Conference PaperPDF Available

Improvement of drag model for non-burning firebrand transport in Fire Dynamics Simulator

Authors:

Abstract

Firebrands play a crucial role in increasing the severity of wildfires by driving fire growth, damaging structures, and starting new fires. Predicting the transport of firebrands and their propensity to ignite new fires is of significant interest to fire communities. Developing an operational firebrand transport sub-model from the field studies is cumbersome, expensive, and has significant associated risks to equipment, community and firefighters. Physics-based models have the potential to assist in the development of such firebrand transport sub-models, which can be utilised to improve the efficacy of existing operational fire models. The present study showcases one of the initial works carried out in the development of such a physics-based firebrand model. The work utilises Fire Dynamics Simulator (FDS), a commonly used open-source physics-based fire model. The Lagrangian particle sub-model of FDS is used to simulate the transport of firebrand particles. The Lagrangian sub-model is generally used to model the transport of droplets and mist, and has been extensively validated. However, the validation of this sub-model for the transport of solid particles such as firebrands is limited. The issue is exacerbated when particles are of a non-spherical shape and can undergo complex reactions over their transport, such as burning. In this work, we utilise a firebrand generator prototype that produces a uniform Lagrangian shower of non-burning idealised firebrands. A set of in-house experiments are conducted to study the transport of three isometric shapes of non-burning firebrands i.e. cubiform, cylindrical and square-disc. These sets of experiments are used to quantify the efficacy of the inbuilt particle drag model of FDS and suggest potential alternative drag models that can be employed without loss of computational speed, major amendment in the fire model, are applicable to a wide range of particles shapes, and potentially improved prediction. In general, it is found that the suggested alternative Haider and Levenspiel drag model improves the estimation of firebrand distribution in terms of peak location, maximum and minimum longitudinal distribution with exception to cubiform particles for peak location. The exception is mainly due to inherent error association with the alternative drag model in overestimating the drag coefficient. For other situations, Haider and Levenspiel drag model shows either an improvement or stays the same. However, the study found that the existing point particle assumption to represent particles in FDS is not suited to estimate the lateral spread of firebrands, especially when the secondary motion of a particle on its axis is involved, such as cylindrical and square-disc particles. Our studies found, the lateral spread is found to be in the range of ~5-15% thinner compared to its experimental width for cylindrical particle distribution. For the square disc, it is not possible to quantify such differences due to the computational limit associated with our present study. It can be qualitatively suggested that it is found to be more than cylindrical particles. A further set of experiments and their numerical validation is required to ascertain the above finding, especially with different sizes, isometric and non-isometric shape, the speed of firebrand particles and burning process to establish the efficacy of a particular drag model for firebrand transport.
Improvement of drag model for non-burning firebrand
transport in Fire Dynamics Simulator
Rahul Wadhwani a,b, Duncan Sutherland a,b,c, Graham Thorpe d and Khalid A Moinuddin a,b
a Centre for Environmental Safety and Risk Engineering, Victoria University, Melbourne, Australia
b Bushfire and Natural Hazards CRC, Melbourne, Australia
c School of Science, University of New South Wales, Canberra, Australia
d College of Engineering and Science, Victoria University, Melbourne, Australia
Email: khalid.moinuddin@vu.edu.au
Abstract: Firebrands play a crucial role in increasing the severity of wildfires by driving fire growth,
damaging structures, and starting new fires. Predicting the transport of firebrands and their propensity to ignite
new fires is of significant interest to fire communities. Developing an operational firebrand transport sub-model
from the field studies is cumbersome, expensive, and has significant associated risks to equipment, community
and firefighters. Physics-based models have the potential to assist in the development of such firebrand
transport sub-models which can be utilised to improve the efficacy of existing operational fire models. The
present study showcases one of the initial works carried out in the development of such a physics-based
firebrand model. The work utilises Fire Dynamics Simulator (FDS), a commonly used open-source physics-
based fire model. The Lagrangian particle sub-model of FDS is used to simulate the transport of firebrand
particles. The Lagrangian sub-model is generally used to model the transport of droplets and mist and has been
extensively validated. However, the validation of this sub-model for the transport of solid particles such as
firebrands is limited. The issue is exacerbated when particles are of a non-spherical shape and can undergo
complex reactions over their transport such as burning.
In this work, we utilise a firebrand generator prototype that produces a uniform Lagrangian shower of non-
burning idealised firebrands. A set of in-house experiments are conducted to study the transport of three
isometric shapes of non-burning firebrands i.e. cubiform, cylindrical and square-disc. These sets of experiments
are used to quantify the efficacy of the inbuilt particle drag model of FDS and suggest potential alternative
drag models that can be employed without loss of computational speed, major amendment in the fire model,
are applicable to a wide range of particles shapes, and potentially improved prediction. In general, it is found
that the suggested alternative Haider and Levenspiel drag model improves the estimation of firebrand
distribution in terms of peak location, maximum and minimum longitudinal distribution with exception to
cubiform particles for peak location. The exception is mainly due to inherent error association with the
alternative drag model in overestimating the drag coefficient. For other situations, Haider and Levenspiel drag
model shows either an improvement or stays the same. However, the study found that the existing point particle
assumption to represent particles in FDS is not suited to estimate the lateral spread of firebrands especially
when the secondary motion of a particle on its axis is involved such as cylindrical and square-disc particles.
Our studies found, the lateral spread is found to be in the range of ~5-15% thinner compared to its experimental
width for cylindrical particle distribution. For the square disc, it is not possible to quantify such differences due
to the computational limit associated with our present study. It can be qualitatively suggested that it is found
to be more than cylindrical particles. A further set of experiments and their numerical validation is required to
ascertain the above finding, especially with different sizes, isometric and non-isometric shape, the speed of
firebrand particles and burning process to establish the efficacy of a particular drag model for firebrand
transport.
Keywords: Fire Dynamics Simulator (FDS), firebrand particles, Lagrangian particle, firebrand generator,
drag models
24th International Congress on Modelling and Simulation, Sydney, NSW, Australia, 5 to 10 December 2021
mssanz.org.au/modsim2021
400
Wadhwani et al., Improvement of drag model for non-burning firebrand transport in Fire Dynamics Simulator
1. INTRODUCTION
Every year billions of dollars are lost to wildfire in countries like the US, Canada, and Australia (Ronchi et al.,
2017). Recent wildfires such as the 2019-20 Black Summer in Australia, the 2019-20 Amazon fire in Brazil,
the 2019 Californian wildfires in the USA, the ongoing 2021 Evia Island fire in Greece are the prominent
incidences causing uncountable damage to the ecosystem with economic damage crossing billions of dollars.
These wildfires are also associated with the massive evacuation of communities which created another
challenge for fire and emergency services. A major contributor to the severity of these wildfires is the
production of firebrands and the resulting ignitions, which assist in increasing the rate of fire spread, ignition
of new fires, and damages the structures at the wildland-urban-interface (WUI).
Firebrands are burning pieces of twig, leaf, seed, and bark material that travel along with the wind and ignite
the vegetation ahead of the fire front. The ignition caused by the transport of firebrands is called spotting and
can be classified based on the distance travelled by firebrands as either: (1) short-range (<750m), (2) medium-
range (1-5 km), and (3) long-range (>5km) spotting (Cruz et al., 2015). In short-range spotting, the firebrands
travel along with the wind with little to no lofting and can travel up to 750m. Most of the short-range firebrands
are in flaming (as opposed to smouldering) states and are just broken material from trees. The effect of the
firebrands is incorporated using the Huygens’ wavelet principle model in operational wildfire models such as
FARSITE, PROMETHEUS, and PHOENIX (Sullivan, 2009b). In medium and long-range spotting, firebrands
lofted into the convective air column may travel from more than a kilometre. These firebrands are typically
burned off and usually in the glowing (smouldering combustion) state (Ellis, 2012).
Most of the research (Ellis, 2012; Fernandez-Pello, 2017; Manzello et al., 2020) in the last decade or so focuses
on long-range spotting which causes significant losses in a bushfire. The short-range firebrands contribute to
increasing the rate of fire spread by causing spotfires that coalesce with the primary fire front. There is very
limited significant research carried out in understanding the phenomena of short-range spotting. It has been
argued that poor prediction of firebrand transport and spot is a major reason for the under-prediction of fire
models used by emergency services (Cruz et al., 2015).
Fire dynamics simulator (FDS) developed by the National Institute of Standards and Technology (NIST) is one
of the most used open-source physics (Computational Fluid Dynamics (CFD))-based fire models (McGrattan
et al., 2015; Sullivan, 2009a). FDS solves a thermally driven Navier-Stokes equation to simulate the
propagation of fire. A Lagrangian particle model, one of the sub-models of FDS, is used to simulate the
transport of particles in the flow field. The model is extensively verified and validated for liquid particles like
droplets and mist (Mahmud et al., 2016). However, verification and validation for the transport of solid
particles are very limited. Wadhwani et al. (2017) attempted to verify the applicability of the inbuilt Lagrangian
particle model of FDS ver. 6.2.0 (McGrattan et al., 2015) for the transport of non-burning firebrands. We
observed that FDS inbuilt model under-predicts the spatial distribution of the cubiform and cylindrical
firebrand particles. It was anticipated that the simulation did not account for the secondary motion of a particle
on its axis due to its shape and thus providing a higher drag force on the particles. The particle model treats
particles as point particles that do not account for any rotation on its axis and are only translated due to shear
force. Furthermore, the inbuilt drag model is limited to only two shapes of particles viz. spherical and
cylindrical which limits its application for firebrand transport research.
The present work focuses on finding an alternative drag model which is effective to a wide range of isometric
shapes and fit in the current framework of FDS. We limited ourselves to the point particle approach due to two
main reasons. Firstly, the three-dimensional (3-D) physics-based fire models are computationally expensive
and are not applicable at a scale of more than a few hundred meters in area (Ronchi et al., 2017). Secondly, the
application fire models like PHOENIX, SPARK, and FARSITE are operated at a square grid size of 15-30m2
providing a sufficient acceptable limit (Ronchi et al., 2017). Hence, various commonly used drag models in
the literature are tested to improve the spatial distribution of cubiform, cylindrical, and square disc non-burning
particles without making a major change in FDS to reduce its computation speed and finding an alternate drag
model effective on different isometric shapes. The drag models tested in the present work are: Haider &
Levenspiel (1989), Ganser (1993), Hölzer & Sommerfeld (2008), Bagheri & Bonadonna (2016).
2. DRAG MODELS
Table 1, gives brief details of the inbuilt FDS drag model for spherical (used for cubiform particles in
(Wadhwani et al., 2017) and in this study) (Equation 1) and cylindrical (Equation 2) particles, and the drag
models available in the literature for isometric particles. The detailed discussion about each alternative drag
model can be found in their respective references. There is no inbuilt drag model in FDS for disc-shaped
particles, and hence only the alternative models are tested for square disc particles for their efficacy.
401
Wadhwani et al., Improvement of drag model for non-burning firebrand transport in Fire Dynamics Simulator
Haider & Levenspiel (1989) developed an empirical model (Equation 3) for non-spherical shapes valid in the
Stokes’ and Newtonian flow region. The Stokes’ region corresponds to laminar flow conditions while the
Newtonian corresponds to turbulent flow conditions. The drag model was developed from the experimental
data published in the literature for different shapes and particle Reynolds’ number(). The particle shape
factor is accounted by sphericity () of the particles. Sphericity is defined as the ratio of the surface area of a
given particle to the surface area of a sphere of the same volume. Ganser extended the work of Haider &
Levenspiel in the Ganser drag model (1993) (Equation 4) which accommodates the shape factor in two different
flow regimes, one for the Stokes’ region and other for the Newtonian region. However, Hölzer & Sommerfeld
(2008) observed that the above two model shows a mean deviation in predicting drag coefficient for disc or
plate particles up to 2000% and for cubiform and cylindrical particles up to 40%, thereby providing two similar
drag models. These account for the effect of the tumbling of isometric and non-isometric particles in terms of
crosswise and lengthwise sphericity. Hölzer & Sommerfeld (2008) showed that a mean error using their model
in predicting the drag coefficient of disc or plate particles was ~17% and ~30% for cubiform and cylindrical
particles. They suggested equation 5 (in Table 1), with little loss of accuracy in predicting the drag coefficient,
because it is complicated to measure lengthwise sphericity when the particle is moving. In addition, Ganser
(1993) suggested that the crosswise sphericity is unity for isometric particles. Hence, in our cases, we used
this value in our Hölzer and Sommerfeld model (Equation 5).
Table 1. List of drag models tested in this work
Sr. No.
Drag model
Drag correlation
1
FDS Spherical drag
model (McGrattan
et al., 2015) C, =

 , Re< 1
..
.
, 1 < Re<1000
1 , Re>1000
(Eq. 1)
2
FDS Cylindrical
drag model
(McGrattan et al.,
2015)
C, =


. , Re< 1
..
.
, 1 < Re<1000
1 , Re>1000
(Eq. 2)
3
Haider &
Levenspiel drag
model (1989)
C, =

1 + ARe
+

 , Re< 2 × 10 (Eq. 3)
where
A = exp (2.3288 6.+ 2.4486), B = 0.0964 + 0.
C = exp (4.905 13.+18.422210.2599)
D = exp (1.4681 +12.20.7322
+15.8855
)
4
Ganser drag model
(1993)
,
=

(1 + 0.1118(ReKK).)+
.
 
 , ReKK 10 (Eq. 4)
where
K1 & K2 is Shape factor in Stoke’s and Newton regimes
For isometric particle, K=[0.3333 + 0.6667
.
]

and K=10
.().
5
Hölzer &
Sommerfeld drag
model (2008)
C, =
+

+
 . + 0.42 10.()
.
, Re10 (Eq. 5))
where
is called as a crosswise sphericity
6
Bagheri &
Bonadonna drag
model (2016)
C, =
1 + 0.125 Re
+ .
 

,Re<3×10 (Eq. 6)
where
k=


, k=10[ ()] , = 0.45 +
 (.()) ,
= 1 
 (()) , apparent density() = , 
,  , F= fe.
F= f e
, fatness(f)= S/I , elongation (e)= I/L
Bagheri and Bonadonna (2016), takes another step forward on top of the Hölzer and Sommerfeld drag model.
Bagheri and Bonadonna averaged the effect of crosswise and lengthwise sphericity to account for the drag
model to keep the drag model similar to the Ganser drag model. Hence, this reduces the requirement of
estimating crosswise and lengthwise sphericity with time. In the Bagheri and Bonadonna model, equation 6,
 is the equivalent diameter of the sphere which has the same volume of the particle. L, I, and S are the
longest, intermediate and shortest length of the particle. Blott and Pye (2008) have discussed how to estimate
L, I, and S which is used in Bagheri and Bonadonna drag model (Equation 6).
402
Wadhwani et al., Improvement of drag model for non-burning firebrand transport in Fire Dynamics Simulator
3. METHOD
3.1. Experimental
The drag models are appraised against laboratory-scale experiments conducted using a firebrand generator
prototype constructed in our facility and detailed in our previous work (Wadhwani et al., 2017). The prototype
produces a uniform flow profile with centreline flow velocity at the mouth of 29.5 m/s to eject non-burning
cubiform, cylindrical, and square disc particles of 10mm nominal size. The measurement grid for particle
collection is of 20 cm size (Fig. 1) which will reduce the tumbling effect of the particles on the distribution.
The particle distribution and their component velocities are measured using a 720p at 120fps camera. The
components of particle velocity (u, v, and w) are measured by displacement of the centroid of particle steak
using particle tracking velocimetry (PTV) (Maas et al., 1993).
The particles tested are (i) cubiform: avg. length of 12.45 mm, and mass of 0.83 g (0.12g std. dev.), (ii)
cylindrical: avg. length 11.6 mm, diameter 6.2 mm, and mass of 0.17 g (0.01g std. dev.), and (iii) square disc:
avg. length 10.18mm, thickness 2.22mm, and mass of 0.12g (0.01g std. dev.). These particles are injected into
the firebrand generator at a rate of 0.33 particles/sec. The spatial distributions of particles are measured using
particle imaging, and their first impact location is noted (Wadhwani et al., 2017). The final distribution of
particles is ignored due to a collision between particles and surface which will be complicated to account for
the simulation.
Figure 1. Experimental rig to study the spatial distribution of particles from the prototype (Wadhwani et al., 2017)
3.2. Numerical Modelling
Simulation of experimental scenario is carried out using FDS 6.2.0. The domain of the simulation is 7 m long,
1.2 m wide, and 2 m high respectively in X-, Y-, and Z- directions. The simulation domain is sub-divided into
four sections (Fig. 2), x=0-0.5, 0.5-1.5, 1.5-2.5, and 2.5-7 m with uniform grid sizes (Δx= Δy= Δz) 5, 10, 20,
and 40 mm respectively. To represent experimental work accurately, six simulation particles densities (
±
4
, ± 4
, and ± 4
; , is the mean density and the standard deviation of particle densities
respectively) are used to represent the particles used in the experiment. The mean and standard deviation of
cubiform, cylindrical, and square disc particle densities are 428.3, 492.9, 512.5, 48.9, 44.3, and 35.9 kg/m3
respectively.
Figure 2. FDS simulation domains to simulate particle distribution divided into four sections (Wadhwani et al., 2017)
4. RESULTS
The experimental distribution of the particles (cubiform, cylindrical, and square disc) at the mouth of pipe in
two orthogonal directions of the flow is shown in Fig. 3. The distribution is almost uniform at the mouth of the
prototype. Although, the distribution is slightly skewed in the Z- direction for cubiform particles due to the
weight of the individual particle. The distribution is approximately a normal distribution due to the very low
loading rate of the particles. The measured components of particle velocities for cubiform particles are 12.5,
0.0, 0.0 m/s respectively, with std. dev. of 0.8, 0.6 and 0.6 m/s respectively. Similarly, for cylindrical particles,
components of particle velocities are 13.4, 0.2, 0.2 m/s respectively, with std. dev. of 0.9, 0.7 and 0.8 m/s
respectively; 13.2, 0.0, 0.0 m/s, respectively, are components of particle velocities for square disc particles
with std. dev. 1.1, 0.9, 1.1 m/s respectively.
403
Wadhwani et al., Improvement of drag model for non-burning firebrand transport in Fire Dynamics Simulator
(b) cylindrical (Wadhwani et al., 2017)
Figure 3. Distribution of the particles at the mouth of the prototype generator
Fig. 4 shows the comparative contours of simulated and experimental spatial distribution for cubiform particles
with different drag models. From Fig. 4(a) we see that for cubiform particles the default FDS drag model for
spherical particles provides the best fit with the experimental observation which can be seen with a lower value
of the difference between experimental and simulated peak and overlap of lateral spread. The cubiform particle
has a minimal tumbling effect due to regular shape in all three directions which has less impact on the drag
coefficient. Moreover, the alternative model used for cubiform particles tends to further under-predict the
distribution as compared to the spherical drag model which may be due to over-estimation of drag coefficients
by respective drag model. Hölzer and Sommerfield (2008) estimated that the drag coefficient computed using
Haider and Levenspiel, Ganser, and their own models have a mean error of 42.3%, 38.4%, and 27.2%
respectively with experimental data. This could be one reason for such observation in our study. The lateral
spread of cubiform particles is found to consistent with all the drag model and found to overlap with the
experimental distribution.
(a) FDS default drag model (Wadhwani et al., 2017)
(b) Haider and Levenspiel drag model
(c) Ganser drag model
(d) Hölzer and Sommerfeld drag model
(e) Bagheri and Bonnadonna drag
model
Figure 4. Spatial distribution of cubiform particles with different drag models
For particles shaped like a cylindrical and square disc, the tumbling and secondary motion plays a significant
role during their transport. It can be seen for cylindrical particles in Fig. 5 and the square disc in Fig. 6 shows
more lateral spread as compared to cubiform particles in Fig. 4 which can be seen from their peak intensity.
For cylindrical particles, the difference between the peak location of simulation and experiment was reduced
when alternative drag models were tested (the minimum was achieved with the Haider and Levenspiel model)
as compared to FDS inbuilt model. However, there was no significant change in improving the lateral spread
which is anticipated as the particles are assumed to be point particles with no secondary motion at their axis.
The lateral spread is found to be ~5-15% thinner than its experimental lateral spread. Similarly, for square disc
particles, only alternative drag models are tested (Fig. 6) as FDS do not have any inbuilt feature to represent
such particles. The alternative drag model found to have a reasonable estimation of the peak location with
404
Wadhwani et al., Improvement of drag model for non-burning firebrand transport in Fire Dynamics Simulator
Bagheri and Bonadonna model shows the minimum difference in the peaks. The computational domain limits
restricts the domain between -0.6 to 0.6m (in Fig. 2), bounds the comparison of lateral spread with their
experimental observations, thus, also affecting the simulated peak intensity.
(a) FDS default drag model (Wadhwani et al., 2017)
(b) Haider & Levenspiel drag model
(c) Ganser drag model
(d) Hölzer and Sommerfeld drag model
(e) Bagheri and Bonnadonna drag
model
Figure 5. Spatial distribution of cylindrical particles with different drag models
(a) Haider and Levenspiel drag model
(b) Ganser drag model
(c) Hölzer and Sommerfeld drag model
(d) Bagheri and Bonnadonna drag model
Figure 6. Spatial distribution of square disc particles with different drag models
5. CONCLUSION
A combined experimental and numerical study is conducted to find an alternative particle drag model which
can be utilised to improve the prediction of physics-based fire models such as FDS with the point particle
assumption for non-burning firebrands. Using a prototype firebrand generator and various measurement
techniques a number of parameters are measured such as fluid velocity, particle velocity components at the
mouth and distribution of particles at the mouth. These data are used as input parameters for FDS. Particle
landing distributions are also measured and compared with a physics-based model. Three different non-burning
405
Wadhwani et al., Improvement of drag model for non-burning firebrand transport in Fire Dynamics Simulator
firebrands (particle) shapes were used: cubiform, cylindrical, and square disc shapes. Four different drag
models were implemented into the physics-based model, FDS: Haider and Levenspiel, Ganser, Hölzer and
Sommerfeld and Bagheri and Bonnadonna, besides FDS’s inbuilt drag model.
The result shows the application of all four alternative drag models generally improve the prediction of non-
burning firebrands’ distribution and keep the simulated error under the tolerance level. The improvement in
landing distribution is achieved without the loss of any computational speed. The alternative drag models,
especially Haider & Levenspiel, reduced the difference between the experimental and simulated experimental
cases except for cubiform particles. This exception is due to the minimal secondary motion effect of cubiform
particles and the inherent inaccuracy of drag models of over-estimating the drag coefficient. For particles like
cylindrical and square disc shapes, where secondary motion is an important part of the motion, a significant
improvement is observed in longitudinal direction. However, it is also observed that the existing point particle
assumption of FDS is not suited to estimate the lateral spread of firebrands especially when the secondary
motion of a particle on its axis is involved such as cylindrical and square-disc particles. In our studies, the
lateral spread is found to be in the range of ~5-15% thinner as compared to its experimental width for cylindrical
particle distribution. Tumbling of the particles is expected to increase lateral variation in the flight path of a
particle, and thus decrease the total distance travelled, resulting in a wider distribution such as observed in
experiments. For the square disc, it is not possible to quantify due to the computational limit associated with
our present study. In future, it will be necessary to explore the sensitivity of our observations to different sizes,
shapes, speeds of particles and burning processes to ascertain the efficacy of a particular drag model for
firebrand transport.
ACKNOWLEDGEMENT
We wish to acknowledge the financial support given by Bushfire and Natural Hazard Cooperative Research
Centre, Australia. The authors wish to thank Mr Lyndon Macindoe and Mr Philip Dunn for their support in
constructing the firebrand generator prototype at Victoria University. Also, the authors wish to thank the
administrative team of the Spartan HPC, the University of Melbourne for their support in running simulations.
REFERENCES
Bagheri, G., & Bonadonna, C. (2016). On the drag of freely falling non-spherical particles. Powder technology,
301, 526-544.
Blott, S. J., & Pye, K. (2008). Particle shape: a review and new methods of characterization and classification.
Sedimentology, 55(1), 31-63.
Cruz, M. G., et al. (2015). Guide to Rate of Fire Spread Models for Australian Vegetation: CSIRO Land and
Water Flagship, Canberra, ACT and AFAC, Melbourne, VIC.
Ellis, P. F. (2012). A review of empirical studies of fireband behaviour. Melbourne, Australia: Bushfire
Cooperative Research Centre.
Fernandez-Pello, A. C. (2017). Wildland fire spot ignition by sparks and firebrands. Fire safety journal, 91, 2-
10.
Ganser, G. H. (1993). A rational approach to drag prediction of spherical and nonspherical particles. Powder
technology, 77(2), 143-152.
Haider, A., & Levenspiel, O. (1989). Drag coefficient and terminal velocity of spherical and nonspherical
particles. Powder technology, 58(1), 63-70.
Hölzer, A., & Sommerfeld, M. (2008). New simple correlation formula for the drag coefficient of non-spherical
particles. Powder technology, 184(3), 361-365.
Maas, H., et al. (1993). Particle tracking velocimetry in three-dimensional flows. Experiments in fluids, 15(2),
133-146.
Mahmud, H., et al. (2016). Study of water‐mist behaviour in hot air induced by a room fire: Model
development, validation and verification. Fire and materials, 40(2), 190-205.
Manzello, S. L., et al. (2020). Role of firebrand combustion in large outdoor fire spread. Progress in Energy
and Combustion Science, 76, 100801.
McGrattan, K., et al. (2015). Fire Dynamics Simulator Technical Reference Guide Volume 1: Mathematical
Model Gaithersburg, Maryland, USA.
Ronchi, E., et al. (2017). e-Sanctuary: Open Multi-Physics Framework for Modelling Wildfire Urban
Evacuation (FPRF-2017-22). Quincy, MA, USA, .
Sullivan, A. L. (2009a). Wildland surface fire spread modelling, 1990–2007. 1: Physical and quasi-physical
models. International Journal of Wildland Fire, 18(4), 349-368. doi:10.1071/WF06143
Sullivan, A. L. (2009b). Wildland surface fire spread modelling, 19902007. 3: Simulation and mathematical
analogue models. International Journal of Wildland Fire, 18(4), 387-403. doi:10.1071/WF06144
Wadhwani, R., et al. (2017). Verification of a Lagrangian particle model for short-range firebrand transport.
Fire safety journal, 91, 776-783.
406
... I When comparing the lateral distribution of firebrands predicted by the Haider and Levenspiel model (Haider and Levenspiel 1989) and the Ganser model (Ganser 1993), it can be seen that the Haider and Levenspiel model shows a consistent overlaps with the experimental spread. This aspect (comparing lateral distribution) becomes profound when other shapes, such as cylindrical and square disc-shaped firebrand particles, are considered, as observed by Wadhwani (2019) and Wadhwani et al. (2021) for non-burning particles. A similar comparison can be drawn for the maximum and minimum spotting distances by comparing the whiskers in Fig. 9. ...
... Our novel firebrand generator can be used to conduct such validation studies of Lagrangian submodels with different shapes, rates of combustion and Reynolds numbers in future studies. Some preliminary comparisons of different shapes appear in Wadhwani et al. (2021). ...
Article
Firebrands (often called embers) increase the propagation rate of wildfires and often cause the ignition and destruction of houses. Predicting the motion of firebrands and the ignition of new fires is therefore of significant interest to fire authorities. Numerical models have the potential to accurately predict firebrand transport. The present study focuses on conducting a set of benchmark experiments using a novel firebrand generator, a device that produces controlled and repeatable sets of firebrands, and validating a numerical model for firebrand transport against this set of experiments. The validation is conducted for the transport of non-burning and burning cubiform firebrand particles at two flow speeds. Four generic drag sub-models used to estimate drag coefficients that are suited for a wide variety of firebrand shapes are verified for their applicability to firebrand transport modelling. The four sub-models are found to be good in various degrees at predicting the transport of firebrand particles.
Technical Report
Full-text available
The present work describes a novel framework for modelling wildfire urban evacuations. The framework is based on multi-physics simulations that can quantify the evacuation performance. The work argues that an integrated approached requires considering and integrating all three important components of WUI evacuation, namely: fire spread, pedestrian movement, and traffic movement. The report includes a systematic review of each model component, and the key features needed for the integration into a comprehensive toolkit.
Article
Full-text available
Firebrands are a harbinger of damage to infrastructure; their effects cause a particularly important threat to people living within the wildland-urban-interface. Short-range firebrands travel with the wind with little or no lofting, and cause spotfires. In this work, the design of a novel firebrand generator prototype is discussed to achieve a uniform shower of firebrands. The transport of short-range firebrand is studied to verify the existing Lagrangian particle model of Fire Dynamics Simulator. Uniform, non-combusting cubiform and cylindrical firebrands are projected using the firebrand generator. The experimentally observed distribution of particles on the ground is compared with a simulated distribution using the fire dynamic simulator. The results show that the existing Lagrangian model gives a good agreement with the experimental data.
Book
Full-text available
The knowledge of a free-burning fire’s potential rate of spread is critical to safe and effective bushfire control and use in Australia. A number of models for predicting the rate of fire spread in various Australian vegetation types have been developed over the past 60 years or so since Alan G. McArthur began his pioneering research into bushfire behaviour. Most of the major vegetation types in Australia have had more than one rate of fire spread model developed for operational use. A better understanding of these rate of fire spread models and their utility appears warranted in light of recent developments in both bushfire research and management in Australia. This publication presents, reviews and discusses these models and their applicability for operational use in prescribed burning and wildfire suppression in grasslands, shrublands, both dry and wet eucalypt forests, and in pine plantation fuel types. Background information and in turn a description of each rate of fire spread model is given, including the data used in the model development that constitute their application bounds. The mathematical equations that form each model are presented along with a discussion of model form and behaviour, the main input variables and their influence, and performance evaluation studies undertaken to date. Accompanying graphs, tables and photos are used throughout to illustrate key concepts. This publication identifies those models that constitute the current state of our knowledge with respect to bushfire behaviour science in Australia. Recommendations are accordingly made on which models should underpin best practices for operational and scientific predictions of rate of fire spread in the near term and those that should now be discounted and the reasons.
Article
Full-text available
We present a new general model for the prediction of the drag coefficient of non-spherical solid particles of regular and irregular shapes falling in gas or liquid valid for sub-critical particle Reynolds numbers (i.e. _Re_ < 3 × 105). Results are obtained from experimental measurements on 300 regular and irregular particles in the air and analytical solutions for ellipsoids. Depending on their size, irregular particles are accurately characterized with a 3D laser scanner or SEM micro-CT method. The experiments are carried out in settling columns with height of 0.45 to 3.60 m and in a 4 m-high vertical wind tunnel. In addition, 881 additional experimental data points are also considered that are compiled from the literature for particles of regular shapes falling in liquids. New correlation is based on the particle Reynolds number and two new shape descriptors defined as a function of particle flatness, elongation and diameter. New shape descriptors are easy-to-measure and can be more easily characterized than sphericity. The new correlation has an average error of ~ 10%, which is significantly lower than errors associated with existing correlations. Additional aspects of particle sedimentation are also investigated. First, it is found that particles falling in dense liquids, in particular at _Re_ > 1000, tend to fall with their maximum projection area perpendicular to their falling direction, whereas in gases their orientation is random. Second, effects of small-scale surface vesicularity and roughness on the drag coefficient of non-spherical particles found to be < 10%. Finally, the effect of particle orientation on the drag coefficient is discussed and additional correlations are presented to predict the end members of drag coefficient due to change in the particle orientation.
Article
Full-text available
Water-mists are emerging as an effective agent for the suppression of fires. However, the mechanisms of suppression are complex and the behaviour of individual water droplets in a smoke layer generated by fires must be quantified. This study investigates the behaviour of individual droplets injected from a nozzle into a hot air environment induced by a room fire. A semi-empirical model has been developed based on the conservation of mass, momentum and energy to evaluate the heat and mass transfer phenomena in an air-water droplet system. The model has considered the effect of change of momentum of an evaporating droplet. A forward finite difference approach is applied to solve the governing time dependent ordinary differential equations. The droplets are considered to be ‘lumped mass’ and variable thermo-physical properties of water and air and the change of Reynolds number of the droplets, due to the change of their diameter and velocity are considered. The effect of high evaporation rate on the mass and heat transfer coefficient and the contribution of radiation emanating by a flame and the surrounding boundary walls are also considered in the model which were not taken into account in the previous studies. Experimental data on terminal velocity and adiabatic saturation temperature are used to validate and verify the model. The validation and verification indicate that the proposed model predicted the terminal velocity within 4% of the experimental data and predicted the saturation temperature within 5% of the adiabatic saturation temperature. This semi-empirical model is also used as a tool to validate a more comprehensive computational fluid dynamics (CFD) based tool, Fire Dynamics Simulator (FDS). It is found that FDS results agree well with the results of the proposed model. Furthermore, the proposed model can be used to evaluate the temperature, velocity, diameter and other physical properties of a droplet travelling through a layer of hot air. Copyright
Article
Large outdoor fires are an increasing danger to the built environment. Wildfires that spread into communities, labeled as Wildland-Urban Interface (WUI) fires, are an example of large outdoor fires. Other examples of large outdoor fires are urban fires including those that may occur after earthquakes as well as in informal settlements. When vegetation and structures burn in large outdoor fires, pieces of burning material, known as firebrands, are generated, become lofted, and may be carried by the wind. This results in showers of wind-driven firebrands that may land ahead of the fire front, igniting vegetation and structures, and spreading the fire very fast. Post-fire disaster studies indicate that firebrand showers are a significant factor in the fire spread of multiple large outdoor fires. The present paper provides a comprehensive literature summary on the role of firebrand mechanisms on large outdoor fire spread. Experiments, models, and simulations related to firebrand generation, lofting, burning, transport, deposition, and ignition of materials are reviewed. Japan, a country that has been greatly influenced by ignition induced by firebrands that have resulted in severe large outdoor fires, is also highlighted here as most of this knowledge remains not available in the English language literature. The paper closes with a summary of the key research needs on this globally important problem.
Article
Wildland and Wildland Urban Interface (WUI) fires are an important problem in many areas of the world and may have major consequences in terms of safety, air quality, and damage to buildings, infrastructure, and the ecosystem. It is expected that with climate changes the wildland fire and WUI fire problem will only intensify. The spot fire ignition of a wildland fire by hot (solid, molten or burning) metal fragments/sparks and firebrands (flaming or glowing embers) is an important fire ignition pathway by which wildfires, WUI fires, and fires in industrial settings are started and may propagate. There are numerous cases reported of wildfires started by hot metal particles from clashing power-lines, or generated by machines, grinding and welding. Once the wildfire or structural fire has been ignited and grows, it can spread rapidly through ember spotting, where pieces of burning material (e.g. branches, bark, building materials, etc.) are lofted by the plume of the fire and then transported forward by the wind landing where they can start spot fires downwind. The spot fire problem can be separated in several individual processes: the generation of the particles (metal or firebrand) and their thermochemical state; their flight by plume lofting and wind drag and the particle thermo-chemical change during the flight; the onset of ignition (smoldering or flaming) of the fuel after the particle lands on the fuel; and finally, the sustained ignition and burning of the combustible material. Here an attempt has been made to summarize the state of the art of the wildfire spotting problem by describing the distinct individual processes involved in the problem and by discussing their know-how status. Emphasis is given to those areas that the author is more familiar with, due to his work on the subject. By characterizing these distinct individual processes, it is possible to attain the required information to develop predictive, physics-base wildfire spotting models. Such spotting models, together with topographical maps and wind models, could be added to existing flame spread models to improve the predictive capabilities of landscape-scale wildland fire spread models. These enhanced wildland fire spread models would provide land managers and government agencies with better tools to prescribe preventive measures and fuels treatments before a fire, and allocate suppression resources and issue evacuation orders during a fire.
Article
In recent years, advances in computational power and spatial data analysis (GIS, remote sensing, etc) have led to an increase in attempts to model the spread and behvaiour of wildland fires across the landscape. This series of review papers endeavours to critically and comprehensively review all types of surface fire spread models developed since 1990. This paper reviews models of a simulation or mathematical analogue nature. Most simulation models are implementations of existing empirical or quasi-empirical models and their primary function is to convert these generally one dimensional models to two dimensions and then propagate a fire perimeter across a modelled landscape. Mathematical analogue models are those that are based on some mathematical conceit (rather than a physical representation of fire spread) that coincidentally simulates the spread of fire. Other papers in the series review models of an physical or quasi-physical nature and empirical or quasi-empirical nature. Many models are extensions or refinements of models developed before 1990. Where this is the case, these models are also discussed but much less comprehensively. Comment: 20 pages + 9 pages references + 1 page figures. Submitted to the International Journal of Wildland Fire
Article
In recent years, advances in computational power and spatial data analysis (GIS, remote sensing, etc) have led to an increase in attempts to model the spread and behaviour of wildland fires across the landscape. This series of review papers endeavours to critically and comprehensively review all types of surface fire spread models developed since 1990. This paper reviews models of a physical or quasi-physical nature. These models are based on the fundamental chemistry and/or physics of combustion and fire spread. Other papers in the series review models of an empirical or quasi-empirical nature, and mathematical analogues and simulation models. Many models are extensions or refinements of models developed before 1990. Where this is the case, these models are also discussed but much less comprehensively. Comment: 31 pages + 8 pages references + 2 figures + 5 tables. Submitted to International Journal of Wildland Fire