}V153I_J RETURNING MATERIALS: Place in book drop to LIBRARJES remove this checkout from —c—. your record. FINES will be charged if book is returned after the date stamped below. EFFECT OF WATER ON IGNITION OF CELLULOSIC MATERIALS BY Mahmood Abu-Zaid A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Mechanical Engineering Michigan State University East Lansing, Michigan 48824 1988 Lfl m KW \\ k L1 3‘ ‘. KW ABSTRACT EFFECT OF WATER ON IGNITION OF CELLULOSIC MATERIALS BY Mahmood Abu-Zaid This experimental study is an attempt to understand and quantify the effect of water on fire extinguishment, thermal decomposition, and piloted ignition of wood. For the extinguishment part, the effect of water as externally applied on hot porous and non-porous ceramic solids was studied. Both solids were instrumented with several surface and in-depth thermocouples. Similarities and differences in their thermal behavior, and the heat transfer from the non-porous solid during the evaporation of a single droplet of pure water were investigated. This study will help determine the minimum water application, and the application strategy to cool as-yet-unburned objects. Thermal decomposition of wood as a function of sample moisture content and externally applied radiation was investigated. Experiments with three different moisture contents were performed at different heat fluxes in a controlled atmosphere combustion wind tunnel. Simultaneous measurements of weight loss rate; surface, bottom, and in-depth temperatures; 0 depletion; production of CO', CO, total hydrocarbons, 2 2 and water were made. It was found that the presence of adsorbed water delayed the decomposition process and diluted the decomposition products. For piloted ignition, experiments on Douglas fir with four different moisture contents were performed at different levels of externally applied radiation. In these experiments, in addition to the quantities measured in the decomposition experiments, the time for piloted ignition was also recorded. An absolute minimum mass flux of approximately 0.22 mg/cmzs was found to be necessary for piloted igntion to occur. It was also found that the presence of moisture increases the ignition time, surface temperature and evolved mass flux at ignition, and the critical heat flux. A single equation was derived to correlate all of the ignition data. This correlation accounts for the moisture-dependent thermal properties and the heat loss from the sample. To my father; Zaal, mother; Nadwa, my wife; Ensherah, and children; Sawsan, Anas, Ala, Esra, whose patient made the work possible. ii ACKNOWLEDGMENTS I express my sincerest gratitude to Professor Arvind Atreya for suggesting to me the fire extinguishment area as my dissertation topic and for his encouragement and guidance. Professor Arvind Atreya has been a source of profound inspiration, and working with him has itself been a rewarding experience. I am also very grateful for comments, suggestions, and encouragement provided by Professors James V. Beck, John R. Lloyd, Indrek S. Wichman and David Yen. I am thankful to Professor John J. McGrath for providing the video system used in the experiments. I am especially thankful to Dr. David D. Evans of NIST for his helpful suggestions and encouragement. I would like to thank my friends and colleagues, especially Mr. Said Nurbakhsh and Mr. Kamel El Mekki, for their valuable help and useful discussions. I am also thankful to the technical staff of the Department of Mechanical Engineering, especially Mr. Robert Rose and Mr. Leonard Eisele, for building the experimental facilities and machining the wood samples. Finally, I would like to thank Ms. Julia Pond for her diligent typing of the manuscript. This work was supported by the National Institute of Standard and Technology under the Grant No. 60NANBSOOS78, my graduate studies were supported by Mu'tah University, Jordan. iii TABLE OF CONTENTS LIST OF TABLES ................................................. vii LIST OF FIGURES ................................................ viii NOMENCLATURE ................................................... xiv 1. INTRODUCTION ................ . ............................... 1 1.1 Description of the Problem ................................. 2 1.2 Related Literature ......................................... 5 1.3 Present Work ............................................... 7 2. EXPERIMENTAL APPARATUS AND PROCEDURE ....................... 9 2.1 Droplet Evaporation Experiments ............................ 9 2.1.1 Apparatus ............................................ 9 2.1.1.1 Heated Ceramic Block ......................... 10 2.1.1.2 Droplet Generating System .................... 15 2.1.1.3 Data Acquisition Equipment ................... 16 2.1.2 Experimental Procedure ............................... 16 2.2 Wood Decomposition and Piloted Ignition Experiments ........ 17 2.2.1 Apparatus ............................................ 17 2.2.1.1 Combustion Wind Tunnel ....................... 17 2.2.1.2 Gas Analysis Equipment ....................... 21 2.2.1.3 Data Acquisition Equipment ................... 21 2.2.2 Procedure ............................................ 23 2.2.2.1 Sample Preparation ........................... 23 2.2.2.2 Calibration .................................. 24 2.2.2.3 Experiments .................................. 25 iv 2.2.3 Data Reduction ....................................... 26 2.2.4 Experimental Error ....... ‘ ............................ 27 3. DROPLET EVAPORATION ON HOT POROUS AND NON-POROUS SOLIDS ..... 28 3.1 Introduction .................. ’ ............................. 28 3.2 Background ................................................. 29 3.3 Literature Review .......................................... 30 3.4 Results and Discussion ..................................... 33 3.4.1 Transient Cooling of Hot Porous and Non-Porous Solids 34 3-4.1.1 Non-Porous Solid ...................... _ ....... 39 3.4.1.2 Porous Solid ................................. 40 3.4.2 Heat Transfer During Droplet Evaporation on the Non-Porous Solid ..................................... 58 4. DECOMPOSITION OF WOOD ....................................... 74 4.1 Background ................................................. 75 4.2 Literature Review .......................................... 76 4.3 Products Analysis .......................................... 78 4.4 Results and Discussion ..................................... 79 4.4.1 Sample Mass Flux ..................................... 80 4.4.2 Sample Temperature ................................... 85 4.4.3 Decomposition Products ............................... 90 5. PILOTED IGNITION ............................................ 102 5.1 Background ................................................. 102 5.2 Previous Literature ........................................ 104 5.3 Apparatus and Procedure .................................... 109 5.4 Results and Discussion ..................................... 112 5.4.1 Ignition Delay Time .................................. 112 5.4.2 Correlation of Results ............................... 114 5.4.3 Surface Temperature .................................. 122 5.4.4 Sample Mass Flux ..................................... 129 5.4.5 Products Evolved ..................................... 132 6. CONCLUSIONS ................................................. 140 6.1 Droplet Evaporation Experiments ............................ 140 6.2 Thermal Decomposition and Piloted Ignition Experiments ..... 142' 6.3 Recommendations for Future Work ............................ 143 APPENDIX A. METHOD OF ESTIMATING LOCATIONS OF IN-DEPTH THERMOCOUPLES AND THERMOPHYSICAL PROPERTIES OF THE CERAMIC ........................................ 145 A.1 In-depth Thermocouples Locations ........................... 145 A.2 Porosity ................................................... 146 A.3 Thermophysical Properties .................................. 146 APPENDIX B. CRITICAL NOZZLES ........................ ........... 152 3.1 Mass Flowrate .............................................. 152 B.2 Calibration ................................................ 153 APPENDIX C. THE GAS ANALYSIS SYSTEM ........................... 157 C.1 H20 - CO2 Analyzer ......................................... 158 0.2 CO Analyzer ................................................ 158 0.3 02 Analyzer ................................................ 159 0.4 Total Hydrocarbon Analyzer ................................. 159 LIST OF REFERENCES ............................................. 160 vi Table Table Table Table 5.1 5.2 1A 2A LIST OF TABLES Comparison of Calculated and Experimental Values of 0* and L ................ ‘. ......................... 122 Comparison of Mass Flux at Ignition for Different Moisture Contents at Various Heat Fluxes ............. 132 In-depth Thermocouples Locations ..................... 148 Magnesium Oxide Ceramic-Refractory Grade ............. 149 vii Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure LIST OF FIGURES Schematic of Experimental apparatus for-Droplet Evaporation ........................................... Ceramic Block Configuration ........................... Apparatus for Thermal Decomposition and Piloted Ignition Experiments .................................. Gas Analysis Equipment ................................ Surface Temperatures vs Time for a Non-Porous Solid During the Evaporation of a 30 pL Droplet ....... Indepth Temperatures vs Time for a Non~Porous Solid During the Evaporation of a 30 pL Droplet ....... Surface Temperatures vs Time for a Porous Solid During the Evaporation of a 30 pL Droplet ............. Indepth Temperatures vs Time for a Porous Solid During the Evaporation of a 30 pL Droplet ...... , ...... Non-Dimensional Surface Temperatures vs Non- Dimensional Time for a Non-Porous Solid at an Initial Surface Temperature of 100°C During the Evaporation of a 51 pL Droplet .................... Non-Dimensional Indepth Temperatures vs Non- Dimensional Time for a Non-Porous Solid at an Initial Surface Temperature of 100°C During the Evaporation of a 51 uL Droplet ........................ Non-Dimensional Indepth Temperatures vs Non- Dimensional Time for a Non—Porous Solid at an Initial Surface Temperature of 100°C During the Evaporation of a 10 pL Droplet ......................... Non-Dimensional Surface Temperatures vs Non- Dimensional Time for a Porous Solid at an Initial Surface Temperature of 100°C During the Evaporation of a 51 pL Droplet ..................................... viii 11 12 20 22 35 36 37 38 41 42 43 44 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure .10 .11 .12 .13 .14 .15 .16 .17 .18 .19 .20 .21 .22 Non-Dimensional Indepth Temperatures vs Non- Dimensional Time for a Porous Solid at an Initial Surface Temperature of 100°C During the Evaporation of a 51 pL Droplet ...................................... 45 Non-Dimensional Surface Temperatures vs Non- Dimensional Time for a Non-Porous Solid at an Initial Surface Temperature of 200‘C During the Evaporation of a 51 pL Droplet ......................... 46 Non-Dimensional Indepth Temperatures vs Non- Dimensional Time for a Non-Porous Solid at an Initial Surface Temperature of 200’C During the Evaporation of a 51 uL Droplet ......................... 47 Non-Dimensional Surface Temperatures vs Non- Dimensional Time for A Porous Solid At An Initial Surface Temperature of 200’C During the Evaporation of a 51 uL Droplet ..................................... 48 Non-Dimensional Indepth Temperatures vs Non- Dimensional Time for A Porous Solid at an Initial Surface Temperature of 200'C During the Evaporation of a 51 uL Droplet ......................... 49 Evaporation Time vs Initial Solid Surface Temperature for Various Droplet Volumes ................ 52 Recovery Time vs Initial Solid Surface Temperature For Various Droplet Volumes ............................ 53 Non-Dimensional Maximum Radial Influence vs Initial Solid Surface Temperature for Various Droplet Volumes ........................................ 54 Non-Dimensional Maximum Axial Influence vs Initial Solid Surface Temperature For Various Droplet Volumes.. 55 Non-Dimensional maximum Volume of Influence vs Initial Solid Surface Temperature For Various Droplet Volumes.. 56 Parameter B vs Initial Solid Surface Temperature For Various Droplet Volumes ................................ 57 Surface Temperatures For A Non-Porous Solid vs Time During the Evaporation of a 51 pL Droplet .............. 60 Indepth Temperatures for a Non-Porous Solid vs Time During the Evaporation of a 51 pL Droplet.........,.... 61 Contact Temperature vs Initial Solid Surface Temperature 64 ix Figure 3.23 Heat Transfer Coefficients at the Surface of the Ceramic block vs Initial Solid Surface Temperature ............. 65 Figure 3.24 Average Evaporative Heat Flux vs Initial Solid Surface Temperature For Various Droplet Volumes ................ 66 Figure 3.25 Average Evaporative Mass Flux vs Initial Solid Surface Temperature for Various Droplet Volumes ................ 67 Figure 3.26 Evaporation Time vs Initial Solid Surface Temperature for Various Droplet Volumes ............................ 68 Figure 3.27 Recovery Time vs Initial Solid Surface Temperature For Various Droplet Volumes ............................ 69 Figure 3.28 Non-Dimensional Maximum Volume of Influence vs Initial , Solid Surface Temperature for Various Droplet Volumes.. 72 Figure 3.29 Parameter B vs Initial Solid Surface Temperature for Various Droplet Volumes ................................ 73 Figure 4.1 Effect of Incident Heat Flux on Time-Dependent Mass Flux For Dry Condition ........................... 81 Figure 4.2 Effect of Incident Heat Flux on Time-Dependent Mass Flux for Dry Condition (Log-Log basis) ........... 81 Figure 4.3 Effect of Incident Heat Flux on Time-Dependent Mass Flux for 11% Moisture Content Condition ........... 82 Figure 4.4 Effect of Incident Heat Flux on Time-Dependent Mass Flux for 11% Moisture Content Condition (Log-Log basis) ....................................... 82 Figure 4.5 Effect of Incident heat Flux on Time-Dependent Mass Flux for 17% Moisture Content Condition .......... 83 Figure 4.6 Effect of Incident Heat Flux on Time-Dependent Mass Flux for 17% Moisture Content Condition (Log-Log basis) ....................................... 83 Figure 4.7 Effect of Moisture Content on Surface Temperature At Various Incident Radiant Flux ....................... 87 Figure 4.8 Temperature Versus Time at Various Locations in the Sample at 2 W/cm2 for 9% Moisture Content Condition ... 88 Figure 4.9 Temperature versus Time at Various Locations in the Sample at 3 W/cm2 for 17% Moisture Content Condition... 89 Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure .10 .ll .12 .13 .14 .15 .16 .17 .18 .19 .20 .21 Time-Dependent Mass Flux of Evolved Products at 2 W/cm2 for 9% Moisture Content Condition ............. 91 Time-Dependent Product as Percent of Mass Flux at 2 W/cm2 for 9% Moisture Content Condition ............. 91 Time-Dependent Mass Flux of Evolved Products at 3 W/cm2 for 9% Moisture Content Condition ............. 92 Time-Dependent Mass Flux of Evolved Products at 3 W/cm2 for 17% Moisture Content Condition ............ 92 Time-Dependent Mass Flux of Evolved Products at 4 W/cm2 for Dry Condition ............................. 93 Time-Dependent Mass Flux of Evolved Products at 4 W/cm2 for 19% Moisture Content Condition ............ 93 Time-Dependent Mass Flux of Evolved Products at 1 W/cm2 for Dry Condition ............................. 9a Time-Dependent Mass Flux of Evolved Products at 1 W/cm2 for 17% Moisture Content Condition ............ 94 Time-Integrated Product Mass and Composition (shown as percentages next to bar) As Function of Incident Heat Flux for 9% Moisture Content Condition .. 97 Time-Integrated Product Mass and Composition (shown as percentages next to each bar) As Function of Moisture Content at a Heat Flux of 2 W/cm2 ......... 98 Time-Integrated Product Mass and Composition (shown as percentages next to each bar) As Function of Incident Heat Flux for 17% Moisture Content Condition 99 Time-Integrated Product Mass and Composition (shown as percentages next to each bar) As Function of Moisture Content at A Heat Flux of 3 W/cm2 ............. 100 Schematic of the Ignition Process ..................... 110 Effect of Moisture on Ignition Delay Time ............. 113 Terms of Equation 5.4 vs Incident Heat Flux ........... 117 Correlation of Ignition Delay Time .......... . ......... 120 xi Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure Figure .10 .11 .12 .13 .14 .15 .16 .17 .18 .19 .20 Effect of Incident Heat Flux on Surface Temperature For Dry Condition ..................................... 124 Effect of Incident Heat Flux on Surface Temperature for 11% Moisture Content Condition .................... 125 Effect of Moisture Content on Surface Temperature at an Incident Heat Flux of 2.65 W/cm2 .................... 126 Effect of Moisture Content on Surface Temperature at Ignition ........................................... 127 Sample Temperatures Histories for 11% Moisture Content at 1.8 W/cmz .......................................... 128 Effect of Incident heat Flux on Time-Dependent Mass Flux for Dry Condition ................................ 130 Effect of Incident Heat Flux on Time-Dependent Mass Flux for 17% Moisture Condition ........................ 131 Mass Flux of Evolved Products at Incident Heat Flux 3.5 W/cm2 for Dry Condition ........................... 133 Products as Percent of Mass Flux at Incident Heat Flux 3.5 W/cm2 for Dry Condition ...................... 133 Mass Flux of Evolved Products at Incident Heat Flux 2.6 W/cm2 for Dry Condition ...................... 134 Products as Percent of Mass Flux at Incident Heat Flux 2.6 W/cm2 for Dry Condition ...................... 134 Mass Flux of Evolved Products at Incident Heat Flux 2.75 W/cm2 for 11% Moisture Content Condition ......... 135 Products as Percent of Mass Flux at Incident Heat Flux 2.75 W/cm2 for 11% Moisture Content Condition .... 135 Mass Flux of Evolved Products at Incident Heat Flux 2.77 W/cm2 for 27% Moisture Content Condition ......... 136 Products as Percent of Mass Flux at Incident Heat Flux 2.77 W/cm2 for 27% Moisture Content Condition .... 136 Mass Flux of Evolved Products at Incident Heat Flux 1.85 W/cm2 for Dry Condition ..................... 137 xii Figure 5.21 Mass Flux of Evolved Products at Incident Heat Figure Figure Figure Figure 1A 2A 13 2B Flux 3.21 W/cm2 for 17% Moisture Content Condition .... 137 Lease Square Fit to the Steady State Temperatures ..... 150 Transient Indepth Temperatures During a Step Increase and Cut-Off of Heat Flux ..................... 151 Configuration of the Critical Nozzles .................. 155 Experimental and Theoretical Flow Rates of Air Through a Critical Nozzle of Diameter 0.0785" ................. 156 xiii ‘ l e \ IEIHMCLATURE a constant D/d specific heat specific heat at constant volume specific heat at constant pressure diameter of droplet before impact maximum diameter of droplet on surface incident heat flux heat flux entering the solid release height, i.e., the distance between the needle tip and the solid surface heat transfer coefficient overall heat transfer coefficient thermal conductivity liter heat flux rate of heat radius of the disk maximum radius of droplet on the surface solid temperature equilibrium temperature xiv sat (X.y.2) (€.n.¢) W a contact temperature initial solid surface temperature in droplet evaporation analysis, solid surface temperature in piloted ignition analysis ambient temperature saturation temperature droplet temperature before impact time averaged surface temperature TWS - Tsat time dimensionless time, t/r initial volume of droplet volume of droplet influence in solid cartesian coordinates oblate spheroidal coordinates water thermal diffusivity Jpck penetration depth critical droplet thickness, i.e., droplet thickness on the non-porous solid at the time the evaporation rate starts to increase sharply emissivity T-T T -T in droplet evaporation analysis, T-Tco in piloted s e ignition analysis Ts-Tan density Stefan-Boltzmann constant evaporation time of droplet CHAPTER 1 INTRODUCTION Cellulosic materials such as wood, cardboard, paper and cotton often constitute the bulk of the fuels in many fires. .A common fire starts in a residential room by a source of ignition. Thermal radiation from the flames and the hot gases may heat and cause ignition of nearby combustible materials allowing the fire to grow. The process of fire spread is closely related to piloted ignition, and may be regarded as a rapid succession of piloted ignitions. Today, in the most technologically advanced civilization, we still face the threat of unwanted fires. A recent estimate shows an annual loss of $10 billion of property and forest in the United States. More important and not expressible in dollars, is the annual loss of thousands of lives, and millions of people sustaining serious burn and smoke inhalation injuries. A fundamental study to understand the chemical and physical process that occur during combustion of cellullosic materials will result in achieving effective control and prevention of unwanted fires. In the burning of Cellulosic materials, water plays an important role from ignition to extinguishment due to their porous and hygroscopic nature and the availability of a large area to absorb moisture from surrounding air. The moisture content of these materials is variable, and is influenced by the climate, season and location. Also, water is the most common agent used to extinguish unwanted fires. Thus, a study of ignition and extinguishment of combustible cellulosic materials is of great importance from the point of view of fire safety because these materials constitute the bulk of fuels in many building fires. This study is an attempt to understand and quantify the effect of water on piloted ignition and extinguishment of cellulosic materials. Piloted ignition has been chosen because it is closely related to fire spread and occurs at lower critical temperatures relative to spontaneous ignition. Consequently, it is more hazardous. This work will be useful in determining the onset of piloted ignition, the fire growth rate under different humidity conditions as well as the minimum water application rate and the application strategy to restrain the fire growth. 1.1 Description of the Problem The problem of extinguishment by water may be divided into two parts; (i) extinguishment of the already burning objects and (ii) prevention from burning of as-yet-unburned objects [Atreya (1985)]. In the cooling of unburned solids, the physical process of water evaporation on hot porous and non-porous solids is different. Results from some early experiments conducted by Atreya (1985) indicate that a water droplet impinging on a hot surface of wood (surface temperatures 0 . in piloted ignition range from 300 - 400 C ) behaves differently than it does on a hot copper plate at the same temperature. On the copper plate, a thin vapor film is formed between the droplet and the hot surface. As a result of this layer, the heat transfer is considerably lower than in the case of direct contact (liquid has much better thermal transport properties than its associated vapor). On the other hand, in wood the droplet is quickly absorbed and re-evaporated. Thus, in-depth cooling of the solid occurs. The investigation of ignition of cellulosic materials must be preceeded by the study of thermal decomposition of such materials. The physical aspects of decomposition are associated with the heat and mass transfer, processes which occur as the solid is heated. The external heat is primarily transferred to the surface of the solid by radiation. This radial energy will be transferred into the solid by conduction process. As the temperature at any point rises, hygroscopic water is first released, followed by decomposition of the solid in a complex fashion. About two hundred compounds have been identified as the products of decomposition of wood [Coos (1952)]. The rate at which these products are generated depends upon the response of the solid phase to the applied heat flux. These products are driven to the surface of the sample by high pressure generated inside the solid where they mix with the surrounding air. A good quantitative understanding of the factors controlling the decomposition of the solid, and the composition of the evolved gases, is important for ignition and fire spread and for fire prevention in general. It is also important in the context of minimizing air pollution from wood burning stoves, Ohlemiller (1987). The ignition mechanism is influenced by factors both external and internal to the sample. Factors external to the sample are the environmental variables, such as temperature, composition and velocity of the surrounding gases, etc. Factors internal to the sample are its thermophysical and thermochemical properties and its moisture content. Experimentally, all variables other than the thermochemical and thermophysical can be controlled. The actual process of ignition is quite complicated. A simplified phenomenological picture of piloted ignition by radiation has been clarified by Atreya (1986) and Kashiwagi (1981). The solid must first chemically decompose to inject fuel gases into the boundary layer. These fuel gases must then mix with the surrounding air, and the local mixture ratio must be near or within the flammability limits. At this instant, a premixed flame, originating from the pilot flame, flashes across the surface of the solid through the fuel-air mixture formed in the boundary layer. Further heating of the solid results in evential ignition and finally the establishment of a diffusion flame in the boundary layer. The adsorbed water in the wood matrix could effectively delay the wood decomposition process, and reduce the flammability of wood (i.e., increase the ignition time) due to the high heat of vaporization of water. In addition, water vapor could complicate decomposition reactions, such as the reaction between water vapor and wood char to produce carbon monoxide and hydrogen at high temperature. In view of the above observations, the research problem is logically divided into two parts: (i) Effect of water as externally applied in the form of droplets, (ii) Effect of adsorbed water on wood decomposition process, and piloted ignition as a function of sample moisture content and externally applied radiation. 1.2 Related Literature In this section, only the general literature related to the problem is discussed. Literature related to the specific component of the problem is discussed in the appropriate chapters. For droplet vaporization a number of studies on hot non-porous metallic solids have been reported in the literature. These studies show that the vaporization mode of a liquid droplet on a hot surface depends on many factors. These are: (1) initial surface temperature of the solid, (ii) isothermal or non-isothermal condition of the solid, (iii) thermal properties of the droplet and the solid, and (iv) the droplet momentum at impact. Few studies [diMarzo and Trehan (1986)] have focused on in-depth cooling of the solid (which is important for predicting the rate at which fuel gases are produced), and none have reported transient in-depth temperature measurements. A substantial portion of the combustible building materials are porous with low thermal conductivity and thermal diffusivity. Vaporization of water droplets on such solids is expected to be different from that for the non-porous metallic solids. However, despite the need to address the cooling of hot porous chars and unburnt wood during fires, the author has been unable to find relevant literature for porous solids. A considerable amount of work has been done on the thermal decomposition of wood and other cellulosic materials. These studies were performed using different materials and under various conditions. Martin (1965) investigated the decomposition of pure cellulose in helium, Atreya (1983) investigated the decomposition from different kinds of wood in air. The decomposition of PMMA and particle board subjected to variable heat flux under an inert atmosphere was investigated by Volvelle et a1. (1984) Kashiwagi (1987) studied the products generated from wood decomposition in 3 atmospheres of different oxygen concentration (N2, 10% O2 - 90% N2, and air). Lee and Diehl (1981) investigated the effect of absorbed water on samples containing 50% of oven-dry weight as water. They indicated that moisture not only changes the solid-phase thermal properties but also substantially dilutes the decomposition products. Atreya (1983) from his recent work concluded that the desorption of moisture, which has been ignored by previous investigators, has a considerable effect on the energetics of the decomposition process. Ignition of cellulosic materials (both spontaneous and piloted) has been an active area of research in the past. Several excellent reviews have been published on this subject [Welker (1970), Kanury (1972), and Steward (1974)]. Although numerous techniques have been developed to investigate piloted ignition phenomena, the experiments essentially consist of exposing a sample to a known external radiant flux and recording the time required to ignite it in the presence of a pilot flame. Of the proposed ignition criteria, critical fuel mass flux at ignition [Bamford et a1. (1946)] seems to be the most correct physically (since it can be related to flammability limits), whereas the critical surface temperature at ignition has proved to be the most useful [Atreya (1983)]. However, only a few investigators [Atreya (1983), Kashiwagi (1981), and Garden (1953)] have actually measured the surface temperature at ignition, and measurements of the critical fuel mass flux at ignition for cellulosic materials are not available due to the complications caused by the simultaneous evolution of absorbed moisture. The importance of moisture content in relation to fire tests has been recognized. Researchers have carefully controlled the moisture content of their samples, but little attention has been devoted to the effect of sample moisture on decomposition and piloted ignition. This study is a systematic investigation of its effect. 1.3 Present work In this work, the effect of water on piloted ignition and extinguishment is investigated experimentally. The apparatus and procedure used, experimental errors incurred, and data reduction procedures are described in Chapter 2. For the droplet experiments, the heated solid was cast from ceramic and instrumented by several surface and in-depth thermocouples. In order to compare the results of porous and non-porous solids under the same conditions two identical solids, porous and non-porous, were cast from the same ceramic powder. Using different droplet sizes, a systematic set of experiments (initial solid temperature ranged from 75-325 0C) were performed. In these experiments, transient surface and in-depth temperatures, and the droplet evaporation process were recorded. The results of these experiments and a discussion is presented in chapter 3. Having recognized the importance and need for understanding the phenomenon of thermal decomposition for the piloted ignition problem, a set of experiments with three different moisture contents were performed at different levels of externally applied radiation. In these experiments simultaneous measurements of several physical and chemical quantities were recorded. The results of these experiments are described in Chapter 4. For piloted ignition, a set of experiments with four different moisture contents were performed at different levels of externally applied radiation. In these experiments, besides the quantities measured in the decomposition experiments, the time for ignition was also recorded. Critical surface temperature and critical fuel mass flux ignition criteria were discussed, and a correlation for ignition data was derived and discussed. Chapter 5 describes the results of these experiments. Finally, conclusions for the various components of this research problem and some recommendations for future work are presented in Chapter 6. CHAPTER 2 EXPERIMENTAL APPARATUS AND PROCEDURE The primary objective of the present work is to understand and obtain useful data for droplet evaporation on hot porous and non-porous ceramic solids, wood decomposition, and piloted ignition of wood. Experiments for these components were performed using different facilities. This chapter describes the apparatus and procedure used for each component, the data processing methods, and errors encountered in these experiments. 2.1 Droplet Evaporation Experilents. 2.1.1 Apparatus Experimental study of cooling of a hot, low-thermal-diffusivity porous solid by water droplets requires an apparatus that satisfies the following conditions: (1) The solid should be nearly isothermal, porous, and made of a low thermal conductivity and diffusivity material, and the temperature of the heated solid must be controllable. (2) The heated solid must be instrumented with surface and in-depth thermocouples. From these transient temperature measurements during droplet evaporation the following information can then be determined: (1) recovery time i.e. the time for the surface to recover to its 10 initial temperature, and (ii) the size of "surface and in-depth zones" affected by the droplet. (3) A system capable of generating different size droplets is also required. This system must be capable of delivering droplets on the solid surface at a specified rate and at a specified release height. An experimental apparatus was built to satisfy the above requirements, a schematic is shown in Figure 2.1. 2.1.1.1 Heated Ceramic Block To obtain a nearly isothermal solid three different shapes of the ceramic blocks were considered. These were (i) Hemi-sphere: Here, to make the flat surface isothermal, a large variation of the heat flux applied to the hemispherical surface is necessary. (ii) Cylinder: It is possible to obtain a flat isothermal surface in this geometry if the cylinder is heated from below with a constant heat flux, and the cylindrical surface is kept well insulated. The main disadvantage is that a very small cylinder height is needed for the solid to be nearly isothermal. (iii) Oblate spheriodal: This geometry has been chosen because it combines the advantages of both cylindrical and spherical geometry. It also provides a natural co-ordinate system for this problem. A castable porous ceramic (MgO) was used to cast a semi-oblate spheroidal solid. This solid was used to simulate an isothermal semi- infinite solid. The configuration of this block along with the calculated temperatures and heat flux is shown in figure 2.2. The block has a major semi-axis of 8.48 cm and a minor semi-axis of 7.22 cm. Theoretical calculations described below show that this ceramic block 11 > oHH 53.33;» 3339 you. monsoon? Hanson—Coax” no causaosom a.” can»: Hmunm um “N30“ ommmm ou mAHz0: /. «mamnho< Hmoumm onH on wqueao onm 12 .musoauuonxo ooanouv page: new swoon «Houses on» ma Aolmv wanna use .ooseaon «any o>ownue ou woumasuauo as: Auodo.mofiv unease: ago no ousuuuonaou one .muoumo: one aoum oomuusm man» as msw>wuum Roam use: one On assoc ea aoaumwvou use souuoo>aoo an commune cocoons on» aouu umoa usau one: one .Aeuov muses en» su .nH.H .35 .36 .o um 09 unannouuou mane—Loom.“ 05. 303333.30». 3001— 38930 ~.~ was»: Aemzou use nmzHa wane u see Aamzoo out mammmaomH amamn uzh Hm: mxmh¢mmmo 3o i—“v (snows svo) 31m aaivan f meme< z nousueuoaaeu. edema—am n6 swam: Aommv oEfi 0mm . 03. 0am 0am 0m. F .38 308.. a .8 8382 29885.0. . ea 3 a. e. caucuses E... a. o . oESoofiafi as a. n . £38535. as n . v . 3:82.55 as c n . snoooosafi es 4 a . £88285. £50 a . €885.25. _ (3°) SSJnloJedUJSi 3:30an .uoaaoua a: On a Mo coaunuonfifi 05 95.3.5 v.30» 39.3." I you 2.3. g nousunuogoa 5955 in can»; Aoomv oEt. oem 00¢ com com 00. o . _ . F b — p — p _ p on .38 328 a 8.. 2882 390863;... .5: Q: 2 . 033853 .. E... in .2 u 03:08:55. IDN as; 3%; 883 «.0333 on .33 2%8853 f .l (3°) seJnmJeduJal HldGPUI 39 For both the porous and the non-porous solids at an initial surface temperature (Ts) 5 164°C, the hot surface immediately cooled to the theoretical contact temperature upon contact with the droplet. For TS > 164°C, the surface temperature, upon contact with the droplet falls considerably below the theoretical contact temperature. Seki et a1. (1978) found that on stainless steel, the measured contact temperature increases with increase in the surface temperature up to 180°C, and was in agreement with the theoretical contact temperature. For 200 5 TS 5 300°C, they found that the measured contact temperature was less than the theoretical contact temperature and.was approximately constant. The present results for the ceramic are in qualitative agreement with Seki's data except that the present data shows a deviation from the theoretical value at 164°C rather than at 180°C. After achieving the contact temperature the solid surface approaches the "equilibrium temperature" (interface temperature during evaporation). During the droplet evaporation period the thermal behavior of the porous and non-porous solids are different. These differences are described below: 3.4.1.1 Ronrporous solid Measurements show that for TS - 75°C, the surface and indepth temperatures remains nearly constant during the droplet evaporation process. This implies that the heat flux during the evaporation time is also nearly constant. For 100 5 T8 5 175°C, the interface temperature * remains nearly constant until a critical drOplet thickness (6 ) is 40 reached on the hot surface. Then, the temperature starts to decrease , sharply until the droplet has completely evaporated as shown in Figure 3.5. After the droplet has completely evaporated, the solid recovers Quickly to its initial surface temperature. This sharp decrease in the surface temperature implies an increase in the evaporation rate. This will be discussed in more details in section 3.4.2. For T - 200°C, the interface temperature drops to a value few 8 degrees above the saturation temperature of water (see Figure 3.10). Then the surface recover to its initial value. The evaporation times at this temperature were very short, since the droplet diameter after impact on the surface was very large (see Figure 3.19). This large droplet diameter is due to lower surface tension. 3.4.1.2 Porous solid For porous materials the interface temperature never clearly attains an equilibrium value. Both the surface and the indepth temperatures continue to decrease until the droplet vanishes from the surface (as can be seen from Figures 3.8 and 3.9). The time taken for the droplet to vanish from the surface is defined as the evaporation time. By this time a part of the droplet has already been evaporated, and the rest has penetrated in both the axial and the radial directions. The effect of the droplet penetration is clear from following two observations: 1) A thermocouple in the porous matrix at the same location as the non-porous matrix cools faster under otherwise indentical conditions. This implies that, for the porous matrix, there must be an energy sink in the neighborhood of the thermocouple. Thus confirming the indepth 61 no 3332—38. onus Hosea-soaao .uoaaoua A: an e no soaueuoaq>m as» mcauso 0.00“ coauuam ueuuusn so us 00000 nsouom.soz a you .soz n> copay-unmask oonuusm Hosea-soaao.Coz n.n ouawum ...”. mEz. _oco_mcmEElcoz e; N; 0; 0.0 0.0 v.0 N0 0.0 N0! — . p . _ . _ . _ . _ p _ . _ O.Nll 6:8 388.8,. .. 3. .8532 338252... . EE «.0 h . 20:08:55. 0 8 no“. .. .88 h 0 . 30:08:55. ES» .n.o_a=o8=5.c. r043: E...“ v. 0.038.505. as v n. 30800535. 1 .5... 2033853.: no.7. 2.50 ._ 238852.... 1.0.0...» 0 . 10.0 a . v I00 2......1.--\ -1 .Flt- this . - a. x 10; 0 seJmoJeduJal aoopn'g louogsuang—UON 42 .uoanoua A: an a mo souuuuomefi on» 05.30 0.00." no assuage—.3. 030.30 .2335 as us 3.3m asouoméoz a won 053. assoausoaaaéoz 2' masseuse—Sn. sudopaa decode—550502 06 9.5»: *0 mEC. _oco_mcmE_olcoz .3 a; .o._ 90 0.0 ....o do 0.0 «.0: _ . — . — . P . p. p — p — — O.NII i=8 388.8: a 3.. 209.8. 39.8853... . EE 0.: = 0 03:83:20. 0 we I on. .. 2&3 223885.... 10.7. 65nd massage... .. SEQ.— «co—aggfi » 88...” :2038593 i.e.—...: + [0.0: rod 10.0 re; a uoN a seJmoJeduJel Lndepm louogsuowgg— 43 .uoaaoua A: 0H e no soauauoau>m onu usausa 0.00. no cuss-nomeoa ooeuusm Heauusu an as vaaom esouom.soz a you 06.9 Assad-606.0.soz u> nouauquoaaoa nuaovsu Hence-606.0.soz s.n ousuwm .... mEc. _oco_mcmE_olcoz 0.. .... a; 0.. who who .30 «“0 £0 «.0: _ . . 0.0.. . i=8 328.80 a .6. 22.8. 20.68505 5. a... .. . 20.6852... 0 8 0.. .. SE 0.0 o. 0 296852.... 1. m. P I 5.. n... 0 . 20.68505 . EE 0.. 0 0.296858... . 85:0 . 0 208852.... ... O. F .... 10.0.. p p n - H . 2 Jr” 10.. l 0. O eseanoJedwai Lndepul louogsuew‘gQ—UON (+4 .00a0000 A: an 0 00 souuauo0o>m on» 00.»:0 0.00. «0 00500000809 0000000 Huuuucu 00 00 0HHom asouom 0 you 05.8 Husoaosoa.0.:oz 0> 00u00000050a.0000030 Hucoaaaoeua.soz 0.n ouswam .... 0E: _o:o_mcmE_olcoz 0.\. 0.0 0.0 0.... 0.0 0.N 0.. 0.0 0. T... . . . _ . _ . _ . _ . 0 . . 0. _..1 .0000 26.00 a .6. 30080. 20080520. 5: 0.0 .. 0 2000852.... 0 a0 10... . .EE 0 0 0 0.0000052... ....w.0.l .5... 000.00.08.52... . :50 v0 0.0080505. Wum 0' .500 22008052.... r.v.O.l SE 0 a 0 0.0008500... 00.8.0... .. 850 ..208852... . TN.O.1 .00 N . 1N.0 10.0 10.0 1.. .. ....w 0 10._. a 1N.— .. 0.. BSGJnlDJedLUGi aooyng louogsuewgg—UON 115 .uoa0000 A: an 0 00 00.000000>m on» mCHuso 0.00H mo 00000000309 0000050 Hquuacm ca 00 0.Hom nsouom 0 now 05.9 Hucoamc0e.0.coz m> 00u=unu0050a £00000H H000.m¢0e.0.coz 0.n ouswam ... 0E... _0co.mcmE_01coz 0.0 0mm 0mm .0? own 0N 0“. oflo 021 0:8 26.00 :0. 2.0.68. 20008505. 0 . O —.l .50... 30208853.. 000... ...lw.O.l 65.0.0.0. 02000805.... . . 5.; 0.208850... ”.0 0| EEQ— ..w§o_na§§fi o l¢.0|1 2030 .uoaSSEip 028- a . .1N.O.1 10.0 1N0 -....0 10.0 10.0 w 111111 2| 0.. .... e SQJnloJedel mdepu; {Duogsuewlg—UON 46 .00H0ou0 A: an a mo coaueuo00>m 0&0 00.000 0.00N mo ousuuu00aoa oouuusm Huauusu :0 00 ouaom 0:0u00.coz 0 now 03.9 ..50..a:050.:oz 0> 00.309.000.00. 0000.30 1:05:030002 0.6 9.50... .... 0E... 600.058.0152 0.0. 0._N— 0M0 0“0 - b 0.0.... ......s 58.8.. . a. 23.8. 200852.... 5.2 .. . 20.38520. 0 .0. .. .... .E... .. 0 0 0.000850... 5.. 0 0 0 2000852.... :5 0 v 0 20000052.... 02:0. n..§:622sa#h E... v N 0 2000805.... .. ... ..rww . m ”1m 2.000 ..296852F 0.0 9 SGJnlDJadUJSl eoopng jDUO1SU8UJ10—UON £17 .0000000 A: an 0 00 00.000000>m 000 00.000 0.000 00 00000000309 0000000 00.0.0. :0 00 00.00 000000.002 0 000 05.9 00000000600.coz 0> 000300000509 £00000. H0000000300.coz HH.n 0030.0 .. 08: 6020000501002 o. 0.0. 0.0. 0.0 0.0 0.0... n — _ , — . — p 0.0 ... 8 30.00.02. 0 .0 2.0 08 0 9.0850 0 = .. .0 . . .0. e... 0.: .. . 20885.0. 0 .0. u 0. . . .5. 0.0 0. . 20385.0. ...... 0.... 0 . 20885.... 1 N0 ...... 0.. a 0 0.968500. 08.5.0 . 0 0.0008520. . 1 ...0 ... 0 .0 1 0.0 0 . 1M / ...: e SGJnlDJadLUGi Lndepul [ouogsuewgg—UON 48 00.30.00 A: an 0 m0 03000000>u 05 900030 0.00“ «0 00300000600. 000.0030 H3300" 00 u< 0.300 030000 0 00m 0.0.3. 100000030002 0> 000900000500. 000003 0000000050002 36 00030 .L 08:. 6020000051002 o.m 0.0 0.0 o.m 0.0 o.n 0N o; 0.0 0.7. . _ . . . . . L 0 — 0 _ 0 _ 0 L . N.OII .2000 3800 0 00.. 0002000— 3000000.0005. 0 5300 .20—088590. 002.. 0. . .0000. 0.0—00000055 p . E0. 0. n 0 3000000005.. IO 0 3&0 1300840500. . 0.0; n00.00800cofi . .00; «0030085800. N TNO 880 22088690. .. 4.0 T06 ..md ..... -00 m.— _e SQJmpJedUJSl eooyng [ouogsuewgg—UON 49 .0000000 .00 0m. 0 .00 00000000020 000 000000 0.000 .00 00000000000. 0000000 0000000 00 00 00000 000000 0 00.0 0000. 00000000300002 0> 000000000500. £00005 00000000200002 00.0.. 00000.0 *0 00.5 6020000051002 0.0 0K. one ohm 0.0 0?... ohm 02 ouo oql . . . . . _ . _ . 0.01 .33 0000000000 200000.0—00800005. EE 0.: : 00.088055. 08— no.0. .. SE in o— 0000008§ _ 1.0.0 00...; 0030080055. .552 0000088590. . 000050 00.800.890.05. ...N.O SP“; .. 0 0 0 L Tao 10.0 r 10.0 “IO.— j‘ll . : r0 0 e SSJrnDJSdLUQl Lndepul louOgsuewig—UON SO penetration of moisture. (ii) The evaporation time for the porous case is lower than the non-porous for the same droplet size and under the same condition. After the droplet vanishes from the surface, the porous solid recovers slowly, as opposed to the quick, pure conduction, recovery for the non-porous solid. The porous indepth temperature profiles during recovery are not as smooth as the non-porous solid, due to the migration of moisture inside the matrix (as can be seen from Figures 3.4 and 3.9). The presense of the moisture changes the thermal properties of the solid. This change makes the recovery behavior not only different but also longer than the non-porous solid. The evaporation time for various droplet sizes versus intial solid surface temperature is shown in Figure 3.14. This time is as defined previously, the time taken for the droplet to vanish from the surface. Figure 3.14 shown that this time is a function of the droplet size and the intial surface temperature. As expected, lower the initial surface temperature and larger the droplet volume, more time is required to evaporate the droplet. Figure 3.14 shows also that for TS 5 164°C, the evaporation time of the non-porous is a larger than the porous. For Ts > 164°C the evaporation time for the porous is larger. For this range the parameter B for the non-porous solid was larger than the porous solid (see Figure 3.19). Generally, the evaporation time on the non-porous solid is larger than the porous solid for the same droplet under the same conditions. This indicates that in the process of cooling hot porous and non-porous solids, the porous solid needs droplets at higher frequency. 51 The recovery time is shown in Figure 3.15. This time is defined as the time for the surface to recover to its initial temperature. This time was determined by evaluating the time it took the surface temperature to recover to 20 percent of the maximum temperature drop. The maximum temperature drop is the difference between the initial surface temperature and the equilibrium temperature of the thermocouple underneath the evaporating droplet. The non-porous solid took slightly less time to recover than the porous solid despite the fact that for the non-porous solid, the evaporation process is steady and the evaporation time is larger. Figures 3.16 and 3.17 show the non-dimensional maximum radial (x/rd) and axial (z/rd) influence distances of a single droplet plotted against the initial surface temperature. The influence distance encloses all locations where the temperature drop due to the droplet evaporation is at lest 20 percent of the maximum temperature drop i.e., the non- dimensional temperature 0 - 0.8. Figures 3.16 and 3.17 show that the influence distance decreases at higher solid temperatures as more energy is available to evaporate the droplet. Also these influence distances are larger for the porous solid. The non-dimensional maximum volume of influence (Vi/V) is shown in Figure 3.18. This quantifies the cooling effect of a single droplet. The volume of influence is defined as the volume of the semi-oblate spheroid formed by the axial and radial influence distances. Due to the low thermal conductivity/diffusivity of the ceramic, the droplet will produce intensive local cooling, because heat recovery from other portions of the solid matrix is limited. Thus, the influence zone is 52 . 3532, unseen «so?!» you 339.3309 ooumusm 33m .2335 a> on; 60388926 36 0563 AOL 9.383th mootam 2.0m BEE now 02 m: cm. 0%. cm? m—... 2.: mm ... druid”. a I | I. t i ' ' -H A I. I - 3L SEEKEB. oootam 2.0m .03.... u 9" - mm. . ofL at . 3. . om. . 3... r at 1' II lo'"" I 1.0 . IIII 1” III, II .. I IQ IO. D. llllllll II I», m I u I I I . I. .8u2.8.. I I; 11 m ...... I. x ,d .2 h is; o I; s .12; . z a .184 a I I. 12.5 . 13.5 . on I 8 (ass) emu uogoJodD/g 53 . .3332, 3398 28.3.3 you 8333309 3556 3.3m H335" u> as: hugged 36. enough 8.. otabotanoH mootam 30m 625 now cm. wt of. .22. .23. .01. .03. mm on 1|).l1» . — . b 0 row // Too. m. 8.. $322.53. 8325 2.8 3...... .0 . o 2...... . . . . / sea 4 . M / / ICON “.4 a x . w m m. .63 a . s.) .... a I m , .. ogunssmueun /// III N) /// soon {3-\ L .I one / .6 I ...“.mnmn / a m .. -3... “Hum”; on. // . V .. > a s 8.. 5h .urasao>.uoanou0 33.25 you 333.038. 33.30 03o» 7.3.25 2r 9233.35 H363. 55x3. announces—3&3. 35 can»: Aoov 9.30..an0... montam 2.0m BEE new. .03. MAL. .0m... .92. .23. .mt. .0n.:. mm 0N. . _ . . in.“ ......nn. H . .. M/ -3. .12.... 0.... /.//u . 12.5 To MEN/MU. 4100.; nun. . N I lwmn H 10.0 .12->01. , IIdI/r/M .. .... 5...... 7» I 0.0 . 9n (pJ/X) aouenuug [ogpoa 'ngow louogsueng—uo 55 .uossHo> uoaaouo nsouue> you ousuauodaoa ooeuusm vqaom asuuunn o> cocosducn anqx< essaxaz Hacodacoe«0.coz s~.n ousmmm 8.9 238380... montam 2.0m .025 new 0? 0m... 0m. 0...; 0m: 0: 0a.: 00 h 0 — p b 0n. .aouoa.coz .souoa sao.n> saa.n> sLonn> 41n¢n> ...nn> aouenuul [0ng ways»; youogsuewgg—UON 56 ..oa:~o> uoanoua nsoauu> uom ousuuuoaaoa ooauuam page» Houuacn u> cocoaaucn «o oasao> asauxdl uncoaucoa«0.:oz 0H.n ousuum 8L. 830..an3 mootam 23w .025 com .02. .0? 3.. .92. .9»... .m... .3... .mm. .o naouoaucox I 059.30 0.16 ....o.n> o .32.; a ....on..> a .32.; o .16.; p ( A/EA) aouanuu; ,Lo awnloA 'xow [ouogsuewgg—UON .35."; nod—85 3325 you 3313939 .0336 3.?» .7313 2. n nova-.33 36 33»: 8L 9.30..an2. 32:5 2.8 .025 2.x . om. . om. . o... . o? . 2.... o.» . 8 2 o _ . - . S7 \»II I... II \h/ \x d\ \ \\ \ %\ \ .4.‘ I... & mu \4\ zlox ... \- \| \ ... \0\ 0.. osouoaucoz I :6»: 9.6 .32.... o .32.; 4 1.8.5 o 4:23.. o .3 .9... p Jagawwod 8 58 expected to be relatively high. The influence zone of the porous solid is larger than the non-porous solid. This is due to the enetration of the droplet into the solid matrix, thus causing more cooling in both the radial and the axial directions. Figure 3.18 shows that the non—porous solid at T < 100°C cools more with a 10 pL droplet than with larger S droplets. This is because the evaporation process for larger droplets attains a steady state and the extra fluid in the larger droplets just increases the evaporation time. This phenomena was not observed in porous solids because the evaporation process never quite attains a steady state. Figure 3.19 shows the parameter 3 plotted against the initial solid surface temperature. This parameter is defined as the ratio of maximum droplet diameter during evaporation to the droplet diameter before impact. The diameter of the droplet on the surface was determined from the measured maximum wetted area. The droplets after impact had a circular shape during the evaporation process except for the non-porous solid at temperatures greater than 164’6. At these temperatures the droplets explode due to the intensive vapor formation between the droplet and the solid surface. As the surface temperature increases, B tends to increase due to the decrease in surface tension. B for the non-porous solid was larger than that for the porous solid due to migration of a part of the droplet into the porous matrix. 3.4.2. Heat Transfer During Droplet Evaporation on the Ron-Porous Solid During the experiments, the evaporation phenomena was recorded by a video camera. The droplet evaporation behavior changed with the initial 59 solid surface temperatures (Ts). For 75 5 T8 5 125°C, upon impact, the droplet have a disk configuration, and direct contact between the droplet and the solid exists. The evaporation takes place from the top and the sides of the droplet. The droplet diameter on the surface remains constant for up to approximately 80 percent of the evaporation time, and then starts to shrink. At Ts - 125°C, after the initial contact, nucleate boiling was visually observed (see Figure 3.20; note that the equilibrium temperature is above the boiling point of water). As Ts increases (i.e. 150 5 T8 5 175°C), the boiling became more dynamic, and the extensive formation of vapor sometimes scatters the droplet. Seki et a1. (1978) reported similar droplet behavior at these temperatures. For Ts > 200°C, the droplets spread very fast on the hot surface, forming a thin water film. The evaporation time at this temperatures was very short. Individual traces of the surface and indepth temperatures for a 51 pL droplet impacting on the solid at T8 - 125°C are shown in Figures 3.20 and 3.21. As discussed in section 3.4.1, the hot surface immediately cooled to the contact temperature upon contact with the droplet. The contact temperature was evaluated at the temperatures studies using equation 3.1. This temperature was compared to the experimental values. For T8 < 164°C, both temperatures were in excellent agreement. For Ts > 164°C, the measured contact temperature is approximately constant (slightly above the boiling point) and is lower than the theoretical contact temperature. The theoretical and experimental contact temperatures for the range studied are shown in 60 9535 03.2. Joan—0.8 A: an a no souuouoegu on» I? ”MAO” gOHOm :32 < ooh sensuouonaoa ooomusm owé 35mg Aoomv oEc. o —... F . 0.0 °_B O—W . .0.” . o. P. . o _. Ilom 6:... 328.8: a 8.. 82.82 3§o8§o§ as n... . p . 258.52... . .. S n ..om u. D 3 u 9 II.— roo F 8 I. w r d m ..o. . m. a m u 9 D p c - s It .1 4 WI! IONF \d I) l O N . ( on... 61 60398 A: a... a mo scavenger» on“. 95.39 3.3. Er 63o» osouoméoz a you nouguuooaoa snoop—nu .36 83»: A03. oEc. om. o... o.» 0.. o... f - . . on as a... z . €385.25 .. 55...... o. . £885.... .5... ....» a . 933853.... low E... a; a . 2&08585. cue...” . . 2:88:55. .. ..oo . fl 4)., ‘1! u no . . , o .. O~ jONF 11 .. UH on... (30) semimadwel mdepu' 62 Figure 3.22. Seki et a1. (1978) found that on stainless steel, the measured contact temperature was in agreement with the theoretical contact temperature up to 180°C. For higher temperatures the measured contact temperature was lower than the theoretical temperature. Figure 3.20 shows that after achieving the contact temperature the solid surface approaches the ”equilibrium temperature," (i.e. interface temperature during evaporation). This equilibrium temperature remains nearly constant until a critical droplet thickness (6*) is reached on the hot surface. Then, the surface temperature starts to decrease sharply until the droplet has completely evaporated. This sharp decrease in the surface temperature implies an increase in the evaporation rate. As the droplet thickness reduces, the heat transfer through the droplet increases allowing more evaporation to take place from the top surface of the droplet. The time for evaporation after the droplet has achieved the critical thickness is longer for larger droplet at a specified initial surface temperature. This is because larger droplets spread more on the surface, and this large volume require longer time to evaporate. As the initial surface temperature increases this time decreases due to higher heat flux. After the droplet has' completely evaporated, the solid recovers smoothly to its initial surface temperature. For T8 2 200°C, the interface temperature (which is equal to the contact temperature) was a few degrees above the boiling temperature of water. Since, the droplet diameter after impact on the surface for this temperature range was very large (see Figure 3.29) the evaporation times were therefore very short. After evaporation, the surface recovers to its initial temperature. 63 Prior to droplet impact, the measured temperatures were compared with steady state calculations. The agreement was within 2°C. This temperature distribution was used to calculate the heat flux inside the solid which was then evaluated at the solid surface. From the surface heat flux, the overall heat transfer coefficient was determined according to equation (3.2), the convective heat transfer was determined according to equation (3.3) q” - ho (Ts - To) (3.2) q" - h (Ts - T”) + co (T: - T2) (3.3) Figure 3.23 shows the heat transfer coefficients at the surface of the ceramic block versus surface temperature. Surface and indepth temperature measurements during droplet evaporation show that the isotherms and the heat flux lines inside the solid matrix may be approximated by oblate spheriodal coordinate system. In such a coordinate system a temperature distribution [adapted from [Keltner (1973)] is given by: 2(T -T ) 2 T no.8 use: 3332—26 gauge 3..” 9.53... EL 838an2. montam 26m 625 on... can emu own. om. 2... on . p . . . . lo 0 a a O 1 m . w 4 O rOON D O O D . D d r.00... . . o D d u o p o r.80 o > 4 o r u a. ..u o. o o ._ a. a .. 4 q o .1 on a com o .3 Q. o . o .3. .n p coop (aw/MM) an meH eAgoJodo/xg efioJeAv 67 £25.25 noun—9.8 26.32, you 333330.... 333» v.30» 1.3.2: -> “3.8 .3: 2,3309% omwuo>< mm. n 0.3»; Gov 230..an8. mootam 26m 62:. _ V on» can emu 8m . 09 8. on m r p n p — . — . — filo J o a a r ,m a r 9 o ”.00.. 3 ... m o NOON d o a . m D noon m... .. A. d i. 9 o -oo¢ n o n .W 4 noon m. > p n H o p a o .Sopouooon a o 41 3 4 .. X o m. .3 on a woos m1 ..u 2 o. ... w P .. B p fl.oom W . .z S ( 68 Jog—30> uoaaoua 28.33, . you 3330960.“. commas» 33m #335 -> 3:. 603.9235 36 as»; Aoov BBEmQEm... mootaw 2.0m BEE can com com com om. cm: on [illicit-4m . . o a a wow r 89 239.3th 8826 2.0m .025 N can . own L can . Ra . om“ . omu P a? . aged 0 o How . ION— o o o 0 Had «A... N “nom— o a a a 1.6 .m < f . a a a . 1 w 52 D N O [0.0 W. t u w a r n .. n a a .3 m. fi3N ‘ a ..a o. o o #0.. \sl IO¢N i 2 « 1 x o u an ... . r2 ( 1.2m 4: 3. . Y r t. l> o o F) (098) awu uogoJodo/g 69 ...-5.25 uni—93 33...; won 03532—18. 35.3» 3.?» H335 u> as: .9263... mm... 353 Aoov 9.30..an0... mootam 2.0m .03.... own 000 com DON OMF 00—. on III}; . . _ . o . a o o m m 0 . d 0 F00 D d r % m -8. 3 8b SEEKS: 823m 2.0m .02.... o m o...” a.» . om - a? , emu . omu . 3a . om. . mso D .. m 0 G o 1 0 Q TOWP M a m m ... . .. p . u... o u u p q we w p o n o v o .l 9 n 4 1.- m 0 com ) a . s 1 o 1 u. 3 ”.3 M" u ém 10mm mw .3 on a . m o . 41 fl? 0 o o. ...-COW .3 B o . . 2 b 70 Te respectively are the solid surface temperatures before and after the step change, and 8 is the penetration depth. This is the solution for a step change in temperature over a disk-shaped area on the surface of a half-space. Using this equation, temperatures were evaluated at various locations and at different times. The results were in fair agreement with the measured values. From equation (3.4) it can be shown that the rate of heat arriving at the surface (Q), is given by the following equation I. Q - 4 R K (Te - T8) (1 + 6) . (3.5) The instantaneous average evaporative heat flux is then determined by dividing by the instantaneous wetted area. The instantaneous evaporation rate may also be determined from Q by dividing it by the heat of evaporation. The average evaporative heat flux and the average evaporative mass flux were correlated with the initial solid surface temperature as shown in Figures 3.24 and 3.25. Both quantities follow the same pattern; the higher the initial surface temperature and the smaller the draplet, the higher is the average evaporative heat flux and the average evaporative mass flux. This is because at higher temperatures, more energy is available for droplet evaporation, and small droplets have larger conductance due to their smaller thickness on the hot surface. Figures 3.24 and 3.25 also show a slight decrease in both quantities for T8 > 300°C. The evaporation time and the recovery time curves (Figures 3.26 and 3.27) also show a slight increase at the same temperature. Since, 71 this temperature corresponds to the maximum average evaporative heat flux and the minimum evaporation time, it is likely the beginning of the transition regime. Unfortunately, the experiments could not be continued for higher temperatures, (i.e. in the transition regime), due to the burn out of the heating mantle; The non-dimensional maximum volume of influence (Vi/V) is shown in Figure 3.28. This quantifies the cooling effect of a single droplet. Figure 3.28 shows that the solid at T8 5 lOO‘C cools more with a 10 pL droplet than with larger droplets. This is because the evaporation process for larger droplets attains a steady state and the extra fluid in the larger droplets just increases the evaporation time. Thus, the smaller droplets are more efficient for cooling materials at this range of temperatures. Figure 3.29 shows the parameter B plotted against the initial solid surface temperature. This parameter is defined as the ratio of maximum droplet diameter during evaporation to the droplet diameter before impact. For T8 2 200 °C, B is quite large, since the droplets spread fast and form a thin water film on the hot surface. For this range B decreases as T8 increases. This is because the droplet is evaporating very fast, and does not have enough time to spread on the hot surface. 72 own .3552, unack— naoauu> pom ouauuonaou. accuser. v.38 iguana .9 00:05.35 mo 25.3., .555: announces—3&3“ 36 936...... A0... 8320an... 83:6 2.0m .03.... 8... SN 0.8 on. 8. on T . . _ . _ r _ . b . . om m m m s a -06. m 4 m e ( w m m 10.0? a o . » Q 10.... D t ..1 o— o > m 06— D .II ..1 3 4 m o ..m on a 4 a .. .. n... o I . ..1 .n p o 0 CNN (AW-A) eouenum ;o ewmoA °xow louogsuewgg—UON 73 .3533 0030.8 30.725 you 33300309 003.50 030m ~03qu 0> a 302.30.. 36 0.33.. A0... 0.30.0083. mootam 2.0m .05.... own . own . OMNL owu . om? p Wm. . 0.0 . so... 0 u . m a m p 10.. m ..o.m we . w m m -3 .w. 1? o w . m w a m M Tod 8 m ._a o. 0 . .... a. 4 c o .... on 0 106 a m .3. n... 0 .. P D ...: PD 9 CHAPTER 4 DECOMPOSITION OF WOOD In Chapter 2 it was noted that, prior to ignition investigation, it is important to understand the response of the solid phase to heat, and the resulting production of decomposition products. In this work, products generated when a wood sample heated under controlled conditions are examined in a systematic set of experiments. The set includes 12 experiments, performed at 4 different heat fluxes (approximately 1,2,3, and 4 w/cmz), and at 3 different moisture contents (dry, 9%, and 17%). Conditions for these experiments were suggested by the specific application of interest. In room fires, the wood burning process may be treated as transient heating and decomposition of a semi-infinite wood sample. The moisture content of wood varies and is influenced by the climate, season and location. The heat flux incident on the wood samples also varies with time and their location relative to an already existing flame in a room fire. The results of the decomposition experiments are presented in this chapter. Section 4.1 describes the physical aspects of the problem and discusses the importance of the moisture content and its relation to the overall decomposition problem. Section 4.2 reviews the previous 74 75 literature. Section 4.3 outlines the procedure used to analyze all the measured species. Section 4.4 includes results and discussion. 4.1 Background Decomposition of wood refers to the process that produces various chemical species and a residual charred solid upon heating. The physical properties of wood are highly dependent upon its microscopic structure. Wood is predominantly made up of a lignocellulosic structure with various infiltrated substances. It is composed of three major constituents: (i) cellulose (50% by wt), (ii) hemicellulose (25%), and lignin (25%) [Stamm (1964)]. The physical aspects of wood decomposition are associated with heat and mass transfer. Consider a semi-infinite wood sample initially at ambient temperature, subjected to a prescribed heat flux. The sample heats up by pure transient conduction, in which moisture starts to evaporate and an evaporation zone begins to travel into the solid. When the sample surface layer becomes sufficiently hot, it starts to decompose. This decomposition zone moves into the interior of the solid. Consequently, the products of decomposition are driven to the surface of the sample, and flow upward due to buoyancy and mix with the surrounding air. As the interior of the sample becomes hotter the decomposition zone penetrates deeper into the virgin solid, leaving behind a thermally insulating layer of char. The rate at which the decomposition products and water vapor are generated depends on the response of the solid-phase to the applied heat flux. As pointed out by Simms and Law (1967) and Lee et al. 76 (1974), the presence of moisture in the solid phase changes its thermophysical properties. This will influence the heat and mass transfer process because of its large heat of vaporization. This heat of vaporization depends upon how the water is held inside the solid matrix. There are four ways by which water can be held in wood: (i) Free or absorbed water: this water is mechanically held in the capillary structure as a result of surface tension forces. Its quantity is limited by the porosity of the wood. The energy required to evaporate this water is only slightly greater than its latent heat. (ii) Bound or adsorbed water: this water is held by hydrogen bonds to the cell walls. Its quantity is limited to 30 percent of the oven-dry weight of the wood. The energy required to evaporate this water increases as moisture content decreases. At the fiber saturation point (~30 percent moisture content) the energy required is the same as for absorbed water. (iii) Water vapor: this is present in the air filling the cell cavities. The quantity of this vapor is normally a very small fraction of the total moisture content. (iv) Water of constitution: this is a part of the molecular structure of cell walls and it is firmly held by chemical bonds. In reality it is not water at all, and is only released upon thermal degradation of wood. A.2 Literature Review Many investigators have studied decomposition of wood and determined the various products. Schwenker and Beck (1963) detected 37 - volatile compounds, and they identified only 18 of them. Coos (1952) 77 listed 213 compounds which have been identified as products of wood decomposition. Martin (1965), investigated the decomposition of pure cellulose in helium. He reported a time-dependent analysis of permanent gases and tar. He indicated that there are at least two fundamentally different ways in which cellulose thermally decomposes; one supplies the bulk of the volatile fuel which supports the flaming combustion of the material, the other produces mainly water and the oxides of carbon. Vovelle et al. (1986) studied the decomposition from PMMA and particle board subjected to a variable radiant heat flux under an inert atmosphere. They indicated that the mass loss rate is directly proportional to the instantaneous value of the heat flux absorbed by the surface of the sample. Lee et al. (1976), investigated wood decomposition in air at two heat fluxes (3.2 and 8.4 w/cmz). They indicated that the decomposition process is dependent upon the external heating rate, the total time of heating and the anisotropic properties of wood. Lee and Diehl (1981) investigated the effect of absorbed water (samples with 50% of oven-dry weight) on the decomposition of wood. They indicated that moisture not only changes the solid-phase thermal properties and delays the decomposition process, but also substantially dilutes the decomposition products. Atreya (1983) , investigated wood decomposition experimentally and theoretically. He performed numerous experiments for 10 different kinds of wood in air, and reported transient measurements of surface temperature and evolved mass flux. In his theoretical investigation he 78 included energetics and kinetics of moisture desorption in his model. He indicated that the desorption of moisture, which has been ignored by previous investigators, has a considerable effect on the energetics of the decomposition process. Kashiwagi et a1. (1987), studied the products generated from wood decomposition in 3 atmospheres of different oxygen concentrations (N2, 10% 0 - 90% N 'and air). They indicated that ambient oxygen 2 2’ significantly increases the desorption mass flux and surface temperature. They showed that decomposition rates in air were nearly double those in nitrogen, and surface temperature in air increased as much as 200°C over that in nitrogen at 4 w/cmz, due to oxidation of char. 4.3 Products Analysis It is important to know the production and the depletion of major species during decomposition. The equipment used for continuous measurements of C02, C0, 02, H20 and total hydrocarbons (THC), as well as the necessary time lag and response time corrections, are described in Chapter 2. As a result, the mole fractions of all measured species were obtained as a function of time. Since the instrument used to measure CO require a dry sample stream, the mole fraction of CO was on a dry basis. This value was converted to a wet basis using the measured mole fraction of H20. The following equation was used: mole fraction on wet basis - measured mole fraction on dry basis (4.1) x (l-mole fraction of water removed by the drier) 79 The evolved mass fluxes of all the measured species were then determined using the total mass flow rate of air. The total mass flow rate of air includes the air flow through the critical nozzles (equation B.l in Appendix B), and leakages from the gap around the sample and top of the tunnel. As explained in Section 2.2.2.3, these leakages were determined by the methane trace method, and they had been calculated from the difference in the methane concentration measurements upstream and down- stream of the tunnel. The instantaneous 02 depletion was determined by subtracting the initial steady state value of 02 from the measured value at any time. The sample mass flux was calculated by dividing the time derivative of the measured transient weight-loss by the initial front surface area of the sample. Although it was not possible to measure the organic condensibles (tar), its instantaneous value was determined by difference [(sample mass flux + 02 depletion) - (sum of the mass evolution rates of C02, CO, H20 and THC)]. Finally, the mass fluxes of all the measured evolved species were normalized by the sum of sample mass flux and O2 depletion to obtain the percent mass flux. 4.4 Results and.Discussion This section describes the results obtained from the set of systematic experiments done on Douglas fir. In these experiments as described in Chapter 2, simultaneous measurements of weight loss, surface, bottom, and indepth temperatures, oxygen depletion, production of C02, C0, total hydrocarbons (THC), and water were made. The 80 experiments were parameterized with the sample moisture content and the external radiation flux. All experiments were performed in air under non-flaming conditions. A duration of 30 minutes was assumed to be long enough to reveal enough information about the decomposition process. For all three moisture contents, spontaneous ignition occurred at 4 W/cmz, and data was continuously recorded even after ignition. 4.4.1 Sample.lass Flux Mass flux histories for dry, 9%, and 17% moisture contents at 1,2,3, and 4 W/cm2 are shown in Figures 4.1, 4.3, and 4.5 respectively. These figures show that higher evolved mass flux corresponds to higher heat flux. The general trend for mass flux at high heat fluxes (4, 3, and 2 W/cmz) is rise to a peak, fall-off and then attaining a nearly quasi- steady state value. Wichman and Atreya (1987) have developed a simplified model for the decomposition of charring materials in an inert atmosphere. They predict a fall-off in mass flux proportional to the negative one-half power of time. Figures 4.2, 4.4, and 4.6 show the mass flux histories plotted on a log-log basis. The fall-off rule agrees only for some cases and for only a short period of time. For example, the curve for the dry case at 4 W/cm2 follows such a time-dependence from 100-240 seconds. This disagreement may occur because of the effect of ambient oxygen concentration on the decomposition process. The behavior of evolved mass flux in the presence of oxygen is expected to be different than in an inert atmosphere, because oxygen not only causes a strongly exothermic reaction at the sample surface, but it could also diffuse MASS FLUX (mg/cm2 3) MASS FLUX (mg/cm2 8) 81 1.4 . 1.2-1 1.0-1 1.. ‘ . Z -, a W/cm2 3, 3', III .0 :0 q. 0 0 . ’9 0.8-‘5. \ '3 ", if. .‘5 315- ":\ ° 2 ' .‘g‘g' 7‘fl€?§ 1‘... o,’ '3' ‘1‘, 5" O.6-: ' °' ' - 'r' a", . : . . 3 w/Cllz . : 3’0 0". . 04-3 .."-- ’3. 1"}. ,.~’;.' '1-513 2. : ":2: WW“ ' ' 002-:‘05 Mm ‘ '3’ t 1 w/cm2 . A ... 0.0 It"!!! If! llrviil IT! 733 III 0 200 400 600 800 1000 1200 1400 1600 1800 TIME (s) ‘ Figure 4.1 Effect of Incident Heat Flux on Time-Dependent Haas Flux for.Dry Condition. 1.00 0..“ .....'0~.. (O w/sz 5"- o.80-E \/\t"m‘{\’?§‘{1"‘a 0.60— ,5 . n....r*""”‘~\“"~\.s}}{5% 0.40- ° “ W 020- 2 ulcmz 0.101 0.081 0.06: 0.04- It .. ’ fw\w‘W-;‘3M§ 0.02— . m2 0001 U U r '2'; I I A ltwlc I I I TU' 10 50 '1' O 500 1000 TIME (s) Figure 4.2 Effect of Incident Heat Flux on Time-Dependent Hess Flux for Dry Condition MASS FLUX (mg/cm2 3) MASS FLUX- (mg/cm’ 8) 82 1.0 0.8- ;; 4 U/cnz 05': - NW w‘w‘. "WWW-.4” «JAM :0. \x: vxflwflhlxflw’ 3"“ I 00‘: o... 3 W/sz 'W .. y 1' b# 1 WICIIIZ .— .;_-'A—-:“ A L4— __ —— .. A TAvA_M_A_.__’ 000'000l101'00000' '0' '0' '1‘ '0' 0" 200 400 600 800 1000 1200 1400 1st 1800 TIME (s) Figure 4.3 Effect of Incident Bent Flux on Tine-Dependent Hus Flux for 112 Moisture Content Condition 1.00 008° ... ‘0. 4 "/m2 0060'“ .. . LWAWVW' 0-40- .° ..-°---:" 3 w/ «X3 va‘W‘J” . 3 ' ° 3 0 ::°.. WW 0.20-a I 1 0.101‘ 0.084 --.-.... 1 111/an .1 ,. ' 0008- ”‘11.. ~11» . .‘ .. «I- . :‘Afgfiwfi 0004 l T l I I I I i g ‘ I U I I 'l 10 50 100 500 1000 TIME (s) Figure 4. 4 Effect of Incident Heat Flux on Time-Dependent Mace Flux for 112 Moisture Content Condition ‘ MASS FLUX (mg/cm2 3) MASS FLUX (mg/cm”1 s) 83 0"] 000 1200 1400 16'00 1800 0.0 IIIIIIFIII'III 0"l000l000l000 0 200 600 300 TIME (s) Figure 4.5 Effect of Incident Bent Flux on Tine-Dependent Mess Flux for 172 .Ioieture Content Condition 1 000 '- 0.80% 0.60- 0.40- 0.20- 0.10: 0.08- 0.061 0.04-- 0.02 I .' U-IYIIUI I I I III! 10 50 100 500 1000 TIME (s) Figure 5.6 Effect of Incident Heat Flux on Time-Dependent Mass Flux for 172 Moisture Content Condition 84 through the porous char to the decomposition front of virgin wood. Thus, to theoretically investigate wood decomposition in air, it is important to include the heat released by char oxidation in such a model. From Figures 4.1-4.6, we note that the mass flux curves do not start. at the origin. This is because the mass flux is an average quantity, determined from the time derivative of the measured mass loss. Since the initial mass evolution from the samples is essentially pure water, the starting point is higher for higher moisture content. The desorption of water from the solid matrix is quickly followed by the evolution of a large amount of tar. This increases the mass flux, and it continues to increase until a thin char layer forms. In this heating stage, the mass flux attains its maximum value. .As the sample heating rate increases at higher heat fluxes, the heat-up time is decreased. Thus, for higher heat fluxes the mass flux peak occurs earlier. Vovelle et al. (1984), observed that this peak increases linearly with heat flux. The production of volatile gases, which escape from the solid, leaves behind a layer of low thermal conductivity char. Formation of the char increases both the absorptivity and emissivity of the surface. But since the thermal conductivity of the char is much lower than that of wood, the net heat transfer through the char is reduced but the net heat loss by radiation is increased. The oxygen in the ambient air also reacts with the char in a highly exothermic reaction and supplies additional heat to the solid. This heat compensates for the reduction of the net heat transfer through the solid due to the formation of char. Furthermore, the char cracks permitting a larger quantity of volatiles 85 to escape. A combination of these factors is believed to produce the quasi-steady value attained by the decomposition rate after attaining the peak. 4.4.2 Sample Temperatures Measurements of surface, bottom, and indepth temperature of the sample were also made during the experiments. These measurements contain the history of heating of the solid matrix, and they reflect the thermal response of the solid to the irradative heat flux. For example, surface temperature measurement is very important with regard to the heat generated by char oxidation. These reactions take place primarily in a thin surface layer, because their rates strongly depend on temperature and the presence of oxygen. Thus, surface temperature is very useful in determining when the heat of char oxidation is being released as well as at the rate that this heat is transferred to the surroundings. Surface temperature profiles for dry and 17% moisture content at 1,2,3 W/cm2 are shown in Figure 4.7. In this figure higher surface temperature corresponds to higher heat flux. The rate of rise for these temperatures is determined by the moisture content and the irradiative heat flux. This rate increases with higher heat flux and lower moisture content. As an example for the dry case, it took 240 seconds exposure 2 for the surface temperature to reach 300°C and Only 16 at l W/cm seconds exposure at 3 W/cmz, whereas for 17% moisture content it took 1600 sec to reach the same temperature at 1 W/cm2 (see Figure 4.7). 86 Char oxidation occurs when the surface temperature is greater than 400°C [Kashiwagi (1987)]. This reaction is strongly exothermic, and is expected to rapidly accelerate the local temperature rise. Figures 4.7 and 4.13 show that, when surface temperature suddenly increases, a significant increase in oxygen depletion is noticed at the same time, as expected. Figure 4.7 shows a sharp rise in-surface temperature occurs for the 17% moisture content case at 245 seconds. Simultaneously oxygen depletion increases significantly at the same time (see Figure 4.13). Thus an inflection point in the surface temperature at a value greater than 400°C indicates the start of char oxidation. Figure 4.7 shows that the start of the oxidation time increases as incident heat flux decreases, and decreases as moisture content decreases at the same incident heat flux. For example, char oxidation 2 occurs for the dry case at 550 seconds when exposed to 2 W/cm , and at 215 seconds for 3 W/cmz. For 17% moisture content the oxidation occurs at 1050 seconds for 2 W/cm2 and at 245 seconds at 3 chm2. This is because the char develops more quickly at higher heat fluxes and lower moisture contents. The effect of moisture is expected to be less significant near the heated surface after char oxidation starts, because for this region the temperature profile is primarily determined by the exothermic heat generated from oxidation of carbon. Temperature profiles at various locations in the sample 2 corresponding to 9% moisture content at 2 W/cm and 17% moisture content at 3 W/cm2 are shown in Figures 4.8 and 4.9, respectively. The indepth temperature rose slightly above the boiling point of water, and remained constant for some time before continuing its rise. This is due to the .xsdm panacea acovaoca naoaun> no unsunuunaoe ooeuusm so usoucoo unsumaoz mo uouuum u.e «woman Amy .2: com. com. 00¢. com. 000. com com 00... DON bub—DIPP_-P—Pbb|— o >mo --.. 2 R up I 87 I L 00000000000000.0000 0.00 C 8 00“.. 00 M ¢ . nub PEI. DP. —-P-bPPL—o nae-fir NEO\3 .— I} 1111' ‘3'“) . 111 .............................:..... NEO\3 H a 4...... wao\3 ‘swts 5 «an; ~\‘\~30\3 n \ 00.10”... . .0113 s. ... a N «J . ‘fi .. EO\3 m oo_. (0.) ElaanaZ-IdWZ-Il BOVJEIDS 88 000m 0 .souuuvsoo acoueoo ousueqoz.no you Nao\3 N as «Haanm on» c. enoaueooa esouue> an «8.9 nouo> oueunuonsoa m.< ouswum 00.. 00.0 3 Us: .00.: 003 000. 000 000 00¢ 00m 0 up. bun—b-n—nnn—ppb—ppn—-p-PP—o 1.111111111111111 3an .838 - 1 aacu ooouuam (0.) aamvaadwai BTdWVS 89 .eoauaveoo usousoo «asunder as. you «aux: n on «Haaam on» e. «coaueooa esoaue> on «sub usouo> ouaunuoaaoa a. e shaman 3 m2: 000.. 000—. 00¢. 00m. 000. 000 000 00¢ 00m 0 DD. :DDDBLInbb-D-L-Ir. b-n b-P —-DL_0PLI—o .3... 333. , I 1. \ 1 En... 00— Hoom m noon m H00¢ moon 8.2%. .....000 aam.~ u 005 (0.) aamvaadwai 31dwvs 90 endothermic vaporization of the adsorbed water. This plateau in temperature-time curves became longer at greater depths in the sample and at higher moisture contents. The upward inflection of the indepth temperatures at 600 seconds in Figure 4.8, is caused by the heat generated from char oxidation (notice the sudden increase in the surface temperature at this time). For different moisture contents it is difficult to compare local indepth temperatures, since thermocouples are not at the same locations for all cases. 4.4.3 Decomposition Products In this study, oxygen depletion, production of 002, CO, total hydrocarbons (THC), and water are closely examined in order to quantify the effect of the adsorbed water on the decomposition process. The data are presented as mass flux of the permanent gases, and as percent of mass flux. Plots for different moisture contents at different heat fluxes are shown in Figures 4.10-4.17. The water has a similar trend at 2 and 3_W/cm2. An initial mass flux peak arises from the desorption water in the solid matrix, followed by a rise to another peak, and eventually a gradual fall-off in the water production. This latter water may include the water of constitution. The water curve at 1 W/cm2 shows only one peak (see Figures 4.16 and 4.17). The quantity of water produced is generally larger for larger heat fluxes and higher moisture contents. The water production is quickly followed by a large amount of tar whose quantity was determined by difference, as mentioned previously. This organic condensate has an extremely complex in composition. MASS FLUX (mg/cm” 3) PRODUCT AS PERCENT OF MASS FLUX 0.24 0.16 0.08 0.00 91 320 o 10“ 0. “H . .. 0‘ 02 DE? . .°'0'o‘o. m c ‘ .1 0. 0.1 002 “-"JV‘L: . ...-«"3.» .1...- __ «9'7: 0 0, CO " ’5”: ' ° " ° . .. m n'. fig?‘ I 0 200 400 600 800 1000 1200 1400 1600 1800 TIME (s) Figure 4. 10 Time-Dependent Mass Flux of Evolved Products at z ale-2 for 9: Moisture Content Condition 100 80- q 60 " 320 O‘." 40 I [d 0‘ “I "1 a 1'- a 1‘. TAR \ 0 .‘0 - CO .e 2 8 ‘ ‘ .2”-’.o..0'-r,w .' ~04“..-..._‘ Lou—....- quY~ 3119' CO # “...,”W’. 7 T110 0 I 0" '0' '0' '0' ll. UIUIUIUIIIU 0 200 400 600 800 1000 1200 1400 1600 1800 TIME (s) Figure 4.11 Time-Dependent Product as Percent of Mass Flux at 2 1J/c1n2 for 92 moisture Content Condition. MASS FLUX (mg/cm2 3) MASS FLUX (mg/cm’ 3) -OoM'III'III'TIr'III II III rII rII'Ifi 0 200 400 600 800 1000 1200 1400 16001800 TIME (s) Figure 4. 12 Time-Dependent Hess Flux of Evolved Products at 3 11]an for 9: Hoieture Content Condition. 0.32 320 0.28 0.24 -‘ 1 0 . g a 1 , 0.20 . 1 o ‘0 '.\ 0 1 o " ~°‘TAR 0.16 .0 \o 0 ‘\. ‘ .... ...‘ 02 DEPLETION . ..°‘O..I‘.‘ "s. . ‘9 I Oe12 . 3 co . ‘0 \ 0| . 0 as ' ' x ' ‘-~"‘""": .331. ~ I A . ......--........~.' - -. 0.04 ., I ' VAT 4 fr” 0.00 :“ -OoO‘IIVI'III'III'III IIiIIfiI III'ITI'II 0 200 400 600 800 10001200 140016001800 TIME (s) 2 Figure 4.13 Time-Dependent Mass Flux of Evolved Products at 3 H/cm for 172 Moisture Content Condition. 93 1.2-3 .1 A I I .0) . 1 1'4 '1 E a - 1. 01 0.8: 00,1. co2 -. .l p 0“ 1'. “A 0' ‘m 's .0-- 3 2 . ‘\ 1"" ...11. : ' “ “x. x 016': '1 0; nmmon D m .‘ . ..J . '1 ' h- 0.4- ; .. ‘ ' I . 1 g : : . : \.. . ... maz . 0" :0 s I, 1 2 0.2:: : ‘3 so 1“”...0.“ “. '0': as“: .. 1‘: . .“.‘ - ' ' co and mo ’. Crc' .,' ,. '.' ‘ ' 0.0 ’ “" ‘ 0 200 400 600 800 1000 1200 1400 TIME (s) Figure 4.14 Time-Dependent Mass Flux of Evolved Products at 4 chmz for Dry Condition. 1.0 .I r-\ i m 1 n 0.8-1 E .1 O '1 g b. 0,1 E ' 4 t. y l X 1 . . “2° 3 0.4-1 \ - . '3. '- '0' '4 fl 'o."‘V Vvv‘vv 12:” la. ' J 1 1 0‘ I. c02:.: \. ,‘M.’ *Vu .. '0.."\.:.'~ m - . 0‘ ~ .. 1" '. '. ... ..W". 0..., 0‘130': ‘0 0.2-1 Ig‘.o"'o' ' 'Q . 1' T“ o 1' . < .I 0 e. 2 .4 , m 0.0 ,g 1‘ .14-m _r00 and THC , 0 200 400 600 800 1000 1200 1400 1600 1800 TIME (s) Figure 4.15 Time-Dependent Mass Flux of Evolved Products at 4 W/cm2 for 91 Moisture Content Condition. 94 MASS FLUX (mg/cm” 3) 0.040 A 0.032 n 0 0.024 I I I a I , ’0: m ' '5“. O 0‘ 0.016 . e 1"...1.‘ 1.. .l' :0"; ".0 I ' '1' . . 1 -",.1‘."1'1. 1- ,{L'°°',: 5.11"." "' 7.1111111." '.' -: ':. ' . °I E 0008‘ . .1 1 THC - . , '1'. .. ,- 02 flananou d o ' rrmm' 1““ Hui '1 "‘7 0.000 .. ’. Jill ‘M‘.. I s 0 200 400 600 800 1000 1200 1400 1600 1800 TIME (s) Figure 4.16 Time-Dependent Mhss Flux of Evolved Products at 1 Wen2 for Dry Condition 0.08 A 2 m N .. E 0.06- 0 .. 320 E 1 E " . V 0.04- :, v, . >< ,1 ,."11An . ‘ s D " 1 ' 1' ’1 ‘1 ‘0 '1 d .I c I .I. . fl .0 3... ".I II ' I. ' ' . g. 1"." H" '4’;'1 III (I) 0.02 ° I. ' ‘I'? a ‘1'”.1 I" ‘ III I I Q I. ”I...” :.'I ' I ... b I .: : 2 A '0‘ '. 1 . . ° II DEPLET 0N 21g. .fll HALL“. -1- 2 $691,311. 0.00 - -1«— ' 200 400 600 800 10001200140016001800 TIME (s) Figure 4.17 Time-Dependent Mass Flux of Evolved Products at 1 II/cm2 for 171 Moisture Content Condition. 95 Kashiwagi et a1. (1987) investigated its composition by means of capillary gas chromatography coupled with mass spectroscopy. They were able to pass only 20% by weight through the gas chromatograph. They identify more then 40 different species. The production of combustible gases (C0 and THC) at 2 and 3 W/cm2 shows similar trends; gradual increase until attaining a steady value. For both gases this steady state value is slightly higher for higher heat fluxes and lower moisture contents. But a major difference is the time it took to attain this value. For example, for the 9% moisture content case the steady state value was attained after 950 seconds at 2W/cm2 (see Figure 4.10), 200 seconds at 3 W/cm2 (see Figure 4.12), whereas for the case with 17% moisture content it took 400 seconds at 3 W/cm2 (see Figure 4.13). At 1 U/cmz, there were only a small amount of combustible gases produced for the dry case (see Fig. 4.16), and a not noticabe amount for the 17% moisture content (see Figure 4.17). Thus the presence of moisture significantly affects the production of combustible gases by both delaying and decreasing their production. The oxygen depletion is the amount of oxygen used to react with carbon to form carbon oxides. As mentioned earlier in Section 4.4.2, the time when a significant increase in oxygen depletion occurs corresponds to the time of an inflection point in surface temperature. Generally, the higher the heat flux and lower the moisture content, the larger the oxygen depletion and C02 production. The mass flux of all the measured species was time-integrated for all three moisture contents and for l, 2, and 3 W/cmz. Since 96 spontaneous ignition occurs at 4 W/cm2 for all cases, this set of experiments will be discussed separately later in this section. The effect of the heat flux and the moisture content on the integrated mass products are shown in Figures 4.18 - 4.21. These Figures show that a larger amount of water production corresponds to higher heat flux and higher moisture content. Also larger oxygen depletion and 602 production correspond to higher heat flux and lower moisture content. Finally, the production of combustible gases is larger for high heat fluxes and lower moisture contents, as expected. From these results, it is clear that the presence of adsorbed water in the solid matrix reduces the average fuel rate evolved during the decomposition process, by delaying their production. At 4 W/cmz, and for all moisture content cases (dry, 9% and 17% moisture content), spontaneous ignition occurred. In reality, these experiments belong to the discussion on ignition, but since they were performed in the absence of a pilot flame like the rest of the decomposition experiments, discussion of these result is presented here. The general trend of the measured species in the three cases is very similar (see Figures 4.14 and 4.15). At time of ignition there is a sudden increase to a sharp peak, then fall-off, and finally attainment of a steady state value. The sharp peak is caused by the large amount of fuel available in the gas phase at time of ignition. From visual observation, the flame then reduces in size, because it is now fed by the fuel evolved from the solid phase. As noticed previously in the decomposition experiments, the fuel produced attained a steady state value even with the char formation. This fuel with a constant value of 97 120 110’ ‘ 100— ‘ . / mc 90— {I -1 A. // CO 5 80- If”, -. 8 ”,1. c. D - 8 7°” l// / 2 n: gl/ 1:. / / m 50.— ’18 02 .- c: I ‘/ 1’ A a: / / ca 50.. .— 1 [1’ 97 m / 5 4°” /4 J ‘ c: / / a’ “ /f/ z’ 30— // / 2 ’ 52 a 0 '- /// 2 20 ' IA?/,I “ ,W .. 29 ’ - 10 5* 77 1 2 3 Figure 4.18 Time-Integrated Product Mass and Composition (shown as percentages next to each bar) as Function of Incident Heat Flux for 9! Moisture Content Condition. 98 120 110 '1 "’ 100 ’3’ 90 ' T 3 {-1 80 " 0 O D c: ‘_, 2 70 - 9 P- m 4‘ \ \\ m — o 60 _ 24 \\9~I_-~ .. 3 P‘ ~ ~11--.— 50 - 18 - j H'- — _l.£ ...-D" 31 - S 40 - 21 _. {-1 y’u' r, 30 " 47 21 ~~ ‘ / 20 — / -1 / I 10 b 29 61 -’ Dry 9: 17: Figure 4.19 Time-Inetegrated Product Mass and Composition (shown as percentages next to each bar) as Function of Moisture Content at 2 "[032. 99 120 1101. d 1110 100 - ’1 « Ill 00 a 90 " C’ / " . I: ,1 CO2 g; 80 ' , I] (>/ ' §70_ {ll/1602 .. / $3 / fl ’ 27 TAR '3} l’ E1 /418 ,rl g 40 - / , 1' [ll/I 30 (I, 35/ 4’ ’ .. I / .. 20 I” , / 27 / 5 10 " " 82 1 *2 3 HEAT FLUX (W/cmz) Figure 4.20 Time-Integrated Product Mass and Composition (shown as percentages next to each bar) as Function of Incident Heat Flux for 17% Moisture Content Condition. 100 120 110- x‘mc - 100- III 7} . ... ’T I I .v 6” life 90" 7 ’a“ / I" 13 ,’ 12 co2 .7 sur- , 13 18 // 15 02 70 ...--- / -‘ ~ / 25 18 // 60 '1 / ‘ 27 TAR .1 *“‘“f 501’ 59 27 I 1 / 40" ,r’ - , 54 1120 30 - // . / 52 20 b 1’ - 10 ' 33 '- Dry 9: 172 MOISTURE CONTENT Figure 4.21 Time-Integrated Product Mass and Composition (shown as percentages next to each bar) as Function of Moisture Content at A Heat Flux of 3 W/cm’i 101 the oxygen concentration in ambient air produces nearly constant values of CO2 and H20. The rate of water release in all three cases is very close. This because the adsorbed water has been evolved prior to the time of ignition and the combustion water is much greater than adsorbed water. The C02 production was larger for the dry case, because more fuel evolved in this case is expected to be more. Oxygen depletion was larger for the dry case, which was needed to produce a large amount of 002 and H20. The time of ignition for the dry, 9%, and 17% cases was 88, 130, and 158 seconds respectively. Thus the presence of moisture increase the time required for spontaneous ignition by diluting and delaying the production of fuel. The average surface temperature at spontaneous ignition was approximately 530°C. This temperature is expected to be [lower for piloted ignition, which is the subject of the next chapter. Chapter 5 Piloted Ignition In the previous chapter, it was shown that the presence of moisture in the wood matrix delayed the decomposition process and diluted the .decomposition products. These factors are expected to have a significant effect on the ignition process. The work presented in this chapter attempts to experimentally quantify the effect of adsorbed water on the piloted ignition of wood. The experiments were performed for four moisture contents (dry, 11%, 17% and 27%) and at various incident heat fluxes ranging from 1.8 to 3.7 W/cmz. This chapter is divided into four major sections. Section 5.1 describes the physical aspects of the problem, factors influencing the ignition process and the relation of moisture to ignition. Section 5.2 reviews the previous literature. Section 5.3 describes the experimental apparatus and procedure. Section 5.4 describes the results of the experiments; a model to correlate the ignition data is also derived and discussed. 5.1 Background Ignition refers to the appearance of flame in the volatile gas stream evolved from a solid exposed to external heating (usually 102 103 radiative). Depending upon whether the ignition occurs with or without the aid of an external ignition source, the result is accordingly classified as spontaneous (auto-) or piloted (forced) [Welker (1970)]. For piloted ignition the actual process, as noted in Chapter 1, is quite complicated. The solid must first decompose to produce combustible gases. The combustible gases then mix with the surrounding air to produce a mixture within the flammability limits of composition. The mixture needs an external source of heat, such as a pilot flame, to initiate combustion. Spontaneous ignition will occur only when two conditions are satisfied at some point in the proximity of the exposed surface. First the air-fuel mixture must attain flammability limits, and second, this mixture must achieve a thermal condition that enables it to automatically react in an exothermic manner to yield a flame without the aid of any external source of heat. It is clear that the attainment of spontaneous ignition is more difficult than piloted ignition. The factors influencing ignition of a sample exposed to a radiant heat flux may be separated into two general categories namely, those external and internal to the sample. Factors external to the sample are environmental variables such as temperature, composition and velocity of the surrounding gases, and the magnitude and spectral quantity of the external radiant flux. Factors internal to the sample are its thermophysical and thermochemical properties, its moisture content, and whether the sample is thermally thick or with negligble temperature gradient (thin materials). Experimentally, except for the sample properties, all other variables can be controlled. 104 A common feature of cellulosic materials is that they are hygroscopic and porous. Thus, a large area is available for them to adsorb moisture from the surrounding air. Moisture content of these materials is variable, and is influenced by the climate, season and location. Normally, wood at room temperature and humidity contains about 10 to 15% water by weight, and may contain up to 30% (fiber saturation point) if it is in equilibrium with air saturated by water. As discussed in Chapter 4, this water is held by hydrogen bonds to the cell walls, and is called agggxhgg_fiatg1. Its presence in the wood matrix is expected to have a significant effect on the ignition process. The obvious effect is to increase the time required for ignition by changing the heat transfer and hence the rate of _ temperature rise. This is because the presence of moisture increases the values of the thermal properties such as the thermal conductivity and the volumetric specific heat. Furthermore, since the specific heat of water vapor is twice that of nitrogen, additional fuel is required to attain the same limit flame temperature. 5.2 Previous Literature The ignition of cellulosic material (both auto- and piloted) has been an active area of research, and several excellent reviews have been published on the subject [Welker (1970), Kanury (1972), and Steward (1974)] . Numerous techniques have been developed to investigate the ignition phenomenon. Various heat sources has been used, namely gas-fired radiant panels [Simms and Co-workers (1960,1963,l967)], high tungsten 105 filament lamps [Smith et al. (1970)], carbon arc [Martin (1969)], 002 laser [Kashiwagi (1981)], electrical heating elements [Atreya (1983)], etc. The experiment essentially constituts of exposing a sample to a known external radiant flux and recording the timee required to ignite the sample in the presence of a pilot flame. The ignition times measured by various investigators have not always been in good agreement. For example, the time for ignition of wood irradiated by a tungsten lamp is about four times larger than by diffusion flames at the same incident irradiance [Wesson et a1. (1971)]. The primary reason for this disagreement was found to be due to the difference in the spectral quality of the incident radiation relative to the spectral absorptance of wood. Thus the spectral nature of the heat source to be used in the experiments must closely match the intended application. For building fires this implies that the heat source should have an effective blackbody radiation temperature of about 1200’K. Several ignition criteria have been proposed such as: critical surface temperature at ignition [Simms (1963)], critical mass flux [Bamford et a1. (1946)], critical char depth [Sauer and Interim (1956)], critical mean solid temperature [Martin (1965)], etc. Of these, critical fuel mass flux at ignition seems to be physically the most correct since it can be related to flammability limits. Surface temperature has proved to be the most useful ignition criterion since it can be conveniently related to fire spread [Atreya (1983)]. However, only a few investigators [Atreya (1983)], Kashiwagi (1981), Garden (1953)] have directly measured the surface temperature at ignition; others have estimated it by extrapolating measured indepth 106- temperatures [Martin (1965)], or by using a linear heat conduction theory [Simms (l960,l963,l967)]. The reported surface temperature at piloted ignition for wood ranges from 300-540'0. Atreya's recent work (1983) has shown that with the correct interpretation of measured surface temperatures a somewhat narrower range [375 i 30°C] is obtained. In this regard, it is important to note that all of the above criteria are based on an indirect quantity which is assumed to be closely related to ignition. The importance of moisture content in relation to fire tests has long been recognized. Researchers have carefully controlled the moisture content of their samples. However, few investigators have addressed its effect on ignition. Martin et a1. (1958), studied auto- ignition of a-cellulose conditioned at different relative humidities using carbon arc as a radiant source. Simms and Law (1967) performed both auto- and piloted ignition tests on samples with different moisture contents using gas-fired radiant panels. In both these studies moisture was found to significantly affect the ignition process. Atreya (1983) investigated (experimentally) piloted ignition for different types of wood. He showed that critical surface temperature is a reasonable and convenient criterion for ignition. He indicates that this temperature increases for low heat flux experiments due to the build up of char. Later Atreya et a1. (1987) investigated the effect of sample orientation on piloted ignition and opposed-wind flame_ spread. They used two types of wood (red oak and mahogany) to investigate two orientations (horizontal and vertical). They reduced the experimental data according to the thermal flame spread theory 107 using the measured surface temperature. They indicated that as long as the temperatures are defined consistently with the thermal theory, the results are orientation independent within the measurement error. The mathematical description of the heat transfer for the ignition problem has been widely investigated. The problem is usually considered to be one dimensional with all heat transfer occurring normal to the sample surface. The cases investigated produce correlations of different forms, because different boundary conditions and various assumptions were used in each case. Butler et a1. (1956), assumed that the sample is inert, undergoes no decomposition, front surface is opaque to the incident radiation, no heat lossess, and the rear surface is insulated. They used the solution of this problem as the basis for correlating their ignition data. They omitted the surface temperature rise at ignition because of the difficulties they encountered in attaching physical significance to it. They also indicated that the temperature rise at ignition was relatively constant. Thus, they claimed no significance error was introduced by omitting it. Simms and Law (1967), included convective cooling losses at the front surface in the boundary condition. However, in order to correlate their ignition data for wet and dry woods, they had to fix two unmeasured parameters, the convective heat transfer coefficient (h) and surface temperature at ignition (Ts). For piloted ignition they used Ts - 360°C and h - 8.6x10'4 cal/cm2.sec°C; for spontaneous ignition they used Ts - 525°C and h - 14x10.4 cal/cmz. sec°C. The 108 convective coefficients used in their work are larger than those used by other investigators. For example, Alvares et a1. (1969) used h - 4 cal/cm2.sec°C based on free convection heat transfer theory. 2.8 x 10' In the opinion of Welker (1970), the reason for a large value of h (required to obtain a correlation) in Simms work may be that the heat losses by reflection of incident radiation and emission of radiation from the surface were ignored. Atreya and Wichman (1987) developed an approximate analytical model for the piloted ignition process. They solved for time, surface temperature and mass flux at ignition, and showed that the predictions were in good agreement with the experimental measurements. Tzeng et a1. (1988) developed a theoretical model for the piloted ignition process. They solved the equations numerically, and examined the location of the ignition source, fuel mass evolution rate from the ‘ surface, and the surface temperature of the solid. They showed that the (1) model adequately explains the pre-ignition flashes that-are often observed experimentally, (ii) provides a rational criterion for positioning the pilot flame, (iii) indicates that the heat losses to the surface play an important role, and shows that the fuel flow rate by itself is insufficient for predicting the onset of piloted ignition. Recently, Mikkola and Wichman (1988) suggested a correlation that relates the incident heat flux (F) to the negative one-half power of ignition time (t). They deduced this relation from the solution of the one-dimensional heat conduction equation for a semi-infinite solid with constant external heat flux, linearized surface heat loss, and a constant initial sample temperature. They used wood ignition data 109 0.5 available in the literature and plotted t' vs F, and showed that the data collapsed onto separate straight lines whose common abscissal intercept is approximately constant (12.5 KW/mz). They explained this constancy for the different types of wood by the similarity in the chemical structure of wood. They indicated that this value (12.5 W/mz) is the minimum heat flux for ignition, and provided additional support for their correlation, by showing that the data for polypropylene and polyurethane, have different slopes and intercepts. They explained that this is because their chemical structure are different from wood. 5.3 Apparatus and Procedure A schematic of the experimental arrangement used for these tests is shown in Figure 5.1. Since the experimental apparatus used is described in detail in Chapter 2, only the relevant aspects are considered here. The external radiation was provided by electrical radiant heaters. These heaters have a maximum backbody temperature of 1200°K and are therefore well suited to simulate external radiation in building fires. A small natural-gas flame was located in the mixing layer, pointing downward. The pilot flame was off the sample edge at a height of 0.5" from the sample surface (see Figure 5.1). The configuration and the position of the pilot were chosen in order to avoid heating the sample. The samples were prepared and instrumented with thermocouples. The samples were then dried by a standard procedure in a temperature- 110 onoooum nowuasmu can no owuqaonom _~.m enough Mooqm mummoo afigooolmu9<§ zH momzmm Rank H53 29:29 —bb-—-b—.-— —--L-P—Lnb—bnn—~p-Lbn...- 09: 00m 00m DON. com com on:~ com com 00? o PB“ V‘s—N 2:: *xxx 0: O 11'1'I‘IU'IIIIIIIUIIITT' m.— 0. N “2 (\l (awe/M) Xn'ld 1V3H lNBGIONI 0. 1’) O. <- 114 shows also a large difference in ignition time. For example, a dry sample ignited at 24 seconds at 3.5 W/cmz, whereas a sample with 27% moisture content ignited at 90 seconds at the same incident heat flux, suggesting that fire spread on wood containing moisture would be slower than on dry wood. 5.4.2 Correlation of results Ignition time data were correlated on the basis of a relationship suggested by the solution of the one-dimensional heat diffusion equation in the solid. Consider a semi-infinite solid initially at a constant temperature (To). At t > 0 the solid is exposed to a prescribed heat flux (F). Assuming that the solid is inert and undergoes no thermal decomposition. Since the heat transfer through an inert solid is by conduction only, the following form of the one-dimensional heat diffusion equation applies 2 2 pcg—f-xgfi- forx>0,t>0 (5.1) where 0 - T - T”. The solid is assumed to be at the temperature of its surroundings prior to the experiment, therefore 0 - 0 for t s 0, x > O The boundary condition is obtained from the energy balance at the solid surface (i.e., heat diffused into the solid - incident heat flux - convective heat loss - radiative heat loss). Thus, the boundary condition at the front surface becomes 115 a; _ _ _ a - a - K 6x x_0- £8 (as, T0, t) e [P h as a((as + To) Tm)] (5.2) where e - the emissivity (or absorptivity) of the surface, 5‘ I heat transfer coefficient/e, Stefan-Boltzmann constant, e I F - the prescribed incident heat flux, 0 - Ts - Tdo where T8 is the surface temperature, f - heat flux going into the semi-infinite solid at the front surface, and p,C, and K are, respectively the density, heat capacity, and thermal conductivity of the solid. Atreya (1983), obtained an analytical solution for the above problem by the use of approximate integral methods. He solved for the ignition time and gave it in the following form; 2 o (2so +r - Jfi)(r + J?) o (r + 230 ) t - M[-§§ + -§7§ 2n[ § ]- —§———1;—-JL], (5.3) 2fs p (25108 + r + /§)(r - J?) 5 £8 where M - g 2E% , r - -(h + 4 a T3), 5 - -2% a T: , 2 AF d f* fig F o + 02 fl - (r - 3). an s - e - + r s s s 116 This equation was simplified by approximating the logarithm term, and substituting the value of f: as the net heat flux diffused into the solid [i.e. f: - (F - L)], where L, is the total heat loss by convection and radiation. The final form of equation 5.3 is 2 . - H 0 ]Q [J - ICEJZ . 2(r/0 + 23) (l-L/F) t 2 -f5- 1 + 2 2 ---§* 2 (5.4) fs 16s 08(1 - r /4Fs) 43(1 - r /4Fs) The terms inside the bracket of equation 5.4, were examined closely, and they were evaluated for a wide range of heat fluxes (P). Let 19 {J-IIE12 , 2(r/0 + 23) (1 . L/F) e , e - 1 165 08(1 - r2/4Fs)2 2 as(1 - r2/4 Fs) e3 - e1 + e2 and ea - e3+ 1. Figure 5.3 shows the value of each term plotted against the incident heat flux. Note that at piloted ignition the surface temperature as is approximately constant. Thus, the only variable is F. For the range of heat fluxes studied (1.8 - 6 W/cm2 which are representative for igntion), two conclusions may be drawn from Figure 5.3: (i) the variation in each of the terms e1 and e2 is small compared to the dominant term which has a value of 1.0 (ii) the value of all the terms in the bracket (ea) can be approximated by a constant with an error of less than i 6% in the range of interest. 117 .xsam use: ucovaocH n> q.n coauosvm mo menus n.n shaman ANEQE x3... .5: 5.362 m m. a. n N _ [Ll- - u — p p p H'- . . — p p p _ p p - n.0' O O O .. O OHoO O O O O m mod .wnd . med 0 I on... o o M e m . O O No 9 0 mum.O my aw Aw mv aw .0 m WNJ .. mm; .qo -- n I .. I I I I I I I ......m... Pd ‘17'9 NOLLVOOB :10 SWHHl 118 By substituting the value of M given under Eq. (5.3), equation (5.4) may be written as 64 pCK 0: t um: . . (5.5) Assuming that the front surface temperature at the time of ignition is constant, so that it is no longer a correlating parameter, the heat flux into the solid at the instant of ignition is given by f -c*c - , (5.6) / pCK where, C* - J 43:2 08 . (5.7) Substitute the value-of f: back into equation 5.2, the following relation between incident heat flux and time for ignition is obtained F - 0* c’0'5 + L , (5.8) where C* is given by equation 5.7, and L represents the heat loss by convection and radiation. L, is given by the following equation 4 L - h as + 0((0s + I”)4 - Tm) . (5.9) 119 Equation 5.8 is the basic relationship for correlating the data of ignition time as a function of incident heat flux. The ignition time to the negative half power was plotted versus incident heat flux. The data for the four moisture contents are shown in Figure 5.4; the measured values were shown by points and the least square fit by the straight lines. As shown in the figure each moisture data collapsed onto a separate straight line, the intercept corresponds to L, and the slope corresponds to 0* in equation 5.7. * The parameters L and C , were calculated for each value of the moisture contents; L was determined from equation 5.9. For 08, the value of the measured surface temperature at ignition was used. The value of 6* was calculated from equation 5.7. The values for the thermophysical properties (pCK) were obtained from the literature [Parker (1988), Simms (1967)]. The oven-dry properties were given in Parker (1988), who used the same type of wood. The properties for the rest of the moisture contents were determined by using the formulas of Simms (1967). A value of 1.75 was used for ca. This is the average value of all the terms in the bracket of equation 5.4 from Figure 5.3. A comparison of the derived and the calculated values of L and C* are presented in Table 5.1. Because surface temperature at ignition was found to increase with moisture content (this will be discussed in the next section) it would be expected that the heat loss increases. The derived and the calculated values of L show an increase, as the moisture content increases. The reason for a low value in the experimental heat loss (L) at 27% moisture content is not known to the 120 .osua mud-n coauacuu uo soda-Houses <.n enema» reabsv x3“. :m: 5.862. m.» .onn. .mum. cum 0.? n p b p b L n n n b n \ .11.. 2 R nu ma \ .1... 2 N t \ 111 2 N Z .11. #3 C) 0| ID II (so-S) 31111 AV'EICI NOLLINOI 121 author; more data is needed to confirm this value. Because the thermophysical properties of wood increase as moisture content * increase, it is expected that the value of C increases. The experimental values of 0* have an average deviation of 10% from the calculated values; this may be because of the error introduced by approximating all terms in the bracket of equation 5.4 by a constant. Such an approximation is expected to introduce ~ 12% error as discussed earlier in this section. The sixth column in Table 5.1. was obtained using a computer program (NLINA), provided by Professor J. Beck (1977). This program can be used to estimate parameters for both linear and non-linear cases. The calculated heat loss values along with the ignition data were used to estimate 0*. The ignition correlation reveals the strong influence of the heat loss and moisture content upon ignition characteristics. It shows that, as the time required for ignition tends to infinity, the heat loss tends to the incident heat flux (see equation 5.8). Hence the critical incident heat flux, below which ignition is impossible, is given by L. This critical flux is, in effect, the rate at which heat is lost from the surface at ignition temperature. Because surface temperature has been found to increase with moisture content, it would be expected that the critical incident heat flux would increase (see Table 5.1). Thus the moisture has a significant effect on piloted ignition; it increases the ignition delay time and more importantly increases the critical incident heat flux. Piloted ignition is closely related to fire spread. In fire spread, the volatiles produced from the heated surface are ignited 122 either by the source of the heat itself or by the aid of external sources . Thus, fire spread may be regarded as a rapid succession of piloted ignitions. piloted ignition is a very important concept for fire spread. This suggests that the critical heat flux for Using this value in building regulations will result in a safe separation distances between housings, and this will reduce the risk of fire spread to neighboring properties. TABLE 5.1 * Comparison of Calculated and Experimental Values of C and L * C Moisture Experimental Using Content Calculated Experi- Calculated Experi- Calculated Heat Loss mental mental Dry 1.01 1.13 8.6 9.78 10.73 11%. 1.05 1.20 13.2 15.53 14.84 17% 1.10 1.32 15.54 13.3 15.02 27% '1.30 1.14 20.86 22.85 20.20 5.4.3 Surface temperature Surface temperature profiles for dry and 11% moisture content at 2,3, and 4 W/cm2 are shown in Figure 5.5 and 5.6 respectively. These Figures show that the rate of rise of surface temperature increases as the incident heat flux increases. Also the rate of rise increases as moisture content decreases at the same incident heat flux as shown in Figure 5.7. 123 At the time of ignition the flame appears in the gas phase, and this causes a sudden increase in the surface temperature. The arrows in Figures 5.5-5.7 indicate the time of ignition. These figures show that ignition time increases as heat flux decreases (see Figures 5.5 and 5.6), and increases as moisture content increases at the same incident heat flux (see Figure 5.7). Data for surface temperature at ignition are summarized in Figure 5.8. There is some scatter in the data, but the trends are clear. At any incident heat flux, the surface temperature at ignition is higher for higher moisture content, and at any moisture content, it increases as heat flux decreases. This increase in ignition temperature is due to the slow decomposition of the solid at the surface, and the resulting build up of the low thermal conductivity char layer prior to ignition. The sample moisture content also had a significant effect on surface temperature at auto-ignition. In the decomposition experiments discussed in Chapter 4, it was pointed out the auto-ignition occurs at 4 W/cm2 for the 3 moisture cases studied. The surface temperature of the sample at ignition in these experiments were 510, 525, 550°C for dry, 9%, and 17% moisture content, respectively. These temperatures are much higher than the piloted ignition temperatures under the same conditions. This is because in the case of piloted ignition, the sample surface temperature must rise only high enough to decompose the sample, provided that decomposition is rapid enough to produce enough combustible gases to form a combustible mixture with the surrounding air at the pilot location. 0n the other hand auto-ignition, must be initiated by heating either the sample surface to the point where it ignites the gases, or by ' 124 cowuavcoo man you ouauauoeaoa oueuusm so seam nee: acovaocu no uoouum n.n enough 3 m2: 000 005 000 000 00¢ [000 00m 009 uh. uh- - n - —[Pf-|n - p 0 as: ZOHHHZUH «an? 8 A 2 mafia. ZOHHHZGH h . _ o h 8. com 8... 8... com com 8s 8m 8m -000? (0.) BaniVHBdWElJ. 3011:1808 coauuvcoo acouaoo assuage: «as new ensueuoaaos ooeuusm so seam use: ucoeqocu mo uooumu o.n «woman 3 m2: com com 00¢ com com cos 0 125 b . . p L . _ . . . L [p . . _ p . p — . . b —. o .. 00—. BAHH. ZOHHHZUH h CON h 00m. ~83 2s 00¢ «so; a; N5; 0 00 3sz 35:22 h 000 005 000 000 (0.) BaanHI-IdWEIl 30113803 - 000_. 126 «so; was Co 5.: use 0532: me an ensueuoeaoa ooeuusm so acousoo assuage: mo uoomuu ~.n «woman 3 m2: .. com com 00. o n b n — n b — n p. — P L b \ 00 F 00m 000 00¢ 000 000 00k. - 000 (0.) 38101101381131 Bowans 127 0.¢ .uoquusmn ue eunuch—oases 3356 so acousoo assuage: 00 33.4.0 0.¢. can»: Anemia x3: 2m: .2862. an o.» 2 em as b n b p b L p b L b n r h p — r b b L] b L p b P . .I 00” won». on .I l .....o wot». Po . oo / / 0 m8... 1.. lllllll 0 .. am.-----:: - -- ..-mn.J, I / womn e -..?me / ... III I [I/ WON¢ « woee z e R .. .....I s 2 N t o .1... woo¢ 2 a : . ...... - ..o 1. >8 0 .ll m m 1 com. (0.) NOLLINSI 1v BHMVHEIdWZ-ll 3011:1808 128 .Nao\3 o.“ u. ueoueoo assuage: «am new ooauouaqm omusueuoaaos oddeem o.n ouemam 3 m2: com com com com com 00¢ com com 00, o uh. DLf—-D.—bL- :b-b-nb D-PL-pP—D-Lo 111 208800 00 _. DON BE 5353 * con Saga 00¢ com 000 00m. - 000 (0.) 3801118381131 3181111115 129 heating the gases themselves to the ignition point. In either case, this would require a higher surface temperature. Sample surface and bottom temperatures during a typical test of piloted ignition are shown in Figure 5.9. This Figure shows that the bottom temperature increased only by 2-3°C before the time of ignition. Thus the assumption that,the samples were thermally thick used in section 5.4.2 is valid. 5.4.4 Sample Iass flux Mass flux histories for dry and 17% moisture content are shown in Figures 5.10 and 5.11 respectively. The arrows in these Figures indicate the time of ignition. These Figures show that the mass flux increased dramatically at the time of ignition and continue to increase. Meanwhile a char layer on the sample surface forms. This char layer is not completely inert but it is much more resistant to the evolved gases than original wood. This is because char has lower thermal conductivity, and the heat penetrating through this thick char would decrease. This consequently decreases the volatile mass flux, which continues to decrease until attaining a nearly steady state condition. Numerical values of the mass flux at the time of ignition for the different moisture contents at various incident heat fluxes were determined from the mass flux histories. These values are presented in Table 5.2. This table shows that the mass flux increases as the incident heat flux increases, and at the same incident heat flux, it increases as moisture content increases. This is because the initial mass evolution from the samples is essentially desorption of water followed by a large amount of tar. These quantities are evolved at 130 soauavsoo an: new asap one: uoovcoaonuoaua so seam aeom ocueaosm mo uoouum o~.n enough 3 m2: 000. .muoxu 000 000 00¢ 000 00m 00w 0 b b p b n b n n n n n n b n - MZHH :onHHZUH a ‘1‘ . mmzHa ZOHHHZUH ao\3 0.N N5; 34 (S ZwO/fiw) x013 ssvw 131 souuuvcoo ousuuwoz aha you Mean and: unaccononloaua no Kean ueom.u=ov«ouu mo uouuum 3 m2: 000— com com com com com oov .omn b.» —:—_-—- ——-—_—__~—— P 0 -CD ~«.n shaman HZHH ZOHHHZUH— N502» cm .H Nau\3 H~.m (s awe/5w) xmj ssvw 132 higher rates as moisture content and incident heat flux increases. Table 5.2 shows that a mass flux of about 0.22 mg/cmzs is necessary for piloted ignition to occur. Bamford et a1. (1945) proposed that piloted igntion of vertical slab of wood is possible when the mass flux of the decomposition products exceeds a critical value.I For deal, oak, and pine wood, they gave this critical mass flux as approximatley 0.25 mg/cm2.s. Table 5 . 2 Comparison of Mass Flux at Ignition For Different Moisture Content At Various Heat Fluxes (F) Moisture F I Mass Flux F I Mass Flux F I Mass Flux Content W/cm2 I mg/cmzs S/cm2 I mg/cmzs W/cm2 I Mg/cmzs | I I Dry 1.85 I 0.22 2.6 I 0.28 3.48 I 0.48 11% 1.80 I 0.28 2.76 I 0.37 3.53 I 0.50 17% 1.84 0.30 2.36 I 0.44 3.21 0.42 27% 1.92 0.37 2.77 0.47 3.62 0.58 5.4.5 Products evolved In this study oxygen depletion, production of C02, CO, total hydrocarbons [THC], and water were measured before and after the time of ignition. The procedure used for the analysis of these products is the same as used for decomposition experiments, as described in section 4.3. The data is presented as mass flux of permanent gases, and as percent of mass flux. Plots for different moisture contents at various heat fluxes are shown in Figures 5.12 - 5.21. 133 /-\ OD N E (J ‘\t (75 E \_I X I33. (I) g 0.0-£1— fi-,--.,.‘-~~.-’ -O.2 UII'III'IIIIUITIIUI" U 100~ 200 300 400 500 T1ME.(s) . Figure 5.12 Mass Flux of Evolved Products at Incident Beet PRODUCT AS PERCENT OF MASS FLUX Figure 5.13 Flux 3.5 Vie-2 for Dry Condition ‘U U U U I I I I I U 200 TIME (3) Products as Psrcent of Mass Flux at Incident Best I I V l 300 400 500 Flux 3.5 H/cm for Dry Condition PRODUCT AS PERCENT OF MASS FLUX MASS FLUX (mg/cm“ 3) 134 : , ‘“~. CO A’ ' O 4- M’ ° 0 I , o2 narrator: 0.2‘3 I ‘ s \e. TAR O O- 3,... ' co and mc ~ . . , . - . I I I2 I I I l I I v I ~ ‘ 0 "JV 100 300 400 500 ' OOTIME (s) Figure 5.14 Mass Flux of Evolved Products at Incident Beat Flux 2. 6 “Ion2 for Dry Condition Figure 5.15 Products as Percent of Mass Flux at Incident Beat Flux 2.6 W/cm2 for dry Condition 135 '5 P- :83“ i}: I“ 30 O G I \ . .0 ‘ ~\.}--“'- ' 02 D PLETION MASS FLUX (mg/cm2 3) TIME (s) Figure 5.16 Mass Flux of olved Products at Incident Beet Flux 2.75 W/cI for 112 Moisture Content Condition 1 00.0 PRODUCT AS PERCENT OF MASS FLUX TIME (s) Figure 5.17 Proudcts as Percent of Mass Flux at Incident Heat Flux 2.75 H/cm for 112 Moisture Content Condition 136 1.2 . ’“~ I m 1.0-:- N 5 0.8-3 E 2 E 0.6-; “2° V - \k . 0 DEPLETIO x .. 2 0 0.4: I: 0.2: 00 I g 00-: ‘ * ' .mc, . -O.2. I I OI I 3 OI—I I '1 0 100 200 300 600 TIME (s) Figure 5.18 Mess Flux of Exolved Products at Incident Heat Flux 2. 77 His. for 271 Moisture Content Condition PRODUCT AS PERCENT OF MASS FLUX -2000 T I I I I I F! I t i I I I I I I I 1'1 0 1 00 200 300 400 500 600 TIME (s) Figure 5.19 Products as Percent of Mass Flux at Incident Beat Flux 2. 77 chmz for 272 Moisture Content Condition MASS FLUX (mg/cm2 3) 137 1.0 . 0.6-3 kl I ‘\*“-I 0.4”: ‘I 320 0.2; '- o n . . II! 0.0-3 - """""”'"’."I' ,°° “‘1 m "092 - I I l I I I I I I I I I I I 1 I l I I I '1 100 200 300 400 500 6 700 TIME (s) Figure 5.20 MASS FLUX (mg/cm” 3) Mass Flux of Evolved Products at Incident Beet Flux 1.85 via-2 for Dry Condition 1120 02 DEPLETION "Vs “\Joz ‘\ s v’ A L ' .I "‘ ‘ ‘0. 200 TIME (s) Figure 5.21 Mass Flux of Evolved Products at Incident Heat Flux 3.21 Vlcm2 for 172 Moisture Content Condition. 138 Before ignition, the gases evolved are products of decomposition. This was the subject of discussion of Chapter 4. These Figures show that water is a major component of the evolved gases. The higher the incident eat flux and moisture content, the higher is the water production rate. As pointed out in Chapter 4, the presence of moisture in the samples delayed and diluted the combustible gases. This is the reason why ignition occurs at later times for samples with higher moisture content. After the time of ignition, the general trend of the measured species is very similar for all moisture content cases at the various incident heat fluxes; a sudden increase to a sharp peak, fall-off, and then attainment of a steady state value. The sharp peak is caused by the burning of the large amount of fuel available in the gas phase at the time of ignition. Meanwhile a char layer starts to form which reduces the rate of the evolved gases including the combustible gases. This results in a reduction of the product rates, which eventually attain a steady state. Figures 5.12 - 5.21 show that the production of water after ignition increases as the incident heat flux increases, and at the same incident heat flux, the water production does not change significantly with the increase in moisture content. This is because the combustion water is much larger than the adsorbed water, and a large quantity of the adsorbed water was evolved prior to the time of igntion. This is supported by the fact that water production prior to ignition is higher for higher moisture contents. These Figures also show that the production of 002 decreases as moisture content increases at the same 139 incident heat flux, and decreases as incident heat flux decreases at the same moisture content. Oxygen depletion follows exactly the same trend as 002, as expected. This is because, as pointed out in Chapter 4, the production rate of combustible gases decreases as the heat flux decreases, and are diluted by the presence of moisture in the sample. Figures 5.12-5.21 also show that the production of combustible gases is almost zero, since these gases burn to produce C02 and water. Figure 5.18 shows that at about 475 sec, as the flame gradually died out, 002 production and 02 depletion gradually decreased, while the production of combustible gases (C0 and THC), started to gradually increase. After the flame died-out completely a steady state was attained. They may now be considered as decomposition products. The tar in some experiments shows a negative value (see Figure 5.12) because it was determined by differences (as discussed in Section 4.3), and any measurement error will show in the tar value. CHAPTER 6 CONCLUSIONS This chapter presents a summary of the important results for different components of this work. Section 6.1 summarizes the conclusions derived from the results of the droplet evaporation on the hot porous and non-porous solids. Section 6.2 summarizes the conclusions obtained from a set of systematic experiments performed to investigate the effect of water on both thermal decomposition and ignition. Some recommendations for future work are presented in Section 6.3. 6.1 Droplet Evaporation Experilents The transient surface and indepth temperature data obtained during droplet evaporation reveal the thermal behavior of both the porous and the non-porous solids. The following conclusions are apparent from this investigation: 1. The only similarity in the thermal behavior of both porous and non- porous solids is the agreement in the experimental contact temperature. This is because it occurs during the initial contact and the draplet sees both solids as a semi-infinite body. 140 141 2. The theoretical and experimental contact temperatures are in good agreement up to the boiling point of water and then diverage with the experimental contact temperature becoming roughly constant at a value slightly greater than the boiling point. 3. During the droplet evaporation process, surface and indepth temperatures for the non-porous solid remain nearly constant whereas for the porous solid there was a continuous decrease in these temperatures. 4. A thermocouple in the porous matrix at the same location as that of the non-porous matrix cools faster under identical conditions. 5. Both, the recovery time and the envelope of droplet influence were larger for the porous solid. These results confirm the indepth cooling Iof the porous solid due to water penetration. 6. The evaporation time is longer for the non-porous solid than for the porous solid for the same droplet diameter and under identical conditions. This is because part of the droplet diffuses into the porous solid matrix. 7. Smaller droplets are more efficient for cooling non-porous solid at or below 100°C. This is because the evaporation process for larger droplets attains a steady state and the extra fluid in the larger droplets just increase the evaporation time. This phenomenon was not 142- observed in the porous solid because evaporation process never quite attains a steady value. 8. For the non-porous solid, the instantaneous evaporation rate, and the instantaneous average evaporative heat fluxes were determined from the transient measurements of surface and indepth solid temperatures, and the droplet diameter on the solid surface. Although the total heat transfer is more for large droplets the average evaporative heat flux is higher for smaller droplets. This is because small droplets have larger conductance due to their smaller thickness on the hot surface. 6.2 Thermal Decomposition and Piloted Ignition Experhments The following conclusions are drawn from this investigation: 1. The presence of moisture in the wood sample has a significant effect on the surface temperature and char oxidation. As the moisture content increases, the rate of increase of surface temperature decreases. Thus, the char oxidation occurs at later times for the same incident heat flux. 2. The presence of moisture in the wood matrix delays the decomposition process and dilutes the decomposition products. 3. The mass flux at the time of ignition increases as the moisture content increases. This is because the initial mass evolution from the sample is essentially desorbed water, and this component evolves at 143 higher rates as the moisture content increases. An absolute minimum mass flux of about 0.22 mg/cm2.s is necessary for piloted ignition to occur . 4. The presence of moisture in the wood matrix increases the total energy needed to ignite the sample by increasing the ignition time. It also increases the critical incident heat flux by increasing the heat loss from the sample surface. 5. A single correlation was derived for all ignition data. This correlation accounts for the moisture dependent thermal properties and the heat loss from the sample. 6. The surface temperature at ignition increases as moisture content increases. This is due to slow decomposition of the solid at the surface and the resulting buildup of the low thermal conductivity char layer prior to ignition. 6.3 Recommendations for Future work Based on this research, it seems that the following areas need further investigation in the future. 1. A study of transient cooling of the porous solid by droplet evaporation of fluids other than pure water. This will hopefully yield the optimum cooling agent. 144 2. A study of the effect of droplet release height on the cooling process will yield the optimum strategy for delivering the coolant agent. 3. A study of the droplet evaporation of common fuels such as methanol and benzene at higher temperatures on solids with different porosity. This may have direct application in the design of combustion chambers for internal combustion engines. 4. Investigation of the effect of oxygen concentration on the ignition process. This will allow determination the minimum oxygen concentration necessary for ignition. 5. Investigation of the effect of air velocity on the ignition process. Air velocity may significantly affect the ignition process, for ignition to occur, the mixture of the products of decomposition generated by the solid and the surrounding air should be within flammability limits. 6. Develop a theoretical model for thermal decomposition of wood and compare with the present experimental data. APPENDIX A APPENDIX A METHOD OF ESTIMATING LOCATIONS 0F IN-DEPTH THERMOCOUPLES AND THERMOPHYSICAL PROPERTIES OF THE CERAMIC .In this appendix, the procedures used to determine the location of in-depth thermocouples in the ceramic blocks and the thermophysical properties of the ceramic are described. Some of the properties were supplied from the Sandia National Laboratories [Taylor and Groot (1985)]. However, since the properties of castable materials change according to the procedure used in casting and drying them, it was necessary to re-evaluate these properties. A program provided by Professor J.V. Beck (1977) was used to estimate the thermal conductivity and the density-specific heat product using transient temperature and heat flux measurements. A.1 Inrdepth.Thermocouples Locations The locations of thermocouples, prior to casting, were measured to within 2 0.1 mm. Due to the uncertainties introduced during the casting process and shrinkage, thermocouple locations were evaluated by a least square fit to the measured steady state temperatures. This criterion yields a unique "best fit" line for the given set of data. Figure 1A shows the least square fit to the steady state temperatures. The evaluated thermocouple locations from the least square fit were close 145 146 to the measured values before casting. Table 1A shows the comparison of the two values. Attempts to determine thermocouple locations radiographically by using X-rays were not successful. A.2 Porosity Porosity is the fractional void volume of the ceramic. A simple and straightforward method was used to determine the porosity. Several samples were dried by a standard procedure in a temperature-controlled chamber at 105 0C until constant weights were achieved. The samples were then rinsed in distilled water until all the samples attained saturated weights. Since there was no change in dimensions of the dry and the wet samples (i.e, no water adsorption), the difference between the weights of the saturated and dry samples was used to determine the amount of water displaced. Then the volume fraction of the cell-wall substance was calculated. The value of the porosity tabulated in Table 2A is an average of four tests. A.3 Thermophysical Properties The thermal conductivity, the product of density and specific heat were estimated by a technique used by Professor J.V. Beck. Two thin electrical circular heaters 3" in diameter were placed between two identical ceramic solid-cylinders. Both solid cylinders were also 3" in diameter. One of them was instrumented by the thermocouples. Two 147 aluminum solids were placed at both ends in order to obtain constant temperature boundary conditions. The solids were allowed to heat up until a steady state condition was achieved and then an electrical step power input was applied to the heaters. Finally, the power was turned off. During the experiment temperature data was collected. The temperature history of the thermocouples in a typical test is shown in Figure 2A. The time for the step power input and the time when power was turned off are also shown. Voltage and current inputs were also measured. Program PROP, provided by Professor J.V. Beck, was used to estimate the thermal conductivity and the density-specifitheat product. This finite difference program solves a transient one-dimensional homogenous partial differential equation with temperature-dependent properties. The two boundary conditions are the heat flux at the upper surface and the constant temperature at the lower surface of the solid. The heat flux was calculated from the voltage and current measurements and from the area of the heaters. Results of the thermophysical properties at four different temperatures are tabulated in Table 2A. From these results it may be concluded that the thermophysical properties of the ceramic are not strongly temperature-dependent. Table 2A also contains the constituents of the refractory ceramic. 148 TABLE 1A In-depth Thermocouples Locations W L Warm-.1 0.0 | 0.0 1.3 | 1.69 2.9 | 3.06 5.5 | 6.63 . 6.3 | 7.33 10.5 | 12.27 26.0 | 26.56 (i) (ii) (iii) 149 TABLE 2A Magnesium Oxide Ceramic-Refractory Grade Constituents or chemical analysis by weight _§emneune.. | Seanceitien... Mg 0 | 96.11% s1 02 I 2.2% Fe2 03 0.36% A12 03 0.19% Cr2 03 0.06% Ca 0 1.08% Porosity 33.28% Thermophysical Properties Tempera- Density Specific Thermal Thermal sure Heat Condustizisxm__nifussizi£x L0G) (gm12m31_u_il£gm_.fil (uggm,k) (gmzzggg.) 65 2.053 1.00 0.02283 0.0111 100 2.053 1.076 0.02261 0.0102 150 2.053 1.307 0.02225 0.0083 240 2.053 1.307 0.02225 0.0083 150 .uousueuodsua uueum museum «:9 ca ugh museum ueuog <~ ouswum AEEV 8:320 0N 0N mp bIn—bL-Irb-bb 0p m 0 .L b - n b b p — n n b - «E 96:00 «moo.- I N¢ 20823302 dEo.» x (0°) mmmsdwei 151 .xsdm use: no uuonuso use oeeouosu doom < wcuusc mousueuoa509 :udovcu usuaeceus <~ shaman Aoomv 08:. on. 8. .2... .om. . 2.... an on 9 cm a s. —.L p p n L p 0* acnhnnzoo 529500 mags—ugh. 2.5200 7 v ..... . o . a 8030“— 9053 muhm mushy—=0 .l O“ mac Guzman. mus?— (30) SOJI‘IIOJOdwOIL q1dapug APPENDIX B APPENDIX B CRITICAL NOZZLES This Appendix presents a description and the calibration procedure of the critical nozzles. These nozzles were used because they are a very accurate way of controlling and measuring mass flow rates of gases over a wide range of flow rates. Critical nozzles offer many advantages over other measuring devices; they do not have any moving parts which produce friction or wear [Aschenbrenner (1983)]; no pressure differential measurement is needed [Arnberg (1962)); simple construction, and simple flow equation that depends on pressure and temperature of upstream.‘ The configuration of the nozzles was adapted from [Arnberg [1962]]. The general design is shown in Figure 13. A set of nozzles with different diameters were manufactured to deliver a flowrate of air anywhere in the range (175-32,000 ft3/hr). An identical set was also manufactured to be used for either 02 or N2. A straight section of pipe approximately 120" was placed in front of the nozzle in order to achieve fully developed conditions at that location. 8.1 Mass Flowrate The theoretical mass flowrate of a given nozzle is directly proportional to upstream pressure and temperature. The static state is 152 153 given in ASME (1959)]. [Holman (1984)] presented the equation in the following form: 2:. -_ _1_ m A9. P1 RTl 1+1[(1+1) ' (3'1) .2. 1'” ., where m - mass flowrate (lbm/sec), A* - area of the nozzle (in.2). P1 - inlet static pressure (Psia), g - gravitational conversion factor (32.2 lbm ft/lb secz), R - gas constant (ft lbf/lbm R), T1 - inlet temperature (°R), 1 - ratio of specific heats of the gas (Cp/Cv)° B.2 Calibration A trace gas technique was used for calibration of the nozzles. A known flowrate of methane (measured by a flowmeter) was-introduced downstream of the critical nozzle. A sample of this mixture passed through the gas analyzer, where the percentage of methane was recorded by the gas chromatograph. The experiment was recorded by the gas chromatograph. The experiment was repeated several times for different upstream pressures at the known fixed amount of methane. 154 The experimental gas flowrate was calculated by writing a mass balance of methane. _flmetuf_methene_ mole fraction of methane total flowrate of gases . (B.2) The mole fraction of methane was determined from the chromatograph reading (A), the flowrate of methane is the measured value of the flowmeter (Q1), the total flowrate is the flowrate of air and methane (Qa + Q1). Thus the flowrate of air (Qa) is given by Q -Q Q J—J-. (3.3) The experimental and the theoretical flowrates are compared in Figure 28. A discharge coefficient of 0.964 is listed in the literature for a similar critical nozzle design. The discharge coefficient is the ratio of the experimental to the theoretical value. Figure 23 shows an error of ~ 4%. The discharge coefficient used here is 0.97. The error is probably due to the combined error in the measurements of the nozzle diameter, flowrate of methane, and the reading of the chromatograph. 155 0.5“ I‘—~~ L.--— —--- :1 4 Figure 18 Configuration of the Critical Nozzles 156 cm ..noac.o nouoauan uo.oauuoz muoauauo a amsouna a: mo .33— 305 acoauouoofi. was #353:an am can»; Agmav 8:381 3282 no on as oe mm on L.Pb.PLb—-Lpb—-ppph-Lbb pr-L omF a 100m a ro¢u r IomN n 8205590 0 town 30:83:... ..I . (Jq/yo) sing MOU APPENDIX C APPENDIX C THE GAS ANALYSIS SYSTEM This Appendix describes the gas analysis system. This system is mainly composed of the individual instruments used to make continuous measurements of H20, C02, C0, 02, and total hydrocarbons (THC) in a sample stream. A schematic of the instruments along with the accessories needed to feed the desired sample stream into each analyzer is shown in Figure 2.4. The gas sample was pumped by a metal bellows pump model MB 601 HT driven by 0.75 HP motor. This pump was originally designed as a two stage pump. The two stages were used as separate pumps; one side was used as a suction pump for H20 and C02 analyzers, the other side was used to pump through the rest of the analyzers. A pressure regulator originally built in the THC analyzer was used to regulate the pressure before the CO and the 02 analyzers. A heating tape was used to heat the sample line in order to prevent condensation of HC and H20. Needle valves were used to control the flow through each analyzer which was monitored by rotameters. The sample gas supplied to the C0 analyzer was dried by cooling it to -5°C, using a Neslab U-cool immersion cooler as required by the instrument. The sample lines were conStructed from teflon tubing and stainless steel swagelock fittings. 157 158 C.1 H20 - CO Analyzer 2 An infrared AR-600 nondispersive infrared dual gas, dual range analyzer was used for H20 (Full scale 0 - 5.0%) and C02 (full scale 0 - 1.0%) analysis. The analyzer has a 100 ms response time and an accuracy of i 1% of full scale (F.S.). The sample gas is drawn through a 0.5 micron inline filter, and into the analyzer by pump suction. This arrangement yields the fastest response time. I This analyzer has certain unique features: (1) the measurements are not affected by the presence of HC in the sample gas, (2) it is equipped with a sophisticated differential signal processing techniques which eliminate zero drifts, (3) it is instrumented with an interconnection network between the C02 and H20 measurements which provides a spectral balance adjustment for the spectral interference range of the two gases, and finally (4) the analyzer is equipped with an internal calibration reference for setting the span control whenever a calibration gas is not available. 6.2 CO Analyzer A Beckman 3153 infrared analyzer was used to continuously analyze the concentration of C0 in the sample gas. The analysis is based on a differential measurement of the absorption of infrared energy. The analyzer has an adjustable full scale sensitivity, 90% response time in 0.5 second at 400 cc/min, and an accuracy of i 1% of full scale. 159 C.3 0 Analyzer A Beckman 778, a polaragraphic analyzer was used for 02 analysis. The concentration of 02 is determined by measuring the partial pressure of oxygen. Because of this characteristic, it is important that the sample gas mixture be kept under the same total pressure as when the instrument was calibrated. The analyzer has three measuring ranges (0- 5, 0-25, 0-100%), 90% response in 10 seconds, and an accuracy of i 1% F.S. 6.4 Total Hydrocarbon (THC) Analyzer A Beckman 400 was used for analysis of non-condensible hydrocarbons. The analyzer utilizes the flame ionization method. The sensor is a burner; a regulated flow of sample gas passes through a flame sustained by regulated flows of a fuel gas (40% H2 - 60 He) and air. The concentration of total hydrocarbons in the original sample is essentially measured by determining the rate at which carbon atoms enter the burner. The analyzer has an adjustable range from 1 ppm CH4 to 2% CH4, response time for 90% CH4 of 0.5 second with sample by-pass flow of 3 liters/minute, and an accuracy of i 1% F.S. LIST 01“ REFERENCES LIST OF REFERENCES Abu-Zaid, M., and Atreya, A. [1988a], ”Transient Cooling of Hot Porous and Non-Porous Ceramic Solids by Droplet Evaporation," Submitted to the J. of Heat Transfer. Abu-Zaid, M., and Atreya, A. [1988b], "Heat Transfer During Evaporation of A Water Droplet on A Heated Non-Porous Ceramic Solid," accepted for presentation at the Eastern Session of the Combustion Institute, Florida. Alvares, N, Blackshear, P., and Murty, K. [1969], "The Influence of Free Convection on the Ignition of Vertical Cellulosic Panels by Thermal Radiation," Central States Section, Combustion Institute Meeting, University of Minnesota. American Society of Mechanical Engineers Research Committee on Fluid Meters [1959], "Fluid Meters, Their Theory and Application,” 5th Ed., Published by ASME, New York. Arnberg, B. [1962], "Review of Critical Flowmeters for Gas Flow Measurements," J. of Basic Engineering, Vol. 84, 447. Aschenbrenner, A. [1983], "The Influence of Humidity on the Flowrate of Air Through Critical Flow Nozzles," Proceeding of Flomeko 1983 Imeko Conference on Flow Measurement. Budapest, Hungary. Atreya, A. [1983], "Pyrolysis, Ignition and Fire Spread on Horizontal Surfaces of Wood, ” Ph.D. Thesis, Harvard University, Cambridge, MA. Atreya, A. [1984], "Fire Growth on Horizontal Surfaces of Wood," Combustion Science and Technology, Vol. 39, 163. Atreya, A. [1985], "Effect of Water on Ignition of Cellulosic Materials,” Research Proposal, submitted to the National Bureau of Standards, Washington, D.C. Atreya, A. Carpentier, C., and Hankleroad, M. [1986], ”Effects of Sample Orientation on Piloted Ignition and Flame Spread," First International Symposium on Fire Safety Science, Vol. 1, 97. Atreya, A., and Wichman, I. [1987], "Heat and Mass Transfer During Piloted Ignition of Cellulosic Solids,” 2nd ASME-JSME Thermal Engineering Joint Conference, Vol. 1, 433. 160 161 Avedisian, C., and Koplik, J. [1987], ”Leidenfrost Boiling of Methanol Droplets on Hot Porous/Ceramic Surfaces," Int. J. Heat Mass Transfer, Vol. 30, No. 2, 379. Bamford, C., Crank, J., and Malan, D. [1946], "The Combustion of Wood," Proc. Cambridge Phil. Soc., Vol. 42, 166. Baumeister, H., and Simon, F. [1973], "Leidenfrost Temperature - Its Correlation for Liquid Metals, Crygens, Hydrocarbons, and Water," J. of Heat Transfer, Vol. 95, 166. Beck, J., and Arnold, K. [1977], ”Parameter Estimation In Engineering and Science," John Wiley and Sons, New York. Beck, J., Blackwell, B., and St. Clair, Jr., C. [1985], "Inverse Heat Conduction Ill-posed Problems,” A Wiley-Interscience Publication, New York. ' Bonancian, C., Comini, G., and Del Giudice, S. [1975], "Evaporation of Atomized Liquids on Hot Surfaces," Letters in Heat and Mass Transfer, Vol. 2, 401. Bonacina, C., Del Giudice, 8., and Comini, G., [1979], "Dropwise Evaporation,” J. of Heat Transfer, Vol. 101, 441. Butler, C., Martin, S., and Lai, W. [1956], "Thermal Radiation Damage to Cellulosic Materials - Part II - Ignition of Alpha-Cellulose by Square Wave Exposure,“ U.S. Naval Radiological Defense Lab. Tech. USNRDL-TR- 135, NS 081-001, AF SWP-906 Carslaw, H., and Jaeger, J. [1959], ”Conduction of Heat in Solids," Oxford University Press, Oxford, 2nd Ed., p. 88. Chapra, 8., and Canale, R. {1985], ”Numerical Methods for Engineers," McGraw-Hill Book Company, New York. di Marzo, M., and Trehan, A. [1986], "Transient Cooling of A Hot Surface by Droplets Evaporation,” National Bureau of Standards Interaency Report, NBS-GCR-86-516. di Marzo, M., Wang, 2., and Meng, W. [1987], ”Transient Cooling of A Hot Surface by Droplets Evaporation," National Bureau of Standards Report, NBS-GCR-87-534. Emmons, H. [1973], "Heat Transfer in A Fire," J. of Heat Transfer, Vol. 99, 145. Evans, 0., and di Marzo, M. [1986], ”Evaporation of'A Water Droplet Deposited on a Hot High Thermal Conductivity Solid Surface, National Bureau of Standards Report, NBSIR 86-3384. Garden, R. [1953], "Temperature Attained in Wood Exposed to High Intensity Thermal Radiation, "Fuel Research Lab., MIT, Tech. Report No. 3. 162 0005, A. [1952], "The Thermal Decomposition of Wood,” Wood Chemistry, Vol. 2, ACS Nomograph Series No. 97, 2nd Ed., Reinhold Pub. Co., New York. Gottfried, B., and Bell, K. [1961], "The Evaporation of Small Drops on A Flat Plate in the Film-Boiling Regime," Paper Press. at 54th Meeting of the A.I.Ch.E. Gottfried, B., Lee, C. and Bell, K. [1966], ”The Leidenfrost Phenomenon: Film Boiling of Liquid Droplets on A Flat Plate, Int. J. Heat Mass Transfer, 9, 1167. Gradshteyn, I., and Ryzhik, I. [1980], "Table of Integrals, Series, and Products, Academic Press. Inc., Orlando. Holman, J. [1984, "Experimental Methods For Engineers," McGraw-Hill Book Company, 4th Ed. Holman, J. [1981], ”Heat Transfer," McGraw-Hill Book Company, 5th Ed., New York. Incropera, F., and Dewitt, D. [1981], "Fundamentals of Heat Transfer," John Wiley and Sons, New York. Kanury, A. [1972], ”Ignition of Cellulosic Solids - A Review," Fire- Research Abstracts and Reviews, Vol. 14, 24. Kashiwagi, T. [1981], "Radiative Ignition Mechanism of Solid Fuels," Fire Safety Journal, Vol. 3, 185. Kashiwagi, T., Ohlemiller, J. and Werner, K. [1987], "Effects of External Radiant Flux and Ambient Oxygen Concentration on Nonflaming Gasification Rates and Evolved Products of White Pine,” J. Combustion and Flame, Vol. 69, 331. Keltner, N. [1973], "Transient Heat Flow in Half-Space Due to an Isothermal Disk on the Surface," J. of,Heat Transfer, Vol. 95, 412. Lee, C., Chaiken, R., and Singer, J. [1976], ”Charring Pyrolysis of Wood in Fires by Laser Simulation," Sixteenth Symposium (International) on Combustion, 1459. Lee, C., and Diehl, R. [1981], "Combustion of Erradiated Dry and Wet Oak," Combustion and Flame, Vol. 42, 123. Lee, L., Chen, J., and Nelson, R. [1985], "LiquidsSolid Contact Measurements Using A Surface Thermocouple Temperature Probe in Atmospheric Pool Boiling Water," Int. J. Heat Mass Transfer, Vol. 28, 1415. - Lee, T., Loftus, J., and Gross, D. [1974], "Effect of Moisture on Surface Flammability of Coated and Uncoated Cellulosic Materials," ASTM Special Technical Publication, No. 385. 163 Makino, K., and Michiyoshi, I. [1979], ”Effects of the Initial Size of Water Droplet on Its Evaporation on Heated Surfaces," Int. J. Heat Mass Transfer, Vol. 22, 979. Martin, S. [1965], "Diffusion-Controlled Ignition of Cellulosic Materials, by Intensive Radiant Energy," Tenth Symposium (International) on Combustion, 877. Martin, 8., Lincoln, R., and Ramstad, R. [1958], "Thermal Radiation Damage to Cellulosic Materials, Part IV. Influence of the Moisture Content and the Radiant Absorptivity of Cellulosic Materials on their Ignition Behavior,” Report No. USNRDL-TR-295, AFSWP-1117. Naval Radiological Defense Laboratory, San Francisco. Michiyoshi, I., and Makino, K. [1978], "Heat Transfer Characteristics of Evaporation of A Liquid Droplet on Heated Surfaces,” Int. J. Heat Mass Transfer, Vol. 21, 605. Mikkola, E., and Wichman, I. [1988], "The Thermal Ignition of Combustible Materials," Accepted for presentation at the Eastern Session of the Combustion Institute, Florida. Ohlemiller, T., Kashiwagi, T., and Werner, R., ”Wood Gasification at Fire Level Heat Fluxes,” Combustion and Flame, Vol. 69, 155. Parker, W. [1988], ”Prediction of the Heat Release Rate of Wood," Ph.D. Thesis, George Washington University. Pedersen, C. [1970], ”An Experimental study of the Dynamic Behavior and Heat Transfer Characteristics of Water Droplets Impinging Upon A Heated Surface,” Int. J. Heat Mass Transfer, Vol. 13, 369. Quintiere, J. [1981a], "A Simplified Theory for Generalizing Results from Radiant Panel Rate of Flame Spread Apparatus," Fire and Materials, Vol. 5, 51. Quintiere, J. [1988], "The Application of Flame Spread Theory to Predict Material Performance,” J. Research of National Bureau of Standards, Vol. 93, 61. Renkisizbulut, M., and Yuen, M. [1983], ”Experimental Study of Droplet Evaporation in A High-Temperature Air Stream," ASME J. of Heat Transfer, Vol. 103, 384. Rizza, J. [1981], "A Numerical Solution to Dropwise Evaporation," J. of Heat Transfer, Vol. 103, 501. Sauer, F. [1956], "The Charring of Wood During Exposure to Thermal Radiation: Correlation Analysis for Semi-infinite Solids," Intrim, T. AFSWP-868 U.S.D.A. Forest Service. 164 Schwenker, R., and Beck, L. [1963], “Study of the Pyrolytic Decomposition of Cellulose by Gas Chromatography," J. Polymer Sci. Part C, 1, No. 2, 331. Seki, M., Kawamura, H., and Sanokawa, K. [1978], ”Transient Temperature Profile of a Hot Wall Due to an Impinging Liquid Droplets,” J. of Heat Transfer, Vol. 100, 167. Simms, D [1960], "Ignition of Cellulosic Materials by Radiation," Combustion and Flame, Vol. 4, 293. Simms, D. [1963], ”On the Pilot Ignition of Wood by Radiation," Combustion and Flame, 7, 253. Simms, D., and Law, m. [1967], ”The Ignition of Wet and Dry Wood by Radiation,“ Combustion and Flame, Vol. 11, 377 Smith W., and King, J. [1970], "Surface Temperature of Materials During Radiant Heating to Ignition," J. Fire and Flammability, Vol. 1, 272. Spiegel, M. [1968], "Mathematical Handbook of Formulas and Tables," Schaum's Outline Series, McGraw-Hill Book Company, New York. Stamm, A. [1964], "Wood and Cellulose Science," The Ronald Press Company, New York. Steward, F. [1974], "Ignition Characteristics of Cellulosic Materials," Heat Transfer in Fires, Scripta Book Company, 379. Taylor, R., and Groot, H. [1985] "Thermophysical Properties of Cast MgO," A report to Sandia National Laboratories, Purdue University. Thomson, H., and Drysdale, D. [1988], "An Experimental Evaluation of Critical Surface Temperature as A Criterion for Piloted Ignition of Solid Fuels," Fire Safety Journal, Vol. 13, 185. Toda, S. [1972], "A Study of Mist Cooling (lst Report: Investigation of Mist Cooling),” Heat Transfer Japanese Research, Vol. 1 (3), 39. Tzeng, L., Wichman, I., and Atreya, A. [1988], ”A Theoretical Study of One-Dimensional Piloted Ignition," Submitted to Combustion and Flame. Vovelle, C., Akrich, R., and Delfau, J. [1984], ”Thermal Degradation of Solid materials Under A Variable Radiant Heat Flux," Twentieth Symposium (International) on Combustion, 1647. Wachters, L., Bonne, H., and Van Nouhuis, H. [1966], "The Heat Transfer from A Hot Horizontal Plate to Sessible Water Drops in the Spherodial State," Chemical Engineering Science, Vol. 21, 923. 165 Wachters, L., and Westerling, N. [1966], "The Heat Transfer from A Hot Wall to Impinging Water Drops in the Spheroidal State," Chemical Engineering Science, Vol. 21, 1047. Welker, J. [1970], "The Pyrolysis and Ignition of Cellulosic materials: A Literature Review," J. of Fire and Flammability, Vol. 1, 12. Wesson, H., Welker, J. and Sliepcevich, C. [1971], "The Piloted Ignition of Wood by Thermal Radiation," Combustion and Flame, Vol. 16, 303. Wichman, I. [1983], "Flame Spread in an Opposed Flow with a Linear Velocity Gradient,” Combustion and Flame, Vol. 50, 287. Wichman, 1., and Atreya, A. [1987], "A Simplified.Mode1 for the Pyrolysis of Charring Materials," Combustion and Flame, Vol. 68, 231. Yuen, M., and Chen, L. [1978], ”Heat-Transfer Measurements of Evaporating Liquid Droplets," Int. J. Heat Mass Transfer, Vol. 21, 537.