[go: up one dir, main page]

 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (812)

Search Parameters:
Keywords = wildfire modeling

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
18 pages, 7239 KiB  
Article
A Lightweight Wildfire Detection Method for Transmission Line Perimeters
by Xiaolong Huang, Weicheng Xie, Qiwen Zhang, Yeshen Lan, Huiling Heng and Jiawei Xiong
Electronics 2024, 13(16), 3170; https://doi.org/10.3390/electronics13163170 - 11 Aug 2024
Viewed by 241
Abstract
Due to extreme weather conditions and complex geographical features, the environments around power lines in forest areas have a high risk of wildfires. Once a wildfire occurs, it causes severe damage to the forest ecosystem. Monitoring wildfires around power lines in forested regions [...] Read more.
Due to extreme weather conditions and complex geographical features, the environments around power lines in forest areas have a high risk of wildfires. Once a wildfire occurs, it causes severe damage to the forest ecosystem. Monitoring wildfires around power lines in forested regions through deep learning can reduce the harm of wildfires to natural environments. To address the challenges of wildfire detection around power lines in forested areas, such as interference from complex environments, difficulty detecting small target objects, and high model complexity, a lightweight wildfire detection model based on the improved YOLOv8 is proposed. Firstly, we enhanced the image-feature-extraction capability using a novel feature-extraction network, GS-HGNetV2, and replaced the conventional convolutions with a Ghost Convolution (GhostConv) to reduce the model parameters. Secondly, the use of the RepViTBlock to replace the original Bottleneck in C2f enhanced the model’s feature-fusion capability, thereby improving the recognition accuracy for small target objects. Lastly, we designed a Resource-friendly Convolutional Detection Head (RCD), which reduces the model complexity while maintaining accuracy by sharing the parameters. The model’s performance was validated using a dataset of 11,280 images created by merging a custom dataset with the D-Fire data for monitoring wildfires near power lines. In comparison to YOLOv8, our model saw an improvement of 3.1% in the recall rate and 1.1% in the average precision. Simultaneously, the number of parameters and computational complexity decreased by 54.86% and 39.16%, respectively. The model is more appropriate for deployment on edge devices with limited computational power. Full article
(This article belongs to the Section Artificial Intelligence)
Show Figures

Figure 1

Figure 1
<p>Overall structure diagram of the improved YOLO model.</p>
Full article ">Figure 2
<p>Ghost-HGNetV2 network architecture diagram.</p>
Full article ">Figure 3
<p>Ghost-HGBlock network architecture diagram.</p>
Full article ">Figure 4
<p>Comparison of C2f and C2f-RVT structures: (<b>a</b>) C2f; (<b>b</b>) C2f-RVT.</p>
Full article ">Figure 5
<p>Resource-friendly Convolutional Detection Head network structure.</p>
Full article ">Figure 6
<p>Typical scenarios of DW-fire: (<b>a</b>) normal fire, (<b>b</b>) early fire, (<b>c</b>) fire disturbance, and (<b>d</b>) smoke disturbance.</p>
Full article ">Figure 7
<p>Visualization statistics of dataset labels. (<b>a</b>) Statistics of label positions relative to images. (<b>b</b>) Statistics of label sizes relative to images.</p>
Full article ">Figure 8
<p>Comparison chart of computational complexity for each layer: (<b>a</b>) original model; (<b>b</b>) improved model.</p>
Full article ">Figure 9
<p>Comparison of detection results between advanced wildfire algorithms and our method: (<b>a</b>) wildfire scene 1, (<b>b</b>) wildfire scene 2, and (<b>c</b>) wildfire scene 3 [<a href="#B24-electronics-13-03170" class="html-bibr">24</a>,<a href="#B25-electronics-13-03170" class="html-bibr">25</a>].</p>
Full article ">Figure 10
<p>Detection results under interference environment.</p>
Full article ">Figure 11
<p>Detection results in low-light conditions.</p>
Full article ">Figure 12
<p>Comparison diagram of the detection effect of the model on small target objects.</p>
Full article ">Figure 13
<p>YOLOv8 and the improved model in heatmaps of different sizes of detection heads: (<b>a</b>) 20 × 20 detection head, (<b>b</b>) 40 × 40 detection head, and (<b>c</b>) 80 × 80 detection head.</p>
Full article ">Figure 14
<p>The curves of precision, recall, and mAP@50.</p>
Full article ">
17 pages, 1946 KiB  
Article
Data-Driven PM2.5 Exposure Prediction in Wildfire-Prone Regions and Respiratory Disease Mortality Risk Assessment
by Sadegh Khanmohammadi, Mehrdad Arashpour, Milad Bazli and Parisa Farzanehfar
Fire 2024, 7(8), 277; https://doi.org/10.3390/fire7080277 - 7 Aug 2024
Viewed by 298
Abstract
Wildfires generate substantial smoke containing fine particulate matter (PM2.5) that adversely impacts health. This study develops machine learning models integrating pre-wildfire factors like weather and fuel conditions with post-wildfire health impacts to provide a holistic understanding of smoke exposure risks. Various [...] Read more.
Wildfires generate substantial smoke containing fine particulate matter (PM2.5) that adversely impacts health. This study develops machine learning models integrating pre-wildfire factors like weather and fuel conditions with post-wildfire health impacts to provide a holistic understanding of smoke exposure risks. Various data-driven models including Support Vector Regression, Multi-layer Perceptron, and three tree-based ensemble algorithms (Random Forest, Extreme Gradient Boosting (XGBoost), and Natural Gradient Boosting (NGBoost)) are evaluated in this study. Ensemble models effectively predict PM2.5 levels based on temperature, humidity, wind, and fuel moisture, revealing the significant roles of radiation, temperature, and moisture. Further modelling links smoke exposure to deaths from chronic obstructive pulmonary disease (COPD) and lung cancer using age, sex, and pollution type as inputs. Ambient pollution is the primary driver of COPD mortality, while age has a greater influence on lung cancer deaths. This research advances atmospheric and health impact understanding, aiding forest fire prevention and management. Full article
(This article belongs to the Special Issue Forest Fuel Treatment and Fire Risk Assessment)
Show Figures

Figure 1

Figure 1
<p>Correlation among weather variables, fuel variables, and wildfire smoke from June 2019 to May 2020. PM<sub>2.5</sub> is the daily mean of fine particulate matter with a diameter ≤2.5 μg/m<sup>3</sup>; M<sub>dl</sub> is fuel moisture content in percentage; T is the ambient temperature that is reported by the weather station in Celsius; S is solar radiation that is reported by the weather station in megajoule per square meter; R is rain that is reported by the weather station in millimeters.</p>
Full article ">Figure 2
<p>The prediction performance of the NGBoost model. The selected data-driven model for predicting PM<sub>2.5</sub> uses the weather and fuel variables (air temperature, relative humidity, solar radiation, rain, season, wind speed, and fuel moisture) as input variables. Plots of the observed PM<sub>2.5</sub> versus the predicted PM<sub>2.5</sub> for (<b>a</b>) spring (<b>b</b>) summer (<b>c</b>) fall (<b>d</b>) winter. The solid black line is the line of perfect agreement. The dashed lines indicate the ±35% error interval. (<b>e</b>) shows the importance of each input variable in the model development.</p>
Full article ">Figure 3
<p>Prediction performance of the NGBoost model. The model output is the number of deaths associated with COPD and TB&amp;L by using input variables (sex, age, household and ambient air pollution, and year of impact). Plots of the observed versus the predicted number of deaths associated with COPD (<b>a</b>) and TB&amp;L (<b>b</b>). The dashed lines indicate the ±35% error interval. Red points are the training dataset (80% of data records), and blue points are the test dataset (20% of data records). SHAP interpretation of the selected model to find the influential variables for COPD-related deaths (<b>c</b>) and TB&amp;L-related deaths (<b>d</b>).</p>
Full article ">Figure 4
<p>Changes in PM<sub>2.5</sub> μg/m<sup>3</sup>.</p>
Full article ">
21 pages, 11248 KiB  
Article
Transferability of Empirical Models Derived from Satellite Imagery for Live Fuel Moisture Content Estimation and Fire Risk Prediction
by Eva Marino, Lucía Yáñez, Mercedes Guijarro, Javier Madrigal, Francisco Senra, Sergio Rodríguez and José Luis Tomé
Fire 2024, 7(8), 276; https://doi.org/10.3390/fire7080276 - 6 Aug 2024
Viewed by 355
Abstract
Estimating live fuel moisture content (LFMC) is critical for assessing vegetation flammability and predicting potential fire behaviour, thus providing relevant information for wildfire prevention and management. Previous research has demonstrated that empirical modelling based on spectral data derived from remote sensing is useful [...] Read more.
Estimating live fuel moisture content (LFMC) is critical for assessing vegetation flammability and predicting potential fire behaviour, thus providing relevant information for wildfire prevention and management. Previous research has demonstrated that empirical modelling based on spectral data derived from remote sensing is useful for retrieving LFMC. However, these types of models are often very site-specific and generally considered difficult to extrapolate. In the present study, we analysed the performance of empirical models based on Sentinel-2 spectral data for estimating LFMC in fire-prone shrubland dominated by Cistus ladanifer. We used LFMC data collected in the field between June 2021 and September 2022 in 27 plots in the region of Andalusia (southern Spain). The specific objectives of the study included (i) to test previous existing models fitted for the same shrubland species in a different study area in the region of Madrid (central Spain); (ii) to calibrate empirical models with the field data from the region of Andalusia, comparing the model performance with that of existing models; and (iii) to test the capacity of the best empirical models to predict decreases in LFMC to critical threshold values in historical wildfire events. The results showed that the empirical models derived from Sentinel-2 data provided accurate LFMC monitoring, with a mean absolute error (MAE) of 15% in the estimation of LFMC variability throughout the year and with the MAE decreasing to 10% for the critical lower LFMC values (<100%). They also showed that previous models could be easily recalibrated for extrapolation to different geographical areas, yielding similar errors to the specific empirical models fitted in the study area in an independent validation. Finally, the results showed that decreases in LFMC in historical wildfire events were accurately predicted by the empirical models, with LFMC <80% in this fire-prone shrubland species. Full article
Show Figures

Figure 1

Figure 1
<p>Study area (black rectangle) within the region of Andalusia (red) in Spain (green), with details on the location of the 27 sampling plots (INIA + INFOCA) and the historical wildfires (burnt areas for each year in different colours) in the pilot areas (Z1, Z2 and Z3).</p>
Full article ">Figure 2
<p>Observed LFMC values in INIA (green) and INFOCA (orange) field sampling plots during the study period (June 2021 to September 2022).</p>
Full article ">Figure 3
<p>Pearson correlation coefficients for spectral indices and LFMC in the calibration dataset (values highlighted in dark blue denote stronger correlations, r &gt; 0.90).</p>
Full article ">Figure 4
<p>Extrapolation of the best empirical models previously fitted for the region of Madrid to the sampling plots in Andalusia, NLR-exp (<b>left</b>) and NLR-sqr (<b>right</b>), depicting validation results of original models (<b>upper</b>) and after recalibration (<b>lower</b>) with the linear regression (Y = a + bX) between observed and predicted values: NLR-exp (a = 17.33, b = 0.712); NLR-sqr (a = 7.659, b = 0.832). Y = recalibrated LFMC; X = predicted LFMC.</p>
Full article ">Figure 5
<p>Observed vs. predicted LFMC values in the best models for each formulation tested with the calibration dataset (n = 224) for field plots in Andalusia.</p>
Full article ">Figure 5 Cont.
<p>Observed vs. predicted LFMC values in the best models for each formulation tested with the calibration dataset (n = 224) for field plots in Andalusia.</p>
Full article ">Figure 6
<p>Observed vs. predicted LFMC values in the independent validation (n = 111) of the best models fitted for field plots in Andalusia.</p>
Full article ">Figure 6 Cont.
<p>Observed vs. predicted LFMC values in the independent validation (n = 111) of the best models fitted for field plots in Andalusia.</p>
Full article ">Figure 7
<p>Performance of the best empirical model for LFMC estimation before the 7 selected historical wildfire events in the pilot areas (Z1, Z2 and Z3), depicting the mean value of reference plots available for each wildfire (blue) and the LFMC range for each date (vertical bars). Headings indicate the wildfire name, with the ignition date in brackets and depicted as red symbols.</p>
Full article ">Figure A1
<p>Changes in LFMC estimation obtained with the best empirical model derived from Sentinel-2 data in the Z2 pilot area before the Almonaster wildfire (ignited on 27 August 2020), which burned an area of 14,957 ha in 12 days (perimeter in black). This wildfire was the biggest event in the region of Andalusia during the study period (2018–2022). LFMC is only shown for pixels corresponding to shrubland. Reference plots used in <a href="#fire-07-00276-f007" class="html-fig">Figure 7</a> for LFMC value calculations are depicted as green triangles.</p>
Full article ">Figure A1 Cont.
<p>Changes in LFMC estimation obtained with the best empirical model derived from Sentinel-2 data in the Z2 pilot area before the Almonaster wildfire (ignited on 27 August 2020), which burned an area of 14,957 ha in 12 days (perimeter in black). This wildfire was the biggest event in the region of Andalusia during the study period (2018–2022). LFMC is only shown for pixels corresponding to shrubland. Reference plots used in <a href="#fire-07-00276-f007" class="html-fig">Figure 7</a> for LFMC value calculations are depicted as green triangles.</p>
Full article ">
28 pages, 20313 KiB  
Article
SHAP-Driven Explainable Artificial Intelligence Framework for Wildfire Susceptibility Mapping Using MODIS Active Fire Pixels: An In-Depth Interpretation of Contributing Factors in Izmir, Türkiye
by Muzaffer Can Iban and Oktay Aksu
Remote Sens. 2024, 16(15), 2842; https://doi.org/10.3390/rs16152842 - 2 Aug 2024
Viewed by 579
Abstract
Wildfire susceptibility maps play a crucial role in preemptively identifying regions at risk of future fires and informing decisions related to wildfire management, thereby aiding in mitigating the risks and potential damage posed by wildfires. This study employs eXplainable Artificial Intelligence (XAI) techniques, [...] Read more.
Wildfire susceptibility maps play a crucial role in preemptively identifying regions at risk of future fires and informing decisions related to wildfire management, thereby aiding in mitigating the risks and potential damage posed by wildfires. This study employs eXplainable Artificial Intelligence (XAI) techniques, particularly SHapley Additive exPlanations (SHAP), to map wildfire susceptibility in Izmir Province, Türkiye. Incorporating fifteen conditioning factors spanning topography, climate, anthropogenic influences, and vegetation characteristics, machine learning (ML) models (Random Forest, XGBoost, LightGBM) were used to predict wildfire-prone areas using freely available active fire pixel data (MODIS Active Fire Collection 6 MCD14ML product). The evaluation of the trained ML models showed that the Random Forest (RF) model outperformed XGBoost and LightGBM, achieving the highest test accuracy (95.6%). All of the classifiers demonstrated a strong predictive performance, but RF excelled in sensitivity, specificity, precision, and F-1 score, making it the preferred model for generating a wildfire susceptibility map and conducting a SHAP analysis. Unlike prevailing approaches focusing solely on global feature importance, this study fills a critical gap by employing a SHAP summary and dependence plots to comprehensively assess each factor’s contribution, enhancing the explainability and reliability of the results. The analysis reveals clear associations between factors such as wind speed, temperature, NDVI, slope, and distance to villages with increased fire susceptibility, while rainfall and distance to streams exhibit nuanced effects. The spatial distribution of the wildfire susceptibility classes highlights critical areas, particularly in flat and coastal regions near settlements and agricultural lands, emphasizing the need for enhanced awareness and preventive measures. These insights inform targeted fire management strategies, highlighting the importance of tailored interventions like firebreaks and vegetation management. However, challenges remain, including ensuring the selected factors’ adequacy across diverse regions, addressing potential biases from resampling spatially varied data, and refining the model for broader applicability. Full article
(This article belongs to the Special Issue Artificial Intelligence for Natural Hazards (AI4NH))
Show Figures

Figure 1

Figure 1
<p>Study area and wildfire inventory.</p>
Full article ">Figure 2
<p>Topographical conditioning factors: (<b>A</b>) elevation, (<b>B</b>) slope, (<b>C</b>) aspect, (<b>D</b>) TWI.</p>
Full article ">Figure 3
<p>Climatic conditioning factors: (<b>A</b>) annual average temperature, (<b>B</b>) annual rainfall, (<b>C</b>) annual solar radiation, (<b>D</b>) annual average wind speed.</p>
Full article ">Figure 4
<p>Anthropogenic conditioning factors: (<b>A</b>) LULC, (<b>B</b>) distance to roads, (<b>C</b>) distance to villages, (<b>D</b>) distance to streams.</p>
Full article ">Figure 5
<p>Vegetation-related conditioning factors: (<b>A</b>) forest type, (<b>B</b>) tree cover density, (<b>C</b>) NDVI.</p>
Full article ">Figure 6
<p>Summary of the research methodology steps utilized in the study.</p>
Full article ">Figure 7
<p>Multicollinearity test results (<b>upper</b>) before recursive elimination and (<b>lower</b>) after recursive elimination and selected factors.</p>
Full article ">Figure 8
<p>Pearson’s correlation coefficient matrix.</p>
Full article ">Figure 9
<p>Confusion matrixes.</p>
Full article ">Figure 10
<p>Classifiers’ performance comparison.</p>
Full article ">Figure 11
<p>ROC curves.</p>
Full article ">Figure 12
<p>SHAP summary plot of RF classifier’s output.</p>
Full article ">Figure 13
<p>Global feature importance by absolute SHAP values.</p>
Full article ">Figure 14
<p>Each factor’s SHAP dependence plots.</p>
Full article ">Figure 15
<p>Generated wildfire susceptibility map for the province of Izmir.</p>
Full article ">Figure 16
<p>Area extent of susceptibility classes and number of wildfire samples corresponding to each susceptibility class.</p>
Full article ">
21 pages, 8219 KiB  
Article
An Improved Fire and Smoke Detection Method Based on YOLOv8n for Smart Factories
by Ziyang Zhang, Lingye Tan and Tiong Lee Kong Robert
Sensors 2024, 24(15), 4786; https://doi.org/10.3390/s24154786 - 24 Jul 2024
Viewed by 358
Abstract
Factories play a crucial role in economic and social development. However, fire disasters in factories greatly threaten both human lives and properties. Previous studies about fire detection using deep learning mostly focused on wildfire detection and ignored the fires that happened in factories. [...] Read more.
Factories play a crucial role in economic and social development. However, fire disasters in factories greatly threaten both human lives and properties. Previous studies about fire detection using deep learning mostly focused on wildfire detection and ignored the fires that happened in factories. In addition, lots of studies focus on fire detection, while smoke, the important derivative of a fire disaster, is not detected by such algorithms. To better help smart factories monitor fire disasters, this paper proposes an improved fire and smoke detection method based on YOLOv8n. To ensure the quality of the algorithm and training process, a self-made dataset including more than 5000 images and their corresponding labels is created. Then, nine advanced algorithms are selected and tested on the dataset. YOLOv8n exhibits the best detection results in terms of accuracy and detection speed. ConNeXtV2 is then inserted into the backbone to enhance inter-channel feature competition. RepBlock and SimConv are selected to replace the original Conv and improve computational ability and memory bandwidth. For the loss function, CIoU is replaced by MPDIoU to ensure an efficient and accurate bounding box. Ablation tests show that our improved algorithm achieves better performance in all four metrics reflecting accuracy: precision, recall, F1, and mAP@50. Compared with the original model, whose four metrics are approximately 90%, the modified algorithm achieves above 95%. mAP@50 in particular reaches 95.6%, exhibiting an improvement of approximately 4.5%. Although complexity improves, the requirements of real-time fire and smoke monitoring are satisfied. Full article
(This article belongs to the Section Sensing and Imaging)
Show Figures

Figure 1

Figure 1
<p>The structure of the YOLOv8n algorithm.</p>
Full article ">Figure 2
<p>The architecture of ConvNeXt V2.</p>
Full article ">Figure 3
<p>The architecture of Simconv, RepConv, and RepBlock.</p>
Full article ">Figure 4
<p>Schematic diagram of MPDIoU.</p>
Full article ">Figure 5
<p>The structure of the improved YOLOv8n for factory fire and smoke detection.</p>
Full article ">Figure 6
<p>The after-process after collection of images using Visual Similarity Duplicate Image Finder.</p>
Full article ">Figure 7
<p>Examples of factory fire disaster images inside.</p>
Full article ">Figure 8
<p>Examples of factory fire disaster images outside.</p>
Full article ">Figure 9
<p>Visualization results of the self-made datasets and labels. (<b>a</b>) The number of labels for fire and smoke labels; (<b>b</b>) the size of the labels; (<b>c</b>) the distribution of labels’ centroid locations of the total image; (<b>d</b>) the distribution of labels’ size of the total image.</p>
Full article ">Figure 10
<p>Improved YOLOv8 vs. the other methods: bar charts of FPS and mAP@0.5.</p>
Full article ">Figure 11
<p>Precision-recall curve and precision-confidence curve.</p>
Full article ">Figure 12
<p>The curve of precision–epochs, recall–epochs, and mAP–epochs.</p>
Full article ">Figure 13
<p>Visible experiments of improved and original algorithms for various indoor environments in factories.</p>
Full article ">Figure 14
<p>Visible experiments of improved and original algorithms for various outdoor environments in factories.</p>
Full article ">
7 pages, 1892 KiB  
Proceeding Paper
Simulating the Aerial Ballet: The Dance of Fire-Fighting Planes and Helicopters
by Juha Alander, Lauri Honkasilta and Kalle Saastamoinen
Eng. Proc. 2024, 68(1), 54; https://doi.org/10.3390/engproc2024068054 - 19 Jul 2024
Viewed by 222
Abstract
This study introduces a simulation model to analyze the efficacy of different aerial firefighting strategies in Finland, focusing on the comparative water production capacity and associated costs of firefighting aircraft versus helicopters of varying sizes. By utilizing publicly available data and direct inquiries, [...] Read more.
This study introduces a simulation model to analyze the efficacy of different aerial firefighting strategies in Finland, focusing on the comparative water production capacity and associated costs of firefighting aircraft versus helicopters of varying sizes. By utilizing publicly available data and direct inquiries, the model evaluates the impact of water collection distance on the volume of extinguishing water procured and its costs. The simulation reveals that firefighting aircraft offer a cost-effective solution, particularly when collecting water from distances of thirteen kilometers, where their cost per liter of water aligns with that of smaller helicopters operating closer to the fire zone. The study underscores the importance of precise data input into the calculator, highlighting the potential of aerial firefighting strategies in enhancing wildfire suppression efforts. Full article
(This article belongs to the Proceedings of The 10th International Conference on Time Series and Forecasting)
Show Figures

Figure 1

Figure 1
<p>View of the counter assembly section.</p>
Full article ">Figure 2
<p>Water point on all air units a kilometer away on all. Duration of operation 120 min.</p>
Full article ">Figure 3
<p>Helicopters fetch water from 1 km away, Air Tractor from 6 km away. Duration of operation 120 min.</p>
Full article ">Figure 4
<p>Helicopters fetch water from 1 km away, Air Tractor from 13 km away. Duration of operation 120 min.</p>
Full article ">
21 pages, 4017 KiB  
Article
Probabilistic Path Planning for UAVs in Forest Fire Monitoring: Enhancing Patrol Efficiency through Risk Assessment
by Yuqin Wang, Fengsen Gao and Minghui Li
Fire 2024, 7(7), 254; https://doi.org/10.3390/fire7070254 - 17 Jul 2024
Viewed by 437
Abstract
Forest fire is a significant global natural disaster, and unmanned aerial vehicles (UAVs) have gained attention in wildfire prevention for their efficient and flexible monitoring capabilities. Proper UAV patrol path planning can enhance fire-monitoring accuracy and response speed. This paper proposes a probabilistic [...] Read more.
Forest fire is a significant global natural disaster, and unmanned aerial vehicles (UAVs) have gained attention in wildfire prevention for their efficient and flexible monitoring capabilities. Proper UAV patrol path planning can enhance fire-monitoring accuracy and response speed. This paper proposes a probabilistic path planning (PPP) module that plans UAV patrol paths by combining real-time fire occurrence probabilities at different points. Initially, a forest fire risk logistic regression model is established to compute the fire probabilities at different patrol points. Subsequently, a patrol point filter is applied to remove points with low fire probabilities. Finally, combining fire probabilities with distances between patrol points, a dynamic programming (DP) algorithm is employed to generate an optimal UAV patrol route. Compared with conventional approaches, the experimental results demonstrate that the PPP module effectively improves the timeliness of fire monitoring and containment, and the introduction of DP, considering that the fire probabilities and the patrol point filter both contribute positively to the experimental outcomes. Different combinations of patrol point coordinates and their fire probabilities are further studied to summarize the applicability of this method, contributing to UAV applications in forest fire monitoring and prevention. Full article
(This article belongs to the Special Issue Drone Applications Supporting Fire Management)
Show Figures

Figure 1

Figure 1
<p>Overview of the paper’s work. Label (<b>1</b>) represents “Filter off”, and Label (<b>2</b>) represents “Filter on”.</p>
Full article ">Figure 2
<p>Correlation matrix of variables.</p>
Full article ">Figure 3
<p>ROC curve.</p>
Full article ">Figure 4
<p>The task area. The white dot icons represent UAV’s patrol points, the blue icon represents UAV, and the question mark icon represents UAV’s path planning process.</p>
Full article ">Figure 5
<p>The path results for different methods applied to the same data: (<b>a</b>) Group 1 (conventional method), (<b>b</b>) Group 2, (<b>c</b>) Group 3, and (<b>d</b>) Group 4 (our method). The numbers adjacent to the coordinate points denote the probability of a fire occurring at each point.</p>
Full article ">Figure 6
<p>Comparison of (<b>a</b>) path lengths and (<b>b</b>) fire spread areas by different methods.</p>
Full article ">Figure 7
<p>Fire spread areas at points reached by different methods. The circles are darker further to the left, representing higher fire probabilities and emphasizing the need for stronger fire prevention measures.</p>
Full article ">Figure 8
<p>The paths derived using different methods for the same data. (<b>a</b>) Group 1 (conventional method), (<b>b</b>) Groups 3 and 4 (filtering method). The numbers next to each coordinate point denote the probability of a fire occurring at that location.</p>
Full article ">Figure 9
<p>Time required to reach each point using different methods.</p>
Full article ">Figure 10
<p>Fire spread area and statistical spread area for each point under dynamic programming with and without probability consideration.</p>
Full article ">Figure 11
<p>Fire spread area and statistical spread area for each point under dynamic programming with and without the use of a filter.</p>
Full article ">
14 pages, 2281 KiB  
Article
Modeling the Potential Habitat Gained by Planting Sagebrush in Burned Landscapes
by Julie A. Heinrichs, Michael S. O’Donnell, Elizabeth K. Orning, David A. Pyke, Mark A. Ricca, Peter S. Coates and Cameron L. Aldridge
Conservation 2024, 4(3), 364-377; https://doi.org/10.3390/conservation4030024 - 15 Jul 2024
Viewed by 523
Abstract
Many revegetation projects are intended to benefit wildlife species. Yet, there are few a priori evaluations that assess the potential efficiency of restoration actions in recovering wildlife habitats. We developed a spatial vegetation–habitat recovery model to gauge the degree to which field planting [...] Read more.
Many revegetation projects are intended to benefit wildlife species. Yet, there are few a priori evaluations that assess the potential efficiency of restoration actions in recovering wildlife habitats. We developed a spatial vegetation–habitat recovery model to gauge the degree to which field planting strategies could be expected to recover multi-factor habitat conditions for wildlife following wildfires. We simulated a wildfire footprint, multiple sagebrush (Artemisia spp.) planting scenarios, and tracked projected vegetation growth for 15 years post-fire. We used a vegetation transition framework to track and estimate the degree to which revegetation could accelerate habitat restoration for a Greater sage-grouse (Centrocercus) population within the Great Basin, western United States. We assessed the amount of habitat 15 years post-fire to estimate the degree to which revegetation could be expected to accelerate habitat restoration. Our results highlight a potential disconnect between the expansive areas required by wide-ranging wildlife such as sage-grouse and the relatively small areas that planting treatments have created. Habitat restorations and planting strategies that are intended to benefit sage-grouse may only speed up localized habitat restoration. This study provides an example of how linked revegetation–habitat modeling approaches can scope the expected return on restoration investment for habitat improvements and support the strategic use of limited restoration resources. Full article
Show Figures

Figure 1

Figure 1
<p>Location of Great Basin, Tuscarora study site (Nevada, USA) used for a proof-of-concept approach demonstrating the utility of linked post-fire revegetation–habitat modeling in restoration planning. Variable big sagebrush (<span class="html-italic">Artemisia tridentata</span>) percent cover (%) is shown in green, average male attendance at Greater sage-grouse (<span class="html-italic">Centrocercus urophasianus</span>) breeding sites (lek) is denoted by the size of the circle, and a simulated fire perimeter (35,362 ha) is represented in red.</p>
Full article ">Figure 2
<p>Conceptual overview of vegetation transitions. Boxes represent discrete vegetation states (e.g., dominant cover, height conditions), arrows indicate transitions, lines represent deterministic changes in states (solid lines; dominant cover precedence), and ages (dashed lines; for percent cover, vegetation type). Transitions follow simulated vegetation loss and regrowth over pre-fire (time since fire, TSF = 0), fire event (TSF = 1), sagebrush (<span class="html-italic">Artemisia</span> spp.) planting (TSF = 2), and post-fire habitat recovery (TSF ≥ 2–15) time periods. Success of planted seedling survival was probabilistic (30%, 70%, 100%) and realized the year of planting (TSF = 2). Precedence: mountain big sagebrush (<span class="html-italic">Artemisia tridentata vaseyana</span>) &gt; big sagebrush (<span class="html-italic">A.t. tridentata</span>) &gt; other sagebrush &gt; annual grass &gt; perennial grass &gt; bare ground. See <a href="#app1-conservation-04-00024" class="html-app">Supplementary Materials</a> (<a href="#app1-conservation-04-00024" class="html-app">Sections S2 and S7</a>—Fire-induced Vegetation Loss, Vegetation Transitions, and Regrowth) for details on simulating vegetation loss. See <a href="#app1-conservation-04-00024" class="html-app">Supplementary Materials</a> (<a href="#app1-conservation-04-00024" class="html-app">Sections S6 and S7</a>—Big Sagebrush Growth Rates, Vegetation Transitions, and Regrowth) for details on simulating vegetation regrowth.</p>
Full article ">Figure 3
<p>Modeled recovery of landscape-level habitat for Greater sage-grouse (<span class="html-italic">Centrocercus urophasianus</span>) from a proof-of-concept modeling approach using linked post-fire vegetation transition and habitat selection models. Recovery reflects habitat suitability within a simulated burn perimeter (red) and sagebrush (<span class="html-italic">Artemisia</span> spp.) transplant locations (circles) over pre-fire (where time since fire [TSF] was 0 years), post-fire (TSF 1), the recovery of herbaceous vegetation (TSF 2), and the recovery of critical spring breeding habitat (TSF 15). Top panel: Multi-year planting effort (1.5 million plants), with planting locations targeted to occur in breeding habitat, using fewer, larger patches (71 ha), a high density of plants (235 plants/30 m pixel), and assuming 30% transplant survival. Bottom panel: Single-year planting effort (350,000 plants), with no targeting of planting occurrence in sage-grouse habitat, multiple small patches (9 ha), a high density of plants, and assuming 30% survival. TSF 0 to 15 displays the habitat and suitability gained over 15 years. TSF 2 to 15 indicates habitat recovery due to sagebrush transplants and other slower-returning vegetation, including sagebrush.</p>
Full article ">Figure 4
<p>Percentage of recovered Greater sage-grouse (<span class="html-italic">Centrocercus urophasianus</span>) habitat from a proof-of-concept modeling approach using linked post-fire vegetation transition and habitat selection models. Recovery reflects habitat suitability within a simulated burn perimeter (35,362 ha) derived from projected habitat selection. In these charts, selection classifications were averaged across sagebrush (<span class="html-italic">Artemisia</span> spp.) planting scenarios (<span class="html-italic">n</span> = 48) for pre-fire (left), the year of the fire (center), and 15 years after the fire relative to spring (breeding; top), summer (middle), and winter (bottom) seasons. TSF = time since fire.</p>
Full article ">
17 pages, 2931 KiB  
Review
A Systematic Approach to Map and Evaluate the Wildfire Behavior at a Territorial Scale in the Northwestern Iberian Peninsula
by Thais Rincón, Laura Alonso, Juan Picos, Domingo M. Molina-Terrén and Julia Armesto
Fire 2024, 7(7), 249; https://doi.org/10.3390/fire7070249 - 13 Jul 2024
Viewed by 711
Abstract
In the current context of extreme wildfires, understanding fire behavior at a territorial level has proven crucial for territory planning. This type of analysis is usually conducted by analyzing past wildfire statistics. In this study, we forego the past information related to wildfires [...] Read more.
In the current context of extreme wildfires, understanding fire behavior at a territorial level has proven crucial for territory planning. This type of analysis is usually conducted by analyzing past wildfire statistics. In this study, we forego the past information related to wildfires and analyze, instead, the behavior of the entire territory in the face of wildfires. This allows for the distribution of ignition points to be systematized and for typical and atypical weather scenarios to be considered. This analysis relies on the use of wildfire simulation software. Ignition points used for the simulations were distributed using a systematic 1 × 1 km grid throughout the whole study area. Wildfires were simulated for each ignition point using eight different weather scenarios representing both typical and atypical weather conditions. The fire behavior on the territory was analyzed using rate of spread and intensity parameters for each simulated wildfire. It was observed that this territory is extremely prone to large wildfires both in typical and atypical weather conditions and that there is a tendency for extreme behaviors to develop. Some features were identified as prevention issues that ought to be addressed. This study develops a strategy to evaluate, in a systematic manner, the response of the territory to the threat of wildfires. Full article
(This article belongs to the Special Issue Nature-Based Solutions to Extreme Wildfires)
Show Figures

Figure 1

Figure 1
<p>Study area: the SW corner of Galicia (northwestern Spain).</p>
Full article ">Figure 2
<p>Map of the distribution of Rothermel fuel models in the study area (<b>left</b>) and a detailed view of a portion of the study area (<b>right</b>).</p>
Full article ">Figure 3
<p>Workflow followed to perform the study.</p>
Full article ">Figure 4
<p>Distribution of ignition points, represented as gray dots, in the study area.</p>
Full article ">Figure 5
<p>Output simulation data obtained for a sample ignition point.</p>
Full article ">Figure 6
<p>Graphical representation of wildfire size, FS (in hectares), for each ignition point considered.</p>
Full article ">Figure 7
<p>Wildfire behavior aggregated maps of the study area: (<b>A</b>) ROS in typical conditions; (<b>B</b>) FLI in typical conditions; (<b>C</b>) ROS in atypical conditions; (<b>D</b>) FLI in atypical conditions.</p>
Full article ">
16 pages, 971 KiB  
Article
Revolutionizing Firefighting: UAV-Based Optical Communication Systems for Wildfires
by Mohammad Furqan Ali, Dushantha Nalin K. Jayakody and P. Muthuchidambaranathan
Photonics 2024, 11(7), 656; https://doi.org/10.3390/photonics11070656 - 11 Jul 2024
Viewed by 588
Abstract
Wildfires are one of the most devastating natural disasters in the world. This study proposes an innovative optical wildfire communication system (OWC) that leverages advanced optical technologies for wildfire monitoring and seamless communication towards the 5G and beyond (5GB) wireless networks. The multi-input–multi-output [...] Read more.
Wildfires are one of the most devastating natural disasters in the world. This study proposes an innovative optical wildfire communication system (OWC) that leverages advanced optical technologies for wildfire monitoring and seamless communication towards the 5G and beyond (5GB) wireless networks. The multi-input–multi-output (MIMO) optical link among communication nodes is designed by gamma–gamma (GG) distribution under consideration of intensity modulation and direct-detection (IM/DD) following an on–off-keying (OOK) scheme. In this study, the performance metrics of the proposed MIMO link that enables unmanned aerial vehicles (UAVs) are analytically derived. The end-to-end (E2E) performance metrics and the novel closed-form expressions for the average BER (ABER) and outage probability (Pout) are investigated for the proposed system models. Furthermore, the simulation results are obtained based on the real experimental data. The obtained results in this study are improved spatial resolution and accuracy, enabling the detection by communication of even small-scale wildfires at their inception stages. In the further perspective of this research, the development of the proposed system holds the potential to revolutionize wildfire prevention and control efforts, making a substantial impact on safeguarding ecosystems, communities, and economies from the devastating effects of fires. Full article
Show Figures

Figure 1

Figure 1
<p>The proposed optical wildfire detection communication (OWDC) system model, where the <span class="html-italic">n</span>th of UAVs share the recorded data with the <span class="html-italic">m</span>th number of base stations (BSs). The GG distribution designs the whole communication system model to optimize channel turbulence.</p>
Full article ">Figure 2
<p>The ABER performances while varying the altitude of an individual UAV. In this figure, it is depicted clearly that with increasing UAV altitude the performance decreases.</p>
Full article ">Figure 3
<p>The ABER performance is obtained and comprises a fluctuating number of communication nodes. It is depicted that by increasing the number of UAVs, ABER performance improves simultaneously.</p>
Full article ">Figure 4
<p>The ABER performance comparison among MIMO, MISO, SIMO, and SISO when the numbers of UAVs and BS vary. Moreover, the best ABER performance is depicted within MIMO and MISO wildfire detection systems.</p>
Full article ">Figure 5
<p>The outage performances are depicted on varying threshold SNR while the real-time data are used for July in Lisbon, Portugal. It is shown that the best performance is obtained at <math display="inline"><semantics> <mrow> <msub> <mi>γ</mi> <mrow> <mi>t</mi> <mi>h</mi> </mrow> </msub> <mo>=</mo> <mn>2</mn> </mrow> </semantics></math> dB.</p>
Full article ">Figure 6
<p>The outage probability performance for the successive altitude of UAV from the ground level while the other parameters are kept constant.</p>
Full article ">
21 pages, 4723 KiB  
Review
Research Trends in Wildland Fire Prediction Amidst Climate Change: A Comprehensive Bibliometric Analysis
by Mingwei Bao, Jiahao Liu, Hong Ren, Suting Liu, Caixia Ren, Chen Chen and Jianxiang Liu
Forests 2024, 15(7), 1197; https://doi.org/10.3390/f15071197 - 10 Jul 2024
Viewed by 598
Abstract
Wildfire prediction plays a vital role in the management and conservation of forest ecosystems. By providing detailed risk assessments, it contributes to the reduction of fire frequency and severity, safeguards forest resources, supports ecological stability, and ensures human safety. This study systematically reviews [...] Read more.
Wildfire prediction plays a vital role in the management and conservation of forest ecosystems. By providing detailed risk assessments, it contributes to the reduction of fire frequency and severity, safeguards forest resources, supports ecological stability, and ensures human safety. This study systematically reviews wildfire prediction literature from 2003 to 2023, emphasizing research trends and collaborative trends. Our findings reveal a significant increase in research activity between 2019 and 2023, primarily driven by the United States Forest Service and the Chinese Academy of Sciences. The majority of this research was published in prominent journals such as the International Journal of Wildland Fire, Forest Ecology and Management, Remote Sensing, and Forests. These publications predominantly originate from Europe, the United States, and China. Since 2020, there has been substantial growth in the application of machine learning techniques in predicting forest fires, particularly in estimating fire occurrence probabilities, simulating fire spread, and projecting post-fire environmental impacts. Advanced algorithms, including deep learning and ensemble learning, have shown superior accuracy, suggesting promising directions for future research. Additionally, the integration of machine learning with cellular automata has markedly improved the simulation of fire behavior, enhancing both efficiency and precision. The profound impact of climate change on wildfire prediction also necessitates the inclusion of extensive climate data in predictive models. Beyond conventional studies focusing on fire behavior and occurrence probabilities, forecasting the environmental and ecological consequences of fires has become integral to forest fire management and vital for formulating more effective wildfire strategies. The study concludes that significant regional disparities in knowledge exist, underscoring the need for improved research capabilities in underrepresented areas. Moreover, there is an urgent requirement to enhance the application of artificial intelligence algorithms, such as machine learning, deep learning, and ensemble learning, and to intensify efforts in identifying and leveraging various wildfire drivers to refine prediction accuracy. The insights generated from this field will profoundly augment our understanding of wildfire prediction, assisting policymakers and practitioners in managing forest resources more sustainably and averting future wildfire calamities. Full article
Show Figures

Figure 1

Figure 1
<p>A summary of the flowchart and study design.</p>
Full article ">Figure 2
<p>Number of publications in each year from 2003 to 2023.</p>
Full article ">Figure 3
<p>Cited journal network collaboration map.</p>
Full article ">Figure 4
<p>National cooperation network diagram.</p>
Full article ">Figure 5
<p>Research institution collaboration network diagram.</p>
Full article ">Figure 6
<p>Paper co-citation network analysis.</p>
Full article ">Figure 7
<p>Keyword co-occurrence network in wildland fire prediction research.</p>
Full article ">Figure 8
<p>Keyword cluster map for wildland fire prediction research.</p>
Full article ">Figure 9
<p>Top 25 burst keywords with the highest emergent intensity in wildland fire prediction research.</p>
Full article ">
28 pages, 16344 KiB  
Article
Operational Forest-Fire Spread Forecasting Using the WRF-SFIRE Model
by Manish P. Kale, Sri Sai Meher, Manoj Chavan, Vikas Kumar, Md. Asif Sultan, Priyanka Dongre, Karan Narkhede, Jitendra Mhatre, Narpati Sharma, Bayvesh Luitel, Ningwa Limboo, Mahendra Baingne, Satish Pardeshi, Mohan Labade, Aritra Mukherjee, Utkarsh Joshi, Neelesh Kharkar, Sahidul Islam, Sagar Pokale, Gokul Thakare, Shravani Talekar, Mukunda-Dev Behera, D. Sreshtha, Manoj Khare, Akshara Kaginalkar, Naveen Kumar and Parth Sarathi Royadd Show full author list remove Hide full author list
Remote Sens. 2024, 16(13), 2480; https://doi.org/10.3390/rs16132480 - 6 Jul 2024
Viewed by 1168
Abstract
In the present research, the open-source WRF-SFIRE model has been used to carry out surface forest fire spread forecasting in the North Sikkim region of the Indian Himalayas. Global forecast system (GFS)-based hourly forecasted weather model data obtained through the National Centers for [...] Read more.
In the present research, the open-source WRF-SFIRE model has been used to carry out surface forest fire spread forecasting in the North Sikkim region of the Indian Himalayas. Global forecast system (GFS)-based hourly forecasted weather model data obtained through the National Centers for Environmental Prediction (NCEP) at 0.25 degree resolution were used to provide the initial conditions for running WRF-SFIRE. A landuse–landcover map at 1:10,000 scale was used to define fuel parameters for different vegetation types. The fuel parameters, i.e., fuel depth and fuel load, were collected from 23 sample plots (0.1 ha each) laid down in the study area. Samples of different categories of forest fuels were measured for their wet and dry weights to obtain the fuel load. The vegetation specific surface area-to-volume ratio was referenced from the literature. The atmospheric data were downscaled using nested domains in the WRF model to capture fire–atmosphere interactions at a finer resolution (40 m). VIIRS satellite sensor-based fire alert (375 m spatial resolution) was used as ignition initiation point for the fire spread forecasting, whereas the forecasted hourly weather data (time synchronized with the fire alert) were used for dynamic forest-fire spread forecasting. The forecasted burnt area (1.72 km2) was validated against the satellite-based burnt area (1.07 km2) obtained through Sentinel 2 satellite data. The shapes of the original and forecasted burnt areas matched well. Based on the various simulation studies conducted, an operational fire spread forecasting system, i.e., Sikkim Wildfire Forecasting and Monitoring System (SWFMS), has been developed to facilitate firefighting agencies to issue early warnings and carry out strategic firefighting. Full article
(This article belongs to the Section Forest Remote Sensing)
Show Figures

Figure 1

Figure 1
<p>The study area location map. (India boundary map (1:16 m) source: Survey of India; Satellite data source: Sentinel 2, 29 January 2023 (standard False Color Composite)).</p>
Full article ">Figure 2
<p>The sampling scheme used to collect the fuel samples.</p>
Full article ">Figure 3
<p>WRF nested domains used for forest-fire spread forecasting (1 depict outermost and 3 depicts innermost domain).</p>
Full article ">Figure 4
<p>Methodology flow chart of forest-fire spread forecasting.</p>
Full article ">Figure 5
<p>Operational forest-fire spread simulation.</p>
Full article ">Figure 6
<p>Fuel map (1. Grassland, Grazing Land, Barren Rocky, Agriculture; 2. Open Scrub; 3. Dense Scrub; 4. Forest, Forest Plantation; 5. (No-Fuel) Built-up (Rural), Built-up (Urban), Core Urban, Gullied/Ravenous, Hamlets and Dispersed Households, Lakes/Ponds, Mixed Settlement, Peri-Urban, Reservoir/Tanks, River/Stream/Drain, Sandy, Snow/Glacial).</p>
Full article ">Figure 7
<p>Hourly forest-fire spread superimposed onto the Sentinel 2 False Color Composite (29 January 2023). The yellow line depicts the forecasted burnt area (1.72 km<sup>2</sup> (24th hour)), whereas the black line depicts the actual burnt area (1.07 km<sup>2</sup>). The red point indicates ignition initiation location.</p>
Full article ">Figure 8
<p>Forecasted incremental and hourly percentages of burnt area over 24 h. The incremental burn is the area growing each hour, whereas the hourly burn is the burn in that particular hour.</p>
Full article ">Figure 9
<p>Fuel map overlaid with contours (100 m intervals) depicted by solid black lines; the blue line depicts the forecasted burnt area (1.72 km<sup>2</sup>) and the dotted thick black line depicts the actual burnt area (1.07 km<sup>2</sup>), the red point depict fire initiation location. (1. Grassland, Grazing Land, Barren Rocky, Agriculture; 2. Open Scrub; 3. Dense Scrub; 4. Forest, Forest Plantation; 5. (No-Fuel) Built-up (Rural), Built-up (Urban), Core Urban, Gullied/Ravenous, Hamlets and Dispersed Households, Lakes/Ponds, Mixed Settlement, Peri-Urban, Reservoir/Tanks, River/Stream/Drain, Sandy, Snow/Glacial).</p>
Full article ">Figure 10
<p>(<b>a</b>–<b>x</b>) Wind simulation from 900 h (UTC) 27 January 2023 to 800 h (UTC) 28 January 2023. The red point depicts the fire ignition initiation location, the black lines depict the elevation contours, and the red lines indicate the hourly fire spread (1.72 km<sup>2</sup>).</p>
Full article ">Figure 10 Cont.
<p>(<b>a</b>–<b>x</b>) Wind simulation from 900 h (UTC) 27 January 2023 to 800 h (UTC) 28 January 2023. The red point depicts the fire ignition initiation location, the black lines depict the elevation contours, and the red lines indicate the hourly fire spread (1.72 km<sup>2</sup>).</p>
Full article ">Figure 11
<p>Windrose diagram for 24 h of simulation in the forecasted burnt region depicting the average wind speed and direction.</p>
Full article ">Figure 12
<p>The fire forecast parameter window.</p>
Full article ">Figure 13
<p>The visualization of fire spread in GIS. The SWFMS provides the flexibility to depict the fire spread in relation with other GIS layers. The open street map can also be opened in the background.</p>
Full article ">Figure 14
<p>Sending of SMS alerts to the stakeholders in the specified buffer zone.</p>
Full article ">
28 pages, 77027 KiB  
Article
A Framework for Conducting and Communicating Probabilistic Wildland Fire Forecasts
by Janice L. Coen, Gary W. Johnson, J. Shane Romsos and David Saah
Fire 2024, 7(7), 227; https://doi.org/10.3390/fire7070227 - 1 Jul 2024
Viewed by 723
Abstract
Fire models predict fire behavior and effects. However, there is a need to know how confident users can be in forecasts. This work developed a probabilistic methodology based on ensemble simulations that incorporated uncertainty in weather, fuel loading, and model physics parameters. It [...] Read more.
Fire models predict fire behavior and effects. However, there is a need to know how confident users can be in forecasts. This work developed a probabilistic methodology based on ensemble simulations that incorporated uncertainty in weather, fuel loading, and model physics parameters. It provided information on the most likely forecast scenario, confidence levels, and potential outliers. It also introduced novel ways to communicate uncertainty in calculation and graphical representation and applied this to diverse wildfires using ensemble simulations of the CAWFE coupled weather–fire model ranging from 12 to 26 members. The ensembles captured many features but spread was narrower than expected, especially with varying weather and fuel inputs, suggesting errors may not be easily mitigated by improving input data. Varying physics parameters created a wider spread, including identifying an outlier, underscoring modeling knowledge gaps. Uncertainty was communicated using burn probability, spread rate, and heat flux, a fire intensity metric related to burn severity. Despite limited ensemble spread, maps of mean and standard deviation exposed event times and locations where fire behavior was more uncertain, requiring more management or observations. Interpretability was enhanced by replacing traditional hot–cold color palettes with ones that accommodate the vision-impaired and adhere to web accessibility standards. Full article
(This article belongs to the Special Issue Probabilistic Risk Assessments in Fire Protection Engineering)
Show Figures

Figure 1

Figure 1
<p>Examples of different types of uncertainty outcomes. Modified from [<a href="#B26-fire-07-00227" class="html-bibr">26</a>].</p>
Full article ">Figure 2
<p>NCEP SREF forecasts for Oklahoma City (OKC) for 2 days with different weather types: 11 August 2022, at 15 UTC, a day with little precipitation (upper row) and 19 August 2022, at 15 UTC, a day with convective precipitation (lower row). The forecast fields are (<b>a</b>) 3 hourly precipitation, (<b>b</b>,<b>e</b>) 3 hourly temperature, (<b>d</b>) total cumulative precipitation, and (<b>c</b>,<b>f</b>) 3 hourly 10 m wind speed. Figures obtained from [<a href="#B43-fire-07-00227" class="html-bibr">43</a>].</p>
Full article ">Figure 3
<p>Example of the NAWFD fuel bed view for the Douglas fir—western hemlock existing vegetation type (EVT), showing computed probability density functions for each fuel bed stratum. Figure was generated using <a href="https://fuels.mtri.org" target="_blank">https://fuels.mtri.org</a>.</p>
Full article ">Figure 4
<p>The (<b>left</b>) fine woody debris (surface) fuel loads and (<b>right</b>) aggregate tree (canopy) fuel loads associated with the existing vegetation types present in the vicinity of the Caldor Fire from the North American Wildland Fuels Database. Each entry’s mean load for the rectangle sampled from (in Mg ha<sup>−1</sup>) is given in the center box. Figures were generated using <a href="https://fuels.mtri.org" target="_blank">https://fuels.mtri.org</a>.</p>
Full article ">Figure 5
<p>CAWFE ensemble members’ simulations of the Mosquito Fire, highlighting the variation among simulations due to different physics parameters, as specified in <a href="#fire-07-00227-t001" class="html-table">Table 1</a>. MQ01 was the control experiment, around which other configurations were varied. The red contour outlines the simulated fire extent at the conclusion of the 34 h simulation, on 9 September 2022, at 11 p.m. local time. The terrain is shown with contour intervals of 54 m. The simulation domain covers an area of 33.3 km by 33.3 km.</p>
Full article ">Figure 6
<p>VIIRS images used to initialize and evaluate progression in simulations of the Mosquito Fire. Panel (<b>a</b>) shows the fire on 8 September 2022 at 2036 UTC and (<b>b</b>) displays the fire’s state on 9 September 2022 at 1035 UTC, overlaid with ensemble member simulated fluxes at that time. The legend for satellite active fire detection data shown in (<b>a</b>) applies also to (<b>b</b>).</p>
Full article ">Figure 7
<p>CAWFE ensemble members’ simulations of the Mosquito Fire, each initialized by a different member of the Short-Range Ensemble Forecast (SREF) large-scale weather ensemble. The top 13 frames show simulations initialized by ARW-based members, while the bottom 13 frames show simulations initialized by NMB-based members. Each frame is labeled with the specific SREF member used for initialization. The red contour outlines the simulated fire extent at the end of the 34 h simulation period, on 9 September 2022, at 11 p.m. local time. Terrain contour intervals are 54 m, and the domain size is 33.3 km by 33.3 km.</p>
Full article ">Figure 8
<p>VIIRS data from the Mosquito Fire from (<b>a</b>) 8 September 2022, at 2036 UTC, the data used to initialize fire in progress, and (<b>b</b>) 9 September 2022, at 1035 UTC, over which lie ensemble member simulated fluxes initialized with 8 September 2022 15Z SREF members at that time.</p>
Full article ">Figure 9
<p>CAWFE ensemble simulations for the Tubbs Fire, initialized with varying SREF large-scale weather ensemble members. The top 13 frames represent ARW-based initializations, while the bottom 13 frames depict NMB-based initializations (details in text). The red contour delineates the simulated fire extent after 9 h of simulation, on 9 October 2017, at 4 a.m. local time. Terrain contours are shown at 31 m intervals in a domain spanning 26.7 km × 26.7 km.</p>
Full article ">Figure 10
<p>VIIRS data from the Tubbs Fire from 9 October 2017, at 1006 UTC (3:06 a.m. local time), over which lie ensemble member’s simulated fluxes at that time. The Tubbs Fire CAWFE ensemble (the 26 members are coded by color in figure legend) was initialized from its ignition point at 0443 UTC (9:43 p.m. local time).</p>
Full article ">Figure 11
<p>CAWFE ensemble members for simulations of the Caldor Fire, where each member is initialized with different fuel information, indicated in each frame. The top left figure uses mean NAWFD fire woody debris and tree fuel loads. The lower left figure is a simulation using the industry standard LANDFIRE fuel data. The top row uses mean fine woody debris fuel loads and (from left to right) minimum, 25th percentile, median, 75th percentile, and maximum tree fuel loads from NAWFD, respectively. The bottom row uses mean tree fuel loads and (from left to right) minimum, 25th percentile, median, 75th percentile, and maximum fine woody debris fuel loads, respectively. The red contour outlines the simulated fire extent after 23 h of simulation, on 17 August 2021, at 10 a.m. local time. Terrain contours are shown at 69 m intervals in a domain spanning 44.4 km × 44.4 km.</p>
Full article ">Figure 12
<p>VIIRS data from the Caldor Fire from (<b>a</b>) 16 August 2021, at 2150 UTC (2:50 p.m. local time), the data used to initialize fire in progress, and (<b>b</b>) 17 August 2021, at 0919 UTC, (2:19 a.m. local time) over which lie ensemble member simulated fluxes at that time. The legend for satellite active fire detection data shown in (<b>a</b>) applies also to (<b>b</b>).</p>
Full article ">Figure 13
<p>The burn probability (the percentage, 0–100%, in intervals of 10%) of ensemble members that burned each cell, shown at the end of the forecast period, for each ensemble: (<b>a</b>) Caldor Fire, (<b>b</b>) Mosquito Fire (physics-varying ensemble, (<b>c</b>) Mosquito Fire (SREF ensemble weather initialization), and (<b>d</b>) Tubbs Fire (SREF ensemble weather initialization). The color palette for (<b>a</b>–<b>d</b>) is shown in (<b>e</b>). This figure illustrates how burn probabilities were visualized, highlighting regions with varying likelihoods of burning based on ensemble forecasts.</p>
Full article ">Figure 14
<p>Montage of (<b>a</b>) mean spread rate, normalized to the maximum, identifying locations of rapid fire spread, and (<b>b</b>) standard deviation of the Mosquito Fire physics-varying ensemble, normalized to the maximum, at the end of the forecast period, indicating areas of uncertainty. (<b>c</b>) Layer shows the 95th percentile values for spread rate, normalized to the maximum, recorded within each cell at each timestep for the Mosquito Fire physics-varying ensemble, showing potential fire spread chutes. (<b>d</b>) The AAA compliance-rated color palette developed for spread rate, ensuring clarity and accessibility used in (<b>a</b>,<b>c</b>), where the range from 0.0 to 1.0 is divided into intervals of 0.1. Subfigure (<b>d</b>) shows three settings of transparency, which may be used as an additional accessibility tool. The standard deviation (<b>b</b>) uses the color palette shown in <a href="#fire-07-00227-f013" class="html-fig">Figure 13</a>e. This figure provides insights into the dynamics of fire spread and identifies regions of high uncertainty and potential rapid expansion.</p>
Full article ">Figure 15
<p>(<b>a</b>) The ensemble mean total (over the forecast) heat flux for the Mosquito Fire physics-varying ensemble, indicating locations of ensemble agreement on intense burning and heat release. (<b>b</b>) The standard deviation in heat flux, showing areas of uncertainty in the fire’s behavior. The values in (<b>a</b>,<b>b</b>) range from 0.0 to 1.0, each normalized to the maximum, divided into intervals of 0.1. Color schemes selected for displaying heat flux mean, representing a temperature effect, and variability, showing how transparency can be used to maintain clarity and accessibility, are presented in (<b>c</b>,<b>d</b>), respectively.</p>
Full article ">
26 pages, 8396 KiB  
Article
Data-Driven Wildfire Spread Modeling of European Wildfires Using a Spatiotemporal Graph Neural Network
by Moritz Rösch, Michael Nolde, Tobias Ullmann and Torsten Riedlinger
Fire 2024, 7(6), 207; https://doi.org/10.3390/fire7060207 - 19 Jun 2024
Viewed by 836
Abstract
Wildfire spread models are an essential tool for mitigating catastrophic effects associated with wildfires. However, current operational models suffer from significant limitations regarding accuracy and transferability. Recent advances in the availability and capability of Earth observation data and artificial intelligence offer new perspectives [...] Read more.
Wildfire spread models are an essential tool for mitigating catastrophic effects associated with wildfires. However, current operational models suffer from significant limitations regarding accuracy and transferability. Recent advances in the availability and capability of Earth observation data and artificial intelligence offer new perspectives for data-driven modeling approaches with the potential to overcome the existing limitations. Therefore, this study developed a data-driven Deep Learning wildfire spread modeling approach based on a comprehensive dataset of European wildfires and a Spatiotemporal Graph Neural Network, which was applied to this modeling problem for the first time. A country-scale model was developed on an individual wildfire time series in Portugal while a second continental-scale model was developed with wildfires from the entire Mediterranean region. While neither model was able to predict the daily spread of European wildfires with sufficient accuracy (weighted macro-mean IoU: Portugal model 0.37; Mediterranean model 0.36), the continental model was able to learn the generalized patterns of wildfire spread, achieving similar performances in various fire-prone Mediterranean countries, indicating an increased capacity in terms of transferability. Furthermore, we found that the spatial and temporal dimensions of wildfires significantly influence model performance. Inadequate reference data quality most likely contributed to the low overall performances, highlighting the current limitations of data-driven wildfire spread models. Full article
Show Figures

Figure 1

Figure 1
<p>Burned area variables within the H3 grid for an example wildfire AOI. (<b>a</b>) Time series of the total burned area (variable “<span class="html-italic">burned</span>”). (<b>b</b>) Time series of the daily new burned cells displaying the wildfire propagation (variable “<span class="html-italic">burned_new</span>”).</p>
Full article ">Figure 2
<p>Datacube representation of a wildfire time series containing all H3 cells (c) (1st dimension), with all static and dynamic features (X) (2nd dimension) over the time steps (t) (3rd dimension). The datacube can also be represented as an ordered sequence of two-dimensional H3 grids over time.</p>
Full article ">Figure 3
<p>Different data representations of the burned area of a wildfire in Portugal. Background: Sentinel-2 RGB image from the 07.08.2020. (<b>a</b>) Burned area perimeter derived by the Sentinel-3 mapping algorithm. (<b>b</b>) Burned area perimeter displayed in H3 cells (resolution 9). (<b>c</b>) Burned area perimeter displayed as a graph.</p>
Full article ">Figure 4
<p>Schematic display of the STGNN wildfire spread model and its subcomponents. (<b>a</b>) Workflow of the STGNN model. (<b>b</b>) Spatial subcomponent of the TGCN model with a schematic representation of the graph convolution process of the GCN (<b>b.1</b>,<b>b.2</b>) and the GCN layer stacking (<b>b.3</b>). (<b>c</b>) Temporal subcomponent of the TGCN with a schematic representation of the GRU cell.</p>
Full article ">Figure 5
<p>Model workflow in training and testing mode.</p>
Full article ">Figure 6
<p>Wildfire spread prediction of the Portugal model for an example test fire in 2019.</p>
Full article ">Figure 7
<p>Overall performance of the Portugal model per fire season.</p>
Full article ">Figure 8
<p>Overall performance of the Mediterranean model per country. The number in each country refers to the respective number of wildfires in the Mediterranean reference dataset.</p>
Full article ">Figure 9
<p>Overall model performance of the Mediterranean model per country and fire season.</p>
Full article ">Figure 10
<p>Model performance per daily wildfire spread size (number of new burned H3 cells). Each predicted wildfire is represented as a point. The blue line represents the trend line with the 95% confidence level interval (grey). (<b>a</b>) Portugal model. (<b>b</b>) Mediterranean model.</p>
Full article ">Figure 11
<p>Model performance per prediction day. (<b>a</b>) Portugal model. (<b>b</b>) Mediterranean model.</p>
Full article ">
15 pages, 3788 KiB  
Article
Wildfire Susceptibility Prediction Based on a CA-Based CCNN with Active Learning Optimization
by Qiuping Yu, Yaqin Zhao, Zixuan Yin and Zhihao Xu
Fire 2024, 7(6), 201; https://doi.org/10.3390/fire7060201 - 16 Jun 2024
Viewed by 498
Abstract
Wildfires cause great losses to the ecological environment, economy, and people’s safety and belongings. As a result, it is crucial to establish wildfire susceptibility models and delineate fire risk levels. It has been proven that the use of remote sensing data, such as [...] Read more.
Wildfires cause great losses to the ecological environment, economy, and people’s safety and belongings. As a result, it is crucial to establish wildfire susceptibility models and delineate fire risk levels. It has been proven that the use of remote sensing data, such as meteorological and topographical data, can effectively predict and evaluate wildfire susceptibility. Accordingly, this paper converts meteorological and topographical data into fire-influencing factor raster maps for wildfire susceptibility prediction. The continuous convolutional neural network (CCNN for short) based on coordinate attention (CA for short) can aggregate different location information into channels of the network so as to enhance the feature expression ability; moreover, for different patches with different resolutions, the improved CCNN model does not need to change the structural parameters of the network, which improves the flexibility of the network application in different forest areas. In order to reduce the annotation of training samples, we adopt an active learning method to learn positive features by selecting high-confidence samples, which contributes to enhancing the discriminative ability of the network. We use fire probabilities output from the model to evaluate fire risk levels and generate the fire susceptibility map. Taking Chongqing Municipality in China as an example, the experimental results show that the CA-based CCNN model has a better classification performance; the accuracy reaches 91.7%, and AUC reaches 0.9487, which is 5.1% and 2.09% higher than the optimal comparative method, respectively. Furthermore, if an accuracy of about 86% is desired, our method only requires 50% of labeled samples and thus saves about 20% and 40% of the labeling efforts compared to the other two methods, respectively. Ultimately, the proposed model achieves the balance of high prediction accuracy and low annotation cost and is more helpful in classifying fire high warning zones and fire-free zones. Full article
Show Figures

Figure 1

Figure 1
<p>Location of Chongqing Municipality, in China, and distribution of wildfires in 2017.</p>
Full article ">Figure 2
<p>The model framework for the wildfire susceptibility prediction.</p>
Full article ">Figure 3
<p>Raster map of the average temperature in Chongqing, China, in 2017.</p>
Full article ">Figure 4
<p>ROC curves and AUC of the five methods.</p>
Full article ">Figure 5
<p>Radar maps of metrics derived from six different models on the validation set.</p>
Full article ">Figure 6
<p>Wildfire susceptibility maps derived from different methods for Chongqing Municipality, China, in 2017. (<b>a</b>) Our method; (<b>b</b>) CNN-based method; (<b>c</b>) RF; (<b>d</b>) Decision Tree; (<b>e</b>) MLP; (<b>f</b>) SVM.</p>
Full article ">Figure 7
<p>Classification accuracy with different percentages of labeled samples.</p>
Full article ">
Back to TopTop