Limiting Oxygen Concentration Trend of ETFE-Insulated Wires under Microgravity

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

Limiting Oxygen Concentration of Flame Resistant Material in Microgravity Environment

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

Microgravity Opposed Flow Flame Spread in Polyvinyl Chloride Tubes

Effect of Environmental Variables on the Flammability of Fire Resistant Materials. Andres Felipe Osorio

GLOWING AND FLAMING AUTOIGNITION OF WOOD

QUIESCENT FLAME SPREAD OVER THICK FUELS IN MICROGRAVITY

Laminar Premixed Flames: Flame Structure

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

Lecture 8 Laminar Diffusion Flames: Diffusion Flamelet Theory

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

Lecture 7 Flame Extinction and Flamability Limits

ELEC9712 High Voltage Systems. 1.2 Heat transfer from electrical equipment

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

Lecture 28. Key words: Heat transfer, conduction, convection, radiation, furnace, heat transfer coefficient

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

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

Power generation properties of Direct Flame Fuel Cell (DFFC)

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

Home-built Apparatus for Measuring Thermal Conductivity of Glass and Polymer Materials

FLAME SPREAD AND EXTINCTION OVER SOLIDS IN BUOYANT AND FORCED CONCURRENT FLOWS: MODEL COMPUTATIONS AND COMPARISON WITH EXPERIMENTS

The Critical Velocity and the Fire Development

Improvements to Wire Bundle Thermal Modeling for Ampacity Determination

Iterative calculation of the heat transfer coefficient

Maximum Heat Transfer Density From Finned Tubes Cooled By Natural Convection

Effects of Convective Heat Transfer Coefficient in Prediction of Materials Properties from Cone Calorimeter Testing

Part II Combustion. Summary. F.A. Williams, T. Takeno, Y. Nakamura and V. Nayagam

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

2 nd Joint Summer School on Fuel Cell and Hydrogen Technology September 2012, Crete, Greece. Hydrogen fires

Solar Flat Plate Thermal Collector

Step-Wake Stabilized Jet Diffusion Flame in a Transverse Air Stream

Lecture 6 Asymptotic Structure for Four-Step Premixed Stoichiometric Methane Flames

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

Thermal Systems Design MARYLAND. Fundamentals of heat transfer Radiative equilibrium Surface properties Non-ideal effects

Quantitative Study of Fingering Pattern Created by Smoldering Combustion

UNIT FOUR SOLAR COLLECTORS

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

Natural convection heat transfer around a horizontal circular cylinder near an isothermal vertical wall

AN EXPERIMENTAL STUDY ON CRIB FIRES IN A CLOSED COMPARTMENT

In a high-temperature environment, such as the exit of the hot-air duct in the current counterflow

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

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

EXPERIMENTS WITH RELEASE AND IGNITION OF HYDROGEN GAS IN A 3 M LONG CHANNEL

Ch 17 Problem Set 31. A toaster is rated at 600 W when connected to a 120-V source. What current does the toaster carry, and what is its resistance?

Effects of radiative heat loss on the extinction of counterflow premixed H 2 air flames

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

COMPUTATIONAL SIMULATION OF SPHERICAL DIFFUSION FLAME DYNAMICS AND EXTINCTION IN MICROGRAVITY TIANYING XIA. A thesis submitted to the

INVESTIGATION OF VAPOR GENERATION INTO CAPILLARY STRUCTURES OF MINIATURE LOOP HEAT PIPES

The Electrodynamics of a Pair of PV Modules with Connected Building Resistance

Experimental and Theoretical Evaluation of the Overall Heat Loss Coefficient of a Vacuum Tube Solar Collector

Thermal Systems Design

Transient Heat Transfer Experiment. ME 331 Introduction to Heat Transfer. June 1 st, 2017

Exergy Analysis of a Compartment Fire

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

Heat Transfer Correlations for Small Closed End Heat Pipe with Special Vapor Chamber

RESEARCH PAPERS FACULTY OF MATERIALS SCIENCE AND TECHNOLOGY IN TRNAVA, SLOVAK UNIVERSITY OF TECHNOLOGY IN BRATISLAVA, 2017 Volume 25, Number 40

APPENDIX A: LAMINAR AND TURBULENT FLAME PROPAGATION IN HYDROGEN AIR STEAM MIXTURES*

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

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 Performance Characterization of Embedded Pulsating Heat Pipe Radiators by Infrared Thermography

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

Convective Mass Transfer

Thermo-Kinetic Model of Burning for Polymeric Materials

Simple Method to Predict Downward Heat Flux from Flame to Floor

Fire Safety in Space Investigating flame spread interaction over wires

Diffuse Thermal Radiation Focusing Device

Correlations between Microscale Combustion Calorimetry and Conventional Flammability Tests for Flame Retardant Wire and Cable Compounds

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

AN EXPERIMENTAL STUDY ON CRIB FIRES IN A CLOSED COMPARTMENT

Physics 196 Final Test Point

Natural convective heat transfer in trapezoidal enclosure of box-type solar cooker

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

Confirmation of Heat Generation during Hydrogenation of Oil

Fluid Mechanics Qualifying Examination Sample Exam 2

Experimental study on the explosion characteristics of methane-hydrogen/air mixtures

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

of Nebraska - Lincoln

Heriot-Watt University

Fall 2014 Qualifying Exam Thermodynamics Closed Book

International Fire Safety Symposium 2015

Well Stirred Reactor Stabilization of flames

Chapter 1: 20, 23, 35, 41, 68, 71, 76, 77, 80, 85, 90, 101, 103 and 104.

IGNITABILITY ANALYSIS USING THE CONE CALORIMETER AND LIFT APPARATUS

PERFORMANCE EVALUATION OF REFLECTIVE COATINGS ON ROOFTOP UNITS

Laminar Diffusion Flame Shapes Under Earth- Gravity and Microgravity Conditions

INVESTIGATION INTO RISE TIME OF BUOYANT FIRE PLUME FRONTS

ANALYTICAL MODEL OF FLAME SPREAD IN FULL- SCALE ROOM/CORNER TESTS (ISO9705)

High Beta Discharges with Hydrogen Storage Electrode Biasing in the Tohoku University Heliac

Experimental Evaluation of Natural Heat Transfer in Façade Integrated Triangular Enclosures

Y.; Kobayashi, H.; Inatani, Y. Citation Physics Procedia (2015), 67: Right article under the CC BY-NC-ND

Thermal Systems. What and How? Physical Mechanisms and Rate Equations Conservation of Energy Requirement Control Volume Surface Energy Balance

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

INDEX. (The index refers to the continuous pagination)

GCE AS/A level 1321/01 PHYSICS PH1 Motion, Energy and Charge

Scale and Transport Considerations on Piloted Ignition of PMMA

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

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

EXPERIMENTAL INVESTIGATION OF DIFFERENT TRACKING MODES OF THE PARABOLIC TROUGH COLLECTOR

January 2017 Qualifying Exam

Basic Study on the Generation of RF Plasmas in Premixed Oxy-combustion with Methane

Transcription:

Int. J. Microgravity Sci. Appl., 35 (1) (2018) 350104 DOI: 10.15011//jasma.35.350104 IIII Original Paper IIII Limiting Oxygen Concentration Trend of ETFE-Insulated Wires under Microgravity Ken MIZUTANI, Kyosuke MIYAMOTO, Nozomu HASHIMOTO, Yusuke KONNO and Osamu FUJITA Abstract This study investigated the flammability of the fire-resistant material ethylene-tetrafluoroethylene (ETFE) as insulation for copper wires under different flow velocity and gravity conditions. The limiting oxygen concentration (LOC) of flame spreading horizontally over the sample was investigated at external opposed flow velocities ranging from 0 to 200 mm/s under normal gravity (1g 0) and microgravity (μg 0). The LOC under μg 0 showed a U-shape, which has been reported in previous studies. A minimum LOC of approximately 26% was found at external flow velocities ranging 50 100 mm/s. An expanded heat balance model and radiation number for wire combustion (R rad,wire ) were proposed considering the heat conduction through the copper core, which is a notable feature of wire combustion. The U-shaped LOC curve was qualitatively explained in the low flow velocity region by this model and in the high flow velocity region by the Damköhler number. We also compared the LOC trend of ETFE with that of polyethylene (PE)-insulated wires reported in a previous study and demonstrated that the drop of LOC in ETFE was much larger than that of PE when the gravitational condition was changed from 1g 0 to μg 0 ( LOC). This large difference was explained by two factors. First, the rate of change of flame temperature with an increasing oxygen concentration is small at high oxygen concentrations. Second, the increase in heat input through the copper core owing to gravity change was larger for ETFE than for PE because of the difference in the rate of change in flame length along the copper core. Keyword(s): Flame spread, Extinction, Ethylene-tetrafluoroethylene (ETFE), Limiting oxygen concentration (LOC) Received 27 September 2017, Accepted 20 December 2017, Published 31 January 2018 1. Introduction As space exploration is being pursued by both national space agencies and private companies, the safety of astronauts has become more and more important. Fire is a key hazard during space missions. A large number of electrical devices and cables are installed in spacecraft, and these are a potential source of fire when short-circuiting or overloading. In fact, incidents involving electrical cables have already been reported in manned space missions 1) although none have led to fatalities. To prevent such incidents, materials used in spacecraft are regulated by NASA s fire safety standard STD-6001B 2). In the case of electrical wiring, Test 4 in this standard applies. However, this test only denotes whether it is a pass or fail and provides no quantitative values. Fujita 3, 4) noted that the test does not address actual flammability in microgravity (μg0) and suggested the necessity for a quantitative evaluation, for example, by applying the limiting oxygen index method, as defined in ISO 4589-2 5) or ASTM D2863 6). Fire in a spacecraft may be different from one on the ground because of the lack of gravity and particularly the lack of natural convection. In the early days, spacecraft materials were assumed to have lower flammability in μg0 because less oxygen is supplied owing to a reduced buoyant flow. However, recent researches 7, 8) have revealed that materials can become more flammable under certain conditions. Takahashi et al. 8, 9) investigated flame propagation over a thermally thin flat PMMA plate and constructed a flammability map that showed the limiting oxygen concentration (LOC) as a function of opposed flow velocity. The minimum oxygen concentration to sustain flame spread was identified for a thin flat PMMA sheet. The flammability map clearly showed a flammable region at a lower flow velocity and lower oxygen concentration than that at ground-level conditions in normal gravity (1g0). Other research has examined flame propagation over electrical wiring insulation. Work conducted by Takahashi 10) revealed that wire insulation under μg0 could be more flammable than under 1g0. They measured the LOC at a range of external opposed flow velocities (60 200 mm/s) using polyethylene (PE) insulation on copper (Cu) and nickel chromium (NiCr) wire. In this test, they observed that the LOC dropped by approximately 2 3% when shifting to the μg0 condition with both the Cu and NiCr cores. More recently, Osorio et al. conducted flame spread experiments with ethylenetetrafluoroethylene (ETFE)-insulated Cu wire 11). This is known as a fire resistant (FR) material and has a high limiting oxygen index (LOI: LOC under the condition specified by ISO 4589-2 5) defined for downward flame spread.) value of 31% 12), compared with 18% for PE 13). In this research, they measured LOC variations with different external radiative heat fluxes (0 25 kw/m 2 ) at a fixed external flow velocity (120 mm/s). They also Division of Mechanical and Space Engineering, Hokkaido University, Kita 13 Nishi 8, Kita-ku, Sapporo, Hokkaido 060-8628, Japan (E-mail: ofujita@eng.hokudai.ac.jp) 0915-3616/2018/35(1)/350104(6) 350104 1 2018 The Jpn. Soc. Microgravity Appl. http://www.jasma.info/journal/

Limiting Oxygen Concentration Trend of ETFE-Insulated Wires under Microgravity observed a drop in LOC with the shift to μg0. Interestingly, the drop in LOC (ΔLOC) was approximately 6%, which was clearly greater than that of PE-insulated wire. This result suggested that the LOC flammability gap of FR materials under 1g0 and μg0 is larger than that of PE despite the difference in insulation thickness: 0.15 mm for PE and 0.3 mm for ETFE. This suggests that we need to consider how the choice of insulation material affects the change in LOC under different gravity conditions. When we consider the extinction of flame propagation across solid fuels, two types of mechanism should be considered: blowoff in high flow velocity regions and quenching in low flow velocity regions. In the blow-off region, LOC decreases as flow velocity decreases. This is explained by the Damköhler number ( Da ) which is the ratio of residence time to characteristic chemical time as shown in Eq. (1) 8, 14, 15). Da t res t chem ~ α g V r 2 ρ gy O Aexp ( E RT f ) (1) where t res and t chem indicate the residence time of fuel and oxidizer in reaction zone and the chemical reaction time and A, E, R, T f, V r, Y O, α g, and ρ g are pre-exponential factor, activation energy, gas constant, flame temperature, velocity relative to the flame (V r = V g + V f, where V f and V g are flame spread rate and opposed-flow velocity), mass fraction of oxygen, thermal diffusivity of gas, and density of gas, respectively. Once Da drops to some critical value (Da < Da crit ), flame is extinguished because the residence time is too short. Conversely, in a quenching region, the LOC increases as flow velocity decreases. The flame extinction here is explained by the Radiation number (R rad ), which is the ratio of heat loss to heat input in a pre-heat region, as shown in Eq. (2) 8, 16, 17). R rad = ε(1 a abs )σ(t v 4 T 4 ) ρ g c g V r (T f T v ) where a abs, c g, T v, T, ε, and σ indicate absorption coefficient of gas, specific heat of gas, vaporization temperature of solid, ambient temperature, surface emissivity, and Stefan Boltzmann coefficient, respectively. The flame extinguishes when R rad reaches the critical value (R rad > R rad,crit ) as the radiative heat loss exceeds the heat input. In theory, these characteristics make the flammability boundary V-shaped. As the external flow velocity changes, the condition that defines the bottom of the V- shape should be a fundamental limit (the minimum limiting oxygen concentration (MLOC)). However, the MLOC estimated by the V-shaped model shows the smaller value than that obtained by the μg0 experiment. Because in reality, the boundary forms a U-shape because the region around the MLOC is affected by both extinction mechanisms 7 10). Takahashi et al. estimated the effect (2) of the Da on the heat transfer from flame to solid phase 9, 18) on the spreading of flame over a thin PMMA plate. They successfully constructed a theoretical model that gave a more accurate flammability limit than the ordinal V-shaped model. In the case of wire insulation combustion, a significant feature is heat transfer through the metal core, which can act either as a heat sink or heat source depending on the surrounding flow and temperature profile. Takahashi et al. 10) discussed the quenching and blow-off of wire combustion by qualitatively treating the wire as a flat plate and considering the core effect, thereby using the PE-insulated samples as laboratory material. Nonetheless, flammability limit estimation for electrical wiring is still remained as difficult problem because the heat transfer between insulation and metal core is unclear. To understand the mechanism, it is indispensable to grasp metal core temperature profile which is briefly mentioned but not discussed enough in the work conducted by Takahashi et al. 10). The present study focuses on the LOC of ETFE-insulated wires with opposed flow velocity variation. First, we carried out LOC measurements at different flow velocities to obtain the flow velocity dependencies of the LOC under 1g0 and μg0 conditions, and these results are presented in Section 3.1. In Section 3.2, we discuss the heat balance model for a wire system and propose the R rad for wire combustion considering the core effect in more detail. In section 3.3, we compare the trend of LOC of ETFE with that of PE 10) and discuss the ΔLOC differences between the materials. 2. Experimental Configuration Figure 1 shows a schematic of the combustion chamber that had a volume of approximately 46 L. The test section in this chamber has a 60 mm diameter and 250-mm-long flow duct made of glass tube and is located in the top section of the chamber. A manifold with a sample supply system on both sides of the duct connected the duct and bottom sections of the chamber. An air suction fan (Sanyo Denki 9GV0612P1H031) was attached at the bottom end of the manifold and produced left-to-right airflow of 0 200 mm/s inside the duct (Fig. 1). Airflow was made uniform by honeycomb flow straighteners at both manifolds. The sample wire (Sugita Densen Co., Ltd.) had a 0.5-mm-diameter copper Fig. 1 Schematic of combustion chamber. 350104 2

Ken MIZUTANI et al. core and 0.15-mm-thick ETFE coating. The sample geometry was the same as that in a previous study 10) to make comparison convenient. The sample wire was placed along the center line of the flow duct and ignited by a 0.5-mm-thick and 8-mm-diameter Kanthal coil downstream of the duct using 92 W (14.7 V, 6.3 A) of power and 12 s of energizing time. A programmable logic controller (Mitsubishi MELSEC FX2N-16MR) and power supply (Takasago EX-375L2) controlled ignition to realize opposedflow flame spreading over the wire insulation. The combustion products (carbon dioxide, water, and hydrogen fluoride) were absorbed by a zeolite filter and hydrogen fluoride filter placed upstream of the suction fan. Chamber pressure, oxygen concentration, and temperature were monitored by a pressure sensor (Nagano Keiki KP15-17G), an oxygen sensor (Jikco JKO- 25LII), and a T-type thermocouple placed in the bottom section of the chamber, respectively. All data were recorded by a data logger (Graphtec GL220 midi LOGGER dual). Before the start of each test run, the total pressure inside the chamber was set to 100 kpa for all conditions, and the oxygen concentration was set to the desired value using partial pressure. Once the ambient gas composition was at the target value, the chamber was completely sealed until the end of the test. The change in oxygen concentration by the combustion reaction during each test was less than 1%. For the μg0 experiments, the combustion chamber, control systems, power supply systems, and gas supply/evacuation systems were installed in a μg0 environment provided by parabolic flights of an airplane (Gulfstream-II) operated by Diamond Air Service Co. Ltd. in Nagoya, Japan. The duration of μg0 was approximately 20 s and acceleration was approximately 10 2 g0. The suction fan was turned on and the mixing fan was turned off 30 s before μg0 was induced. The igniter was energized 10 s before μg0 because of the ignition delay time, thus, allowing the flame spreading to begin immediately after μg0 conditions began. We defined propagation as the condition under which flame propagated by more than 100 mm under 1g0, and the condition under which the flame was sustained throughout the μg0 duration of 20 s. As the 20 s of μg0 duration was insufficient for 100 mm propagation, this distinction was necessary. These two different criteria were applied in a previous study and did not produce any significant differences 10, 19). When the extinction limit was determined according to the above definitions, the propagation and no propagation results overlapped near the flammability limit. 20) This was attributed to the stochastic nature of the extinction phenomenon, as reported by Olson et al. 21). In the present study, we defined an intermediate value between the maximum oxygen concentration with no propagation (NP max ) and the minimum oxygen concentration with propagation (P min ), as the LOC for each flow velocity. The gap between the NP max and the P min was plotted as an error bar. Fig. 2 LOC of ETFE-insulated copper wire (0.5 mm core diameter and 0.15 mm insulation thickness) under 1g0 and μg0. The triangles and squares represent the 1g0 and μg0 LOCs, respectively. Fig. 3 Typical flame shape of PE- and ETFE-insulated wires under 1g0. The oxygen concentration was 19% for PE and 38.5% for ETFE. The external flow velocity was 150 mm/s for both samples. Both conditions represented propagation. Fig. 4 Typical flame shapes of PE- and ETFE-insulated wires under μg0. The oxygen concentrations were 18% and 29% for PE and ETFE, respectively. The external flow velocity was 150 mm/s for both samples. Both conditions represented propagation. 350104 3

Limiting Oxygen Concentration Trend of ETFE-Insulated Wires under Microgravity 3. Results and Discussion 3.1 Experimental Results Figure 2 shows the LOC of ETFE for each external opposed flow velocity under 1g0 and μg0. The horizontal axis represents the external opposed flow velocity, and the vertical axis represents the ambient oxygen concentration. No U-shaped curve or MLOC was observed in the 1g0 test. The LOC had a constant value of approximately 37%, except at a 0 mm/s flow velocity, where the value was approximately 42%. Figure 3 compares the typical flame shapes of PE and ETFE under 1g0. In both cases, the flame wrapped around the wire and stretched upwards owing to buoyancy-induced flow. In contrast to the previous study 11), no drop in molten insulation was observed because the insulation used in this study was thinner (0.15 mm) than that in the previous study (0.3 mm). The LOCs at each external flow velocity in the μg0 experiment are shown in Fig. 2; a clear U-shape and MLOCs were observed. The MLOC was present at external flow velocities range 50 100 mm/s, with a value of approximately 26%. Note that this value is much lower than the LOI of ETFE, i.e., 31% 12). Figure 4 compares the typical flame shape of PE with ETFE under μg0. In contrast to 1g0, the flame for both PE and ETFE elongated downstream along the copper core. 3.2 U-Shaped Trend in Wire Combustion As discussed in Section 1, Takahashi et al. 10) have suggested a radiation number for wire combustion using the core effect. The radiation number R loss is given by Eq. (3): R loss = rad + sc gs = r sε s σ(t 4 v T 4 ) + λ s (T v T c )/ ln(r s r c ) λ g (T f T v )/ ln{1 + α g /(r s V r )} where, r, T c, and λ are the amount of heat transport per unit time, radius, the core temperature, and the thermal conductivity, respectively. Subscripts rad, sc, gs, c and s represent radiative heat loss, heat conduction from insulation to metal core, and heat conduction from a gas phase to insulation, the metal core, and the insulation, respectively. However, in this model, T c and heat transfer from insulation to core sc need to be specified to draw flammability map for electrical wire. This term is strongly affected by the overlap length of the flame and burned core wire 22). At a longer overlap length sc may increase and vice versa. These characteristics suggest that the ordinal control volume in the gas phase pre-heat region is insufficient to explain the heat balance of wire combustion. Therefore, we reconsidered the expanded heat balance with two additional regions: a pyrolysis region and a flame-wire overlap region, as region, as shown in Fig. 5. Subscripts gc heat conduction from gas phase to metal core. The terms pr and py following these subscripts represent (3) Fig. 5 Heat balance model for electrical wire with flame propagation. the pre-heat and pyrolysis regions, whereas cc1 and cc2 indicate heat conduction from the burned bare wire core to the pyrolysis region through the wire core and heat conduction from the pyrolysis region to the pre-heat region through the wire core, respectively. In this model, flame spreading at a constant velocity is assumed. When the right edge of the wire core is assumed to be the adiabatic boundary, based on the work done by Hu et al. 22), gc and cc1 can be equated. We also substituted sc,pr = cc2 and cc2 = sc,py + cc1. Finally, the overall heat balance can be expressed by Eq. (4): gc + gs,py + gs,pr rad,py rad,pr = req,pr + req,py where req,pr and req,py are the heat required to maintain preheat and pyrolysis length at a steady flame spread velocity of V f. We then normalized Eq. (4) by heat input to the pre-heat region, gc + gs,py + gs,pr req,py, following Takahashi s study for flat plates 8). We obtained the following expression: where (4) R rad,wire = 1 η wire (5) rad,py + rad,pr R rad,wire = and gc + gs,py + gs,pr req,py req,pr η wire = gc + gs,py + gs,pr req,py are the radiation number for wire and a non-dimensional flame spread velocity, respectively. When R rad,wire reached unity (or when η wire became zero), the flame was assumed to be extinguished, same as that in the flat-plate case. The observed U- shaped trend can then be explained consistently by Eqs. (1) and (5) as follows. In regions with a low flow velocity (less than ~100 mm/s in this study), radiative heat loss from the pre-heat region, rad,pr, 350104 4

Ken MIZUTANI et al. becomes relatively larger than the heat input from the increase in pre-heat length. Therefore, R rad,wire becomes larger and then the LOC becomes higher as the flow velocity decreases. In wire combustion, it is expected that the heat input to the wire core, gc, enhances flammability by decreasing R rad,wire. However, this term may not affect the shape of the LOC trend in the low flow velocity region because flame length does not significantly vary with flow velocity. Instead, gc may just shift down the LOC curve when the gravity condition changes from 1g0 to μg0. On the other hand, in the high flow velocity region (more than ~100 mm/s in this study), the effect of blow-off owing to the short residence time becomes dominant according to Eq. (1). Therefore, LOC increases with flow velocity. The LOC trend observed in the μg0 experiment is consistently explained by the theory above. 3.3 Comparison of ΔLOC for PE and ETFE Figure 6 represents the LOC curves for PE (from Takahashi et al. 10) ) and ETFE under 1g0 and μg0. The most notable feature of Fig. 6 the large ΔLOC of the ETFE. For example, the LOC at 100 mm/s of external flow velocity was 10% for ETFE compared with 3% for PE. A possible explanation for this difference is the oxygen concentration dependence of the flame temperature. As shown in Section 3.1, when flame spreads under μg0, it wraps around and stretches along the burned bare copper core, which has high thermal conductivity. This results in an increase of flame wire overlap length so that gc is increased under μg0 and the net heat input to the pre-heat region becomes larger than that under 1g0. We suggest that this difference (Δ in) causes ΔLOC. In particular, in the quenching (low flow velocity) region, a decrease in oxygen concentration is needed to cancel the Δ in by a decrease in flame temperature to maintain the net heat input in the pre-heat region (extinction occurs at a certain heat input according to Eq. (5)). The rate of temperature increase with the increase of oxygen concentration is known to become smaller as the oxygen concentration becomes larger, as shown in Fig. 7. This reflects the relation between the adiabatic flame temperature and the oxygen concentration, which is in line with the chemical equilibrium calculation from the NASA equilibrium program (CEA2) 23). As a result, material that has a higher LOC under 1g0 (ETFE in this study) requires a larger oxygen concentration drop to achieve the same temperature decrease ( ΔT f ). From experimental observations, the downstream flame stretching was much larger for ETFE than for PE. Thus, both Δ in and ΔT f of ETFE may be larger than those of PE, which may result in a larger ΔLOC for ETFE than for PE. In addition, if the external flow velocity is high enough to make buoyant flow negligible, the flame shape under 1g0 becomes similar to that under μg0, and the LOCs should be close to one another (as the ΔLOC becomes smaller). The decrease in ΔLOC with the increase in the flow velocity, observed in Fig. 6, supports this hypothesis. Fig. 6 Comparison of the LOCs of ETFE and PE under 1g0 and μg0. The LOC of PE is taken from Takahashi et al. 10). The cross shows the case where the flame does not go to extinction. Fig. 7 Example of adiabatic flame temperature variation with ambient oxygen concentration (calculated by CEA2 23) only chemical equilibrium was considered) 4. Conclusions This study investigated the LOC variation, i.e., the flammability boundary, of ETFE at different external flow velocities (0 200 mm/s) under 1g0 and μg0. Under μg0, the LOC curve showed a U-shape, often found in other flammability maps. The MLOC of the ETFE, at approximately 25%, was also observed at external flow velocities ranging 50 100 mm/s. The 25% of MLOC is much lower than ETFE s LOI value so that it supports its importance of flammability analysis on electrical wiring in μg0. The heat balance model for wire combustion has been developed to duplicate core effect by considering two additional control volumes: the pyrolysis and flame wire overlap regions. 350104 5

Limiting Oxygen Concentration Trend of ETFE-Insulated Wires under Microgravity The U-shaped LOC trend for wire samples has been explained by the radiation number in the new model ( R rad,wire ) and the Da from the extinction theory proposed by previous researchers 7-10), 14,17). Core temperature estimation included to this model can be a key of quantitative flammability assessment. A significant large LOC difference between 1g0 and μg0 (ΔLOC) was observed; at 100 mm/s of external flow velocity, ΔLOC was approximately 10% for ETFE and it was approximately 3% for PE. This difference has been explained by the relation between the flame temperature and the ambient oxygen concentration. It was predicted that high-loc materials (which are usually fireresistant) would have a higher ΔLOC than low-loc materials such as PE. The results suggest that flammability is enhanced by changes in gravity, particularly in the case of polymer metal composites, such as electrical wiring, because of a change in flame shape. Although this research has only provided a qualitative assessment, follow-up studies will contribute to the development of novel fire-safety standards for use in space. Acknowledgements This research is supported by JAXA as a candidate experiment for the third stage use of JEM/ISS titled Evaluation of gravity impact on combustion phenomenon of solid material towards higher fire safety (called as FLARE ). References 1) T. Limero, S. Wilson, S. Perlot and J. James: The Role of Environmental Health System Air Quality Monitors in Space Station Contingency Operations, 1992. 2) NASA STD 6001B, Flammability, Offgassing and Compatibility Requirements and Test Procedures, 2010. 3) O. Fujita: Proc. Combust. Inst., 35 (2015) 2487. 4) O. Fujita: Int. J. Microgravity Sci. Appl., 32 (2015). 5) ISO 4589-2, Plastic Determination of Burning Behaviour by Oxygen Index Part 2, Ambient Temperature Test. 6) ASTM D2863-13, Standard Test Method for Measuring the Minimum Oxygen Concentration to Support Candle-Like Combustion of Plastics (Oxygen Index), ASTM International, West Conshohocken, PA, 2013. 7) S.L. Olson: Combust. Sci. Technol. 76 (1991) 233. 8) S. Takahashi, T. Ebisawa, S. Bhattacharjee and T. Ihara: Proc. Combust. Inst., 35 (2015) 2535. 9) K. Maruta, K. Tsuboi and S. Takahashi: Int. J. Microgravity Sci. Appl., 34 (2017). 10) S. Takahashi, H. Ito, Y. Nakamura and O. Fujta: Combust. Flame, 160 (2013) 1900. 11) A.F. Osorio, K. Mizutani, C. Fernandez-Pello and O. Fujta: Proc. Combust. Inst., 35 (2015) 2683. 12) DuPont, DuPont Tefzel Fluoropolymer Resin Properties Handbook, 2003. 13) J.G. Quintiere: Fundamentals of Fire Phenomena, John Wiley & Sons, Ltd., West Sussex, England, 2006, p.441. 14) R.A. Altenkirch and M. Vedha-Nayagam: Symp. Combust., 22 (1989) 1495. 15) A.C. Fernandez-Pello, S.R. Ray and I. Glassman: Symp. Combust., 18 (1981) 579. 16) J.S. T ien: Combust. Flame, 65 (1986) 31. 17) S. Bhattacharjee, R. Ayala, K. Wakai and S. Takahashi: Proc. Combust. Inst., 30 (2005) 2279. 18) S. Takahashi, K. Tsuboi and T. Ihara: Int. J. Microgravity Sci. Appl., 31 (2014) 22. 19) O. Fujita: Int. J. Microgravity Sci. Appl., 28 (2011) 20) M. Kikuchi: Int. J. Microgravity Sci. Appl., 32(2015) 21) S.L. Olson, H.D. Beeson and C. Fernandez-Pello: in 44th Int. Conf. Environ. Syst., AIAA, Tucson, AZ, 2014, ICES 2014 179. 22) L. Hu, Y. Zhang, K. Yoshioka, H. Izumo and O. Fujta: Proc. Combust. Inst., 35 (2015) 2607. 23) S. Gordon and B.J. McBride: Computer Program for Calculation of Complex Chemical Equilibrium Compositions and Applications I. Analysis, 1994. 350104 6