Available online at ScienceDirect

Similar documents
HEAT RELEASE RATES FOR NUCLEAR POWER PLANT ELECTRICAL ENCLOSURE FIRES

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

Available online at ScienceDirect. Procedia Engineering 121 (2015 )

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

IGNITABILITY ANALYSIS USING THE CONE CALORIMETER AND LIFT APPARATUS

Documentation of the Solutions to the SFPE Heat Transfer Verification Cases

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

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

Ampacity simulation of a high voltage cable to connecting off shore wind farms

Simple Method to Predict Downward Heat Flux from Flame to Floor

Thermal Cube. Custom-built Heat Flux Sensor

Available online at ScienceDirect. Procedia Engineering 157 (2016 )

Simulating Fuel Spill Fires Under the Wing of an Aircraft

Available online at ScienceDirect. Energy Procedia 69 (2015 )

Thermal Cube. Custom-built Heat Flux Sensor

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

GLOWING AND FLAMING AUTOIGNITION OF WOOD

An Experimental Study on Glass Cracking and Fallout by Radiant Heat Exposure

ELEC9712 High Voltage Systems. 1.2 Heat transfer from electrical equipment

Fire scenarios modelling for the safe design of a passenger rail carriage

Pin Fin Lab Report Example. Names. ME331 Lab

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

Safety Assessment on the Storage of Irradiated Graphite Waste Produced from the Decommissioning of KRR-2

Diffuse Thermal Radiation Focusing Device

SEMICONDUCTOR THERMAL MEASUREMENT PROCEDURE

Autumn 2005 THERMODYNAMICS. Time: 3 Hours

Available online at ScienceDirect. XVII International Colloquium on Mechanical Fatigue of Metals (ICMFM17)

Laboratory 12: Three Thermodynamics Experiments

MECH 375, Heat Transfer Handout #5: Unsteady Conduction

Available online at ScienceDirect. Physics Procedia 69 (2015 )

Thermal Calculation for Linear Regulator

Fires in Vehicles - FIVE 2014

EXPERIMENTAL FLAME HEAT TRANSFER CORRELATIONS FOR A STEEL COLUMN ADJACENT TO AND SURROUNDED BY A POOL FIRE

THERMAL ANALYSIS OF UMBILICAL CABLES

Title Process of Ethanol in Activated Car. Citation Physics Procedia (2015), 69:

EPR Insulated, Heavy Duty HOFR Sheathed Flexible Cables

Technical Notes. Introduction. PCB (printed circuit board) Design. Issue 1 January 2010

Theoretical and Experimental Studies on Transient Heat Transfer for Forced Convection Flow of Helium Gas over a Horizontal Cylinder

HEAT FLOW SENSORS MF-180 MF-180M MF-190 MF-200 INSTRUCTION MANUAL

ENSC 388. Assignment #8

N. Lemcoff 1 and S.Wyatt 2. Rensselaer Polytechnic Institute Hartford. Alstom Power

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

TECHNICAL INFORMATION

Pyrolysis Modelling of PVC Cable Materials

THERMAL ANALYSIS OF A SPENT FUEL TRANSPORTATION CASK

INFLUENCE OF ZINC COATING TO A TEMPERATURE OF STEEL MEMBERS IN FIRE

Coolant. Circuits Chip

METHOD OF IN-SITU MEASUREMENT OF THERMAL INSULATION PERFORMANCE OF BUILDING ELEMENTS USING INFRARED CAMERA

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

Available online at ScienceDirect. Procedia Engineering 90 (2014 )

Peltier Application Note

MASSACHUSETTS INSTITUTE OF TECHNOLOGY DEPARTMENT OF MECHANICAL ENGINEERING Thermal-Fluids Engineering I

Available online at ScienceDirect. Energy Procedia 57 (2014 ) ISES Solar World Congress

Edinburgh Research Explorer

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

TEST METHOD FOR STILL- AND FORCED-AIR JUNCTION-TO- AMBIENT THERMAL RESISTANCE MEASUREMENTS OF INTEGRATED CIRCUIT PACKAGES

Florida Institute of Technology College of Engineering Department of Chemical Engineering

Characterization of high temperature solar thermal selective absorber coatings at operation temperature

University of New Mexico Mechanical Engineering Spring 2012 PhD qualifying examination Heat Transfer

Single Core XLPE/LSF/AWA/LSF Power Cable

SOLUTIONS. Question 6-12:

ISO INTERNATIONAL STANDARD

Available online at ScienceDirect. Energy Procedia 75 (2015 ) Multiphysics Simulations of a Thermoelectric Generator

Available online at ScienceDirect. Reynaldo Zoro*

I, THERMAL ANALYSIS OF BUS-BAR FOR SWITCH BOARD

Superconducting Magnet with a Minimal Steel Yoke for the Future Circular Collider Detector

Available online at ScienceDirect. Procedia Engineering 133 (2015 ) th Fatigue Design conference, Fatigue Design 2015

Thermocouple Calibrations and Heat Transfer Coefficients

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

AN EXPERIMENTAL INVESTIGATION OF BOILING HEAT CONVECTION WITH RADIAL FLOW IN A FRACTURE

Introduction to Heat and Mass Transfer. Week 5

Available online at ScienceDirect. Andrej Godina*, Bojan Acko

INVESTIGATION INTO RISE TIME OF BUOYANT FIRE PLUME FRONTS

Sapphire Fiber-Optic Temperature Sensor Based on Black-Body Radiation Law

Experimental and numerical investigations of heat transfer and thermal efficiency of an infrared gas stove

Mechanical Engineering Division June 1, 2010

Fire Resistant Armoured Power Cable

Direct fired strip preheating

International Fire Safety Symposium 2015

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

CS Series. Capacitance Type. Type. Proximity Sensor. Features

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

Characterisation of Deposits on Membrane Walls of Steam Generators by Heat Flux Density Measurement

L 305/8 Official Journal of the European Union (Acts whose publication is not obligatory) COMMISSION

( ) PROBLEM C 10 C 1 L m 1 50 C m K W. , the inner surface temperature is. 30 W m K

A FINITE VOLUME-BASED NETWORK METHOD FOR THE PREDICTION OF HEAT, MASS AND MOMENTUM TRANSFER IN A PEBBLE BED REACTOR

Safety Precautions WARNING If critical situations that could lead to user s death or serious injury is assumed by mishandling of the product.

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

Heat and Mass Transfer Unit-1 Conduction

11. Advanced Radiation

THERMAL PROPERTY MEASUREMENT TECHNIQUES. AND SOME RESULTS FOR CdTe. H. F. Poppendiek D. J. Connelly R. K. Livett

Available online at ScienceDirect. Physics Procedia 67 (2015 ) Calibration of a HTS based LOX 400mm level sensor

Proximity Sensor Ideal for High Temperatures and Cleaning Processes. Applicable to 120 C

SHRI RAMSWAROOP MEMORIAL COLLEGE OF ENGG. & MANAGEMENT B.Tech. [SEM V (ME-51, 52, 53, 54)] QUIZ TEST-1 (Session: )

Examination Heat Transfer

Available online at ScienceDirect. Physics Procedia 67 (2015 ) Superfluid helium heat pipe. P.

Temperature and Field Dependence of Field Grading Tubes for Medium Voltage XLPE Cable Joints

* GRAM, Inc., Albuquerque, New Mexico, United States of America

Finite-Element Evaluation of Thermal Response Tests Performed on U-Tube Borehole Heat Exchangers

Saving Energy with Space Age High Emissivity Ceramic Coatings

Transcription:

Available online at www.sciencedirect.com ScienceDirect Procedia Engineering 62 ( 213 ) 829 836 The 9 th Asia-Oceania Symposium on Fire Science and Technology model evaluation for cables used in nuclear plants in Japan Marc L. Janssens a, *, Stephen L. Turner b, Susumu Tsuchino c a Southwest Research Institute, 622 Culebra Road, San Antonio, TX 78238, USA b Science Applications International Corporation, 31 Laboratory Road, Oak Ridge, TN 37831, USA c Japan Nuclear Energy Safety Organization, Toranomon Towers Office Building 1-28, Toranomon 4-chome, Minato-ku, Tokyo, Japan Abstract Many nuclear power plants in the U.S. and elsewhere are transitioning from a prescriptive to a risk-informed performance-based fire protection program. The acceptance of a proposed performance-based fire protection program is partly based on the results of a probabilistic risk assessment (PRA), which quantifies the likelihood of core damage and harmful radiological release in the event of a fire. The fire PRA relies on engineering analyses to determine the effect of fires on safety-related equipment and components, such as electrical cables critical for safe plant shutdown. The report of the CAROLFIRE research program published in 8 provides thermal and electrical performance data for a broad range of cable types used in U.S. nuclear power plants. The report also describes the thermally-induced electrical failure () model, which estimates the temperature response of a cable for a specified thermal exposure. The purpose of the research described in this paper was to obtain thermal and electrical performance data and to evaluate the predictive capability of the model for some cables that are used in nuclear plants in Japan. Two types of cables were tested, designated as CCV and PSHV. The average fire-induced electrical failure temperature under the jacket of the CCV cable was approximately C. The average failure temperature of the PSHV cable was approximately 37 C. Both Japanese cables therefore performed like the thermoset cables in the CAROLFIRE tests. The model predictions are in remarkably good agreement with the measured cable temperatures for the tests in which electrical failure was observed. However, the model overestimates the cable temperature at the end of the tests in which failure was not observed. 213 Published International by Elsevier Association Ltd. Selection for Fire Safety and/or Science. peer-review Published under responsibility by Elsevier Ltd. of the Open Asia-Oceania access under Association CC BY-NC-ND for Fire license. Science and Selection Technology. and peer-review under responsibility of the Asian-Oceania Association of Fire Science and Technology Keywords: Electrical cables; Nuclear plants; Penlight; Performance-based fire protection; model Nomenclature c specific heat capacity (kj/kg K) Greek symbols h convection coefficient (kw/m 2 K) ε surface emissivity k thermal conductivity (W/m K) ρ density (kg/m 3 ) q " heat flux (kw/m 2 ) σ Boltzmann constant (5.67 1-11 kw/m 2 K 4 ) r radial coordinate (m) Subscripts t time (s) g gauge R cable radius (m) net net T temperature ( C or K) s shroud * Corresponding author. Tel.: +1 21 522 6655; fax: +1 21 522 3377. E-mail address: mjanssens@swri.org. 1877-758 213 International Association for Fire Safety Science. Published by Elsevier Ltd. Open access under CC BY-NC-ND license. Selection and peer-review under responsibility of the Asian-Oceania Association of Fire Science and Technology doi: 1.116/j.proeng.213.8.132

83 Marc L. Janssens et al. / Procedia Engineering 62 ( 213 ) 829 836 1. Introduction An increasing number of nuclear power plants in the U.S. and elsewhere are transitioning from a prescriptive to a riskinformed performance-based fire protection program. The acceptance of a proposed performance-based fire protection program depends on the ability of the plant operator to provide reasonable assurance that a fire will not result in damage to the fuel core and harmful radiological release. This assurance is partly based on the results of a probabilistic risk assessment (PRA), which quantifies the likelihood of core damage and large radiological release in the event of a fire. The fire PRA in turn is based on engineering analyses to determine the effect of fires in each area on structures, equipment and components. Electrical cables that are critical for the safe shutdown of the plant are the primary targets that need to be considered in these analyses. In 8 the U.S. Nuclear Regulatory Commission published the results of the CAROLFIRE research program. The first two volumes of the CAROLFIRE report provide thermal and electrical performance data for a broad range of electrical cable types used in U.S. nuclear power plants [1, 2]. CAROLFIRE testing comprised 78 small-scale and 18 intermediatescale open burn tests. The small-scale tests involved exposure of one or several lengths of cable to radiant heating in a cylindrical exposure chamber called Penlight. A broad range of thermoplastic and thermoset insulated cables were tested. All Penlight tests measured the cable thermal response using thermocouples embedded within the target cables, and cable electrical performance was monitored using an insulation resistance measuring system (IRMS). The temperature under the jacket of thermoplastic cables at electrical failure (development of short between two conductors or between a conductor and ground) ranged from approximately -25 C. Failure temperatures for thermoset cables ranged from approximately -45 C. Volume 3 of the CAROLFIRE report documents the thermally-induced electrical failure () model, which was developed and validated based on the aforementioned Penlight test data [3]. The model takes, as input, an estimate of the radiative and convective heat flux to a cable during a fire and predicts, as output, the temperature response of the cable. The time to electrical failure is then based on an assumed failure threshold temperature characteristic of the cable of interest. The purpose of the research described in this paper was to obtain thermal and electrical performance data for cables that are used in nuclear plants in Japan and to evaluate the predictive capability of the model for these cables. Two types of cables, designated as CCV and PSHV, were tested in a Penlight apparatus at Southwest Research Institute (SwRI ) in San Antonio, Texas. The experimental setup and heat flux calibrations are described in detail in the next section, as well as the cables that were tested and the results that were obtained. The model calculations for the two cables are discussed in the next section. Conclusions and recommendations for future work are presented at the end of the paper. 2. Experimental 2.1. Test setup An apparatus was constructed at SwRI similar to the Penlight chamber used in the small-scale CAROLFIRE tests at Sandia National Laboratories (SNL) in Albuquerque, New Mexico. The apparatus consists of a cylindrical ring of rodshaped.61-m long quartz heating lamps, each held in a water-cooled aluminum fixture with a reflector to direct the heat toward the center of the lamp array. An Inconel cylindrical shroud, approximately.51 m in diameter and.81 m long, is installed within the array of heating lamps. The shroud is painted with high-temperature flat black paint. The quartz lamps are used to heat the shroud to a desired (and controlled) temperature. The shroud in turn acts as a grey-body radiator heating any target object located within it. Twenty-eight grounded junction Inconel sheathed thermocouples (1 mm diameter) are installed to monitor the temperature distribution along the length and the perimeter of the shroud. A separate centrally located thermocouple provides the input signal for the power controller. In addition, a system was built at SwRI to measure the electrical insulation resistance between pairs of conductors and between each conductor and ground during exposure in the Penlight chamber. The system was based on the insulation resistance measurement system (IRMS) at SNL described in Appendix B of Volume 1 of the CAROLEFIRE report [1]. The IRMS at SwRI is capable of providing electrical resistance data for cables with up to eight conductors. For a cable with eight conductors, a complete set of resistance measurements can be obtained in less than 5 seconds. For a three-conductor cable the fastest scan rate reduces to less than 1 seconds. A functional check showed than resistances between 1 and 2 kω measured with the IRMS are better than 1% of the actual values.

Marc L. Janssens et al. / Procedia Engineering 62 ( 213 ) 829 836 831 2.2. Heat flux calibration A series of calibrations were performed to establish a relationship between the shroud temperature and the incident heat flux in the Penlight apparatus at SwRI. The calibrations involved measuring the heat flux to a 13-mm diameter Schmidt- Boelter heat flux gauge at different shroud temperatures. The gauge was mounted in a 13-mm calcium-silicate dummy board backed by 25-mm thick ceramic fiber blanket. The gauge was located at the center of the shroud and faced the upper half of the shroud. The dummy board was mounted on a ladder tray to closely simulate the thermal environment that a cable specimen is exposed to in a test. The results of the heat flux measurements with the shroud closed off at one end are given in Table 1. The following empirical relationship was established to determine the heat flux as a function of the shroud temperature: q " net s 4 s 4 g ( T ) = ε σt σt + h T (1) The best fit was obtained for ε s =.9 and h =.8 kw/m 2 K and is shown in Fig. 1 below. The resulting calculated heat fluxes at the control thermocouple temperatures at which the flux calibrations were performed are included in Table 1. s g Table 1. Measured and calculated heat flux as a function of Penlight shroud temperature Control TC temperature ( C) Measured heat flux (kw/m 2 ) Calculated heat flux (kw/m 2 ) 434 14.5 14.7 532 24 24 616 35 35 687 49 47 77 65 65 839 83 83 886 98 98 Fig. 1. Best fit of gauge heat flux as a function of Penlight shroud control thermocouple temperature. To determine how the heat flux varies along the cable specimens, the heat flux profile was measured at control thermocouple temperatures of 47 C and 665 C. The results are given in Table 2. The heat flux is within ±.5 kw/m 2 over the central.25-m section of the shroud at a control thermocouple temperature of 47 C, and within ±1.3 kw/m 2 at 665 C. The average heat flux over the five interior locations is 15.1 kw/m 2 and 37.9 kw/m 2 for the two shroud temperatures, respectively. These values are slightly higher than the heat fluxes at these shroud temperatures calculated by SNL, i.e. 14.1 kw/m 2 and 35.8 kw/m 2 respectively (see page 3 of Reference [1]). This discrepancy might, for example, be due to heat flux measurement errors or a lower than actual shroud emissivity in the SNL calculations. Finally, the heat flux calibrations showed that the shroud temperature rises from ambient to 47 C in approximately 5 s and to 665 C in about 9 s.

832 Marc L. Janssens et al. / Procedia Engineering 62 ( 213 ) 829 836 Table 2. Heat flux profile at shroud control thermocouple temperatures of 47 C and 665 C Distance from open end (m) Heat flux for T s = 47 C (kw/m 2 ) Heat flux for T s = 665 C (kw/m 2 ).25 7.1 18.7.152 12.4 31.2.279 16. 39.3.46 17. 41.9.533 16.4 41..66 13.6 36..787 11.2 31.1 2.3. Test specimens Table 3 provides some important characteristics of the two cables that were tested. In each test, two specimens of the cable were exposed to the radiant heat from the shroud of the Penlight apparatus. One cable specimen was instrumented with five thermocouples embedded into the cable (two directly under the jacket and three at various depths below the jacket). The other cable specimen was connected to the IRMS. The two cable specimens were fixed with clamps to the rungs of a standard -mm wide ladder-backed pre-galvanized steel cable tray, as shown in Fig. 2. Care was taken to ensure that none of the cable thermocouples were shielded by a rung. The cable tray was identical to that used in the CAROLFIRE program (see page 31 in reference [2] for a detailed description) except that it was 3 m instead of 3.6 m long. The exposed part of the cable specimens was positioned at 38 mm on either side of the center of the cable tray resulting in a spacing of 76 mm between the two cable specimens. Prior to the test the cable tray was routed through the Penlight shroud and positioned horizontally so that TC 2 was at equal distances from both ends of the shroud and positioned vertically so that the top of the rungs was approximately 3 mm below the horizontal centerline of the Penlight shroud (see Fig. 2). Table 3. Characteristics of the two cables that were tested in the Penlight apparatus Designation Insulation/Jacket Diameter (mm) Jacket thickness (mm) Mass per length (kg/m) Copper mass fraction CCV XLPE/PVC 14.4 1.6.29.5 PSHV EPR/PVC 15.8 1.7.44.44 2.4. Results A total of nine Penlight tests were conducted; a preliminary test on the CCV cable and four additional tests on each of the two types of cable. Baseline data were collected for a period of two minutes at the start of a test. Power was then turned on to heat the Penlight apparatus to the preset shroud temperature. A PID controller was used to minimize the heating time without excessive overshoot of the target shroud temperature. All tests except the preliminary CCV test were terminated after one hour of exposure, or sooner if electrical failure of all conductors was observed. The primary purpose of the preliminary CCV test was to identify problems or issues that needed to be addressed before the actual Penlight tests were conducted. In this test the shroud temperature was initially set at C. After 45 and 6 minutes the shroud temperature was increased to 34 C and C, respectively. Finally, after 75 minutes the shroud temperature was increased to 47 C and all cables started to fail at approximately 85 minutes (initially conductor to conductor and on the next scan, 45 s later, conductor to ground). Electrical failure coincided with ignition of the cable. The cable specimen connected to the IRMS ignited first. The maximum cable temperature immediately prior to ignition was approximately 382 C. In the next three tests on the CCV cable specimens, electrical failure was not observed during 6 minutes of exposure. The shroud temperature in these three tests was 38 C, 41 C, and 44 C, respectively. The shroud temperature during the final CCV test was 47 C. In this test the cable started failing between 965 and 125 seconds into the test (Conductor 1 to Ground short). Ignition was observed shortly thereafter. The maximum cable temperature immediately prior to ignition was approximately 442 C as shown in Fig. 3.

Marc L. Janssens et al. / Procedia Engineering 62 ( 213 ) 829 836 833 In the first PSHV test the cable failed between 952 and 956 seconds, i.e. after nearly 14 minutes of exposure to the radiant heat flux from the Penlight shroud heated at 47 C. The highest cable thermocouple temperature at that time was 368 C. The first electrical failure occurred between Conductor 1 and Ground. Consistent with the values used in the CAROLFIRE program, the failure threshold was set at 1, Ω while the insulation resistance of undamaged cable was assigned a value of kω (e.g., see pages 32 and B-5 in Reference [1], respectively). Fig. 2. Tray with cable specimens positioned inside the Penlight chamber prior to testing. 4.E+5 3.E+5 Cable TC 1 Cable TC 2 2.E+5 Cable TC 3 Cable TC 4 Cable TC 5 1.E+5 1.E+ 1 1 2 Failure ~ 121-125 s Fig. 3. Temperature and electrical resistance data obtained in the final CCV test. The shroud temperature in the second PSHV test was 44 C. Electrical failure was not observed, but resistances started to decrease (without ever dropping below 1, Ω) between 1114 and 113 seconds. The highest cable thermocouple temperature at that time was approximately 339 C. The remaining PSHV tests were both performed at a shroud temperature of 455 C and the results provide some indication of the repeatability of the Penlight tests. Temperature and resistance data obtained in these two tests are shown in Figs. 4 and 5, respectively. In the third PSHV test electrical failure was first observed between Conductor 2 and Ground. In the fourth test electrical failure was first recorded between Conductor 1 and Ground. The highest cable thermocouple temperature at the time of electrical failure was approximately 386 C and 358 C in the third and fourth test, respectively. 1 Cable TC 1 Cable TC 2 Cable TC 3 Cable TC 4 Cable TC 5 Resistance C2-G Failure ~ 1252-1254 s 4.E+5 3.E+5 2.E+5 1.E+5.E+ 1 1 2 Fig. 4. Temperature and electrical resistance data obtained in the third PSHV test.

834 Marc L. Janssens et al. / Procedia Engineering 62 ( 213 ) 829 836 1 Cable TC 1 Cable TC 2 Cable TC 3 Cable TC 4 Cable TC 5 4.E+5 3.E+5 2.E+5 1.E+5.E+ 1 1 2 Fig. 5. Temperature and electrical resistance data obtained in the fourth PSHV test. 3. model calculations 3.1. model Andersson and Van Hees proposed that the thermally-induced electrical failure () of PVC insulated cables can be predicted via a simple one-dimensional heat transfer calculation, under the assumption that the cable can be treated as a homogenous cylinder [4]. McGrattan used the Penlight data from the CAROLFIRE program to extend the validity of the model to a wide range of cable types. A brief description of the model based on McGrattan s discussion in Reference [3] follows. The simplifying assumptions underlying the model are as follows: 1. The heat penetration into a cable of circular cross section is largely in the radial direction. 2. The cable is homogenous in composition. 3. The thermal properties conductivity, specific heat, and density of the assumed homogenous cable are independent of temperature. 4. It is assumed that no decomposition reactions occur within the cable during its heating, and ignition and burning are not considered in the model. 5. Electrical failure occurs when the temperature just inside the cable jacket reaches an experimentally determined value. Given these assumptions, the governing equation for the cable temperature, T(r,t), is given by: T 1 T ρ c = kr (2) t r r r where, c, and k are the effective density, specific heat, and conductivity of the solid, all assumed constant. The boundary condition at the exterior boundary, r = R, is given by: T k = r " ( R t) q, net (3) " where q net is the assumed axially-symmetric net heat flux at the exterior surface of the cable. The model employs a single value for the specific heat and the thermal conductivity, 1.5 kj/kg K and.2 W/m/K, respectively, for both thermoset and thermoplastic cables. The emissivity of the cable jacket is assumed to be.95. These property values are typical of several types of commonly used cable jacket and insulation materials [5]. The bulk density of the cable, ρ, can be calculated by dividing the mass per unit length by the cross sectional area. 3.2. model calculations for tests in which electrical failure was observed Figures 6-9 compare the model predictions to the temperature measured under the jacket for the Penlight tests in which electrical failure was observed. The agreement is quite remarkable, given the simplicity of the model. For the final CCV test, the net heat flux in Equation (3) was determined based on Equation (1), with the exterior surface temperature of the cable replacing the gauge temperature and an emissivity of the cable of.95 instead of 1 for the gauge. However, in the final CCV test the cable specimens ignited and there was extensive melting and dripping. As a result, the

Marc L. Janssens et al. / Procedia Engineering 62 ( 213 ) 829 836 835 interior of the Penlight shroud had to be cleaned and in the process part of the paint was inadvertently removed. It is assumed that this adversely affected the heat transfer characteristics of the shroud surface. Consequently, the calculations for the PSHV tests were performed with ε s =.82 and h =.5 kw/m 2 K. Additional heat flux calibrations are needed to confirm these values. 3.3. model calculations for tests in which electrical failure was not observed Figure 1 shows an example of a Penlight test in which electrical failure was not observed. Except for the first six minutes of Penlight exposure, the model significantly overpredicts the temperature under the jacket. In this case the model actually predicts cable failure while this was not observed in the test. Hence, the model errs on the conservative side. A possible explanation for the discrepancy is that the model does not account for heat losses along the conductors (see assumption #1). 4. Conclusions Two types of eight-conductor cables that are used in nuclear plants in Japan were tested in the Penlight apparatus at SwRI. The first cable, referred to as CCV, has cross-linked polyethylene (XLPE) insulation and a heat resistant vinyl (PVC) jacket. The second cable, referred to as PSHV, has fire retardant ethylene propylene rubber (EPR) insulation and a fire retardant low hydrochloric heat resistant vinyl (PVC) jacket. The CCV cable temperature under the jacket at the time of electrical failure was 42 ± 2 C. The PSHV cable temperature under the jacket at the time of electrical failure was 371 ± 14 C. For both types of cables and the tests in which electrical failure was observed, the model predicts the temperature under the jacket with remarkable accuracy. However, the model overestimates the cable temperature at the end of the tests in which failure was not observed. A possible explanation for this discrepancy is that the model does not account for heat losses along the conductors. The intent is to perform Penlight tests on six other types of cables and to extend the validity of the model to a wider range of cables used in nuclear plants in Japan. 4.E+5 3.E+5 2.E+5 1.E+5 1.E+ 9 1 1 1 Failure ~ 121-125 s Fig. 6. model predictions and selected measurements obtained in the final CCV test. 1 4.E+5 3.E+5 2.E+5 1.E+5.E+ 9 1 1 1 Failure ~ 952-956 s Fig. 7. model predictions and selected measurements obtained in the first PSHV test.

836 Marc L. Janssens et al. / Procedia Engineering 62 ( 213 ) 829 836 4.E+5 3.E+5 2.E+5 1.E+5 1.E+ 9 1 1 1 Failure ~ 1252-1254 s Fig. 8. model predictions and selected measurements obtained in the third PSHV test. 4.E+5 3.E+5 2.E+5 1.E+5 1 Failure ~ 1112-1114 s.e+ 9 1 1 1 Fig. 9. model predictions and selected measurements obtained in the fourth PSHV test. 1 9 1 1 1 Fig. 1. Temperature and electrical resistance data obtained in the third CCV test following the preliminary test. Acknowledgements The support by the Japan Nuclear Energy Safety Organization for the work described in this paper is acknowledged and greatly appreciated. References [1] Knowlen, S., Wyant, F., Cable Response to Live Fire (CAROLFIRE), Volume 1: Test Descriptions and Analysis of Circuit Response Data, NUREG/CR-6931, Vol. 1, Sandia National Laboratories, Albuquerque, NM, 8. [2] Knowlen, S., Wyant, F., Cable Response to Live Fire (CAROLFIRE), Volume 2: Cable Fire Response Data for Fire Model Improvement, NUREG/CR-6931, Vol. 2, Sandia National Laboratories, Albuquerque, NM, 8. [3] McGrattan, K., Cable Response to Live Fire (CAROLFIRE), Volume 3: Thermally-Induced Electrical Failure () Model, NUREG/CR-6931, Vol. 3, National Institute of Standards and Technology, Gaithersburg, MD, 8. [4] Andersson, P., Van Hees, P., 5. Performance of Cables Subjected to Elevated Temperatures, Fire Safety Science - Proceedings of the 8th International Symposium, International Association for Fire Safety Science, pp. 1121-1131. [5] Hamins, A., Maranghides, A., Johnsson, R., Donnelly, M., Yang, J., Mulholland, G. Anleiter, R., Report of Experimental Results for the International Fire Model Benchmarking and Validation Exercise #3, NIST Special Publication 113-1, National Institute of Standards and Technology, Gaithersburg, MD, 6.