Next Article in Journal
Urban Comprehensive Carrying Capacity and Urbanization in Northeast China
Previous Article in Journal
Study on Carbon Emission Measurement and Influencing Factors for Prefabricated Buildings at the Materialization Stage Based on LCA
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modified Quasi-Physical Grassland Fire Spread Model: Sensitivity Analysis

by
Esmaeil Mohammadian Bishe
,
Hossein Afshin
* and
Bijan Farhanieh
School of Mechanical Engineering, Sharif University of Technology, Azadi Ave., Tehran 1458889694, Iran
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(18), 13639; https://doi.org/10.3390/su151813639
Submission received: 29 June 2023 / Revised: 17 August 2023 / Accepted: 7 September 2023 / Published: 12 September 2023

Abstract

:
Developing models for predicting the rate of fire spread (ROS) in nature and analyzing the sensitivity of these models to environmental parameters are of great importance for fire study and management activities. A comprehensive sensitivity analysis of a general and modified quasi-physical model is provided in the current study to predict parameters that affect grassland fire propagation patterns. The model considers radiative heat transfer from the flame and fuel body and convective heat transfer to predict the fire’s rate of spread and the grassland fire patterns. The model’s sensitivity to ten main parameters that affect fire propagation, including temperature, humidity, wind speed, specifications of vegetable fuel, etc., is studied, and the results are discussed and analyzed. The model’s capability is validated with experimental studies and a comprehensive physical model WFDS. The model’s capability, as quasi-physical, faster than the real-time model, shows high consistency in fire propagation parameters compared with experimental real data from the Australian grassland fire Cases C064 and F19. The comprehensive sensitivity analysis provided in this study resulted in a modified equation for the corrected rate of fire spread which shows quite an improvement in ROS prediction from 5% to 65% compared with the experimental results. The study could be a base model for future studies, especially for those researchers who aim to design experiments and numerical studies for grassland fire spread behavior.

1. Introduction

Wildfires are a growing concern for sustainable development [1]. Records indicate that in recent times, the number of wildfires has increased due to population growth, land use change, and climate change. Part of this increase is likely due to improved detection and reporting of fires [2,3,4]. As wildfires increase, integrated strategies for forests, climate, and sustainability modeling become more essential. With this concern, studies should be conducted to link the wildfire effects to sustainable development goals [5,6]. Unfortunately, climate change in the Middle East has affected the growing trend of forests fires and grassland fires which consequently destroy the vegetal cover for a couple of years, affect plant diversity, and promote soil erosion in the forest ecosystems [7,8,9,10,11]. In recent years, there have been frequent fires in the Hur-al-Aazim Wetland in the southern part of the Iran–Iraq border and the Anzali wetland located on the Caspian Sea.
Analyzing forest fire behavior is very important because of all the environmental repercussions [12]. The development and application of fire models has become a priority for many countries as well as international unions of firefighters and environmental organizations. The purpose of modeling fire behavior in grasslands and pastures is to understand and predict possible fire behaviors [13,14,15]. Wildland fire behavior modeling has been active since the 1920s when Hawley presented theoretical considerations regarding factors that influence forest fires [16], and Gisborne presented the objectives of forest fire–weather research and complicated controls of fire behavior [17,18].
Fire behavior models fall into three broad categories based on the way they are constructed: physical and quasi-physical models, empirical and quasi-empirical models, and simulation and mathematical analogue models [19]. Physical models represent both the physics and chemistry of fire spread, such as the WFDS [2], FIRETEC [20], FIRESTAR [21], Weber [22], and Grishin [23]; but combustion chemistry is not directly considered in quasi-physical models, such as ADFA I and II [24], Albini [25], and UoC-B [26]. The physics of fire dynamics is not considered at all in empirical models, such as CSIRO Grass [27], CSIRO Forest [28], and Helsinki [29], while quasi-empirical models, such as TRW [30], NBRU [31], and Nelson [32], do use some form of physical concepts. Simulation studies have also been conducted on the fire spread in wildland–urban interface fires [33,34]. In these studies, various physical, computational, or statistical methods have been used [35]. The WFDS model, known as the 3D physical model, is used for various purposes, including numerical modeling of fire spread through grasslands and shrubs [36] or modeling fire behavior in areas with multiple tree arrangements [37]. A newer version of the WFDS, called WFDS svn9977, based on the FDS6 has been used frequently in recent studies [38,39,40].
Simulation and mathematical analogue models implement existing fire behavior models, such as Firemap [41], SiroFire [42], FARSITE [43], and FlamMap [44], or utilize a mathematical precept rather than a physical one to model the spread of wildland fires [45,46]. In contrast, physical models use mathematical relationships generated from the governing fluid dynamics equations, combustion chemistry, and heat transfer mechanisms including convection and radiation using fuel characteristics implemented on terrain topography and weather conditions [47]. The main features that differentiate an operational model from a physical model is the lack of combustion chemistry and a reliance on the heat flame geometry and flame temperature [19].
Knowing the different categories of wildfire behavior prediction models and the features, advantages, and limitations of each of these categories is the first step in fire behavior modeling and studying towards the improvement of these models. The primary heat transfer mode used in wildfire behavior modeling significantly impacts the model approach [48]. Wildfire propagation is generally driven by radiation and convection; in some studies, radiation is considered the dominant heat transfer mechanism and the convection heat transfer mechanism is less discussed [25]. Radiative models were initially proposed to describe the uniform propagation of a planar fire front. These models assume there is no ambient wind, and expressions are obtained for the forward rate of the flame front [24]. Despite full-radiative wildfire models methods, some authors suggest that convection terms of heat transfer should be considered when calculating fire spread based on flame dynamics in specific fuel beds [26,47,49]. Since convective heat transfer has the greatest impact on fire propagation under windy conditions, high slope, and low fuel loads [30,50], some studies suggest considering both convective heating and cooling in calculations to improve the accuracy of results [51]. For low fuel load, fire can be in boundary layer mode even when the driving wind is not that strong [52]. Predicting the thermal radiation emitted by a flame also plays an essential role in wildfire modeling [53]. The Solid Flame Model (SFM) assumes that the flame absorbs and emits radiation as a surface. The Volume Flame Models (VFMs) refer to models where radiation is emitted from inside the actual flame volume. Studies have shown that both these models provide reasonable results in wildfire modeling, which is a wide-ranging study category [54]. In solid flame models, each fuel element’s emissivity, absorptivity, and configuration factor are parameterized to calculate thermal radiation [55]. Two-dimensional quasi-physical fire propagation models are suitable for faster than real-time operational use [50,56,57].
Considering the listed importance of physical and quasi-physical models of fire behavior prediction, a series of quasi-physical wildfire modeling studies have been mentioned in this section. Koo et al. proposed a modified simple physical model and tested it in a series of laboratory and field experiments at the University of California, Berkeley [26]. Morvan et al. proposed a 3D physical model to study the behavior of vegetation fires at a laboratory scale, concentrating on a detailed physical fire model of forest fuels [58]. Studies also showed that increasing ambient wind speed significantly increases the forward spread of the simulated fires [59,60]. A comparison with the wind tunnel experiment showed that the physical characteristics of the model successfully evaluate the influence of wind and slope on the fire spread rate. Fuel bed slope is another parameter that must be considered in fire spread studies [47,61]. A summary of the main quasi-physical models published in the literature between 1990 and 2020 is presented in Table 1.
To evaluate and improve fire behavior modeling, there is a need for new methods such as genetic algorithms, which have been considered in recent years [63,64]. Machine learning has also helped experts to increase the accuracy of the wildfire simulation while reducing the speed of calculations [65]. Faster prediction gives the fire management systems a better tool to prevent fire damage [66]. Recent models try to reduce the calculation time for faster fire prediction by reducing the number of model parameters [66,67]. For this purpose, studies have focused on the sensitivity analysis of parameters in recent years [68,69,70].
The current study proposes a modified quasi-physical fire spread model to predict fire propagation patterns in grasslands that considers radiative heat transfer from the flame and fuel body and convective heat transfer. This model considers the characteristics of the vegetative fuel bed in three geometric dimensions. The proposed modified quasi-physical model objectives are model simplicity, adaptability to the Earth’s topographic conditions, and designed to be faster than real-time for operational purposes. The model can also be generalized to various plant fuels. By carrying out a comprehensive sensitivity analysis on all the parameters affecting the fire spread rate and improving the physics of the quasi-physical models developed, especially in the convective heat transfer section, the current research has improved the prediction results of the fire spread rate and the shape and pattern of the fire propagation. To the authors’ knowledge, such a comprehensive sensitivity analysis has not been provided on wildfire quasi-physical modeling in previous studies.

2. Materials and Methods

The proposed model consists of a typical two-dimensional network with circular cells of equal size. Figure 1 presents the schematic of the model network and the relationship between cells. The network of combustible cells, or in other words, the pasture bed that is being modeled, can be created using a schematic of the model network to present a vegetation map. It should be noted that burning cells represent the vegetation pasture in this model. The model was applied to a set of reference fires conducted in controlled conditions (Australian grassland experiments) and cases modeled by WFDS [2].

2.1. Hypotheses

The first step to develop a two-dimensional quasi-physical fire spread model is to determine the correct and pre-proven principles and assumptions. These physical or geometric assumptions can include the characteristics of plant fuels or environmental conditions. A description of these hypotheses is summarized in the following paragraphs:
  • Each combustible cell (j) is defined as a cylinder of height (Hj) and diameter (Dj) (Figure 2).
  • The combustible cell is considered a healthy cell as long as the cell temperature (Tj) is equal to the ambient temperature (T∞).
  • The energy absorbed by the flammable cell when exposed to the fire front is used to increase the temperature of the wet fine fuel cells to the boiling point of water (373 K).
  • By evaporating the cell moisture (humidity as fuel) temperature of the fine fuel cells, the temperature of the dry cell is raised to combustion temperature.
  • The combustible cell then enters the flame combustion phase for the duration of tc, the flame residence time, while simultaneously transferring heat to adjacent combustible cells by radiation and convection mechanisms.
  • In the Solid Flame Model, the visible flame is considered a solid body with uniform radiation in the form of a cylinder from the surface of the emitted thermal radiation.

2.2. Equations

Combustion cell temperature is calculated using the energy conservation equation. The energy received by the flammable cell moves toward the target cell through the two mechanisms of displacement and radiation. The most important heat dissipation mechanism in a flammable cell is radiation [57]. The total heat transfer, qt, can be calculated as follows:
q t = q c o n v + + q r a d + + q r a d
where qt is the total heat transfer, qconv is the convection heat transfer, qrad+ is the radiation to cell j, and qrad is the radiation losses of cell j. Equation (1) is extended in Equation (2) to obtain the total heat transfer emitted from the burning cell. Then, the radiation and convection parameters are added in Equations (3) and (4).
q i j = q i j s r + q i j i r + q i j r l + q i j s c + q i j i c
q r a d + + q r a d = q i j s r + q i j i r + q i j r l
q c o n v + = q i j s c + q i j i c
where q i j s r is the surface radiation, q i j i r is the internal radiation, q i j r l is the radiative loss from the combustible cell (the radiation parameters), and q i j s c is the surface convection and q i j i c is the internal convection (the convection parameters). Figure 3 is a schematic of how heat is transferred from the flammable cell to the target cell through the two mechanisms of displacement and radiation [26,47].

2.2.1. Radiation

In the process of fire spread, radiant heat transfer plays a major role in heat transfer mechanisms and fire spread [71]. It is assumed that flammable cell (j) is at the dij distance from the burning cell (i). The total heat transfer emitted from the burning cell (i) received by the flammable cell (j) is denoted by qij, which is calculated in (2). The right-hand side of (2) is a summation of all possible heat transfer mechanisms, including radiation from the upper surface of cell j, internal radiation from the ember zone of cell j, convection above the cell j, convection above the head of cell j, internal convection inside the fuel bed, and radiation losses at the upper surface of cell j.
Since the air in grasslands is often unstable and the wind speed is non-zero, other preheating heat transfer mechanisms, i.e., turbulent diffusion and phase conduction, convective cooling, and energy absorbed during pyrolysis before combustible cell ignition point, can be ignored [50]. In modeling, only thin, thermally delicate plants burn (such as grasses, thin branches, and the leaves of trees and shrubs with a thickness of less than 6 mm), and only this type of burning determines the fire spread rate [72]. Other fuels, such as tree trunks, burn more slowly and therefore play no role in spreading the fire [72]. All the parameters of the above equations are explained in order below. Radiation to the upper surface of the fuel bed received from the flame to cell j is calculated from (5).
q i j s r = a f b ε f l σ T f l 4 H j F i j
ε f l = 1 e x p ( 0.6 L f l )
where εfl is the flame emissivity, afb is the fuel bed absorptivity, σ is the Stefan–Boltzmann constant (5.67 × 10−8 W/m2 K), and Fij is the view factor or visibility [26].
The visible flame is considered to have a uniform surface of radiation with cylindrical geometry (Figure 2). Fij visibility is the fraction of radiation absorbed by cell j due to radiation from cell i which is explained in Appendix A. The View Factor or Configuration Factor is comprehensively described in Henkinson’s study [73,74]. This method is similar to the Monte Carlo method used in Ansys Fluent software version 14.0 or upper. The surface-to-surface (SOS) radiation model calculates the net heat flows for given temperatures of the plane, gray, diffuse surfaces [75]. Analytical expressions to calculate this view factor were introduced in various studies. In a porous fuel bed, unburned cells receive radiant heat flux from the fuel bed environment and the flame front. This internal radiation from cell i to cell j at the distance dij is obtained from (7) [26,50].
q i j i r = 0.25   A f b ε b σ T b 4 exp ( 0.25 A f b d i j )
where εb is the ember emissivity equal to 1, Tb is the ember temperature equivalent to ignition temperature, and Afb is the total fuel particle surface area per fuel bed volume. Correspondingly, the heat loss of unburned cells due to radiation to the environment above the fuel bed surface is calculated from (8). The expression q i j r l is equivalent to qrad in (1).
q i j r l = ε f b σ ( T j 4 T 4 ) H j
where ε f b , the diffusion coefficient of the (8) is related to the fuel bed. Further explanation about flame height based on Heskestad equations is presented in Appendix B [76].

2.2.2. Convection

In fire spread modeling, the heat transfer calculations only consider the forced heat transfer due to ambient wind. The heat transfer over the surface of cell j is calculated from the following equation [26,77]:
q i j s c = 0.565 k f l R e d i j 1 2 P r 1 2 d i j H j ( T f l T j ) e x p ( 0.3 d i j / L f l ) β i j
q i j i c = 0.911 k b A f b R e D 0.385 P r 1 3 D j ( T b T j ) e x p ( 0.25 A f b d i j ) β i j
where Pr is the Prandtl number, Redij stands for the Reynolds number, Uw is the ambient wind speed, and Kfl is the thermal conductivity. The heat transfer coefficient for a cylinder in a cross-flow used for the interior of the fuel bed is derived from (9) [26].
R e D j = U f b D j ϑ g
U f b = ( 1 j ) U w
β i j = 1       i j   U   0       i j     U
The βij coefficient shows that the internal convection will be considered in calculations only if the connecting vector from cell burning cell i to combustible cell j is aligned with the wind vector.

2.3. Methodology

The ordinary differential equation system must be solved before the ignition point. The system of equations is as follows:
f o r   0 t t c d T j d t = 1 ρ j C P j j   l = s r   o   r l q i j l                       f o r   T j 373   K d W j d t = 1 ρ j h v a p j   l = s r   o   r l q i j l         f o r   T j = 373   K
where Tj(0) is equal to the ambient temperature (T∞) and Wj(0) is the initial water mass fraction. Each cell has its interactive range using the current method and considering the preheating mechanisms [26].
All the coefficients used in the above equations in each simulation and in accordance with the physical conditions of the environment to be simulated are presented in separate tables. Simulation results of the WFDS model conducted by Mell et al., the developers of WFDS, are used to compare the results of a two-dimensional simulation of the proposed model [1]. The conditions used for the simulation using the WFDS model are listed in Table 2.

3. Results and Discussion

In the current study, the fire spread model was validated with two scenarios: grassland fire experiments conducted in the Northern Territory of Australia, F19, and a complete 3D physical wildland fire dynamics simulator called WFDS. A summary of the main parameters used in the modeling of the current study is listed in Table 3.

3.1. Model Validation

A photograph of the fire perimeter taken during the Australian experiment F19 is shown in Figure 4. The time step had a significant effect on the accuracy of the results of the numerical calculations in the present model. Although reducing the time step increases the accuracy of the calculations, it leads to an increase in computational time and cost, which contradicts the initial goal of developing a simple model. A series of calculations were performed under the same physical conditions for six time steps ranging from 2.5 s to 0.1 s to determine the optimal time step, see Figure 5. The number of burned cells and the average temperature of 3600 cells in the simulation domain were chosen as a base for proper time step selection. The fire spreads in a 200 × 200 m grassland plot with the fuel properties of experiment F19 shown in Figure 6. One of the most important reasons for the oscillations in the fire head location is the interactions of the fire plume and the local wind field, which is visible in Figure 6.
According to calculation results, a time step of 0.25 s is generally used in simulations based on network conditions. Results show a reasonable agreement between the experimental fire propagation line and the 3D physical simulations of the WFDS Model (Figure 6). Comparative results of fire spread rate and fire spread perimeter are included in Table 4. A 2.5% to 19.7% difference in fire perimeter prediction and a 7.8% to 20.7% difference in the spread prediction rate seems reasonable for a faster-than-real-time prediction model. It can significantly help fire management and firefighting organizations with any preventative measures. Therefore, the results show the current model’s remarkable accuracy compared to the WFDS model’s validated results.
Another excellent indicator of fire behavior is the average width of the fire front is the fire front head width. The head fire width can be directly determined from the model results. The product of the rate of spread and the flame residence time can also calculate this characteristic length.
Where dhead is the head fire front average width, ROS is the rate of fire spread, and tc is the fire residence time. The value of head fire front average width is calculated by the product of ROS at fire front by tc in a certain period.
To strengthen the validity of the model results, achieve more accurate evaluation, and improve the quality of the physical model, validation of the model with experimental data was carried out in the case of studies C064 and F19. For case C 064, the rate of fire spread and the shape of the fire front is drawn for three time periods of 56, 86, and 138 s. Similarly, for the case of F 19, the mentioned parameters are fixed for the period of 27, 53, and 100 s (Figure 7).

3.2. Modified Model

It is necessary to search for the existing shortcomings regarding the prediction of heat transfer parameters to improve the results of the general quasi-physical model of fire spread in wildfires. Recent studies on the role of buoyant turbulent flow in the spread of wildfires show that buoyant instabilities lead to the formation of transverse waves toward the center of the fire [78]. As shown in experimental studies, especially those mentioned by Finney et al., streamwise vortex flow instabilities push convective heat transfer through the firing domain [78]. Therefore, unlike previous wildfire spread quasi-physical models, in addition to the cells in the main wind direction, the convective heat transfer term is considered for the induced wind gradient created towards the center of the fire front.
As shown in Figure 7, the model modification leads to a significant improvement in predicting the fire spread contour and fire spread rate. Table 5 shows that the model is significantly improved in predicting the fire front line and fire rate of spread by modifying the model’s behavior in predicting convective heat transfer. The considerable reduction of the model error in predicting fire spread for two real and famous case studies, C064 and F19, is listed in Table 5.
Figure 8 compares the results of the developed semi-physical model and the results of other models and extracted relationships. As shown in Figure 8, the result of the prediction of the fire spread rate according to the wind speed for the present model compared to the mathematical model of the Rothermel Model [79], McArthurs’ Model [80], and other prediction equations [27,81] for two cases, C064 and F19, are shown.
Results show that the predicted rate of fire spread in the modified model improves from 5% to 65% versus the previous general model, both in comparison with experimental results.

3.3. Sensitivity Analysis

This section presents the sensitivity analysis of the model to each of the parameters affecting the fire spread rate (ROS). Difficulties in measuring these parameters produce uncertainties in the model results. Therefore, the model’s predictive capability depends on the accuracy of determining these parameters. The range of parameter variations was chosen to cover typical vegetable and meteorological conditions, see Table 6.
The sensitivity analysis makes it possible to estimate responses as functions of these parameters and to make proper correlations between the parameter effects and the system response. The range of changes selected for the parameters of environmental conditions and the physical characteristics of vegetable fuel mentioned in Table 6 are chosen based on the typical range of changes of these parameters in nature. The rate of spread can then be corrected with the following equation:
R O S c = R O S 0 + i = 1 10 α X i + i , j , ( i n t e r a c t i o n s   f r o m   o r d e r   2   t o   o r d e r   10 )
where ROSc and ROS0 are the corrected and predicted rate of spread, respectively, α is the parameter sensitivity, and Xi is the result of each of the parameters analyzed. High orders of parameter interaction are not noticeable; therefore, only low order parameter interactions are considered [70,84]. By omitting the interaction of parameters in Equation (12), Equation (13) is obtained based on the sensitivity analysis of parameters in Table 6:
R O S c = R O S 0 1.225 H f 5 + 0.291 L f 5 + 0.0033 T f l 1083 0.292 T i g n 500 + 0.0097 T 307 + 0.083 U w 5 7.57 W 0 0.058 0.00025 c p 1110 3.7 T 0.0029 ρ s 512 1248 ( 0.0012 )
Figure 9 shows the fire propagation contour for case F19 with 8 m and 25 m fire ignition lines at two moments (t = 60 s and t = 120 s). The comparative chart of the relative sensitivity analysis of 10 main parameters to the rate of fire spread (ROS) is also presented in Figure 10.
As shown in Figure 10, the flame temperature has the most significant effect on the rate of fire spread, and increasing the ignition temperature has the most effective relative impact on reducing the rate of fire spread.
Another essential component in evaluating fire propagation models is studying the heat transfer rate for each radiative heat transfer and displacement mechanism and its effect on fire spread. In this regard, Figure 11 depicts a diagram of heat transfer mechanisms from the ignition moment to the ignition moment across with a temperature diagram for the same point on the simulation environment.
As shown in Figure 11, the contribution of internal radiation and internal convection to the fire moment is minimal compared to other heat transfer mechanisms. Therefore, to better separate the charts, these two mechanisms have been removed in Figure 12, and the time interval is limited to 10 s when the fire front approaches the desired point.
The effect of wind speed on the rate of fire spread in three fuel bed packing ratios was investigated, and the results are shown in Figure 13. The ratio of initial moisture content of vegetable fuel on the rate of fire spread at three wind speeds of 2, 5, and 10 m/s was also investigated. Figure 14 shows that the ROS is essentially much slower than dry vegetable fuels for very high fuel moisture content. The range of plant fuel moisture in pasture fires is usually below 0.45.

4. Conclusions

In this study, a comprehensive modified quasi-physical model was developed to predict fire propagation patterns on gentle terrain. Preheating energy transfer mechanisms, including surface radiation from the flame, internal radiation from the ember part, radiative loss, surface convection from the top of the flame, internal convection, and cooling convection from the swamp or wet soil, are considered in the model to predict the fire contour and rate of fire spread on grassland and pastures. The model also includes the effects of vegetation specification, topography and terrain characteristics, and climatic parameters. The proposed model was validated with the WFDS model, and the results showed that it makes the model valuable compared to other physical models. In this model, vegetation is considered a network of combustible cells on a macroscopic fuel map. To make it suitable for fire management and support systems, the model development pattern is considered such that it provides considered results faster than real-time computation period.
In the sensitivity analysis, the most important section of this study, the model’s sensitivity to ambient conditions and vegetative fuel properties that affect fire propagation was explained for academic considerations. Ambient temperature, ambient humidity, and wind speed were considered the most effective ambient conditions. Moreover, the sensitivity of vegetable fuel properties, including specific heat, vegetable height, packing ratio, initial moisture of vegetative fuel, flame temperature, and flame length, were studied separately. Results show that the flame temperature has the most positive, and the ignition temperature has the most negative effect on the rate of fire spread (ROS).
In summary, a significant share of heat transfer occurs around the surface radiation, surface convection, and radiation losses. Therefore, internal radiation and convection could be ignored to reduce the volume of the simulations and increase the computational speed. Based on the analytical results obtained, it can be concluded that before 373 K (the boiling point of water and evaporation of all the moisture in the vegetative fuel), the contribution of the surface convection, and after this temperature, the contribution of the surface radiation to the total heat transfer is the largest among all heat transfer mechanisms.
This sensitivity analysis leads to a general equation for the rate of fire spread which is based on 10 main effecting parameters. Results show that the predicting rate of fire spread in modified model improved from 5% to 65% versus the previous general model, both in comparison with experimental results.

Author Contributions

Writing—original draft, E.M.B.; Supervision, H.A.; Project administration, B.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

The authors would like to acknowledge the assistance of Mohammad Abdolmaleki, Ali Mostafazadeh, and Nategheh Najafpour at the Sharif University of Technology, for their contributions and discussions about the topic.

Conflicts of Interest

Authors hereby declared that they do not have any conflicts of interest. All authors have participated in (a) conception and design, or analysis and interpretation of the data; (b) drafting the article or revising it critically for important intellectual content; and (c) approval of the final version. This manuscript has not been submitted to, nor is under review at, another journal or other publishing venue. The authors have no affiliation with any organization with a direct or indirect financial interest in the subject matter discussed in the manuscript. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Nomenclature

TTemperature
tTime
tcFlame Residence Time
QHeat Release Rate
HCell Height
DCell Diameter
cpSpecific Heat Capacity
LvapEnthalpy of Vaporization
VFMVolume Flame Model
SFMSolid Flame Model
FMCFuel Moisture Content
ROSRate of Fire Spread
GSAGlobal Sensitivity Analysis
WFDSWildland-Urban Interface Fire Dynamics Simulator
qconvConvection Heat Transfer
qrad+Radiation to Combustible Cell
qradRadiation Losses of Combustible Cell
εEmissivity
aAbsorptivity
σStephan Boltzmann constant
FView factor
PrPrandtl Number
ReReynolds Number
UwAmbient Wind Speed
KThermal Conductivity
dijDistance from the burning cell i to combustible cell j
vKinematic Viscosity
Subscripts:
jindex of combustible cell
iindex of burning cell
ignignition
fflame
fbfuel bed
ambient
ccombustion
srsurface radiation
irinternal radiation
rlradiative loss
scsurface convection
icinternal convection

Appendix A. View Factor

Figure A1. View factor schematic.
Figure A1. View factor schematic.
Sustainability 15 13639 g0a1
This section presents how to calculate the coefficient of visibility based on Hankinson relationships [73]. The radiant surface of the flame is divided into small parallelograms. The coefficient of visibility between these parallelogram surfaces of AP and these infinitesimally small cells is obtained from Equation (A1).
F i j = A l l   E l e m e n t s W h e r e cos θ 1 > 0 a n d cos θ 2 > 0 cos θ 1 cos θ 2 A P π r 2
To calculate the visibility factor, vectors must be defined in Cartesian coordinates. In the following equations, the coordinates of the centers of cell O and the center of the radiant surface P are considered.
r = r x , r y , r z = [ X o X p , Y o Y p , Z o Z p ]
r = r = r x 2 + r y 2 + r z 2 1 2
To determine the θ1 and θ2, unit vectors must be calculated first.
r ^ o p = [ X o X p r , Y o Y p r , Z o Z p r ]
r ^ p o = [ X p X o r , Y p Y o r , Z p Z o r ]
In addition, the normal vectors of cell O and surface P are calculated as follows. Due to the nature of networking in the present model and the use of the solid flame model with a cylindrical shape, this normal vector of surface P is shown schematically in Figure A1.
n ^ o = [ n o x , n o y , n o z ]
n ^ p = [ n p x , n p y , n p z ]
Therefore, cos θ1 and cos θ2 are calculated from the following equations:
cos θ 1 = n ^ o . r ^ p o = X p X o r 0 + Y p Y o r 0 + Z p Z o r 1 = Z p Z o r
cos θ 2 = n ^ p . r ^ o p = X o X p r c o s θ + Y o Y p r sin θ + Z o Z p r 0 = X o X p c o s θ + Y o Y p s i n ( θ ) r
When the angles θ1 and θ2 are greater than 90 degrees with the target cell O, or cosθ1 and cosθ2 are less than zero, these elements are out of sight and radiant heat transfer does not occur between the two elements.
Xo = Xj
Yo = Yj
Zo = Zj
Xp = Xi + R × cos(θ)
Yp = Yi + R × sin(θ)
Zp = Zi
Figure A2. Normal vector.
Figure A2. Normal vector.
Sustainability 15 13639 g0a2

Appendix B. Flame Properties

The details of the flame properties have been separated from the method section and placed separately in the appendix [76]. Based on the influence of wind, the luminous flame height under the influence of wind is estimated from Equation (A10):
H f = H f 0 ( 1 + 4 U 2 g H f 0 )
Hf0 is the luminous flame height without wind given by Heskestad correlation.
H f 0 = 0.0148 Q ˙ 0.4 1.02 D
where:
D: Burning Cell Diameter (m)
Q ˙ : Heat Release Rate (W)
The heat release rate of flame combustion corresponds to the chemical reaction of volatiles with air. Therefore, the associated heat release rate, Q ˙ , may be expressed in terms of mass loss rate [57]:
Q ˙ = m D F h c π D 2 4 t c
where:
m D F : initial mass of dry fine fuel per unit area
h c : heat of combustion of volatiles
tc: residence time.

References

  1. Lang, Y.; Moeini-Meybodi, H. Wildfires—A Growing Concern for Sustainable Development. UN Department of Economic and Social Affairs (DESA) Policy Briefs, 22 October 2021; 4. [Google Scholar]
  2. Mell, W.; Jenkins, M.A.; Gould, J.; Cheney, P. A Physics-Based Approach to Modelling Grassland Fires. Int. J. Wildland Fire 2007, 16, 1–22. [Google Scholar] [CrossRef]
  3. Kalabokidis, K.; Palaiologou, P.; Gerasopoulos, E.; Giannakopoulos, C.; Kostopoulou, E.; Zerefos, C. Effect of Climate Change Projections on Forest Fire Behavior and Values-at-Risk in Southwestern Greece. Forests 2015, 6, 2214–2240. [Google Scholar] [CrossRef]
  4. Lan, Z.; Su, Z.; Guo, M.; Alvarado, E.C.; Guo, F.; Hu, H.; Wang, G. Are Climate Factors Driving the Contemporary Wildfire Occurrence in China? Forests 2021, 12, 392. [Google Scholar] [CrossRef]
  5. Martin, D.A. Linking Fire and the United Nations Sustainable Development Goals. Sci. Total Environ. 2019, 662, 547–558. [Google Scholar] [CrossRef] [PubMed]
  6. Kline, K.L.; Dale, V.H.; Rose, E.; Tonn, B. Effects of Production of Woody Pellets in the Southeastern United States on the Sustainable Development Goals. Sustainability 2021, 13, 821. [Google Scholar] [CrossRef]
  7. Matallana, R.G.; Lucas-Borja, M.E.; Gómez-Sánchez, M.E.; Uddin, S.M.M.; Zema, D.A. Post-Fire Restoration Effectiveness Using Two Soil Preparation Techniques and Different Shrubs Species in Pine Forests of South-Eastern Spain. Ecol. Eng. 2022, 178, 106579. [Google Scholar] [CrossRef]
  8. Halofsky, J.E.; Peterson, D.L.; Harvey, B.J. Changing Wildfire, Changing Forests: The Effects of Climate Change on Fire Regimes and Vegetation in the Pacific Northwest, USA. Fire Ecol. 2020, 16, 1–26. [Google Scholar]
  9. Kraaij, T.; Baard, J.A.; Arndt, J.; Vhengani, L.; Van Wilgen, B.W. An Assessment of Climate, Weather, and Fuel Factors Influencing a Large, Destructive Wildfire in the Knysna Region, South Africa. Fire Ecol. 2018, 14, 1–12. [Google Scholar] [CrossRef]
  10. Masyagina, O.V.; Evgrafova, S.Y.; Menyailo, O.V.; Mori, S.; Koike, T.; Prokushkin, S.G. Age-Dependent Changes in Soil Respiration and Associated Parameters in Siberian Permafrost Larch Stands Affected by Wildfire. Forests 2021, 12, 107. [Google Scholar] [CrossRef]
  11. Mohammadian Bishe, E.; Norouzi, M.; Afshin, H.; Farhanieh, B. A Case Study on the Effects of Weather Conditions on Forest Fire Propagation Parameters in the Malekroud Forest in Guilan, Iran. Fire 2023, 6, 251. [Google Scholar] [CrossRef]
  12. Cave, H.; Adams, M.; Jaeger, T.; Peet, T.; Staats, L.; Garneau, D.; Lesser, M. Wildlife Response to Wildfire in a Northern New York Jack Pine Barrens. Forests 2021, 12, 676. [Google Scholar] [CrossRef]
  13. Perrakis, D.D.B.; Lanoville, R.A.; Taylor, S.W.; Hicks, D. Modeling Wildfire Spread in Mountain Pine Beetle-Affected Forest Stands, British Columbia, Canada. Fire Ecol. 2014, 10, 10–35. [Google Scholar] [CrossRef]
  14. Salis, M.; Arca, B.; Alcasena, F.; Arianoutsou, M.; Bacciu, V.; Duce, P.; Duguy, B.; Koutsias, N.; Mallinis, G.; Mitsopoulos, I. Predicting Wildfire Spread and Behaviour in Mediterranean Landscapes. Int. J. Wildland Fire 2016, 25, 1015–1032. [Google Scholar] [CrossRef]
  15. Norouzi, M.; Mohammadian Bishe, E.; Afshin, H.; Farhanieh, B.; Jahdi, R. The Numerical Investigation of Air Humidity Effects on the Fire Spread in Forests (Case Study: Malekrod-Siahkal Forest). For. Wood Prod. 2023, 76, 55–64. [Google Scholar]
  16. Hawley, L.F. Theoretical Considerations Regarding Factors Which Influence Forest Fires. J. For. 1926, 24, 756–763. [Google Scholar]
  17. Gisborne, H.T. The Objectives of Forest Fire-Weather Research. J. For. 1927, 25, 452–456. [Google Scholar]
  18. Gisborne, H.T. The Complicated Controls of Fire Behaviour. J. For. 1929, 27, 311–312. [Google Scholar]
  19. Sullivan, A.L. Wildland Surface Fire Spread Modelling, 1990–2007. 1: Physical and Quasi-Physical Models. Int. J. Wildland Fire 2009, 18, 349–368. [Google Scholar] [CrossRef]
  20. Linn, R.; Reisner, J.; Colman, J.J.; Winterkamp, J. Studying Wildfire Behavior Using FIRETEC. Int. J. Wildland Fire 2002, 11, 233–246. [Google Scholar] [CrossRef]
  21. Morvan, D.; Dupuy, J.-L.; Rigolot, E.; Valette, J. FIRESTAR: A Physically Based Model to Study Wildfire Behaviour. For. Ecol. Manag. 2006, 234, S114. [Google Scholar] [CrossRef]
  22. Weber, R.O. Modelling Fire Spread through Fuel Beds. Prog. Energy Combust. Sci. 1991, 17, 67–82. [Google Scholar] [CrossRef]
  23. Grishin, A.M.; Gruzin, A.D.; Gruzina, E.E. Aerodynamics and Heat Exchange between the Front of a Forest Fire and the Surface Layer of the Atmosphere. J. Appl. Mech. Tech. Phys. 1984, 25, 889–894. [Google Scholar] [CrossRef]
  24. De Mestre, N.J.; Catchpole, E.A.; Anderson, D.H.; Rothermel, R.C. Uniform Propagation of a Planar Fire Front without Wind. Combust. Sci. Technol. 1989, 65, 231–244. [Google Scholar] [CrossRef]
  25. Albini, F.A. Iterative Solution of the Radiation Transport Equations Governing Spread of Fire in Wildland Fuel. Combust. Explos. Shock Waves 1996, 32, 534–543. [Google Scholar] [CrossRef]
  26. Koo, E.; Pagni, P.; Stephens, S.; Huff, J.; Woycheese, J.; Weise, D.R. A Simple Physical Model for Forest Fire Spread Rate. Fire Saf. Sci. 2005, 8, 851–862. [Google Scholar] [CrossRef]
  27. Cheney, N.P.; Gould, J.S.; Catchpole, W.R. Prediction of Fire Spread in Grasslands. Int. J. Wildland Fire 1998, 8, 1–13. [Google Scholar] [CrossRef]
  28. Gould, J.S.; McCaw, W.L.; Cheney, N.P.; Ellis, P.F.; Knight, I.K.; Sullivan, A.L. Project Vesta: Fire in Dry Eucalypt Forest: Fuel Structure, Fuel Dynamics and Fire Behaviour; Csiro Publishing: Clayton, Australia, 2008; ISBN 0643102779. [Google Scholar]
  29. Tanskanen, H.; Granström, A.; Larjavaara, M.; Puttonen, P. Experimental Fire Behaviour in Managed Pinus Sylvestris and Picea Abies Stands of Finland. Int. J. Wildland Fire 2007, 16, 414–425. [Google Scholar] [CrossRef]
  30. Wolff, M.F.; Carrier, G.F.; Fendell, F.E. Wind-Aided Firespread across Arrays of Discrete Fuel Elements. II. Experiment. Combust. Sci. Technol. 1991, 77, 261–289. [Google Scholar] [CrossRef]
  31. Beer, T. The Speed of a Fire Front and Its Dependence on Wind-Speed. Int. J. Wildland Fire 1993, 3, 193–202. [Google Scholar] [CrossRef]
  32. Nelson Jr, R.M. An Effective Wind Speed for Models of Fire Spread. Int. J. Wildland Fire 2002, 11, 153–161. [Google Scholar] [CrossRef]
  33. Kganyago, M.; Shikwambana, L. Assessing Spatio-Temporal Variability of Wildfires and Their Impact on Sub-Saharan Ecosystems and Air Quality Using Multisource Remotely Sensed Data and Trend Analysis. Sustainability 2019, 11, 6811. [Google Scholar] [CrossRef]
  34. Molina, A.; Little, J.; Drury, S.; Jandt, R. Homeowner Preferences for Wildfire Risk Mitigation in the Alaskan Wildland Urban Interface. Sustainability 2021, 13, 11754. [Google Scholar] [CrossRef]
  35. Leuenberger, M.; Parente, J.; Tonini, M.; Pereira, M.G.; Kanevski, M. Wildfire Susceptibility Mapping: Deterministic vs. Stochastic Approaches. Environ. Model. Softw. 2018, 101, 194–203. [Google Scholar] [CrossRef]
  36. Mell, W.E.; Manzello, S.L.; Maranghides, A. Numerical Modeling of Fire Spread through Trees and Shrubs. For. Ecol. Manag. 2006, 234, S82. [Google Scholar] [CrossRef]
  37. Contreras, M.A.; Parsons, R.A.; Chung, W. Modeling Tree-Level Fuel Connectivity to Evaluate the Effectiveness of Thinning Treatments for Reducing Crown Fire Potential. For. Ecol. Manag. 2012, 264, 134–149. [Google Scholar] [CrossRef]
  38. Sutherland, D.; Sharples, J.J.; Moinuddin, K.A.M. The Effect of Ignition Protocol on Grassfire Development. Int. J. Wildland Fire 2019, 29, 70–80. [Google Scholar] [CrossRef]
  39. Mueller, E.V.; Skowronski, N.S.; Clark, K.L.; Gallagher, M.R.; Mell, W.E.; Simeoni, A.; Hadden, R.M. Detailed Physical Modeling of Wildland Fire Dynamics at Field Scale-An Experimentally Informed Evaluation. Fire Saf. J. 2021, 120, 103051. [Google Scholar] [CrossRef]
  40. Ex, S.A.; Ziegler, J.P.; Tinkham, W.T.; Hoffman, C.M. Long-Term Impacts of Fuel Treatment Placement with Respect to Forest Cover Type on Potential Fire Behavior across a Mountainous Landscape. Forests 2019, 10, 438. [Google Scholar] [CrossRef]
  41. Vasconcelos, M.J.; Guertin, D.P. FIREMAP-Simulation of Fire Growth with a Geographic Information System. Int. J. Wildland Fire 1992, 2, 87–96. [Google Scholar] [CrossRef]
  42. Coleman, J.R.; Sullivan, A.L. A Real-Time Computer Application for the Prediction of Fire Spread across the Australian Landscape. Simulation 1996, 67, 230–240. [Google Scholar] [CrossRef]
  43. Finney, M.A. FARSITE, Fire Area Simulator—Model Development and Evaluation; US Department of Agriculture, Forest Service, Rocky Mountain Research Station: Fort Collins, CO, USA, 1998. [Google Scholar]
  44. Frost, S.M.; Alexander, M.E.; Jenkins, M.J. The Application of Fire Behavior Modeling to Fuel Treatment Assessments at Army Garrison Camp Williams, Utah. Fire 2022, 5, 78. [Google Scholar] [CrossRef]
  45. Hargrove, W.W.; Gardner, R.H.; Turner, M.G.; Romme, W.H.; Despain, D.G. Simulating Fire Patterns in Heterogeneous Landscapes. Ecol. Model. 2000, 135, 243–263. [Google Scholar] [CrossRef]
  46. Watt, S.D.; Roberts, A.J.; Weber, R.O. Dimensional Reduction of a Bushfire Model. Math. Comput. Model. 1995, 21, 79–83. [Google Scholar] [CrossRef]
  47. Balbi, J.H.; Chatelon, F.J.; Morvan, D.; Rossi, J.L.; Marcelli, T.; Morandini, F. A Convective–Radiative Propagation Model for Wildland Fires. Int. J. Wildland Fire 2020, 29, 723–738. [Google Scholar] [CrossRef]
  48. Coen, J. Some Requirements for Simulating Wildland Fire Behavior Using Insight from Coupled Weather—Wildland Fire Models. Fire 2018, 1, 6. [Google Scholar] [CrossRef]
  49. Chatelon, F.J.; Balbi, J.H.; Morvan, D.; Rossi, J.L.; Marcelli, T. A Convective Model for Laboratory Fires with Well-Ordered Vertically-Oriented Fuel Beds. Fire Saf. J. 2017, 90, 54–61. [Google Scholar] [CrossRef]
  50. Adou, J.K.; Brou, A.D.V.; Porterie, B. Modeling Wildland Fire Propagation Using a Semi-Physical Network Model. Case Stud. Fire Saf. 2015, 4, 11–18. [Google Scholar] [CrossRef]
  51. Catchpole, W.R.; Catchpole, E.A.; Tate, A.G.; Butler, B.; Rothermel, R.C. A Model for the Steady Spread of Fire through a Homogeneous Fuel Bed. In Proceedings of the Forest Fire Research and Wildland Fire Safety: Proceedings of IV International Conference on Forest Fire Research 2002 Wildland Fire Safety Summit, Luso, Coimbra, Portugal, 18–23 November 2002; Millpress Science Publishers: Rotterdam, The Netherlands, 2002. [Google Scholar]
  52. Apte, V.B.; Bilger, R.W.; Green, A.R.; Quintiere, J.G. Wind-Aided Turbulent Flame Spread and Burning over Large-Scale Horizontal PMMA Surfaces. Combust. Flame 1991, 85, 169–184. [Google Scholar] [CrossRef]
  53. Rossi, J.-L.; Chetehouna, K.; Collin, A.; Moretti, B.; Balbi, J.-H. Simplified Flame Models and Prediction of the Thermal Radiation Emitted by a Flame Front in an Outdoor Fire. Combust. Sci. Technol. 2010, 182, 1457–1477. [Google Scholar] [CrossRef]
  54. Collin, A.; Boulet, P. Evaluation of Simple Models of Flame Radiation in the Frame of Fire Propagation. Int. J. Heat Mass Transf. 2013, 59, 83–92. [Google Scholar] [CrossRef]
  55. Santoni, P.A.; Balbi, J.H.; Dupuy, J.-L. Dynamic Modelling of Upslope Fire Growth. Int. J. Wildland Fire 1999, 9, 285–292. [Google Scholar] [CrossRef]
  56. Santoni, P.A.; Balbi, J.H. Modelling of Two-Dimensional Flame Spread across a Sloping Fuel Bed. Fire Saf. J. 1998, 31, 201–225. [Google Scholar] [CrossRef]
  57. Drissi, M. Modelling the Spreading of Large-Scale Wildland Fires. arXiv 2014, arXiv:1402.6187. [Google Scholar]
  58. Morvan, D.; Accary, G.; Meradji, S.; Frangieh, N.; Bessonov, O. A 3D Physical Model to Study the Behavior of Vegetation Fires at Laboratory Scale. Fire Saf. J. 2018, 101, 39–52. [Google Scholar] [CrossRef]
  59. Cunningham, P.; Linn, R.R. Dynamics of Fire Spread in Grasslands: Numerical Simulations with a Physics-Based Fire Model. For. Ecol. Manag. 2006, 234, S92. [Google Scholar] [CrossRef]
  60. Jahdi, R.; Darvishsefat, A.A.; Etemad, V.; Mostafavi, M.A. Wind Effect on Wildfire and Simulation of Its Spread (Case Study: Siahkal Forest in Northern Iran). J. Agric. Sci. Technol. 2014, 16, 1109–1121. [Google Scholar]
  61. Simeoni, A.; Salinesi, P.; Morandini, F. Physical Modelling of Forest Fire Spreading through Heterogeneous Fuel Beds. Int. J. Wildland Fire 2011, 20, 625–632. [Google Scholar] [CrossRef]
  62. Vaz, G.C.; André, J.C.S.; Viegas, D.X. Fire Spread Model for a Linear Front in a Horizontal Solid Porous Fuel Bed in Still Air. Combust. Sci. Technol. 2004, 176, 135–182. [Google Scholar]
  63. HomChaudhuri, B.; Kumar, M.; Cohen, K. Genetic Algorithm Based Simulation–Optimization for Fighting Wildfires. Int. J. Comput. Methods 2013, 10, 1350035. [Google Scholar] [CrossRef]
  64. Pereira, J.; Mendes, J.; Júnior, J.S.S.; Viegas, C.; Paulo, J.R. A Review of Genetic Algorithm Approaches for Wildfire Spread Prediction Calibration. Mathematics 2022, 10, 300. [Google Scholar] [CrossRef]
  65. Jain, P.; Coogan, S.C.P.; Subramanian, S.G.; Crowley, M.; Taylor, S.; Flannigan, M.D. A Review of Machine Learning Applications in Wildfire Science and Management. Environ. Rev. 2020, 28, 478–505. [Google Scholar] [CrossRef]
  66. Guelpa, E.; Sciacovelli, A.; Verda, V.; Ascoli, D. Faster Prediction of Wildfire Behaviour by Physical Models through Application of Proper Orthogonal Decomposition. Int. J. Wildland Fire 2016, 25, 1181–1192. [Google Scholar] [CrossRef]
  67. Valero, M.M.; Jofre, L.; Torres, R. Multifidelity Prediction in Wildfire Spread Simulation: Modeling, Uncertainty Quantification and Sensitivity Analysis. Environ. Model. Softw. 2021, 141, 105050. [Google Scholar] [CrossRef]
  68. Ujjwal, K.C.; Aryal, J.; Garg, S.; Hilton, J. Global Sensitivity Analysis for Uncertainty Quantification in Fire Spread Models. Environ. Model. Softw. 2021, 143, 105110. [Google Scholar]
  69. Salvador, R.; Pinol, J.; Tarantola, S.; Pla, E. Global Sensitivity Analysis and Scale Effects of a Fire Propagation Model Used over Mediterranean Shrublands. Ecol. Model. 2001, 136, 175–189. [Google Scholar] [CrossRef]
  70. Liu, Y.; Hussaini, M.Y.; Ökten, G. Global Sensitivity Analysis for the Rothermel Model Based on High-Dimensional Model Representation. Can. J. For. Res. 2015, 45, 1474–1479. [Google Scholar] [CrossRef]
  71. Howell, J.R.; Mengüç, M.P.; Daun, K.; Siegel, R. Thermal Radiation Heat Transfer; CRC Press: Boca Raton, FL, USA, 2020; ISBN 0429327307. [Google Scholar]
  72. Trollope, W.S.W.; Trollope, L.A.; Potgieter, A.L.F.; Zambatis, N. SAFARI-92 Characterization of Biomass and Fire Behavior in the Small Experimental Burns in the Kruger National Park. J. Geophys. Res. Atmos. 1996, 101, 23531–23539. [Google Scholar] [CrossRef]
  73. Hankinson, G. A Method for Calculating the Configuration Factor between a Flame and a Receiving Target for a Wide Range of Flame Geometries Relevant to Large Scale Fires. Fire Saf. Sci. 1986, 1, 197–206. [Google Scholar] [CrossRef]
  74. Hankinson, G. Geometries Relevant to Large Scale Fires. In Proceedings of the Fire Safety Science: Proceedings of the First International Symposium; CRC Press: Boca Raton, FL, USA, 1986; Volume 1, p. 197. [Google Scholar]
  75. Frank, A.; Heidemann, W.; Spindler, K. Modeling of the Surface-to-Surface Radiation Exchange Using a Monte Carlo Method. In Proceedings of the Journal of Physics: Conference Series; IOP Publishing: Bristol, UK, 2016; Volume 745, p. 32143. [Google Scholar]
  76. Heskestad, G. Luminous Heights of Turbulent Diffusion Flames. Fire Saf. J. 1983, 5, 103–108. [Google Scholar] [CrossRef]
  77. Bejan, A. Convection Heat Transfer; John Wiley & Sons: Hoboken, NJ, USA, 2013; ISBN 1118330080. [Google Scholar]
  78. Finney, M.A.; Cohen, J.D.; Forthofer, J.M.; McAllister, S.S.; Gollner, M.J.; Gorham, D.J.; Saito, K.; Akafuah, N.K.; Adam, B.A.; English, J.D. Role of Buoyant Flame Dynamics in Wildfire Spread. Proc. Natl. Acad. Sci. USA 2015, 112, 9833–9838. [Google Scholar] [CrossRef]
  79. Rothermel, R.C. A Mathematical Model for Predicting Fire Spread in Wildland Fuels; Intermountain Forest & Range Experiment Station, Forest Service, US: Washington, DC, USA, 1972; Volume 115. [Google Scholar]
  80. Noble, I.R.; Gill, A.M.; Bary, G.A. V McArthur’s Fire-danger Meters Expressed as Equations. Aust. J. Ecol. 1980, 5, 201–203. [Google Scholar] [CrossRef]
  81. Cheney, N.P.; Gould, J.S. Fire Growth in Grassland Fuels. Int. J. Wildland Fire 1995, 5, 237–247. [Google Scholar] [CrossRef]
  82. Cheney, N.P.; Gould, J.S.; Catchpole, W.R. The Influence of Fuel, Weather and Fire Shape Variables on Fire-Spread in Grasslands. Int. J. Wildl. Fire 1993, 3, 31–44. [Google Scholar] [CrossRef]
  83. Rothermel, R.C. How to Predict the Spread and Intensity of Forest and Range Fires. In US Department of Agriculture, Forest Service, Intermountain Forest and Range Experiment Station; USDA: Washington, DC, USA, 1983; Volume 143. [Google Scholar]
  84. Rabitz, H.; Aliş, Ö.F.; Shorter, J.; Shim, K. Efficient Input—Output Model Representations. Comput. Phys. Commun. 1999, 117, 11–20. [Google Scholar] [CrossRef]
Figure 1. A schematic of the model network and the relationship between cells in the fuel bed.
Figure 1. A schematic of the model network and the relationship between cells in the fuel bed.
Sustainability 15 13639 g001
Figure 2. Model of the solid flame used in the proposed model.
Figure 2. Model of the solid flame used in the proposed model.
Sustainability 15 13639 g002
Figure 3. Schematic diagram of heat transfer and fire propagation mechanisms.
Figure 3. Schematic diagram of heat transfer and fire propagation mechanisms.
Sustainability 15 13639 g003
Figure 4. Image of the grassland fire experiments conducted in the Northern Territory of Australia, F19, at t = 56 s [2].
Figure 4. Image of the grassland fire experiments conducted in the Northern Territory of Australia, F19, at t = 56 s [2].
Sustainability 15 13639 g004
Figure 5. Time step independence graph (time step: 0.1, 0.15, 0.25, 0.5, 1.0, 2.0, and 2.5 s).
Figure 5. Time step independence graph (time step: 0.1, 0.15, 0.25, 0.5, 1.0, 2.0, and 2.5 s).
Sustainability 15 13639 g005
Figure 6. Time evolution of fire perimeters, shown as shaded contours of temperature contour, for Uw = 5 m/s and ignition line lengths (Lig) of 8 m and 25 m. The fire perimeters are plotted at times t = 0, 60, and 120 s. Upper diagrams belong to current study and lower diagrams belong to simulation with WFDS.
Figure 6. Time evolution of fire perimeters, shown as shaded contours of temperature contour, for Uw = 5 m/s and ignition line lengths (Lig) of 8 m and 25 m. The fire perimeters are plotted at times t = 0, 60, and 120 s. Upper diagrams belong to current study and lower diagrams belong to simulation with WFDS.
Sustainability 15 13639 g006
Figure 7. Validation of the general and improved models for two prominent experimental tests, Case C064 and Case F19, for three time periods.
Figure 7. Validation of the general and improved models for two prominent experimental tests, Case C064 and Case F19, for three time periods.
Sustainability 15 13639 g007aSustainability 15 13639 g007b
Figure 8. Comparison diagram of the rate of fire spread based on the ambient wind speed of the developed model with the results of other models [27,80,81,82,83]. ((a): C064 and (b): F19).
Figure 8. Comparison diagram of the rate of fire spread based on the ambient wind speed of the developed model with the results of other models [27,80,81,82,83]. ((a): C064 and (b): F19).
Sustainability 15 13639 g008
Figure 9. Fire propagation contour; case F19 with 8 m and 25 m fire ignition line. (at t = 60 s and t = 120 s).
Figure 9. Fire propagation contour; case F19 with 8 m and 25 m fire ignition line. (at t = 60 s and t = 120 s).
Sustainability 15 13639 g009
Figure 10. Comparative chart of relative sensitivity analysis of 10 parameters to rate of fire spread (ROS).
Figure 10. Comparative chart of relative sensitivity analysis of 10 parameters to rate of fire spread (ROS).
Sustainability 15 13639 g010
Figure 11. Diagram of heat transfer mechanisms on one of the points in the path of the fire front until the ignition moment, and the temperature diagram for the same point in the simulation environment.
Figure 11. Diagram of heat transfer mechanisms on one of the points in the path of the fire front until the ignition moment, and the temperature diagram for the same point in the simulation environment.
Sustainability 15 13639 g011
Figure 12. Graph of heat transfer rate before the burning moment (50 < t < 60 s)—(x = 60 and y = 0).
Figure 12. Graph of heat transfer rate before the burning moment (50 < t < 60 s)—(x = 60 and y = 0).
Sustainability 15 13639 g012
Figure 13. Numerical simulations of the rate of spread (ROS) against wind velocity (Case F19) for different packing ratios (PRs).
Figure 13. Numerical simulations of the rate of spread (ROS) against wind velocity (Case F19) for different packing ratios (PRs).
Sustainability 15 13639 g013
Figure 14. Relative decay of the rate of spread vs. fuel initial moisture. (Case F19).
Figure 14. Relative decay of the rate of spread vs. fuel initial moisture. (Case F19).
Sustainability 15 13639 g014
Table 1. Quasi-physical models published in the literature between 1990 and 2020.
Table 1. Quasi-physical models published in the literature between 1990 and 2020.
Authors (Year)OriginHeat
Transfer Mechanism
Flame
Radiation Type
Flame
Equivalent Temperature
DimensionPlane
de Mestre et al. [24]AustraliaRadiationSFM7951x
Carrier et al. (1991) [30]USAConvection/DiffusionVFM-2xy
Albini et al. (1996) [25]USARadiationSFM10502xz
Santoni and Balbi (1998) [56]FranceRadiation/ConvectionSFM9702xy
Catchpole et al. (2002) [51]Australia and USARadiation/ConvectionSFM-2xz
Vaz et al. (2004) [62]PortugalMixSFM11002xy
Koo et al. (2005) [26]USARadiation/ConvectionSFM10832xz
Mohamed Drissi (2015) [57]FranceRadiation/ConvectionSFM10832xy
J.K. Adou et al. (2015) [50]Ivory CoastRadiation/ConvectionSFM10832xy
Balbi et al. (2020) [47]FranceRadiation/ConvectionSFM-1x
SFM: Surface Flame Model; VFM: Volume Flame Model.
Table 2. Simulation parameters used in the WFDS model (Case F19).
Table 2. Simulation parameters used in the WFDS model (Case F19).
ParameterSymbolValue
Domain Sizexyz200 m × 200 m × 25 m
Minimum Grid Resolution-0.1 m
Enthalpy of CombustionΔhc15,600 KJ·Kg−1
Wind SpeedU∞uniform—5 m·s−1
Ambient TemperatureT∞307 K
Solid Phase Densityρs512 kg·m−3
Vegetable Surface to Volume Ratioσs12,240·m−1
Height of the Vegetative (Cell j)Hj0.51
Maximum Simulation Period-138 s
Table 3. Model parameters.
Table 3. Model parameters.
ItemParameterSpecificationCase F19Case F19
1σStefan–Boltzmann constant (W/m2 K)5.67 × 10−85.67 × 10−8
2αfbfuel bed absorptivity0.60.6
3Lflflame length (m)5.05.0
4Djlength scale inside the fuel bed (m)2.52.5
5εflflame emissivity--
6εfbfuel bed emissivity0.60.6
7εbember emissivity1.01.0
8Hflflame height (m)5.05.0
9Hjheight of the vegetative cell j (m)0.510.21
10RSFMSolid Flame Model radius (m)1.251.25
11Tflflame temperature (K)10831083
12Tbignition temperature (K)500500
13Tpyrpyrolysis temperature (K)500500
14Tinfambient temperature (K)307305
15UWambient wind (m/s)4.84.6
16Ufbvelocity of wind inside the fuel bed (m/s)11
17Kflthermal conductivity of air at flame temperature (W/m·K)70 × 10−370 × 10−3
18Kbthermal conductivity of air at bed ambient (W/m·K)40 × 10−340 × 10−3
19ϑgkinematic viscosity of air at the flame temperature13 × 10−513 × 10−5
20Pr.Prandtl number0.70.7
21Afbtotal fuel particle surface area per fuel bed volume12,2409770
22W0initial vegetable moisture0.0580.063
23Cpwetfuel particle wet fuel particle wet specific heat capacity1110 + 3.7 × T(K)1110 + 3.7 × T(K)
24packing ratio0.00120.0026
Table 4. Simulation results in comparison with simulations on the WFDS model; Case Study F19.
Table 4. Simulation results in comparison with simulations on the WFDS model; Case Study F19.
ItemIgnition Line (m)Time (s)ROS (WFDS) (m/s)ROS (Current Study) (m/s)ROS Difference (%)Fire Perimeter (WFDS) (m)Fire Perimeter (Current Study) (m)Fire Perimeter Difference (%)
18.060.00.850.927.8121.7118.72.5
28.0120.00.971.1720.7265.8316.218.9
325.060.01.351.1415.9195.0156.619.7
425.0120.01.321.199.7376.4362.13.8
Table 5. Results of the general model and modified model in comparison with experimental cases F19 and C064.
Table 5. Results of the general model and modified model in comparison with experimental cases F19 and C064.
ItemCaseTime (s)Interval (s)ROS (Experiment) (m/s)ROS (General Model) (m/s)ROS Difference (%)ROS (Modified Model) (m/s)ROS Difference (%)
1F1956561.341.0330%1.293%
2F1986301.421.1721%1.5811%
3F19138521.591.2032%1.590%
4C06427270.700.745%0.9324%
5C06453261.120.6766%1.065%
6C064100471.190.6972%1.127%
Table 6. Variation range and sensitivity of parameters effect on the rate of spread (Case F19).
Table 6. Variation range and sensitivity of parameters effect on the rate of spread (Case F19).
ItemParameterReference ValueVariation Range1st LevelROS Change @ 1st Level2nd LevelROS Change @ 2nd Level
HfVegetative height (m)0.5120%0.408+0.1250.612−0.125
LfFlame length520%4−0.2916+0.291
TflFlame temperature (K)108310%975−0.3331191+0.375
TignIgnition temperature (K)50010%450+0.167550−0.125
TAmbient temperature (K)3075%292−0.125322+0.167
UwWind speed (m/s)520%4−0.0836+0.083
W0Initial vegetable moisture0.05838%0.036+0.2080.08−0.125
cpFuel specific heat (kJ/kg·K)1.11 + 3.7 × T22%6.10 + 3.7 × T+0.1251.610 + 3.7 × T−0.125
ρsSolid phase density (kg/m3)51225%384+0.458640−0.291
Packing ratio0.001225%0.0009+0.4580.0015−0.291
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mohammadian Bishe, E.; Afshin, H.; Farhanieh, B. Modified Quasi-Physical Grassland Fire Spread Model: Sensitivity Analysis. Sustainability 2023, 15, 13639. https://doi.org/10.3390/su151813639

AMA Style

Mohammadian Bishe E, Afshin H, Farhanieh B. Modified Quasi-Physical Grassland Fire Spread Model: Sensitivity Analysis. Sustainability. 2023; 15(18):13639. https://doi.org/10.3390/su151813639

Chicago/Turabian Style

Mohammadian Bishe, Esmaeil, Hossein Afshin, and Bijan Farhanieh. 2023. "Modified Quasi-Physical Grassland Fire Spread Model: Sensitivity Analysis" Sustainability 15, no. 18: 13639. https://doi.org/10.3390/su151813639

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop