Turbulence and fire-spotting effects into wild-land fire simulators

Turbulence and fire-spotting effects into wild-land fire simulators

Accepted Manuscript Turbulence and fire-spotting effects into wild-land fire simulators Inderpreet Kaur, Andrea Mentrelli, Fred ´ eric ´ Bosseur, Jea...

2MB Sizes 0 Downloads 69 Views

Accepted Manuscript

Turbulence and fire-spotting effects into wild-land fire simulators Inderpreet Kaur, Andrea Mentrelli, Fred ´ eric ´ Bosseur, Jean-Baptiste Filippi, Gianni Pagnini PII: DOI: Reference:

S1007-5704(16)30075-2 10.1016/j.cnsns.2016.03.003 CNSNS 3804

To appear in:

Communications in Nonlinear Science and Numerical Simulation

Received date: Revised date: Accepted date:

4 December 2015 29 February 2016 4 March 2016

Please cite this article as: Inderpreet Kaur, Andrea Mentrelli, Fred ´ eric ´ Bosseur, Jean-Baptiste Filippi, Gianni Pagnini, Turbulence and fire-spotting effects into wild-land fire simulators, Communications in Nonlinear Science and Numerical Simulation (2016), doi: 10.1016/j.cnsns.2016.03.003

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

Highlights • A mathematical approach to model the random processes like turbupresented.

CR IP T

lence and fire-spotting into a DEVS based wild-land fire simulator is

• The performance of a Lagrangian and an Eulerian based wildfire sim-

ulators is compared in context of their application to simulate the ran-

AN US

dom behavior associated with the wildfires.

• The formulation is successful in reproducing the realistic situations associated with wildfires eg. flank fires and back fires, faster spread due to the effect of hot air and firebrands, fire crossing fuel break zones

M

and development of isolated secondary fires due to fire-spotting. • The results show the potential for implementation of the formulation

AC

CE

PT

Fire.

ED

into operational wildland fire simulators like WRF-SFIRE and Fore-

1

ACCEPTED MANUSCRIPT

Turbulence and fire-spotting effects into wild-land fire simulators

CR IP T

Inderpreet KAURa,∗, Andrea MENTRELLIb,a , Fr´ed´eric BOSSEURc , Jean-Baptiste FILIPPIc , Gianni PAGNINIa,d

AN US

a BCAM–Basque Center for Applied Mathematics Alameda de Mazarredo 14, E-48009 Bilbao, Basque Country – Spain b Department of Mathematics & 2 (AM) –Alma Mater Research Center on Applied Mathematics University of Bologna, via Saragozza 8, I-40123 Bologna, Italy c Laboratoire SPE–Sciences Pour l’Environnement CNRS - UMR6134 University of Corsica, BP 52, F-20250 Corte, Corsica – France d Ikerbasque–Basque Foundation for Science Calle de Mar´ a D´ıaz de Haro 3, E-48013 Bilbao, Basque Country – Spain

M

Abstract

This paper presents a mathematical approach to model the effects and the

ED

role of phenomena with random nature such as turbulence and fire-spotting into the existing wildfire simulators. The formulation proposes that the propagation of the fire-front is the sum of a drifting component (obtained from an

PT

existing wildfire simulator without turbulence and fire-spotting) and a random fluctuating component. The modelling of the random effects is embod-

CE

ied in a probability density function accounting for the fluctuations around the fire perimeter which is given by the drifting component. In past, this

AC

formulation has been applied to include these random effects into a wildfire simulator based on an Eulerian moving interface method, namely the Level ∗

Corresponding author Email address: [email protected] (Inderpreet KAUR)

Preprint submitted to Commun. Nonlinear Sci. Numer. Simulat.

March 12, 2016

ACCEPTED MANUSCRIPT

Set Method (LSM), but in this paper the same formulation is adapted for a wildfire simulator based on a Lagrangian front tracking technique, namely

CR IP T

the Discrete Event System Specification (DEVS). The main highlight of the present study is the comparison of the performance of a Lagrangian and an Eulerian moving interface method when applied to wild-land fire propagation.

Simple idealised numerical experiments are used to investigate the potential

applicability of the proposed formulation to DEVS and to compare its be-

AN US

haviour with respect to the LSM. The results show that DEVS based wildfire

propagation model qualitatively improves its performance (e.g., reproducing flank and back fire, increase in fire spread due to pre-heating of the fuel by hot air and firebrands, fire propagation across no fuel zones, secondary fire generation, . . . ) when random effects are included according to the present

M

formulation. The performance of DEVS and LSM based wildfire models is comparable and the only differences which arise among the two are due to

ED

the differences in the geometrical construction of the direction of propagation. Though the results presented here are devoid of any validation exercise

PT

and provide only a proof of concept, they show a strong inclination towards an intended operational use. The existing LSM or DEVS based operational

CE

simulators like WRF-SFIRE and ForeFire respectively can serve as an ideal basis for the same. Keywords: Wildland fire propagation, fire simulators, Level Set Method,

AC

Discrete Event System Specification, ForeFire, random phenomena, turbulence, fire-spotting

3

ACCEPTED MANUSCRIPT

1

1. Introduction and motivations Modelling of wild-land fire propagation serves as a crucial tool to combat

3

the high economic and environmental damage associated with the wildfires.

4

An effective and swift modelling can aid in building up an efficient plan for

5

fire suppression and restrict the life and property damage to a minimum.

6

Modelling the propagation of wild-land fire is a complex, multi-scale, multi-

7

physics and multi-discipline process, and the crux of this simulation lies in

8

delivering either a versatile or a specialised model which is easy to imple-

9

ment and is capable of providing timely information. With the availability

10

of higher computational power, various new and improved modelling ap-

11

proaches have been developed over the last decade. An extensive review by

12

Sullivan [1, 2, 3] provides a comprehensive overview of the wide spectrum of

13

physical, empirical, statistical and mathematical analogue models available

14

in literature.

ED

M

AN US

CR IP T

2

Most of the statistical and mathematical models available in literature

16

comprise of a Rate of Spread (ROS) formulation and a moving interface

17

technique. The analytical formulation of the ROS is developed independent

18

of the moving interface scheme and can be characterised in terms of the

19

wind speed, slope, fuel characterisation, combustion properties, along with

20

other experimental data. Various formulations for the ROS are available in

21

literature [4, 5, 6, 7, 8], and they are usually versatile enough to be used

AC

CE

PT

15

22

with most moving interface schemes. Among the different moving interface

23

methods available, Eulerian Level Set Method (LSM) is extensively used in

24

wildfires to represent the propagating fire-line [9, 6, 10]. LSM is a scheme

25

which represents a moving interface on a simple Cartesian grid. This method 4

ACCEPTED MANUSCRIPT

is particularly appropriate for wildfire simulations, as it permits an accurate

27

computation of the front normal vector which is used to propagate the front.

28

Other approaches available in literature which focus on wildfire propaga-

29

tion modelling include the Lagrangian Discrete Event System Specification

30

(DEVS) [11, 12, 13, 14] and Huygens’ principle [15, 16, 17, 18, 19, 20]. The

31

Lagrangian DEVS approach simulates the propagation of the wildfire with-

32

out defining any underlying mesh to represent the burning state. DEVS

33

being an event driven simulation considers time as a continuous variable and

34

permits faster simulations over higher resolution. On the other hand, wildfire

35

models based on Huygens’ principle utilise an elliptical spread at each point

36

of the fire-front and in some cases also benefit from some analytical results

37

[15, 16, 17, 18, 19, 20].

AN US

CR IP T

26

Wild-land fire propagation involves the heat transport to the surroundings

39

in accordance with the atmospheric wind, and other factors like fuel distribu-

40

tion, elevation, and orography. Different statistical and mathematical models

41

utilized for wild-fire propagation in operational simulators include a part of

42

the physical processes based on the experimental data or its mathematical

43

analogues, but are usually limited to the average value without taking into

44

account the random fluctuations. These models accommodate the variabil-

45

ity introduced by the changes in fuel type, wind conditions or orography but

46

fail in quantifying a sudden erratic output of the fire due to random effects,

AC

CE

PT

ED

M

38

47

e.g., due to turbulence or fire-spotting. The generation of heat introduces a

48

turbulent flow in the vicinity of the fire which in turn allows a wider contact

49

of the hot air with the unburned fuel [21, 22, 23, 24, 25, 10]. This convective

50

and radiative transfer of heat causes a rapid ignition of the unburned fuel 5

ACCEPTED MANUSCRIPT

bed and leads to a faster spread of the fire. Turbulence can also facilitate

52

the generation of spot fires when the firebrands are transported away from

53

the main fire due to advection [26, 27, 28, 29, 30]. The effect of such ran-

54

dom phenomena can range from subdued to catastrophic depending upon the

55

concurrent weather conditions and fuel characteristics. Existing operational

56

wildfire simulators are limited by their formulation to decipher these random

57

phenomena, and require implementation of additional schemes/variables to

58

represent such processes.

AN US

CR IP T

51

In past, few approaches have been proposed to model different random

60

processes within the wildfire propagation [31, 32, 33, 30] in order to provide

61

a more realistic fire spread. In Ref. [33], the authors incorporate the stochas-

62

tic fire-spotting phenomenon based on a continuous-time Markov chain on a

63

lattice. The state of the lattice site is controlled according to the local tran-

64

sition rate functions, where the physical processes like - fire spread, spotting

65

and burnout compete against each other. In Ref. [30], the authors integrate

66

the mathematical models with a cellular automata scheme to demonstrate

67

the variability in the landing patterns of firebrands. In Ref. [34], the authors

68

discuss a DEVS based model which incorporates the uncertainty by sampling

69

certain variables from arbitrary probability distributions.

ED

PT

The present authors in a number of their recent works [35, 36, 37, 38, 39,

40] have proposed a new approach to include the effects of random processes

AC

71

CE

70

M

59

72

in some operational wildfire propagation models. They derive a formulation

73

to modify the modelled position of the fire-front to include random processes.

74

The motion of the propagating fire-front is randomised by the addition of a

75

random displacement distributed according to a probability density function 6

ACCEPTED MANUSCRIPT

(PDF) corresponding to the turbulent transport of heat and to the fire-

77

spotting landing distance. The driving equation of the resulting averaged

78

process is analogous to an evolution equation of the reaction-diffusion type,

79

where the ROS controls the source term. If the motion is considered without

80

such random processes, the diffusive part disappears, and the spread of the

81

fire-line is identical to the one given by the chosen operational simulator

82

and driven only by the ROS. The proposed formulation to include random

83

processes holds true for any method of determination of the ROS and the

84

statistical spread is determined by the PDF of displacements of the random

85

contour points marked as active burning points. The ensemble averaging

86

of the displacements smoothens out of the noisy fluctuations and results

87

in a smooth fire-line contour. This formulation can be implemented into

88

existing wildfire models as a post-processing scheme at each time step without

89

calling for any major changes in the original framework. The efficacy of this

90

formulation by using a LSM based fire propagation model has already been

91

shown [35, 36, 38, 39, 40].

ED

M

AN US

CR IP T

76

In this study, the authors proceed with their preliminary attempt [41]

93

to include the proposed formalism for random processes (turbulence and

94

fire-spotting), already discussed and implemented into an Eulerian LSM

95

based wild-land fire simulator [35, 36, 37, 38, 39, 40], into a Lagrangian

96

DEVS based wild-land fire simulation model: ForeFire (https://github.

AC

CE

PT

92

97

com/forefireAPI/firefront) [13]. ForeFire is an open-source wildfire sim-

98

ulator developed by University of Corsica, France [42] to serve as a basis

99

for an operational simulator and provides flexibility to include new physics

100

options. ForeFire is based on a Lagrangian technique and integrates the so7

ACCEPTED MANUSCRIPT

called Balbi model for the ROS [5] with a DEVS [43] based front tracking

102

method to simulate the wildfire propagation. Coupling the front tracking

103

method with DEVS permits the utilisation of time as a continuous vari-

104

able; and hence removes the limitation of establishing a trade off between

105

the temporal resolution and the scale of the simulation according to the

106

Courant–Friedrichs–Lewy (CFL) condition. ForeFire model is developed de-

107

void of any description of random processes such as turbulent heat transfer

108

and fire-spotting and therefore hinders the evolution of the fire across fuel

109

break zones.

AN US

CR IP T

101

The aim of this article is the implementation of the mathematical for-

111

mulation for turbulence and fire-spotting in a DEVS based wildfire model

112

(ForeFire) to improve its efficiency. Operational utilisation of ForeFire for

113

active wildfire propagation is the main driving factor for improving its output

114

specially while dealing with realistic situations. This article tries to provide

115

a proof of concept through a series of idealised numerical simulations to

116

show the potential efficacy of the method in adapting the fire-line under the

117

effect of turbulence and fire-spotting. Besides this, the investigation also

118

provides a unique scope to compare the performance of Eulerian and La-

119

grangian approach based moving interface methods. Although a number of

120

studies on fluid flows focus on the comparison of performance of Eulerian and

121

Lagrangian methods for particle transport [44, 45, 46], none of the studies

AC

CE

PT

ED

M

110

122

concentrate on their comparison in context of wildfire propagation. In the

123

present article, various test cases are described to compare the performance

124

of these two schemes in application to wildfires.

125

The remaining paper is organised as follows. The salient features of DEVS 8

ACCEPTED MANUSCRIPT

and LSM based wildfire simulators are described in section 2. The descrip-

127

tion of the formulation for random effects is provided in section 3 and the

128

numerical set-up to include this formulation into the DEVS based simulator

129

is described in section 4. Section 5 describes the different simulation set-ups

130

for evaluation. A discussion of the numerical behaviour of the two methods

131

and a detailed comparison of their performance is provided in section 6. A

132

brief sensitivity analysis of the spatial distribution of the fire-brand landing

133

patterns is also presented in section 7. Concluding remarks are provided in

134

section 8.

135

2. Overview of wildfire simulators

AN US

CR IP T

126

Modelling of wildfire spread focusses on the development of a phenomeno-

137

logical and theoretical formula for a suitable representation of the ROS of

138

fire and the simulation of the propagation of the fire perimeter to recon-

139

struct the shape of fire over time. Various ROS models available in literature

140

[4, 5, 6, 7, 8] provide an analytical formulation of the ROS for given wind con-

141

ditions, slope and fuel characteristics. The available ROS models are usually

142

developed independent of the moving interface methods and user discretion

143

is advised in selecting one according to the requirements. Since this study

144

aims at studying the effect of the random processes using wildfire simulators

145

based over different moving interface methods, an analysis of different ROS

AC

CE

PT

ED

M

136

146

models is not presented in this article.

147

Wildfire simulators based on the DEVS and the LSM techniques are in-

148

herently different, but the concept of Lagrangian markers and the constant

149

perimeter resolution in the DEVS front tracking method can be considered 9

ACCEPTED MANUSCRIPT

analogous to the concept of active fuel points and constant action arc length

151

in LSM tracking method as discussed in Ref. [39]. This section provides a

152

brief overview of the salient features of the DEVS and LSM based wildfire

153

simulators.

154

2.1. DEVS based wildfire simulator

CR IP T

150

The temporal scheme used to simulate the fire perimeter in wild-land

156

fire simulator ForeFire is based on the DEVS [47]. DEVS handles the time

157

advancement in terms of the increment of physical quantities instead of a

158

discrete time step. The resulting front is a polygon whose marker points

159

have real (i.e.. non-discrete) coordinates instead of being located on nodes

160

of a regular mesh or grid. This computationally in-expensive Lagrangian

161

technique simulates the evolution of an interface without any underlying grid

162

representing the state of the system. Each fire-front is discretised into a series

163

of markers connected to the next marker by a piecewise linear segment. The

164

fire-break zones like roads and rivers are also represented as arbitrary shaped

165

polygons. Each marker is associated with a propagation speed and direction,

166

and each time a marker moves, the intersection with the neighbourhood is

167

checked to take care of collision and topological changes.

M

ED

PT

The propagation speed of the marker is defined by the ROS, while the

CE

168

AN US

155

propagation direction is defined by a front normal function. While spline

170

interpolation may be used to estimate this normal, the bisector angle made

AC

169

171

by the marker with its immediate left and right neighbours is used as a

172

computationally effective approximation of this outward normal.

173

In DEVS, time is treated as a continuous parameter and each marker

174

evolves according to its own independent time step. The time advancement 10

ACCEPTED MANUSCRIPT

of the markers is event based and all markers do not share the same time

176

step. Due to different time advances for each marker, the CFL condition

177

applies only locally and the markers move asynchronously. All events, trig-

178

gering marker movement, are time sorted in an event list and processed by

179

a scheduler. This self adjusting temporal resolution in Lagrangian formu-

180

lation is computationally efficient and provides an efficient way to simulate

181

the spatially in-homogeneous problem. The simulation in DEVS advances as

182

new events are generated; the generation of new events is managed by the

183

following two criteria:

184

Collision criterion: A collision happens when a marker moves into a differ-

185

ent area, e.g., from an unburned to a burned area or from an active fire

186

to a fuel break. Each collision generates a modification of the shape.

AN US

M

187

CR IP T

175

Quantum distance criterion: Quantum distance ∆q is defined as the maximum distance each marker is allowed to cover during advancement.

189

The actual resolution of the simulation is limited by this parameter and

190

details smaller than the quantum distance ∆q may not be accounted

191

for.

PT

Three type of events can be defined to control the front propagation:

CE

192

ED

188

decomposition, regeneration and coalescence. The decomposition function is

194

activated when a marker enters a different area (e.g., while approaching a fuel

AC

193

195

break zone). As soon as the marker enters a new area, two new markers are

196

created on the boundary of the new area. In regeneration, the markers are

197

redistributed to refine the shape; if two markers are separated by a distance

198

greater than the perimeter resolution ∆c, a new marker is generated between 11

ACCEPTED MANUSCRIPT

the two markers. The coalescence function reconstructs the fire-perimeter by

200

merging the markers. All markers with separation less than ∆r are merged

201

together. For stability and to avoid cross over of two markers, ∆r = ∆c/2

202

is assumed and the condition ∆c ≥ 2∆q should be respected. The precision

203

of the method is highly dependent on the choice of ∆q and ∆c. Quantum

204

distance ∆q, should be of a much higher resolution than the wind data for

205

minimal error. A detailed description of the DEVS front tracking method

206

can be found in Ref. [13].

AN US

CR IP T

199

ForeFire library is developed with C/C++/Java and Fortran bindings

208

for a UNIX compatible environment and compiled using a SWIG (http:

209

//swig.org/) build platform. NetCDF library with legacy C++ interface

210

is required to build ForeFire, and SCons (http://www.scons.org/) python

211

tool is used to make the library. A python based interface is used to launch

212

the simulations with software calls to the main code. A command line mode

213

interpreter is also available to run simulations.

214

2.2. LSM based wildfire simulator

ED

M

207

LSM is one of the widely used and successful tools for tracking fronts

216

in any dimension [48, 9]. It is particularly suitable for problems where the

217

speed of the evolving interface is dependent on the interface properties and

218

the boundary conditions at the interface. Wild-land fire propagation is one

219

of problems where LSM finds a frequent application to represent the evolving

AC

CE

PT

215

220

fire-fronts [6, 49, 10]. In particular, it is the basis for the operational software

221

WRF-SFIRE (http://github.com/jbeezley/wrf-fire/).

222

Let Γ be a simple closed curve representing the fire-front interface in

223

two-dimensions, and let Ω be the region bounded by the fire-front Γ. If 12

ACCEPTED MANUSCRIPT

the interface is made up of more than one closed surface, the domain Ω is

225

not simply connected and represents more than one independently evolving

226

bounded areas. Let γ : S ×[0, +∞[→ R be an implicit function defined on the

CR IP T

224

227

domain of interest S ⊆ R2 such that the level set γ(x, t) = γ∗ coincides with

228

the evolving front, i.e., Γ(t) = {x ∈ S | γ(x, t) = γ∗ }. If Γ is an ensemble

229

of n surfaces, the ensemble of the n interfaces is considered as an interface.

230

The evolution of the isocontour γ = γ∗ in time is governed by

232

233

(1)

If the motion of the interface is assumed to be directed towards the front b normal n then, Eq. (1) reduces to

b=− n

∇γ , k ∇γ k

M

231

AN US

Dγ ∂γ dx Dγ∗ = + · ∇γ = = 0. Dt ∂t dt Dt

(3)

ED

∂γ = V(x, t) k ∇γ k . ∂t

(2)

In application to wild-land fire modelling, an indicator function φ(x, t)

235

is introduced to allow a convenient identification of the burned area and the

236

fire perimeter

238

CE

φ(x, t) =

  1, x ∈ Ω(t) ,

 0, x 6∈ Ω(t) .

(4)

The boundary of Ω can be identified as the front-line contour of the wildfire

AC

237

PT

234

and V denotes the ROS of the fire-front.

239

The LSM code makes use of a general-purpose library which aims at pro-

240

viding a robust and efficient tool for studying the evolution of co-dimensional

241

fronts propagating in one-, two- and three-dimensional system. The library, 13

ACCEPTED MANUSCRIPT

written in Fortran2008/OpenMP, along with standard algorithms useful for

243

the calculation of the front evolution by means of the classical LSM, in-

244

cludes Fast Marching Method algorithms. All the results presented in this

245

paper are obtained by making use of the above-mentioned software; Matlab

246

(www.mathworks.com/products/matlab/) and open source softwares such

247

as SciPy [50] and Matplotlib [51] in the IPython framework [52] have been

248

used for visualisation purposes.

249

3. Random effects formulation

AN US

CR IP T

242

The proposed approach is based on the idea to split the motion of the

251

front position into a drifting part and a fluctuating part [35, 36, 37, 38,

252

39, 40, 53, 54]. This splitting allows specific numerical and physical choices

253

which can be useful to improve the algorithms and the models. In particular,

254

the drifting part can be related to existing methods for moving interfaces,

255

for example the Eulerian LSM [9] or the Lagrangian DEVS [11, 12]. The

256

fluctuating part is the result of a comprehensive statistical description of the

257

system which includes the random effects in agreement with the physical

258

properties of the system. As a consequence, the fluctuating part can have a

259

non-zero mean, implying that the drifting part does not correspond to the

260

average motion. Due to this fact, the present splitting is distinct from the

261

well-known Reynolds decomposition frequently adopted in turbulent flows.

AC

CE

PT

ED

M

250

262

The motion of the each burning point can be random due to the effect

263

of turbulence and/or fire-spotting. Let X ω (t, x0 ) be the ω realisation of

264

the random trajectory of an active burning point at time t with the initial

265

condition X ω (0, x0 ) = x0 . The trajectory of a single interface particle can be 14

ACCEPTED MANUSCRIPT

described by the one-particle PDF f ω (x; t) = δ(x − X ω (t, x0 )), where δ(x)

267

is the Dirac delta function. The random front contour can be derived by

268

using the sifting property of the Dirac delta function peaking at a random

269

position. This random position represents the stochastic trajectory which

270

describes the random motion of the front line and includes both drifting and

271

fluctuating part of the front position. In the ω realisation, the evolution in

272

time of the function γ ω (x, t), which embeds the random fire-line Γω is given

273

by [35, 39, 53, 54] γ (x, t) =

Z

S

274

γ(x, t)δ(x − X ω (t, x) dx .

(5)

The effective indicator of the burned area enclosed by the random front φe (x, t) : S × [0, +∞[→ [0, 1] may be defined as Z  ω δ(x − X (t, x)) dx φe (x, t) = Ω(t) Z = hδ(x − X ω (t, x))i dx Ω(t) Z = f (x; t|x) dx Ω(t) Z = φ(x, t)f (x; t|x) dx ,

PT

ED

M

275

AN US

ω

CR IP T

266

(6) (7)

S

where f (x; t|x) = hδ(x − X ω (t, x))i is the PDF of the displacement of the

277

active burning points around the position x which is obtained from existing

278

wildfire simulators.

AC

CE

276

279

280

281

An arbitrary threshold value φth e is chosen to serve as the criterion to mark  the region as burned or unburned, i.e., Ωe (x, t) = x ∈ S | φe (x, t) > φth e . The evolution of the effective indicator φe (x, t) can be obtained by using

15

ACCEPTED MANUSCRIPT

283

284

the Reynold transport theorem to Eq. (7) [55, 35] Z Z ∂f ∂φe = dx + ∇x · [V (x, t)f (x; t|x)] dx . ∂t Ω(t) ∂t Ω(t)

(8)

Since f (x; t|x) is a PDF, its evolution in time is here assumed to be governed by an equation of the type ∂f = εf , ∂t

CR IP T

282

(9)

where ε is an operator acting on space variable x, and not on x and t.

286

For example, the operator ε is the Laplacian when f (x; t|x) is the Gaussian

287

AN US

285

density function. Using (9), equation (8) can be written as Z ∂φe = εφe + ∇x · [V (x, t)f (x; t|x)] dx . ∂t Ω(t)

(10)

The velocity of the fire-line propagation is controlled by the ROS V and

289

defined through V = V(x, t)b n, while turbulence and fire-spotting phenomena

290

are modelled by modifying the PDF function. This method is compatible

291

with the estimation of the ROS by any technique.

ED

M

288

The modelling of the effects of the random processes is handled by the

293

PDF f (x; t|x), accounting for the two independent random variables repre-

294

senting turbulence and fire-spotting respectively. It should be noted that for

295

brevity, here fire-spotting is assumed to be an independent downwind phe-

296

nomenon; hence the effect of fire-spotting is accounted only for the leeward

297

part of the fire-line. Taking into account these assumptions the PDF for the

CE

random processes can be defined as Z ∞   b U ; t)q(`; t) d` , n b ·n bU ≥ 0 , G(x − x − ` n     0 f (x; t|x) =      G(x − x; t), otherwise ,

AC

298

PT

292

16

(11)

ACCEPTED MANUSCRIPT

300

301

b U is the unit vector aligned with the mean wind direction. where n

Turbulent diffusion is assumed to be isotropic and modelled by a bi-

variate Gaussian PDF 1 G(x − x; t) = exp 4πDt





CR IP T

299

(x − x)2 + (y − y)2 4Dt

,

(12)

where D is the turbulent diffusion coefficient such that h(x − x)2 i = h(y − y)2 i =

303

2Dt. In the present model, the whole effect of the turbulent processes over

304

different scales is assumed to be parametrised only by the turbulent diffu-

305

sion coefficient. Also, only the turbulent fluctuations with respect to the

306

mean transport are considered. Hence, the characterisation of turbulence

307

is independent of the mean characteristics of the motion, and in particular

308

the estimation of the turbulent diffusion coefficient D is independent of the

309

mean wind. Moreover, since the simulations are performed with a flat terrain

310

and without boundaries in the spatial domain, horizontal isotropy is also as-

311

sumed. Within this framework, the mean wind does not break the isotropy,

312

because only the fluctuations are considered, but a non-trivial orography can

313

indeed break the isotropy, e.g., valleys or slopes. Even if an exact estima-

314

tion of D is out of the scope of the present study, D is one of the necessary

315

parameters for simulations and a quantitative estimation is required. The

316

desired diffusion coefficient D corresponds to the turbulent heat convection

317

generated by the fire. In literature, the ratio between the total heat trans-

318

fer and the molecular conduction of heat is known as the Nusselt number:

319

Nu = (D + χ)/χ. Where, χ is the thermal diffusivity of the air at ambi-

320

ent temperature and is well-known in literature: χ = 2 × 10−5 m2 s−1 . It is

321

also experimentally known that the relation between the Nusselt number Nu

322

and the Rayleigh number Ra is given by Nu ' 0.1 Ra1/3 [56]. The Rayleigh

AC

CE

PT

ED

M

AN US

302

17

ACCEPTED MANUSCRIPT

number is the ratio between the convection and the conduction of heat and

324

it is defined as Ra = γ ∆T g h3 /(νχ), where γ is the thermal expansion co-

325

efficient, ∆T the temperature difference between the bottom and the top of

326

the convective cell, h the dimension of the convection cell, g the acceleration

327

gravity and ν the kinematic viscosity. From above, D can be computed by

328

the formula:

CR IP T

323

D ' 0.1 χ [γ ∆T g h3 /(νχ)]1/3 − χ,

(13)

where values γ = 3.4 × 10−3 K−1 , g = 9.8 ms−2 and ν = 1.5 × 10−5 m2 s−1

330

are taken from literature. Since the heat transfer is considered in the hori-

331

zontal plane, perpendicular to the vertical “heating wall” embodied by the

332

fire, the length scale of the convective cell is assumed to be h = 100 m with

333

a temperature difference ∆T = 100 K. Considering these values, the scale of

334

the turbulent diffusion coefficient turns out to be approximately 104 times

335

the thermal diffusivity of air at ambient temperature (χ).

M

AN US

329

The jump length of the firebrands varies according to the meteorological

337

conditions, the intensity of the fire and the fuel characteristics [57, 58, 59].

338

A precise formulation of the landing distributions is difficult due to the lim-

339

ited small scale experimental results [60] and the complexities involved in

340

characterising the detailed aspect of firebrand generation and landing. The

341

firebrands land only in the positive direction and their landing distribu-

342

tion increases with distance to reach a maximum before dropping to zero

AC

CE

PT

ED

336

343

[61, 62]. In past, researchers have utilized different statistical distributions

344

to fit/reproduce the distribution pattern of particles deposited on ground

345

from a given source, e.g., the lognormal distribution is used to depict the dis-

346

tribution of the short-range firebrands (up to 4500 m) [27] while other studies 18

ACCEPTED MANUSCRIPT

347

also indicate the use of a Rayleigh distribution [63] or a Weibull distribution

348

[28] to describe the landing distributions.

350

In this study, a lognormal distribution is used to describe the downwind

CR IP T

349

distribution of the firebrands:

  1 (ln `/`0 − µ(t))2 q(`; t) = √ exp − , 2 σ(t)2 2π σ(t) `

(14)

where µ(t) = hln `/`0 i and σ(t) = h(ln `/`0 − µ(t))2 i are the mean and the

352

standard deviation of ln `/`0 respectively, and `0 is a unit reference length.

AN US

351

Since the fuel ignition due to hot air and firebrands is not an instanta-

354

neous process, a suitable criterion related to an ignition delay is introduced

355

for application to wildfire modelling. This ignition delay was previously con-

356

sidered as a heating-before-burning mechanism due to the hot air [35, 36]

357

and generalised to include fire-spotting [39]. In particular, since the fuel can

358

burn because of two pathways, i.e., hot-air heating and firebrand landing,

359

the resistance analogy suggests that the resulting ignition delay can be ap-

360

proximately computed as resistances acting in parallel. Hence, representing

361

τh and τf as the ignition delay due to hot air and firebrands respectively, the

362

joint ignition delay τ is

ED

PT

CE

1 1 1 τh + τf = + = . τ τh τf τh τf

(15)

In this study, the delay effect is portrayed as a heating-before-burning

AC

363

M

353

364

mechanism and described by an accumulative process of the effective fire-

365

front over time, i.e., ψ(x, t) =

Z

t

φe (x, η)

0

19

dη , τ

(16)

ACCEPTED MANUSCRIPT

where, ψ(x, 0) = 0 corresponds to the initial unburned fuel. This accumula-

367

tion can be understood as an accumulation of heat that results in an increase

368

of the fuel temperature T (x, t). Finally, the function ψ(x, t) may be related

369

to the fuel temperature T (x, t) as follows: ψ(x, t) ∝

T (x, t) − Ta (x) , Tign − Ta (x)

CR IP T

366

(17)

where Tign is the ignition temperature, T ((x), 0) = Ta (x) and T (x, t) ≤ Tign .

371

Let ∆t be the ignition delay due to the effects of the random processes,

372

then after the elapse of the ignition delay time T (x, ∆t) = Tign , and with

373

the assumptions Ta  Tign and the constant of proportionality equal to 1,

374

Eq. (17) reduces to

AN US

370

ψ(x, ∆t) = 1 .

(18)

Hence, as a cumulative effect of hot air and firebrands on the unburned fuel,

376

an ignition point x at time t is created when ψ(x, t) = 1, and the condition

377

φ(x, t) = 1 is also imposed.

378

4. Turbulence and fire-spotting effects into DEVS based wildfire

PT

ED

M

375

simulator

380

ForeFire is a DEVS based wild-land fire model developed bereft of any

CE

379

additional terms or parameters to model turbulence or fire-spotting. As

382

described earlier, DEVS is a Lagrangian method where each marker (active

383

burning point in the fire-line) advances by the quantum distance and the

384

advancement of each marker is managed by decomposition, regeneration or

385

coalescence functions. When one such marker arrives to a new region, the

386

marker decomposes to generate new markers at the interface between the two

AC

381

20

ACCEPTED MANUSCRIPT

regions. If this new region belongs to a fuel break zone, the new markers are

388

assigned null propagation velocity and their further evolution is inhibited.

389

But additional information about turbulence and fire spotting can allow the

390

fire-line to cross the fuel breaks. In this article, the DEVS based ForeFire

391

model is extended to include the effects of turbulence and fire-spotting by

392

post-processing the output obtained from the model at each time step. The

393

present method does not call for any major changes in the original framework

394

of DEVS, and can be versatilely used to reconstruct its output to include

395

these two processes. The detailed steps of the numerical procedure to re-

396

construct the output of DEVS based ForeFire are described as follows:

397

Step 1. Beginning with an initial fire-line, the DEVS front tracking method

398

is used to estimate the propagation of the markers to build up a new

399

fire perimeter for the next time step. This output is modified to

400

include the effects of turbulence and fire spotting by post-processing

401

numerical enrichment. This post-processing step is independent of

402

the definition of ROS.

ED

M

AN US

CR IP T

387

Step 2. The fire perimeter obtained from the active burning points is used to

404

construct the indicator function φ(x, t) (4). The spatial information

405

contained in φ(x, t) is sufficient to modify the fire-line with respect

to turbulence and fire-spotting and serves as the input to the post-

AC

407

CE

406

PT

403

408

processing step. Indicator function φ(x, t) has a value 0 over the region of the available fuel and 1 over the burned area.

409

Step 3. The spatial coordinates of the fire perimeter obtained from DEVS are

410

discrete values spread all over the domain and cannot be represented

411

by a specific mesh/grid. For brevity, the effective indicator function 21

ACCEPTED MANUSCRIPT

φe (x, t) (7) is generated over a Cartesian grid to facilitate the com-

413

putation of the function ψ(x, t) (16) over the same grid. Point in

414

polygon is used to generate the indicator function φ(x, t) from the lo-

415

cus of points representing the fire perimeter in DEVS. It is remarked

416

that φ(x, t), and φe (x, t) can also be computed from the real coor-

417

dinates representing the fire perimeter in DEVS, but the inherent

418

construction of the function ψ(x, t) mandates its computation over

419

a regular grid.

AN US

CR IP T

412

Step 4. The value of the effective indicator φe (x, t) is computed through the

421

numerical integration of the product of the indicator function φ(x, t)

422

and the PDF of fluctuations according to Eq. (7). The effect of

423

turbulence or fire-spotting is included by choosing the corresponding

424

PDF (11).

425

M

420

Step 5. The function ψ(x, t) is updated for each grid point by integration in time with the current value of φe (x, t) (16).

ED

426

Step 6. All points which satisfy the condition ψ(x, t) ≥ 1 are labelled as new

428

ignition points. The post-processing enrichment of the input fire-

429

front is completed at this step and the new markers describing the

430

enriched fire-front are used to define a new front valid at the next time step in DEVS framework. With the inclusion of heating-beforeburning mechanism, the new fire-front perimeter represents a larger

AC

432

CE

431

PT

427

433

area than the input fire perimeter.

434

Step 7. At the next time step, the new markers progress according to the

435

DEVS front tracking method, and the updated perimeter is again

436

subjected to the post-processing to enrich the fire front with the 22

ACCEPTED MANUSCRIPT

random fluctuations pertaining to turbulence and fire-spotting. The

438

sequence is repeated till the final “event time” step or till the fire

439

reaches the end of the domain.

CR IP T

437

Step 8. The DEVS models time as a continuous parameter and updates the

441

system in accordance with the events. The system is updated at user

442

defined “event times”. In between the consecutive events, no change

443

in the system occurs and a list of pending events is maintained in

444

chronological order. The event list is updated at the next “event

445

time”. Since, DEVS provides freedom to select an “event time”, it is

446

chosen to coincide with the maximum allowable time step according

447

to the CFL criterion. It is remarked that this choice of “event time”

448

limits the efficiency of DEVS computation (increase in the compu-

449

tation time due to excessive computations), but is indispensable in

450

order to introduce the heating-before-burning mechanism (16) in the

451

current formulation.

ED

M

AN US

440

It is remarked that the direction of the propagation vector for DEVS

453

is decided by the bisector of the angle formed between a marker and its

454

immediate neighbouring markers on left and right. This direction vector is a

455

weak approximation of the normal and can be very different from the actual

456

normal direction in the case of a non-circular profile.

AC

CE

PT

452

457

5. Simulation set-up

458

To estimate the performance of the DEVS front tracking method with the

459

inclusion of turbulence and fire-spotting, a series of numerical experiments

460

are performed. Similar set of experiments are also carried out with the LSM 23

ACCEPTED MANUSCRIPT

based simulator to enable a fair comparison between the two techniques. A

462

formulation developed in Ref. [6] and Ref. [10] is followed for LSM, while

463

for the DEVS method, ForeFire fire simulator [13] is used. To allow a direct

464

comparison between the two, both models are parametrised in an identical

465

set-up.

CR IP T

461

In the present study, for brevity, no particular type of vegetation is de-

467

fined and simulations are carried out with a pre-defined constant value of

468

ROS. It is assumed that the ROS remains constant for a particular domain.

469

The parametrisation of the numerical simulations is oversimplified and the

470

test cases are chosen with the purpose of highlighting the potential appli-

471

cability of the mathematical formulation. It is remarked that the chosen

472

parametrisations for the numerical simulations do not reproduce any of the

473

forest-fire or prescribed burn experiments but to emphasise the applicability

474

of the formulation, the values of various parameters are chosen to approxi-

475

mately lie in the valid range. The present scope of this work is to provide a

476

first look into the investigation of the performance of LSM and DEVS based

477

fire simulators in the context of inclusion of the effects due to turbulence and

478

fire-spotting.

PT

ED

M

AN US

466

A flat area of hypothetical homogeneous vegetation spread over a domain

480

size of 5000 m × 5000 m is selected for simulations. Different values of the

ROS are utilised for different test cases. The ROS is assumed to be 0.05 ms−1

AC

481

CE

479

482

483

484

b in no wind conditions, while in the presence of along the normal direction n wind, it is estimated by the 3% model [14], i.e.,

b, ROS = 0.03 U · n

(19)

where U is the mean wind velocity. Since, 3% model limits the propagation 24

ACCEPTED MANUSCRIPT

only towards the mean wind direction, a ROS formulation provided by Mallet

486

et al. [6] is followed to prescribe non-zero spread velocities for the flank and

487

rear fires: ROS(U, θ) =

 √  εo + a U cosn θ,

CR IP T

485

if |θ| ≤

 εo (α + (1 − α)| sin θ|), if |θ| >

π 2

,

π 2

,

(20)

where U 2 = U · U , εo is the flank velocity, (εo α) is the rear velocity with

489

α ∈ [0, 1], and θ is defined as the angle between the normal to the front

490

and the wind direction. In the present set-up, the parameters are defined as

491

α = 0.8, n = 3, a = 0.5 m1/2 s−1/2 , εo = 0.2 ms−1 , and U = 3 ms−1 [6].

AN US

488

In the case of the LSM based simulator, the domain is discretised with

493

a Cartesian grid of 20 m both in x and y directions, while for DEVS the

494

resolution of the simulation is established in the terms of quantum distance

495

∆q and perimeter resolution ∆c [13]. Quantum distance ∆q is defined as

496

the maximum allowable distance to be covered by a particle at each advance,

497

while a measure of ∆c is used to decompose/regenerate/coalesce two particles

498

on propagation. The choice of ∆q and ∆c is dependent on the type of problem,

499

and in the present study, two sets of values are utilized. The simulations

500

are performed with ∆q = 4 m, ∆c = 18 m for zero wind conditions, and

501

∆q = 2 m, ∆c = 8 m in the presence of wind. In DEVS based simulator, the

502

locus of the points constituting the initial front are wind up in the clockwise

503

direction to represent an expanding front (anticlockwise order represents a

504

contracting front). For each numerical simulation presented here, the initial

505

fire-front is represented by 200 markers. A brief study focussing on the

506

numerical behaviour of DEVS and LSM based techniques is also presented

507

in the next section.

AC

CE

PT

ED

M

492

25

ACCEPTED MANUSCRIPT

The mean wind, wherever used, is assumed to be constant in magnitude

509

and direction. The turbulent diffusion coefficient D, and ignition delays cor-

510

responding to hot air and firebrands heating are also assumed to be constant

511

throughout the simulations. In this study, the ignition delay correspond-

512

ing to the hot air τh and to the firebrand τf are assumed to be 10 min and

513

1 min respectively. Physically, a measure of the ignition delay can be inferred

514

from the fire intensity, fuel properties and the length scale at which the fuel

515

is exposed to the heat source [64]. According to Ref. [64], for the same

516

fuel characteristics and low fire intensities the ignition delay due to hot air is

517

higher than the firebrand landing, but in the cases of high fire intensities, the

518

ignition delay at different length scales can be comparable due to the func-

519

tional relationship between the ability of the fuel surface to absorb energy

520

and the ability of the fuel to conduct heat inward. Flammability classifica-

521

tion of different natural fuels available in literature [65, 66, 67] can aid in the

522

selection of an appropriate value for ignition delay.

ED

M

AN US

CR IP T

508

Appropriate values of D are chosen by using formula 13 to assess the

524

effect of increasing level of turbulence and to highlight its role in comparison

525

to the firebrand landing. In this study, three different values of D are utilised:

526

D = 0.30 m2 s−1 , D = 0.15 m2 s−1 , D = 0.075 m2 s−1 . Also, with reference to the lognormal distribution (14), the unit reference

length `0 in feet (1 ft = 0.30 m).

AC

528

CE

527

PT

523

529

Multiple idealised simulation tests are performed both in the presence and

530

the absence of wind by neglecting and considering the effects of the random

531

phenomena. The first case evaluates an isotropic growth of the fire-line in

532

the absence of wind by neglecting all the random processes. A fire-break 26

ACCEPTED MANUSCRIPT

zone 60 m wide is also considered in this case. In the second test, the spread

534

of fire-line for different ROS in different directions is studied. The third test

535

discusses the propagation of the fire-line with wind when the ROS is defined

536

according to the 3% of the mean wind, formula (19), and when it is defined

537

according to formula (20) . The random processes are neglected for the

538

first three test cases. The pure LSM and pure DEVS based wildfire models

539

fail while managing the realistic cases of fire propagation across the fire-

540

break zones, hence the fourth test evaluates the performance of DEVS when

541

turbulent processes are also considered in the model. The effect of turbulent

542

processes is studied both in the presence and the absence of the wind. The

543

fifth test evaluates the performance of the two simulators when fire-spotting

544

is also included along with turbulence. Fire-break zones are also introduced

545

in the last two tests (the fourth and the fifth) to observe the propagation of

546

the fire-line while encountering areas of null fuel. To simplify the simulation

547

and maintain equivalence between the two simulators, the region across and

548

behind the centre of the initial fire-line is demarcated as the leeward side and

549

the windward side respectively.

PT

ED

M

AN US

CR IP T

533

As an extension of the study, a brief analysis of the firebrand landing

551

distribution is also made. With reference to the lognormal distribution (14)

552

for firebrand landing, different values of µ and σ are chosen motivated by a

553

possible correlation with some physical aspects of wildfire like changing fire

AC

CE

550

554

intensity and meteorological conditions. For these simulations, only the LSM

555

based simulator is used and the wind speed is assumed to be 3 ms−1 , while

556

the ROS is given by the Rothermel model [4, 10].

557

All the simulations have been performed at BCAM–Basque Center for Ap27

ACCEPTED MANUSCRIPT

plied Mathematics (www.bcamath.org) in Bilbao, Basque Country – Spain.

559

6. Discussion

560

6.1. Numerical behaviour of DEVS and LSM

CR IP T

558

The stability of DEVS method relies on the choice of quantum distance

562

∆q and the perimeter resolution ∆c but the numerical simulations involving

563

a uniform ROS in all directions (e.g., a constant 0.05 ms−1 irrespective of

564

wind or direction) are unaffected by the choice of these two parameters. In a

565

homogenous domain, a uniform ROS forces all the markers to advance with

566

an identical speed at a same time step and leads to generation of synchronous

567

events. In this case, all the markers are driven by the same CFL condition,

568

and hence, the motion of the markers is analogous to an Eulerian technique.

569

On the contrary, markers with different speed generate different number of

570

discrete events and the resolution of ∆q and ∆c comes into picture. In con-

571

text of the idealised test cases involving wind, an evaluation of performance

572

of the method concerning different values of ∆q and ∆c is demonstrated in

573

Fig. 1. The numerical simulations show the evolution of the fire-line in the

574

presence of an east wind of 3 ms−1 with ROS as 3% of the normal wind, for

575

different pairs of ∆q and ∆c. Each contour represents the fire-front at 134

576

min. The perimeter resolution should be at least twice of the quantum dis-

577

tance (∆c ≥ 2∆q) to ensure that two markers do not cross-over [68]. Taking

AC

CE

PT

ED

M

AN US

561

578

into account this restriction, the upper plot of Fig. 1 highlights the differences

579

that arise in the evolution of the fire-front with respect to different values

580

of ∆c. DEVS employs a neighbourhood check for each marker to identify

581

its two immediate neighbours and utilises these neighbours to regulate the 28

ACCEPTED MANUSCRIPT

shape of the front by merging and generating new markers. A large value of

583

∆c can consider disconnected markers as neighbours and can lead to a unde-

584

sired generation or merging of markers. The loss of information by unsought

585

merging can cause the front to collapse. On the other hand, decreasing

586

the perimeter resolution ∆c without increasing the resolution of the marker

587

jumps causes the regeneration function to be evoked frequently and leads to

588

creation of a large number of markers with small separations. Over time,

589

too many markers at a coarse spatial scale leads to excessive unnecessary

590

computations and small numerical errors introduced by the computation can

591

escalate quickly over time. A finer scale of ∆q is desired with small values

592

of ∆c for a complete optimisation of the numerical simulation. The lower

593

plot in Fig. 1 shows the variation of the fire-front with respect to increasing

594

resolution of ∆q. Though a finer resolution of ∆q resolves the fire map at a

595

better resolution, it has no significant impact on the simulation results; on

596

the other hand, it increases the computation cost of the simulation. In the

597

present scenario, the homogeneous vegetation and topography of the present

598

domain causes the Lagrangian movement of markers to be driven predom-

599

inantly by regeneration and the fine scale information about the domain is

600

redundant. But on the other hand, in an inhomogeneous domain (involving

601

diverse vegetation, irregular topography, etc), where both decomposition and

602

coalescence play an important role, a finer marker jump length is necessary

CE

PT

ED

M

AN US

CR IP T

582

to account for the finer details of the domain. Usually, the value of ∆q is

604

taken to be around 2-3 m to ensure a correct representation of the vari-

605

able vegetation, changing topography and fire break zones like roads, rivers

606

etc. Concerning a pure quantitative analysis of the total error in DEVS, the

AC 603

29

ACCEPTED MANUSCRIPT

607

interested readers are referred to [68] and [69]. The numerical behaviour of the LSM is dependent on the time step (dt),

609

the grid resolution (dx, dy) and the maximum absolute speed of all points

610

on the grid. The iterative time step involved while solving the differential

611

equations needs to satisfy the CFL criteria for a converging solution. The

612

CFL condition defines the maximum limit of the time step in order to limit

613

the evolution of the contour to at most one grid cell at each time step. This

614

upper limit on the time step is important for the stability of the system and

615

a converging solution. In one-dimension, the CFL condition can be described

616

as:

AN US

CR IP T

608

dt ≤ k

min(dx, dy) vmax

(21)

where, vmax is the maximum absolute speed, and 0 < k < 1 is a constant. To

618

demonstrate the numerical behaviour of LSM, similar numerical simulations

619

with 3% model and an east wind of 3 ms−1 are repeated for different values

620

of dt and dx . The upper figure in Fig. 2 shows the fire-line contours at 380

621

min for various values of dt, and when dx = dy = 20 m. With values of dt

622

smaller than the threshold defined by CFL condition, the fire-line evolves

623

without any instabilities. Similar results are achieved with further decrease

624

in the time step but with an increase in the computation time. On the

625

other hand, when for the same speed function, dt is chosen to violate the

626

CFL criteria, the fire-line evolves with an undesirable behaviour and this

AC

CE

PT

ED

M

617

627

obscurity magnifies over time. In the same way, CFL condition can also

628

be interpreted as the constraint on the grid resolution. The lower figure in

629

Fig. 2 shows the evolution of the fire-line with respect to different values of

630

dx. The increase in the grid resolution over an identical time step introduces 30

ACCEPTED MANUSCRIPT

numerical instability and causes the fire-lines to evolve erroneously. For the

632

same velocity field, shorter time steps are required over small grid spacings

633

in order to respect the CFL condition, though it leads to an increase in the

634

computation costs. Hence, accounting for the numerical behaviour of the

635

LSM and to avoid the excessive computation, in the present study, dt is

636

defined according to (21), with k = 0.5.

CR IP T

631

It is remarked here that since the mathematical formulation for turbulence

638

and fire-spotting is implemented in DEVS and LSM based wildfire simulators

639

as a post-processing scheme, it preserves the original framework of the wildfire

640

simulators and does not affect the numerical stability and convergence of both

641

these methods.

642

6.2. Comparison of DEVS and LSM

M

AN US

637

In the first three test scenarios, the evolution of the fire-line without the

644

impact of random processes in both LSM and DEVS approaches is analysed.

645

Fig. 3 presents the propagation of the fire-line in no wind conditions for both

646

the simulators. The advancement of the initial circular fire-line follows an

647

isotropic growth and with identical initial conditions the two different moving

648

interface schemes provide a similar evolution of the fire-line. But the fire-line

649

fails to propagate across the fire-break zone in both the techniques. The

650

evolution is shown only up to 140 min, but an extended run up to 250 min

651

(not shown) indicates the limitation of the two approaches to reproduce the

AC

CE

PT

ED

643

652

653

fire jump across the no fuel zone. Fig. 4 presents the growth of an initial spot fire but with a non-homogeneous

654

ROS in absence of any wind. The different values of the ROS can be at-

655

tributed to different fuel types, but here no effort has been made to describe 31

ACCEPTED MANUSCRIPT

the fuel type. The fire-line propagates with different speed in the three dif-

657

ferent directions, but the isotropy is preserved. The results from these two

658

test cases indicate that in the absence of wind, the two advancing schemes:

659

LSM and DEVS, show an identical behaviour in simulating situations with

660

constant ROS and hence, this observation provides a crucial background and

661

scope for the comparison of situations with increasing variability and com-

662

plexity.

CR IP T

656

Fig. 5 shows the evolution of the fire-front of a circular initial profile with

664

radius 300 m in case of a weak wind of 3 ms−1 directed in the east direction.

665

The isochronous fronts are plotted at every 20 min and follow an oval shape

666

for both the simulators. DEVS fire contours diverge slightly from the mean

667

wind direction and develop into an increasing flanking fire over time. This

668

divergence in the evolution of fire-front occurs due to differences in the com-

669

putation of the normal between the two approaches. This fact can be very

670

well appreciated when an initial square profile is considered. Fig. 6 shows the

671

progress of the fire when the initial fire perimeter is considered to be a square

672

with side 600 m. Under the effect of a constant zonal wind and restricted 3%

673

model (19), the evolution in LSM strictly follows the initial square shape,

674

but in DEVS the active burning markers at the corners advance spuriously

675

to provide an additional flanking spread. The DEVS markers at the corners

676

are affected by the approximate normal and this leads to a larger flank fire.

AC

CE

PT

ED

M

AN US

663

677

Whereas, the normal direction for markers pointing towards the direction of

678

wind is identical for both the methods, hence the head fires arrive to same

679

location in both the simulations. The flank fires generated by the approxi-

680

mate construction of the normal seem to be in a more qualitative agreement 32

ACCEPTED MANUSCRIPT

681

with a realistic case, but within the stated analytical formulation of the ROS

682

(19), the lateral progress of the flank fires is an anomalous behaviour. The ROS formulation developed by Mallet et al. [6], here formula (20),

684

provides a simple way to study the evolution of flank fire by introducing

685

different ROS for the head, flank and rear directions. In the original paper,

686

θ is defined as the angle between the normal to the front and the wind

687

direction. Since, the normal computation in DEVS approach is approximate,

688

two separate tests are performed to evaluate the effect of such approximation

689

on the spread: firstly when θ is computed according to the original definition

690

[6], and secondly when θ is assumed to be the angle between the line joining

691

the front and the wind direction to enforce an identical definition of θ in

692

both the methods. Fig. 7 shows the evolution of fire when θ is computed

693

according to the original definition; the advancement of the head and rear

694

fires are identical, but the flank fire spread is spuriously larger in DEVS

695

because of the approximate computation of the front normal direction. On

696

the other hand, it is evident from Fig. 8 that in the second case, when the

697

computation of θ is not affected by the approximate computation of the

698

front normal in DEVS, the fire-line advancement is identical to LSM in all

699

directions.

AN US

M

ED

PT

As shown in Fig. 3, in case of fire-break zones, pure LSM and DEVS

based simulators are inherently unable to simulate the realistic situations of

AC

701

CE

700

CR IP T

683

702

fire overcoming a fire break. But Fig. 9 shows that with the introduction of

703

turbulence, the simulators can model the effect of hot air to overcome fire-

704

break zones. Here the value of turbulent diffusion coefficient D is assumed

705

to be 0.15 m2 s−1 . The evolution of the fire-line is almost similar for both 33

ACCEPTED MANUSCRIPT

the simulators, though a slight underestimation can be visibly observed in

707

the DEVS approach. A cross section of φ(x, t) and φe (x, t) at y = 2500 m

708

(Fig. 10) provides a more detailed illustration of the differences in the evolu-

709

tion of the fire-line between the two approaches. The introduction of marginal

710

differences with the inclusion of turbulence can be explained by the contrast

711

in the construction of the indicator function between the two simulators. In

712

LSM, the indicator function is defined over a Cartesian grid by definition,

713

while on the other hand, in DEVS, a gridded indicator function is constructed

714

from the original real coordinates. The point in polygon technique is used to

715

transform the real coordinates to a Cartesian grid. This technique provides

716

only an approximate measure of the same; small errors can be introduced

717

by the improper classification of the points lying very close to the boundary

718

[70], and can lead to a slight underestimation/overestimation of the original

719

fire-line.

M

AN US

CR IP T

706

Fig. 11 shows the results for the two simulators when the turbulent dif-

721

fusion coefficient is D = 0.30 m2 s−1 . Stronger turbulence causes a rapid

722

propagation of the fire-line and an earlier ignition across the fire-break zone.

723

A detailed analysis of the effect of varying turbulence over long-term propa-

724

gation with the LSM can be found in [35, 36].

PT

Fig. 12 presents the effect of inclusion of turbulence with a non-zero wind

profile and 3% model (19). The effect of turbulence is most pronounced in

AC

726

CE

725

ED

720

727

the direction of the wind and it can be clearly observed that with turbulence,

728

the spread over the head, flank and back fire-lines is faster and the head fire

729

is also able to overcome the fuel break. The quantum of increase in the

730

fire spread can also be appreciated in comparison to Fig. 5, which presents 34

ACCEPTED MANUSCRIPT

the ideal case with the same initial conditions. Both simulators show almost

732

similar characteristics in the spread of fire, though the flank fire has a slightly

733

larger spread in DEVS.

CR IP T

731

Another aspect contributing towards the increase in the fire spread due

735

to new fire ignitions generated by the firebrands is presented in Fig. 13. With

736

the inclusion of fire-spotting along with turbulence, the evolution of the fire-

737

front is faster in comparison to the effect of turbulence alone (Fig. 12). The

738

flank fire and the head fire are also well simulated in both the approaches,

739

and again a larger spread out the flanking fires is observed in the DEVS

740

scheme. The proposed mathematical formulation of spot-fires along with

741

turbulence is effective in mimicking the rapid advancement of the fire-line

742

due to the generation of new ignition points by the firebrands. It is also

743

effective in simulating the new ignition points developed across the no fuel

744

zones. Within the current parametrisation, the firebrand landing distances

745

lie close to the main fire and are not long enough to develop a new secondary

746

fire, but in the next section, a sensitivity study of the spatial pattern of the

747

firebrand landing distances is also presented.

PT

ED

M

AN US

734

Fig. 14 shows the time evolution of the fire-front in presence of a mean

749

wind velocity of 3 ms−1 and two fire-break zones. The results are obtained

750

when only the effects of turbulence are included in the formulation and the

751

diffusion coefficient D = 0.075 m2 s−1 . Similarly, Fig. 15 shows the time evo-

AC

CE

748

752

lution of fire-front but when both turbulence and fire-spotting are included

753

in the formulation. Both approaches follow an identical evolution pattern as

754

observed in other cases. As expected, the spread of the fire-front is faster and

755

is able to overcome both the fire-break zones when turbulence is included; 35

ACCEPTED MANUSCRIPT

but inclusion of fire-spotting along with turbulence causes the fire to propa-

757

gate at a much faster rate and results in a larger spread. Again, the major

758

differences which arise in the evolution of flank fires between the two meth-

759

ods are due to the dissimilarity in the construction of the direction of the

760

propagation of the active burning points.

761

7. Insight on firebrand distribution

CR IP T

756

In the simulations discussed previously, the phenomenon of fire-spotting

763

contributes towards an increase in the burning area and a faster propagation

764

of fire, but the landing distances are not sufficiently long enough to develop

765

new secondary fires detached from the primary fire.

AN US

762

Generally, the firebrands generated in low intensity fires fall close to the

767

primary fire and although they contribute towards a faster fire spread, the

768

separation from the primary fire is not large enough for the new ignitions

769

to develop a new secondary fire. But high intensity fires and extreme mete-

770

orological conditions such as strong wind/extremely high temperatures/low

771

humidity can facilitate the transport of embers over longer distances and the

772

area under firebrand attack can be sufficiently far away from the primary

773

fire [71, 72, 73, 74, 75]. In Ref. [76] and references therein, a brief account

774

of the impact of various historical forest fires is provided emphasising the

775

importance of strong winds, and extreme meteorological conditions on the

AC

CE

PT

ED

M

766

776

fire-spotting phenomenon. They describe different historical fires where low

777

intensity fires coupled with strong winds and low relative humidity caused

778

a large scale fire-spotting, while on the other hand, fires with high intensi-

779

ties in calm meteorological conditions had a negligible contribution from the 36

ACCEPTED MANUSCRIPT

780

firebrands. In lognormal distribution (14), a suitable choice of the parameters µ and

782

σ can be utilized to modify the firebrand landing distributions to represent

783

different fire intensities and wind conditions. In this study, the main interest

784

lies in the mathematical modelling of the firebrands landing distributions,

785

hence constant meteorological conditions are assumed for all the simulations

786

presented here. It is also assumed that each active ignition point is capable

787

of generating firebrands and all embers are controlled by identical transport

788

forces.

AN US

CR IP T

781

Fig. 16 and Fig. 17 show the variation of the area under the firebrand

790

attack with different values of µ and σ. A constant σ and an increasing µ

791

points towards a decrease in the effectiveness of the firebrands, and slows

792

down the fire propagation. A slight deviation from the pattern is observed

793

for µ = 15 and σ = 5, when the firebrands travel far enough to create a new

794

secondary fire.

ED

M

789

For an increasing value of σ (Fig. 17a-c) with constant µ, the fire evolu-

796

tion increases as larger number of firebrands land away from the source, but

797

the landing distances are still not long enough to delineate the new ignitions

798

from the main fire. With further increase in σ, a secondary fire emerges away

799

from the main source. Equal contribution to the firebrands by each flaming

800

point leads to a regular distribution pattern of the secondary fire. As time

AC

CE

PT

795

801

progresses, the primary fire also advances and covers up the gap with the

802

secondary fire. With further increase in σ, the probability of the firebrands

803

landing at farther distances increases, and this can lead to a potentially ex-

804

plosive firebrands attack. Fig. 17f shows the effect of fire-spotting at shorter 37

ACCEPTED MANUSCRIPT

scales contributing to a faster rate of spread, but the development of new

806

secondary fires is not observed. For this particular value of σ, the statistical

807

distribution of the long range landing distances lies outside the domain and

808

is impossible to resolve them in the limited time and spatial scales considered

809

in this simulation.

CR IP T

805

As a consequence of the idealised set-up for simulations, the new sec-

811

ondary fires evolve with the same intensity as the primary fire and also serve

812

as a source of generation of newer fire brands and henceforth the tertiary

813

fires. Fig. 18a-f demonstrates the generation of new secondary fires by the

814

primary fire, and tertiary fires by the secondary fires. As time progresses,

815

the each of the fire-line evolves independently along with a small contribution

816

from its parent fire-line, and eventually they merge among each other.

AN US

810

In a lognormal distribution, an increasing value of the µ causes a right

818

shift of the maximum, indicating towards a larger jump length, but the sim-

819

ulations suggest a decreasing trend in the landing distance. This can be

820

explained by the fact that the ignition of the unburned fuel is not an instan-

821

taneous process, but rather an accumulative process. The fuel starts burning

822

only when the temperature is high enough to allow spontaneous combustion.

823

Though a larger µ allows the firebrands to fall at larger distances but it limits

824

the frequency of landing. Insufficient number of firebrands landing at farther

825

distances diminishes the efficacy of their attack to cause combustion; hence

AC

CE

PT

ED

M

817

826

leads to a smaller contribution from fire-spotting. For an effective combus-

827

tion, sufficient number of firebrands should be available for an adequate heat

828

exchange with the unburned fuel. On the contrary, an increasing σ causes the

829

lognormal distribution to have a lower maximum and an elongated right tail 38

ACCEPTED MANUSCRIPT

of the distribution, but with a high probability of landing far away from the

831

main source. The increase in the number of firebrands landing away from

832

the maximum leads to a rapid ignition of the unburned fuel and a faster

833

propagation of the fire.

CR IP T

830

According to the increasing and decreasing trend of the landing distance

835

with respect to increasing σ and decreasing µ respectively, within the log-

836

normal characterisation assumed here for the firebrand landing distribution,

837

the reference length scale L of the landing distance can be stated as:

AN US

834

L∝ 838

σα , µβ

(22)

where α > 0 and β > 0 are exponents to be established.

From relation (22) it can be deduced that neither the maximum nor the

840

mean of the lognormal distribution can be considered as an estimation of

841

the landing distance length scale L. Physically, the landing distance can be

842

described in terms of the dimensions of the convective column, wind condi-

843

tions, fire intensity and fuel characteristics. Further analysis towards a phys-

844

ical parametrisation of fire-spotting is postponed for a future paper, where

845

firebrand distributions different from the lognormal will be also considered.

846

8. Conclusions

ED

PT

CE

847

M

839

This article describes a mathematical formulation to model the effects

of turbulence and fire-spotting in wild-land fire propagation models. Previ-

849

ously, this formulation has been applied to an Eulerian LSM based wildfire

850

propagation method, and this article describes the implementation of the

851

mathematical formulation into a Lagrangian DEVS based wildfire propaga-

852

tion method. This formulation splits the motion of the front into a drifting

AC

848

39

ACCEPTED MANUSCRIPT

part and a fluctuating part. The drifting part is independent of the fluctuat-

854

ing part and is given by the fire perimeter determined from one of the various

855

wildfire propagation methods that exist in literature (DEVS and LSM in the

856

present study); while the fluctuating part is generated by a comprehensive

857

statistical description of the system and includes the effects of random pro-

858

cesses in agreement with the physical properties of the process. This study

859

also provides a unique opportunity for the comparison of two simple wildfire

860

simulators based on the Eulerian LSM and on the Lagrangian DEVS scheme.

861

Numerical simulations based on identical set-ups are used to compare the

862

performance of the two approaches.

AN US

CR IP T

853

A series of idealised numerical simulations show that the proposed ap-

864

proach for random effects emerges to be suitable for both LSM and DEVS

865

based simulators to manage the real world situations related to random char-

866

acter of fire, e.g., increase in fire spread due to pre-heating of the fuel by hot

867

air, vertical lofting and transporting of firebrands, fire overcoming no fuel

868

zones, and development of new isolated secondary fires. The simulations

869

from both LSM and DEVS techniques follow nearly identical patterns, and

870

the only difference which distinguishes the outputs from the two techniques

871

is due to the distinction in the treatment of the direction of propagation. In

872

DEVS, the active burning points propagate in a direction given by the bisec-

873

tor of the angle made by each marker with its immediate left and right mark-

AC

CE

PT

ED

M

863

874

ers, in contrast to the normal direction of propagation used in LSM. Such

875

differences result in “spurious” flanking fires in DEVS, which are anomalous

876

with respect to the definition of the ROS, but qualitatively provide a more

877

“realistic” representation of the fire contour. Due to this fact, the two meth40

ACCEPTED MANUSCRIPT

ods can be considered complementary to each other for simple situations,

879

but with increasing complexity and the introduction of random processes a

880

validation exercise is necessary to single out the best representation of the

881

fire propagation.

CR IP T

878

A brief study on the sensitivity of the firebrand landing distribution to

883

the changes in the shape of the lognormal distribution indicates that the

884

shape parameters (µ and σ) can control the spread of the firebrand landing

885

and the lognormal distribution is capable of simulating new secondary fires

886

separated from the main fire perimeter. In future, the formulation for the

887

firebrand landing distance would be modified to include information about

888

the wind speed, fire intensity, radius of the firebrands and fuel characteristics.

889

Though without a validation test case, these simulations do provide an

890

insight into the flexibility of this formulation to incorporate effects of random

891

processes such as the turbulent heat diffusion and fire-spotting, specially

892

while characterising the landing distance of firebrands. It can be deduced

893

that the mathematical formulation is independent of the choice of the method

894

used to ascertain the drifting part and can serve as a versatile addition to

895

the existing fire spread simulators to include the effects of random spread.

896

The results from this study provide a support to the implementation of the

897

proposed approach for the effects of random phenomena into operational

898

softwares such as WRF-SFIRE and ForeFire. The source code utilized for all

AC

CE

PT

ED

M

AN US

882

899

the simulations in this article is hosted at https://github.com/ikaur17/

900

firefronts.

41

ACCEPTED MANUSCRIPT

901

Acknowledgement This research is supported by MINECO under Grant MTM2013-40824-P,

903

by Bizkaia Talent and European Commission through COFUND programme

904

under Grant AYD-000-226, and also by the Basque Government through the

905

BERC 2014-2017 program and by the Spanish Ministry of Economy and

906

Competitiveness MINECO: BCAM Severo Ochoa accreditation SEV-2013-

907

0323. The authors would also like to thank the two anonymous reviewers for

908

their insightful comments which helped in improving the manuscript both in

909

presentation and scientific content.

910

References

AN US

CR IP T

902

[1] A. L. Sullivan, Wildland surface fire spread modelling, 1990–2007. 1:

912

Physical and quasi-physical models, Int. J. Wildland Fire 18 (2009) 349–

913

368.

ED

M

911

[2] A. L. Sullivan, Wildland surface fire spread modelling, 1990–2007. 2:

915

Empirical and quasi-empirical models, Int. J. Wildland Fire 18 (2009)

916

369–386.

918

Simulation and mathematical analogue models, Int. J. Wildland Fire 18 (2009) 387–403.

AC

919

[3] A. L. Sullivan, Wildland surface fire spread modelling, 1990–2007. 3:

CE

917

PT

914

920

[4] R. C. Rothermel, A mathematical model for predicting fire spread in

921

wildland fires, Research Paper INT-115, USDA Forest Service, Inter-

922

mountain Forest and Range Experiment Station, Ogden, Utah 84401,

923

available at: http://www.treesearch.fs.fed.us/pubs/32533 (1972). 42

ACCEPTED MANUSCRIPT

924

925

[5] J. H. Balbi, F. Morandini, X. Silvani, J. B. Filippi, F. Rinieri, A physical model for wildland fires, Combust. Flame 156 (2009) 2217–2230. [6] V. Mallet, D. E. Keyes, F. E. Fendell, Modeling wildland fire propaga-

927

tion with level set methods, Comput. Math. Appl. 57 (2009) 1089–1101.

928

[7] M. Finney, Fire growth using minimum travel time methods, Can. J.

930

931

932

933

For. Res. 32 (2002) 420–1424.

[8] M. Finney, Calculation of fire spread rates across random landscapes,

AN US

929

CR IP T

926

Int. J. Wildland Fire 12 (2003) 167–174.

[9] J. A. Sethian, P. Smereka, Level set methods for fluid interfaces, Ann. Rev. Fluid Mech. 35 (2003) 341–372.

[10] J. Mandel, J. D. Beezley, A. K. Kochanski, Coupled atmosphere-

935

wildland fire modeling with WRF 3.3 and SFIRE 2011, Geosci. Model.

936

Dev. 4 (2011) 591–610.

ED

M

934

[11] H. Karimabadi, J. Driscoll, Y. A. Omelchenko, N. Omidi, A new asyn-

938

chronous methodology for modeling of physical systems: breaking the

939

curse of Courant condition, J. Comp. Phys. 205 (2005) 755–775.

CE

PT

937

[12] Y. A. Omelchenko, H. Karimabadi, HYPERS: A unidimensional asyn-

941

chronous framework for multiscale hybrid simulations, J. Comp. Phys.

AC

940

942

231 (2012) 1766–1780.

943

[13] J. B. Filippi, F. Morandini, J. H. Balbi, D. Hill, Discrete event front

944

tracking simulator of a physical fire spread model, Simulation 86 (2009)

945

629–646. 43

ACCEPTED MANUSCRIPT

[14] J. B. Filippi, V. Mallet, B. Nader, Evaluation of forest fire models on

947

a large observation database, Nat. Hazards Earth Syst. Sci. 14 (2014)

948

3077–3091.

950

951

952

[15] D. Anderson, E. Catchpole, N. de Mestre, T. Parkes, Modelling the spread of grass fires., J. Aus. Math. Soc. 23 (1982) 451–466.

[16] G. D. Richards, An elliptical growth model of forest fire fronts and its numerical solution, Int. J. Numer. Meth. Eng. 30 (1990) 1163–1179.

AN US

949

CR IP T

946

953

[17] G. D. Richards, The properties of elliptical wildfire growth for time

954

dependent fuel and meteorological conditions, Combust. Sci. Technol.

955

95 (1993) 357–383.

958

959

M

957

[18] G. D. Richards, A general mathematical framework for modelling twodimensional wildland fire spread, Int. J. Wildland Fire 5 (1995) 63–72. [19] J. Glasa, L. Halada, On elliptical model for forest fire spread modeling

ED

956

and simulation, Math. Comput. Simulat. 78 (2008) 76–88. [20] O. Rios, W. Jahn, G. Rein, Forecasting wind-driven wildfires using an

961

inverse modelling approach, Nat. Hazards Earth Syst. Sci. 14 (2014)

962

1491–1503.

CE

[21] T. L. Clark, M. A. Jenkins, J. Coen, D. Packham, A coupled

AC

963

PT

960

964

atmospheric-fire model: convective feedback on fire-line dynamics, J.

965

Appl. Meteor. 35 (1996) 875–901.

966

967

[22] B. E. Potter, A dynamics based view of atmosphere-fire interactions, Int. J. Wildland Fire 11 (2002) 247–255. 44

ACCEPTED MANUSCRIPT

[23] B. E. Potter, Atmospheric interactions with wildland fire behaviour - I.

969

Basic surface interactions, vertical profiles and synoptic structures, Int.

970

J. Wildland Fire 21 (2012) 779–801.

971

972

CR IP T

968

[24] B. E. Potter, Atmospheric interactions with wildland fire behaviour - II. Plume and vortex dynamics, Int. J. Wildland Fire 21 (2012) 802–817.

[25] C. B. Clements, S. Zhong, X. Bian, W. E. Heilman, D. W. Byun, First

974

observations of turbulence generated by grass fires, J. Geophys. Res. 113

975

(2008) D22102.

AN US

973

976

[26] N. Sardoy, J. L. Consalvi, B. Porterie, A. C. Fernandez-Pello, Modeling

977

transport and combustion of firebrands from burning trees, Combust.

978

Flame 150 (2007) 151–169.

[27] N. Sardoy, J. L. Consalvi, A. Kaiss, A. C. Fernandez-Pello, B. Porterie,

980

Numerical study of ground-level distribution of firebrands generated by

981

line fires, Combust. Flame 154 (2008) 478–488.

ED

M

979

[28] S. Kortas, P. Mindykowski, J. L. Consalvi, H. Mhiri, B. Porterie, Exper-

983

imental validation of a numerical model for the transport of firebrands,

984

Fire Safety J. 44 (2009) 1095–1102.

CE

PT

982

[29] H. A. Perryman, A mathematical model of spot fires and their manage-

986

ment implications, Master’s thesis, Humboldt State University, Arcata,

AC

985

987

CA (2009).

988

[30] H. A. Perryman, C. J. Dugaw, J. M. Varner, D. L. Johnson, A cellular

989

automata model to link surface fires to firebrand lift-off and dispersal,

990

Int. J. Wildland Fire 22 (2013) 428–439. 45

ACCEPTED MANUSCRIPT

992

993

994

[31] C. Favier, Percolation model of fire dynamic, Phys. Lett. A 330 (2004) 396–401. [32] H. Hunt, A new conceptual model for forest fires based on percolation

CR IP T

991

theory, Complexity 13 (2007) 12–17.

[33] D. Boychuk, W. J. Braun, R. J. Kulperger, Z. L. Krougly, D. A. Stan-

996

ford, A stochastic forest fire growth model, Environ. Ecol. Stat. 16 (2009)

997

133–151.

998

999

1000

AN US

995

[34] L. Ntaimo, B. P. Zeigler, M. J. Vasconcelos, B. Khargharia, Forest fire spread and supression in DEVS, Simulation 80 (2004) 479–500. [35] G.

Pagnini,

fire

Massidda,

method,

1003

dinia,

1004

http://publications.crs4.it/pubdocs/2012/PM12a/pagnini massidda-

1005

levelset.pdf and arXiv:1408.6129 (July 2012).

revised

version:

randomized

CRS4,

August

Pula 2014,

level-set

(CA),

Sar-

available

at:

PT

ED

Italy,

2012/PM12a,

the

effects

1002

Rep.

by

turbulence

in

Tech.

propagation

Modelling

1001

M

wildland

L.

[36] G. Pagnini, L. Massidda, The randomized level-set method to model

1007

turbulence effects in wildland fire propagation, in: D. Spano, V. Bacciu,

1008

M. Salis, C. Sirca (Eds.), Modelling Fire Behaviour and Risk. Proceedings of the International Conference on Fire Behaviour and Risk. ICFBR

AC

1009

CE

1006

1010

2011, Alghero, Italy, October 4–6 2011, 2012, pp. 126–131, ISBN 978-

1011

88-904409-7-7.

1012

[37] G. Pagnini, A model of wildland fire propagation including random ef-

1013

fects by turbulence and fire spotting, in: Proceedings of XXIII Congreso 46

ACCEPTED MANUSCRIPT

1014

de Ecuaciones Diferenciales y Aplicaciones XIII Congreso de Matem´atica

1015

Aplicada. Castell´o, Spain, 9–13 September 2013, 2013, pp. 395–403. [38] G. Pagnini, Fire spotting effects in wildland fire propagation, in:

1017

F. Casas, V. Mart´ınez (Eds.), Advances in Differential Equations and

1018

Applications, Vol. 4 of SEMA SIMAI Springer Series, Springer Interna-

1019

tional Publishing Switzerland, 2014, pp. 203–214, dOI 10.1007/978-3-

1020

319-06953-1 20. ISBN: 978-3-319-06952-4 (eBook: 978-3-319-06953-1).

AN US

CR IP T

1016

[39] G. Pagnini, A. Mentrelli, Modelling wildland fire propagation by track-

1022

ing random fronts, Nat. Hazards Earth Syst. Sci. 14 (2014) 2249–2263.

1023

[40] G. Pagnini, A. Mentrelli, The randomized level set method and an as-

1024

sociated reaction-diffusion equation to model wildland fire propagation,

1025

in: Proceedings of The 18th European Conference on Mathematics for

1026

Industry, ECMI 2014, Taormina, Italy, June 9–13, 2014, ECMI book

1027

subseries of Mathematics in Industry, Springer Heidelberg, accepted.

ED

M

1021

[41] I. Kaur, A. Mentrelli, F. Bosseur, J. B. Filippi, G. Pagnini, Wildland

1029

fire propagation modelling: A novel approach reconciling models based

1030

on moving interface methods and on reaction-diffusion equations, in:

1031

ˇıstek, T. Vejchodsk´ J. Brandts, S. Korotov, M. Kˇr´ıˇzek, K. Segeth, J. S´ y (Eds.), Proceedings of the International Conference on Applications of Mathematics 2015; Prague, Czech Republic, 18–21 November (2015),

AC

1033

CE

1032

PT

1028

1034

Institute of Mathematics – Academy of Sciences, Czech Academy of

1035

Sciences, Prague, Czech Republic, 2015, pp. 85–99.

1036

[42] J. B. Filippi, F. Bosseur, D. Grandi, ForeFire: open-source code for 47

ACCEPTED MANUSCRIPT

wildland fire spread models, in: D. X. Viegas (Ed.), Advances in forest

1038

fire research, Imprensa da Universidade de Coimbra, 2014, pp. 275–282.

1039

URL https://digitalis.uc.pt/handle/10316.2/34013

1040

1041

CR IP T

1037

[43] B. P. Zeigler, T. G. Kim, H. Praehofer, Theory of Modeling and Simulation, 2nd Edition, Academic Press, Inc., Orlando, FL, USA, 2000.

[44] J. C. Carvalho, M. T. Vilhena, D. M. Moreira, Comparison between Eu-

1043

lerian and Lagrangian semi-analytical models to simulate the pollutant

1044

dispersion in the PBL, Appl. Math. Model. 31 (2007) 120–129.

AN US

1042

[45] Z. Zhang, Q. Chen, Comparison of the Eulerian and Lagrangian methods

1046

for predicting particle transport in enclosed spaces, Atmos. Environ. 41

1047

(2007) 5236–5248.

M

1045

[46] M. Saidi, M. Rismanian, M. Monjezi, M. Zendehbad, S. Fatehiboroujeni,

1049

Comparison between Lagrangian and Eulerian approaches in predicting

1050

motion of micron-sized particles in laminar flows, Atmos. Environ. 89

1051

(2014) 199–206.

1054

object-oriented environment, Simulation 49 (1987) 219–230. [48] S. Osher, R. Fedkiw, Level Set Methods and Dynamic Implicit Surfaces, Springer–Verlag, New York, Berlin, Heidelberg, 2003.

AC

1055

PT

1053

[47] B. P. Zeigler, Hierarchical, modular discrete-event modelling in an

CE

1052

ED

1048

1056

1057

[49] R. G. Rehm, R. J. McDermott, Fire-front propagation using the level set method, Tech. Note 1611, Natl. Inst. Stand. Technol. (April 2009).

48

ACCEPTED MANUSCRIPT

1058

[50] E. Jones, T. Oliphant, P. Peterson, et al., SciPy: Open source scientific tools for Python (2001–).

1060

URL http://www.scipy.org/

1061

1062

1063

CR IP T

1059

[51] J. D. Hunter, Matplotlib: A 2D graphics environment, Comput. Sci. Eng. 9 (3) (2007) 90–95.

[52] F. P´erez, B. E. Granger, IPython: a system for interactive scientific computing, Comput. Sci. Eng. 9 (3) (2007) 21–29.

1065

URL http://ipython.org

1066

1067

AN US

1064

[53] A. Mentrelli, G. Pagnini, Random front propagation in fractional diffusive systems, Commun. Appl. Ind. Math. 6 (2) (2015) e–504. [54] A. Mentrelli, G. Pagnini, Front propagation in anomalous diffusive me-

1069

dia governed by time-fractional diffusion, J. Comp. Phys. 293 (2015)

1070

427–441.

1073

1074

ED

[56] J. J. Niemela, K. R. Sreenivasan, Turbulent convection at high Rayleigh numbers and aspect ratio 4, J. Fluid Mech. 557 (2006) 411–422.

[57] F. A. Albini, Potential spotting distance from wind-driven surface fires,

AC

1075

combustion, Phys. Rev. Lett. 107 (2011) 044503.

PT

1072

[55] G. Pagnini, E. Bonomi, Lagrangian formulation of turbulent premixed

CE

1071

M

1068

1076

Research Paper INT-309, U.S. Department of Agriculture Forest Service

1077

Intermountain Forest and Range Experiment Station (1983).

1078

[58] G. A. Morris, A simple method for computing spotting distances from

1079

wind-driven surface fires, Research Note INT-374, U.S. Department of 49

ACCEPTED MANUSCRIPT

1080

Agriculture Forest Service Intermountain Forest and Range Experiment

1081

Station (1987). [59] W. Catchpole, Fire properties and burn patterns in heterogeneous

1083

landscapes, in: R. A. Bradstock, J. E. Williams, M. A. Gill (Eds.),

1084

Flammable Australia: The Fire Regimes and Biodiversity of a Conti-

1085

nent, Cambridge University Press, 2002, Ch. 3, pp. 49–75.

CR IP T

1082

[60] M. E. Houssami, E. Mueller, J. C. Thomas, A. Simeoni, A. Filkov,

1087

N. Skowronski, M. R. Gallagher, K. Clark, R. Kremens, Experimental

1088

procedures characterising firebrand generation in wildfires, Fire Tech-

1089

nol.(Available online). doi:10.1007/s10694-015-0492-z.

1091

[61] K. D. Hage, On the dispersion of large particles from a 15-m source in the atmosphere, J. Appl. Meteor. 18 (1961) 534–539.

M

1090

AN US

1086

[62] S. L. Manzello, J. R. Shields, T. G. Cleary, A. Maranghides, W. E.

1093

Mell, J. C. Yang, Y. Hayashi, D. Nii, T. Kurita, On the development

1094

and characterization of a firebrand generator, Fire Safety J. 43 (2008)

1095

258–268.

1097

PT

[63] H. H. Wang, Analysis on downwind distribution of firebrands sourced

CE

1096

ED

1092

from a wildland fire, Fire Technol. 47 (2011) 321–340.

[64] H. E. Anderson, Forest fuel ignitibility, Fire Technol. 6 (1970) 312–319.

1099

[65] A. Dimitrakopoulos, K. K. Papaioannou, Flammability assessment of

AC

1098

1100

Mediterranean forest fuels, Fire Technol. 37 (2) (2001) 143–152.

50

ACCEPTED MANUSCRIPT

[66] K. J. M. Dickinson, J. B. Kirkpatrick, The flammability and energy

1102

content of some important plant species and fuel components in the

1103

forests of Southeastern Tasmania, J. Biogeogr. 12 (1985) 121–134.

1104

[67] D. M. J. S. Bowman, B. A. Wilson, Fuel characteristics of coastal mon-

1105

soon forests, Northern Territory, Australia, J. Biogeogr. 15 (1988) 807–

1106

817.

CR IP T

1101

[68] J. B. Filippi, F. Bosseur, C. Mari, C. Lac, P. L. Moigne, B. Cuenot,

1108

D. Veynante, D. Cariolle, J. H. Balbi, Coupled atmosphere-wildland fire

1109

modelling, J. Adv. Model. Earth Syst. 1 (2009) art. #11.

1112

1113

1114

tinuous systems, Simulation 78 (2002) 76–89.

M

1111

[69] E. Kofman, A second order approximation for DEVS simulation of con-

[70] K. Hormann, A. Agathos, The point in polygon problem for arbitrary polygons, Comp. Geom.-Theo. Appl. 20 (2001) 131–144.

ED

1110

AN US

1107

[71] R. Wilson, The Devil Wind and Wood Shingles: The Los Angeles Conflagration of 1961, National Fire Protection Assn., 1962.

1116

URL https://books.google.es/books?id=bWFCywAACAAJ

1118

sin Magazine of History 54 (1971) 246–272. URL http://www.jstor.org/stable/4634648

AC

1119

[72] P. Pernin, The great Peshtigo fire: An eyewitness account, The Wiscon-

CE

1117

PT

1115

1120

1121

[73] P. J. Pagni, Causes of the 20 October 1991 Oakland Hills conflagration, Fire Safety J. 21 (1993) 331–339.

51

ACCEPTED MANUSCRIPT

[74] J. Trelles, P. J. Pagni, Fire-induced winds in the 20 October 1991 Oak-

1123

land Hills fire, in: Y. Hasemi (Ed.), Proc. of Fifth International Sympo-

1124

sium on Fire Safety Science, Melbourne, Australia, International Asso-

1125

ciation of Fire Safety Science: Boston, MA, 1997, pp. 911–922.

1126

1127

CR IP T

1122

[75] J. G. Quintiere, Principles of Fire Behavior, Career Education Series, Delmar Publishers, Albany, N.Y., 1998.

[76] E. Koo, P. J. Pagni, D. R. Weise, J. P. Woycheese, Firebrands and

1129

spotting ignition in large-scale fires, Int. J. Wildland Fire 19 (2010)

1130

818–843.

AC

CE

PT

ED

M

AN US

1128

52

ACCEPTED MANUSCRIPT

3500

Y (m)

"c = 4 "c = 8 "c = 12 "c = 18

1500 "q = 1, "q = 2, "q = 3, "q = 4,

"c = 8 "c = 8 "c = 8 "c = 8

2500

AN US

500 0 2000

CR IP T

"q = 2, "q = 2, "q = 2, "q = 2,

2500

3000

3500

4000

4500

X (m)

Figure 1: The fire-lines showing the numerical behaviour of DEVS for different pairs of ∆c and ∆q. Each contour represents the fire-line at 134 min. The initial fire is a circle of

AC

CE

PT

ED

M

radius 300 m and represented by 200 markers in the model.

Figure 2: The fire-lines showing the numerical behaviour of LSM for different pairs of dt

and dx. Each contour represents the fire-line at 380 min and the initial fire is a circle of radius 300 m.

53

ACCEPTED MANUSCRIPT

(a)

120 200 280

Y (m)

80

40 0

1000 1000

2000

3000

0 16

X (m)

3000

2000

4000

1000 1000

2000

3000

4000

X (m)

AN US

2000

Time [min]

0 40 80 120 160 180 220 240

Y (m)

240

3000

(b)

4000

CR IP T

4000

Figure 3: Evolution in time of the fire-line contour without random processes in absence of wind for the a) LSM and b) DEVS based simulators. The initial fire-line is a circle of

M

radius 300 m. The fire-break zone is 60 m wide.

(a)

ED

8000

6000

CE

2000

4000 X (m)

4000

140

6000

2000

AC

20000

20 40

0 20 40 60 80 100 120 140 160 180

Y (m)

PT

Y (m)

60

160 120 80 180

Time [min]

6000

100

4000

(b)

8000

0

2000

4000 X (m)

6000

Figure 4: Evolution in time of the fire-line contour without random processes in absence of wind with a non-homogeneous ROS for a) LSM and b) DEVS based simulators. The initial fire-line is a circle of radius 30 m.

54

ACCEPTED MANUSCRIPT

(a)

3500

(b)

3000

60

80 120

20

0 40

Y (m)

160

2500 140

100

2000

2500

2000

2500

3000 X (m)

3500

4000

1500 2000

2500

3000 X (m)

3500

4000

AN US

1500 2000

0 20 40 60 80 100 120 140 160

CR IP T

3000 Y (m)

Time [min]

3500

Figure 5: Evolution in time of the fire-line contour without random processes with an east wind of 3 ms−1 for a) LSM and b) DEVS based simulators. The initial fire-line is a circle

ED

M

of radius 300 m.

(a)

PT

3500

(b)

3000

40 60

3000

AC

2500

Y (m)

2500

120 140

2000

1500 2000

0 20 40 60 80 100 120 140 160

160

0

CE

2500

100

80

Y (m)

20

Time [min]

3500

2000

3000 X (m)

3500

4000

1500 2000

2500

3000 X (m)

3500

4000

Figure 6: Same as Fig. 5, but when the initial fire-line is a square of side 600 m.

55

ACCEPTED MANUSCRIPT

(a)

6000

40 0 20 0 1 30 7050

60

Y (m)

Y (m)

6000 4000 2000

4000

2000

2000

4000 X (m)

6000

8000

0

0

Time [min]

2000

4000 X (m)

6000

0 10 20 30 40 50 60 70

8000

AN US

00

(b)

8000

CR IP T

8000

Figure 7: Evolution in time of the fire-line contour without random processes with ROS given by formula (20) and when θ is the angle between the outward normal in a contour point and the mean wind direction for a) LSM and b) DEVS based simulators. The mean

ED

M

wind velocity is 3 ms−1 in the positive x-direction.

(a)

8000

PT

0 10 20 30 40 50 60 70

6000

Y (m)

20

0

10

Time [min]

60

4000

30

50

Y (m)

6000

(b)

8000

4000

CE

40

AC

2000

00

2000

2000

70

4000 X (m)

6000

8000

0

0

2000

4000 X (m)

6000

8000

Figure 8: Same as Fig. 7, but when θ is the angle between the line joining a contour point

and the mean wind direction.

56

ACCEPTED MANUSCRIPT

Time [min]

4000

4000

3000

2000

2000

1000

1000

0

1000

2000 3000 X (m)

4000

5000

0

0

1000

2000 3000 X (m)

AN US

0

Time [min]

0 20 40 60 80 100 120 140

Y (m)

Y (m)

3000

(b)

5000

0 20 40 60 80 100 120 140

CR IP T

(a)

5000

4000

5000

Figure 9: Evolution in time of the fire-line contour with turbulence in absence of wind for a) LSM and b) DEVS based simulators. The initial fire-line is a circle of radius 300 m.

ED

M

The turbulent diffusion coefficient is D = 0.15 m2 s−1 .

1

AC

CE

?, ?e

PT

0.8

0.6

0.4 ? - DEVS ? e - DEVS ? - LSM ? e - LSM

0.2

0

0

1000

2000

3000

4000

5000

X (m)

Figure 10: Cross section of the indicator function φ(x, t) and φe (x, t) at y = 2500 m and

t = 140 min, corresponding to the front evolution showed in Fig. 9.

57

ACCEPTED MANUSCRIPT

Time [min]

4000

4000

3000

2000

2000

1000

1000

0

1000

2000 3000 X (m)

4000

5000

0

0

1000

2000 3000 X (m)

AN US

0

Time [min]

0 20 40 60 80 100 120 140

Y (m)

Y (m)

3000

(b)

5000

0 20 40 60 80 100 120 140

CR IP T

(a)

5000

4000

5000

0 20 40 60 80 100 120 140 160

ED

4000

3000

PT

Y (m)

Time [min]

CE

2000

0

1000

2000 3000 X (m)

Time [min]

0 20 40 60 80 100 120 140 160

4000

3000

2000

4000

1000

5000

AC

1000

(b)

5000

Y (m)

(a)

5000

M

Figure 11: Same as Fig. 9, but with turbulent diffusion coefficient D = 0.30 m2 s−1 .

0

1000

2000 3000 X (m)

4000

5000

Figure 12: Evolution in time of the fire-line contour with turbulence and a north wind of 3 ms−1 for a) LSM and b) DEVS based simulators. The turbulent diffusion coefficient is D = 0.15 m2 s−1 .

58

ACCEPTED MANUSCRIPT

Time [min]

Y (m)

4000

3000

3000

2000

1000

2000 3000 X (m)

4000

1000

5000

0

1000

2000 3000 X (m)

4000

5000

AN US

0

Time [min]

0 20 40 60 80 100 120 140 160

4000

2000

1000

(b)

5000

Y (m)

0 20 40 60 80 100 120 140 160

CR IP T

(a)

5000

Figure 13: Same as Fig. 9, but when both turbulence and fire-spotting are included with

M

D = 0.15 m2 s−1 , µ = 2.69 and σ = 1.25.

ED

PT

Y (m)

4000

3000

Time [min]

0 20 40 60 80 100 120 140 160 180 200 220

CE

2000

0

1000

2000 3000 X (m)

Time [min]

0 20 40 60 80 100 120 140 160 180 200 220

4000

3000

2000

4000

1000

5000

AC

1000

(b)

5000

Y (m)

(a)

5000

0

1000

2000 3000 X (m)

4000

5000

Figure 14: Evolution in time of the fire-line contour with turbulence, a north wind of 3 ms−1 and two fire-break zones for a) LSM and b) DEVS based simulators. The initial fire-line is a circle of radius 300 m. The turbulent diffusion coefficient is D = 0.075 m2 s−1 .

59

AN US

CR IP T

ACCEPTED MANUSCRIPT

(a)

Time [min]

0 20 40 60 80 100 120 140 160 180 200 220

3000

2000

0

1000

2000 3000 X (m)

4000

5000

ED

1000

M

Y (m)

4000

5000

(b)

Time [min]

0 20 40 60 80 100 120 140 160 180 200 220

4000

Y (m)

5000

3000

2000

1000

0

1000

2000 3000 X (m)

4000

5000

Figure 15: Same as Fig. 14, but when both turbulence and fire-spotting are included with

AC

CE

PT

D = 0.075 m2 s−1 , µ = 2.69 and σ = 1.25.

60

ACCEPTED MANUSCRIPT

10000

10000

8000

8000 150

00

2000

4000 6000 X (m)

00

8000 10000

(a)

2000

25

0

2000

4000 6000 X (m)

00

8000 10000

ED

00

2000

(c)

(d)

00

2000

100

0

4000 6000 X (m)

(e)

150

4000 2000

0

00

8000 10000

5

6000

17

Y (m)

175

125

150

PT

8000

0 10 25 50

2000

8000 10000

10000

75 25 50

CE

Y (m)

8000

4000 6000 X (m)

75 125

10000

175

0

2000

M

2000

125

4000

0 10 25 50

50

75

6000

1 75 50

75

125

Y (m)

100

4000

AC

8000 10000

8000

6000

4000

4000 6000 X (m)

AN US

10000

8000

6000

150

0

(b)

10000

Y (m)

50 100

2000

CR IP T

Y (m)

125

50 100 0

2000

4000

75 25

75 25

4000

6000 125

Y (m)

6000

2000

4000 6000 X (m)

8000 10000

(f)

Figure 16: Series of figures showing the variation in the fire propagation due to increasing value of µ. From a-f, µ varies from 13 to 18, in increments of 1, while σ = 5.

61

ACCEPTED MANUSCRIPT

10000

10000

8000

8000

50

25

2000

4000 6000 X (m)

0

2000 00

8000 10000

2000

(a) 8000

25

25

50

0

2000

4000 6000 X (m)

00

8000 10000

2000

4000 6000 X (m)

425

565

150

PT

Y (m)

4000

360

360 Y (m)

CE

5

100

X (m)

32

250 50

AC

8000

0

0

4000

475 400

0

50

530 50 4

35 0

12000

200

2000

50

16000

350

10050

250

00

300 200 1

5

32

4000

20000

350

16000

8000

(d)

ED

(c)

20000

8000 10000

300

00

0

2000

M

2000

5 12 71500

4000

50

75

6000

150

75

125

Y (m)

100

4000

12000

0

8000

15

AN US

10000

6000

8000 10000

(b)

10000

Y (m)

4000 6000 X (m)

375

00

25

50 100

4000

0

2000

100

5

75

25

75

CR IP T

4000

150 1

6000

Y (m)

200 150 12

175

Y (m)

6000

6000

8000

00

10000

(e)

4000

8000

12000 X (m)

16000

20000

(f)

Figure 17: Series of figures showing the variation in the fire propagation due to increasing value of σ. From a-f, σ varies from 4 to 6.5, in increments of 0.5, while µ = 15.

62

CR IP T

7000

Y (m)

5000

75

5000

75

5000

AN US

7000 Y (m)

7000

9000

100

9000

95

9000

75

Y (m)

ACCEPTED MANUSCRIPT

95

M

ED

1000 2000

4000 6000 X (m)

(b)

8000

1000 2000

25 75

25 75

25 75

(a)

8000

0

4000 6000 X (m)

50

1000 2000

3000

0

12

0

50

3000

50

3000

0

4000 6000 X (m)

8000

(c)

Figure 18: Series of figures showing the generation of secondary fires due to fire-spotting

AC

CE

PT

and their subsequent merging with the primary fire when µ = 15 and σ = 5.

63