Experimental Investigation and Numerical Simulation of the Mechanical and Thermal Behavior of a...

9
This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright

Transcript of Experimental Investigation and Numerical Simulation of the Mechanical and Thermal Behavior of a...

This article appeared in a journal published by Elsevier. The attachedcopy is furnished to the author for internal non-commercial researchand education use, including for instruction at the authors institution

and sharing with colleagues.

Other uses, including reproduction and distribution, or selling orlicensing copies, or posting to personal, institutional or third party

websites are prohibited.

In most cases authors are permitted to post their version of thearticle (e.g. in Word or Tex form) to their personal website orinstitutional repository. Authors requiring further information

regarding Elsevier’s archiving and manuscript policies areencouraged to visit:

http://www.elsevier.com/copyright

Author's personal copy

Experimental investigation and numerical simulation of a furnished office fire

Chien-Jung Chen a, Wei-Dong Hsieh b, Wei-Chieh Hu b, Chi-ming Lai c, Ta-Hui Lin b,*

aArchitecture and Building Research Institute, Ministry of Interior, TaiwanbDepartment of Mechanical Engineering, National Cheng Kung University, Tainan, TaiwancDepartment of Civil Engineering, National Cheng Kung University, Tainan, Taiwan

a r t i c l e i n f o

Article history:Received 28 December 2009Received in revised form2 June 2010Accepted 3 June 2010

Keywords:FDS (Fire Dynamics Simulator)Office fireFire load

a b s t r a c t

Experiments were conducted in a full-scale model room equipped with both movable and fixed fire loadsto explore fire growth and spread via heat release rates, indoor air temperature and species concen-tration. The room space is a brick structure that measures 5.7 m in interior length, 4.7 m in width and2.4 m in ceiling height. The northeast and southeast corners each feature a 2.1 m � 0.9 m open doorway.Numerical simulations with parameter adaptation were carried out using FDS software to predict the firefeatures and were compared with the experimental results. In this study, the material properties andoxygen limit settings in the FDS software were tested to explore their influence on the tendency of heatrelease rate and on the total amount of heat release. The results show that the heat release rate from theFDS simulations is comparable to the full-scale experiment results during the fire growth period.Temperature profile near ceiling can be modeled well. In the full-involvement burning and decayingperiods, the qualitative trends were identical, although the simulated value differed greatly from theexperimental result.

� 2010 Elsevier Ltd. All rights reserved.

1. Introduction

Experimental measurement and numerical analysis are twocomplementary methods used to study fire dynamics and speciestransport in the enclosure fire tests. These two methods have theirrespective benefits and deficiencies; they are complementary toeach other. In a full-scale experiment, the focus is on the repro-duction of the real fire scenario, which requires a lot of manpower,a huge amount of budget and a large space for performing theexperiments. However, targets of the full-scale experiment arelimited, the repeatability is also poor due to the unstable nature offire and it is difficult to extend the tests or draw solid conclusionsfrom limited tests. However, data from full-scale fire experimentscan be compared with empirical formulas and numerical simula-tion results to validate or improve the model. Numerical simulationanalyses often serve to explore complicated full-scale fires. Thedevelopment of simulation software has seen rapid progression.However, due to an insufficient material database and incompletephysical and chemical simulation models on the material burningand pyrolysis, there is still room for significant improvement onnumerical simulation. Thus, full-scale fire experiments can

examine the accuracy of numerical results in order to improve theirapplication and precision. Of the numerical simulation softwareavailable, FDS (Fire Dynamics Simulator) [1] from NIST (NationalInstitute of Standards and Technology) is widely used among thefire groups for simulating building fires. And it has been used fora large number of studies and applications in recent years, fromsmoke and flow characteristics under fire scenario [2e5], room firesimulation [6,7], arson fire reconstruction [8] to fire fightingmodeling [9].

Many researchers have also used FDS to simulate compartmentfires. Kim and Ryou [9] used FDS to simulate methyl and hexane firesuppression using a water mist in a 4.0 m � 4.0 m � 2.3 m room.The simulated temperature was very accurate, with a 10 �C differ-ence from the experiment. Due to the lack of a fire extinguishingmodel, the fire burnout time was not accurately predicted. Zou andChow [7] conducted ISO 9705 room tests and FDS simulations forcomparison. In the experimental tests, pool fires were adjusted tomimic different heat release rates and ventilation conditions werethen changed to estimate flashover time. The results showed thatFDS simulation corresponds to the temperature or heat radiationobtained from experimental results. The simulated temperaturenear the ceiling was within 10% of that of the experiment. Thisimproves the judgment of flashover time.

For the simulation of a full-scale experiment result, the FDSshowed great dependence on the input parameters [10,11]. Jahn

* Corresponding author. Tel.: þ886 6 2757575x62167; fax: þ886 6 2352973.E-mail address: [email protected] (T.-H. Lin).

Contents lists available at ScienceDirect

Building and Environment

journal homepage: www.elsevier .com/locate/bui ldenv

0360-1323/$ e see front matter � 2010 Elsevier Ltd. All rights reserved.doi:10.1016/j.buildenv.2010.06.003

Building and Environment 45 (2010) 2735e2742

Author's personal copy

et al. [10] studied the sensitivity of FDS modeling to a set of inputparameters related to fire growth, such as fire size and location,convection, radiation and combustion parameters. It is identifiedthe simulations of fire growth are significantly sensitive to locationor the heat release rate, fire area, flame radiative fraction andmaterial thermal and ignition properties. Rein et al. [11] performeda round-robin study on the modeling prediction of the DalmarnockFire One Test. Comparison of the round-robin modelingresults shows a large scatter and considerable disparity amongthe predictions, and between predictions and experimentalmeasurements.

FDS is widely used in fire simulations of different aspects, andproviding valuable information, such as smoke movement, localtemperature, local gas components and velocity. But there are somelimitations due to its modeling. For example, FDS only deals thecombustion of one single gaseous fuel that leads to bad predictionof scenarios of multiple fuels. After the work in Edinburgh in [11]and [12], it has become patent that FDS cannot predict well firein small compartments and under ventilated fires. The generalconclusion is that the modeling of realistic enclosure fires must beconducted with the aid of experimental measurements of fire testsdirectly related to the event under study.

The aim of this research is to improve the accuracy of numericalsimulations. By comparing the simulation results with the full-scaleexperimental results, influences of parameters in FDS were exam-ined. In this paper, we first introduced the setups of the full-scaleexperiment. After that, the numerical simulation and the properties

we have examined in the simulations were presented. In the resultsand discussions section, FDS simulation and experimental results ofheat release rate, oxygen concentration and temperaturemeasurements were presented. Effects of FDS parameters on theprediction were also discussed. At last, some conclusions weredrawn from this study.

2. Full-scale experiment and results

A practical full-scale room is used to illustrate the actual fireconditions. The interior plan dimensions are 5.7 m� 4.7 mwith thenet ceiling height 2.4 m. The walls of the model room are consistedof reinforced bricks with 0.26m in thickness and the ceiling is madeof a light rigid frame and gypsum board. There are two2.1 m � 0.9 m doors kept opening throughout the experiment onthe northeast and southeast corner of the room. Experiments wereconducted in this full-scale model room equipped with bothmovable and fixed fire loads, as shown in Fig. 1.

Four wall furnishings types are used as the fixed fire loads in theexperiments. They are no furnishing but the plywood wall only (asitem (I) in Fig. 1); a top-hanging wood cabinet with lower plywoodwall (II); a bottom short cabinet with upper plywood wall (III) anda tall wood cabinet (IV). The top wood cabinet with lower plywoodwall (II) and tall wood cabinet (IV) are combined, as they havesimilar scopes of thermal plume influence, and they are installed onwalls (2) and (4). Similarly, the plywood wall (I) and the bottomshort cabinet with upper plywood wall (III) are combined and

Fig. 1. Schematic diagram of the investigated model room (not to scale).

Fig. 2. Photos of the interior furnishing and the movable fire load.

C.-J. Chen et al. / Building and Environment 45 (2010) 2735e27422736

Author's personal copy

installed on walls (1) and (3). The plywood walls are 1.6 m wide,2.4 m high and 0.04 m thick. The combination of type (II) and (IV)furnishings (as used at walls (2) and (4)) have a total weight of102.1 kg. The total weight of type (I) and (III) furnishings (as used atwalls (1) and (3)) is 92.1 kg. The total fire load is approx. 15.4 kg/m2

Fig. 2(a) shows the photos of the wall (3) and the movable fire loadare shown.

A wooden crib consists of 200 stacked 2.4 cm � 3.6 cm � 90 cmwood strips, is used as the fire source in the center of the room,shown as item (a) in Fig.1 and Fig. 2(b). Stripswerefirst conditionedat 105 �C for 24 h and then used in the experiments. Each layer iscomposed of six wooden strips that are stacked and fixed with nailsto avoid collapse due to thermal deformation. The wooden crib wasignited by burning 1 L ethanol paste in a circular pan beneath it.

Fig. 3 illustrated the overall layouts of the experimentmeasurements. Measurements of temperature and gas composi-tions are performed by thermocouple and gas analyzer. The roundsymbols marked by the numbers 1e21 represent thermocoupletrees TC1 to TC21. Sets of K-type thermocouples, referred to asthermocouple trees, were tied together vertically to thin ironchains and then wrapped by thermal insulation material. Eachthermocouple tree consists of eight measurement pointsnumbering from 0 (10 cm from the ceiling) to 7 (near the floor). No.0 to No. 4 was 10 cm apart to have better vertical temperature

resolution near the ceiling, others were 50 cm apart. For example,TC1-1 is the thermocouple located on TC1 at ceiling. The totallength of each thermocouple tree is 200 cm.

The triangle symbol marked by the letter G in Fig. 3 is a gasanalyzer, which measures the O2, CO and CO2 concentration of ata height of 180 cm. The investigated model room is located belowa smoke collection hood of a 10-MW fire test facility (as shown inFig. 4). Hot gas, smoke and combustion products are collected bythe smoke collection hood and then pass through a measurementsection. The continuous online gas analysis system is located in themeasurement section and consists of (1) the gas analysis system(including O2, CO, CO2, NOx and HC analyzers, as well as a gassampling/calibration system), (2) an optical density analyzer, (3)a flow rate/temperature monitor and (4) a data-processing system.This study mainly uses the oxygen consumption (OC) method [13]to calculate the heat release rates.

3. Numerical simulation

The CFD package used in this study is the FDS 5.2.3 developed byNIST. For complete simulation of the wood strips burning in thesimulation, the grid size need to be very small, and more compu-tation efforts will be needed. Besides, the main issue of this study isabout the fixed fire load, not the central movable fire load. There-fore, the central movable fire load in the full-scale experiment wasreplaced by a gaseous burner in the simulation. The heat releaserate profile of the gaseous burner was set according to the calcu-lated wooden crib heat release rate, Q_WC, which will be furtherelaborated in Section 4.2.

The geometry of the openings and layout of the fixed fire loadwere identical to the full-scale experimental test. In the numericalsimulations, the calculation domain was 6.7 m � 5.4 m � 2.7 m,a little larger than the experimental space, in order to observe theair flow through the openings, as shown in Fig. 5. The wallfurnishing thickness is set to 0.04 m. Two different materials wereused in the simulation: spruce and plywood. The properties of thespruce were obtained from the FDS 4 database and it is used as thedefault properties for the wall lining and cabinet in the FDS simu-lation, while the properties of plywood were directly measuredfrom the furnishings of full-scale experiment. The density is 450 kg/m3 for the spruce, 730 kg/m3 for the plywood. The properties of thematerials constitute one of the major parameters affecting simu-lation results. The default setting from the FDS 4 database indicatesthe amount of ash produced from the spruce is 35%. However, thesimulated heat release rate is notably different from that ofthe experiment. After examining the ash residue after fire in theexperiment, the amount of ash in the setting of the material

Fig. 3. Overall layout of the measurement apparatus.

Fig. 4. 10-MW fire test facility and the model room.

C.-J. Chen et al. / Building and Environment 45 (2010) 2735e2742 2737

Author's personal copy

properties was altered to 15% for the simulations. The comparisonof materials listed in Table 1.

The default inflammability setting in the FDS indicates thatmaterial cannot be ignited when the oxygen concentration is lowerthan 15% and when the temperature is below 1427 K. This setupgreatly affects the growth of the heat release rate and total heat.Simulations were performed at oxygen limits of 15% and 12%respectively in this study to discuss the effect of oxygen limits onthe simulation results.

It is well known that the size of the calculation grid can directlyaffect the simulation results. Ma and Quintiere [14] used FDS tosimulate axially symmetric flames. A characteristic length, definedas the ratio of the grid size to the characteristic flame length, isincorporated. A characteristic length of around 0.05 has been shownto produce the best simulation results for the flame length.McGrattan et al. [15] and Kwon [16] showed that, by consideringexperimental conditions and computing time, a grid size of 0.1 isacceptable for determining the characteristic flame length.According to these studies, the grid size in our studywasdeterminedby the following equation proposed by Ma and Quintiere [14].

Z* ¼ max�dx; dy; dz

���_Q=rNCpTNg1=2

�2=5(1)

Where Q_ is the heat release rate (kW), rN, Cp, and TNare propertiesof ambient air. According to their study, Z* of 0.05 corresponds to anoptimal grid size for best resolution of fire plume. The heat releaserate here referred to the fire load in the room, which is estimated tobe 5200MJ. Therefore, we use grid size (di) of 0.05m correspondingto Z* ¼ 0.05 in the numerical simulation.

Though a grid size of 0.05 mwas determined by Eq. (1), a test ofgrid size sensitivity was performed. The heat release rate profiles ofthree grid sizes (0.1, 0.05, and 0.0375 m) were showed in Fig. 6. Itshows that altering the grid size from 0.05 m to 0.0375 m resulted

in little change in the heat release rate profile, inferring conver-gence at the 0.05 m grid spacing. Therefore, the grid size is set to0.05 m in the following study.

4. Results and discussions

4.1. Full-scale fire test results

The entire fire process (see Table 2 and Fig. 7), from ignition toextinction, occurs within approximately 840 s. At 75 s, usingalcohol paste ignition, the movable fire load was ignited. Atapproximately 284 s, the upper layer of thick smoke started to exitfrom the southeast door; the southeast fixed fire loadwas ignited at330 s. At 363 s, the top of the tall cabinet in the north was alsoignited. At 400 s, the central movable fire load reached 600 �C, anda large plume of fire emerged from the opening, at which timeflashover occurred. At 690 s, only four wooden corners were stillburning. The experiment ends at 840 s due to the cease of burning.

4.2. Experimental heat release rate

In Fig. 8(a) and (b), the black lines (marked with ‘exp.’) representhistory of the experimental heat release rate. The circle labeled ‘E’on the x-axis represents the time at which the southeast fixed fireload begins to burn. ‘N’ is the time at which the northern fixed fireload begins burning. ‘F’ is the observed flashover time. The dottedlines represent the heat release rate of the central movable fire load(wooden crib) Q_WC(MW) which was calculated by the mass lossrate (measured from a platform scale) times the heat value ofPhilippine lauan wood strips, 12.06 MJ/kg, as Eq. (2) shows.

_QWC ¼ �dmdt

� 12:06 (2)

Where m is the mass of the movable fire load (kg), and t is the time(s).

As shown in Fig. 8, the heat release curve has two peaks in theexperimental process. The first peak results from the movable fireload. The overall heat release rate (HRR) reached 7.5 MW; after-wards, as the fixed fire load was not completely ignited and the

Fig. 5. Simulation model.

Table 1Properties of materials.

Properties Spruce Plywood

Density (kg/m3) 450 730Ash Content (%) 15 21.5Energy per unit oxygen (kJ/kg) 11 020 12 075Conductivity (W/m-K) Temperature dependent,

properties from the FDSdatabase were used

Specific Heat (kJ/kg-K)

Fig. 6. The heat release rate profiles at different grid sizes.

C.-J. Chen et al. / Building and Environment 45 (2010) 2735e27422738

Author's personal copy

combustion of the central movable fire load was in decline, theoverall heat release dropped. Before the occurrence of the firstpeak, the movable fire load dominated the heat release. At around650 s, the surrounding fixed fire loads were totally ignited and ledto the second HRR peak. The total heat releasewas around 2300MJ.The model room had two openings for ventilation and the10-MWfire test facility can only analyzed the flue gas escaped from thesoutheast door right beneath the hood. Therefore, a portion of thesmoke escaped from the northeast opening may not be analyzed.From the visual observation of the smoke leakage through thenortheast opening, it was estimated that about 10e15% of thesmoke was not collected by 10 MW hood. Therefore, the measuredHRR would be 85e90% of the actual HRR.

4.3. Comparison of heat release rate

Comparison between the heat release rate obtained from theFDS simulation and experiment can also be observed on Fig. 8. InFig. 8(a), ‘spruce/0.15’ and ‘plywood/0.15’ represent the simulationresults for the using spruce and plywood as the material properties,

respectively. The number ‘0.15’ is the oxygen limit at 15%. Althoughthe properties of the spruce and plywood are different, there aretwo peak values in both results.

Using spruce as the properties for the furnishings, the simulatedheat release rate increases more rapidly than for the plywood caseand is closer to the experimental results. The maximum heatrelease rate is approximately 7 MW. Although the simulatedheat release rate growth is slower, the time of the first peak in theheat release rate is near 400 s, matchingwell with the experimentalresult. The second heat release rate peak in the simulation occursearlier, and its duration is also shorter. Subsequently (at 660 s), theheat release rate drops quickly to the value of the gaseous burner.This means that the fixed fire load in the simulationmay have burntout; whereas the movable and fixed fire loads in the experimentwere both still burning at this moment. The overall heat releases ofthe spruce case differ greatly from that of experimental one.

Using plywood as the properties for the furnishings in thesimulation, the heat release rate growsmore slowly than that in theexperiment. The first heat release peak was postponed to around500 s. The highest heat release rate was the same as that in thespruce case, 7 MW. After reaching the highest heat release rate,there was a slight decline. At 650 s, it dropped to a value around3.5 MW equal to experimental one; the second heat release peakarrived much later, at 850 s, with the same heat release rate as inthe experiment. The total heat released in the plywood simulationwas higher than the experimental one.

During the fire growth period, the spruce has a larger firegrowth rate of heat release than the plywood does. The main factoris the amount of ash produced. Spruce produces 15% ash aftercombustion, while plywood leaves a greater amount of 21.5%. It ismore difficult to ignite materials with more ash, which (as in theplywood case) prolongs the time to reach the first heat release ratepeak. However, the peak value of the heat release rate is the same at

Table 2Documentary of the investigated fire.

Time (t) Phenomena Reference pictures

75 s Bottom of the wooden crib ignited284 s Black smoke from southeast door Fig. 5 (a)330 s Top of southeast wall material ignited Fig. 5 (b)363 s Top of northern tall cabinet ignited379 s Flames out of south entrance Fig. 5 (c)400 s Flashover690 s Wall material and cabinets burnt

out with only wooden corners still burningFig. 5 (d)

840 s End of experiment

Fig. 7. Documental photos of the fire progress at different instant.

C.-J. Chen et al. / Building and Environment 45 (2010) 2735e2742 2739

Author's personal copy

7 MW. The total heat releases for both spruce and plywood aredifferent from that in the experiment. The total heat releasedepends on the material density and ash content. At the samedimensions, plywood releases more heat for its higher density.Even with higher ash content, the total heat released in theplywood case in the simulation was still higher. In addition to thematerial properties, another possible reason for the differences intotal heat release is the parameter of radiative fraction. The defaultvalue is 0.35 in FDS database, and about 0.3e0.4 in real fires. But there-absorption of energy is not considered in FDS. Though theparameter was not tested in this study, a proper value of radiativefraction may be required for a better match of the experimentalheat release.

To have better simulation results, we further tuned the oxygenlimit in the FDS setting. Fig. 8(b) shows a comparison of the heatrelease rates obtained from the FDS simulations of oxygen limitssetting of 15% and 12% with the experimental results. The overallheat release changes for different oxygen limits resulted in a doublepeak variation and were comparable with those of the experiment.With a default oxygen limit of 15%, the period of fire growth issimilar to that of the experiment; the highest heat release ratereached 7 MW.

With an oxygen limit of 12%, the overall heat release, time toignition and time to burnout all get closer to the experimentalresults. The peak value was around 7.5 MW, which was higher thanthe 15% oxygen limit case and closer to the experimental results.Following the first heat release peak, the decrease in the heatrelease occurred 150 s later than the experimental result, failing tosimulate the experimental phenomena completely. However, theheat release peak increased; the total heat release was higher thanthat of 15% oxygen limit case. Thus, a decrease in the oxygen limitcan increase the rate of fire growth and reduce the underestimationof heat release.

4.4. Comparison of oxygen consumption

In Fig. 9(a), lower oxygen concentration corresponds to higherheat release rate in the discussion above. When using plywood,there was another oxygen concentration increase at 650 s, whichcorresponds to the minimum heat release rate. The minimumoxygen concentrationwas 10% in the experiment and 2% in the FDSsimulation indicating a poor ventilated environment. Despite thematerial properties setting, the FDS oxygen consumption wasclearly higher than that in the experiment. It is possible that the

Fig. 8. Comparison of the heat release rates obtained from the experiments and simulations.

Fig. 9. Comparison of the oxygen concentrations of the experiments and simulations.

C.-J. Chen et al. / Building and Environment 45 (2010) 2735e27422740

Author's personal copy

combustion model in the FDS does not apply to a situation withinsufficient air intake.

Fig. 9(b) shows a comparison of oxygen concentrations ofexperiment and FDS simulation with different oxygen limitsettings. As the oxygen limit setting is decreased, the oxygenconsumption and heat release both increase. The oxygenconsumption in the FDS simulation was marginally affected by theoxygen limit settings.

4.5. Comparison of temperature

Fig. 10 and Fig. 11 show the temperature variations measured byTC7-TC9 and TC13-TC15 thermocouple trees (shown in Fig. 3) in thefull-scale experiment (left column) and the temperature variationsof FDS simulation (right column). Some of the thermocouple treesmalfunctioned during the experiment is labeled by ‘failure’.Numbering �1e�8 are the eight points on the thermocouple treelocated at 10, 20, 30, 40, 50, 100, 150 and 200 cm from the ceiling.Plywood and oxygen limit of 15% was used in this discussion. Thedotted lines at 400 s represent the occurrence of flashover in theexperiment and the FDS simulation. After the flashover, the simu-lated temperature continued to increase up to 1200 �C, higher thanin the experiment (1000 �C).

At the upper layer (10e50 cm from ceiling), the temperaturestarted to increase after ignition in both experiment and simula-tion. But the simulated temperatures are 100e600 �C higher thanthe experiment. Unlike the experimental results, the differencesbetween the simulated temperatures at the five points under theceiling (labeling TC-1e-5) are relatively small, which show a well-mixed status at upper layer in the simulation. At the middle andlower layer (100, 150 and 200 cm from ceiling), a great difference is

observed. In the experiment, the temperatures increase constantlyand reached a maximum of 800e900 �C at about 650 s whichcorresponds to the second peak of heat release rate. However, in thesimulation, there is a ramp up in the temperatures when theflashover occurred. The simulated temperatures at middle andlower layer showed two peaks at 500 s and 850 s which corre-sponded to the heat release rate peaks. At these peaks, a largeamount of smoke was produced and descended toward the floor.The rise of temperatures at middle layer may indicate that the heattransferred from smoke was overestimated.

In general, based on the measurements of TC1 to TC21, thequalitative trends of the temperature changes in the simulationwere similar to the experimental results at the upper layer.However the temperatures were 100e600 �C over predicted in thesimulation.

5. Conclusions

Experiments were conducted in a full-scale model room equip-ped with both movable and fixed fire loads to explore fire growthand spread. Numerical simulationswith parameter adaptationwerethen carried out using FDS software to predict the fire features andwere compared with the experimental results. During the firegrowth period, the experimental results correlate with the numer-ical simulations, including flashover time, the highest heat releaseand required time to reach the heat release peak. With respect tocontinuous burning, the difference between the experimentalresults and simulation increased. The material properties are influ-ential factors for the simulation results. Spruce and plywood wereused for comparison to investigate the influence of density and ashcontent on the growth and spread of fires. The oxygen limit directly

Fig. 10. Comparison of temperature variations of thermocouple trees TC7, TC8 and TC9for the experimental results and simulation.

Fig. 11. Comparison of temperature variations of thermocouple trees TC13, TC14 andTC15 for the experimental results and simulation.

C.-J. Chen et al. / Building and Environment 45 (2010) 2735e2742 2741

Author's personal copy

affects the heat release trend and total heat release. By settinga lower oxygen limit, the overall simulated heat release rate is closerto the experimental value, and the total heat release is higher;however, the oxygen consumption is not affected. Regarding thetemperature distribution, the pattern of temperature variation atupper layer can be simulated. And the prediction of flashover con-sisted with the experiment. The simulated temperatures are closerto the experiment near the ceiling, but it is about 100e600 �C higherthan the experiment. At the middle and lower layers, the tempera-tures differed greatly from the experiment.

Acknowledgements

Support from the Architecture and Building Research Institute,Ministry of Interior, Taiwan is gratefully acknowledged.

References

[1] McGrattan KB, Hostikka S, Floyd JE, Baum HR, Rehm RG. Fire dynamicssimulator (Version 5), technical reference guide. NIST special publication1018-5. Gaithersburg, Maryland: National Institute of Standards and Tech-nology; 2007.

[2] Chow WK, Zou GW. Numerical simulation of pressure changes in closedchamber fires. Building and Environment 2009;44:1261e75.

[3] Zhang JY, Lu WZ, Huo R, Feng R. A new model for determining neutral-planeposition in shaft space of a building under fire situation. Building and Envi-ronment 2008;43:1101e8.

[4] Chow WK, Zou GW. Correlation equations on fire-induced air flow ratesthrough doorway derived by large eddy simulation. Building and Environ-ment 2005;40:897e906.

[5] Sun XQ, Hu LH, Li YZ, Huo R, Chow WK, Fong NK, et al. Studies on smokemovement in stairwell induced by an adjacent compartment fire. AppliedThermal Engineering 2009;29:2757e65.

[6] Gutiérrez-Montes C, Sanmiguel-Rojas E, Viedma A, Rein G. Experimental dataand numerical modelling of 1.3 and 2.3 MW fires in a 20 m cubic atrium.Building and Environment 2009;44:1827e39.

[7] Zou GW, Chow WK. Evaluation of the field model, fire dynamics simulation,for a specific experimental scenario. Journal of Fire Protection Engineering2005;15:77e92.

[8] Shen T-S, Huang Y-H, Chien S- W. Using fire dynamic simulation (FDS) toreconstruct an arson fire scene. Building and Environment 2008;43:1036e45.

[9] Kim SC, Ryou HS. An experimental and numerical study on fire suppressionusing a water mist in an enclosure. Building and Environment2003;38:1309e16.

[10] Jahn W, Rein G, Torero JL. The effect of model parameters on the simulation offire dynamics. Fire Safety Science 2008;9:1341e52.

[11] Rein G, Torero JL, Jahn W, Stern-Gottfried J, Ryder NL, Desanghere S, et al.Round-robin study of a priori modelling predictions of the Dalmarnock FireTest One. Fire Safety Journal 2009;44:590e602.

[12] Rein G, Empis CA, Carvel R. The Dalmarnock fire tests: experiments andmodelling. School of Engineering and Electronics, University of Edinburgh;2007.

[13] Huggett C. Estimation of rate of heat release by means of oxygen-consump-tion measurements. Fire and Materials 1980;4:61e5.

[14] Ma TG, Quintiere JG. Numerical simulation of axi-symmetric fire plumes:accuracy and limitations. Fire Safety Journal 2003;38:467e92.

[15] McGrattan KB, Baum HR, Rehm RG. Large eddy simulations of smoke move-ment. Fire Safety Journal 1998;30:161e78.

[16] Kwon J. Evaluation of FDS V.4: upward flame spread. MA, USA: WorcesterPolytechnic Institute; 2006.

C.-J. Chen et al. / Building and Environment 45 (2010) 2735e27422742