRONT SPEED One such point is Q, , where the left-hand side of Eq. (2.27) becomes H (ip) = (T, - 1)/Q , so that it reduces to a simple Abel integral equation, whose solu- tion is (§>-—2- g” T"1 (230) 87' p -‘/7—t l_§p QVR- . . But from Eq. (2.17.ii) we also have = __2- M ' 2 31 37%”) EB" erfc(-MVP?) ’ ( ' ) so that the requirement of a continuous heat flux into the surface gives, after equating Eqs. (2.30) and (2.31) and reintroducing physical variables, 2 e a e e e e 2 U3, - U; p313; T: - T; erfc (-M‘1Pr ) ' Here the flame temperature is defined as T; = T; + Q‘ Y5,/VC;.,. Note the similar- ity between this result and the Loh and Fernandez-Pello (1984) formula, which becomes even more striking when we consider small M , and observe that U;/(U:. - U;)=U;/U:, under most conditions. 2.4.2. FLAME FRONT SPEED This is found by using the fact that there is no fuel or oxygen at the flame reat- tachment point. From Eq. (2.4) we have, by definition, By (§f ,0)/B” = (Y0 ,Jv)/(Y0,,/v + YFS), which we equate to Eq. (2.21,) evaluated at g = Q, 11 = 0, to 38 find Y0, = _2_°J:8_s,erfc(mfs - MW)dS (2 33) Y0..+VYps~ ‘56 0 erfc(-MW) ’ . where (of = (0(§f ). In the general case this integral can be evaluated in terms of an infinite series (Appendix 2.D) so that the root éf can be determined by using a stan- dard bisection method. For the case of low mass flux, however, it can be evaluated in closed form to give, after some working, the very useful formula 2 1: Y5../VY;S 2 1+ 13¢ng U f" = U;cosec (2.34) Hence U; is always greater than U;, as expected. In the special case Y5,/vY,§S << 1, which is reasonable for low oxygen concentrations, this reduces further to ,_ 4UP , , _2 U f z 1:2 (You/VYps) . (2.35) 2.4.3. FLAME LOCATION The flame position in the pyrolysis zone is found by putting Y0 = Y; = 0 in Eq. (2.17.i) and in the preheat zone by applying the same conditions to Eq. (2.21). It may be compared with experimental flame shapes if and when they are ever measured. 39 2.5. DISCUSSION ' In this section the predictions of the previous formulas are examined in detail by thorough calculation of a numerical example and by comparing the flame spread for- mulas to others already in the literature. For the numerical example it turns out that the single most important parameter is the ratio QYmJL = Q ' Y5../L' , which increases with Q * and the oxidizer mass fraction and decreases with the vaporization enthalpy. Therefore we study its effect by fixing the solid phase parameters for PMMA (Wichman, 1983; Anonymous, 1988) (p; = 1.19 g/cm3, c; = 0.5 cal/gK, 1.; = 6.4x10'4 cal/cm s K, T: = 643 K, x =1, L" = 380 cal/g PMMA) and by fixing the ambient gas phase parameters for air at T 3 (p3, = 5.36x10‘4 g/cm3, ‘ (2;, = 0.25 cal/gK, A; = 1.2x10‘4 cal/cm s K, T; = 295 K, Y5, = 0.233, Pr = 0.7, v = 1.92). For the present we choose U 1 = 90 cm Is and put Q’ = 4500 cal lg fuel. These parameter choices give R = 4.22x10‘5, C = 0.5, i Q = 61, L = 5.15, Q/L = 11.84, and B = 0.83 [from Eq. (2.20)]. The minimum mass flux from the formula h( M ‘(P—r ) = You/(Vx) is M = 0.0636. Shown in Fig. 2.4 is a plot of M and the surface fuel mass fraction Yps versus QYO ,JL. Note that in Eq. (2.20) the quantities Q and Y0, are fixed and L is varied through large and small values. As L/QYO” -) 0, Eq. (2.20) gives h(Mt/P_r- ) ~ QCYOJV(T,, - 1), which produces the horizontal asymptote. The limiting value of h (M W ) likewise produces, from Eq. (2.19), the asymptote for Yps. Figure 2.5 shows a plot of the nondimensional pyrolysis and flame front speeds versus the free-stream oxygen mass fraction. We observe that the pyrolysis front speed varies linearly with the oxygen mass fraction, not with its square as predicted by Loh and Fernandez-Pello formula (1984). In our formula the increase in oxygen concentra- tion increases the fuel mass flux M which in turn decreases the factor exp(-—M2Pr )lerfc (—M1/P7 ), apparently as Y 07.0.5 . Consequently the effect of the oxy- gen is to change this quantity into a mass flux correction factor of the type that 40 0-0 I ' l ' i ' 1 0 10 20 3D YoooQ/ L Figure 2.4 Variation of mass flux parameter (M) and surface fuel mass fractiOn (Yps) as a function of QYO ,JL 41 SPREAD RATES P O 8 l 1.1.rmlml\. 0.0 0:2 ' 0.4 Figure 2.5 Variation of pyrolysis front speed ( function of free stream oxygen mass 0.0 ' 0.8 ' 1.0 and flame fi'ont speed (fif) as a § ) ITBCIIOII (Y0 g.) 42 frequently appears in physical problems with combined heat and mass transfer (Atreya and Wichman, 1989; Bird et al., 1960). Although our analysis assumes an Oseen flow it still appears because the mass flux distribution is of the boundary-layer type (Emmons, 1956). The pyrolysis front speed is found to decrease linearly with M (Fig. 2.6). This occurs because the volatile mass flux into the gas can be increased in only two ways, by increasing either M or ép. In the former case the fuel entering the gas increases tirroughout the pyrolysis region whereas in the latter an increment of mass flux is added in the interval from 5,, to Q, + Ag? . Thus as M becomes larger the incremental fuel vapor produced by changing Q, can actually be negative, i.e., 5,, can decrease; correspondingly, Fig. 2.6 and Eq. (2.32) show that as Q, increases M must decrease. Consequently it is not unusual that Eq. (2.32) should predict a smaller pyrolysis zone spread. From Fig. 2.6 we also see that the flame front speed initially increases linearly but then levels off and decreases with further increase of M . The reasons for this are clear from the preceding discussion, with the assistance of Eq. (2.34) [or (2.35)] and Fig. 2.4. As QYOJ L increases Fig. 2.4 shows that M and YFW both increase. Therefore U; decreases (as already discussed) and by either of Eqs. (2.34) or (2.35) the factor multiplying U; increases. This dominates the decrease of U; so that U ; irri- tially increases. But, since Y W increases to its asymptotic maximum faster than M (see Fig. 2.4), the multiplicative factors of Eqs. (2.34) and (2.35) level off while U; continues its decrease. . Thus, for larger values of M the flame front velocity U ; even- tually decreases. Figure 2.7 shows that although 0; and U; both finally decrease with M, their ratio U ; IU; increases linearly. This result may be useful for experimental correlations since it is much easier to measure U; (by surface temperature) and the total mass flux (by species concentrations) than the flame speed, which requires fast photography, image processing or other sophisticated techniques. 43 0.010 r I m 1 . 1 ' i ' l ' i ' l ' 0.0084 0.0031 1 0.0041 SPREAD RATES ,°-° 0.1 0.2 ' 0.3 0.4 ' 0.5 ' 0.6 ' 017 0.0 Y00° Figure 2.6 Spread rates versus mass flux parameter. HereE is the non- dimensional pyrolysis front speed, Ef is the nondimensionalp flame front speed, %ms the approximate (zero-M assumption) nondimensional pyrolysis speed and E, is the approximate (zero-M) nondimensional flame front speed. The freason the curves do not coincide is that E and Eare evaluated for M= 0 whereas in the present theory the minimum v no of M is 0.0636 30 . I a I ' f ' I ' I ' I r I 20% i ‘ m5" \.. w 10% "‘ O ' I ' I ' I ' I ' I ' I .' I ' 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 M Figure 2.7 Ratio of spread rates (Ep/ Ef) versus mass flux parameter M 45 A note of caution, however; the higher values of M in Figs. 2.6 and 2.7 are very near to the maximum permissible blowing rates for which boundary layer analyses of this kind are valid. For if 2 0.62 the analysis of Emmons (1956) predicts flame blowoff. We note from Fig. 2.4 that the maximum value of M = 0.7. Since our M W is Emmons’ 117, this produces a maximum value of M = 0.74. Also shown in Fig. 2.6 are the predictions of the Loh and Fernandez-Pello for- mula (1984), EI, and Eq. (2.34), Ef, which are valid for small M only. Both produce upper bounds for the respective spread rates. We note that the Loh and Fernandez- Pello limit of Eq. (2.32) is justifiable when M ~ 0 (10*), which by Fig. 2.4 is possi- ble only when Q IL ~ 0 (1). For PMMA burning in air, however, we have Q/L ~ 0 (10), which gives M = 0.37 and thereby invalidates the Loh and Femandez— Pello’s (1984) formula by introducing a non-negligible correction factor, [exp(-M2Pr )/ erfc(-M11173]2 = 0.44. For illustrative purposes the flame location was calculated according to Sec. 2.4.3. A plot of flame height versus streamwise position is shown in Fig. 2.8. We can compare our pyrolysis front formula, Eq. (2.32), to the recent predictions of Carrier, et. al. (1990) for their solution of practically the same problem (they do not invoke the Oseen approximation). Using Eqs. (2.18) and (2.32) with M417; 13 = 0.508 In (1+3 )IB ”’54 , ose $20 (Carrier et al., 1990) gives 2 U. . :0.” * CD 2 +7 =1r(0.508)2 9”} f, , L y + f ”“150?” , (2.36) U, - Up psAsc, cp..(Tv —T,,) cp, B ' which is exactly the same (with a different constant multiplicative factor) as the result of Carrier et. al. (1990). 46 X 0.4 r I I I ' I I ' r ' I -— 1.80.5 . -- t= '- t=10 0.3“ .......... [I \\\\ I \ N 0.24 I, \‘\ II ‘ \\ u 'I \\ 'i \\ 0.1~ L ‘t - i' l r‘ \ 1‘ \ l. i ‘1 ‘ i ‘r 0-0 I ' 'I ' I ‘ I ' I r I ' I ' I ' I L ' 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 Figure 2.8 Flame locations at different nondimensional time instants. The x and z coor- dinates are nondimensional Howarth coordinates 47 The other of our two formulas that can be compared with previous results is Eq. (2.34) [or Eq. (2.35)] for the flame front velocity. We first rewrite this in a more fam- iliar form by subtracting U P. from both sides and using some simple trigonometric identities to find Y5 .. IVY;S 1 + Y5 ,, Ivy}, (1;: (U; - 0;) tanzd) ; <12 = 12‘- (2.37) If on the left-hand side we put 0; = dx;/dt' and on the right-hand side we write U; =xf'lt‘ and U; = xglt' and then define r' = tiltan2 we obtain the well-known spread rate formula of Quintiere (1981), dx‘ x'-x _i:_= f . P , (2.38) dt ‘t which has been extensively used to predict spread rates and to correlate experimental data (see, e.g., Quintiere et al., 1986). It is here, however, that our analysis and theirs diverge. Whereas we have solved a particular problem and derived a specific formula relating U f' to the known pyrolysis front speed [from Eq. (2.32)] and therefore the relevant thermophysical parameters [for example, Eq. (2.35) gives U ; ~ Yfisz and U; ~ 5...], they on the other hand have used Eq. (2.38) as the starting point from which to derive (hopefully) more general results . This is done by postulating vari- ous functional forms for, e.g., the flame location x;(t') in terms of the heat flux to the surface and the mass flux from the surface (Saito et al., 1986; Kulkarni and Fisher, 1988) and the time t" (t‘ = (1t/4)p;c;A;[(T: - T; )/q,: 12 is commonly used [Quin- tiere, 1981]). Recent studies have attempted to model the local heat flux at the surface 48 in terms of fourth-order polynomials in the time t' to simulate the mass flux distribu- tion for more complex materials such as wood (Kulkarni and Fisher, 1988). Our result [Eq. (2.37)] bears no relation to these formulations although it does justify the use of Eq. (2.38). Whether or not this simple result can then be legitimately extended to describe more general cases can only be answered by examining them in the same detail as here. An important omission from our analysis is finite-rate chemical kinetics. Although this should not influence the pyrolysis front speed calculation it will doubt- less affect the flame speed calculation. The flame will never completely reattach at the surface, even if Y5... = 1. On the contrary it will break open when the Damkohler number, D = A "' Y5_exp[—E " IR ‘ T;]/(p;U; 2N3), obtained by nondimensionalizing Eqs. (2.1) and (2.2), is less than some extinction value Dan. The value of Em at which this occurs is found by first calculating Tf'm from the formula for D , then put- ' ting Y0 = 0 in Eq. (2.22) and determining the point where T; has decreased to this value. Clearly, Em must be greater than Ep and less than Ef . This finite-rate analysis, although complete in principle, requires knowledge of g7 (E), which in turn requires the solution of Eq. (2.27); we will contend with this difficult problem in the next chapter. A better sense of the unreality introduced by the assumption of infinite- rate chemistry can be obtained by rcexamining the steps leading to Eq. (2.33). From Eq. (2.4), BY(E,0) = — Yoalv — Yps (E) for E less than Ef; equating this to Eq. (2.21) evaluated at 11 = 0 yields Yo... + VYFs(§) 2 °‘ _ 2ng(t0(E)5 - MW) = __ s 2. You + vrps 45);" erfc(—MW ) d3 ’ ( 39) from which Eq. (2.33) follows directly by putting Y ps (Ef) = 0 and (0(Ef) = (of. Now the maximum value for both sides is unity; it occurs when E: E, , where 49 Yrs (E) = Y ,3 and (0(Ep) = 0. The minimum possible value of the left-hand side occurs when Y rs (é) = 0. This generally requires large (0(E), since the right-hand side is small only when this is true; note that for large 0) the right-hand side is 0((0'1). Consequently if one defines the location at which Y R,- (E) = 0 as the "flame reattache- ment" point, it follows that 00(E) must be "large", giving Ef = Ep(l + (0}); even for (of ~ 3, Ef is already ten times greater than Ep. If, however, the flame terminates somewhere in the gas because of finite-rate chemistry the fuel mass fraction beneath it, at the surface (Yps (E, )) can be substantially larger than zero. Then the left side of Eq. (2.39) is much closer to unity and (0(E) must be small, giving :, = g, (1 + (0}); for (of < 1 the flame tip and the pyrolysis fi'ont may approach one another very closely indeed. The only remaining requirement in this finite-rate case is that Y [:5 (E) must vanish at or before E = 1, i.e., all of the fuel liberated by pyrolysis must eventually be consumed; unlike the previous (infinite-rate) case this need not occur at a reattachment surface flame. In sum, finite-rate chemistry effects could produce a flame front velo- city only slightly larger than the pyrolysis front velocity. The distinction between flame front and pyrolysis front would then become almost irrelevant. Another omission that is generally much more important than finite-rate chemical kinetics is gas-phase radiation. It is well known that flame spread is not possible over materials such as wood without a substantial background external radiant flux level (Saito et al., 1987). We also know from this and previous theoretical analyses (Carrier et al., 1983) that it is necessary for acceleratory spread over simple materials of the kind analyzed here. Nevertheless the proper means for including radiation are not clear and further study beyond the preliminary scope of this work is warranted. 2.6. CONCLUSIONS Analytical expressions have been derived in this study for the pyrolysis front and flame front speeds. This was made possible by the simplifying assumption of Sec. 2.3, 50 namely, a steady state in the pyrolysis zone. The formulation in temrs of the simi- larity variables (E11) makes these spread rates constant. Thus, physical quantities that cannot be reduced to self-similar form in these variables, such as the radiant flux from the flame for example, are necessary to produce acceleratory spread. The pyrolysis front speed is found to depend linearly on the free stream velocity and the ambient oxygen mass fraction, while the flame front speed is proportional to the inverse of the oxygen mass fraction. The flame front to pyrolysis front velocity ratio varies linearly with the mass flux. We have shown that Loh and Femandez-Pello’s formula for the pyrolysis front speed can only be used in the limit of small mass flux; the discussion of Sec. 2.5 sug- geSts this assumption is inappropriate even for PMMA under ordinary burning condi- tions. We have also shown that the well-known spread rate correlation of Quintiere [Eq. (2.38)] has a basis in this more detailed theory of ours. It is not clear, however, whether it is legitimate to extend this simple result to cases of far greater generality as some authors have done. A more comprehensive theory of the kind developed here, perhaps including radiation, perhaps charring materials, etc., will be necessary to answer this question. In next chapter the numerical solution of Eq. (2.27) is obtained, along with the surface heat flux distribution. An expression for the excess pyrolyzate is also obtained. 51 APPENDIX 2.A: Green’s Function Solution of Gas-Phase Preheat Zone Equation Here the equations for By and B1 are of the form B, + BJr = (l/Pr)B,,, B = 0 (initial condition), B = B2 = O (in the far field, 2 = co) and B, = 0 [at the surface, see e.g., Eq. (2.10.i)]. In addition, to patch the preheat zone solution for B to that in the pyrolysis zone we. apply Eqs. (2.17.i,ii) at 11p = (JP—r/2)z NE; , i.e., B = F (Ep) .=. Bserfcmp — MVP?)/erfc(-M\/Fr-) at xp = Ept. To solve we define the new variables x = (x - Ept)/(1 - Ep) and p. =52 - t (Baum, 1980). Since E= x/t s l we have 1.1 s. 0, so that from the definition of It the range of t is now lul s t S T, where T is some arbitrary upper bound. The coordi- nate x is now eliminated from the equations and boundary conditions, which reduce to B, =(1/Pr)Bu, Bz =0 at 2 =0 and z =oo and B=F(11p), where 11,, = (Way/«127W, since it = 0 gives t = ‘11 = lul. We therefore write B = B(z,t;|1rl) since u is only a parameter, so that the patching condition, written for- mally, becomes B(z,|ul;|u|) = F[(\(P7/2)z/\IEP_W]. The solution in terms of Green’s functions is now straightforward. Define G (2 ,t IC,1:) as the solution to the equation G1 + (l/Pr)G§§ = 5(C — z)5(1: - t) subject to G = 0 when 1: = T, G; = 0 at C = 0 and G = G); =0 at C: co, whereby G(z,t| 12,0 = (2.A.1) V a _ ‘11—); —Pr (g—z )2 —Pr (§+z )2 — 40 4) + e 4U '4’) , t _ 2‘1 1t(t -1:) and therefore Ben; ii”) = —jG’de=2(—1'§— j tan2”0d0 0 n=0 "- 0 no (_1)nc?.n n—l (_l)mA2(n-m)-l = ——-—'——-2 +(—1)"tan-1A (2.D.3) Substitution of Eq. (2.D.3) into Eq. (2.D.2) gives the desired solution. CHAPTER 3 HEAT FLUX DISTRIBUTIONS SYNOPSIS In this chapter the heat flux distributions in idealized wind-aided flame spread are determined. The surface heat flux and the surface temperature can be found by numer- ically solving an integral equation. Algebraic expressions independent of the free stream velocity and the oxygen mass fraction are obtained for the surface heat fluxes. The ratio of total surface heat flux in the under-flame zone to that in the pre-flame zone is found as a function of the pyrolysis and flame-front speeds. In addition, the fractions of the total flame heat release that appear as heat fluxes to the pyrolysis region and to the preheating region are determined. An analytical expression is obtained for the excess pyrolyzate. 57 58 3.1. INTRODUCTION Tire self-sustaining feedback mechanism for wind-aided flame spread is well known: the heat flux from the flame must be sufficient to produce a spread rate sufficient to produce a mass flux sufficient to support the heat flux from the flame, etc. In order to theoretically determine the rate of wind-aided flame spread over solid fuels it is therefore necessary to know something about the heat flux to the fuel surface, and also about the relationship between the heat flux (or the surface temperature) and the fuel mass flux. For the various engineering analyses that use integral methods to solve highly simplified forms of the conservation equations, the heat flux along the entire surface must be specified before solving for the spread rate. In theories employing more complicated forms of the conservation equations it is necessary, for analytical solution, to specify the heat flux only in certain regions, such as downstream of the advancing flame (Fernandez-Pello, 1979), or in (and slightly downstream of) the pyro- lysis zone (Carrier et al., 1990), or only in the pyrolysis zone (Wichman and Agrawal, 1991; Chapter 2); in Carrier et al. (1990) and Wichman and Agrawal (1991) (Chapter 2) the heat and mass fluxes in the pyrolysis zone are directly proportional to one another (they have the same functional dependence). For the numerical studies using still more complicated equations (Carrier et al., 1980; DiBlasi et al., 1987), no condi- tions are needed on the heat flux; it is determined along with the flame spread rate dur- ing the course of the solution, although some restrictions on the mass flux may Still be necessary (DiBlasi et al., 1987). In this chapter we will examine the heat flux and surface temperature distributions arising in our idealized model of wind-aided flame spread. The previous chapter was devoted primarily to deriving formulas for the pyrolysis and flame-front speeds. For practical reasons, obtaining flame spread formulas is certainly a desirable goal. In order to better understand flame spread, however, it becomes necessary to examine the heat flux distributions in the physical plane, for it is the heat fluxes that determine the 59 rate at which flame spreading actually occurs. In other words, the flame heat release and the ensuing heat fluxes to the fuel surface are the principal causes of flame spread. For this reason it is crucial that the heat flux distributions be determined for our (or any other) model. But there are also other important reasons. The knowledge of the heat flux distributions may result in the development of revised, perhaps more useful, interpretations of wind-aided flame spread. Knowledge of heat flux distribu- tions also forces us to recognize the weaknesses and limitations in our current models, that newer and better models might follow. It is of no ultimate use, therefore, to cloak a model’s intrinsic defects by emphasizing only its most favorable and most visible features. In section 3.2 we discuss the model and the surface heat-flux formulas that can be derived fi-om it. In section 3.3 the heat fluxes reaching the preheated surface are cal- culated, as well as the preheat zone surface temperatures. In section 3.4 we first dis- cuss the normalized heat fluxes (the heat fluxes found in sec. 3.3 depend upon Yo_; they are normalized to factor out Y0 ,.) and then develop expressions for the fractions of the total flame heat release that appear as heat fluxes to the pyrolysis region, to the under flame preheat region, and to the pre-flame preheat region. An expression for the excess pyrolyzate is also developed. In section 3.5 we present our conclusions and suggest future work. 3.2. THE MODEL Shown in Fig. 3.1 is the schematic illustration of the model problem. In labora- tory experiments the ceiling configuration is the easiest to examine (Mekki et al., 1990; Chapter 4) because the adverse influences of buoyancy (which, in the floor configuration, lift the flame away from the fuel surface) are minimized; in this case the flame much more closely resembles the model configuration shown in Fig. 3.1, where buoyancy and finite-rate chemistry are neglected. Note, however, that in the ceiling GAS (oxidizer) I'__—” : U” . §—>Up §-> Uf YROLYSIS—t §= PREHEAT ZONE = DEAD ZONE - s g ZONE Difi’usion flame x :=U°°t ’ "p SOLID Figure 3.1 Model configuration 61 configuration buoyancy may produce instability, since cold gas flows between the flame and the surface (Agrawal and Wichman, 1992b; Chapter 8). Since gravity is neglected in our model, we are simply drawing attention here to the fact that the actual tests in the ceiling configuration resemble the illustrations of Fig. 3.1 much more closely than do tests in the floor configuration. After ignition, the tip of the surface-pyrolysis zone propagates downstream with velocity U;, while the flame tip propagates downstream with a higher velocity U I? , burning the "excess" (Pagrri and Shih, 1976) fuel species released from the pyrolyzing surface and not already burnt there. The wind speed, U L, is greater than both U; and U f . As shown in Fig. la, the flame is embedded inside a thermal boundary layer. Tire thermal boundary layer is slightly thicker than the velocity boundary layer, since the Prandtl number for gases is generally less than unity. In the model problem shown in Fig. 3.1 there are four distinct zones to be con- sidered. The first is the dead zone, x'ZULt' , where there is no influence of any of the processes occurring in the region 09:" - >2 éRU-énl-é: 1.1-e) (3.6) = i=j+l “ _ 3/2 3/2 Rta.b.x)=2(bx_ zflx-Z- -\/x -b 2(2): _b)(x a) -(x-a) -a b—a 2 (x-a )5’2—(x -—b )5” 5 b-a ' (3.7) The solution vector G is now obtained by using Gauss elimination to solve the linear system (343-111 The accuracy of this numerical scheme was tested by using (3' (§)=§”', with 092156. The function I?(§,§P) was then evaluated from Eq. (3.3). By using this mag), 6 (g) was calculated from Eqs. (3.5)-(3.7). For m=0 and m=1, (7’ was com- pletely recovered, as expected (linear splines reproduce linear functions exactly); for larger values of m the error norm was found to be given approximately by HG (§)-5(§)II“SC (§)/(N -—1)2. Thus as N ->oo the error becomes vanishingly small. 67 Unfortunately the actual problem [H (§,§p) given by h(§,§p)/\IT:E;, as seen from Eq. (3.2) et seq.] is n0t so well behaved, see Fig. 3.2. There are two distinct regions, one where 2‘, ~ 0.(§f)=0(10-2), the other where g ~ 0 ( 1). By using several stepsizes, . ranging from A: ~ 10‘4 to Ag ~ 10‘2, it was found that the solutions for (32, . . . 13m changed very little (<5%), but that (31 varied somewhat (10%). This is to be anticipated because pivoting in Gaussian elimination piles the greatest error on the last element calculated, which is GI. To check this effect, 62, (i3 and (34 were used to extrapolate values for GI that were compared to analytically calculated values at the pyrolysis front using (Wichman and Agrawal, 1991) 61:6 (gp )= 3T 1 2 Eap [TV-1] = — — . (3.8) gnu—é,» gnu—e.) VT: l-ép QV'R' Agreement was found to be within 5%. Figure 3.3a shows the normalized flux gT (g) for various values of Y0... It is clear that figT increases uniformly with Y0... Also, the maximum of gy(§) occurs at the flame reattachment location, éf, where g7 (é) is piecewise continuous. This is caused by additional heat reaching the surface because of flame reattachment. The function gT (é) drops sharply after fif. The abrupt changes of slope in the curves for §=3§f occur because of the change of grid size; for §S3§f we used A§=O.1§f, while for §>3§f we used A§=0.025. Figure 3.3b shows a plot of the actual (nondimensional) heat flux, g=~lP—rgT/2‘E ; we have put t=1, so that §=x. We have also used2 Pr=0.7. We note by comparison to Fig. 3.3a that although g is piecewise continuous at Q there is no conspicuous local maximum of the kind seen for gr. 2 This value of Prandtl number is chosen because the characteristic length in the gas-phase is still the viscous diffusion length, V/U... even though an Oseen velocity profile is assumed (see, e.g., Carrier, et al., 1980). ‘ 68 Figure 3.3a Plot of the normalized nondimensional heat flux gr (é) versus é, obtained from the solution of Eq, (3.4) using G-splines. The stepsize A§ is increased by the factor 10 at 39. The maximum of the curve represents the location Ef. 69 0.06 0.03 ' 0.10 Figure 3.3b Plot of the nondimensional heat flux, g =(‘1757 /2)g7~ N; , versus é. Here g(§f) is always less than g(§p), even though gT(§f) is always greater than 87' (gp) 70 3.3.2. SURFACE TEMPERATURE A normalized nondimensional surface temperature in the prehea't zone is calcu- lated using (Wichman and Agrawal, 1991; Chapter 2) l . Tho-T; (1.5 )6 (s) T 5 =45 ————ds . 3.9 ,(§) QWE/(zficg...) ! \ls—z; (. ) Note that the surface temperature function in Eq. (3.9) is actually the second integral on the right-hand side of Eq. (3.2). The profiles of T} are plotted in Fig. 3.4 for vari- ous Y a -. We observe first that the surface temperature is uniformly higher for larger Y0... and second that it is always smooth and above ambient at E, . The same abrupt changes of slope occur for T} as g7 (or g) because of the switch to a coarser grid downstream of éf. 71 H N .1 — Yo..=1 -- 0.31 --- 0.432 . - - 0.332 - 0.331 —- 0.232 _ NORMALIZED SURFACE TEMP. (TS) F1gure 3.4 Plot of the reduced temperature 7:, given by Eq. (3.10), versus g. The first kink (from left) in the graph (for Y 0 ”=1) represents the location of the flame front. The second and third kinks (numerical error) occur because of step size changes that make the computation faster. This is also seen in Fig. 3.3a for Yo..=1 72 3.4. SOLUTIONS FOR NORMALIZED FLUXES and OVERALL FLUXES It is of interest to determine whether the many gr (6) profiles of Fig. 3.33 can be arranged to [fall on a single curve, thereby eliminating the explicit appearance of the oxidizer mass fraction from these graphs. We found that this was indeed possible, although separate' correlations were necessary in the under-flame preheat zone (£1, <§<§f) and in the pre-flame preheat zone ( fif <§<1 ). In the region Q <§<§f the g1 (é) curves collapsed when gr VE/gyp Vi: was plotted against the modified coordinate (fi—épméf fip). The resulting single curve, shown in Fig. 3.53, can be accurately represented (i4 % accuracy) by the function gT<§WE =1+5[ H, , n=0.6. (3.10) 81‘, (Q NE; gf ‘E-tp ] In the region é, <§<1 the curves collapse when g1 fi/grlfif was plotted as a function of Uéf. The resulting curve shown in Fig 3.5b, is accurately correlated (13 %) by the function3 8T(§)§ {if 13 . (3.11) 81,(§f)§f - é As a consequence of Eqs. (3.10) and (3.11) we can calculate g7 (é) at any streamwise position, 5, in the preheat zone for any oxidizer mass fraction Y a ,,, once the nondi- mensional pyrolysis and flame front speeds (Q, and g, ), and the normalized» nondi- mensional heat flux at the pyrolysis front, gT (fip ), have all been calculated from Eqs. (3.1),(2.48) and (3.8), respectively. We will return to this later, when we compare the 3 Note that Eq. (3.11) shows algebraic decay, not exponential (as generally assumed). 73 l I 0.0 0.2 0.4 0.3 . 013 1.0 E-S, / 6:5, Figure 3.53 Heat flux correlation forathe under-flame preheat zone, showing the rise of the heat flux as the downstream flame tip (reattachment point) is approached 74 1.0 l ' I V l l l l I I I I I 0.84 d u}; 0.34 _ III m -t \ .. “1‘- no 0.4T - 0.2+ - 1 I l I I I I I —:I=I¥ I I I I O 4 8 ~ 12 Figure 3.5b Heat flux correlation for the pre-flame preheat zone. Note how rapidly the decay occurs 75 fi-dependence of the physical (g (§)) and normalized (gT (§)) nondimensional heat fluxes. At present it is instructive to make comparisons of our results to those of Carrier et a1. (1990), based on the under-flame correlation given by Eq. (3.10). In physical terms, Carrier et al. (1990) postulate that the preheat zone flux, g =8T/ay , is weakened downstream of the pyrolysis front, 1,, , by a factor f“2 because of vertical dilution and by an additional factor of Jc’l’2 because of the functional dependence of army on the self-similar coordinate, ‘n=y/\/J_t- . This means that gx (or gT‘E for the normalized flux) should be constant in the preheat zone. From Fig. 3.5a we find that this is indeed the case, at least when fi-ép is small. As é-ép increases, however, this corre- lation function will remain at unity while ours rises to a value of 6 at éf . This rise is caused by impending flame reattachment, which bends the flame back toward the sur- face downstream of the pyrolysis front. Although strict reattachment is impossible, the "excess pyrolyzate" calculations of Pagni and Shih (1976) show that the flame approaches the surface because the fuel vapors are all on the surface side. The com- peting effect (at least in the floor-spread case) is buoyancy, which lifts the weakening flame away from the surface. In ceiling flow, buoyancy pushes the flame closer to the surface. Impending flame reattachment, along with the correlation given by Eq. (3.10), can also be used to explain the difference between Figs. 3.3a and 3.3b. In Fig. 3.3a we plot gT(§) versus i (or g7 (x) vs x when t=1) for which Eq. (3.10) gives 8T,(§f)/ST,(§p) =6VFP7EF, indicating that .between fip and Q the gT(§) curve must rise. For Fig. 3.3b, however, we recall that g=(‘IF;/2)gT/‘1; , so that when t=1 we find the ratio g f (x, )lgp (xp )=6xp le , which is smaller than unity, demonstrating that between xp and xf the g (x) curve must decrease. Note that g(x) is still only piece- wise continuous at xf , where it has a small local maximum. 76 We now observe that the plot of the heat flux versus é, as given in Figs. 3.3a, is somewhat vague, since coordinate §=x It is only a similarity variable. In order to obtain a clearer picture of the heat flux variation it helps to fix either time, and vary only x (as in Fig. 3.3b), or to fix the location and vary only I. When 1: is fixed, say x=xo, and t is allowed to vary, the dead zone passes at t=xo, and the flame reattach- ment and pyrolysis fronts pass at t=x0/§f and taco/é], , respectively. Since Q, and Q are quite small, this has the effect of compressing the pre-flame preheating zone and lengthening both the under-flame preheating zone and the pyrolysis zone. This is shown in figs. 3.6a, for the heat flux g7 of Fig. 3.33, and 3.6b, for the surface tem- perature of Fig. 3.4; we have used x0¥1 for these graphs. In Fig. 3.6b the region C-A represents the pre-flame preheating zone and A-B represents the under-flame preheat- ing zone. Comparison to experimentally- measured temperature profiles for PMMA (Loh and Fernandez-Pello, 1984; Mekki et al., 1990; Mekki, 1991) shows that the predicted under-flame preheating zone exists over a considerably larger time interval. The reason for this discrepancy (Mekki et al., 1990; see Chapters 4 and 6) is that the theoretical analysis employs a mass flux w=M ‘5 that has attained its steady state (maximum) value whereas in the experiments this maximum value is asymptotically approached only for large times. Consequently more fuel volatiles are released to the gas in the theory, producing more excess pyrolyzate and thus a longer flame. This will be discussed in the Conclusion section in greater detail. For fixed x=xo and variable t it is not difficult to calculate the total heat fluxes to the pre-flame and under-flame surfaces. We have X,/§, x0]; 871’)? = I gdt ; .901? = J- gdt . (3.12) I, X,/§f respectively, where g (rte/04W /2)gT(x0/t)/‘/.? . The separate values of g are of no 77 I 6 I r r T 1 I I _ You-=1 " 0.51 "" 0.432 ' ' 0.332 ‘- 0.331 ‘" 0.232 \ \ .-,~..~ ~ . ‘\ “~ 4 N‘-:‘?~:to\ -—-—_._;'_’:"s= ------------------------- ‘ Méi’ ................... ' I ' ' ' l V V I l r I I I I 400 800 1200 1300 TIME Figure 3.63 Heat flux of Figure 3.3a for fixed x (x0=1) versus time t ‘1_ SURFACE TEMERATURE (K) 78 -- YO..=0.232 - - 0.331 - - 0.333 -- 0.432 1 ' T I l — 1.0 a . . . . . . , . . a 400 800 1200 1600 2000 TIME Figure 3.6b Surface temperature of Figure 3.4 for fixed x (to =1) versus time t 79 great interest to us -- the ratio is. To obtain this, we use Eq. (3.11) for g” and Eq. (3.10) with n=0.5 for g'UF; using n=0.5 instead of n=0.6 allows us to evaluate the integral in Eq. (3.12.ii) exactly while introducing. an error of less than 2% into [we when evaluated with n=0.6. By neglecting Q5 and using gT,\/E}T/gr,‘/E; = 6, we find fur _ 5 (6-1) E: — '5'[03I2 —1 “PS-037] I (3°13) where 0' = Q I§p. As a —-) 1 the ratio vanishes, since the under-flame region has disappeared. As 0 -> oo we find g'UF/g'pp=503’2/9, showing that the under-flame heat flux completely dominates the surface preheating. In this case the pre-flame preheating may be discarded. A plot of Eq. (3.13) is given in Fig. 3.7, which also provides the second abscissa, Yo... (non-linear); the data for the 0 vs. Y0... relationship is taken from Fig. 3.3 of Wichman and Agrawal (1991) (Chapter 2). We see that o for PMMA is large enough over the entire range of Y0, for the large-o approximation, g’UF lg'pp~ Sam/9, to be quite accurate. Consequently, our model states that for laboratory-scale wind-aided flame spread over PMMA (in the ceiling configuration) the fuel surface preheating occurs almost exclusively under the advancing flame. 80 30 I I ' I I [ I 20"- . Eh Q. DD \ . 4 Ch not) 10" - 0 . r ' l ' I ' 0.0 4.0 8.0 12.0 0' Figure 3.7 Plot of the ratio g’Up/g‘PF given by Eq. (3.13). Also shown is the graph of 5/90‘”2 (dotted line), for comparison. In this figure the parameters are for PMMA with a wind speed of 1 m/s. In our model, because of the Oseen flow assumption, the wind speed appears as a non-dimensional quantity, and hence, will not change the above ratio. 81 3.4.1. EXCESS PYROLYZATE Pagni and Shih (1976) have suggested that excess pyrolyzate is responsible for fire growth. Excess pyrolyzate burns far downstream from its source and preheats the virgin solid. Depending upon free stream conditions (speed and oxygen concentration) both convection and radiation heat transfer to the surface could contribute significantly, or could be neglected individually, depending on the situation. An increase in wind speed increases the convection heat transfer to the surface by decreasing the flame height. It also extends the preheat region. Radiation from the flame depends on flame temperature and emissivity, which primarily depend on free stream oxygen mass frac- tion, and fuel type. Part of the flame radiation is lost to. the environment, whereas the rest reaches the surface. For compartment fires all of the flame radiation reaches one wall or another. For a high-convection system in air the flame radiation is negligible (Mekki et al., 1990), but it becomes much more important for higher oxygen concen- trations. Even for low oxygen concentrations it becomes important under quiescent conditions (e.g. microgravity); hence, it is obviously important to include both radia- tion and convection (with boundary layer terms, i.e., relaxing the Oseen-flow assump- tion) for a complete analysis. Solution of the resultant governing system of equations, if not impossible, is extremely difficult. One way to retain the influences of radiation in an approximate form is to include its effects only in the region where preheating of the solid is a maximum. To find the region we first determine the excess pyrolyzate. This is the fuel that could not burn in the pyrolysis region. It is given by (Pagni and Shih, 1976) Y}: nae = j(p*u*r;),; dx“ , (3.14) 0 where y}, is the location of the flame at the pyrolysis front xp‘. The flame lies along 82 the locus 11:1] f, (where Y;=Y;=O) which is given implicitly by (Wichman and Agrawal, 1991; Chapter 2), Byserfcmfl ”MVP—r) _ £1 erfc (-M\/7’7 ) v Converting Eq. (3.14) to similarity variables (Wichman and Agrawal, 1991; Chapter . 2), we obtain 11]: . Inmwn e 0 Y3 = =1+ ——ex [Mp7 2— 34W 2], 3.15 "3p M ‘17,? B P ( ) (11 fl ) ( ) x; where nip is the total mass loss from the solid, mp=wa'dx*. 0 We now perform an energy analysis of the idealized wind-aided flame spread problem to determine the relative importance of each zone (pyrolysis, under—flame preheat and preflame-preheat). The total heat lost from the flame in the pyrolysis zone (gem) is the total heat flux conducted away from the flame in the transverse direction (both above and below the flame) since there is no streamwise conduction because of the boundary-layer approximation. Thus, 1‘; gm, .—_- j [1* £5] dx‘ . (3.16) 0 y TI}: Note that g with a hat (g‘ ) represents the total heat loss from the flame and g with an overbar (g’) represents the total heat received by the solid surface. Now, the total heat 83 produced by the fuel (g‘mT) is the product of the heat of combustion and the total mass loss fi'om the gasifying solid. Hence, x, 81m“ = Qt'hp =Qwa'dx‘ =§Py +§PR . (3-17) 0 where gm is the total heat lost from the flame in the preheat zone. Converting Eq. (3.16) to similarity variables we obtain (Wichman and Agrawal, 1991; Chapter 2), g" = GILT; ” (318) am 2MP? Q " dn “" ’ ' where ' exPl—(Tl ‘M W r )2] Jr I - 2L f’ (in. (3.19) "E 11,, ' agar“ erfc (-M\/1-D; ) The above equation includes heat losses from the flame to the regions above and below it. From Eqs. (3.18) and (3.19), and using B = fiM‘lP—rerfd-M‘JF; )exp(M2Pr), one obtains gPRE/gm (same as Eq. (3.15)). Still unknown are the fi'actions of the total flame heat release that appear as heat fluxes to solid surface in the pyrolysis zone (g‘py/gTOT), in the under-flame preheat zone (gap @701). and in the preflame preheat zone (gnu/gm;- ). The ratio g'py/g‘mT is found by using Eq. (3.18) and evaluating fi- at n=0. Also, only the heat loss below the flame is considered (heat flux reaching the surface). Thus, 84 g-PY = I315 —B . g (3.20) 870T From Eqs. (3.13), (3.17) and (3.20), using gTOT=§py+§UF+§PF, we obtain fup/irm arid EFF/8101'- These results are tabulated in Table 3.1 as a function of free-stream oxygen mass fraction. The dependent variable (Y0 ..) is in column one. Column 2 represents the excess pyrolyzate (fraction of total mass loss). It decreases with increase in Y0 ,,. This suggests that the flame length decreases as the oxygen mass fraction increases. For higher oxygen mass fi'actions the fuel does not have to travel far ahead in order to find oxidizer. For air we found the excess pyrolyzate to be 0.586. Pagni and Shih (1976) found it to be 0.57. Since their analysis used the boundary-layer momentum equa- tions, this result is yet another indication that the Oseen flow approximation is indeed a good approximation for the wind-aided flame spread problem. Column 3 presents the ratio of the flame heat release in the pyrolysis zone to that in the preheat zone. For lower oxygen mass fractions the flame heat release in the preheat zone is larger. For 100% oxygen flow, both zones are equally important. Column 4 to 6 presents the fractions of the total flame heat release that appears as surface heat fluxes to the pyro- lysis region, the under-flame preheat region and the preflame preheat region, respec- tively. It clearly indicates that the under-flame preheat zone is the most important as far as preheating is concerned. Since the flame is flat and relatively close to the surface, it could be assumed that the influences of radiation are limited to the under-flame zones. The streamwise radiation flux can probably be neglected~ (Beier, 1982). 85 Table 3.1 Ratios of the heat released by the flame, or received by the surface in each zone, to the total heat released, as a function of Y0... Here the is the excess pyrolyzate, 711,, is the total mass lost from the solid, g" and gm are the total flame heat release in the pyrolysis zone and the preheat region, respec- tively, and g'py, gm. and g'pp are the total surface heat flux in the pyrolysis zone, the under-flame preheat zone and the preflame preheat zone, respec- tively. The quantity gm, is the total heat released by the flame. You file/"3p . gray/8m fry/firm Eur/8701 in: @101 0.232 0.586 0.706 0.123 0.824 0.053 0.331 0.578 0.730 0.123 0.810 0.067 0.383 0.570 0.754 0.123 0.802 0.075 0.432 0.565 0.770 0.123 0.794 0.083 0.610 0.542 0.845 0.123 0.765 0.112 1.000 0.499 1.004 0.123 0.711 0.166 86 Finally we examine the connection between our heat-flux distribution (Fig. 3.3) and certain previously - assumed distributions. A nondimensional plot of the heat flux at a fixed streamwise location versus time for our model is shown in Fig. 3.63. Quali- tatively it compares well with the assumed distribution of Annamalai and Sibulkin (1979a), there being only two minor differences. The first is that their heat flux jumps from zero to a constant under-flame value; we have a non-zero, and in many cases non-negligible, preheat zone heat flux, depending upon the value of a. Second, they assume that at xp the heat flux drops to a new constant value, a phenomenon explained as heat-flux blockage by volatile fuel vapors. This process has not been explicitly included in our model, although it appears implicitly in the mass flux factor containing the B-number (see Eq. (3.1)). The heat flux distribution of Fig. 3.63 also compares well qualitatively with that of Quintiere et al. (1986), although they study acceleratory upward flame spread. Nevertheless, they find that pre-flame preheating is negligibly small, which corresponds to the limit a -9 co in Eq. (3.13). It is also encouraging to observe that their surface heat-flux measurements correlate decently when plotted against x/xf, which is the same as E,/§f for fixed t. Although they did not use our flux function [given by ng/gTIxf = gx 3’2/gfxf3’2 (for fixed t)], they did extended their correlation up past the pre-flame zone (x Ixf >1), and also extended it downward, at least for PMMA, to 1:09 :0.03. We observe that a limited downward extension of the abscissa of our correlation from Fig. 3.5b to Fig. 3.53 is also possible, because (é—ép )/(§f -§p) = §I§f when §>§p; note, however, that the ordinate has changed from gvgféf to 815/ng3 3.5. CONCLUSIONS In this chapter the heat flux distributions occurring in idealized wind-aided flame spread have been examined. By numerically solving an integral equation [Eq. (3.3)] we were able to determine normalized (gr (5,» and physical (g (§)) heat fluxes, as well 87 as the surface temperature, 123(5). We then eliminated the Y 0,, -dependence from the fluxes, producing the analytical correlations of Figs. 3.53,b (see Eqs. (3.10) and (3.11)). The pre-flame surface flux was found to decay algebraically, not exponentially (see Eq. (3.11)). The correlations were then used to derive expressions for the total heat fluxes to the pre-flame and under-flame regions; their ratio is given by Eq. (3.13), which shows that as o = g, lép —) co the heat flux from the flame to the surface occurs primarily in the under-flame preheat zone. It was found that the excess pyrolyzate is nearly 50% of the total, and that it decreases with increase in the free stream oxygen mass fraction. An energy analysis suggests that most of the preheating occurs in the under-flame preheat zone and that nearly 80% of the total heat produced by the flame falls onto the sample in this zone. This fraction also decreases with increase in the free stream oxygen mass fraction. These results are in direct contrast to the experimental measurements (Mekki et al., 1990; Chapter 4), which show that the flame tip and the pyrolysis front are closer than predicted, and that most of the preheating of the surface occurs in front of the flame tip, in the pre-flame preheat zone. There are several reasons for this contrast, one of which has already been briefly discussed, namely the difference between the theoretical mass flux profile in the pyrolysis zone and the "actual" (experimentally measured) profile. The latter never becomes as large as the former, so that much less pyrolyzate is released for gas-phase combustion in the experiments. This produces shorter flames, less under-flame preheating, and closer flame and pyrolysis fronts. It suggests that putting w=M\5t- is an idealization that is achieved for PMMA only for very large burning times [which are beyond the capacity of current measurement equipment to study (Mekki, 1991)]. Consequently, if we define an "x‘m-solid fuel" as one that quickly establishes the self-similar w=M ‘1; pyrolysis-zone mass-flux distribu- tion we can safely conclude that PMMA is not an "ideal x‘m-solid fuel". This does not mean that no such material exists, or that materials do not exist that approximate 88 better this dependence; it only means that PMMA, the material chosen for the particu- lar set of experiments of Mekki et al.(l990) (Chapter 4), does not show this ideal behavior. This observation may lead to significant engineering insight when we ask the following question - what characteristic features of PMMA (and similar solid fuels) produce the discrepancies between the theoretical and practical results? Once these features are known and understood one may design materials that maximize the non- ideality; one may produce materials that exhibit slow spread rates, short flame overhang, low volatile release rates, etc. This is a desirable fire-safety design goal. It may be necessary, for other applications (not Fire Safety) to design materials that optimize the spread rate; in this case one would try to exploit as many features of the "ideal x’m-solid fuel" as possible. Another reason for the discrepancy is finite-rate chemistry. In the actual case there is never flame reattachment; the flame always extinguishes before it once again touches the surface. As discussed in Wichman and Agrawal (1991) (Chapter 2), under the correct conditions this can bring flame tip and pyrolysis front into rather close proximity. Consequently what was once an under- flame preheat flux now becomes a pre-flame preheat flux, thereby upsetting the balance given in Table 3.1. Finally, discrepancies also arise because our discussion of preheat- ing is limited to flame heat fluxes since "hot gas wash" and radiation have been ignored in both this and our previous analysis (Wichman and Agrawal, 1991; Chapter 2). The simplest way to treat "hot gas wash" (Carrier, et al., 1990) is to require that the incoming flow (across the leading edge at x=0) have a uniform initial temperatme T,- greater than ambient; for large vertical distances from the surface the temperature of the gas in the region 0 < é <1 will be T,- > T a» where T“ is the "ambient" or "room" temperature.4 Thus, depending on how much greater T,- is than T”, the surface can be significantly preheated. In the Steiner—tunnel apparatus this preheating is important. 4 Our analysis considers the simplest case, T,- = T”. 89 For this reason, the pioneering study of Carrier, et al. (1980) (whose motivation was to develop a detailed theoretical model of the Steiner test) has an extended discussion of "hot gas wash" preheating effect. Despite these limitations we conclude by making an observation that nearly all of the heat transfer to the sample occurs in front of the pyrolysis zone. Our model sug— gests 88%, of which the under-flame surface receives the far greater fraction, between 71% and 82% of the total. Since for high oxygen concentrations actual laboratory- scale flames are quite luminous it seems appropriate in future work to focus attention on the influences of soot and .flame radiation on wind-aided flame spread. In next chapter detailed experiments are performed in order to determine the pyrolysis front speed and the flame front speed. The solid behavior is analyzed through species measurement. CHAPTER 4 AN EXPERIMENTAL INVESTIGATION SYNOPSIS This chapter presents the results of 3 detailed experimental investigation on lam- inar forced flow wind-aided flame spread over PMMA in the ceiling configuration. The speed of propagation of the pyrolysis front and flame front and the production rates of major chemical species are measured as a function of time. The objectives are to study the dependence of the flame and pyrolysis fi'ont speeds on the free stream velocity and the oxygen mass fraction and to infer the local fuel pyrolysis rates from the measured production rates of major chemical species. 90 91 4.1. INTRODUCTION This work presents a detailed experimental investigation of laboratory-scale lam- inar wind-aided flame spread along a ceiling-mounted PMMA slab. The gas flow along the slab is forced and its speed and composition are controlled. Apart from the ASTM E-84 tunnel test (Anonymous, 1987), most experimental investigations of wind-aided flame spread are for the buoyancy-controlled upward flame spread (Quintiere et al., 1986; Saito et al., 1986; Kulkarni and Fisher, 1988). However, in these investigations, independent control of environmental variables such as the ambient gas speed. gas composition and external radiation was not possible. This capability has proved very helpful in identifying the controlling mechanisms of wind-opposed flame spread (Fernandez-Pello and Hirano, 1983). For wind-aided flame spread, there is only one such detailed experimental study on PMMA (Loh and Fernandez-Pello, 1984). Several theoretical models of wind-aided flame spread over non-charting solids have been presented in the literature (Fernandez-Pello, 1979; Annamalai and Sibulkin, 1979b; Carrier et al., 1980; DiBlasi et 31, 1987). Recent models for non-charting materials (Chapter 2, Carrier et al, 1990) have sparked a considerable controversy since they seemingly disagree with the detailed experimental study of Loh and Fernandez-Pello (1984). The models suggest that the speed of the pyrolysis front varies linearly with the free stream oxygen mass fraction, whereas the experiments suggest a quadratic dependence. This clearly points out the need for further experi- mental investigation to determine which of the many assumptions made during the development of the models or possible experimental errors are responsible for the discrepancy. Thus, the objectives of this work are: (i) To provide a physical under- standing that will serve as the basis for the development or refinement of theoretical models (Chapter 6); (ii) To provide additional measurements for non-charting materials like PMMA. 92 4.2. EXPERIMENTAL SETUP A schematic of the experimental apparatus is shown in Fig. 4.1. It consists of three main parts: (i) A small-scale combustion wind tunnel which is capable of produc- ing a desired external radiant flux on the sample surface along with a flat-plate boun- dary layer flow at a given free-stream velocity and composition. (ii) A set of continu- ous gas analyzers for measuring the depletion of oxygen and production of C0 2, C0 , H 20 and total unburned hydrocarbons. (iii) Data acquisition equipment to collect, store and process the data. This apparatus was designed by Dr. Arvind Atreya and was built under his guidance by his previous students. 4.2.1. COMBUSTION WIND TUNNEL The combustion wind tunnel is the heart of this apparatus. It consists of three main sections, namely inlet, test and exhaust sections. Gases are supplied to the inlet section by two sources: (i) pressure-regulated laboratory air supply, and (ii) high pres- sure gas cylinders (02 or N 2). The gas composition and speed are controlled and metered by critical (sonic) nozzles. The resultant mixture enters the tunnel test section through a large settling chamber and a set of flow conditioners. 4.2.1.1. Inlet Section The desired air flow and composition results from the mixture of air and oxygen (or nitrogen) after flowing through sonic orifices, as illustrated in Fig. 4.1. The air supplied by the building compressor first enters a large tank in order to dampen the mechanical oscillations. The air then flows through a sonic orifice, of known diameter, at a certain upstream pressure. This pressure should be higher than the critical sonic pressure, whereas the downstream pressure is nearly atmospheric. However, oxygen and nitrogen are supplied by pressurized gas cylinders to a large tank before flowing through another sonic orifice. When the nozzle diameter, upstream pressure and flow 93 RADIANT HEATERS Moewples Heat Flux Gages DWI-Observation Window PAoProbe Acme: TM-Wenee Mampulators HL-Hemd Line SN-Sontc Neale Figure 4.1 Schematic of the experimental apparatus 94 density and temperature are known, the mass flow rate can be computed as (Holman, 1984) . 28c Y 2 7'1 = e 4. m A” P‘VRTI y+1[y+l] ( 1) where 77': = mass flow rate (lbm/sec) AN = area of the nozzle (in 2) P,- = inlet static pressure (psia) gc = gravitational conversion factor (32.2 lbm.ft/lb.sec 2) R = gas constant (ftlbf/lme) T,- = inlet temperature (R) 7 = ratio of specific heats of the gas (op/cw) The sonic nozzles were calibrated using a tracer-gas technique, where a known flow rate of methane is introduced downstream of the orifice, and the mole fraction of the methane-air mixture was measured. The discharge coefficient CD of each nozzle was carefully determined from the flow measurements and was found to be 0.97 (Abu-Zaid, 1988). The mixture flows through a 20 foot long, 3 inch inside diameter tube before entering the settling chamber through 8 pipes, which branch out of the manifold at equal angles. 4.2.1.2. Turbulence Manipulation Section The flow exits the settling chamber with large-scale and small-scale eddies. Hence, the flow is first conditioned in the turbulence manipulation section before 95 running through the test section. Combinations of honeycombs, glass beads and fine screens were tested for a Blasius type boundary layer flow in the test section. The honeycombs are used to damp the large scale eddies and the fine screens are used to dissipate the small-scale eddies. A two-inch space full of glass beads provides an inlet uniform flow. The high velocity near the walls of the settling chamber was decreased using a screen frame. Final results for velocity distributions were very close to the Blasius boundary-layer over 3 flat plate. Details are described in Mekki (1991). 4.2.1.3. Test Section The usable portion of the tunnel test section is 0.81 m long and 0.153 m wide. The test sample (.76 m long, .076 m wide and .0127 m thick) is placed horizontally along the tunnel ceiling while the bottom of the. tunnel is hinged at the inlet, making the tunnel depth adjustable. The tunnel depth at the inlet is 0.1 m and that at the exhaust end can be increased to 0.13 m . This provides a maximum 30% increase in the cross-sectional area at the exhaust end to compensate for acceleration of the gas core because of boundary layer growth and gas expansion due to heat release. The damper on the exhaust fan and the exit tunnel depth were adjusted to provide atmos- pheric pressure in the tunnel test section to within l><10"4 torr. This was necessary to prevent gas leakage in or out of the tunnel for chemical measurements and also to maintain a nearly constant free stream velocity. A maximum 10% increase in the free-stream velocity at the exit was observed. To further reduce the effect of variation in the free-stream gas velocity, data for only the firsr 0.5 m were used. Measurements of the velocity profile inside the tunnel showed that the flow was nearly laminar. Also, the RMS level of turbulent fluctuations was found to be less than 1% of the free-stream velocity. External radiation on the sample surface can be provided from below by an electrically-operated radiant heater with a water-cooled shutter (not used in the present 96 study). The hinged bottom portion of the tunnel contains a 0.153 m wide and 0.76 m long infrared optical glass window which transmits about 70% of the infrared radiation to the sample surface. The heaters were arranged such that the radiation measured at the sample surface was uniform to within :1: 3% over the entire length. The entire tun- nel test section was maintained between 315 K and 335 K by cooling water. 4.2.1.4. Exhaust Section The stratified products of combustion are mixed in the electrically-heated exhaust section of the tunnel in order to obtain a representative gas sample for transient chemi- cal analysis. The chemical measurements, along with the measured mass flow rate inside the tunnel, yield the production and destruction rates of chemical species. These data were then reduced in terms of the mass production rate of the species at the instant they were produced. A detailed description of the gas analysis equipment along with the time lag and response-time corrections is given elsewhere (Atreya, 1983). 4.2.2. GAS ANALYSIS EQUIPMENT For the species concentration measurement, a constant flow rate of a representa- tive sample from the tunnel was supplied to the gas analyzers by a metal bellow vacuum pump. To reduce errors due to condensation, the lines running to the total hydrocarbon [THC] and H 20 analyzers were heated by either electrical tape or hot water as shown in Figure 4.2. The gas was then passed through a cold-trap at -—5° C and dried before passing through the 02 and the C0 -C02 analyzers. Prior to every experiment, the gas analyzers were first zeroed with nitrogen gas, then adjusted to the proper reading for the known concentration of every specie of the calibration gas. 97 .................................................................. Dew Point Meter Heated line CO-CO 2 Analyzer Flow meter Figure 4.2 Schematic of the gas analysis equipment 98 4.2.2.1. Total Hydrocarbon Analyzer The total hydrocarbons were measured by 3 Flame Ionization Detector (FID) gas chromatograph GC —3BF , which used a 40%1-1 2—60%He mixture for fuel. The FID had a very good response time and a time constant of 1.0 second. 4.2.2.2. Water Analyzer The water vapor concentration history was measured by a condensation Dew- Point hygrometer (General Eastern 1200APS ). The dew-point temperature was meas- ured by optically detecting the condensation on a temperature-controlled mirror sur- face. The instrument had an accuracy of i0.2°C and a time constant of 1.0 second. 4.2.2.3. CO-C02 Analyzer An infrared IR 702 nondispersive dual gas analyzer was used for C0 and C02 concentration measurements. The flow was dried before entering the analyzer. The meter had a good accuracy of :l:I% of full scale (CO : 0-3%, 0-12%; C02 : 0-6%, 0— 20%) and a time constant of 2.6 seconds. 4.2.2.4. Oxygen Analyzer The oxygen concentration was measured by a Beckman 0M -11 polarographic analyzer. This meter also required dry flow. The meter had an accuracy of i0.l% and a time constant of 1.5 seconds. More details regarding experimental setup are described in Abu-Zaid (1988), Nur- bakhsh (1989) and Mekki (1991). 99 4.2.3. DATA ACQUISITION The analog signals from the thermocouples, gas analyzers and the radiometer were fed to a data acquisition/control unit (HP 3497A). The data were then acquired by an HP 486 personal computer using LabWindows software (National Instruments). The handshaking commands and data transfer between the computer and the HP data acquisition unit were assured by the use a GPIB card with a standard IEEE 488 inter- face cable. The data were taken at 12 readings per second with 0.1% accuracy. 4.2.4. SAMPLE PREPARATION The samples used for these experiments were obtained from white sheets of PMMA. These were instrumented with nine thermocouples on the surface 0.05 m apart (Chromel-Alumel; 76 um diameter). The first 0.15 m of the sample was not . instrumented to avoid errors due to ignition. 4.2.5. EXPERIMENTAL PROCEDURE Once all the desired conditions were set, the sample was ignited with a small methane porous-metal burner placed at the tunnel inlet with its face parallel to the sample surface as shown in Fig. 4.3. The fuel flow rate to the igniter was controlled such that the flame overhang on the sample surface was about 0.02 m . The data were collected by the computer-controlled data acquisition system and the flame spread pro- cess was recorded by a video camera. 4.3. RESULTS After ignition, the physical process that occurs inside the tunnel is schematically Shown in Fig. 4.3. Here, the fuel vapors generated in the pyrolysis region, A which extends from x = 0 to x = x are burned in the diffusion flame, which extends from p! x = 0 to x = xf, with xf > x The hot combustion products that flow downstream p0 100 X PMMA sample ‘ p : gniter l f ‘ Xf Flame : U <—— Gravity 4— Water cooled plate Figure 4.3 Schematic of wind-aided flame spread in thetunnel over 3 ceiling-mounted sample of PMMA 101 of xf , and the flame extension (xf — xp ), help to convectively and radiatively heat the pristine solid to a temperature (TV) at which it begins to vigorously pyrolyze and con- tribute fuel to the flame. Thus, the flame spread process consists of the spread of the pyrolysis front. Clearly, the rate of flame spread will depend on how fast the surface temperature of the solid is raised to its pyrolysis temperature. This, in its turn, depends on the radiative and convective heat flux from the flame. The convective heat flux is proportional to the gas phase temperature gradient at the surface, which depends on both the boundary layer thickness and the flame temperature. Thus, the two most important variables that affect this mode of flame spread are the wind speed, which controls the boundary layer thickness, and the oxygen concentration, which controls the flame temperature. The results of a systematic study of these two variables are described below. 4.3.1. TEMPERATURE MEASUREMENTS Figure 4.4 shows the measured surface temperatures for PMMA as the flame pro— pagates along the ceiling-mounted sample. These temperatures are typical of all our measurements. Measurements for PMMA show a plateau at about 643 K, which is taken as the melting or vaporization temperature. Figure 4.4 also shows the rate of change of surface temperature for PMMA. From visual observations it was found that the peak rate of change in the surface temperature after ignition occurs at the instant the flame tip arrives at the thermocouple location for both wood and PMMA. By writing a surface energy balance it can also be demon- strated that this peak corresponds to a sharp increase in the incident heat flux which is caused by the flame tip arrival. Such peaks are shown in Fig. 4.4. Thus, the flame tip location, xf (Fig. 4.3), may be determined as a function of time from the measured temperature profiles by calculating the maximum value of dis ldt . Results of such cal- culations show excellent agreement with 1:, determined from video records (Fig. 4.5). 102 1000 20 A 800- '10 M v a 600-1 '12 D E u E 400-1 '8 E13 . [-0 . 200- -4 0 . r .‘ , . . . o 0 100 200 300 400 500 600 700 800 900 1000 TIME (see) Figure 4.4 Measured surface temperatures and their rate of change for PMMA during flame spread (U... = 0.9 m/s, Y0, = 0.331). Here surface temperatures are plotted for different streamwise locations on the sample dT/dt (K/s) 103 The pyrolysis front location xp is found by defining a constant surface temperature T, at which the solid begins to vigorously pyrolyze and contribute fuel to the flame. For PMMA this is the "vaporization" temperature (643 K). Unfortunately, at these tem- peratures the thermocouples often detach from the surface due to melting. 'Ihus, opera- tor judgement is required to determine xp . To eliminate this difficulty, A: between xf and xp was consistently determined by (T, - Txl)/(d1'sldt)mu. An example of this method of determining xf and xp is illustrated in Fig. 4.4. It is also important to note that the surface temperatures at xf are much higher than ambient. They range from 533 K to 573 K for PMMA. This implies that most of the temperature rise has occurred in the preheat zone ahead of the flame. This observation is in sharp contrast with the previous assumption (Fernandez-Pello, 1979; Annamalai and Sibulkin, 1979b; Loh and Fernandez-Pello, 1984) where the surface temperature is taken as ambient until the arrival of the flame tip. 4.3.2. FLAME SPREAD RATES Once xf and xp are determined as a function of time, the pyrolysis-front speed, Up (defined as dxp Idt), and the flame-front speed, U f (defined as dxf ldt), are obtained from the slope of the least-square-fit line. Data for only the first 0.5 m were used to minimize errors from changes in the free stream velocity (U m). In the first 0.5 m of the sample, both xp and xf increase nearly linearly with time, whereas in the last 0.15 771 they exhibit slight acceleration (Figs. 4.5 and 4.6). Figure 4.7 shows Up and U, plotted against the free stream velocity; it is seen that Up and Uf increase linearly with U .. Figure 4.8 shows Up and U f plotted against the oxygen mass fraction (19.). Note that several different measurements of U p and U f for PMMA (from reflective to blackened water-cooled bottom aluminum plate) are presented. It is also noteworthy that Up and Uf are nearly equal (see Figs. 4.7 and 4.8). 104 . l . . o from Temp. - - from VCR - FLAME FRONT LOCATION. xf (cm) v ( . e 23 _ v 3 .33 o .9 .33 ‘ A .3 .43 - D .9 .61 . I o .8 1 . . . . , . . . 1 . . . 800 1200 1600 2000 TIME (sec) 1. Figure 4.5 Flame-front location as a function of time. Note the slight acceleration for x, 240 cm . Also note the negligible error [O(mm)], encountered in deter- mination of flame front location using VCR 105 60 50-1 l ' I ' O 1000 2000 3000 4000 TIME (see) Figure 4.6 Pyrolysis-front location as a function of time 106 O '6 4E-04" ~ 6 .. . 3E-04-1 -i 3E-04-i - SPREAD RATES (m/s) (0 ‘7’ O .5 1 1 2E-04 - t . l . j . I 0.4 0.8 1.2 1.6 2.0 ' FREE STREAM VELOCITY U,° (In/s) Figure 4.7 Dependence of the pyrolysis and flame front speeds on the free-stream velocity 107 6E-03 . I I I I I I I )K Reflective Aluminum Plate o—o Dull Aluminum Plate . . 1:1 Blackened Aluminum /’ . A Blackened Water-cooled /" -- Loh at Fernandez-Pello (1984) 413-034 - - Theory Eq (2 32) )K B - SPREAD RATE (m/s) PMMA 0.2 ' 0:4 ' 0:6 2 0:8 1:0 OXYGEN MASS FRACTION You. Figure 4.8 Dependence of pyrolysis and flame front speeds on the free-stream oxygen mass fraction (blank symbols represent Up and crossed symbols represent U,) 108 Figure 4.8 also shows that Up for PMMA depends upon the surface finish of the aluminum plate directly under the burning sample. The aluminum plate was used only for PMMA since external radiation was not required. As is evident from Fig. 4.8, the surface condition of this plate (and to a lesser degree, the two vertical sides of the tun- nel that contain the observation windows) considerably alters the flame spread rate because it reflects the flame radiation back to the sample surface. This effect magnifies as Y 0.. increases because the flame radiation increases. Measurements show that the flame radiation at the plate surface increases linearly from zero at Y0. = 0.2 to 3.5 W/cm2 at Y0. = 1. Thus, for a reflectivity as small as 0.1, the reflected radiation becomes substantial for higher oxygen. The spread rate measured with a reflective aluminum foil for Y0... = 0.6 is 2.7 times the spread rate for blackened and water- cooled aluminum plate and 1.6 times the spread rate for a dull aluminum plate. Thus, the flame spread rate depends upon the reflections inside the tunnel and the result closest to the truth is that of the blackened and water-cooled plate. For PMMA, these measurements show that U p ~ Y 01;? . This differs from the previous measurements for PMMA (Loh and Fernandez-Pello, 1984; dotted line in Fig. 4.8) which show that U, ~ 1’}... However, the present results for a dull aluminum plate agree well with those of Loh and Fernandez-Pello (1984), indicating that reflection of radiation inside the tunnel may be the reason for the disagreement between theory and experiment. Results for a water-cooled plate agree well with the recent theoretical predictions for PMMA (Chapter 2), which are shown as dashed line in Fig. 4.8. 43.3. SPECIES PRODUCTION RATES Figure 4.9 shows the species production rates for PMMA plotted against the pyro- lysis length xp. Production rates of C0 and total unburned hydrocarbons are not presented because they are two orders of magnitude smaller than the others. The data for small xp are omitted since the measurements are affected by the igniter flame and 109 100.01 ' .1 300-: 1 ’a‘ 1 1 \ 40.0- - M - . v N . O '7 10 0- X 320-5 -‘ 0 '1 I '35 40-? ‘ 0: ' . -. U) m . 1 c0 _____ a .......... 1.01 .......... 1 03‘: 1..»2" - - 02-depl. -; I ' I r _IC012 I I ‘ 10 20 40 60 30 100 . Pyrolysis Front Xp (cm) I Figure 4.9 Dependence of the species production rates on pyrolysis front-location (U... = 0.9 m/s; 1— Y0” = 0.23; 2 — Y0... = 0.43; 3 - Y0, = 1.0) 110 henCe they are subjected to a large percentage error. Also, the soot production rate was not measured even though a substantial amount of soot was formed, especially during experiments at high Y 0 ..- Species data were collected even after the flame had spread over the entire sam- ple, i.e., in the boundary-layer burning zone. These data are presented only for Y0... = 0.43 (for purposes of clarity). Vertical straight lines are obtained (see Fig. 4.9) because 1:, does not change during this period since the pyrolysis front has already reached the end of the sample. For PMMA the mass production rate more than dou- bles during this period before achieving a steady-state value. The fact that the species production rates continue to change substantially even after- the pyrolysis front has reached the end of the sample shows that steady-state mass loss is not achieved in the boundary layer burning zone. For complete combustion, the production rates of C0 2 and H 20 and the deple- tion rates of 02 are directly related to the total fuel production rate between 1: = 0 and x = x For the local mass flux to vary as x‘05, the total fuel production rate any I" - 1 instant must vary as xp ’2. Incompleteness of combustion will only serve to further reduce this power. However, the data for PMMA show that this power is greater than 0.5 and closer to l; for PMMA at Y0... = 1.0 it is even greater than one. Also note that the production rate of H 20 (rhflzo) follows the depletion rate of 02 (77°10 2) for all the experiments. However, the C0 2 production rate (rhea 2) is slower. For Y0... = 0.23, rilco2 > 75102, whereas for Y0... = 1.0, n'tco2 < ”'10: for PMMA. Crossover between film2 and 77102 can be seen at xp = 0.5 m for PMMA (Y0... = 0.43). This indicates that the chemistry is changing as xp and Y0... are increased. Carbon in the fuel is converted to unburned soot leading to lower Inc-02. This agrees with the physi- cal observation that both larger flames and flames at higher Yo... are brighter. 111 4.4. DISCUSSION It is interesting to note that there are several areas of agreement and disagreement between these experimental results and the recent theoretical models developed for non-charring materials (Chapter 2, Carrier et al., 1990). As predicted by previous theories (Chapter 2), both Up and U f are found to vary linearly with U .. (Fig. 4.7). However, these theories also predict that Uf is significantly larger than Up , which is in contrast with the experimental results (Figs. 4.7 and 4.8), which show that U, 2 Up. Likewise, xf is found to be only slightly larger than xp (between 5 and 10%) for PMMA, regardless of the free stream velo- city or the oxygen mass fraction. It clearly suggests that the flame reattachement point (treated as location of xf in chapter 2) can not be taken as flame tip location. This discrepancy arises because the previous theoretical models utilize the steady-state "Emmons" (1956) solution in the boundary layer burning zone. It will be discussed in much more detail in the next chapter, and is also the basis for that study. As dis- cussed earlier (Fig. 4.9), the burning zone is unsteady in the solid phase during the flame spread process. This leads to lower instantaneous fuel mass production rates, resulting in smaller-than-predicted values of xf and Uf. The case for an unsteady burning zone is further supported by the fact that the species production rates vary roughly as xp rather than as xp05. As is evident from Fig. 4.8, the theoretically predicted (chapter 2) dependence of Up on Y0... agrees well with that determined experimentally. An increase in Y0- increases Up , primarily by increasing the heat flux from the flame to the as-yet- unburned solid surface. This heat flux increases because the flame temperature increases in proportion with Y0... in the absence of flame radiation and incompleteness of combustion. This heat flux also decreases because of increased shielding of the fuel surface from the flame by the evolved fuel mass flux (Note that the fuel mass flux increases roughly proportionally with Yo n; Fig. 4.9). As discussed earlier, the previous 112 theoretical models overestimate the evolved fuel mass flux in the burning zone by using a steady-state solution there. This results in an overestimation of the heat block- age factor. It appears to be compensated by overestimatin g the flame temperature, by neglecting flame radiation and incompleteness of combustion. 4.5. CONCLUSIONS Several significant conclusions can be derived from the experimental results presented here: (i) The pyrolysis-front and the flame-front speeds for PMMA vary nearly linearly with the free stream oxygen mass fraction and the free stream velocity. (ii) The pyrolysis-front and the flame-front are much closer to each other than predicted by recent theoretical models. This is also true for the pyrolysis-front and the flame-front speeds regardless of the free stream velocity and/or the oxygen mass frac- tion, primarily because unsteady conditions exist in the burning zone. (iii) Species measurements suggests that the pyrolysis mass flux is roughly constant for PMMA, rather than varying as x'o's. Although these experiments bring several interesting facts to light, it is worth remembering that the theoretical models were constructed in the absence of informa- tion. And despite their limitations, the steady-pyrolysis-zone models comprise self- contained and self-consistent descriptions of the spread process. They prOvided valu- able clues for the subsequent experimental work. More importantly, as with all "clean" models, these descriptions (Carrier et al., 1990; Wichman and Agrawal, 1991; Chapter 2) provide an orderly basis for thinking about the more cOmplex problem. In sum, the theoretical models for non char-ring materials predict the pyrolysis front speed quite accurately despite the steady-state solid-phase assumption in the burning zone. This underscores the fact that the flame spread rate depends primarily on local heating of the solid by the flame tip in the adjacent preheat zone. 113 In next chapter we investigate the transient behavior of a vaporizing solid. An expression is developed for instantaneous mass loss which together with the present conclusions forms the basis for Chapter 6. CHAPTER 5 THERMAL DEGRADATION OF A VAPORIZIN G SOLID SYNOPSIS This chapter presents a simple model for the gasification of a vaporizing solid under an external heat flux. An expression is obtained for the transient mass flux as a function of the net heat flux. This expression was corroborated by transient pyrolysis experiments conducted on PMMA for heat fluxes ranging from 1.6 to 5 W/cm2 in nitrogen atmospheres. 114 115 5.1. INTRODUCTION The previous chapter has demonstrated that the thermal degradation of a solid plays an important role in fire development. During early stages of fire growth, solid pyrolysis is transient, and the flame length (height) depends on the availability of fuel. To incorporate this feature into models of wind-aided flame spread (Chapter 6) or zone-type compartment-fire models (Peacock and Bukowski, 1990), simple and accu- rate expressions for the transient mass loss rate are needed. The present work is lim- ited to gasification of vaporizing solids like PMMA. A proper understanding of the rate of polymer gasification in a fire environment is essential to the mathematical pred- iction of fire growth on such materials. Depending on heating rates (exposure), two types of thermal degradation of PMMA can be observed (Khalturinski and Berlin, 1983) (i) isothermal bulk degrada- tion (chemical regime) and (ii) linear pyrolysis (ablation regime). Our interest lies in linear pyrolysis because the PMMA samples used in this study are thermally thick. Under these conditions a narrow layer near the heated surface liquifies when the sam- ple is exposed to a radiant flux. Small clusters of degradation products must escape from the surface. Since the boiling point of the degradation products is much less than the polymer degradation temperature and since a significant quantity of only partially degraded polymer is still available at the surface, the degradation products are superheated. Consequently they nucleate and form bubbles which grow by mass diffusion through the molten layer, mass vaporization and thermal expansion (Kashiwagi and Ohlemiller, 1982). Furthermore, because of the higher temperatures, the degree of superheating and bubble nucleation rate increase while surface tension decreases as the surface is approached. Transport of bubbles from the interior of the polymer towards the surface is also affected by the existence of viscosity gradients (Kashiwagi et al., 1990). In general the degradation of PMMA occurs according to the following main stages: depolymerization initiation, propagation of reaction chain and 116 termination. Depolymerization is believed to initiate at Chain ends, at random points along the chain, or at isolated " weak links" (Lengelle, 1970). Free radicals, formed during initiation, give rise to the so-called "unzipping", that is the propagation of the reaction chain. Chemical processes at this stage may be characterized by the length of the reaction chain, that is, the number of monomer units produced on average for one initiation. Termination reactions account for the stabilization of free radicals which occurs through the combination with an H atom. Modeling Studies of PMMA degradation can be divided into two main categories (3) Kinetic and (b) Thermal. Kinetic modeling can be further classified as (i) one-step global models which employ a one-step global Arrhenius-rate reaction to account for all chemical processes. Such an approximation has been used with the description of physical processes to model solid-phase processes (Krishnamurthy and Williams, 1973; Vovelle et al., 1987). (ii) Detailed degradation models where kinetic schemes, accounting for chain initiation reactions, depropagation reactions and termination reac- tions, have also been proposed (Kashiwagi et al., 1985, 1986; Hirata et al., 1985; Inaba et al., 1986, 1987). Such models have never been coupled to the description of physi- cal processes. Even the one-step global model of Vovelle et al., 1987 is too complex to incor- porate in the wind-aided flame spread problem which requires simultaneous solution of the gas and solid-phase problems. Vovelle’s model involves numerical solutions of partial differential equations; the model is transient and includes a one-step Arrhenius- type degradation reaction in the sub-surface regions that contribute to the gasification rate. What is needed is a thermal model which can provide an explicit expression for the transient mass loss. The model that now exists is too simple (Tewarson and Pion, 1976), suggesting gasification rates proportional to the net applied heat flux. Recently, Steckler et al. (1991) and Delichatsios et al. (1991) proposed a thermal model which assumes that gasification occurs only at the surface and at a constant "gasification 117 temperature". Thus the surface regresses at a constant temperature. Results were obtained by numerical solutions of ordinary differential equations. An extensive literature exists for solid-phase modeling (V ovelle et al., 1984a, 1984b, 1987; see the references therein and those cited above) but no direct expression is available for the transient mass loss. Thus, the objectives of this chapter are to pro- vide an explicit expression for the transient mass loss as a function of constant and variable applied heat fluxes. 5.2. GOVERNING EQUATIONS As shown schematically in Fig. 5.1, during heating the solid undergoes two different heating stages: (i) The inert preheating stage, in which the surface tempera- ture increases until it reaches the "vaporization" temperature, and (ii), the vaporization . stage, in which the solid surface regresses at a constant temperature, liberating fuel vapor. The assumption of a constant predefined "vaporization" temperature is widely used in solid modeling (Steckler et al., 1991; Delichatsios et al., 1991), see the experi- mental results presented in Fig. 5.2. Plotted are the surface temperature and total mass loss as a function of time for PMMA exposed to a radiant heat flux of 2.1 W/cm 2. The governing equation for one dimensional heat conduction during the preheat stage is 2 £31.91, 0o-, (5.1) dyz Otdt subject to the initial condition, T(y ,0)=T,., and the boundary conditions "k ['3] :‘qnet (t )=QHV _ qloss = QIN-es C(Ts4-To4o )-hs (T3 -Too)r y =09 t >03 (5'2) y=0 118 Surface Temp / Surface Location l 50) Preheating i Vaporization Time Figure 5.1 Schematic of the solid phase model 17 119 0.8 I I l 1 T r I l r 400 s ----------- - - — ----..I *3 i \. 0.6- : a-300 E : no . . . v l to : , 8 0.44 . --200 r—o I u : VJ - I .. ‘5 l E 0.2-i E 4-100 '8 . 4.) I O . : ., 0.0- . . , . . . . I . . , . . o 0 tv 400 300 1200 1600 2000 Time (sec) Temperature (C) Figure 5.2 Surface temperature and total mass loss as a function of time for PMMA. Heat flux = 2.1 W/cm2 120 In the vaporizing stage, the regressing top surface moves downward at a rate ds(t )ldt. Let tv represent the time when the surface starts regressing, i.e., when the surface reaches TV . The governing equation is d2T 1 dT .___-.——, r< <00,t>tv, 5.3 dyz a d: S( ) y ( ) subject to the interface condition: Jr qm(t) + k —— =pL d3 (' ) , T(s (t ),t )=T,, (5.4) dy y=s(t) d‘ where L is the. latent heat of gasification. These equations were solved both numeri- cally and analytically. FOr the analytical solution presented below, a semi-infinite solid was assumed with T(oo,t )=T.., where T g. is the "ambient temperature. However, for the numerical solution, a thick solid, with convective and radiative heat losses fi'om the bottom surface, was employed. 5.3. SOLUTION An extensive literature exists on the pyrolysis of "vaporizing" solids (Steckler et al., 1991; Delichatsios et al., 1991; Vovelle et al., 1987), but explicit expression for transient mass loss is not available. Recently Delichatsios et al. (1991) presented an ‘ integral analysis with an exponential profile, similar to the present work. This, how- ever, resulted in ODE’S and was solved numerically. Thus, a simple but fairly accu- rate integral model is developed here. The problem is divided into two parts, preheat- ing and vaporizing. 121 5.3.1. PREHEATING STAGE During preheating the surface temperature is a function of the applied heat flux, hence Eq. (5.1) must be solved along with the boundary condition (5.2). A very useful and accurate expression for the preheating time is obtained by Atreya (1983) for semi- infinite solids exposed to a constant heat flux. After adjusting the multiplicative con- stant the expression becomes: “:30ka [ 7 g t In (ZEN-VEXMB) _ 3112330 ’ (5.5) 6 e2 2 f2 83/2 (2§m+v'6xr—~IB) Bf where y=T,—T,., er=—(h McoT}, ), §=I-250’T.2./3, 046—43312) and f =(d;,/8mt§yz). Note that qw is constant here. Equation (5.5) gives Iv (time to reach T v) very accurately. In Fig. 5.3 results are plotted for temperature as a function of time. Numerical solution of Eq. (5.1) com- pares very favorably with Eq. (5.5). The sample thickness was taken as 0.127 cm. The other parameters used were (Vovelle et al., 1987) k=0.185W/mzs, Cps=2150KJIkgK , p=ll90kg /m3, T v=635K , Tg,=295K , L=l6501lgm, 8=0.92, and h=0.4 W/mzK was chosen to give (2,3,, = 0.82W/cm2. We will explain this choice later in the experiments section. For varying external heat flux Eq. (5.1) must be solved numerically because in this case Eq. (5.2) becomes a function of time. 5.3.2. VAPORIZING STAGE The integral method is used to solve Eq. (5.3). We choose TO .t')=f;expl-a (7)0 -s (2' ))1 (5.6) 122 TV. , ,,,, 300% tr: 2.9 ,x q"=1.err/cm2 “é ' m 2004 z I 1223 1 Pa :5 8 100 35. 0:. :3 V3 -- NUMERICAL 0 — APPROX l ‘ l ‘ l r I 0 100 200 300 400 TIME (see) Figure 5.3 Surface temperatures as a function‘of time, during preheating of the solid 123 where f=r-T,,, t'=t-t, and a (0)=rj,,';,(t'=0)/k7"', (from Eq. (5.4)). 5.3.2.1. In-depth Temperature Profiles at the Vaporization Temperature It is important to check that in-depth temperature profiles obtained by the integral method, at the onset of vaporization (i.e. surface reaches TV, Eq. (5.6) evaluated at tv) be very close to those obtained by the numerical solution of Eq. (5.1). Results are presented in Fig. 5.4. The match is reasonably good. It improves as the applied heat flux decreases. Moreover the transient effects are much more prominent at lower heat fluxes. 5.3.2.2. Transient Mass Loss Rate Integrating Eq. (5.3) from ymd‘) to y=oo gives ds/dt‘=[ti,{;,(2')-ka(t')f,]/pz. Using this in Eq. (5.4) leads to an ODE in a (t‘): 7 IO"— 59: qw( )a2 - [0. + —v J03. dt pL pL Here we have used Eq. (5.6) to evaluate (dT/dx)s(,). For the variable net heat flux, the above equation is solved numerically to give a value of a which can be substituted into the equation for d5 Id}- to give the transient mass loss. However, for a constant net heat flux the above equation can be solved exactly to give the transient mass loss rate as: --—_-, (5.7) 124 400 I I I I I I T I T If U l I . — NUMERICAL --~ INTEGRAL , 300-4 s.‘ .. “'5‘13? W/sz 200- --------- ..- - Q“.-. .‘~~.‘.‘~ J 1 00 q \‘s k..."' - \ \ ~ \ “ \ s ‘Q Q s... .Q IN DEPTH TEMPERATURE (C) 0 I l I I I l I I I I 0.000 0.002 0.004 0.006 0.006 0.010 0.012 0.014 DISTANCE FROM THE SURFACE (m) Figure 5.4 In-depth temperatures as a function of time, during preheating of the solid 125 where A (t‘)—1 exp[A (t‘) ]_ -Cp.é.;37 —— —exp _ Ao-1 A0 (PkLXL +Cp, To) The results are plotted in Fig. 5.5 for four different heat fluxes along with the corresponding experimental results. Equations (5.1) and (5.3) were also solved numer- ically to predict the surface temperature during preheating and the surface regression rate during solid vaporization. They are also plotted for the same heat fluxes in Fig. 5.5. The agreement is good, especially for higher heat fluxes. For a varying external heat flux (Fig. 5.6) the results are plotted in Fig. 5.7. Also plotted are the numerical and experimental results. The agreement is good. Numerical solutions where obtained by using a program developed by Nurbakhsh (1989) for moving boundary problems. In order to check the effect of sample thickness, the integral solution was com- pared with the numerical solution obtained for various sample thicknesses. The results are plotted in Fig. 5.8 for various constant external heat fluxes. As the sample thick- ness increases the numerical results approach the integral result. Depending upon the time duration of the experiment, a correct sample thickness can be chosen from the above graph to ensure that sample can be treated as thick. 5.4. EXPERIMENTAL SETUP Experiments were performed on PMMA under various heat fluxes to obtain its degradation characteristics. These were done in a controlled nitrogen environment cone heater apparatus shown in Fig. 5.9. Although it has been shown that mass loss rate for PMMA is different in oxygen environments (Kashiwagi and Ohlemiller, 1982), nitrogen was used because in actual fire situations there is no oxygen between the MASS LOSS RATE (lrg/nn2 sec) 126 200 I 400 ' 600 I 600 ' 10'00 ' TIME (see) Figure 5.5 Mass loss rate as a function of time 12'00 ' 1400 127 4IIIII7IIIIIIIITI A —HeetFlux m d 8 1 { 3-1 .. E __ N. .2 ISL. 4.32- 2 .. a! Q) m d '73 5:1- 1 - I... 0) 4.: x q LII-J OIIIIEIIIIIIIIIIII 0. 200 ,400 600 800 1000 1200 1400 1800 1800 TIME (see) Figure 5.6 Varying external heat flux 128 0.0016 . G—O E'xptil . A-A Exptfiz q — Integral 0.001%. -- Numerical .1 0.0008-1 Mass Loss Rate (gm/cmzsec) ’ I I I I ' I I I fl I ' 0 200 400 600 800 1000 1200 1400 1800 TIME (see) Figure 5.7 Mass loss rate as a function of time under varying heat flux 129 I I 1 ' I ’1 I I ’0‘ I x . an we we 2.3 mm“ I —- i=.766m o _ - NW q"=5W/cm2 , _ $3111 8 0-002‘ ~--- acclaim)“ Q — Integral 8 no .t v m I I ’ - \ .... e E. PI’T-fl—‘F -_ ........ .. . é‘ a 0.001 ! - SW/cmz - (I) I l S 2 g 1.5W/cm 0.000 t . ’ . . . . . . . r 0 1000 - 2000 3000 4000 5000 TIME(sec) Figure 5.8 Effect of sample thickness on mass loss rate 130 Cone Heater a Water cooled cover / Viewing window RN l l\\\ \ I. N 2 Water cooled Sample , $ housing Thermocouples ~\\\‘ \ ‘V w \ § § s s s ow \ § .0 straightners \ \ \\‘ _\ \ \ ——s s . Si \ m '////////////////////////////////////' \\\\\\\\\V [/5 7/////////// Load cell Y-Traverse ’//////////////////. Pressurized chamber Figure 5.9 Schematic of the apparatus Sonic . nozzle 131 diffusion flame and the pyrolyzing surface. During operation, the top part, which houses the cone heater, is bolted to the bottom part, which houses the sample, weigh- ing scale and air flow straighteners. The water-cooled bottom housing is connected to a pressurized chamber which is fed by two independent metered gas lines. A data acquisition system is used to collect, store and process the voltage outputs from the weighing scale and the thermocouples. Once the cone heater reaches the desired heat flux, the water-cooled plate is pulled out, exposing the PMMA sample which was 10 cm x 10 cm and 1.27 cm thick. The top surface of the sample was covered by a reflective aluminum foil with a circular opening (31 cm 2) in the center to eliminate edge effects. These samples A were subjected to four different (1.6, 2.9, 3.4 and 5 W/cm 2) heat fluxes. The results for the mass loss rate are plotted in Fig. 5.5. From the steady state value of the mass flux and the applied heat flux it was determined that surface heat loss (convection and radiation) was 0.82W/cm 2. (Orloff et a1. (1974) found a value of 0.86W/cm2). The literature (Vovelle et a1, 1987) value of surface emissivity of PMMA is 0.92. Thus, the convective heat transfer coefficient (h =0.4W/m 2K ) was calculated accordingly. For variable heat flux (Fig. 5.6) the results for mass loss are presented in Fig. 5.7 5.4.1. EFFECT OF SAMPLE THICKNESS AND IN-DEPI‘H ABSORPTION. To check for the sample thickness effect two samples were glued using acrylite glue making the sample twice as thick. The combined sample was then subjected to a high heat flux (3.5 W/cm 2). The mass flux remained within the experimental error. During this experiment the in-depth temperature profiles were also obtained. They match nicely with the numerical results obtained from Eqs. (5.1) and (5.3) (Fig. 5.10). In order to check for surface absorption a sample was painted black and was subjected to low heat flux (1.6 W/cm 2). The results obtained for the mass loss rate were identi- cal with the untreated, sample. Hence, it was concluded that for the white PMMA 132 500 T e 1 I . . . , . . — Expt. ‘ --- T6635 K 4004 TOP - - races 1: a far-{____________________:_,_ v - x I I 135' ' . I m 300-1 0' "5/ D ’9'” E . MIDDLE .,I A? La zoo-1 .x” s Er: ‘ ’/" E-l ,/ BOTTOM 100% ,v’ /' ,/'/ / d //" 4/ o - . , . . , . - ' . , . . . . . . p 0 400 800 1200 1600 2000 TIME (sec) Figure 5.10 In-depth temperatures as a function of time, during vaporization of the solid under q' =3.4W/cm2 133 sample used in this study, there was negligible in-depth absorption of radiation. 5.5. COMPARISON WITH THE LITERATURE As a further check of the transient mass loss rate model (Eq. 5.7), it was com- pared with the data of Vovelle et. a1 (1987). These comparisons are shown in Fig. 5.11. Vovelle et al’s results are also compared with numerical predictions, as shown in Fig. 5.12. For this case we chose L=10071 lgm as suggested by them. The decrease in the mass loss rate is due to sample thickness. Also, the predictions show some disagreement because of inaccuracy in predicting the back boundary condition. The predictions for bath cases agree well with the data. For varying heat flux we compare Vovelle et al. (1984a) with the integral model. The heat flux and mass loss rates are shown in Fig. 5.13. Clearly, for all cases the predictions of the integral model are encouraging. MASS Loss RATE (kg/m2 sec) 134 0.010-1 CI 0.008“ . 0.006-1 q 0.004~ ' I ' I ' I H Vovelle at 0.1.0987) -- INTGL. -- "’- ‘— a’ o‘ " ’ 400 ’ ” 600 800 TIME (sec) ' 1600 ----'----I 1500 ' 1400 Figure 5.11 Mass loss rate as a function of time. Comparison of integral model with the data of Vovelle et al. (1987). MASS Loss RATE (gm/cmzsec) 135 0.0016-i 1 0.00124 0.0008fi . , . , . . O Vovelle ct al.(1984b) - . -- Numerical q"=3.1 W/crnz r l l 000000 0002.5 . a .. o0 - .1 ~ °°-° . 1.7 °°o. ‘ I. 0. 000° 0 o 00 ‘ O O .=. 0 0° 0 — . 0° - . .. 00° 0 ONOQOOOOOONOOOOOQ 861’ 0 00 ° O O 1.4 / . r . w 1000 2000 3000 4000 TIME (sec) Figure 5.12 Mass loss rate as a function of time. Comparison of numerical results with the data of Vovelle et al (1984b). 136 A O ' 1 I l 3000 0 . 0 a 0 0.0016-1 0 ””2 0°° n a an ‘ _ mm :0 a " ”" maxim o I { - — Hoot Flux / a 0.0012: ./ -2000 ’3 :9 . ' . g o . V *3 0.00084 9 c: . ~ 8 % ~1000 F 3 03004-1 m . m d a . 0.0000 I , a we... 1 2 4 5 External Heat. Flux (W/cmz) Figure 5.13 Mass loss rate as a function of time. Comparison of integral model with the data of Vovelle et al. ( 1984a). CHAPTER 6 FLAME SPREAD OVER AN UNSTEADILY VAPORIZIN G SOLID SYNOPSIS This chapter presents a simple model for the flame front speed during wind-aided flame spread on a thick vaporizing solid like PMMA. Model predictions compare favorably with the experiments conducted in the ceiling configuration. Effects of the blowing, surface heat loss and the unsteadiness of the solid are analyzed. 137 138 6.1. INTRODUCTION Wind-aided flame spread is the most hazardous type of flame spread in building fires, and has been the most difficult to understand and quantify. An improved under- standing of wind-aided flame spread will result in better material hazard classification and greater fire safety. For fire safety, the knowledge of the burning rate as a function of time is important because it controls the flame size or the flame length. During wind-aided flame spread, the burning rate depends upon the local burning rate and the burning area (or in the 2-D ceiling configuration), which changes with time. Previous studies (Fernandez-Pello, 1979; Carrier et al., 1980, 1990; Loh and Fernandez-Pello, 1984; Wichman and Agrawal, 1991; Atreya and Mekki, 1992) of wind-aided flame spread have primarily focussed on determining the length of the burning (or pyrolyz- ing) zone and have assumed the local pyrolysis rate to be steady. As a result, the pyrolysis front location and speed (which primarily depends on the ease of ignition of the as-yet-unburned material) have been extensively analyzed but predictions of the flame front speed or flame length do not show close agreement with the limited experi- mental data. This is especially true for the ceiling configuration. Thus, this chapter focuses on determining the flame front speed. Previous experimental investigation (Mekki et al., 1990; Chapter 4) on wind-aided flame spread over a vaporizing solid (PMMA) has shown that the flame front speed is essentially controlled by solid vaporization. More importantly, it was found that though the vaporization rate was unsteady throughout the flame spread process, the spread rate was directly proportional to the free-stream velocity, implying a quasi- steady behavior of the gas-phase thermal boundary layer. Thus the characteristic time for the development of a steady thermal boundary layer is much shorter than the characteristic time for the solid-phase to achieve steady state. Recent experiments 112 have shown that the convective heat transfer under the flame is proportional to x‘ , exhibiting steady behavior of the thermal boundary layer (Atreya and Mekki, 1992). 139 Also, for most building materials the surface emissivity is nearly unity (0.92 for PMMA) resulting in a substantial (~ 1W/cm2) radiative heat loss. This further compli- cates the problem. Even for the laminar case, the problem is inherently transient and necessitates the use of approximate solution methods. Thus, the objectives of this study are: (i) To develop a simple but workable steady laminar gas phase model; (ii) To couple it with the transient solid phase model (obtained in the previous chapter) using appropriate boundary conditions at the surface to predict the flame front location (xf) as a function of the pyrolysis front location and time, and (iii) To provide physi- cal understanding which could serve as a basis for material hazard evaluation or development of more complex models. 6.2. MODEL PROBLEM In the gas-phase energy balance, the heat supplied by the flame to the solid is the difference between the heat generated by combustion and that convected downstream by the flow. The heat convected downstream depends upon the thermal boundary layer thickness and the temperature distribution within it, which is affected by the flame location. Estimates of the thermal boundary layer thickness show that, for typi- cal velocities (~ lm/s), it soon becomes substantially larger than 0.1 m while experi- ments (Mekki et al., 1990; Atreya and Mekki, 1992) show that in the ceiling configuration the flame stand-off distance is only about 3 mm. Thus, the error in deter- mining the heat convected downstream by assuming the flame to lie on the surface is expected to be small. It will result in a slight underestimate of the heat convected downstream by the flow and hence a slight over estimate of the heat supplied to the solid. This is conservative from the fire-safety perspective. Hence, to simplify the gas-phase analysis, we assume at the outset that the flame lies at the solid-gas inter- face and has a constant temperature T f. This flame is maintained by diffusion of I oxygen to the solid-gas interface. A schematic of this problem is shown in Fig. 6.1. 140 Solid Figure 6.1 Schematic of the model 141 A surface flame was employed by Wichman and Williams (1983) for opposed-flow flame spread in their analysis of spread into an Oseen flow of oxidizer. In their case, the flame actually lay on the surface, because of the infinite reaction rate, the high rates of downstream convection and because the mass flux boundary condition was linearized. Their surface-flame theory, in other words, was justified by, and consistent with, the remainder of the analysis. This is not the case here because the gas flow is not Oseen (it is a boundary-layer flow) and the mass flux is large near the leading edge (infact, it is infinite at the origin). Hence, from the standpoint of the equations and boundary conditions, the surface-flame hypothesis is strictly not possible. The observed closeness of the flame to the surface arises because of three factors: (i) The solid is transient, hence the mass flux is small. Also, the boundary layer approxima- tion breaks down near the leading edge, making the mass flux much smaller (details in the next chapter). The location of the flame primarily depends on the stoichiometric ratio. Since the fuel is much less than the corresponding steady-state value, the flame moves to the surface. (ii) High streamwise convection and (iii) Buoyant forces also push flame closer to the sample surface. We take the gas-phase governing equations to be the steady isobaric system Bu 8v 3x + By =0, (6.1) a“ a" -2. 3_u Pug-H” a), - a), [it By]. (6.2) ar ar -2. a'_r pCp UFV-a—y']-ay [k y], (6.3) 8Y0 8Y0 _ a D 8Y0 64 “ewe“Epe ‘” 142 We introduce the following non-dimensional parameters, ' x*-_-x/xf, “Eu/u”, Ref=uwxoolvw T'=(Tf-T)/(Tf-T,,), Y;=Y0/Yo.., y y y e =(Ref1’2/xf )Ip/p,dy , v' =Rcf1/2(pv+uI-3%dy )/p,,u ... p2v=p3v” and Le =1, 0 0 The governing equations in the non-dimensional form are 3“, + 3", =0, (6.5) 8x 8y i t 2 I at Bu +V. Bu _a u (6.6) ax‘ ay' ‘ay‘z’ + - _1 axl v 8”. Pr ayiz 1. e 2 e u' 3T .8T _1 8T (6.7) . 3Y3 . BY; 1 321’; —.+V —.—=— I. . 3x a); Pr 8); 2 (6.8) The boundary conditions are: aty’=0: u‘=0; T'=0 for 0 < x*x;; Yg=0 for 0oo: Note that the transient behavior of the solid comes through mass flux, mp ”(x ,t ). 6.3. SOLUTION 6.3.1. SURFACE HEAT FLUX UNDER THE FLAME Th integral method is used to solve Eqs. (6.5)-(6.8). Integrating the continuity equation (Eq. (6.5)) from the surface to some location Y outside the boundary layer gives Y d t * . ud , (5.9) a! y v‘=M(x*)— where M (x’ )=v' (x' ,0) is the non-dimensional mass flux from the solid surface divided by the density. For M (x' )=0 and negligible unheated starting length, the integral solution using cubic profiles is given in the literature (Kays and Crawford, 1980). An exact solution for the mass flux from a steadily vaporizing solid under a boundary layer diffusion flame is also available (Emmons, 1956). It shows that the mass flux decreases as x‘m. Since this solution neglects the radiative heat loss from the sample surface and assumes steady vaporization of the solid, it gives the maximum mass flux. Thus, M (x‘ )=0 and M (x' )=cx"’1’2 represent the minimum and maximum limits of the mass flux from the sample surface which vaporizes unsteadily during the flame spread process (Mekki et al., 1990). 144 The integral solution using cubic profiles gives the hydrodynamic boundary layer thickness as: (i) 6" (x"‘)=(2bx"/a)1’2 for M(x“)=0, and (ii) 8"(x‘) = (2[b+c]x'/a)m for M(x*)=cx*"1’2._ Here a=39/280 , b=3/2 and c is the non-dimensional blowing constant. Also, the thermal boundary layer thickness (8;) is related - to 5' as: 5H8" =0.977Pr'1’3. Clearly, the boundary layer thickness increases because of blow- ing represented by c. Note that egg/83;,“ = (1+c /b)“2 is a constant that depends on the value of c . Thus, the heat flux to the sample in the presence of maximum blowing may be obtained by simply dividing the heat flux for zero blowing with this constant. However, for unsteady solid pyrolysis considered here, the above "blowing correc- tion" is inappropriate. Instead, the local blowing correction recommended by deRis (1979) is used (see Appendix 6.A for the derivation). This correction has the form: 5(a) = B/(eB—l), where [3 = tagger” (T, -T,)/ 4.2.... (6.10) where qcom, is the convective heat flux to the sample for zero blowing. For com- * parison with experiments, ma; (x) is the local average mass flux evolved during the experiment (see Appendix 6.A for the effects of transient mass flux), uu, mm (x) = I a, (x ,t)dt. xP/UP Here I is the length. of the sample. Formulas for mp"(x,t) and q'c’;m(x) are derived below. The convective heat flux from the flame to the surrounding is qf_,.. = -k(dT/dy )y=0. Using the variables defined previously for M = 0, we obtain: 145 1/2 x ,. R chase)=0.332k(Tf—T..)Pr1’3 x . (6.11) where the adiabatic flame temperature T f is given by Y0... Y0, (AH—L) — (-—v—C,,,+ Cpg )(T,-T,,) V T =T + . (6.12) f v Cpga + YOJV) Similarly, the oxidizer mass flux from the free stream to the flame located at the sam- ple surface is given by .. Rel’2 rho (x)=o.332-"—-Pr1’3 " r0... (6.13) Cpg x The convective heat flux from the. flame to the solid for M = 0 is obtained from the difference between the amount generated and convected to the ambient, rh;(x)AH 4.2m) = — 4;’_...(x) . (6.14) Thus, the net heat flux into the solid in the presence of blowing (M at 0) can now be obtained as 41;: (I) = [qc’dnv (x )]E (B(x )) — (Ir'e'rad - (6°15) Since the surface temperature T, is constant, an,“ is also constant. Thus, q'w(x) is 146 time-independent in the pyrolysis zone (x,-Y0 “Iv must be satisfied. 148 6.4. RESULTS AND DISCUSSION Previous experiments and theories (Fernandez-Pello, 1979; Carrier et al., 1980, 1990; Loh and Fernandez-Pello, 1984; Mekki. et al., 1990; Wichman and Agrawal, . 1991; Atreya and Mekki, 1992) have suggested a constant pyrolysis front speed Up. Hence, when the pyrolysis front is at xp , a location x upstream of xp has already been exposed to a constant heat flux q}; (x) for (xp —x )/Up seconds. Thus for a moving pyrolysis front, t in Eq. (6.17) is equal to (xi, -x )/Up. Note that the flame length will continue to increase even if the pyrolysis front has reached the end of the sample (Mekki et al., 1990), i.e. xp = I. For this case, I is equal to At+(l-x)/Up. Here At represents the time elapsed after the pyrolysis front has reached the end (I ). To compare the predictions of the present model with our previous experiments (Mekki et al. 1990; Chapter 4), we briefly describe the ignition method employed. This is essential for determining the location of the leading edge of the pyrolysis front x,- used in Eq. (6.17). The sample was ignited by a pilot flame which was left on during the experiments. The igniter was 3.25 cm long and produced a pilot flame with a 3 cm overhang across the test sample. Thus, we assume that the start of the maximum pyrolysis rate is 3 cm downstream of the sample’s leading edge (x;=3 cm) or 6.25 cm downstream of the inception of the thermal boundary layer. Between the start of the sample and the tip of the flame the net heat flux is assumed to follow a quadratic power law. It is assumed that heat flux is zero at the start of the sample and is max- imum at the flame tip with its derivative being zero there. This is justified since the solid demands a two-dimensional solution near the leading edge. The solid region next to the water-cooled igniter is nearly at ambient temperature, which makes the streamwise conduction large; hence it cannot be neglected. This is discussed in greater detail in Agrawal (1992b) (Chapter 7). Due to incompleteness of combustion, oxygen depletion data was considered most appropriate for comparison with the model (Mekki I, et al., 1990; Chapter 4). For any x , the total oxygen consumed is vwx Im;(x)dx, It 149 where w is the width of the sample (5.7 cm). The predictions of the above expression . for PMMA burning in different oxygen concentration environments are presented in Flg. 6.2 along with the experimental results. Given the simplicity of the model, the agreement is quite good except for higher Y 0,, and when the flame becomes large. Under these conditions incomplete combustion and the resulting flame radiation has a significant effect (Mekki et al., 1990; Chapter 4) which is not included in the model. For calculations presented in Fig. 6.2, the following data were used: AH =26.56KJIgm, v=1.92 and Cpg=l.3KJ/kg in Eq. (6.12) (Sibulkin et al., 1981). Also, k and Pr for Eqs. (6.11) and (6.13) were calculated at the film temperature, (T, +T,.)/2 for air (Incropera and Dewitt, 1985); Cp was based on the integrated average shown below . T! C, = (140.99”: + YWCPOI, 9:1] cpdr. (6.19) Cp~,=9-3355 - 122.560'15 + 256.3%2 — 196.080'3 cal/gm -moleK , 6,058.9465 + 00048044615 - 42.6799-1-5 + 56.61594, where 0=0.01T(Kelvins) (V anWylen and Sonntag, 1973). The property values for various oxygen mass fractions are given in Table 6.1. Here Up is the pyrolysis front speed obtained in Chapter 4. The flame temperature, T, , and C” are obtained from Eqs. (6.11) and (6.19) respectively, 46;“)? is obtained fi'om Eq. (6.14). 150 0.4a 0.3-1 0.24 OXYGEN DEPLETION RATE (gm/s) Figure 6.2 Dependence of total oxygen depletion (total fuel production x v) on the pyrolysis front location. Symbols represent experimental results. l-Yo..=0.233; 2—Y0..=0.43; 3—Y0..=0.61. 151 Table 6.1. Property values Y0- U, T, k C, Pr claims/E m/s K W/mK J/kgK W/m1~s 0.233 0.00025 2342 0.083 1000 0.7 6807 0.383 0.00066 3441 0.126 1098 0.68 14749 0.432 0.00092 3769 0.140 1114 0.67 18450 ' 0.61 0.00175 4856 0.225 1152 0.61 41497 Figure 6.3 shows the calculated values of 8(6) for different oxygen mass frac- tions. Fig. 6.4 shows 6,2,, and q); for different values of the oxygen mass fraction. Clearly the net heat flux into the solid is far less than qconv and also is relatively flat. Fig. 6.5 shows the calculated oxygen consumption rates for the experiment in ambient oxygen condition (#1) with different assumptions. This experiment was chosen because the flame was blue and incompleteness of combustion and flame radiation were not significant factor. The experimental data is plotted as open circles and the solid curve (1) represents the model predictions, as in Fig. 6.2. The calculated curves (1a) to (1e) are for different conditions that are described below. 6.4.1. EFFECT OF BLOWING In Fig. 6.5, curves (1a) and (1b) are calculated exactly as curve (1), except that the steady state rather than the average mass flux was used in the blowing correction [153(0)] for curve (1a) and no blowing correction was used for curve (1b ). For this experiment under ambient 02 conditions, the evolved mass flux is small, hence the 152 BLOWING CORRECTION E(fl) ' Figure 6.3 Blowing correction 153 1.4E-l-06 . , . , , , , - — 9:00” . 1.2E-l-05- .... q not .. 0’2‘ ‘ 1- omar-0.233 - E 1.0E+O5-l 2- 0,-0.432 - 3 ‘ 3- o,,,,=0.810 ' v 8.0r:+044 E'. 6.OE+04-1 E—o -( . a , Figure 6.4 Comparison between the convective heat flux and the net heat flux 154 ‘ m‘ 004 I I I I I I I l I I I I I o" -( I". "o' a DD o" '0 v '0' 0" 0" m 0 .‘ 'o' - B I" 'l o d o’ o 'o o m It " O I """" " 1e x «I l '0 o '0‘ z ....... 1 ,,,,,, c " 0' O —I ooooo '0' H ,,,, "o' «- E-1 t' ...... ” oooo ' '0 ‘o a o o 1' """"" o" """""" . . 1d ...... '0 ''''' cl ” """ ’0 ooooo ’I ''''' I’ 00000 . ’0 ''''''' 0. 1 ~ .I ..... _ ’ 0' 0' l a '0 z 1' """"" m I " 0’" . I - ,. 1b u--“‘" I, 0’ i "' --'-—w . . O C . . . . I ' 0' .00....” O O ....... one... a '0' ”a” . O O __ 1 a o I l I I I I F Figure 6.5 Dependence of oxygen depletion rate on the pyrolysis front location (Parametric dependence). Y0 ..=0.233. Symbols represent experimental results. Solid curve represents the model predictions. Calculated Curves (1a) to (1e) are for different conditions as explained in Secs. 6.4.1-6.4.3 155 influence of blowing is small. To clearly see the effect of blowing, consider curve (1d ), which represents steady-state purely convective burning of the solid as discussed in previous models (Carrier et al., 1990; Wichman and Agrawal, 1991; Chapter 2). Here, the evolved mass flux is much larger than measured during the experiments. Thus, the blowing correction is substantial. This is shown in curve (1e), which is cal- culated for the same conditions as curve (1d) except with E ([3) = 1, i.e. no blowing. The effect of blowing was also substantial for higher Y0... cases (Fig. 6.2) because the evolved mass flux was much larger. More generally it is seen from Fig. 6.3 that blow- ing has a substantial influence on cases where the evolved mass flux is large. 6.4.2. EFFECT OF RADIANT LOSSES FROM THE SURFACE It is interesting to analyze the effect of heat loss from the vaporizing surface. This is constant since T, is constant, and is primarily due to re-radiation. The curve (1c) in Fig. 6.5 was calculated for the same conditions as curve (1), except $2,“ = 0. Neglecting surface re-radiation significantly increases the evolved mass flux and hence the 0 2 consumption rate. Re-radiation also has another effect on the flame spread process. It eventually causes the pyrolysis front speed Up to become zero. For simplicity, consider the steadily pyrolyzing case with x,-=O in Eq. (6.17). Also, to examine the eventual dis- tance the flame will propagate, xp M, the effect of blowing can be legitimately ignored because the sample at this location is barely pyrolyzing. Thus, with mp’M = 6,;«1. +Cp (T, -T,,)) one obtains from Eqs. (6.11)-(6.17): 2 AI-IY v-C T,—T.. ] ,whereB= OJ ”A ) L+Cp, (T, -T,,) . (6.20) xf 1 érerad‘fx—p- v(1+B) 1 , .. xP 2 qcoanx— v+Yo~ 156 This equation shows that xf/xp decreases as xp increases. This decrease will continue until the not heat flux into the solid at the pyrolysis front becomes zero (6,2,, =q',:,d). This location defines the maximum pyrolysis front length (xp M) of a semi-infinite sample and is primarily controlled by surface heat loss. Here, xf /xp is one fourth of the steady state value without re-radiation (Eq. (6.18)). This is in agreement with the experimental results. As shown in Fig. 6.2, for the experiment under ambient 02 conditions, the pyrolysis front stopped at point B and never reached the end of the sample (66 cm). Similar results were also obtained for experiments with 18% oxygen and free stream velocity of 0.9 m/s . Since no sample is infinitely thick, after some time the leading edge will burn out and result in an increase in the convective heat flux at x, (since xp -x,- decreases). The pyrolysis front will then move ahead and stop somewhere downstream where q}; =0. Clearly, for this case the flame spread occurs as fast as the burnout front navels. 6.4.3. INFLUENCE OF UNSTEADY PYROLYSIS To examine the effect of unsteady solid pyrolysis, curve (1d ) in Fig. 6.5 was cal- culated under exactly the same conditions as curve (1c) except the solid was assumed to be steadily pyrolyzing and the corresponding E (B) was used. Thus, curve (1d) represents steady state purely convective burning of the solid as assumed by previous models (Carrier et al., 1990; Wichman and Agrawal, 1991). Two points are noteworthy: (i) a comparison of curve (1d ) with the experimental data which includes both re-radiation and unsteady effects shows significant disagreement, and (ii) a com- parison with curve (10) shows that even after eliminating the effect of re-radiation, the unsteady effects are substantial. The fact that the sample undergoes transient pyrolysis throughout the wind-aided flame spread process is evident in Fig. 6.2. The point A (for Y o ..=O-43) represents the end of the sample. Here, the pyrolysis front reaches the end but the flame continues to grow until a steady state is achieved. This increase in 157 the flame length is represented by the vertical line commencing at A. Now let us consider the effect of unsteadiness of the solid when E (B)=1 and d,’,',,,,=0. From Eq. (6.16) one obtains ' " / C T" A _ . :nP (1 xp) = [1+ p: v][1_ o- ] ’ (6.21) meme/e) L 40) where A (t-)Efunction ($.37). Without surface heat loss 4;, goes as x’“2 (Eq. (6.15)) and t_=(xp -x )/Up, this implies that A (7)5function (x /xp). On integration of Eq. (6.21) from zero to x/xp one obtains mp =const. xmpmmy . Thus, the total mass loss rate from a sample at any instant is a constant fraction of its steady state value. This con- stant fraction remains the same throughout the flame spread process. It increases with increase in wind speed and oxygen mass fiaction and decreases with increase in pyro- lysis front speed. Since the flame front location is directly related to total oxygen depleted [Eq. (6.17)], agreement between experimental and theoretical predictions of total oxygen depletion (Fig. 6.2) will also result in agreement of the flame front location. Hence, the present model also predicts U f :Up as obtained experimentally in chapter 4. How- ever, this still disagrees with the predictions of the steady- state model developed in chapters 2 and 3 where U f 2UP. The main reasons are (i) steady pyrolysis as dis- cussed in this section and (ii) finite rate chemistry. In the actual case there is never flame reattachement, the flame always extinguishes before it once again touches the surface, thus decreasing the flame overhang. This is discussed in greater details in the Conclusions of Chapter 3. 158 6.6. CONCLUSIONS In this study a simple but realistic model for the transient flame front location during wind-aided flame spread across a ceiling-mounted thick vaporizing solid has . been developed. The model predictions are in good agreement with the experimental results. This model is based on the experimental observation that the solid—phase is transient while the gas-phase is quasi-steady. The experimental fact that in the ceiling configuration, the flame stand-off distance is much smaller than the thermal boundary layer thickness is used to considerably simplify the gas-phase analysis. It yields an expression for the flame heat flux as a function only of the streamwise distance. This enables including surface re-radiation and blowing, which were found to have a large influence on the burning process. To account for transient pyrolysis, an analytical expression is developed for transient mass loss from the sample subjected to a constant heat flux. It compares favorably with the numerical and experimental results obtained in this study and those available in the literature. This expression is combined with the gas-phase heat flux to yield the flame front location. It was found that because of the surface heat loss, for cases where the flow is laminar and flame radiation is small, the pyrolysis fi'ont speed will eventually become zero for a thick solid. Also, as a result of re-radiation, blowing and transient pyrolysis, the net heat flux into the solid 1’2, as assumed by previous models. To summarize, the determina- does not vary as x' tion of Up does not result in a complete understanding of wind-aided flame spread. The process is inherently transient and the determination of the flame front location and speed is perhaps more important. In next chapter surface heat flux in the burning zone (pyrolysis zone) is deter- mined from burned PMMA samples. A simple model is developed which gives the regressed surface location as a function of applied heat flux and time. 159 APPENDIX 6.A. Blowing Correction Since the flame is very close to the surface, the streamwise velocity is very small between flame and solid surface. Also the temperature gradients in the streamwise direction are much lower than temperature gradients in the transverse direction. Hence udl‘ldx Xp PMMA sample T5 = cilia) i ‘ W i i q’mm Flow qmmd=const qmmd(x,t) —-> Flame i ’ Xf Underflame Preflame Pyrolysis Zone Preheat Zone > Figure 7.1 Schematic of the gas-phase model 167 7.2. NET SURFACE HEAT FLUX IN THE BURNING ZONE In our previous experimental investigation (Mekki et al., 1990; Chapter 4) on wind-aided flame spread over a vaporizing solid (PMMA) it was found that while the vaporization rate was unsteady throughout the flame spread process, the spread rate was directly proportional to the free-stream velocity, implying quasi-steady behavior of the gas-phase thermal boundary layer. Thus the characteristic time for the develop- ment of a steady thermal boundary layer is much shorter than the characteristic time for the solid-phase to achieve a steady state. Recent experiments of Atreya and Mekki (1992) have shown that the convective heat transfer to the surface under the flame is proportional to x’l’z, exhibiting steady behavior of the thermal boundary layer. Under experimental conditions where the flame radiation is negligible (lower oxy- gen concentrations), the net heat flux into the solid under the flame in the pyrolysis zone is time-independent, since convective heat transfer is only a function of location and the surface reradiation is constant (the surface temperature being the constant "vaporization" temperature). However in the underfl ame preheat zone the surface tem- perature is not constant, hence the surface reradiation is a function of time and location making the net surface heat flux transient. In the preflame preheat zone the convective heat flux is a function of time and location (Agrawal and Wichman, 1991; Atreya and Mekki, 1992). Thus in the burning (pyrolysis) zone, a location x0 faces a net heat flux 4,; (x0) for time to during vaporization (see Fig. 7.1). Here to is the maximum (total duration of experiment) time minus the time taken for the pyrolysis front x, to reach x0 (to =tmu—t, ). Thus the solid has regressed a distance so at x0 in time to under the net heat flux 4,; (x0 ). These observations are used in order to develop a one- dimensional thermal degradation model. It requires to and so and gives the value q.n£t (x0 )- 168 7.3. REGRESSED SURFACE LOCATION For developing a simple solid-phase model of "vaporizing" materials like PMMA, we note that in the pyrolyzing zone, the surface temperature remains constant (vapori- zation temperature) with respect to the streamwise coordinate x . Also, the maximum temperature gradient in the x-direction in the preheat zone is at least an order of mag- nitude smaller than the temperature gradient in the y -direction. Thus, the solid-phase heat transfer is approximated as one-dimensional normal to the surface. Fig. 7.2 schematically shows this heating process. The solid undergoes two stages: (i) The inert preheating stage in which the surface temperature increases until it reaches the vaporization temperature, and (ii) The vaporization stage in which the solid surface regresses at a ‘ constant temperature liberating fuel vapor. The model and its predic- tions for the preheating stage are presented elsewhere (Chapter 5). In the present chapter I, is determined experimentally (explained later in flame spread experiments section). In the vaporizing stage, the regressing top surface moves downward at a speed ds(t )ldt. Here t, represents the time when the surface starts regressing, i.e., when the temperature of the top surface reaches T, . The governing energy equation is: (FT 1 err —=——, t< t ( ) subject to the interface conditions: 4;. + 435] =pti-‘fl , T(s(t).t)=T. . (7.2) dy r=r(t) d‘ Also T(oo,t )=T.. in the far field. The above solid-phase model has been widely used 169 Surface Temp / Surface Location Preheating Vaporization <——t°—> tv tmax Time Figure 7.2 Schematic of the solid-phase model 170 to model vaporizing solids like PMMA (Delichatsios et al., 1991; Steckler et al., 1991; Agrawal, 1992; Chapter 5). The integral method is used to solve Eq. (7.1) subject to Eq. (7.2). We choose the profile T(x,t-) = iexp[-a (t_)(x -s (t—))], where f=T-.T,., T=t-t, and a(0)=q,;;,/k7-‘-, (from Eq. (7.2)). Integrating Eq. (7.1) from s(t) to no one obtains ds/dt"=[q,,';,—ka(i')f,]/pz.. Using this in Eq. (7.2) leads to an ODE in a (7) which can be solved exactly to give the regressed surface location: _ :- kT‘, - - C T, s“) = A .71 1_A(_t) ; Ao=__P"__-’ (7.3) qnet l-AO L+CpsTv where (A (?)-1)/(A.—1>exp(4 (2)/A. )=cxpi—C,...q‘..'3?/(pkL )(L+c,,r', )1. Knowledge of s and t- can produce 4,2,. This particular integral method was recently used to predict the transient mass loss during gasification of PMMA. The results agreed favorably with the experiments and i with the numerical results (Agrawal, 1992; Chapter 5). 7.4. PYROLYSIS EXPERIMENTS: DETERMINATION OF SQ To confirm the. predictions of Eq. (7 .1), experiments were performed on 10 cm x 10 cm and 1.27 cm thick samples of PMMA under various heat fluxes. These experiments were conducted in a water-cooled controlled-environment cone heater apparatus in a N 2 atmosphere (Fig. 5.9). White pigment PMMA was used to minimize in-depth absorption of external radiation. The top surface of the sample was 171 covered by a reflective aluminum foil with a circular opening (31 cm 2) in the center to eliminate edge effects and a water-cooled plate was used to protect the sample during heater transients. These samples were subjected to various heat fluxes. The total mass loss was measured as a function of time, as plotted in Fig. 7.3. Here the mass loss is divided by the surface area and density to give the regressed surface location. The predictions of Eq. (7.4) are also shown in Fig. 7.3. The match is favorable. The sam- ple thickness was taken as 0.127 cm. The other parameters used were (Agrawal, 1992; Chapter 5) k=0.185W/m2s, C,,=2150KJ/kgK. p=ll90kg/m3, T,=635K, T..=295K, L=1650 J/gm and 4,2,“ = 0.82W/cm2. 7.5. FLAME SPREAD EXPERIMENTS: DETERMINATION OF ta The experimental apparatus (Fig. 4.1) used consists of a small-scale combustion wind tunnel which is capable of providing a flat plate boundary layer flow at a specified free stream velocity and chemical composition across the sample surface. A detailed description of this apparatus is given in Chapter 4. The test sample (0.76m long, 0.076m wide and 0.0127m thick) is placed horizon- tally along the tunnel top. The samples used for these experiments were obtained from white sheets of PMMA. These were instrumented with nine thermocouples on the sur- face 0.05 m apart (Chromel-Alumel; 76 pm diameter). The first 0.15 m of the sample was not instrumented to avoid errors due to ignition. Once all the desired conditions were set (flow speed and composition), the sample was ignited with a small methane porous-metal burner placed at the tunnel inlet with its face parallel to the sample sur- face. The fuel flow rate to the igniter was controlled such that the flame overhang on the sample surface was about 0.03 m . The data were collected by the computer- controlled data acquisition system and the flame spread process was recorded by a video camera. The flame is extinguished with the nitrogen before it reaches the end of the sample. It is well known that PMMA vaporizes at a constant temperature, and this 172 03 +1.. J .1; J n l l 00 J_ l l l REGRESSED DISTANCE s (mm) 1 ‘l f —- Experimental -CD L Theoretical ’ , w .. ,” "=2.75 W/crnz - I I I . I I d I I, u . 2.20 ,’ ’ I ’ 4 I’ 1.82 ,’ l/ I I’ 1.52 1.32 I”” ’Igv’I’ ””””””” 1 z ', ”1"” , I . I . I j 500 1000 1500 2000 TIME (sec) Figure 7 .3 Regressed surface location as a function of time 2500 173 fact has been extensively used in the literature to characterize the onset of pyrolysis in PMMA (Loh and Fernandez-Pello, 1984; Mekki et al., 1990; Chapter 4). Thus, from the measured surface temperature profile at x0 , tv represents the time required by the surface to reach the vaporization temperature. Also, to =tmu-tv. 7.6. RESULTS AND DISCUSSION The burnt sample sheets are out along the longitudinal center line and its thick- ness measured with a micrometer (accuracy of $0.0254 mm) at various positions along the center line. The net regressed distance 3 is obtained by subtracting the final thick- ness from the initial sample thickness. Results for the net surface heat flux in the pyrolysis zone as determined from Eq. (7.4) with the use of measured so and to are presented in Fig. 7.4. The net heat flux increases with free stream velocity and decreases with the Streamwise distance. Note that the decrease does not follow Jc’l’2 dependence. A typical Jc'l’2 profile is also shown in Fig. 7.4. Some of the reasons for this discrepancy were discussed in Chapters 4 and 6 (solid-phase unsteadiness and. surface re-radiation). Note that in all cases the maximum net surface heat flux is around 3 em down- stream from the leading edge. No data are presented for region 1: < 3 cm because this region is prone to gas-phase finite reaction-rate effects and solid—phase cooling (see Fig. 7.5 for the surface heat flux near the leading edge). In this region the net heat flux decreased with the decrease in x . Other then the influence of finite gas-phase reaction-rates (Chen and Tien, 1986; DiBlasi et al., 1987) the solid edge (touching the igniter) was cooled because of the water-cooled i gniter. To quantify this effect numer- ical solutions were obtained for (11‘ Id: =a(d7’I'/dx 2+d 2T ldy 2). The boundary condition representing the water-cooled side of the solid was chosen as T (x =O,t )=T... Far from the leading edge the solid was assumed as one dimensional (dT (x —-)oo,t)/dt = 0). The bottom side was assumed adiabatic. The sample was subjected to a constant heat flux q"m(W/cm') 174 l ' I ' I O U A x (cm) Figure 7.4 Net surface heat flux in the burning zone q"...(W/cm') 175 2.0 ‘. l I ' . 't O U..=2.0rn/r 1.8... t“ D U..=1.5m fl 1 3 . §-- 1/\/X Dependence A U-=0.9m;fi 0 fl “ d 1.4; ° ° ° _' q ‘\‘ O O O O 0 . 1.2-j O Ox‘q D O O -.: x‘ o o a 1.04 o a ““““ a a o d . ‘ ““““““ ' o u '- o. - ° ~~~~~ ‘ 81 o A ............... . ° . 0.6+ D A ‘ ¢ ........ “M ...... 'M; _. A A A A A 0.41 g d 0.2- " 0.0 I ' l ' 0 5 10 15 x (cm) Figure 7.5 Net surface heat flux near the leading edge 176 of the order shown in Fig. 7.4. It was found that the leading edge effects are limited to the first 2-3 cms of the sample. This, coupled with surface radiative heat losses and solid-phase unsteadiness seem to be the main reasons for the short flame overhang (Chapter 4). This technique (using burned samples) has been previously used to predict to the regression rate of a PMMA sample burning under turbulent conditions (Zhou and Fernandez-Pello, 1991). The regression rate was determined by dividing the regressed distance by effective time (so/t0 ). Note that in doing so, the solid is essentially assumed steady. Since the solid thermal processes are transient (Mekki et al., 1990; Chapter 4; Agrawal, 1992; Chapter 5; see also Fig. 7.3), this produces an under- estimate of the maximum regression rate and subsequently the net surface heat flux. For a thick sample, the maximum regression rate is directly proportional to the net sur- face heat flux (steady state condition). 7.7. CONCLUSIONS The net surface heat flux in the burning zone during wind-aided flame spread is determined using measurements of the final sample thickness. A simple solid-phase model is developed which gives the net surface heat flux based on the regressed dis- tance and the effective burning time. The model agrees favorably with the pyrolysis experiments. It was found that the surface heat flux increases with the increase in free stream velocity and decreases with the streamwise distance. This study is limited to vaporizing solids and for environmental conditions where flame radiation is negligible. CHAPTER 8 STABILITY ANALYSIS SYNOPSIS A hydrodynamic stability analysis is performed for the wind-aided flame spread problem in the ceiling configuration. The effect of the reaction zone temperature profile on stability is analyzed. It is found that the Froude number is the only parame- ter affecting the pulsation frequency of the flame once the temperature and velocity profiles are specified. The infinite-reaction-rate flame temperature profile is the most unstable: generally, the sharper the temperature profile at the flame, the higher is its characteristic oscillation frequency. It is also found that the oscillation frequency increases with the B -number. 177 178 8.1. INTRODUCTION It is well known that a light fluid beneath a heavy fluid is generally in a buoy- antly unstable condition (Taylor, 1950). This condition occurs for wind-aided flame spread across ceilings (Fig. 8.1a). Despite the apparent importance of the diffusion flame stability problem for determining conditions leading to turbulence, for altering heating rates, etc., there are relatively few studies available in the literature. A brief review of the literature follows. Large jet diffusion flames have been observed to flicker at a frequency of about 12 Hz, which is insensitive to flow rate, burner size or gas composition (Buckmaster and Peters, 1986). This flickering is apparently related to the stability properties of the flame. Since the 12 Hz frequency is largely independent of flame geometry the insta- bility mechanism is presumed to have hydrodynamic origins. The study of Kimura (1965) for tube burner diffusion flames raises a fundamental question as to the choice of undisturbed velocity and temperature profiles because these change significantly throughout the flame. He adopts the bell-shaped profiles measured well above the flame. Since the instability seen within the body of the flame is caused by the profiles at much lower heights, and these profiles are far from being bell shaped, doubts were raised regarding Kimura’s (1965) choice of profiles by Buckmaster and Peters (1986). Grant and Jones (1975) studied the stability of the free diffusion flames for different temperature profiles and suggested that the invariance of the frequency for small-scale free diffusion flames cannot be obtained from a linear stability analysis where the dis- turbances are purely convective, i.e., where all dissipative process such as diffusion of momentum are neglected. Although the stability of free diffusion flames (Buckmaster and Peters, 1986; Ellzey and Oran, 1990; Grant and Jones, 1975; Kimura, 1965; Puri and Hammins, 1990; Toong et al., 1965) has been studied, there are no available stu- dies of diffusion flame stability for flows with contiguous walls. The present study is an attempt to understand the stability of such flames. 179 SAMPLE \\\\\\\\\\\{\\\\\\\\\\\\\\‘ ——I> Up _> Fuel ’. Wind Direction Thermal B.L. Gravity Figure 8.1a Physical configuration of the wind-aided flame spread U Thermal B.L. -le Flame Flow TW ———-> X Figure 8.1b Schematic diagram of the model .180 The questions that this study addresses are: (1) when can instability occur, (2) what are its principal causes? Accepting that the temperature (density variation) may be the driving force (or principal cause) for the instability mechanism, we ask further; , (3) how significantly does the temperature profile affect the stability characteristics of the flame? We wish to determine, if possible, the relevant non-dimensional parameter(s) on which the most unstable frequency depends and the frequency range of instability under representative experimental conditions. The problem we will examine is illusu'ated in Fig. 8.1a. We see that since the density decreases from the wall to the flame the region in between is potentially unstable. Because of large density variations we perform a hydrodynamic stability analysis of the conservation equations. Analysis of the disturbance equation (Appen- dix 8.C) shows that an inflection point in the velocity profile is a necemary but not sufficient condition for instability to occur and that the unstable region exists between the wall and the flame. The velocity profile possess an inflection point because of the blowing of fuel from the "vaporizing" surface. This condition exists only in the pyro- lyzing zone of the flame spread (Fig. 8.1a). In the preheat zone there is no blowing at the surface; hence, instability there can only occur if pyrolysis zone is unstable. Thus, only the pyrolysis zone of the wind-aided flame spread problem will be analyzed. Numerous temperature profiles are examined to identify the principal mechanisms for the instability. This is the most important part of our analysis for it goes to the heart . of question (3). If the instability mechanism is essentially independent of the detailed flame temperature profile, a comprehensive analysis of the diffusion flame su'ucture is unnecessary: one could simply employ the relevant infinite reaction rate profile and produce a generic, universally-valid stability analysis for a given class of flows such as jets, ceiling fire spread, etc. If, however, the stability characteristics depend strongly on the detailed nature of the flame temperature profile, the stability analysis becomes more complex because the reaction-zone temperature profile depends on the chemical 181 rate parameters therein (like the pre-exponential factor and activation energy for a single-step reaction). We note that outside the reaction zone the infinite and finite-rate profiles are virtually identical. Thus, our question is whether the temperature profiles in an extremely thin region of the flow field can affect the stability characteristics of the global problem. We will demonstrate that the answer to this question is affirmative. The governing equations for the wind-aided flame spread problem are presented in the next section. Basic state velocity and temperatm'e profiles are obtained therefrom. In Sec. 8.3 the disturbance equations are identified and a single equation is obtained that uses the steady values obtained in Sec. 8.2 as its basic state. We note that our analysis is for an assumed inviscid parallel flow. Despite the important work done on non-parallel stability analyses for boundary layer flows (Saric and N ayfeh, 1977; also see refs therein) we do not feel these additional complications presently warrant inclusion in this preliminary investigation: our study is focused on the princi- pal thermal causes of the instability, not its purely fluid-mechanical subtleties. Results are presented in Sec. 8.4. Appendices 8.A and 8.B deal with the inner zone solutions of the temperature profiles. A classical Rayleigh-Fjortoft analysis of the disturbance equation is conducted in Appendix 8.C. 8.2. STEADY FLAME SPREAD Shown in Fig. 8.1a is the schematic illustration of the model. After ignition, the tip of the surface-pyrolysis zone propagates downstream with velocity U I: , while the flame tip propagates downstream with a higher velocity U}, burning "excess" fuel released from the pyrolyzing surface and not already burnt there. The pyrolysis zone lies between x' =0 and x;=U;t' . The preheat zone extends from x; to the location ULt‘; there is no effect of leading edge beyond this location. It is in the pyrolysis region of the flow field that we wish to examine the flame stability, for reasons already 182 mentioned in the Introduction. In the analyses of Wichman and Agrawal (1991), Chapter 2 and Carrier et al. (1990) for the wind-aided flame spread problem, the pyrolysis zone is assumed steady, which requires the fuel mass flux to be proportional to the incident heat flux. In the gas-phase the chemical reaction is assumed to occur though the simple one-step irreversible reaction F + v0 -> products , where v is the mass-based stoichiometric coefficient. Other simplifying assumptions are negligible streamwise diffusion (boun- dary layer approximation), equal binary discussion coefficients, constant mixture specific heat, unity Lewis and Prandtl numbers, low Mach number and negligible solid-phase regression rate. The flame radiation is neglected, which is valid for low ambient oxygen concentrations (Mekki et al., 1990; Chapter 4). Limitations and res- u'ictions of the model are discussed at length in Chapter 3 (Agrawal and Wichman, 1992). The steady boundary layer equations describing wind-aided flame spread across a ceiling in a cartesian coordinate system (see Fig. 8.1b for the schematic) are ags‘u‘)+a(p’v‘)=0 (8.1) ax‘ ax“ ’ mass: in sau' a sau‘ 3P. 8 r sau' x-mom: u . v . = t + t t )' , (8.2) p 3:: +9 ay 8x ay “1 ay P‘ t t it y-mom: a , =-(p..-p )g . (8.3) 8y . . .ar“ .aT‘ ~ 3 .ar‘ .2 . . . .ym'r energy: p C [u ——T+v .]= . 2t .]+Ap Q YoYpe , (8.4) ” 9x 3y 3y 3y , P , BY; , ar'l P, airM . . . oxidizer. p u —-;-+v f = 3* p D‘ f —vAp*2Y;Y;e-5 ’R T , (8.5) . ax ay l 8y . ay . ’ a e I ti ,,ay .ar “or fuel: p u —f—+v f]: a. p D f —Ap 2Y0Y1t~e45mT , (8.6) . at 3y J 8y .' 3y . 183 state: P I =p' R ' T' . (8.7) 3 By using the Howarth transformation, with the mass coordinate 2.: I p'dy' and the o transverse mass flux w'=p' v' “,1 (ap' /8x")dy* and with zero pressure gradient (streamwise) we can easily obtain the corresponding incompressible flow boundary layer equations (Wichman and Agrawal, 1991; Chapter 2). These equations are non- dimensionalized by defining u=u'/U:,, x=x'/L', z=z'/p;L', w=w'lp;U:., 7:171"; and L'=v;/U;, ter/1;, Q=Q‘/C;,,T;, Y,=Y,.‘. Here v is the mass-based stoichiometric coefficient for the one- step reaction (F +v0 -—)P ); we have Vp=1 and v0=v. Using the stream function val—27 f (n) and the similarity variable 11:2 Nix- and u =aw/az and w=-8\|I/3x , the x-momentum equation reduces to f + ff " = 0 . (8.8) Here u=f' and w=(1'1f'—f )N2_x_. The boundary conditions are f '(0)=0, f '(oo)=l and f (O)=~—f"(O)B, where B is the "8 number for mass transfer from the surface". Details of the derivation of Eq. (8.8) and the boundary conditions can be found in Emmons (1956). We note here that buoyancy is neglected in the derivation of Eq. (8.8) Wind-aided flame spread experiments conducted in air in the ceiling configuration have shown that for a free stream velocity as low as 0.6 m/s , the flame was stable (Mekki et al., 1990, Chapter 4), although this may not hold true for flames spreading across a floor or flames spreading in much higher oxygen concentrations, where the density variation is large. In addition to the effects of buoyancy, flame radi- ation has to be included for higher oxygen concentrations. Hence, as already stated the present analysis is limited to flame spreading across a ceiling in low oxygen 184 concentrations, where flame radiation can- be neglected. In what follows a prime () as superscript denotes the derivative of that variable with respect to n; x' = dx/dn, x" = dzx /dn2, etc. By choosing Lewis and Prandtl numbers of unity we obtain the follow- - ing energy and species equations: -fT' = r"+2xQR' , . (8.9) —fY,,' = Y,,"+2xv§ , ' (8.10) -fY,.1 = Y£+2xE , (8.11) where 18:03:, YFexp[-B(1-T)/(1—ot(1-T))], with B = E‘(T;-T;)/(R‘r;2) and a = =(Tf-T;)/T;. The quantity D=A" p‘v;U;-2 exp(-E‘/R‘r;), the Damkohler number, is the ratio of the thermal diffusion time to the chemical time. For solution we employ the Schvab—Zeldovich coupling functions 70 =T IQ +l’0 Iv and 7p=TlQ +10.» to obtain 73+fyé=0 and y£+f 713:0- Suppose now that u(n)=f' can be obtained from the solution of Eq. (8.8); the solutions for 70 and 7;- can then be obtained as linear functions of u (Emmons, 1956), Yo=aou+bo . (8.12) Y): mpu'i'bp . (8.13) By then imposing the boundary conditions at 11:0 (u =0, T=Tw, Yo =0, Y F =pr) and at 11->°° (u =1. T=T,., Y0 =Y0.., Yp =0) we determine the integration constants as do: ko—(Tw -T..)/Q, b0 =Tw IQ, a1.- =—YFW —(Tw —T,,)/Q and 12,. =Tw/Q +Ym. 185 8.2.1. BASIC-STATE VELOCITY PROFILE In order to solve for u (1]) only one parameter is required, namely B . We used the values B=0.275, B=1.125 and B=1.6, for which, the heat losses from the flame towards the surface and towards the free sueam are identical, see Appendix 8.A. This preserves the symmetry of the reaction zone, which would otherwise be skewed to one side or the. other, severely complicating the stability calculation. We employ the Runge-Kutta-Gill method with a stepsize of 0.001 to determine u , u' and u" from Eq.- (8.8). The results are plotted in Fig. 8.2 for B=1.125, from which it is clear that the velocity profile has an inflection point at 11:11. . The velocity at the inflection point, . ns=l.405, is u. =0.448. 8.2.2. BASIC-STATE TEMPERATURE PROFILE From the known solutions for u , yo and 7;, the mass fractions Y0 and Y I: can be easily expressed as linear functions of T , T Y = b0-— , 8014 a (a0u+ Q) ( ) Yp=v(apu+bF-%) . (8.15) Equations (8.14) and (8.15) are valid throughout the physical domain (i.e., not simply on one or the other side of the flame). Thus, the reaction term if in Eq. (8.9) can be written in terms of T; consequently Eq. (8.9) can now be numerically solved for cer- tain values of D , B and 0t, for the boundary conditions specified previously. If we wish to develop an asymptotic solution in the inner (reaction) zone we have recourse to asymptotic methods (Buckmaster and Ludford, 1982; Linan, 1974; Williams, 1985). In this case the outer zone solutions are given by the homogeneous forms of Eqs. (8.9)-(8.11) with the condition T=Tf applied at the flame, or by Eqs. (8.9)-(8.11) with 186 Figure 8.2 Basic-state temperature and velocity profiles. here is for the infinite-reaction-rate Note that T-profile plotted 187 5 replaced by a 8—function heat release term at the flame sheet location. Since the objective of this study is to examine the effect of the inner (reaction) zone on the sta- bility we will eventually subject the energy equation to the high-activation energy asymptotic analysis. The reaction zone calculations are described in Appendix 8.A. We presently focus on the outer zone calculations for T. First we determine the location 11f of the flame. Here Y0 =Yp =0. Let us now expand T' near any n-value using Taylor’s series. Thus TQM“ =Tn +AnT; +0 (A102. Now from Eq. (8.9) T"=—fT' (for outer solution) and from our expansion, we obtain TQM“ >Tn when f <0 and 7,3,,“ 0. Since f is negative at and near the fuel surface and posi- tive in the far free stream, T must increase in the fuel region and decrease in the oxy- gen region. Continued reasoning in this way indicates T is a maximum at f =0, which occurs at the inflection point for the streamwise velocity profile. Thus, the flame lies at the inflection point nf =11. . Henceforth we refer to the inflection point with the subscript f . Between the wall and the flame there is no oxygen, and we obtain from Eq. (8.14), using T=1 and u=uf at nf, r=rw+(1-T,,)i o $11311; , (8.16) “r with Y .. l-T 1-u Tw g—°—= ”‘94 f ) . (8.17) V 11f Between the flame and the free stream there is no fuel, and we obtain from Eq. (8.15), using T=T.. and u=1 at n—)oo, l-u “‘“f T=T--u—f-+Tf [ l—uf J 11f511<°° , (8.18) with 1-T..uf —(1-uf )Tw l-llf . QYFW= (8.19) Here subscripts w , f and co represent conditions at the wall, flame and free stream, respectively. From Eqs. (8.17) and (8.19) one obtains the stoichiometric index 4) = va/Yo, = uf/(l—uf ). Solutions for T (Eqs. 8.16 and 8.18) and T' are also plotted in Fig. 8.2. We see that T, is discontinuous at n=r|f since it is the "outer" tempera- ture profile, to which the inner (reaction zone) profile must be matched. Solutions for inner-zone temperature profiles are plotted in Fig. 8.3 for various B. Parameters chosen and calculated are tabulated in Table 8.1. Values of B chosen to obtain the various temperature profiles of Fig. 8.3 are presented in Table 8.2. This completes the solution for velocity and temperature profiles, which are employed as mean or "basic-state" values in the disturbance equation described in the next section. We use T; =295 K , T; =630 K ("vaporization" temperature for PMMA) and Tf=2065 K (adiabatic flame temperature for PMMA burning in air (chapter 2) in Eqs. (8.16) and (8.18). This completes the solution for velocity and temperature profiles, which are input as mean or "basic-state" values in the disturbance equation described in the next section. 189 1.0 0.9-l NON—DIMENSIONAL TEMPERATURE Figure 8.3a Inner zone (reaction region) temperature profiles, B =0.275. Curves 1-6 represent profiles for different B’s 190 NON-DIMENSIONAL TEMPERATURE Figure 8.3b Inner zone (reaction region) temperature profiles, B =1.125. Curves 1-6 represent profiles for different B’s 191 1.00 0.951 0.90% 0.86~ 6 an; 0375‘ 0.70% NON -DIMENSIONAL TEMPERATURE 0.65 v 1 I l ' l 1.00 1.40 1.80 2.20 Figure 8.3c Inner zone (reaction region) temperature profiles, B =1.6. Curves 1-6 represent profiles for different B’s 192 Table 8.1 Table of parameters used. Parameters a -b and QYO ,Jv are obtained from Eqs. (8.A.2) and (8.17) respectively. 7;”; B , uf u,’ n, a-b o Ql'gg/v 3 0.275 0.298 0.414 0.733 2.54 0.424 1.33 7 1.125 0.448 0.354 1.405 1.82 0.811 1.71 15 1.600 0.479 0.340 1.630 1.65 0.920 1.87 T;=295 K, 13:63:; K, u=1, Pr=l, g=9.8m/s2, v..=16x10-6 m2/s, somb=7 Table 8.2 Values of the nondimensional activation energy B used for the different B numbers. Temp. Profile # B =0.275 B =1.125 B =l.6 1 Infinite“ Infinite“ Infinite‘l 2 4.0 4.0 4.0 3 3.0 4.0” 3.0 4 2.0 2.5 2.0 5 1.5 1.3 1.0 6 1.0 0.8 0.75 7 - 0.6 - 8 - 0.6” - “ Temperature Profiles calculated using Eqs. (8.16) and (8.18). b Temperature Profiles calculated as described in Appendix 8.B. Remainder of the temperature profiles calculated as described in Appendix 8.B. 193 8.3. STABILITY ANALYSIS As already mentioned, experiments on wind-aided flame spread suggest that buoy- ancy plays an important role in flame spreading. The wind-aided spread of a flame across a ceiling is generally buoyantly unstable. We wish to determine the conditions under which this instability can be suppressed. We perform a parallel inviscid stability analysis with spatially growing disturbances. Michalke (1965) has suggested that spa- tially growing disturbances accurately describe the instability properties of a disturbed shear layer because some of these properties, such as amplification of disturbances, could not be described by linear inviscid temporal stability analysis. Previous diffusion-flame stability analyses (Buckmaster and Peters, 1986; Ellzey and Oran, 1990; Grant and Jones, 1975; Kimura, 1965; Puri and Hammins, 1990; Toong et al., 1965) have all used parallel inviscid flow. We believe that several important aspects of our problem can be uncovered by an inviscid parallel-flow stability analysis, with the temperature and velocity (mean values) determined from the steady viscous-flow analysis of Sec. 8.2. Nevertheless, viscous and diffusive effects may be important under many conditions. For instance, Weeratunga et al. (1990) suggest that hydro- dynamic effects are responsible for the instability of plane flames and diffusion of heat and mass are responsible for the cellular instability. In the first case diffusion effects can be neglected but the density variation has to be retained (see Buckmaster and Peters, 1986; Ellzey and Oran, 1990; Grant and Jones, 1975; Kimura, 1965; Puri and Hammins, 1990; Toong et al., 1965). In the second limit the hydrodynamics (density variation) can be neglected, but the diffusion terms must be retained. We proceed with our inviscid parallel-flow analysis. According to standard prac- tice (Drazin and Howard, 1966) the governing equations are 89' J ag'u' . ap‘v‘ =0 . 8.20 at. 8x 3x ( ) t Bu' tut all. +psvt all“I = 31’. 3"“ Bx‘ By. ax’ , (8.21) 194 48V. 0 eaV‘ a 03V}. BP’ 4: 4- at = _ .. , 8.22 at,.,+pu ax,+pvay, 8y, (P—P)g ( ) p ar‘ W. 31" “.87" . . . =0. (8.23) at 3): 3y Pe=paRaTt (824) a t e a at: s a 1"2 297+“ a —;.—+V* 8%: 1 BF. 8p. BP“ 30* +[_p+:_]g t ’ (8.25) at 81: By p 2 8y 3x 3x 3)’ P 2 where CEGv/ax —au lay) is the z-component of the vorticity. Equation (8.25) was obtained by subu'acting the y-derivative of Eq. (8.21) from the x-derivative of Eq. (8.22). The instability is analyzed in terms of a mean quantity (obtained from the steady state boundary layer flow of the previous section) and a "disturbance", generally much smaller than the mean flow quantity but still finite. Thus we define new vari- ables r*(x*.y*.z*)=u‘0’>+£’a‘ ; v*(x*,y*.:*)=e*v* ; E‘tx*.y‘,r‘)=c‘0‘)+e*t' ; f‘=T*+e*p* ; fi*(x*.y*.t*)=w*0‘>+e*é‘ ; F'(x’.y',t')=P'(y*)+€‘fi'. where 6* =35" exp(i0t'x'-ito*t'). From Eqs. (8.20)—(8.26) one obtains the following equation for the disturbance su'eam function 0‘ d 1 d6' .. a” 1 d 1 du“ 1 " d1“ s[ It a ]—9 42+ in it" a[ a at ]+ t at 2 8‘2 i: =0’ (8‘27) dy T dy T u -uf dy T dy (u —uf) T dy 195 where c'=0)*/a'. The above equation can also be obtained by from the full govern- ing eqns (like Eqs. (8.1)-(8.7)), after discarding the diffusion terms and higher order (2 and more) nonlinear terms in 8'. We note that when T = constant =1 Eq. (8.27) reduces to the classical "Rayleigh equation" for the stability of inviscid parallel flow (Michalke, 1965). Using non-dimensional parameters c=c'/U:,, (u=co‘L'lUL, a=or'L‘ and 0:0'IUL, z’= p'dy’, n'=z*/‘J2x' and the parameters defined previ- V O ously (after Eq. 8.7) we obtain the nondimensional disturbance equation 9' ' 2 1 u" l 7" —-60t+——+ —=O, 8.28 [ ] ° u-C[T2] Fro(u-c)2 T2 ( ) where 09:015. (00:0)‘127 , Fro—142x704, where Fr is the Froude number (=U12/g‘L‘ ). The disturbances must vanish in the free stream and at the wall. Since ti‘=dé"/dy‘ (which is directly related to 0') and the streamwise disturbance in the velocity vanishes both at the surface and in the free stream, we have the following boundary conditions for Eq. (8.28): 9(0)=e(~)=e'(0)=e' . H a: 0.50 1 5.. (ll [:1 tr: 0.1 at: h , 0.00l . , . I . . . , 0 1 2 3 4 STREAMWISE DISTANCE (m) Figure 8.4 Froude number F r0 as a function of free sueam velocity and streamwise distance 197 In the limit Fro —>oo, where there are no buoyant influences, the disturbance Eq. (8.28) is similar to the equation obtained by Buckmaster and Peters (1986); when Fro _,.. and T’=0 (isothermal fluid) it is identical to Michalke’s equation (1965). The solution of Eq. (8.28) requires existing solutions for the basic-state quantities u, u', u", T and T'. From Fig. 8.3 it is seen that for the infinite-reaction-rate temperature profile (#1), T ' is undefined at the flame location (11f). For the neutral stability case, c’=u; [Buckmaster and Peters, 1986; Michalke, 1965], 0' is not continuous at 11, although 0 is. In this case the eigenvalue 010 is obtained by integrating Eq. (8.28) from both sides, n=0 and n—>oo, until 1],; enforcement of the condition 0;! =0,}; (continuity of 0) gives (10 . The integration is performed using the Runge-Kutta-Gill method; the solu- tion procedtue was tested by reproducing the results of Buckmaster and Peters (1986) and Michalke (1965). This solution procedure is applied for temperature profiles #1-6 for B=0.275 and for B=l.6 and #1,2,4-7 for B=1.125. Since in all these cases the last term of Eq. (8.28) becomes infinite as ln-nfl -—)0, as discussed in Appendix 8.B, they cannot be integrated continuously from one side to the other across the flame. For profile numbers 3 and 8 (B =1.125) the singularity does not exist (see Appendix 8.B) and hence the disturbance equation can be integrated across the flame. As a check for Eq. (8.28), temperature profile #8 is integrated from T]—)°°, which we take to be 11 =10, to 11:0 and a0 is obtained by enforcing the conditions 0(0)=0 and 0:0. This procedure gave the same O-profile and 010 as by integrating from the wall and the free stream to n, and matching values there. We note that, strictly speaking, the proper means for finding a solution for 0 is to integrate for complex a0 , then to progressively pass to the limit do has —)0. The solution should converge to the 010 “mg -+0 limit. It is found, however, that solutions obtained in this way are iilglgigfl to the ones obtained by patching the two separate solutions at nf. Hence, this limiting process is not employed here, nor in fact do any of the previous authors employ it (Buckmaster and Peters, 1986, Michalke, 1965). 198 Figure 8.5 presents the stability curves for different Froude numbers for various -B -numbers. Plotted is the non-dimensional frequency 0) versus F r0 . Generally in pre- vious stability analyses a single profile is analyzed, such as the infinite-reaction-rate profile in (Buckmaster and Peters, 1986), or a representative inner-outer profile as in (Puri and Hammins, 1991); the numerical results are then compared to the experi- ments. Our approach, as already stated, is somewhat different. We wish to understand how the differences in the T-profile influence the final stability properties. Figure 8.5 suggests the existence of three regions, purely stable, purely unstable and an overlap region, which could be stable or unstable depending on the specific temperature profile. Thus, experimental results lying in this region could be predicted differently depending on the choice of the temperature profile. In Fig. 8.5b the infinite-reaction-rate temperature profile (1) is the most unstable. Profile 8, which is flat across the reaCtion zone, is the most stable. This is also obvious if we exanrine profiles 3 and 4. Profile 3 has a higher temperature than profile 4 but is flatter and hence more stable. Also, from Fig. 8.5 we observe that the infinite-rate profile has a much higher frequency for the onset of the stable region than the most stable flame (profile 8). Moreover, for experimental design purposes the stability criteria obtained for the infinite-rate temperature profile could be used as a lower bound. Comparison between different B -numbers suggests flame flickering increases with increase in B - number. There has been a question as to what the numerical value of the streamwise coor- dinate it should be to quantify the flickering frequency of the flame (Buckmaster and Peters, 19860. In Fig. 8.6 we present the dimensional frequency as a function of the streamwise distance. The frequency is given by (0* =t00 Prof/U; . The streamwise distance is obtained from the definition of the Froude number (x ' =0.5U,,",,s lv..g '2Fr02). We choose U i=1 m/s and put g*=9.8 m/sz. The results are presented up to x1 =7 m. After that the governing equations are not valid because the flow becomes 199 N L stable H l unstable ' r . l ' l ' J ' l . r r 0 p 5 10. 15 20 25 30 35 FROUDE NUMBER Fro O NON -DIMENSIONAL FREQUENCY coo Figure 8.5a Stability curve; non-dimensional frequency versus Froude number, B =0.275. Curves 1-6 are for different inner temperature profiles of Fig. 8.3a 3°7 ‘ B ‘l 26‘ £53 ~l 1 CV5"? 2 E -‘. 3 I :4“ 4 stable It: 5 . 2. g o _l '9' ‘4 3 E2~i‘\ 5. 9O 8 T . z o ~~-_. z I . t . t . l 20 25 30 35 FROUDE NUMBER Fro Figure 8.5b Stability curve; non-dimensional frequency versus Froude number, B =1.125. Curves 1-8 are for different inner temperature profiles of Fig. 8.3b 201 NON —DIMENSIONAL FREQUENCY coo l ' I ' l 0'5'10‘15'210'25 30 35 FROUDE NUMBER Fro Figure 8.5c Stability curve; non—dimensional frequency versus Froude number, B =1.6. Curves 1-6 are for different inner temperature profiles of Fig. 8.3c 202 FREQUENCY (0 (Hz) x (In) Figure 8.6a Frequency curve; neutral frequency versus streamwise distance, B =0.275. Curves 1-6 are for different inner temperature profiles of Fig. 8.3a 203. FREQUENCY at) (Hz) h OI G O O O l l 1 L I I5 03 5 30‘ . 7 M 20" 8 fi 10 l I I ' I I I 0 1 2 3 4 5 6 X(m)l Figure 8.6b Frequency curve; neutral frequency versus streamwise distance, B =1.125. Curves 1-8 are for different inner temperature profiles of Fig. 8.3b 204 125 H O O l W . “l 75- 50+ 25 ' I ' I I I x (In) Figure 8.6c Frequency curve; neutral frequency versus streamwise distance, B =1.6. Curves 1-6 are for different inner temperature profiles of Fig. 8.30 205 turbulent. For higher B -numbers the disturbance frequency essentially remains con- stant with the streamwise distance but changes substantially with the choice of tem- perature profile (density variation at the flame). For lower B -numbers, where there is n0t much density variation, the frequency decreases with the increase in streamwise distance. Thus characteristic frequency strongly depends on the density variation in the vicinity of the flame. To obtain the growth rate curve, the real and imaginary parts of do are obtained for the frequencies to below the neutral frequency. Figure 8.7 shows a plot of Imtxo versus 0) for B =1.125. It is observed that at lower Froude numbers the growth rate is higher, corresponding to our previous observations. Results are plotted for profile #5 for Fro = 2 and 4. For the more-buoyant flame (Fro =2) the maximum growth rate occurs at a larger frequency ((00 =03, —Im (10:0.238) than for the less-buoyant flame (Fro=4), where (00:0.14 and -Im eta =.18. In Appendix 8.C the disturbance Eq. (8.28) is analyzed. An inflection point in the velocity profile is a necessary condition for instability. Hence, in the wind-aided flame spread problem, the flame overhang (underflame preheat zone, see chapter 2) will be stable if the pyrolysis zone flame is stable, i.e, the flame will be most unstable near the leading edge, in the pyrolysis zone. The instability starts at the flame locus; the unstably suatified region between flame and wall is the source of the instability. 8.5. CONCLUSIONS In this study we have analyzed the hydrodynamic stability of the wind-aided flame spread problem in the ceiling configuration. The basic state solution employs x‘m blowing, the boundary layer assumption, and the restriction to self-similar profiles of T; this last condition requires that the flame temperature along the entire length of the flame be constant, which is valid for Re—roo. The inviscid parallel flow analysis introduced the Froude number as the sole global nondimensional parameter. For 206 0.3 . I ' I ' I ' I I I ' I Temp profile# 5 31.125 0.2‘ fl "' is" . . ' / 0.1—J . .. Fro=4 Fr°=2 ‘ 0.0 t . l ' I ' I ' I ' I ' I 0.0 0.2 0.4 0.5 0.8 1.0 1.2 ' 1.4 Figure 8.7 Variation of frequency ((00) with —Im(oro ). Curves plotted are for tempera- ture profile #5 for Fro equal to 2 and 4; B=1.125. . 207 higher Froude numbers ( > 8) the flames are essentially stable and the detailed tem- perature profile has little effect on the stability of the flame. For lower Froude numbers (< 8) the buoyancy effect is appreciable and the form of the temperature profile is crucial to the stability calculation. The infinite-reaction rate flame tempera- ture profile predicts very high frequency flames which increases with sueamwise dis- tance. Thus, for low Froude numbers a parameter representing the detailed flame tem- I perature distribution is necessary; the exuemely localized flame vicinity dramatically influences the stability of the whole problem. 208 APPENDIX 8.A: REACTION ZONE SOLUTION FOR TEMPERATURE A brief discussion is given here. Details can be found in Linan (1974). By using suetched variables §=B(n—nf )la and T=(9+c §)/Bb Eq. (8.9) reduces to 1m O§§=8[92—§2]e b , (8.A.1) as B—too, where de/d §=:tl as C—>ioo and 8(x )=2va2xD /(bQ B3) is the reduced Dam- kohler number. Here _-(ao+ap) _£( _ ) ._ -——(ao_ap) , a—b (10 OF d“ (0f) -1 dn ' and B is the nondimensional activation energy, B=E ' (Tf—T; )/(R I T ,3 2). When c=0 the heat losses from the flame toward both sides are equal. Since the objective here is to examine the effect of temperature profiles on stability, in a qualitative sense, we will assume for simplicity that c=0. This gives (using definitions of a0, ap, b0 and 1),.- given in Sec. 8.2) l—Tw "f " 2—Tw -—T,, ' For Tw=0.3046 (since T;=635 K for PMMA), T,.=l/7 (since 7.1/T; ~ 1n)we get uf =0.448, which in turn requires B = 1.125. For the present case a=1.82/b with Q=l4.12 (from Eq. (8.17) for Yo..=0.233). Now using the first term of the expansion 8 5=50+__1_+... and putting 86’3b =7 and using the inner variables defined above, we 5 209 obtain a single graph for 9 versus é. Now using B = 0.6 to 4 we obtain various graphs in Fig. 8.3. Solutions for profile numbers 3 and 8 are discussed in Appendix 8.B. 210 APPENDIX 8.B: ASYMPTOTIC APPROXIMATION FOR T' AS n—mf Near n=nf the most important term in Eq. (8.28) is the last one since (u--uf)‘2 there becomes very small. The solution for 6 in this region thus depends strongly on T'(u —uf )‘2. Using Taylor series and neglecting (u —u ”)4 and higher order terms we obtain (u -uf)-2 = (fl-nf)-2(f”)_2 since f”'(nf )=f (hf)=0 Use of the inner variables defined in appendix 8.A gives T’ _ -92 are WWLa] We observe that 9 is symmetric in § since we have assumed c=0 (Appendix 8.A); hence, in the expansion for 6 = 90 + 9,: + 92E2 +---, we have 61: 93= 95:0. Thus the term in Eq. (88.1) becomes infinite as §—>O. When 92:0, the above term becomes firs—tang as §—>0. Thus when 6 = 90 + 94:4 +--- Eq. (8.B.1) is continuous near nf; 9 (see Eq. (8.28)) will be continuous for 0Sn0, and u'>o, we have u"'0, lu—c, |>O, (u -c, )2+c,-2>O, and sgn (T ')>0 when n<11f (in the fuel region). Thus, in our case we expect the fuel region, in which den- sity increases as we move upwards toward the wall (away from the flame), to be the source of the instability. When g —> - g (F r0 <0), the oxidizer region will be the source of the instability. CHAPTER 9 CONCLUSIONS SYNOPSIS In this chapter some conclusions are drawn and possible future work is suggested. 214 215 9.1. CONCLUSIONS In this study a detailed theoretical and experimental investigation of wind-aided flame spread over vaporizing solids is performed. Effects of free stream velocity and oxygen mass fraction, unsteadiness of the solid, surface radiative heat loss and surface heat flux near the leading edge on pyrolysis and the flame front speed (location) are investigated. A detailed investigation of the thermal degradation of a vaporizing solid is performed. Also performed is a stability analysis of wind-aided flame spread across a ceiling. In Chapter 2 a simplified heat-transfer model of wind-aided flame spread was developed. An explicit expression for the pyrolysis front speed was obtained: 2 l e e t t t 2 U; - U; p,‘).;c; T: - TI. erfc(-M W) It was found that the pyrolysis front speed varies linearly with both free stream velo- city and oxygen mass fraction. It compares favorably with the experiments of Chapter 4. Also an expression was obtained for flame tip location (flame reattachment point): , , Y’ale‘ Uf = Upcosec2 £- 0 ,I PS. . (9.2) 2 1 + Y0a/VYFS In Chapter 3 the model of Chapter 2 was further examined. Numerical results were obtained for excess pyrolyzate and surface heat flux in the preheat zone. The numerical results were correlated to provide expressions for the surface heat flux in the underflame preheat zone: 216 0.6 grain/E =1+5[§gp_] , (9.3) gTP(§p)\]E; gf-{P and in the preflame preheat zone: g (a): P. 3 T f = — , 9.4 81‘,(§f)§f [ a] ‘ ’ The excess pyrolyzate was nearly 50% for all oxygen concentrations. In Chapter 4 the pyrolysis and flame front locations and production rates of major chemical species were measured as functions of time for coflow flame spreading over PMMA samples for various flow speeds and oxygen mass fractions. It was found that both the pyrolysis front and the flame front speeds vary linearly with the free stream velocity, as predicted by the model of Chapter 2. With respect to the free stream oxy- gen mass fraction, the pyrolysis front speed was also affected by internal reflections at the higher oxygen concentrations. Under completely adiabatic conditions (no reflections) the dependence approaches as predicted by Eq. (9.1). It was also observed that the flame front location was very close to pyrolysis front location. In Chapter 5, mass loss rate and in-depth temperature measurements were obtained during the pyrolysis of PMMA samples in a nitrogen environment for various heat fluxes. An explicit expression was obtained for the mass loss rate: . » 4;;(1) A0 z: 1— _ , 9.5 mp (X) L [ 14(1)] ( ) 217 In Chapter 6 a model for wind-aided flame spread over an unsteadily vaporizing solid was developed. An explicit expression for the surface convective heat flux under the flame was obtained: , .. k U3 Re,“2 Yo... qcom, (x ) = 0332-qu 'x— AH T-C pg (Tf -T..) (9.6) Also obtained was an expression for the maximum flame length as a) function of the pyrolysis front location: x_r_ __1:B_ " xp -[1+Y0,Jv] ’ (9'7) where n is equal to 2 for laminar flow and 1.25 for turbulent flow. It was found that inclusion of surface radiative heat loss is necessary to result in a phenomena observed during the experiments of Chapter 4, flame spread stoppage before reaching the end of the sample. I In Chapter 7, burned PMMA samples were used to determine the surface heat flux distribution in the burning zone using a regression model. An explicit expression for the regressed surface location was also obtained: _ "" ki" _ '- .(.) = 4m". 4M] . (9.3) It was determined that surface convective heat flux near the upstream leading edge decreases instead of rising in proportion to 16"“2 as predicted by boundary layer 218 approximation. This, coupled with the surface radiative heat loss and the unsteadiness of the solid are the main reasons for the shorter flame overhang (not much excess pyrolyzate) observed during experiments of Chapter 4. In Chapter 8 a detailed inviscid stability analysis was performed which suggests that in the ceiling configuration the flame is stable because of high rates of forced con- vection. For smaller Froude numbers it is essential to include gas-phase finite reaction rates in the analysis. Thus, for development of comprehensive models it is essential to relax the boun- dary layer approximation in the gas-phase, at least near the leading edge. It is also important to include detailed surface boundary conditions and transient effects in the solid. Till now in the Ostudy we have not mentioned anything about the choice of ther- mophysical properties. As can be seen from Eq. (9.1) the correct choice of gas-phase properties plays an important role in comparisons with the experiments. Generally they are kept constant for all free stream conditions. In Fig. 9.1 is plotted the product of density, thermal conductivity and specific heat as a function of temperature for vari- ous free stream oxygen mass fractions. Clearly in this temperature range an error of 20% can be encountered because of the choice of conditions under which the proper- ties are selected. When comparing experiments to theory, an error of this order can thus be treated as normal and can be attributed to inaccurate properties values. kpcp/(kpcp)¢ . m 219 IL05 P o ‘t’ l I I l .° a: ‘1” I ' I I I ' I ' I ‘\ (kpcp). “I. 30.76 kP/m‘K'see lJllJlllLlllljllj IJLAILI 0.30 ' T r I ' r ' I I I I 400 600 800 1000 1200 TEMPERATURE (K) Figure 9.1 Variation of kpCp as a function of temperature 1400 220 9.2. SUGGESTIONS FOR FUTURE WORK In the present study we have not included the effects of soot radiation, gas-phase kinetics, turbulence and chairing of the solid. Because of the inclusion of so many non-linear effects the next stage of modeling should be oriented towards numerical modeling. Elliptic equations should be used in the gas-phase which could be simplified to parabolic equations later downstream. Because of this gas-phase kinetics should also be included. Modeling should not be limited to determination of spread rates but should be able to predict species produc- tion and depletion rates, which is very important from the pollution point of view. Experimental efforts should also head in the same direction. LIST OF REFERENCES 221 LIST OF REFERENCES Abramowitz, M. and Stegun, I. A. (1972), Handbook of Mathematical Functions, Dover, New York. Abu-Zaid, M. (1988), Effects of water on ignition of cellulosic materials, PhD Thesis, Michigan State University, East Lansing, MI. Agrawal, S. (1992a),"0n thermal degradation of a vaporizing solid", Spring Technical Meeting of Central States of the Combustion Institute, Columbus, OH, April. Agrawal, S. (1992b), "Determination of surface heat flux in the burning zone during wind-aided flame spread using burned PMMA samples", submitted. Agrawal, S. and Atreya, A. (1992), "Wind-aided flame spread over an unsteadily vaporizing solid", Twenty-Fourth Symposium (International) on Combustion. The Combustion Institute, Pittsburgh, in press. Agrawal, S. and Wichman, I. S. (1992a), "Heat flux distributions in wind—aided flame spread", Combust. Sci. T echnol., vol. 81, pp. 25-43. Agrawal, S. and Wichman, I. S. (1992b), "Effect of reaction zone temperature profiles on pulsation frequency of flames spreading with the flow across a ceiling", Com- bust. Flame, in press. Annamalai, K. and Sibulkin, M. (1979a), "Flame spread over combustible sm'faces in laminar flow systems: 1. Excess fuel and heat flux", Combust. Sci. Technol., vol. 19. pp. 167-183. Annamalai, K. and Sibulkin, M. (1979b), "Flame spread over combustible surfaces for laminar flow systems Part II: Flame height and fire spread rates", Combust. Sci. Technol., vol. 9, pp. 185-193. Anonymous (1987), Annual Book of ASTM Standards, Vol 04.07, p.332, Philadelphia, PA, USA. Anonymous (1988), Physical Properties of Acrylite GP, CYRO Industries, N. 1., USA. Atreya, A. (1983), Pyrolysis, ignition and fire spread on horizontal surfaces of wood, PhD Thesis, Harvard University, Cambridge. 222 Atreya. A. and Wichman, I. S. (1989), "Heat and mass transfer during piloted ignition of cellulosic solids", J. Heat Transfer, vol. 111, pp. 719-725. Atreya, A. (1990), Personal communication. Atreya, A. and Mekki, K. (1992), "Heat transfer during wind-aided flame spread on a ceiling mounted sample", Twenty-Fourth Symposium (International) on Combus- tion. The Combustion Institute, Pittsburgh, in press. Baum, H. R. (1980), "An alternative derivation of some flame spread integral equa- tions", Combust. Sci. Technol. vol. 23, pp. 79- 82. Beier, R. A. (1982), Soot and radiation in combusting boundary layers, PhD Thesis, University of California, Berkeley, CA, USA. Bird, R. 3., Stewart, W. E. and Lightfoot, E. N. (1960), Transport Phenomena, Wiley and Sons, Inc., New York. Buckmaster, J. D. and Ludford, G. S. S. (1982), Theory of Laminar Flames, Cam- bridge University Press. Buckmaster, J. and Peters, N. (1986), "The infinite candle and its stability- A paradigm for flickering diffusion flames", Twenty First Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, pp. 1829-1836. Carrier, G. F., Fendell, F. E. and Feldman, P. S. (1980), "Wind-aided flame spread along a horizontal fuel sla ", Combust. Sci. Technol., vol. 23, pp. 41-78. Carrier, G. F., Fendell, F. E. and Fink, S. (1983), "Towards wind-aided flame spread along a horizontal charring slab: The steady-flow problem", Combust. Sci. Tech- nol., vol. 32, pp. 161-209. Carrier, 6., Fendell, F., Butler, G., Cook, 8. and Morton, C. (1990), "laminar wind- aided flame spread across a thick horizontal fuel slab", Combust. Sci. Technol., vol. 72, pp. 17-36. Chen, C. H. and Tien, J. S. (1986), "Diffusion flame stabilization at the leading edge of a fuel plate", Combust. Sci. Technol., vol. 50, pp. 283-306. Delichatsios, M. M., Mathews, M. K. and Delichatsios, M. A. (1991), "An upward fire spread and growth simulation", Fire Safety Science - Proceedings of the Third International Symposium, pp. 207-216. deRis, J. N. (1969), "Spread of a laminar diffusion flame", Twelfth Symposium (Inter- national) on Combustion, The Combustion Institute, Pittsburgh, pp. 241-249. deRis, J. N. (1979), "Fire radiation - A review", Seventeenth Symposium (Interna- tional) on Combustion, The Combustion Institute, Pittsburgh, pp. 1003-1016. 223 DiBlasi, C., Crescitelli, S. and Russo, G. (1987), "Numerical modeling of flow assisted flame spread", Proceedings of the Eight Intl. Conf. on Compt. Methods in Applied Sci. and Engg., Versailles, pp. 473-485. DiBlaSi, C., Crescitelli, S. and Russo, G. (1988a), "Near limit flame spread over thick fuels in a concurrent forced flow," Combust. Flame, vol 72, pp. 205-215. DiBlasi, C., Crescitelli, S., Russo, G. and Fernandez-Pello, A. C. (1988b), "Model of the flow assisted spread of flames over a thin chaning combustible", Twenty- second Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, pp. 1205-1212. Drazin, P. G. and Howard, L. N. (1966), Advances in Applied Mechanics, vol 9, Academic Press, New York. Emmons, H. R. (1956), "The film combustion of liquid fuel", Zeit. fur Angew. Math. and Mech., vol 36(2), pp. 60-71. Emmons, H. R. (1980), "Scientific progress on fire", Ann. Rev. Fluid Mechanics, vol. 12. pp. 223-236. Fernandez-Pello, A. C. and Williams, P. A. (1975), "Laminar flame spread over PMMA surfaces", Fifteenth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, pp. 217-231. Fernandez-Pello, A. C. (1978), "A theoretical model for the upward laminar spread of flames over vertical fuel surfaces", Combust. Flame, Vol. 31, pp. 135-148. Fernandez-Pello, A. C. (1979), "Flame spread in a forward forced flow", Combust. Flame, vol. 36, pp. 63-78. Fernandez-Pello, A. C. and Mao, C. P. (1981), "A unified analysis of concurrent modes of flame spread", Combust. Sci. T echnol., vol. 26, pp. 147-155. Fernandez-Pello, A. C. and Quintiere, J. G. (1982), "A simplified model of radiating- turbulent upward flame spread over the surfaces of a charring combustible", Fall meeting of Eastern Section of the Combustion Institute, Atlantic City, N. 1., December. Fernandez-Pello, A. C. and Hirano, T. (1983), "Controlling mechanisms of flame spread", Combust. Sci. Technol., vol. 32, pp. 1-31. Femandez-Pello, A. C. (1984), "Flame spread modeling," Combust. Sci. T echnol., vol 39, pp. 119-134. Fjortoft, R. (1950), "Application of integral theorems in deriving criteria of stability of laminar flow and for the baroclinic circular vortex", Geofys Publ., vol. 17(6), pp. 1-52. 224 Friedman, R. A. (1968), "A survey of knowledge about idealized fire spread over sur- faces", Fire Res. Abs. Review, vol. 10, p. 1. Grant, A. J. and Jones, J. M. (1975), "Low-frequency diffusion flame oscillations", Combust. Flame vol. 25, pp. 153-160. Hasemi, Y. (1986), "Thermal modeling of upward wall flame spread", Fire Safety Sci- ence - Proceedings of the First International Symposium. pp. 87-96. Hines, A. L. and Maddox, R. N. (1982), Mass Transfer, Prentice-Hall, Inc. Hirano, T., .and Akita, K. (1979), "Ignition of polymethylmethacrylate and flame spread over its surface", Joint meeting ACS/ JCS, Honolulu, April. Hirata, T., Kashiwagi, T., and Brown, J. E. (1985), "Thermal and oxidative degrada- tion of poly(methylmethacrylate): weight loss", Macromolecules, vol. 18, pp. 1410-1418. Holman, J. (1987 ), Experimental methods for engineers, McGraw-Hill Book Company. Inaba, A. and Kashiwagi, T. (1986), "A calculation of thermal degradation initiated by random scission. 1. Steady-state radical concentration", Macromolecules, vol. 19, pp. 2412-2419. Inaba, A. and Kashiwagi, T. (1987), "A calculation of thermal degradation initiated by random scission. unsteady radical concentration", Eur. Polym. Journal, vol. 23, pp. 871-881. Incropera, F. and DeWitt, D. P. (1985), Introduction to Heat Transfer, John Wiley and Sons. Kashiwagi, T. and Ohlemiller, T. J. (1982), "A study of oxygen effects on flamming transient gasification of PMMA and PE during thermal irradiation", Nineteenth Symposium (International) on Combustion, The Combustion Institute, pp. 815- 823. Kashiwagi, T., Hirata, T. and Brown, J. E. (1985), "Thermal and oxidative degradation of poly(methylmethacrylate): molecular weight", Macromolecules, vol. 18, pp. 131-138. - Kashiwagi, T., Inaba, A., Brown, J. E., Hatada, K., Kityama, T. and Masuda, E. (1986), "Efl'ects of weak linkages on the thermal and oxidative degradation of poly(methylmethacrylates)", Macromolecules, vol. 19, pp. 2160-2168. Kashiwagi, T., Omuri, A. and Nanbu, H. (1990), "Effect of melt viscosity and thermal stability on polymer gasification", Combust. Flame, vol. 81,‘ pp. 188-201. ' Kays, W. M. and Crawford, M. E. (1980), Convective Heat and Mass Transfer, McGraw-Hill Book Company. 225 Khalturinski, N. A. and Berlin A1. (1983), "Degradation and Stabilization", A series of Comprehensive Reviews, p. 1, Ed. J ellinek, H. H. G., Elsevier, Amsterdam. Kimura, I. (1965), "Stability of laminar-jet flames", Tenth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, pp. 1295-1300. Krishnamurthy, L. and Williams, F. A. (1973), "On the temperatures of regressing PMMA surfaces", Combust. Flame, vol. 20, pp. 163-169. Kulkarni, A. K. and Fisher, S. (1988), "A model for upward flame spread on vertical wall", Fall meeting of Eastern Section Combustion Institute, Clearwater Beach, December. Lengelle, G. (1970), "Thermal degradation kinetics and surface pyrolysis of vinyl poly- mers", AIAA Journal, vol. 8, pp. 1989-1996. Loh, H. T. and Fernandez-Pello, A. C. (1984), "A study of the controlling mechanisms of flow assisted flame sprea ", Twentieth Symposium (International) on Combus- tion, Pittsburgh, pp. 1575-1582. Linan, A. (1974), "The asymptotic structure of counterflow diffusion flames for large activation energies", Acta Astronautica, vol. 1, pp. 1007-1039. Linz., P. (1989), Personal communication, University of California, Davis. Magee, R. S., and McAlevy, R. F 111. (1971), "The mechanism of flame spread", J. Fire Flammability, vol. 2, p. 271. Markstein, G. H. and deRis, J. (1972), "Dynamics of textile fires: A study of the mechanisms- theory and experiment", Factory Mutual Research Corporation, Report No. 20588. Markstein, G. H., and deRis, J. (1973), "Upward fire spread over textiles", Fourteenth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, p. 1015. Mekki, K., Atreya, A., Agrawal, S. and Wichman, I. (1990), "Wind-aided flame spread over charring and non-chairing solids: An experimental investigation", Twenty Third Symposium (International) on Combustion, The Combustion Insti- tute, Pittsburgh, pp. 1701-1707. Mekki, K. (1991), Wind-aided flame spread over chaning and vaporizing solids, PhD Thesis, Michigan State University, East Lansing, MI. Michalke, A. (1965), "On spatially growing disturbances in an inviscid shear layer", J. Fluid Mechanics vol. 23, pp. 521-544. Nurbakhsh, S. (1989), Thermal decomposition of charring materials, PhD Thesis, Michigan State University, East Lansing, MI. 226 Orloff, L., deRis, J. and Markstein, G. H. (1974), "Upward turbulent fire spread and burning of fuel surfaces", Fifteenth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, pp. 183-192. Pagni, P. J., and Shih, T. M. (1976), "Excess pyrolyzate", Sixteemh Symposium (Inter- national) on Combustion, The Combustion Institute, Pittsburgh, pp. 1329-1340. Parker, W. (1977), "An investigation of the fire environment in the ASTM E 84 tunnel test", NBS Technical Note 945, NBS, Washington, DC. Puri, I. K. and Hammins, A. (1991), "Prediction of the pulsation frequency of pool fires", Fall Technical Meeting of the Eastern Section of the Combustion Institute, Ithaca, N. Y., paper no. 92. Quintiere, J. (1981), "A simplified theory for generalizing results fi’om' a radiant panel rate of flame spread apparatus", Fire and Materials, vol. 5(2), pp. 52-60. Quintiere, J., Harkleroad, M. and Hasemi, Y. (1986), "Wall flames and implications for upward flame spread", Combust. Sci. Technol., vol 48, pp. 191-222. Rayleigh, J. W. S. (1880), "On the stability, or instability of certain fluid motions", Proc. Land Math Soc., vol. 9, pp. 57-70. Roberts, G. E. and Kaufman, H. (1966), Tables of Laplace Transforms, W. B. Saunders Co., Philadelphia. Saito, K., Quintiere, J. G. and Williams, F. A. (1986), "Upward turbulent flame spread," Fire Safety Science - Proceedings of the First International Symposium. pp. 75-86. Saito, K., Williams, F. A., Wichman, I. S. and Quintiere, J. G. (1987), "Upward tur- bulent flame spread on wood under external radiation", 1987 ASME/AIChE National Heat Transfer Conference, Pittsburgh. Saric, W. S. and Nayfeh, A. H. (1977), "Nonparallel stability of boundary layers with pressure gradients and suction", Agard Symposium, Technical University of Den- mark, Copenhagen, Denmark, May, paper no. 6. Sibulkin, M. and Kim, J. (1977), "The dependence of flame propagation on surface heat transfer 11: Upward burning, Combust. Sci. Technol., vol. 17, pp. 39-49. Sibulkin, M., Kulkarni, A. K. and Annamalai, A. K. (1981), "Effects of radiation on the burning of vertical fuel surfaces", Eighteenth Symposium (International) on Combustion, The Combustion Institute, pp. 611-617. Sirignano, W. A. (1972), "A critical discussion of theories of flame spread across solid and liquid fuels", Combust. Sci. Technol., vol. 6, pp. 95-105. 227 Sirignano, W. A. (1974), "Theory of flame spread above solids", Acta Astronautica, vol. 1. PP. 1285-1299. Steckler, K. D., Kashiwagi, T., Baum, H. R., and Kanemaru, K. (1991), "Analytical model for transient gasification of noncharring thermoplastic materials", Fire Safety Science - Proceedings of the Third International Symposium, pp. 895-904. Taylor, G. I. (1950), "The instability of liquid surfaces when accelerated in a direction perpendicular to their planes. 1", Proc. Roy Soc. London, vol. A 201, pp. 192-194. Tewarson, A. and Pion, R. F. (1976), "Flammability of plastics, I. Burning intensity", Combust. Flame, vol. 26, pp. 85-103. Toong, T. Y., Salant, R. F., Stopford, J. M. and Anderson, G. Y. (1965), "Mechanisms of combustion instability", Tenth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh, pp. 1301-1313. Van Dyke, M. (197 5), Perturbation Methods in Fluid Mechanics, The Parabolic Press, Stanford, California. ’ VanWylen, G. J. and Sonntag, R. E. (1973), Fundamentals of Classical Thermodynam- ics, John Wiley and Sons. . Vovelle, C., Akrich, R. and Delfau, J. (1984a), "Thermal degradation of solid materials under a variable radiant heat flux", Twentieth Symposium (International) on Combustion, The Combustion Institute, pp. 1647-1654. Vovelle, C., Akrich, R. and Delfau, J. (1984b), "Mass loss rate measurements on solid materials under radiative heating", Combust. Sci. Technol., vol. 36, pp. 1-18. Vovelle, C., Delfau, J., Reuillon, M., Bransier, J. and Laraqui, N. (1987), "Experimen- tal and numerical study of the thermal degradation of PMMA", Combust. Sci. Technol., vol. 53, pp. 187-201. Weeratunga, S., Buckmaster, J. and Johnson, R. E. (1990), "A flame-bubble analogue and its stability", Combust. Flame, vol. 79, pp. 100-109. Wichman, I. S. (1983), Studies of flame spread in an opposed flow over surfaces of solid fuels, PhD Thesis, Princeton University, Princeton, N. J., 1983 Wichman, I. S. and Williams, F. A. (1983), "A simplified model of flame-spread in an opposed flow along a flat surface of a semi-infinite solid", Combust. Sci. Technol., V01 32’ ppo 91’123. Wichman, I. S. and Baum, H. R. (1988), "An integral analysis of two simple model problems of wind-aided flame spread," J. Heat Transfer, vol 110, pp. 437-441. Wichman, I. S. and Agrawal, S. (1991), "Wind-aided flame spread over thick solids", Combust. Flame. vol 83, pp. 127-145. 228 Williams, F. A. (1976), "Mechanisms of fire spread", Sixteemh Symposium (Interna- tional) Combustion, The Combustion Institute, Pittsburgh, pp. 1281-1294. Williams, F. A. (1985), Combustion Theory, The Benjamin] Cummings Publishing Company, Inc., Menlo Park, California. Zhou, L. and Fernandez-Pello, A. C. (1991), "Turbulent bunting of a flat plate sur- face", Fire Safety Science - Proceedings of the Third International Symposium, pp. 415-424.