Open access peer-reviewed chapter

Remote Sensing and GIS Applications in Wildfires

Written By

Georgios Zagalikis

Submitted: 22 January 2023 Reviewed: 06 February 2023 Published: 24 October 2023

DOI: 10.5772/intechopen.111616

From the Edited Volume

Geographic Information Systems - Data Science Approach

Rifaat Abdalla

Chapter metrics overview

269 Chapter Downloads

View Full Metrics

Abstract

Wildfires are closely associated with human activities and global climate change, but they also affect human health, safety, and the eco-environment. The ability of understanding wildfire dynamics is important for managing the effects of wildfires on infrastructures and natural environments. Geospatial technologies (remote sensing and GIS) provide a means to study wildfires at multiple temporal and spatial scales using an efficient and quantitative method. This chapter presents an overview of the applications of geospatial technologies in wildfire management. Applications related to pre-fire conditions management (fire hazard mapping, fire risk mapping, fuel mapping), monitoring fire conditions (fire detection, detection of hot-spots, fire thermal parameters, etc.) and post-fire condition management (burnt area mapping, burn severity, soil erosion assessments, post-fire vegetation recovery assessments and monitoring) are discussed. Emphasis is given to the roles of multispectral sensors, lidar and evolving UAV/drone technologies in mapping, processing, combining and monitoring various environmental characteristics related to wildfires. Current and previous researches are presented, and future research trends are discussed. It is wildly accepted that geospatial technologies provide a low-cost, multi-temporal means for conducting local, regional and global-scale wildfire research, and assessments.

Keywords

  • wildfires
  • GIS
  • remote sensing
  • multispectral sensors
  • lidar
  • UAV
  • mapping

1. Introduction

Forest fires (term used in Europe to designate the unwanted fires burning forests and wildlands) are considered as one of the main environmental hazards worldwide. Forest fires are periodic disturbance events that dramatically affect the structure and distribution of global forest ecosystems, altering soil erosion, causing loss of biodiversity, habitat, production and productivity, endangering human life, and disrupting livelihoods [1].

The impacts of climate change are becoming more evident each year. Forest fires are more frequent and extreme, as also confirmed by the recent report of the International Panel on Climate Change (IPCC) “Climate Change 2022: Impacts, adaptation and vulnerability” [2] and the recent United Nations Environment Program report “Spreading like wildfire: The rising threat of extraordinary landscape fires” [3].

In Mediterranean basin, natural forest fires are common and are an integral part of the terrestrial ecosystems (Figure 1). Fire has been used by man as management tool since early times. Mediterranean region is the most affected area from wildfires in Europe. Approximately 85% of the total burnt area occurs in the EU Mediterranean region [4]. This vulnerability of Mediterranean basin to be affected by fire is linked to its climate, which is characterized by rainy and mild winters followed by warm and dry summers [5]. Extreme weather conditions in summer (high temperature, strong wind, low relative humidity and drought) are a key factor in the ignition and spread of large forest fires [6, 7, 8].

Figure 1.

Forest fires in the pan-European region in 2021 © EU, 2022.

Geospatial technologies such as remote sensing (RS) and geographic information systems (GIS) have been used for the management of forest fires several years ago with the use of aerial photographs and have been increasing as new type of sensors become available varying from spaceborne sensors to the newly era of UAVs sensors. Remote sensing provides the data used through the GIS or other decision support systems.

Three main levels can be distinguished, in which remote sensing and GIS provide results that can be applied directly to the subject of forest fires: pre-fire conditions (fire risk assessment, vegetation mapping, topography, risk of fire spreading, etc.), monitoring fire conditions (fire detection, detection of hotspots, fire thermal parameters, etc.) and post-fire (burnt area mapping, success of forest regeneration, flood risk, etc.). Different types of remote sensing platforms and sensors are used for data acquisition as satellites to UAVs and active sensors (SAR, Lidar) to passive sensors (optical, thermal infrared, multispectral).

The purpose of this chapter is to present a wide outline of the use of remote sensing and GIS in wildfire management. The chapter examines research in the three main levels and within each level, and sub-level, current research trends and case examples are presented, divided by the platform and sensor type used. The goal of this chapter is to provide the reader with a broad view of the current applications of remote sensing and GIS technologies and techniques in the field of wildfire management.

Advertisement

2. Pre-fire condition management

A forest fire results from a complex interaction of biological, meteorological, physical and social factors that influence the likelihood of a forest fire breaking out, its propagation and intensity, duration and extent, and its potential to cause damage to economies, the environment, and society [3]. Terms such as fire hazard, fire risk or danger and severity have been introduced to define risk, hazard and the characteristic (or uncharacteristic) nature of forest fires [9].

Fire risk or danger is a term used to represent the vulnerability of an area to the ignition and spread of fire. This fire vulnerability is influenced by various factors that are dynamic and static in both spatial and temporal dimensions; therefore, it is a complex concept to measure [10]. Furthermore, fire risk factors are both biotic and abiotic [11]. Biotic conditions are aimed at studying the morphological (i.e., fuel biomass load, species composition, height and density) and physiological (i.e., moisture status and chemical properties) characteristics of vegetation, while abiotic indicators are related to the study of external conditions such as topography and meteorological factors on the other hand [11, 12].

Fire risk assessments involve identifying the potentially contributing variables and integrate them into mathematical expression known as “index” [13]. Several risk indices have been developed for fire risk evaluation based on either the temporal scale [14, 15] or the variable data sources [16].

Mapping is used to visualize the fire risk assessments, and this is accomplished through two primary methods: (i) point-wise meteorological data-based operating systems and (ii) the use of remote sensing technologies and geographic information systems (GIS) [17].

The early methods, which assess fire danger, were based on meteorological factors since they have significant role in the occurrence of severe fire episodes [18, 19], including the Nesterov index for use in the former Soviet Union [20], the Forest Fire Danger Index (FFDI) for eastern Australia [21], the National Fire Danger Rating System for the USA [22] and the Canadian Forest Fire Weather Index (FWI) System [23]. These systems use meteorological data such as temperature, precipitation, humidity and wind speed to assess fire danger over extent geographic regions. However, these systems suffer from several limitations as the need for an extent network of weather stations or the need for interpolation to generate fire danger maps. Since the geoinformatic technologies as remote sensing/GIS have the ability to obtain continuous data for an area, the ability to analyze those data is recognized as an effective alternative or addition for the creation of fire risk maps [15, 24].

2.1 Fire hazard mapping

Fire hazard divers from fire danger as it does not include a meteorological component in the assessment [9]. Fire hazard or susceptibility maps are important in land use planning [25, 26], and that maps are useful as a valuable reference to reduce vulnerability and could help to improve decision-making planning in ecological risk prevention [27]. Another potential use is in the insurance markets [28].

Several pre-fire conditions used in the fire hazard maps can be monitored using remote sensing [29]. Vegetation mapping (including fuel loads) is one of the first pre-fire conditions contributing to fire hazard models that are produced usually from high spatial resolution optical or radar images [30, 31]. The fuel maps may range from global-continental products derived from coarse-resolution sensors (AVHRR, Modis, Vegetation) to regional-local inventories, based on higher-resolution data (Landsat TM/ETM, radar or even Ikonos) [32].

Description of fuel properties is critical in all phases of fire management and fire science: prevention, fire danger estimation; suppression: fire behavior modeling; fire effect assessment: trace gas emissions and vegetation recovery after fire [33]. Several fuel characteristics are critical for fire propagation studies: crown bulk density, crown base height, canopy height, canopy closure, surface area to volume ratio, vertical and horizontal continuity, dead and live fuel loads, live woody loads and size of particles referring to vegetation geometry to the fuel type, which can be mapped, like classical vegetation mapping, from high spatial resolution optical or radar images [34].

Depending on the remote sensing data sources used different vegetation characteristics can be interoperated because of the different spatial, spectral and temporal resolution (Table 1).

SensorSpatial resolutionTemporal resolutionAdvantagesDisadvantagesInfo
Landsat MSS,TM, ETM+, OLI15–30 m16 daysFree and easily accessibleLack of canopy penetration, low temporal resolutionhttps://landsat.gsfc.nasa.gov/satellites/landsat-8/
Sentinel-210–60 m5 daysFree, relatively high spatial and temporal resolution, multiple near-infrared (NIR) bandsLack of canopy penetrationhttps://sentinels.copernicus.eu
MODIS250 m–1 km1 to 2 daysFree and easily accessible, high temporal resolution, large area analysisLack of canopy penetration, coarse spatial resolution limits analysis of smaller areashttps://modis.gsfc.nasa.gov/about/
Pleiades Neo0.3–1.2 m1–3 per dayHigh spatial resolutionHigh costhttps://earth.esa.int/eogateway/missions/pleiades-neo
AVIRIS4–20 mAirborneHigh spatial resolution, hyperspectral sensorHigh cost, complicated data processinghttps://aviris.jpl.nasa.gov/
AVIRIS-NG0.3–4.0 mAirborneHigh spatial resolution, hyperspectral sensorHigh cost, complicated data processinghttps://avirisng.jpl.nasa.gov/aviris-ng.html
Hyperion10–30 m16 daysFree and easily accessible, hyperspectral sensorDecommissioned 2017https://www.usgs.gov/centers/eros/science/usgs-eros-archive-earth-observing-one-eo-1-hyperion
UAVHigh centimeter-levelAirborneHigh spatial resolution, hyperspectral sensor, lidar sensorsCost it depends from the sensor, low endurance for areas 1–10 km2

Table 1.

List of sensors commonly used or with potential usage in forest fires research.

Multispectral remote sensing provides fundamental inputs for the mapping of fire hazard [15, 35]. Land cover type plays an important role in evaluating the risk of a given area to fire events as it is related to fuel types and characteristics; usually, classification is fulfilled using vegetation indices. The simplest form of vegetation index is simply a ratio between two digital values from independent spectral bands. The Normalized Difference Vegetation Index (NDVI=NIR R/NIR + R), the most commonly used ratio transformation for vegetation studies, is used to monitor photosynthetic activity and provides information on vegetation biomass and phenology [36, 37, 38]. Biomass is also estimated using various remote sensing data [39, 40].

Usually low or medium spatial resolution multispectral sensors such as Landsat-TM [41, 42] or MODIS-ASTER [43, 44], and hyperspectral sensors, such as Hyperion [45], AVIRIS [46] or MIVIS (Multispectral Infrared Visible Imaging Spectrometer) [47] and lately the Sentinel-2A [48].

High resolution commercial satellites, such as Quickbird, Pleiades-1A, IKONOS, GeoEye-1, WorldView-2,3, have become operational over the past decades, offering data at finer than 10-m spatial resolution have also been used for fuel mapping [44, 49]. Arroyo et al. [50] explored the utility of high resolution (4 m2) satellite imagery for the production of fuel type maps using a different data processing approach and object-oriented classification, which allows for considering spatial context of adjacent pixels in the eventual generation of the fuel map. It was concluded that this approach produced higher accuracies than would have been generated using traditional maximum-likelihood classifiers. Such data can be used for pre-fire studies, but hurdles such as high cost per scene and the limitation common to most of optical sensors, the inability to penetrate forest canopies [51].

Beyond optical remote sensing satellites active sensors such as SAR and Lidar have been used to estimate vegetation characteristics that are critical for fuel type mapping, such as foliar biomass, tree volume, tree height and canopy closure [52, 53]. Lidar data is used in combination with other remote sensing data airborne or spaceborne to estimate vegetation characteristics [49, 54].

Remote sensing provides also topographic data as slope, aspect and elevation that all influence the risk of an area to fire ignition and/or spread [55]. In terms of terrain factors, altitude and slope can affect the moisture loss of vegetation, and in contrast, the aspect affects the amount of solar radiation received by vegetation, directly affecting the vegetation drying degree. For forests, terrain differences lead to differences in wind, water balance and heat transfer between different areas, affecting the spread of fire [56].

Another pre-fire condition that remote sensing systems provide is the historical statistical data about the date and time of the fire, fire location (administrative district), fire size by land use category and fire cause [57, 58]. Historical data can be used to evaluate the accuracy of the fire hazard maps or the fire risk maps and the models used. Examples of remote-sensed databases of fire activity include the World Along Track Scanning Radiometer (ATSR) Fire Atlas [59], the MODIS and the VIIRS Active Fire Products [60] and the LSA SAF Fire Products [61].

In the last decade, unmanned aerial vehicles (UAVs) have become more affordable and one of the most important remote sensing data sources in various natural resource management applications [62, 63, 64]. UAV can use a variety of sensors, including visible light, near infrared (NIR), shortwave infrared (SWIR), thermal infrared (TIR), SAR and Lidar sensors. UAV optical sensors, including visible, NIR and SWIR, also record data as multispectral or hyperspectral bands [65, 66]. Since the sensor technology advances, increasingly smaller, lighter and cheaper sensors have become available for UAV remote sensing applications. Manfreda et al. [67] and Constantine and Rey [68] provide detailed introductions to various UAV remote sensing systems. Also, Dainelli et al. [69] provide a systematic review analysis of unmanned aerial vehicle (UAV)-based forestry research papers.

Remote sensing provides data about proximity to human settlement/infrastructure to identify areas at risk of fire occurrence. Forest fires in the Mediterranean Europe and in Mediterranean basin in general are mostly related to human activities. More than 90% of fires originate from either deliberate or involuntary causes [70] Mapping ignition risk is an important tool for managers, helping to improve the effectiveness of fire prevention, detection and firefighting resource allocation [71, 72].

In Mediterranean basin, in California and in Australia, electric power infrastructure has ignited several of the most destructive wildfires in recent history, even if there is a gap in published research work on wildfire prediction when the root cause of the fire is linked to the power grid infrastructure—despite the fact that some of the most devastating wildfires were sparked by power grids [73]. Remote sensing (RS) and geographic information system (GIS) technologies and their derived applications are effective tools for power line surveillance and disaster prevention [74]. UAVs can be used for detailed map of electric power grid and for identifying utility systems that pose a fire risk [75, 76].

Several methods have been used widely for forest fire occurrence. Based on the literature, there are two main methods and techniques for this aim, namely, data-based (machine learning (ML)) and knowledge-based methods [77]. Numerous studies used knowledge-based methods for forest fire susceptibility mappings, such as fuzzy logic [78], the analytic hierarchy process (AHP) [79] and the analytical network process [80]. In recent years, an obvious trend in the use of machine learning algorithms is in a wide range of natural hazard assessment disciplines [81, 82]. Machine learning (ML) methods, such as logistic regression [83], artificial neural networks [84, 85] and Random Forest [86, 87], are also used for forest fire hazard mapping.

An example of fire hazard mapping is Fire Hazard Severity Zones Maps in California USA and it is produced by State of California and the Department of Forestry and Fire Protection. The Fire Hazard Severity Zone (FHSZ) maps are developed using a science-based and field-tested model that assigns a hazard score based on the factors that influence fire likelihood and fire behavior. Many factors are considered such as fire history, existing and potential fuel (natural vegetation), predicted flame length, blowing embers, terrain, and typical fire weather for the area. There are three levels of hazard in the State Responsibility Areas: moderate, high and very high (Figure 2) [88].

Figure 2.

California fire Hazard severity zone viewer web mapping application.

2.2 Fire risk mapping

Fire risk mapping also includes dynamic variables such as weather and vegetation conditions. The result is a map that displays varying degrees of fire risk ranging from very low to very high. Fire risk maps in short term normally provide risk estimates that are only appropriate for a short period (days-weeks) after their creation. They use many of the same inputs as fire hazard maps but also include variables that are continuously changing such as fuel moisture content, weather conditions and vegetation conditions [9, 89]. Live fuel moisture content (LFMC) is one of the critical factors affecting fire ignition and fire propagation, and therefore is taken into account in most fire danger and fire behavior modeling systems [90]. Dead fuel moisture conditions also affects fire ignition and fire propagation and the computation is easily estimated from weather data and fuel characteristics, because dead fuel moisture is in balance with that of the surrounding atmosphere [91, 92]. Several studies conducted in a wide range of ecosystems have found a significant correlation between burned area and LFMC [93, 94]. Accurate and comprehensive spatial and temporal estimations of LFMC are essential to assess wildfire danger [95] and to develop early warning systems for the evolution of critical conditions [96]. Field-based estimations of LFMC are labor and time intensive as well as costly, and they cover small areas and are difficult to be frequent.

Remote sensing methods cover large areas and this is why remotely sensed images have been used extensively for estimation of the LFMC using various methods, which may be generally classified into statistical (empirical) [97, 98], physical model-based approaches [99, 100] and recently machine learning approaches [101, 102].

The use of satellite data in LFMC estimation has been focused on coarse spatial resolution remote sensing data such as moderate resolution Imaging Spectroradiometer (MODIS) and Advanced Very High-Resolution Radiometer (AVHRR). Combinations of spectral indices have been successfully employed to estimate LFMC [94, 97, 103] usually including the reflectance in the near infrared and the reflectance in the 1.6-μm region [104, 105]. Other studies implement also thermal data to estimate LFMC [106], different satellite data (Sentinel-2) in combination with meteorological variables or without [107, 108] and microwave remote sensing [102, 109, 110]. Yebra et al. [111] reviewed the use of remotely sensed data for estimating LFMC with particular interesting toward the operational use of LFMC products for fire risk assessment.

Fire Weather Index (FWI) module of the Canadian Forest Fire Danger Rating System (CFFDRS) is widely used in several countries such as Alaska, USA [112]; Indonesia and Malaysia [113]; Portugal [114]; Spain [115]; and others. It is also part of the European Forest Fire Information System (EFFIS) (Figure 3) [116]. The FWI System has six components rating fuel moisture content and potential fire behavior in a common fuel type (i.e., mature pine stand) and in no slope conditions.

Figure 3.

Forest fire danger forecast in the pan-European region for 02 Jun 2022 © EU, 2022.

Calculations are based on daily noon measurements of air temperature, relative humidity, wind speed and previous 24-h precipitation. Despite the global acceptance of the FWI, it has an inherent problem in delineating the spatial dynamics of the danger conditions, as it employs geographic information system (GIS)-based interpolation techniques [117]. Sirca et al. [118] assessing the performance of several fire danger indexes in the Mediterranean area concluded that FWI yielded better predictions than the other indexes.

In Italy, the Civil Protection adopted the RISICO model, which is a fire danger rating system that was developed specifically for the vegetation cover of the Mediterranean [119]. RISICO integrates meteorological observations and forecasts from an NWP (Numerical Weather Predication) Limited Area Model (LAM) and ECMWF-Integrated Forecasting System (IFS) with vegetation cover and topography data. Another project developed in Sardinia, Italy called Daily Fire Hazard Index (DFHI) [120], the index is computed using satellite images and the meteorological data.

Advertisement

3. Fire detection and monitoring fire conditions

Fire detection is the earliest critical stage of forest fires management, which is aimed at either fighting or monitoring the fire. For firefighting the early detection is essential; so far, fire detection is based on human observation, the use of fixed optical cameras to monitor the surrounding environment or aerial survey [29]. The temporal and spatial analysis capabilities provided by current satellite sensors are not considered sufficient for firefighting operations by forest fire mangers.

3.1 Spaceborne multispectral sensors

Several of geo-stationary satellite sensors have been used for active fire detection and monitoring including but not only: Advanced Very High-Resolution Radiometer (AVHRR), Meteosat Second Generation, Spinning Enhanced Visible and Infra-red Imager (MSG-SEVIRI), Himawari-8, Geostationary Operational Environmental Satellite (GOES), DMSP OLS (US Air Force Defence Meteorological Satellite Programme – Operational Line-scan System), Visible Infrared Imaging Radiometer Suite (VIIRS) and moderate resolution imaging spectroradiometer (MODIS) [121, 122, 123]. Early attempts of active fire detection and monitoring started during 80′ with the US Forest Service using NOAA satellites to identify forest fires in the western US [124]. AVHRR was the primary sensor system for active fire detection and monitoring until the eventual launch during 1999 of the MODIS sensors onboard the Terra and Aqua platforms [125]. MODIS Terra and Aqua have become the primary sensors at regional to global scales for active fire detection and monitoring due to their high temporal resolution and special channels designed for fire monitoring. The Aqua MODIS instrument acquires data twice daily (1:30 PM and AM), as does the Terra MODIS (10:30 AM and PM). These four daily MODIS fire observations serve to advance global monitoring of fire processes and their effects on ecosystems, the atmosphere and climate [126]. Serval operational fire monitoring systems using MODIS active fire detection include the Canadian Wildland Fire Information System (CWFIS) (http://cwfis.cfs.nrcan.gc.ca), the USA Active Fire Mapping Service or the European Forest Fire Information System (EFFIS) (http://effis.jrc.ec.europa.eu). In the case of EFFIS, post-processing filters based on landcover ancillary data are applied to the MODIS product to reduce the number of false alarms produced by non-fire hot surfaces (e.g., industrial areas, hot ground soils) and therefore increase the reliability of the active fire detection [116]. Figure 4 shows active fires monitoring in the wider Mediterranean region.

Figure 4.

Active fires from MODIS during June 2022 © EU, 2022.

Several studies found that MODIS fire detection algorithms had trouble detecting smaller/cooler fires and frequently detected false alarms [127, 128]. Validation of MODIS and AVHRR active fire products used higher spatial resolution sensors such as Landsat TM and ASTER to evaluate the performance of the MODIS and AVHRR products [129, 130]. The accuracy of fire detection is calculated in terms of commission and omission errors as well as error distributions over various locations. Commission and omission errors have been reported by many studies errors of omission represent “failure to detect fires” and errors of commission represent “false alarms or false positives” pixels identified as fires [130]. Schroeder et al. [127] found errors of commission of approximately 35% over areas of active deforestation, and similarly, Forghani et al. [130] report that overall commission errors of MODIS and AVHRR hotspots over the 5% sample data were 15 and 68%, respectively, and overall omission errors of MODIS and AVHRR hotspots were 17 and 23%, respectively. An interesting comparison of the suitability of some of geo-stationary satellite sensors (AVHRR, MODIS, DMSPOLS) for fire detection was carried out by Cahoon et al. [131]. They found in their study that the minimum size for a detectable fire was approximately 213 m2 for MODIS, 435 m2 for AVHRR and 45 m2 for DMSP-OLS, for a nominal fire front temperature of 1000 K and a background temperature of 305 K. Maier et al. [132] determined a minimum fire size of 100–300 m2 for the detection by the MODIS algorithm. Evaluation of the MODIS active fire product to quantify detection rates of both Terra and Aqua MODIS sensors was carried out by Hawbaker et al. [133]. Finally, MODIS had the ability to capture large fires in the US, but may under represent fires in areas with high cloud, rapid burning or small- and low-intensity fires that are often undetected.

To address these limitations found during evaluation of the MODIS fire detection algorithm, NASA’s Earth Observing System periodically reprocesses the raw instrument data archive using newer fire detection algorithms. In the MODIS version 3 active fire detection algorithm’s sensitivity to small fires was sacrificed to reduce false alarms over certain surface types during the day [134]. The MODIS 4, contextual fire detection algorithm was enhanced, which increased the ability to detect small cool fires [60]. For the recently released Collection 6 MODIS fire products, bands 1 (red), 2 (near infrared), 7 (short-wave infrared), 21 (thermal infrared over a wide-range centered on 4 μm), 22 (thermal infrared in a narrow-range centered on 4 μm), 31 (thermal infrared at 11 μm) and 32 (thermal infrared at 12 μm) are used to identify 1-km “fire pixels” that contain at least one active fire at time of acquisition [128].

Further than MODIS, recent studies have used data from the Landsat 8 Operational Land Imager (OLI) [135] and Sentinel-2 sensors [136]. Also, the use of geostationary weather satellites and their ability to detect active fires have been a research topic of great interest for several years [137, 138, 139].

Active fires can be detected by a wide variety of spectral sensors, but as active fire products are usually used to monitor ongoing fires and the ideal sensors for detecting them will possess a high temporal resolution; at the same time that kind of sensor have low spatial resolution. While sensors with higher spatial resolution like Landsat satellites and Sentiel-2 satellites work well for detecting active fires, their low temporal resolution makes them less than ideal. This is the reason why most active fire products are generated from sensors such as MODIS and VIIRS, as these sensors have a very short revisit time (<2 days) allowing for the active fire products to be rapidly updated [17]. Geostationary satellites have even higher temporal resolution, with updates every few hours or less. However, these high temporal resolutions come at the cost of lower accuracies due to the omission of smaller/cooler fires and commission errors caused by activities such as deforestation [127, 132, 140].

3.2 UAVs

The MODIS satellite imagery with the high temporal resolution (1–2 days revisiting time) is commonly applied in forest fires detection and monitoring. However, the low spatial resolution of MODIS is insufficient for this task at local scales [65]. During the last decade, we have seen a great progress in the field of using unmanned aerial vehicles (UAVs) for forest fire monitoring, detection and even fighting [141]. UAVs have become smaller, more affordable and now have better computation capabilities than in the past making them reliable tools for remote sensing missions in hostile environments [142]. UAVs provide rapid, mobile and cost-effective sensor systems that are currently being adopted for fire detection and monitoring [143, 144]. In their review of UAVs applications for forest fire monitoring, detection and fighting, Yuan et al. [144] define fire monitoring as involving the active search for possible fire occurrences, while fire detection involves the identification of fires in progress. UAVs used for fire monitoring and detection rely on sensors operating in the visible and thermal infrared wavelengths. In a more recent review of UAVs usage in forest fires, Akhloufi et al. [145] present a survey of different approaches for the development of UAV fire assistance systems. They conclude that UAVs can play an important role in the fight against wildland fires in large areas and with the decrease in their prices and their wider commercial availability, new applications in this field will emerge. However, limitations remain such as autonomy, reliability and fault tolerance; also security issues are a concern, as there are risks associated with having UAVs flying over firefighters or close to aircraft carrying water and fire retardants [145].

Advertisement

4. Post-fire condition management

The application of remotely sensed imagery to monitor and assess the impacts of fire on local and regional environments can be broadly divided into several applications such as burnt area mapping, burn severity assessments, vegetation recovery monitoring, effects on soil erosion, and general fire effects on air quality, soil, vegetation and fauna.

4.1 Burnt area mapping

Remotely sensed data have been extensively used for burnt area mapping at local, regional and global scales [146, 147]. Burned area mapping is critical importance for forest managers, climate scientist and policy makers; they provide accurate spatial representations of fire extents and perimeters. Accurate maps of the areas affected by forest fire are needed for rehabilitation planning, calculating the economic and environmental cost of fires, and for regional and global scale estimates in gas and particulate emissions [148, 149].

Fires produce a significant change in the structure and the reflectance of vegetation and the soil properties within the burnt area that are noticeable in the microwave, visible and especially the infrared part of the electromagnetic spectrum [29]. The sensors used for burned area mapping differ depending on scale and purpose of the assessment. Since the 1980s, the majority of techniques have been developed for data acquired from the AVHRR sensor, and as such were restricted to a limited number of reflectance and thermal bands; most commonly method used is multi-temporal comparison of NDVI [150, 151]. Although data from the AVHRR sensor have low spatial resolution (i.e., 1.1 km), global data have been obtained from a series of different satellites since the 1980s allowed the long-term monitoring of large-scale fires in remote and isolated areas. During 2000s, other global burnt area datasets were derived from SPOT Vegetation and the ATSR-2 on board of Envisat [152, 153, 154]. Partial validations of the global burnt area products were performed by Roy et al. [155], Boschetti et al. [156] and Roy and Boschetti [157]. Pereira et al. [150] showed that the accuracy of the results for mapping burnt areas with AVHRR data in the Mediterranean region of Europe was about 80% for large fires. The methods were considered suitable only for fires larger than 1000 ha and reliable for fires larger than 2000 ha. However, the mapping of those fires would correspond only to approximately 30 and 21%, respectively, of the total yearly burnt area in the European Mediterranean region.

Until the launch of the MODIS sensor on board of the TERRA and AQUA satellites, there was not a space-based system design specifically to “look” at terrestrial Earth. MODIS sensors add a new capability for regional mapping of burnt areas, the availability of free data of medium spatial resolution from the MODIS sensors since 2000 provided a definite impulse for the use of remote sensing at the regional and global scales [134]. The higher spectral information and the better radiometry of the MODIS sensor in comparison with the previous sensor systems (e.g., AVHRR—an atmospheric mission), provided the right data for the mapping of burnt areas at these scales. At the global scale, the MODIS program has used to produce a standard product on burned areas based on a multi-temporal change detection approach to analyze differences between modeled and actual reflectance, using an algorithm based on Bidirectional Reflectance Distribution Function (BRDF) change detection approach [155]. Giglio et al. [147] created an algorithm for burned area mapping using 1 km MODIS data, the normalized burn ratio (NBR) and temporal texture. The resulting algorithm (MCD64) was assessed based on Landsat-derived burned area maps for central Siberia, the western US and southern Africa. They found that the algorithm performed well overall, except in a closed canopy subregion in southern Africa where it underestimated burned area. In general, global burned area products have been shown to exhibit relatively large errors of omission and commission. The accuracies of several of these burned area products were compared by Padilla et al. [158]. Six global burned area products were compared using stratified random sampling in the first attempt to implement a statistically designed sample to validate burned area products on a global scale. The products used in this study included MCD45, MCD64, Geoland2, MERGED_cci, MERIS_cci and VGT_cci. The study found MCD64 to be the most accurate, followed by MCD45; however, all the products possessed burned area commission errors above 40% and omission errors above 65%.

At regional scale, MODIS is operationally used in systems such as Canadian CWFIS and the European EFFIS. Two full mosaics of MODIS data are received and processed daily in EFFIS to provide near-real time monitoring of wildfires and map burnt areas. The systems are thus updated up to two times daily, providing accurate information of fire impacts in Europe [159]. Rapid Damage Assessment (RDA) module of EFFIS was initially implemented in 2003 to map burned areas during the fire season, by analyzing MODIS daily images at 250-m spatial resolution. Since the year 2016, the RDA incorporates the mapping of active fires and burnt areas from the VIIRS Sensor, onboard the NASA Suomi National Polar-orbiting Partnership (SNPP) and the NOAA-20, which allows the update of burnt areas maps one more time, every day. The EFFIS from 2018 is started using also Sentinel-2 imagery at 20-m resolution, which allows the detection of fires below the 30-ha threshold and it is estimated that the areas mapped in EFFIS represent about 95% of the total area that burns in the EU every year. Figure 5 shows the extent of burnt areas as they were mapped from MODIS and Sentinel-2 imagery during the last fire season and Figure 6 shows details about the area burnt (dates, location, area, type of vegetation affected).

Figure 5.

Burnt area mapping MODIS/SENTINEL2 from the last fire season © EU, 2022.

Figure 6.

Burnt area mapping from a fire occurred at July 16, 2022 near Cutro, Crotone, Italy © EU, 2022.

At national to local scales, the wide variety of remotely sensed products at high to moderate resolution (1-m to 30-m ground spatial resolution) makes it possible the accurate mapping of burnt areas. Optical image satellites such as the Landsat-8 and Sentinel-2 constitute the latest generation Earth observation missions of the United States Geological Survey and European Space Agency, respectively. The Landsat-8 satellite, launched on February 11, 2013, has 11 bands ranging in spatial resolution from 15 m (panchromatic) to 30 m (visible, near infrared, short-wave infrared) and then to 100 m (thermal). Within 16 days’ temporal resolution, Landsat-8 provides 30-m spatial resolution optical imagery on eight spectral bands via the Operational Land Imager sensor, which can be accessed freely [160, 161]. Landsat-8 30-m products can map small and spatially fragmented burned areas greater than 40 ha in size witch its more detail than other satellite sensors, such as the MODIS 500 m [162]. Sentinel 2 is a high-resolution multispectral imaging satellite composed of Sentinel-2A and Sentinel-2B satellites. The Sentinel-2 satellite covers 13 spectral bands with ground resolutions of 10 m, 20 m and 60 m and provides red-edge spectral bands, the satellite temporal resolution 10 days [163]. Filipponi [164] using threshold-based classification for burnt area mapping of on the 2017 Italy forest fires based on Sentinel-2 time series data reported the methodology generated commission error of around 25% and an omission error of around 40%.

Since the Sentinel-2 and Landsat-8 have similar wavelengths and the same geographic coordinate system, several studies combined both sensors for burnt area mapping [162, 165, 166]. Syifa et al. [166] using SVM and SVM–ICA algorithms for burned area mapping for the Camp Fire (California, USA) report that the SVM–ICA produced a better accuracy (overall accuracies of 83.8 and 83.6% for pre- and post-wildfire using Landsat-8, respectively; 90.8 and 91.8% for pre- and post-wildfire using Sentinel-2, respectively), compared to SVM without optimization (overall accuracies of 80.0 and 78.9% for pre- and post-wildfire using Landsat-8, respectively; 83.3 and 84.8% for pre- and post-wildfire using Sentinel-2, respectively). Different methods of satellite-based burned area detection have been developed, including threshold-based methods using multi-spectral bands or spectral indices such as Normalized Difference Vegetation Index (NDVI), the Soil-adjusted Vegetation Index (SAVI), and the Normalized Burned Ratio (NBR), supervised classification, logistic regression, random forest, etc. [162, 165, 167]. Reviews of burned area mapping algorithms are found in the studies by Boschetti et al. [168] and Chuvieco et al. [169]. Synthetic Aperture Radar (SAR) was also used for burned area mapping mainly in boreal or tropical regions [170, 171], but some examples for the Mediterranean area exist [172, 173]. Also, SAR data have been used in combination with optical data for burned area mapping [174].

4.2 Burn severity and soil erosion assessments

The term fire or burn severity was born out of the need to provide a description of how fire intensity affected ecosystems, particularly following forest fires where direct information on fire intensity was absent and effects are often quite variable within and between different ecosystems [175]. Burn severity impacts vegetation mortality and soil nutrient composition, and causes increased run-off due to decreased of organic matter in the soil and the aboveground vegetation [176, 177]. Burn severity is usually measured in the field using the composite burn index (CBI), which involves an optical assessment of burned areas to determine the fire impacts on ecological conditions. Due to the need for a systematic approach to estimate burn severity across different environments, this approach uses a measure called the composite burn index (CBI) designed to provide a single index to represent many different phenomena of interest to land managers [178]. The CBI combines fire severity metrics and ecosystem responses that include resprouting of herbs, shrubs and hardwood trees, and seedling colonization. The CBI was created to allow for visual estimates to be conducted by rating the degree of damage done by the fire, as well as the estimated vegetation recovery for the area, on a 0 to 3 scale [179].

Various remote sensors (e.g., MODIS, AVIRIS) have been tested for their ability to match field measurements of severity and the Landsat Thematic Mapper sensor was widely accepted as most appropriate for this task [180, 181, 182].

Spectral indices, such as the Normalized Burn Ratio (NBR) and NDVI, have been widely used for assessing burn severity through remote sensing in boreal regions [183, 184, 185, 186]. The NBR has mostly replaced the NDVI as the standard index for burn severity assessment and been widely used as a means for approximating the burn severity and burned area using satellite imagery also in the Mediterranean basin [187, 188]. The NBR is calculated using NIR and SWIR data, with the SWIR wavelength interval generally within the 2.08–2.35-μm range [189]. A numerous differenced indices for burn severity assessment using multi-temporal satellite data and bi-temporal image differencing techniques have been used in boreal regions, such as the Differenced Normalized Burn Ratio (dNBR) [186, 190], the Differenced Normalized Difference Vegetation Index (dNDVI) [191, 192] and a relative version of the dNBR (RdNBR) [193]. The majority of this indices used NIR and SWIR data, where the NIR wavelengths are sensitive to the leaf structure of live vegetation, while the SWIR is sensitive to moisture content and some soil conditions [189, 194]. Vegetation affected by a fire show decreased NIR reflectance and increased SWIR reflectance [187].

Temporal differences in land surface albedo (LSA) and land surface temperature (LST) have been used to estimate burn severity [195, 196]. Quintano et al. [195] found that changes in LST showed high agreement with field-measured CBI when used to map burn severity for an ecosystem dominated by maritime pine (Pinus pinaster) in Sierra del Teleno, Spain. Zheng et al. [196] conclude that the results indicated that the deltaLST/EVI performed relatively better than some commonly used burn severity indices, and its performance was stable and steady across regional forest areas of different eco-type after fires. However, Veraverbeke et al. [197] and Quintano et al. [195] found that while changes in LSA and LST were highly related to burn severity they were highly dependent on seasonality.

With the launch of the Sentinel-2 sensors by the European Space Agency (ESA), with specific spectral bands to record data in the vegetation red-edge spectral domain, which is one of the best radiance-based descriptors of chlorophyll content [198], allowed for the development and application of new spectral indices to discriminate burn severity.

Fernández-Manso et al. [199] used Sentinel-2A imagery to examine capabilities of red-edge spectral indices in the context of burn severity in central-western Spain. The area is a mixed of shrubs and trees, with maritime pine (Pinus pinaster) and Pyrenean oak (Quercus pyrenaica) being the dominate species. Results indicated that the red-edge spectral indices calculated using red-edge 1(B5) and red-edge 3(B7)/NIR(B8) were most suitable for burn severity assessment. Filipponi [164, 200] tested Burned Area Index for Sentinel-2 (BAIS2, BAIS2α), based on Sentinel-2 spectral bands on various study cases in Italy for summer 2017 fires, and results show a good performance of the index and highlighted critical issues related to the Sentinel-2 data processing. Burn severity assessments require imagery with a high to moderate (<100 m) spatial resolution such as Landsat series and the Sentinel-2 sensors, and the limitation of these sensors is the long temporal resolution, which can make the rapid acquisition of data on post-fire conditions difficult.

With the use of more advanced remote sensing technologies, such as hyperspectral imagery, lidar, radar and UAVs, important fire effects may be more accurately and consistently inferred from imagery with higher spectral, spatial and temporal resolutions [201]. Lidar sensors provide a relative new technology, which can be used for burn severity assessments. Lidar data can provide pre- and post-fire vegetation structure profiles, which can be used for severity of burning estimations. Wang and Glenn [202] showed the potential for lidar-derived burn severity estimates is sagebrush steppe rangelands using average vegetation height change. This method outperformed the dNBR index estimations, with an overall accuracy of 84%, and proved to be sensitive to differences between moderate and high severity burns. Wulder et al. [203] compared changes in boreal forest structure, obtained by lidar returns, to post-fire conditions, estimated using spectral indices for the Boreal Plains ecozone in Alberta, Canada. The researchers found that absolute and relative changes in post-fire forest structure exhibited a high correlation with post-fire conditions. The synergy of lidar with multispectral sensor data for burn severity assessment is a topic of ongoing research [204, 205, 206]. Viedma et al. [206] crossed high-resolution lidar data, acquired from an UAV after the fire and fire severity levels based on the relativized burnt ratio (RBR) derived from Sentinel 2A images acquired a few months before and after fire. They conclude that lidar metrics derived from vertical canopy profiles (VCPs) demonstrated promising potential for characterizing fine-grained post-fire plant structures and fire damage when crossed with satellite-based fire severity metrics, turning into a promising approach for better characterizing fire impacts at a resolution needed for many ecological processes.

UAV technologies have seen an important progression in the last decade and they are now used in a wide range of applications. This technology has recently been applied to research focusing on burn severity assessment using various sensors [206, 207, 208, 209, 210, 211, 212, 213]. Ye et al. [211] used object-oriented method to determine the optimal segmentation scale for forest burn severity and a multilevel rule classification and extraction model is established to achieve the automatic identification and mapping. The imagery was obtained by a small, multi-rotor near-ground UAV and the results show that the method mentioned above can recognize different types of forest burn severity: unburned, damaged, dead and burnt with overall accuracy been 87.76%. Carvajal-Ramírez et al. [207] employed an UAV carrying a high-resolution multispectral sensor including green, red, near-infrared and red-edge bands. Flights were carried out pre- and post-controlled fire in a Mediterranean forest. The products obtained from the UAV were a Digital Surface Model (DSM) and multispectral images orthorectified in both periods, temporal differences (d) between pre- and post-fire values of the Excess Green Index (EGI), Normalized Difference Vegetation Index (NDVI) and Normalized Difference Red Edge (NDRE) index, where used to after reclassification to produce fire severity classes. It was concluded that dNDVI was the index that best estimated the fire severity according to the UAV flight conditions and sensor specifications.

Forest fires can negatively impact water catchments, contaminating, increasing soil erosion, changing soil composition and slope stability for extensive periods. A major reason for post-fire assessments of fire or burn severity is that it is used as an important indicator of the potential water runoff and erosion [214, 215, 216], the loss of vegetation exposes soil to erosion and makes burned areas more vulnerable to runoff and then susceptible to flood [217].

Several studies used fire or burn severity data derived from remote sensing applications or/and GIS applications for post-fire soil erosion assessment [217, 218, 219]. Lanorte et al. [218] used a distributed model based on the Revised Universal Soil Loss Equation (RUSLE) to estimate potential post-fire soil loss for four different fire events occurred in Basilicata region (Southern Italy) during 2017. GIS techniques and remote sensing data have been adopted to build a prediction model of post-fire soil erosion risk. Results show that this model is not only able to quantify post-fire soil loss but also to identify the complexity of the relationships between fire severity and all the factors that influence soil susceptibility to erosion. Fox et al. [217] used POSTFIRE a GIS-based model that maps the burn scar and quantifies fire impacts on runoff and soil erosion. The model used pre-fire and post-fire satellite images, calculates the impact of a fire on total rainfall event runoff, and maps soil erosion rates. Additionally, the model requires a Digital Elevation Model (DEM), a mask of the general contour of the fire, land cover map, and tables of runoff coefficients and sediment concentration values for the land covers. Preliminary results from two fires in SE France indicate the model is simple to use, adaptable to local conditions, and provides realistic outputs for the burn scar, discharge and soil erosion maps after a forest fire. UAV imagery has also been used for soil erosion modeling in various types of landscape [220, 221, 222].

4.3 Post-fire vegetation recovery monitoring

Vegetation recovery after a fire event is an important measurement for the determination of the long-term impacts that fire has on an ecosystem. During the process of recovery, vegetation may also be influenced by several environmental factors, such as fire severity/damage and climatological extreme events. Vegetation recovery depends on several climate factors, such as the occurrence of droughts, which may inhibit vegetation growth, but also on precipitation of high intensity, which contributes to nutrient loss and erosion by runoff [223]. Anthropogenic factors such as grazing can also affect vegetation recovery [224, 225].

Monitoring post-fire vegetation recovery is crucial, as it provides valuable information for analyzing ecosystem resilience, for determining landscape dynamics and for forest management purposes. Post-fire monitoring of vegetation recovery can be conducted on the field using plots sampling where they measure seedling germination, plant survival and restoration, and vegetation characteristics [226, 227, 228], or with the use of remote sensing technologies. Since field measurements of post-fire vegetation recovery are a very challenging and expensive task, remote sensing techniques are a time- and cost-effective way to monitor post-fire ecosystem recovery [229]. Remote sensing techniques for estimating vegetation recovery can be grouped into three categories: (1) image classification, (2) vegetation indices (VIs) and (3) spectral mixture analysis (SMA) [230].

The most practical way to monitor changes over large areas and periods is through image-processing techniques based on change detection or classification techniques [231]. Stueve et al. [232] used supervised classification to identify patterns of alpine tree recovery in Mount Rainer National Park, Washington, USA. They performed classification analysis of 1970 satellite imagery and 2003 aerial photography to delineate establishment. Local site conditions were calculated from a lidar-based DEM, ancillary climate data and 1970 tree locations in a GIS. The supervised classification method proved successful, which can be credited to the very high spatial resolution of the data used in this study. Salvia et al. [233] used a combination of unsupervised classification and field data to successfully examine the influence of burn severity on vegetation cover and soil property recovery in the wetlands of the Paraná River Delta, Argentina.

Geographic object-based image analysis (GEOBIA) is an alternative method for image classification and uses geographic objects instead of pixels as the spatial unit of analysis [234]. GEOBIA are techniques that use both spectral response and contextual information to assess post-fire vegetation characteristics in groups of pixels (geographic objects generated by image segmentation). Polychronaki et al. [235] performed GEOBIA-based classification on multi-temporal Système Pour l’Observation de la Terre (SPOT) and European Remote-Sensing (ERS) (C-band VV) images covering the time period from 1993 to 2007 of Thasos, Greece, to estimate forest regeneration after two different forest fires occurred 19 and 23 years ago. Using data from field-stratified sampling, they found that the classification results were validated, achieving an overall accuracy of 90.5%. In the same area, Mitri and Gitas [236] used GEOBIA on a combination of very high spatial (VHS) resolution (QuickBird) and hyperspectral (EO-1 Hyperion) imagery to investigate Pinus brutia and Pinus nigra regeneration. After classification of the segmented imagery, validation showed an overall accuracy of 83.7%.

In relation to post-fire vegetation recovery, these indicators used generally rely on greenness measurements of red-near-infrared (R-NIR) vegetation indices [237, 238]. Several spectral indices based on the NIR is used for post-fire vegetation recovery estimations; nevertheless, the Normalized Difference Vegetation Index (NDVI), the Enhanced Vegetation Index (EVI) and the Normalized Burn Ratio (NBR) are the ones used most frequently [239]. Veraverbeke et al. [240] tested the utility of 13 red-near infrared (R–NIR) vegetation indices (VIs) spectral indices to detect and estimate vegetation recovery. The study found that the soil-adjusted vegetation index (SAVI) outperformed the NDVI in areas with a single type of vegetation, NDVI outperformed the SAVI in areas with heterogeneous vegetation cover and a single soil type, and overall, the NDVI was the most robust VI for assessing vegetation recovery. Several studies of post-fire vegetation responses are based on the discrimination of spectral bands and vegetation indices (mostly NDVI, dNBR and EVI) by MODIS, Landsat, SPOT and Sentinel multi-temporal imagery in different regions and forest ecosystems of the world [241, 242, 243, 244, 245].

Spectral mixture analysis (SMA) have also used in vegetation recovery monitoring [246, 247]. SMA studies have shown high correlation between mapped vegetation recovery and field sampling [248, 249]. There is disagreement on whether the SMA or spectral index-based approach is most effective at estimating vegetation recovery [17]. Specific nature of the vegetation’s spectral response and the different components of forest recovery, the use of a single method may not be the best option, requiring multiple methods to be deployed in each ecosystem. Riaño et al. [248] and Vila and Barbosa [250] compared Spectral Mixture Analysis (SMA) against quantitative vegetation indices (NDVI and Modified Soil Adjusted Vegetation Index (MSAV)), in different types of ecosystems, such as chaparral shrub communities SMA performed better in the contrary NDVI was most accurate applicable to coniferous forests with maquis.

Further than passive remote sensing active synthetic aperture radar (SAR) has been used for post-fire vegetation recovery assessments [30, 251, 252, 253, 254]. Laurin et al. [30] used Cosmo-SkyMed data in a Mediterranean protected area covered by maquis to detect the burnt area extension and to conduct a mid-term assessment of vegetation regrowth. The positive results obtained in this research highlight the importance of the very high-resolution continuous acquisitions and the multi-polarization information provided by COSMO-SkyMed for monitoring fire impacts on vegetation. Several studies also confirmed that the combination of SAR and optical information was essential for estimation of regeneration stages in different forest ecosystems [255, 256, 257] and for post-fire vegetation regrowth [258, 259]. De Luca et al. [259] integrating the use of Synthetic Aperture Satellite Radar (SAR) (Sentinel-1) and optical (Sentinel-2) image time series in a Mediterranean ecosystem. They conclude that the use of optical short-wave infrared (SWIR) and SAR C-band-based data revealed that some ecological characteristics, such as the woody biomass and structure, recovered at slower rates, comparing to those suggested by using near-infrared (NIR) and red-edge data and they proposed the optimized burn recovery ratio (BRR) for the estimation and mapping the spatial distribution of the degree of vegetation recovery.

Lidar sensors are capable of penetrating through vegetation and recording forest structural characteristics [260] makes the system ideal for monitoring vegetation recovery. Sato et al. [261] estimated post-fire changes in forest canopy height and biomass using airborne Lidar in western Amazonia. Comparing burned and unburned as control sites found that even 10 years after the occurrence of an understory fire event, burned forests had significantly lower biomass and height than control sites. Gordon et al. [262] used lidar data to measure post-fire mid-story vegetation regrowth of open forests in Australia using also field surveys. They found that the metrics computed with the lidar data had moderate to strong positive associations with field-derived metrics and provided a suitable representation of post-fire vegetation cover. Several studies also have examined the combined use of optical sensor imagery and lidar data [205, 263, 264, 265, 266]. Viana-Soto et al. [266] combine lidar with Landsat imagery to extrapolate forest structure variables over a 30-year period (1990–2020) to provide insights on how forest structure has recovered after fire in Mediterranean pine forests. They found that less than 50% of the area completely recovered to the pre-fire structure within 26 years, and the area subjected to fire recurrence showed signs of greater difficulty in initiating the recovery. They conclude that the use of Landsat data provides a unique opportunity to analyze the evolution through decades, but the 30-m pixel size may conceal a larger structural variability within. Even so, cover and height are important indicators of forest recovery that can be derived to larger areas, providing useful information to support post-fire restoration activities [267].

In addition to satellite or the conventional airborne sensors, the value of using unmanned aerial vehicles (UAV) for estimation and monitoring of post-fire forest recovery has also been incised with the use of a variety of sensors [206, 210, 268, 269, 270, 271, 272, 273, 274]. Pádua et al. [268] evaluated effectiveness for Post-Fire Monitoring of Sentinel-2 time series data in comparison with high-resolution UAV-based data in an area affected by a fire in north-eastern Portugal. Sentinel-2 images with 10-m resolution, different spatial resolutions of the UAV-based data (0.25, 5 and 10 m) were used and compared to determine their similarities. The results demonstrated the effectiveness of satellite data for post-fire monitoring, even at a local scale as more cost-effective than UAV data. Talucci et al. [273] evaluate the ability of two vegetation indices derived from UAV imagery, one based on the visible spectrum (GCC; Green Chromatic Coordinate) and one using multispectral data (NDVI; Normalized Difference Vegetation Index), to predict field-based vegetation measures collected across post-fire landscapes of high-latitude Cajander larch forests. Findings show the utility of UAV data for NDVI in this region as a tool for quantifying and monitoring the post-fire vegetation dynamics in Cajander larch forests.

UAV systems can also be used to deliver temporal digital surface models (DSM) to detect post-fire changes in vegetation recovery, and the DSMs are produced either from lidar or using photogrammetric approach from digital imagery. Qi et al. [210] used drone laser scanning (DLS) and mobile laser scanning (MLS) to describe post-fire forest structures. There results demonstrate that fused DLS-MLS point clouds can be effective in quantifying post-fire tree structures, which facilitates foresters to develop site-specific management plans. Aicardi et al. [272] used change detection analysis through a time sequence (2008–2015) of DSMs obtained from lidar and digital images (UAV), for a forest stand composed primarily of Scots pine (Pinus sylvestris) in the Aosta Valley Region, Italy. They conclude that these preliminary results highlight the usefulness of low cost, highly flexible UAV systems within areas affected by natural or anthropogenic disturbances, given the possibility to fly on-demand (e.g., immediately after the event) and quickly repeat measures when requested for environmental monitoring.

Advertisement

5. Discussion and conclusions

This chapter aimed to provide a wide review of remote sensing and GIS applications in the field of forest fires—wildfires. The topics include fire hazard and fire risk mapping, fuel mapping, active fire detection and monitoring, burned area estimations, burn severity assessments and mapping, soil erosion assessments and post-fire vegetation recovery assessments and monitoring. Fire hazard mapping can be improved with better mapping and monitoring of manmade Infrastructure like power lines using UAVs with active (lidar) or passive sensors. For fire risk mapping alternative geocomputational techniques as neural networks, classification and regression trees (CARTs), fuzzy modeling, and other machine learning techniques are currently a research subject [275]. Locally fuel mapping can be delivered from UAVs with active (lidar) or passive sensors in more detailed resolution than satellite data. Active fire detection and monitoring is limited from the low temporal or spatial resolution of the available satellite sensors. Smaller and cooler fires have proven difficult to detect in global active fire datasets and by geostationary sensors [132, 140]. Probably, in the near future, geostationary satellites with higher spatial resolutions may become available, allowing for near real-time detection and monitoring of small/cool fires, already new generation satellites for forest and grassland fire detection have been launched such as the Himawari-8 launched by Japan, GOES-R/S/T by the United States and Geo-Kompsat-2A (GK-2A) by South Korea, compared with the previous satellites, and these new-generation geostationary meteorological satellites have been improved in spatial resolution and observation frequency [276]. UAVs equipped with infrared or thermal or RGB cameras can be used where data can provide live feeds and be used to predict information such as propagation of a fire [277]. For burned area estimates mostly Landsat, MODIS, Sentinel, AVHRR and SPOT satellites are used, with the NDVI and dNBR the most frequently used vegetation indices compared to other indices. Relatively newer sensors as Sentinel-2 have improved the high rates of omission errors, common when using other satellite sensors, and allow the detection of small fires (< 100 ha) accounting for a significant proportion of total burned area globally [278]. Currently, burn severity assessments and mapping are mostly based using spectral indices such us dNBR (difference of the Normalized Burn Ratio) or relative differenced Normalized Burn Ratio (RdNBR), and classified according to the ground reference values of the CBI (Composite Burn Index). UAVs equipped with lidar or hyperspatial sensors allowed the rapid burn severity assessments and further research is needed to utilize the hyperspatial data provided from UAV sensors. Also, lidar technology derived from UAVs are able to provide forest structure parameters and become increasingly available. Similar erosion assessments can improve with the use of UAVs since they are able to provide higher spatial data. The use of remote sensing to analyze vegetation recovery is expected to grow even further in application and prominence as new sensors become available (i.e., UAVs and new satellites) and bring enhanced spatial, spectral and temporal resolutions to the observations [239]. Current and future trends are the combination of remote sensing data from passive and active sensors either spaceborne or airborne. With the availability of several spaceborne sensors (passive and active) and the emerging improvements of UAVs technologies (flight endurance, sensors, affordability), the development of algorithms that can monitor changes through time irrespective of the characteristics of each platform [279] will improve the overall accuracy of the assessments in almost all the reviewed topics. Finally, recent advances in artificial intelligence (AI) will play a growing role in geospatial technologies as remote sensing (RS) applications [280, 281, 282].

References

  1. 1. Food and Agriculture Organization. United Nations Environment Programme. Forests, Biodiversity and People. The State of the World’s Forests 2020 [Internet]. 2020 [cited 2023 Jan 17]. pp. 227-49. Available from: www.fao.org/publications%0Ahttp://www.fao.org/documents/card/en/c/ca8642en
  2. 2. Intergovernmental Panel on Climate Change (IPCC). Technical summary. In: Climate Change 2022 – Impacts, Adaptation and Vulnerability: Working Group II Contribution to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge: Cambridge University Press; 2023. pp. 37-118. DOI: 10.1017/9781009325844.002
  3. 3. Kurvits T, Popescu A, Paulson A, Sullivan A, Ganz D, Burton C, et al. Spreading Like Wildfire: The Rising Threat of Extraordinary Landscape Fires. Nairobi: United Nations Environment Programme; 2022. Available from: https://policycommons.net/artifacts/2390241/spreading-like-wildfire/3411461/
  4. 4. San-Miguel-Ayanz J, Camia A. Forest fires. In ‘mapping the impacts of natural hazards and technological accidents in Europe: An overview of the last decade’. European Environment Agency Technical Report N. 2010;13:47-53
  5. 5. Pyne S. Eternal flame: An introduction to the fire history of the mediterranean. In: Earth Observation of Wildland Fires in Mediterranean Ecosystems. Berlin, Heidelberg: Springer; 2009. pp. 11-26. DOI: 10.1007/978-3-642-01754-4_2
  6. 6. Amraoui M, Liberato MLR, Calado TJ, DaCamara CC, Coelho LP, Trigo RM, et al. Fire activity over Mediterranean Europe based on information from Meteosat-8. Forest Ecology and Management. 2013;294:62-75
  7. 7. Pereira MG, Calado TJ, DaCamara CC, Calheiros T. Effects of regional climate change on rural fires in Portugal. Climate Research. 2013;57:187-200 [Internet] [cited 2022 Nov 29]. Available from: https://www.int-res.com/abstracts/cr/v57/n3/p187-200/
  8. 8. Ruffault J, Moron V, Trigo RM, Curt T. Daily synoptic conditions associated with large fire occurrence in Mediterranean France: Evidence for a wind-driven fire regime. International Journal of Climatology. 2017;37:524-533 [Internet] [cited 2023 Jan 17]. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/joc.4680
  9. 9. Hardy CC. Wildland fire hazard and risk: Problems, definitions, and context. Forest Ecology and Management. 2005;211:73-82
  10. 10. Chuvieco E, Aguado I, Yebra M, Nieto H, Salas J, Martín MP, et al. Development of a framework for fire risk assessment using remote sensing and geographic information system technologies. Ecological Modelling. 2010;221:46-58
  11. 11. Fischer MA, di Bella CM, Jobbágy EG. Influence of fuel conditions on the occurrence, propagation and duration of wildland fires: A regional approach. Journal of Arid Environments. 2015;120:63-71
  12. 12. Calle A, Casanova JL. Forest fires and remote sensing. In: Coskun HG, Cigizoglu HK, Maktav MD, editors. Integration of Information for Environmental Security. NATO Science for Peace and Security Series C: Environmental Security. Dordrecht: Springer; 2008. pp. 261-304. DOI: 10.1007/978-1-4020-6575-0_19
  13. 13. San-Miguel-Ayanz J, Carlson JD, Alexander M, Tolhurst K, Morgan G, Sneeuwjagt R, et al. Current methods to assess fire danger potential. In: Chuvieco E, editor. Wildland Fire Danger Estimation and Mapping - The Role of Remote Sensing Data. Singapore: World Scientific Publishing Co. Pte. Ltd.; 2003. pp. 21-61, 44 p
  14. 14. Gerdzheva AA. A comparative analysis of different wildfire risk assessment models (a case study for Smolyan district, Bulgaria). European Journal of Geography. 2014;5:22-36
  15. 15. Adab H, Kanniah KD, Solaimani K. Modeling forest fire risk in the northeast of Iran using remote sensing and GIS techniques. Natura. 2013;65:1723-1743 [Internet] [cited 2022 Dec 4]. Available from: https://link.springer.com/article/10.1007/s11069-012-0450-8
  16. 16. Gabban A, San-Miguel-Ayanz J, Viegas DX. A comparative analysis of the use of NOAA-AVHRR NDVI and FWI data for forest fire risk assessment. International Journal of Remote Sensing. 2008;29:5677-5687
  17. 17. Szpakowski D, Jensen J. A review of the applications of remote sensing in fire ecology. Remote Sensing (MDPI AG). 2019;11:2638 [Internet]. DOI: 10.3390/rs11222638
  18. 18. Finney MA. The challenge of quantitative risk analysis for wildland fire. Forest Ecology and Management. 2005;211:97-108
  19. 19. Hernandez C, Drobinski P, Turquety S. How much does weather control fire size and intensity in the Mediterranean region? Annales de Geophysique. 2015;33:931-939 [Internet] [cited 2022 Dec 4]. Available from: https://angeo.copernicus.org/articles/33/931/2015/
  20. 20. Nesterov VG. Combustibility of the Forest and Methods for its Determination. Moscow, Russia: Goslesbumizdat, USSR State Industry Press; 1949. 76 p
  21. 21. McArthur AG. Fire behaviour in eucalypt forests. Forestry and Timber Bureau. 1967;26:26
  22. 22. Deeming JE, Burgan RE, Cohen JD. The National Fire-Danger Rating System. Gen. Tech. Rep. INT-GTR-39. Ogden, UT: U.S. Department of Agriculture, Forest Service, Intermountain Forest and Range Experiment Station; 1978. 63 p
  23. 23. van Wagner CE. Development and Structure of the Canadian Forest Fire Weather Index System. Forestry Technical Report 35. Ottawa: Canadian Forestry Service, Headquarters; 1987. 35 p
  24. 24. Yu B, Chen F, Li B, Wang L, Wu M. Fire risk prediction using remote sensed products: A case of Cambodia. Photogramm Eng Remote Sensing. 2017;83:19-25
  25. 25. Bax V, Francesconi W. Environmental predictors of forest change: An analysis of natural predisposition to deforestation in the tropical Andes region, Peru. Applied Geography. 2018;91:99-110
  26. 26. Kocher SD, Butsic V, Smith AMS, Lutz J, Hoffman C, Williamson G, et al. Governance of Land Use Planning to Reduce Fire Risk to Homes Mediterranean France and California. 2017. Available from: www.mdpi.com/journal/land
  27. 27. Hong H, Tsangaratos P, Ilia I, Liu J, Zhu AX, Xu C. Applying genetic algorithms to set the optimal combination of forest fire related variables and model forest fire susceptibility based on data mining models. The case of Dayu County, China. Science of The Total Environment. 2018;630:1044-1056
  28. 28. de Leeuw J, Vrieling A, Shee A, Atzberger C, Hadgu KM, Biradar CM, et al. The potential and uptake of remote sensing in insurance: A review. Remote Sensing. 2014;6:10888-10912
  29. 29. Leblon B, Bourgeau-Chavez L, San-Miguel-Ayanz J, Leblon B, Bourgeau-Chavez L, San-Miguel-Ayanz J. Use of remote sensing in wildfire management. Current international perspectives on wildland fires. In: Curkovic S, editor. Sustainable Development - Authoritative and Leading Edge Content for Environmental Management. IntechOpen; 2012
  30. 30. Laurin GV, Avezzano R, Bacciu V, del Frate F, Papale D, Virelli M. Cosmo-skymed potential to detect and monitor mediterranean maquis fires and regrowth: A pilot study in capo Figari, Sardinia, Italy. iForest. 2018;11:389-395
  31. 31. Arroyo LA, Pascual C, Manzanera JA. Fire models and methods to map fuel types: The role of remote sensing. Forest Ecology and Management. 2008;256:1239-1252
  32. 32. Allgöwer B, Carlson JD, van Wagtendonk JW. Introduction to fire danger rating and remote sensing — Will remote sensing enhance wildland fire danger rating? In: Chuvieco E, editor. Wildland Fire Danger Estimation and Mapping: The Role of Remote Sensing Data. Chapter: 1. World Scientific; 2003
  33. 33. Chuvieco E, Riaño D, van Wagtendok J, Morsdof F. Fuel Loads and Fuel Type Mapping. 2003. pp. 119-42
  34. 34. Chuvieco E, Kasischke ES. Remote sensing information for fire management and fire effects assessment. Journal of Geophysical Research – Biogeosciences. 2007;112:G01S90. DOI: 10.1029/2006JG000230
  35. 35. Vadrevu KP, Eaturu A, Badarinath KVS. Fire risk evaluation using multicriteria analysis—A case study. Environmental Monitoring and Assessment. 2009;166:223-239 [Internet] [cited 2022 Dec 9]. Available from: https://link.springer.com/article/10.1007/s10661-009-0997-3
  36. 36. Qin R, Liu T. A review of landcover classification with very-high resolution remotely sensed optical images—Analysis unit, model scalability and transferability. Remote Sensing. 2022;14(3):646. DOI: 10.3390/RS14030646
  37. 37. Yankovich KS, Yankovich EP, Baranovskiy NV. Classification of vegetation to estimate forest fire danger using Landsat 8 images: Case study. Mathematical Problems in Engineering. 2019;2019:1-14. DOI: 10.1155/2019/6296417
  38. 38. Shaik RU, Giovanni L, Fusilli L. New approach of sample generation and classification for wildfire fuel mapping on hyperspectral (prisma) image. In: IEEE International Geoscience and Remote Sensing Symposium IGARSS, Brussels, Belgium. 2021. pp. 5417-5420. DOI: 10.1109/IGARSS47720.2021.9554652
  39. 39. Pizaña JMG, Romero JMNH, NC. Remote sensing-based biomass estimation. In: Marghany M, editor. Environmental Applications of Remote Sensing [Internet]. Rijeka: IntechOpen; 2016. Ch. 1. DOI: 10.5772/61813
  40. 40. Lu D. The potential and challenge of remote sensing-based biomass estimation. International Journal of Remote Sensing. 2006;27:1297-1328
  41. 41. Brandis K, Jacobson C. Estimation of vegetative fuel loads using Landsat TM imagery in New South Wales, Australia. International Journal of Wildland Fire. 2003;12:185-194 [Internet] [cited 2022 Dec 13]. Available from: https://www.publish.csiro.au/wf/wf03032
  42. 42. Reich RM, Lundquist JE, Bravo VA, Reich RM, Lundquist JE, Bravo VA. Spatial models for estimating fuel loads in the Black Hills, South Dakota, USA. International Journal of Wildland Fire. 2004;13:119-129 [Internet] [cited 2022 Dec 13]. Available from: https://www.publish.csiro.au/wf/WF02049
  43. 43. Alonso-Benito A, Arroyo LA, Arbelo M, Hernández-Leal P, González-Calvo A, Alonso-Benito A, et al. Pixel and object-based classification approaches for mapping forest fuel types in Tenerife Island from ASTER data. International Journal of Wildland Fire. 2012;22:306-317 [Internet] [cited 2022 Dec 13]. Available from: https://www.publish.csiro.au/wf/WF11068
  44. 44. Bajocco S, Dragoz E, Gitas I, Smiraglia D, Salvati L, Ricotta C. Mapping Forest fuels through vegetation phenology: The role of coarse-resolution satellite time-series. PLoS One. 2015;10:e0119811 [Internet] [cited 2022 Dec 13]. Available from: https://journals.plos.org/plosone/article?id=10.1371/journal.pone.0119811
  45. 45. Mallinis G, Galidaki G, Gitas I. A comparative analysis of EO-1 Hyperion, Quickbird and Landsat TM imagery for fuel type mapping of a typical Mediterranean landscape. Remote Sensing. 2014;6:1684-1704 [Internet] [cited 2022 Dec 13] Available from: https://www.mdpi.com/2072-4292/6/2/1684/htm
  46. 46. Jia GJ, Burke IC, Goetz AFH, Kaufmann MR, Kindel BC. Assessing spatial patterns of forest fuel using AVIRIS data. Remote Sensing of Environment. 2006;102:318-327
  47. 47. Lasaponara R, Lanorte A, Pignatti S. Multiscale fuel type mapping in fragmented ecosystems: Preliminary results from hyperspectral MIVIS and multispectral Landsat TM data. International Journal of Remote Sensing. 2006;27:587-593
  48. 48. Arellano-Pérez S, Castedo-Dorado F, López-Sánchez C, González-Ferreiro E, Yang Z, Díaz-Varela R, et al. Potential of Sentinel-2A data to model surface and canopy fuel characteristics in relation to crown fire hazard. Remote Sensing. 2018;10:1645. DOI: 10.3390/rs10101645
  49. 49. Alonso-Benito A, Arroyo L, Arbelo M, Hernández-Leal P. Fusion of WorldView-2 and LiDAR data to map fuel types in the Canary Islands. Remote Sensing (MDPI AG). 2016;8:669 [Internet]. DOI: 10.3390/rs8080669
  50. 50. Arroyo LA, Healey SP, Cohen WB, Cocero D, Manzanera JA. Using object-oriented classification and high-resolution imagery to map fuel types in a Mediterranean region. Journal of Geophysical Research. 2006;111:4-04 [Internet]. Available from: http://www.firegrowthmodel
  51. 51. Keane RE, Burgan R, van Wagtendonk J. Mapping wildland fuels for fire management across multiple scales: Integrating remote sensing, GIS, and biophysical modeling. International Journal of Wildland Fire. 2001;10:301-319 [Internet] [cited 2022 Dec 9]. Available from: https://www.publish.csiro.au/wf/wf01028
  52. 52. García M, Popescu S, Riaño D, Zhao K, Neuenschwander A, Agca M, et al. Characterization of canopy fuels using ICESat/GLAS data. Remote Sensing of Environment. 2012;123:81-89
  53. 53. Perko R, Raggam H, Deutscher J, Gutjahr K, Schardt M. Forest assessment using high resolution SAR data in X-band. Remote Sensing. 2011;3:792-815 [Internet] [cited 2013 Dec 30]. Available from: http://www.mdpi.com/2072-4292/3/4/792/
  54. 54. D’este M, Elia M, Giannico V, Spano G, Lafortezza R, Sanesi G. Machine learning techniques for fine dead fuel load estimation using multi-source remote sensing data. Remote Sensing. 2021;13:1658 [Internet] [cited 2022 Dec 10]. Available from: https://www.mdpi.com/2072-4292/13/9/1658/htm
  55. 55. Chuvieco E, Congalton RG. Application of remote sensing and geographic information systems to forest fire hazard mapping. Remote Sensing of Environment. 1989;29:147-159
  56. 56. Pimont F, Dupuy J-L, Linn RR, Pimont F, Dupuy J-L, Linn RR. Coupled slope and wind effects on fire spread with influences of fire size: A numerical study using FIRETEC. International Journal of Wildland Fire. 2012;21:828-842 [Internet] [cited 2022 Dec 16]. Available from: https://www.publish.csiro.au/wf/WF11122
  57. 57. Abdollahi M, Islam T, Gupta A, Hassan Q . An advanced forest fire danger forecasting system: Integration of remote sensing and historical sources of ignition data. Remote Sensing (MDPI AG). 2018;10:923 [Internet]. DOI: 10.3390/rs10060923
  58. 58. Vilar L, Camia A, San-Miguel-Ayanz J. A comparison of remote sensing products and forest fire statistics for improving fire information in mediterranean Europe. European Journal of Remote Sensing. 2015;48:345-364
  59. 59. Arino O, Plummer S, Defrenne D, Arino O, Plummer S, Defrenne D. Fire disturbance: The ten years time series of the ATSR world fire atlas. ESASP. 2005;597:30.1 [Internet] [cited 2022 Dec 17]. Available from: https://ui.adsabs.harvard.edu/abs/2005ESASP.597E..30A/abstract
  60. 60. Giglio L, Descloitres J, Justice CO, Kaufman YJ. An enhanced contextual fire detection algorithm for MODIS. Remote Sensing of Environment. 2003;87:273-282
  61. 61. Trigo IF, Dacamara CC, Viterbo P, Roujean JL, Olesen F, Barroso C, et al. The satellite application facility for land surface analysis. International Journal of Remote Sensing. 2011;32:2725-2744
  62. 62. Shahbazi M, Théau J, Ménard P. Recent applications of unmanned aerial imagery in natural resource management. GIScience & Remote Sensing. 2014;51:339-365
  63. 63. Colomina I, Molina P. Unmanned aerial systems for photogrammetry and remote sensing: A review. ISPRS Journal of Photogrammetry and Remote Sensing. 2014;92:79-97
  64. 64. Lechner AM, Foody GM, Boyd DS. Applications in remote sensing to Forest ecology and management. One Earth [Internet]. 2020;2:405-412. DOI: 10.1016/j.oneear.2020.05.001
  65. 65. Tang L, Shao G. Drone remote sensing for forestry research and practices. Journal of Forestry Research. 2015;26:791-797. DOI: 10.1007/s11676-015-0088-y
  66. 66. Guimarães N, Pádua L, Marques P, Silva N, Peres E, Sousa JJ. Forestry remote sensing from unmanned aerial vehicles: A review focusing on the data, processing and potentialities. Remote Sensing (MDPI AG). 2020;12:1046 [Internet]. DOI: 10.3390/rs12061046
  67. 67. Manfreda S, McCabe M, Miller P, Lucas R, Pajuelo Madrigal V, Mallinis G, et al. On the use of unmanned aerial systems for environmental monitoring. Remote Sensing (MDPI AG). 2018;10:641 [Internet]. DOI: 10.3390/rs10040641
  68. 68. Atole RR, Bello LC, Lirag JR. Eyes in the sky: A review of civilian unmanned aerial vehicles (UAVs). International Journal of Computers and Applications. 2017;173:36-41
  69. 69. Dainelli R, Toscano P, di Gennaro SF, Matese A. Recent advances in unmanned aerial vehicle Forest remote sensing—A systematic review. Part I: A general framework. Forests. 2021;12:327 [Internet] [cited 2022 Dec 20]. Available from: https://www.mdpi.com/1999-4907/12/3/327/htm
  70. 70. Vilar L, Camia A, San-Miguel-Ayanz J. Modelling socio-economic drivers of forest fires in the Mediterranean Europe. Advances in Forest Fire Research. 2014:1874-1882. DOI: 10.14195/978-989-26-0884-6_209
  71. 71. Catry FX, Rego FC, Bação FL, Moreira F. Modeling and mapping wildfire ignition risk in Portugal. International Journal of Wildland Fire. 2009;18:921-931
  72. 72. Romero-Calcerrada R, Novillo CJ, Millington JDA, Gomez-Jimenez I. GIS analysis of spatial patterns of human-caused wildfire ignition risk in the SW of Madrid (Central Spain). Landscape Ecology. 2008;23:341-354
  73. 73. Arab A, Khodaei A, Eskandarpour R, Thompson MP, Wei Y. Three lines of Defense for wildfire risk Management in Electric Power Grids: A review. IEEE Access. 2021;9:61577-61593
  74. 74. Xu K, Zhang X, Chen Z, Wu W, Li T. Risk assessment for wildfire occurrence in high-voltage power line corridors by using remote-sensing techniques: A case study in Hubei Province, China. International Journal of Remote Sensing. 2016;37:4818-4837
  75. 75. Park A, Rajabi F, Weber R. Slash or Burn: Power Line and Vegetation Classification for Wildfire Prevention [Internet]. 2021. Available from: http://arxiv.org/abs/2105.03804
  76. 76. Su C, Wu X, Guo Y, Lai CS, Xu L, Zhao X. Automatic multi-source data fusion technique of powerline corridor using UAV Lidar. In: 2022 IEEE International Smart Cities Conference (ISC2), Pafos, Cyprus. 2022. pp. 1-5. DOI: 10.1109/ISC255366.2022.9922293
  77. 77. Ghorbanzadeh O, Blaschke T, Gholamnia K, Aryal J. Forest fire susceptibility and risk mapping using social/infrastructural vulnerability and environmental variables. Fire. 2019;2:1-27
  78. 78. Pourghasemi H, Beheshtirad M, Pradhan B. A comparative assessment of prediction capabilities of modified analytical hierarchy process (M-AHP) and Mamdani fuzzy logic models using Netcad-GIS for forest fire susceptibility mapping. Geomatics, Natural Hazards and Risk. 2016;7:861-885
  79. 79. Lamat R, Kumar M, Kundu A, Lal D. Forest fire risk mapping using analytical hierarchy process (AHP) and earth observation datasets: A case study in the mountainous terrain of Northeast India. SN Applied Sciences. 2021;3:1-15 [Internet] [cited 2022 Dec 26]. Available from: https://link.springer.com/article/10.1007/s42452-021-04391-0
  80. 80. Serda M, Becker FG, Cleary M, Team RM, Holtermann H, The D, et al. Wildfire susceptibility evaluation by integrating an analytical network process approach into GIS-based analyses. G. Balint, Antala B, Carty C, Mabieme J-MA, Amar IB, Kaplanova A, editors. International Journal on Advanced Science, Engineering and Technology. 2018;7:48-53 [Internet] [cited 2022 Dec 26]. Available from: https://uni-salzburg.elsevierpure.com/en/publications/wildfire-susceptibility-evaluation-by-integrating-an-analytical-n-2
  81. 81. Mohajane M, Costache R, Karimi F, Bao Pham Q , Essahlaoui A, Nguyen H, et al. Application of remote sensing and machine learning algorithms for forest fire mapping in a Mediterranean area. Ecological Indicators. 2021;129:107869 [Internet] [cited 2022 Dec 16]. DOI: 10.1016/j.ecolind.2021.107869
  82. 82. Karpatne A, Ebert-Uphoff I, Ravela S, Babaie HA, Kumar V. Machine learning for the geosciences: Challenges and opportunities. IEEE Transactions on Knowledge and Data Engineering. 2017;31:1544-1554 [Internet] [cited 2022 Dec 28]. Available from: https://arxiv.org/abs/1711.04708v1
  83. 83. Rishickesh R, Shahina A, Nayeemulla KA. Predicting forest fires using supervised and ensemble machine learning algorithms. International Journal of Recent Technology and Engineering. 2019;8:3697-3705
  84. 84. Dutta R, Das A, Aryal J. Big data integration shows Australian bush-fire frequency is increasing significantly. Royal Society Open Science. 2016;3(2). DOI: 10.1098/RSOS.150241
  85. 85. de Souza FT, Koerner TC, Chlad R. A data-based model for predicting wildfires in Chapada das mesas National Park in the state of Maranhão. Environ. Earth Science. 2015;74:3603-3611 [Internet] [cited 2022 Dec 26]. Available from: https://link.springer.com/article/10.1007/s12665-015-4421-8
  86. 86. Kim SJ, Lim C-H, Kim G, Lee J, Geiger T, Rahmati O, et al. Multi-temporal analysis of Forest fire probability using socio-economic and environmental variables. Remote Sensing. 2019;11:86
  87. 87. Oliveira S, Oehler F, San-Miguel-Ayanz J, Camia A, Pereira JMC. Modeling spatial patterns of fire occurrence in Mediterranean Europe using multiple regression and random Forest. Forest Ecology and Management. 2012;275:117-129 [Internet] [cited 2022 Dec 26]. Available from: https://www.researchgate.net/publication/236781184_Modeling_spatial_patterns_of_fire_occurrence_in_Mediterranean_Europe_using_Multiple_Regression_and_Random_Forest
  88. 88. Miller RK, Field CB, Mach KJ. Factors influencing adoption and rejection of fire hazard severity zone maps in California. International Journal of Disaster Risk Reduction. 2020;50. DOI: 10.1016/J.IJDRR.2020.101686
  89. 89. Chuvieco Emilio. Wildland fire danger: estimation and mapping : the role of remote sensing data. 2003 [cited 2022 Dec 28]. pp. 264. Available from: https://books.google.com/books/about/Wildland_Fire_Danger_Estimation_and_Mapp.html?hl=el&id=a2LF_1a8IPcC
  90. 90. Rothermel RC, Wilson RA, Morris GA, Sackett SS. Modeling Moisture Content of Fine Dead Wildland Fuels: Input to the BEHAVE Fire Prediction System. Research Paper INT-359. Ogden, UT: U.S. Department of Agriculture, Forest Service, Intermountain Research Station; 1986. 61 p. DOI: 10.2737/INT-RP-359
  91. 91. Piñol J, Filella I, Ogaya R, Peñuelas J. Ground-based spectroradiometric estimation of live fine fuel moisture of Mediterranean plants. Agricultural and Forest Meteorology. 1998;90:173-186
  92. 92. Chuvieco E, Deshayes M, Stach N, Cocero D, Riaño D. Short-term fire risk: Foliage moisture content estimation from satellite data. Remote Sensing of Large Wildfires. 1999:17-38. DOI: 10.1007/978-3-642-60164-4_3
  93. 93. Luo K, Quan X, He B, Yebra M. Effects of live fuel moisture content on wildfire occurrence in fire-prone regions over Southwest China. Forests (MDPI AG). 2019;10:887 [Internet]. DOI: 10.3390/f10100887
  94. 94. Nolan RH, Boer MM, Resco De Dios V, Caccamo G, Bradstock RA. Large-scale, dynamic transformations in fuel moisture drive wildfire activity across southeastern Australia. Geophysical Research Letters. 2016;43:4229-4238 [Internet] [cited 2022 Dec 28]. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/2016GL068614
  95. 95. Chuvieco E, Aguado I, Salas J, García M, Yebra M, Oliva P. Satellite remote sensing contributions to wildland fire science and management. Current Forestry Reports. 2020;6:81-96 [Internet] [cited 2022 Dec 28]. Available from: https://link.springer.com/article/10.1007/s40725-020-00116-5
  96. 96. Boer MM, Nolan RH, Resco De Dios V, Clarke H, Price OF, Bradstock RA. Changing weather extremes call for early warning of potential for catastrophic fire. Earths. Futures. 2017;5:1196-1202 [Internet] [cited 2022 Dec 28]. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/2017EF000657
  97. 97. Caccamo G, Chisholm LA, Bradstock RA, Puotinen ML, Pippen BG, Caccamo G, et al. Monitoring live fuel moisture content of heathland, shrubland and sclerophyll forest in South-Eastern Australia using MODIS data. International Journal of Wildland Fire. 2011;21:257-269 [Internet] [cited 2022 Dec 29]. Available from: https://www.publish.csiro.au/wf/WF11024
  98. 98. Argañaraz JP, Landi MA, Bravo SJ, Gavier-Pizarro GI, Scavuzzo CM, Bellis LM. Estimation of live fuel moisture content from MODIS images for fire danger assessment in southern Gran Chaco. IEEE Journal of Selected Topics in Applied Earth Observations and Remote Sensing. 2016;9:5339-5349
  99. 99. Jurdao S, Yebra M, Guerschman JP, Chuvieco E. Regional estimation of woodland moisture content by inverting radiative transfer models. Remote Sensing of Environment. 2013;132:59-70
  100. 100. Yebra M, Quan X, Riaño D, Rozas Larraondo P, van Dijk AIJM, Cary GJ. A fuel moisture content and flammability monitoring methodology for continental Australia based on optical remote sensing. Remote Sensing of Environment. 2018;212:260-272
  101. 101. Zhu L, Webb GI, Yebra M, Scortechini G, Miller L, Petitjean F. Live fuel moisture content estimation from MODIS: A deep learning approach. ISPRS Journal of Photogrammetry and Remote Sensing. 2021;179:81-91
  102. 102. Rao K, Williams AP, Flefil JF, Konings AG. SAR-enhanced mapping of live fuel moisture content. Remote Sensing of Environment. 2020;245:111797
  103. 103. Marino E, Yebra M, Guillén-Climent M, Algeet N, Tomé JL, Madrigal J, et al. Investigating live fuel moisture content estimation in fire-prone shrubland from remote sensing using empirical modelling and RTM simulations. Remote Sensing (MDPI AG). 2020;12:2251 [Internet]. DOI: 10.3390/rs12142251
  104. 104. Hunt ER, Rock BN. Detection of changes in leaf water content using near- and middle-infrared reflectances. Remote Sensing of Environment. 1989;30:43-54
  105. 105. Seelig HD, Hoehn A, Stodieck LS, Klaus DM, Adams WW, Emery WJ. The assessment of leaf water content using leaf reflectance ratios in the visible, near-, and short-wave-infrared. International Journal of Remote Sensing. 2008;29:3701-3713
  106. 106. Cunill Camprubí À, González-Moreno P, de Dios VR. Live fuel moisture content mapping in the Mediterranean Basin using random forests and combining MODIS spectral and thermal data. Remote Sensing (MDPI AG). 2022;14:3162 [Internet]. DOI: 10.3390/rs14133162
  107. 107. Costa-Saura JM, Balaguer-Beser Á, Ruiz LA, Pardo-Pascual JE, Soriano-Sancho JL. Empirical models for spatio-temporal live fuel moisture content estimation in mixed Mediterranean vegetation areas using sentinel-2 indices and meteorological data. Remote Sensing (MDPI AG). 2021;13:3726 [Internet]. DOI: 10.3390/rs13183726
  108. 108. Shu Q , Quan X, Yebra M, Liu X, Wang L, Zhang Y. Evaluating the Sentinel-2A satellite data for fuel moisture content retrieval. In: IGARSS 2019 - 2019 IEEE International Geoscience and Remote Sensing Symposium, Yokohama, Japan. 2019. pp. 9416-9419. DOI: 10.1109/IGARSS.2019.8900104
  109. 109. Fan L, Wigneron JP, Xiao Q , Al-Yaari A, Wen J, Martin-StPaul N, et al. Evaluation of microwave remote sensing for monitoring live fuel moisture content in the Mediterranean region. Remote Sensing of Environment. 2018;205:210-223
  110. 110. Wang L, Quan X, He B, Yebra M, Xing M, Liu X. Assessment of the dual Polarimetric sentinel-1A data for Forest fuel moisture content estimation. Remote Sensing. 2019;11:1568 [Internet] [cited 2022 Dec 30]. Available from: https://www.mdpi.com/2072-4292/11/13/1568/htm
  111. 111. Yebra M, Dennison PE, Chuvieco E, Riaño D, Zylstra P, Hunt ER, et al. A global review of remote sensing of live fuel moisture content for fire danger assessment: Moving towards operational products. Remote Sensing of Environment. 2013;136:455-468
  112. 112. Alexander ME, Cole FV. Rating fire danger in Alaska ecosystems: CFFDRS provides an invaluable guide to systematically evaluating burning conditions. Fireline (Fort Wainwright, Alaska: USDI Bureau of Land Management). 2001;12(4):2-3
  113. 113. Groot WJD, Field RD, Brady MA, Roswintiarti O, Mohamad M. Development of the Indonesian and Malaysian fire danger rating systems. Mitigation and Adaptation Strategies for Global Change. 2007;12:165-180
  114. 114. San-Miguel-Ayanz J, Barbosa P, Libertá G, Schmuck G, Schulte E, Bucella P. The European forest fire information system: A European strategy towards forest fire management. In: Proceedings of the 3rd International Wildland Fire Conference, Sydney, Australia. Washington, DC: US Department of the Interior, Bureau of Land Management CD-ROM; 2003
  115. 115. Viegas DX, Bovio G, Ferreira A, Nosenzo A, Sol B. Comparative study of various methods of fire danger evaluation in southern Europe. International Journal of Wildland Fire. 2000;9:235-246
  116. 116. San-Miguel-Ayanz J, Schulte E, Schmuck G, Camia A, Strobl P, Liberta G, et al. Comprehensive monitoring of wildfires in Europe: The European forest fire information system (EFFIS). In: Tiefenbacher J, editor. Approaches to Managing Disaster - Assessing Hazards, Emergencies and Disaster Impacts. 2012
  117. 117. Chowdhury EH, Hassan QK. Development of a new daily-scale forest fire danger forecasting system using remote sensing data. Remote Sensing. 2015;7:2431-2448
  118. 118. Sirca C, Salis M, Arca B, Duce P, Spano D. Assessing the performance of fire danger indexes in a Mediterranean area. iForest. 2018;11:563 [Internet] [cited 2022 Dec 9]. Available from: https://iforest.sisef.org/abstract/?id=ifor2679-011
  119. 119. Fiorucci P, Gaetani F, Minciardi R. Development and application of a system for dynamic wildfire risk assessment in Italy. Environmental Modelling & Software. 2008;23:690-702
  120. 120. Laneve G, Pampanoni V, Shaik RU. The daily fire Hazard index: A fire danger rating method for Mediterranean areas. Remote Sensing. 2020;12:2356 [Internet] [cited 2022 Dec 31]. Available from: https://www.mdpi.com/2072-4292/12/15/2356/htm
  121. 121. Liu X, He B, Quan X, Yebra M, Qiu S, Yin C, et al. Near real-time extracting wildfire spread rate from Himawari-8 satellite data. Remote Sensing. 2018;10:1-15
  122. 122. Ying L, Shen Z, Yang M, Piao S. Wildfire detection probability of MODIS fire products under the constraint of environmental factors: A study based on confirmed ground wildfire records. Remote Sensing (MDPI AG). 2019;11:3031 [Internet]. DOI: 10.3390/rs11243031
  123. 123. Lin Z, Chen F, Niu Z, Li B, Yu B, Jia H, et al. An active fire detection algorithm based on multi-temporal FengYun-3C VIRR data. Remote Sensing of Environment. 2018;211:376-387
  124. 124. Flannigan MD, Vonder Haar TH. Forest fire monitoring using NOAA satellite AVHRR. Canadian Journal of Forest Research. 1986;16:975-982
  125. 125. Justice CO, Korontzi SA. A review of satellite fire monitoring and the requirements for global environmental change research. In: Ahem F, Goldammer G, Justice CO, editors. Global and Regional Vegetation Fire Monitoring From Space: Planning a Coordinated International Effort. The Hague, The Netherlands: SPB Academic Publishing; 2001. pp. 1-18
  126. 126. Giglio L, Csiszar I, Justice CO. Global distribution and seasonality of active fires as observed with the Terra and Aqua moderate resolution imaging spectroradiometer (MODIS) sensors. Journal of Geophysical Research: Biogeosciences. 2006;111(2). DOI: 10.1029/2005JG000142
  127. 127. Schroeder W, Prins E, Giglio L, Csiszar I, Schmidt C, Morisette J, et al. Validation of GOES and MODIS active fire detection products using ASTER and ETM+ data. Remote Sensing of Environment. 2008;112:2711-2726
  128. 128. Giglio L, Schroeder W, Justice CO. The collection 6 MODIS active fire detection algorithm and fire products. Remote Sensing of Environment. 2016;178:31-41
  129. 129. Morisette JT, Giglio L, Csiszar I, Justice CO. Validation of the MODIS active fire product over southern Africa with ASTER data. International Journal of Remote Sensing. 2005;26:4239-4264
  130. 130. Forghani A, Reddy S, Thankappan M, Cechet B. Validation of MODIS and AVHRR fire detections in Australia. International Journal of Geoinformatics. 2021;17:117-131
  131. 131. Cahoon DR, Stocks BJ, Alexander ME, Baum BA, Goldammer JG. Wildland Fire Detection from Space: Theory and Application. 2000 [cited 2023 Jan 5]. pp. 151-69. Available from: https://link.springer.com/chapter/10.1007/0-306-47959-1_9
  132. 132. Maier SW, Russell-Smith J, Edwards AC, Yates C. Sensitivity of the MODIS fire detection algorithm (MOD14) in the savanna region of the Northern Territory, Australia. ISPRS Journal of Photogrammetry and Remote Sensing. 2013;76:11-16
  133. 133. Hawbaker TJ, Radeloff VC, Syphard AD, Zhu Z, Stewart SI. Detection rates of the MODIS active fire product in the United States. Remote Sensing of Environment. 2008;112:2656-2664
  134. 134. Justice CO, Giglio L, Korontzi S, Owens J, Morisette JT, Roy D, et al. The MODIS fire products. Remote Sensing of Environment. 2002;83:244-262
  135. 135. Schroeder W, Oliva P, Giglio L, Quayle B, Lorenz E, Morelli F. Active fire detection using Landsat-8/OLI data. Remote Sensing of Environment. 2016;185:210-220
  136. 136. Gargiulo M, Dell’ Aglio DAG, Iodice A, Riccio D, Ruello G. A CNN-based super-resolution technique for active fire detection on Sentinel-2 data. In: 2019 PhotonIcs & Electromagnetics Research Symposium - Spring (PIERS-Spring), Rome, Italy. 2019. pp. 418-426. DOI: 10.1109/PIERS-Spring46901.2019.9017857
  137. 137. Xu G, Zhong X. Real-time wildfire detection and tracking in Australia using geostationary satellite: Himawari-8. Remote Sensing Letters. 2017;8:1052-1061
  138. 138. Maeda N, Tonooka H. Early stage Forest fire detection from Himawari-8 AHI images using a modified MOD14 algorithm combined with machine learning. Sensors. 2022;23:210 [Internet] [cited 2023 Jan 5]. Available from: https://www.mdpi.com/1424-8220/23/1/210
  139. 139. Hall JV, Zhang R, Schroeder W, Huang C, Giglio L. Validation of GOES-16 ABI and MSG SEVIRI active fire products. International Journal of Applied Earth Observation and Geoinformation. 2019;83:101928
  140. 140. Wang W, Qu JJ, Hao X, Liu Y, Sommers WT. An improved algorithm for small and cool fire detection using MODIS data: A preliminary study in the southeastern United States. Remote Sensing of Environment. 2007;108:163-170
  141. 141. Zajkowski T. Selection of appropriate class UAS/sensors to support fire monitoring: Experiences in the United States. Handbook of Unmanned Aerial Vehicles. 2015:2723-2754. DOI: 10.1007/978-90-481-9707-1_73/COVER
  142. 142. Nex F, Remondino F. Preface: Latest developments, methodologies, and applications based on UAV platforms. Drones. 2019;3:26 [Internet] [cited 2023 Jan 6]. Available from: https://www.mdpi.com/2504-446X/3/1/26/htm
  143. 143. Watts AC, Ambrosia VG, Hinkley EA. Unmanned aircraft systems in remote sensing and scientific research: Classification and considerations of use. Remote Sensing. 2012;4:1671-1692
  144. 144. Yuan C, Liu Z, Zhang Y. UAV-based forest fire detection and tracking using image processing techniques. 2015 international conference on unmanned aircraft systems. ICUAS. 2015;2015:639-643
  145. 145. Akhloufi MA, Couturier A, Castro NA. Unmanned aerial vehicles for wildland fires: Sensing, perception, cooperation and assistance. Drones. 2021;5:15 [Internet] [cited 2023 Jan 6]. Available from: https://www.mdpi.com/2504-446X/5/1/15/htm
  146. 146. Randerson JT, Chen Y, van der Werf GR, Rogers BM, Morton DC. Global burned area and biomass burning emissions from small fires. Journal of Geophysical Research – Biogeosciences. 2012;117:4012 [Internet] [cited 2023 Jan 7]. Available from: https://onlinelibrary.wiley.com/doi/full/10.1029/2012JG002128
  147. 147. Giglio L, Loboda T, Roy DP, Quayle B, Justice CO. An active-fire based burned area mapping algorithm for the MODIS sensor. Remote Sensing of Environment. 2009;113:408-420
  148. 148. Langmann B, Duncan B, Textor C, Trentmann J, van der Werf GR. Vegetation fire emissions and their impact on air pollution and climate. Atmospheric Environment. 2009;43:107-116
  149. 149. Yang W, Jiang X. High-resolution estimation of air pollutant emissions from vegetation burning in China (2000-2018). Frontiers in Environmental Science. 2022;10:1333
  150. 150. Pereira JMC, Sousa AMO, Sá ACL, Martín MP, Chuvieco E. Regional-scale burnt area mapping in Southern Europe using NOAA-AVHRR 1 km data. Remote Sensing of Large Wildfires. 1999:139-155. DOI: 10.1007/978-3-642-60164-4_8
  151. 151. Al-Rawi KR, Casanova JL, Calle A. Burned area mapping system and fire detection system, based on nueral networks and NOAA-AVHRR imagery. International Journal of Remote Sensing. 2001;22:2015-2032
  152. 152. Grégoire JM, Tansey K, Silva JMN. The GBA2000 initiative: Developing a global burnt area database from SPOT-VEGETATION imagery. International Journal of Remote Sensing. 2003;24:1369-1376
  153. 153. Tansey K, Grégoire JM, Stroppiana D, Sousa A, Silva J, Pereira JMC, et al. Vegetation burning in the year 2000: Global burned area estimates from SPOT VEGETATION data. Journal of Geophysical Research: Atmospheres. 2004;109(D14). DOI: 10.1029/2003JD003598
  154. 154. Simon M, Plummer S, Fierens F, Hoelzemann JJ, Arino O. Burnt area detection at global scale using ATSR-2: The GLOBSCAR products and their qualification. Journal of Geophysical Research: Atmospheres. 2004;109(D14). DOI: 10.1029/2003JD003622
  155. 155. Roy DP, Jin Y, Lewis PE, Justice CO. Prototyping a global algorithm for systematic fire-affected area mapping using MODIS time series data. Remote Sensing of Environment. 2005;97:137-162
  156. 156. Boschetti L, Roy D, Barbosa P, Boca R, Justice C. A MODIS assessment of the summer 2007 extent burned in Greece. International Journal of Remote Sensing. 2008;29:2433-2436
  157. 157. Roy DP, Boschetti L. Southern Africa validation of the MODIS, L3JRC, and GlobCarbon burned-area products. IEEE Transactions on Geoscience and Remote Sensing. 2009;47:1032-1044
  158. 158. Padilla M, Stehman SV, Ramo R, Corti D, Hantson S, Oliva P, et al. Comparing the accuracies of remote sensing global burned area products using stratified random sampling and estimation. Remote Sensing of Environment. 2015;160:114-121
  159. 159. San-Miguel-Ayanz J, Pereira JMC, Boca R, Strobl P, Kucera J, Pekkarinen A. Forest fires in the European Mediterranean region: Mapping and analysis of burned areas. Earth Observation of Wildland Fires in Mediterranean Ecosystems. 2009:189-203. DOI: 10.1007/978-3-642-01754-4_13
  160. 160. Roy DP, Wulder MA, Loveland TR, Woodcock CE, Allen RG, Anderson MC, et al. Landsat-8: Science and product vision for terrestrial global change research. Remote Sensing of Environment. 2014;145:154-172
  161. 161. Wang Q , Blackburn GA, Onojeghuo AO, Dash J, Zhou L, Zhang Y, et al. Fusion of Landsat 8 OLI and Sentinel-2 MSI data. IEEE Transactions on Geoscience and Remote Sensing. 2017;55:3885-3899
  162. 162. Roy DP, Huang H, Boschetti L, Giglio L, Yan L, Zhang HH, et al. Landsat-8 and Sentinel-2 burned area mapping - A combined sensor multi-temporal change detection approach. Remote Sensing of Environment. 2019;231. DOI: 10.1016/j.rse.2019.111254
  163. 163. Drusch M, del Bello U, Carlier S, Colin O, Fernandez V, Gascon F, et al. Sentinel-2: ESA’s optical high-resolution Mission for GMES operational services. Remote Sensing of Environment. 2012;120:25-36
  164. 164. Filipponi F. Exploitation of Sentinel-2 time series to map burned areas at the national level: A case study on the 2017 Italy wildfires. Remote Sensing (MDPI AG). 2019;11:622 [Internet]. DOI: 10.3390/rs11060622
  165. 165. Quintano C, Fernández-Manso A, Fernández-Manso O. Combination of Landsat and Sentinel-2 MSI data for initial assessing of burn severity. International Journal of Applied Earth Observation and Geoinformation. 2018;64:221-225
  166. 166. Syifa M, Panahi M, Lee C-W. Mapping of post-wildfire burned area using a hybrid algorithm and satellite data: The case of the camp fire wildfire in California, USA. Remote Sensing (MDPI AG). 2020;12:623 [Internet]. DOI: 10.3390/rs12040623
  167. 167. Katagis T, Gitas IZ, Mitri GH. An object-based approach for fire history reconstruction by using three generations of landsat sensors. Remote Sensing. 2014;6:5480-5496
  168. 168. Boschetti L, Roy DP, Justice CO, Humber ML. MODIS–Landsat fusion for large area 30 m burned area mapping. Remote Sensing of Environment. 2015;161:27-42
  169. 169. Chuvieco E, Mouillot F, van der Werf GR, San Miguel J, Tanasse M, Koutsias N, et al. Historical background and current developments for mapping burned area from satellite earth observation. Remote Sensing of Environment. 2019;225:45-64
  170. 170. French NHF, Bourgeau-Chavez LL, Wang Y, Kasischke ES. Initial observations of radarsat imagery at fire-disturbed sites in interior Alaska. Remote Sensing of Environment. 1999;68:89-94
  171. 171. Siegert F, Ruecker G. Use of multitemporal ERS-2 SAR images for identification of burned scars in south-east Asian tropical rainforest. International Journal of Remote Sensing. 2000;21:831-837
  172. 172. Gimeno M, San-Miguel-Ayanz J. Evaluation of RADARSAT-1 data for identification of burnt areas in southern Europe. Remote Sensing of Environment. 2004;92:370-375
  173. 173. Gimeno M, San-Miguel-Ayanz J, Schmuck G. Identification of burnt areas in Mediterranean forest environments from ERS-2 SAR time series. International Journal of Remote Sensing. 2004;25:4873-4888
  174. 174. de Luca G, Silva JMN, Modica G. Regional-scale burned area mapping in Mediterranean regions based on the multitemporal composite integration of Sentinel-1 and Sentinel-2 data. GIScience & Remote Sensing. 2022;59:1678-1705 [Internet] [cited 2023 Jan 8]. Available from: https://www.tandfonline.com/doi/abs/10.1080/15481603.2022.2128251
  175. 175. Keeley JE. Fire intensity, fire severity and burn severity: A brief review and suggested usage. International Journal of Wildland Fire. 2009;18:116-126
  176. 176. Benavides-Solorio J, MacDonald LH. Post-fire runoff and erosion from simulated rainfall on small plots, Colorado front range. Hydrological Processes. 2001;15:2931-2952 [Internet] [cited 2023 Jan 9]. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/hyp.383
  177. 177. Martin DA, Moody JA. Comparison of soil infiltration rates in burned and unburned mountainous watersheds. Hydrological Processes. 2001;15:2893-2903
  178. 178. Key CH, Benson NC. Landscape assessment (LA) sampling and analysis methods. In: Lutes DC, Keane RE, Caratti JF, Key CH, Benson NC, Sutherland S, et al., editors. FIREMON: Fire Effects Monitoring and Inventory System. Gen. Tech. Rep. RMRS-GTR-164-CD. Fort Collins, CO: U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station; 2006. p. LA-1-55
  179. 179. Kasischke ES, Turetsky MR, Ottmar RD, French NHF, Hoy EE, Kane ES, et al. Evaluation of the composite burn index for assessing fire severity in Alaskan black spruce forests. International Journal of Wildland Fire. 2008;17:515-526 [Internet] [cited 2023 Jan 9]. Available from: https://www.publish.csiro.au/wf/WF08002
  180. 180. Chuvieco E, Riaño D, Danson FM, Martin P. Use of a radiative transfer model to simulate the postfire spectral response to burn severity. Journal of Geophysical Research: Biogeosciences. 2006;111(G4):4-9. DOI: 10.1029/2005JG000143
  181. 181. Kokaly RF, Rockwell BW, Haire SL, King TVV. Characterization of post-fire surface cover, soils, and burn severity at the Cerro Grande fire, New Mexico, using hyperspectral and multispectral remote sensing. Remote Sensing of Environment. 2007;106:305-325
  182. 182. Roy DP, Boschetti L, Trigg SN. Remote sensing of fire severity: Assessing the performance of the normalized burn ratio. IEEE Geoscience and Remote Sensing Letters. 2006;3:112-116
  183. 183. Barrett K, Kasischke ES, McGuire AD, Turetsky MR, Kane ES. Modeling fire severity in black spruce stands in the Alaskan boreal forest using spectral and non-spectral geospatial data. Remote Sensing of Environment. 2010;114:1494-1503
  184. 184. Hoy EE, French NHF, Turetsky MR, Trigg SN, Kasischke ES. Evaluating the potential of Landsat TM/ETM+ imagery for assessing fire severity in Alaskan black spruce forests. International Journal of Wildland Fire. 2008;17:500-514
  185. 185. Verbyla D, Lord R. Estimating post-fire organic soil depth in the Alaskan boreal forest using the normalized burn ratio. International Journal of Remote Sensing. 2008;29:3845-3853
  186. 186. Hall RJ, Freeburn JT, de Groot WJ, Pritchard JM, Lynham TJ, Landry R. Remote sensing of burn severity: Experience from western Canada boreal fires. International Journal of Wildland Fire. 2008;17:476-489
  187. 187. Veraverbeke S, Verstraeten WW, Lhermitte S, Goossens R, Veraverbeke S, Verstraeten WW, et al. Evaluating Landsat thematic mapper spectral indices for estimating burn severity of the 2007 Peloponnese wildfires in Greece. International Journal of Wildland Fire. 2010;19:558-569 [Internet] [cited 2023 Jan 9]. Available from: https://www.publish.csiro.au/wf/WF09069
  188. 188. Veraverbeke S, Lhermitte S, Verstraeten WW, Goossens R. Evaluation of pre/post-fire differenced spectral indices for assessing burn severity in a mediterranean environment with landsat thematic mapper. International Journal of Remote Sensing. 2011;32:3521-3537
  189. 189. Miller JD, Thode AE. Quantifying burn severity in a heterogeneous landscape with a relative version of the delta normalized burn ratio (dNBR). Remote Sensing of Environment. 2007;109:66-80
  190. 190. Soverel NO, Coops NC, Perrakis DDB, Daniels LD, Gergel SE. The transferability of a dNBR-derived model to predict burn severity across 10 wildland fires in western Canada. International Journal of Wildland Fire. 2011;20:518-531
  191. 191. Wu Z, He HS, Liang Y, Cai L, Lewis BJ. Determining relative contributions of vegetation and topography to burn severity from LANDSAT imagery. Environmental Management. 2013;52:821-836
  192. 192. Epting J, Verbyla D, Sorbel B. Evaluation of remotely sensed indices for assessing burn severity in interior Alaska using Landsat TM and ETM+. Remote Sensing of Environment. 2005;96:328-339
  193. 193. Soverel NO, Perrakis DDB, Coops NC. Estimating burn severity from Landsat dNBR and RdNBR indices across western Canada. Remote Sensing of Environment. 2010;114:1896-1909
  194. 194. Slaton MR, Hunt ER, Smith WK. Estimating near-infrared leaf reflectance from leaf structural characteristics. American Journal of Botany. 2001;88:278-284 [Internet] [cited 2023 Jan 10]. Available from: https://onlinelibrary.wiley.com/doi/full/10.2307/2657019
  195. 195. Quintano C, Fernández-Manso A, Calvo L, Marcos E, Valbuena L. Land surface temperature as potential indicator of burn severity in forest Mediterranean ecosystems. International Journal of Applied Earth Observation and Geoinformation. 2015;36:1-12
  196. 196. Zheng Z, Zeng Y, Li S, Huang W. A new burn severity index based on land surface temperature and enhanced vegetation index. International Journal of Applied Earth Observation and Geoinformation. 2016;45:84-94
  197. 197. Veraverbeke S, Verstraeten WW, Lhermitte S, van de Kerchove R, Goossens R, Veraverbeke S, et al. Assessment of post-fire changes in land surface temperature and surface albedo, and their relation with fire–burn severity using multitemporal MODIS imagery. International Journal of Wildland Fire. 2012;21:243-256 [Internet] [cited 2023 Jan 10]. Available from: https://www.publish.csiro.au/wf/WF10075
  198. 198. Horler DNH, Dockray M, Barber J. The red edge of plant leaf reflectance. International Journal of Remote Sensing. 2007;4:273-288 [Internet] [cited 2023 Jan 10]. Available from: https://www.tandfonline.com/doi/abs/10.1080/01431168308948546
  199. 199. Fernández-Manso A, Fernández-Manso O, Quintano C. SENTINEL-2A red-edge spectral indices suitability for discriminating burn severity. International Journal of Applied Earth Observation and Geoinformation. 2016;50:170-175
  200. 200. Filipponi F. BAIS2: Burned area index for Sentinel-2. Proceedings. 2018;2:364 [Internet] [cited 2023 Jan 10]. Available from: https://www.mdpi.com/2504-3900/2/7/364/htm
  201. 201. Morgan P, Keane RE, Dillon GK, Jain TB, Hudak AT, Karau EC, et al. Challenges of assessing fire and burn severity using field measures, remote sensing and modelling. International Journal of Wildland Fire. 2014;23:1045-1060
  202. 202. Wang C, Glenn NF. Estimation of fire severity using pre- and post-fire LiDAR data in sagebrush steppe rangelands. International Journal of Wildland Fire. 2009;18:848-856 [Internet] [cited 2023 Jan 10]. Available from: https://www.publish.csiro.au/wf/WF08173
  203. 203. Wulder MA, White JC, Alvarez F, Han T, Rogan J, Hawkes B. Characterizing boreal forest wildfire with multi-temporal Landsat and LIDAR data. Remote Sensing of Environment. 2009;113:1540-1555
  204. 204. Fernandez-Manso A, Quintano C, Roberts DA. Burn severity analysis in Mediterranean forests using maximum entropy model trained with EO-1 Hyperion and LiDAR data. ISPRS Journal of Photogrammetry and Remote Sensing. 2019;155:102-118
  205. 205. Kane VR, North MP, Lutz JA, Churchill DJ, Roberts SL, Smith DF, et al. Assessing fire effects on forest spatial structure using a fusion of Landsat and airborne LiDAR data in Yosemite National Park. Remote Sensing of Environment. 2014;151:89-101
  206. 206. Viedma O, Almeida DRA, Moreno JM. Postfire tree structure from high-resolution LiDAR and RBR sentinel 2A fire severity metrics in a pinus halepensis-dominated burned stand. Remote Sensing. 2020;12:1-21
  207. 207. Carvajal-Ramírez F, da Silva JRM, Agüera-Vega F, Martínez-Carricondo P, Serrano J, Moral FJ. Evaluation of fire severity indices based on pre- and post-fire multispectral imagery sensed from UAV. Remote Sensing (MDPI AG). 2019;11:993 [Internet]. DOI: 10.3390/rs11090993
  208. 208. McKenna P, Erskine PD, Lechner AM, Phinn S. Measuring fire severity using UAV imagery in semi-arid Central Queensland, Australia. International Journal of Remote Sensing. 2017;38:4244-4264 [Internet] [cited 2023 Jan 11]. Available from: https://www.tandfonline.com/doi/abs/10.1080/01431161.2017.1317942
  209. 209. Woo H, Acuna M, Madurapperuma B, Jung G, Woo C, Park J. Application of maximum likelihood and spectral angle mapping classification techniques to evaluate forest fire severity from UAV multi-spectral images in South Korea. Sensors and Materials. 2021;33:3745-3760
  210. 210. Qi Y, Coops NC, Daniels LD, Butson CR. Assessing the effects of burn severity on post-fire tree structures using the fused drone and mobile laser scanning point clouds. Frontiers in Environmental Science. 2022;10:949442. DOI: 10.3389/FENVS.2022.949442/BIBTEX
  211. 211. Ye J, Cui Z, Zhao F, Liu Q. Automated extraction of Forest burn severity based on light and small UAV visible remote sensing images. Forests. 2022;13
  212. 212. Fraser RH, van der Sluijs J, Hall RJ. Calibrating satellite-based indices of burn severity from UAV-derived metrics of a burned boreal forest in NWT, Canada. Remote Sensing. 2017;9
  213. 213. Shin J, Seo W, Kim T, Park J, Woo C. Using UAV multispectral images for classification of forest burn severity-A case study of the 2019 Gangneung forest fire. Forests (MDPI AG). 2019;10:1025 [Internet]. DOI: 10.3390/f10111025
  214. 214. Wilson CJ, William Carey J, Beeson PC, Gard MO, Lane LJ. A GIS-based hillslope erosion and sediment delivery model and its application in the Cerro Grande burn area. Hydrological Processes. 2001;15:2995-3010
  215. 215. Ruiz-Gallardo JR, Castaño S, Calera A. Application of remote sensing and GIS to locate priority intervention areas after wildland fires in Mediterranean systems: A case study from South-Eastern Spain. International Journal of Wildland Fire. 2004;13:241-252
  216. 216. Lewis SA, Wu JQ , Robichaud PR. Assessing burn severity and comparing soil water repellency, Hayman fire, Colorado. Hydrological Processes. 2006;20:1-16
  217. 217. Fox DM, Laaroussi Y, Malkinson LD, Maselli F, Andrieu J, Bottai L, et al. POSTFIRE: A model to map forest fire burn scar and estimate runoff and soil erosion risks. Remote Sensing Applications. 2016;4:83-91
  218. 218. Lanorte A, Cillis G, Calamita G, Nolè G, Pilogallo A, Tucci B, et al. Integrated approach of RUSLE, GIS and ESA Sentinel-2 satellite data for post-fire soil erosion assessment in Basilicata region (southern Italy). Geomatics, Natural Hazards and Risk. 2019;10:1563-1595
  219. 219. Meléndez-Pastor I, Pedreño JN, Lucas IG, Zorpas AA. A model for evaluating soil vulnerability to erosion using remote sensing data and a fuzzy logic system. In: Modern Fuzzy Control Systems and its Applications. London, UK: InTech; 2017
  220. 220. Pijl A, Reuter LEH, Quarella E, Vogel TA, Tarolli P. GIS-based soil erosion modelling under various steep-slope vineyard practices. Catena (Amst). 2020;193:104604. DOI: 10.1016/J.CATENA.2020.104604
  221. 221. D’Oleire-Oltmanns S, Marzolff I, Peter KD, Ries JB. Unmanned aerial vehicle (UAV) for monitoring soil erosion in Morocco. Remote Sensing. 2012;4:3390-3416 [Internet] [cited 2023 Jan 11]. Available from: https://www.mdpi.com/2072-4292/4/11/3390/htm
  222. 222. Eltner A, Mulsow C, Maas H-G, Eltner A, Mulsow C, Maas H-G. Quantitative measurement of soil erosion from Tls and Uav data. ISPAr. 2013;XL1:119-124 [Internet] [cited 2023 Jan 11]. Available from: https://ui.adsabs.harvard.edu/abs/2013ISPAr.XL1b.119E/abstract
  223. 223. de Luis M, González-Hidalgo JC, Raventós J. Effects of fire and torrential rainfall on erosion in a Mediterranean gorse community. Land Degradation and Development. 2003;14:203-213
  224. 224. Connor SE, Araújo J, Boski T, Gomes A, Gomes SD, Leira M, et al. Drought, fire and grazing precursors to large-scale pine forest decline. Diversity and Distributions. 2021;27:1138-1151
  225. 225. Savadogo P, Sawadogo L, Tiveau D. Effects of grazing intensity and prescribed fire on soil physical and hydrological properties and pasture yield in the savanna woodlands of Burkina Faso. Agriculture, Ecosystems and Environment. 2007;118:80-92
  226. 226. Pausas JG, Ribeiro E, Vallejo R. Post-fire regeneration variability of Pinus halepensis in the eastern Iberian Peninsula. Forest Ecology and Management. 2004;203:251-259
  227. 227. Pausas JG. Resprouting of Quercus suber in NE Spain after fire. Journal of Vegetation Science. 1997;8:703-706 [Internet] [cited 2023 Jan 12]. Available from: https://onlinelibrary.wiley.com/doi/full/10.2307/3237375
  228. 228. Ne’eman G, Lahav H, Izhaki I. Recovery of vegetation in a natural East Mediterranean pine forest on Mount Carmel, Israel as affected by management strategies. Forest Ecology and Management. 1995;75:17-26
  229. 229. Fernández-Guisuraga JM, Suárez-Seoane S, Calvo L. Modeling Pinus pinaster forest structure after a large wildfire using remote sensing data at high spatial resolution. Forest Ecology and Management. 2019;446:257-271
  230. 230. Gitas I, Mitri G, Veraverbeke S, Polychronaki A, Gitas I, Mitri G, et al. Advances in remote sensing of post-fire vegetation recovery monitoring - A review. Remote Sensing of Biomass - Principles and Applications. 2012;1:334. Available from: https://www.intechopen.com/state.item.id
  231. 231. Bullock EL, Woodcock CE, Olofsson P. Monitoring tropical forest degradation using spectral unmixing and Landsat time series analysis. Remote Sensing of Environment. 2020;238:110968
  232. 232. Stueve KM, Cerney DL, Rochefort RM, Kurth LL. Post-fire tree establishment patterns at the alpine treeline ecotone: Mount Rainier National Park, Washington, USA. Journal of Vegetation Science. 2009;20:107-120 [Internet] [cited 2023 Jan 12]. Available from: https://onlinelibrary.wiley.com/doi/full/10.1111/j.1654-1103.2009.05437.x
  233. 233. Salvia M, Ceballos D, Grings F, Karszenbaum H, Kandus P. Post-fire effects in wetland environments: Landscape assessment of plant coverage and soil recovery in the paraná river delta marshes, Argentina. Fire Ecology. 2012;8:17-37 [Internet] [cited 2023 Jan 12]. Available from: https://fireecology.springeropen.com/articles/10.4996/fireecology.0802017
  234. 234. Addink EA, van Coillie FMB, de Jong SM. Introduction to the GEOBIA 2010 special issue: From pixels to geographic objects in remote sensing image analysis. International Journal of Applied Earth Observation and Geoinformation. 2012;15:1-6
  235. 235. Polychronaki A, Gitas IZ, Minchella A. Monitoring post-fire vegetation recovery in the Mediterranean using SPOT and ERS imagery. International Journal of Wildland Fire. 2013;23:631-642 [Internet] [cited 2023 Jan 12]. Available from: https://www.publish.csiro.au/wf/WF12058
  236. 236. Mitri GH, Gitas IZ. Mapping post-fire forest regeneration and vegetation recovery using a combination of very high spatial resolution and hyperspectral satellite imagery. International Journal of Applied Earth Observation and Geoinformation. 2013;20:60-66
  237. 237. Morresi D, Vitali A, Urbinati C, Garbarino M. Forest spectral recovery and regeneration dynamics in stand-replacing wildfires of central Apennines derived from Landsat time series. Remote Sensing. 2019;11:308 [Internet] [cited 2023 Jan 12]. Available from: https://www.mdpi.com/2072-4292/11/3/308/htm
  238. 238. Villarreal ML, Norman LM, Buckley S, Wallace CSA, Coe MA. Multi-index time series monitoring of drought and fire effects on desert grasslands. Remote Sensing of Environment. 2016;183:186-197
  239. 239. Pérez-Cabello F, Montorio R, Alves DB. Remote sensing techniques to assess post-fire vegetation recovery. Current Opinion in Environmental Science & Health. 2021;21:100251
  240. 240. Veraverbeke S, Gitas I, Katagis T, Polychronaki A, Somers B, Goossens R. Assessing post-fire vegetation recovery using red–near infrared vegetation indices: Accounting for background and vegetation variability. ISPRS Journal of Photogrammetry and Remote Sensing. 2012;68:28-39
  241. 241. Lasaponara R, Abate N, Fattore C, Aromando A, Cardettini G, Di Fonzo M. On the use of Sentinel-2 NDVI time series and Google earth engine to detect land-use/land-cover changes in fire-affected areas. Remote Sensing (MDPI AG). 2022;14:4723 [Internet]. DOI: 10.3390/rs14194723
  242. 242. Rahman S, Chang HC, Magill C, Tomkins K, Hehir W. Spatio-temporal assessment of fire severity and vegetation recovery utilising Sentinel-2 imagery in New South Wales, Australia. International Geoscience and Remote Sensing Symposium (IGARSS). 2019;9960-3
  243. 243. Alegria C. Vegetation monitoring and post-fire recovery: A case study in the Centre inland of Portugal. Sustainability (Switzerland). 2022;14
  244. 244. Bright BC, Hudak AT, Kennedy RE, Braaten JD, Henareh KA. Examining post-fire vegetation recovery with Landsat time series analysis in three western North American forest types. Fire Ecology. 2019;15
  245. 245. Li X, Zhang H, Yang G, Ding Y, Zhao J. Post-fire vegetation succession and surface energy fluxes derived from remote sensing. Remote Sensing (MDPI AG). 2018;10:1000 [Internet]. DOI: 10.3390/rs10071000
  246. 246. Veraverbeke S, Somers B, Gitas I, Katagis T, Polychronaki A, Goossens R. Spectral mixture analysis to assess post-fire vegetation regeneration using Landsat thematic mapper imagery: Accounting for soil brightness variation. International Journal of Applied Earth Observation and Geoinformation. 2012;14:1-11
  247. 247. Röder A, Hill J, Duguy B, Alloza JA, Vallejo R. Using long time series of Landsat data to monitor fire events and post-fire dynamics and identify driving factors. A case study in the Ayora region (eastern Spain). Remote Sensing of Environment. 2008;112:259-273
  248. 248. Riaño D, Chuvieco E, Ustin S, Zomer R, Dennison P, Roberts D, et al. Assessment of vegetation regeneration after fire through multitemporal analysis of AVIRIS images in the Santa Monica Mountains. Remote Sensing of Environment. 2002;79:60-71
  249. 249. Sankey TT, Moffet C, Weber K. Postfire recovery of sagebrush communities: Assessment using Spot-5 and very large-scale aerial imagery. Rangeland Ecology & Management. 2008;61:598-604
  250. 250. Solans Vila JP, Barbosa P. Post-fire vegetation regrowth detection in the Deiva Marina region (Liguria-Italy) using Landsat TM and ETM+ data. Ecological Modelling. 2010;221:75-84
  251. 251. Fernandez-Carrillo A, McCaw L, Tanase MA. Estimating prescribed fire impacts and post-fire tree survival in eucalyptus forests of Western Australia with L-band SAR data. Remote Sensing of Environment. 2019;224:133-144
  252. 252. Zhou Z, Liu L, Jiang L, Feng W, Samsonov SV. Using long-term SAR backscatter data to monitor post-fire vegetation recovery in tundra environment. Remote Sensing (MDPI AG). 2019;11:2230 [Internet]. DOI: 10.3390/rs11192230
  253. 253. Chhabra A, Rüdiger C, Yebra M, Jagdhuber T, Hilton J. RADAR-vegetation structural perpendicular index (R-VSPI) for the quantification of wildfire impact and post-fire vegetation recovery. Remote Sensing (MDPI AG). 2022;14:3132 [Internet]. DOI: 10.3390/rs14133132
  254. 254. Chen W, Jiang H, Moriya K, Sakai T, Cao C. Monitoring of post-fire forest regeneration under different restoration treatments based on ALOS/PALSAR data. New Forest (Dordr). 2018;49:105-121
  255. 255. Attarchi S, Gloaguen R. Classifying complex mountainous forests with L-band SAR and Landsat data integration: A comparison among different machine learning methods in the Hyrcanian Forest. Remote Sensing. 2014;6:3624-3647 [Internet] [cited 2023 Jan 13]. Available from: https://www.mdpi.com/2072-4292/6/5/3624/htm
  256. 256. Carreiras JMB, Jones J, Lucas RM, Shimabukuro YE. Mapping major land cover types and retrieving the age of secondary forests in the Brazilian Amazon by combining single-date optical and radar remote sensing data. Remote Sensing of Environment. 2017;194:16-32
  257. 257. Kuplich TM. Classifying regenerating forest stages in Amazônia using remotely sensed images and a neural network. Forest Ecology and Management. 2006;234:1-9
  258. 258. Polychronaki A, Gitas IZ, Minchella A. Monitoring post-fire vegetation recovery in the Mediterranean using SPOT and ERS imagery. International Journal of Wildland Fire. 2014;23:631-642
  259. 259. de Luca G, Silva JMN, Modica G. Short-term temporal and spatial analysis for post-fire vegetation regrowth characterization and mapping in a Mediterranean ecosystem using optical and SAR image time-series. Geocarto International. 2022;37(27):15428-15462. DOI: 10.1080/10106049.2022.2097482
  260. 260. Lefsky MA, Cohen WB, Parker GG, Harding DJ. Lidar remote sensing for ecosystem studies: Lidar, an emerging remote sensing technology that directly measures the three-dimensional distribution of plant canopies, can accurately estimate vegetation structural attributes and should be of particular interest to forest, landscape, and global ecologists. Bioscience [Internet]. 2002;52:19-30. DOI: 10.1641/0006-3568(2002)052[0019:LRSFES]2.0.CO
  261. 261. Sato L, Gomes V, Shimabukuro Y, Keller M, Arai E, dos-Santos M, et al. Post-fire changes in forest biomass retrieved by airborne LiDAR in Amazonia. Remote Sensing (MDPI AG). 2016;8:839 [Internet]. DOI: 10.3390/rs8100839
  262. 262. Gordon CE, Price OF, Tasker EM. Mapping and exploring variation in post-fire vegetation recovery following mixed severity wildfire using airborne LiDAR. Ecological Applications. 2017;27:1618-1632 [Internet] [cited 2023 Jan 14]. Available from: https://onlinelibrary.wiley.com/doi/full/10.1002/eap.1555
  263. 263. Bolton DK, Coops NC, Wulder MA. Characterizing residual structure and forest recovery following high-severity fire in the western boreal of Canada using Landsat time-series and airborne lidar data. Remote Sensing of Environment. 2015;163:48-60
  264. 264. Meng R, Wu J, Zhao F, Cook BD, Hanavan RP, Serbin SP. Measuring short-term post-fire forest recovery across a burn severity gradient in a mixed pine-oak forest using multi-sensor remote sensing techniques. Remote Sensing of Environment. 2018;210:282-296
  265. 265. Martín-Alcón S, Coll L, de Cáceres M, Guitart L, Cabré M, Just A, et al. Combining aerial LiDAR and multispectral imagery to assess postfire regeneration types in a Mediterranean forest. Canadian Journal of Forest Research. 2015;45:856-866
  266. 266. Viana-Soto A, García M, Aguado I, Salas J. Assessing post-fire forest structure recovery by combining LiDAR data and Landsat time series in Mediterranean pine forests. International Journal of Applied Earth Observation and Geoinformation. 2022;108:102754. DOI: 10.1016/J.JAG.2022.102754
  267. 267. de las Heras J, Moya D, Vega JA, Daskalakou E, Vallejo VR, Grigoriadis N, et al. Post-Fire Management of Serotinous Pine Forests. 2012 [cited 2023 Jan 14]. pp. 121-50. Available from: https://link.springer.com/chapter/10.1007/978-94-007-2208-8_6
  268. 268. Pádua L, Guimarães N, Adão T, Sousa A, Peres E, Sousa JJ. Effectiveness of Sentinel-2 in multi-temporal post-fire monitoring when compared with UAV imagery. ISPRS International Journal of Geo-Information. 2020;9:225 [Internet] [cited 2023 Jan 14]. Available from: https://www.mdpi.com/2220-9964/9/4/225/htm
  269. 269. Fernández-Guisuraga JM, Sanz-Ablanedo E, Suárez-Seoane S, Calvo L. Using unmanned aerial vehicles in Postfire vegetation survey campaigns through large and heterogeneous areas: Opportunities and challenges. Sensors. 2018;18:586 [Internet] [cited 2023 Jan 12]. Available from: https://www.mdpi.com/1424-8220/18/2/586/htm
  270. 270. Samiappan S, Hathcock L, Turnage G, McCraine C, Pitchford J, Moorhead R. Remote sensing of wildfire using a small unmanned aerial system: Post-fire mapping, vegetation recovery and damage analysis in grand bay, Mississippi/Alabama, USA. Drones. 2019;3:1-18
  271. 271. Lazzeri G, Frodella W, Rossi G, Moretti S. Multitemporal mapping of post-fire land cover using multiplatform PRISMA hyperspectral and sentinel-UAV multispectral data: Insights from case studies in Portugal and Italy. Sensors (MDPI AG). 2021;21:3982 [Internet]. DOI: 10.3390/s21123982
  272. 272. Aicardi I, Garbarino M, Lingua A, Lingua E, Marzano R, Piras M. Monitoring post-fire forest recovery using multi-temporal digital surface models generated from different platforms. EARSeL eProceedings. 2016;15(1):1-8
  273. 273. Talucci AC, Forbath E, Kropp H, Alexander HD, DeMarco J, Paulson AK, et al. Evaluating post-fire vegetation recovery in Cajander larch forests in Northeastern Siberia using UAV derived vegetation indices. Remote Sensing. 2020;12:2970 [Internet] [cited 2023 Jan 14]. Available from: https://www.mdpi.com/2072-4292/12/18/2970/htm
  274. 274. Larrinaga AR, Brotons L. Greenness indices from a low-cost UAV imagery as tools for monitoring post-fire Forest recovery. Drones. 2019;3:6 [Internet] [cited 2023 Jan 14]. Available from: https://www.mdpi.com/2504-446X/3/1/6/htm
  275. 275. Jain P, Coogan SCP, Subramanian SG, Crowley M, Taylor S, Flannigan MD. A review of machine learning applications in wildfire science and management. Environmental Reviews. 2020;28:478-505
  276. 276. Chen J, Zheng W, Wu S, Liu C, Yan H. Fire monitoring algorithm and its application on the geo-Kompsat-2A geostationary meteorological satellite. Remote Sensing. 2022;14:2655 [Internet] [cited 2023 Jan 16]. Available from: https://www.mdpi.com/2072-4292/14/11/2655/htm
  277. 277. Bailon-Ruiz R, Bit-Monnot A, Lacroix S. Real-time wildfire monitoring with a fleet of UAVs. Robotics and Autonomous Systems. 2022;152:104071. DOI: 10.1016/J.ROBOT.2022.104071
  278. 278. Roteta E, Bastarrika A, Padilla M, Storm T, Chuvieco E. Development of a Sentinel-2 burned area algorithm: Generation of a small fire database for sub-Saharan Africa. Remote Sensing of Environment. 2019;222:1-17
  279. 279. Crowley MA, Cardille JA. Remote Sensing’s recent and future contributions to landscape ecology. Current Landscape Ecology Reports. 2020;5:45-57 [Internet] [cited 2023 Jan 17]. Available from: https://link.springer.com/article/10.1007/s40823-020-00054-9
  280. 280. Chen Y-N, Fan K-C, Chang Y-L, Moriyama T. Special issue review: Artificial intelligence and machine learning applications in remote sensing. Remote Sensing (MDPI AG). 2023;15:569 [Internet]. DOI: 10.3390/rs15030569
  281. 281. Koubaa A, Ammar A, Abdelkader M, Alhabashi Y, Ghouti L. AERO: AI-enabled remote sensing observation with onboard edge computing in UAVs. Remote Sensing (MDPI AG). 2023;15:1873 [Internet]. DOI: 10.3390/rs15071873
  282. 282. Zhang L, Zhang L. Artificial intelligence for remote sensing data analysis: A review of challenges and opportunities. IEEE Geoscience and Remote Sensing Magazine. 2022;10:270-294

Written By

Georgios Zagalikis

Submitted: 22 January 2023 Reviewed: 06 February 2023 Published: 24 October 2023