Vous êtes sur la page 1sur 9

Building and Environment 46 (2011) 1065e1073

Contents lists available at ScienceDirect

Building and Environment


journal homepage: www.elsevier.com/locate/buildenv

A posteriori modelling of the growth phase of Dalmarnock Fire Test One


Wolfram Jahn, Guillermo Rein*, Jose L. Torero
School of Engineering, University of Edinburgh, Edinburgh, Scotland, EH93JL, UK

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 5 September 2010
Received in revised form
8 November 2010
Accepted 11 November 2010
Available online 18 November 2010

The challenge of a posteriori (i.e. after the event) modelling of a well characterized full-scale re test using
computational uid dynamics is illustrated in this work. Test One of The Dalmarnock Fire Tests was
conducted in a 3.5 m by 4.75 m by 2.5 m concrete enclosure with a real residential fuel load. It provides
measured data at the highest spatial resolution available from a re experiment to date. Numerical
simulations of the growth phase have been conducted with the numerical code Fire Dynamics Simulator
(FDSv4) while having full access to all the measurements. This includes description of the fuel load,
compartment layout, temperature and heat ux data, and camera footage. No previous re simulation had
this large amount of data available for comparison. Simulations were compared against average and local
measurements. The heat release rate is reconstructed from additional laboratory tests and upper and
lower bounds for the re growth are found. Within these bounds, the evolution of the average hot layer
temperatures in time could be reproduced with 10e50% error. Worse agreement with local measurements
of gas temperatures was achieved (between 20% and 200% error). Wall temperatures were predicted with
less than 20% error and wall incident heat uxes within 50e150% error. Largest discrepancy is seen close to
the ames. This level of agreement with the measured data was achieved after exploratory simulations
were conducted and several uncertain parameters were adjusted and readjusted while having full access
to all the measurements.
2010 Elsevier Ltd. All rights reserved.

Keywords:
Compartment re
Experiment
CFD
LES
FDS

1. Introduction
Modelling of compartment res using computational uid
dynamics (CFD) has been a research topic since the introduction of
computational techniques in re science in the 1980s [1]. Only in
the last decade the available computational power and knowledge
of re dynamics have grown sufciently to carry out simulations
in real-size building enclosures, using grids that are ne enough
to reproduce re-driven ows reasonably well [2]. Since then, CFD
has been used extensively to model enclosure re dynamics [3e5]
both in research and in industrial applications. There are two
common industrial uses of CFD. One is for design of the re safety
strategies in the built environment (which results are rarely made
available for public scrutiny) and the reconstruction of accidental
res as part of forensic investigations (recent examples are the
2001 WTC [6] and the 2003 Station Nightclub [7] investigations).
The state of the art of re modelling is such that given a re of
known size and power (evolution of the heat release rate (HRR)),
CFD calculates the resulting temperature and smoke concentration
elds. The re source is therefore treated as an input into the model

* Corresponding author. Tel.: 44 131 650 7214; fax: 44 131 651 3576.
E-mail address: reingu@gmail.com (G. Rein).
0360-1323/$ e see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.buildenv.2010.11.001

by means of a prescribed HRR as a function of time (e.g. [8,9]). This


poses a problem in the study of accidental res where the HRR is
unknown. Predicting the evolution of the HRR (i.e., spread rate and
growth pattern) instead of measuring it is among the most challenging pending issues in re research [9,10].
Most modelling work in the literature corresponds to scenarios
with simple re sources, like pool res or a single burning item of
constant or near constant HRR. This type of scenario avoids the
more complex processes of ame spread and re growth observed
in real res. Little research has been done comparing simulations
with real-scale re tests that use realistic fuel loads. Some of most
important examples using pool or crib res are presented here.
Reneke et al. [11] conducted a posteriori simulations of re tests
involving crib res in a full-scale single compartment with a zone
model obtaining reasonable agreement with the measured average
temperatures when the measured HRR is used as an input. Miles
et al. [12] obtain good results in average temperatures when performing a posteriori simulations of a series of re tests involving
wood cribs. Although local measurements where available, neither
of these papers compared results at the eld or local level. Pope
et al. [8] conducted a posteriori simulations of the 1999e2000 BRE
large compartment test series [13] using the CFD code FDS. The
simulations provided reasonable agreement with the measured
temperature eld within the context of structural re safety. Rinne

1066

W. Jahn et al. / Building and Environment 46 (2011) 1065e1073

Fig. 1. Room layout with location of furniture and sensors [15,16].

et al. [14] found temperature proles, smoke layer heights and gas
species concentrations in a 10  10  5 m compartment predicted
by FDSv4 to be in good agreement with experimental data for the
burning of pool res (heptane or toluene) and single cribs (PMMA
or wood). The measured mass loss rate of the fuel was used to
calculate the HRR, which was then input into the model.
The evaluation of the entire process of re modelling, in which the
mathematical model is only a component, is an important task for the
advancement of re safety engineering. The state of the art of re
modelling is reected not only in the mathematical models capabilities, but also on how it is implemented throughout the different
stages of re modelling [9]. The assumptions made by the user, the
collection of data for input and the selection of the parameter values
are crucial components leading to the creation of the input le. This is
particularly important when advanced and complex computational
tools are used (e.g. CFD and evacuation modelling). Under the current
state of the art, there are many ill-dened and uncertain parameters
within the models which cannot be rigorously and uniquely determined. Thus, there is plenty of space for uncertainty and doubt to
unravel, and for curve tting and arbitrary parameter value selection
to take place. This is summarised best in the words attributed to the
German scientist Carl Friedrich Gauss (1777e1855): Give me four
parameters, and I will draw an elephant for you; with ve I will have him
raise and lower his trunk and his tail.
This paper reports a series of CFD simulations (with FDSv4) conducted a posteriori to reproduce the large-scale Dalmarnock Fire Test
One that involved several real burning items leading to a complex re
spread process. The HRR of the growth phase was not measured. But
because the compartment was densely instrumented, it provides
a very rich set of data that could be used to aid dening the HRR of the
re source and uncertain parameters to reproduce the main features
of the re. A key difference between previous modelling studies of
large-scale tests and the present work is that the heavily instrumented full-scale Dalmarnock Fire Tests provides for the rst time
measurements at sufcient spatial resolution to be suitable for
comparison with eld models at the grid size level.
2. Dalmarnock Fire Test One
Detailed information about the experimental set-up and the
chain of events that occurred during the Test One can be found

elsewhere [15,16], but a short summary is given here. Test One was
held in a two-bedroom single family at, with the living room
set-up as the main experimental compartment. This compartment
was 3.50 m by 4.75 m wide and 2.45 m high and made of concrete
walls. It had two-pane windows as shown in Fig. 1. While the main
source of fuel was a two-seat sofa stuffed with exible polyurethane foam, the compartments also contained two ofce work
desks with computers, each with its own foam-padded chair, three
tall wooden bookcases, a short plastic cabinet, three small wooden
coffee tables, a range of paper items and two tall plastic lamps. A
plastic waste paper basket lled with crumpled newspaper and
300e500 ml of heptane was the ignition source. The re spread to
a blanket on the sofa hanging into the basket, igniting the seating
area of the sofa. After about 275 s the bookshelf next to the sofa
ignited and was rapidly engulfed in re. Within 25 s after ignition of
the bookshelf the compartment reached ashover conditions (see
Table 1 for a time line of the events).
Fig. 2 shows the HRR, the average temperature of the hot layer
and the vertical distribution of temperatures along one specic
thermocouple rack measured in Test One. A detailed presentation
of the data can be found in [15,16]. The HRR in the re tests was
estimated by the principle of oxygen depletion, using the
measurements from the gas velocity probes at all the ventilation
openings and assuming total O2 consumption [16]. An estimation
for the post-ashover re results in about 3 MW before the rst
window breakage, and about 5 MW after the second breakage as
shown in Fig. 2. During the growth phase the assumption of total
O2 consumption does not hold, as the re is not ventilation
controlled, thus the evolution of the HRR during re growth is
unknown.
Table 1
Time line of main events during Dalmarnock Test One [15,16].
Event

Time from ignition (s)

Ignition
Cushions ignite
Bookcase ignites
Fire engulfs bookcase (ashover)
Compartment window breakage (NW Pane)
Compartment window breakage (SW Pane)
Extinction

0
9
275
300
800
1100
1140

W. Jahn et al. / Building and Environment 46 (2011) 1065e1073

1067

Fig. 2. Measurements from the Dalmarnock Test One: a) estimated HRR (using oxygen consumption), b) average temperature of the hot layer in the main compartment (corrected
for radiation) and c) vertical temperature distribution in front of the compartment door at 600 s [15,16] (rack 4, see Fig. 1).

Prior to Test One, laboratory experiments were conducted in


order to determine the HRR curves for the sofa and the bookshelves inside the furniture calorimeter (using an exact replica of
those used in Test One) [15]. A second series of laboratory
experiments was conducted after Test One in order to analyse
variations to the ignition protocol [17]. These two sets of experiments were the only direct source of information on the HRR in
Test One.
Fig. 3 shows the evolution of the HRR of the sofa and the effect of
variations to the ignition source. The experiment conducted prior to
Test One (referred to as Set 1) consisted of a sofa with two cushions,
and a waste paper basket located next to the sofa. Accelerant was
poured into the paper basket and the basket was ignited. The re
then spread over the armrest to the sofa and was stopped about
800 s after ignition when approximately one third of the sofa had
been burnt. In the experiments conducted after Test One (referred to
as Set 2), the exact ignition protocol of Dalmarnock Test One was
replicated, which was like Set 1 but including a blanket that had
been placed over the armrest of the sofa, and the accelerant
distributed between basket and blanket. The presence of the blanket
in Set 2 allowed the re to bypass the armrest re barrier and led to
a faster growth rate involving the cushions. The re growth rate for
Set 2 is equivalent to a t-squared re with growth constant 1.6 W/s2
(a medium re [18]). The constant in Set 1 is 0.5 W/s2 (corresponding
to a slow re [18]). The two tests show that the uncertainty in the
growth rate of the sofa re during the early stages is signicant and
varies between a slow and a medium re.
Fig. 3 suggests two patterns in the burning behaviour; an initial
peak that rapidly decreases is followed by a growing re that
resembles a t-squared curve. A similar behaviour can be seen for Set
2. It is conjectured that the initial peak corresponds to the waste
paper basket, accelerant and the blanket (in the case of Set 2), while
the t-squared re corresponds to the sofa itself.

By measuring ame heights from visual recordings, it can be


established that the sofas HRR in Set 1 grew slower than the
observed re in Test One. Flashover in Test One occurred at 300 s.
According to Set 1, this is only 100 s after the period when the basket
re dominated the HRR. Comparison of the re development using
the video available of both res conrms this. At the initial laboratory
experiment no blanket was used, so that the re could not spread
across the armrest to the cushions as fast. The re in Dalmarnock Test
One spread quickly from the basket over the armrest to the sofa,
whereas in Set 1, the ames took considerably longer to spread over
the armrest.
3. A priori vs. a posteriori modelling
Before the Dalmarnock Tests were carried out, a round-robin
study of blind predictions [9] was conducted in order to explore the
a priori predictive capabilities of re modelling in realistic scenarios.
The aim of the exercise was to forecast the re development as
accurately as possible and compare the results. Comparison of the
modelling results showed a large scatter and considerable disparity
among the predictions, and between predictions and experimental
measurements. The scatter of the simulations was much larger than
the error and variability expected in the experiments. The study
emphasized on the inherent difculty of modelling re dynamics in
complex re scenarios like Dalmarnock, and showed that the
accuracy of blind prediction of re growth (i.e. evolution of the heat
release rate) is poor.
The present work revisits the modelling of the Dalmarnock Fire
Test One, this time using the large set of measurements available.
That is, the work is conducted a posteriori. Many different simulations are conducted and many parameters are adjusted and readjusted until acceptable agreement is reached. Work is focused only
on the growth phase of the Dalmarnock Fire Test One.

Fig. 3. Measured HRR for variations to the ignition source and equivalent t-squared res: a) measured in the laboratory prior to Test One including sofa and waste basket (Set 1);
and b) after Test One including the sofa, waste basket and blanket (Set 2).

1068

W. Jahn et al. / Building and Environment 46 (2011) 1065e1073

As mentioned before, during the growth phase of Test One


the evolution of the HRR is unknown. Moreover, the laboratory
experiments of the sofa and similar ignition sources show signicant uncertainty. Indeed, this is one of the most important issues
addressed in this article. The challenge is to be able to reproduce
the HRR such that the re environment (temperature and smoke)
is simulated correctly. The work solves a large and complex inverse
problem by trial and error.
4. Computational model
The code Fire Dynamics Simulator v4.07 [19,20], one of the most
commonly used re CFD codes, is used here. FDS solves a form of
the NaviereStokes equations adequate for low-speed thermally
driven ows. While large eddies are solved directly, turbulences at
subgrid scale are modelled using Smagorinskys approach [21].
FDS is a tool still under active development, and improved
versions are released with some frequency. At the time this paper
goes to press, FDSv5 had been released and FDSv6 is expected
soon. But research has a characteristic time that seems to be longer
that the time between version releases of FDS. This means that by
the time a research project has been completed and conclusions
have been reached, newest versions might be available. While not
all of the ne results of this paper might apply directly to FDSv5 or
FDSv6, the merit of the study is that the bulk of the conclusions
relate to the signicant amount of papers in the technical literature where FDSv4 was used, and to current re safety engineering.
This last is particularly important for infrastructure designed,
approved and built, and to forensic ndings agreed on with the aid
of FDSv4.
The computational domain used reproduces in detail the main
Dalmarnock compartment (Fig. 4) and includes the vent openings
(two-pane windows to the exterior), the nearby kitchen (with
another window to the exterior) and the hallway connecting to
the apartment entry. It was considered that the other rooms of
the apartment did not contribute signicantly to the re dynamics
and therefore were not included. The fuel load in the main

compartment was reduced to the fuel elements involved in the


growth phase; the sofa, the ignition source and two bookshelves in
the corner behind the sofa. The other fuel items, the computers,
coffee table, chairs and desks contributed to the re only after
ashover, and therefore are outside the scope of this work.
The compartment walls and ceiling were modelled as 10 cm
thick concrete, and the compartment oor as covered with a carpet.
Open boundary conditions were used at the exterior of the
compartment at some distance from the building facades.
FDS models combustion based on a mixture fraction approach
[19] with a single-step instantaneous reaction chemistry. While the
Dalmarnock Fire Test One contained several dozen different
fuels burning at the same time, FDSv4 only allows for one single
fuel. Thus, as an overall representative, propane was assumed as
fuel (FDS default fuel) with combustion reaction:

C3 H8 5O2 /3CO2 4H2 O

(1)

The time variation of the HRR is imposed as a ramp of the fuel


injection, while the re area is assumed constant at 2/3 of the seating
area of the sofa. Because the re area affects the air entrainment into
the ame, in principle, it should be changed in time as well. This has
been shown in previous sensitivity studies with FDSv4 [22].
However, the uncertainty in the actual rate of change of the re area
is large. Therefore, it has not been considered here for the sake of
simplicity but its effects can be investigated in future studies
together with other uncertain combustion, turbulence and radiation
parameters within FDS.
The grid size is one of the most critical parameters in numerical
simulations. In the scope of this work a large number of simulations
had to be run in order to converge toward good agreement with the
measured data and it was therefore necessary to use a sufciently
coarse grid allowing for an efcient use of computational resources.
However, too coarse a grid could induce signicant numerical
errors in the solution. It has been proposed [23,24] that to resolve
the re plume properly, the ratio between the characteristic re
diameter, D*, and the grid size should be at least 5 to 10, where D* is
dened as:

Fig. 4. Computational domain and fuel items used in the FDS simulations.

W. Jahn et al. / Building and Environment 46 (2011) 1065e1073

D*

$
2=5
Q
:
p
rN cp TN g

(2)

During the growth phase, the peak of the HRR is 300 kW, and
the characteristic diameter of the sofa is around 0.6 m. Hence the
grid size should be smaller than 11 cm for an adequate plume
resolution. This value is used as an order of magnitude reference. In
order to select an adequate grid size, simulation were run with
a wide range of different grid sizes: 5 cm, 10 cm, 15 cm and 20 cm
edge cubes. The 10 cm grid size resulted in 120,000 cells, and the
number of cells in other grids scales with the inverse of the cube of
the grid size. The HRR was prescribed according to the laboratory
experiment Set 1.
Fig. 5 shows the temperature vs. height distribution at two
different locations in the experimental compartment, one near the
burning sofa, and one near the window away from the re. See Fig. 1
for rack locations. The time of comparison is 140 s, and data are
averaged in time (10 s) in order to lter out inconsistent uctuations.
The simulations using coarser grids (15 and 20 cm) actually showed
better agreement when compared to the measured data, both
qualitatively and quantitatively than the ner grids (5 and 10 cm).
Near the re (Fig. 5a), the 20 cm and 15 cm grids are in agreement
between them and reproduce well the lower zone temperatures and
the smoke layer height, but overpredict the temperatures of the
smoke layer by 17%. The 10 cm grid underpredicts the smoke layer
height by around 15%, but is in good agreement regarding the smoke
layer temperature. The 5 cm grid underpredicts the measurements
by more than 35% at all heights. Away from the re (Fig. 5b), the 10,
15 and 20 cm grids are in reasonable agreement to the measurements, while the 5 cm grid shows divergence both quantitatively
(>30%) and qualitatively.
Based on this grid dependency study, the 10 cm grid was chosen
because it showed good comparison to the experiments, allowed
for fast computations and complied with the recommendation
associated to the characteristic re diameter [23,24]. Given the
present results, there are no good reasons to choose a different grid,
but most validation studies by the FDS developers had been conducted with similar grid sizes to this.
We cannot explain why the coarser grids in this case provide
lower global errors. While it is generally accepted for numerical
integration that a smaller grid yields more accurate results, FDS
results are known to depend signicantly on the size of the numerical
grid due to the LES approximation [25,24]. A possible explanation is
that the LES Smagorinsky constant used by default in FDS was calibrated to predict smoke movement in a very large enclosure and

1069

Fig. 6. Addition of the blanket to the HRR measured in Set 1 (Measured sofa re) to
obtain Set 1b (combined re).

away from the ames [26]. This is a very different scenario to the
Dalmarnock Test One and the constant might need recalibration.
All other parameters not discussed here were left with the
default values in FDSv4.
5. Hot layer average temperatures
Two distinct levels of detail are analysed. This section predicts
the average hot layer temperature while the next section looks at
the distribution of local eld temperature and wall heat uxes.
Comparison at detailed level requires a good agreement at averaged level rst.
As seen in the Set 1 and Set 2 experiments, the blanket signicantly modies the HRR. A possible HRR of the blanket alone can
be estimated assuming quadratic growth and decay phases. The
blanket, made of cotton, weighed 1 kg, hence the total combustible
energy stored in the blanket including 100 ml of accelerant can be
estimated to be around 21 MJ [22], assuming a heat of combustion
of 16.5 MJ/kg for the cotton. The resulting peak HRR would be
150 kW. This is not negligible compared to the measured HRR of
the sofa in the early stages, and should therefore be included
in the input HRR for the model. Fig. 6 shows the reconstructed HRR
of the blanket alone and its addition to the HRR of Set 1. The

Fig. 5. Comparison of results with different grid sizes at 140 s: a) distribution at the north wall next to the bookshelf (rack 1); and b) distribution near the window (rack 17).

1070

W. Jahn et al. / Building and Environment 46 (2011) 1065e1073

Fig. 7. Predicted average hot layer temperatures; a) Set 1; b) Set 1b and c) Set 2.

resulting HRR is referred to as Set 1b. Note that here the data is
shifted by 150 s compared to Fig. 3, so that the ignition occurs at 0 s.
The predicted hot layer average temperature using different
HRR and comparison to measurements are shown in Fig. 7. For the
Set 1 (Fig. 7a), the simulated hot layer temperature rise agrees with
the measurements during the rst 100 s. The predicted average
temperatures after t 100 s are lower that measured (up to 50%
error). For Set 1b (Fig. 7b), the simulated hot layer temperatures are
in good agreement with the measured temperatures (within the
instrumental uncertainty) until 200 s into the re. After that, the
simulated temperature decreases in contrast to the measured
temperature which rises continually until ashover. For Set 2
(Fig. 7c), the average hot layer temperature is overpredicted by
about 50% between 25 s and 100 s, and by about 35% between 150 s
and 250 s.
Overall, it is seen that Set 1 results in unrealistically low average
temperatures for the hot layer, but Set 1b and Set 2 provide predictions closer to the measurements. Thus, it is concluded that Set 1b
provides a lower bound to the HRR curve during the growth phase,
while Set 2 is an upper bound. This range captures the intrinsic
uncertainty of re growth in real complex scenarios and also includes
experimental variability. The uncertainty in the average temperature
predictions using these HRR bounds is between 0 and 50%. It conrms
that to predict temperatures with reasonable agreement, the HRR
must be well characterized.

6. Comparison of local measurements


The density of measurements in Test One presents an opportunity to assess eld level simulations. Fig. 8 shows the temperature vs. height distributions at different locations (three vertical
thermocouple racks) in the experimental compartment using the
two bounding HRR curves Set 1b and Set 2. Two different times,
150 s (HRR plateau) and 200 s (rapid HRR rise), are chosen arbitrarily for illustration purposes. In general the simulations are in
reasonable agreement with the measurements, although it can be
seen that the further away from the burning sofa the better the
agreement between measurements and simulations.
As expected the lower and upper HHR bounds result in upper and
lower bounds for the smoke temperatures. It is observed that the
thickness of the hot layer is not signicantly affected by the HRR
within the bounds, but the upper HHR leads to higher temperatures
in smoke. Within the cold layer, the simulations underpredict the
temperature even for the upper HHR bound. This indicates that the
Dalmarnock Fire Test had a less well dened hot layer than predicted.
At thermocouple rack 7, roughly 1.5 m from the re, the
temperatures are underpredicted by 20e200% at 150 s (Fig. 8a)
when using both the lower HRR bound (Set 1b) or the upper HRR
bound (Set 2). At rack 11, near the centre of the compartment, and
150 s into the re (Fig. 8b) the simulations with both input HRR
curves are in good agreement with the measured data lying within

Fig. 8. Temperature vs. height distribution at different locations in the compartment at 150 s (a, b and c) and 200 s (d, e and f). Thermocouple rack 7 (a, d) is located near the re,
rack 11 (b, e) is located near the centre of the compartment, and rack 19 (c, f) is located near the window. See Fig. 1 for rack location.

W. Jahn et al. / Building and Environment 46 (2011) 1065e1073

1071

Fig. 9. Wall temperature vs. height (a and c) and Heat ux vs. height (b and d) distribution on the wall of the re compartment against the kitchen at times 150 s (a and b) and 200 s
(c and d).

the experimental error in the hot layer, but underpredicting the


temperatures in the cold layer by about 40%. At rack 19, near the
window, and 150 s (Fig. 8c), the lower HRR bound produces
temperatures that lie within the experimental error in the hot layer,
while the upper HRR bound results in overprediction by about 30%.
At 200 s the hot layer temperatures at rack 7 are slightly overestimated (although close to the experimental error, Fig. 8d) when
using the upper HRR bound, and underestimated by around 35%
when using the lower HRR bound. The cold layer temperatures are
underestimated by 50e100% for both input HRR curves. At rack 11
temperatures simulated with the lower HRR bound are underpredicted by less than 20% in the hot layer (Fig. 8e), having similar
shape to the measured distribution. With the upper HRR bound as
input, the temperatures are overpredicted by about 25% in the hot
layer, but underpredicted by about 35% in the cold layer, thus indicating higher temperature differences between hot and cold layer.
Near the window, at rack 19 (Fig. 8f), the temperatures (specially in
the hot layer) are in good agreement with the lower HRR bound as
input, although the predicted hot layer height is around 0.5 m
higher than that observed. But the upper HHR bound overpredicts
the hot layer temperature by 40%, which is in accordance to the
overprediction by around 40% of the hot layer average temperature
resulting from the upper HRR bound at 200 s (Fig. 7c).
In addition to gas phase temperatures, the simulated wall
temperatures and wall heat uxes are compared to the measurements in Fig. 9 at two different times. The ux for both measurement
and predictions is the total incident heat arriving to the location.
Using the lower HRR bound, the wall temperature is underpredicted
by about 20% at 150 s, and within (or very close to) the experimental
error range at 200 s. For the upper HRR bound, the simulated wall
temperature lies within the experimental error at both times. But this
apparently best agreement of the upper HRR bound for wall
temperatures contradicts the corresponding results for the wall heat

uxes. While the lower HRR bound produces heat uxes in the hot
layer close to the experimental error at 150 s, the upper HRR bound
overpredicts heat uxes in the hot layer by 20% and the cold layer by
100%. Fig. 9 shows that at 200 s the predicted wall heat ux with the
lower HRR bound is in reasonable agreement with the measurements, both qualitatively and quantitatively. The upper HRR bound
overpredicts of the heat uxes by 60e150%.
Fig. 10 shows the comparison of the distribution of heat uxes
on the ceiling along two perpendicular directions at two different
times. The simulated heat ux to the ceiling agrees with the
measured heat ux when using the lower HRR bound (within 20%
error). Using the upper HRR bound, the heat uxes to the ceiling are
consistently overpredicted (20e100%). The better predictions of the
ceiling heat ux compared to the wall heat ux (see Fig. 9) for the
lower HRR bound could be explained by the fact that the radiation
from the ame to the ceiling is reduced by the presence of the
smoke layer. The inaccuracies when modelling the ame could
have a smaller effect on the ceiling heat ux.
The difculty of modelling the near eld of the re on one hand,
and the relatively good agreement of the results away from the re
on the other, is consistent throughout this paper and with previous
work [9].
7. Conclusions
This work presents a detailed account of the a posteriori
modelling of Dalmarnock Test One. A very large pool of scenario
descriptions and measurements is used together with possible
HRR curves. Simulations were then compared against detailed
measurements, and uncertain input parameters were adjusted in
order to t the results. It was shown that even with full access to
the measurements, it is not easy to satisfactorily reproduce the
course of the re. Although averaged quantities were in reasonable

1072

W. Jahn et al. / Building and Environment 46 (2011) 1065e1073

Fig. 10. Heat ux to the ceiling compartment; a) and c) heat ux vs. longitudinal distance in the direction from the west window to the kitchen wall along the middle of the
compartment at 150 s and 200 s respectively. b) and d) heat ux vs. transversal distance in the direction from the south bedroom wall to the north wall along the middle of the
compartment at 150 s and 200 s.

agreement with the measurements when the HRR is known (<10%


error for Set 1b and <50% for Set 2), comparison of the simulations
with the measurements at a local level showed larger errors, from
20% to 200% error in the gas phase temperatures, around 20% for
wall temperatures, and between 50e100% for wall incident heat
uxes.
Hot layer temperatures were consistently overpredicted when
using an upper HRR bound (25e40%), and underpredicted when
using a lower HRR bound (less than 20%). However, it was observed
that the cold layer temperatures were underpredicted by both
upper and lower HRR bounds. The higher HRR input raised the hot
layer temperatures without signicantly increasing the cool layer
temperatures, thus predicting a larger gradient between hot and
cold layers than measured.
Comparison of the eld temperatures in the experimental
compartment suggested that far away from the re, the simulations
are capable of capturing the temperature vs. height distribution of
the gas phase (provided an acceptable HRR is input), while close to
the re important differences between simulations and measurements were consistently observed, both qualitatively and quantitatively (from 20 to 200% error).
Although it was possible to reconstruct a range of suitable HRR
curves that provided upper and lower bounds, the incapability of
predicting re growth is shown to be a fundamental constraint
to re modelling. This is an important issue for the application of
re modelling to real scenarios when the HRR is unknown
(ie, forensic investigations and assumed design scenarios).
A signicantly long list of parameters in the turbulence,
combustion, and radiation submodels and numerics within FDS
remain to be studied in this work, both independently and combined.
Some of these parameters are known to carry a signicant amount

of uncertainty and to be important for re modelling [9,10,22,27].


However, they have not been considered here for the sake of
simplicity but their effects should be investigated in future studies.
The work is not intended to be a validation study but serves in
describing and showing explicitly the challenges of modelling
a complex re scenario. As illustrated here, when the HRR is
unknown as it is in most practical cases, the use of lower and upper
HRR bounds should be included as to reect in the predictions the
effect of uncertainty in the HRR.
Acknowledgements
The support for W. Jahn from the European Union Programme
AlBan of High Level Scholarships for Latin America (Grant No.
E06D100038CL) and BRE Trust is gratefully acknowledged. This
work forms part of FireGrid, www.regrid.org, and has been
partially funded by the UK Technology Strategy Boards Collaborative Research and Development Programme. The authors are
grateful for the funding and the help of the many organisations
involved in the Dalmarnock Fire Tests, BBC Horizon programme,
Engineering and Physical Sciences Research Council, The University
of Edinburgh, Glasgow Housing Authority, Strathclyde Fire &
Rescue Service, Glasgow Caledonian University, BRE Trust, Xtralis
and Powerwall Systems. Our gratitude to the Building and Fire
Research Laboratory at NIST for developing and making it available
the computer re model FDS.
References
[1] Emmons H. The further history of re science. Combust Sci Technol 1984;40
(1):167e74 (reprinted in Fire Technology 21(3), pp. 230e238, 1985. doi:
10.1007/BF01039976).

W. Jahn et al. / Building and Environment 46 (2011) 1065e1073


[2] McGrattan K. Fire modelling: where are we? where are we going? Fire Saf Sci
2005;8:53e68. doi:10.3801/IAFSS.FSS.8-53.
[3] Ma T, Quintiere J. Numerical simulation of axi-symmetric re plumes: accuracy and limitations. Fire Saf J 2003;38(5):467e92.
[4] Hasib R, Kumar, Shashi R, Kumar S. Simulation of an experimental compartment re by CFD. Build Environ 2007;42(9):3149e60.
[5] Lattimer B, Hunt S, Wright M, Sorathia U. Modeling re growth in a combustible corner. Fire Saf J 2003;38(8):771e96.
[6] McGrattan K, Bouldin C, Forney G. Computer Simulation of the Fires in the
WorldTrade Center Towers (Draft). Technical report. NIST; 2005.
[7] Grosshandler W, Bryner N, Madrzykowski D. Report of the Technichal
Investigation of The Station Nightclub Fire. Technical report. NIST; 2005.
[8] Pope N, Bailey C. Quantitative comparison of FDS and parametric re curves
with post-ashover compartment re test data. Fire Saf J 2006;41(2):
99e110.
[9] Rein G, Torero JL, Jahn W, Stern-Gottfried J, Ryder NL, Desanghere S, et al.
Rounderobin study of a priori modelling predictions of the Dalmarnock Fire
Test One. Fire Saf J 2009;44(4):590e602. doi:10.1016/j.resaf.2008.12.008.
[10] Kwon JW, Dembsey N, Lautenberger C. Evaluation of FDS v.4: upward ame
spread. Fire Technol 2007;43(4):255e84. doi:10.1007/s10694-007-0020-x.
[11] Reneke P, Peatross M, Jones W, Beyler C, Richards R. A comparison of CFAST
predictions to USCG real-scale re tests. J Fire Protect Eng 2001;11:43e68.
doi:10.1106/HH4D-0CKM-J53X-FQK1.
[12] Miles S, Kumar S, Cox G. Comparisons of blind predictions of a CFD model
with experimental data. Fire Saf Sci 2002;6:543e54. doi:10.3801/IAFSS.FSS.6543.
[13] Welch S, Jowsey A, Deeny S, Torero JL. BRE large compartment re tests e
characterising post-ashover res for model validation. Fire Saf J 2007;42(8):
548e67.
[14] Rinne T, Hietaniemi J, Hostikka S. Experimental validation of the FDS simulations of smoke and toxic gas concentrations. VTT Working Papers, 66; 2007,
http://www.vtt./inf/pdf/workingpapers/2007/W66.pdf; 2007.

1073

[15] Rein G, Abecassis-Empis C, Carvel R, editors. The Dalmarnock re tests:


experiments and modelling. The University of Edinburgh, ISBN 978-09557497-0-4; 2007, <www.era.lib.ed.ac.uk/handle/1842/2037>; 2007.
[16] Abecassis-Empis C, Reszka P, Steinhaus T, Cowlard A, Biteau H, Welch S, et al.
Characterisation of Dalmarnock re test one. Exp Therm Fluid Sci 2008;32
(7):1334e43. doi:10.1016/j.expthermusci.2007.11.006.
[17] Steinhaus T, Jahn W. Full-scale furniture tests. Internal Publication BRE Centre
for Fire Safety Engineering, University of Edinburgh; 2007.
[18] Drysdale D. An introduction to re dynamics. 2nd ed. New York: Wiley &
Sons, ISBN 0-471-97290-8; 1998.
[19] McGrattan K. Fire dynamics simulator (Version 4) e Technical Reference
Guide. 6783. NISTIR; 2003.
[20] McGrattan K. Fire dynamics simulator (Version 4) e Users Manual. 6784.
NISTIR; 2003.
[21] Smagorinsky J. General circulation experiments with the primitive equations,
the basic experiment. Mon Weather Rev 1963;91(3):99e164. doi:10.1175/
1520-0493(1963)091<0099:GCEWTP>2.3.CO;2.
[22] Jahn W, Rein G, Torero JL. The effect of model parameters on the simulation of
re dynamics. Fire Saf Sci 2008;9:1341e52. doi:10.3801/IAFSS.FSS.9-1341,
<http://hdl.handle.net/1842/2696>.
[23] Dreisbach J, McGrattan K. Verication and validation of selected re models
for nuclear power plant applications, vol. 6. FDS. Technical report, U.S. Nuclear
Regulatory Commission, Ofce of Nuclear Regulatory Research; 2006.
[24] McGrattan K, Rehm R, Baum H. Large eddy simulations of smoke movement.
Fire Saf J 1998;30(2):161e78.
[25] Pope S. 10 Questions Concerning the large-eddy simulation of Turbulent
ows. New J Phys 2004;6(35).
[26] McGrattan K, Hostikka S, Floyd J, Baum H, Rehm R. Fire dynamics simulator
(Version 5) e Technical Reference Guide. 1018e5. NIST Special Publication; 2008.
[27] Yang D, Hu LH, Jiang YQ, Huo R, Zhu S, Zhao XY. Comparison of FDS predictions by different combustion models with measured data for enclosure res.
Fire Saf J August 2010;45(5):298e313. doi:10.1016/j.resaf.2010.06.002.

Vous aimerez peut-être aussi