. - . ‘4; h fifimtf {‘“r‘, $3.: ~.‘ ‘Q i ‘ “ 3“”; “1.4" 3 “ "' a” vi ‘ o- ‘.. 37".. “c" ‘7 JP... q, .- ‘W W. WM, ma; we, .‘u 1.4 "3“,; - .xzs.‘ ~ w ‘M, 1 g. x "if .fi‘ 4' —, 3 ’5: 133*»,- {3‘7 ‘ w urn , m ”.mp4": ‘ lv-I‘ fiJLEi‘TrQiE -. a i ‘3'}.‘3g-NJ: IICJCJ m mums. ', '2'21'tgrll ~~ rt, w w 9m '<~ Qua-.1. 31". 5'; .. m... - . . 1c, :3 r ‘1‘"!‘1; . . 'r. :13??? 'I . , .‘ >H' ,'...‘ ._ ‘ " " ‘ . ..‘.;,‘.".',v‘5 w ' ‘ . x -- -. ' v , “ * 5.23.5,“ “ ‘- u , K , .. , . w . '3‘ . ~ . '0 T , mung-gr“; . ..’ ' n "'2.“ 12"“ .13” xi“ 1' . 7‘ “.5 J v A “htow MICHIGAN TATE UNIVERSITY L imuumzni mumummunm’ril‘lliil I! J 3 93 009117106 This is to certify that the dissertation entitled WIND—AIDED FLAME SPREAD OVER CHARRING AND VAPORIZING SOLIDS presented by KAMEL EL MEKKI has been accepted towards fulfillment Of the requirements for DOCTOR OF PHILOSOPHY degreein MECHANICAL ENGINEERING flzmfl 417% E Date 6/xx/q/ MS U is an Affirmative Action/Equal Opportunity Institution 0- 12771 , \ LIBRARY Michigan State University K r PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. DATE DUE DATE DUE DATE DUE "4 L‘ 3 a A.’ ’ in g 2 £13}; ,_ I :l l- fil_|l——l l J MSU Io An Affirmative Action/Equal Opportunity Institution ' omens-pt WIND-AIDED FLAME SPREAD OVER CHARRING AND VAPORIZING SOLIDS By Kamel El Mekki A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering 1991 ABSTRACT WIND-AIDED FLAME SPREAD OVER CHARRING AND VAPORIZING SOLIDS By Kamel El Mekki This work presents the results of a detailed experimental investigation and analysis of laminar forced flow wind-aided flame spread over wood and PMMA in the ceiling configuration. This investigation includes measurements of flame spread rate and heat and mass transfer. In the flame spread work, the speed of propagation of the pyrolysis front and the flame front and the production rates of the major chemical species were measured as a function of time. The effect Of the free stream velocity and the oxygen mass fraction on the flame spread rate and the production and depletion rates Of major chemical species were investigated for both wood and PMMA. The effect of external radiation was investigated only for wood, because use of external radiation for PMMA results in excessive melting and dripping. It was found that all these environmental parameters, and especially the oxygen mass fraction, control the flame spread rate and the chemical species production rates. In the heat transfer part of this work, the effect of wind speed and ambient oxygen mass fraction on heat transfer to the surface underneath and ahead of the flame tip during wind-aided flame spread were investigated experimentally. The data were correlated according to a simple theoretical model. The experiments were also performed in the ceiling configuration. High-temperature ceramic solids containing surface and in-depth thermocouples were used downstream of the bumin g PMMA sample to measure the net heat transfer to the surface. Simultaneous measurements of the PMMA surface temperature (to determine the pyrolysis front), ceramic solid- phase temperatures (to determine the conduction heat transfer), gas-phase temperatures (to determine the convective heat transfer), and the flame tip location were made. Since the excess pyrolyzate (which produced a flame front that was only slightly ahead of the pyrolysis front) was low for all experiments, the heat transfer measurements underneath the flame were conducted with a natural gas diffusion flame in the boundary layer over the ceramic solids mounted in the ceiling configuration. For these laminar flames, it was found that convection by the hot gases is the dominant mode of heat transfer to the surface ahead of and underneath the flame. The radiation component becomes more significant as the free stream oxygen mass fraction increases. Surface temperatures and flame spread rate equations were derived from these heat transfer correlations. It was found that the flame spread rate is linearly with free stream velocity, as expected, and varies to the 1.5-power of the free stream oxygen mass fraction. To my father; Mohammed, my mother; Zohra, my wife; Dalila, and my children; Wafa and Karim for their confidence, devotion, and understanding in helping me attain this important milestone of my life. iv Acknowledgement The author is greatly thankful to his advisor Professor Arvind Atreya. His constant guidance, encouragement and direct involvement in every step of the work was of invaluable help in the completion Of this work.I deeply appreciate his patience and his friendship. I would also like to thank the members of my Committee, Dr. I. S. Wichman, Dr. J. Beck, Dr. K. Mukherjee and Dr. P. Hunt for their input and guidance. I would like to thank my colleagues and friends; especially Sanjay Agrawal for his help in conducting the flame spread experiment on PMMA. This work was supported by USDA under contract number 86-FSTY-9-90192. Table Of Contents LIST OF TABLES .............................................................................................. LIST OF FIGURES ........................................................................................... NOMENCLATURE ............................................................................................ Chapter One: Introduction and Literature Review ................................... 1.1 Background ................................................................................................... 1.1.1 Pyrolysis ............................................................................................ 1.1.1.1 Pyrolysis Of non-charting solids (PMMA) ................................... 1.1.1.2 Pyrolysis of charting solids (WOOD) .......................................... 1.1.2 Flame spread ...................................................................................... 1.1.3 Heat transfer to the preheat zone ......................................................... 1.2 Prospect Of this work ...................................................................................... Chapter Two: Experimental Apparatus and Procedure .......................... 2.1 Small-scale combustion wind tunnel ................................................................ 2.1.1 Inlet Section ....................................................................................... 2.1.2 Turbulence manipulation section ......................................................... 2.1.3 Test section ........................................................................................ 2.1.4 Extemal radiation source .................................................................... 2.1.5 Exhaust section .................................................................................. 2.2 Gas analysis equipment .................................................................................. 2.2.1 Total Hydrocarbons [THC] ................................................................ 2.2.2 H20 meter ......................................................................................... 2.2.3 C0—C02 meter ................................................................................. 2.2.4 02 meter ............................................................................................ 2.3 Data acquisition ............................................................................................. Q’i. O‘MUJNN 12 14 16 16 18 19 23 23 25 28 28 28 28 30 30 2.4 WOOD and PMMA samples preparation ........................................................ 30 2.5 Ceramic samples preparation .......................................................................... 31 2.6 Flame spread experimental procedure ............................................................. 31 2.7 Heat transfer experimental procedure .............................................................. 33 2.8 Error analysis ................................................................................................ 36 2.8.1 Mass balance ...................................................................................... 36 2.8.2 Pyrolysis-front spread rate .................................................................. 40 2.8.3 Heat flux ............................................................................................ 40 Chapter Three: Flame Spread over Vaporizing Solids .......................... 42 3.1 Results .......................................................................................................... 43 3.1.1 Temperature measurements ................................................................. 43 3.1.2 Flame spread rates .............................................................................. 45 3.1.3 Species production rates ..................................................................... 52 3.1.3.1 Burning zone behavior during flame spread ................................ 55 3.1.3.2 Incompleteness Of combustion .................................................... 57 3.2 Discussion ..................................................................................................... 59 Chapter Four: Flame Spread over Charring Solids ................................. 60 4.1 Results .......................................................................................................... 61 4.1.1 Temperature measurements ................................................................. 61 4.1.2 Pyrolysis and flame fronts .................................................................. 63 4.1.3 Spread rates ....................................................................................... 66 4.1.4 Species production rates ..................................................................... 68 Chapter Five: Convective and Radiative Heat Transfer to the Solid ........................................................................................................................... 79 5.1 Formulation ................................................................................................... 80 5.2 Results and Discussion .................................................................................. 82 5.2.1 Transient Heat Flux Measurements ..................................................... 82 5.2.2 Steady-State Heat Flux Measurements ................................................. 90 vii 5.2.3 Surface Temperature ........................................................................... 94 5.2.4 Flame Spread Rate ............................................................................. 97 Chapter Six: Conclusions .................................................................................. 100 Appendix A: Study of Number and Location of Thermocouples in a Ceramic Solid for Heat Flux Computations ......... s ..................................... 103 Appendix B: Data of Flame Spread Experiments on PMMA ................ 115 Appendix C: Data of Flame Spread Experiments on WOOD ............... 120 Appendix D: Data of Heat Transfer Experiments ..................................... 134 List Of References ................................................................................................ 139 viii Table 2.1 Table 2.2 Table 4.1 LIST OF TABLES Mass balance for CH4 diffusion flame. Mass balance for PMMA diffusion flame. Empirical formula Of Poplar wood [Atreya 1984]. Figure 1.1 Figure 1.2 Figure 1.3 Figure 1.4 Figure 1.5 Figure 1.6 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 2.9 Figure 2.10 Figure 2.11 Figure 2.12 Figure 2.13 Figure 2.14 Figure 2.15 LIST OF FIGURES PMMA surface temperature and mass flux histories during pyrolysis experiment (Y o..=0.233. q"=3W/cm 2) [Vovelle 1987]. Mass flux behavior of PMMA sample under different rates of external radiation [Vovelle 1984]. Typical mass flux history Of a wood sample during pyrolysis [Nurbakhsh 1989]. Schematics Of different modes of flame spread relative to gravity. ............... Schematic of upward flame spread on wood sample. PMMA surface temperature history during flame spread experiment. ........... Schematic diagram Of the experimental apparatus. Velocity profile at the leading edge Of the sample (x=0) before and after installing the frame metal screen. Frame metal screen used in the turbulence manipulation section. ................. Velocity profiles at the leading edge Of the sample. Optimum combination Of glass beads, honeycombs, and metal screens used in the turbulence manipulation section. Cross section of the test section and the heater assembly. Screens configuration for radiation scattering, resulting into almost uniform incident radiation along the sample surface. Measured heat flux profiles along the sample (x=0 corresponds to the leading edge Of the sample). Schematic diagram Of the mixing chamber. Gas analysis equipment. Schematic Of wind-aided flame spread on wood. Schematic of wind-aided flame spread on PMMA with ceramic solids mounted downstream for transient heat transfer measurements. Schematic Of natural gas diffusion flame over ceramic solids for steady-state heat transfer measurements. Diagram Of the porous-metal burner used in heat transfer experiments. ’I‘henaturalgascomesonlyfromthesidesofthembes. Mass balance test experiment on CH1. diffusion flame. 20 20 21 22 27 29 32 35 37 Figure 2.16 Figure 2.17 Figure 2.18 Figure 3.1 Figure 3.2 Figure 3.3 Figure 3.4 Figure 3.5 Figure 3.6 Figure 3.7 Figure 3.8 Figure 3.9 Figure 3.10 Figure 3.11 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Mass production rates histories during wind-aided flame spread on PMMA (U..=O.9m/s. and Yo..=0.233). Pyrolysis-front history on PMMA sample (U..=O.9m/s and Yo..=0.233). ........... Measured incident and convected heat fluxes repeatability (U..=O.9m/s, Yo..=0.233, x=321n.). Measured surface temperatures during flame spread on PMMA (U..=O.9m/s, Y...=O.233). Flame-front history comparison between video records (dashed lines) and surface temperature measurements (solid lines). Pyrolysis-front and flame-front histories during flame spread experiments on PMMA. Dependence Of the pyrolysis and flame-from speeds on the free stream velocity for PMMA (Yo..=0.233). Dependence Of the pyrolysis and flame-front speeds on the free stream oxygen mass fraction for PMMA (U..=O.9m/s). Dependence Of the measured flame radiation at the center Of the test section bottom for different oxygen mass raction for PMMA (xF70cm). Picture of flame spreading on PMMA (a) U..=O.9m/s, Yo..=0.233 (b) U..=O.9m/s, Yo..=0.6l. Dependence of the species mass production rates on the pyrolysis-front location for different wind speed (Y o..=0.23: l-U..=O.6m/s; 2-U..=0.9m/s; 3-U..=l.5m/s). Dependence of the species mass production rates on the pyrolysis-front location for different oxygen mass fraction (U..=O.9m/s; l-Yo..=0.233; 2-Yo..=0.43; 3-Yo..=1.0). Normalized meg/mm: with the corresponding mass based stoichiomeu‘ic fraction (i.e. the value of 1 corresponds to complete combustion). ' Normalized mega/mo with the corresponding mass based stoichiometric fraction (i.e. the value Of 1 corresponds to complete combustion). Measured surface temperatures and their rate Of change with time for wood during flame spread (U..=O.6m/s, Y,.=O.233, £1"=0.5W/cm2). Measured surface temperatures for wood during flame spread (U..=O.9m/s, Y,_=O.233, q"=1.3W/cm2). Pyrolysis and flame front histories (Y 0.50.233, £1"=0.5W/cm 2). Pyrolysis and flame front histories «050233, q"=O.9wIcm 2). 39 41 41 46 47 49 50 53 S4 58 58 62 62 Figure 4.5 Pyrolysis and flame front histories (Yo..=0.233, O"=l.3W/cm 2). ........................ 65 Figure 4.6 Pyrolysis and flame front histories (U..=O.9m/s, q"=0.5W/cm 2). ........................ 65 Figure 4.7 Dependence Of the pyrolysis and flame front speeds on the free stream oxygen mass fraction (U..=O.9m/s, (1"=0.5W/cm2). 67 Figure 4.8 Dependence of the pyrolysis and flame front speeds on the external radiation (Yo..=0.233). 69 Figure 4.9 Dependence Of the pyrolysis and flame front speeds on the free stream velocity (Y o..=0.233). 69 Figure 4.10 Species production rates history during the 3 phases of wind-aided flame spread on wood (U..=O.6m/s, Y,..=O.233, q"=l.3W/cm2). 70 Figure 4.11 Dependence of the species mass production rates on the pyrolysis-front location for different wind speeds (Y,.=0.233. d"=o.SW/cm 2). ........................ 71 Figure 4.12 Dependence Of the species mass production rates on the pyrolysis-front location for different speeds (Yo..=0.233, Q"=O.9W/cm2). - 71 Figure 4.13 Dependence Of the species mass production rates on the pyrolysis-front location for different wind speeds (Y,.=O.233, q"=l.3W/cm 2). ........................ 72 Figure 4.14 Dependence Of the species mass production rates on the pyrolysis-front for different oxygen mass fraction (U..=O.9m/s, q"=0.5W/cm 2). ....................... 72 Figure 4.15 Normalized Elect/limo: with the corresponding mass based stoichiometric fraction (i.e., the value Of 1 corresponds to complete combustion). 75 Figure 4.16 Picture Of a spreading flame on wood sample (U..=O.9 m/s, Yo..=0.233. c'f'=0.5W/cm2). 75 Figure 4.17 Instantaneous pictures taken of a spreading flame on a wood sample (U..=O.9m/s. Yo.=0.61,t'1"=0.5W/cm2). 77 Figure 4.18 Normalized mcoytnmo with the corresponding mass based stoichiometric fraction (i.e. the value Of 1 correspond to complete combustion). - 78 Figure 5.1 Schematic diagram of the model. 80 Figure 5.2 Comparison of the measured and the curve fitted gas-phase temperature during flame spread on PMMA (U..=O.9m/s, Yo..=l.0 , x=33cm, film). 84 Figure 5.3 Heat flux measurements during flame spread on PMMA (U..=O.9mls,Y.,..=0.233). 85 Figure 5.4 Heat flux measurements during flame spread on PMMA (U..=O.9m/s, Y...=1.0). 85 xii 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 5.14 Appendix A Figure A.l Figure A.2 Figure A.3 Figure A.4 Figure A.5 Figure A.6 Figure A.7 Figure A.8.1 Figure 16.8.1 Nondimensional measured and predicted convective heat transfer during flame spread on PMMA. Schematic diagram of an expanding plane underneath the sample surface. Measlned and predicted flame radiation for different Yo... Net, total and surface radiation heat fluxes histories (U.=l.2m/s. xFZS.4cm, x=17.7cm). Measured gas-phase temperature profiles Of a methane diffusion flame (U..=O.9mls, Yo..=0.233. xF2S.4cm) Measured convective and radiative heat fluxes underneath and downstream of steady porous-metal burner diffusion flames (Y 0..=0.211). Nondimensional measured convective heat flux underneath and downstream Of porous-metal burner diffusion flames (Y...=O.211). ........ Nondimensional measured and predicted PMMA surface temperatures. Measured and predicted pyrolysis-front spread rate dependence on ........... Measured and predicted pyrolysis-flout spread rate dependence on Y... (U..=O.9mIs). Experimental setup. Power step input history. Temperature histories at the surface and indepth Comparison of measured and smoothed temperature data. Computed heat flux history using one in-depth thermocouple measurements. Temperature residuals when using one in-depth thermocouple measurements for heat flux calculations. Computed heat flux history using 2 in-depth thermocouple mearnrrements. Temperature residuals when using 2 in-depth thermocouple measurements for heat flux calculations. Temperature residual of the 2lul thermocouple when using xiii 88 88 89 91 91 93 93 96 98 98 105 105 106 106 107 107 109 109 Figure A.9 Figure A.10 Figure A.ll Figure A.12 Figure A.13 Figure A.14 Appendix B Figure B.Txx Figure B.prx Appendix C Figure C.Txx Figure C.prx Appendix D Figure Door 2 in-depth thermocouple measurements for heat flux calculations ................ Computed heat flux history when using the thermocouple at x=0 and another in-depth. Computed heat flux history when using the thermocouple at x=0 and another 2 in-depth. Effect Of the number Of thermocouples used on the computed heat flux. Effect Of 2mm location error Of the in-depth thermocouple on the heat flux results (using 2 thermocouples). Effect Of 2mm location error Of the in-depth thermocouple on the heat transfer results (using one thermocouple). Effect of properties error on the computed heat flux (using all the thermocouple measurements). Species mass production rate histories for PMMA. Species mass production rate plotted against the pyrolysis-front for PMMA. Species mass production rate histories for wood. Species mass production rate plotted against the pyrolysis-front for wood. Heat flux measurements during flame spread on PMMA. xiv 110 110 111 111 113 113 114 NOMENCLATURE C = Specific Heat f = Mass-based stoichiometric coefficient h = Convective heat transfer coefficient AH = Heat Of combustion per unit mass of PMMA k m" = Thermal conductivity = Mass flux Nu = Nusselt number Pr = Ptandtl number c1" = Heat flux Re = Reynolds number i = Time T = Temperamre U,V _= Velocity x = Streamwise coordinate y = Transverse coordinate Y = Mass fraction Greek = Thermal diffusivity p = Density 9 = T—T. Subscripts f = Flame g = Gas 0 = Oxygen p = Pyrolysis s = Surface T = Total x = at x co = free stream CHAPTER ONE Introduction and Literature Review Wood is, and has been, for centuries a common construction material. But unfor- tunately, wood is combustible and is involved in hazardous building fires. In residen- tial rooms or buildings, once wood is ignited, fire spreads in different directions depending on the location of the wood with respect to gravity (floor, ceiling, wall). The spread of fire in a building goes through complex interactions between physical, heat and mass transfer, and chemical reaction processes. Thermal radiation from a ceiling layer Of hot gases and hot walls is the primary mode of energy transfer in a fire to the as-yet-unburned materials. The flame-spread rate su'ongly depends upon the magnitude of this incident heat flux, which causes the surface temperature to reach the ignition (or pyrolysis) temperature. At this temperature, the solid starts to pyrolyze and feed combustible gases into the fire plumes. As the fire burns in a closed room, the oxygen concentration decreases to a point where the flame can no longer exist. Hence, the oxygen concentration is also expected to significantly afiect the rate of flame spread Finally, as windows and doors are broken or burned, air currents develop, which along with buoyant currents cause the flame to spread in difl'erent directions. This results in wind-opposed (fire spread and incoming air in opposite directions) and wind-aided (fire spread and incoming air in the same direction) modes of flame spread. The three environmental conditions identified above, namely; incident heat flux, ambient oxygen concenu'ation, and wind speed; are the primary external parameters that conuol the flame-spread rate over combustible mamials like wood [Emmons 1974]. The room-fire problem has been approached in two difl'erent ways: some investi- gators have focused on understanding the effect Of building geometry on fire growth [Emmons 1980, Quintiere 1980], while others have conducted experiments searching for the flammability characterisrics and behavior of fire growth on different building materials, such as polymethylmethacrylate (PMMA), paper, wood, etc. However, the results of some of these experiments are different because the investigators failed to control some of the environmental parameters [Emmons 1980]. Such independent con- trol of wind speed and oxygen mass fraction have proved to be very helpful in identi— fying the controlling mechanisms of wind-opposed flame spread [Fernandez-Pello et a1. 1981]. Only one similar experimental investigation of wind-aided flame-spread, on PMMA, has been reported in the literature [Loh and Fernandez-Fella 1984]. This investigation did not consider external radiation as one of the variables. Experimental studies of wind-aided flame spread on (thick) charting materials like wood have not been reported. In this work, wind-aided flame-spread over wood is extensively studied for all the above mentioned environmental parameters. Similar experiments are also conducted on PMMA for comparison with the available results of Loh and Fernandez-Fella (1984). 1.1 Backgmund 1.1.1 Pyrolysis To study flame spread over combustible solids, one has to understand the solid- phase decomposition (pyrolysis) of the material, because the incident heat flux from the existing flame leads to gasification of the material producing the fuel, which in turn sustains the flame. Thus, the rate of the solid gasification is central to the rate of flame spread. The pyrolysis Of non-charting and chaning solids represents two extreme cases of solid-phases degradation. For PMMA, the solid is totally consumed and essentially "vaporizes". However, for wood, a char layer is formed at the sm'face while the pyrolysis fiont propagates into the solid. ' 1.1.1.1 Pyrolysis Of vaporizing solids (PMMA) As the surface temperature reaches the pyrolysis temperature, the solid starts to pyrolyze and surface regression is observed as gasification continues. For a constant incident heat flux, the mass flux was found to reach a steady state as illustrated in Fig- ure 1.1 [V ovelle et al. 1987]. Tewarson and Plan (1976) observed that the mass loss rate correlates as m = —-— (1.1) where O = incident flux (W/cmz) Ow = heat loss from the surface (W/cmz) LG = heat Of gasification 0/3) + heat conducted into the solid This linear dependency on the incident heat flux was also confirmed by Vovelle et al. (1984). However in a real fire, the incident heat flux at the surface increases as the flame approaches, therefore a transient mass loss rate would result. Such a study has been carried out by Vovelle et al. (1984) when they examined the effect of known variable incident heat flux on the mass loss rate of both PMMA and particle board in an inert atmosphere. As expected, the mass loss rate of the PMMA sample rises with the heat flux as shown in Figure 1.2. 400 Ts (°C) 10 4 (g/szs) —4 E3 X 'E *2 Olvr'fi""lr T'T'tro O 500 1000 1500 2000 Time (see) Figure 1.1 PMMA surface temperature and mass flux histories during pyrolysis experiment (Y...=0.233, q"=3W/cm 2) [Vovelle 1987]. 2E-03 .; 0 const. heat flux expt. .0 e 1.66X1O’3W/cm’.s El 3.33X10'3W/cm’s . o 2E—03— ”a? N . E o \ 113—034 3 "E ESE—04— OE+OO 1 5 ll 2 q (W/ cm ) Figure 1.2 Mass flux behavior of PMMA sample under different rates Of external radiation [V ovelle 1984]. 1.1.1.2 Pyrolysis Of charring solids (WOOD) As soon as the sample is exposed to an external heat, the so-called inert heating stage starts. During this stage, the temperature rises without decomposition. However, some moisture evaporation occurs. This phenomenon holds true until the pyrolysis temperature is reached at the surface. At this moment the pyrolysis stage and the release of volatiles are initiated. As the surface temperature continues to increase, the volatile mass flux rapidly increases until it reaches a maximum when a thin char layer starts to form at the surface. Then as the char layer thickness increases, the volatile mass flux gradually decreases. This phenomenon, shown in Figure 1.3, has been Observed by a number of investigators [Vovelle et a1. 1984, Au'eya 1983, Nurbakhsh 1989]. 2E-O3 O '1 . . 0 . 0 g 0 O O ”a? 15—034 , 'o .. N g o ' E ‘ \s n. < O .0 ./ ~..~ on o .o f“ l “E 5E-O4— : O O OE+OO += , , . I O 100 200 300 400 Time (sec) ' Figure 1.3 Typical mass flux history of a wood sample during pyrolysis [Nurbakhsh 1989]. Due to its lower thermal conductivity than the original wood, char becomes a bar- rier to the incoming heat flux and prevents it from passing into the inward propagating pyrolysis zone. In addition, the re-emitted radiant flux by the char surface increases owing to its high emissivity. This causes the conduction heat transfer to the in~depth virgin wood to drop in time. Thus, for these two different combustible materials, wood and PMMA, one expects that during wind-aided flame spread: For PMMA, the flame will stay attached to the leading edge as the flame tip propagates downstream. For wood, however, an extinction-front would propagate behind the pyrolysis-front due to the lack of fuel caused by the thick char-layer. 1.1.2 Flame Smead Pyrolysis-front spread rate and heat transfer to the solid ahead and underneath the flame have been investigated, both experimentally and theoretically, for difl'erent materials and at different angles relative to gravity [Loh and Femandez-Pello 1984, Atreya 1983, Saito et a1. 1987, Fernandez-Polio and Williams 1975, Femandez-Pello and Hirano 1983, Saito et a1. 1986]. Figure 1.4 shows qualitatively the effects of grav- ity and air flow on propagating flames. Clearly, the most hazardeous cases are wind- aided flame spread in the ceiling configuration and upward (buoyancy driven) flame spread. The flame in these configurations is pushed closer toward the unwa sur- faces, thus increasing the heat transfer by the hot gases to the as-yet-unbumed sur- faces. Several theoretical papers have been published on the upward buoyancy driven flame spread [Saito et al. 1986, Femandez-Pello 1978, Annamalai and Sibulkin 1979, Markstein and deRis 1973, Orloff and deRis 1976, Sibulkin and Kim 1977, Fernandez—Pello and Quintiere 1982, Kulkarni and Fisher 1988]. All of these models are in qualitative agreement with the experimental results. Fernandez-Pello (197 8) developed a model that predicred flame heights larger than the experimental data. <— —> s\ a) Horizontal flame spread —> - _> N —> -—> —> —> s\ .3 —> b W-A fl spread fl ) ame on oor d) W-A flame spread on ceiling _>. L” \‘ —> <— —-> <— s 3’ c) W—O flame spread on floor I . . e) W-O flame spread on ceiling is s i f) Upward flame spread g) Downward flame spread Figure 1.4 Schematics of different modes of flame spread relative to gravity. This larger predicted flame height increases the heat flux to the surface, causing higher flame-spread rate predictions. Saito et al. (1986) conducted experiments on both PMMA and Douglas-fir and developed a model for upward flame spread. They predicted that the pyrolysis front spread rate goes as = . (12) where the characteristic ignition time 1: depends on the fuel properties, and the heat flux (assumed constant) from the flame to the surface underneath, and xf and x13 are the flame-front and the pyrolysis-front locations, respectively, as shown in Figure 1.5. The experimental results agree very well with the predictions and previous theories for PMMA. However for wood, the authors did not use external radiant heaters; and, therefore, no well defined continued flame spread occured. Thus, their prediction of an initial acceleratory spread, before reaching steady state, for charting material could not be confirmed. Figure 1.5 Schematic of upward flame spread on wood sample. A number of models for wind-aided flame spread have also been rier et al. 1980, Carrier et a1. 1983, DiBlasi 1987, Wichman and At these models have focused on vaporizing solids. Recently Wichman an. (1991) and Carrier et. a1. (1990) have also developed such models. These models predict steady flame spread rates, and assume a steady state in the pyrolyzing zone. Wichman and Agrawal (1991) predicted the flame spread rate as 2 2 Vp ___ p..Cp.L. Tr-Tp exp(-M2Pr) . (1.3) This forrnulaticn shows the dependency on solid and gas properties, flame temperature and wind speed. The experimental study and the model of Loh and Fernandez-Pena (1984) for PMMA were in agreement, and the pyrolysis front spread rate was found to be: (1.4) v n __w.- [my 9 °‘ p,C,3., Tp—T, ' The above equations look similar, but discrepancies exists between the results Of Loh and Femandez-Pello (1984) and those of Wichman and Agrawal (1991) and the com- plicated mathematical analysis and numerical solution Of integral equations by Carrier et a1. (1980). These two latter models predict that Vp~ Yo”; however, Loh and Femandez-Pello concluded from their analysis and experiments, that Vp~ Y3... This certainly points out the complexity of the wind-aided flame spread process and the need to experimentally control the external conditions carefully. Also it seems essen- tial to evaluate the validity of the several simplifying assumptions made in the theoreti- cal models. Transient solutions, with finite rate gas phase kinetics, were obtained by DiBlasi et aL (1987, 1988a) using a finite-difference scheme. They concluded that as the oxy- gen mass fraction decreases, the flame temperature and the heat fluxes to the unburned fuel surface drop. This causes a slower fuel production rate, and consequently, a slower spread rate. At a very low oxygen mass fiaction, the effects of finite kinetics 10 appear mainly at the leading edge, where extinction begins. Previous investigators have primarily focused on non-charting materiala because of the difficulty in simultanously analyzing the solid-phase processes for charting solids. Canier et a1. (1983) made the first and only attempt toward the development of models for wind-aided flame spread over charring solids. They treated the char and the virgin wood as solids with different thermal properties. The surface temperature was considered to rise to th pyrolysis temperature at the arrival of the pyrolysis front, then to keep rising until some temperature T, at which time char starts to chemically erode to form gas. The transient problem was formulated, but only the steady state part Of the solution was obtained. The numerical solution for both Blasius-type and Oseen-type flow were also obtained. While the flame temperature was found to be invariant under both flow-field approximations, the char-layer thickness and the flame position under the Blasius-type flow were about twice as large as those under the Oseen-type flow. The char layer thickness was also found to be sensitive to both the temperature T,, and the latent heat. DiBlasi et al. (1988b) solved numerically the problem of wind-aided flame spread. over thin charring solids. Their results show that the pyrolysis front spread rate ini- tially accelerates while the flame foot is still at the leading edge. As the burn-out front starts to propagate, the pyrolysis front spread rate decelerates. Steady state is then reached when both the pyrolysis front and the bum-out front spread rates are equal. If the burn-out front spread rate keeps accelerating, extinction occurs. These results were in agreement with previous experimental data. However, the dependency of the pyro- lysis front spread rate on the flow velocity was not fully resolved. Although the experiments indicate that the spread rate becomes constant for relatively high flow speeds. the model predicts that the spread rate keeps increasing. All the above models for wind-aided flame spread are for the laminar case and can be classified as thermal models where one step infinitely fast gas-phase chemical reaction (va+v°O——)products) is assumed. Finite rate chemical reactions were found 11 to have an effect only at the leading edge and the flame tip, since flame temperature drops at these locations [DiBlasi et a1. 1987, 1988a, 1988b]. The flame spread rate is determined by the rate at which the surface temperature is raised to the solid pyrolysis temperature (i.e. Vp=dxpldt where T,=Tp (or Tig) at xp). The flame spread rate depends on how fast the solid surface temperature is raised to its pyrolysis temperature, which depends on the convective and radiative heat flux from the flame. The convective heat flux is proportional to the gas-phase temperature gra- dient at the sm'face, which depends on both the flame temperature and the boundary layer thickness. Therefore, the most important variables that affect this mode of flame spread are the wind speed. which controls the boundary layer thickness, and the oxy- gen mass fraction, which controls the flame temperature. The above equations show the flame spread rate dependency on these two parameters. Flame spread experiments over PMMA samples are conducted to compare the results with those of Loh and Fernandez-Fella (1984) and to seek the reason(s), if any, for the above-mentioned discrepancies. Previous experiments on PMMA have proved that the heat flux by flame radiation and convection by the hot gases is sufficient to bring the solid surface temperature to its ignition temperature and allow the flame to spread. Therefore, only the wind speed and the ambient oxygen mass fraction have been studied for PMMA. For the experi- ments on wood, however, a minimum external radiant flux level is necessary for the spread to occur [Saito et a1. 1987]. The magnitude of this incident heat flux and the preheat time have a significant effect on the flame spread rate. The incompleteness of combustion as well as flame radiation,which both affect the flame temperature, were not previously studied and are investigated in detail in the present study. The objec- tive is to provide a complete physical understanding of this mode of fire spread over charring and non-charting solids. 12 The combustion products were only measured and analyzed by Atreya (1983) in his pyrolysis, ignition, and fire spread on horizontal surfaces of wood, and by Abu- Zaid (1988) and Nurbakhsh (1989) in their pyrolysis experiments on wood. The flame spread results of Atreya (1983) show that the measured average mass flux is initially proportional to the 4/3 power of the fire radius, then stays constant due to the balance between the attenuation due to charting and increase in fire size, and finally becomes inversely proportional to the fire radius due to char build-up. In this work, the mass production rate of all the major chemical species are measured and analyzed. 1.1.3 Heat transfer to the solid Figure 1.6 shows a typical surface temperature hiStory during flame spread on PMMA. The rise of the solid surface temperature from the ambient temperature T. to the pyrolysis temperature T1) at the pyrolysis front is due both to both the preheating upstream of the flame tip x>xf and the preheating underneath the flame xpmxm< DUED LE Q w ....... . I a E 92:85 30 ....... m x m m awn—>250 1H. w A— way—Em m m m n mambwm: PZSQ \c; 009 Ohm O_© O_.v CW 0 e e paeeeee o t o lwd o o to; o o I 0000 OOO iN.—. 000000000 0 00 000000 (S/w) n 22 3...; g E iiiilINUlllllllllllllllllllllllllllHHHHi 5:3 g niillIlHllIIIIIIIHIIIIIIIHIIIIIIHIniii} gag J 1k Optimum combination of glass beads, honeycombs, and metal screens used in the turbulence manipulation section. Figure 2.5 23 2.1.3. Test section The useable portion of the tunnel test section is 0.81 m long and 0.153 m wide. The test sample (0.76 m long, 0.076 m wide and 0.019 m thick) is placed horizontally along the tunnel top while the bottom of the tunnel is hinged at the inlet allowing the tunnel depth to be adjusted. The tunnel depth at the inlet is 0.1 m but can be increased to 0.13 m at the exhaust end. This provides a maximum 30% increase in the cross-sectional area at the exhaust end to compensate for the acceleration of the gas core because of boundary layer grthh 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 1x10‘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 first 0.5 m were used. The RMS level of turbulent fluctuations inside the tunnel was found to be less than 1% of the free- stream velocity. Furthermore, the measurements of the velocity profile inside the tun- nel (see Figure 2.4), show that the flow is laminar. 2.1.4. External radiation source External radiation on the sample surface was provided from below by two types of electrically-operated radiant heaters. The first type consists of three high-temperatm'e (with maximum filament temperature of 1230 K) quartz electrical heaters (10 in x 10 in) placed at the bottom of the heater assembly shown in Figure 2.6. The second type consists of six U-shaped Chromlox coil heaters (3/8 in diameter, Incoloy sheath, type UTU, each 1.8 KW) installed above the quartz heaters. The heaters are separately con- trolled by two 3-phase 440-Volt variable transformers. These heaters are housed in an insulated box with the inside frame sides being highly reflective aluminum sheets. This housing is covered by water-cooled shutters which slide out at the beginning of the experiment. 24 Kawool insulation W009 sample Observation window Hinged bottom Water-cooled chutter Coil heaters Quartz electric heater Ceramic fiber ' Heater temperature insulation probe Figure 2.6 Cross section of the test section and the heater assembly. 25 This heaters assembly is suited to simulate external radiation in building fires. The incident radiation from the heaters passes through an optical glass window (0.153 m wide and 0.76 m long) contained in the hinged bottom portion of the tunnel. About 70% of the infrared radiation is transmitted to the sample surface. However, due to the view factor. the radiation is a maximum at the center of the sample and drops by about 30% at the two ends of the sample. To overcome this difficulty, three screens have been installed between the heaters and the infrared optical glass window to scatter the radiation. After a number of trials with the width and number of screens, an optimum combination (see Figure 2.7) was configured such that the radiation meas- ured at the sample surface was uniform to within i3% over the entire length of the sample, as shown in Figure 2.8. The entire tunnel test section was maintained between 315 K and 335 K by cooling water. 2.1.5. Exhaust section The main part of the exhaust section is the mixing chamber. In order to obtain a representative gas sample for transient chemical analysis, the stratified products of combustion have to be well mixed with the core flow. This mixing process is acheived through the use of a combination of baffles, a series of electric tapes for large-scale mixing, and a net of electric resistance wires for small-scale mixing. A metal louver is added at the outlet of the mixing chamber to assure the mixing of gases as shown in Figure 2.9. The electric tapes and the net of electric resistance are heated to avoid condensation of heavy hydrocarbons or water. A well mixed representative sample of gases is then extracted through the sam- pling probe for chemical analysis. The rest of the flow then goes through a large chamber before getting sucked by the exhaust fan. This large chamber is used to suppress the mechanical oscillations created by the exhaust fan. 26 30” 1” 3” 1” .1 0.5” l 1 Cooling water 10” 70 Figure 2.7 Screens configuration for radiation scattering, resulting into almom uniform incident radiation along the sample surface. 1.0 0.9 a - t ‘ / \ 9 0.8+ // \ E U \ 0.7-1 E / 3 \ ‘cr 0.5- M 0-4 I ' I ' I ' I I I ' T O 10 20 3O 4O 50 60 X (cm) Figure 2.8 Measured heat flux profiles along the sample (x=0 corresponds to the leading edge of the sample). 27 .3220 «:38 05 he Ease 3.2.2.8 a." «Sufi Eon been» new on.» men? 2.3 ”.58: 03.5 683: V\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ \\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\_\\ \I W \ ‘l f L ‘Mul / T / / T hege— 2.3 ”:32 ”5.8: 28 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 H20 analyzers were heated by either electrical tape or hot water as shown in Figure 2.10. The gas was then passed through a cold-trap at -5°C and dried before passing through the 02 and the CO—COZ analyzers. Prior to every experiment, the gas analyzers were first zeroed with nitrogen gas flowing through, then they were adjusted to the proper reading for the known concen- tration of every specie of the calibration gas running through. 2.2.1. Total Hydrocarbons analyzer [THC] The total hydrocarbons were measured by a Flame Ionization Detector (FID) gas chromatograph GC—3BF, which used a 40%H2—60%He mixture for fuel. The FID had a very good response time and a time constant of 1.0 second. 2.2.2. Water analyzer The water vapor concenu'ation 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 10.2 °C and a time constant of 1.0 second. 2.2.3. CO-CO2 analyzer An infrared IR702 nondispersive dual gas analyzer was used for CO and C02 concentration measurements. The flow was dried before entering the analyzer. The meter had a good accuracy of 11% of full scale (CO : 0-3%, 0-12%; C02 : 0-6%, 0- 20%) and a time constant of 2.6 seconds. 29 Dew Point Meter Heated line Figure 2.10 Gas analysis equiprneut. 30 2.2.4. 02 analyzer The oxygen concentration was measured by a Beckman OM-ll polarographic analyzer. This meter also required dry flow. The meter had an accuracy of 10.1% and a time constant of 1.5 seconds. 2.3. Data acquisition The analog signals from thermocouples, gas analyzers and the radiometer were fed to a data acquisition/control unit (HP3497A). The data was 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 was taken at 12 readings per second with a 10‘6 Volts mm. 2.4. WOOD and PMMA sample preparation The samples used for the flame spread experiments were obtained from clear boards of poplar and white sheets of PMMA. These were first conditioned at room humidity and temperature (50% RH and 294 K) and then instrumented with nine ther- mocouples on the surface 0.05 m apart. These thermocouples were made fi'om fine chromel and alumel wires 76 pm in diameter. The method employed to install the thermocouples on wood surface was developed by Atreya (1983). Two very fine layers of wood were skinned off from the surface using a sharp razor blade, then a portion of the thermocouple wire of each side on the junction was secured underneath each layer using wood glue. A very fine drop of this glue was also put on the thermocouple junc- tion. The assembly was then allowed to dry under a heavy weight for about 2 hours. For the PMMA samples, however, an electric heat gun was used to heat the surface so that the thermocouple wires stick. 31 2.5. Ceramic samples preparation The available tools for measuring the heat flux at the surface are water—cooled wide-angle radiometers. However, these radiometers are inappropriate because they provide total heat flux measurements to a water-cooled surface rather than to an already heated wall whose surface temperature is changing with time. Since wood chars and PMMA melts and drips dming flame spread, indepth temperature measure- ments are difficult and inadequate for incident heat flux computations. To overcome this problem, ceramic detectors with surface and indepth thermocouples were developed. These high temperature ceramic samples were cast with surface and indepth ther- mocouples (6 in length by 3 in width and l in thickness). The optimum design for the number and location of the thermocouples was experimentally determined. It was found that a total of 3 thermocouples, one at every boundary and one indepth near the surface exposed to external heat flux, provide a sufficient number of temperature his- tories for accurate incident heat flux computations. For accurate results, all the ceramic samples have one thermocouple at each boundary and three indepth. The exact location of the indepth thermocouples and thermal properties, at different temperatures, of each ceramic sample, were carefully determined [Beck and Arnold 1977]. See Appendix A for details. These ceramic detectors were also painted black to make their emissivities near unity. 2.6 Flame spread experimental procedure The test samples (30 in long, 3 in wide and 3/4 in thick) were placed horizontally along the tunnel top. The wood samples were obtained from clear boards of poplar. and the PMMA samples were obtained from white sheets of PMMA. They were first conditioned at room temperature and humidity and then instrumented with nine ther- mocouples two inches apart on the surface. Once all the desired conditions were set, i the sample was ignited with a small methane porous metal burner placed at the inlet 32 with its face parallel to the sample surface as shown in Figure 2.11. The fuel flow rate to the igniter was controlled such that the flame overhang on the sample surface was about 0.02m. After ignition, the physical process that occurs inside the tunnel is schematically shown in Figure 2.11. Surface temperature and species concentration histories were collected and stored in the computer. In addition, video records of the flame tip were taken. The surface temperature measurements provided the pyrolysis front arrival, and the video records showed the transient flame tip position, while the species concentration measurements provided the mass flux during the flame spread process. This mass flux and the pyrolysis front spread rate would reveal the burning zone condition and the incompleteness of combustion. woon Xf ‘ SAMPLE xp¢ Igniter I Thermal B.L. WIND ‘— DIRECTION <— EXTERNAL RADIATION Figure 2.11 Schematic of wind-aided flame spread on wood. 33 2.7 Heat transfer experimental procedure To study the heat transfer in the preheat zone ahead of the flame tip, a spreading flame on 10 inch and 16 inch PMMA samples upstream of the ceramic detectors was used. Time varying solid and gas-phase temperatures were measured. Video records of the flame tip were also collected. Figure 2.12 shows a schematic of a spreading flame on PMMA sample upstream of the ceramic detectors. The heat transfer underneath the flame, however, will be studied for the case of a natural gas diffusion flame existing in the boundary layer over the solid surface in the ceiling configuration, as illustrated in Figure 2.13. The porous-metal burner used in the latter case was increased in size to 4 inches and specially constructed (see Figure 2.14) to reduce the exit velocity at the porous metal surface for long flame experiments. A lower exit velocity allows the flame to be more buoyantly dominated. This will certainly cause the porous metal burner diffusion flame to have a stand-off distance close to that of the excess pyro- lyzate in the overfire region during an actual flame spread experiment under similar environmental conditions. Inverse heat conduction calculations were used to obtain the net (incident) heat flux as a function of time. The convective heat transfer component was derived from the temperature gradient at the solid- gas interface. The re-radiation heat flux com- ponent was computed from the knowledge of the surface and surrounding temperatures and the surface emissivity. Finally, the flame radiation component of the total heat flux was computed by applying an energy balance at the solid-gas interface. The effect of the free stream velocity and the ambient oxygen concentration on the flame convective and radiative components of the total heat transfer to the solid was studied. 34 Ceramic solids Water cooled plate Schematic of wind-aided flame spread on PMMA with ceramic solids Figure 2.12 mounted downstream for transient heat transfer measurements. Burner e V Ceramic 3 g solids ' g x é! .— Flame 3 ; i _ U y ‘— Water cooled plate Figure 2.13 Schematic of natural gas diffusion flame over ceramic solids for steady-state heat transfer measurements. 35 3.5” Cooling- water tube \ \\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\\ UT .7, ”‘J 3” Figure 2.14 Diagram of the porous-metal burner used in heat transfer experiments. Thenaturalgascomesonlyfromthesidesofthetubes. 36 2.8. Error analysis The experimental determination of any parameter is based upon measurements, which usually contain errors. Two kinds of errors exist: uncertainty or random errors and systematic errors. The uncertainty errors in the measurements can be neglected since the data were collected by the computer through a data acquisition unit with a 10"6 Volts accuracy. Systematic errors, however, fall into one of two categories: 1. Calibration errors in the measurement devices. 2. Neglecting significant outside influences. Since these systematic errors can exist in the measurements, a great deal of atten- tion has been focused on calibrating the gas analyzers (errors of the 1“ kind) and keep- ing atmospheric pressure in the test section to prevent leakage (errors of the 2“d kind). All the experiments were repeated to confirm the results. Accuracy of the species mass balance, the pyrolysis-front spread rate and the heat flux measurements are described in the following sections. 2.8.1. Mass balance 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 at which they were produced. Investigation of the existence of systematic errors was performed by checking the mass balance of all the species. A known mass flow rate of methane was introduced into the burner prior to igni- tion. To reduce errors due to incompleteness of combustion, the methane flow rate was originally set such that a blue flame was established after ignition. The results of such experimentation are shown in Figure 2.15, where the methane mass flow rate was set at 0.016 g/s. 37 0.08 x THC 0 C02 0 02-depl. A H20 6'“ ememmae A 0 0 < 3 on g ‘3 V O a, gymnawwmqbo 6 0.04-— 0 0° " feMmpmman-f (I) s O (n o no g 8 a 0.024 f ‘9 0 i too 200 300 400 500 600 700 Time (sec) Figure 2.15 Mass balance test experiment on CH4 diffusion flame. The following overall chemical reaction then describes the process (neglecting CO for- mation; as shown in Figure 2.15, CO concentration is small) CH4 + 202 —> C02 4" 2H20 The predicted and measured mass flow rate of all the species are summarized in Table 2.1. The largest error in the species measurements is about 3%, which is acceptable. 38 Table 2.1 Mass balance for CH4 diffusion flame. Specie Prediction Experiment %Error (9/8) (9/8) CH4 0.016 0.0155 02-dep1. 0.064 0.063 C02 0 .044 0 .043 H20 0.036 0.035 2.5 The accuracy of the species mass production rate measurements during an actual flame spread experiment can also be checked for the case of room air condition. Figure 2.16 shows the species mass production rates for the case of flame spread on PMMA where U..=O.9m/s and Y°=0.233. For PMMA, the overall chemical reaction is described by: C5H802 + 6 02 —> 5C02 + 4 H20 . The predicted and measured mass flow rates at t=1500 sec. (see Figure 2.16) are sum- marized in Table 2.2. Since the total mass consumption rate of the sample is not known, all the predictions are based on the oxygen depletion rate. The very low per- cent errors shown in Table 2.2 show the reliability of all the gas analyzers and provide confidence in the species measurements. 39 Table 2.2 Mass balance for PMMA diffusion flame. Specie Prediction Experiment %Error (9/8) (9/3) OZ-depl. 0.05 0.05 -- C02 0.06 0.058 3.3 H20 0.0188 0.019 1.3 . 0.10 X THC '1 0 C02 4' C0 0'08“ A 02-depl. A U H20 (0 E v 0.06“ B O 0: m 0.04“ (I) O J 2 o ' 500 1000 1500 2000 2500 Time (sec) Figure 2.16 Mass production rates histories during wind-aided flame spread on PMMA (U..=O.9 m/s and Y...=O.233). 40 2.8.2. Pyrolysis-front spread rate The accuracy of the pyrolysis-front spread rate was determined by the repeatabil- ity of the experiments. Hence, all of the flame spread experiments were repeated to confirm the results. Figure 2.17 shows three pyrolysis-front histories for three separate experiments conducted for the case where U..=O.9m/s and Yo..=0.233. The deviation in the pyrolysis-front spread rate is about 2% from its average value. 2.8.3. Heat flux The accuracy of the incident heat flux measurements can be analyzed in two ways: (i) the low level of the errors (RMS) of the output results of the inverse heat conduction calculations, (ii) and the repeatability of the experiments. The low RMS values show the good agreement between the measured and predicted temperature histories, which in turn shows that there is low error in the cal- culated heat flux. The repeatability, however, can be investigated by looking at the heat flux results of experiments repeated for the same condition. Figure 2.18 shows the incident heat flux with the corresponding convected heat flux for two separate tests where U..=O.9m/s, Yo..=0.233 and x=32in. The results of the two experiments are in very good agreement, with an average difference of 4% for the net heat flux and 5% for the convected heat flux. The good repeatability of the results for species measurements, pyrolysis-front spread rate, and heat flux measurements provide confidence that all the measured quan- tities are within acceptable experimental error bounds. 41 60 A E 50~ 8/ x0. +J 404 c O .f: | .2) 30” (f7 3‘ 8 Vp (cm/s) 3 20‘1 V 2.75E—2 O 2.73E-2 10 .fifi j Tar 2.72E-2 l I ' l ‘ T ' l l ' l H 400 600 800 1000 1200 1400 1600 1800 2000 2200 ~ Time (sec) Figure 2.17 Pyrolysis-front history on PMMA sample (U..=O.9 m/s and Y...=O.233). 0.7 <1“ (W/cmz) Figure 2.18 Measured incident and convected heat fluxes repeatability (U..-0.9m/s, Y...=0.233, x=32in.). CHAPTER THREE Flame Spread over Vaporizing Solids This chapter presents a detailed experimental investigation on laboratory-scale laminar wind-aided flame spread along a ceiling-mounted slab. The gas flow along the slab is forced, and its speed and composition are controlled. Recent models for non-charting materials [W ichman and Agrawal 1991, Carrier et 'al. 1990] have sparked a considerable controversy since they seemingly disagree with the detailed experimental study of Loh and Femandez—Pello (1984). The models sug- gest that the speed of the pyrolysis front varies linearly with the fiee stream oxygen mass fraction, while the experiments suggest a quadratic dependence. This clearly points out the need for further experimental investigation to determine whether any of the assumptions made during the development of the models or possible experimental errors are responsible for the discrepancy. Thus, the objectives of this work are to (i) provide a physical understanding that will serve as the basis for the development or refinement of theoretical models; and (ii) provide additional measurements for non- charring materials like PMMA. 42 43 3.1. Results As shown in Figure 2.11, the fuel vapors generated in the pyrolysis region which extends from x = 0 to x = XI, are burned in the diffusion flame, which extends from x = 0 to x = x; with xf > xp. The hot combustion products that flow downstream of x f and the flame extension (xf - xp) help to convectively and radiatively heat the pris- tine solid to a temperature (Tp) at which it begins to vigorously pyrolyze and contri- bute 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. 3.1.1. Temperature Measurements Figure 3.1 shows the measured surface temperatures for PMMA as the flame pro- pagates along the ceiling-mounted sample. These temperatures are typical of all the measurements. These measurements show a temperature plateau at about 643 K, which is taken as the melting or vaporization temperature. From visual observations it was found that the peak rate of change in the surface temperature after ignition occurs is at the instant the flame tip arrives at the thermo- couple location for both wood and PMMA. By writing a surface energy balance it can also be demonstrated that this peak corresponds to a sharp increase in the incident heat flux, which is caused by the flame tip arrival. Thus, the flame tip location, Xf, may be determined as a function of time from the measured temperature profiles by calculating the maximum value of dT,/dt. Results of such calculations show excellent agreement with xf determined from video records for both wood and PMMA. Figure 3.2 shows such agreement for PMMA where the solid lines correspond to the temperature meas- urements and the dashed lines correspond to the video records. 800 Q 700— / i, ‘3 B 600- O .1 L (I) O- 500— E 2 a) 400— U .9 S m 300— 200 ' ' ' T ' ' ' T ' ' ' l ' ' fl 1 ' ' ' 0 400 800 1200 1600 2000 Time (sec) Figure 3.1 Measured surface temperatures during flame spread on PMMA (U..=O.9m/s. Y,.=0.233). 60 50- E - U V 40— >2- .. U YO 5 301 dx 47 NN Annmarie <22“. .8 £8? 63.5.... 3.: 05 ca 38% 2:00-08“: 23 $32.3 05 we 8:8:800 1n 9:5...— Am\Ev 362m; Eombm 00E 01m. 0; 0; a; N; 0; 0.0 0.0 ed p _ . _ p _ _ _ b — . _ p _ . _ . F0.0 C> O a> e IN00 100.0 d 140.0 00.0 (S/LUO) 010.1 [303st 48 Figure 3.5 shows VI, and Vf plotted against the oxygen mass fraction (Y 0,,,.). Note that several different measurements of VP and Vf for PMMA (from reflective to blackened water-cooled bottom aluminum plate) are presented. It is once again noteworthy that Vp and Vf are nearly equal (see Figures 3.4 and 3.5). Figure 3.5 also shows that Vp depends upon the surface finish of the aluminum plate directly under the bunting sample. The aluminum plate was used only for PMMA since external radiation was not required. For the experiments on wood sam- ples described in the next chapter, this plate was replaced with an infrared optical glass window to allow external radiation. As is evident from Figure 3.5, the surface condi- tion of this plate (and to a lesser degree, the two vertical sides of the tunnel that con- tain the observation windows) considerably alters the flame spread rate because it reflects the flame radiation back to the sample surface. This effect magnifies as Yo. increases because the flame radiation increases. Measurements show that the flame radiation at the plate surface increases linearly from nearly zero at Y0... = 0.2 to 3.5 W/cmz at Y0. = 1 (Figure 3.6). Thus, for a reflectivity as small as 0.1, the reflected radiation becomes comparable with the external radiation used for the wood samples discussed in chapter 4. 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 Vp ~ Yggf. This differs from the previous measurements for PMMA [Loh and Fernandez-Per 1984] (dotted line in Figure 3.5), which show that vP ~ Y3... How- ever, the present results involving 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 disagreement (this corresponds to the systematic errors of the 2“‘1 kind described in chapter 2). Results for a water-cooled plate agree well with the recent theoretical predictions for PMMA [Wichman and Agrawal 1991, Carrier et al. 1990], which are shown plotted by the dashed line in Figure 3.5. 49 Eamoudv <25 .8 Sues as: 5:6 Each» 8b 05 so 88% 88-08.“: 23 £523 05 do 8:3:88 Wm «Sufi 80> c0500: 30:: cmmxxo o; No Ho eho whim. 23233.2. m>_Hom_.Cwm P b —y n — b _ h . Cgv Become. om coEcoS) - u $3: o_.ee|NeeceEe.._ e :3 Own—OOUKMBQB QMZMXUSm 23233.2 QmZMXUSm .. . m:<|_n_ EDZ_§DI_< 4:50 )KOCIQ [molmm Tmolmv molmo (s/w) aim pDGJdS 50 F; .302.le 5233 e8 .5th mng 5996 .555 ecu 883 .838 .8. on. .«o 383 05 3 cows—o8 2:3 @28on 05 do coco—ocean ed 0.53..— 80> c0500: mmoE cmmxxo 04 md md 50 md md To 00 md . _ b r . _ b b p p L _ p _ . _ .X 0 I— X 1N rm. <— LUZ) N\ pDJHb UOIlDIpDJ GLUDIj Z . . 51 The effect of the oxygen mass fraction on the flame spread rate can be further exam- ined by looking at the chemisrry and the flame temperature. The following one step chemical reaction can be used to describe the combustion process during flame spread: CXHyO + (x+-§--;-)02——) xC02 + %H20 (til) (mam) (thcoz) (dingo) + [CO,THC,SOOT] (that) (Incomplete Combustion) + XAfAHu—XR) . (3.1) (Heat Liberated) where xA is the correction factor due to incompleteness of combustion (x =1 for com- plete combustion), and am is the correction factor due to flame radiation losses (xR=0 for no flame radiation). If f is the stoichiometric fuel to oxygen ratio by mass, then fro. gm of fuel requires Yo. gm of 02 and liberates [waAAHa-xnn J of heat. This energy is utilized to (i) raise the temperature of a unit mass of (02 , N2) mixture from T, to Tf: [gar-Th]; ' (ii) raise the temperature of fYo, gms of fuel vapor from TI, to Tf: [CPCTf-TQwa]; (iii) provide for the heat required by the solid to produce fuel at the surface [QfYoa]. Thus, an energy balance would then lead to waAa-«xnmn = cpcrf—T“) + gut-rpm... + QtYo, (3.2) 52 After rearranging the terms the flame temperature can be expressed as fYoatxAAHa-xe-Ql—cpcrp—Te (Tr—T?) = Cp (1 + We») (3.3) where Q = L+Cp.(Tp-T..), where L is the latent heat of the solid. For complete combustion and no flame radiation [xA(1-xR)=l] Tf - Tp ~ Y0... which leads to vp ~ Y3“, since vp ~ (Tr-T92. However, as will be explained in section 3.1.3.2, the data show that xA decreases and 7m increases with Yo... This makes the flame temperature less dependent on the oxygen mass fiaction, and so also the pyrolysis-front spread rate. The results show that Vp ~ Ygfi. Figure 3.7 shows 2 pictures taken of the flame during the flame spread experiments (a) U..=O.9 m/s, Yo..=0.233, and (b) U..=O.9 m/s, Yo..=0.61. Notice the difference in the flame brightness which indicates the change in flame radiation inten- sity. 3.1.3. Species Production Rates The species mass production and depletion rate histories are presented in Appen— dix B for all the experiments performed. Figures 3.8 and 3.9, respectively, show the species production rates for different free stream velocities and different oxygen mass fractions plotted against the pyrolysis length, xp. Production rates of CO and total unburned hydrocarbons are not presented because they are two orders of magnitude smaller than the others; however, they are extremely important in fire research since they were the primary reasons for many lost lives in building and other fires. 53 (b) Figure 3.7 Picture of flame spreading on PMMA (a) U..=O.9m/s. Y°_--0.233 (b) U..=O.9m/s, Y°_=0.6l. 54 1E-01 ABE-02- m \ - 0‘ V4E—024 B . O (I C .9 . *5 1E-021 . we" . , ,- 385—031 13'," _ ‘o , - O L . 0- 45—03e (D J 1,’ tn 0 2 - ' H20 -u- 02—dem. -— CO 1E-O3 1 I l l n l7 l I 10 20 30 4O 50 60 708090100 Pyrolysis front Xp (cm) Figure 3.8 Dependence of the species mass production rates on the pyrolysis-front location for different wind speed (Y ,.=0.23; 1-U..=0.6m/s; 2-U..d).9tn/s; 3-U..=1.5m/s). CDC 9.0 00 [LL LlllL _LLL J_Lllll 1 till Mass Production Rate X102 (9/5) 8 O J 10 U 40 I ' 80 100 Pyrolysis Front Xp (cm) Figure 3.9 Dependence of the species mass [Induction rates on the pyrolysis-front location for different oxygen mus fraction (U..=O.9m/s; l-Y...=O.233; 2-Y...=0.43; 3-Y...=l.0). 55 The data for small x9 is omitted since the measurements are affected by the igniter flame and, hence, they are subjected to a large percentage error. Also, the soot pro- duction rate was not measured even though a substantial amount of soot was formed, especially during experiments at high Yo... 3.1.3.1. Burning zone behavior during flame spread 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 Yo..=0.43 for purposes of clarity (the rest of the data are in Appendix B). Vertical straight lines are obtained (see Figure 3.9). because xp does not change during this period since the pyrolysis front has already reached the end of the sample. The mass production rate more than doubles during this period before achieving a steady-state value for the case Yo..=0.43. The fact that the species production rate continues to change substantially even after the pyrolysis front has reached the end of the sample shows that the steady-state condition is not achieved in the boundary layer burning zone. Vovelle et al (1987) showed that the mass loss rate kept increasing even after the surface had started pyrolyzing as shown in Figure 1.1. Looking at this figure, the surface temperature reaches steady state at about 300 sec, however the mass flux reaches steady state at about 700 sec. Thus, the fact that the surface temperature in the burning zone has become steady does not imply that the mass loss rate has become steady. The measured species mass rate can be expressed as am Ih(t) = w j m”(x,t) dx , (3.4) 0 where W is the width of the sample. 56 Taking the time derivative of the above equation gives d1:1'l__(_t)= prj)a lt-[m”(x, t)]dx + W m"(xp,t)-d- dip, (3.5) 0 dxp-V where -— - dt p = constant. If a steady-state condition exists in the burning zone, (m”(x,t)=m”(x)), then [dm/dxp] rhll(xp) = W (3.6) For complete combustion, the production rates of C02 and H20 and the depletion rates of 02 are directly related to the total fuel production rate between x = O and x = xp. For the local mass flux to vary as x'05, the total fuel production rate at any instant must vary as xp’z. Incompleteness of combustion will only serve to further reduce this power. However, the data in Figure 3.9 shows that this power is greater than 0.5 and closer to l; for PMMA at Y0... = 1.0 it is even greater than one. In addi- tion, if xp=constant (end of the sample) then dxp/dt=0. Then, any changes in the mass loss rate mm are due only to transient effects, with . 0) $de- = w x1 %-[m”(x,t)] dx . (3.7) Consider now the experimental results of Figure 3.9. These results show that the mass production rate keeps increasing even after the pyrolysis-front has reached the end of the sample (or xp has become constant). This proves that the burning zone is in a transient state during flame spread. This argument was further confirmed by con- ducting flame spread experiments on 10 inch long PMMA samples. Video records of these experiments showed that the flame length kept increasing even after the pyrolysis-front had reached the end of the sample. 57 3.1.3.2. Incompleteness of combustion Referring to Figure 3.9, we note that the production rate of H20 (tilHZO) seems to follow that of C02 (taco) for all the experiments. The mass production rate of C02 is always larger than that of H20. However, this is not true for the comparison between the mass production rate of C02 and the mass depletion rate of 02 (rilez). For Y0, = 0.23, rilco2 > onz, whereas for Y0, = 1.0, riico2 < ’hAOa The crossover between rilco2 and film, can be seen at xp = 0.5 m for Y6” = 0.43. This indicates that the combustion chemistry is changing as X? and Y0... are increased. Carbon in the fuel is converted to unburned soot leading to lower mcoz. Reaction equation 3.1 implies that: [i] < [i] mC02 comp. comb. InC02 incornp. comb. This agrees with the physical observation that both larger flames and flames at higher Yo... are brighter. Figures 3.10 and 3.11 show mega/saw: and taco/tum, respectively, normalized with the corresponding mass based stoichiometric fi’action (i.e. a value of 1 corresponds to complete combustion), where PMMA is C15H40. The data show that meg/mm: is slightly decreasing with x1) and that it drops significantly with Yo... However, meg/map seems to be constant for all Yo... The data in Figure 3.8 shows that the major chemical species mass rates increase with the wind speed, as expected. since the heat transfer to the solid increases with U, as will be explained in detail in Chapter 5. 58 m(:oz/moz O O) l .I vvvv 0.5- va'vvvvvvvvvv v J VVvvvvvvvvvva 0.4“ 0.2-‘ 0.0 . I I I ' I ' I ' I ' 10 20 30 40 50 50 70 Pyrolysis—Front Xp (cm) Figure 3.10 Normalized rhea/mm: with the corresponding mass based 0.9 1.0 0.9 0.43 1.5 0.233 0.9 0.233 0.6 0.233 ¥O$>< stoichiometric fraction (i.e. the value of 1 corresponds to complete combustion). 1.6 1.44 or O n .W' 2 fl (£1 1.2" & .lv“ v‘.. ' p \ 1 0- vIrv ’ “H- k V N , vvvvvv ’v o . v" v v 0 May v v p E 0.8‘ AAAAAAAAAAAA 3: 0.6-i 0.4 I I ' l T I ' I I I I 10 20 30 4° 5° 60 70 Pyrolysis—Front Xp (cm) Figure 3.11 Normalized Margo with the corresponding mass based Ugo You, (m/S) 0.9 1.0 0.9 0.43 1.5 0.233 0.9 0.233 0.6 0.233 X0894 stoichiometric fraction (i.e. the value of 1 corresponds to complete combustion). 59 3.2. 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 [Wichman and Agrawal 1991, Carrier et al. 1990]. As predicted by previous theories[Wichman and Agrawal 1991, Carrier et al. 1990], both VI, and Vf are found to be linear with U, (Figure 3.4). However, these theories also predict that usually Vf is significantly larger than Vp' which is in contrast with the experimental results (Figures 3.4 and 3.5), which show that Vf = Vp' Like- wise, Xf is found to be only slightly larger than xp (between 5 and 10%), regardless of the free -stream velocity or the oxygen mass fraction. This discrepancy arises because the theoretical models utilize the steady-state Emmons’ (1956) solution in the boundary layer burning zone. As discussed earlier (Figure 3.9), the burning zone is unsteady in the solid phase during the flame spread process. This leads to lower instantaneous fuel mass production rates that result in smaller-than-predicted value of xf and Vf. The existence of this unsteady bunting zone is further confirmed by the fact that the species production rates vary roughly as Itp rather than as x35. As is evident fi'om Figure 3.5, the theoretically predicted [Wichman and Agrawal 1991, Carrier et al. 1990] dependence of V1) on Yo... agrees well with that determined experimentally. Increase in Y0, increases Vp 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 proportionally with Yo, in the absence of flame radia- tion 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. As discussed earlier, the theoretical models overestimate the evolved fuel mass flux in the burning zone by assuming steady state Emmons’ solution. This results in an overestimation of the heat blockage factor. Fortunately, it is compensated by overes- timating the flame temperature by neglecting flame radiation and incompleteness of combustion. CHAPTER FOUR Flame Spread over C harring Solids Previous experimental work on charting materials has been done on particle boards in the upward (buoyancy-driven) flame spread mode [Saito et al. 1986, Kul- karni and Fisher 1988, Quintiere et al. 1986]. However, studies in the forced convec- tive mode have been primarily limited to non-chairing solids, such as PMMA [Fernandez-Pello et al. 1981]. As discussed in Chapter 3, even for PMMA, uncertain- ity exists regarding the prediction of flame spread rates under different environmental conditions. The problem is further complicated for charrin g solids due to the forma- tion of an insulating char layer. In addition, since wood requires preheating [Mekki et al. 1990, Saito et al. 1986]. This problem is further complicated becauSe it introduces two new parameters namely the external radiation and the preheat time. Atreya (1983) studied the effect of the preheat time for certain incident heat fluxes on the horizontal flame spread rate on wood. The results of his experimental work indicated that the flame spread rate on wood increases with increasing either or both the external radia- tion and the preheat time. His work also showed the dependency of the flame spread rate on both the internal solid properties and the external environmental conditions. 60 61 This Chapter presents an experimental study of wind-aided flame spread over wood slabs in the ceiling configuration. This study covers the three external conditions that control the flame spread rate: wind speed (U..=0.2 to 1.5 m/s), oxygen concentra- tion (Yo..=o.233 to 1.0), and external radiation (q"=0.5 to 1.3 W/cmz). The objectives of this work are to (i) provide a physical understanding that will serve as the basis for the development or refinement of theoretical models; and (ii) provide measurements for chaning materials like wood. 4.1. Results 4.1.1. Temperature measurements Figure 2.11 shows the schematic of the wind-aided flame spread in the tunnel over a ceiling mounted sample of wood. The flame spread rate depends on how high is the surface temperature at the end of the preheat (i.e. prior to ignition). For a certain prescribed external radiation, both the surface temperature and the char layer thickness increase in time. A high external radiation with a long preheat time will bring the sample surface temperature close to its ignition temperature. This makes the flame spread rate approach, at a critical point (q"(U,,, Yo... tww), that of the premixed-gas flame propagation (flash through). In addition, the preheat time has to be chosen such that the surface would undergo only a minor change due to external radiation. There- fore, after a number of experimental trials, a 600 sec preheat time was chosen for low external radiation (0.5 and 0.9 W/cmz) and 300 secs for high external radiation (1.3 W/cmz). Under these conditions, as shown in Figures 4.1 and 4.2, the surface tem- perature nearly equilibrates with the surface heat loss; and its rate of change with time becomes almost zero -i.e. prior to ignition. Thus, any changes in the surface tempera- ture due to external radiation, during the flame spread process, can be ignored as shown in Figures 4.1 and 4.2. These temperature histories also show that the external radiation is uniform along the wood sample surface as discussed in section 2.1.4. 62 800 700~ T25 92 6004 t 500— 400~ Surface Temperature ( K) .300 r . )""_._‘.’_é_ -; —T w 0 200 I ' I I 400 600 800 1000 Time (sec .) Figure 4.1 Measured surface temperatures and their rate of change with time for wood during flame spread (U..=0.6m/s. Y,.=0.233, Q"=05W/cm’). 600 .1-..“ —-- ___-.. --, _ 500— 4004 EGO-i Temperature (°C) M O O L r v I I 1 ' I I I’ ' I r 100 200 300 400 500 600 700 Time (sec) Figure 4.2 Measured surface temperatures for wood during flame spread (U..=O.9m/s, Y,.=o.233, ('I"=l.3W/cm2). 63 The effect of wind speed on the flame spread rate was investigated for all these sets of external radiation and preheat time. In studing the effect of the oxygen mass fraction on the flame spread rate, however, a low level of external radiation (0.5 W/cmz) and a large preheat time (600 secs) were used. This low external radiation was selected so that a clear flame spread process is observed at high Y0... and to facilitate comparison with PMMA (which burns without external radiation). 4.1.2. Pyrolysis and flame fronts As previously mentioned in Chapter 3, visual observations led to the conclusion that the peak rate of change in the surface temperature after ignition occurs at the instant the flame tip arrives at the thermocouple location (Figure 4.1). Thus, the flame tip location, x f, and the pyrolysis-front location, xp, may be determined as a function of time from the measured temperature profiles as explained in section 3.1.1. The methodology is illustrated on Figure 4.1. Results of such calculations agree exactly with x f determined from video records. The pyrolysis temperature is taken as the piloted ignition temperature (375 °C), at which the solid begins to vigorously pyrolize and contribute fuel to the flame. Results of such calculations agree exactly with xf determined form video records. The three steps of preheating can now be distinguished during the flame spread process. First the preheat by external radiation, then, after ignition, by the post- combustion gases, and then by the diffusion flame extending downwind of the burning zone. This extended flame is caused by the excess pyrolyzate. It is worth noting that the surface temperature at the flame tip arrival ranges from 300 to 340°C, which is higher by almost 120°C than at the conclusion of the preheating by the external radia- tion. This indicates that there is major preheating by the hot gases ahead of the flame. Figures 4.3 to 4.6 show both the pyrolysis-front and the flame-front histories, after ignition, for different conditions. Both xP and xf have an almost exactly linear dependency on time and are very close to each other in all cases. Xp 8c Xf (cm) (14 O L r N O l O l O r r ‘ I ' 600 700 800 900 1000 Time (sec) Figure 4.3 Pyrolysis and flame front histories (Y o..=0.233, q"=O.SW/cm 2). 70 a 0.6 i C] 0.9 50— A 15 204 10 ' I T I ' I ' ”I ' I ' I T I ' I ' 600 610 620 630 640 650 660 670 680 690 Time (see) Figure 4.4 Pyrolysis and flame front histories (Y 0..=0.233, q"=0.9W/cm 2). 65 60 50— A J E \o/ 40- L.— X *5 304 0,, XQ (m/S) 20— 9 ‘ c: A 1.5 CI 0.9 O 0.5 10'I'I'I'I'T'F'I'I'I' 330 332 334 336 338 340 342 344 346 348 350 Time (sec) Figure 4.5 Pyrolysis and flame front histories (Y,..=o.233. q"=l.3W/cm 2). 60 ’0 50% E \0/ 40— >;- . (’8 3Oe Yo O. X v 1.0 204 0 0.61 A 0.43 i' a 0.32 o 0.233 10 "'I"'I"'1"'I+j' 620 660 700 740 780 820 Time (sec) Figure 4.6 Pyrolysis and flame front histories (U..=O.9m/s. Q"=0.5W/Cm 2). 66 The excess pyrolyzate (fuel that is n0t burned in the burning zone 0 < x < X?) causes the flame front to be always ahead of the pyrolysis front. As the ambient oxygen mass fraction increases, oxygen rich case, most of the fuel is consumed in the burning zone causing the excess pyrolyzate to drop. Therefore the distance (Xf-Xp) should decrease with increasing the ambient oxygen mass fraction Y0... The data shown in Figure 4.6 (as well as in Figure 3.4 for the case of PMMA) certainly agree with this hypothesis. 4.1.3. Spread rate Once x f and x p are determined as a function of time, the pyrolysis-front speed V p (defined as dxp/dt) and the flame-front speed V f (defined as dx f/dt) are obtained from the slope of the least square fit line. Data for only the first 0.5 m was considered to minimize errors due to changes in the free stream velocity. Figure 4.7 shows Vp and V f plotted against the ambient oxygen mass fraction Y 0,, where both the free stream velocity and the external radiation were kept constant at 0.9 m/s and 0.5 W/cm 2, respectively. For these environment condition, both V p and V f show a nearly linear dependency on Y 0... As discussed in Chapter 3 for the flame spread study on vaporizing solids, both the decrease in combustion efficiency xA and the increase in the flame radiation cause the flame temperature to be much less than that of the adiabatic case as the ambient oxygen mass fi‘action increases. This is also true for charting solids. Brighter flames (higher radiation) and soot deposition on the walls were observed as Yo“ was increased. In fact, these two parameters are related since the brightness of the flame is due to the hot soot particles. The data shows that Vp ~ YJJ. 67 .AnEo\3n.ourv £256an 5:25 ”8.: 5908 588 out 05 so 33% =8: 25: can $32.8 05 .o 8:85qu he 23E 000% 04 md 0.0 v.0 Nd — _ _ _ _ . _ _ 00.0 Led Imod m; a > D no.0 /\ d (S/w) JA 78> 68 Figure 4.8 shows that both Vp and V f sharply increase with the external radiation for all the free stream velocities. Therefore, if the wind speed and the preheat time are kept constant, increasing further the external radiation would result in a "flash-through" at some critical point (q;(U,., Yo“, tpnhag). This very rapid flame spread would occur when the surface temperature, at the end of the preheating -i.e. prior to ignition, is at or above the ignition temperature of the solid (375°C). Figure 4.9 shows the linear dependency of both Vp and Vf on U“, for all (1". Figures 4.7, 4.8, and 4.9 show how close the flame spread rate Vf is to the pyrolysis-front spread rate Vp for all experiments. It is interesting to see the effect of the wind speed on the flame spread rate as shown in Figure 4.9 where the flame spread rate for U..=0.2 m/s and q"=1.3 W/cm2 is almost equal to that of U..=1.5 m/s and q"=o.5 W/cm 2. However, the effect of Y0, can be seen when noting that Vp for Yo..=1.0 with q"=0.5 W/cm2 (Figure 4.7) is very close to that of Yo..=0.233 and q"=0.9 W/cm2 (Figure 4.8). This shows that the flame radiation for high Y 0.. may be thought of as external radiation. Stable flame spread below 1.3 W/cm2 for 0.2 m/s was not possible. 4.1.4. Species production rates The species production and depletion mass rate histories are presented in Appen- dix C. A typical species production rate history during the three processes namely: preheat, flame spread, and extinction, is shown in Figure 4.10 (U..=0.6 m/s, Yo“: 0.233, q"=1.3 W/cmz). Since in all the experiments performed, the total unburned hydrocarbons (THC) and CO were negligible, only the major chemical species (pro- duction of CO 2 and H 20, and depletion of O 2) mass rates are plotted versus the pyrolysis-front xp and are presented in Figures 4.11 to 4.14 for all the conditions stu- died. The data for xp<12 cm is not included because it is expected to have large errors due to the ignition process. These figures are plotted in log-log scale to show the relationship between the mass production rate and the pyrolysis-front. 69 12 . . . r . . . . 4 , . , . . G—eU=1.5m/s ia—aU=O.9m/s IO—A—AUO _ v 00 A s Q 8~ _ E _ . 8/ a. 6" _ > -i 025 ‘0'- 44 _‘ > -I 2_. —4 04.,...,4..,...,.e 0.4 0.6 0.8 1.0 1.2 1.4 (1"(W/Cm2) Figure 4.8 Dependence of the pyrolysis and flame front speeds on the external radiation (Y...=0.233). 12 .J 104 Q 8-4 E 8, 0. SJ > d «=25 ‘+— 4“ > .1 2— 0 0.0 Figure 4.9 Dependence of the pyrolysis and flame front speeds on the free stream velocity (Y ...=0.233). 70 OI» basin—ta £89....» £8538 Bee 8 82% 063 can??? do 88:9 m 05 mama—e roam: 83.. couoseoa «£0on Aommv 0E; 8... use... 0 food Iwod IN P .6 T0 F .O H cozoczxm emopwchw Second r _ _ u _ e p . b p _ . b . l— . (8/6) 9105 UOllOnpOJd ssaw 71 A 10.0 CD \ 8.0‘ on V 6.0-4 N O X 4.0'i B a? 2.0-3 C .9 U 8 8 1.0“ Q_ 0.8“ 3)) 0.6“ [3‘8 H20 0 0‘0 OZ‘dei. E W CO 0.4 I I T I I .7 I I 10 20 40 60 80 100 Pyrolysis Front Xp (cm) Figure 4.11 Dependence of the species mass production rates on the pyrolysis-front location for different wind speeds (Y ..=0.233. q”=0.5W/cm 2). 20 A m \ 3 N TO-l O .— 81 x .9 6~ 0 CK c 4T .9 23' 3 .0 e 2.. 0. [3'8 H20 0') 8 G—OOz-depl. V—V CO 2 1 l l T 2 l 10 20 4O 60 80 100 Pyrolysis Front Xp (am) “Inn 4.12 Dependence of the species mass production rates on the pyrolysis-front location for different speeds (Y...=0.233, d"=0.9W/cm’). 72 1,7 :50 E If . ‘. 0.9 (\l 20~ 1’ £933 0 o X .33 O m 0.9 C .9 73 3 .0 O L. Q. 5)) 4- 8—8 H20 g 9-9 02-depl. 3 V—V C07 l I I I 10 20 40 50 80 100 Pyrolysis Front Xp (cm) Figure 4.13 Dependence of the species mass production rates on the pyrolysis-front location for different wind speeds CY...=O.233, (1"==l.3W/cm 2). 100 80q 60% 0, 4o« 20— 2" B—E] H20 ' " G‘Q 02-depl. - V47 CO; I I I —[ F 10 20 4O 60 80 100 Moss Production Rote XlO2 (g/s) Pyrolysis Front Xp (cm) Figure 4.14 Dependence of the species mass production rates on the pyrolysis-front for different oxygen mass fraction (U..=O.9m/s. ti”=0.5W/cm 2). 73 The chemical species mass rate are related to the total quantities over the entire bum- ing zone by 19.0) m (t) = I ah (x,t)dx x=0 with the local mass flux being higher near the pyrolysis front due to low char depth. As the fire spreads, the mass rate increases roughly proportional to the pyrolysis- front. When the pyrolysis-front reaches the end of the sample, the mass rate slightly increases before reaching a plateau, then sharply drops (see Appendix C). The duration of this plateau depends on the environment condition, it increases with increasing the external radiation or the oxygen mass fraction. Therefore a thick char layer is expected in these cases. The drop in the mass rate, however, is due to (i) the local mass flux drop as shown in Figure 1.3, and (ii) the propagation of the extinction-front (flame foot) behind the pyrolysis-front as observed during the experiments. However, the flame does often flash back, as observed in the experiments, when some fuel has occumulated enough to support a premixed flame upstream of the extinction-front. The flame stand-off distance from the combustible surface in the ceiling configuration mainly depends on the buoyancy, the wind speed, and the ambient oxy- gen mass fraction. The buoyancy causes the flame (high temperature zone) to rise. The wind speed controls the oxygen mass transfer rate to the flame. And the ambient oxygen mass fraction controls the stoichiometric plane (or the flame) stand-off distance from the surface and the flame temperature. The lower Y 0,, is, the higher the flame stand-off distance and flame lenght are. Therefore, a combination of high wind speed and high oxygen mass fraction would result into a short flame stand-off distance from the surface, which in turn increases the net heat flux at the surface causing higher mass flux. The data shows that the mass production rates of the major chemical species increase with U... Y0... and (1". This is expected since the flame approaches the sur- face with increase in U... the flame temperature increases with Y0... and the pyrolysis- 74 front propagates deeper into the solid with increase in (1". For complete combustion, the production rates of CO 2 and H 20 and the deple— tion rate of O 2 should be related to the total fuel production rate between x=0 and x=xp (the burning zone). For high Y 0... due to soot formation and deposition of heavy hydrocarbons on the tunnel walls, the depletion rate of O 2 is not expected to follow the production rates of CO 2. As discussed in the previous chapter for PMMA, for the local mass flux to vary as x435, the total fuel production rate at any instant must vary as x“. However, the data shows that this power is about 0.4 at Yo..=0.233, and increases with Yo... up to 1 at Yoo.=l. The steady-state solid-phase can not be addressed in the charring solids like wood due to the char build-up which continuously attenuates the production of the pyrolysis products as shown in Figure 1.3. Using the one step chemical reaction described in chapter three, the ratio of the production rate of CO 2 to the depletion rate of O 2 is (equation 3.1) tilC02 44x mm: ’ 8(4x+y—2) Where the chemical composition of the volatiles for poplar wood is shown in Table 4.1 taken from Atreya (1984). The above ratio should approach the value 1.25 for the complete combustion case (Y 0.50333)- Figure 4.15 shows meg/mm? normalized with its mass based stoichiometric fraction, plotted versus xp. Most of the room air oxygen concentration experiments are close to 1 (complete combustion). However, as Y 0.. increases, the normalized ratio {nah/mm2 drops to an almost 0.4 at 100% oxygen environment. This indicates that the chemistry has changed as Yo. is increased. Car- bon in the fuel is converted to unburned soot leading to lower mm: as discussed in section 3.1.3.2. This agrees with the physical observation that both larger and flames at higher Y0. are brighter. Figure 4.16 shows a picture of the flame taken during the flame spread experiment (U..=O.9 m/s, Yo..=0.233, £1"=0.5 W/cmz). 75 1.2 . um You 1.0~ (m/s) g g ”*9 . e 'o 80”” V 0.9 1.0 8 0'8 , u‘W- 0| 12°?" ' V 0.9 0.61 E ,2,” “U..-..‘ A 0.9 0.43 \ O 6“ V V v V O 0.9 0.32 8‘ VV"""" ao.90.233 o v ' v A 0.9 0.233 ' ea 06 0.233 E04“ ”""H"" 00.60.2315 0 0.6 0.233 x 0.2 0.233 0’2 v 1.5 0.233 v 15 0.233 0 0.9 0.233 0-0 l 1 v r r l . , fire—T—fiefi *i 0 to 20 30 40 50 60 7O PyrolySis Front XD (err) Figure 4.15 Normalized moo/mm with the corresponding mass based stoichiometric fraction (i.e., the value of 1 corresponds to complete combustion). Figure 4.16 Picture of a spreading flame on wood sample (U..=O.9 m/s. Yo..=0.233, Q"=0.5W/cm2). q (W/cm’) 0.5 0.5 0.5 0.5 0.9 0.5 1.3 0.9 0.5 1.3 1.3 0.9 1.3 76 Table 4.1 Empirical formula of Poplar wood [Atreya 1984]. Char Empirical Lower Heat of Yield Formula Combustion (KI/gm) Poplar 0.33 (avg) C H O 19.33 1.66 2.43 Char - C H O 26.86 5.44 3.12 Volatile - C H O 15.62 0.99 2.31 Figure 4.17 shows, in chronological order, pictures of the flame as it spreads (U..=O.9 m/s, Yo..=0.61, £1"=0.5 W/cmz). The flame color in these pictures certainly indicates the increase in the flame radiation with Yo... Figure 4.18 shows ducal/map, normal- ized with its mass based stoichiometric fration 2 (according to Table 4.1), for all the cases studied. The data shows that all the experiments fall close the value of l. (4) Figure 4.17 Instantaneous pictures taken of a spreading flame on a wood sample (IJ..aO.9m/s. Yo..=0.61, q*=o.SW/cm’). 78 A8255 339:8 9 388:8 _ no 32> 05 6.5 SE oEoEoEuBfi n. 2.0 No x nd nmd 90.0 md nNd od D 3 Rd so a nd nNd md Q no no no a n; mud ad 0 no as m... > 3 8.0 3 e no and no e 0.0 n¢d ad I no 5.0 no 9 no 3 no > peoba Am}; ..a so» a: 83.. age gaseous 2.. 55 one? 33oz AEoV ax E8... m_m>_o._>n_ on on on om 9. on om 'P _ . _ p . _ . _ . _ § a... 83E S 0.0 flmd i¢.o rod am r 0 Z / w H Z O CHAPTER FIVE Convective and Radiative Heat Transfer to the Solid During wind-aided flame spread over combustible solids, radiative and convective energy transfer occur from the flame and the hot gases to the as-yet-unburned solid surface. The rise of the solid surface temperature from ambient temperature far down- stream of the flame to the ignition temperature at the pyrolysis front is due to (i) the preheating downstream of the flame tip x>xf and (ii) the preheating underneath the flame xpxf, the integral can only be solved if the sur- face temperature is known. Since the resultant integral term in the preheat zone is only a correction to the computed heat flux at the surface, and especially that dT/dz is only significant near the flame tip, the temperature is assumed uniform (F(x)= a,,) in the rest of the analysis. Hence, the heat flux to the surface in the preheat zone (x > xf) takes the form __ 0.332 kg Pr”3 Real/2 I z 3/ -1 q — x \Tf‘Tw) -1+ 1- ; (5.3) 01' Nux 2 34 -1 W = 0.332 -1 + 1 - (x )- I] 3] (5.4) ex 82 where Nux = q” x/(Tf - T...) kg and x*=x/xf is the related distance from the flame tip. 5.2. Results and discussion 5.2.1. Transient heat flux measurements Figure 2.12 shows a schematic of a spreading flame on PMMA sample upsu'eam of the ceramic detectors used in this study. The procedure to compute the heat transfer modes from the transient measurements of the solid and gas phase temperature is as follows. Inverse heat conduction calculations were used to obtain the incident heat flux as a function of time [Beck et al. 1985]. The heat flux algorithm used in the inverse heat conduction analysis is sequential, where few future temperatures, associ- ated with future times, are used to compute the heat flux at every instant. The main aspect of this algorithm is the use of the least squares criterion to estimate the heat flux at the surface qm from the measured temperatures Y1, ...Yj , at times t‘“, rm”, ...,t‘“”H. The parameter r is the number of future time step and J is the number of thermocouples. The criterion is to minimize S with respect to qm K I +' l m+i-l 2 S = §§[ij r- _ Tj ] (5.5) 1: j: where ij+H is the measured temperature at the j‘h sensor and at time t’m'i’l, and iji'1 is the corresponding calculated temperature. The convective heat transfer rate from the surface was computed from the temperature gradient at the solid-gas interface as tic.” = - k — (5.6) where its is the gas thermal conductivity evaluated at the surface temperature. In the transient flame spread experiments on PMMA, four thermocouples were maintained in the gas-phase at some known distances from the surface. This small number of pOints makes it difficult to rely with confidence on the curve fit results from which the 83 gradient is computed. Thus, the temperature gradient for these experiments is approxi- mated as dT/dy = (Tg—Ts)/Ay, where 'I‘g is the gas temperature at distance Ay from the surface. This is a good approximation since the thermocouple used is very close to the surface (Ay<<8,, where St is the thermal boundary layer). First, the thermocouples were constructed in such a way that measurements error due to heat conduction in the thermocouple wires are reduced. Then, one thermocouple was placed about 1 mm from the surface. An error of 0.1 mm on this distance causes a 10% error in evaluating the heat convection. Therefore, the distance y has to be accurately found for every experi- ment. At far distances from the flame, x/xf>>1, the radiation from the flame to the surface located at distance x is negligible, then an energy balance at the surface leads to err" = as". the total and convection heat fluxes using all the data points N for x/xf>2 Hence the distance Ay can be well estimated using least square fit of N 2 kg 2‘; ATi Ay = N ‘= (5.7) 2 ATj (113" j=l This distance is then used to determine the convected heat flux during the whole experiment. The radiative heat flux from the surface was computed from the knowledge of the surface and surrounding temperatures and the surface emissivity. Finally, the flame radiation was computed by applying an energy balance at the solid- gas interface. In order to compare and correlate the results of different conditions, the data has to be plotted vs. x'=x/xf, which is the related distance fiom the flame tip. However, to get the heat flux a function of x' for every experiment, the time dependent tempera- tures and flame tip have to be curve fitted by a very low filter and outputcd with the same time increment. Figure 5.2 shows such curve fitting for the gas-phas tempera- ture. The different modes of heat transfer were then computed and are shown in Fig— ures 5.3 (U..=O.9m/s, Yo..=0.233) and 5.4 (U..=O.9m/s, Yo..=l.0) and Appendix D. 84 ASE?» .Eommnx . agate» .m>=o.cnre 5233 .5 32% 0:3: mats—o 2382—89 829mm» 3.5 256 05 new 3532: 05 no __SEQEOU «d 95$..— Aoomv cc: 5 ON P 00 P om om ca ON 0 _ — p L _ _ . b . — F b O icom ioov fix \. loom x x xxx . .x x x .03 X - x E l. com (Q) dine), Q”(W/CN3) ¢'(W/cw3) 85 Figure 5.3 Heat flux measurements during flame spread on PMMA (U..=O.9m/s,Yo..=O.233). 3.5 1 .4 3K cond K O conv 30 fl x [:1 (Cd ‘* ‘A re—rod 2 .5 48>: X/Xf Heat flux measurements during flame spread on PMMA (U..=O.9m/s, Yo..=1.0). Figure 5.4 86 In all the 21% oxygen experiments, the flame color was blue (see Figures 3.7 and 4.16), which indicates low flame radiation to the surface ahead of the flame. This is confirmed by the results in Figure 5.3. The total heat conducted in the solid equals the heat convected by the hot gases. Both radiation from the flame to the surface ahead of the flame tip and the surface radiation to the surrounding are negligible. However, as the oxygen mass fraction increases, the flame becomes brighter (see Figtnes 3.7 and. 4.17), which indicates a significant radiation to the surrounding. This observation is confirmed by the results shown in Figure 5.4 for the case of 100% oxygen. As the flame tip approaches, the radiation from the flame to the surface becomes more significant. It was also observed that the deposition of soot on the walls increases with increasing oxygen mass fraction. This introduces an error of the gas-phase temperature measurements very close to the flame tip for high oxygen mass fraction, which in turn under-estimates the computed heat convection. Again, this is only true for high oxygen mass fraction and very close to the flame tip. Since the energy lost by the flame to evaporate the fuel from the pyrolyzing sur-_ face [Cpflf-TproJ is not considered in the model (the flame is on the surface), then the following adiabatic flame temperature is used in correlating the results f Yo. (Tr-T...) = —C;— (AH-Q) (5.8) where Cp, the gas specific heat, is computed at the average temperature by the integral 7: j Cpm dT = f Yo. (AH-Q) . (5.9) T. This makes the flame temperature go nearly (because Cp increases with Yo.) linear with Yo... The flame temperature is then over—estimated because the real Tf is reduced due to incompleteness of combustion and flame radiation. As discussed in both Chapter 3 and Chapter 4, it was found that complete combustion holds only for ambient air condition. As the oxygen mass fraction increases, the combustion process becomes less efficient (xA decreases with Ya). In addition, the flame gets brighwr as 87 Yo. increases, (xR increases with Y”). Clearly, these two factors reduce the flame temperature more significantly as You increases. To approximate this incompleteness of combustion and radiation losses depen- dence on Yo”, the convective heat flux measurements for different oxygen mass frac- tion were normalized with (YOJO.233) “. This normalization makes the flame tempera- ture adiabatic at ambient air condition, and less efficient as Yo. increases. The power n was then approximated by using least square analysis on all the results. Expressing the normalized convective heat flux measurements as in equation 5.4, the following equation results N u" 1“ = (2 :6m (5.10) Y us 112 _2; P’ R“ [0233 where C = 0.5, m = -2.3, and n = -0.25. The gas properties used in the above equa- tions are evaluated at the film temperann'e Tm=(TrI-3T.)/4. The experimental results normalized according to the left hand side of the above equation for all the conditions tested are presented in Figure 5.5. In the same figure, the theoretical results of equa- tion 5.4 are also presented. The flame radiation, however, is treated as in the case of an expanding plane in the flame spread direction with known temperature and emissivity and at a constant stand-off distance from the surface (Figtue 5.6). The view factor is then determined utilizing contour integration (Stoke’s theorem) and is 1 Xi" _ a Fat-2 = '1; 31m 1[ ] L\](xf--x)5+dz \Rxf—x): -I-dz a -1 X -1 x_xf + tan -tan V 82+d2 L ["1 a2+d2] [ Va2+d2] + x ‘1 a 5.11 W” [m] ‘ ’ 88 Ln (Nux/PrmRe}’2(Yo../0.233)‘0'25) U X m/s cm Theory 0.9 48 0.9 32 0.9 48 0.9 32 0.9 32 1.5 48 1.5 32 1.2 48 1.2 32 0.9 32 0.6 32 O$DEBI>DOQ<> 120 8 0.44 a—a 120 6 J A--A 120 6 E83 90 1O 0.2e [3E] 90 8 ‘ $69 90 6 0.0 T, _ .- . GD 90 6 0.0 0.5 1.0 1.5 2.0 2.5 3.0 X/Xf Figure 5.10 Measured convective and radiative heat fluxes underneath and downstream of steady porous-metal burner diffusion flames (Y...=0.211). s U Xf A —1.0-: . : . 0 ° .5 U o cm/s in § 3’ ' ° 3 '1 . 2° 3 -1.4~ 'XD 6! . + 150 10 Q8 . o o: v 150 10 v V 150 8 g: ‘1 8: to; v 150 6 .3, j o 120 10 8‘ . v o 120 10 ._. -2.2-‘ ug , O 120 8 Q” j , A 120 6 2 _ A 120 6 v -25- v 33 90 10 5 4 El 90 8 j 0 63 90 6 O -30 ' l j I T 1 ' l ' I 7 i ' i r l ' 2 O 90 6 ~08 —0.6 -0.4 —0.2 0.0 0.2 0.4 0.6 0.8 1.0 Ln(X/Xf) Figure 5.11 Nondimensional measured convective heat flux underneath and downstream of porous-metal burner diffusion flames (Y...=0.211). 94 5.2.3. Surface temperature To obtain the sample’s surface temperature during the flame spread, the one dimensional heat conduction in a semi-infinite solid is examined. The solid is initially at ambient temperature, then experiences a transient heat flux at the surface 95 = Ts-T, = 0 y20 t=0 d." = 51”(9,.T...t) y=0 t>0 Since the heat flux applied at the surface in equation 5.10 can be expressed as (xFVft) b 61” = 10.) -t- 0 < 3— < 1 (5.1221) ‘1 if it” = for) 1< i < i (5.12b) tf it (where tf and tp are the times xf and xp reach x, respectively, and b=2.3), the problem can be solved using the Green’s funcrion or Duhamel’s integral [Atreya 1983]. The surface temperatme is then expressed as 6 = —-§tf)—I(tb) (513) ‘ 1:41:ka ’ ' where t Tb I b = — . .14 0.) loft-‘1‘“ (5 ) Ahead of the flame (0vp (all the heating is accom- plished under the flame). Since Vf is unknown, D is approximated by fitting equation 5.18 to the experimental results and found to be 1.3. Once again, the air properties are evaluated at the film temperature (T rl-3T..)/4. Equation 5.18 reveals the pyrolysis-front spread rate dependency on solid and gas thermal properties and environment conditions. Figures 5.13 and 5.14 show experimen- tal and predicted Vp dependency on U... and Y0.” respectively. The flame spread rate is, therefore, linear with the free stream velocity and is proportional to Y}: for PMMA as given by equation 5.18. This dependency is in agreement with the experi- mental results where vp was found to be linear with U... and is proportional to Ygfi. Loh and Fernandez-Pello (1984) found Y3 dependency on their floor configuration assisted flow flame spread experiments. However, it was concluded that reflection 98 5E—O4 4E—O4~ 3E-O4‘ ' vp (m/s) 25—04a ‘ 1E-O4 T r j I I r I l I I I I I—I I I r r r I r (15 (17 (19 lJ 1.3 L5 Ua. (m/S) Figure 5.13 Measured and predicted pyrolysis-front spread rate dependence on U. (Yo..=0.233). 4E—03 3E—03- 2E-03‘ vp (m/s) 1E-03~ 0E+OO I l I T T I f r I I I l I I r l I (12 (13 (14 (15 (16 (17 (18 (19 ‘L0 1.1 You, Figure 5.14 Measured and predicted pyrolysis-front spread rate dependence on Y... (U..=O.9mIs). 99 from the surroundings, in addition to the buoyant flame radiation, to the sm'face ahead of the flame tip is the reason behind this higher dependency on Yo... Equation 5.18 is a typical wind-aided flame spread formula as for those derived in previous theories [W ichman and Agrawal 1991, Carrier et al. 1990] except for the difference in the constant coefficient. Wichman and Agrawal (1991) who employed the Oseen flow assumption, have a constant coefficient of 1 (see equation 1.3). Carrier et al. (1990), however, who employed the boundary-layer assumption from the onset and later invoked the Oseen flow approximation, have a coefficient of 0.25. The high coefficient in Wichman and A grawal (1991) derivation explains their high predicted ratio VWP. This coefficient is therefore dependent on the pyrolysis zone part of the problem which determines the flame length and consequently the relation between VP and Vf. CHAPTER SIX Conclusions The focus of this study was to understand the wind-aided flame spread on wood and PMMA. This included understanding and determening (i) the pyrolysis-front and flame-front spread rates and (ii) the major chemical species mass production rates. In the process, it was necessary to determine the the convective and radiative heat transfer to the solid both ahead of and underneath the flame. The conclusions of this work are summarized in the following two sections. Sec- tion 6.1 summarizes the conclusions derived from the flame spread experiments on chairing and non-charrin g solids. Section 6.2 summarizes the conclusions derived from the results of the heat transfer study. 6.1 Flame Spread Experiments The wind-aided flame spread experiments were conducted on both chairing (poplar wood) and non-charring (PMMA) solids mounted in the ceiling configuration. The pyrolysis front and the flame front spread rates were determined from the pyrolysis-front and the flame tip histories, respectively. The mass production rates of the major chemical species were derived from the transient species concentration meas- urements during the experiments. 100 101 Several significant conclusions can be derived from the experimental results: 1- The pyrolysis-front and the flame-front are much closer to each other than that predicted by theoretical models. This is also true for the pyrolysis-front and flame-front speeds regardless of the free stream velocity, external radiation (applied only for wood), and/or the oxygen mass fraction for both wood and PMMA. This is because unsteady conditions exist in the burning zone. The pyrolysis-front and the flame-front spread rates vary linearly with the free stream velocity for b0th wood and PMMA. The theoretical models for non- charring materials predict the pyrolysis-front speed quite accurately despite the solid phase assumption in the burning zone. This underscores the fact that the flame spread rate depends primarely on local heating of the solid by the flame tip in the adjacent preheat zone. The pyrolysis-front and the flame-front spread rates sharply increase with the external radiation for wood. The pyrolysis-front and the flame-front spread rates vary with the free stream oxygen mass fraction as Yolg.‘ for wood and Yolg.‘ for PMMA. The species measurements show that the pyrolysis mass flux is roughly constant both for wood and PMMA rather than vary as x'05 as predicted by Emmons theory. The species measurements also show that the chemistry is changing as :1, and Y0. are increased because of incompleteness of combustion. 6.2 Heat Transfer Experiments The heat transfer to the preheat zone ahead of the pyrolysis front in wind-aided flame spread has been addressed experimentally and theoretically. Even though the model is simple, it provided the necessary heat transfer nondimentional parameters to correlate the results b0th ahead of and underneath the flame. 102 Several key results emerge from this study: 1.. The convection is the dominant mode of heat transfer for near room air oxygen concentration condition both ahead and underneath the flame. The flame radiation becomes more significant underneath and near the flame tip as the ambient oxygen mass fraction increases. The steady-state convective heat flux measurements agree with those of the tran- sient case ahead of the flame tip. This shows that the gas—phase reaches steady- state rather fast (compared to the solid-phase) Correlations have been developed for the convective heat transfer, the surface temperature and the flame spread rate. Ahead of the flame, the convection is transient (q” a :23) and goes as {2.8. Underneath the flame, however, the convection is steady and goes as x‘m (boundary-layer dependency). The predicted pyrolysis-front spread rate varies linearly with U... and varies to the 1.5—power of Y0... These predictions are in agreement with the experimental results. Appendix A Appendix A Study of number and location of thermocouples in a ceramic solid for heat flux computations The minimum error can be reached by using many thermocouples indepth and at the surface in order to collect as much information as possible. However, extra infor- mation is sometimes irrelevant and not worth the trouble of placing thermocouples in the cast ceramic solid. Thus, the sufficient number of thermocouples needed for accu- rate heat flux computation was investigated. In addition, for a given number of thermocouples, the location of each thermo- couple plays a role in the computed heat flux accuracy. The closer the thermocouple is to the surface where heat flux is applied, the more sensitive it is to changes in heat flux. Thus, the thermocouple locations were also investigated. After casting a ceramic sample with indepth thermocouples, one needs to know the exact thermocouple locations after the ceramic solid has dried out. The effect of such error on heat flux results was also investigated. Error in solid properties is also reported. 103 104 A.l. Experiment Two thin resistance circular heaters were placed between two identical solid cylinders of ceramic. Thermocouples were placed in the top solid as shown in Figure A1 The heaters were set at 6.45 volts and 1.531 amps for more than 3 hours in order to reach steady state temperatures in the solid. The power was then increased to 12.8 volts and 3.156 amps for 92 seconds and then turned off. This power step input is shown in Figure A.2. During this time, temperature data was collected from all the thermocouples, and is presented in Figure A.3. As the data acquisition unit scanned over those 8 thermo- couples, the timestep varied in time from 0.7 to 0.9 seconds for each thermocouple. Therefore, for simplicity, a low filter was used to approximate the temperature data at equal timestep of 1 second. A comparison of the temperature at x=0 before and after smoothing is presented in Figure A.4. Then the program CONTPC was run to compute the transient heat flux at the sur- face. The number of future times used in all the computations was kept constant at a value of 4. A.2. Results and discussion A.2.l. Usingonly one thermocouple indepth The case of each thermocouple that was studied is presented in Figure A.5. As the indepth distance of the thermocouple increases, the thermocouple becomes less sensitive to heat flux changes at the surface, and so, the error in the heat flux, as seen in Figure A.5, increases. The case of the deepest thermocouple, at x=2.7cm, has oscil- lations with high magnitude. Therefore it is not reported. Figure A.6 shows the residual of those cases presented in Figure A.5. Unlike the heat flux; as the thermocouple location from the surface increases; the error between the calculated and the actual temperature decreases. 105 100 m m _ h. , w m film Experimental setup. Aluminum plate 60 80 40 Time (sec) Power step input history. 20. mm - . - .. 5r...” 32 000 1 I I 1 Figure A.l —20 Figure A2 10 0.84 E one 04— A. o}: ._o o.2~ 0.0 -40 106 80 Y I Y I V I I l r T T U 70-I _l 0 L. :3 4.1 E 0.) 60" _l O- W E —/ 0) " .4 l— . 50" "l l 7 40 I I u I I I f I I r I I I I I 0 20 4O 60 80 100 120 1 4O 1 60 Time (s) Figure A3 Temperature histories at the surface and in-depth I I I I I I T I I I ‘+ .4 72-1 .4 - a ‘4 3 I 8 68-1 .. 3 d l O " . l... 3 64‘ d 1 _ E . . Q) I— 4 .1 . 60-1 ~ ‘ o smoothed ' — real 56 I I I I I I I I r I fl r I W‘ I 0 20 40 60 80 100 120 140 160 Time (s) Figure AA Comparison of measured and smoothed temperature data. 107 .1 _. A d (\l E _ U \ .1 3’» v x .. 2 .1 Li. 4.; 4 0 fi Edi) ‘ ~ 5 v 1.14cm 1 I o 0.53cm 0'0 I 0.28cm d o 0.0cm 4 -— ll TC ‘0-2 I I I I r I I I I I I I a I 0 20 40 60 80 1 00 1 20 1 40 Time (see) Figure A.5 Computed heat flux history using one in-depth thermocouple measurements. 0'6 T l l I l l I d . ‘ -I 0.4-1 0 .1 4 ° 1 A '1 8 l B 3 -1 .‘S.’ In . d 0 (I: 0 .4 .1 .0 . .4 _ e V 1 .1 4cm OJH o 0.53cm I I 0.28am 4 __0 6- o 0.0cm ' ' I I I I I I I I I I I I O 20 40 60 80 1 00 120 1 40 Time (see) Figure A.6 Temperatme residuals when using one in-depth thermocouple measurements for heat flux calculations. 108 A.2.2. Usingtwo thermocouples indepth Several cases were studied and are presented in Figure A.7. As the distance to the first thermocouple from the surface increases, the error in the heat flux increases. The residual (Y -T) of the cases in Figure A.7 are shown in Figures A81 and A.8.2 for the first and second thermocouple, respectively. Figure A.8.l shows that a maximum error of 0.6" C exists for the first thermocouple. However Figure A.8.2 shows a maximum error of 2 ° C for the second one because the minimization of the root mean square is heavily weighted on the first thermocouple. A.2.3. Usingone thermocouple at the surface and others indepth The thermocouple at the surface is more sensitive to changes in the heat flux than the indepth ones. Thus, its effect dominates over the rest of thermocouples. Several cases were studied on using one or two thermocouples indepth. The cases of using one thermocouple indepth are shown in Figure A9. All the results are very close, and a maximum error of 2% in the heat flux is seen for the case of x=l. 14cm. The domination of the thermocouple at the surface, mentioned earlier, can be observed in the case of having a thermocouple at x=1.l4cm. The addition of the thermocouple at the surface brought the results from 10% oscillations to stable results with low percent error. The results of having two thermocouples indepth, in addition to the one at the Surface, are very close, as shown in Figure A.10. The cases of using 1, 2, and 3 thermocouples indepth with the one at the surface are shown in Figure A.11. A maximum error of about 0.02W/cm2 (2%) is observed. For comparison, the case of using all 7 thermocouples to compute the heat flux is Pusented in Figures A.9, A.10, and A.11. 109 1.0 l I l l T I . ...... I l ' u A t. (El 0.8“ _ , '1 ”J \o ‘. .- ~;~ x " ‘ 3x . x'.‘ 3' x 0.6-I ‘: " “i 3 ‘u '0. [1: xII ‘- I l 8- .‘e' “ 4-J I - 8 : .1 0". .53,2.7cm I 0.4-4 3 . ‘3.“ 8.1.14cm - -’ 7 .53cm 1 ll 2.7cm 4 0 .53cm --- ll 0'2 I I ' l ' l f l ' a [c 0 20 4O 60 80 100 120 140 Time (sec) Figure A.7 Computed heat flux history using 2 ill-depth thermocouple measurements. 0.6 r f I I I TI T '0 4 O I _‘ 0.4-1. .. O "' .3“ A J -- _s-II, - a ’- 3» 0.2-1 I A ‘ '6 3’” l 1?, .0 ea —4 a) QC . . . I, an .53.2.7cm -O.2- r: v .28.1.14cm ‘1 o .28..53cm .1 I .12.2.7em —0.4 f . r t f r fl 0 .12.;53ctn Time (see) Figure A.8.l Temperature residuals of the 1“ thermocouple when using 2 in-depth thermocouple measurements for heat flux calculations. 110 B l B 3 '-4 39 (I) a) 4 o: -2-1 Vf .53.2.7cm -1 .28.1.14cm . III .28,.53cm o .12.2.7cm “1 It .12,.53em \J fl I T I 1 r Tr r I l T T I 0 20 40 60 80 100 120 140 Time (see) Figure A.8.2 Temperature residuals of the 2'“l thermocouple when using 2 in—depth thermocouple measurements for heat flux calculations. 0.8—l 0.6-1 0.4-4 Heot Flux (W/cmZ) 0.2 I I I I r 0 20 4O 60 80 1 OO 1 20 1 40 Time ( sec) Figure AS Computed heat flux history when using the thermocouple at x=0 and another in-depth. 111 1.0 . , . , l I I I ‘ 1 A 0.84 _4 (\l E 1 cl 0 ; 0.6-i .. '1 .4 ’3‘ LT. 0.4% _. +J o . . a) I 0.24 _ l: 0.12cm. 1.1 cm 0 0.12cm. 0.2 cm 0.0 . , . , . , . , . f“. °" to . O 20 40 60 80 100 120 140 Time (sec) Figure A.10 Computed heat flux history when using the thermocouple at x=0 and another 2 in-depth. 1.0 I I I I 1 . 1 A i N _ E a o \ a 5 . x —-I 3 01 LI. ‘6 . al I III .12..53.2.7c I .12.1.14cm - o .28em 0.2 . - all TO I ’ I I 1 r l T 1 0 10 20 30 4O 50 Time (see) Figure A.11 Effect of the number of thermocouples used on the computed heat flux. 112 A.3. Thermocouple locations As mentioned in the previous section, the closer the thermocouples to the Stu-face are, the better the results are. This is shown in Figures A.7 and A.8 for heat flux and error, respectively. A.4. Errors in the thermocouple location For a given temperature history, an error in the thermocouple location would lead to an error in the heat flux results. This is shown in Figures A.12 and A.13 for the cases of two and one thermocouple indepth, respectively, with a location error of 2mm. Comparing these results with those of Figure A.7, it is seen that as the thermo— couples are closer to the surface, the results show more sensitivity to location errors. A.5 Errors in thermal properties All the above results were computed by using published ceramic thermal proper- ties. Just recently, the real properties of each ceramic block were computed by using the parameter estimation code developed by Dr. J. Beck. For comparison, the heat flux results of both cases are presented in Figure A.14 where all the thermocouples were used to compute the heat flux at the surface. The error is more significant at the beginning and the end of the step input. A.6. Conclusions - Besides the thermocouple at the Other boundary, it has been shoWn that the addi- tion of another thermocouple gives approximate results if it is very close to the surface where the heat flux is applied. - Two thermocouples give even better results. For higher accuracy, one of the two thermocouples must be at the surface. - The location of the 2 "d thermocouple can be anywhere indepth as long as the l“ is very close to the surface. 113 r I l l 7 j if ' l 1 .6‘1 —1 A . I N - E 1.2. i \ 1 -l E . . :5 0.8-4 3" 7 , 1 4 LI— 4 . ..O I I. .‘ +1 4 . I. . ' . -4 O .4 ' ' . g 0 44 I: I O. _ . -< t. .' 0 hr . : ~m\52.1.05cm .73.2.7 0.0 I’ I V l I I I l I l o 1' r cm 0 20 4O 60 80 100 120 140 Time (see) Figure A.12 Effect of 2mm location error of the in-depth thermocouple on the heat flux results (using 2 thermocouples). 1 .2 1 l I r f T . J R? E - o \ 3 _ v x 3 -1 LL J .‘J o a a) I o 1.1.4cm a 0.53am - . o 0.28cm 0 ,, ' —- 0!! TC _ 0‘. 1 r I r I ‘r T f 1* f I 0 2O 4O 60 80 100 120 1 40 Time (sec) Figure A.13 Effect of 2mm location error of the in-depth thermocouple on the heat transfer results (using one thermocouple). 114 1' 0 I I T l l . l T l f p 0.8" . —J ’8? E « : ' \U f .v . g 0.6'* c ' m V I 0' x 1 $ ' '1 :3 , 0' E: 0.4“J '0 ' '4 ,4 . \ O 4 ' o . 0 ' . I 0 . 0.2 : o ... v ubliehed pro 4 o e ' oted pr“ — step input 0 0 T l T t T I T r r r r T -20 0 20 4O 60 80 100 1 20 1 40 Time (sec) Figure A.14 Effect of properties error on the computed heat flux (using all the thermocouple measurements). An error in the thermocouple location has more effect on those that are close to the surface, since they are have higher sensitivity coefficients. A maximum error of 2 0 C in the temperature was observed for the deepest ther- mocouple, in case of using 2 thermocouples indepth. Errors in' the solid properties have high effects on the results when sharp changes happen in the heat flux applied at the surface. In addition, the computed thermal pmperties of each ceramic solid give better results than the published ones. Appendix B Results offiame-spread experiments on PMMA (106 115 (105- (104- Moss Production Rate (9/5) 5E—O3 ~4E-03 ~3E-03 ~2E-03 -1E-03 THC 8c CO Moss Production Rate (9/3) 0.00 “ , . , 0E+00 0 1000 2000 3000 4000 Time (see) Figure B.Tl Species mass production rate: history during flame spread on PMMA (U..=0.6m/s, Y...=0.233). 1E-O1 , fa? .. \\ 55-02~ O) '1 v e *5 J (I C :2 1E-021 0 i 3 . 8 55-034 L. a 0. A (D A (I) C 4 A H20 2 Cl 02-depl. C 1 E-03 I I I r 6 I O? I I 10 20 4O 60 80 100 Pyrolysis—Front Xp (cm) Figure B.Xpl Species mass production rates plotted against the pyrolysis-front location for PMMA (U..=0.6m/s, Y,.=0.233). 116 0.10 4E—O3 Q o 002 3 E 3 323““ 2 o» 008-1 ,6 “3C é: v + C0 , ' + _ “ ”3E-03 B . . . g o , . _ 2;; 05 0.06- . .- + g C . 4' '0 g * . - -2E-03 g ‘5 . $3 0. D 0.04- «5’ . . . 8 9 g" 0 OJ; _ - i it —1£-—03 5 0.02- . * ' O (8 . f‘»; 0 E _ . + «Art 08 0.00 . e , f , . j . 05+00 (I) 0 500 1 000 1 500 2000 2500 1— Time (see) Figure B.T2 Species mass production rate history during flame spread on PMMA (U..=O.9m/s, Yo..=0.233). ,0? 10.0 4 1' I l 1T m r t t _ \ 8.0+ _ 3 .1 .1 6.0d _ (\l o . ‘ ‘ § 4.0‘ -i m 4—3 " O " o 0: C 2.0 '3 _ .9 1 ° . 4" a g 0“.“ ‘O ‘A‘ 8 1.01 a“ 0- .: ‘ ‘ l -l g 0.8 d ‘ ‘ A A H20 0 0.6‘ D 02-depl. E 9 C09 | l l ' I I I l f 10 20 30 40 50 60 7O 80 90100 Pyrolysis—Front Xp (cm) Figure 1!.sz Species mass production rates history during flame spread on PMMA (U..=O.9m/s, Y...=0.233). 117 0.10 45-03 { o 002 ~ 3 ’(‘n‘ J D 02-depl. _ a) \ V H20 , f *6 c» 0.081 v 1': 23° x115! -3I-:—0:s 0: a) 1 . ,3! - c *5 ‘ f .9 0‘ 0.064 ~ ' g S " " 1’ 2%; 4 .. ,' V g _2E*03 08- '8 0.04- ‘, ‘1 I.“- J a) O ‘ ‘ (D L’ -I “ M. o t x PIE-03 2 ~ 0 8 0.02 )g? . 0 2 ~ MAéflfi % 0.00 ”‘W rs .fi , . , . . . 05+00 0 0 " m " 800 1200 1600 2000 5 Time (see) Figure 8.13 Species mass production rates plotted against the pyrolysis-front location for PMMA (U..=l.5m/s, Y...=0.233). ft]? 1E-011 \ _ .. 01 8E 021 v -I a) . "5 [r 4E-02- c i .9 U :3 .0 O L ”- 1E-021 a 815-03- : A H20 2 I U 02-depl. 4E-03 I ' I I T QICOI? I 1 10 20 3O 40 50 60 70 80 90100 Pyrolysis-Front Xp (cm) Figure B.Xp3 Species mass production rates history during flame spread on PMMA (U..=1.5m/s, Y...=0.233). 118 Mass Production Rate (9/3) 0.4 8E-03 { v 002 A? - 3 J O 02-depl. o (D 1:] H20 3 b *5 0 THC a: 0'3“ x 00 : ”55‘03 c .' . .7 _ g I 'o 0.24 . —4E-03 9 a? - 0. 1’ °’ 0‘) , - - o 0.1- l" . ~2E—03 2 . - o ,wew ‘ o 0.0 " ' . T . I . , 1 1 . . . 0E+00 I 0 1 00 200 300 400 500 600 700 800 1— Time (see) Figure B.T4 Species mass production rates history during flame spread on PMMA (U..=O.9m/s. YsO.43). 100.0 1 I ' I ' l ' l 80.01 60.0: 40.0- lllllL 1 20.04 10.01 8.0-: 6.01 4.0- 4 llllllL l 2.0- 11111 Moss Production Rote X102 (g/s) PYronsis—Front Xp (cm) Figure B.Xp4 Species mass production rates plotted against the pyrolysis-front location for PMMA (U..=O.9m/s, Y...=0.43). 119 1.0 25—03 Q 0 C02 Q l- 3 ”U? A 02-depl. Q) \ U H20 it _ 45 3 01H x THC z; 0: _ O c 0.6 Q o . 0 8 .3 Q 0 —BE-04 9 .8 0.4— It 9 O 3} 0: .I X x o O P O (I) X K O o 0 * F4E—04 E 0 020 xxx. Q Do .55” . 8 ° Xix 683°° W i 2 x g 6 6 9 R 325 x* * 0° 5' ga-Uig** T 0.04 , . . . , 0 , . 0E+00 (i) 0 20 40 60 80 100 1 20 1 40 160 I— Time (see) Figure 3.13 Species mass production rates history during flame spread on PMMA (U..=O.9m/s, Y...=l.0). ’0‘ 100.0. ' ' w ' ‘— E 0.0: no ,. 5 v : 000° 0.. " 0° .O' ‘ 65 40.0- 000° .... -‘ ;2 ' 00000...." at 1 Duo .0. A““ d 0.) a o. a“ 4... O ”a: '... A‘ ‘ m 10.0“ 0. A‘ _I 80— ‘A . C . J ‘0‘. _ a '3 4 04 e“. ‘ O ' A‘ — 3 . ‘A‘ . "U a O L D. W 1.01 A H20 - ‘8 0.81 D 02-depl. j :2 63 C09 T l I t I l I I 10 40 80 100 Pyrolysis—Front Xp (cm) Figure B.Xp5 Species mass production rates plotted against the pyrolysis-front location for PMMA (U..=O.9m/s, Yo..=l.0). Appendix C Results of flame-spread experiments on wood Moss Production Rote X102 (g/s) Moss Production Rate (9/8) 20 1 120 0.12 v 002 O 02-depl. 0.10- U H20 0 THC 1 % CO 0.084 0.06“ 0.04- «I 0.02‘ 0.00 r 0E+00 ' I I T I ' 100 200 300 400 500 600 700 Time (sec) Figure C.Tl Species mass production rates history during flame spread on wood (U..=O.2mls, Y...=0.233, t'f'=l.3W/cm’). v (:02 [1 H20 0 02- depl. 10 Figure C.Xpl l I T l 20 40 60 80 100 Pyrolysis Front Xp (cm) Species mass production rates plotted against the pyrolysis-front location for wood (U..=O.2m/s, Y...=0.233, é”=l.3W/cm’). THC 8: CO Moss production Rate (9/8) 121 0.05 SSE-03 V 002 "m‘ 0 02-depl. \\ C1 H20 3 0.04fi o THC / '6 /¥ "2E-03 01 0.03- / c -' .9 . /: .5 If I -\.’ g 0.020 0 3.1-Lg. ‘ _. 8 t :::- ' ~1E—03 0. 1 0 (0 ° “ .0 m 0.01—1 .- I ”I“, ”A: 0 % .‘A-fi'kl'.‘ :2 o f? 1f .3: (100 . ' .J- . 1 . I .r 0E+OO 0 200 ‘00 600 800 1000 1200 Time (see) Figure C.T2 Species mass production rates history during flame spread on wood (U..=0.6m/s, Y...=o.233. q"=o.swlcm2). 10.0 8.0n 6.04 4.0— 002 02-depl. 1304 2.0“ 1.0“ 0.8-4 0.6-1 O.4-+ 0.2n Moss Production Rote X102 (g/s) 0.1 T I i l 10 20 40 50 80 100 Pyrolysis Front Xp (cm) Figure C.Xp2 Species mass production rates plotted against the pyrolysis-front location for wood (U..=O.6m/s, Y...=o.233, o"=o.SWIcm2). THC & CO Moss Production Rate (9/5) 122 0.12 1 I . I . I ' I V 002 A O 02-depl. U) _ ... \ 0.100 0 H20 g 15 02 3 1 an THC , Q) X CO ' .H 2 00: 0.08‘1 v C i 9 ~8E-03 .9 0.06— a ‘g , 0 i . U 9 0.04- 9!: L 0. . $0 —4I—:—03 a ’ - O 0.02- 9 _ 2 A 0.00 . , . 0E+00 0 200 400 600 800 1000 Time (see) Figure C.T3 Species mass production rates history during flame spread on wood (U..=0.6m/s, Y°.=0.233. (1"=0.9W/cm2). U) V 002 F, O 02-depl. V C] H20 N 104 9.. g; X 7.. .‘L’ 6~ (is) 5d #0050 c 4—0 000099.09 0 0' 'V ‘8 v v 8 2‘ V UDDDUDDDDCIP 0' 0000 U) (D U 2 1 1 I i r 10 20 40 50 80 100 Pyrolysis Front Xp (cm) Figure C.Xp3 Species mass production rates plotted against the pyrolysis-front location for wood (U..=0.6m/s, Y,.=0.233, ti"=0.9W/cm2). THC & CO Mass Production Rate (9/8) 123 ”a? 0.20 - _ v €02 * 2E 02 E» ru? 0 02-depl. g." ‘ V \ 016- D H20 1: - 3 3 . q o THC 9 ‘ ‘ZE—OZ CE 3 as CO , w ‘ C O - o - 9 a: 012. v ~1E—02 T5 5 ‘ - 3 '8 . 3 9 ' Q § 0.08:- ~8E-03 m u ’ U) 0L- : ' °°°° ‘51.: g 3 0.04- “‘3 -4E-O3 o o - é - 0 2 4 05 Q; 0.00” . , . , i , . . -OE+00 O o 100 200 300 400 500 500 760 800 E Time (see) Figure C.T4 Species mass pmduction rates history during flame spread on wood (U..=O.6m/s, Y...=o.233, q"=1.3wlcm2). A 20 Q v (:02 on O 02-depl. V v V ' o D H20 V 8 ' o o N v o o O 3 ° ; 10~ 3 3 5 .33 8o 0 9 V (r a D D U D c 5‘i U a a O D a '5 o D U C] 5 4~ 0 L. Q (I) (n O 2 2 I i T I 10 20 40 60 80 100 Pyrolysis Front Xp (cm) Figure C.Xp4 Species mass production rates plotted against the pyrolysis-front location for wood (U..=0.6m/s, Yo..=0.233, ti"=l.3W/cm2). 124 0.12 7E-03 J V C02 ’0? O (Dz-depl. E 0.10% D ”20 M _ v T : THC ' om) _5E_03 0’ l- *5 0.08- 0‘ l c .g 0.06— —3E—03 U i 3 . “o - 8 0.04~ 0' ~1E-03 ‘3 . o 0.02“ 2 “M * W , 0.00 . I I r —1E—03 600 700 800 900 1000 Time (sec) Figure C.T5 Species mass production rates history during flame spread on wood (U..=O.9m/s, Y...=0.233, Q"=0.SW/cm’). 10 ’8‘ 9“ V C02 \ 84 O 02—depl. 3 7~ [3 H20 " N 6_‘ '7'" V 2 SJ 'vV'v oocpo x "' 00° 4 _‘ ' V V ' 00 .2 v ' 000 O O o O 000 0: 3g 0 o c .9 U "" D g 2 4 000 U o .30 L 0. 0° C! (CID, o O 0 o E 1 9 I I I j I I I I 10 20 30 4O 50 6O 7O 80 90100 Pyrolysis Front Xp (cm) Figure C.Xp5 Species mass production rates plotted against the pyrolysisofront location for wood (U..=O.9m/s, Y...=O.233, ti"=0.5W/cm’). THC 8c CO Mass Production Rate (9/3) 125 1E-02 l v—v coz . 1; 0154 9—0 Cz-depl. f 4 B-El H O ' 3‘ . 04> mic ,'. -8F_—03 m . H co l *5 0.12— l' 01 ' ' -6E-O3 c d .9 a 1;) 0.08~ L ‘O -l 4E-03 o . L Q .I O 2 0.00 v . l . t OE+00 O 200 400 6 U 0 800 1000 Time (sec) Figure C.T6 Species mass production rates history during flame spread on wood (U..=O.9m/s, Y...=O.233, £1"=O.9W/cm2). A 20 U) V C02 E O 02—depl. v V D H20 ""' N ' ' 0000 O ' ' ' 0°C «— _4 v o x 18d ' V V o O o O .4 O O: 7 v V o o 00 6“ o o 000 c: ‘,o .9 5 —I a D D 6 c. 0 ° 3 o '0 4d 0 a 8 o D O. 3.- 0 a (I) (D o 2— 2 F l l i 10 20 4O 60 80 100 Pyrolysis Front Xp (cm) Figure C.Xp6 Species mass production rates plotted against the pyrolysis-front location for wood (U..=O.9m/s, Y...=0233,q"=o.9wIcm1). THC & CO Moss Production Rotes (g/s) 126 0.3 1E—02 v 002 O 02-depl. - ~ U “20 “i. ~1E—02 O THC at II 3K CO fi'v " h 0.24 waxy -8E—03 0" P 0v 0' Moss Production Rate (9/5) ~6E—03 ~4E—03 ”b b : _ ° -2E—03 t " ' . . I . r 0E+OO 0 1 00 200 300 400 500 600 Time (sec) Figure C.T7 Species mass production rates history during flame spread on wood (U..=O.9m/s, Y...=0.233, q"=1.3wlcm1). A 30 fl *7 002 8 O 02—depl. 8 H a O a '- a x a a) a *6 8 c1 0 O: 1 O J 5 o D C 9 -4 U U a . D 2 a U ‘1 c1 8 6 —— 0 O 0 g .- O. ‘J W 0 g} 4 1 o E 3 T I I I 10 20 4O 60 80 100 Pyrolysis Front Xp (cm) Figure C.Xp7 Species mass production rates plotted against the pyrolysis-front location for wood (U..=O.9m/s, Yo..=0.233, d"=l.3W/cm2). THC 8c CO Moss Production Rate (9/3) 127 0.12 ' 8E—03 J v—v coz . ’3 r E 0.10a EHEJ H20 , l o—o mc .93 _ O 0.08“ 0: C . .g 0.064 —4E-O3 U . . 3 U E 0.04~ . « ~2E-03 a . O 0.02- 2 d 0.00 ,. ~— , . . . 05+00 0 200 400 600 800 1 000 1 200 THC & CO Mass Production Rate (9/8) Time (sec) Figure C.T8 Species mass production rates history during flame spread on wood (U..=l.5m/s, Y...=0.233, Q”=;O.SW/cm’). 10.0 8.0-l 6.0-1 ' v v 0 4.0“ o a 2.0% D 0 1.0“ 0.8“ O.6~ '3 H20 0 02—depl. V CO 0'4 I ' i I I 1 l7 I l 10 20 30 40 50 60 70 80 90100 Moss Production Rote XlO2 (g/s) Pyrolysis Front Xp (cm) Figure C.Xp8 Species mass production rates plotted against the pyrolysis-front location for wood (U..=l.5m/s, Y...=o.233,d"=o.sw1cm2). Moss Production Rate (9/3) Moss Production Rote X102 (9/3) 128 «W002 0,16d 9—0 02-depl. 'lG-EIH20 «HTHC qX—XCO 0.12- I“ " ‘rr ‘ 8E—03 ~6E-O3 \”1—4E-03 0E+OO TOO THC 8c CO Moss Production Rate (9/8) 0.00 .,.1. 100 200 300 400 500 600 700 800 900 ' Time (sec) Figure C.'l‘9 Species mass production rates history during flame spread on wood (U..=l.5m/s, Y...=o.233, q"=0.9wlcnt1). 20 V C02 0 02—depl. [:1 H20 v ' 0 10 -< o 9 a , ; ° 8 ._. ' o 74 ° .3 6'1 D D _ a D 0 b " a 0 o 4_ D D :5 .4 2 T l I l 10 20 4O 60 80 Pyrolysis Front Xp (cm) Figure C.Xp9 Species mass production rates plotted against the pyrolysis-front location for wood (U..=l.5m/s, Y...=0233. d"=0.9W/cm2). 129 0.3 0.2-l 0.1q Moss Production Rate (9/3) 0.0 O 1E—02 V—V C02 9-0 02-depl. ' - B—El H20 . 0—0 THC 3 ‘ ; “BE-03 H CO ' .. _ FEE-03 _4E-03 “25—03 "' OE+00 T I I l 1 I —U I I 1 00 200 300 400 500 600 700 800 Time (sec) Figure C.T10 Species mass production rates history din-ing flame spread 50 40 30 20 Moss Production Rote X102 (9/5) on wood (U..=1.5m/s, Y...=o.233. ('1"31.3WIcm’). V C02 0 02—depl. “ D H20 ‘ . l I T I 10 20 40 60 80 100 Pyrolysis Front Xp (cm) Figure C.Xp10 Species mass production rates plotted against the pyrolysis—front location for wood (U..=l.5m/s, Y...=0.233, d"=l.3W/em2). THC 8: CO Mass Production Rate (9/5) 130 53 t3 1 x0004 Moss Production Rate (9/3) . I. . O ' ‘- O n" . , K . I Q: " s. wt 4 .0 ‘ I"' ~ . x.‘ . e. i o- . a . 45-03 002 02-depl. H20 THC CO r3E—03 rza-os ~1E—O3 ' r r l l m I t OE+OO 600 650 700 750 800 850 Time (sec) Figure C.Tll Species mass production rates history during flame spread (mvwndGLsOSmwtYmsailQEQJWMmfiL 20 r“? V 002 E O 02-depl. V D H20 N ‘l — (D O .— 8‘ X 38 Q) 6‘ v'ggaz *5 . v 0 0: . 3 ° ° c 4“ ' 3° ' o .9 ° 00‘3 +J DD 0 0° 3 0° '0 0° 9. 2 - 0 ° 0. D 0 D U m m 0 2 1 I I I I 10 20 4O 60 80 100 Pyrolysis Front Xp (cm) Figure C.Xpll Species mass production rates plotted against the pyrolysis-front lumumlRuummdaLFQSmfltYgs0334r=QSVMmm§. THC & CO Moss Production Rate (9/3) 131 0.3 35-03 ‘V 002 ”a O 02-depl. \\ < [3 H20 . 3 o THC 0) 316 CO '6 Q2— -2E-03 0: C .9 . — 4.) U :3 .0 ..:'; 9 0.1- . ~IE-03 Q. 9 .:§"' ' a) tn ‘J‘ _ o .11" 0.0 i I I I 1 I I I I I T OE‘I‘OO 600 620 640 660 680 700 720 740 Time (s) Figure C.T12 Species mass Induction rates history during flame spread on wood (U..=O.9m/s. Y...=0.43, ti"=0.5W/cm2). 100 80 v 002 60 — O 02-depl. Cl HO 40n 2 20‘ Moss Production Rote XiO2 (g/s) 5.5 I I I 10 20 40 60 80 100 Pyrolysis Front Xp (cm) Figure C.Xp12 Species mass production rates plotted against the pyrolysis-front location for wood (U..=O.9m/s. Yo..=0.43, t'I”=0.5W/cm2). THC & CO Mass Production Rate (9/5) 132 Moss Production Rate (9/8) . I .7: . I l I v 'I I r I I OE+OO 600 620 640 660 680 700 720 Time (sec) Figure C.T13 Species mass production rates history dm'ing flame spread on wood (U..=O.9m/s, Y...=0.6l, ti"=0.5W/cm2). 100 V C02 0 02—depL CI H20 45 07 CI: 0 O C l 1 | N O l o o o e a c Moss Production Rote XiO2 (g/s) ES 1 I i i I F 10 20 4O 60 80 100 . Pyrolysis Front Xp (cm) Figure C.Xp 13 Species mass production rates plotted against the pyrolysis-front location for wood (U..=O.9m/s, Yo..=0.6l,t'1"=0.5W/cm2). THC & CO Mass Production Rate (9/3) Moss Production Rate (9/5) 133 1.0 3E-03 V C02 ‘ O 02-depl. CI H20 I. 0-8‘ o THC OOOWO X CO ° Q3000 ‘ o q me W 000 ~2E-03 0°6" o W o W " 0 WV * o a" If ' v m .— 0.4H 0° ”if. " vvvvvvvvvvvvvvvvv ' 0° ' PIE—03 q o ' ‘ I v 0.2-i ............. . THC 8c CO Mass Production Rate (9/3) 0.0 ' I r . OE-I-OO 600 640 660 680 Time (sec) Figure C.Tl4 Species mass Iroduction rates history during flame spread on wood (U..=O.9m/s, Y...=l.o, d"=0.5W/cm2). A 100 _ < 80% v 002 o0o O, 50— o 02-depl. o o 0 °° V 4 U H20 0 0 v N 40-1 o O ' ' v Q o O o , v ' ' 1'- 0 ' x V 8 20“ v ' ' ' 0000.300 0 v ' 0 o a O: 101 o 0 0 s 81 . ° 1.: 6— (3) i ‘o 4‘ O L O. U) 2 2 U) o 2 1 T 1 l I I I I I 10 20 4O 60 80 100 Pyrolysis Front Xp (cm) Figure C.Xpl4 Species mass production rates plotted against the pyrolysis-front location for wood (U..=O.9mls, Yo..=l.0, d"=0.5W/cm2). Appendix D Results of heat transfer experiments q" (W600?) 134 0.5 X cond. . O conv. U re-roo. O 4~ai A. rod. Q A ‘ E N g 034 a U \\. . E t 0.2- U' 0.i~ “In Dig." t ’ 'I.. 'II II'u-uugulioll III II 0() “BEH!HWQ..~nfilUflfiSKIDOOKIIOIOENMMUDDG‘K‘Iflint. . . T’ ' l ' FT ' l 2 3 4 X/xf Figure D.l Heat flux measurements during flame spread on PMMA (U..=0.6m/s, Y,.=0.233, x=33cm). cond. conv. re-rod. rod. DDOBK 0:01 "Beleeoeaeuaag a a . . . i 2 3 4 .5 X / Xf Figure 0.2 Heat flux measurements during flame spread on PMMA (U..=O.9m/s. Yo..=0.233, x=33cm). 0" (Wm?) <1" (W/cm2) 0.0 135 Figure D3 a 2 3 4 X / Xf Heat flux measurements during flame spread on PMMA (U..=1.2m/s, Yo..=0.233, x=33). Heat flux measurements during flame spread on PMMA (U..=l.2m/s, Y,.=0.233, x=40cm). 136 q" (W/cm") Figure D.5 Heat flux measurements during flame spread on PWA ('U.=1.Sm/s, Y...=0.233, x=33cm). q" (W/cm’) X/Xr Heat flux measurements during flame spread on PMMA (U..=O.9m/s. Yo..=0.33, x=33cm). Figure 0.6 137 ¢'(W/cn3) X/Xr Figure 0.7 Heat flux measurements during flame spread on PMMA (U..=O.9m/s, Y°.=O.43, x=33cm). q”(W/CW3> to Its 20 :25 30 :15 4c» 4&5 50 X/Xf Figure D.8 Heat flux measurements during flame spread on PMMA (U..=O.9m/s, Yo..=0.61, 33cm). ("it‘ll/ C m 2y 138 5.5“” ’ " T 3K cond 1 C O conv .).0 n 0 (Cd n A re-rod 2 54%;; 2.04 E 1.5— ' J 1.0 II'l *n 0.5—0CEJJ "u 40 ".""'..IIIIIIIIIIIII 0.0 1 2 3 4 5 X/Xr Figure 0.9 Heat flux measurements during flame spread on PMMA (U..=O.9m/s, Y...=1.0, x=33cm). List of References LIST OF REFERENCES Abu-Zaid, M.Z., PhD Thesis, Michigan State University, East Lansing (1988). Ahmad, T., and Faeth, G. M., Journal of Heat Transfer, 100, p 112 (1978). Ahmad, T., and Faeth, G.M., Seventeenth Symp. anti.) Comb. (1979). Annamalai, K., and Sibulkin, M., Comb. Sci. Techn.. 9, p.185 (1979). Atreya, A., mo. Thesis, Harvard University, Cambridge, MA. (1983). Also published as a National Bureau of Standards report, NBS—GCR-83-449, March 1984. Beck, J.V., and Arnold, K.J., Parameter Estimation in Engineering and Science, John Wiley and Sons (1977). Beck, J.V., Blackwell, B., St. Clair, C.R., Inverse Heat Conduction, John Wiley and Sons (1985). Carrier, G.F., Fendell, FF... and Feldman, P.S., Comb. Sci. Techn., 23, p.41 (1980). Carrier, G.F., Fendell, F.E., and Fink, 8., Comb. Sci. Techn.. 32. p.161 (1983). Carrier, G.F., Fendell, F.E., Butler, (3., Cook, 8., and Morton, C., Comb. Sci. Techn.. (1990). DiBlasi, C., Crescitelli, S., and Russo. 6., Proc. of the Eight Intl. Conf. on Comp. Methods in Applied Sci. and Engg., Versailles, France, (1987). DiBlasi, C., Crescitelli, S., and Russo, G., Comb. Flame, 72, p.205 (1988a). DiBlasi, C., Crescitelli, S., Russo, (3., and Fernandez-Pello, A.C., Twenty-second Symp.(lntl.) on Comb., Seattle, (1988b). Emmons, H.W., Zeit. fur Angew. Math. and Mech. 36(2), p.60 (1956). Emmons, H.W., Scientific American, 231(1), p.21 (1974). Emmons, H.W., Ann. Rev. Fluid Mech., 12. p.223 (1980). 139 140 Fernandez-Pello, A.C., and Williams, F.A., Fifteenth Symp. (1nd.) on Comb., p.217 (1975). Fernandez-Fella, A.C., Comb. Flame, 31, p.131 (1978). Fernandez-Pello, A.C., and Mac. C.P., Comb. Sci. Tech., 26, p.147 (1981). Fernandez-Pena, A.C., and Quintiere, 1.6., Fall meeting of Eastern Sect. of the Comb. Institute, Atlantic City, December (1982). Fernandez-Pello, A.C., and Hirano, T., Comb. Sci. Techn. 32, pl (1983). Groff, E.G., and Faeth, G.M., Comb. Flame, 32, p.139 (1978). Holman, J ., Experimental Methods for Engineers, McGraw-Hill Book Company (1984). Kashiwagi, T., Ohlemiller, T.J., and Werner, K, Comb. Flame, 69, p. 331 (1987). Kays, W.M., and Crawford, M.E., Convective Heat and Mass Transfer, McGraw-Hill Book Company (1980). Kulkami, AK, and Fisher, 8., Fall meeting of Eastern Sect. Comb. Institute, Clearwater Beach, (1988). Loh, H.T., and Fernandez-Penn, A.C., Twentieth Symp. (1nd.) on Comb. p 1575 (1984). Markstein, G.H., and deRis, J., Fourteenth Symp. (1nd.) Comb., p. 1015 (1973). Mekki,. K., Atreya, A., Agrawal, 8., and Mchman, 1.8., Twenty-third Symp. (1nd.) on Comb., (1990). Nurbakhsh, 8., PhD Thesis, Michigan State University, East Lansing (1989). Orloff, L., deRis, J., and Markstein, G.H., Fifteenth Symp. (Intl.) on Comb., p. 183 (1975). Quintiere, J ., Design of Buildings for Fire Safety, ASTM STP 685, p. 139 (1980). Quintiere, J., Harkleroad, M., and Hasemi, Y., Comb. Sci. Tech. 48, p. 191 (1986). Saito, K., Quintiere, 1.6., and Williams, F.A., Proc. First Intl. Symp. on Fire Safety 8ci., p. 75 (1986). Saito, K., Williams, F.A., Vlflchman, 1.8., and Quintiere, 1.6., National Heat Transfer Conf, Pittsburg (1987). Sibulkin, M., and Kim, J., Comb. Sci. Tech., 17, p.39 (1977). Tewarson, A., and Pion, R.F., Comb. Flame, 26, p.85 (1976). Vovelle, C., Akrich, R., and Delfau, J.L., Twentieth Symp. (1nd.) Comb., p. 1674 (1984). 141 Vovelle. C., Delfau, J.L., Reuillon, M., Bransier, J., Larqui, N., Comb. Sci. Tech., 53, p.187 (1987). Wichman, 1.8, and Agrawal, 8., Comb. Flame, 83, p.127 (1991).