Flame Spread and Extinction over Thermally Thick PMMA in Low Oxygen Concentration Flow

Similar documents
EXPERIMENTAL AND NUMERICAL STUDIES FOR FLAME SPREAD OVER A FINITE-LENGTH PMMA WITH RADIATION EFFECT

GLOWING AND FLAMING AUTOIGNITION OF WOOD

Microgravity Opposed Flow Flame Spread in Polyvinyl Chloride Tubes

Effect of Gravity on Radiative Heat Feedback on Small-Scale Pool Fires Using the Radiative Absorption Model

Limiting Oxygen Concentration Trend of ETFE-Insulated Wires under Microgravity

ONE-DIMENSIONAL MODEL OF PYROLYSIS AND IGNITION OF MEDIUM DENSITY FIBERBOARD SUBJECTED TO TRANSIENT IRRADIATION

Burner Tubing Specification for the Turbulent Ethylene Non-Premixed Jet Flame

A Methodology for Estimation of Local Heat Fluxes in Steady Laminar Boundary Layer Diffusion Flames

Opposed Flow Impact on Flame Spread Above Liquid Fuel Pools

QUIESCENT FLAME SPREAD OVER THICK FUELS IN MICROGRAVITY

Smoldering combustion of incense sticks - experiments and modeling

THE EFFECTS OF AIR-BORNE WATER MIST ON THE LOCAL BURNING RATE OF A PMMA PLATE IN BOUNDARY LAYER COMBUSTION

Density Field Measurement by Digital Laser Speckle Photography

Flame-Spread Characteristics of n-decane Droplet Arrays at Different Ambient Pressures in Microgravity

S. Kadowaki, S.H. Kim AND H. Pitsch. 1. Motivation and objectives

Critical Conditions for Water-based Suppression of Plastic Pool Fires. H. Li 1, A. S. Rangwala 1 and J.L. Torero 2

Interactions between oxygen permeation and homogeneous-phase fuel conversion on the sweep side of an ion transport membrane

EFFECTS OF FINITE SAMPLE WIDTH ON TRANSITION AND FLAME SPREAD IN MICROGRAVITY

IR temperature measurement, upward flame spread, vertical flat and corner walls, heat flux, heat transfer, empirical power law.

If there is convective heat transfer from outer surface to fluid maintained at T W.

STUDY OF HEAT TRANSFER MECHANISMS DURING THE LENS TM PROCESS

Theoretical Analysis of Overall Heat Loss Coefficient in a Flat Plate Solar Collector with an In-Built Energy Storage Using a Phase Change Material

Thermal Coatings for In-vacuum Radiation Cooling LIGO-T C R. Abbott, S. Waldman, Caltech 12 March, 2007

Rapid yet accurate measurement of mass diffusion coefficients by phase shifting interferometer

NATURAL CONVECTION HEAT TRANSFER CHARACTERISTICS OF KUR FUEL ASSEMBLY DURING LOSS OF COOLANT ACCIDENT

Fluid Flow and Heat Transfer of Combined Forced-Natural Convection around Vertical Plate Placed in Vertical Downward Flow of Water

Large scale diesel oil burns

The Planck Blackbody Equation and Atmospheric Radiative Transfer

FLAME PHOTOMETRY AIM INTRODUCTION

a. Fourier s law pertains to conductive heat transfer. A one-dimensional form of this law is below. Units are given in brackets.

IHTC DRAFT MEASUREMENT OF LIQUID FILM THICKNESS IN MICRO TUBE ANNULAR FLOW

LIQUID FILM THICKNESS OF OSCILLATING FLOW IN A MICRO TUBE

Quantitative Study of Fingering Pattern Created by Smoldering Combustion

INFLUENCE OF SURFACE EMISSIVITY AND OF LOW EMISSIVITY SHIELDS ON THE THERMAL PROPERTIES OF LOW DENSITY INSULATING MATERIALS

IGNITABILITY ANALYSIS USING THE CONE CALORIMETER AND LIFT APPARATUS

Yiguang Ju, Hongsheng Guo, Kaoru Maruta and Takashi Niioka. Institute of Fluid Science, Tohoku University, ABSTRACT

Department of Mechanical Engineering, University Visvesvaraya College of Engineering, Bangalore University, Bangalore, Karnataka, India

MCS 7 Chia Laguna, Cagliari, Sardinia, Italy, September 11-15, 2011

Lecture 8 Laminar Diffusion Flames: Diffusion Flamelet Theory

Experimental and Theoretical Study of the Ignition and Smoldering of Wood Including Convective Effects

Evaporation and heat transfer from thin water films on vertical panels in simulated fire conditions

of Nebraska - Lincoln

Supplemental Information. Storage and Recycling of Interfacial. Solar Steam Enthalpy

Experimental Study of 2D-Instabilities of Hydrogen Flames in Flat Layers

Combustion and Flame. An embedded upward flame spread model using 2D direct numerical simulations

3 - Atomic Absorption Spectroscopy

Chapter 9 NATURAL CONVECTION

Effects of Variation of the Flame Area and Natural Damping on Primary Acoustic Instability of Downward Propagating Flames in a Tube

Heat and Mass Transfer Unit-1 Conduction

Experimental Validation of Real Gas CO 2 Model near Critical Conditions

Chamber Development Plan and Chamber Simulation Experiments

THE EFFECT OF LIQUID FILM EVAPORATION ON FLOW BOILING HEAT TRANSFER IN A MICRO TUBE

Chapter 5 Test. Directions: Write the correct letter on the blank before each question.

Advanced Heat and Mass Transfer by Amir Faghri, Yuwen Zhang, and John R. Howell

Comparative Measurement in Speckle Interferometry using Fourier Transform

Dropwise Evaporative Cooling of a low Thermal Conductivity Solid

I. Measurements of soot - Laser induced incandescence, LII. spectroscopy, LIBS

Simple Method to Predict Downward Heat Flux from Flame to Floor

Florida Institute of Technology College of Engineering Department of Chemical Engineering

Critical mass flux for flaming ignition of dead, dry wood as a function of external radiant heat flux

Temperature time-history measurements in a shock tube using diode laser absorption of CO 2 near 2.7 µm

GRAVITY EFFECT ON THE DISTRIBUTION OF REFRIGERANT FLOW IN A MULTI-CIRCUITED CONDENSER

A Numerical Study of Solid Fuel Pyrolysis under Time Dependent Radiant Heat Flux Conditions

Development of measurement technique to evaluate thermal conductivity of thermoelectric Bi 2 Te 3 submicron thin films by photothermal radiometry

The energy performance of an airflow window

2. Illustration of Atmospheric Greenhouse Effect with Simple Models

Transactions on Engineering Sciences vol 5, 1994 WIT Press, ISSN

TIR100-2 Measurement of Thermal Emissivity

Digital Holographic Measurement of Nanometric Optical Excitation on Soft Matter by Optical Pressure and Photothermal Interactions

Heat Flux Distribution and Flame Shapes on the Inert Facade

Available online at ScienceDirect. Nuclear power plant explosions at Fukushima-Daiichi. Takashi Tsuruda*

Extensions to the Finite Element Technique for the Magneto-Thermal Analysis of Aged Oil Cooled-Insulated Power Transformers

The Effect of Nozzle Height on Cooling Heat Transfer from a Hot Steel Plate by an Impinging Liquid Jet

Numerical Simulation for Freeze Drying of Skimmed Milk with Moving Sublimation Front using Tri- Diagonal Matrix Algorithm

Earth s Energy Balance and the Atmosphere

EXPERIMENTAL STUDY OF MULTICELLULAR NATURAL CONVECTION IN A TALL AIR LAYER

Experimental Investigation of Adiabatic Film Cooling Effectiveness and Heat Transfer Coefficients over a Gas Turbine Blade Leading Edge Configuration

Analysis of Thermal Diffusivity of Metals using Lock-in Thermography

Warehouse Commodity Classification from Fundamental Principles. Part II: Flame Heights and Flame Spread

Iterative calculation of the heat transfer coefficient

KEYNOTE PAPER LIQUID FILM THICKNESS IN MICRO CHANNEL SLUG FLOW

Experimental and numerical study of the initial stages in the interaction process between a planar shock wave and a water column

TOPICAL PROBLEMS OF FLUID MECHANICS 97

Scale and Transport Considerations on Piloted Ignition of PMMA

PROBLEM 1.2 ( ) 25 C 15 C dx L 0.30 m Ambient air temperature, T2 (C)

Department of Energy Science & Engineering, IIT Bombay, Mumbai, India. *Corresponding author: Tel: ,

Fluorescence tracer technique for simultaneous temperature and equivalence ratio measurements in Diesel jets

17/02/02 IMAPS Thermal Management La Rochelle

CORRELATION BETWEEN HOT PLATE EMISSIVITY AND WAFER TEMPERATURE AT LOW TEMPERATURES

Comparison of competitive and non-competitive char formation in polymer combustion

arxiv: v1 [physics.chem-ph] 6 Oct 2011

Heat Flux and Thermal Response Measurements for Designing a Propane Fuel Fast Cook Off Test Apparatus

Experiment 1. Measurement of Thermal Conductivity of a Metal (Brass) Bar

Thermal Stress Cracking of Sliding Gate Plates

Advances in Fluid Mechanics and Heat & Mass Transfer

ABSTRACT A GENERALIZED MODEL FOR WALL FLAME HEAT FLUX DURING UPWARD FLAME SPREAD ON POLYMERS. Kevin T. Korver, Master of Science, 2015

Onset of Flow Instability in a Rectangular Channel Under Transversely Uniform and Non-uniform Heating

External Forced Convection :

International Fire Safety Symposium 2015

Thermodynamics and Energy. First Law of Thermodynamics and Energy Transfer Mechanisms. Applications of Thermodynamics

Transcription:

Flame Spread and Extinction over Thermally Thick PMMA in Low Oxygen Concentration Flow Y. KUDO, M. ITAKURA, Y. FUJITA, and A. ITO Faculty of Science and Technology Hirosaki University 3 Bunkyo-cho Hirosaki, 36-8561, Japan ABSTRACT This report discusses the mechanisms of flame spread and extinction over a thermally thick PMMA surface in a low-oxygen-concentration flow, based on experimental investigations. The temperature distribution in the condensed phase when the oxygen concentration in an air current decreases was measured in detail using holographic interferometry and thermography. The energy balance in the condensed phase was investigated based on the detailed temperature distribution, and the dominant heat transfer path in a low-oxygen-concentration flow was clarified. The net heat transfer rate in the direction normal to the surface, Q y is little changed with a decrease in oxygen concentration, while the heat transfer rate in the direction parallel to the surface, Q x decreases with oxygen concentration. The total heat transfer rate, Q T also decreases, due to the decrease in Q X. The radiative heat loss from unburned surface to surroundings, Q R increases with decreasing oxygen concentration. The balance of heat in the unburned fuel falls drastically due to the increase in the ratio of radiative heat loss to the total heat transfer rate, Q R /Q T. This leads to flame extinction. KEYWORDS: flame spread over solids, low oxygen concentration flow, flame extinction, holographic interferometry INTRODUCTION To understand the propagation and extinction mechanisms of a spreading flame along a combustible solid is important in predicting the behavior of the initial stage of building fires. The controlling mechanism of flame spread appears to differ with the surrounding conditions, such as the amount of ambient oxygen, gas flow velocity, direction of flame spread, and so on [1-4]. Recently, the influence of gravity in flame spread and extinction has also been investigated. It is reported that the extinction of a thermally thin solid fuel under microgravity conditions is significantly influenced by radiative heat loss from the unburned fuel surface [5]. The flame spread rate in air is controlled by the rate of thermal energy feedback from the flame to the unburned fuel surface [6,7]. Our latest thermally thick fuel experiments in opposed-airflow conditions [4] showed that the heat transfer rate from the gas phase is about 6% of the net total heat transfer rate in the non-opposed flow condition, while the heat transfer rate through the condensed-phase was over 8% near the extinction limit. Moreover, radiative heat loss was about 1% of the total heat transfer in flame spread over thermally thick solid fuel. When airflow velocity was increased, radiative heat loss was small enough not to have a major effect on extinction. From these results, it can be concluded that the extinction mechanism of a solid fuel changes depending on the surrounding conditions. FIRE SAFETY SCIENCE PROCEEDINGS OF THE EIGHTH INTERNATIONAL SYMPOSIUM, pp. 457-468 COPYRIGHT 25 INTERNATIONAL ASSOCIATION FOR FIRE SAFETY SCIENCE 457

Most recently constructed buildings are quite airtight and have efficient insulation. When a fire occurs in such a highly airtight compartment, it would seem that the oxygen concentration in the involved compartment is decreased. However, few studies have been published dealing with flame spread over a solid fuel surface under low-oxygenconcentration conditions, and at present the details of the effect of oxygen concentration on flame spread over a solid fuel are not adequately understood [8]. In particular, the effect of the oxygen concentration on heat transfer from flame to unburned surface is unknown. Flame spread in microgravity conditions may be simulated by restricting the supply of oxygen with a low oxygen concentration, because flame spread in microgravity is controlled by the supply of oxygen. In order to elucidate the heat transfer mechanism and determine the heat transfer path, we must know the detailed temperature structure both in the gas and condensed phases. Usually, thermocouples have been used for measuring temperature distributions [9-11]. However, with this technique it is difficult to determine the temperature gradient accurately due to limited spatial resolution. We recently used holographic interferometry (HI) combined with thermography to determine the temperature structure within a transparent material [3,4,12]. Careful use of these techniques can determine the heat flux distribution with high spatial resolution. In this study, the detailed temperature distribution in a thick slab of polymethyl methacrylate (PMMA) during horizontal flame spread in low-oxygen-concentration opposed-flow conditions was measured. Based on the temperature structure in the condensed phase, the dominant heat transfer path and the extinction mechanism under low-oxygen conditions are discussed. EXPERIMENTAL PROCEDURE AND APPARATUS The wind tunnel used in this study, and details of the test section, are illustrated in Fig. 1. The wind tunnel is 4 mm long by 17 mm high and 34 mm wide, and is made of Pyrex. A rectification tube was installed ahead of the test section in order to provide uniform, laminar flow. The low-oxygen-concentration gas was made by mixing compressed dry air and nitrogen gas from a pressure tank, and supplied upstream from the rectification tube. The gas supplied to the wind tunnel was sampled upstream from the rectification tube using a gas-tight syringe. The oxygen concentration in the sampled gas was measured by gas chromatography. The opposed-flow rate was measured with a Pitot tube. By traversing the Pitot tube from top to bottom across the center of the wind tunnel, the velocity profile and the average velocity, u _, were obtained. Average airflow velocity in this experiment was fixed at.27 m/s (Reynolds number is about 47). We selected samples that were 1 mm wide, 25 mm thick, and 1 mm long. Pyrex plates 1 mm wide were installed on both sides of the PMMA sample with a small space between the sample and the Pyrex plate. This space is close to the quenching distance, about 2 mm, which prevents flames passing over both sides of the sample and ensures twodimensional flame spread. These Pyrex plates affect the fringe shift only minimally because their thermo-optic coefficient and thermal expansion coefficient are much lower than those of PMMA. The temperature distribution on the PMMA surface was recorded by thermography through an NaCl window in the top of the test section. The IR camera, with a 7.5-13 µm spectral range, has a temperature resolution of.1 C and a time resolution of 3 ms. The emissivity of the PMMA at 7.5-13 µm was determined to be.95 by calibrating the 458

temperature measured by thermography against that measured by a Chromel-Alumel thermocouple. Its emissivity is almost independent of its surface temperature, ranging from.9 to.99 below the glass transition temperature (about 15 C). The optical setup and the holographic procedure were described in Ref. [13]. Gas chromatography Pyrex Plate Sample Sampling Thermography NaCl Window 2mm 1mm N 2 Air Chamber Laser (for Holography) Sample(PMMA) Rectification tube Fig. 1. Schematic illustration of the experimental apparatus. The details of the HI measuring technique have been described previous papers [3,4,11], so only a brief description is presented here. The relationship between the temperature and the fringe number, N, can be expressed as B T T T T ( dnp / dt) dt + αb( T T ) [ ( dn p / dt ) dt + n p na ] = λ(2n 1) / 2 α (1) where B is the width of the sample. The term dn p /dt is the thermo-optic coefficient of the material; λ is the linear thermal expansion coefficient, n p is the refractive index of the material at the ambient temperature, n a is the refractive index of air at the ambient temperature, and k is the wavelength of the light. The thermo-optic coefficient and the linear thermal expansion coefficient of PMMA have been recorded up to a little over the glass transition temperature [12]. Above the glass transition temperature, accurate data for these coefficients are not currently available. Therefore, the surface temperature distribution measured by the IR camera was used as the reference temperature to analyze interference fringes above the glass transition temperature. For recording of holograms, the real-time method [14] was used. The optical setup and the holographic procedure were described in Ref. [13]. RESULTS AND DISCUSSION Temperature Distribution in Condensed Phase A schematic diagram of flame spread, including the coordinates, is shown in Fig. 2. The vaporization front on the sample surface is defined as x = in this study, where x is a coordinate parallel to the sample surface. The sample surface is defined as y =, where y is the normal coordinate. The flame leading edge gradually retreats and the blue flame becomes small as the oxygen concentration decreases. With further reduction in the oxygen concentration, the flame retreated and extinguished. The oxygen concentration at 459

the extinction limit, Y O2limit was 18.4 vol.% in this test. Typical interferograms and isotherm contours within the condensed phase are shown in Fig. 3. In these interferograms, it can be seen that part of the fringe pattern appears to be different from the rest. This is caused by the glass transition in the PMMA sample. The thermo-optic coefficient of PMMA decreases with increasing temperature (negative value), but the thermal expansion coefficient increases with increasing temperature [11]. Near the glass transition temperature, about 15 C, the term including these coefficients becomes zero. Therefore, the fringe pattern appears to be at a singular point at the glass transition temperature of PMMA. Except for this layer, the fringes shown in Fig. 2 correspond to isotherms. The interval of the x direction of isotherm contours increases as oxygen concentration decreases. In other words, the interval of fringes increases in the x-direction. This means that the temperature gradient becomes small and the heat flux conducted to the unburned area through the condensed phase decreases. The interferogram photograph near the blow-off limit (u _ =.65 m/s) and isotherm contours are shown in Fig. 3e for comparison with those of oxygen concentration near the extinction limit (Fig. 3d). When airflow velocity is increased, the interval of the x direction of isotherm contours increases only slightly and it is suggested that the temperature gradient and the heat flux likewise barely increase. This is quite different than the effect of oxygen concentration. u=.27m/s x= V q y Flame -x y= y L PMMA q x Fig. 2. Flame spread model. Interferogram photograph Gas phase (cm).2 Surface -.2 -.4 Condensed phase 4 ( o C) 38 3 26 22 18 14 1 -.6 5 6 8 -.8 -.5.3 (cm) Isotherm contours Fig. 3a. Typical interferogram photograph and isotherm contours at Y O2 =21. vol.% and u _ =.27 m/s case. 46

Gas phase Surface Interferogram photograph Gas phase (cm) 1.2 8 1418 ( o C) Surface -.2 38 3 -.4 26 -.6 4 5 6 22 -.8-1.5-1. -.5.3 (cm) Isotherm contours Fig. 3b. Typical interferogram photograph and isotherm contours at Y O2 =19.4 vol.% and u _ =.27 m/s case. Gas phase Surface Interferogram photograph Gas phase (cm).2 1 14 18 ( o C) Surface -.2 38 3 -.4 -.6 26 5 6 8 22 -.8-2. -1.5-1. -.5.5 Isotherm contours (cm) Fig. 3c. Typical interferogram photograph and isotherm contours at Y O2 =19. vol.% and u _ =.27 m/s case. 461

Surface Gas phase Condensed Surface phase Gas phase (cm).2 Surface -.2 -.4 Interferogram photograph Interferogram photograph 1 14 ( o C) 38 3 26 22 18 -.6 5 6 8 -.8-2. -1.5-1. -.5.5 Isotherm contours (cm) Fig. 3d. Typical interferogram photograph and isotherm contours at Y O2 =18.4 vol.% and u _ =.27m/s case (Extinction limit condition). Interferogram photograph Gas phase (cm).2 Surface -.2 ( o C) 38 -.4 26 22 -.6 18 5 6 14 -.8 -.5.3 1 8 (cm) Isotherm contours Fig. 3e. Typical interferogram photograph and isotherm contours at Y O2 =21. vol.% and u _ =.65 m/s case (Blow-off condition). Surface temperature profiles across the sample width at its center, measured by thermography at various oxygen concentrations, are shown in Fig. 4. The surface temperature distribution in the preheat zone tends to increase with decreasing oxygen concentration. As the oxygen concentration decreases, the chemical reaction rate is 462

decreased and the flame temperature is decreased. The heat feedback from the flame to the unburned fuel is decreased, and the flame spread rate, V, is also decreased. Due to the decrease in V, the flame residence time required for heat feedback into the unburned fuel from the flame is increased. The increased residence time increases the heat conducted to the front of the unburned fuel. Therefore the preheating time becomes longer because the flame stays longer in front of the preheat zone due to the slow flame spread rate, and this causes the changed temperature distribution shown in Fig. 4. 4 Y O2 =19.4vol.% Surface temperature T S ( o C) 3 2 1 Y O2 =21.vol.% Y O2 =19.vol.% Y O2 =18.4vol.% -5-1 -15 x(mm) Fig. 4. Surface temperature distribution measured by thermography. Heat Flux and Heat Transfer Rate For quantitative evaluation of the heat feedback, the conductive heat flux in the direction normal to the surface at y =, q y, the conductive heat flux in the direction parallel to the surface at x = (the vaporization front on the sample surface), q x, and radiative heat flux from surface to surroundings, q r, are calculated based on the flame spread model shown in Fig. 2. q T T λ + q q (2) y = s = λg y y y= + y= rf r T qx = λ (3) s x x= q r 4 s 4 = ε σ ( T T ) (4) s where λ s and λ g are the thermal conductivities of the sample and gas, respectively. q rf is the radiative heat flux from the flame to the unburned preheating surface, ε s is the emissivity of the sample surface, and σ is the Stefan-Boltzmann constant. T s is the surface temperature of the sample, and T is the ambient temperature. The first term on the right-hand side of Eq. 2 represents the convective/conductive heat flux from the gas 463

phase. The third term represents the radiative heat loss from the surface to surroundings. Therefore, q y is the net heat flux from the gas phase at the sample surface. By taking the spatial derivative of the temperature distributions, conductive heat flux in the direction normal to the surface, q y, and the direction parallel to the surface, q x, can be calculated. Since holographic interferometry has very high spatial resolution, the exact values of dt/dx and dt/dy were obtained, and they were compared with the values obtained from the temperature distribution measured using the thermocouple. The calculated values of q y are plotted in Fig. 5. The distribution of heat flux q y became smooth and the absolute value at the vaporization front, x = also became small as the oxygen concentration was reduced. The net heat flux from the gas phase, q y at x = in Y O2 = 21. vol.% was about 95 kw/m 2, becoming about 6 kw/m 2 in Y O2limit = 18.4 vol.%, which is near the extinction limit. The heat conducted through the sample at the x = plane, q x is plotted in Fig. 6. In the Y O2 = 21 vol.% case, q x increased rapidly near the surface. The values of q x on the surface were about 7 kw/m 2. The value of q x on the surface of the solid decreased as oxygen concentration decreased. The q x at y = in oxygen concentrations near the extinction limit became about 5 kw/m 2. The radiative heat flux from the sample surface to the surroundings, q r, which was calculated from the measured temperature distribution, is plotted in Fig. 7. It was found that the radiative heat flux far from the vaporization front was larger at lower oxygen concentrations. 1 Heat flux q y (kw/m 2 ) 8 6 4 2 Y O2 =21.vol.% Y O2 =19.4vol.% Y O2 =18.4vol.% Y O2 =19.vol.% -.1 -.2 -.3 -.4 -.5 -.6 -.7 -.8 -.9 x (cm) Fig. 5. Distribution of heat flux in the direction normal to the surface, q y. 1 Heat flux q x (kw/m 2 ) 8 6 4 2 Y O2 =19.4vol.% Y O2 =18.4vol.% Y O2 =21.vol.% Y O2 =19.vol.% -.9 -.8 -.7 -.6 -.5 -.4 -.3 -.2 -.1 y(cm) Fig. 6. Distribution of heat flux in the direction parallel to the surface, q x. 464

Heat flux q r (kw/m 2 ) 1 8 6 4 2 Y O2 =21.vol.% Y O2 =19.4vol.% Y O2 =19.vol.% Y O2 =18.4vol.% -.1 -.2 -.3 -.4 -.5 -.6 -.7 -.8 -.9 x (cm) Fig. 7. Distribution of heat flux from surface to surroundings, q r. The energy balance in the condensed phase is constructed by using the temperature and heat flux distributions. The control volume is the layer from x = to - and y = to -L (L is the depth of the sample). The total heat transfer rate into the control volume in the direction normal to the surface, Q y, is T Qy = λ = s dx q ydx (5) y y= and the amount of heat transfer into the control volume through the plane x = in the direction parallel to the surface, Q x,is T Qx = λ =. L s dy q L x dy (6) x x= The rate of heat loss from the rear sample surface is Q L T = λ s dx. (7) y y = L Then, the total rate of heat transferred into the control volume is Q T = Q + Q Q (8) y x L Q L is small in comparison with the other values, so it is assumed to be negligible. The rate of radiative heat loss from the control volume is Q = 4 4 R sσ ( Ts T ) dx. ε (9) 465

The Q R given in Eq. 9 is already included in the Q y given in Eq. 5 because the given q y in Eq. 2 included the radiative heat loss from the surface to the surroundings. The following thermal properties were used to calculate the above integrals: thermal conductivity of the sample k s = 2.1x1-1 W/m/K [14] and density ρ s = 119 kg/m 3 [16]. The calculated values of Q y, Q x, Q T, and Q R for each oxygen concentration are plotted in Fig. 8. When oxygen concentration decreases below Y O2 = 21. vol.%, Q y is changed, while Q x decreases near Y O2 = 19. vol.% concentration. Q T also decreases due to the decrease in Q X. On the other hand, Q R increases with decreasing oxygen concentration. The relationship between the ratio Q R /Q T and oxygen concentration is shown in Fig. 9..1 Q T Heat transfer rate Q (kw/m).8.6.4.2 Extinction limit Q y Q R Q x 1..8 18 19 2 21 22 Oxygen concentration Y O2 (vol.%) Fig. 8. Relationship between heat transfer rate and oxygen concentration. Extinction limit Q R /Q T (-).6.4.2 18 19 2 21 22 Oxygen concentration Y O2 (vol.%) Fig. 9. Ratio Q R /Q T as a function of oxygen concentration. In the case of Y O2 = 21.vol.%, the radiative heat loss, Q R was about 13% of the total heat transfer, Q T. The ratio Q R /Q T increases as oxygen concentration is decreased. At Y O2limit = 18.4 vol.% oxygen (near the extinction limit), Q R is about 6% of Q T. At the low oxygen concentration near the extinction limit, it is difficult for the sample to receive sufficient heat feedback to maintain flame spread. The balance of heat at the sample surface falls 466

drastically due to the increased radiative heat loss, so eventually the flame must be extinguished. CONCLUSIONS The detailed temperature distribution near the flame leading edge during flame spread over PMMA in low-oxygen-concentration opposed-flow conditions was experimentally examined. Based on the temperature distribution, the energy balance in the condensed phase was investigated. The conclusions obtained from this study are summarized here. 1. The net heat transfer rate normal to the surface, Q y, the net heat transfer rate parallel to the surface, Q x, and the radiative heat loss from the surface, Q R were obtained from measured isotherm contours and surface temperature profiles. The Q y is little changing with decreasing oxygen concentration, while Q x decreases near oxygen concentration, Y O2 =19. vol.%. The total heat transfer rate, Q T also decreases due to the decrease in Q x. 2. The radiative heat loss from surface to surroundings, Q R increases with decreasing oxygen concentration because the surface temperature of the sample far from the flame front is higher at a lower oxygen concentration. 3. Near the extinction limit, there is barely enough heat feedback to the sample to support flame spread under low- oxygen-concentration conditions. The heat balance collapses as the ratio of radiative heat loss to the total heat transfer rate (Q R /Q T ) increases, so eventually the flame must be extinguished. REFERENCES [1] Magee, R.S., and McAlevy III, R.F., The Mechanism of Flame Spread, J. Fire and Flammability, 2, pp. 271, (1971). [2] Ray, S.R., and Glassmann, I., The Detailed Processes involved in Flame Spread over Solid Fuels, Combust. Sci. Technol, 32:33-48 (1983). [3] Ito, A. and Kashiwagi, T., Characterization of Flame Spread over PMMA using Holographic Interferometry Sample Orientation Effects, Combust. Flame, 71, pp. 189-24, (1988). [4] Kudo, Y., and Ito, A., Propagation and Extinction Mechanisms of Opposed- Flow Flame Spread over PMMA, Proc. Combust. Inst., 29, pp. 237-243, (22). [5] Takahashi, S., Kondou, M., Wakai, K., and Bhattacharjee, Effect of Radiation Loss on Flame Spread over Thin PMMA Sheet in Microgravity, S., Proc. Combust. Inst., 29, pp. 2579-2586, (22). [6] Williams, F. A., Mechanism of Fire Spread, Proc. Combust. Inst., 16, pp. 1281-1294, (1977). [7] Fernandez-Pello, A.C., and Hirano, T., Controlling Mechanisms of Flame Spread, Combust. Sci. Technol., 32, pp. 1-31, (1983). [8] Ross, H.D., Microgravity Combustion: Fire in Free Fall, Academic Press, 21, pp. 326-329. [9] Fernandez-Pello, A.C. and Williams, F.A., Laminar Flame Spread over PMMA Surfaces, Proc. Combust. Inst., 15, pp. 217-231, (1975). 467

[1] Hirano, T., Koshida, T., and Akita, K., Flame Spread Mechanism over PMMA Surface, Bull. Japan Assoc. Fire Sci. Eng., 27, pp. 33-39, (1977). [11] Ray, S.R., Fernandez-Pello, A.C. and Glassmann, I., A Study of the Heat Transfer Mechanisms in Horizontal Flame Propagation, J. Heat Transfer, ASME, 12, pp. 357-363, (198). [12] Ito, A. and Kashiwagi, T., Measurement Technique for Determining the Temperature Distribution in a Transparent Solid Using Holographic Interferometry, Appl. Opt., 29:954-958, (1987). [13] Ito, A., and Kashiwagi, T., Temperature Measurements in PMMA during Downward Flame Spread in Air using Holographic Interferometry, Proc. Combust. Inst., 21, pp. 65-74, (1986). [14] Vest, C.M., Holographic Interferometry, Willy, New York, 1979, Chapter 6. [15] Hand, S.D. and Horsfall, F.J., A Thermal Conductivity Apparatus for Solid and Molten Polymers, J. Phy. E: Sci. Instrum., 8:687-69, (1975). [16] Bares, V. and Wunderlich, B., Heat Capacity of Molten Polymers, J. Polym. Sci., 11:861-873, (1973). 468